Pathwise Stochastic Integrals For Model Free Finance
Pathwise Stochastic Integrals For Model Free Finance
Nicolas Perkowski
CEREMADE & CNRS UMR 7534
Universite Paris-Dauphine
[email protected]
David J. Promel
Humboldt-Universitat zu Berlin
Institut f
ur Mathematik
[email protected]
November 4, 2014
Abstract
We present two different approaches to stochastic integration in frictionless model
free financial mathematics. The first one is in the spirit of Itos integral and based on
a certain topology which is induced by the outer measure corresponding to the minimal
superhedging price. The second one is based on the controlled rough path integral. We
prove that every typical price path has a naturally associated Ito rough path, and
justify the application of the controlled rough path integral in finance by showing that
it is the limit of non-anticipating Riemann sums, a new result in itself. Compared to
the first approach, rough paths have the disadvantage of severely restricting the space of
integrands, but the advantage of being a Banach space theory.
Both approaches are based entirely on financial arguments and do not require any
probabilistic structure.
Contents
1 Introduction
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
5
5
7
9
10
11
Supported by the Fondation Sciences Mathematiques de Paris (FSMP) and by a public grant overseen by
the French National Research Agency (ANR) as part of the Investissements dAvenir program (reference:
ANR-10-LABX-0098).
14
15
18
22
27
B Davies criterion
28
Introduction
This also motivates our second approach, which is more in the spirit of [Lyo95b, DOR13,
DS14]. While in the first approach we do have a continuous operator, it is only continuous
with respect to a sequence of pseudometrics and it seems impossible to find a Banach space
structure that is compatible with it. However, using the model free Ito integral we are able to
show that every typical price path has a natural Ito rough path associated to it. Since in
financial applications we can always restrict ourselves to typical price paths, this observation
opens the door for the application of the controlled rough path integral [Lyo98, Gub04] in
model free finance. Controlled rough path integration has the advantage of being an entirely
linear Banach space theory which simultaneously extends
the Riemann-Stieltjes integral of S against functions of bounded variation which was
used by [DS14];
the Young integral [You36]: typical price paths have finite p-variation for every p > 2,
and therefore
for every F of finite q-variation for q < 2 (so that 1/p + 1/q > 1), the
R
integral F dS is defined as limit of non-anticipating Riemann sums;
Follmers [F
ol81] pathwise It
o integral which was used by [Lyo95b, DOR13]. That this
last integral is a special case of the controlled rough path integral is, to the best of
our knowledge, proved rigorously for the first time in this paper, although also [FH14]
contains some preliminary observations in that direction.
In other words, our second approach covers all previously known techniques of integration in
model free financial mathematics, while the first approach is much more general but at the
price of leaving the Banach space world.
There is only one pitfall: the rough path integral is usually defined as a limit of compensated Riemann sums which have no obvious financial interpretation. This sabotages our
entire philosophy of only using financial Rarguments. That is why we show that under some
weak condition every rough path integral F dS is given as limit of non-anticipating Riemann
sums that do not need to be compensated the first time that such a statement is shown for
general rough path integrals. While this will not change anything in concrete applications, it
is of utmost importance from a philosophical point of view. Indeed, the justification for using
the Ito integral in classical financial mathematics is crucially based on the fact that it is the
limit of non-anticipating Riemann sums, even if in every day applications one never makes
reference to that; see for example the discussion in [Lyo95b].
Plan of the paper
Below we present a very incomplete list of solutions to the stochastic integration problem
under model uncertainty and in a discrete time model free context (both a priori much simpler
problems than the continuous time model free case), and we introduce some notations and
conventions that will be used throughout the paper. In Section 2 we briefly recall Vovks
game-theoretic approach to mathematical finance and introduce our outer measure. We also
construct the topology induced by the outer measure. Section 3 is devoted to the construction
of the model free It
o integral. Section 4 recalls some basic results from rough path theory,
and continues by constructing rough paths associated to typical price paths. Here we also
prove that the rough path integral is given as a limit of non-anticipating Riemann sums. We
also compare F
ollmers pathwise It
o integral with the rough path integral and prove that the
latter is an extension of the former. Appendix A recalls Vovks pathwise Hoeffding inequality.
3
In Appendix B we show that a result of Davie which also allows to calculate rough path
integrals as limit of Riemann sums is a special case of our results in Section 4.
Stochastic integration under model uncertainty
The first works which studied the option pricing problem under model uncertainty were
[ALP95] and [Lyo95b], both considering the case of volatility uncertainty. As described above,
[Lyo95b] is using F
ollmers pathwise Ito integral, while in [ALP95] the problem is reduced to
the classical setting by deriving a worst case model for the volatility.
A powerful tool in financial mathematics under model uncertainty is Karandikars pathwise
construction of the It
o integral [Kar95, Bic81] which allows to construct the Ito integral of
a c`adl`ag integrand simultaneously under all semimartingale measures. The crucial point
that makes the construction useful is that the Ito integral is a continuous operator under
every semimartingale measure. While its pathwise definition would allow us to use the same
construction also in a model free setting, it is not even clear what the output should signify
in that case (for example the construction depends on a certain sequence of partitions and
changing the sequence will change the output). Certainly it is not obvious whether the
Karandikar integral is continuous in any topology once we dispose of semimartingale measures.
A more general pathwise construction of the Ito integral was given in [Nut12], but it suffers
from the same drawbacks with respect to applications in model free finance.
A general approach to stochastic analysis under model uncertainty was put forward in
[DM06], and it is based on quasi sure analysis. While this approach is extremely helpful when
working under model uncertainty, it also does not allow us to define stochastic integrals in a
model free context.
In a related but slightly different direction, in [CDGR11] non-semimartingale models are
studied (which do not violate arbitrage assumptions if the set of admissible strategies is
restricted). While the authors work under one fixed probability measure, the fact that their
price process is not a semimartingale prevents them from using Ito integrals, a difficulty which
is overcome by working with the Russo-Vallois integral [RV93].
Of course all these technical problems disappear if we restrict ourselves to discrete time,
and indeed in that case [BHLP13] develop an essentially fully satisfactory duality theory for
the pricing of derivatives under model uncertainty.
Notation and conventions
Throughout the paper we fix T (0, ) and we write := C([0, T ], Rd ) for the space of
d-dimensional continuous paths. The coordinate process on is denoted by St () = (t),
t [0, T ]. For i {1, . . . , d}, we also write Sti () = i (t), where = ( 1 , . . . , d ). The
filtration (Ft )t[0,T ] is defined as Ft := (Ss : s t), and we set F := FT . Stopping times
and the associated -algebras F are defined as usually.
Unless explicitly stated otherwise, inequalities of the type Ft Gt , where F and G
processes on , are supposed to hold for all , and not modulo null sets, as it is usually
assumed in stochastic analysis.
The indicator function of a set A is denoted by 1A .
A partition of [0, T ] is a finite set of time points, = {0 = t0 < t1 < < tm = T }.
Occasionally, we will identify
with the set of intervals {[t0 , t1 ], [t1 , t2 ], . . . , [tm1 , tm ]}, and
P
write expressions like [s,t] .
4
For f : [0, T ] Rn and t1 , t2 [0, T ], denote ft1 ,t2 := f (t2 ) f (t1 ) and define the pvariation of f restricted to [s, t] [0, T ] as
kf kpvar,[s,t] := sup
m1
X
|ftk ,tk+1 |p
1/p
: s = t0 < < tm = t, m N ,
p > 0,
(1)
k=0
(possibly taking the value +). We set kf kpvar := kf kpvar,[0,T ] . We write T = {(s, t) :
0 s t T } for the simplex and define the p-variation of a function g : T Rn in the
same manner, replacing ftk ,tk+1 in (1) by g(tk , tk+1 ).
