math0608783/M19.tex
1: 
2: \documentclass{article}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \usepackage{amsmath}
7: 
8: \setcounter{MaxMatrixCols}{10}
9: %TCIDATA{OutputFilter=LATEX.DLL}
10: %TCIDATA{Version=5.00.0.2552}
11: %TCIDATA{<META NAME="SaveForMode" CONTENT="1">}
12: %TCIDATA{Created=Monday, March 13, 2006 14:39:55}
13: %TCIDATA{LastRevised=Thursday, August 31, 2006 14:32:01}
14: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
15: %TCIDATA{<META NAME="DocumentShell" CONTENT="Standard LaTeX\Blank - Standard LaTeX Article">}
16: %TCIDATA{Language=American English}
17: %TCIDATA{CSTFile=40 LaTeX article.cst}
18: 
19: \newtheorem{theorem}{Theorem}
20: \newtheorem{acknowledgement}[theorem]{Acknowledgement}
21: \newtheorem{algorithm}[theorem]{Algorithm}
22: \newtheorem{axiom}[theorem]{Axiom}
23: \newtheorem{case}[theorem]{Case}
24: \newtheorem{claim}[theorem]{Claim}
25: \newtheorem{conclusion}[theorem]{Conclusion}
26: \newtheorem{condition}[theorem]{Condition}
27: \newtheorem{conjecture}[theorem]{Conjecture}
28: \newtheorem{corollary}[theorem]{Corollary}
29: \newtheorem{criterion}[theorem]{Criterion}
30: \newtheorem{definition}[theorem]{Definition}
31: \newtheorem{example}[theorem]{Example}
32: \newtheorem{exercise}[theorem]{Exercise}
33: \newtheorem{lemma}[theorem]{Lemma}
34: \newtheorem{notation}[theorem]{Notation}
35: \newtheorem{problem}[theorem]{Problem}
36: \newtheorem{proposition}[theorem]{Proposition}
37: \newtheorem{remark}[theorem]{Remark}
38: \newtheorem{solution}[theorem]{Solution}
39: \newtheorem{summary}[theorem]{Summary}
40: \newenvironment{proof}[1][Proof]{\noindent\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
41: \input{tcilatex}
42: 
43: \begin{document}
44: 
45: \title{The Burkholder-Davis-Gundy Inequality for Enhanced Martingales}
46: \author{Peter Friz \and Nicolas Victoir}
47: \maketitle
48: 
49: \begin{abstract}
50: Multi-dimensional continuous local martingales, enhanced with their
51: stochastic area process, give rise to geometric rough paths with a.s. finite
52: homogenous $p$-variation, $p>2$. Here we go one step further and establish
53: quantitative bounds of the $p$-variation norm in the form of a BDG
54: inequality. Our proofs are based on old ideas by L\'{e}pingle. We also
55: discuss geodesic and piecewise linear approximations.
56: \end{abstract}
57: 
58: \section{\protect\bigskip Introduction}
59: 
60: The theory of rough paths provides a new and robust way to drive
61: differential equations by multi-dimensional stochastic processes in a 
62: \textit{deterministic way}. In most cases, this is achieved by taking into
63: account a certain stochastic area process and by establishing fine
64: regularity properties of the resulting \textit{enhanced} process. The object
65: of study in this paper is a $d$-dimensional continuous local martingale $M$
66: null at $0$ for which the area is defined by iterated stochastic
67: integration; the area process $A_{t}$ is simply the anti-symmetric part of
68: the iterated Stratonovich integral,%
69: \begin{equation*}
70: \mathbf{M}_{t}^{2}\equiv \int_{0}^{t}\int_{0}^{s}dM_{r}\otimes \circ
71: dM_{s}\in \mathbb{R}^{d}\otimes \mathbb{R}^{d}.
72: \end{equation*}%
73: Note that the symmetric part of $\mathbf{M}_{t}^{2}$ is given by $\frac{1}{2}%
74: M_{t}\otimes M_{t}$ and hence redundant if one knows $\mathbf{M}%
75: _{t}^{1}\equiv M_{t}$. It follows that the enhanced process $\mathbf{M}%
76: \equiv \left( 1,\mathbf{M}^{1},\mathbf{M}^{2}\right) \in \mathbb{R\oplus }%
77: \mathbb{R}^{d}\oplus \mathbb{R}^{d}\otimes \mathbb{R}^{d}$ lives in
78: submanifold, namely in $G^{2}\left( \mathbb{R}^{d}\right) \equiv \exp \left( 
79: \mathbb{R}^{d}\oplus so\left( d\right) \right) $, where $\exp :\left(
80: x,a\right) \mapsto \left( 1,x,a+\frac{1}{2}x\otimes x\right) .$ The space $%
81: \mathbb{R}^{d}\oplus so\left( d\right) $ carries a Lie algebra structure and
82: induces a (Lie-)group structure on $G^{2}\left( \mathbb{R}^{d}\right) $. The
83: interest in this algebraic exercise is that the resulting product operation
84: on $G^{2}\left( \mathbb{R}^{d}\right) $ is exactly what one needs to patch
85: together "iterated integral increments" over adjacent intervals. $%
86: G^{2}\left( \mathbb{R}^{d}\right) $ is also a metric (in fact,\ Polish)
87: space under the Carnot-Caratheodory metric $d$. Intuitively, the distance of
88: two points under this metric is the length of the shortest path in $\mathbb{R%
89: }^{d}$ which wipes out a prescribed area. When $d=2$, geodesics are seen to
90: be parts of circles. $G^{2}\left( \mathbb{R}^{d}\right) $ carries a dilation
91: induced by $\left( x,a\right) \mapsto \left( \lambda x,\lambda ^{2}a\right) $
92: for real $\lambda $. In fact, the CC-metric is induced by a sub-additive
93: norm, homogenuous w.r.t. dilation. Since all continuous homogenous norms are
94: Lipschitz equivalent, computations are often carried out w.r.t. $\left\vert
95: \left\vert \left\vert \left( x,a\right) \right\vert \right\vert \right\vert =
96: $ $\left\vert x\right\vert +\left\vert a\right\vert ^{1/2}$. We refer to 
97: \cite{Ly98, LQ02} for background on rough paths, \cite{FV1} contains a more
98: detailed discussion of the relevant geometry and algebra. The notion of a
99: (weak) geometric $p$-rough path \cite{FV2} becomes quite elegant: by
100: definition, one requires that the $G^{2}\left( \mathbb{R}^{d}\right) $%
101: -valued path $\mathbf{M}$ has finite $p$-variation%
102: \begin{equation*}
103: \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}[0,T]}=\left( \sup_{0\leq
104: t_{1}\leq ...\leq t_{n}\leq T}\sum d\left( \mathbf{M}_{t_{i}}\mathbf{,M}%
105: _{t_{i+1}}\right) ^{p}\right) ^{1/p}<\infty .
106: \end{equation*}%
107: Is is known\ \cite{CL} that this holds for a.e. $\mathbf{M}=\mathbf{M}\left(
108: \omega \right) $ when $p>2$. The first main topic of this paper is to
109: establish quantitative bounds of the $p$-variation norm in the form of a
110: two-sided BDG inequality: for any moderate function $F$ such as $x\mapsto
111: x^{r}$ for $r>0$,%
112: \begin{equation*}
113: \mathbb{E}\left( F\left( \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}%
114: [0,T}\right) \right) \sim \mathbb{E}\left( F\left( \left\vert \left\langle
115: M\right\rangle _{T}\right\vert ^{1/2}\right) \right) .
116: \end{equation*}%
117: The algebraic and geometric preparations made above prove crucial to recycle
118: many of the arguments given in L\'{e}pingle's seminal paper \cite{L} from
119: 1976. Secondly, we discuss approximations and show $L^{q}$-convergence (at
120: least for $q>1$) of lifted piecewise linear approximations of a continuous $%
121: L^{q}$-martingale w.r.t. homogenous $p$-variation topology.
122: 
123: The authors would like to thank D. L\'{e}pingle for a helpful email
124: exchange. 
125: 
126: \section{Preliminaries}
127: 
128: We write $\mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( [0,\infty ),\mathbb{R%
129: }^{d}\right) $ or $\mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( \mathbb{R}%
130: ^{d}\right) $ for the class of $\mathbb{R}^{d}$-valued continuous local
131: martingales $M:[0,\infty )\rightarrow \mathbb{R}^{d}$ null at $0$. The
132: bracket process $\left\langle M\right\rangle :[0,\infty )\rightarrow \mathbb{%
133: R}^{d}$ is defined component-wise, the $i^{th}$ component is given by the
134: usual bracket $\left\langle M^{i}\right\rangle =\left\langle
135: M^{i},M^{i}\right\rangle .$
136: 
137: The area-process $A:[0,\infty )\rightarrow so\left( d\right) $ is defined by
138: It\^{o}- or Stratonovich stochastic integration. As the matrix $\left\langle
139: M^{i},M^{j}\right\rangle $ is symmetric both lead to the \textit{same} area,%
140: \begin{eqnarray*}
141: A_{t}^{i,j} &=&\frac{1}{2}\left(
142: \int_{0}^{t}M^{i}dM^{j}-\int_{0}^{t}M^{j}dM^{i}\right) \\
143: &=&\frac{1}{2}\left( \int_{0}^{t}M^{i}\circ dM^{j}-\int_{0}^{t}M^{j}\circ
144: dM^{i}\right) .
145: \end{eqnarray*}%
146: We note that the area-process is a vector-valued continuous martingale. By
147: disregarding a null-set we can and will assume that $M$ and $A$ are
148: continuous.
149: 
150: \begin{definition}
151: Set $S_{2}\left( M\right) :=\mathbf{M}:=\exp \left( M+A\right) $ so that $%
152: \mathbf{M}\in C\left( [0,\infty ),G^{2}\left( \mathbb{R}^{d}\right) \right) $%
153: . The resulting class of enhanced (continuous, local) martingales is denoted
154: by $\mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( G^{2}\left( \mathbb{R}%
155: ^{d}\right) \right) $. We refer to the operation $S_{2}:M\mapsto \mathbf{M}$
156: as lift.
157: \end{definition}
158: 
159: The lift is compatible with the stopping and time-changes.
