An Introduction To Rough Paths
An Introduction To Rough Paths
Antoine Lejay
1 Introduction
This article is an introduction to the theory of rough paths, which has been
developed by T. Lyons and his co-authors since the early ’90s. The main results
presented here are borrowed from [32, 36]. This theory concerns differential
equations controlled by irregular paths and integration of differential forms
against irregular trajectories. Here, x is a continuous function from [0, 1] to Rd ,
and the notion of irregularity we use is that of p-variation, as defined by
N. Wiener. This means that for some p 1,
k−1
sup |xti+1 − xti |p < +∞.
k1, 0t0 ···tk 1 i=0
partition of [0,1]
As we will see, the integer p plays an important role in this theory.
In probability theory, most stochastic processes are not of finite variation,
but are of finite p-variation for some p > 2. We show in Sect. 10 how to apply
this theory to Brownian motion. But the theory of rough paths could be used
for many other types of processes, as presented in Sect. 12.
Firstly, we give a meaning to the integral
t
f (xs ) dxs , or equivalently, f (1.1)
0 x([0,t])
d
∂
f (y) = fi (y) .
i=1
∂xi
This will be done using Picard’s iteration principle, from the result on integra-
tion of one-forms. Using the terminology of controlled differential equations,
x is called a control.
The theory of rough paths also provided some results on the continuity of
the map x → y, where y is given either by (1.1) or (1.3).
The theory of rough paths may be seen as a combination of two families
of results:
(1) Integration of functions of finite q-variation against functions of finite
p-variation with 1/p + 1/q > 1 as defined by L.C. Young in [52].
(2) Representation of the solutions of (1.3) using iterated integrals of x: this
approach is in fact an algebraic one, much more than an analytical one.
Let us give a short review of these notions.
k−1 d d
∂fj1 i,(j ,j )
lim fj (xti )(xjti+1 − xjti ) + (xti )xti ,t2i+11
mesh(Π)→0
i=0 j=1 j ,j =1
∂xj2
1 2
d
∂ p−1 fj1 p,(j ,...,j1 )
+ ···+ (xti )xti ,ti+1p (1.4)
j1 ,...,jp =1
∂xjp
· · · ∂xj2
with formally
i,(j ,...,j1 )
xs,t i = dxjsii · · · dxjs11 . (1.5)
ssi ···s1 t
This expression (1.4) is provided by the Taylor formula on f and the more x
is irregular, i.e., the larger p is, the more regular f needs to be.
What makes the previous definition formal is that the “iterated integrals”
of x have to be defined, and there is no general procedure to construct them,
nor are they unique. The terms xk,(i1 ,...,ik ) for k = 2, . . . , p are limits of
iterated integrals of piecewise smooth approximations of x, but they are sen-
sitive to the way the path x is approximated. Due to this property, the general
principle in the theory of rough paths is:
d t
The integral j=1 s fj (xr ) dxjr is not driven by x but, if it exists, by
x = (x1,(i1 ) , x2,(i1 ,i2 ) , . . . , xp,(i1 ,...,ip ) )i1 ,...,ip =1,...,d corresponding
formally to (1.5).
As first proved by K.T. Chen in [6], Φ([s, t], x) fully characterizes the path x,
and for all s u t,
This relation between iterated integrals is also used to prove that the limit
in (1.4) exists. If exp is the non-commutative exponential (defined by a power
series), then there exists a formal series Ψ ([s, t], x) such that Φ([s, t], x) =
exp(Ψ ([s, t], x)) and
k,(i ,...,ik )
Ψ ([s, t], x) = F(i1 ,...,id ) (X 1 , . . . , X d )xs,t 1
k1 (i1 ,...,ik )∈{ 1,...,d }k
where F(i1 ,...,id ) (X 1 , . . . , X d) belongs to the Lie algebra generated by the in-
determinates X 1 , . . . , X d , i.e., the smallest submodule containing X 1 , . . . , X d
and closed under the Lie brackets [Y, Z] = Y Z − ZY .
If f = (f1 , . . . , fd ) and each of the fi is linear, i.e., fi (y) = Ci y where Ci
is a matrix, then the solution y of (1.3) is equal to
yt = exp Ψ([s, t], x) ys ,
where Ψ ([s, t], x) is equal to Ψ ([s, t], x) in which X i was replaced by the ma-
trix Ci . If f is not linear, but is for example a left-invariant vector field on
a Lie group, then a similar relation holds, where X i is replaced by fi , and
the Lie brackets [·, ·] are replaced by the Lie bracket between vector fields.
Here, the exponential is replaced by the map defining a left-invariant vector
field from a vector in the Lie algebra, i.e., the tangent space at 0 (see for
example [13]).
This result suggests that when one knows x, he can compute its iterated
integrals and then formally solve (1.3) by replacing the indeterminates by f .
In fact, when x is irregular, the solution y of (1.3) will be constructed using
Picard’s iteration
t principle, i.e., as the limit of the sequence y n defined by
n+1
yt = y0 + 0 f (yr ) dxr . But it corresponds, if (xδ )δ>0 is a family of piecewise
n
Thus for k = 1, . . . , p, one can compute xks,t from xis,u and xiu,t when these
quantities are known for i = 1, . . . , k.
The objects x that could be reached as an extension of the truncated
Chen series Φ([s, t], x) and satisfying (1.7) are called geometric multiplicative
functionals.
Our goal is to construct from x new geometric multiplicative functionals z.
For example, the integral f (xr ) dxr will itself be constructed as a geometric
t
multiplicative functional. Remark that for zs,t = s f (xr ) dxr , (1.8) is no more
than the Chasles relation.
The machinery we use to construct z is the following: We construct first
an approximation y of z. For example, if x is of finite p-variation t with p < 2,
we define ys,t1
by f (xs )(xt − xs ), which is an approximation of s f (xr ) dxr .
The object y is a non-commutative polynomial, but y does not satisfy re-
lation (1.7) in general. Thus, if Π = { ti 0 t1 · · · tk 1 } is a partition
of [0, 1], we set
Π
ys,t = ys,ti yti ,ti+1 · · · ytj−1 ,tj ytj ,t .
where i and j are the smallest and the largest integers such that [ti , tj ] ⊂ [s, t].
At the first level of path, this relation reads
d
Π,1 1,(i) 1,(i) 1,(i) 1,(i)
ys,t = X i ys,ti + yti ,ti+1 + · · · + ytj−1 ,tj + ytj ,t .
i=1
Π Π\{ t }
for a nice ε, then one could consider the difference between ys,t and ys,t j
for an element tj in Π ∩(s, t). If tj is well chosen, and the choice of tj is similar
to the one done for the construction of Young’s integral (see above), one could
show that |ys,tΠ
| C(s, t), for a function C(s, t) that does not depend on the
partition Π. One could then pass to the limit as the mesh of the partition Π
decreases to 0. Of course, it has to be proved that under reasonable conditions,
the limit, which we denote by z, is unique and is a geometric multiplicative
functional. Moreover, the work has to be done iteratively at each level of
iterated integrals. Thus, if z1 , . . . , zk are already constructed, one gains the
fact that (z1 , . . . , zk ) satisfies (1.7), and zk+1 is constructed using the previous
machinery.
f, ∂f, . . . , ∂ p f
and focus on x. In the previous section, we have seen how to construct new
geometric multiplicative functionals from x, but we have not said how x is
constructed. We have already said that x may be difficult to construct. The
most natural approach is to choose a piecewise smooth approximation xδ of x
k,(i ,...,ik )
and to define xs,t 1 as the limit of
dxiskk,δ · · · dxis11,δ .
ssk ···s1 t
Motivations
This article does not give a full treatment of the theory of rough paths. But
its aim is to give the reader sufficient information about this theory to help
him to have a general view of it, and maybe to apply it. The reader who is
interested in this theory can read either [32] or [36] to go further.