For > 0, the space C consists of those functions that are bc times continuously
differentiable, with ( bc)-H
older continuous partial derivatives of order bc. The space
Cb consists of those functions in C that are bounded, together with their partial derivatives,
and we define the norm kkCb by setting
kf kCb :=
bc
X
k=0
2
2.1
In a recent series of papers, Vovk [Vov08, Vov11, Vov12] has introduced a model free, hedging
based approach to mathematical finance that uses arbitrage considerations to examine which
properties are satisfied by typical price paths. This is achieved with the help of an outer
measure given by the cheapest superhedging price.
Recall that T (0, ) and = C([0, T ], Rd ) is the space of continuous paths, with
coordinate process S, natural filtration (Ft )t[0,T ] , and F = FT . A process H : [0, T ] Rd
is called a simple strategy if there exist stopping times 0 = 0 < 1 < . . . , and Fn -measurable
bounded functions Fn : Rd , such that for every we have n () = for all but
finitely many n, and such that
Ht () =
n=0
Fn ()(Sn+1 t () Sn t ()) =
n=0
Fn ()Sn t,n+1 t ()
n=0
is well defined for all , t [0, T ]. Here Fn ()Sn t,n+1 t () denotes the usual inner
product on Rd . For > 0, a simple strategy H is called -admissible if (H S)t () for
all , t [0, T ]. The set of -admissible simple strategies is denoted by H .
Definition 2.1. The outer measure of A is defined as the cheapest superhedging price
for 1A , that is
n
o
n
n
P (A) := inf > 0 : (H )nN H s.t. lim inf ( + (H S)T ()) 1A () .
n
Clearly Pe P . To see the opposite inequality, let Pe(A) < . Let (H n )nN H be a sequence
of simple strategies such that lim inf n supt[0,T ] ( + (H n S)t ) 1A , and let > 0. Define
n := inf{t [0, T ] : ++(H n S)t 1}. Then the stopped strategy Gnt () := Htn ()1t<n ()
is in H H+ and
lim inf ( + + (Gn S)T ()) lim inf 1{++supt[0,T ] (H n S)t 1} () 1A ().
n
Therefore P (A) + , and since > 0 was arbitrary P Pe, and thus P = Pe.
Lemma 2.3 ([Vov12], Lemma 4.1). P is in fact an outer measure, i.e. a nonnegative function
defined on the subsets of such that
- P () = 0;
- P (A) P (B) if A B;
S
P
- if (An )nN is a sequence of subsets of , then P ( n An ) n P (An ).
Proof. Monotonicity and P () = 0 are obvious. So let (An ) be a sequence of subsets of .
Let > 0, n N, and let (H n,m )mN be a sequence of (P (An ) + 2n1 )-admissible simple
lim inf
m
X
P (An ) + + (G S)T
n=0
k
X
n=0
1Sk
n=0
An
for all k N. Since the left hand side does not depend on k, we can replace 1Sk An by
n=0
1Sn An and the proof is complete.
Maybe the most important property of P is that there exists an arbitrage interpretation
for sets with outer measure zero:
Lemma 2.4. A set A is a null set if and only if there exists a sequence of 1-admissible
simple strategies (H n )n H1 such that
lim inf (1 + (H n S)T ) 1A (),
n
(2)
k
X
n=0
Since the left hand side does not depend on k, the sequence (Gm ) satisfies (2).
In other words, if a set A has outer measure 0, then we can make infinite profit by investing
in the paths from A, without ever risking to lose more than the initial capital 1.
This motivates the following definition:
Definition 2.5. We say that a property (P) holds for typical price paths if the set A where
(P) is violated is a null set.
The basic idea of Vovk, which we shall adopt in the following, is that we only need to
concentrate on typical price paths. Indeed, non-typical price path can be excluded since
they are in a certain sense too good to be true: they would allow investors to realize infinite
profit while at the same time taking essentially no risk.
2.2
Before we continue, let us discuss different notions of arbitrage and argue that our outer
measure is an interesting object to study. We start by observing that P is an outer measure
which simultaneously dominates all local martingale measures on .
Propostion 2.6 ([Vov12], Lemma 6.3). Let P be a probability measure on (, F), such that
the coordinate process S is a P-local martingale, and let A F. Then P(A) P (A).
Proof. Let > 0 and let (H n )nN H be such that lim inf n ( + (H n S)T ) 1A . Then
P(A) EP [lim inf ( + (H n S)T )] lim inf EP [ + (H n S)T ] ,
n
where in the last step we used that + (H n S) is a nonnegative P-local martingale and thus
a P-supermartingale.
This already indicates that P -null sets are quite degenerate, in the sense that they are
null sets under all local martingale measures. However, if that was the only reason for our
definition of typical price paths, then a definition based on model free arbitrage opportunities
would be equally valid. A map X : [0, ) is a model free arbitrage opportunity if X is not
identically 0 and if there exists c > 0 and a sequence (H n ) Hc such that lim inf n (H n
S)T () = X() for all . See [DH07, ABPS14] where (a similar) definition is used in the
discrete time setting.
It might then appear more natural to say that a property holds for typical price paths if
the indicator function of its complement is a model free arbitrage opportunity, rather than
working with Definition 2.5. This arbitrage definition would also imply that any property
which holds for typical price paths is almost surely satisfied under every local martingale
measure. Nonetheless we decidedly claim that our definition is the correct one. First of all
the arbitrage definition would make our life much more difficult, since it seems not very easy to
work with. But of course this is only a convenience and cannot possibly serve as justification
of our approach. Instead, we argue by relating the two notions to classical mathematical
finance.
For that purpose recall the fundamental theorem of asset pricing [DS94]: If P is a probability measure on (, F) under which S is a semimartingale, then there exists an equivalent
measure Q such that S is a Q-local martingale if and only if S admits no free lunch with
vanishing risk (NFLVR). But (NFLVR) is equivalent to the two conditions no arbitrage (NA)
(intuitively: no profit without risk) and no arbitrage opportunities of the first kind (NA1)
(intuitively: no very large profit with a small risk). The (NA) property holds if for every
c > 0 and every sequence (H n ) Hc for which limn (H n S)T () exists for all we have
P(limn (H n S)T < 0) > 0 or P(limn (H n S)T = 0) = 1. The (NA1) property holds if
{1 + (H S)T : H H1 } is bounded in P-probability, i.e. if
lim sup P(1 + (H S)T c) = 0.
c HH1
Strictly speaking this is (NA1) with simple strategies, but as observed by [KP11] (NA1) and
(NA1) with simple strategies are equivalent; see also [IP11]. In the case of continuous S, the
equivalence of (NA1) and (NA1) with simple strategies had previously been shown by [Ank05],
Corollary 8.3.2, although here the result is formulated in a slightly different language.
Now the arbitrage definition of typical price paths corresponds to (NA), while our definition corresponds to (NA1):
Propostion 2.7. Let A F be a null set, and let P be a probability measure on (, F) such
that the coordinate process satisfies (NA1). Then P(A) = 0.
