ArticlePDF Available

Existence of L\'evy's area and pathwise integration

Authors:

Abstract

Rough path analysis can be developed using the concept of controlled paths and with respect to a topology in which besides the uniform distance L\'evy's area plays a role. For vectors of rough paths we investigate the relationship between the property of being controlled and the existence of associated L\'evy areas. Given a path which is controlled by another one, the L\'evy area in the It\^o or Stratonovitch sense can be defined pathwise, eventually provided the existence of quadratic variation along a sequence of partitions is guaranteed. This leads us to a study of the pathwise change of variable (It\^o) formula in the spirit of F\"ollmer, from the perspective of controlled paths.
arXiv:1404.3645v1 [math.PR] 14 Apr 2014
Existence of L´evy’s area and pathwise integration
Peter Imkeller and David J. Pr¨omel1
Humboldt-Universit¨at zu Berlin
Institut f¨ur Mathematik
May 7, 2014
Abstract
Rough path analysis can be developed using the concept of controlled paths and with
respect to a topology in which besides the uniform distance L´evy’s area plays a role.
For vectors of rough paths we investigate the relationship between the property of being
controlled and the existence of associated L´evy areas. Given a path which is controlled
by another one, the L´evy area in the Itˆo or Stratonovitch sense can be defined pathwise,
eventually provided the existence of quadratic variation along a sequence of partitions is
guaranteed. This leads us to a study of the pathwise change of variable (Itˆo) formula in
the spirit of F¨ollmer, from the perspective of controlled paths.
Key words: Controlled path, F¨ollmer integration, Itˆo’s formula, Levy’s area, rough path,
Stratonovich integral.
MSC 2010 Classification: 26A42, 60H05.
1 Introduction
The theory of rough paths (see [LCL07], [Lej09] or [FH13]) has established an analytical
frame in which stochastic differential- and integral calculus beyond Young’s classical notions
is traced back to properties of the trajectories of processes involved without reference to a
particular probability measure. For instance, in the simplest non-trivial setting it provides a
topology on the set of continuous functions enhanced with an ”area”, with respect to which the
(Itˆo) map associating the trajectories of a solution process of a stochastic differential equation
driven by trajectories of continuous martingale is continuous. In this topology, convergence of
a sequence of functions Xn= (X1,n,··· , Xd,n)nNdefined on the time interval [0, T ] involves
besides uniform convergence also the convergence of the L´evy areas associated to the vector
of trajectories, formally given by
Li,j,n
t=Zt
0
(Xi,n
sdXj,n
sXj,n
sdXi,n
s),1i, j d, t [0, T ].
1We are grateful to Randolf Altmeyer and Nicolas Perkowski for helpful comments and discussions on the
subject matter. D.J.P. was financially supported by a Ph.D. scholarship of the DFG Research Training Group
1845 ”Stochastic Analysis with Applications in Biology, Finance and Physics”.
Email: imkeller@math.hu-berlin.de
Email: proemel@math.hu-berlin.de
1
In probability theory the concept of L´evy’s area was already studied in the 1940s. It
was first introduced by P. L´evy in [L´ev40] for a two dimensional Brownian motion (B1, B2).
For time Tfixed, and any trajectory of the process it is defined as the area enclosed by the
trajectory (B1, B 2) and the chord given by the straight line from (0,0) to (B1
T, B2
T), and may
be expressed formally by
1
2ZT
0
B1
tdB2
tZT
0
B2
tdB1
t,
provided the integrals make sense.
More recently, an alternative calculus with a more (Fourier) analytic touch has been de-
signed (see [GIP14], [Per14]) in which an older idea by Gubinelli [Gub04] is further developed.
It is based on the concept of controlled paths. In this calculus, rough integrals are described
in terms of Fourier series for instance in the Haar-Schauder wavelet, and are seen to decom-
pose into different parts, one of them representing L´evy’s area. The existence of a stochastic
integral in this approach is seen to be linked to the existence of the corresponding L´evy area,
and both can be approximated along a Schauder development in which H¨older functions are
limits of their finite degree Schauder expansions. In its simplest (one-dimensional) form a
path of bounded variation Yon [0, T ] is controlled by another path Xof bounded variation
on [0, T ], if the associated signed measures µX, µYon the Borel sets of [0, T ] satisfy that µY
is absolutely continuous with respect to µX. In its version relevant here two rough (vector
valued) functions Xand Yon [0, T ] are considered, both with finite p-variation for some
p1. In the simplest setting, Yis controlled by Xif there exists a rough function Yof
finite p-variation such that the first order Taylor expansion errors
RY
s,t =YtYsY
s(XtXs)
are bounded in r-norm, i.e. P[s,t]π|RY
s,t|ris bounded over all possible partitions πof [0, T ].
Here 1
r=2
p.Since for a path Xolder continuity of order 1
pis closely related to finite p-
variation, the control relation can be seen as expressing a type of fractional Taylor expansion
of first order: the first order Taylor expansion error of Ywith respect to X- both of H¨older
order 1
pand ”derivative” Y- is of double H¨older order 2
p.In its para-controlled refinement
as developed by Gubinelli et. al. in [GIP13] this notion has been seen to give an alternative
approach to classical rough path analysis and is essentially suitable for the application to
singular PDEs. In the comparison of the two approaches, to make the Itˆo map continuous,
information stored in the Levy areas of vector paths has to be replaced by information con-
veyed by path control or vice versa. This raises the problem about the relationship between
the existence of L´evy’s area and the control relationship between vector trajectories or the
components of such.
In the first part of this paper (section 2) we shall deal with this question. In fact we shall
show that for a vector Xof functions a particular version of control which we will call self-
control (roughly, for pairs of components (Xi, Xj) one has to control the other one) is sufficient
for the existence of the L´evy areas, and thus for the existence of related pathwise integrals.
We shall also show by an example that missing control can lead to the failure of possessing a
evy area. In the second part (section 3) we shall study the question how control concepts
and the existence of different kinds of integrals (Itˆo type, Stratonovich type) are related, and
in particular in which way control leads to versions of F¨ollmer’s pathwise Itˆo formula. We will
decompose Riemann approximations of different versions of integrals into a symmetric and
2
an antisymmetric component. While the convergence of the antisymmetric component, very
closely related to L´evy areas, does not require to go along prescribed sequences of partitions,
this will be the case for the symmetric part. In F¨ollmer’s approach the entire integral is
constructed along a given sequence of partitions, and the existence of quadratic variation
along this sequence is assumed. We shall show that for the classical Stratonovich integral just
the antisymmetric Riemann sums have to converge, while for more general Stratonovich or
Itˆo type integrals the existence of limits for the symmetric part has to be guaranteed along
fixed sequences of partitions, as in F¨ollmer’s approach. The versions of pathwise Itˆo’s formula
derived therefrom will exhibit this variety of integrals.
2 evy’s area and controlled paths
We know that both the control of a (one-dimensional) path X1with respect to another
path X2, as well as the existence of L´evy’s area for (X1, X2) ([Lyo98]) entails the existence
of the rough path integral of X1with respect to X2. This raises the question about the
strengths of the hypotheses leading to the existence of the integral. This question will be
answered here. We will show that control entails the existence of L´evy’s area. The analysis
we present, as usual, is based on ddimensional rough paths, and corresponding notions of
areas. For a continuous path X: [0, T ]Rd, say X= (X1, ..., Xd), we recall that evy’s
area L(X) = (Li,j (X))i,j is given by
L(X)i,j := ZT
0
Xi
tdXj
tZT
0
Xj
tdXi
t,1i, j d,
where Xdenotes the transpose of the vector X, if the respective integrals exist. There are
pairs of H¨older continuous paths X1and X2for which L´evy’s area does not exist (see example
7 below). To answer this question, we need the basic setup of rough path analysis, starting
with the notion of power variation.