160: 
161: \begin{lemma}
162: \bigskip (i) Let $\tau $ be a stopping time. Then $\mathbf{M}^{\tau
163: }=S_{2}\left( M^{\tau }\right) $. (ii) Let $\phi $ be a time-change, that
164: is, a family $\phi _{s},\,s\geq 0,$ of stopping times such that the maps $%
165: s\mapsto \phi _{s}$ are a.s. increasing and right-continuous. Assume that $M$
166: is constant on each interval $\left[ \phi _{t-},\phi _{t}\right] $. Then $%
167: M\circ \phi $ is a continuous local martingale and%
168: \begin{equation*}
169: \mathbf{M}\circ \phi =S_{2}\left( M\circ \phi \right) .
170: \end{equation*}
171: \end{lemma}
172: 
173: \begin{proof}
174: Stopped processes are special cases of time-changed processes (take $\phi
175: _{t}=t\wedge \tau $) so it suffices to show the second statement. To this
176: end, recall the compatibility of a time change $\phi $ and stochastic
177: integration w.r.t. a continuous local martingale, constant on each interval $%
178: \left[ \phi _{t-},\phi _{t}\right] $, Proposition V.1.5. (ii)\ of \cite{RY}.
179: The lift is a special case of stochastic integration.
180: \end{proof}
181: 
182: The lift is also compatible with respect to scaling and concatenation of
183: (local martingale) paths.
184: 
185: \begin{lemma}
186: (i) If $\delta _{c}:$ $G^{2}\left( \mathbb{R}^{d}\right) \rightarrow
187: G^{2}\left( \mathbb{R}^{d}\right) $ is the dilation operator given by $%
188: \delta _{c}\exp \left( x+a\right) =\exp \left( cx+c^{2}a\right) $ then $%
189: \delta _{c}\mathbf{M}=S_{2}\left( cM\right) $. (ii) We have%
190: \begin{equation*}
191: S_{2}\left( M\right) _{0,t}=S_{2}\left( \left. M\right\vert _{\left[ 0,s%
192: \right] }\ast \left. M\right\vert _{\left[ s,t\right] }\right)
193: _{0,t}=S_{2}\left( M\right) _{0,s}\otimes S_{2}\left( M\right)
194: _{s,t},\,\,\,0\leq s\leq t<\infty
195: \end{equation*}
196: \end{lemma}
197: 
198: \begin{proof}
199: (i) follows is trivial consequence of linearity of stochastic integrals.
200: (ii) is true whenever a first order calculus underlies the lift. It now
201: suffices to note that $S_{2}$ is (equivalently) defined as Stratonovich
202: lift, 
203: \begin{equation*}
204: S_{2}\left( M\right) _{t}=\exp \left( M_{t}+A_{t}\right)
205: =1+M_{t}+\int_{0}^{t}M_{s}\otimes \circ dM_{s}\text{.}
206: \end{equation*}
207: \end{proof}
208: 
209: \begin{definition}
210: $F:\mathbb{R}^{+}\rightarrow \mathbb{R}^{+}$ is \textit{moderate} if (i) $F$
211: is continuous and increasing, (ii) $F\left( x\right) =0$ if and only if $x=0$
212: and (iii) for some (and then for every) $\alpha >1$,%
213: \begin{equation*}
214: \sup_{x>0}\frac{F\left( \alpha x\right) }{F\left( x\right) }<\infty .
215: \end{equation*}
216: \end{definition}
217: 
218: The following result is found, for instance, in \cite[p93]{RW2}.
219: 
220: \begin{theorem}[Burkholder-Davis-Gundy ]
221: Let $F$ be a moderate function, $M\in \mathcal{M}_{0,\text{\textrm{loc}}%
222: }^{c}\left( \mathbb{R}\right) $. Then there exists a constant $C=C\left(
223: F,d,\left\vert \cdot \right\vert \right) $ so that 
224: \begin{equation*}
225: C^{-1}\mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle
226: _{\infty }\right\vert ^{1/2}\right) \right) \leq \mathbb{E}\left( F\left(
227: \sup_{s\geq 0}\left\vert M_{s}\right\vert \right) \right) \leq C\mathbb{E}%
228: \left( F\left( \left\vert \left\langle M\right\rangle _{\infty }\right\vert
229: ^{1/2}\right) \right) .
230: \end{equation*}
231: \end{theorem}
232: 
233: We collect a few properties of moderate functions.
234: 
235: \begin{lemma}
236: (i) $x\mapsto F\left( x\right) $ moderate iff $\mapsto F\left(
237: x^{1/2}\right) $ moderate.\newline
238: (ii) Given $c,A,B>0$ $:c^{-1}A\leq B\leq cA\implies \exists C=C\left(
239: c,F\right) :$%
240: \begin{equation*}
241: C^{-1}F\left( A\right) \leq F\left( B\right) \leq CF\left( A\right) .
242: \end{equation*}%
243: \newline
244: (iii) $\exists C:\forall x,y>0:F\left( x+y\right) \leq C\left[ F\left(
245: x\right) +F\left( y\right) \right] .$
246: \end{lemma}
247: 
248: \begin{proof}
249: (i),(ii) are left to the reader. Ad (iii): W.l.o.g. \thinspace $x<y$, then $%
250: F\left( x+y\right) \leq F\left( 2y\right) \leq CF\left( y\right) $ by
251: moderate growth of $F$.
252: \end{proof}
253: 
254: \begin{corollary}
255: Let $F$ be a moderate function, $M\in \mathcal{M}_{0,\text{\textrm{loc}}%
256: }^{c}\left( \mathbb{R}^{d}\right) $ and $\left\vert \cdot \right\vert $ a
257: norm on $\mathbb{R}^{d}$. Then there exists a constant $C=C\left(
258: F,d,\left\vert \cdot \right\vert \right) $ so that 
259: \begin{equation*}
260: C^{-1}\mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle
261: _{\infty }\right\vert ^{1/2}\right) \right) \leq \mathbb{E}\left( F\left(
262: \sup_{s\geq 0}\left\vert M_{s}\right\vert \right) \right) \leq C\mathbb{E}%
263: \left( F\left( \left\vert \left\langle M\right\rangle _{\infty }\right\vert
264: ^{1/2}\right) \right) .
265: \end{equation*}
266: \end{corollary}
267: 
268: \begin{proof}
269: When $\left\vert a\right\vert =\max \left\{ \left\vert a^{1}\right\vert
270: ,...,\left\vert a^{d}\right\vert \right\} $ this is a simple consequence of
271: BDG for $\mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( \mathbb{R}\right) $,
272: applied componentwise. The lemma above shows that one can switch to
273: Lipschitz equivalent norms.
274: \end{proof}
275: 
276: From L\'{e}pingle \cite{L}, $\sup_{s\geq 0}\left\vert M_{s}\right\vert $
277: above can be replaced by the $p$-variation norm\footnote{%
278: In the next section, we will see a more general version of this.}. Noting
279: that the $p$-variation of a discrete-time martingale $\left( Y_{n}\right) $
280: is naturally defined as%
281: \begin{equation*}
282: \left\vert Y\right\vert _{p\text{-var}}\equiv \left[ \sup_{\left(
283: n_{k}\right) \nearrow }\sum_{k}\left\vert Y_{n_{k+1}}-Y_{n_{k}}\right\vert
284: ^{p}\right] ^{1/p},
285: \end{equation*}%
286: the following lemma is best viewed as a BDG-type upper bound for
287: discrete-time martingales.
288: 
289: \begin{lemma}
290: \label{Prop2bLep76} Let $F$ be moderate. If $1<q<p\leq 2$ or $1=q=p$ then
291: there exists a constant $c$ such that for all, possibly $\mathbb{R}^{d}$%
292: -valued, discrete-time martingales $\left( Y_{n}:n\in \mathbb{Z}^{+}\right) $%
293: \begin{equation*}
294: \mathbb{E}\left( F\left( \left\vert Y\right\vert _{p\text{-var}}\right)
295: \right) \leq c\mathbb{E}\left[ F\left( \left[ \sum_{n}\left\vert
296: Y_{n+1}-Y_{n}\right\vert ^{q}\right] ^{1/q}\right) \right] .
297: \end{equation*}
298: \end{lemma}
299: 
300: \begin{proof}
301: For $d=1$ we can use Proposition 2.b in \cite{L} with the remark that a
302: discrete-time martingale can be viewed as a particular case of a
303: continuous-time martingale with purely discontinuous sample paths. As above,
304: the extension to $d>1$ does not pose any difficulty.
305: \end{proof}
306: 
307: \section{\protect\bigskip BDG on the group}
308: 
309: \begin{lemma}[{Good$\,\protect\lambda \,$inequality, \protect\cite[p.94]{RW2}%
310: }]
311: Let $X,Y$ be nonnegative random variables, and suppose there exists $\beta
312: >1 $ such that for all $\lambda >0,\delta >0,$%
313: \begin{equation*}
314: \mathbb{P}\left( X>\beta \lambda ,Y<\delta \lambda \right) \leq \psi \left(
315: \delta \right) \mathbb{P}\left( X>\lambda \right)
316: \end{equation*}%
317: where $\psi \left( \delta \right) \searrow 0$ when $\delta \searrow 0$.
318: There, for each moderate function $F,$ there exists a constant $C$ depending
319: only on $\beta ,\psi ,F$ such that%
320: \begin{equation*}
321: \mathbb{E}\left( F\left( X\right) \right) \leq C\mathbb{E}\left( F\left(
322: Y\right) \right) .
323: \end{equation*}
324: \end{lemma}
325: 
326: \begin{proposition}
327: \label{TschebTypeBound}Let $\left\vert \cdot \right\vert ,\,\left\Vert \cdot
328: \right\Vert $ continuous homogonous norm on $\mathbb{R}^{d},G^{2}\left( 
329: \mathbb{R}^{d}\right) $ respectively. Then there exists a constant $%
330: A=A\left( d,\left\vert \cdot \right\vert ,\left\Vert \cdot \right\Vert
331: \right) $ such that 
332: \begin{equation*}
333: \forall \mathbf{M}\in \mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left(
334: G^{2}\left( \mathbb{R}^{d}\right) \right) \forall \lambda >0:\mathbb{P}%
335: \left( \sup_{u,v\geq 0}\left\Vert \mathbf{M}_{u,v}\right\Vert \geq \lambda
336: \right) \leq A\frac{\mathbb{E}\left( \left\vert \left\langle M\right\rangle
337: _{\infty }\right\vert \right) }{\lambda ^{2}}.