Proof. Let (H n )nN H1 be such that 1 + lim inf n (H n S)T 1A . Then for every c > 0
P(A) = P A lim inf (H n S)T > c sup P({(H S)T > c}).
n
HH1
2.3
It will be very useful to introduce a topology on functionals on . For that purpose let us
identify X, Y : R if X = Y for typical price paths. Clearly this defines an equivalence
relation, and we write L0 for the space of equivalence classes. We then introduce the analog
of convergence in probability in our context: (Xn ) converges in outer measure to X if
lim P (|Xn X| > ) = 0
As for completeness, let (Xn ) be a Cauchy sequence with respect to d. Then there exists
a subsequence (Xnk ) such that d(Xnk , Xnk+1 ) 2k for all k, so that
i X
hX
X
d(Xnk , Xnk+1 ) < ,
E
(|Xnk Xnk+1 | 1)
E[|Xnk Xnk+1 | 1] =
k
which means that (Xnk ) converges for typical price paths. Define X := lim inf k Xnk . Then
we have for all n and k
X
d(Xn` , Xn`+1 ) d(Xn , Xnk ) + 2k .
d(Xn , X) d(Xn , Xnk ) + d(Xnk , X) d(Xn , Xnk ) +
`k
2.4
Our definition of the outer measure P is not exactly the same as Vovks [Vov12]. We find our
definition more intuitive and it also seems to be easier to work with. However, since we rely
on some of the results established by Vovk, let us compare the two notions.
For > 0, Vovk defines the set of processes
X
X
k
k
S :=
H : H Hk , k > 0,
k = .
k=0
For every G =
k0 H
k=0
(G S)t () :=
(H k S)t () =
k0
(k + (H k S)t ())
k0
is well defined and takes values in [, ]. Vovk then defines for A the cheapest
superhedging price as
Q(A) := inf > 0 : G S s.t. + (G S)T 1A .
This definition corresponds to the usual construction of an outer measure from an outer
content (i.e. an outer measure which is only finitely subadditive and not countably subadditive); see [Fol99], Chapter 1.4, or [Tao11], Chapter 1.7. Here, the outer content is given
by the cheapest superhedging price using only simple strategies. It is easy to see that P is
dominated by Q:
Lemma 2.10. Let A . Then P (A) Q(A).
P
P
Proof. P
Let G = k H k , with H k Hk and k k = , and assume that + (G S)T 1A .
Then ( nk=0 H k )nN defines a sequence of simple strategies in H , such that
lim inf +
n
n
X
Hk S
= + (G S)T 1A .
T
k=0
Corollary 2.11. For every p > 2, the set Ap := { : kS()kpvar = } has outer
measure zero, that is P (Ap ) = 0.
Proof. Theorem 1 of Vovk [Vov08] states that Q(Ap ) = 0, so P (Ap ) = 0 by Lemma 2.10.
It is a remarkable result of [Vov12] that if = C([0, ), R) (i.e. if the asset price process
is one-dimensional), and if A is invariant under time changes and such that S0 () = 0
for all A, then A F and Q(A) = P(A), where P denotes the Wiener measure. This can
be interpreted as a pathwise Dambis Dubins-Schwarz theorem.
Model free It
o integration
The present section is devoted to the construction of a model free Ito integral. The main
ingredient is a (weak) type of model free Ito isometry, which allows us to estimate the integral
against a step function in terms of the amplitude of the step function and the quadratic
variation of the price path. Based on the topology introduced in Section 2.3, it is then easy
to extend the integral to c`
adl`
ag integrands by a continuity argument.
Since we are in an unusual setting, let us spell out the following standard definitions:
Definition 3.1. A process F : [0, T ] Rd is called adapted if the random variable
7 Ft () is Ft -measurable for all t [0, T ].
The process F is said to be c`
adl`
ag if the sample path t 7 Ft () is c`adl`ag for all .
To prove our weak It
o isometry, we will need an appropriate sequence of stopping times:
Let = ( 1 , . . . , d ) and n N. For each i = 1, . . . , d define inductively
n,i
() := inf t kn,i : | i (t) i (kn,i )| 2n , k N.
k+1
0n,i () := 0,
Since we are working with continuous paths and we are considering entrance times into closed
sets, the maps ( n,i ) are indeed stopping times, despite the fact that (Ft ) is neither complete
nor right-continuous. Denote n,i := {kn,i : k N}. To obtain an increasing sequence of
partitions, we take the union of the ( n,i ), that is we define 0n := 0 and then
n
()
k+1
:= min t >
kn ()
:t
d
[
n,i
() ,
k N,
(3)
i=1
2
n,i
i (k+1
t) i (kn,i t) ,
t [0, T ],
n N,
k=0
2n
Ntn ()
d
X
i=1
Proof. For i {1, ..., d} define Ntn,i := max{k N : kn,i t}. Since i is continuous, we
n,i
n,i
have | i (k+1
) i (kn,i )| = 2n as long as k+1
T . Therefore, we obtain
Ntn ()
d
X
n,i
Ntn,i ()
i=1
()1
d Nt X
X
i=1
k=0
n,i
(k+1
)
22n
2
(kn,i )
2n
d
X
Vtn,i ().
i=1
We will start by constructing the integral against step functions, which are defined similarly as simple strategies, except possibly unbounded: A process F : [0, T ] Rd is called
a step function if there exist stopping times 0 = 0 < 1 < . . . , and Fn -measurable functions
Fn : Rd , such that for every we have n () = for all but finitely many n, and
such that
X
Ft () =
Fn ()1[n (),n+1 ()) (t).
n=0
For notational convenience we are now considering the interval [n (), n+1 ()) which is closed
on the right hand side. This allows us define the integral
(F S)t :=
Fn Sn t,n+1 t =
n=0
Fn Sn t,n+1 t ,
t [0, T ].
n=0
The following lemma will be the main building block in the construction of our integral.
Lemma 3.4 (Model free It
o isometry). Let a > 0 and let F be a step function. Then for all
a, b, c > 0 we have
sup Fnk Sa nk t,a nk+1 t a d2n .
t[0,T ]
Hence, the pathwise Hoeffding inequality, Lemma A.1 in Appendix A, yields for every R
the existence of a 1-admissible simple strategy H ,n H1 such that
2 (n )
,n
2n 2
1 + (H S)t exp (F S)a t (Nt
+ 1)2
a d =: E,n
a t
2
12
Nt
( )
:= max{k : nk t} Ntn + Nt
Et,n + Et,n
1
2 2
lim inf sup
exp ab c ca d .
n t[0,T ]
2
2
2
The argument inside the exponential is maximized for = b/(a cd), in which case we obtain
1/2 exp(b2 /(2d)). The statement now follows from Remark 2.2.
Of course, we did not actually establish an isometry but only an upper bound for the
integral. But this estimate is the key ingredient which allows us to extend the model free
Ito integral to more general integrands, and it is this analogy to the classical setting that the
term model free It
o isometry alludes to.
Let us extend the topology of Section 2.3 to processes: we identify X, Y : [0, T ] Rm
if for typical price paths we have Xt = Yt for all t [0, T ], and we write L0 ([0, T ], Rm ) for the
resulting space of equivalence classes which we equip with the distance
d (X, Y ) := E[kX Y k 1].
Ideally, we would like the stochastic integral on step functions to be continuous with
respect to d . However, using Proposition 2.6 it is easy to see that P (k((1/n) S)k > ) = 1
for all n N and > 0. This is why we also introduce for c > 0 the pseudometric
dc (X, Y ) := E[(kX Y k 1)1hSiT c ] d (X, Y ),
and then
dcpct (X, Y ) :=
n=1
dc (F, G)
+
+ ab c
a
b2 d (F, G)
c
2 exp
+
+ ab c
2d
a
p
p
whenever a, b > 0. Setting a = dc (F, G) and b = d| log a|, we deduce that
n=1
13
adl`
ag process with values in Rd . Then there exists
RTheorem 3.5. Let F be an adapted, c`
n
n
F dS L0 ([0, T ], R) such thatR for every sequence of step functions
R (F ) with limn d (F , F ) =
n
0 we have limn dcpct ((F S), F dS) = 0. The integral
process F dS is continuous for typR
ical price paths, and there exists a representative F dS which is adapted, although it may
Rt
R
R
take the values .We usually write 0 Fs dSs := F dS(t), and we call F dS the model
free Ito integral of RF with respect to S.