Apartition π:= {[ti1, ti]|i= 1, ..., N}of an interval [0, T ] is a family of essentially
disjoint intervals such that SN
i=1[ti1, ti] = [0, T ]. For any 1 p < , a continuous function
X: [0, T ]Rdis of finite p-variation if
||X||p:= sup
π∈P X
[s,t]π
|Xs,t|p
1
p
<,
where the supremum is taken over the set Pof partitions of [0, T ] and Xs,t := XtXsfor
s, t [0, T ], st. We write Vp([0, T ],Rd) for the set (linear space) of continuous functions of
finite p-variation. Let, more generally, R: [0, T ]2Rd×dbe a continuous function, 1 r <
. In this case we consider the functional
||R||r:= sup
π∈P X
[s,t]π
|Rs,t|r1
r
.
An equivalent way to characterize the property of finite p-variation is by the existence of a
control function. Denoting by ∆T:= {(s, t)[0, T ]2|0stT}, we call a continuous
3
function ω: ∆TR+vanishing on the diagonal control function if it is superadditive, i.e. if
for (s, u, t)[0, T ]3such that 0 sutT
ω(s, u) + ω(u, t)ω(s, t).
Note that a function is of finite p-variation if and only if there exists a control function ω
such that |Xs,t|pω(s, t) for (s, t)T. For a more detailed discussion of p-variation and
control functions see Chapter 1.2 in [LCL07].
A fundamental insight due to Gubinelli [Gub04] was that an integral RY dX exists if “Y
looks like Xin the small scale”. This leads to the concept of controlled paths which we recall
in its general form.
Definition 1. Let p, q , r R+be such that 2/p + 1/q > 1 and 1/r = 1/p +1/q. Suppose X
Vp([0, T ],Rd). We call Y∈ Vp([0, T ],Rd)controlled by Xif there exists Y∈ Vq([0, T ],Rd×d)
such that the remainder term RYgiven by the relation Ys,t =Y
sXs,t +RY
s,t satisfies ||RY||r<
. In this case we write YCq
X, and call YGubinelli derivative.
See Theorem 1 in [Gub04] for the case of H¨older continuous paths, or Theorem 24 in
[PP13] for precise existence results of RY dX. Let us now modify this concept to a notion of
control of a path by itself.
Definition 2. Let p, q, r R+be such that 2/p + 1/q > 1 and 1/r = 1/p + 1/q. We call
X∈ Vp([0, T ],Rd)self-controlled if we have XiCq
Xjor XjCq
Xifor all 1 i, j dwith
i6=j.
With this notion we are now able to deal with the main task of this section, the construc-
tion of the L´evy area of a self-controlled path X. In fact, the integrals arising in L´evy’s area
will be obtained via left-point Riemann sums as
L(X)i,j =ZT
0
Xi
tdXj
tZT
0
Xj
tdXi
t:= lim
|π|→0X
[s,t]π
(Xi
sXj
s,t Xj
sXi
s,t),(1)
for 1 i, j d, where |π|denotes the mesh of a partition π. Our approach uses the abstract
version of classical ideas due to Young [You36] comprised in the so-called sewing lemma.
Lemma 3. [Corollary 2.3, Corollary 2.4 in [FD06]] Let Ξ : ∆TRdbe a continuous function
and K > 0some constant. Assume that there exist a control function ωand a constant θ > 1
such that for all (s, u, t)[0, T ]3with 0sutTwe have
|Ξs,t Ξs,u Ξu,t| ≤ K ω(s, t)θ.(2)
Then there exists a unique function Φ: [0, T ]Rdsuch that Φ(0) = 0 and
|Φ(t)Φ(s)Ξs,t| ≤ C(θ)ω(s, t)θ,0stT ,
where C(θ) := K(1 21θ)1, and we have
lim
|π(s,t)|→0X
[u,v]π(s,t)
Ξu,v = Φ(t)Φ(s),
where π(s, t)denotes a partition of [s, t].
4
With this tool we now derive the existence of L´evy’s area for self-controlled paths of finite
p-variation with p1.
Theorem 4. Let 1p < and suppose that X∈ Vp([0, T ],Rd)is self-controlled, then
evy’s area as defined in (1) exists.
Proof. Let X∈ Vp([0, T ],Rd) for 1 p < be self-controlled and fix 1 i, j d, i 6=j.
We may assume without loss of generality that XiCq
Xj, i.e. Xi
s,t =X
s(i, j)Xj
s,t +Ri,j
s,t and
||X(i, j)||q,||Ri,j ||r<. In order to apply Lemma 3, we set Ξi,j
s,t := Xi
sXj
s,t Xj
sXi
s,t for
(s, t)Tand observe that for (s, u, t)[0, T ]3with 0 sutTwe have
Ξi,j
s,t Ξi,j
s,u Ξi,j
u,t =Xi
sXj
s,t Xj
sXi
s,t Xi
sXj
s,u +Xj
sXi
s,u Xi
uXj
u,t +Xj
uXi
u,t
=Xi
s,uXj
u,t Xj
s,uXi
u,t
= (X
s(i, j)Xj
s,u +Ri,j
s,u)Xj
u,t Xj
s,u(X
u(i, j)Xj
u,t +Ri,j
u,t)
=Ri,j
s,uXj
u,t Xj
s,uRi,j
u,t + (X
s(i, j)X
u(i, j))Xj
s,uXj
u,t.
Since the finite sum of control functions is again a control function, we can choose the same
control function ωfor X, X(i, j ) and Ri,j, and therefore we get
|Ξi,j
s,t Ξi,j
s,u Ξi,j
u,t| ≤ ω(s, t)
1
p+1
r+ω(s, t)
1
p+1
r+ω(s, t)
2
p+1
q3ω(s, t)θ
with θ:= 2
p+1
q>1.
We will next show that Riemann sums with arbitrary choices of base points for the inte-
grand functions lead to the same L`evy area as just constructed.
Lemma 5. Let X∈ Vp([0, T ],Rd)for some 1p < . Suppose Xis self-controlled. Denote
by s[s, t]an arbitrary point chosen in a partition interval [s, t]π. Then L´evy’s area from
the preceding theorem is also given by
L(X)i,j = lim
|π|→0X
[s,t]π
(Xi
sXj
s,t Xj
sXi
s,t),1i, j d.
Proof. From Theorem 4 we already know that the left-point Riemann sums converge. Hence,
we only need to show that
X
[s,t]πn
(Xi
sXj
s,t Xj
sXi
s,t)X
[s,t]πn
(Xi
sXj
s,t Xj
sXi
s,t) (3)
tends to zero along every sequence of partitions (πn) such that the mesh |πn|converges to
zero. Indeed, we may write for a partition interval [s, t]
Xi
sXj
s,t Xj
sXi
s,t (Xi
sXj
s,t Xj
sXi
s,t) = Xi
s,sXj
s,t +Xj
s,sXi
s,t
=(X
s(i, j)Xj
s,s+Ri,j
s,s)Xj
s,t +Xj
s,s(X
s(i, j)Xj
s,t +Ri,j
s,t)
=Ri,j
s,sXj
s,t +Xj
s,sRi,j
s,t.
Taking again the same control function ωfor Xand Ri,j , we estimate
|Xi
sXj
s,t Xj
sXi
s,t (Xi
sXj
s,t Xj
sXi
s,t)|=| − Ri,j
s,sXj
s,t +Xj
s,sRi,j
s,t| ≤ 2ω(s, t)θ
5
with θ:= 2
p+1
p>1. Recalling the superadditivity of ω, we get for nN
X
[s,t]πn
(Xi
s,sXj
s,t Xj
s,sXi
s,t)
X
[s,t]πn
ω(s, t)θmax
[s,t]πn
ω(s, t)θ1ω(0, T ),
which means that (3) tends to zero as n→ ∞.
