This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Stochastic integrals and Brownian motion on abstract nilpotent Lie groups

Tai Melcher Department of Mathematics
University of Virginia
Charlottesville, VA 22903 USA
[email protected]
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 inequality
2010 Mathematics Subject Classification:
Primary 60J65 28D05; Secondary 58J65 22E65
11footnotemark: 1 This research was supported in part by NSF Grants DMS-0907293 and DMS-1255574.

1. 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 (𝔤,𝔤CM,μ)(\mathfrak{g},\mathfrak{g}_{CM},\mu) denote an abstract Wiener space, where 𝔤\mathfrak{g} is a Banach space equipped with a centered non-degenerate Gaussian measure μ\mu and associated Cameron–Martin Hilbert space 𝔤CM\mathfrak{g}_{CM}. We will assume that 𝔤CM\mathfrak{g}_{CM} additionally carries a nilpotent Lie algebra structure [,]:𝔤CM×𝔤CM𝔤CM[\cdot,\cdot]:\mathfrak{g}_{CM}\times\mathfrak{g}_{CM}\rightarrow\mathfrak{g}_{CM}, 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 𝔤CM\mathfrak{g}_{CM} with an explicit group operation under which 𝔤CM\mathfrak{g}_{CM} becomes an infinite-dimensional group. When thought of as a group, we will denote this space by GCMG_{CM}. We equip GCMG_{CM} with the left-invariant Riemannian metric which agrees with the inner product on 𝔤CMTeGCM\mathfrak{g}_{CM}\cong T_{e}G_{CM}, and we denote the Riemannian distance by dd. Note that, despite the use of the notation 𝔤\mathfrak{g}, we do not assume that the Lie bracket structure extends to 𝔤\mathfrak{g}, and so this space is not necessarily a Lie algebra or a Lie group. Still, when 𝔤\mathfrak{g} is playing a role typically played by the group, we will denote 𝔤\mathfrak{g} by GG.

If 𝔤\mathfrak{g} were a Lie algebra (and thus carried an associated group structure), we could construct a Brownian motion on GG as the solution to the Stratonovich stochastic differential equation

δgt=gtδBt:=LgtδBt, with g0=𝐞=(0,0),\delta g_{t}=g_{t}\delta B_{t}:=L_{g_{t}*}\delta B_{t},\text{ with }g_{0}=\mathbf{e}=(0,0),

where LxL_{x} is left translation by xGx\in G and {Bt}t0\{B_{t}\}_{t\geq 0} is a standard Brownian motion on 𝔤\mathfrak{g} (as a Banach space) with Law(B1)=μ\mathrm{Law}(B_{1})=\mu. In finite dimensions, the solution to this stochastic differential equation may be obtained explicitly as a formula involving the Lie bracket. In particular, for t>0t>0 and nn\in\mathbb{N}, let Δn(t)\Delta_{n}(t) denote the simplex in n\mathbb{R}^{n} given by

{s=(s1,,sn)n:0<s1<s2<<sn<t}.\{s=(s_{1},\cdots,s_{n})\in\mathbb{R}^{n}:0<s_{1}<s_{2}<\cdots<s_{n}<t\}.

Let 𝒮n\mathcal{S}_{n} denote the permutation group on (1,,n)(1,\cdots,n), and, for each σ𝒮n\sigma\in\mathcal{S}_{n}, let e(σ)e(\sigma) denote the number of “errors” in the ordering (σ(1),σ(2),,σ(n))(\sigma(1),\sigma(2),\cdots,\sigma(n)), that is, e(σ)=#{j<n:σ(j)>σ(j+1)}e(\sigma)=\#\{j<n:\sigma(j)>\sigma(j+1)\}. Then the Brownian motion on GG could be written as

(1.1) gt=n=1r1σ𝒮n((1)e(σ)/n2[n1e(σ)])Δn(t)[[[δBsσ(1),δBsσ(2)],],δBsσ(n)],g_{t}=\sum_{n=1}^{r-1}\sum_{\sigma\in\mathcal{S}_{n}}\left((-1)^{e(\sigma)}\bigg{/}n^{2}\begin{bmatrix}n-1\\ e(\sigma)\end{bmatrix}\right)\int_{\Delta_{n}(t)}[[\cdots[\delta B_{s_{\sigma(1)}},\delta B_{s_{\sigma(2)}}],\cdots],\delta B_{s_{\sigma(n)}}],

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 𝔤\mathfrak{g} that includes the expression above. Additionally, we show that one may make sense of the above expression when the Lie bracket on 𝔤CM\mathfrak{g}_{CM} does not necessarily extend to 𝔤\mathfrak{g}. Thus we are able to define a “group Brownian motion” {gt}t0\{g_{t}\}_{t\geq 0} on GG via (1.1). We let νt:=Law(gt)\nu_{t}:=\mathrm{Law}(g_{t}) be the heat kernel measure on GG.

In particular, the integrals above are defined as a limit of stochastic integrals on finite-dimensional subgroups GπG_{\pi} of GCMG_{CM}. We show that these GπG_{\pi} are nice in the sense that they approximate GCMG_{CM} 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 hGCMh\in G_{CM}, let Lh,Rh:GCMGCML_{h},R_{h}:G_{CM}\rightarrow G_{CM} denote left and right translation by hh, respectively. Then LhL_{h} and RhR_{h} define measurable transformations on GG, and for all T>0T>0, νtLh1\nu_{t}\circ L_{h}^{-1} and νtRh1\nu_{t}\circ R_{h}^{-1} are absolutely continuous with respect to νt\nu_{t}. Let

Jtl(h,):=d(νtLh1)dνt and Jtr(h,):=d(νtRh1)dνtJ_{t}^{l}(h,\cdot):=\frac{d(\nu_{t}\circ L_{h}^{-1})}{d\nu_{t}}\qquad\text{ and }\qquad J_{t}^{r}(h,\cdot):=\frac{d(\nu_{t}\circ R_{h}^{-1})}{d\nu_{t}}

be the Radon-Nikodym derivatives, kk be the uniform lower bound on the Ricci curvatures of the finite-dimensional approximation groups GπG_{\pi} and

c(t):=tet1, for all t,c(t):=\frac{t}{e^{t}-1},\qquad\text{ for all }t\in\mathbb{R},

with the convention that c(0)=1c(0)=1. Then, for all p[1,)p\in[1,\infty), Jtl(h,),Jtr(h,)Lp(νt)J_{t}^{l}(h,\cdot),J_{t}^{r}(h,\cdot)\in L^{p}(\nu_{t}) and both satisfy the estimate

Jt(h,)Lp(νt)exp(c(kt)(p1)2td(𝐞,h)2),\|J_{t}^{*}(h,\cdot)\|_{L^{p}(\nu_{t})}\leq\exp\left(\frac{c(kt)(p-1)}{2t}d(\mathbf{e},h)^{2}\right),

where =l*=l or =r*=r.

The fact that one may define a measurable left or right action on GG by an element of GCMG_{CM} is discussed in Section 3. The lower bound on the Ricci curvature is proved in Proposition 3.10.

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 𝔳\mathfrak{v} by an infinite-dimensional abstract Wiener space (see Example 3.3). At several points in the analysis there, the fact that dim(𝔳)<\mathrm{dim}(\mathfrak{v})<\infty was used in a critical way. In particular, it was used to show that the stochastic integrals defining the Brownian motion on GG 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 𝔤\mathfrak{g}, it is not assumed that the Lie bracket structure on 𝔤CM\mathfrak{g}_{CM} extends to 𝔤\mathfrak{g}, and so 𝔤\mathfrak{g} 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 𝔤\mathfrak{g}. However, it turns out that the group construction on 𝔤CM\mathfrak{g}_{CM} 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 𝔤\mathfrak{g}, its critical analysis depends more on the structure of 𝔤CM\mathfrak{g}_{CM}. 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 [,]:𝔤×𝔤𝔤[\cdot,\cdot]:\mathfrak{g}\times\mathfrak{g}\rightarrow\mathfrak{g}. For example, in the event that 𝔤=W×𝔳\mathfrak{g}=W\times\mathfrak{v} where 𝔳\mathfrak{v} is a finite-dimensional Lie algebra and [,]:𝔤×𝔤𝔳[\cdot,\cdot]:\mathfrak{g}\times\mathfrak{g}\rightarrow\mathfrak{v}, it is well-known that this implies that the restriction of the bracket to 𝔤CM=H×𝔳\mathfrak{g}_{CM}=H\times\mathfrak{v} is Hilbert-Schmidt. (For any continuous bilinear ω:W×WK\omega:W\times W\rightarrow K where KK is a Hilbert space, one has that ω(H2)K<\|\omega\|_{(H^{\otimes 2})^{*}\otimes K}<\infty; 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 𝔤CM\mathfrak{g}_{CM} be a Lie subalgebra of 𝔤\mathfrak{g}, that is, for the restriction of the Lie bracket to 𝔤CM\mathfrak{g}_{CM} to preserve 𝔤CM\mathfrak{g}_{CM}. 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 𝔤\mathfrak{g} which preserves 𝔤CM\mathfrak{g}_{CM}.

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 WW is a real separable Banach space and W\mathcal{B}_{W} is the Borel σ\sigma-algebra on WW.

Definition 2.1.

A measure μ\mu on (W,W)(W,\mathcal{B}_{W}) is called a (mean zero, non-degenerate) Gaussian measure provided that its characteristic functional is given by

μ^(u):=Weiu(x)𝑑μ(x)=e12q(u,u), for all uW,\hat{\mu}(u):=\int_{W}e^{iu(x)}d\mu(x)=e^{-\frac{1}{2}q(u,u)},\qquad\text{ for all }u\in W^{*},

for q=qμ:W×Wq=q_{\mu}:W^{*}\times W^{*}\rightarrow\mathbb{R} a symmetric, positive definite quadratic form. That is, qq is a real inner product on WW^{*}.

Theorem 2.2.

Let μ\mu be a Gaussian measure on WW. For wWw\in W, let

wH:=supuW{0}|u(w)|q(u,u),\|w\|_{H}:=\sup\limits_{u\in W^{*}\setminus\{0\}}\frac{|u(w)|}{\sqrt{q(u,u)}},

and define the Cameron–Martin subspace HWH\subset W by

H:={hW:hH<}.H:=\{h\in W:\|h\|_{H}<\infty\}.

Then HH is a dense subspace of WW, and there exists a unique inner product ,H\langle\cdot,\cdot\rangle_{H} on HH such that hH2=h,hH\|h\|_{H}^{2}=\langle h,h\rangle_{H} for all hHh\in H, and HH is a separable Hilbert space with respect to this inner product. For any hHh\in H, hWChH\|h\|_{W}\leq C\|h\|_{H} for some C<C<\infty.

Alternatively, given WW a real separable Banach space and HH a real separable Hilbert space continuously embedded in WW as a dense subspace, then for each wWw^{*}\in W^{*} there exists a unique hwHh_{w^{*}}\in H such that h,w=h,hwH\langle h,w^{*}\rangle=\langle h,h_{w^{*}}\rangle_{H} for all hHh\in H. Then WwhwHW^{*}\ni w^{*}\mapsto h_{w^{*}}\in H is continuous, linear, and one-to-one with a dense range

(2.1) H:={hw:wW},H_{*}:=\{h_{w^{*}}:w^{*}\in W^{*}\},

and WwhwWW^{*}\ni w^{*}\mapsto h_{w^{*}}\in W is continuous. A Gaussian measure on WW is a Borel probability measure μ\mu such that, for each wWw^{*}\in W^{*}, the random variable ww,ww\mapsto\langle w,w^{*}\rangle under μ\mu is a centered Gaussian with variance hwH2\|h_{w^{*}}\|_{H}^{2}.

Suppose that P:HHP:H\rightarrow H is a finite rank orthogonal projection such that PHHPH\subset H_{*}. Let {hj}j=1m\{h_{j}\}_{j=1}^{m} be an orthonormal basis for PHPH. Then we may extend PP to a (unique) continuous operator from WHW\rightarrow H (still denoted by PP) by letting

(2.2) Pw:=j=1mw,hjHhjPw:=\sum_{j=1}^{m}\langle w,h_{j}\rangle_{H}h_{j}

for all wWw\in W.

Notation 2.3.

Let Proj(W)\mathrm{Proj}(W) denote the collection of finite rank projections on WW such that PWHPW\subset H_{*} and P|H:HHP|_{H}:H\rightarrow H is an orthogonal projection, that is, PP has the form given in equation (2.2).

Let {Bt}t0\{B_{t}\}_{t\geq 0} be a Brownian motion on WW with variance determined by

𝔼[Bs,hHBt,kH]=h,kHmin(s,t),\mathbb{E}\left[\langle B_{s},h\rangle_{H}\langle B_{t},k\rangle_{H}\right]=\langle h,k\rangle_{H}\min(s,t),

for all s,t0s,t\geq 0 and h,kHh,k\in H_{*}, where HH_{*} is as in (2.1). Note that for any PProj(W)P\in\mathrm{Proj}(W), PBPB is a Brownian motion on PHPH. In the rest of this section, we will verify the existence of martingales defined as certain iterated stochastic integrals with respect to BtB_{t}.

The following is Proposition 4.1 of [32]. Note that again this was stated in the context where H=𝔤CMH=\mathfrak{g}_{CM} 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 {Pm}m=1Proj(W)\{P_{m}\}_{m=1}^{\infty}\subset\mathrm{Proj}(W) such that Pm|HIHP_{m}|_{H}\uparrow I_{H}. Then, for ξL2(Δn(t),Hn)\xi\in L^{2}(\Delta_{n}(t),H^{\otimes n}) a continuous mapping, let

Jnm(ξ)t\displaystyle J_{n}^{m}(\xi)_{t} :=Δn(t)Pmnξ(s),dBs1dBsnHn\displaystyle:=\int_{\Delta_{n}(t)}\langle P_{m}^{\otimes n}\xi(s),dB_{s_{1}}\otimes\cdots\otimes dB_{s_{n}}\rangle_{H^{\otimes n}}
=Δn(t)ξ(s),dPmBs1dPmBsnHn.\displaystyle=\int_{\Delta_{n}(t)}\langle\xi(s),dP_{m}B_{s_{1}}\otimes\cdots\otimes dP_{m}B_{s_{n}}\rangle_{H^{\otimes n}}.