The map F 7 F dS is linear, satisfies
Z
Z
dcpct
F dS, GdS . d (F, G)1/2
for all > 0, and the model free It
o isometry extends to this setting:
n Z
o
forRall m N. So, if (cm log m) converges to 0, then for typical price paths (F m S) converges
to F dS.
Proof. For c > 0 the model free It
o isometry gives
Z
p
1
m
P
k(F S) F dSk cm 4d log m c {hSiT c} 2 .
m
Since this is summable in m, the claim follows from Borel Cantelli (which only requires
countable subadditivity).
Remark 3.7. The speed of convergence (5) is better than the one that can be obtained using
the arguments in [Kar95], where the summability of (cm ) is needed.
Our second approach to model free stochastic integration is based on the rough path integral,
which has the advantage of being a continuous linear operator between Banach spaces. The
disadvantage is that we have to restrict the set of integrands to those locally looking like S,
modulo a smoother remainder. Our two main results in this section are that every typical
price path has a naturally associated Ito rough path, and that the rough path integral can be
constructed as limit of Riemann sums.
Let us start by recalling the basic definitions and results of rough path theory.
14
4.1
Here we follow more or less the lecture notes [FH14], to which we refer for a gentle introduction
to rough paths. More advanced monographs are [LQ02, LCL07, FV10]. The main difference
to [FH14] in the derivation below is that we use p-variation to describe the regularity, and not
Holder continuity, because it is not true that all typical price paths are Holder continuous.
Also, we make an effort to give reasonably sharp results, whereas in [FH14] the focus lies
more on the pedagogical presentation of the material. We stress that in this subsection we
are merely collecting classical results.
Definition 4.1. A control function is a continuous map c : T [0, ) with c(t, t) = 0 for
all t [0, T ] and such that c(s, u) + c(u, t) c(s, t) for all 0 s u t T .
Observe that if f : [0, T ] Rd satisfies |fs,t |p c(s, t) for all (s, t) T , then the pvariation of f is bounded from above by c(0, T ).
Definition 4.2. Let p (2, 3). A p-rough path is a map S = (S, A) : T Rd Rdd such
that Chens relation
S i (s, t) = S i (s, u) + S i (u, t)
and
holds for all 1 i, j d and 0 s u t T and such that there exists a control function
c with
|S(s, t)|p + |A(s, t)|p/2 c(s, t)
(in other words S has finite p-variation and A has finite p/2-variation). In that case we call
A the area of S.
Remark 4.3. Chens relation simply states that S is the increment of a function, that
is S(s, t) = S(0, t) S(0, s) =: S(t) S(s), and that for all i, j there exists a function
j
f i,j : [0, T ] R such that Ai,j (s, t) = f i,j (t) f i,j (s) S i (s)Ss,t
. Indeed, it suffices to set
j
i,j
i,j
i
f (t) = A (0, t) + S (0)S0,t .
Remark 4.4. The (strictly speaking incorrect) name area stems from the fact that if S is
smooth and two-dimensional and if
Z t
Z t Z r2
i
Ai,j (s, t) =
dS i (r1 )dS j (r2 ) =
Ss,r
dS j (r2 ),
2
s
then the antisymmetric part of A(s, t) corresponds to the algebraic area enclosed by the curve
(S(r))r[s,t] . It is a deep insight of Lyons [Lyo98], proving a conjecture of F
ollmer, that the
area is exactly the additional information which is needed to solve differential equations driven
by S in a pathwise continuous manner, and to construct stochastic integrals as continuous
maps. Actually, [Lyo98] solves a much more general problem and proves that if the driving
signal is of finite p-variation for some p > 1, then it has to be equipped with the iterated
integrals up to order bpc 1 to obtain a continuous integral map. The for us relevant case
p (2, 3) was already treated in [Lyo95a].
Example 4.5. If S is a continuous semimartingale and if we set S(s, t) = Ss,t as well as
Z t Z r2
Z t
i,j
i
j
i
A (s, t) =
dS (r1 )dS (r2 ) =
Ss,r
dS j (r2 ),
2
s
15
where the integral can be understood either in the Ito or in the Stratonovich sense, then almost
surely S = (S, A) is a p-rough path for all p (2, 3). This is shown in [CL05], and we will
give a simplified model free proof below (indeed we will show that every typical price path is
a p-rough path for all p (2, 3), from where the statement about continuous semimartingales
easily follows).
From now on we fix p > 2 and we assume that S is a p-rough path. Gubinelli [Gub04]
observed that for every rough path there is a naturally associated Banach space of integrands,
the space of controlled paths. Heuristically, a path F is controlled by S, if it locally looks
like S, modulo a smooth remainder. The precise definition is:
Definition 4.6. Let q > 0 be such that 2/p + 1/q > 1. Let F : [0, T ] Rn and F 0 : [0, T ]
Rnd . We say that the pair (F, F 0 ) is controlled by S if the derivative F 0 has finite q-variation,
and the remainder RF : T Rn , defined by
RF (s, t) = Fs,t Fs0 Ss,t ,
has finite r-variation for 1/r = 1/p + 1/q. In this case, we write (F, F 0 ) CSq = CSq (Rn ), and
define
k(F, F 0 )kC q := kF 0 kqvar + kRF krvar .
S
|F00 |
where c is a control function for S. As the image of the continuous path S is compact, it is
not actually necessary to assume that is bounded. We may always consider a C 1+ function
of compact support, such that agrees with on the image of S.
This example shows that in general RF (s, t) is not a path increment of the form RF (s, t) =
G(t) G(s) for some function G defined on [0, T ], but really a function of two variables.
16
Example 4.8. Let G be a path of finite r-variation for some r with 1/p + 1/r > 1. Setting
(F, F 0 ) = (G, 0), we obtain a controlled path in CSq , where 1/q = 1/r 1/p. In combination
with Theorem 4.9 below, this example shows in particular that the controlled rough path
integral extends the Young integral and the Riemann-Stieltjes integral.
R
The basic idea of rough path integration is that if we already know how
R to define SdS,
and if F looks like S on small scales, then we should be able to define F dS as well. The
precise result is given by the following theorem:
Theorem 4.9 (Theorem 4.9 in [FH14], see also [Gub04], Theorem 1).
R Let q be such that
2/p + 1/q > 1. Let (F, F 0 ) CSq . Then there exists a unique function F dS C([0, T ], Rn )
which satisfies
Z t
Fu dSu Fs Ss,t Fs0 A(s, t) . kSkpvar,[s,t] kRF krvar,[s,t] + kAkp/2var,[s,t] kF 0 kqvar,[s,t]
s
for all (s, t) T . The integral is given as limit of the compensated Riemann sums
Z
Fu dSu = lim
0
m
[s1 ,s2
Fs1 Ss1 ,s2 + Fs01 A(s1 , s2 ) ,
(6)
] m
Remark 4.10. To the best of our knowledge, there is no publication in which the controlled
path approach to rough paths is formulated using p-variation regularity. The references on
the subject all work with H
older continuity. But in the p-variation setting, all the proofs work
exactly as in the H
older setting, and it is a simple exercise to translate the proof of Theorem 4.9
in [FH14] (which is based on Youngs maximal inequality which we will encounter below) to
obtain Theorem 4.9.
There is only one small pitfall: We did not require F or F 0 to be continuous. The rough
path integral for discontinuous functions is somewhat tricky, see [Wil01]. But here we do not
run into any problems, because the integrand S = (S, A) is continuous. The construction based
on Youngs maximal inequality works as long as integrand and integrator have no common
discontinuities, see the Theorem on page 264 of [You36].