338: \end{equation*}
339: \end{proposition}
340: 
341: \begin{proof}
342: We note that $\sup_{u,v\geq 0}\left\Vert \mathbf{M}_{u,v}\right\Vert \leq
343: 2\sup_{t\geq 0}\left\Vert \mathbf{M}_{t}\right\Vert $. By equivalence of
344: homogeneous norm, 
345: \begin{equation*}
346: \left\Vert \mathbf{M}_{t}\right\Vert ^{2}\leq C\left( \left\vert
347: M_{t}\right\vert ^{2}+\left\vert A_{t}\right\vert \right) .
348: \end{equation*}%
349: From BDG, $\mathbb{E}\left( \sup_{u\geq 0}\left\vert M_{u}\right\vert
350: ^{2}\right) \leq C\mathbb{E}\left( \left\vert \left\langle M\right\rangle
351: _{\infty }\right\vert \right) .$ Note that $u\mapsto \left\vert \left\langle
352: M\right\rangle \right\vert _{u}:=\sum_{i=1}^{d}\left\langle
353: M^{i}\right\rangle _{u}$ is increasing (in fact, there is no loss in
354: generality in assuming that $\left\vert \cdot \right\vert $ on $\mathbb{R}%
355: ^{d}$ is given given by $\left\vert a\right\vert =\sum \left\vert
356: a^{i}\right\vert $ ...). Then, using BDG again, 
357: \begin{eqnarray*}
358: \mathbb{E}\left( \sup_{u\geq 0}\left\vert A_{u}\right\vert \right) &\leq &C%
359: \mathbb{E}\left( \left\vert \int_{0}^{\infty }\left\vert M_{u}\right\vert
360: ^{2}d\left\vert \left\langle M\right\rangle \right\vert _{u}\right\vert
361: ^{1/2}\right) \\
362: &\leq &C\mathbb{E}\left( \sup_{u\geq 0}\left\vert M_{u}\right\vert
363: .\left\vert \left\langle M\right\rangle \right\vert _{\infty }^{1/2}\right)
364: \\
365: &\leq &C\sqrt{\mathbb{E}\sup_{u\geq 0}\left\vert M_{u}\right\vert ^{2}}\sqrt{%
366: \mathbb{E}\left[ \left\vert \left\langle M\right\rangle \right\vert _{\infty
367: }\right] } \\
368: &\leq &C\mathbb{E}\left( \left\vert \left\langle M\right\rangle \right\vert
369: _{\infty }\right) .
370: \end{eqnarray*}%
371: An application of Chebyshev's inequality finishes the proof.
372: \end{proof}
373: 
374: \begin{theorem}
375: \label{BDGgroupUniform}Let $F$ be a moderate function, $\mathbf{M}\in 
376: \mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( G^{2}\left( \mathbb{R}%
377: ^{d}\right) \right) ,$ and $\left\vert \cdot \right\vert ,\,\left\Vert \cdot
378: \right\Vert $ continuous homogonous norm on $\mathbb{R}^{d},G^{2}\left( 
379: \mathbb{R}^{d}\right) $ respectively. Then there exists a constant $%
380: C=C\left( F,d,\left\vert \cdot \right\vert ,\left\Vert \cdot \right\Vert
381: \right) $ so that%
382: \begin{equation*}
383: C^{-1}\mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle
384: _{\infty }\right\vert ^{1/2}\right) \right) \leq \mathbb{E}\left( F\left(
385: \sup_{s,t\geq 0}\left\Vert \mathbf{M}_{s,t}\right\Vert \right) \right) \leq C%
386: \mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle _{\infty
387: }\right\vert ^{1/2}\right) \right) .
388: \end{equation*}
389: \end{theorem}
390: 
391: \begin{proof}
392: The lower bound comes from%
393: \begin{equation*}
394: \left\Vert \mathbf{M}_{s,t}\right\Vert \geq \left\vert M_{s,t}\right\vert
395: \end{equation*}%
396: the monotonicity of $F$ and the classical BDG lower bound. We prove the
397: upper-bound: we fix $\lambda ,\delta >0$ and $\beta >1,$ and we define the
398: stopping times 
399: \begin{eqnarray*}
400: S_{1} &=&\inf \left\{ t>0,\sup_{u,v\in \left[ 0,t\right] }\left\Vert \mathbf{%
401: M}_{u,v}\right\Vert >\beta \lambda \right\} , \\
402: S_{2} &=&\inf \left\{ t>0,\sup_{u,v\in \left[ 0,t\right] }\left\Vert \mathbf{%
403: M}_{u,v}\right\Vert >\lambda \right\} , \\
404: S_{3} &=&\inf \left\{ t>0,\left\vert \left\langle M\right\rangle
405: _{t}\right\vert ^{1/2}>\delta \lambda \right\} ,
406: \end{eqnarray*}%
407: with the convention that the infimum of the empty set if $\infty .$ Define
408: the local martingale $N_{t}=M_{S_{3}\wedge S_{2},\left( t+S_{2}\right)
409: \wedge S_{3}}$noting that $N_{t}\equiv 0$ on $\left\{ S_{2}=\infty \right\} $%
410: . It is easy to see that 
411: \begin{equation}
412: \sup_{u,v\in \left[ 0,S_{3}\right] }\left\Vert \mathbf{M}_{u,v}\right\Vert
413: \leq \sup_{u,v\in \left[ 0,S_{3}\wedge S_{2}\right] }\left\Vert \mathbf{M}%
414: _{u,v}\right\Vert +\sup_{u,v\geq 0}\left\Vert \mathbf{N}_{u,v}\right\Vert .
415: \end{equation}%
416: By definition of the relevant stopping times, 
417: \begin{equation*}
418: \mathbb{P}\left( \sup_{u,v\geq 0}\left\Vert \mathbf{M}_{u,v}\right\Vert
419: >\beta \lambda ,\left\vert \left\langle M\right\rangle _{\infty }\right\vert
420: ^{1/2}\leq \delta \lambda \right) =\mathbb{P}\left( S_{1}<\infty
421: ,S_{3}=\infty \right) .
422: \end{equation*}%
423: On the event $\left\{ S_{1}<\infty ,S_{3}=\infty \right\} $ one has%
424: \begin{equation*}
425: \sup_{u,v\in \left[ 0,S_{3}\right] }\left\Vert \mathbf{M}_{u,v}\right\Vert
426: >\beta \lambda
427: \end{equation*}%
428: and, since $S_{2}\leq S_{1}$, one also has $\sup_{u,v\in \left[
429: 0,S_{3}\wedge S_{2}\right] }\left\Vert \mathbf{M}_{u,v}\right\Vert =\lambda
430: . $ Hence, on $\left\{ S_{1}<\infty ,S_{3}=\infty \right\} ,$%
431: \begin{equation*}
432: \sup_{u,v\geq 0}\left\Vert \mathbf{N}_{u,v}\right\Vert \geq \sup_{u,v\in %
433: \left[ 0,S_{3}\right] }\left\Vert \mathbf{M}_{u,v}\right\Vert -\sup_{u,v\in %
434: \left[ 0,S_{3}\wedge S_{2}\right] }\left\Vert \mathbf{M}_{u,v}\right\Vert
435: \geq \left( \beta -1\right) \lambda .
436: \end{equation*}%
437: Therefore, using Proposition \ref{TschebTypeBound},%
438: \begin{eqnarray*}
439: \mathbb{P}\left( \sup_{u,v\geq 0}\left\Vert \mathbf{M}_{u,v}\right\Vert
440: >\beta \lambda ,\left\vert \left\langle M\right\rangle _{\infty }\right\vert
441: ^{1/2}\leq \delta \lambda \right) &\leq &\mathbb{P}\left( \sup_{u,v\geq
442: 0}\left\Vert \mathbf{N}_{u,v}\right\Vert \geq \left( \beta -1\right) \lambda
443: \right) \\
444: &\leq &\frac{A}{\left( \beta -1\right) ^{2}\lambda ^{2}}\mathbb{E}\left(
445: \left\vert \left\langle N\right\rangle _{\infty }\right\vert \right) .
446: \end{eqnarray*}%
447: From the definition of $N$, for every $t\in \left[ 0,\infty \right] $,%
448: \begin{equation*}
449: \left\langle N\right\rangle _{t}=\left\langle M\right\rangle _{S_{3}\wedge
450: S_{2},\left( t+S_{2}\right) \wedge S_{3}}.
451: \end{equation*}%
452: On $\left\{ S_{2}=\infty \right\} $ we have $\left\langle N\right\rangle
453: _{\infty }=0$ while on $\left\{ S_{2}<\infty \right\} $ we have, from
454: definition of $S_{3}$,%
455: \begin{equation*}
456: \left\vert \left\langle N\right\rangle _{\infty }\right\vert =\left\vert
457: \left\langle M\right\rangle _{S_{3}\wedge S_{2},S_{3}}\right\vert
458: =\left\vert \left\langle M\right\rangle _{S_{3}}-\left\langle M\right\rangle
459: _{S_{3}\wedge S_{2}}\right\vert \leq 2\left\vert \left\langle M\right\rangle
460: _{S_{3}}\right\vert =2\delta ^{2}\lambda ^{2}.
461: \end{equation*}%
462: It follows that%
463: \begin{equation*}
464: \mathbb{E}\left( \left\vert \left\langle N\right\rangle _{\infty
465: }\right\vert \right) \leq 2\delta ^{2}\lambda ^{2}\mathbb{P}\left(
466: S_{2}<\infty \right) =2\delta ^{2}\lambda ^{2}\mathbb{P}\left( \sup_{u,v\geq
467: 0}\left\Vert \mathbf{M}_{u,v}\right\Vert >\lambda \right)
468: \end{equation*}%
469: and we have the estimate%
470: \begin{equation*}
471: \mathbb{P}\left( \sup_{u,v\geq 0}\left\Vert \mathbf{M}_{u,v}\right\Vert
472: >\beta \lambda ,\left\vert \left\langle M\right\rangle _{\infty }\right\vert
473: ^{1/2}\leq \delta \lambda \right) \leq \frac{2A\delta ^{2}}{\left( \beta
474: -1\right) ^{2}}\mathbb{P}\left( \sup_{u,v\geq 0}\left\Vert \mathbf{M}%
475: _{u,v}\right\Vert >\lambda \right) .
476: \end{equation*}%
477: An application of the good $\lambda $-inequality finishes the proof.