Then {Jnm(ξ)t}t0\{J_{n}^{m}(\xi)_{t}\}_{t\geq 0} is a continuous L2L^{2}-martingale, and there exists a continuous L2L^{2}-martingale {Jn(ξ)t}t0\{J_{n}(\xi)_{t}\}_{t\geq 0} such that

(2.3) limm𝔼[supτt|Jnm(ξ)τJn(ξ)τ|2]=0\lim_{m\rightarrow\infty}\mathbb{E}\left[\sup_{\tau\leq t}|J_{n}^{m}(\xi)_{\tau}-J_{n}(\xi)_{\tau}|^{2}\right]=0

and

(2.4) 𝔼|Jn(ξ)t|2ξL2(Δn(t),Hn)2\mathbb{E}|J_{n}(\xi)_{t}|^{2}\leq\|\xi\|^{2}_{L^{2}(\Delta_{n}(t),H^{\otimes n})}

for all t<t<\infty. The process Jn(ξ)J_{n}(\xi) is well-defined independent of the choice of increasing orthogonal projections {Pm}m=1\{P_{m}\}_{m=1}^{\infty} into HH_{*}, and so will be denoted by

Jn(ξ)t=Δn(t)ξ(s),dBs1dBsnHn.J_{n}(\xi)_{t}=\int_{\Delta_{n}(t)}\langle\xi(s),dB_{s_{1}}\otimes\cdots\otimes dB_{s_{n}}\rangle_{H^{\otimes n}}.

Now we may use this result to define stochastic integrals taking values in another Hilbert space KK.

Proposition 2.5.

Let KK be a Hilbert space and FL2(Δn(t),(Hn)K)F\in L^{2}(\Delta_{n}(t),(H^{\otimes n})^{*}\otimes K) be a continuous map. That is, F:Δn(t)×HnKF:\Delta_{n}(t)\times H^{\otimes n}\rightarrow K is a map continuous in ss and linear on HnH^{\otimes n} such that

Δn(t)F(s)22𝑑s=Δn(t)j1,,jn=1F(s)(hj1hjn)K2ds<.\int_{\Delta_{n}(t)}\|F(s)\|_{2}^{2}\,ds=\int_{\Delta_{n}(t)}\sum_{j_{1},\ldots,j_{n}=1}^{\infty}\|F(s)(h_{j_{1}}\otimes\cdots\otimes h_{j_{n}})\|_{K}^{2}\,ds<\infty.

Then

Jnm(F)t:=Δn(t)F(dPmBs1dPmBsn)J_{n}^{m}(F)_{t}:=\int_{\Delta_{n}(t)}F(dP_{m}B_{s_{1}}\otimes\cdots\otimes dP_{m}B_{s_{n}})

is a continuous KK-valued L2L^{2}-martingale, and there exists a continuous KK-valued L2L^{2}-martingale {Jn(F)t}t0\{J_{n}(F)_{t}\}_{t\geq 0} such that

(2.5) limm𝔼[supτtJnm(F)τJn(F)τK2]=0,\lim_{m\rightarrow\infty}\mathbb{E}\left[\sup_{\tau\leq t}\|J_{n}^{m}(F)_{\tau}-J_{n}(F)_{\tau}\|_{K}^{2}\right]=0,

for all t<t<\infty. The martingale Jn(F)J_{n}(F) is well-defined independent of the choice of orthogonal projections, and thus will be denoted by

Jn(F)t=Δn(t)F(dBs1dBsn).J_{n}(F)_{t}=\int_{\Delta_{n}(t)}F(dB_{s_{1}}\otimes\cdots\otimes dB_{s_{n}}).
Proof.

Let {ej}j=1\{e_{j}\}_{j=1}^{\infty} be an orthonormal basis of KK. Since F(s)(),ej\langle F(s)(\cdot),e_{j}\rangle is linear on HnH^{\otimes n}, for each ss there exists ξj(s)Hn\xi_{j}(s)\in H^{\otimes n} such that

(2.6) ξj(s),k1kn=F(s)(k1kn),ej.\langle\xi_{j}(s),k_{1}\otimes\cdots\otimes k_{n}\rangle=\langle F(s)(k_{1}\otimes\cdots\otimes k_{n}),e_{j}\rangle.

If ξj:Δn(t)Hn\xi_{j}:\Delta_{n}(t)\rightarrow H^{\otimes n} is defined by equation (2.6), then clearly ξjL2(Δn(t),Hn)\xi_{j}\in L^{2}(\Delta_{n}(t),H^{\otimes n}) and in particular

FL2(Δn(t)×Hn,K)2\displaystyle\|F\|_{L^{2}(\Delta_{n}(t)\times H^{\otimes n},K)}^{2} =j=1ξjL2(Δn(t),Hn)<.\displaystyle=\sum_{j=1}^{\infty}\|\xi_{j}\|_{L^{2}(\Delta_{n}(t),H^{\otimes n})}<\infty.

Thus, for Jn(ξj)J_{n}(\xi_{j}) as defined in Proposition 2.4,

𝔼[j=1|Jn(ξj)t|2]\displaystyle\mathbb{E}\left[\sum_{j=1}^{\infty}|J_{n}(\xi_{j})_{t}|^{2}\right] tnn!𝔼[Δn(t)j=1|ξj(s),dBs1dBsnHn|2]\displaystyle\leq\frac{t^{n}}{n!}\mathbb{E}\left[\int_{\Delta_{n}(t)}\sum_{j=1}^{\infty}|\langle\xi_{j}(s),dB_{s_{1}}\otimes\cdots\otimes dB_{s_{n}}\rangle_{H^{\otimes n}}|^{2}\right]
=tnn!j=1ξjL2(Δn(t),Hn)2<,\displaystyle=\frac{t^{n}}{n!}\sum_{j=1}^{\infty}\|\xi_{j}\|^{2}_{L^{2}(\Delta_{n}(t),H^{\otimes n})}<\infty,

and so we may write

j=1Jn(ξj)tej\displaystyle\sum_{j=1}^{\infty}J_{n}(\xi_{j})_{t}e_{j} =j=1Δn(t)ξj(s),dBs1dBsnHnej\displaystyle=\sum_{j=1}^{\infty}\int_{\Delta_{n}(t)}\langle\xi_{j}(s),dB_{s_{1}}\otimes\cdots\otimes dB_{s_{n}}\rangle_{H^{\otimes n}}e_{j}
=Δn(t)j=1F(s)(dBs1dBsn),ejKej\displaystyle=\int_{\Delta_{n}(t)}\sum_{j=1}^{\infty}\langle F(s)(dB_{s_{1}}\otimes\cdots\otimes dB_{s_{n}}),e_{j}\rangle_{K}e_{j}
=Δn(t)F(s)(dBs1dBsn).\displaystyle=\int_{\Delta_{n}(t)}F(s)(dB_{s_{1}}\otimes\cdots\otimes dB_{s_{n}}).

Thus, taking Jn(F)t:=j=1Jn(ξj)tejJ_{n}(F)_{t}:=\sum_{j=1}^{\infty}J_{n}(\xi_{j})_{t}e_{j}, we also have that

𝔼Jn(F)tJnm(F)tK2\displaystyle\mathbb{E}\|J_{n}(F)_{t}-J_{n}^{m}(F)_{t}\|_{K}^{2} =𝔼[j=1|Jn(ξj)tJnm(ξj)t|2]0\displaystyle=\mathbb{E}\left[\sum_{j=1}^{\infty}|J_{n}(\xi_{j})_{t}-J_{n}^{m}(\xi_{j})_{t}|^{2}\right]\rightarrow 0

as mm\rightarrow\infty by (2.3) and dominated convergence since

𝔼|Jn(ξj)tJnm(ξj)t|24ξjL2(Δn(t),Hn)2\mathbb{E}|J_{n}(\xi_{j})_{t}-J_{n}^{m}(\xi_{j})_{t}|^{2}\leq 4\|\xi_{j}\|_{L^{2}(\Delta_{n}(t),H^{\otimes n})}^{2}

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 WW.

Proposition 2.6.

Let VV be an arbitrary finite-dimensional subspace of HH, and let π:HV\pi:H\rightarrow V denote orthogonal projection onto VV. Then for any Hilbert space KK and FL2(Δn(t),(Hn)K)F\in L^{2}(\Delta_{n}(t),(H^{\otimes n})^{*}\otimes K) a continuous map, the stochastic integral

Jnπ(F)t:=Δn(t)F(dπBs1dπBsn)J_{n}^{\pi}(F)_{t}:=\int_{\Delta_{n}(t)}F(d\pi B_{s_{1}}\otimes\cdots\otimes d\pi B_{s_{n}})

is well-defined, and {Jnπ(F)t}t0\{J_{n}^{\pi}(F)_{t}\}_{t\geq 0} is a continuous KK-valued L2L^{2}-martingale. Moreover, if VmV_{m} is an increasing sequence of finite-dimensional subspaces of HH such that the corresponding orthogonal projections πmIH\pi_{m}\uparrow I_{H}, then

limm𝔼[supτtJnπm(F)τJn(F)τ2]=0,\lim_{m\rightarrow\infty}\mathbb{E}\left[\sup_{\tau\leq t}\|J_{n}^{\pi_{m}}(F)_{\tau}-J_{n}(F)_{\tau}\|^{2}\right]=0,

where Jn(F)J_{n}(F) is as defined in Proposition 2.5.

Proof.

First consider the case that K=K=\mathbb{R}, and thus F(s)=ξ(s),F(s)=\langle\xi(s),\cdot\rangle for a continuous ξL2(Δn(t),Hn)\xi\in L^{2}(\Delta_{n}(t),H^{\otimes n}). Since πnξL2(Δn(t),Hn)\pi^{\otimes n}\xi\in L^{2}(\Delta_{n}(t),H^{\otimes n}), the definition of Jnπ(ξ)=Jn(πnξ)J_{n}^{\pi}(\xi)=J_{n}(\pi^{\otimes n}\xi) follows from Propostion 2.4. Moreover, by equation (2.4),

𝔼|Jnπm(ξ)tJn(ξ)t|2\displaystyle\mathbb{E}|J_{n}^{\pi_{m}}(\xi)_{t}-J_{n}(\xi)_{t}|^{2} =𝔼|Jn(πmnξ)tJn(ξ)t|2=𝔼|Jn(πmnξξ)t|2\displaystyle=\mathbb{E}|J_{n}(\pi_{m}^{\otimes n}\xi)_{t}-J_{n}(\xi)_{t}|^{2}=\mathbb{E}|J_{n}(\pi_{m}^{\otimes n}\xi-\xi)_{t}|^{2}
πmnξξL2(Δn(t),Hn)0\displaystyle\leq\|\pi_{m}^{\otimes n}\xi-\xi\|_{L^{2}(\Delta_{n}(t),H^{\otimes n})}\rightarrow 0

as mm\rightarrow\infty. Now the proof for general FF follows just as in Proposition 2.5. ∎

3. Abstract nilpotent Lie algebras and groups

Definition 3.1.

Let (𝔤,𝔤CM,μ)(\mathfrak{g},\mathfrak{g}_{CM},\mu) be an abstract Wiener space such that 𝔤CM\mathfrak{g}_{CM} is equipped with a nilpotent Hilbert-Schmidt Lie bracket. Then we will call (𝔤,𝔤CM,μ)(\mathfrak{g},\mathfrak{g}_{CM},\mu) an abstract nilpotent Lie algebra.

The Baker-Campbell-Hausdorff-Dynkin formula implies that

log(eAeB)=A+B+k=1r1(n,m)kan,mkadAn1adBm1adAnkadBmkA,\log(e^{A}e^{B})=A+B+\sum_{k=1}^{r-1}\sum_{(n,m)\in\mathcal{I}_{k}}a_{n,m}^{k}\mathrm{ad}_{A}^{n_{1}}\mathrm{ad}_{B}^{m_{1}}\cdots\mathrm{ad}_{A}^{n_{k}}\mathrm{ad}_{B}^{m_{k}}A,

for all A,B𝔤CMA,B\in\mathfrak{g}_{CM}, where

an,mk:=(1)k(k+1)m!n!(|n|+1),a_{n,m}^{k}:=\frac{(-1)^{k}}{(k+1)m!n!(|n|+1)},

k:={(n,m)+k×+k:ni+mi>0 for all 1ik}\mathcal{I}_{k}:=\{(n,m)\in\mathbb{Z}_{+}^{k}\times\mathbb{Z}_{+}^{k}:n_{i}+m_{i}>0\text{ for all }1\leq i\leq k\}, and for each multi-index n+kn\in\mathbb{Z}_{+}^{k},

n!=n1!nk! and |n|=n1++nk;n!=n_{1}!\cdots n_{k}!\quad\text{ and }\quad|n|=n_{1}+\cdots+n_{k};

see, for example, [18]. If 𝔤CM\mathfrak{g}_{CM} is nilpotent of step rr, then

adAn1adBm1adAnkadBmkA=0if |n|+|m|r.\mathrm{ad}_{A}^{n_{1}}\mathrm{ad}_{B}^{m_{1}}\cdots\mathrm{ad}_{A}^{n_{k}}\mathrm{ad}_{B}^{m_{k}}A=0\quad\text{if }|n|+|m|\geq r.

for A,B𝔤CMA,B\in\mathfrak{g}_{CM}. Since 𝔤CM\mathfrak{g}_{CM} 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 𝔤CM\mathfrak{g}_{CM} as both a Lie algebra and Lie group, and one may verify that

(3.1) gh\displaystyle g\cdot h =g+h+k=1r1(n,m)kan,mkadgn1adhm1adgnkadhmkg\displaystyle=g+h+\sum_{k=1}^{r-1}\sum_{(n,m)\in\mathcal{I}_{k}}a_{n,m}^{k}\mathrm{ad}_{g}^{n_{1}}\mathrm{ad}_{h}^{m_{1}}\cdots\mathrm{ad}_{g}^{n_{k}}\mathrm{ad}_{h}^{m_{k}}g

defines a group structure on 𝔤CM\mathfrak{g}_{CM}. Note that g1=gg^{-1}=-g and the identity 𝐞=(0,0)\mathbf{e}=(0,0).

Definition 3.2.

When we wish to emphasize the group structure on 𝔤CM\mathfrak{g}_{CM}, we will denote 𝔤CM\mathfrak{g}_{CM} by GCMG_{CM}.

Lemma 3.3.

The Banach space topology on 𝔤CM\mathfrak{g}_{CM} makes GCMG_{CM} into a topological group.

Proof.

