Stochastic integrals and Brownian motion on abstract nilpotent Lie groups
Abstract.
We construct a class of iterated stochastic integrals with respect to Brownian motion on an abstract Wiener space which allows for the definition of Brownian motions on a general class of infinite-dimensional nilpotent Lie groups based on abstract Wiener spaces. We then prove that a Cameron–Martin type quasi-invariance result holds for the associated heat kernel measures in the non-degenerate case, and give estimates on the associated Radon–Nikodym derivative. We also prove that a log Sobolev estimate holds in this setting.
Key words and phrases:
Heat kernel measure, infinite-dimensional Lie group, quasi-invariance, logarithmic Sobolev inequality2010 Mathematics Subject Classification:
Primary 60J65 28D05; Secondary 58J65 22E651. Introduction
The construction of diffusions on infinite-dimensional manifolds and the study of the regularity properties of their induced measures have been a topic of great interest for at least the past 50 years; see for example [13, 26, 27, 19, 20, 3, 2, 30], although many other references exist. The purpose of the present paper is to construct diffusions on a general class of infinite-dimensional nilpotent Lie groups, and to show that the associated heat kernel measures are quasi-invariant under appropriate translations. We demonstrate that this class of groups is quite rich. We focus here on the elliptic setting, but comment that, as nilpotent groups are standard first models for studying hypoellipticity, examples of infinite-dimensional versions of such spaces are important for the more general study of hypoellipticity in infinite dimensions. This is an area of great interest, and is of particular relevance in the study of stochastic PDEs and their applications [1, 5, 8, 22, 29, 31]. The present paper studies heat kernel measures for elliptic diffusions on these spaces, which is a necessary precursor to understanding the degenerate case.
1.1. Main results
Let denote an abstract Wiener space, where is a Banach space equipped with a centered non-degenerate Gaussian measure and associated Cameron–Martin Hilbert space . We will assume that additionally carries a nilpotent Lie algebra structure , and we will further assume that this Lie bracket is Hilbert-Schmidt. We will call such a space an abstract nilpotent Lie algebra. Via the standard Baker-Campbell-Hausdorff-Dynkin formula, we may then equip with an explicit group operation under which becomes an infinite-dimensional group. When thought of as a group, we will denote this space by . We equip with the left-invariant Riemannian metric which agrees with the inner product on , and we denote the Riemannian distance by . Note that, despite the use of the notation , we do not assume that the Lie bracket structure extends to , and so this space is not necessarily a Lie algebra or a Lie group. Still, when is playing a role typically played by the group, we will denote by .
If were a Lie algebra (and thus carried an associated group structure), we could construct a Brownian motion on as the solution to the Stratonovich stochastic differential equation
where is left translation by and is a standard Brownian motion on (as a Banach space) with . In finite dimensions, the solution to this stochastic differential equation may be obtained explicitly as a formula involving the Lie bracket. In particular, for and , let denote the simplex in given by
Let denote the permutation group on , and, for each , let denote the number of “errors” in the ordering , that is, . Then the Brownian motion on could be written as
(1.1) |
where this sum is finite under the assumed nilpotence. An obstacle to the development of a general theory of stochastic differential equations on infinite-dimensional Banach spaces is the lack of smoothness of the norm in a general Banach space which is necessary to define a stochastic integral on it. Still, in Section 2, we prove a general result to define a class of iterated stochastic integrals with respect to Brownian motion on the Banach space that includes the expression above. Additionally, we show that one may make sense of the above expression when the Lie bracket on does not necessarily extend to . Thus we are able to define a “group Brownian motion” on via (1.1). We let be the heat kernel measure on .
In particular, the integrals above are defined as a limit of stochastic integrals on finite-dimensional subgroups of . We show that these are nice in the sense that they approximate and that there exists a uniform lower bound on their Ricci curvatures.
Using these results, we are able to prove the following main theorem.
Theorem 1.1.
For , let denote left and right translation by , respectively. Then and define measurable transformations on , and for all , and are absolutely continuous with respect to . Let
be the Radon-Nikodym derivatives, be the uniform lower bound on the Ricci curvatures of the finite-dimensional approximation groups and
with the convention that . Then, for all , and both satisfy the estimate
where or .