The theory of rough paths is suitable for trajectories of stochastic pro-
cesses, since there are many types of stochastic processes for which it is possi-
ble to construct their “iterated integrals”. Yet each application to a particular
type of probabilistic problem may require a specific approach. As randomness
8 Antoine Lejay
plays no role in this theory, probability theory takes only a small place in this
article. The reader is refered to Sect. 12 and to the bibliography for applica-
tions to stochastic analysis.
Note.
At first reading, the reader may go directly from the end of Sect. 3 to Sect. 10
for an application to the Brownian motion.
Acknowledgement
δ
−1
Π
zs,t = ys,tδj + ytδ ,t + ytδi ,tδi+1 ,
i=j
where j and are such that Π δ ∩ (s, t) = tδj , . . . , tδ .
10 Antoine Lejay
δ
Π
Proposition 2.1. Under Assumption 2.2, zs,t admits a limit denoted by zs,t
for all 0 s t 1. Furthermore, (s, t) ∈ ∆+ → zs,t is continuous, and
zs,u + zu,t = zs,t (Chasles’ relation) for all 0 s u t 1.
Finally, there exists some constant K depending only on f , p and ω(0, 1)
such that |zs,t |p Kω(s, t) for all (s, t) ∈ ∆+ . This implies that z has finite
p-variation.
t
Thus, one may define s f (xr ) dxr to be zs,t . The proof relies on the fol-
lowing Lemmas.
Lemma 2.1. There exists a constant C depending only on f such that for all
0 s u t 1,
1+α
|ys,t − ys,u − yu,t | Cω(s, t)θ , with θ = > 1. (2.4)
p
Proof. The result is clear if k = 2, since ω(t1 , t2 ) ω(s, t). Assume that k 3.
As ω is super-additive, ki=1 ω(ti−1 , ti+1 ) 2ω(s, t). So, at least one of the
ω(ti−1 , ti+1 )’s is smaller than 2ω(s, t)/k.
by suppressing the point tδk . We use the convention that tδj−1 = s and tδ+1 = t.
According to Lemma 2.2, the point tδk is chosen so that
An Introduction to Rough Paths 11
2
ω(tk−1 , tk+1 ) ω(s, t).
|Π δ ∩ (s, t)|
where K = C + 2θ C n1 1/nθ , and i and i are such that [tδi(δ) , tδi (δ) ] is the
smallest interval containing [s, t]. In particular, tδi(δ) increases to s and tδi (δ)
decreases to t as δ decreases to 0.
Let 0 s u t 1. Set Π δ ∩ (s, u) = tδj , . . . , tδk and Π δ ∩ [u, t) =
δ
tj , . . . , tδk . So,
Πδ Πδ Πδ
zs,u + zu,t = zs,t − ytδk ,tδ − ytδk ,u − yu,tδ . (2.6)
j j
k−1
k−1
|∆zs,t | |∆zti ,ti+1 | 2K ω(ti , ti+1 )θ
i=1 =1
δ
Π
So, the limit of (zs,t )δ>0 is unique.
With (2.5) and the boundedness of f , |zs,t | Kω(s, t)θ +|ys,t |. But |ys,t |
f ∞ |xs,t | f ∞ ω(s, t)1/p . Thus, |zs,t |p (Kω(0, 1)α + f p∞ )ω(s, t) and
z is of finite p-variation. The proposition is then proved.
In the preceding proof, the regularity of x plays in fact no role. The only
condition required is (2.2). Note that (2.2) implies that for any partition
Π = { t0 , . . . , tk } of [s, t],
k−1
k−1
|xti+1 − xti |p ω(ti , ti+1 ) ω(s, t).
i=0 i=0
Remark 2.1. When one considers x(t) = t and p > 1, it is immediate that
for
k−1 any partition 0 t0 · · · tk 1, the following inequality holds:
i=0 |ti+1 − ti | supi=0,...,k−1 |ti+1 − ti |
p p−1
. The later quantity converges
to 0 with the mesh of the partition. But Varp,[s,t] (x) = 1. This means that in
the definition of the p-variation, we have really to consider a supremum on all
the partitions, and not only on those whose mesh converges to 0.
and δp,[s,t] (x) = δp,[s,t] (x, 0). Set also δp (x, y) = δp,[0,1] (x, y) and δp (x) =
δp,[0,1] (x). Note that δp,[0,1] is a not a distance, excepted when restricted to
functions for which x0 is fixed.
Let x be a function such that δp (x) is finite. If there exists a function
ω : ∆+ → R+ satisfying Assumption 2.1 and such that
|xt − xs |p ω(s, t)
for all (s, t) ∈ ∆+ , then x is said to be of finite p-variation controlled by ω. It
is clear that the function ω defined by ω(s, t) = δp,[s,t] (x)p satisfies Assump-
tions 2.1 and that for all (s, t) ∈ ∆+ , |xt − xs |p ω(s, t).
The following lemma is related to sequences of functions of finite p-
variation.
Lemma 2.3. Let (xn )n∈N be a sequence of functions of finite p-variation and
let x be a function of finite p-variation such that δp (xn , x) converges to 0.
Then there exists a subsequence (xnk )k∈N and some function ω satisfying
Assumption 2.1 such that xnk and x are of finite p-variation controlled by ω.
Moreover, for any ε > 0, there exists an integer k for which
∀ k, δp,[s,t] (xn , x) εω(s, t) for all (s, t) ∈ ∆+ .
Proof. There exists a subsequence (nk )k∈N such that δp (x, xnk ) 4−k . Hence,
we set
+∞
ω(s, t) = 2p−1 δp,[s,t] (x)p + 2k δp,[s,t] (xnk , x)p .
k=0
By our choice of the subsequence, this function ω is well defined for all (s, t) ∈
∆+ . As δp,[s,t] (xn , x) δp (xn , x) −−−−→ 0, and δp,[s,t] (xn , x) is continuous near
+∞ n→∞
the diagonal, (s, t) → k=0 2k δp,[s,t] (xnk , x)p is continuous near the diagonal.
Similarly, (s, t) → δp,[s,t] (xn )p is continuous near the diagonal. Clearly, ω is
super-additive and satisfies Assumption 2.1.
Since δp,[s,t] (xnk )p 2p−1 δp,[s,t] (xnk , x)p + 2p−1 δp,[s,t] (x)p , x and all of the
nk
x ’s are controlled by ω. Furthermore,
1
δp,[s,t] (xnk , x)p ω(s, t),
2k
and the lemma is proved.
2.3 Continuity
For any bounded and α-Hölder continuous function f with a Hölder constant
α > p−1, we have constructed a map Kf : x → z, where x is a function on [0, 1]
t
with finite p-variation
(with 1 p < 2) and z is the function s f (xr ) dxr ;
0st1 .
We have seen in Proposition 2.1 that Kf (x) is also of finite p-variation.
We are now interested in the continuity of Kf . Let x and x be two func-
tions of finite p-variation, both satisfying Assumption 2.2 with respect to the
same ω.
Proposition 2.2. Assume that there exists some ε > 0 such that for all 0
s t 1,
|(xt − x
t ) − (xs − x
s )|p εω(s, t),
and that x0 = x 0 . Then there exists a function κ(ε) decreasing to 0 as ε
decreases to 0 and depending only on f and p such that
|Kf (x)s,t − Kf (
x)s,t | κ(ε)ω(s, t)1/p
Proof. The proof is similar to the one of Proposition 2.1. Using the same
Πδ Πδ
notations, define zs,t and zs,t . As previously, we create a new partition Π by
suppressing a carefully chosen point of Π δ . Hence, to estimate
Π δ
Π
zs,t − zs,t
Πδ Π
− zs,t − zs,t ,
and that
|f (xu ) − f (xs ) − f ( xs )| 2Cω(s, t)α/p .
xu ) + f (
Choosing β ∈ (0, 1) such that βα + 1 > p, one gets
Denote by G p (Rd ) the space of continuous functions in C([0, 1]; Rd) of finite
p-variation and starting at the same given point x0 . Denote by V p the topology
that the distance δp defines on the space G p (Rd ).
for every n nε and all (s, t) ∈ ∆+ . From Proposition 2.2, for all (s, t) ∈ ∆+
and any n ∈ N,
The previous convergences are proved at least along a subsequence, but using
the way Lemma 2.3 is proved, the limit of Kf (xn ) is in fact unique. The
Corollary is then proved.