478: \end{proof}
479: 
480: \section{Path regularity and $p$-variation BDG}
481: 
482: \bigskip Let $p>2$. From \cite{CL} it is known that for every $\mathbf{M}\in 
483: \mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( G^{2}\left( \mathbb{R}%
484: ^{d}\right) \right) $ and every $T>0$ 
485: \begin{equation}
486: \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[ 0,T\right] }<\infty 
487: \text{ a.s.}  \label{FinitePvarForEnhMart}
488: \end{equation}%
489: Here, we go one step further and provided quantitative bounds for the $p$%
490: -variation of the enhanced martingale in terms of $\left\langle
491: M\right\rangle _{T}$. En passant, we give a simplified proof of (\ref%
492: {FinitePvarForEnhMart}).
493: 
494: \begin{proposition}
495: Let $\mathbf{M}\in \mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( G^{2}\left( 
496: \mathbb{R}^{d}\right) \right) $. Then, for every $T>0,$%
497: \begin{equation*}
498: \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[ 0,T\right] }<\infty 
499: \text{ a.s.}
500: \end{equation*}
501: \end{proposition}
502: 
503: \begin{proof}
504: There exists a sequence of stopping times $\tau _{n}\rightarrow \infty $
505: a.s. such that $M^{\tau _{n}}$ and $\left\langle M^{\tau _{n}}\right\rangle $
506: are bounded (for instance, $\tau _{n}=\inf \{t:\left\vert M_{t}\right\vert
507: >n $ or $\left\vert \left\langle M\right\rangle _{t}\right\vert >n\}$ will
508: do.) Since%
509: \begin{equation*}
510: \mathbb{P}\left( \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[ 0,T%
511: \right] }\neq \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[
512: 0,T\wedge \tau _{n}\right] }\right) \leq \mathbb{P}\left( \tau _{n}<T\right)
513: \rightarrow 0\text{ as }n\rightarrow \infty
514: \end{equation*}%
515: it suffices to consider the lift of a bounded continuous martingale with
516: bounded quadratic variation. We can work with the $l^{1}$-norm on $\mathbb{R}%
517: ^{d}$, $\left\vert a\right\vert =\sum_{i=1}^{d}\left\vert a_{i}\right\vert . 
518: $ The time change $\phi \left( t\right) :=\inf \left\{ s:\left\vert
519: \left\langle M\right\rangle _{s}\right\vert >t\right\} $ may have jumps but
520: continuity of $\left\vert \left\langle M\right\rangle \right\vert $ ensures
521: that $|\left\langle M\right\rangle _{\phi \left( t\right) }|\ =t$. From
522: definition of $\phi $ and the BDG inequality on the group, both $\left\vert
523: \left\langle M\right\rangle \right\vert $ and $\mathbf{M}$ are constant on
524: the intervals $\left[ \phi _{t-},\phi _{t}\right] $. It follows that $%
525: \mathbf{X}_{t}=\mathbf{M}_{\phi \left( t\right) }$ defines a continuous%
526: \footnote{%
527: From Lemma 2, $\mathbf{X}=S_{2}\left( M\circ \phi \right) ,$ the lift of a
528: continuous local martingale. In particular, this is another way to see
529: continuity of $\mathbf{X}$.} path from $\left[ 0,\left\vert \left\langle
530: M\right\rangle _{T}\right\vert \right] $ to $G^{2}\left( \mathbb{R}%
531: ^{d}\right) $ and it is easy to see that%
532: \begin{equation*}
533: \left\Vert \mathbf{X}\right\Vert _{p\text{-var},\left[ 0,\left\vert
534: \left\langle M\right\rangle _{T}\right\vert \right] }=\left\Vert \mathbf{M}%
535: \right\Vert _{p\text{-var},\left[ 0,T\right] }.
536: \end{equation*}%
537: As argued in the beginning of the proof, we may assume that $\left\vert
538: \left\langle M\right\rangle _{T}\right\vert \leq R$ for some deterministic $%
539: R $ large enough. Therefore, 
540: \begin{eqnarray}
541: \mathbb{P}\left( \left\Vert \mathbf{M}\right\Vert _{p\text{-var},\left[ 0,T%
542: \right] }>K\right) &=&\mathbb{P}\left( \left\Vert \mathbf{X}\right\Vert _{p%
543: \text{-var},\left[ 0,\left\vert \left\langle M\right\rangle _{T}\right\vert %
544: \right] },\left\vert \left\langle M\right\rangle _{T}\right\vert \leq
545: R\right)  \label{MltX} \\
546: &\leq &\mathbb{P}\left( \left\Vert \mathbf{X}\right\Vert _{p\text{-var},%
547: \left[ 0,R\right] }>K\right) .  \notag
548: \end{eqnarray}%
549: We go on to show that $\mathbf{X}$ is in fact H\"{o}lder continuous. For $%
550: 0\leq s\leq t\leq R$, we can use the BDG\ inequality on the group, theorem %
551: \ref{BDGgroupUniform}, to obtain%
552: \begin{equation*}
553: \mathbb{E}\left( \left\Vert \mathbf{X}_{s,t}\right\Vert ^{2q}\right) =%
554: \mathbb{E}\left( \left\Vert \mathbf{M}_{\phi \left( s\right) ,\phi \left(
555: t\right) }\right\Vert ^{2q}\right) \leq C_{q}\mathbb{E}\left( \left\vert
556: \left\langle M\right\rangle _{\phi \left( t\right) }-\left\langle
557: M\right\rangle _{\phi \left( s\right) }\right\vert ^{q}\right) .
558: \end{equation*}%
559: Observe that 
560: \begin{eqnarray*}
561: \left\vert \left\langle M\right\rangle _{\phi \left( t\right) }-\left\langle
562: M\right\rangle _{\phi \left( s\right) }\right\vert &=&\sum_{i}\left(
563: \left\langle M^{i}\right\rangle _{\phi \left( t\right) }-\left\langle
564: M^{i}\right\rangle _{\phi \left( s\right) }\right) \\
565: &=&\left\vert \left\langle M\right\rangle _{\phi \left( t\right)
566: }\right\vert -\left\vert \left\langle M\right\rangle _{\phi \left( s\right)
567: }\right\vert =t-s.
568: \end{eqnarray*}%
569: Thus, for all $q<\infty $ there exists a constant $C_{q}$ s.t.%
570: \begin{equation*}
571: \mathbb{E}\left( \left\Vert \mathbf{X}_{s,t}\right\Vert ^{2q}\right) \leq
572: C_{q}\left\vert t-s\right\vert ^{q}.
573: \end{equation*}%
574: Knowing that $\mathbf{X}$ is continuous, we can apply GRR\footnote{%
575: There is no modification of $\mathbf{X}$ needed.} for paths in $\left(
576: G^{2}\left( \mathbb{R}^{d}\right) ,d\right) $ to see that $\left\Vert 
577: \mathbf{X}\right\Vert _{1/p\text{-H\"{o}lder},\left[ 0,R\right] }\in L^{q}$
578: for all $q\in \lbrack 1,\infty )$ and 
579: \begin{equation*}
580: \mathbb{P}\left( \left\Vert \mathbf{X}\right\Vert _{p\text{-var},\left[ 0,R%
581: \right] }>K\right) \leq \frac{\mathbb{E}\left( \left\Vert \mathbf{X}%
582: \right\Vert _{p\text{-var},\left[ 0,R\right] }\right) }{K}\leq \frac{\mathbb{%
583: E}\left( \left\Vert \mathbf{X}\right\Vert _{1/p\text{-H\"{o}lder},\left[ 0,R%
584: \right] }\right) }{K}
585: \end{equation*}%
586: tends to zero as $K\rightarrow \infty $. Together with (\ref{MltX}) we see
587: that $\left\Vert \mathbf{M}\right\Vert _{p\text{-var},\left[ 0,T\right]
588: }<\infty $ with probability $1$ as claimed.
589: \end{proof}
590: 
591: We are now going to prove a $p$-variation version of BDG. For $\mathbb{R}$%
592: -valued martingales this result is due to L\'{e}pingle, \cite{L}. With the
593: preparations made, our proof follows the same lines.
594: 
595: \begin{lemma}
596: There exists a constant $A$ such that for all continous local martingales $M$%
597: , for all $\lambda >0,$ 
598: \begin{equation*}
599: \mathbb{P}\left( \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}[0,\infty
600: )}>\lambda \right) \leq A\frac{\mathbb{E}\left( \left\vert \left\langle
601: M\right\rangle _{\infty }\right\vert \right) }{\lambda ^{2}}.
602: \end{equation*}
603: \end{lemma}
604: 
605: \begin{proof}
606: If suffices to prove the statement when $\lambda =1$ (the general case
607: follows by considering $M/\lambda $ with lift $\delta _{1/\lambda }\mathbf{M}
608: $). The statement then reduces to 
609: \begin{equation*}
610: \exists A:\forall M:\mathbb{P}\left[ \left\Vert \mathbf{M}\right\Vert _{p%
611: \text{-var;}[0,\infty )}>1\right] \leq A\,\mathbb{E}\left( \left\vert
612: \left\langle M\right\rangle _{\infty }\right\vert \right) .
613: \end{equation*}%
614: Assume this is false. Then for every $A$, and in particular for $A\left(
615: k\right) \equiv k^{2}$,there exists $M\equiv M^{\left( k\right) }$ with lift 
616: $\mathbf{M}^{\left( k\right) }$ s.t. the condition is violated,%
617: \begin{equation*}
618: k^{2}\,\mathbb{E}\left[ \left\vert \left\langle M^{\left( k\right)
619: }\right\rangle _{\infty }\right\vert \right] <\mathbb{P}\left[ \left\Vert 
620: \mathbf{M}^{\left( k\right) }\right\Vert _{p\text{-var;}[0,\infty )}>1\right]
621: \leq 1.
622: \end{equation*}%
623: Set $u_{k}=\mathbb{P}\left[ \left\Vert \mathbf{M}^{\left( k\right)
624: }\right\Vert _{p\text{-var;}[0,\infty )}>1\right] $, $n_{k}=\left[ 1/u_{k}+1%
625: \right] \in \mathbb{N}$ and note that $1\leq n_{k}u_{k}\leq 2$. Take $n_{k}$
626: copies of each $M^{\left( k\right) }$ and get a sequence of martingales of
627: form%
628: \begin{equation*}
629: \left( \tilde{M}\right) \equiv (\underset{n_{1}}{\underbrace{M^{\left(
630: 1\right) },...,M^{\left( 1\right) }}};\underset{n_{2}}{\underbrace{M^{\left(
631: 2\right) },...,M^{\left( 2\right) }}};M^{\left( 3\right) },...).