1.2. Discussion
The present paper builds on the previous work in [14] and [32], significantly generalizing these previous works in several ways. In particular, the paper [32] considered analogous results for “semi-infinite Lie groups”, which are infinite-dimensional nilpotent Lie groups constructed as extensions of finite-dimensional nilpotent Lie groups by an infinite-dimensional abstract Wiener space (see Example 3.3). At several points in the analysis there, the fact that was used in a critical way. In particular, it was used to show that the stochastic integrals defining the Brownian motion on as in equation (1.1) were well-defined. In the present paper, we have removed this restriction, as well as removing the “stratified” structure implicit in the construction as a Lie group extension.
Again, we note that, despite the use of the notation , it is not assumed that the Lie bracket structure on extends to , and so itself is not necessarily a Lie algebra or Lie group. In [14] and [32], it was assumed that the Lie bracket was a continuous map defined on . However, it turns out that the group construction on is the only necessary structure for the subsequent analysis. As is usual for the infinite-dimensional setting, while the heat kernel measure is itself supported on the larger space , its critical analysis depends more on the structure of . Still, as was originally done in [14] and then in [32], one may instead define an abstract nilpotent Lie algebra starting with a continuous nilpotent bracket . For example, in the event that where is a finite-dimensional Lie algebra and , it is well-known that this implies that the restriction of the bracket to is Hilbert-Schmidt. (For any continuous bilinear where is a Hilbert space, one has that ; this follows for example from Corollary 4.4 of [28].) More generally, in order for the subsequent theory to make sense, one would naturally need to require that be a Lie subalgebra of , that is, for the restriction of the Lie bracket to to preserve . As the proofs in the sequel rely strongly on the bracket being Hilbert-Schmidt, it would be then necessary to add the Hilbert-Schmidt hypothesis as it does not follow immediately if one only assumes a continuous bracket on which preserves .
Additionally, the spaces studied in the present paper are well-designed for the study of infinite-dimensional hypoelliptic heat kernel measures, and there has already been progress on proving quasi-invariance and stronger smoothness properties for these measures in the simplest case of a step two Lie algebra with finite-dimensional center; see [7] and [15]. More generally, the paper [35] explores related interesting lines of inquiry for heat kernel measures on infinite-dimensional groups, largely in the context of groups of maps from manifolds to Lie groups.
Acknowledgements. The author thanks Bruce Driver and Nathaniel Eldredge for helpful conversations during the writing of this paper.
2. Iterated Itô integrals
Recall the standard construction of abstract Wiener spaces. Suppose that is a real separable Banach space and is the Borel -algebra on .
Definition 2.1.
A measure on is called a (mean zero, non-degenerate) Gaussian measure provided that its characteristic functional is given by
for a symmetric, positive definite quadratic form. That is, is a real inner product on .
Theorem 2.2.
Let be a Gaussian measure on . For , let
and define the Cameron–Martin subspace by
Then is a dense subspace of , and there exists a unique inner product on such that for all , and is a separable Hilbert space with respect to this inner product. For any , for some .
Alternatively, given a real separable Banach space and a real separable Hilbert space continuously embedded in as a dense subspace, then for each there exists a unique such that for all . Then is continuous, linear, and one-to-one with a dense range
(2.1) |
and is continuous. A Gaussian measure on is a Borel probability measure such that, for each , the random variable under is a centered Gaussian with variance .
Suppose that is a finite rank orthogonal projection such that . Let be an orthonormal basis for . Then we may extend to a (unique) continuous operator from (still denoted by ) by letting
(2.2) |
for all .
Notation 2.3.
Let denote the collection of finite rank projections on such that and is an orthogonal projection, that is, has the form given in equation (2.2).
Let be a Brownian motion on with variance determined by
for all and , where is as in (2.1). Note that for any , is a Brownian motion on . In the rest of this section, we will verify the existence of martingales defined as certain iterated stochastic integrals with respect to .
The following is Proposition 4.1 of [32]. Note that again this was stated in the context where was a “semi-infinite Lie algebra”, but a brief inspection of the proof shows that this is a general statement about stochastic integrals on Hilbert spaces.
Proposition 2.4.
Let such that . Then, for a continuous mapping, let
Then is a continuous -martingale, and there exists a continuous -martingale such that
(2.3) |
and
(2.4) |
for all . The process is well-defined independent of the choice of increasing orthogonal projections into , and so will be denoted by
Now we may use this result to define stochastic integrals taking values in another Hilbert space .