Since 𝔤CM\mathfrak{g}_{CM} is a topological vector space, gg1=gg\mapsto g^{-1}=-g and (g1,g2)g1+g2(g_{1},g_{2})\mapsto g_{1}+g_{2} are continuous by definition. The map (g1,g2)[g1,g2](g_{1},g_{2})\mapsto[g_{1},g_{2}] is continuous in the 𝔤CM\mathfrak{g}_{CM} topology by the boundedness of the Lie bracket. It then follows from (3.1) that (g1,g2)g1g2(g_{1},g_{2})\mapsto g_{1}\cdot g_{2} is continuous as well. ∎

3.1. Measurable group actions on GG

As discussed in the introduction, given a Hilbert-Schmidt Lie bracket on 𝔤CM\mathfrak{g}_{CM} and a subsequently defined group operation on GCMG_{CM}, one may define a measurable action on GG by left or right multiplication by an element of GCMG_{CM}.

In particular, let {en}n=1\{e_{n}\}_{n=1}^{\infty} be an orthonormal basis of 𝔤CM\mathfrak{g}_{CM}. For now, fix nn and consider the mapping 𝔤CM𝔤CM\mathfrak{g}_{CM}\rightarrow\mathfrak{g}_{CM}^{*} given by hadh,enh\mapsto\langle\mathrm{ad}_{h}\cdot,e_{n}\rangle. Then this is a continuous linear map on 𝔤CM\mathfrak{g}_{CM} and in the usual way we may make the identification of 𝔤CM𝔤CM\mathfrak{g}_{CM}^{*}\cong\mathfrak{g}_{CM} so that we define the operator An:𝔤CM𝔤CMA_{n}:\mathfrak{g}_{CM}\rightarrow\mathfrak{g}_{CM} given by

Anh,k=adhk,en;\langle A_{n}h,k\rangle=\langle\mathrm{ad}_{h}k,e_{n}\rangle;

in particular, Anh=adhenA_{n}h=\mathrm{ad}_{h}^{*}e_{n}. Note that, for any h,k𝔤CMh,k\in\mathfrak{g}_{CM}

Anh,k=Ank,h=adken,h=en,adkh=en,adhk=adhen,k\langle A_{n}^{*}h,k\rangle=\langle A_{n}k,h\rangle=\langle\mathrm{ad}_{k}^{*}e_{n},h\rangle=\langle e_{n},\mathrm{ad}_{k}h\rangle=-\langle e_{n},\mathrm{ad}_{h}k\rangle=-\langle\mathrm{ad}_{h}^{*}e_{n},k\rangle

and thus An=AnA_{n}^{*}=-A_{n}. Now fix hGCM=𝔤CMh\in G_{CM}=\mathfrak{g}_{CM}. Then for adh:𝔤CM𝔤CM\mathrm{ad}_{h}:\mathfrak{g}_{CM}\rightarrow\mathfrak{g}_{CM} we may write

adhk=nadhk,enen=nAnh,ken.\mathrm{ad}_{h}k=\sum_{n}\langle\mathrm{ad}_{h}k,e_{n}\rangle e_{n}=\sum_{n}\langle A_{n}h,k\rangle e_{n}.

Since each Anh,𝔤CM\langle A_{n}h,\cdot\rangle\in\mathfrak{g}_{CM}^{*} has a measurable linear extension to GG such that Anh,L2(μ)=Anh𝔤CM\|\langle A_{n}h,\cdot\rangle\|_{L^{2}(\mu)}=\|A_{n}h\|_{\mathfrak{g}_{CM}} (see, for example, Theorem 2.10.11 of [10]), we may extend adh\mathrm{ad}_{h} to a measurable linear transformation from G=𝔤G=\mathfrak{g} to GCM=𝔤CMG_{CM}=\mathfrak{g}_{CM} (still denoted by adh\mathrm{ad}_{h}) given by

adhg:=nAnh,gen.\mathrm{ad}_{h}g:=\sum_{n}\langle A_{n}h,g\rangle e_{n}.

Note that here we are using the fact that

nAnh,L2(μ)2=nAnhH2\displaystyle\sum_{n}\|\langle A_{n}h,\cdot\rangle\|_{L^{2}(\mu)}^{2}=\sum_{n}\|A_{n}h\|_{H}^{2} =n,mAnh,em2\displaystyle=\sum_{n,m}\langle A_{n}h,e_{m}\rangle^{2}
=n,madhem,en2h2[,]HS2\displaystyle=\sum_{n,m}\langle\mathrm{ad}_{h}e_{m},e_{n}\rangle^{2}\leq\|h\|^{2}\|[\cdot,\cdot]\|_{HS}^{2}

which implies that

nAnh,g2<g-a.s.\sum_{n}\langle A_{n}h,g\rangle^{2}<\infty\quad g\text{-a.s.}

Similarly, we may define

adgh:=adhg=nAnh,gen.\mathrm{ad}_{g}h:=-\mathrm{ad}_{h}g=-\sum_{n}\langle A_{n}h,g\rangle e_{n}.

In a similar way, note that we may write, for h,kGCMh,k\in G_{CM} and m<rm<r,

adhmk\displaystyle\mathrm{ad}_{h}^{m}k =1m(b=1m1Abh,eb+1)Amh,ke1\displaystyle=\sum_{\ell_{1}}\cdots\sum_{\ell_{m}}\left(\prod_{b=1}^{m-1}\langle A_{\ell_{b}}h,e_{\ell_{b+1}}\rangle\right)\langle A_{\ell_{m}}h,k\rangle e_{\ell_{1}}
=(1)m1m(b=1m1Abeb+1,h)Amk,he1.\displaystyle=(-1)^{m}\sum_{\ell_{1}}\cdots\sum_{\ell_{m}}\left(\prod_{b=1}^{m-1}\langle A_{\ell_{b}}e_{\ell_{b+1}},h\rangle\right)\langle A_{\ell_{m}}k,h\rangle e_{\ell_{1}}.

and thus for h,k,xGCMh,k,x\in G_{CM} and n+m<rn+m<r

adknadhmx=(1)nj1jn1m(a=1n1Ajaeja+1,k)Ajne1,k(b=1m1Abh,eb+1)Amh,xej1.\mathrm{ad}_{k}^{n}\mathrm{ad}_{h}^{m}x=(-1)^{n}\sum_{j_{1}}\cdots\sum_{j_{n}}\sum_{\ell_{1}}\cdots\sum_{\ell_{m}}\left(\prod_{a=1}^{n-1}\langle A_{j_{a}}e_{j_{a+1}},k\rangle\right)\langle A_{j_{n}}e_{\ell_{1}},k\rangle\\ \left(\prod_{b=1}^{m-1}\langle A_{\ell_{b}}h,e_{\ell_{b+1}}\rangle\right)\langle A_{\ell_{m}}h,x\rangle e_{j_{1}}.

More generally for |n|+|m|<r|n|+|m|<r

adkn1adhm1adknsadhmsk\displaystyle\mathrm{ad}_{k}^{n_{1}}\mathrm{ad}_{h}^{m_{1}}\cdots\mathrm{ad}_{k}^{n_{s}}\mathrm{ad}_{h}^{m_{s}}k
=(1)|n|j11mssc=1s{(ac=1nc1Ajaccejac+1c,k)Ajncce1c,k\displaystyle=(-1)^{|n|}\sum_{j^{1}_{1}}\cdots\sum_{\ell^{s}_{m_{s}}}\prod_{c=1}^{s}\Bigg{\{}\left(\prod_{a_{c}=1}^{n_{c}-1}\langle A_{j^{c}_{a_{c}}}e_{j^{c}_{a_{c}+1}},k\rangle\right)\langle A_{j^{c}_{n_{c}}}e_{\ell^{c}_{1}},k\rangle
×(bc=1mc1Abcch,ebc+1c)}{c=1s1Amcch,ej1c+1}Amssh,kej11\displaystyle\qquad\qquad\times\left(\prod_{b_{c}=1}^{m_{c}-1}\langle A_{\ell^{c}_{b_{c}}}h,e_{\ell^{c}_{b_{c}+1}}\rangle\right)\Bigg{\}}\left\{\prod_{c=1}^{s-1}\langle A_{\ell^{c}_{m_{c}}}h,e_{j^{c+1}_{1}}\rangle\right\}\langle A_{\ell^{s}_{m_{s}}}h,k\rangle e_{j^{1}_{1}}

Thus, for hGCMh\in G_{CM} and |n|+|m|<r|n|+|m|<r, we may define a measurable action on GG given by

Gg\displaystyle G\ni g\mapsto\, adgn1adhm1adgnsadhmsg\displaystyle\mathrm{ad}_{g}^{n_{1}}\mathrm{ad}_{h}^{m_{1}}\cdots\mathrm{ad}_{g}^{n_{s}}\mathrm{ad}_{h}^{m_{s}}g
:=(1)|n|j11mssc=1s{(ac=1nc1Ajaccejac+1c,g)Ajncce1c,g\displaystyle:=(-1)^{|n|}\sum_{j^{1}_{1}}\cdots\sum_{\ell^{s}_{m_{s}}}\prod_{c=1}^{s}\bigg{\{}\left(\prod_{a_{c}=1}^{n_{c}-1}\langle A_{j^{c}_{a_{c}}}e_{j^{c}_{a_{c}+1}},g\rangle\right)\langle A_{j^{c}_{n_{c}}}e_{\ell^{c}_{1}},g\rangle
×(bc=1mc1Abcch,ebc+1c)}{c=1s1Amcch,ej1c+1}Amssh,kej11GCM.\displaystyle\times\left(\prod_{b_{c}=1}^{m_{c}-1}\langle A_{\ell^{c}_{b_{c}}}h,e_{\ell^{c}_{b_{c}+1}}\rangle\right)\Bigg{\}}\left\{\prod_{c=1}^{s-1}\langle A_{\ell^{c}_{m_{c}}}h,e_{j^{c+1}_{1}}\rangle\right\}\langle A_{\ell^{s}_{m_{s}}}h,k\rangle e_{j^{1}_{1}}\in G_{CM}.

(For |n|+|m|r|n|+|m|\geq r, we define this mapping to be 0, which is certainly measurable.) Again, we are using that

j11j21mssc=1s{(ac=1nc1Ajaccejac+1c,)Ajncce1c,×(bc=1mc1Abcch,ebc+1c)}{c=1s1Amcch,ej1c+1}Amssh,kej11L2(μ)2<.\sum_{j^{1}_{1}}\Bigg{\|}\sum_{j^{1}_{2}}\cdots\sum_{\ell^{s}_{m_{s}}}\prod_{c=1}^{s}\Bigg{\{}\left(\prod_{a_{c}=1}^{n_{c}-1}\langle A_{j^{c}_{a_{c}}}e_{j^{c}_{a_{c}+1}},\cdot\rangle\right)\langle A_{j^{c}_{n_{c}}}e_{\ell^{c}_{1}},\cdot\rangle\\ \times\left(\prod_{b_{c}=1}^{m_{c}-1}\langle A_{\ell^{c}_{b_{c}}}h,e_{\ell^{c}_{b_{c}+1}}\rangle\right)\Bigg{\}}\left\{\prod_{c=1}^{s-1}\langle A_{\ell^{c}_{m_{c}}}h,e_{j^{c+1}_{1}}\rangle\right\}\langle A_{\ell^{s}_{m_{s}}}h,k\rangle e_{j^{1}_{1}}\Bigg{\|}_{L^{2}(\mu)}^{2}<\infty.

This holds by straightforward but tedious computations — in fact, iterative applications of Cauchy-Schwarz combined with the fact that n,mAnem2=[,]HS2<\sum_{n,m}\|A_{n}e_{m}\|^{2}=\|[\cdot,\cdot]\|_{HS}^{2}<\infty. Thus,

j11(j21mssc=1s{(ac=1nc1Ajaccejac+1c,g)Ajncce1c,g×(bc=1mc1Abcch,ebc+1c)}{c=1s1Amcch,ej1c+1}Amssh,kej11)2<,\sum_{j^{1}_{1}}\Bigg{(}\sum_{j^{1}_{2}}\cdots\sum_{\ell^{s}_{m_{s}}}\prod_{c=1}^{s}\Bigg{\{}\left(\prod_{a_{c}=1}^{n_{c}-1}\langle A_{j^{c}_{a_{c}}}e_{j^{c}_{a_{c}+1}},g\rangle\right)\langle A_{j^{c}_{n_{c}}}e_{\ell^{c}_{1}},g\rangle\\ \times\left(\prod_{b_{c}=1}^{m_{c}-1}\langle A_{\ell^{c}_{b_{c}}}h,e_{\ell^{c}_{b_{c}+1}}\rangle\right)\Bigg{\}}\left\{\prod_{c=1}^{s-1}\langle A_{\ell^{c}_{m_{c}}}h,e_{j^{c+1}_{1}}\rangle\right\}\langle A_{\ell^{s}_{m_{s}}}h,k\rangle e_{j^{1}_{1}}\Bigg{)}^{2}<\infty,

gg-a.s. and adgn1adhm1adgnsadhmsg\mathrm{ad}_{g}^{n_{1}}\mathrm{ad}_{h}^{m_{1}}\cdots\mathrm{ad}_{g}^{n_{s}}\mathrm{ad}_{h}^{m_{s}}g as given above is defined a.s. Thus, we have the following result.

Proposition 3.4.

For hGCMh\in G_{CM}, the mapping GgghGG\ni g\mapsto g\cdot h\in G defined analogously to (3.1) is a measurable right group action by GCMG_{CM} on GG, and similarly for the left action ghgg\mapsto h\cdot g.

3.2. Examples of abstract nilpotent Lie algebras

Example 3.1 (Free nilpotent Lie algebras).

Starting with an abstract Wiener space (W,H,μ)(W,H,\mu), one may construct in the standard way the abstract free nilpotent Lie algebra of step rr with generators {hi}i=1\{h_{i}\}_{i=1}^{\infty} an orthonormal basis of HH. See for example Section 0 of [21].

Example 3.2 (Heisenberg-like algebras).

Let (Wi,Hi,μi)(W_{i},H_{i},\mu_{i}) for i=1,2i=1,2 be abstract Wiener spaces. Then for any ω:H1×H1H2\omega:H_{1}\times H_{1}\rightarrow H_{2} a Hilbert-Schmidt map, we may define a Lie bracket on 𝔤CM=H1×H2\mathfrak{g}_{CM}=H_{1}\times H_{2} by

[(h1,h2),(h1,h2)]:=(0,ω(h1,h1)),[(h_{1},h_{2}),(h_{1}^{\prime},h_{2}^{\prime})]:=(0,\omega(h_{1},h_{1}^{\prime})),

and 𝔤=W1×W2\mathfrak{g}=W_{1}\times W_{2} 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 𝔳\mathfrak{v} and 𝔥\mathfrak{h} be Lie algebras, and let Der(𝔳)\mathrm{Der}(\mathfrak{v}) denote the set of derivations on 𝔳\mathfrak{v}; that is, Der(𝔳)\mathrm{Der}(\mathfrak{v}) consists of all linear maps ρ:𝔳𝔳\rho:\mathfrak{v}\rightarrow\mathfrak{v} satisfying Leibniz’s rule:

ρ([X,Y]𝔳)=[ρ(X),Y]𝔳+[X,ρ(Y)]𝔳.\rho([X,Y]_{\mathfrak{v}})=[\rho(X),Y]_{\mathfrak{v}}+[X,\rho(Y)]_{\mathfrak{v}}.