Example 6. Let (Bt;t[0, T ]) be a standard Brownian motion on a probability space
(Ω,F,P) and fC2(R,R) a twice continuously differentiable function. The trajectories of
Bare of finite p-variation for all p > 2 outside a null set Nand thus we can deduce from
Theorem 4 that L´evy’s area of (B , f(B)) exists outside the same null set N.
The following example illustrates that for p2 things are essentially different. It will in
particular show that in this case self-control of a path is necessary for the existence of L´evy’s
area.
Example 7. Let us consider the function X: [1,1] R2with components given by
X1
t:=
X
k=1
aksin(2kπt) and X2
t:=
X
k=1
akcos(2kπt),
where ak:= 2αk and α[0,1]. For mNlet Xm: [1,1] R2be the mth finite partial
sum of the infinite series with components
X1,m
t:=
m
X
k=1
aksin(2kπt) and X2,m
t:=
m
X
k=1
akcos(2kπt),
for t[1,1]. These functions are α-H¨older continuous uniformly in m. Indeed, let s, t
[1,1] and choose kNsuch that 2k1≤ |st| ≤ 2k. Then we can estimate as follows
|X1,m
tX1,m
s|=
m
X
l=1
al(sin(2lπt)sin(2lπs))
=
m
X
k=1
al2 cos(2l1π(s+t)) sin(2l1π(st))
= 2
k
X
l=1
|al|| sin(2l1π(st))|+ 2
X
l=1
|al|
2
k
X
l=1
|al|2l1π|st|+ 2
X
l=k+1
|al|
k
X
l=1
2lαlπ|st|+ 2α(k+1)+1 1
12α
2(k+1)(1α)1
21α1π|st|+21α
12α|st|α
2(k+1)(1α)1
21α1π2k(1α)|st|α+21α
12α|st|αC|st|α.
6
for some constant C > 0 independent of mN. Analogously, we can get the α-H¨older
continuity of X2,m . Furthermore, it can be seen with the same estimate that (Xm) converges
uniformly to Xand thus also in α-H¨older topology. The limit function Xis not β-H¨older
continuous for every β > α. In order to see this, choose s= 0 and t=tn= 2nfor nN
and observe that
|X1
tnX1
0|
|tn0|β=
n1
X
k=1
2αk+βn sin(2knπ)2(βα)n+α,
which obviously tends to infinity with n. Since α-H¨older continuity is obviously related to
finite 1
α-variation, we can conclude that X V 1
α([1,1],R2),and X6∈ Vγ([1,1],R2) for
γ < 1
α.Let us now show that Xpossesses no L´evy area. For this purpose, fix α[0,1] and
mN. Then L´evy’s area for Xmis given by
Z1
1
X1,m
sdX2,m
sZ1
1
X2,m
sdX1,m
s
=
m
X
k,l=1
akalZ1
1sin(2kπs) sin(2lπs)2lπ+ cos(2lπs) cos(2kπs)2kπds
=
m
X
k,l=1
akal2lπZ1
1
1
2(cos((2k2l)πs)cos((2k+ 2l)πs)) ds
+ 2kπZ1
1
(cos((2k2l)πs) + cos((2k+ 2l)πs)) ds
=2
m
X
k=1
a2
k2kπ=2
m
X
k=1
2(12α)kπ.
This quantity diverges as mtends to infinity for 1
α2. Since (Xm) converges to Xin
the α-H¨older topology, we can use this result to choose partition sequences of [1,1] along
which Riemann sums approximating the L´evy area of Xdiverge as well. This shows that
Xpossesses no L´eva area. In return Theorem 4 implies that Xcannot be self-controlled.
However, it is not to hard to see directly that no regularity is gained by controlling X1with
X2. For this purpose, note that for 1st1, and 0 6=X
sR
|X1
s,t X
sX2
s,t|=
X
k=1
ak[(sin(2kπt)sin(2kπs)) X
s(cos(2kπt)cos(2kπs))]
=
2
X
k=1
aksin(2k1π(st)) cos(2k1π(s+t)) + X
ssin(2k1π(s+t)) sin(2k1π(st))
=
2
X
k=1
aksin(2k1π(st))p1 + (X
s)2sin(2k1π(s+t) + arctan((X
s)1))
.
Let us now consider the H¨older regularity for s= 0. First, assume X
0>0, and take t= 2n
to obtain
|X1
0,2nX
0X2
0,2n|
2βn = 2βn
2
n
X
k=1
aksin(2k1nπ)q1 + (X
0)2sin(2k1nπ+ arctan((X
0)1))
2(βα)nsin π
2+ arctan((X
0)1).
7
For X
0<0 the same estimates work for tn=2ninstead. Therefore, the H¨older regularity
at 0 cannot be better than αand in particular Xcannot be self-controlled for 1
α>2.
The sewing lemma (Lemma 3) not only guarantees the existence of L`evy’s area. As we
will now show, it also allows to conclude that the area is regular as a function of time.
Corollary 8. Let Ξ : ∆TRdbe a continuous function and K > 0some constant. Assume
that there exist a control function ω, a constant θ > 1, and r > 1such that
|Ξs,t|rω(s, t)and |Ξs,t Ξs,u Ξu,t | ≤ Kω(s, t)θ,0sutT .
Then there exists a unique continuous function Φ: [0, T ]Rdsuch that Φ(0) = 0 and
|Φ(t)Φ(s)Ξs,t| ≤ C(θ)ω(s, t)θ,0stT ,
with C(θ)from Lemma 3. Φalso satisfies for (s, t)T|Φ(t)Φ(s)|rC2(r, θ , ω)ω(s, t)
with some constant C2(r, θ , ω)>0.
Proof. Under the given assumptions, Lemma 3 provides a function Φ: [0, T ]Rdsuch that
Φ(0) = 0 and |Φ(t)Φ(s)Ξs,t | ≤ C(θ)ω(s, t)θfor (s, t)T. In order to prove the
r-variation inequality for Φ, we estimate for (s, t)T
|Φ(t)Φ(s)|r|Ξs,t|+C(θ)ω(s, t)θr
2r1|Ξs,t|r+ 2r1C(θ)rω(s, t)2r1+ 2r1C(θ)rω(0, T ) 1ω(s, t).
Remark 9. For consistency we state Corollary 8 only for a continuous function Ξ: ∆TRd.
Yet, it still holds true without the continuity assumption and for a general Banach space
replacing Rd, since the sewing lemma also holds under these weaker conditions. See Theorem
1 and Remark 3 in [FDM08].
For instance, Corollary 8 implies that for p < 2 the mapping
L:Vp([0, T ],R2)→ Vp([0, T ],R2),
X7→ Lt(X) := lim
|π(0,t)|→0X
[u,v]π(0,t)
(X1
uX2
u,v X2
uX1
u,v)
is locally Lipschitz continuous since it satisfies the estimate
||L·(X)||rC2(r, θ, ||X||p)||X||p.(4)
3 ollmer integration
In his well-known paper F¨ollmer [F¨ol79] considered one dimensional pathwise integrals. He
was able to give a pathwise meaning to the limit
ZT
0
DF(Xt) dπnXt:= lim
n→∞ X
[s,t]πn
hDF(Xs), Xs,ti,
8
provided FC2(Rd,R). His starting point was the hypothesis that quadratic variation of
XC([0, T ],Rd) exists along a sequence of partitions (πn)nNwhose mesh tends to zero.
Here ,·i denotes the usual inner product on Rd. This is today known as ollmer integration.
As indicated and discussed below, this construction of an integral depends strongly on the
chosen sequence of partitions (πn)nN.
Before coming back to an approach of F¨ollmer’s integral, we shall construct a Stratonovich
type integral, thereby discussing the problem of dependence on a chosen sequence of partitions.