632: \end{equation*}%
633: Then%
634: \begin{equation*}
635: n_{k}k^{2}\mathbb{E}\left[ \left\vert \left\langle M^{\left( k\right)
636: }\right\rangle _{\infty }\right\vert \right] \leq n_{k}\mathbb{P}\left[
637: \left\Vert \mathbf{M}^{\left( k\right) }\right\Vert _{p\text{-var;}[0,\infty
638: )}>1\right] =n_{k}u_{k}\leq 2.
639: \end{equation*}%
640: and%
641: \begin{equation*}
642: \sum_{k}\mathbb{P}\left[ \left\Vert \mathbf{\tilde{M}}^{\left( k\right)
643: }\right\Vert _{p\text{-var;}[0,\infty )}>1\right] =\sum_{k}n_{k}u_{k}=+\infty
644: \end{equation*}%
645: while%
646: \begin{equation*}
647: \sum_{k}\mathbb{E}\left[ \left\vert \left\langle \tilde{M}^{\left( k\right)
648: }\right\rangle _{\infty }\right\vert \right] =\sum_{k}n_{k}\mathbb{E}\left[
649: \left\vert \left\langle M^{\left( k\right) }\right\rangle _{\infty
650: }\right\vert \right] \leq \sum_{k}\frac{2}{k^{2}}<\infty .
651: \end{equation*}%
652: Thus, if the claimed statement is false, there exists a sequence of
653: martingales, we now revert to write $M^{\left( k\right) },\mathbf{M}^{\left(
654: k\right) }$ instead of $\tilde{M}^{\left( k\right) },\mathbf{\tilde{M}}%
655: ^{\left( k\right) }$ respectively, each defined on some filtered probability
656: space $\left( \Omega ^{k},\left( \mathcal{F}_{t}^{k}\right) ,\mathbb{P}%
657: ^{k}\right) $ with the two properties%
658: \begin{equation*}
659: \sum_{k}\mathbb{P}^{k}\left[ \left\Vert \mathbf{M}^{\left( k\right)
660: }\right\Vert _{p\text{-var;}[0,\infty )}>1\right] =+\infty \text{ and }%
661: \sum_{k}\mathbb{E}^{k}\left[ \left\vert \left\langle M^{\left( k\right)
662: }\right\rangle _{\infty }\right\vert \right] <\infty .
663: \end{equation*}%
664: Define the probability space $\Omega =\bigotimes_{k=1}^{\infty }\Omega ^{k},$
665: the probability $\mathbb{P}=\bigotimes_{k=1}^{\infty }\mathbb{P}^{k},$ and
666: the filtration $\left( \mathcal{F}_{t}\right) $ on $\Omega $ given by%
667: \begin{equation*}
668: \mathcal{F}_{t}=\left( \bigotimes_{i=1}^{k-1}\mathcal{F}_{\infty
669: }^{i}\right) \otimes \mathcal{F}_{g\left( k-t\right) }^{k}\otimes \left(
670: \bigotimes_{j=k+1}^{\infty }\mathcal{F}_{0}^{k}\right) \text{ \ for }k-1\leq
671: t<k.
672: \end{equation*}%
673: where $g\left( u\right) =1/u-1$ maps $\left[ 0,1\right] \rightarrow \left[
674: 0,\infty \right] $. Then, a continous martingale on $\left( \Omega ,\left( 
675: \mathcal{F}_{t}\right) ,\mathbb{P}\right) $ is defined by concatenation, \ 
676: \begin{equation*}
677: M_{t}=\sum_{i=1}^{k-1}M_{\infty }^{\left( i\right) }+M_{g\left( k-t\right)
678: }^{\left( k\right) }\,\,\,\text{for \ \ }k-1\leq t<k.
679: \end{equation*}%
680: which implies%
681: \begin{equation*}
682: \mathbf{M}_{t}=\left( \bigotimes_{i=1}^{k-1}\mathbf{M}_{\infty }^{\left(
683: i\right) }\right) \otimes \mathbf{M}_{g\left( k-t\right) }^{\left( k\right)
684: }.
685: \end{equation*}%
686: We also observe that, again for $k-1\leq t<k$, 
687: \begin{equation*}
688: \left\langle M\right\rangle _{t}=\sum_{i=1}^{k-1}\left\langle M^{\left(
689: i\right) }\right\rangle _{\infty }+\left\langle M^{\left( k\right)
690: }\right\rangle _{g\left( k-t\right) }.
691: \end{equation*}%
692: In particular, $\left\langle M\right\rangle _{\infty }=\sum_{k}\left\langle
693: M^{\left( k\right) }\right\rangle _{\infty }$ and, using the second property
694: of the martingale sequence, $\mathbb{E}\left( \left\vert \left\langle
695: M\right\rangle _{\infty }\right\vert \right) <\infty $. Define the events%
696: \begin{equation*}
697: A_{k}=\left\{ \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[ k-1,k%
698: \right] }>1\right\} .
699: \end{equation*}%
700: Then, using the first property of the martingale sequence, 
701: \begin{equation*}
702: \sum_{k}\mathbb{P}\left( A_{k}\right) =\sum_{k}\mathbb{P}^{k}\left(
703: \left\Vert \mathbf{M}^{k}\right\Vert _{p\text{-var;}[0,\infty )}>1\right)
704: =\infty .
705: \end{equation*}%
706: Since the events $\left\{ A_{k}:k\geq 1\right\} $ are independent, the
707: Borel-Cantelli lemma implies that $\mathbb{P}\left( A_{k}\text{ i.o.}\right)
708: =1$. Thus, almost surely, for all $K>0$ there exists a finite number of
709: increasing times $t_{0},\cdots ,t_{n}\in \lbrack 0,\infty )$ so that%
710: \begin{equation*}
711: \sum_{i=1}^{n}\left\Vert \mathbf{M}_{t_{i-1},t_{i}}\right\Vert >K
712: \end{equation*}%
713: and $\left\Vert \mathbf{M}\right\Vert _{p\text{-var;}[0,\infty )}$ must be
714: equal to $+\infty $ with probability one. We now define a martingale $N$ by
715: time-change, namely via $f\left( t\right) =t/\left( 1-t\right) $ for $0\leq
716: t<1$ and $f\left( t\right) =\infty $ for $t\geq 1$,%
717: \begin{equation*}
718: N:t\mapsto M_{f\left( t\right) }.
719: \end{equation*}%
720: Note that $\mathbb{E}\left( \left\vert \left\langle M\right\rangle _{\infty
721: }\right\vert \right) <\infty $ so that $M$ can be extended to a (continuous)
722: martingale indexed by $\left[ 0,\infty \right] $ and $N$ is indeed a
723: continuous martingale with lift $\mathbf{N}$. Since lifts interchange with
724: time changes, $\left\Vert \mathbf{N}\right\Vert _{p\text{-var;}%
725: [0,1]}=\left\Vert \mathbf{M}\right\Vert _{p\text{-var;}[0,\infty )}=+\infty $
726: with probability one. But this contradicts to $p$-variation regularity
727: result above.
728: \end{proof}
729: 
730: The very same argument that was used in the proof of Theorem \ref%
731: {BDGgroupUniform}\ now leads to the following BDG inequality for enhanced
732: continuous local martingales w.r.t. homogenuous $p$-variation norm.
733: 
734: \begin{theorem}
735: \label{pVarBDG}Let $F$ be a moderate function, $\mathbf{M}\in \mathcal{M}_{0,%
736: \text{\textrm{loc}}}^{c}\left( G^{2}\left( \mathbb{R}^{d}\right) \right) ,$
737: and $\left\vert \cdot \right\vert ,\,\left\Vert \cdot \right\Vert $
738: continuous homogonous norm on $\mathbb{R}^{d},G^{2}\left( \mathbb{R}%
739: ^{d}\right) $ respectively and $p>2$. Then there exists a constant $%
740: C=C\left( p,F,d,\left\vert \cdot \right\vert ,\left\Vert \cdot \right\Vert
741: \right) $ so that%
742: \begin{equation*}
743: C^{-1}\mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle
744: _{\infty }\right\vert ^{1/2}\right) \right) \leq \mathbb{E}\left( F\left(
745: \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}[0,\infty )}\right) \right)
746: \leq C\mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle
747: _{\infty }\right\vert ^{1/2}\right) \right) .
748: \end{equation*}
749: \end{theorem}
750: 
751: \begin{remark}
752: When $p\in \left( 2,3\right) $ and $N\in \left\{ 3,4,...\right\} $, $\mathbf{%
753: M}$ lifts uniquely to a $G^{N}\left( \mathbb{R}^{d}\right) $-valued path
754: with finite homogenuous $p$-variation regularity, denoted by $S_{N}\left( 
755: \mathbf{M}\right) $, which is identified with the first $N$ iterated
756: Stratonovich integrals of $M$. A basic theorem of Lyons asserts that%
757: \begin{equation*}
758: \left\Vert S_{N}\left( \mathbf{M}\right) \right\Vert _{p\text{-var}}\leq
759: C\left( N\right) \left\Vert \mathbf{M}\right\Vert _{p\text{-var}}
760: \end{equation*}%
761: and BDG inequalities for the $p$-variation of this step-$N$ lift are an
762: immediate corollary of Theorem \ref{pVarBDG}.
763: \end{remark}
764: 
765: \section{Approximations}
766: 
767: We now only consider (lifted) local martingales on $\left[ 0,T\right] $,
768: defined or identified with local martingales stopped at $T>0.$
769: 
770: \subsection{Geodesic approxiations}
771: 
772: The $p$-variation norm of geodesics approximations is uniformly controlled
773: by the original $p$-variation norm. Therefore%
774: \begin{equation*}
775: \mathbb{E}\left( F\left( \sup_{D}\left\Vert \mathbf{M}^{\left[ D\right]
776: }\right\Vert _{p\text{-var;}[0,T]}\right) \right) \leq C\mathbb{E}\left(
777: F\left( \left\vert \left\langle M\right\rangle _{T}\right\vert ^{1/2}\right)
778: \right)
779: \end{equation*}%
780: where $\mathbf{M}^{\left[ D\right] }$ denotes the geodesics approxiation to $%
781: \mathbf{M}$ based on some dissection $D$ of $\left[ 0,T\right] $. Note that
782: this is stronger than%
783: \begin{equation*}
784: \sup_{D}\mathbb{E}\left( F\left( \left\Vert \mathbf{M}^{\left[ D\right]
785: }\right\Vert _{p\text{-var;}[0,T]}\right) \right) \leq C\mathbb{E}\left(
786: F\left( \left\vert \left\langle M\right\rangle _{T}\right\vert ^{1/2}\right)
787: \right)
788: \end{equation*}%
789: which is what we are going to show for piecewise linear approximations.