Now suppose there are a linear mapping α:𝔥Der(𝔳)\alpha:\mathfrak{h}\rightarrow\mathrm{Der}(\mathfrak{v}) and a skew-symmetric bilinear mapping ω:𝔥×𝔥𝔳\omega:\mathfrak{h}\times\mathfrak{h}\rightarrow\mathfrak{v}, satisfying, for all X,Y,Z𝔥X,Y,Z\in\mathfrak{h},

(B1) [αX,αY]α[X,Y]𝔥=adω(X,Y)[\alpha_{X},\alpha_{Y}]-\alpha_{[X,Y]_{\mathfrak{h}}}=\mathrm{ad}_{\omega(X,Y)}

and

(B2) cyclic(αXω(Y,Z)ω([X,Y]𝔥,Z))=0.\sum_{\text{cyclic}}\left(\alpha_{X}\omega(Y,Z)-\omega([X,Y]_{\mathfrak{h}},Z)\right)=0.

Then, one may verify that, for X1+V1,X2+V2𝔥𝔳X_{1}+V_{1},X_{2}+V_{2}\in\mathfrak{h}\oplus\mathfrak{v},

[X1+V1,X2+V2]𝔤:=[X1,X2]𝔥+ω(X1,X2)+αX1V2αX2V1+[V1,V2]𝔳[X_{1}+V_{1},X_{2}+V_{2}]_{\mathfrak{g}}:=[X_{1},X_{2}]_{\mathfrak{h}}+\omega(X_{1},X_{2})+\alpha_{X_{1}}V_{2}-\alpha_{X_{2}}V_{1}+[V_{1},V_{2}]_{\mathfrak{v}}

defines a Lie bracket on 𝔤:=𝔥𝔳\mathfrak{g}:=\mathfrak{h}\oplus\mathfrak{v}, and we say 𝔤\mathfrak{g} is an extension of 𝔥\mathfrak{h} over 𝔳\mathfrak{v}. That is, 𝔤\mathfrak{g} is the Lie algebra with ideal 𝔳\mathfrak{v} and quotient algebra 𝔤/𝔳=𝔥\mathfrak{g}/\mathfrak{v}=\mathfrak{h}. The associated exact sequence is

0𝔳ι1𝔤π2𝔥0,0\rightarrow\mathfrak{v}\overset{\iota_{1}}{\longrightarrow}\mathfrak{g}\overset{\pi_{2}}{\longrightarrow}\mathfrak{h}\rightarrow 0,

where ι1\iota_{1} is inclusion and π2\pi_{2} is projection. In fact, these are the only extensions of 𝔥\mathfrak{h} over 𝔳\mathfrak{v} (see, for example, [4]).

Now suppose that (W,H,μ)(W,H,\mu) is a real abstract Wiener space, and (𝔳,𝔳CM,μ0)(\mathfrak{v},\mathfrak{v}_{CM},\mu^{0}) is an abstract nilpotent Lie algebra. Motivated by the previous discussion, we may consider HH as an abelian Lie algebra and construct extensions of HH over 𝔳CM\mathfrak{v}_{CM}. In this case, we need a linear mapping α:HDer(𝔳CM)\alpha:H\rightarrow\mathrm{Der}(\mathfrak{v}_{CM}) and a skew-symmetric bilinear mapping ω:H×H𝔳CM\omega:H\times H\rightarrow\mathfrak{v}_{CM}, such that ω\omega and α:H×𝔳CM𝔳CM\alpha:H\times\mathfrak{v}_{CM}\rightarrow\mathfrak{v}_{CM} are both Hilbert-Schmidt and together ω\omega and α\alpha satisfy (B1) and (B2), which in this setting become

[αX,αY]=adω(X,Y)[\alpha_{X},\alpha_{Y}]=\mathrm{ad}_{\omega(X,Y)}

and

αXω(Y,Z)+αYω(Z,X)+αZω(X,Y)=0,\alpha_{X}\omega(Y,Z)+\alpha_{Y}\omega(Z,X)+\alpha_{Z}\omega(X,Y)=0,

for all X,Y,ZHX,Y,Z\in H. Then we may define a Lie algebra structure on 𝔤CM=H𝔳CM\mathfrak{g}_{CM}=H\oplus\mathfrak{v}_{CM} via the Lie bracket

[(X1,V1),(X2,V2)]𝔤CM:=(0,ω(X1,X2)+αX1V2αX2V1).[(X_{1},V_{1}),(X_{2},V_{2})]_{\mathfrak{g}_{CM}}:=(0,\omega(X_{1},X_{2})+\alpha_{X_{1}}V_{2}-\alpha_{X_{2}}V_{1}).

Again, these are in fact the only extensions of HH over 𝔳CM\mathfrak{v}_{CM}.

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 β:H𝔳CM\beta:H\rightarrow\mathfrak{v}_{CM} be any Hilbert-Schmidt map, and for X,YHX,Y\in H and V𝔳CMV\in\mathfrak{v}_{CM} define

ω(X,Y):=[β(X),β(Y)]𝔳CM\omega(X,Y):=[\beta(X),\beta(Y)]_{\mathfrak{v}_{CM}}
αXV:=adβ(X)V=[β(X),V]𝔳CM.\alpha_{X}V:=\mathrm{ad}_{\beta(X)}V=[\beta(X),V]_{\mathfrak{v}_{CM}}.

Then the Lie bracket on 𝔤CM\mathfrak{g}_{CM} is given by

[(X,V),(Y,U)]=(0,[β(X),β(Y)]𝔳CM+[β(X),U]𝔳CM[β(Y),V]𝔳CM+[V,U]𝔳CM).[(X,V),(Y,U)]\\ =(0,[\beta(X),\beta(Y)]_{\mathfrak{v}_{CM}}+[\beta(X),U]_{\mathfrak{v}_{CM}}-[\beta(Y),V]_{\mathfrak{v}_{CM}}+[V,U]_{\mathfrak{v}_{CM}}).

Note that, if 𝔳CM\mathfrak{v}_{CM} is nilpotent of step rr, then 𝔤CM\mathfrak{g}_{CM} will automatically be nilpotent of step rr.

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 dim(𝔳CM)<\mathrm{dim}(\mathfrak{v}_{CM})<\infty trivializes Example 3.4. We elaborate in the following remark.

Remark 3.5.

Returning to Example 3.4, since β\beta is linear and continuous, we have the decomposition

H=Nul(β)Nul(β),H=\mathrm{Nul}(\beta)\oplus\mathrm{Nul}(\beta)^{\perp},

where dim(Nul(β))dim(𝔳CM)\mathrm{dim}(\mathrm{Nul}(\beta)^{\perp})\leq\mathrm{dim}(\mathfrak{v}_{CM}). Thus, for X,YHX,Y\in H we may write X=X1+X2,Y=Y1+Y2Nul(β)Nul(β)X=X_{1}+X_{2},Y=Y_{1}+Y_{2}\in\mathrm{Nul}(\beta)\oplus\mathrm{Nul}(\beta)^{\perp} and

[(X1+X2,0),(Y1+Y2,0)]\displaystyle[(X_{1}+X_{2},0),(Y_{1}+Y_{2},0)] =[β(X1+X2),β(Y1+Y2)]=[β(X2),β(Y2)],\displaystyle=[\beta(X_{1}+X_{2}),\beta(Y_{1}+Y_{2})]=[\beta(X_{2}),\beta(Y_{2})],

and ω\omega is a map on Nul(β)×Nul(β)\mathrm{Nul}(\beta)^{\perp}\times\mathrm{Nul}(\beta)^{\perp}. Thus, [Nul(β),Nul(β)]={0}[\mathrm{Nul}(\beta),\mathrm{Nul}(\beta)]=\{0\} and similarly [Nul(β),𝔳]={0}[\mathrm{Nul}(\beta),\mathfrak{v}]=\{0\}. So

𝔤CM=H𝔳CM=Nul(β)Nul(β)𝔳CM,\mathfrak{g}_{CM}=H\oplus\mathfrak{v}_{CM}=\mathrm{Nul}(\beta)\oplus\mathrm{Nul}(\beta)^{\perp}\oplus\mathfrak{v}_{CM},

and, in particular, when dim(𝔳CM)<\mathrm{dim}(\mathfrak{v}_{CM})<\infty, 𝔤CM\mathfrak{g}_{CM} is just an extension of the finite-dimensional vector space Nul(β)\mathrm{Nul}(\beta)^{\perp} by the finite-dimensional Lie algebra 𝔳CM\mathfrak{v}_{CM}, and the construction is not truly infinite-dimensional.

3.3. Properties of 𝔤CM\mathfrak{g}_{CM}

This section collects some results for topological and geometric properties of 𝔤CM\mathfrak{g}_{CM} that we’ll require for the sequel.

Proposition 3.6.

For all m2m\geq 2, [[[,],],]:𝔤CMm𝔤CM[[[\cdot,\cdot],\ldots],\cdot]:\mathfrak{g}_{CM}^{\otimes m}\rightarrow\mathfrak{g}_{CM} is Hilbert-Schmidt.

Proof.

For m=2m=2, this follows from the definition of GG. Now assume the statement holds for all m=m=\ell, and consider m=+1m=\ell+1. Writing [[hi1,hi2],,hi+1]𝔤CM[[h_{i_{1}},h_{i_{2}}],\cdots,h_{i_{\ell}+1}]\in\mathfrak{g}_{CM} in terms of the orthonormal basis {ej}j=1\{e_{j}\}_{j=1}^{\infty} and using Hölder’s inequality gives

[[[,\displaystyle\|[[[\cdot, ],],]22=[[[,],],](𝔤CM)+1𝔤CM\displaystyle\cdot],\ldots],\cdot]\|_{2}^{2}=\|[[[\cdot,\cdot],\ldots],\cdot]\|_{(\mathfrak{g}_{CM}^{*})^{\otimes\ell+1}\otimes\mathfrak{g}_{CM}}
=i1,,i+1=1[[[hi1,hi2],,hi],hi+1]2\displaystyle=\sum_{i_{1},\ldots,i_{\ell+1}=1}^{\infty}\|[[[h_{i_{1}},h_{i_{2}}],\cdots,h_{i_{\ell}}],h_{i_{\ell+1}}]\|^{2}
=i1,,i+1=1j=1[ej,hi+1]ej,[[hi1,hi2],,hi]2\displaystyle=\sum_{i_{1},\ldots,i_{\ell+1}=1}^{\infty}\left\|\sum_{j=1}^{\infty}[e_{j},h_{i_{\ell+1}}]\langle e_{j},[[h_{i_{1}},h_{i_{2}}],\cdots,h_{i_{\ell}}]\rangle\right\|^{2}
i1,,i+1=1(j=1[ej,hi+1]2)(j=1|ej,[[hi1,hi2],,hi]|2)\displaystyle\leq\sum_{i_{1},\dots,i_{\ell+1}=1}^{\infty}\left(\sum_{j=1}^{\infty}\|[e_{j},h_{i_{\ell+1}}]\|^{2}\right)\left(\sum_{j=1}^{\infty}|\langle e_{j},[[h_{i_{1}},h_{i_{2}}],\cdots,h_{i_{\ell}}]\rangle|^{2}\right)
=(i+1=1j=1[ej,hi+1]2)(i1,,i=1[[hi1,hi2],,hi]2)\displaystyle=\left(\sum_{i_{\ell+1}=1}^{\infty}\sum_{j=1}^{\infty}\|[e_{j},h_{i_{\ell+1}}]\|^{2}\right)\left(\sum_{i_{1},\dots,i_{\ell}=1}^{\infty}\|[[h_{i_{1}},h_{i_{2}}],\cdots,h_{i_{\ell}}]\|^{2}\right)
[,](𝔤CM)2𝔤CM2[[[,],],](𝔤CM)𝔤CM2\displaystyle\leq\|[\cdot,\cdot]\|_{(\mathfrak{g}_{CM}^{*})^{\otimes 2}\otimes\mathfrak{g}_{CM}}^{2}\|[[[\cdot,\cdot],\ldots],\cdot]\|_{(\mathfrak{g}_{CM}^{*})^{\otimes\ell}\otimes\mathfrak{g}_{CM}}^{2}

and the last line is finite by the induction hypothesis. ∎

Next we will recall that the flat and geometric topologies on 𝔤CM\mathfrak{g}_{CM} are equivalent. First we set the following notation.

Notation 3.7.

For gGCM,g\in G_{CM}, let Lg:GCMGCML_{g}:G_{CM}\rightarrow G_{CM} and Rg:GCMGCMR_{g}:G_{CM}\rightarrow G_{CM} denote left and right multiplication by gg, respectively. As GCMG_{CM} is a vector space, to each gGCMg\in G_{CM} we can associate the tangent space TgGCMT_{g}G_{CM} to GCMG_{CM} at gg, which is naturally isomorphic to GCMG_{CM}. For f:GCMf:G_{CM}\rightarrow\mathbb{R} a Frechét smooth function and v,xGCMv,x\in G_{CM} and h𝔤CMh\in\mathfrak{g}_{CM}, let

f(x)h:=hf(x)=ddt|0f(x+th),f^{\prime}(x)h:=\partial_{h}f(x)=\frac{d}{dt}\bigg{|}_{0}f(x+th),

and let vxTxGCMv_{x}\in T_{x}G_{CM} denote the tangent vector satisfying vxf=f(x)vv_{x}f=f^{\prime}(x)v. If σ(t)\sigma(t) is any smooth curve in GCMG_{CM} such that σ(0)=x\sigma(0)=x and σ˙(0)=v\dot{\sigma}(0)=v (for example, σ(t)=x+tv\sigma(t)=x+tv), then

Lgvx=ddt|0gσ(t).L_{g*}v_{x}=\frac{d}{dt}\bigg{|}_{0}g\cdot\sigma(t).
Notation 3.8.