As in the previous section, our approach is based on the notion of controlled paths. This will
also lead us on a route which does not require the existence of iterated integrals as in the
classical rough paths approach. We fix a γ[0,1], to discuss Stratonovich limits for Riemann
sums where integrands are taken at convex combinations γXs+ (1 γ)Xtof the values of
Xat the extremes of a partition interval [s, t]. We start by decomposing these sums into
symmetric and antisymmetric parts. For p, q [1,), X, Y ∈ Vp([0, T ],Rd) , YCq
Xwe
have
γ-ZT
0
YtdXt:= lim
|π|→0X
[s,t]π
hYs+γYs,t , Xs,ti
=1
2lim
|π|→0X
[s,t]π
(hYs+γYs,t , Xs,t +hXs+γXs,t, Ys,t i)
+1
2lim
|π|→0X
[s,t]π
(hYs+γYs,t , Xs,ti − hXs+γXs,t , Ys,ti)
=1
2γ-ZT
0
YtdXt+γ-ZT
0
XtdYt+1
2γ-ZT
0
YtdXtγ-ZT
0
XtdYt
=1
2SγhX, Y i+1
2AγhX, Y i.(5)
Note that γ= 0 corresponds to the classical Itˆo integral and γ=1
2to the classical Stratonovich
integral.
If the variation orders of Xand Yfulfil 1/p + 1/q > 1, we are in the framework of Young’s
integration theory. Below 1, either the existence of the rough path or control is needed. To
illustrate this, we go back to Example 7.
Example 10. Let X= (X1, X 2) be given according to Example 7. In this case, we have
seen that X1and X2are of finite 1
α-variation. With decomposition (5), we see that
1
2-Z1
0
X2
tdX1
t:= lim
|π|→0X
[s,t]π
hX2
s+1
2X2
s,t, X 1
s,ti=1
2S1
2hX1, X2i+1
2A1
2hX1, X2i
=1
2lim
|π|→0X
[s,t]π
(hX2
s+X2
t, X1
tX1
si+hX1
s+X1
t, X2
tX2
si) + 1
2L1,2(X)
=1
2lim
|π|→0X
[s,t]π
hX1, X2is,t +1
2L1,2(X)
=1
2(X1
1X2
1X1
0X2
0) + 1
2L1,2(X).
9
Therefore, the integral exists if and only if L´evy’s area exists, which is not the case for instance
if α=1
2.So beyond Young’s theory, the existence of the Stratonovich-1
2integral is closely
linked to the existence of L´evy’s area.
Using a suitable control concept, we will next construct the Stratonovich integral described
above, but not just with restriction to a particular sequence of partitions. This time, due to
the fact that for twice continuously differentiable Fthe Hessian matrix D2Fis symmetric,
the usual concept of controlled paths is sufficient.
Theorem 11. Let γ[0,1],X∈ Vp([0, T ],Rd)and YCq
X. If Y
tis a symmetric matrix
for all t[0, T ], then the antisymmetric part
AγhX, Y i:= lim
|π|→0X
[s,t]πhYs+γYs,t , Xs,ti − hXs+γXs,t , Ys,ti,(6)
exists and satisfies
AγhX, Y i=AhX, Y i:= lim
|π|→0X
[s,t]πhYs, Xs,ti − hXs, Ys,t i
for every choice of points s[s, t]π.
Proof. It is easy to verify that by definition the antisymmetric part, if it exists as a limit of
the Riemann sums considered, has to satisfy the second formula of the claim at least with the
choice s=s, for all intervals [s, t] belonging to a partition. To prove that this limit exists,
we use Lemma 3. For this purpose, we set Ξs,t := hYs, Xs,t i − hXs, Ys,tifor (s, t)T. Since
Yis controlled by X, we obtain
Ξs,t Ξs,u Ξu,t =hYs, Xs,ti − hXs, Ys,ti − hYs, Xs,u i+hXs, Ys,ui − hYu, Xu,ti+hXu, Yu,t i
=hYu,t, Xs,u i − hXu,t, Ys,ui
=hY
uXu,t +RY
u,t, Xs,u i − hXu,t, Y
sXs,u +RY
s,ui
=hRY
u,t, Xs,u i − hXu,t, RY
s,ui+hY
uXu,t, Xs,u i − hXu,t, Y
sXs,ui
=hRY
u,t, Xs,u i − hXu,t, RY
s,ui+hXu,t , Y
uXs,u Y
sXs,ui
for 0 s < u < t T, where we used hY
uXu,t, Xs,u i=hXu,t, Y
uXs,uiin the last line thanks
to symmetry. With the same control ωfor all functions involved as above, this gives
|Ξs,t Ξs,u Ξu,t| ≤ ω(s, t)
1
p+1
r+ω(s, t)
1
p+1
r+ω(s, t)
2
p+1
q3ω(s, t)θ
with θ:= 2
p+1
q>1. So from Lemma 3 we conclude that the left-point Riemann sums
converge. It remains to show that
X
[s,t]πnhYs, Xs,ti − hXs, Ys,t iX
[s,t]πnhYs, Xs,ti − hXs, Ys,ti(7)
tends to zero along every sequence of partitions (πn) such that the mesh |πn|converges to
zero. Applying the symmetry of Y, we get
hYs, Xs,ti − hXs, Ys,t i−hYs, Xs,ti − hXs, Ys,ti=hYs,s, Xs,ti − hXs,s, Ys,t i
=hY
sXs,s+RY
s,s, Xs,ti − hXs,s, Y
sXs,t +RY
s,ti
=hRY
s,s, Xs,ti − hXs,s, RY
s,ti,
10
and thus
hYs, Xs,ti − hXs, Ys,t i − hYs, Xs,ti − hXs, Ys,t i
ω(s, t)θ
with θ:= 1
p+1
r>1, where we choose the same control function ωfor Xand RY. Therefore,
the properties of ωimply
X
[s,t]πnhYs,s, Xs,ti − hXs,s, Ys,ti
X
[s,t]πn
ω(s, t)θmax
[s,t]πn
ω(s, t)θ1ω(0, T ),
which means that (7) tends to zero as |πn|tends to zero.
Remark 12. 1. The proof of Theorem 11 analogously works under the assumption that X
is controlled by Yand X
tis a symmetric matrix for all t[0, T ].
2. If Yis controlled by Xand Y
tis an antisymmetric matrix for all t[0, T ], then an
analogous result to Theorem 11 holds true for the symmetric part SγhX, Y i.
In case γ=1
2as in the example above, the symmetric part simplifies considerably, and
therefore the preceding Theorem will already imply the existence of the 1
2-Stratonovich inte-
gral.
Corollary 13. Let X∈ V p([0, T ],Rd),YCq
Xand suppose Y
tis a symmetric matrix for all
t[0, T ]. Then, the Stratonovich integral
ZT
0
YtdXt:= lim
|π|→0X
[s,t]π
hYs+1
2Ys,t, Xs,t i(8)
exists and satisfies
1
2-ZT
0
YtdXt=ZT
0
YtdXt=1
2hYT, XTi − hY0, X0i+1
2AhX, Y i.
Proof. By equation (5) we may separately deal with the symmetric part S1
2hX, Y iand the
antisymmetric part A1
2hX, Y iof the integral 1
2-RT
0YtdXt. The existence of the antisymmetric
part A1
2hX, Y ifollows from Theorem 11. For the symmetric part, note that as in Example 7
for (s, t)T
2hYs+1
2Ys,t, Xs,t i+hXs+1
2Xs,t, Ys,t i=hYs+Yt, Xs,ti+hXs+Xt, Ys,t i
=hYs, Xti − hYs, Xsi+hYt, Xti − hYt, Xsi+hXs, Yti − hXs, Ysi+hXt, Yti − hXt, Ysi
= 2hY, X is,t.