Proposition 2.5.
Let be a Hilbert space and be a continuous map. That is, is a map continuous in and linear on such that
Then
is a continuous -valued -martingale, and there exists a continuous -valued -martingale such that
(2.5) |
for all . The martingale is well-defined independent of the choice of orthogonal projections, and thus will be denoted by
Proof.
Let be an orthonormal basis of . Since is linear on , for each there exists such that
(2.6) |
If is defined by equation (2.6), then clearly and in particular
Thus, for as defined in Proposition 2.4,
and so we may write
Thus, taking , we also have that
as by (2.3) and dominated convergence since
by (2.4). Then equation (2.5) holds by Doob’s maximal inequality. ∎
Note that the preceding results then imply that one may define the above stochastic integrals with respect to any increasing sequence of orthogonal projections – that is, we need not require that the projections extend continuously to .
Proposition 2.6.
Let be an arbitrary finite-dimensional subspace of , and let denote orthogonal projection onto . Then for any Hilbert space and a continuous map, the stochastic integral
is well-defined, and is a continuous -valued -martingale. Moreover, if is an increasing sequence of finite-dimensional subspaces of such that the corresponding orthogonal projections , then
where is as defined in Proposition 2.5.
3. Abstract nilpotent Lie algebras and groups
Definition 3.1.
Let be an abstract Wiener space such that is equipped with a nilpotent Hilbert-Schmidt Lie bracket. Then we will call an abstract nilpotent Lie algebra.
The Baker-Campbell-Hausdorff-Dynkin formula implies that
for all , where
, and for each multi-index ,
see, for example, [18]. If is nilpotent of step , then
for . Since is simply connected and nilpotent, the exponential map is a global diffeomorphism (see, for example, Theorems 3.6.2 of [38] or 1.2.1 of [12]). In particular, we may view as both a Lie algebra and Lie group, and one may verify that
(3.1) |
defines a group structure on . Note that and the identity .
Definition 3.2.
When we wish to emphasize the group structure on , we will denote by .
Lemma 3.3.
The Banach space topology on makes into a topological group.
Proof.
Since is a topological vector space, and are continuous by definition. The map is continuous in the topology by the boundedness of the Lie bracket. It then follows from (3.1) that is continuous as well. ∎
3.1. Measurable group actions on
As discussed in the introduction, given a Hilbert-Schmidt Lie bracket on and a subsequently defined group operation on , one may define a measurable action on by left or right multiplication by an element of .
In particular, let be an orthonormal basis of . For now, fix and consider the mapping given by . Then this is a continuous linear map on and in the usual way we may make the identification of so that we define the operator given by
in particular, . Note that, for any
and thus . Now fix . Then for we may write
Since each has a measurable linear extension to such that (see, for example, Theorem 2.10.11 of [10]), we may extend to a measurable linear transformation from to (still denoted by ) given by
Note that here we are using the fact that
which implies that
Similarly, we may define
In a similar way, note that we may write, for and ,
and thus for and
More generally for
Thus, for and , we may define a measurable action on given by
(For , we define this mapping to be 0, which is certainly measurable.) Again, we are using that
This holds by straightforward but tedious computations — in fact, iterative applications of Cauchy-Schwarz combined with the fact that . Thus,
-a.s. and as given above is defined a.s. Thus, we have the following result.
Proposition 3.4.
For , the mapping defined analogously to (3.1) is a measurable right group action by on , and similarly for the left action .
3.2. Examples of abstract nilpotent Lie algebras
Example 3.1 (Free nilpotent Lie algebras).
Starting with an abstract Wiener space , one may construct in the standard way the abstract free nilpotent Lie algebra of step with generators an orthonormal basis of . See for example Section 0 of [21].
Example 3.2 (Heisenberg-like algebras).
Let for be abstract Wiener spaces. Then for any a Hilbert-Schmidt map, we may define a Lie bracket on by
and may be thought of as an abstract Heisenberg-like algebra as in [14].
These abstract Heisenberg-like algebras are central extensions of one abstract Wiener space by another abstract Wiener space. The next example generalizes this construction.
Example 3.3 (Extensions of Lie algebras).