Let C1([0,1],GCM)C^{1}([0,1],G_{CM}) denote the collection of C1C^{1}-paths g:[0,1]GCMg:[0,1]\rightarrow G_{CM}. The length of gg is defined as

CM(g):=0TLg1(s)g(s)𝔤CM𝑑s.\ell_{CM}(g):=\int_{0}^{T}\|L_{g^{-1}(s)*}g^{\prime}(s)\|_{\mathfrak{g}_{CM}}\,ds.

The Riemannian distance between x,yGCMx,y\in G_{CM} is then defined as

d(x,y):=inf{CM(g):gCh1([0,1],GCM) such that g(0)=x and g(1)=y}.\displaystyle d(x,y):=\inf\{\ell_{CM}(g):g\in C_{h}^{1}([0,1],G_{CM})\text{ such that }g(0)=x\text{ and }g(1)=y\}.

The following proposition was proved as Corollary 4.13 in [32]. This was proved under the conditions that 𝔤CM\mathfrak{g}_{CM} was a “semi-infinite Lie algebra”, that is, under the assumption that 𝔤CM\mathfrak{g}_{CM} was a nilpotent Lie algebra extension (as in Example 3.3) of a finite-dimensional nilpotent Lie algebra 𝔳\mathfrak{v} 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 𝔤CM=H×𝔳\mathfrak{g}_{CM}=H\times\mathfrak{v} or the fact that the image of the Lie bracket was a finite-dimensional subspace.

Proposition 3.9.

The topology on GCMG_{CM} induced by dd is equivalent to the Hilbert topology induced by 𝔤CM\|\cdot\|_{\mathfrak{g}_{CM}}.

We may find a uniform lower bound on the Ricci curvature of all finite-dimensional subgroups of GCMG_{CM}. Finite-dimensional subgroups of GCMG_{CM} may be obtained by taking the Lie algebra generated by any finite-dimensional subspace of 𝔤CM\mathfrak{g}_{CM} (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

k:=12sup{[,X]𝔤CM𝔤CM2:X𝔤CM=1}.k:=-\frac{1}{2}\sup\left\{\|[\cdot,X]\|^{2}_{\mathfrak{g}_{CM}^{*}\otimes\mathfrak{g}_{CM}}:\,\|X\|_{\mathfrak{g}_{CM}}=1\right\}.

Then k>k>-\infty and kk is the largest constant such that

RicπX,X𝔤πkX𝔤π2, for all X𝔤π,\langle\mathrm{Ric}^{\pi}X,X\rangle_{\mathfrak{g}_{\pi}}\geq k\|X\|^{2}_{\mathfrak{g}_{\pi}},\quad\text{ for all }X\in\mathfrak{g}_{\pi},

holds uniformly for all 𝔤π\mathfrak{g}_{\pi} finite-dimensional Lie subalgebras of 𝔤CM\mathfrak{g}_{CM}.

Proof.

For 𝔤\mathfrak{g} any nilpotent Lie algebra with orthonormal basis Γ\Gamma,

RicX,X\displaystyle\langle\mathrm{Ric}\,X,X\rangle =14YΓadYX212YΓadYX212YΓ[Y,X]2\displaystyle=\frac{1}{4}\sum_{Y\in\Gamma}\|\mathrm{ad}^{*}_{Y}X\|^{2}-\frac{1}{2}\sum_{Y\in\Gamma}\|\mathrm{ad}_{Y}X\|^{2}\geq-\frac{1}{2}\sum_{Y\in\Gamma}\|[Y,X]\|^{2}

for all X𝔤X\in\mathfrak{g}. Thus, for 𝔤π\mathfrak{g}_{\pi} any finite-dimensional Lie algebra

RicπX,X𝔤πkπX𝔤π2, for all X𝔤π,\langle\mathrm{Ric}^{\pi}X,X\rangle_{\mathfrak{g}_{\pi}}\geq k_{\pi}\|X\|^{2}_{\mathfrak{g}_{\pi}},\quad\text{ for all }X\in\mathfrak{g}_{\pi},

where

(3.2) kπ:=12sup{[,X]𝔤π𝔤π2:X𝔤π=1}12[,]22>.k_{\pi}:=-\frac{1}{2}\sup\left\{\|[\cdot,X]\|^{2}_{\mathfrak{g}_{\pi}^{*}\otimes\mathfrak{g}_{\pi}}:\,\|X\|_{\mathfrak{g}_{\pi}}=1\right\}\geq-\frac{1}{2}\|[\cdot,\cdot]\|_{2}^{2}>-\infty.

Taking the infimum of kπk_{\pi} over all 𝔤π\mathfrak{g}_{\pi} completes the proof. ∎

4. Brownian motion on GG

Suppose that BtB_{t} is a smooth curve in 𝔤CM\mathfrak{g}_{CM} with B0=0B_{0}=0, and consider the differential equation

g˙t=gtB˙t:=LgtB˙t, with g0=𝐞.\dot{g}_{t}=g_{t}\dot{B}_{t}:=L_{g_{t}*}\dot{B}_{t},\quad\text{ with }g_{0}=\mathbf{e}.

The solution gtg_{t} may be written as follows (see [36]): For t>0t>0, let Δn(t)\Delta_{n}(t) denote the simplex in n\mathbb{R}^{n} given by

{s=(s1,,sn)n:0<s1<s2<<sn<t}.\{s=(s_{1},\cdots,s_{n})\in\mathbb{R}^{n}:0<s_{1}<s_{2}<\cdots<s_{n}<t\}.

Let 𝒮n\mathcal{S}_{n} denote the permutation group on (1,,n)(1,\cdots,n), and, for each σ𝒮n\sigma\in\mathcal{S}_{n}, let e(σ)e(\sigma) denote the number of “errors” in the ordering (σ(1),σ(2),,σ(n))(\sigma(1),\sigma(2),\cdots,\sigma(n)), that is, e(σ)=#{j<n:σ(j)>σ(j+1)}e(\sigma)=\#\{j<n:\sigma(j)>\sigma(j+1)\}. Then

(4.1) gt=n=1rσ𝒮n((1)e(σ)/n2[n1e(σ)])×Δn(t)[[B˙sσ(1),B˙sσ(2)],,]B˙sσ(n)]ds,g_{t}=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}\left((-1)^{e(\sigma)}\bigg{/}n^{2}\begin{bmatrix}n-1\\ e(\sigma)\end{bmatrix}\right)\times\\ \int_{\Delta_{n}(t)}[\cdots[\dot{B}_{s_{\sigma(1)}},\dot{B}_{s_{\sigma(2)}}],\ldots,]\dot{B}_{s_{\sigma(n)}}]\,ds,

where the n=1n=1 term is understood to be 0t𝑑Bs=Bt\int_{0}^{t}dB_{s}=B_{t}. Using this as our motivation, we first explore stochastic integral analogues of equation (4.1) where the smooth curve BB is replaced by Brownian motion on 𝔤\mathfrak{g}.

4.1. Brownian motion and finite-dimensional approximations

We now return to the setting of an abstract Wiener space (𝔤,𝔤CM,μ)(\mathfrak{g},\mathfrak{g}_{CM},\mu) endowed with a nilpotent Hilbert-Schmidt Lie bracket on 𝔤CM\mathfrak{g}_{CM}. Again, let BtB_{t} denote Brownian motion on 𝔤\mathfrak{g}. By equation (4.1), the solution to the Stratonovich stochastic differential equation

δgt=LgtδBt, with g0=𝐞,\delta g_{t}=L_{g_{t}*}\delta B_{t},\quad\text{ with }g_{0}=\mathbf{e},

should be given by

(4.2) gt=n=1rσ𝒮ncnσΔn(t)[[[δBsσ(1),δBsσ(2)],],δBsσ(n)],g_{t}=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}c^{\sigma}_{n}\int_{\Delta_{n}(t)}[[\cdots[\delta B_{s_{\sigma(1)}},\delta B_{s_{\sigma(2)}}],\cdots],\delta B_{s_{\sigma(n)}}],

for coefficients cnσc_{n}^{\sigma} determined by equation (4.1).

To understand the integrals in (4.2), consider the following heuristic computation. Let {Mn(t)}t0\{M_{n}(t)\}_{t\geq 0} denote the process in 𝔤n\mathfrak{g}^{\otimes n} defined by

Mn(t):=Δn(t)δBs1δBsn.M_{n}(t):=\int_{\Delta_{n}(t)}\delta B_{s_{1}}\otimes\cdots\otimes\delta B_{s_{n}}.

By repeatedly applying the definition of the Stratonovich integral, the iterated Stratonovich integral Mn(t)M_{n}(t) may be realized as a linear combination of iterated Itô integrals:

Mn(t)=m=n/2n12nmα𝒥nmItn(α),M_{n}(t)=\sum_{m=\lceil n/2\rceil}^{n}\frac{1}{2^{n-m}}\sum_{\alpha\in\mathcal{J}_{n}^{m}}I^{n}_{t}(\alpha),

where

𝒥nm:={(α1,,αm){1,2}m:i=1mαi=n},\mathcal{J}_{n}^{m}:=\left\{(\alpha_{1},\ldots,\alpha_{m})\in\{1,2\}^{m}:\sum_{i=1}^{m}\alpha_{i}=n\right\},

and, for α𝒥nm\alpha\in\mathcal{J}_{n}^{m}, Itn(α)I_{t}^{n}(\alpha) is the iterated Itô integral

Itn(α)=Δm(t)𝑑Xs11𝑑XsmmI_{t}^{n}(\alpha)=\int_{\Delta_{m}(t)}dX^{1}_{s_{1}}\otimes\cdots\otimes dX^{m}_{s_{m}}

with

dXsi={dBsif αi=1j=1hjhjdsif αi=2;dX_{s}^{i}=\left\{\begin{array}[]{cl}dB_{s}&\text{if }\alpha_{i}=1\\ \sum_{j=1}^{\infty}h_{j}\otimes h_{j}\,ds&\text{if }\alpha_{i}=2\end{array}\right.;

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 F1:𝔤CM𝔤CMF_{1}:\mathfrak{g}_{CM}\to\mathfrak{g}_{CM} by F1(k)=kF_{1}(k)=k, and for n{2,,r}n\in\{2,\cdots,r\} define Fn:𝔤CMn𝔤CMF_{n}:\mathfrak{g}_{CM}^{\otimes n}\rightarrow\mathfrak{g}_{CM} by

(4.3) Fn(k1kn):=[[[[k1,k2],k3],],kn].F_{n}(k_{1}\otimes\cdots\otimes k_{n}):=[[[\cdots[k_{1},k_{2}],k_{3}],\cdots],k_{n}].

For each fixed nn and σ𝒮n\sigma\in\mathcal{S}_{n}, define Fnσ:𝔤CMn𝔤CMF_{n}^{\sigma}:\mathfrak{g}_{CM}^{\otimes n}\rightarrow\mathfrak{g}_{CM} by

(4.4) Fnσ(k1kn):=Fn(kσ(1)kσ(n))=[[[kσ(1),kσ(2)],],kσ(n)].\begin{split}F_{n}^{\sigma}(k_{1}\otimes\cdots\otimes k_{n})&:=F_{n}(k_{\sigma(1)}\otimes\cdots\otimes k_{\sigma(n)})\\ &=[[\cdots[k_{\sigma(1)},k_{\sigma(2)}],\cdots],k_{\sigma(n)}].\end{split}

Then we may write

gt=n=1rσ𝒮ncnσFnσ(Mn(t))=n=1rσ𝒮nm=n/2ncnσ2nmα𝒥nmFnσ(Itn(α)),g_{t}=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}c^{\sigma}_{n}F^{\sigma}_{n}(M_{n}(t))=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}\sum_{m=\lceil n/2\rceil}^{n}\frac{c^{\sigma}_{n}}{2^{n-m}}\sum_{\alpha\in\mathcal{J}_{n}^{m}}F^{\sigma}_{n}(I^{n}_{t}(\alpha)),

presuming we can make sense of the integrals Fnσ(Itn(α))F_{n}^{\sigma}(I_{t}^{n}(\alpha)).

For each α\alpha, let pα=#{i:αi=1}p_{\alpha}=\#\{i:\alpha_{i}=1\} and qα=#{i:αi=2}q_{\alpha}=\#\{i:\alpha_{i}=2\} (so that pα+qα=mp_{\alpha}+q_{\alpha}=m when α𝒥nm\alpha\in\mathcal{J}_{n}^{m}), and let

𝒥n:=m=n/2n𝒥nm.\mathcal{J}_{n}:=\bigcup_{m=\lceil n/2\rceil}^{n}\mathcal{J}_{n}^{m}.

Then, for each σ𝒮n\sigma\in\mathcal{S}_{n} and α𝒥n\alpha\in\mathcal{J}_{n},

Fnσ(Itn(α))=Δpα(t)fα(s,t)F^nσ,α(dBs1dBspα),F_{n}^{\sigma}(I_{t}^{n}(\alpha))=\int_{\Delta_{p_{\alpha}}(t)}f_{\alpha}(s,t)\hat{F}_{n}^{\sigma,\alpha}(dB_{s_{1}}\otimes\cdots\otimes dB_{s_{p_{\alpha}}}),

where F^nσ,α\hat{F}_{n}^{\sigma,\alpha} and fαf_{\alpha} are as follows.

The map F^nσ,α:𝔤pα𝔤\hat{F}_{n}^{\sigma,\alpha}:\mathfrak{g}^{\otimes p_{\alpha}}\rightarrow\mathfrak{g} is defined by

(4.5) F^nσ,α(k1kpα):=j1,,jqα=1Fnσ(k1kpαhj1hj1hjqαhjqα),\hat{F}_{n}^{\sigma,\alpha}(k_{1}\otimes\cdots\otimes k_{p_{\alpha}})\\ :=\sum_{j_{1},\ldots,j_{q_{\alpha}}=1}^{\infty}F_{n}^{\sigma^{\prime}}(k_{1}\otimes\cdots\otimes k_{p_{\alpha}}\otimes h_{j_{1}}\otimes h_{j_{1}}\otimes\cdots\otimes h_{j_{q_{\alpha}}}\otimes h_{j_{q_{\alpha}}}),

for {hj}j=1\{h_{j}\}_{j=1}^{\infty} an orthonormal basis of 𝔤CM\mathfrak{g}_{CM} and σ=σ(α)𝒮n\sigma^{\prime}=\sigma^{\prime}(\alpha)\in\mathcal{S}_{n} given by σ=στ1\sigma^{\prime}=\sigma\circ\tau^{-1}, for any τ𝒮n\tau\in\mathcal{S}_{n} such that

τ(dXs11dXsmm)=j1,,jqα=1dBs1dBspαhj1hj1hjqαhjqαds1dsqα.\tau(dX^{1}_{s_{1}}\otimes\cdots\otimes dX^{m}_{s_{m}})\\ =\sum_{j_{1},\cdots,j_{q_{\alpha}}=1}^{\infty}dB_{s_{1}}\otimes\cdots\otimes dB_{s_{p_{\alpha}}}\otimes h_{j_{1}}\otimes h_{j_{1}}\otimes\cdots\otimes h_{j_{q_{\alpha}}}\otimes h_{j_{q_{\alpha}}}ds_{1}\cdots ds_{q_{\alpha}}.