As an application, let Π δ = tδi 0 tδ0 · · · tδkδ 1 be a family of
partitions of [0, 1] whose meshes go to 0 with δ. Then it is easily seen that the
piecewise linear approximation xδ of some path x ∈ G q (Rd ) for some q ∈ [1, 2)
given by
t − tδi
xδt = xtδi + (x δ − xtδi ) when t ∈ [tδi , tδi+1 ]
tδi+1 − tδi ti+1
converges uniformly to x.
Let 0 s0 . . . s 1 be a partition of [0, 1]. Then,
δ
k −1
−1
δ q
|xδsi+1 − xδsi |q = x − xδsi
si+1
i=1 j=0 i s.t. si ∈[tδj ,tδj+1 ]
δ
k −1 δ q
+ xs − xδsi .
i+1
j=0 i s.t. tδj ∈(si ,si+1 )
However, δ q
x − xδsi |xtj+1 − xtj |q
si+1
i s.t. si ∈[tδj ,tδj+1 ]
for any (s, t) ∈ ∆+ . Clearly, x = x. Define also I(y, x) = y. For any integer
n 1, set y n = I(y n−1 , x). Of course, if y n converges to some function y in
G p (Rm ), then y is solution to (2.9).
Step 1: Existence. Assume that two paths x and y of finite p-variation are
controlled respectively by ω and γω on a time interval [S, T ], for some con-
stant γ > 0.
A slight modification of the proof of Proposition 2.1 shows that there exists
some constant K, depending only on f and p, such that
Ascoli’s theorem and Lemma 2.4, there exists some y of finite p-variation
such that a subsequence of (y n )n∈N converges to y in q-variation for some
q > p. But y → I(y, x) is also continuous on G q (Rm ). So, we deduce that y is
solution to (2.9) with the initial condition y0 .
Step 2: Uniqueness. In this step, assume that f is continuous, bounded with
a bounded α-Hölder continuous derivative with α > p − 1.
Let y and y be two paths of finite p-variation controlled by ω and starting
from the same point, that is y0 = y0 . Assume also that x is of finite p-variation
controlled by ω, and that y − y is of finite p-variation controlled by γω for
some γ > 0. It is clear that γ may be chosen smaller than 2p .
We are interested in I(y, x) − I( y , x). With our construction, this dif-
k−1
ference is approximated by i=1 (f (yti ) − f ( yti ))xti ,ti+1 on some partitions
Π = { ti 0 t1 . . . tk 1 } whose meshes go to 0. We follow the proof
of Proposition 2.1 and we set for all s u t,
εs,u,t = |(f (ys ) − f (
ys ))xs,t − f (ys ) − f (ys ) xs,u − f (yu ) − f (yu ) xu,t |
= f (ys ) − f (yu ) − f (ys ) − f (yu ) xu,t
1
∇f ys + τ ( ys − ys ) · (
ys − ys ) dτ
0
1
− ∇f yu + τ ( yu − yu ) · ( yu − yu ) dτ ω(s, t)1/p .
0
From (2.10) and (2.11), one can select a time T small enough depending on
α, p, f and ω such that
γ 1/p
|Is,t (y, x) − Is,t (
y , x)| ω(s, t)1/p .
2
In other words, I(y, x) − I( y , x) is controlled by 2−p γω on [0, T ].
If both y and y are solutions to (2.9), then I(y, x) − I( y , x) = y − y. So,
iterating the procedure, one deduces that y − y is controlled by 2−np ω on
the time interval [0, T ] for each integer n. This proves that y = y on [0, T ].
Similarly, it is possible to construct iteratively a finite sequence of increasing
times Tk for k = 1, . . . , n with T1 = 0, T2 = T and such that Tn = 1 and
y = y on [Tk , Tk+1 ] as soon as yTk = yTk . For that, these times are constructed
so that ω(Tk , Tk+1 ) is smaller than a given constant c small enough, which
explains why this set is finite. We deduce that the solution of (2.9) is unique
on [0, 1].
Step 3: Continuity. Denote by If,y0 the map which at x gives the solution y
to (2.9) with the given initial condition y0 .
For a given y 0 , one may iteratively construct for each integer n 1 a
path y n by setting y n = I(y n−1 , x). In Step 1, we have seen that (y n )n∈N
admits a convergent subsequence, and in Step 2, under stronger hypotheses
on f , that the limit If,y0 (x) of (y n )n∈N is unique. Furthermore, if y 0 , y 1 ,
y 1 − y 0 and x are of finite p-variation controlled by ω, y n − y n−1 are of finite
p-variation controlled by 2−(n−1)p ω. So, If,y0 (x) − y n is of finite p-variation
controlled by 2−(n−2)p ω.
Now, consider two paths x and x both of finite p-variation controlled by
ω, and such that x − x is of finite p-variation controlled by εω for some ε > 0.
Let (y n )n∈N and ( y n )n∈N be two sequences of functions of finite p-variation
controlled by ω with y 0 = y0 and constructed by setting y n = I(y n−1 , x) and
yn = I( ). From Proposition 2.2 it is clear that for each n 0, there
y n−1 , x
exists a function ϕn (ε) converging to 0 with ε such that y n − yn is of finite
p-variation controlled by ϕn (ε)ω. But y − y n and y − yn are both of finite
p-variation controlled by 2−(n−1)p ω.
Thus, for all η > 0, there exists n0 large enough so that both If,y0 (x) − y n
and If,y0 ( x) − yn is controlled by ηω for all n n0 . Besides, if ε is small
enough and x − x is controlled by εω, then y n0 − yn0 is controlled by ηω.
This means that for ε small enough, If,y0 (x) − If,y0 ( x) is controlled by 3p ηω,
if ε is also chosen smaller than η. With Lemma 3.1, this means that If,y0 is
continuous from G p (Rn ) to G p (Rm ).
20 Antoine Lejay
Remark 2.3. The previous proof is slightly different from the original proof
of [32], where f was required to be differentiable with a α-Hölder continuous
derivative to prove the existence of a solution.
for some constant C and some θ > 1. The fact that θ > 1 iscrucial, since the
proof of Proposition 2.1 involves the Zeta function ζ(θ) = n1 1/nθ .
In Lemma 2.1, we used the fact that f is α-Hölder continuous, that α > p
and that |xt − xs |p ω(t, s). With only (3.2a) if p 2, this no longer works
even if f has a bounded derivative, i.e., α = 1.
We are then forced to use a better estimate. If x is smooth, then for
i = 1, . . . , d,
d t
∂f i
f i (xt ) = f i (xs ) + (xr ) dxjr ,
i=1 s ∂xj
and then
t t d t
r1
∂f i
f i (xr1 ) dxir1 = f i (xs ) dxir1 + (xr2 ) dxjr2 dxir1 .
s s j=1 s s ∂xj
∂f j
∇f (y), ei ⊗ ej = (y).
∂xi
t
A first approximation of the integral f (xr ) dxr will be given by
s
t
ys,t = f (xs )(xt − xs ) + ∇f (xs ) dx ⊗ dx. (3.3)
s
d
∂f i
= f i (a) + (a)(bj − aj ) + Ri (a, b) (3.5)
j=1
∂xj
with
d 1
∂f i ∂f i
|R (a, b)| =
i
(a) − a + (b − a)r (b − a ) dr
j j
0 ∂xj ∂xj (3.6)
j=1
1+α
N (f ) b − a
since the derivative of f is α-Hölder continuous (the quantity N (f ) has been
defined in (3.4)). Set R = (R1 , . . . , Rd ).