Therefore, S1
2hX, Y iis given by
S1
2hX, Y i= lim
|π|→0X
[s,t]πhYs+1
2Ys,t, Xs,t i+hXs+1
2Xs,t, Ys,t i=hYT, XTi − hX0, Y0i.(9)
11
The treatment of γ-Stratonovich integrals above has shown that the corresponding asym-
metric component can be treated by means of the concept of path control. In the case γ6=1
2,
a symmetric term is left to consider. This does not seem to be possible by means of the
ideas used for the asymmetric component. And this brings us back to F¨ollmer’s approach.
Our treatment of the symmetric part is much related to F¨ollmer’s treatment of quadratic
variation, and will therefore be strongly dependent on partition sequences. For this purpose
we define the quadratic variation in the sense of F¨ollmer (cf. [F¨ol79]) and call a sequence of
partitions (πn)increasing if for all [s, t]πnthere exist [ti, ti+1 ]πn+1,i= 1, ..., N , such
that [s, t] = SN
i=1[ti, ti+1 ].
Definition 14. Let (πn) be an increasing sequence of partitions such that limn→∞ |πn|= 0.
A continuous function f: [0, T ]Rhas quadratic variation along (πn) if the sequence of
discrete measures on ([0, T ],B([0, T ])) given by
µn:= X
[s,t]πn
|fs,t|2δs(10)
converges weakly to a measure µ, where δsdenotes the Dirac measure at s[0, T ]. We write
[f]tfor the “distribution function” of the interval measure associated with µ. A continuous
path X= (X1, ..., Xd) has quadratic variation along (πn) if (10) holds for all Xiand Xi+Xj,
1i, j d. In this case, we set
[Xi, Xj]t:= 1
2([Xi+Xj]t[Xi]t[Xj]t), t [0, T ].
Remark 15. Since in our situation the limiting distribution function is continuous, the weak
convergence is equivalent to the uniform convergence of the distribution function. Hence,
X= (X1, ..., Xd)C([0, T ],Rd)has quadratic variation in the sense of F¨ollmer if and only
if
[Xi, Xj]n
t:= X
[u,v]πn
Xi
ut,vtXj
ut,vt
convergences uniformly to [Xi, Xj]·in C([0, T ],R)for all 1i, j d, where ut:= min{u, t}.
See Lemma 32 in [PP13].
Remark 16. Let us emphasize here that quadratic variation should not be confused with the
notion of 2-variation: quadratic variation depends on the choice of a partition sequence (πn),
2-variation does not. In fact, for every continuous function fC([0, T ],R)there exits a
sequence of partitions (πn)with limn→∞ |πn|= 0 such that [f , f]t= 0 for all t[0, T ]. See
for instance Proposition 70 in [Fre83].
The existence of quadratic variation guaranteed, F¨ollmer was able to prove a pathwise
version of Itˆo’s formula. In his case, the construction of the integral is closely linked to the
partition sequence chosen for the quadratic variation. We will now aim at combining the
techniques of controlled paths with the quadratic variation hypothesis, and derive a pathwise
version of Itˆo’s formula for paths with finite variation, in which the quadratic variation term
may depend on a partition sequence, but the integral does not. As a first step, we derive the
existence of γ-Stratonovich integrals for any γ[0,1].To do so, we will need the following
technical lemma the easy proof of which is left to the reader.
12
Lemma 17. Let p1,(πn)be an increasing sequence of partitions such that limn|πn|=
0,X∈ Vp([0, T ],Rd)with quadratic variation along (πn)and YCq
X. In this case the
quadratic covariation of Xand Yexists and is given by
[Y, X ]t:= lim
n→∞ X
[s,t]πn
hXs,t, Ys,t i=X
1i,jdZT
0
Y
t(i, j) dπn[Xi, X j]t,
where Y
t= (Y
t(i, j))1i,j d,0tT.
Theorem 18. Let X∈ Vp([0, T ],Rd),YCq
Xand suppose Y
tis a symmetric matrix for
all t[0, T ]. Let (πn)be a increasing sequence of partitions such that limn→∞ |πn|= 0 and
Xhas quadratic variation along (πn). Then for all γ[0,1] the γ-RYtdπnXtintegral exists
and is given by
γ-ZT
0
YtdπNXt=ZT
0
YtdXt+1
2(2γ1) X
1i,jdZT
0
Y
t(i, j) dπn[Xi, X j]t,
where Y
t= (Y
t(i, j))1i,j d.
Proof. Fix γ[0,1]. As before we split the sum in (5) into its symmetric and antisymmetric
parts:
X
[s,t]πn
hYs+γYs,t , Xs,ti=1
2X
[s,t]πnhYs+γYs,t , Xs,ti+hXs+γXs,t , Ys,ti
+1
2X
[s,t]πnhYs+γYs,t , Xs,ti − hXs+γXs,t , Ys,ti.
The second sum converges for every sequence of partitions (πn) such that limn→∞ |πn|= 0
and is independent of γthanks to Theorem 11. Taking γ= 1/2 we can apply Corollary 13
to see that 1
2AhX, Y i=ZT
0
YtdXt1
2hXT, YTi − hX0, Y0i.(11)
For the symmetric part, we note for (s, t)T
hYs+γYs,t , Xs,ti+hXs+γXs,t , Ys,ti
=(1 γ)hYs, Xs,ti+hXs, Ys,t i+γhYt, Xs,ti+hXt, Ys,t i
=(1 γ)hYt, Xti − hYs, Xsi+hYs, Xti − hYt, Xti+hXs, Ys,ti
+γhYt, Xti − hYs, Xsi+hYs, Xsi − hYt, Xsi+hXt, Ys,ti
=(1 γ)hYt, Xti − hYs, Xsi − hXs,t, Ys,ti+γhYt, Xti − hYs, Xsi+hXs,t, Ys,ti
=hYt, Xti − hYs, Xsi+ (2γ1)hXs,t, Ys,t i.
Thus the first sum reduces to
1
2X
[s,t]πnhYs+γYs,t , Xs,ti+hXs+γXs,t, Ys,ti
=1
2hYT, XTi − hY0, X0i+2γ1
2X
[s,t]πn
hXs,t, Ys,t i.
Therefore, the symmetric part converges along (πn), and the assertion follows by (11) and
Lemma 17.
13
An application of Theorem 18 to the particular case Y= DF(X) for Fsmooth enough
provides the classical Stratonovich formula.
Lemma 19. Let 1p < 3,X∈ Vp([0, T ],Rd)and FC2(Rd,R). Suppose that the second
derivative D2Fis α-H¨older continuous of order α > max{p2,0}. Then the Stratonovich
integral RhDF(Xt)dXtiexists and is given by
ZT
0
DF(Xt)dXt=F(XT)F(X0).
Proof. Let X= (X1, ..., Xd)∈ Vp([0, T ],Rd) for 1 p < 3. Then, with r=p
2in the
definition of controlled paths we easily see that DF(X)Cp
X. Thus by Corollary 13 the
(1
2-)Stratonovich integral is well-defined and independent of the chosen sequence of partitions
(πn) along which the limit is taken. Now choose an increasing sequence of partitions (πn) such
that limn→∞ |πn|= 0 and [X]t= 0 along (πn) for t[0, T ] (cf. Proposition 70 in [Fre83]).