To define fαf_{\alpha}, first consider the polynomial of order qαq_{\alpha}, in the variables sis_{i} with ii such that αi=1\alpha_{i}=1 and in the variable tt, given by evaluating the integral

(4.6) fα((si:αi=1),t)=Δqα(t)i:αi=2dsi,f_{\alpha}^{\prime}((s_{i}:\alpha_{i}=1),t)=\int_{\Delta^{\prime}_{q_{\alpha}}(t)}\prod_{i:\alpha_{i}=2}ds_{i},

where Δqα(t)={si1<si<si+1:αi=2}\Delta^{\prime}_{q_{\alpha}}(t)=\{s_{i-1}<s_{i}<s_{i+1}:\alpha_{i}=2\} with s0=0s_{0}=0 and sm+1=ts_{m+1}=t. Then fαf_{\alpha} is fαf_{\alpha}^{\prime} with the variables reindexed by the bijection {i:αi=1}{1,,pα}\{i:\alpha_{i}=1\}\rightarrow\{1,\ldots,p_{\alpha}\} that maintains the natural ordering of these sets. (For example, for α=(1,2,1,2)𝒥64\alpha=(1,2,1,2)\in\mathcal{J}_{6}^{4},

fα(s1,s3,t)={s1<s2<s3,s3<s4<t}𝑑s2𝑑s4=(ts3)(s3s1),f_{\alpha}^{\prime}(s_{1},s_{3},t)=\int_{\{s_{1}<s_{2}<s_{3},s_{3}<s_{4}<t\}}ds_{2}\,ds_{4}=(t-s_{3})(s_{3}-s_{1}),

so that fα(s1,s2,t)=(ts2)(s2s1)f_{\alpha}(s_{1},s_{2},t)=(t-s_{2})(s_{2}-s_{1}).)

This explicit realization of fαf_{\alpha} is not critical to the sequel. It is really only necessary to know that fαf_{\alpha} is a polynomial of order qαq_{\alpha} in s=(s1,,spα)s=(s_{1},\ldots,s_{p_{\alpha}}) and tt, and thus may be written as

fα(s,t)=a=0qαbαataf~α,a(s),f_{\alpha}(s,t)=\sum_{a=0}^{q_{\alpha}}b_{\alpha}^{a}t^{a}\tilde{f}_{\alpha,a}(s),

for some coefficients bαab_{\alpha}^{a}\in\mathbb{R} and polynomials f~α,a\tilde{f}_{\alpha,a} of degree qαaq_{\alpha}-a in ss. Now, if F^nσ,α\hat{F}_{n}^{\sigma,\alpha} is Hilbert-Schmidt on 𝔤CMpα\mathfrak{g}_{CM}^{\otimes p_{\alpha}}, then

Δpα(t)f~α,a(s)F^nσ,α22𝑑s=f~α,aL2(Δpα(t))2F^nσ,α22<,\int_{\Delta_{p_{\alpha}}(t)}\left\|\tilde{f}_{\alpha,a}(s)\hat{F}_{n}^{\sigma,\alpha}\right\|_{2}^{2}\,ds=\left\|\tilde{f}_{\alpha,a}\right\|^{2}_{L^{2}(\Delta_{p_{\alpha}}(t))}\left\|\hat{F}_{n}^{\sigma,\alpha}\right\|_{2}^{2}<\infty,

and

(4.7) Fnσ(Itn(α))\displaystyle F_{n}^{\sigma}(I_{t}^{n}(\alpha)) =a=0qαbαataJn(f~α,aF^nσ,α)t\displaystyle=\sum_{a=0}^{q_{\alpha}}b_{\alpha}^{a}t^{a}J_{n}(\tilde{f}_{\alpha,a}\hat{F}_{n}^{\sigma,\alpha})_{t}

may be understood in the sense of the limit integrals in Proposition 2.5. (In particular, if αm=1\alpha_{m}=1, then fα=fα(s)f_{\alpha}=f_{\alpha}(s) does not depend on tt, and Proposition 2.5 implies that Fnσ(Itn(α))F_{n}^{\sigma}(I_{t}^{n}(\alpha)) is a 𝔳\mathfrak{v}-valued L2L^{2}-martingale.)

The above computations show that, if for all nn, σ𝒮n\sigma\in\mathcal{S}_{n}, and α𝒥n\alpha\in\mathcal{J}_{n}, F^nσ,α\hat{F}_{n}^{\sigma,\alpha} is Hilbert-Schmidt, then we may rewrite (4.2) as

gt=n=1rσ𝒮nm=n/2ncnσ2nmα𝒥nma=0qαbαataJn(f~α,aF^nσ,α)t,g_{t}=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}\sum_{m=\lceil n/2\rceil}^{n}\frac{c^{\sigma}_{n}}{2^{n-m}}\sum_{\alpha\in\mathcal{J}_{n}^{m}}\sum_{a=0}^{q_{\alpha}}b_{\alpha}^{a}t^{a}J_{n}(\tilde{f}_{\alpha,a}\hat{F}_{n}^{\sigma,\alpha})_{t},

where JnJ_{n} is as defined in Proposition 2.5. The next proposition shows that F^nσ,α\hat{F}_{n}^{\sigma,\alpha} is Hilbert-Schmidt as desired, and thus gtg_{t} in (4.2) is well-defined.

Proposition 4.1.

Let n{2,,r}n\in\{2,\ldots,r\}, σ𝒮n\sigma\in\mathcal{S}_{n}, and α𝒥n\alpha\in\mathcal{J}_{n}. Then F^nσ,α:𝔤CMpα𝔤CM\hat{F}_{n}^{\sigma,\alpha}:\mathfrak{g}_{CM}^{\otimes p_{\alpha}}\rightarrow\mathfrak{g}_{CM} is Hilbert-Schmidt.

Proof.

For the whole of this proof, all sums will be taken over an orthonormal basis of 𝔤CM\mathfrak{g}_{CM}.

Now, for FnF_{n} and FnσF_{n}^{\sigma} as defined in equations (4.3) and (4.4), we may write

Fnσ(k1kpαh1h1hqαhqα)=Fn(Aσ(1)Aσ(n))\displaystyle F_{n}^{\sigma}(k_{1}\otimes\cdots\otimes k_{p_{\alpha}}\otimes h_{1}\otimes h_{1}\otimes\cdots\otimes h_{q_{\alpha}}\otimes h_{q_{\alpha}})=F_{n}(A_{\sigma(1)}\otimes\ldots\otimes A_{\sigma(n)})

where

Ab:={kbif b=1,,pαh(bpα)/2if b=pα+1,,n.A_{b}:=\left\{\begin{array}[]{ll}k_{b}&\text{if }b=1,\ldots,p_{\alpha}\\ h_{\lceil(b-p_{\alpha})/2\rceil}&\text{if }b=p_{\alpha}+1,\ldots,n\end{array}\right..

If qα=0q_{\alpha}=0, then pα=np_{\alpha}=n, each Aσ(j)=kiA_{\sigma(j)}=k_{i} for some i=1,,ni=1,\ldots,n, and

F^nσ,α22\displaystyle\|\hat{F}_{n}^{\sigma,\alpha}\|_{2}^{2} =k1,,knFnσ(k1kn)𝔤CM2\displaystyle=\sum_{k_{1},\ldots,k_{n}}\|F_{n}^{\sigma}(k_{1}\otimes\cdots\otimes k_{n})\|_{\mathfrak{g}_{CM}}^{2}
=k1,,knFn(k1kn)𝔤CM2=[[[[,],],](𝔤CM)n𝔤CM2\displaystyle=\sum_{k_{1},\ldots,k_{n}}\|F_{n}(k_{1}\otimes\cdots\otimes k_{n})\|_{\mathfrak{g}_{CM}}^{2}=\|[[[\ldots[\cdot,\cdot],\ldots],\cdot]\|_{(\mathfrak{g}_{CM}^{*})^{\otimes n}\otimes\mathfrak{g}_{CM}}^{2}

which is finite by Proposition 3.6.

Now, if qα=1q_{\alpha}=1, let N=N(σ)N=N(\sigma) denote the second jj such that σ(j){n1,n}\sigma(j)\in\{n-1,n\}; that is,

Fnσ(k1\displaystyle F_{n}^{\sigma}(k_{1}\otimes\cdots\otimes kn1h1h1)\displaystyle k_{n-1}\otimes h_{1}\otimes h_{1})
=Fn(Aσ(1)Aσ(N1)h1Aσ(N+1)Aσ(n))\displaystyle=F_{n}(A_{\sigma(1)}\otimes\cdots\otimes A_{\sigma(N-1)}\otimes h_{1}\otimes A_{\sigma(N+1)}\otimes\ldots\otimes A_{\sigma(n)})
=[[[[[[Aσ(1),Aσ(2)],],Aσ(N1)],h1],Aσ(N+1)],],Aσ(n)]\displaystyle=[[[[[\ldots[A_{\sigma(1)},A_{\sigma(2)}],\ldots],A_{\sigma(N-1)}],h_{1}],A_{\sigma(N+1)}],\ldots],A_{\sigma(n)}]

where exactly one of Aσ(1),,Aσ(N1)A_{\sigma(1)},\ldots,A_{\sigma(N-1)} is h1h_{1} and, for all j>Nj>N, Aσ(j)=kiA_{\sigma(j)}=k_{i} for some

iI:=I(σ):={σ(j):j=N+1,,n}{1,,pα}={1,,n2}.i\in I:=I(\sigma):=\{\sigma(j):j=N+1,\ldots,n\}\subseteq\{1,\ldots,p_{\alpha}\}=\{1,\ldots,n-2\}.

Thus, writing

𝒜(h1,ki:iIc):=[[[Aσ(1),Aσ(2)],],Aσ(N1)],\mathcal{A}(h_{1},k_{i}:i\in I^{c}):=[[\ldots[A_{\sigma(1)},A_{\sigma(2)}],\ldots],A_{\sigma(N-1)}],

we have that

Fnσ(k1kn1h1h1)=e1𝒜(h1,ki:iIc),e1𝔤CM[[[e1,h1],Aσ(N+1)],,Aσ(n)],F_{n}^{\sigma}(k_{1}\otimes\cdots\otimes k_{n-1}\otimes h_{1}\otimes h_{1})\\ =\sum_{e_{1}}\,\langle\mathcal{A}(h_{1},k_{i}:i\in I^{c}),e_{1}\rangle_{\mathfrak{g}_{CM}}[[\ldots[e_{1},h_{1}],A_{\sigma(N+1)}],\ldots,A_{\sigma(n)}],

and so

F^nσ,α22\displaystyle\|\hat{F}_{n}^{\sigma,\alpha}\|_{2}^{2} =k1,,kn1h1Fnσ(k1kn1h1h1)𝔤CM2\displaystyle=\sum_{k_{1},\ldots,k_{n-1}}\left\|\sum_{h_{1}}\,F_{n}^{\sigma}(k_{1}\otimes\cdots\otimes k_{n-1}\otimes h_{1}\otimes h_{1})\right\|_{\mathfrak{g}_{CM}}^{2}
k1,,kn1(h1,e1|𝒜(h1,ki:iIc),e1𝔤CM|2)\displaystyle\leq\sum_{k_{1},\ldots,k_{n-1}}\left(\sum_{h_{1},e_{1}}|\langle\mathcal{A}(h_{1},k_{i}:i\in I^{c}),e_{1}\rangle_{\mathfrak{g}_{CM}}|^{2}\right)
×(h1,e1[[[e1,h1],Aσ(N+1)],,Aσ(n)]𝔤CM2)\displaystyle\qquad\qquad\qquad\quad\times\left(\sum_{h_{1},e_{1}}\|[[[e_{1},h_{1}],A_{\sigma(N+1)}],\ldots,A_{\sigma(n)}]\|_{\mathfrak{g}_{CM}}^{2}\right)
=(ki:iIc,h1,e1|𝒜(h1,ki:iIc),e1𝔤CM|2)\displaystyle=\left(\sum_{k_{i}:i\in I^{c},h_{1},e_{1}}|\langle\mathcal{A}(h_{1},k_{i}:i\in I^{c}),e_{1}\rangle_{\mathfrak{g}_{CM}}|^{2}\right)
×(ki:iI,h1,e1[[[[e1,h1],Aσ(N+1)],],Aσ(n)]𝔤CM2)\displaystyle\qquad\qquad\qquad\quad\times\left(\sum_{k_{i}:i\in I,h_{1},e_{1}}\|[[\ldots[[e_{1},h_{1}],A_{\sigma(N+1)}],\ldots],A_{\sigma(n)}]\|_{\mathfrak{g}_{CM}}^{2}\right)
[[[,],],](𝔤CM)N1𝔤CM2[[[,],],](𝔤CM)nN+1𝔤CM2.\displaystyle\leq\|[[\ldots[\cdot,\cdot],\ldots],\cdot]\|_{(\mathfrak{g}_{CM}^{*})^{\otimes N-1}\otimes\mathfrak{g}_{CM}}^{2}\|[[\ldots[\cdot,\cdot],\ldots],\cdot]\|_{(\mathfrak{g}_{CM}^{*})^{\otimes n-N+1}\otimes\mathfrak{g}_{CM}}^{2}.

Now more generally when qα2q_{\alpha}\geq 2, we may similarly “separate” the pairs of hjh_{j}’s as above. More precisely, define

Φ(b):=Φα(b):={bif b=1,,pαbpα2+pαif b=pα+1,,n.\Phi(b):=\Phi_{\alpha}(b):=\left\{\begin{array}[]{ll}b&\text{if }b=1,\ldots,p_{\alpha}\\ \lceil\frac{b-p_{\alpha}}{2}\rceil+p_{\alpha}&\text{if }b=p_{\alpha}+1,\ldots,n\end{array}\right..