Using (3.5) and (3.1),
t
|ys,t − ys,u − yu,t | ∇f (xu ) − ∇f (xs ) dx ⊗ dx
u
+ |R(xu , xs )(xt − xu )|.
t
As ∇f is α-Hölder continuous, x satisfies (3.2a) and s dx⊗dx satisfies (3.2b),
2+α
|ys,t − ys,u − yu,t | Cω(s, t)θ with θ = > 1.
p
The Lemma is then proved.
view of integral driven by rough paths. This issue is discussed in Sect. 10.3
for Brownian motion.
In the previous sections, we have given all the elements to construct the in-
tegral of a differential one-form along a path x of finite p-variations with
p ∈ [2, 3), given that one also knows a geometric multiplicative x lying above x.
Once Lemma 3.1 has been proved, then one could use the same machinery
as in the proof of Proposition 2.1, to prove that
δ
k −1
zs,t = lim ytδi ,tδi+1 (3.8)
δ→0
i=0
exists and is unique for all partition tδ0 , . . . , tδkδ of [s, t] when y is given by
(3.3).
Proposition 3.1 below summarizes this result. However, we will give in the
next section a more complete construction of the integral of a one-form along
the path x. In this new definition, the integral belongs to the set of geometric
multiplicative functionals G p (W). This means that this integral could also be
used as a path along which a another differential one-form is integrated.
t
3.5 The iterated integrals of s f (xr ) dxr
d
Let us consider the differential form f (x) = i=1 fi (x)dxi , where the f i ’s are
t
functions from V = Rd into W = Rm . The integral z1s,t = s f (xr ) dxr takes
An Introduction to Rough Paths 25
one may first define the integration of one-forms, and then use Picard’s it-
eration principle. However, integrating with respect to a control of finite p-
variation with p ∈ [2, 3) requires an element in G p (V). So, to use a fixed point
t
theorem, we need to construct s f (xr ) dxr not only as an element of W, but
also as an element of G p (W).
To this end, set
1
ys,t = f (xs )x1s,t + ∇f (xs )x2s,t ∈ W, (3.9a)
2
ys,t = f (xs ) ⊗ f (xs ) · x2s,t ∈ W ⊗ W, (3.9b)
1 2
and ys,t = (ys,t , ys,t ). In the definition of y2 , we used a shorthand, which
means in fact that
d
2
ys,t = f i (xs ) ⊗ f j (xs )x2,i,j
s,t .
i,j=1
t
Denote by z1s,t the element of W given by s f (xr ) dxr .
Let 1 denote an element of a one-dimensional space. We use the following
computation rule: If y belongs to W⊗k for some integer k = 1, 2, then 1 ⊗ y =
y ⊗ 1 = y ∈ W⊗k . If y and z belong to W, then y ⊗ z belongs to W⊗2 .
If y belongs to W and z belongs to W⊗2 , then y ⊗ z = z ⊗ y = 0. Set
T2 (W) = 1 ⊕ W ⊕ W⊗2 . By the definition of the tensor product, if x, y and z
belong to T2 (W), then for all α, β ∈ R, (αx + βy) ⊗ z = αx ⊗ z + βy ⊗ z and
z ⊗ (αx + βy) = αz ⊗ x + βx ⊗ y.
Let Π = { ti t0 · · · t } be a partition of [s, t]. Set
The computation rules previously given mean that we keep only the elements
in T2 (W), and not those in W⊗k for k > 2. From Proposition 3.1, the projec-
tion zΠ,1 of zΠ ∈ T2 (W) on W is equal to z1s,t .
The proof that zΠ has a limit when the mesh of the partition Π decreases
to 0 is similar to the proof of Proposition 2.1. But we have also to estimate
2 2 2
the “error” when ys,t is split into ys,u and yu,t .
Lemma 3.2. For all 0 s u t 1, set
2
ε(s, u, t) = ys,t − ys,u
2
− yu,t
2
− ys,u
1
⊗ yu,t
1
.
There exists some constant C depending only on N (f ), ω(0, 1) and α such that
|ε(s, u, t)| Cω(s, t)θ
with θ = (2 + α)/p > 1.
26 Antoine Lejay
1
ys,u ⊗ yu,t
1
= f (xs )x1s,u ⊗ f (xu )x1u,t
= f (xs ) ⊗ f (xs ) · x1s,u ⊗ x1u,t + f (xs ) ⊗ (f (xu ) − f (xs )) · x1s,u ⊗ x1u,t ,
we obtain that
ε(s, u, t) = f (xs ) ⊗ f (xs ) − f (xu ) ⊗ f (xu ) · x2s,t
− f (xs ) ⊗ f (xu ) − f (xs ) · x1s,u ⊗ x1u,t .
But
Moreover, |x1s,u ⊗ x1u,t | |x1s,u | · |x1u,t | ω(s, t)2/p . The Lemma is now easily
proved by combining all the previous estimates.
Proof. Assume that Π ∩(s, t) has more than one element. Let tk be an element
of Π ∩ (s, t) such that ω(tk−1 , tk+1 ) 2ω(s, t)/|Π ∩ (s, t)| (see Lemma 2.2).
We use the convention that tk−1 = s if Π ∩ (s, tk ) = ∅, and that tk+1 = t is
Π ∩ (tk , t) = ∅. Using the computations’ rules on 1 ⊕ W ⊕ W⊗2 provided in
Sect. 3.5, one has
Set
|z1r,u − yr,u
1
| Kω(r, u)(2+α)/p
where the constants θ > 1 and C depend only on f , α and p. The end of
the proof is similar to the one of Proposition 2.1.
For the uniqueness of the limit, remark that if z and z are two multiplica-
tive functionals of finite p-variation such that z1s,t =
z1s,t for all (s, t) ∈ ∆+ ,
then
ψ(s, t) = z2s,t −
z2s,t
is additive, i.e., ψ(s, u) + ψ(u, t) = ψ(s, t) for all 0 s u t 1. Let z
δ
and z be two cluster points of (zΠ )δ>0 for a family (Π δ )δ>0 of partitions of
[0, 1]. By construction, z1 = z1 . Moreover, for all integer n 1,
2
n−1
zs,t −
z2s,t z2tn ,tn z2tni ,tni+1 2Kω(s, t)
− sup ω(tni , tni+1 )θ−1 ,
i i+1
i=1 i=1,...,n−1
where tni = s + i(t − s)/n. Since ω is continuous near its diagonal, letting n
increase to infinity proves that z2 =
z2 , and the limit is unique.
Corollary 3.1. The map
t
x ∈ G p (V) −→ f (x0 + x10,r ) dxr ; (s, t) ∈ ∆+ ∈ G p (W)
s
In this section and the next one, we consider iterated integrals of piecewise
smooth paths. We present some results, mainly due to K.-T. Chen (see [6]
and related articles), which allows to perform some manipulations on smooth
paths which could be expressed using algebraic computations. These results
provides us with a very powerful tool. The first main result expresses that a
piecewise smooth path x can be uniquely defined by a power series involving
its iterated integrals.
Theorem 4.1 (K.-T. Chen). If Φ([0, t], x) = Φ([0, t], y) for two paths x
and y, then x = y on [0, t] up to a translation.
Another interesting property arises when one considers the product of two
iterated integrals:
i i
Ψs,t = dxiu11 . . . dxiujj dxu11 . . . dxujj .
s<u1 <···<uj <t s<u1 <···<uj <t
We present now some aspects of Lie algebras and enveloping algebras. The
relation with Φ([0, t], x) is developed in Sect. 5.3. On this topic, see for example
[40, Chap. 1].
Let A = { a1 , . . . , an } be some letters. In Sect. 5.3, these letters will be
identified with the indeterminates X i . The letters may be used to construct
some words ai1 · · · aik for some multi-index I = (i1 , . . . , ik ) of length k. The
set of words with letters in A for which an empty word 1 is added is denoted
by A∗ . Let K be a ring containing Q. A non-commutative polynomial P is a
linear combination over K of words on A :
30 Antoine Lejay
P = Pw w, Pw ∈ K with only a finite number of terms.
w∈A∗
An element of L
K (A) is called a Lie polynomial. A formal series S that may
be written S = n0 Sn where the Sn ’s are Lie polynomials is called a Lie
series.