790: 
791: \subsection{Piecewise linear approximations}
792: 
793: Let $D=\left( t_{i}\right) $ be a subdivision of $\left[ 0,T\right] .$ Given 
794: $x\in C\left( \left[ 0,T\right] ,\mathbb{R}^{d}\right) $ we define $x^{D}$
795: to be the piecewise linear approximation of $x$ which coincides with $x$ on $%
796: D.$ Since $x^{D}$ is of bounded variation, it admits a canonical lift to a $%
797: G^{2}\left( \mathbb{R}^{d}\right) $-valued path, denoted by $\mathbf{x}^{D}.$
798: This notation applies path-by-path to $M\in \mathcal{M}_{0,\text{\textrm{loc}%
799: }}^{c}\left( \mathbb{R}^{d}\right) $, we write $\mathbf{M}^{D}=\mathbf{M}%
800: ^{D}\left( \omega \right) $ for the lifted piecewise linear approximation to 
801: $M\left( \omega \right) $. The next lemma involves no probabilty.
802: 
803: \begin{lemma}
804: \label{pVarOfLiftedPWApprox}Set $\mathbf{x}^{D}=S_{2}\left( x^{D}\right) $
805: where $x^{D}$ is linear between the points of $D$. Then there exists a
806: constant $C=C=C\left( d,\left\vert \cdot \right\vert ,\left\Vert \cdot
807: \right\Vert \right) $ such that%
808: \begin{equation*}
809: \left\Vert \mathbf{x}^{D}\right\Vert _{p-var;\left[ 0,T\right] }\leq C\left(
810: \max_{\left( s_{k}\right) \subset D}\sum_{k}\left\Vert \mathbf{x}%
811: _{s_{k},s_{k+1}}^{D}\right\Vert ^{p}\right) ^{1/p}+C\left\vert x\right\vert
812: _{p\text{-var};\left[ 0,T\right] }.
813: \end{equation*}
814: \end{lemma}
815: 
816: \begin{proof}
817: First we note that $\left\Vert \mathbf{x}_{s,t}^{D}\right\Vert ^{p}\leq
818: 3^{p-1}\left[ \left\vert x_{s,s^{D}}^{D}\right\vert ^{p}+\left\Vert \mathbf{x%
819: }_{s^{D},t_{D}}^{D}\right\Vert ^{p}+\left\vert x_{t_{D},t}^{D}\right\vert
820: ^{p}\right] $. Now let $\left( u_{k}\right) $ be a dissection of $\left[ 0,T%
821: \right] $, unrelated to $D$. Recall that $u^{D}$ resp. $u_{D}$ refers to the
822: right- resp. left-neighbours of $u$ in $D$. 
823: \begin{eqnarray*}
824: \sum_{k}\left\Vert \mathbf{x}_{u_{k},u_{k+1}}^{D}\right\Vert ^{p} &\leq
825: &3^{p-1}\sum_{k}\left\Vert \mathbf{x}_{u_{k}^{D},u_{k+1,D}}^{D}\right\Vert
826: ^{p}+3^{p-1}\sum_{k}\left[ \left\vert x_{u_{k},u_{k}^{D}}^{D}\right\vert
827: ^{p}+\left\vert x_{u_{k+1,D},u_{k}}^{D}\right\vert ^{p}\right] \\
828: &\leq &3^{p-1}\left( \max_{\left( s_{k}\right) \subset D}\sum_{k}\left\Vert 
829: \mathbf{x}_{s_{k},s_{k+1}}^{D}\right\Vert ^{p}\right) +3^{p-1}\left\vert
830: x^{D}\right\vert _{p\text{-var;}\left[ 0,T\right] } \\
831: &\leq &3^{p-1}\left( \max_{\left( s_{k}\right) \subset D}\sum_{k}\left\Vert 
832: \mathbf{x}_{s_{k},s_{k+1}}^{D}\right\Vert ^{p}\right) +C3^{p-1}\left\vert
833: x\right\vert _{p\text{-var;}\left[ 0,T\right] }.
834: \end{eqnarray*}
835: \end{proof}
836: 
837: \begin{theorem}
838: \label{UpperBDGboundUniformOverDissections}Let $F$ be a moderate function, $%
839: \mathbf{M}$ $\in \mathcal{M}_{0,\text{\textrm{loc}}}^{c}\left( G^{2}\left( 
840: \mathbb{R}^{d}\right) \right) ,$ and $\left\vert \cdot \right\vert
841: ,\,\left\Vert \cdot \right\Vert $ continuous homogonous norm on $\mathbb{R}%
842: ^{d},G^{2}\left( \mathbb{R}^{d}\right) $ respectively. Then there exists a
843: constant $C=C\left( p,F,d,\left\vert \cdot \right\vert ,\left\Vert \cdot
844: \right\Vert \right) $ so that for all dissections $D$ of $\left[ 0,T\right]
845: , $%
846: \begin{equation*}
847: \mathbb{E}\left( F\left( \left\Vert \mathbf{M}^{D}\right\Vert _{p-var\text{;}%
848: [0,T]}\right) \right) \leq C\mathbb{E}\left( F\left( \left\vert \left\langle
849: M\right\rangle _{T}\right\vert ^{1/2}\right) \right) .
850: \end{equation*}
851: \end{theorem}
852: 
853: \begin{proof}
854: From Lemma \ref{pVarOfLiftedPWApprox}, $\left\Vert \mathbf{M}^{D}\right\Vert
855: _{p\text{-var;}\left[ 0,T\right] }$ is bounded by $C\left\vert M\right\vert
856: _{p\text{-var;}\left[ 0,T\right] }$ plus%
857: \begin{eqnarray*}
858: &&C\left( \max_{\left( s_{k}\right) \subset D}\sum_{k}\left\Vert \mathbf{M}%
859: _{s_{k},s_{k+1}}^{D}\right\Vert ^{p}\right) ^{1/p}\leq C\left( \max_{\left(
860: s_{k}\right) \subset D}\sum_{k}\left\Vert \mathbf{M}_{s_{k},s_{k+1}}\right%
861: \Vert ^{p}\right) ^{1/p} \\
862: &&+C\left( \max_{\left( s_{k}\right) \subset D}\sum_{k}d\left( \mathbf{M}%
863: _{s_{k},s_{k+1}},\mathbf{M}_{s_{k},s_{k+1}}^{D}\right) ^{p}\right) ^{1/p}.
864: \end{eqnarray*}%
865: Trivially, $\left\vert M\right\vert _{p\text{-var;}\left[ 0,T\right] }\leq
866: \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[ 0,T\right] }$ and
867: with a new constant $C$, 
868: \begin{equation*}
869: \left\Vert \mathbf{M}^{D}\right\Vert _{p\text{-var;}\left[ 0,T\right] }\leq
870: C\left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[ 0,T\right] }+C\left(
871: \max_{\left( s_{k}\right) \subset D}\sum_{k}d\left( \mathbf{M}%
872: _{s_{k},s_{k+1}},\mathbf{M}_{s_{k},s_{k+1}}^{D}\right) ^{p}\right) ^{1/p}%
873: \text{.}
874: \end{equation*}%
875: For fixed $k,$ there are $i<j$ so that $s_{k}=t_{i}$ and $s_{k+1}=t_{j}$.
876: Then%
877: \begin{eqnarray*}
878: \mathbf{M}_{s_{k},s_{k+1}} &=&\bigotimes_{l=i}^{j-1}\exp \left(
879: M_{t_{l},t_{l+1}}+A_{t_{l},t_{l+1}}\right) \\
880: \,\mathbf{M}_{s_{k},s_{k+1}}^{D}\, &=&\bigotimes_{l=i}^{j-1}\exp \left(
881: M_{t_{l},t_{l+1}}\right) .
882: \end{eqnarray*}%
883: Hence, $d\left( \mathbf{M}_{s_{k},s_{k+1}},\mathbf{M}_{s_{k},s_{k+1}}^{D}%
884: \right) $ equals%
885: \begin{equation}
886: \left\Vert \mathbf{M}_{s_{k},s_{k+1}}^{-1}\otimes \mathbf{M}%
887: _{s_{k},s_{k+1}}^{D}\right\Vert =\left\Vert \exp \left(
888: \sum_{l=i}^{j-1}A_{t_{l},t_{l+1}}\right) \right\Vert \leq C\left\vert
889: \sum_{l=i}^{j-1}A_{t_{l},t_{l+1}}\right\vert ^{1/2}.  \label{SumOfSmallAreas}
890: \end{equation}%
891: The key idea is to introduce the (vector-valued) discrete-time martingale%
892: \begin{equation*}
893: Y_{j}=\sum_{l=0}^{j-1}A_{t_{l},t_{l+1}}\in so\left( d\right) .
894: \end{equation*}%
895: From (\ref{SumOfSmallAreas}) and equivalence of homogenous norms we have%
896: \begin{equation*}
897: \max_{\left( s_{k}\right) \subset D}\sum_{k}d\left( \mathbf{M}%
898: _{s_{k},s_{k+1}},\mathbf{M}_{s_{k},s_{k+1}}^{D}\right) ^{p}\leq
899: C\max_{\left\{ i_{1},...,i_{n}\right\} \subset \left\{ 1,...,\#D\right\}
900: }\sum_{k}\left\vert Y_{i_{k+1}}-Y_{i_{k}}\right\vert ^{p/2},
901: \end{equation*}%
902: which leads to%
903: \begin{eqnarray*}
904: \left\Vert \mathbf{M}^{D}\right\Vert _{p\text{-var}} &\leq &C\left\Vert 
905: \mathbf{M}\right\Vert _{p\text{-var}}+C\sqrt{\left( \max_{\left\{
906: i_{1},...,i_{n}\right\} \subset \left\{ 1,...,\#D\right\}
907: }\sum_{k}\left\vert Y_{i_{k+1}}-Y_{i_{k}}\right\vert ^{p/2}\right) ^{2/p}} \\
908: &=&C\left\Vert \mathbf{M}\right\Vert _{p\text{-var}}+C\sqrt{\left\vert
909: Y\right\vert _{p/2\text{-var}}}.