If now Cb1+ for some > 0, then using a Taylor expansion one can show that there
exist p > 2 and q > 0 with 2/p + 1/q > 0, such that (F, F 0 ) 7 ((F ), 0 (F )F 0 ) is a locally
bounded map from CSp to CSq . Combining this with the fact that the rough path integral is a
bounded map from CSq to CSp , it is not hard to prove the existence of solutions to the rough
differential equation
dXt = (Xt )dSt ,
X0 = x0 ,
(7)
p R
t [0, T ], where X CS , (Xt )dSt denotes the rough path integral, and S is a typical price
path. Similarly, if Cb2+ , then the map (F, F 0 ) 7 ((F ), 0 (F )F 0 ) is a locally Lipschitz
continuous from CSp to CSq , and this yields the uniqueness of the solution to (7) at least
among the functions in the Banach space CSp . See Section 5.3 of [Gub04] for details.
17
pvar
CM |F0 F0 | + |F00 F00 | + kF 0 F 0 kqvar
pvar + kA Ak
p/2var ,
+ kRF RF krvar + kS Sk
as long as
pvar , kAk
p/2var } M.
max{|F00 | + k(F, F 0 )kC q , |F00 | + k(F , F 0 )kCq , kSkpvar , kAkp/2var , kSk
S
In other words, the rough path integral depends on integrand and integrator in a locally
Lipschitz continuous way, and therefore it is no surprise that the solutions to differential
equations driven by rough paths depend continuously on the signal.
4.2
Our second approach to stochastic integration in model free financial mathematics is based
on the rough path integral. Here we show that for every typical price path, the pair (S, A) is
a p-rough path for all p (2, 3), where
Z t
Z t
Z t
j
Sri dSrj Ssi Ss,t
.
A(s, t) =
Ss,r dSr :=
Ss,r dSr :=
s
1i,jd
The main ingredient in the proof will be our speed of convergence (5).
18
Let p > 2. Then for typical price paths, A = (Ai,j )1i,jd has finite p/2-variation, and in
particular S = (S, A) is a p-rough path.
Proof. Define the dyadic stopping times (kn )n,kN by 0n := 0 and
n
k+1
:= inf{t kn : |St Skn | = 2n },
P
n Sk
n . Accorcing to (5), for typical
n ) (t), so that kS
and set Stn := k Skn 1[kn ,k+1
2
price paths there exists C() > 0 such that
Z
p
n
(S S)() SdS()
C()2n log n.
Fix such a typical price path , which is also of finite q-variation for all q > 2 (recall from
Corollary 2.11 that this is satisfied by typical price paths). Let us show that for such , the
process A is of finite p/2-variation for all p > 2.
We have for (s, t) T , omitting the argument of the processes under consideration,
Z t
|As,t |
Sr dSr (S n S)s,t + |(S n S)s,t Ss Ss,t |
s
p
C()2n log n + |(S n S)s,t Ss Ss,t | . C()2n(1) + |(S n S)s,t Ss Ss,t |
for every n N, > 0. The second term on the right hand side can be estimated, using an
argument based on Youngs maximal inequality (see [LCL07], Theorem 1.16), by
max{2n c(s, t)1/q , (#{k : kn [s, t]})12/q c(s, t)2/q + c(s, t)2/q },
(8)
where c(s, t) is a control function with |Ss,t |q c(s, t) for all (s, t) T . Indeed, if there exists
no k with kn [s, t], then |(S n S)s,t Ss Ss,t | 2n c(s, t)1/q , using that |Ss,t | c(s, t)1/q .
This corresponds to the first term in the maximum in (8).
Otherwise, note that at the price of adding c(s, t)2/q to the right hand side, we may
suppose that s = kn0 for some k0 . Let now kn0 , . . . , kn0 +N 1 be those (kn )k which are in [s, t).
Without loss of generality we may suppose N 2, because otherwise (S n S)s,t = Ss Ss,t .
Abusing notation, we write kn0 +N = t. The idea is now to successively delete points (kn0 +` )
from the partition, in order to pass from (S n S) to Ss Ss,t . By super-additivity of c, there
must exist ` {1, . . . , N 1}, for which
c(kn0 +`1 , kn0 +`+1 )
2
c(s, t).
N 1
Deleting kn0 +` from the partition and subtracting the resulting integral from (S n S)s,t , we
get
|Skn +`1 Skn +`1 ,kn +` + Skn +` Skn +` ,kn +`+1 Skn +`1 Skn +`1 ,kn +`+1 |
0
0
0
0
0
0
0
0
0
2
2/q
= |Skn +`1 ,kn +` Skn +` ,kn +`+1 | c(kn0 +`1 , kn0 +`+1 )2/q
c(s, t)
.
0
0
0
0
N 1
19
Successively deleting all the points except kn0 = s and kn0 +N = t from the partition gives
N
2/q
X
2
. N 12/q c(s, t)2/q ,
|(S S)s,t Ss Ss,t |
c(s, t)
k1
n
k=2
and therefore (8). Now it is easy to see that #{k : kn [s, t]} 2nq c(s, t) (compare also the
proof of Lemma 3.3), and thus
|As,t | . C()2n(1) + max{2n c(s, t)1/q , (2nq c(s, t))12/q c(s, t)2/q + c(s, t)2/q }
= C()2n(1) + max{2n c(s, t)1/q , 2n(2q) c(s, t) + c(s, t)2/q }.
(9)
This holds for all n N, > 0, q > 2. Let us suppose for the moment that c(s, t) 1 and let
> 0 to be determined later. Then there exists n N for which 2n1 < c(s, t)1/(1) 2n .
Using this n in (9), we get
n
o
|As,t | .,, c(s, t) + max c(s, t)1/(1) c(s, t)/q , c(s, t)(2q)/(1)+ + c(s, t)2/q
q+(1)
2q+(1)
2/q
q(1)
1
= c(s, t) + max c(s, t)
, c(s, t)
+ c(s, t)
.
We would like all the exponents in the maximum on the right hand side to be larger or equal
to 1. For the first term, this is satisfied as long as < 1. For the third term, we need q/2.
For the second term, we need (q 1 )/(1 ). Since > 0 can be chosen arbitrarily
close to 0, it suffices if > q 1. Now, since q > 2 can be chosen arbitrarily close to 2, we
see that can be chosen arbitrarily close to 1. In particular, we may take = p/2 for any
p > 2, and we obtain
|As,t |p/2 ., c(s, t)(1 + c(s, t) ) c(s, t)(1 + c(0, T ) )
for a suitable > 0.
It remains to treat the case c(s, t) > 1, for which we simply estimate
Z
Z
p/2
p/2
Sr dSr
+ kSkp
Sr dSr
+ kSkp c(s, t).
|As,t |p/2 .p
0
So for every interval [s, t] we can estimate |As,t |p/2 .,p c(s, t), and the proof is complete.
Remark 4.13. To the best of our knowledge, this is one of the first times that a non-geometric
rough path is constructed in a non-probabilistic setting, and certainly we are not aware of any
works where rough paths are constructed using financial arguments.
We also point out that, thanks to Proposition 2.6, we gave a simple, model free, and
pathwise proof for the fact that a local martingale together with its It
o integral defines a
rough path. While this seems intuitively clear, the only other proof that we are aware of is
somewhat involved: it relies on a strong version of the Burkholder-Davis-Gundy inequality, a
time change, and Kolmogorovs continuity criterion; see [CL05] or Chapter 14 of [FV10].
The following auxiliary result will allow us to obtain the rough path integral as a limit of
Riemann sums, rather than compensated Riemann sums which are usually used to define it.