Applying Taylor’s theorem to F, we observe that
F(XT)F(X0) = 1
2X
[s,t]πn(F(Xt)F(Xs)) (F(Xs)F(Xt))
=1
2X
[s,t]πn
hDF(Xs), Xs,ti+1
2X
1i,jd
Di,j(Xs)Xi
s,tXj
s,t +R(Xs, Xt)
+1
2X
[s,t]πn
hDF(Xt), Xs,ti − 1
2X
1i,jd
Di,j(Xt)Xi
s,tXj
s,t +˜
R(Xs, Xt)
=X
[s,t]πn
h1
2DF(Xs) + 1
2DF(Xt), Xs,ti+1
4X
[s,t]πnX
1i,jd
(Di,j (Xs)Di,j (Xt))Xi
s,tXj
s,t
+X
[s,t]πn
(R(Xs, Xt) + ˜
R(Xs, Xt)),
where |R(x, y)|+|˜
R(x, y)| ≤ φ(|xy|)|xy|2,for some increasing function φ: [0,)R
such that φ(c)0 as c0. Since Xis continuous and has zero quadratic variation along
(πn), the last two terms converge to 0 as n→ ∞, and we obtain
ZT
0
DF(Xt)dXt= lim
n→∞ X
[s,t]πn
hDF(Xs) + 1
2(DF(Xt)DF(Xs)), Xs,ti=F(XT)F(X0).
We can now state and prove the announced version of the pathwise formula by ollmer
(cf. [F¨ol79]).
Corollary 20. Let 1p < 3,γ[0,1] and (πn)be an increasing sequence of partitions such
that limn→∞ |πn|= 0. Assume FC2(Rd,R)with α-H¨older continuous second derivative
D2Ffor some α > max{p2,0}. If X∈ Vp([0, T ],Rd)has quadratic variation along (πn),
then the formula
F(XT) = F(X0) + γ-ZT
0
DF(Xt) dπnXt1
2(2γ1) X
1i,jdZT
0
DiDjF(Xs) dπn[Xi, Xj]s
14
holds.
Proof. Combine the results of Theorem 18 and Lemma 19.
The assumptions, that Xis of finite p-variation for some 1 p < 3 and that the second
derivative D2Fis α-H¨older continuous for some α > max{p2,0}can be considered the price
we have to pay for obtaining an integral the antisymmetric part of which does not depend on
the chosen partition sequence. ollmer [F¨ol79] does not need these hypotheses. However, his
integral is only defined along the sequence of partitions for which the existence of quadratic
variation is claimed.
So far we have always assumed that the Gubinelli derivative of a controlled path is sym-
metric. This assumption can be avoided if the paths involved control each other.
Definition 21. Let X, Y ∈ Vp([0, T ],Rd). We say that Xand Yare similar if there exist
X, Y ∈ V q([0, T ],Rd×d) such that XCq
Ywith Gubinelli derivative X,YCq
Xwith
Gubinelli derivative Y, and ((X
t))1=Y
tfor all t[0, T ]. In this case we write YSq
X.
Let us give an very simple example of two paths X, Y ∈ Vp([0, T ],Rd) such that YSq
X
but neither YCq
Xwith Ysymmetric nor XCq
Ywith Xsymmetric.
Example 22. For p[2,3) take X1∈ Vp([0, T ],R) and X2, X3 V p
2([0, T ],R). If we set
X= (X1, X2, X3) and Y= (X1,0,0), we obviously have X, Y ∈ Vp([0, T ],R3). In this case
we could choose Xand Yidentical to
1 0 0
0 0 1
0 1 0
.
We see that YSp
X, but Xand Yare not symmetric matrices.
For similar paths the antisymmetric part of the integral also exists and is independent of
the choice of a particular sequence of partitions along which Riemann sums are defined.
Theorem 23. Let 1p < and X∈ Vp([0, T ],Rd). If YSq
X, then for every γ[0,1]
the antisymmetric part of the integral as defined in (6) exists and satisfies
AγhX, Y i=AhX, Y i= lim
|π|→0X
[s,t]πhYs, Xs,ti − hXs, Ys,t i,
where sis an arbitrary point in [s, t]for each [s, t]π.
Proof. By Lemma 3 it is again sufficient to check condition (2) for Ξs,t := hYs, Xs,ti−hXs, Ys,ti.
Because Xand Yare similar, we obtain
Ξs,tΞs,u Ξu,t =hYs, Xs,t i − hXs, Ys,ti − hYs, Xs,ui+hXs, Ys,u i − hYu, Xu,ti+hXu, Yu,ti
=hXs,u, Yu,t i − hYs,u, Xu,ti
=hX
sYs,u +RX
s,u, Y
uXu,t +RY
u,ti − hYs,u , Xu,ti
=hX
sYs,u, RY
u,ti+hRX
s,u, Y
uXu,ti+hRX
s,u, RY
u,ti+hX
sYs,u, Y
uXu,ti − hYs,u , Xu,ti
15
for 0 sutT. The second last term in the preceding formula can be rewritten as
hX
sYs,u, Y
uXu,ti − hYs,u , Xu,ti=hYs,u ,(X
t)Y
uYu,t Xu,ti=hYs,u ,(X
t)(Y
uY
t)Xu,ti,
where we applied ((X
t))1=Y
t. Since the finite sum of control functions is again a control
function, we can choose the same control function ωfor X, X, RXand Y, Y , RY, and therefore
|Ξs,t Ξs,u Ξu,t|
≤||X||ω(s, u)
1
pω(u, t)1
r+||Y||ω(s, u)1
rω(u, t)
1
p+ω(s, u)1
rω(u, t)1
r
+||X||ω(s, u)
1
qω(u, t)
1
pω(s, u)
1
p
≤||X||ω(s, t)
1
p+1
r+||Y||ω(s, t)
1
r+1
p+ω(s, t)1
r+1
r+||X||ω(s, t)
1
q+2
p,
where || · ||denotes the supremum norm. Setting θ:= 2/p + 1/q > 1, we conclude that
|Ξs,t Ξs,u Ξu,t| ≤ 2||X||+||Y||+ω(0, T )
1
qω(s, t)θ.
We therefore have shown that the left-point Riemann sums converge. It remains to prove
that (7) goes to zero along every sequence of partitions (πn) such that the mesh |πn|tends to
zero. Since Xand Yare similar, we observe that for (s, t)T, and s[s, t]
hYs, Xs,ti − hXs, Ys,t i − hYs, Xs,ti − hXs, Ys,ti=hYs,s, Xs,t i − hXs,s, Ys,t i
=hY
sXs,s+RY
s,s, X
sYs,t +RX
s,ti − hXs,s, Ys,ti
=hY
sXs,s, X
sYs,ti+hY
sXs,s, RX
s,ti+hRY
s,s, X
sYs,ti+hRY
s,s, RX
s,ti − hYs,t , Xs,si
=hY
sXs,s, RX
s,ti+hRY
s,s, X
sYs,ti+hRY
s,s, RX
s,ti.
To obtain the last line, we once again use ((X
s))1=Y
s. Taking again the same control
function ωfor X, X, RXand Y, Y , RY, we estimate
hYs, Xs,ti − hXs, Ys,ti − hYs, Xs,t i − hXs, Ys,ti
≤ ||Y||ω(s, t)
1
p+1
r+||X||ω(s, t)
1
p+1
r+ω(s, t)1
r+1
r+||Y||ω(s, t)
1
p+1
p+1
q
||X||+ 2||Y||+ω(0, T )
1
qω(s, t)θ,
with θ:= 2
p+1
p>1. Superadditivity of ωfinally gives
X
[s,t]πnhYs,s, Xs,ti − hXs,s, Ys,ti
||X||+ 2||Y||+ω(0, T )
1
qmax
[s,t]πn
ω(s, t)θ1ω(0, T ),
which means that (7) tends to zero as |πn|tends to zero.
We conclude this section by presenting an analogue of Theorem 18 for similar paths.
Theorem 24. Let (πn)be an increasing sequence of partitions such that limn→∞ |πn|= 0,
X∈ Vp([0, T ],Rd)and YSq
X.
16
1. Then the Stratonovich integral as defined in (8) exists and fulfills
1
2-ZT
0
YtdXt=ZT
0
YtdXt=1
2hXT, YTi − hX0, Y0i+1
2AhX, Y i.