Let and be Lie algebras, and let denote the set of derivations on ; that is, consists of all linear maps satisfying Leibniz’s rule:
Now suppose there are a linear mapping and a skew-symmetric bilinear mapping , satisfying, for all ,
(B1) |
and
(B2) |
Then, one may verify that, for ,
defines a Lie bracket on , and we say is an extension of over . That is, is the Lie algebra with ideal and quotient algebra . The associated exact sequence is
where is inclusion and is projection. In fact, these are the only extensions of over (see, for example, [4]).
Now suppose that is a real abstract Wiener space, and is an abstract nilpotent Lie algebra. Motivated by the previous discussion, we may consider as an abelian Lie algebra and construct extensions of over . In this case, we need a linear mapping and a skew-symmetric bilinear mapping , such that and are both Hilbert-Schmidt and together and satisfy (B1) and (B2), which in this setting become
and
for all . Then we may define a Lie algebra structure on via the Lie bracket
Again, these are in fact the only extensions of over .
Combining Examples 3.2 and 3.3 shows that these constructions are iterative, in that one may construct new abstract nilpotent Lie algebras as Lie algebra extensions of another abstract nilpotent Lie algebra. The next example builds on the previous one to give one precise way to construct some Lie algebra extensions.
Example 3.4.
Let be any Hilbert-Schmidt map, and for and define
Then the Lie bracket on is given by
Note that, if is nilpotent of step , then will automatically be nilpotent of step .
The examples above demonstrate that the space of abstract nilpotent Lie algebras is quite rich, and there are many natural examples with a straightforward construction. This significantly improves the results of [32], which studied heat kernel measures on nilpotent extensions of abstract Wiener spaces over finite-dimensional nilpotent Lie algebras. For example, the restriction trivializes Example 3.4. We elaborate in the following remark.
Remark 3.5.
Returning to Example 3.4, since is linear and continuous, we have the decomposition
where . Thus, for we may write and
and is a map on . Thus, and similarly . So
and, in particular, when , is just an extension of the finite-dimensional vector space by the finite-dimensional Lie algebra , and the construction is not truly infinite-dimensional.
3.3. Properties of
This section collects some results for topological and geometric properties of that we’ll require for the sequel.
Proposition 3.6.
For all , is Hilbert-Schmidt.
Proof.
For , this follows from the definition of . Now assume the statement holds for all , and consider . Writing in terms of the orthonormal basis and using Hölder’s inequality gives
and the last line is finite by the induction hypothesis. ∎
Next we will recall that the flat and geometric topologies on are equivalent. First we set the following notation.
Notation 3.7.
For let and denote left and right multiplication by , respectively. As is a vector space, to each we can associate the tangent space to at , which is naturally isomorphic to . For a Frechét smooth function and and , let
and let denote the tangent vector satisfying . If is any smooth curve in such that and (for example, ), then
Notation 3.8.
Let denote the collection of -paths . The length of is defined as
The Riemannian distance between is then defined as
The following proposition was proved as Corollary 4.13 in [32]. This was proved under the conditions that was a “semi-infinite Lie algebra”, that is, under the assumption that was a nilpotent Lie algebra extension (as in Example 3.3) of a finite-dimensional nilpotent Lie algebra by an abstract Wiener space. However, a cursory inspection of the proofs there will show that they only depended on the fact that the Lie bracket was Hilbert-Schmidt, and not on the “stratified” structure of or the fact that the image of the Lie bracket was a finite-dimensional subspace.
Proposition 3.9.
The topology on induced by is equivalent to the Hilbert topology induced by .
We may find a uniform lower bound on the Ricci curvature of all finite-dimensional subgroups of . Finite-dimensional subgroups of may be obtained by taking the Lie algebra generated by any finite-dimensional subspace of (which is again necessarily finite-dimensional by the nilpotence of the bracket) and endowing it with the standard group operation via the Baker-Campbell-Hausdorff-Dynkin formula. An analogue of the following proposition was proved as Proposition 3.23 and Corollary 3.24 in [32]. It follows directly from the form of the Ricci curvature on nilpotent groups (endowed with a left invariant metric) and the assumption that the Lie bracket is Hilbert-Schmidt. This proof is essentially the same as in [32], but it is quite brief and so is included for completeness.
Proposition 3.10.