20
Lemma 4.14. Let (cn )nN be a sequence of positive numbers such that (cn log n) converges
n
n
n
n
to 0 for all > 0. For
P n N define 0 := 0 and k+1 := inf{t k : |Snt Sk | = cn },
n
n ) (t). Then for typical price paths, ((S
k N, and set St = k Skn 1[kn ,k+1
S)) converges
R
uniformly to SdS. Moreover, for p > 2 and for typical price paths there exists a control
function c = c(p, ) such that
|(S n S)kn ,`n () Skn ()Skn ,`n ()|p/2
1.
c(kn , `n )
n k<`
R
Proof. The uniform convergence of ((S n S)) to SdS follows from Corollary 3.6. For the
second claim, fix n N and k < ` such that `n T . Then
Z
n
n
|(S S)kn ,`n Skn Skn ,`n | .
(S S)
Ss dSs
+ Akn ,`n
0
p
n n 2/p
. cn log n + vp/2 (k , ` ) . cn1 + vp/2 (kn , `n )2/p , (10)
sup sup
where > 0 and the last estimate holds by our assumption on the sequence (cn ), and where
p/2
vp/2 (s, t) := kAkp/2var,[s,t] for (s, t) T . Of course, this inequality only holds for typical
price paths and not for all .
On the other side, the same argument as in the proof of Theorem 4.12 (using Youngs
maximal inequality and successively deleting points from the partition) shows that
|(S n S)kn ,`n Skn Skn ,`n | . cn2q vq (kn , `n ),
(11)
2q+(1)
(1)
21
4.3
Theorem 4.12 shows that we can apply the controlled rough path integral in model free
financial mathematics, since every typical price path is a rough path. But there remains
a
R
philosophical problem: As we have seen in Theorem 4.9, the rough path integral F dS is
given as limit of the compensated Riemann sums
Z t
X
Fs dSs = lim
Fr1 Sr1 ,r2 + Fr01 A(r1 , r2 ) ,
m
[r1 ,r2 ] m
where ( m ) is an arbitrary sequence of partitions of [0, t] with mesh size going to 0. The
term Fr1 Sr1 ,r2 has an obvious financial interpretation as profit made by buying Fr1 units of
the traded asset at time r1 and by selling them at time r2 . However, for the compensator
Fr01 A(r1 , r2 ) there seems to be no financial interpretation, and therefore it is not clear whether
the rough path integral can be understood as profit obtained by investing in S.
However, we observed in Section 3 that along suitable stopping times (kn )n,k , we have
Z t
X
n t .
Ss dSs = lim
Skn Skn t,k+1
n
By the philosophy of controlled paths, we expect that also for F which looks like S on small
scales we should obtain
Z t
X
n t ,
Fs dSs = lim
Fkn Skn t,k+1
n
Assumption (rie). Let n = {0 = tn0 < tn1 < < tnNn = T }, n N, be a given sequence of
partitions such that sup{|Stnk ,tnk+1 | : k = 0, . . . , Nn 1} converges to 0, and let p (2, 3). Set
Stn :=
NX
n 1
k=0
(S n
S) converge uniformly to
(12)
Remark 4.15. The coarse-grained regularity condition (12) has recently been independently
rediscovered in [Kel14]; see also [GIP14].
Our proof that the rough path integral is given as limit of Riemann sums is somewhat indirect. We translate everything from Ito type integrals to related Stratonovich type integrals,
for which the convergence follows from the continuity of the rough path integral, Proposition 4.11. Then we translate everything back to our Ito type integrals. To go from Ito to
Stratonovich, we need the quadratic variation:
22
Sri dSrj
Srj dSri .
hS , S it = lim hS
n
, S j int
NX
n 1
(Stin
= lim
k+1 t
Stin t )(Stjn
k+1 t
Stjn t ).
(13)
k=0
NX
n 1
Stin
k+1
j
t Stn
k+1
j
i
t Stn t Stn t
k
k=0
k+1 t
Stjn
k+1 t
k+1 t
k+1 t
+ Stin t,tn
k
k+1 t
Stjn t,tn
k
k+1 t
R
so that the convergence in (13) is a consequence of the convergence of (S n S) to SdS.
To see that hS i , S j i is of bounded variation, note that
2
2
1 i
j
i
j n
i
j n
n
n
Stn t,tn t Stn t,tn t =
(S + S )tk t,tk+1 t (S S )tk t,tk+1 t
k
k+1
k
k+1
4
(read hS i , S j i = 1/4(hS i +S j ihS i S j i)). In other words, the n-th approximation of hS i , S j i
is the difference of two increasing functions, and its total variation is bounded from above by
NX
n 1
(S i + S j )tnk ,tnk+1
2
NX
m 1
2
+ (S i S j )tnk ,tnk+1
. sup
(Stim ,tm )2 + (Stjm ,tm )2 .
m
k=0
k+1
k+1
k=0
Since the right hand side is finite, also the limit hS i , S j i is of bounded variation.
Given the quadratic variation, the existence of the Stratonovich integral is straightforward:
Lemma 4.17. Under Assumption (rie), define Sn |[tnk ,tnk+1 ] as the linear interpolation of Stnk
R
and Stn for k = 0, . . . Nn 1. Then ( Sn dSn ) converges uniformly to
k+1
Sr dSr :=
s
Rt
s
1
Sr dSr + hSis,t .
2
n dS
n for (s, t) T , we have supn kAn kp/2var < .
Ss,r
r
23
(14)
Proof. Let n N and k {0, . . . , Nn 1}. Then for t [tnk , tnk+1 ] we have
Stn = Stnk +
so that
Z
tn
k+1
tn
k
t tnk
Stn ,tn ,
tnk+1 tnk k k+1
1
Srn dSrn = Stnk Stnk ,tnk+1 + Stnk ,tnk+1 Stnk ,tnk+1 ,
2
(15)
from where the uniform convergence and the representation (14) follow by Lemma 4.16.
To prove that An has uniformly bounded p2 -variation, consider (s, t) T . If there exists
k such that tnk s < t tnk+1 , then we estimate
p/2
Z t
p/2 Z t
|Stn ,tn |2
n
p/2
n
n
|A (s, t)| =
Ss,r dSr (r s) n k k+1n 2 dr
|
t
|t
s
s
k
k+1
=
2p/2
|t s|
|Stnk ,tnk+1 |p
|tnk+1 tnk |p
|t s|
kSkppvar,[tn ,tn ] .
k k+1
|tnk+1 tnk |
(16)
Otherwise, let k0 be the smallest k such that tnk (s, t), and let k1 be the largest such k. We
decompose
n
tnn
n S
An (s, t) = An (s, tnk0 ) + An (tnk0 , tnk1 ) + An (tnk1 , t) + Ss,t
k0
n
k0 ,tk1
n
tnn ,t .
n S
+ Ss,t
k1
k1
,tn
k
Stnk Stnk
0
,tn
k
|p/2 + (hSintn
n
k0 ,tk1
)p/2 ,
where hSin denotes the n-th approximation of the quadratic variation. By (12) and Lemma 4.16,
there exists a control function c so that the right hand side is bounded from above by c(tnk0 , tnk1 ).
n
n
n n
Combining this with (16) and a simple estimate for the terms Ss,t
n Stn ,tn and Ss,tn Stn ,t ,
k0
k0 k1
k1
k1
we deduce that kAn kp/2var . c(0, T ) + kSk2
, and the proof is complete.
pvar
Fs dSs = lim
0
NX
n 1
k=0
n
respect to Sn is given by R
T
s s,t
F n
F n
has finite r-variation for 1/r = 1/p + 1/q.