For a formal series of the form S = 1 + T where T have a zero constant term,
the logarithm is defined to be
(−1)n−1
log(S) = log(1 + T ) = T n.
n
n1
Thus, the series Φ([0, t], x), which belongs to KA, may be seen as an
element of R ⊕ V ⊕ V⊗2 ⊕ · · · . More precisely, for all integer k, the element
t ⊗k
I=(i1 ,...,ik ) ei1 ⊗ · · · ⊗ eik s dI x belongs to V . In Sect. 4.2, we have seen
that the concatenation of x : [0, s] → V and x : [s, t] → V creates a new paths
x : [0, t] → V characterized by the series Φ([0, t], x) given by the product of
Φ([0, s], x) and Φ([s, t], x). With our convention, (4.1) is rewritten Φ([0, t], x) =
Φ([0, s], x) ⊗ Φ([s, t], x).
32 Antoine Lejay
Remark 5.2 (Another notation for the iterated integrals). If e1 , . . . , ed is the
canonical basis, identified
t with e1 , . . . , ed , of the dual V∗ of V, then define a
multi-linear form s dx ⊗ · · · ⊗ dx on V∗ × · · · × V∗ by
t
dx ⊗ · · · ⊗ dx, (ei1 , . . . , eik ) = dxis11 · · · dxiskk .
s s<s1 <···<sk <t
t
Thus, s dx ⊗ · · · ⊗ dx is an element of the dual of V∗ × · · · × V∗ , which is
t
identified with V⊗k , and s dx ⊗ · · · ⊗ dx is identified with
t
ei1 ⊗ · · · ⊗ eik dI x.
s
I=(i1 ,...,ik )∈{ 1,...,d }k
The following Proposition may be proved using the properties of the shuffle
product and the Campbell–Hausdorff formula, and links the series constructed
in Sect. 4 with our constructions of objects related to Lie algebras.
Proposition 5.2 ([5]). For any irreducible, piecewise smooth path x : [0, t] →
Rd , Φ([0, t], x) is a group-like element when the elements ei of the basis of Rd
are identified with the indeterminates X i . Moreover,
t
log Φ([0, t], x) = ΘI dI x, (5.1)
I multi-index s
where ΘI belongs to LK (A) = 0 ⊕ V ⊕ [V, V] ⊕ [V, [V, V]] ⊕ · · · and does not
involve more that length(I) Lie brackets.
Remark 5.3. There are very nice algebraic properties that can be considered
on the series of type Φ([0, t], x). In particular, two structures of bi-algebra may
be considered, one corresponding to concatenation of paths, the other one to
product of the series and then using shuffle products (see for example [40,
Sect. 1]). See also [47] for an example of use of the Poincaré-Birkhoff-Witt
Theorem that gives a basis of Lie algebras.
lying above a path x corresponds to x together with its first “iterated inte-
grals”, and such that the iterated integrals xδ, = dxδ ⊗· · ·⊗dxδ of piecewise
smooth approximations xδ of x converge to x for = 2, . . . , k.
An Introduction to Rough Paths 33
If x is a piecewise smooth path, we have seen that one can construct its
Chen series Φ([0, t], x), which fully characterize x. Moreover, given a Chen
series of a path x, one could formally solve a differential equation controlled
by x or consider the integral of a one-form along the path x, by writing these
objects with the Chen series of x (For example, see Sect. 6.3 below).
If x is irregular, then knowing x is not sufficient to define its iterated
integrals (see Remark 3.1 and Sect. 6.2 for example). However, when one
knows a (geometric) multiplicative functional x = (1, x1 , x2 , . . . , xk ) lying
above a path x of finite p-variation with k = p, then it will be proved that
there exists a procedure to extend x in a (geometric) multiplicative functional
(1, x1 , x2 , . . . ), and that this extension has some nice properties, especially
with respect to the topology of generated by the norm of p-variation. Thus,
one can construct an extension of the notion of Chen series for irregular paths,
provided enough information is known on the path, i.e., its first “iterated in-
tegrals”. And, in view of the results of Sects. 2 and 3, one will not be surprised
by the results of Sects. 8, where integrals of one-form along irregular paths are
constructed, and 9, where differential equations controlled by irregular paths
are solved.
For any integer k 1, set
which is a truncated tensor algebra. Let also Ak (V) ⊂ Tk (V) containing all
the elements of A(V) = 0 ⊕ V ⊕ [V, V] ⊕ [V, [V, V]] ⊕ . . . , where all the terms
involving more than k Lie brackets are set to 0. Similarly, computations on
Tk (V) are done by setting to 0 all tensor products involving more than k
terms.
The norm we choose on V⊗k is such that
This norm is also denoted by | · |, and there are different possibilities for
constructing such a norm (see for example [41]).
and x0 = 1.
34 Antoine Lejay
Proposition 6.1 ([32, Lemma 2.3.1, p. 259]). The space G p (V) is the closure
of S p (V) (i.e., geometric multiplicative functionals lying above piecewise
smooth trajectories) with respect to ·p .
Remark 6.2. In particular, (6.4a) and (6.4b) are true if (xn0 )n∈N is bounded
and there exists a function ω satisfying Assumption 2.1 such that xn q,[s,t]
ω(s, t) for all (s, t) ∈ ∆+ and any n ∈ N.
Remark 6.3. Thanks to the Ascoli theorem and the relation (6.1b), the con-
dition (6.4a) is also equivalent to saying that the sequences of functions
(t → xn,i
0,t )n∈N are relatively compact in the space of continuous functions
for i = 1, . . . , k. This is a fact we use in the proof of this Lemma.
Hence,
36 Antoine Lejay
|y0,t
i
− y0,s
i
| |y0,s
k
| · |ys,t
|
k+=i,1
i
C y∞ |ys,t
| C 2 y∞ yq,[s,t] .
=1
for some real number q 1. Then (X)n is tight in (Tk (V), ·p ) (hence in
Mp (V) if k = p) for all p > q.
Remark 6.4. Since (Tk (V), ·p ) is not separable, a sequence (Xn )n∈N may be
tight in this space but fails to satisfy (6.6).
Remark 6.5. Owing to (6.1b), the tightness of (t → Xn,i 0,t )n∈N for all i ∈
{ 1, . . . , k } is equivalent to saying that for all ε > 0 and any C > 0, there
exists some η > 0 small enough such that
i,n
sup sup P sup |Xs,t | > C ε.
n∈N i=1,...,k |t−s|<η
Proof. The proof is immediate from Lemma 6.2 and Remark 6.5, since the
subsets K of Tk (V) of the form K0 ∩ K1 for a given C > 0 are relatively
compact in (Tk (V), ·p ), where the sets of functions (t → xi0,t )x∈K0 are equi-
continuous for i = 1, . . . , k and K1 contains the multiplicative functionals such
that supx∈K1 xq < C for some q < p and a given constant C.
An Introduction to Rough Paths 37
S(x) and A(x) are respectively the symmetric part and the antisymmetric
part of x2 . Moreover, S(x) depends only on x1 , and if x1 lies above a path x,
then S(x) = S(x). Moreover, if x is of finite p-variation, the map x → S(x)
is continuous for the topology generated by Varp/2,[s,t] (·) + ·∞ .
On the other side, if x is a natural geometric multiplicative functional in
S 2 (V) lying above a smooth path x, then Ai,js,t (x) is the area contained between
the curve (xir , xjr )r∈[s,t] and the chord (xis , xjs )(xit , xjt ). Denote also this area
by Ai,j
s,t (x).
Lemma 6.3. For all 0 s t 1, the map x ∈ S 2 (Rd ) → As,t (x) is not
continuous with respect to the uniform norm (except if d = 1, in which case
As,t (x) = 0).
Proof. Assume that d = 2 and identify R2 with the complex plane C. Set
2
xnt = n−1 ein t . Then, A0,2π (xn ) = π, but xn converges uniformly to 0.