Let N0=1N_{0}=1, and set

Ω1j:={Φ(σ()):=N0,,j1} and N1:=min{j>N0:Φ(σ(j))Ω1j},\Omega_{1}^{j}:=\{\Phi(\sigma(\ell)):\ell=N_{0},\ldots,j-1\}\quad\text{ and }\quad N_{1}:=\min\{j>N_{0}:\Phi(\sigma(j))\in\Omega_{1}^{j}\},
Ω2j:={Φ(σ()):=N1,,j1} and N2:=min{j>N1:Φ(σ(j))Ω2j}.\Omega_{2}^{j}:=\{\Phi(\sigma(\ell)):\ell=N_{1},\ldots,j-1\}\quad\text{ and }\quad N_{2}:=\min\{j>N_{1}:\Phi(\sigma(j))\in\Omega_{2}^{j}\}.

Similarly, we define

Ω2m+1j:={Φ(σ()):=N2m,,j1},\Omega_{2m+1}^{j}:=\{\Phi(\sigma(\ell)):\ell=N_{2m},\ldots,j-1\},
N2m+1:=min{j>N2m:Φ(σ(j))i=0m1Ω2i+1N2i+1Ω2m+1j},N_{2m+1}:=\min\left\{j>N_{2m}:\Phi(\sigma(j))\in\bigcup_{i=0}^{m-1}\Omega_{2i+1}^{N_{2i+1}}\cup\Omega_{2m+1}^{j}\right\},
Ω2mj:={Φ(σ()):=N2m1,,j1}, and\Omega_{2m}^{j}:=\{\Phi(\sigma(\ell)):\ell=N_{2m-1},\ldots,j-1\},\text{ and}
N2m:=min{j>N2m1:Φ(σ(j))i=1m1Ω2mN2mΩ2mj}.N_{2m}:=\min\left\{j>N_{2m-1}:\Phi(\sigma(j))\in\bigcup_{i=1}^{m-1}\Omega_{2m}^{N_{2m}}\cup\Omega_{2m}^{j}\right\}.

Then there is an M<qαM<q_{\alpha} such that the sets {Aσ(Ni),,Aσ(Ni+11)}i=0M1\{A_{\sigma(N_{i})},\ldots,A_{\sigma(N_{i+1}-1)}\}_{i=0}^{M-1} and {Aσ(NM),,Aσ(n)}\{A_{\sigma(N_{M})},\ldots,A_{\sigma(n)}\} separate the hjh_{j}’s in the sense that no hjh_{j} is repeated inside any one of these sets, and moreover the union of “even” sets contains exactly one copy of each hjh_{j} and similarly with the “odd” sets. We can write

Fnσ\displaystyle F_{n}^{\sigma} (k1kpαh1h1hqαhqα)\displaystyle(k_{1}\otimes\cdots\otimes k_{p_{\alpha}}\otimes h_{1}\otimes h_{1}\otimes\cdots\otimes h_{q_{\alpha}}\otimes h_{q_{\alpha}})
=e1,,eM{e1,[[[Aσ(1),Aσ(2)],],Aσ(N11)]\displaystyle=\sum_{e_{1},\ldots,e_{M}}\bigg{\{}\langle e_{1},[[[A_{\sigma(1)},A_{\sigma(2)}],\ldots],A_{\sigma(N_{1}-1)}]\rangle
×(i=1M1ei+1,[[[ei,Aσ(Ni)],],Aσ(Ni+11)])[[[eM,Aσ(NM)],],Aσ(n)]}.\displaystyle\qquad\times\left(\prod_{i=1}^{M-1}\langle e_{i+1},[[[e_{i},A_{\sigma(N_{i})}],\ldots],A_{\sigma(N_{i+1}-1)}]\rangle\right)[[[e_{M},A_{\sigma(N_{M})}],\ldots],A_{\sigma(n)}]\bigg{\}}.

If MM is even, then

𝒜=e1,[[[Aσ(1),Aσ(2)],],Aσ(N11)]×(i=1M/21e2i+1,[[[e2i,Aσ(N2i)],],Aσ(N2i+1)1])[[[eM,Aσ(NM)],],Aσ(n)]\mathcal{A}=\langle e_{1},[[[A_{\sigma(1)},A_{\sigma(2)}],\ldots],A_{\sigma(N_{1}-1)}]\rangle\\ \times\left(\prod_{i=1}^{M/2-1}\langle e_{2i+1},[[[e_{2i},A_{\sigma(N_{2i})}],\ldots],A_{\sigma(N_{2i+1})-1}]\rangle\right)[[[e_{M},A_{\sigma(N_{M})}],\ldots],A_{\sigma(n)}]

is a function of h1,,hqα,e1,,eMh_{1},\ldots,h_{q_{\alpha}},e_{1},\ldots,e_{M} and kik_{i} for iIi\in I some subset of {1,,n}\{1,\ldots,n\}, and

=i=1M/2e2i,[[[e2i1,Aσ(N2i1)],],Aσ(N2i)1]\mathcal{B}=\prod_{i=1}^{M/2}\langle e_{2i},[[[e_{2i-1},A_{\sigma(N_{2i-1})}],\ldots],A_{\sigma(N_{2i})-1}]\rangle

is a function of h1,,hqα,e1,,eMh_{1},\ldots,h_{q_{\alpha}},e_{1},\ldots,e_{M} and kik_{i} for iIci\in I^{c}. Thus,

F^nσ,α22(ki:iI,h1,,hqα,e1,,eM𝒜𝔤CM2)(ki:iIc,h1,,hqα,e1,,eM||2)\|\hat{F}_{n}^{\sigma,\alpha}\|_{2}^{2}\leq\left(\sum_{k_{i}:i\in I,h_{1},\ldots,h_{q_{\alpha}},e_{1},\ldots,e_{M}}\|\mathcal{A}\|_{\mathfrak{g}_{CM}}^{2}\right)\left(\sum_{k_{i}:i\in I^{c},h_{1},\ldots,h_{q_{\alpha}},e_{1},\ldots,e_{M}}|\mathcal{B}|^{2}\right)

which is finite again by Proposition 3.6. Similarly, if MM is odd,

𝒜=e1,[[[Aσ(1),Aσ(2)],],Aσ(N11)]×i=1(M1)/2e2i+1,[[[e2i,Aσ(N2i)],],Aσ(N2i+1)1]\mathcal{A}=\langle e_{1},[[[A_{\sigma(1)},A_{\sigma(2)}],\ldots],A_{\sigma(N_{1}-1)}]\rangle\\ \times\prod_{i=1}^{(M-1)/2}\langle e_{2i+1},[[[e_{2i},A_{\sigma(N_{2i})}],\ldots],A_{\sigma(N_{2i+1})-1}]\rangle

and

=(i=1(M1)/2e2i,[[[e2i1,Aσ(N2i1)],],Aσ(N2i)1])[[[eM,Aσ(NM)],],Aσ(n)]\mathcal{B}=\left(\prod_{i=1}^{(M-1)/2}\langle e_{2i},[[[e_{2i-1},A_{\sigma(N_{2i-1})}],\ldots],A_{\sigma(N_{2i})-1}]\rangle\right)[[[e_{M},A_{\sigma(N_{M})}],\ldots],A_{\sigma(n)}]

are both functions of h1,,hqα,e1,,eMh_{1},\ldots,h_{q_{\alpha}},e_{1},\ldots,e_{M} and some kik_{i}, and

F^nσ,α22(ki:iI,h1,,hqα,e1,,eM|𝒜|𝔤CM2)(ki:iIc,h1,,hqα,e1,,eM𝔤CM2)\|\hat{F}_{n}^{\sigma,\alpha}\|_{2}^{2}\leq\left(\sum_{k_{i}:i\in I,h_{1},\ldots,h_{q_{\alpha}},e_{1},\ldots,e_{M}}|\mathcal{A}|_{\mathfrak{g}_{CM}}^{2}\right)\left(\sum_{k_{i}:i\in I^{c},h_{1},\ldots,h_{q_{\alpha}},e_{1},\ldots,e_{M}}\|\mathcal{B}\|_{\mathfrak{g}_{CM}}^{2}\right)

is again finite by Proposition 3.6 and this completes the proof. ∎

Propositions 2.5 and 4.1 now allow us to make the following definition.

Definition 4.2.

A Brownian motion on GG is the continuous GG-valued process defined by

gt=n=1rσ𝒮nm=n/2ncnσ2nmα𝒥nmΔpα(t)fα(s,t)F^nσ,α(dBs1dBspα),g_{t}=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}\sum_{m=\lceil n/2\rceil}^{n}\frac{c^{\sigma}_{n}}{2^{n-m}}\sum_{\alpha\in\mathcal{J}_{n}^{m}}\int_{\Delta_{p_{\alpha}}(t)}f_{\alpha}(s,t)\hat{F}_{n}^{\sigma,\alpha}(dB_{s_{1}}\otimes\cdots\otimes dB_{s_{p_{\alpha}}}),

where

cnσ=(1)e(σ)/n2[n1e(σ)],c_{n}^{\sigma}=(-1)^{e(\sigma)}\bigg{/}n^{2}\begin{bmatrix}n-1\\ e(\sigma)\end{bmatrix},

F^nσ,α\hat{F}_{n}^{\sigma,\alpha} is as defined in equation (4.5), and fαf_{\alpha} is as defined below equation (4.6). For t>0t>0, let νt=Law(gt)\nu_{t}=\mathrm{Law}(g_{t}) be the heat kernel measure at time tt, a probability measure on GG.

Proposition 4.3 (Finite-dimensional approximations).

For GπG_{\pi} a finite-dimensional Lie subgroup of GCMG_{CM}, let π\pi denote orthogonal projection of GCMG_{CM} onto GπG_{\pi} and let gtπg^{\pi}_{t} be the continuous process on GπG_{\pi} defined by

gtπ=n=1rσ𝒮nm=n/2ncnσ2nmα𝒥nmΔpα(t)fα(s,t)F^nσ,α(dπBs1dπBspα),g_{t}^{\pi}=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}\sum_{m=\lceil n/2\rceil}^{n}\frac{c^{\sigma}_{n}}{2^{n-m}}\sum_{\alpha\in\mathcal{J}_{n}^{m}}\int_{\Delta_{p_{\alpha}}(t)}f_{\alpha}(s,t)\hat{F}_{n}^{\sigma,\alpha}(d\pi B_{s_{1}}\otimes\cdots\otimes d\pi B_{s_{p_{\alpha}}}),

where the stochastic integrals are defined as in Proposition 2.6. Then gtπg_{t}^{\pi} is Brownian motion on GπG_{\pi}. In particular, for G=GπG_{\ell}=G_{\pi_{\ell}} an increasing sequence of finite-dimensional Lie subgroups such that the associated orthogonal projections π\pi_{\ell} are increasing to I𝔤CMI_{\mathfrak{g}_{CM}}, let gt=gtπg^{\ell}_{t}=g_{t}^{\pi_{\ell}}. Then, for all t<t<\infty,

(4.8) lim𝔼gtgt𝔤2=0.\lim_{\ell\rightarrow\infty}\mathbb{E}\left\|g^{\ell}_{t}-g_{t}\right\|_{\mathfrak{g}}^{2}=0.
Proof.

First note that gtπg_{t}^{\pi} solves the Stratonovich equation δgtπ=LgtπδπBt\delta g_{t}^{\pi}=L_{g_{t}^{\pi}*}\delta\pi B_{t} with g0π=𝐞g_{0}^{\pi}=\mathbf{e}, see [9, 11, 6] where πBt\langle\pi B\rangle_{t} is a standard 𝔤π\mathfrak{g}_{\pi}-valued Brownian motion. Thus, gtπg_{t}^{\pi} is a GπG_{\pi}-valued Brownian motion.

By equation (4.7) and its preceding discussion,

gt=n=1rσ𝒮nm=n/2ncnσ2nmα𝒥nma=0qαbαataJn(f~αF^nσ,α)t,g_{t}^{\ell}=\sum_{n=1}^{r}\sum_{\sigma\in\mathcal{S}_{n}}\sum_{m=\lceil n/2\rceil}^{n}\frac{c^{\sigma}_{n}}{2^{n-m}}\sum_{\alpha\in\mathcal{J}_{n}^{m}}\sum_{a=0}^{q_{\alpha}}b_{\alpha}^{a}t^{a}J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t},

and thus, to verify (4.8), it suffices to show that

lim𝔼πBtBt𝔤2=0\lim_{\ell\rightarrow\infty}\mathbb{E}\|\pi_{\ell}B_{t}-B_{t}\|_{\mathfrak{g}}^{2}=0

and

lim𝔼Jn(f~αF^nσ,α)tJn(f~αF^nσ,α)t2=0,\lim_{\ell\rightarrow\infty}\mathbb{E}\left\|J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}-J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}\right\|^{2}=0,

for all n{2,,r}n\in\{2,\ldots,r\}, σ𝒮n\sigma\in\mathcal{S}_{n} and α𝒥n\alpha\in\mathcal{J}_{n}.

So let μt=Law(Bt)\mu_{t}=\mathrm{Law}(B_{t}). Then it is known that, if VV is a finite-dimensional subspace of 𝔤CM\mathfrak{g}_{CM} and πV\pi_{V} is the orthogonal projection from 𝔤CM\mathfrak{g}_{CM} to VV, then πV\pi_{V} admits a μt\mu_{t}-a.s. unique extension to 𝔤\mathfrak{g}. Moreover, if VnV_{n} is an increasing sequence of finite-dimensional subspaces, then

limn𝔼πVnBtBt𝔤2=0;\lim_{n\rightarrow\infty}\mathbb{E}\|\pi_{V_{n}}B_{t}-B_{t}\|_{\mathfrak{g}}^{2}=0;

see for example Section 8.3.3 of [37].

By Proposition 4.1, F^nσ,α\hat{F}_{n}^{\sigma,\alpha} is Hilbert-Schmidt, and recall that f~α\tilde{f}_{\alpha} is a deterministic polynomial function in ss. Thus Jn(f~αF^nσ,α)J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha}) and Jn(f~αF^nσ,α)J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha}) are 𝔤CM\mathfrak{g}_{CM}-valued martingales as defined in Proposition 2.6, and Proposition 2.6 gives the desired convergence as well (in 𝔤CM\mathfrak{g}_{CM} and thus in 𝔤\mathfrak{g}). ∎

Remark 4.4.