Let
Then and is the largest constant such that
holds uniformly for all finite-dimensional Lie subalgebras of .
Proof.
For any nilpotent Lie algebra with orthonormal basis ,
for all . Thus, for any finite-dimensional Lie algebra
where
(3.2) |
Taking the infimum of over all completes the proof. ∎
4. Brownian motion on
Suppose that is a smooth curve in with , and consider the differential equation
The solution may be written as follows (see [36]): For , let denote the simplex in given by
Let denote the permutation group on , and, for each , let denote the number of “errors” in the ordering , that is, . Then
(4.1) |
where the term is understood to be . Using this as our motivation, we first explore stochastic integral analogues of equation (4.1) where the smooth curve is replaced by Brownian motion on .
4.1. Brownian motion and finite-dimensional approximations
We now return to the setting of an abstract Wiener space endowed with a nilpotent Hilbert-Schmidt Lie bracket on . Again, let denote Brownian motion on . By equation (4.1), the solution to the Stratonovich stochastic differential equation
should be given by
(4.2) |
for coefficients determined by equation (4.1).
To understand the integrals in (4.2), consider the following heuristic computation. Let denote the process in defined by
By repeatedly applying the definition of the Stratonovich integral, the iterated Stratonovich integral may be realized as a linear combination of iterated Itô integrals:
where
and, for , is the iterated Itô integral
with
compare with Proposition 1 of [9]. This change from multiple Stratonovich integrals to multiple Itô integrals may also be recognized as a specific case of the Hu-Meyer formulas [23, 24], but we will compute more explicitly to verify that our integrals are well-defined.
Define by , and for define by
(4.3) |
For each fixed and , define by
(4.4) |
Then we may write
presuming we can make sense of the integrals .
For each , let and (so that when ), and let
Then, for each and ,
where and are as follows.
The map is defined by
(4.5) |
for an orthonormal basis of and given by , for any such that
To define , first consider the polynomial of order , in the variables with such that and in the variable , given by evaluating the integral
(4.6) |
where with and . Then is with the variables reindexed by the bijection that maintains the natural ordering of these sets. (For example, for ,
so that .)
This explicit realization of is not critical to the sequel. It is really only necessary to know that is a polynomial of order in and , and thus may be written as
for some coefficients and polynomials of degree in . Now, if is Hilbert-Schmidt on , then
and
(4.7) |
may be understood in the sense of the limit integrals in Proposition 2.5. (In particular, if , then does not depend on , and Proposition 2.5 implies that is a -valued -martingale.)
The above computations show that, if for all , , and , is Hilbert-Schmidt, then we may rewrite (4.2) as
where is as defined in Proposition 2.5. The next proposition shows that is Hilbert-Schmidt as desired, and thus in (4.2) is well-defined.
Proposition 4.1.
Let , , and . Then is Hilbert-Schmidt.
Proof.
For the whole of this proof, all sums will be taken over an orthonormal basis of .
Now, if , let denote the second such that ; that is,
where exactly one of is and, for all , for some
Thus, writing
we have that
and so
Now more generally when , we may similarly “separate” the pairs of ’s as above. More precisely, define
Let , and set
Similarly, we define
Then there is an such that the sets and separate the ’s in the sense that no is repeated inside any one of these sets, and moreover the union of “even” sets contains exactly one copy of each and similarly with the “odd” sets. We can write
If is even, then
is a function of and for some subset of , and
is a function of and for . Thus,
which is finite again by Proposition 3.6. Similarly, if is odd,
and
are both functions of and some , and
is again finite by Proposition 3.6 and this completes the proof. ∎
Definition 4.2.
Proposition 4.3 (Finite-dimensional approximations).
For a finite-dimensional Lie subgroup of , let denote orthogonal projection of onto and let be the continuous process on defined by
where the stochastic integrals are defined as in Proposition 2.6. Then is Brownian motion on . In particular, for an increasing sequence of finite-dimensional Lie subgroups such that the associated orthogonal projections are increasing to , let . Then, for all ,
(4.8) |
Proof.
First note that solves the Stratonovich equation with , see [9, 11, 6] where is a standard -valued Brownian motion. Thus, is a -valued Brownian motion.
By equation (4.7) and its preceding discussion,
and thus, to verify (4.8), it suffices to show that
and
for all , and .