Lemma 6.4. If x is in G p (V) for p ∈ [2, 3), and ϕ = (ϕi,j )i,j=1,...,d is a func-
tion from [0, 1] to d × d-antisymmetric matrices, and of finite p/2-variation.
Then x defined by
is also in G p (V).
Then x
d
d
xs,t = 1 + ei x1,i
s,t + ei ⊗ ej x2,i,j
s,t .
i=1 i,j=1
The quantity log(xs,t ), which belongs to A2 (V) (see (6.1c)), may be explicitly
computed:
38 Antoine Lejay
d
1
d
log(xs,t ) = ei x1,i
s,t + [ei , ej ]Ai,j
s,t (x),
i=1
2 i,j=1
1
d
xs,t ) = log(xs,t ) +
log( [ei , ej ](ϕi,j (t) − ϕi,j (s)).
2 i,j=1
belongs to G p (V).
xs,t ) belongs to A2 (V) for all (s, t) ∈ ∆+ . Then, x
Thus, log(
In Sect. 10.2 below, these results will be used to compare the theory of integra-
tion given by this theory and the Stratonovich integral for Brownian motion
(see especially (10.2) and (10.3)).
The fact that ϕ was taken additive (i.e., ϕ(s, u) + ϕ(u, s) = ϕ(s, t) for
all 0 s u t 1) in Lemma 6.4 is justified by the following Lemma.
Lemma 6.5 ([32, Lemma 2.2.3, p. 250]). If x and x are two multiplicative
functionals in Mp (V) which agree for all order smaller than k = p (i.e.,
is,t for i = 0, 1, . . . , k − 1), then ϕ(s, t) = xks,t − x
xis,t = x ks,t is additive, and
is of finite p/k-variation.
d
dyt = Ci yt dxit . (6.7)
i=1
As y appears in the right-hand side of (6.7), one may replace it by its value
given by (6.7). Thus,
d t d
yt = ys + Ci ys dxis + Ci Cj ys dxir1 dxjr2 .
i=1 s i,j=1 s<r1 <r2 <t
by setting
yt = Id + CI xlength
s,t
I,I
ys . (6.10)
I multi-index
Inequality (6.2) implies that the series that appears in this expression is con-
vergent, in the sense of the norm of operators.
40 Antoine Lejay
ω(s, t)i/p
|xis,t | (6.11)
βΓ (i/p)
Proof (Sketch of the proof ). The idea is to construct y() ∈ T (V) recursively,
by setting y(k) = x ∈ Tk (V) and, once y
()
has been defined, set z(+1)
in T+1 (V) by z(+1)
= y + (i1 ,...,i+1 ) 0 · ei1 ⊗ e+1 . Thus, y(+1) is defined
()
by
(+1) (+1) (+1) (+1)
ys,t = lim ztδ ,tδ ⊗ ztδ ,tδ ⊗ · · · ⊗ ztδ ,tδ ,
δ→0 0 1 1 2 iδ −1 iδ
where Π =δ
s ···
tδi tδ0 t is a partition of [s, t] whose mesh
tδiδ
goes to 0 as δ decreases to 0. Thanks to the multiplicative property of y() ,
remark that y(+1),i = y(),i for i = 1, . . . , . In fact, this idea was already
used in the proof of Proposition 3.2, and will be used later in the proof of
Theorem 7.1. In T (V), the extension of x is then y() .
Theorem 6.2 ([32, Theorem 2.2.2]). Let x and y be two multiplicative func-
tionals in Tk (V) satisfying the hypotheses of Theorem 6.1 for the same ω.
Assume that there exists a constant β large enough (depending only on p)
and a constant 0 < ε < 1 such that
An Introduction to Rough Paths 41
ω(s, t)i/p
|xis,t − ys,t
i
|ε for i = 1, . . . , k. (6.12)
βΓ (i/p)
In view of Lemma 3.1, this means that the map giving the extension
of a multiplicative functional in Tp (V) is continuous with respect to the
norm ·p .
Combined with Proposition 6.1, this means that the extension of a geo-
metric multiplicative functional in G p (V) to T (V) is also a geometric multi-
plicative functional in this space for all > p.
Using Theorem 6.1, if x is a geometric multiplicative functional in Tp (V)
of finite p-variation (i.e., xp is finite: there is no difficulty to find a function ω
such that x is controlled by ω), then one may solve (6.9) by density using (6.10)
and the previous Theorem.
The idea behind Theorem 6.1 is that the more irregular is a trajectory
(“rough”), the more “iterated integrals” have to be considered. But on the
other side, once one knows enough iterated integrals, then the whole set of
iterated integrals could be known. Hence, when one deals with a path x of finite
p-variation, then he needs to know a geometric multiplicative functional x
lying above x and belonging to the truncated tensor algebra Tp (V).
Of course, (ys,t )(s,t)∈∆+ fails to satisfy (7.1), but the error εs,u,t = ys,t − ys,u −
yu,t was easily controlled. And the estimate on εs,u,t was the key of the proof
of Proposition 2.1. So, (ys,t )(s,t)∈∆+ may be called an almost multiplicative
functional.
42 Antoine Lejay
Proof (Sketch of the proof ). The idea was already used in the proofs of Propo-
sition 3.2 and Theorem 6.1: Construct z by setting z0 = z(0) = 1 and recur-
() (−1)
sively for = 1, . . . , k, ys,t = zs,t + xs,t , where z(−1) = (z0 , z1 , . . . , z−1 ) ∈
T−1 (V). Hence, define z ∈ T (V) by
()
where Π δ = tδi s tδ0 · · · tδiδ t is a partition of [s, t] whose mesh
goes to 0 as δ decreases to 0. The multiplicative functional z is then z(k) .
The notion of Lipschitz functions we use is that of E.M. Stein (see for example
the book [44]).
An Introduction to Rough Paths 43
Definition 8.1. Let F be a closed subset of the normed space U, and α > 0.
Let W be a separable Banach space, and f a function with values in W.
Assume that k < α k + 1. Then f belongs to Lip(α, F, W) if there exist
some functions f J where J is a multi-index of length length(J) k and some
functions RJ : F × F → W such that for all x, y ∈ F,
1
f J (x) − f J (y) = f (J,L) (y)(x1 − y 1 ) · · · (xm − y m )
1 ! · · · m !
L=(1 ,...,m ),
length(J,L)k
+ RJ (x, y),
where (J, L) denotes the concatenation of the multi-indexes J and L. By
definition, f ∅ = f . The functions f J shall satisfy
|f J (x)| M and |RJ (x, y)| M |x − y|α−length(J) (8.1)
for all x, y ∈ F and any J with length(J) k. Denote by f Lip the small-
est M such that (8.1) is true. With this norm, Lip(α, F, W) is a Banach
space.
This definition requires some comments. If F = V = Rd , then the functions
in Lip(α, Rd , W) are from Rd to W with bounded derivatives up to order
α. Moreover, f (i1 ,...,i ) = ∂xi ∂···∂x
f
i
, and f (i1 ,...,i α) is (α − α)-Hölder
1
continuous.
If F is a strict subset of Rd , then a function f ∈ Lip(α, F, W) may be
extended continuously to a function in Lip(α, Rd , W), but the family of the
f J ’s is not necessarily unique. In this case, by a function f in Lip(α, F, W),
we denote not only f , but the whole family (f J )J multi-index, length(J)α .
8.2 Integration
To start with, let f = (f1 , . . . , fd ) be a smooth function defined on the Banach
t
space V = Rd . The idea to define 0 f (xs ) dxs is to construct an almost multi-
t
plicative function y such that ys,t gives a first approximation of s f (xr ) dxr ,
and then to transform y to a geometric multiplicative functional using Theo-
rem 7.1.
The value of ys,t will be computed as previously using a Taylor expansion
of f when it is assumed that x is smooth. So, in a first approach, assume that x
is piecewise smooth and that x is the geometric multiplicative functional given
by its iterated integrals.