910: \end{eqnarray*}%
911: Using basic properties of moderate functions we have%
912: \begin{eqnarray*}
913: \mathbb{E}\left[ F\left( \left\Vert \mathbf{M}^{D}\right\Vert _{p\text{-var}%
914: }\right) \right] &\leq &C\mathbb{E}\left[ F\left( \left\Vert \mathbf{M}%
915: \right\Vert _{p\text{-var}}\right) \right] +C\mathbb{E}\left[ F\left( \sqrt{%
916: \left\vert Y\right\vert _{p/2\text{-var}}}\right) \right] \\
917: &=&C\mathbb{E}\left[ F\left( \left\Vert \mathbf{M}\right\Vert _{p\text{-var}%
918: }\right) \right] +C\mathbb{E}\left[ F\circ \sqrt{\cdot }\left( \left\vert
919: Y\right\vert _{p/2\text{-var}}\right) \right] .
920: \end{eqnarray*}%
921: Note that $F\circ \sqrt{\cdot }$ is moderate since $F$ is moderate. Let $%
922: 2<p^{\prime }<p<3$. Then $1<p^{\prime }/2\leq p/2\leq 2$ and and Lemma \ref%
923: {Prop2bLep76} yields 
924: \begin{eqnarray*}
925: \mathbb{E}\left[ F\circ \sqrt{\cdot }\left( \left\vert Y\right\vert _{p/2%
926: \text{-var}}\right) \right] &\leq &\mathbb{E}\left[ F\circ \sqrt{\cdot }%
927: \left( \left( \sum_{l}\left\vert Y_{l+1}-Y_{l}\right\vert ^{p^{\prime
928: }/2}\right) ^{2/p^{\prime }}\right) \right] \\
929: &=&\mathbb{E}\left[ F\circ \sqrt{\cdot }\left( \left( \sum_{l}\left\vert
930: A_{t_{l},t_{l+1}}\right\vert ^{p^{\prime }/2}\right) ^{2/p^{\prime }}\right) %
931: \right] \\
932: &\leq &\mathbb{E}\left[ F\left( \left( \sum_{l}\left\Vert \mathbf{M}%
933: _{t_{l},t_{l+1}}\right\Vert ^{p^{\prime }}\right) ^{1/p^{\prime }}\right) %
934: \right] \\
935: &\leq &\mathbb{E}\left[ F\left( \left\Vert \mathbf{M}\right\Vert _{p^{\prime
936: }\text{-var;}\left[ 0,T\right] }\right) \right] .
937: \end{eqnarray*}%
938: Combing the last two estimates and using Theorem \ref{pVarBDG} (with $%
939: p^{\prime }=1+p/2>2$ and $p$ respectively) gives%
940: \begin{eqnarray*}
941: \mathbb{E}\left[ F\left( \left\Vert \mathbf{M}^{D}\right\Vert _{p\text{-var;}%
942: \left[ 0,T\right] }\right) \right] &\leq &C\mathbb{E}\left[ F\left(
943: \left\Vert \mathbf{M}\right\Vert _{p\text{-var;}\left[ 0,T\right] }\right) %
944: \right] +C\mathbb{E}\left[ F\left( \left\Vert \mathbf{M}\right\Vert
945: _{p^{\prime }\text{-var;}\left[ 0,T\right] }\right) \right] \\
946: &\leq &2C\mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle
947: _{T}\right\vert ^{1/2}\right) \right) .
948: \end{eqnarray*}
949: \end{proof}
950: 
951: \begin{remark}
952: We don't expect a lower BDG bound uniformly over all dissections $D$ of $%
953: \left[ 0,T\right] $. For instance,%
954: \begin{equation*}
955: C^{-1}\mathbb{E}\left( F\left( \left\vert \left\langle M\right\rangle
956: _{T}\right\vert ^{1/2}\right) \right) \leq \mathbb{E}\left( F\left(
957: \left\vert M^{D}\right\vert _{\infty \text{;}[0,T]}\right) \right)
958: \end{equation*}%
959: can't hold since $D=\left\{ 0,T\right\} $ implies $M_{\infty \text{;}%
960: [0,T]}^{D}=\left\vert M_{T}\right\vert $ and for $F\left( x\right) =x$ we
961: would control%
962: \begin{equation*}
963: \mathbb{E}\left( \left\vert M\right\vert _{\infty \text{;}[0,T]}\right) \sim 
964: \mathbb{E}\left( \left\vert \left\langle M\right\rangle
965: _{T}^{1/2}\right\vert \right)
966: \end{equation*}%
967: in terms of $\mathbb{E}\left( \left\vert M_{T}\right\vert \right) $ which is
968: Doob's $L^{q}$ maximal inequality with $q=1$. But, as is well known, one
969: needs $q>1$ for Doob's $L^{q}$-inequality to hold true.
970: \end{remark}
971: 
972: Let us now bound the supremum distance between $\mathbf{M}$ and $\mathbf{M}%
973: ^{D}:$
974: 
975: \begin{lemma}
976: \label{convergencedinfty}Assume that $M$ is a martingale such that%
977: \begin{equation}
978: \left\vert M\right\vert _{\infty ;[0,T]}\in L^{q}\left( \Omega \right) \text{
979: for some }q\geq 1\text{.}  \label{MinfinityLqcondition}
980: \end{equation}%
981: \ \ If $D^{n}$ is a sequence of subdivisions whose time steps tends to $0$
982: when $n$ tends to $\infty $, then $d_{\infty ;\left[ 0,T\right] }\left( 
983: \mathbf{M,M}^{D_{n}}\right) $ converges to $0$ in $L^{q}$.
984: \end{lemma}
985: 
986: \begin{remark}
987: If $q>1$, Doob's maximal inequality implies that (\ref{MinfinityLqcondition}%
988: ) holds for any $L^{q}$-martingale.
989: \end{remark}
990: 
991: \begin{proof}[Proof of Lemma \protect\ref{convergencedinfty}]
992: As in the proof of Theorem \ref{UpperBDGboundUniformOverDissections},
993: equation (\ref{SumOfSmallAreas}) more specifically, we have that when $%
994: t=t_{i}\in D$%
995: \begin{equation*}
996: d\left( \mathbf{M}_{t},\mathbf{M}_{t}^{D}\right) \leq C\left\vert
997: \sum_{k=0}^{i-1}A_{t_{k},t_{k+1}}\right\vert ^{1/2}.
998: \end{equation*}%
999: Next, consider $t\in \left[ t_{i},t_{i+1}\right] $ for some $i$. The path $%
1000: M_{\cdot }^{D}$ restricted to $\left[ t_{i},t_{i+1}\right] $ is a straight
1001: line with no area, hence
1002: 
1003: \begin{equation*}
1004: \mathbf{M}_{t_{i},t}^{D}=\exp \left( \frac{t-s}{t_{i+1}-t_{i}}%
1005: M_{t_{i},t_{i+1}}\right) .
1006: \end{equation*}
1007: 
1008: and%
1009: \begin{eqnarray*}
1010: d\left( \mathbf{M}_{t,},\mathbf{M}_{t}^{D}\right) &=&d\left( \mathbf{M}%
1011: _{t_{i}}\otimes \mathbf{M}_{t_{i},t},\mathbf{M}_{t_{i}}^{D}\otimes \mathbf{M}%
1012: _{t_{i},t}^{D}\right) \\
1013: &=&\left\Vert \left( \mathbf{M}_{t_{i},t}^{D}\right) ^{-1}\otimes \left( 
1014: \mathbf{M}_{t_{i}}^{D}\right) ^{-1}\otimes \mathbf{M}_{t_{i}}\otimes \mathbf{%
1015: M}_{t_{i},t}\right\Vert \\
1016: &\leq &\left\Vert \left( \mathbf{M}_{t_{i},t}^{D}\right) \right\Vert
1017: +\left\Vert \left( \mathbf{M}_{t_{i}}^{D}\right) ^{-1}\otimes \mathbf{M}%
1018: _{t_{i}}\right\Vert +\left\Vert \mathbf{M}_{t_{i},t}\right\Vert \\
1019: &\leq &2\sup_{u,v\in \left[ t_{i},t_{i+1}\right] }\left\Vert \mathbf{M}%
1020: _{u,v}\right\Vert +C\max_{i,j}\left\vert
1021: \sum_{l=i}^{j-1}A_{t_{l},t_{l+1}}\right\vert ^{1/2}.
1022: \end{eqnarray*}
1023: 
1024: For the $L^{q}$ convergence, because $\mathbf{M}$ is almost surely
1025: continuous (in fact, uniformly continuous on the compact $\left[ 0,T\right] $%
1026: )%
1027: \begin{equation*}
1028: \max_{i=0,...,\#D-1}\sup_{s,t\in \left[ t_{i}^{n},t_{i+1}^{n}\right]
1029: }\left\Vert \mathbf{M}_{t_{i},t_{i+1}}\right\Vert \rightarrow 0\text{ a.s.}
1030: \end{equation*}%
1031: Hence, by dominated convergence, 
1032: \begin{equation*}
1033: \lim_{\left\vert D_{n}\right\vert \rightarrow 0}\mathbb{E}\left(
1034: \max_{i=0,...,\#D-1}\sup_{s,t\in \left[ t_{i}^{n},t_{i+1}^{n}\right]
1035: }\left\Vert \mathbf{M}_{t_{i},t_{i+1}}\right\Vert ^{q}\right) =0.