So let . Then it is known that, if is a finite-dimensional subspace of and is the orthogonal projection from to , then admits a -a.s. unique extension to . Moreover, if is an increasing sequence of finite-dimensional subspaces, then
see for example Section 8.3.3 of [37].
Remark 4.4.
In fact, for each of the stochastic integrals , it is possible to prove the stronger convergence that, for all ,
for all , and . Again, Proposition 2.6 gives the limit for and thus for . For , Doob’s maximal inequality implies it suffices to show that
Since each and has chaos expansion terminating at degree , a theorem of Nelson (see Lemma 2 of [34] and pp. 216-217 of [33]) implies that, for each , there exists such that
In a similar way, one may prove the following convergence for the Brownian motions under right translations by elements of .
Proposition 4.5.
Remark 4.6.
Note that, while the present paper focuses on the case where is non-degenerate and is Brownian motion on , the above construction and finite-dimensional approximations would all follow with essentially no modification if one considered instead a Gaussian measure whose support was, for example, a subspace of such that generates the span of via the Lie bracket.
4.2. Quasi-invariance and log Sobolev
We are now able to prove Theorem 1.1, which states that the heat kernel measure is quasi-invariant under left and right translation by elements of and gives estimates for the Radon-Nikodym derivatives of the “translated” measures. Given the results so far, the proof could be given as an application of Theorem 7.3 and Corollary 7.4 of [16]. However, we provide here a full proof for the reader’s convenience.
Proof of Theorem 1.1. Fix and an orthogonal projection onto a finite-dimensional subspace of . Let , and be an increasing sequence of projections such that for all and . Let denote the Radon-Nikodym derivative of with respect to . Then for each and for any , we have the following integrated Harnack inequality
where is the uniform lower bound on the Ricci curvature as in Proposition 3.10 and is Riemannian distance on ; see for example Theorem 1.6 of [16].
By Proposition 4.5, we have that for any , the class of bounded continuous functions on
(4.9) |
where denotes the inclusion map. Note that for any
where is the conjugate exponent to . Allowing in this last inequality yields
(4.10) |
by equation (4.9) and the fact that the length of a path in can be approximated by the lengths of paths in the finite-dimensional projections. That is, for any and with , there exists an increasing sequence of orthogonal projections such that , , and
To see this, let be a path in . Then one may show that
for appropriate coefficients ; see for example Section 3 of [32]. Thus, we have proved that (4.10) holds for and . As this union is dense in by Proposition 3.9, dominated convergence along with the continuity of in implies that (4.10) holds for all .
Since the bounded continuous functions are dense in (see for example Theorem A.1 of [25]), the inequality in (4.10) implies that the linear functional defined by
has a unique extension to an element, still denoted by , of which satisfies the bound
for all . Since , there then exists a function such that
(4.11) |
for all , and
Now restricting (4.11) to , we may rewrite this equation as
(4.12) |
Then a monotone class argument (again use Theorem A.1 of [25]) shows that (4.12) is valid for all bounded measurable functions on . Thus, exists and is given by , which is in for all and satisfies the desired bound.
A parallel argument gives the analogous result for . Alternatively, one could use the right translation invariance just proved along with the facts that inherits invariance under the inversion map from its finite-dimensional projections and that .
The following also records the straightforward fact that the heat kernel measure does not charge .
Proposition 4.7.
For all , .
Proof.
This follows trivially from the fact that is the sum of a Brownian motion on with a finite sequence of stochastic integrals taking values in . ∎
Thus, maintains its role as a dense subspace of of measure 0 with respect to the distribution of the “group Brownian motion”.
Definition 4.8.
A function is said to be a (smooth) cylinder function if for some finite-dimensional projection and some (smooth) function . Also, is a cylinder polynomial if for a polynomial function on .
Theorem 4.9.
Given a cylinder polynomial on , let be the gradient of , the unique element of such that
for all . Then for as in Proposition 3.10,
Proof.