Set
k t
yt − ys = ys,t =
1
DI (f )(xs ) dI x
j=1 I multi-index, I=(i1 ,...,ij ) s
∂ j−1 fi1
with DI (f )(xs ) = (xs ).
∂xij · · · ∂xi2
44 Antoine Lejay
j
Once this is done, one may define y forjt = 2, . . . , k using the iterated
j
integrals of y: ys,t = I multi-index, I=(i1 ,...,ij ) s dI y. But this involves expres-
sions such as
s1 s
S(J1 , . . . , J ) = d dJ1 x · · · d dJ x ,
ss1 ···s t 0 0
(x)eij ⊗ · · · ⊗ ei1 .
(i ,...,i2 )
D(i1 ,...,ij ) (f )(x) = fi1 j
Theorem 8.1 ([32, Theorem 3.2.1, p. 285]). Definition 8.2 is valid, i.e., the
definition of y gives rise to an almost multiplicative functional, which is con-
trolled by Kω if x is controlled by ω, where the constant K depends only
on α, p, f Lip and sup(s,t)∈∆+ ω(s, t). Moreover, x → f (x0 + x10,s ) dxs is
continuous from G p (V) to G p (W).
Remark 9.1. To prove the existence of a solution under the assumption that f
belongs to Lip(α, W, W) with α > p − 1, one only has to act as in Step 2 in
the proof of Theorem 2.1: In Step 3 of the proof of Theorem 4.1.1 in [32], it is
proved that the paths y n given by the Picard iteration principle are of finite
p-variations controlled by the same ω. Hence Lemma 6.2 can be used.
But we know that: (i) The limit of B2,i,j,δ depends on the choice of the
approximation. (ii) When it converges, B2,i,j,δ does not converge almost surely
but only in probability or in L2 (P) (however, it is proved that for dyadic
partitions, the convergence may be almost sure. See [24] for example).
Point (ii) is contained in the classical result from E. Wong and M. Zakai
in [51] for some piecewise linear approximation of the Brownian motion, while
point (i) is related to the extensions of such a result (see [19, Sect. VI,-7,
p. 392] or [22, Chap. 5.7, p. 274] for example). In fact, the problems with
(i) are similar to the results given in Sect. 6.2: There are different geometric
multiplicative functionals lying above the same path B.
Btδ (ω) = Bti (ω) + (ti+1 − ti )−1 (t − ti )(Bti+1 (ω) − Bti (ω)) (10.1)
the convergence holding with respect to both the uniform norm and the norm
of p-variation. But from a theoremof Wong-Zakai type, it is also known that
·
X δ converges in probability to x0 + 0 f (Bs )◦dBs . Thus, this integral is almost
surely equal to X. There is in fact a deep relation between the Stratonovich
integral and the one given by the theory of rough paths.
We develop in this section the link between stochastic integrals given by the
theory of rough paths and Stratonovich integrals for Brownian motion. It also
explains the influence of the term B2 , where B is a geometric multiplicative
functional lying above the Brownian motion, but different from the one given
by the “natural” construction of Sect. 10.1. Note that such geometric multi-
plicative functionals may arise naturally. For example (see e.g., [26, 28] in the
homogenization theory), there exist some families (X ε )ε>0 of semi-martingales
converging, thanks to a central limit theorem, to a Brownian motion B, but
such that As,t (X ε ) converges to As,t (B) + c(t − s) for some matrix c.
t
By definition, the Stratonovich integral 0 fj (Bs )◦dBsj is the limit in prob-
ability of
k−1
(fj (Bti+1 ) + fj (Bti )) j
δ def
I = (Bti+1 − Btji )
i=1
2
k−1
k−1
(fj (Bti+1 ) − fj (Bti )) j
= fj (Bti )(Btji+1 − Btji ) + (Bti+1 − Btji ),
i=1 i=1
2
Rj (x, y) |y − x| . Thus,
α
k−1
Iδ = fj (Bti )(Btji+1 − Btji )
i=1
1 ∂fj
k−1 d
+ (Bti )(Bti+1 − Bti )(Btji+1 − Btji ) + εδ ,
2 i=1 ∂x
=1
with
48 Antoine Lejay
k−1
εδ = Rj (Bti , Bti+1 )(Btji+1 − Btji )
i=1
k−1
C |ti+1 − ti |(α+1)/p Ct sup |ti+1 − ti |(α+1−p)/p −−−→ 0.
i=1,...,k−1 δ→0
i=1
This constant C is such that |Bt − Bs | C 1/(1+α) |t − s|1/p for the considered
trajectory of the Brownian motion.
Now, let B be a geometric multiplicative functional lying above B. There
is no necessity to choose the previous one, and we have seen in Sect. 6.2 how
to construct as many areas as we want. Then
1
(B − Bti )(Btji+1 − Btji ) = B2,,j ,j
ti ,ti+1 − Ati ,ti+1 (B),
2 ti+1
where A(B) is the antisymmetric part of B2 . Moreover, we have seen that
d k−1 j j 1 ∂fj 2,,j
j=1 i=1 fj (Bti )(Bti+1 − Bti ) + 2 ∂x (Bti )Bti ,ti+1 converges almost surely
to Xt − X0 , where X is the path above which f (Bs ) dBs ∈ G p (W) lies. So,
we deduce that
d t
X0 + fj (Bs )◦dBsj = Xt − Qt (B)
j=1 0
d k
∂fj δ ,j
with Qt (B) = lim (Bti )Ati ,ti+1 (B).
δ→0
i=1
∂x
j,=1
d k−1
1 ∂fj ∂f
Qt (B) = lim − (Btδi )A,j
ti ,ti+1 (B).
δ→0 2 ∂x ∂xj
i=1 j,=1
∂fj
Thus, if ∂x − ∂f
∂xj = 0, Qt (B) = 0, then X depends only on B and not
2 ∂F
on the choice of B . In particular, this is true if fi = ∂x i
for some function
δ
F . In such a case, this could be shown directly, if (B )δ>0 is a family of
geometric multiplicative function lying above an approximation B δ of B and
converging to G p (V) to B, then the change of variables’ formula reads: F (Btδ )−
t
F (B0δ ) = 0 fi (Bsδ ) dBsδ . Thus, F (Btδ ) − F (B0δ ) converges to F (Bt ) − F (B0 ),
t
while 0 fi (Bsδ ) dBsδ converges to Xt .
Now, if Bnat is the “natural” rough path lying above B (see Sect. 10.1),
then A,j
s,t (B
nat
) is the Lévy area A,js,t (B) of the 2-dimensional Brownian mo-
tion (Br , Brj )r∈[s,t] , i.e., the area enclosed between the curve of (B , B j ) and
its chord:
An Introduction to Rough Paths 49
t t
1
A,j
s,t (B) = (Br − Bs )◦dBrj − (Brj − Bsj )◦dBr .
2 s s
The result given at the end of Sect. 10.1 implies that Qt (Bnat ) = 0 almost
surely.
Moreover, one knows from Sect. 6.2 that there exists a function ϕ =
(ϕi,j )i,j=1,...,d from [0, 1] to the space of antisymmetric matrices (i.e., ϕi,j (t) =
−ϕj,i (t) for all t ∈ [0, 1]) and of finite p/2-variation such that As,t (B) =
As,t (B) + ϕ(t) − ϕ(s). We deduce that
d
t
∂fj
Qt (B) = (Bs ) dϕ,j (s).
0 ∂x
j,=1
To summarize,
if B δ is a piecewise smooth approximation of B such that
(1, B , dB ⊗ dB δ ) converges to the geometric multiplicative functional B
δ δ
in G p (V), and As,t (B) = As,t (B) + ϕ(t) − ϕ(s), then we obtain directly that
t t d
t
probability ∂fj
f (Bsδ ) dBsδ −−−−−−−→ f (Bs )◦dBs + (Bs ) dϕ,j (s)
0 δ→0 0 0 ∂x
j,=1
d
t
1 t ∂fj ∂f
= f (Bs )◦dBs + − (Bs ) dϕ,j (s). (10.2)
0 2 0 ∂x ∂xj
j,=1
Here, the drift term is different from the one in (10.2), since it comes from
the cross iterated integrals of the type dY ⊗ dB, which may be computed
first for smooth paths, and then by passing to the limit.