24
|t s|
kRF krvar,[tnk ,tnk+1 ] + kF 0 kqvar,[tnk ,tnk+1 ] + kSkpvar,[tnk ,tnk+1 ] ,
n
tk |
|tnk+1
(17)
where in the last step we used that 1/r = 1/p + 1/q, and thus r/q + r/p = 1.
Otherwise, there exists k {1, . . . , Nn 1} with tnk (s, t). Let k0 and k1 the smallest
and largest such k, respectively. Then
0
n n (tn , t)|r +|Fs,t
n n (tn , tn )|r +|R
n n (s, t)|r .r |R
n n (s, tn )|r +|R
n Stn
|R
k1
k0 k1
k0
k
F
F
F
F
k0
,tn
k
0
r
n Stn ,t | .
|r +|Fs,t
k
k1
n (tn , tn ) = RF (tn , tn ), and therefore we can use (17), the assumption on RF , and
Now R
k0 k1
F n k0 k1
the fact that 1/r = 1/p + 1/q (which is needed to treat the last two terms on the right hand
side), to obtain
n n krvar .r kRF krvar + kF 0 kqvar + kSkpvar .
kR
F
n ) converges uniformly to (F, RF ).
On the other side, since F and RF are continuous, (F n , R
F n
Now for continuous functions, uniform convergence with uniformly bounded p-variation implies convergence in p0 -variation for every p0 > p. See Exercise 2.8 in [FH14] for the case of
Holder continuous functions.
Thus, using Lemma 4.17, we see that if p0 > p and q 0 > q are such that 2/p0 +1/q 0 > 0, then
n
((S , An )n ) converges in (p0 , p0 /2)-variation to (S, A ), where A (s, t) = A(s, t) + 1/2hSis,t .
n )) converges in (q 0 , p0 , r0 )-variation to (F, F 0 , RF ), where 1/r0 = 1/p0 +
Similarly, ((F n , F 0 , R
F n
0
1/q .
R
R
Proposition 4.11 now yields the uniform convergence of F n dSn to F dS, by which we
denote the rough path integral of the controlled path (F, F 0 ) against the rough path (S, A ).
But for every t [0, T ] we have
Z t
X 1
lim
Fsn dSsn = lim
(Ftn + Ftnk+1 )Stnk ,tnk+1
n 0
n
2 k
k:tn
t
k+1
X
1 X
n
n
n
n
n
n
n
Ftk ,tk+1 Stk ,tk+1 .
= lim
Ftk Stk ,tk+1 +
n
2 n
n
k:tk+1 t
k:tk+1 t
R
where the integral 0 F (Ss )dSs is given as limit of Riemann sums along that same sequence
of partitions.
Friz and Hairer [FH14] observe that if for p (2, 3) the function S is of finite p-variation
and hSi is an arbitrary continuous function of finite p/2-variation, then setting
1 i j
Sym(A)(s, t) := (Ss,t
Ss,t + hSis,t )
2
one obtains a reduced rough path (S, Sym(A)). They continue to show that if F is controlled
R
by S with symmetric derivative F 0 , then it is possible to define the rough path integral F dS.
This is not surprising since then we have Fs0 As,t = Fs0 Sym(A)s,t for the compensator term in
the definition of the rough path integral. They also derive
R an Ito formula for reduced rough
paths, which takes the same form as (18), except that now F (S)dS is a rough path integral
(and therefore defined as limit of compensated Riemann sums).
So both the assumption and the result of [FH14] are slightly different from the ones
in [Fol81], and while it seems intuitively clear, it is still not shown rigorously that Follmers
pathwise It
o integral is a special case of the rough path integral. We will now show that
Follmers result is a special case of Theorem 4.18. For that purpose we only need to prove
that Follmers condition on the convergence of the quadratic variation is a special case of the
assumption in Theorem 4.18, at least as long as we only need the symmetric part of the area.
Definition 4.19. Let f C([0, T ], R) and let n = {0 = tn0 < tn1 < < tnNn = T }, n N
be such that sup{|ftnk ,tnk+1 | : k = 0, . . . , Nn 1} converges to 0. We say that f has quadratic
variation along ( n ) in the sense of F
ollmer if the sequence of discrete measures (n ) on
([0, T ], B[0, T ]), defined by
NX
n 1
|ftnk ,tnk+1 |2 tnk ,
n :=
(19)
k=0
t [0, T ].
Lemma 4.20 (see also [Vov11], Proposition 6.1). Let p (2, 3), and let S = (S 1 , ..., S d )
C([0, T ], Rd ) have finite p-variation. Let n = {0 = tn0 < tn1 < < tnNn = T }, n N, be a
sequence of partitions such that sup{|Stnk ,tnk+1 | : k = 0, . . . , Nn 1} converges to 0. Then the
following conditions are equivalent:
1. The function S has quadratic variation along ( n ) in the sense of F
ollmer.
2. For all 1 i, j d, the discrete quadratic variation
hS
, S j int
:=
NX
n 1
Stin t,tn
k
k+1 t
Stjn t,tn
k=0
k+1 t
PNn 1 i
Stn 1[tnk ,tnk+1 ) , i {1, . . . , d}, n N, the Riemann sums (S n,i S j ) +
3. For S n,i := k=0
k
R
R
(S n,j S i ) converge uniformly to a limit S i dS j + S j dS i . Moreover, the symmetric
part of the approximate area,
Sym(An )i,j (s, t) =
1
j
i
(S n,i S j )s,t +(S n,j S i )s,t Ssi Ss,t
Ssj Ss,t
, 1 i, j d, (s, t) T ,
2
Stin t,tn
k
1
((S i + S j )tnk t,tnk+1 t )2 (Stin t,tn t )2 (Stjn t,tn t )2 .
k
k+1
k
k+1
2
Thus, the uniform convergence of hS i , S j in and the fact that hS i , S j i = [S i , S j ] follow once we
show that F
ollmers weak convergence of the measures (19) implies the uniform convergence of
their distribution functions. But since the limiting distribution is continuous by assumption,
this is a standard result.
Next, assume 2. The uniform convergence of the Riemann sums (S n,i S j ) + (S n,j S i ) is
shown as in Lemma 4.16. To see that Sym(An ) has uniformly bounded p/2-variation along
( n ), note that for 0 k ` Nn and 1 i, j d we have
|(S n,i S j )tnk ,tn` + (S n,j S i )tnk ,tn` Ssi Stjn ,tn Ssj Stin ,tn |p/2 = |Stin ,tn Stjn ,tn hS i , S j intn ,tn |p/2
k
In the construction of the pathwise Ito integral for typical price processes we needed the
following result, a pathwise formulation of the Hoeffding inequality which is due to Vovk.
Here we present a slightly adapted version.
Lemma A.1 ([Vov12], Theorem A.1). Let (n )nN be a strictly increasing sequence of stopping
times with 0 = 0, such that for every we have n () = for all but finitely many
n N. Let for n N the function hn : Rd be Fn -measurable, and suppose that there
exists a Fn -measurable bounded function bn : R, such that
sup |hn ()Sn t,n+1 t ()| bn ()
t[0,T ]
27
(20)
for all . Then for every R there exists a simple strategy H H1 such that
X
Nt
2 X
2
bn
1 + (H S)t exp
hn Sn t,n+1 t
2
n=0
n=0
hn ()Sn t,n+1 t ()
exp b2n ()
2
2bn ()
=: fn ()Sn t,n+1 t ().
(21)
P
This inequality is shown in (A.1) of [Vov12]. We define Ht := n Fn 1(n ,n+1 ] (t), with Fn
that have to be specified. We choose F0 := f0 , which is bounded and F0 -measurable, and on
[0, 1 ] we obtain
2
1 + (H S)t exp h0 Sn t,n+1 t b20 .