Following the method of Bakry and Ledoux applied to (see Theorem 2.9 of [17] for the case needed here) shows that
for as in equation (3.2). Since the function is decreasing and for all finite-dimensional projections , this estimate also holds with replaced with . Now applying Proposition 4.3 to pass to the limit as gives the desired result. ∎
References
- [1] Hélène Airault and Paul Malliavin, Backward regularity for some infinite dimensional hypoelliptic semi-groups, Stochastic analysis and related topics in Kyoto, Adv. Stud. Pure Math., vol. 41, Math. Soc. Japan, Tokyo, 2004, pp. 1–11. MR 2083700 (2005h:58062)
- [2] Hélène Airault and Paul Malliavin, Quasi-invariance of Brownian measures on the group of circle homeomorphisms and infinite-dimensional Riemannian geometry, J. Funct. Anal. 241 (2006), no. 1, 99–142. MR 2264248
- [3] Sergio Albeverio, Alexei Daletskii, and Yuri Kondratiev, Stochastic equations and Dirichlet operators on infinite product manifolds, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 6 (2003), no. 3, 455–488. MR 2016321
- [4] Dmitri Alekseevsky, Peter W. Michor, and Wolfgang A.F. Ruppert, Extensions of Lie algebras, 2000.
- [5] Yuri Bakhtin and Jonathan C. Mattingly, Malliavin calculus for infinite-dimensional systems with additive noise, J. Funct. Anal. 249 (2007), no. 2, 307–353. MR 2345335 (2008g:60173)
- [6] Fabrice Baudoin, An introduction to the geometry of stochastic flows, Imperial College Press, London, 2004. MR MR2154760 (2006f:60003)
- [7] Fabrice Baudoin, Maria Gordina, and Tai Melcher, Quasi-invariance for heat kernel measures on sub-Riemannian infinite-dimensional Heisenberg groups, Trans. Amer. Math. Soc. 365 (2013), no. 8, 4313–4350. MR 3055697
- [8] Fabrice Baudoin and Josef Teichmann, Hypoellipticity in infinite dimensions and an application in interest rate theory, Ann. Appl. Probab. 15 (2005), no. 3, 1765–1777. MR 2152244 (2006g:60080)
- [9] Gérard Ben Arous, Flots et séries de Taylor stochastiques, Probab. Theory Related Fields 81 (1989), no. 1, 29–77. MR MR981567 (90a:60106)
- [10] Vladimir I. Bogachev, Gaussian measures, Mathematical Surveys and Monographs, vol. 62, American Mathematical Society, Providence, RI, 1998. MR MR1642391 (2000a:60004)
- [11] Fabienne Castell, Asymptotic expansion of stochastic flows, Probab. Theory Related Fields 96 (1993), no. 2, 225–239. MR MR1227033 (94g:60110)
- [12] Lawrence J. Corwin and Frederick P. Greenleaf, Representations of nilpotent Lie groups and their applications. Part I, Cambridge Studies in Advanced Mathematics, vol. 18, Cambridge University Press, Cambridge, 1990, Basic theory and examples. MR MR1070979 (92b:22007)
- [13] Ju. L. Daleckiĭ and Ja. I. Šnaĭderman, Diffusion and quasiinvariant measures on infinite-dimensional Lie groups, Funkcional. Anal. i Priložen. 3 (1969), no. 2, 88–90. MR 0248888
- [14] B. Driver and M. Gordina, Heat kernel analysis on infinite-dimensional Heisenberg groups, J. Funct. Anal. 255 (2008), no. 2, 2395–2461.