2. If Xhas quadratic variation along (πn), then the limit γ-RT
0YtdXtas defined in (5)
exists for all γ[0,1] and satisfies
γ-ZT
0
YtdXt=ZT
0
YtdXt+1
2(2γ1) X
1i,jdZT
0
Y
t(i, j) d[Xi, X j]t,
where Y
t= (Y
t(i, j))1i,j d.
Proof. The first statement follows as Corollary 13 by replacing Theorem 11 with Theorem 23
in the proof. The second statement is proved in analogy to the proof of Theorem 18.
References
[FD06] Denis Feyel and Arnaud De La Pradelle, Curvilinear integrals along enriched paths.,
Electron. J. Probab. 11 (2006), 860–892.
[FDM08] Denis Feyel, Arnaud De La Pradelle, and Gabriel Mokobodzki, A non-commutative
sewing lemma., Electron. Commun. Probab. 13 (2008), 24–34.
[FH13] Peter Friz and Martin Hairer, A short course on rough paths, in preparation (2013).
[F¨ol79] Hans F¨ollmer, Calcul d’Itˆo sans probabilit´es, eminaire de Probabilit´es XV 80
(1979), 143–150 (French).
[Fre83] David Freedman, Brownian motion and diffusion, second ed., Springer-Verlag, New
York, 1983. MR 686607 (84c:60121)
[GIP13] Massimiliano Gubinelli, Peter Imkeller, and Nicolas Perkowski, Paracontrolled dis-
tributions and singular PDEs, preprint arXiv:1210.2684v2 (2013).
[GIP14] , A Fourier Approach to pathwise stochastic integration, in preparation,
2014.
[Gub04] Massimiliano Gubinelli, Controlling rough paths, J. Funct. Anal. 216 (2004), no. 1,
86–140.
[LCL07] Terry J. Lyons, Michael Caruana, and Thierry L´evy, Differential equations driven
by rough paths, Lecture Notes in Mathematics, vol. 1908, Springer, Berlin, 2007.
MR 2314753 (2009c:60156)
[Lej09] Antoine Lejay, eminaire de Probabilit´es XLII, Lecture Notes in Math. 1979 (2009).
[L´ev40] Paul L´evy, Le mouvement brownien plan, Amer. J. Math. 62 (1940), 487–550
(French). MR 0002734 (2,107g)
17
[Lyo98] Terry J. Lyons, Differential equations driven by rough signals, Rev. Mat. Iberoam.
14 (1998), no. 2, 215–310.
[Per14] Nicolas Simon Perkowski, Studies of robustness in stochastic analysis and mathe-
matical finance, Ph.D. thesis, 2014.
[PP13] Nicolas Perkowski and David J. Pr¨omel, Pathwise stochastic integrals for model free
finance, Preprint arXiv:1311.6187 (2013).
[You36] Laurence C. Young, An inequality of the H¨older type, connected with Stieltjes inte-
gration, Acta Math. 67 (1936), no. 1, 251–282.
18
... By this, we mean an R d valued function whose components are univariate Weierstrass functions with possibly different parameters. A two dimensional Weierstrass function has been studied in [10,11]. In [10] the authors consider the two dimensional Weierstrass function ...
... A two dimensional Weierstrass function has been studied in [10,11]. In [10] the authors consider the two dimensional Weierstrass function ...
... However, the same arguments work with little difference if we replace every cos with sin. However, as shown in [10] and further studied in [11], if we replace one cos with sin, the arguments above no longer work. ...
Preprint
Full-text available
Rough paths theory allows for a pathwise theory of solutions to differential equations driven by highly irregular signals. The fundamental observation of rough paths theory is that if one can define ``iterated integrals" above a signal, then one can construct solutions to differential equations driven by the signal. The typical examples of the signals of interest are stochastic processes such as (fractional) Brownian motion. However, rough paths theory is not inherently random and therefore can treat irregular deterministic driving signals such as a (multivariate) Weierstrass function. To the authors' best knowledge, no explicit construction of a rough path (the ``iterated integrals") above a multivariate Weierstrass function has been constructed, nor has there been an explicit solution to a differential equation driven by a multivariate Weierstrass function. This note supplies a construction of a rough path above a multivariate Weierstrass function. We conclude with some illustrations and some examples of solving differential equations driven by rough Weierstrass functions.
... This Fourier analytic concept generalizes the original control concept introduced by Gubinelli [5]. In search of a good example of a pair of functions not controlling each other, in [12] the authors came up with the pair of Weierstrass functions defined by ...
... Hence when the first one has minimal increments, the second one has maximal ones, and vice versa. This can be seen to mathematically rigorously underpin the fact that they are mutually not controlled (see [12,Example 2.8]). It also implies that the Lévy areas of the approximating finite sums of the representing series do not converge. ...
Preprint
Full-text available
We compute the Hausdorff dimension of a two-dimensional Weierstrass function, related to lacunary (Hadamard gap) power series, that has no L\'evy area. This is done by interpreting it as a pullback attractor of a dynamical system based on the Baker transformation. A lower bound for the Hausdorff dimension is obtained by investigating the pushforward of the Lebesgue measure on the graph along scaled neighborhoods of stable fibers of the underlying dynamical system following the graph. Scaling ideas are crucial. They become accessible by self similarity properties of a mapping whose increments coincide with vertical distances on the stable fibers.
... This Fourier analytic concept generalizes the original notion of control introduced by Gubinelli [8]. In search of a good example of two-dimensional functions for which no component is controlled by the other one, in [15] we come up with a pair of Weierstrass functions W = (W 1 , W 2 ). One of them fluctuates on all dyadic scales in a sinusoidal manner, the other one in a cosinusoidal one. ...
Preprint
We investigate Takagi-type functions with roughness parameter $\gamma$ that are H\"older continuous with coefficient $H=\frac{\log\gamma}{\log \eh}.$ Analytical access is provided by an embedding into a dynamical system related to the baker transform where the graphs of the functions are identified as their global attractors. They possess stable manifolds hosting Sinai-Bowen-Ruelle (SBR) measures. We show that the SBR measure is absolutely continuous for large enough $\gamma$. Dually, where duality is related to time reversal, we prove that for large enough $\gamma$ a version of the Takagi-type curve centered around fibers of the associated stable manifold possesses a square integrable local time.
... We thus do not seek further extensions of the setup, e.g., to cover time-dependent functions f , cf. [12], path-dependent functions, cf. [6,18,34], nor to develop higher order local times in the spirit of [7] for càdlàg paths. These, while interesting, would distract from the main focus of the paper and are left as avenues for future research. ...
... This Fourier analytic concept generalizes the original notion of control introduced by Gubinelli [8]. In search of a good example of two-dimensional functions for which no component is controlled by the other one, in [14] we come up with a pair of Weierstrass functions W = (W 1 , W 2 ). One of them fluctuates on all dyadic scales in a sinusoidal manner, the other one in a cosinusoidal one. ...
Preprint
Full-text available
We investigate Weierstrass functions with roughness parameter $\gamma$ that are H\"older continuous with coefficient $H={\log\gamma}/{\log \frac12}.$ Analytical access is provided by an embedding into a dynamical system related to the baker transform where the graphs of the functions are identified as their global attractors. They possess stable manifolds hosting Sinai-Bowen-Ruelle (SBR) measures. We show that the SBR measure is absolutely continuous for large enough $\gamma$ with square-integrable density. For this purpose, we use a telescoping property of associated measures and the transversality of the flow related to the mapping describing the stable manifold. This smoothness property of the SBR measure can be used to compute the Hausdorff dimension of the graphs of the original Weierstrass functions.
... We thus do not seek further extensions of the setup, e.g., to cover time-dependent functions f , cf. [FZ06], path-dependent functions, cf. [CF10,IP15,Sap18], nor to develop higher order local time in the spirit of [CP19] for càdlàg paths. These, while interesting, would distract from the main focus of the paper and are left as avenues for future research. ...