2
Observe also that 1 + (H S)1 = 1 + f0 S0 ,1 is bounded, because by assumption h0 S0 ,1 is
bounded by the bounded random variable b0 .
Assume now that Fk has been defined for k = 0, . . . , m 1, that
X
Nt
2 X
2
bn
1 + (H S)t exp
hn Sn t,n+1 t
2
n=0
n=0
for all t [0, m ], and that 1 + (H S)m is bounded. We define Fm := (1 + (H S)m )fm ,
which is Fm -measurable and bounded. From (21) we obtain for t [m , m+1 ]
1 + (H S)t = 1 + (H S)m + (1 + (H S)m )fm Sm t,m+1 t
2 2
t
2 X
b2n ,
exp
hn Sn t,n+1 t
2
n=0
n=0
where the last step follows from the induction hypothesis. From the first line of the last
equation we also obtain that 1 + (H S)m+1 is bounded because fm Sm ,m+1 is bounded for
the same reason that f0 S0 ,1 is bounded.
Davies criterion
It was already observed by Davie [Dav07] that in certain situations the rough path integral
can be constructed as limit of Riemann sums and not just compensated Riemann sums. Davie
shows that under suitable conditions, the usual Euler scheme (without area compensation)
28
converges to the solution of a given rough differential equation. But from there it is easily
deduced that then also the rough path integral is given as limit of Riemann sums. Here we
show that Davies criterion implies our assumption (rie).
Let p (2, 3) and let S = (S, A) be a 1/p-Holder continuous rough path, that is |Ss,t | .
|t s|1/p and |A(s, t)| . |t s|2/p . Write := 1/p and let (1 , 2). Davie assumes
that there exists C > 0 such that the area process A satisfies
`1
X
A(jh, (j + 1)h) C(` k) h2 ,
(22)
j=k
whenever 0 < k < ` are integers and h > 0 such that `h T . Under these conditions,
Theorem 7.1 of [Dav07] implies that for F C with > p and for tnk = kT /n, n, k N, the
Riemann sums
n1
X
F (Stnk )Stnk t,tnk+1 t , t [0, T ],
k=0
converge uniformly to the rough path integral. But it can be easily deduced from (22) that
the area process A is given as limit of non-anticipating Riemann sums along (tn )n . Indeed,
letting h = T /n,
Z t
n1 Z tn t
n1
X
X
k+1
n
n
n
n
n
n
S
dS
S
S
=
S
dS
S
S
s
s
t
t
t,t
t
s
s
t
t
t
t,t
t
k
k
k+1
k
k
k+1
n
0
k=0
tk t
k=0
bt/hc1
n1
X
X
n
n
Akh,(k+1)h + |A(bt/hc, t)|
=
A(tk t, tk+1 t)
k=0
. Cbt/hc h
k=0
2
+ h kAk2 . Cth
+ h2 kAk2 .
Since < 2, the right hand side converges to 0 as n goes to (and thus h goes to 0).
Futhermore, (22) implies the uniformly bounded p/2-variation condition (12):
Z tn
X
Z tn
n
`
`1
j+1
(S S)tn ,tn Stn Stn ,tn
Ss dSs Stnj Stnj ,tnj+1
n Ss dSs Stnk Stnk ,tn` +
k `
k
k `
n
tk
kAk2 |tn`
kAk2 |tn`
tnk |2
tnk |2
j=k
tj
X
`1
+
Atnk ,tnk+1 kAk2 |tn` tnk |2 + C(` k) h2
+
j=k
C|tn`
tnk |2 .
References
[ABPS14] Beatrice Acciaio, Mathias Beiglbock, Friedrich Penkner, and Walter Schachermayer, A model-free version of the fundamental theorem of asset pricing and the
super-replication theorem, Math. Finance (2014+).
[ALP95]
Marco Avellaneda, Arnon Levy, and Antonio Paras, Pricing and hedging derivative
securities in markets with uncertain volatilities, Appl. Math. Finance 2 (1995),
no. 2, 7388.
29
[Ank05]
[CDGR11] Rosanna Coviello, Cristina Di Girolami, and Francesco Russo, On stochastic calculus related to financial assets without semimartingales, Bull. Sci. Math. 135
(2011), no. 6, 733774.
[CL05]
Laure Coutin and Antoine Lejay, Semi-martingales and rough paths theory,
Electron. J. Probab. 10 (2005), no. 23, 761785 (electronic). MR 2164030
(2006i:60042)
[Dav07]
[DH07]
Mark HA Davis and David G Hobson, The range of traded option prices, Math.
Finance 17 (2007), no. 1, 114.
[DM06]
Laurent Denis and Claude Martini, A theoretical framework for the pricing of
contingent claims in the presence of model uncertainty, Ann. Appl. Probab. 16
(2006), no. 2, 827852. MR 2244434 (2007j:60100)
[DOR13]
[DS94]
[DS14]
Yan Dolinsky and H. Mete Soner, Robust Hedging and Martingale Optimal Transport in Continuous Time, Probab. Theory Related Fields 160 (2014), no. 12,
391427.
[FH14]
Peter Friz and Martin Hairer, A Course on Rough Paths: With an Introduction
to Regularity Structures, Springer, 2014.
[Fol81]
Hans F
ollmer, Calcul dIt
o sans probabilites, Seminar on Probability, XV (Univ.
Strasbourg, Strasbourg, 1979/1980) (French), Lecture Notes in Math., vol. 850,
Springer, Berlin, 1981, pp. 143150. MR 622559 (82j:60098)
[Fol99]
Gerald B. Folland, Real analysis, second ed., Pure and Applied Mathematics (New
York), John Wiley & Sons Inc., New York, 1999, Modern techniques and their
applications, A Wiley-Interscience Publication. MR 1681462 (2000c:00001)
[FV10]
[GIP14]
Massimiliano Gubinelli, Peter Imkeller, and Nicolas Perkowski, A Fourier approach to pathwise stochastic integration, arXiv preprint arXiv:1410.4006 (2014).
[Gub04]
Massimiliano Gubinelli, Controlling rough paths, J. Funct. Anal. 216 (2004), no. 1,
86140.
[IP11]
Peter Imkeller and Nicolas Perkowski, The existence of dominating local martingale measures, arXiv preprint arXiv:1111.3885 (2011).
[Kar95]
[Kel14]
[KK07]
Ioannis Karatzas and Constantinos Kardaras, The numeraire portfolio in semimartingale financial models, Finance Stoch. 11 (2007), no. 4, 447493.
[KP11]
[LCL07]
Terry J. Lyons, Michael Caruana, and Thierry Levy, Differential equations driven
by rough paths, Lecture Notes in Mathematics, vol. 1908, Springer, Berlin, 2007.
MR 2314753 (2009c:60156)
[Lej12]
[LQ02]
Terry Lyons and Zhongmin Qian, System control and rough paths, Oxford University Press, 2002.
[Lyo95a]
[Lyo95b]
[Lyo98]
[Nut12]
[Ruf13]
Johannes Ruf, Hedging under arbitrage, Math. Finance 23 (2013), no. 2, 297317.
[RV93]
Francesco Russo and Pierre Vallois, Forward, backward and symmetric stochastic
integration, Probab. Theory Related Fields 97 (1993), no. 3, 403421. MR 1245252
(94j:60113)
31
[Tao11]
Terence Tao, An introduction to measure theory, Graduate Studies in Mathematics, vol. 126, American Mathematical Society, Providence, RI, 2011. MR 2827917
(2012h:28003)
[Vov08]
[Vov11]
[Vov12]
[Wil01]
[You36]
32