- [15] Bruce K. Driver, Nathaniel Eldredge, and Tai Melcher, Hypoelliptic heat kernels on infinite-dimensional Heisenberg groups, Trans. Amer. Math. Soc. 368 (2016), no. 2, 989–1022. MR 3430356
- [16] Bruce K. Driver and Maria Gordina, Integrated Harnack inequalities on Lie groups, J. Differential Geom. 83 (2009), no. 3, 501–550. MR MR2581356
- [17] Bruce K. Driver and Terry Lohrenz, Logarithmic Sobolev inequalities for pinned loop groups, J. Funct. Anal. 140 (1996), no. 2, 381–448. MR MR1409043 (97h:58176)
- [18] J. J. Duistermaat and J. A. C. Kolk, Lie groups, Universitext, Springer-Verlag, Berlin, 2000. MR MR1738431 (2001j:22008)
- [19] K. D. Elworthy, Measures on infinite-dimensional manifolds, Functional integration and its applications (Proc. Internat. Conf., London, 1974), 1975, pp. 60–68. MR 0501086
- [20] Denis Feyel and Arnaud de La Pradelle, Brownian processes in infinite dimension, Potential Anal. 4 (1995), no. 2, 173–183. MR 1323825
- [21] Mikhael Gromov, Carnot-Carathéodory spaces seen from within, Sub-Riemannian geometry, Progr. Math., vol. 144, Birkhäuser, Basel, 1996, pp. 79–323. MR 1421823 (2000f:53034)
- [22] Martin Hairer and Jonathan C. Mattingly, A theory of hypoellipticity and unique ergodicity for semilinear stochastic PDEs, Electron. J. Probab. 16 (2011), no. 23, 658–738. MR 2786645 (2012e:60170)
- [23] Y. Z. Hu and P.-A. Meyer, Chaos de Wiener et intégrale de Feynman, Séminaire de Probabilités, XXII, Lecture Notes in Math., vol. 1321, Springer, Berlin, 1988, pp. 51–71. MR MR960508 (89m:60193)
- [24] by same author, Sur les intégrales multiples de Stratonovitch, Séminaire de Probabilités, XXII, Lecture Notes in Math., vol. 1321, Springer, Berlin, 1988, pp. 72–81. MR MR960509 (89k:60070)
- [25] Svante Janson, Gaussian Hilbert spaces, Cambridge Tracts in Mathematics, vol. 129, Cambridge University Press, Cambridge, 1997. MR 1474726 (99f:60082)
- [26] Hui Hsiung Kuo, Integration theory on infinite-dimensional manifolds, Trans. Amer. Math. Soc. 159 (1971), 57–78. MR 295393
- [27] by same author, Diffusion and Brownian motion on infinite-dimensional manifolds, Trans. Amer. Math. Soc. 169 (1972), 439–459. MR 309206
- [28] by same author, Gaussian measures in Banach spaces, Lecture Notes in Mathematics, Vol. 463, Springer-Verlag, Berlin, 1975. MR MR0461643 (57 #1628)
- [29] Paul Malliavin, Hypoellipticity in infinite dimensions, Diffusion processes and related problems in analysis, Vol. I (Evanston, IL, 1989), Progr. Probab., vol. 22, Birkhäuser Boston, Boston, MA, 1990, pp. 17–31. MR 1110154 (93b:60132)
- [30] by same author, Invariant or quasi-invariant probability measures for infinite dimensional groups: II. Unitarizing measures or Berezinian measures, Jpn. J. Math. 3 (2008), no. 1, 19–47. MR MR2390182 (2009b:60170)
- [31] Jonathan C. Mattingly and Étienne Pardoux, Malliavin calculus for the stochastic 2D Navier-Stokes equation, Comm. Pure Appl. Math. 59 (2006), no. 12, 1742–1790. MR 2257860 (2007j:60082)
- [32] Tai Melcher, Heat kernel analysis on semi-infinite Lie groups, Journal of Functional Analysis 257 (2009), no. 11, 3552–3592. MR MR2572261
- [33] Edward Nelson, The free Markoff field, J. Functional Analysis 12 (1973), 211–227. MR MR0343816 (49 #8556)
- [34] by same author, Quantum fields and Markoff fields, Partial differential equations (Proc. Sympos. Pure Math., Vol. XXIII, Univ. California, Berkeley, Calif., 1971), Amer. Math. Soc., Providence, R.I., 1973, pp. 413–420. MR MR0337206 (49 #1978)
- [35] Doug Pickrell, Heat kernel measures and critical limits, Developments and trends in infinite-dimensional Lie theory, Progr. Math., vol. 288, Birkhäuser Boston, Inc., Boston, MA, 2011, pp. 393–415. MR 2743770
- [36] Robert S. Strichartz, The Campbell-Baker-Hausdorff-Dynkin formula and solutions of differential equations, J. Funct. Anal. 72 (1987), no. 2, 320–345. MR MR886816 (89b:22011)
- [37] Daniel W. Stroock, Probability theory, second ed., Cambridge University Press, Cambridge, 2011, An analytic view.
- [38] V. S. Varadarajan, Lie groups, Lie algebras, and their representations, Graduate Texts in Mathematics, vol. 102, Springer-Verlag, New York, 1984, Reprint of the 1974 edition. MR 85e:22001