In fact, for each of the stochastic integrals Jn(f~αF^nσ,α)J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha}), it is possible to prove the stronger convergence that, for all p[1,)p\in[1,\infty),

lim𝔼[supτtJn(f~αF^nσ,α)τJn(f~αF^nσ,α)τp]=0,\lim_{\ell\rightarrow\infty}\mathbb{E}\left[\sup_{\tau\leq t}\left\|J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{\tau}-J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{\tau}\right\|^{p}\right]=0,

for all n{2,,r}n\in\{2,\ldots,r\}, σ𝒮n\sigma\in\mathcal{S}_{n} and α𝒥n\alpha\in\mathcal{J}_{n}. Again, Proposition 2.6 gives the limit for p=2p=2 and thus for p[1,2]p\in[1,2]. For p>2p>2, Doob’s maximal inequality implies it suffices to show that

lim𝔼Jn(f~αF^nσ,α)tJn(f~αF^nσ,α)tp=0.\lim_{\ell\rightarrow\infty}\mathbb{E}\left\|J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}-J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}\right\|^{p}=0.

Since each Jn(f~αF^nσ,α)J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha}) and Jn(f~αF^nσ,α)J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha}) has chaos expansion terminating at degree nn, a theorem of Nelson (see Lemma 2 of [34] and pp. 216-217 of [33]) implies that, for each jj\in\mathbb{N}, there exists cj<c_{j}<\infty such that

𝔼Jn(f~αF^nσ,α)tJn(f~αF^nσ,α)t2jcj(𝔼Jn(f~αF^nσ,α)tJn(f~αF^nσ,α)t2)j.\mathbb{E}\left\|J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}-J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}\right\|^{2j}\leq c_{j}\left(\mathbb{E}\left\|J_{n}^{\ell}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}-J_{n}(\tilde{f}_{\alpha}\hat{F}_{n}^{\sigma,\alpha})_{t}\right\|^{2}\right)^{j}.

In a similar way, one may prove the following convergence for the Brownian motions under right translations by elements of GCMG_{CM}.

Proposition 4.5.

For any yGCMy\in G_{CM},

lim𝔼gtygty𝔤2=0.\lim_{\ell\rightarrow\infty}\mathbb{E}\|g_{t}^{\ell}y-g_{t}y\|^{2}_{\mathfrak{g}}=0.

where gtyg_{t}y is the measurable right group action of yGCMy\in G_{CM} on gtGg_{t}\in G, as in Proposition 3.4.

Remark 4.6.

Note that, while the present paper focuses on the case where μ\mu is non-degenerate and BB is Brownian motion on GG, the above construction and finite-dimensional approximations would all follow with essentially no modification if one considered instead a Gaussian measure μ\mu whose support was, for example, a subspace 𝔥\mathfrak{h} of 𝔤\mathfrak{g} such that 𝔥\mathfrak{h} generates the span of 𝔤\mathfrak{g} 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 νt=Law(gt)\nu_{t}=\mathrm{Law}(g_{t}) is quasi-invariant under left and right translation by elements of GCMG_{CM} 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 t>0t>0 and π0\pi_{0} an orthogonal projection onto a finite-dimensional subspace G0G_{0} of 𝔤CM\mathfrak{g}_{CM}. Let hG0h\in G_{0}, and {πn}n=1\{\pi_{n}\}_{n=1}^{\infty} be an increasing sequence of projections such that G0πnGCMG_{0}\subset\pi_{n}G_{CM} for all nn and πn|GCMIGCM\pi_{n}|_{G_{CM}}\uparrow I_{G_{CM}}. Let Jtn,r(h,)J^{n,r}_{t}(h,\cdot) denote the Radon-Nikodym derivative of νtnRh1\nu_{t}^{n}\circ R_{h}^{-1} with respect to νtn\nu_{t}^{n}. Then for each nn and for any q[1,)q\in[1,\infty), we have the following integrated Harnack inequality

(Gn(Jtn,r(h,g))q𝑑νtn(g))1/qexp((q1)k2(ekt1)dn(e,h)2)\left(\int_{G_{n}}\left(J^{n,r}_{t}(h,g)\right)^{q}d\nu_{t}^{n}(g)\right)^{1/q}\leq\exp\left(\frac{(q-1)k}{2(e^{kt}-1)}d_{n}(e,h)^{2}\right)

where kk is the uniform lower bound on the Ricci curvature as in Proposition 3.10 and dnd_{n} is Riemannian distance on GnG_{n}; see for example Theorem 1.6 of [16].

By Proposition 4.5, we have that for any fCb(G)f\in C_{b}(G), the class of bounded continuous functions on GG

(4.9) Gf(gh)𝑑νt(g)=𝔼[f(gth)]=limn𝔼[f(gtnh)]=limnGn(fin)(gh)𝑑νtn(g),\begin{split}\int_{G}f(gh)d\nu_{t}(g)&=\mathbb{E}[f(g_{t}h)]\\ &=\lim_{n\rightarrow\infty}\mathbb{E}[f(g_{t}^{n}h)]=\lim_{n\rightarrow\infty}\int_{G_{n}}(f\circ i_{n})(gh)\,d\nu_{t}^{n}(g),\end{split}

where in:GnGi_{n}:G_{n}\rightarrow G denotes the inclusion map. Note that for any nn

Gn|(fin)(gh)|𝑑νtn(g)\displaystyle\int_{G_{n}}|(f\circ i_{n})(gh)|\,d\nu_{t}^{n}(g) =GnJtn,r(h,g)|(fin)(g)|𝑑νtn(g)\displaystyle=\int_{G_{n}}J^{n,r}_{t}(h,g)|(f\circ i_{n})(g)|d\nu_{t}^{n}(g)
finLq(Gn,νtn)exp(k(q1)2(ekt1)dn(e,h)2),\displaystyle\leq\|f\circ i_{n}\|_{L^{q^{\prime}}(G_{n},\nu_{t}^{n})}\exp\left(\frac{k(q-1)}{2(e^{kt}-1)}d_{n}(e,h)^{2}\right),

where qq^{\prime} is the conjugate exponent to qq. Allowing nn\rightarrow\infty in this last inequality yields

(4.10) G|f(gh)|𝑑νt(g)fLq(G,νt)exp(k(q1)2(ekt1)d(e,h)2),\int_{G}|f(gh)|\,d\nu_{t}(g)\leq\|f\|_{L^{q^{\prime}}(G,\nu_{t})}\exp\left(\frac{k(q-1)}{2(e^{kt}-1)}d(e,h)^{2}\right),

by equation (4.9) and the fact that the length of a path in GCMG_{CM} can be approximated by the lengths of paths in the finite-dimensional projections. That is, for any π0\pi_{0} and φC1([0,1],GCM)\varphi\in C^{1}([0,1],G_{CM}) with φ(0)=𝐞\varphi(0)=\mathbf{e}, there exists an increasing sequence {πn}n=1\{\pi_{n}\}_{n=1}^{\infty} of orthogonal projections such that π0πn\pi_{0}\subset\pi_{n}, πn|𝔤CMI𝔤CM\pi_{n}|_{\mathfrak{g}_{CM}}\uparrow I_{\mathfrak{g}_{CM}}, and

CM(φ)=limnGπn(πnφ).\ell_{CM}(\varphi)=\lim_{n\rightarrow\infty}\ell_{G_{\pi_{n}}}(\pi_{n}\circ\varphi).

To see this, let φ\varphi be a path in GCMG_{CM}. Then one may show that

Gπn(πnφ)\displaystyle\ell_{G_{\pi_{n}}}(\pi_{n}\circ\varphi) =01πnφ(s)+=1r1cadπnφ(s)πnφ(s)𝔤CM𝑑s\displaystyle=\int_{0}^{1}\left\|\pi_{n}\varphi^{\prime}(s)+\sum_{\ell=1}^{r-1}c_{\ell}\mathrm{ad}_{\pi_{n}\varphi(s)}^{\ell}\pi_{n}\varphi^{\prime}(s)\right\|_{\mathfrak{g}_{CM}}\,ds

for appropriate coefficients cc_{\ell}; see for example Section 3 of [32]. Thus, we have proved that (4.10) holds for fCb(G)f\in C_{b}(G) and hπGπh\in\cup_{\pi}G_{\pi}. As this union is dense in GG by Proposition 3.9, dominated convergence along with the continuity of d(e,h)d(e,h) in hh implies that (4.10) holds for all hGCMh\in G_{CM}.

Since the bounded continuous functions are dense in Lq(G,νt)L^{q^{\prime}}(G,\nu_{t}) (see for example Theorem A.1 of [25]), the inequality in (4.10) implies that the linear functional φh:Cb(G)\varphi_{h}:C_{b}(G)\rightarrow\mathbb{R} defined by

φh(f)=Gf(gh)𝑑νt(g)\varphi_{h}(f)=\int_{G}f(gh)\,d\nu_{t}(g)

has a unique extension to an element, still denoted by φh\varphi_{h}, of Lq(G,νt)L^{q^{\prime}}(G,\nu_{t})^{*} which satisfies the bound

|φh(f)|fLq(G,νt)exp(k(q1)2(ekt1)d(e,h)2)|\varphi_{h}(f)|\leq\|f\|_{L^{q^{\prime}}(G,\nu_{t})}\exp\left(\frac{k(q-1)}{2(e^{kt}-1)}d(e,h)^{2}\right)

for all fLq(G,νt)f\in L^{q^{\prime}}(G,\nu_{t}). Since Lq(G,νt)Lq(G,νt)L^{q^{\prime}}(G,\nu_{t})^{*}\cong L^{q}(G,\nu_{t}), there then exists a function Jtr(h,)Lq(G,νt)J_{t}^{r}(h,\cdot)\in L^{q}(G,\nu_{t}) such that

(4.11) φh(f)=Gf(g)Jtr(h,g)𝑑νt(g),\varphi_{h}(f)=\int_{G}f(g)J_{t}^{r}(h,g)\,d\nu_{t}(g),

for all fLq(G,νt)f\in L^{q^{\prime}}(G,\nu_{t}), and

Jtr(h,)Lq(G,νt)exp(k(q1)2(ekt1)d(e,h)2).\|J_{t}^{r}(h,\cdot)\|_{L^{q}(G,\nu_{t})}\leq\exp\left(\frac{k(q-1)}{2(e^{kt}-1)}d(e,h)^{2}\right).

Now restricting (4.11) to fCb(G)f\in C_{b}(G), we may rewrite this equation as

(4.12) Gf(g)𝑑νt(gh1)=Gf(g)Jtr(h,g)𝑑νt(g).\int_{G}f(g)\,d\nu_{t}(gh^{-1})=\int_{G}f(g)J_{t}^{r}(h,g)\,d\nu_{t}(g).

Then a monotone class argument (again use Theorem A.1 of [25]) shows that (4.12) is valid for all bounded measurable functions ff on GG. Thus, d(νtRh1)/dνtd(\nu_{t}\circ R_{h}^{-1})/d\nu_{t} exists and is given by Jtr(h,)J_{t}^{r}(h,\cdot), which is in LqL^{q} for all q(1,)q\in(1,\infty) and satisfies the desired bound.

A parallel argument gives the analogous result for d(νtLh1)/dνtd(\nu_{t}\circ L_{h}^{-1})/d\nu_{t}. Alternatively, one could use the right translation invariance just proved along with the facts that νt\nu_{t} inherits invariance under the inversion map gg1g\mapsto g^{-1} from its finite-dimensional projections and that d(e,h1)=d(e,h)d(e,h^{-1})=d(e,h). \square

The following also records the straightforward fact that the heat kernel measure does not charge GCMG_{CM}.

Proposition 4.7.

For all t>0t>0, νt(GCM)=0\nu_{t}(G_{CM})=0.

Proof.

This follows trivially from the fact that gtg_{t} is the sum of a Brownian motion BtB_{t} on 𝔤\mathfrak{g} with a finite sequence of stochastic integrals taking values in 𝔤CM\mathfrak{g}_{CM}. ∎

Thus, GCMG_{CM} maintains its role as a dense subspace of GG of measure 0 with respect to the distribution of the “group Brownian motion”.

Definition 4.8.

A function f:Gf:G\rightarrow\mathbb{R} is said to be a (smooth) cylinder function if f=Fπf=F\circ\pi for some finite-dimensional projection π\pi and some (smooth) function F:GπF:G_{\pi}\rightarrow\mathbb{R}. Also, ff is a cylinder polynomial if f=Fπf=F\circ\pi for FF a polynomial function on GπG_{\pi}.

Theorem 4.9.

Given a cylinder polynomial ff on GG, let f:G𝔤CM\nabla f:G\rightarrow\mathfrak{g}_{CM} be the gradient of ff, the unique element of 𝔤CM\mathfrak{g}_{CM} such that

f(g),h𝔤CM=h~f(g):=f(g)(Lgh𝐞),\langle\nabla f(g),h\rangle_{\mathfrak{g}_{CM}}=\tilde{h}f(g):=f^{\prime}(g)(L_{g*}h_{\mathbf{e}}),

for all h𝔤CMh\in\mathfrak{g}_{CM}. Then for kk as in Proposition 3.10,

G(f2lnf2)𝑑νt(Gf2𝑑νt)ln(Gf2𝑑νt)21ektkGf𝔤CM2𝑑νt.\int_{G}(f^{2}\ln f^{2})\,d\nu_{t}-\left(\int_{G}f^{2}\,d\nu_{t}\right)\cdot\ln\left(\int_{G}f^{2}\,d\nu_{t}\right)\leq 2\frac{1-e^{-kt}}{k}\int_{G}\|\nabla f\|_{\mathfrak{g}_{CM}}^{2}\,d\nu_{t}.
Proof.

Following the method of Bakry and Ledoux applied to GPG_{P} (see Theorem 2.9 of [17] for the case needed here) shows that

𝔼[(f2lnf2)(gtπ)]𝔼[f2(gtπ)]ln𝔼[f2(gtπ)]21ekπtkπ𝔼(πf)(gtπ)𝔤π2,\mathbb{E}\left[\left(f^{2}\ln f^{2}\right)\left(g^{\pi}_{t}\right)\right]-\mathbb{E}\left[f^{2}\left(g^{\pi}_{t}\right)\right]\ln\mathbb{E}\left[f^{2}\left(g_{t}^{\pi}\right)\right]\leq 2\frac{1-e^{-k_{\pi}t}}{k_{\pi}}\mathbb{E}\left\|(\nabla^{\pi}f)\left(g^{\pi}_{t}\right)\right\|^{2}_{\mathfrak{g}_{\pi}},

for kπk_{\pi} as in equation (3.2). Since the function x(1ex)/xx\mapsto(1-e^{-x})/x is decreasing and kkπk\leq k_{\pi} for all finite-dimensional projections π\pi, this estimate also holds with kπk_{\pi} replaced with kk. Now applying Proposition 4.3 to pass to the limit as πI\pi\uparrow I 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