Thus, the theory of rough paths provides us with some new light on the
results presented in Sect. VI-7 in [19, p. 392] (see also Historical Note 5 below).
In this book, the results concern the case where ϕ,j (t) = c,j where c =
(c,j ),j=1,...,d is an antisymmetric matrix, whose terms are given by
δ δ
1 1 ,j
δ
c,j = lim E Bs dBs −
,δ j,δ j,δ ,δ
Bs dBs = lim E A0,δ B ,
δ→0 2δ 0 0 δ→0 2δ
50 Antoine Lejay
For any integer n and any k = 0, . . . , 2n , set tnk = i/2n, that is (tnk )k=0,...,2n is
the dyadic partition of [0, 1] at level n. Let (s, t) belongs to ∆+ , and construct
recursively a sequence (sm )m∈Z, m=0 of elements in (tnk )n∈N, k=0,...,2n by the
following way: Let n0 be the smallest integer such that [tnk 0 , tnk+10
] ⊂ [s, t] for
some integer k. set s−1 = tnk 0 and s1 = tnk+1 0
. Hence, construct sm for m 1
by setting, if sm < t,
n
nm = inf n nm−1 ∃k ∈ N, tnk m = tk m−1 , tm k+1 t .
n
j
xjs,t = xrs1m1 ,sm1 +1 ⊗ · · · ⊗ xrsim ,smi +1 .
i
i=1 r1 ,...,ri =1,...,j m1 <···<mi
r1 +···+ri =j m1 ,...,mi ∈Z∗
m∈Z∗ m∈Z∗
r/p
j
C m |xsm ,sm+1 |
nβp/r r p/r
,
r=1 m∈Z∗
52 Antoine Lejay
(p−r)/p
−βp/(p−r)
where C(r) = m∈Z ∗ n m and C = supr=1,...,k C(r). Our
choice of β ensures that C is finite. Then, there exists a constant K depending
only on k and C such that
j/p
j
|xjs,t | K m |xsm ,sm+1 |
nβp/r r p/r
r=1 m∈Z∗
where n(s, t) is the smallest integer n such that there exists some integer k
for which [tnk , tnk+1 ] ⊂ [s, t].
Inequality (11.1) is useful, since it allows to estimate both
which satisfies
j/p
j n
2 −1
Var (xj ) K nβp/r |xrtni ,tni+1 |p/r (11.2)
p/j,[0,1]
r=1 n0 i=0
provided one knows xtni ,tni+1 for all dyadic point tni = i/2n .
estimating Bδs,t , where s, t are dyadic points (s, t) = (k/2m , (k + 1)/2m ) for
all m 1 and any k ∈ { 0, . . . , 2n }, it could be shown that if Π δ is the dyadic
partition, then Bδ converges almost surely to B (see [24]).
Let q be a real number in (2, p). If m n, then according to the Doob
inequality, there exists a constant C depending only on q such that for a =
1, . . . , d,
q ! !
E Bta,δ m − Bta,δ
m
E Btam − Btam q C(tm
k+1 k k+1 − tk )
m q/2
C2−mq/2 .
k+1 k
If m > n, then again by the Doob inequality, there exists a constant C de-
pending only on q such that for a = 1, . . . , d,
q
q ! tm − tm !
E Bta,δ
m − Bta,δ
m
k+1
k
E Bta,δ a,δ q
n − Btn
k+1 k ti+1 − ti
n n i i
Let Aa,b δ
s,t (B (ω)) be the area enclosed between the curve defined by
(Br (ω), Br (ω))srt and its chord for a, b ∈ { 1, . . . , d }.
a,δ b,δ
for a = 1, . . . , d, since Btδni = Btni for all i ∈ { 0, . . . , 2n }. The Doob and the
convexity inequalities imply that there exists a constant C depending only on
q such that for a = 1, . . . , d,
a,δ
E sup Br − Btm a,δ q
C2q−1 E |B a tni+1 − B a tni |q/2
k
r∈[tm m
k ,tk+1 ]
+ E |B a tni − B a tm |q/2
k (11.3)
C2q−1 (2−nq/2 + 2−mq/2 )
C2q 2−mq/2 .
So, the inequality (11.4) implies that there exists some constant C depending
only on q such that
!
E Aa,b δ q/2
tm ,tm (B ) C2−mq/2 for all m n. (11.5)
k k+1
C2−mq/2
m
2 −1
m β
2−mq/2 = mβ 2m(1−q/2) < +∞
m1 i=0 m1
[u, v] for some integer k. This quantity is deterministic and depends only on
u − v. Owing to (11.1), for β large enough and for all η > 0,
An Introduction to Rough Paths 55
q 2,δ q/2
sup E sup B1,δ
s,t
+ E sup B s,t
C mβ 2−m(1−q/2)
δ>0 |t−s|<η |t−s|<η
mm(η)
for
some constant C that depends only on q and β. Consequently, the series
β −m(1−q/2)
mm(η) m 2 may be arbitrary small if η is chosen small enough.
Corollary 6.1 proves that (Bδ )δ>0 is tight in G p (R2 ) for all p > q > 2. But
we already know from the Wong-Zakai theorem that Bδs,t converges in proba-
bility to Bs,t uniformly in (s, t) ∈ ∆+ . Thus, (Bδ )δ>0 converges in probability
to B in G p (Rd ).
References
Link to an article with a doi on the publisher site: <dx.doi.org/doi>.
Link to an article on ArXiv: <front.math.ucdavis.edu/arkiv>.
[1] Bass, R.F., Hambly, B., Lyons, T.J. (2002): Extending the Wong-Zakai theorem
to reversible Markov processes. J. Eur. Math. Soc., 4:3, 237–269. <doi:10.1007/
s100970200040>.
[2] Ben Arous, G. (1989): Flots et séries de Taylor stochastiques. Probab. Theory
Related Fields, 81:1, 29–77.
[3] Capitaine, M., Donati-Martin, C. (2001): The Lévy area process for the free
Brownian motion. J. Funct. Anal., 179:1, 153–169. <doi:10.1006/jfan.2000.
3679>.
[4] Castell, F., Gaines, J. (1995): An efficient approximation method for stochastic
differential equations by means of the exponential Lie series. Math. Comput.
Simulation, 38:1-3, 13–19. Probabilités numériques (Paris, 1992).
[5] Chen, K.-T. (1957): Integration of paths, geometric invariants and a generalized
Baker-Hausdorff formula. Ann. of Math. (2), 65, 163–178.
[6] Chen, K.-T. (1958): Integration of paths–a faithful representation, of paths by
noncommutative formal power series. Trans. Amer. Math. Soc., 89:2, 395–407.
[7] Chistyakov, V.V., Galkin, O.E. (1998): On maps of bounded p-variations with
p > 1. Positivity, 2, 19–45.
[8] Cohen, S., Estrade, A. (2000): Non-symmetric approximations for manifold-
valued semimartingales. Ann. Inst. H. Poincaré, Probab. Stat., 36:1, 45–70.
[9] Coutin, L., Qian, Z. (2000): Stochastic differential equations for fractional Brow-
nian motions. C. R. Acad. Sci. Paris Sér. I Math., 331:1, 75–80.
[10] Coutin, L., Qian, Z. (2002): Stochastic analysis, rough path analysis and
fractional brownian motions. Probab. Theory Related Fields, 122:1, 108–140.
<doi:0.1007/s004400100158>.
An Introduction to Rough Paths 57
[51] Wong, E., Zakai, M. (1965): On the convergence of ordinary integrals to stochas-
tic integrals. Ann. Math. Statist., 36, 1560–1564.
[52] Young, L.C. (1936): An inequality of the Hölder type, connected with Stieltjes
integration. Acta Math., 67, 251–282.