Preprint
Three concepts of local times for deterministic c{\`a}dl{\`a}g paths are developed and the corresponding pathwise Tanaka--Meyer formulae are provided. For semimartingales, it is shown that their sample paths a.s. satisfy all three pathwise definitions of local times and that all coincide with the classical semimartingale local time. In particular, this demonstrates that each definition constitutes a legit pathwise counterpart of probabilistic local times.
... Our results thus complement previous results on the pathwise approach to Itô calculus [2,5,7,9,9,10,15,21,25,29] by identifying a set of paths for which these results are robust to the choice of the sequence of partitions involved in the construction. In contrast to the constructions in [22,23,24], our construction does not involve path-dependent partitions and does not rely on any probabilistic tools. ...
Preprint
Full-text available
We study the concept of quadratic variation of a continuous path along a sequence of partitions and its dependence with respect to the choice of the partition sequence. We define the quadratic roughness of a path along a partition sequence and show that, for Holder-continuous paths satisfying this roughness condition, the quadratic variation along balanced partitions is invariant with respect to the choice of the partition sequence. Paths of Brownian motion are shown to satisfy this quadratic roughness property almost-surely. Using these results we derive a formulation of Föllmer's pathwise integration along paths with finite quadratic variation which is invariant with respect to the partition sequence.
... The existence of quadratic variation in the sense of Theorem 3.2 is equivalent to the existence of quadratic variation in the sense of Föllmer (see [Vov15, Proposition 3]). Therefore, Theorem 3.2 opens the door to apply Föllmer's pathwise Itô formula [Föl79] for typical non-negative price paths and in particular to define the pathwise integral f ′ (S s ) dS s for f ∈ C 2 or more general for path-dependent functionals as shown by Cont and Fournié [CF10], Imkeller and Prömel [IP15], and Ananova and Cont [AC16]. ...
Article
Full-text available
Using Vovk's outer measure, which corresponds to a minimal superhedging price, the existence of quadratic variation is shown for "typical price paths" in the space of non-negative c\`adl\`ag functions. In particular, this implies the existence of quadratic variation in the sense of F\"ollmer quasi surely under all martingale measures. Based on the robust existence of quadratic variation and a certain topology which is induced by Vovk's outer measure, model-free It\^o integration is developed on the space of continuous paths, of non-negative c\`adl\`ag paths and of c\`adl\`ag paths with mildly restricted jumps.
Article
Full-text available
We approach the problem of integration for rough integrands and integrators, typically representing trajectories of stochastic processes possessing only some Hölder regularity of possibly low order, in the framework of para-control calculus. For this purpose, we first decompose integrand and integrator into Paley–Littlewood packages along the Haar–Schauder system. By careful estimation of the components of products of packages of the integrand and derivatives of the integrator we obtain a characterization of Young’s integral. For the most interesting case of functions with Hölder regularities that sum up to an order below 1 we have to employ the concept of para-control of integrand and integrator with respect to a reference function for which a version of antisymmetric Lévy area is known to exist. This way we obtain an interpretation of the rough path integral. Lévy areas being known for most frequently used stochastic processes such as (fractional) Brownian motion, this integral serves as a basis for pathwise stochastic calculus, as the integral in classical rough path analysis.
Article
Full-text available
We show that Lyons' rough path integral is a natural tool to use in model free financial mathematics by proving that it is possible to make an arbitrarily large profit by investing in those paths which do not have a rough path associated to them. We also show that in certain situations, the rough path integral can be constructed as a limit of Riemann sums, and not just as a limit of compensated Riemann sums which are usually used to define it. This proves that the rough path integral is really an extension of F\"ollmer's pathwise It\^o integral. Moreover, we construct a "model free It\^o integral" in the spirit of Karandikar.
Article
Full-text available
We introduce an approach to study certain singular PDEs which is based on techniques from paradifferential calculus and on ideas from the theory of controlled rough paths. We illustrate its applicability on some model problems like differential equations driven by fractional Brownian motion, a fractional Burgers type SPDE driven by space-time white noise, and a non-linear version of the parabolic Anderson model with a white noise potential.
Thesis
Diese Dissertation behandelt Fragen aus der stochastischen Analysis und der Finanzmathematik, die sich unter dem Begriff der Robustheit zusammenfassen lassen. Zunächst betrachten wir finanzmathematische Modelle mit Arbitragemöglichkeiten. Wir identifizieren die Abwesenheit von Arbitragemöglichkeiten der ersten Art (NA1) als minimale Eigenschaft, die in jedem finanzmathematischen Modell gelten muss, und zeigen, dass (NA1) äquivalent zur Existenz eines dominierenden lokalen Martingalmaßes ist. Als Beispiel für Prozesse, die (NA1) erfüllen, studieren wir stetige lokale Martingale, die darauf bedingt werden nie Null zu treffen. Anschließend verwenden wir eine modellfreie Version der (NA1) Eigenschaft, die es erlaubt, qualitative Eigenschaften von “typischen Preistrajektorien” zu beschreiben. Hier konstruieren wir ein pfadweises Itô-Integral. Dies deutet an, dass sich typische Preispfade als rough-path-Integratoren verwenden lassen. Nun entwickeln wir mittels Fourierentwicklungen einen alternativen Zugang zur rough-path-Theorie. Wir zerlegen das Integral in drei Operatoren mit verschiedenen Eigenschaften. So wird offensichtlich, dass Integratoren mit der Regularität der Brownschen Bewegung mit ihrer Lévy-Fläche versehen werden müssen, um ein pfadweise stetiges Integral zu erhalten. Daraufhin bemerken wir, dass die Integration zweier Funktionen gegeneinander äquivalent dazu ist, eine Funktion mit der Ableitung einer anderen (im Allgemeinen eine Distribution) zu multiplizieren. In höheren Dimensionen ist das Multiplikationsproblem jedoch allgemeiner. Wir verwenden Littlewood-Paley-Theorie, um unseren Fourier-Zugang zur rough-path-Theorie auf Funktionen mehrdimensionaler Variablen zu erweitern. Wir konstruieren einen Operator, der für Funktionen mit dem punktweisen Produkt übereinstimmt und in einer geeigneten Topologie stetig ist. Nun lassen sich stochastische partielle Differentialgleichungen lösen, die bisher aufgrund von Nichtlinearitäten nicht zugänglich waren.
Article
Inspired by the fundamental work of T.J. Lyons (loc.cit), we develop a theory of curvilinear integrals along a new kind of enriched paths in ℝ d. We apply these methods to the fractional Brownian Motion, and prove a support theorem for SDE driven by the Skorohod fBM of Hurst parameter H > 1/4.
Article
A long time ago I started writing a book about Markov chains, Brownian motion, and diffusion. I soon had two hundred pages of manuscript and my publisher was enthusiastic. Some years and several drafts later, I had a thot:sand pages of manuscript, and my publisher was less enthusiastic. So we made it a trilogy: Markov Chains Brownian Motion and Diffusion Approximating Countable Markov Chains familiarly - Me, B & D, and ACM. I wrote the first two books for beginning graduate students with some knowledge of probability; if you can follow Sections 3.4 to 3.9 of Brownian Motion and Diffusion you're in. The first two books are quite independent of one another, and completely independent of the third. This last book is a monograph, which explains one way to think about chains with instantaneous states. The results in it are supposed to be new, except where there are spe­ cific disclaimers; it's written in the framework of Markov Chains. Most of the proofs in the trilogy are new, and I tried hard to make them explicit. The old ones were often elegant, but I seldom saw what made them go. With my own, I can sometimes show you why things work. And, as I will argue in a minute, my demonstrations are easier technically. If I wrote them down well enough, you may come to agree.