1036: \end{equation*}%
1037: With $Y$ defined as in the proof of Theorem \ref%
1038: {UpperBDGboundUniformOverDissections}, 
1039: \begin{equation*}
1040: \max_{i,j}\left\vert \sum_{l=i}^{j-1}A_{t_{l},t_{l+1}}\right\vert ^{1/2}\leq
1041: C\left[ \left( \max_{\left\{ i_{1},...,i_{n}\right\} \subset \left\{
1042: 1,...,\#D\right\} }\sum_{k}\left\vert Y_{i_{k+1}}-Y_{i_{k}}\right\vert
1043: ^{p/2}\right) ^{2/p}\right] ^{1/2}
1044: \end{equation*}%
1045: the computation given therein with $F\left( x\right) =x^{q}$ shows%
1046: \begin{eqnarray*}
1047: \mathbb{E}\left( \max_{i,j}\left\vert
1048: \sum_{l=i}^{j-1}A_{t_{l},t_{l+1}}\right\vert ^{q/2}\right) &\leq &C\mathbb{E}%
1049: \left[ F\circ \sqrt{}\left( \max_{\left\{ i_{1},...,i_{n}\right\} \subset
1050: \left\{ 1,...,\#D\right\} }\sum_{k}\left\vert
1051: Y_{i_{k+1}}-Y_{i_{k}}\right\vert ^{p/2}\right) ^{2/p}\right] \\
1052: &\leq &C\mathbb{E}\left[ F\left( \left( \sum_{l:t_{l}\in D_{n}}\left\Vert 
1053: \mathbf{M}_{t_{l},t_{l+1}}\right\Vert ^{q}\right) ^{1/q}\right) \right] \\
1054: &=&\mathbb{E}\left[ \left( \sum_{l:t_{l}\in D_{n}}\left\Vert \mathbf{M}%
1055: _{t_{l},t_{l+1}}\right\Vert ^{q}\right) \right] .
1056: \end{eqnarray*}%
1057: Bu this last expression tends to zero, combining the bounded convergence
1058: theorem with a.s. convergence%
1059: \begin{equation*}
1060: \lim_{n\rightarrow \infty }\sum_{l:t_{l}\in D_{n}}\left\Vert \mathbf{M}%
1061: _{t_{l},t_{l+1}}\right\Vert ^{q}=0.
1062: \end{equation*}%
1063: Indeed, this follows from $\mathbf{M}\in C^{0,q\text{-var}}$ since $q>2$ and
1064: using the usual squeezing argument. To show $L^{q\text{ }}$convergence with
1065: respect to $d_{\infty \text{ }}=d_{\infty ;\left[ 0,T\right] }$, we also
1066: write $\left\Vert \mathbf{\cdot }\right\Vert _{\infty }=\left\Vert \mathbf{%
1067: \cdot }\right\Vert _{\infty ;\left[ 0,T\right] }$ here, recall that 
1068: \begin{equation*}
1069: d_{\infty }\left( \mathbf{M,M}^{D}\right) \leq \sup_{t\in \left[ 0,T\right]
1070: }d\left( \mathbf{M}_{t}\mathbf{,M}_{t}^{D}\right) +c\left\vert \left\Vert 
1071: \mathbf{M}\right\Vert _{\infty }\sup_{t\in \left[ 0,T\right] }d\left( 
1072: \mathbf{M}_{t}\mathbf{,M}_{t}^{D}\right) \right\vert ^{1/2}.
1073: \end{equation*}%
1074: We just showed that $\sup_{t\in \left[ 0,T\right] }d\left( \mathbf{M}_{t}%
1075: \mathbf{,M}_{t}^{D}\right) \rightarrow 0$ in $L^{q}$. Then%
1076: \begin{eqnarray*}
1077: &&\mathbb{E}\left( \left\vert \left\Vert \mathbf{M}\right\Vert _{\infty
1078: }\sup_{t\in \left[ 0,T\right] }d\left( \mathbf{M}_{t}\mathbf{,M}%
1079: _{t}^{D}\right) \right\vert ^{q/2}\right) \\
1080: &\leq &\left( \mathbb{E}\left( \left\vert \left\Vert \mathbf{M}\right\Vert
1081: _{\infty }\right\vert ^{q}\right) \right) ^{1/2}\left( \mathbb{E}\left(
1082: \left\vert \sup_{t\in \left[ 0,T\right] }d\left( \mathbf{M}_{t}\mathbf{,M}%
1083: _{t}^{D}\right) \right\vert ^{q}\right) \right) ^{1/2}.
1084: \end{eqnarray*}%
1085: (Note that by the our BDG\ inqualities%
1086: \begin{equation*}
1087: \mathbb{E}\left( \left\vert \left\Vert \mathbf{M}\right\Vert _{\infty ;\left[
1088: 0,T\right] }\right\vert ^{q}\right) \leq C\mathbb{E}\left( \left\vert
1089: \left\langle M\right\rangle _{T}\right\vert ^{q}\right) \leq C\mathbb{E}%
1090: \left( \left\vert \left\vert M\right\vert _{\infty ;\left[ 0,T\right]
1091: }\right\vert ^{q}\right)
1092: \end{equation*}%
1093: and the last expression is finite by assumption.)
1094: \end{proof}
1095: 
1096: \begin{theorem}
1097: Let $M$ be as in Lemma \ref{convergencedinfty}. Then, $d_{p\text{-var;}\left[
1098: 0,T\right] }\left( \mathbf{M}^{D},\mathbf{M}\right) $ converges to $0$ in $%
1099: L^{q}.$ If $M$ is a local martingale, then convergence holds in probability.
1100: \end{theorem}
1101: 
1102: \begin{proof}
1103: The result for the local martingale will hold if the first result holds, by
1104: a localisation argument that we leave to the reader. We already saw that $%
1105: L^{q}$-convergence holds w.r.t. $d_{\infty }=d_{\infty ;\left[ 0,T\right] }$%
1106: . To go further, writing $d_{p\text{-var}}\equiv d_{p\text{-var;}\left[ 0,T%
1107: \right] }$, we use the interpolation formula 
1108: \begin{equation*}
1109: d_{p\text{-var}}\left( \mathbf{M},\mathbf{M}^{D}\right) \leq Cd_{\infty
1110: }\left( \mathbf{M,M}^{D}\right) ^{1-\frac{p^{\prime }}{p}}\left( \left\Vert 
1111: \mathbf{M}\right\Vert _{p^{\prime }-var}^{\frac{p^{\prime }}{p}}+\left\Vert 
1112: \mathbf{M}^{D}\right\Vert _{p^{\prime }-var}^{\frac{p^{\prime }}{p}}\right)
1113: ,\,\,\,2<p^{\prime }<p.
1114: \end{equation*}%
1115: Hence,%
1116: \begin{equation*}
1117: \mathbb{E}\left( \left\vert d_{p-var}\left( \mathbf{M}^{D},\mathbf{M}\right)
1118: \right\vert ^{q}\right) \leq C\mathbb{E}\left( \left( \left\Vert \mathbf{M}%
1119: \right\Vert _{p^{\prime }-var}^{q\frac{p^{\prime }}{p}}+\left\Vert \mathbf{M}%
1120: ^{D}\right\Vert _{p^{\prime }-var}^{q\frac{p^{\prime }}{p}}\right) d_{\infty
1121: }\left( \mathbf{M,M}^{D}\right) ^{q\left( 1-\frac{p^{\prime }}{p}\right)
1122: }\right)
1123: \end{equation*}%
1124: Using H\"{o}lder with conjugate exponents $1/\left( p^{\prime }/p\right) $
1125: and $1/\left( 1-p^{\prime }/p\right) $ gives%
1126: \begin{equation*}
1127: \mathbb{E}\left( \left\vert d_{p-var}\left( \mathbf{M}^{D},\mathbf{M}\right)
1128: \right\vert ^{q}\right) \leq C\mathbb{E}\left( \left\Vert \mathbf{M}%
1129: \right\Vert _{p^{\prime }-var}^{q}+\left\Vert \mathbf{M}^{D}\right\Vert
1130: _{p^{\prime }-var}^{q}\right) ^{p^{\prime }/p}\left[ \mathbb{E}\left(
1131: d_{\infty }\left( \mathbf{M,M}^{D}\right) ^{q}\right) \right] ^{1-p^{\prime
1132: }/p}.
1133: \end{equation*}%
1134: But now it suffices to remark, using our BDG estimates, that%
1135: \begin{equation*}
1136: \mathbb{E}\left( \left\Vert \mathbf{M}\right\Vert _{p^{\prime }-var;\left[
1137: 0,T\right] }^{q}\right) ,\mathbb{E}\left( \left\Vert \mathbf{M}%
1138: ^{D}\right\Vert _{p^{\prime }-var;\left[ 0,T\right] }^{q}\right) \leq C%
1139: \mathbb{E}\left( \left\vert \left\langle M\right\rangle _{T}\right\vert
1140: ^{q/2}\right) \leq C\mathbb{E}\left( \left\vert \left\vert M\right\vert
1141: _{\infty ;\left[ 0,T\right] }\right\vert ^{q}\right)
1142: \end{equation*}%
1143: and the last term is finite by assumption.
1144: \end{proof}
1145: 
1146: \begin{thebibliography}{9}
1147: \bibitem{CL} Coutin Laure, Lejay Antoine: Semi-martingales and rough paths
1148: theory, Electronic Journal of Probability, Vol. 10, Paper 23, 2005.
1149: 
1150: \bibitem{FV1} Friz,\ Peter; Victoir, Nicolas: Approximations of the Brownian
1151: Rough Path with Applications to Stochastic Analysis, Annales de l'Institut
1152: Henri Poincare (B), Probability and Statistics, Volume 41, Issue 4, 2005.
1153: 
1154: \bibitem{FV2} Friz, Peter; Victoir, Nicolas:\ On the notion of Geometric
1155: Rough Paths. To appear in Probab. Theory Relat. Fields, 2006.
1156: 
1157: \bibitem{LLP} Lenglart \'{E}rik, L\'{e}pingle Dominique, Pratelli Maurizio:
1158: Pr\'{e}sentation unifi\'{e}e de certaines in\'{e}galit\'{e}s de la th\'{e}%
1159: orie des martingales. LNM 1404, 1980.
1160: 
1161: \bibitem{L} L\'{e}pingle Dominique: La variation d'ordre p des
1162: semi-martingales, Z. Wahrscheinlichkeitstheorie und Verw. Gebiete, Volume
1163: 36, Issue 4,1976.
1164: 
1165: \bibitem{Ly98} Lyons, Terry: Differential equations driven by rough signals.
1166: Rev. Mat. Iberoamericana 14, no. 2, 215--310, 1998.
1167: 
1168: \bibitem{LQ02} Lyons, Terry; Qian, Zhongmin: System Control and Rough Paths,
1169: OUP, 2002.
1170: 
1171: \bibitem{RY} Revuz Daniel, Yor Marc: Continuous Martingales and Brownian
1172: Motion, 3rd edition, Springer, 1999.
1173: 
1174: \bibitem{RW2} Rogers LCG, Williams David: Diffusions, Markov Processes, and
1175: Martingales : It\^{o} Calculus, CUP,\ 2000.
1176: \end{thebibliography}
1177: 
1178: \end{document}
1179: