Part III Stochastic Calculus and Applications
Based on lectures by R. Bauerschmidt
Notes taken by Dexter Chua
Lent 2018
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
Brownian motion. Existence and sample path properties.
Stochastic calculus for continuous processes. Martingales, local martingales, semi-
martingales, quadratic variation and cross-variation, Itˆo’s isometry, definition of
the stochastic integral, Kunita–Watanabe theorem, and Itˆo’s formula.
Applications to Brownian motion and martingales. L´evy characterization of
Brownian motion, Dubins–Schwartz theorem, martingale representation, Gir-
sanov theorem, conformal invariance of planar Brownian motion, and Dirichlet
problems.
Stoc hastic differential equations. Strong and weak solutions, notions of existence
and uniqueness, Yamada–Watanabe theorem, strong Markov property, and
relation to second order partial differential equations.
Pre-requisites
Knowledge of measure theoretic probability as taught in Part III Advanced Probability
will be assumed, in particular familiarity with discrete-time martingales and Brownian
motion.
Contents
0 Introduction
1 The Lebesgue–Stieltjes integral
2 Semi-martingales
2.1 Finite variation processes
2.2 Local martingale
2.3 Square integrable martingales
2.4 Quadratic variation
2.5 Covariation
2.6 Semi-martingale
3 The stochastic integral
3.1 Simple processes
3.2 Itˆo isometry
3.3 Extension to local martingales
3.4 Extension to semi-martingales
3.5 Itˆo formula
3.6 The L´evy characterization
3.7 Girsanov’s theorem
4 Stochastic differential equations
4.1 Existence and uniqueness of solutions
4.2 Examples of stochastic differential equations
4.3 Representations of solutions to PDEs
0 Introduction
Ordinary differential equations are central in analysis. The simplest class of
equations tend to look like
˙x(t) = F (x(t)).
Stochastic differential equations are differential equations where we make the
function
F
“random”. There are many ways of doing so, and the simplest way
is to write it as
˙x(t) = F (x(t)) + η(t),
where
η
is a random function. For example, when modeling noisy physical
systems, our physical bodies will be subject to random noise. What should we
expect the function
η
to be like? We might expect that for
|t s|
0, the
variables
η
(
t
) and
η
(
s
) are “essentially” independent. If we are interested in
physical systems, then this is a rather reasonable assumption, since random noise
is random!
In practice, we work with the idealization, where we claim that
η
(
t
) and
η
(
s
) are independent for
t 6
=
s
. Such an
η
exists, and is known as white noise.
However, it is not a function, but just a Schwartz distribution.
To understand the simplest case, we set F = 0. We then have the equation
˙x = η.
We can write this in integral form as
x(t) = x(0) +
Z
t
0
η(s) ds.
To make sense of this integral, the function
η
should at least be a signed measure.
Unfortunately, white noise isn’t. This is bad news.
We ignore this issue for a little bit, and proceed as if it made sense. If the
equation held, then for any 0 = t
0
< t
1
< ···, the increments
x(t
i
) x(t
i1
) =
Z
t
i
t
i1
η(s) ds
should be independent, and moreover their variance should scale linearly with
|t
i
t
i1
|. So maybe this x should be a Brownian motion!
Formalizing these ideas will take up a large portion of the course, and the
work isn’t always pleasant. Then why should we be interested in this continuous
problem, as opposed to what we obtain when we discretize time? It turns out
in some sense the continuous problem is easier. When we learn measure theory,
there is a lot of work put into constructing the Lebesgue measure, as opposed
to the sum, which we can just define. However, what we end up is much easier
it’s easier to integrate
1
x
3
than to sum
P
n=1
1
n
3
. Similarly, once we have set
up the machinery of stochastic calculus, we have a powerful tool to do explicit
computations, which is usually harder in the discrete world.
Another reason to study stochastic calculus is that a lot of continuous
time processes can be described as solutions to stochastic differential equations.
Compare this with the fact that functions such as trigonometric and Bessel
functions are described as solutions to ordinary differential equations!
There are two ways to approach stochastic calculus, namely via the Itˆo
integral and the Stratonovich integral. We will mostly focus on the Itˆo integral,
which is more useful for our purposes. In particular, the Itˆo integral tends to
give us martingales, which is useful.
To give a flavour of the construction of the Itˆo integral, we consider a simpler
scenario of the Wiener integral.
Definition
(Gaussian space)
.
Let (Ω
, F, P
) be a probability space. Then a
subspace
S L
2
(Ω
, F, P
) is called a Gaussian space if it is a closed linear
subspace and every X S is a centered Gaussian random variable.
An important construction is
Proposition.
Let
H
be any separable Hilbert space. Then there is a probability
space (Ω
, F, P
) with a Gaussian subspace
S L
2
(Ω
, F, P
) and an isometry
I
:
H S
. In other words, for any
f H
, there is a corresponding random
variable
I
(
f
)
N
(0
,
(
f, f
)
H
). Moreover,
I
(
αf
+
βg
) =
αI
(
f
) +
βI
(
g
) and
(f, g)
H
= E[I(f)I(g)].
Proof.
By separability, we can pick a Hilbert space basis (
e
i
)
i=1
of
H
. Let
(Ω
, F, P
) be any probability space that carries an infinite independent sequence
of standard Gaussian random variables
X
i
N
(0
,
1). Then send
e
i
to
X
i
, extend
by linearity and continuity, and take S to be the image.
In particular, we can take H = L
2
(R
+
).
Definition
(Gaussian white noise)
.
A Gaussian white noise on
R
+
is an isometry
W N
from
L
2
(
R
+
) into some Gaussian space. For
A R
+
, we write
W N
(
A
) =
W N(1
A
).
Proposition.
For A R
+
with |A| < , W N (A) N(0, |A|).
For disjoint
A, B R
+
, the variables
W N
(
A
) and
W N
(
B
) are indepen-
dent.
If A =
S
i=1
A
i
for disjoint sets A
i
R
+
, with |A| < , |A
i
| < , then
W N(A) =
X
i=1
W N(A
i
) in L
2
and a.s.
Proof. Only the last point requires proof. Observe that the partial sum
M
n
=
n
X
i=1
W N(A)
is a martingale, and is bounded in L
2
as well, since
EM
2
n
=
n
X
i=1
EW N(A
i
)
2
=
n
X
i=1
|A
i
| |A|.
So we are done by the martingale convergence theorem. The limit is indeed
W N(A) because 1
A
=
P
n=1
1
A
i
.
The point of the proposition is that
W N
really looks like a random measure
on
R
+
, except it is not. We only have convergence almost surely above, which
means we have convergence on a set of measure 1. However, the set depends on
which
A
and
A
i
we pick. For things to actually work out well, we must have a
fixed set of measure 1 for which convergence holds for all A and A
i
.
But perhaps we can ignore this problem, and try to proceed. We define
B
t
= W N([0, t])
for t 0.
Exercise.
This
B
t
is a standard Brownian motion, except for the continuity
requirement. In other words, for any
t
1
, t
2
, . . . , t
n
, the vector (
B
t
i
)
n
i=1
is jointly
Gaussian with
E[B
s
B
t
] = s t for s, t 0.
Moreover,
B
0
= 0 a.s. and
B
t
B
s
is independent of
σ
(
B
r
:
r s
). Moreover,
B
t
B
s
N(0, t s) for t s.
In fact, by picking a good basis of L
2
(R
+
), we can make B
t
continuous.
We can now try to define some stochastic integral. If
f L
2
(
R
+
) is a step
function,
f =
n
X
i=1
f
i
1
[s
i
,t
i
]
with s
i
< t
i
, then
W N(f) =
n
X
i=1
f
i
(B
t
i
B
s
i
)
This motivates the notation
W N(f) =
Z
f(s) dB
S
.
However, extending this to a function that is not a step function would be
problematic.
1 The Lebesgue–Stieltjes integral
In calculus, we are able to perform integrals more exciting than simply
R
1
0
h
(
x
) d
x
.
In particular, if
h, a
: [0
,
1]
R
are
C
1
functions, we can perform integrals of
the form
Z
1
0
h(x) da(x).
For them, it is easy to make sense of what this means it’s simply
Z
1
0
h(x) da =
Z
1
0
h(x)a
0
(x) dx.
In our world, we wouldn’t expect our functions to be differentiable, so this is
not a great definition. One reasonable strategy to make sense of this is to come
up with a measure that should equal “da”.
An immediate difficulty we encounter is that
a
0
(
x
) need not be positive all the
time. So for example,
R
1
0
1 d
a
could be a negative number, which one wouldn’t
expect for a usual measure! Thus, we are naturally lead to think about signed
measures.
From now on, we always use the Borel
σ
-algebra on [0
, T
] unless otherwise
specified.
Definition
(Signed measure)
.
A signed measure on [0
, T
] is a difference
µ
=
µ
+
µ
of two positive measures on [0
, T
] of disjoint support. The decomposition
µ = µ
+
µ
is called the Hahn decomposition.
In general, given two measures
µ
1
and
µ
2
with not necessarily disjoint
supports, we may still want to talk about µ
1
µ
2
.
Theorem.
For any two finite measures
µ
1
, µ
2
, there is a signed measure
µ
with
µ(A) = µ
1
(A) µ
2
(A).
If
µ
1
and
µ
2
are given by densities
f
1
, f
2
, then we can simply decompose
µ
as (
f
1
f
2
)
+
d
t
+ (
f
1
f
2
)
d
t
, where
+
and
denote the positive and negative
parts respectively. In general, they need not be given by densities with respect to
dx, but they are always given by densities with respect to some other measure.
Proof.
Let
ν
=
µ
1
+
µ
2
. By Radon–Nikodym, there are positive functions
f
1
, f
2
such that µ
i
(dt) = f
i
(t)ν(dt). Then
(µ
1
µ
2
)(dt) = (f
1
f
2
)
+
(t) · ν(dt) + (f
1
f
2
)
(t) · ν(dt).
Definition
(Total variation)
.
The total variation of a signed measure
µ
=
µ
+
µ
is |µ| = µ
+
+ µ
.
We now want to figure out how we can go from a function to a signed measure.
Let’s think about how one would attempt to define
R
1
0
f
(
x
) d
g
as a Riemann
sum. A natural option would be to write something like
Z
t
0
h(s) da(s) = lim
m→∞
n
m
X
i=1
h(t
(m)
i1
)
a(t
(m)
i
) a(t
(m)
i1
)
for any sequence of subdivisions 0 =
t
(m)
0
< ··· < t
(m)
n
m
=
t
of [0
, t
] with
max
i
|t
(m)
i
t
(m)
i1
| 0.
In particular, since we want the integral of
h
= 1 to be well-behaved, the
sum
P
(a(t
(m)
i
) a(t
(m)
i1
)) must be well-behaved. This leads to the notion of
Definition
(Total variation)
.
The total variation of a function
a
: [0
, T
]
R
is
V
a
(t) = |a(0)| + sup
(
n
X
i=1
|a(t
i
) a(t
i1
)| : 0 = t
0
< t
1
< ··· < t
n
= T
)
.
We say a has bounded variation if V
a
(T ) < . In this case, we write a BV .
We include the
|a
(0)
|
term because we want to pretend
a
is defined on all of
R with a(t) = 0 for t < 0.
We also define
Definition
(C`adl`ag)
.
A function
a
: [0
, T
]
R
is c`adl`ag if it is right-continuous
and has left-limits.
The following theorem is then clear:
Theorem. There is a bijection
n
signed measures on [0, T ]
o
c`adl`ag functions of bounded
variation a : [0, T ] R
that sends a signed measure
µ
to
a
(
t
) =
µ
([0
, t
]). To construct the inverse, given
a, we define
a
±
=
1
2
(V
a
± a).
Then a
±
are both positive, and a = a
+
a
. We can then define µ
±
by
µ
±
[0, t] = a
±
(t) a
±
(0)
µ = µ
+
µ
Moreover, V
µ[0,t]
= |µ|[0, t].
Example. Let a : [0, 1] R be given by
a(t) =
(
1 t <
1
2
0 t
1
2
.
This is c`adl`ag, and it’s total variation is
v
0
(1) = 2. The associated signed
measure is
µ = δ
0
δ
1/2
,
and the total variation measure is
|µ| = δ
0
+ δ
1/2
.
We are now ready to define the Lebesgue–Stieltjes integral.
Definition
(Lebesgue–Stieltjes integral)
.
Let
a
: [0
, T
]
R
be c`adl`ag of
bounded variation and let
µ
be the associated signed measure. Then for
h
L
1
([0, T ], |µ|), the Lebesgue–Stieltjes integral is defined by
Z
t
s
h(r) da(r) =
Z
(s,t]
h(r)µ(dr),
where 0 s t T , and
Z
t
s
h(r) |da(r)| =
Z
(s,t]
h(r)|µ|(dr).
We also write
h · a(t) =
Z
t
0
h(r) da(r).
To let T = , we need the following notation:
Definition
(Finite variation)
.
A c`adl`ag function
a
: [0
,
)
R
is of finite
variation if a|
[0,T ]
BV [0, 1] for all T > 0.
Fact. Let a : [0, T ] R be c`adl`ag and BV, and h L
1
([0, T ], |da|), then
Z
T
0
h(s) da(s)
Z
b
a
|h(s)| |da(s)|,
and the function
h · a
: [0
, T
]
R
is c`adl`ag and BV with associated signed
measure h(s) da(s). Moreover, |h(s) da(s)| = |h(s)| |da(s)|.
We can, unsurprisingly, characterize the Lebesgue–Stieltjes integral by a
Riemann sum:
Proposition.
Let
a
be c`adl`ag and BV on [0
, t
], and
h
bounded and left-
continuous. Then
Z
t
0
h(s) da(s) = lim
m→∞
n
m
X
i=1
h(t
(m)
i1
)
a(t
(m)
i
) a(t
(m)
i1
)
Z
t
0
h(s) |da(s)| = lim
m→∞
n
m
X
i=1
h(t
(m)
i1
)
a(t
(m)
i
) a(t
(m)
i1
)
for any sequence of subdivisions 0 =
t
(m)
0
< ··· < t
(m)
n
m
=
t
of [0
, t
] with
max
i
|t
(m)
i
t
(m)
i1
| 0.
Proof. We approximate h by h
m
defined by
h
m
(0) = 0, h
m
(s) = h(t
(m)
i1
) for s (t
(m)
i1
, t
(m)
i
].
Then by left continuity, we have
h(s) = lim
n→∞
h
m
(s)
by left continuity, and moreover
lim
m→∞
n
m
X
i=1
h(t
(m)
i1
)(a(t
(m)
i
) a(t
(m)
i1
)) = lim
m→∞
Z
(0,t]
h
m
(s)µ( ds) =
Z
(0,t]
h(s)µ(ds)
by dominated convergence theorem. The statement about
|
d
a
(
s
)
|
is left as an
exercise.
2 Semi-martingales
The title of the chapter is “semi-martingales”, but we are not going even meet
the definition of a semi-martingale till the end of the chapter. The reason is that
a semi-martingale is essentially defined to be the sum of a (local) martingale and
a finite variation process, and understanding semi-martingales mostly involves
understanding the two parts separately. Thus, for most of the chapter, we
will be studying local martingales (finite variation processes are rather more
boring), and at the end we will put them together to say a word or two about
semi-martingales.
From now on, (Ω
, F,
(
F
t
)
t0
, P
) will be a filtered probability space. Recall
the following definition:
Definition
(C`adl`ag adapted process)
.
A c`adl`ag adapted process is a map
X : Ω × [0, ) R such that
(i) X is c`adl`ag, i.e. X(ω, ·) : [0, ) R is c`adl`ag for all ω Ω.
(ii) X is adapted, i.e. X
t
= X( ·, t) is F
t
-measurable for every t 0.
Notation.
We will write
X G
to denote that a random variable
X
is measur-
able with respect to a σ-algebra G.
2.1 Finite variation processes
The definition of a finite variation function extends immediately to a finite
variation process.
Definition
(Finite variation process)
.
A finite variation process is a c`adl`ag
adapted process
A
such that
A
(
ω, ·
) : [0
,
)
R
has finite variation for all
ω Ω. The total variation process V of a finite variation process A is
V
t
=
Z
T
0
|dA
s
|.
Proposition.
The total variation process
V
of a c`adl`ag adapted process
A
is
also c`adl`ag, finite variation and adapted, and it is also increasing.
Proof.
We only have to check that it is adapted. But that follows directly
from our previous expression of the integral as the limit of a sum. Indeed, let
0 =
t
(m)
0
< t
(m)
1
< ··· < t
n
m
=
t
be a (nested) sequence of subdivisions of [0
, t
]
with max
i
|t
(m)
i
t
(m)
i1
| 0. We have seen
V
t
= lim
m→∞
n
m
X
i=1
|A
t
(m)
i
A
t
(m)
i1
| + |A(0)| F
t
.
Definition
((
H ·A
)
t
)
.
Let
A
be a finite variation process and
H
a process such
that for all ω and t 0,
Z
t
0
H
s
(ω)| |dA
s
(ω)| < .
Then define a process ((H · A)
t
)
t0
by
(H · A)
t
=
Z
t
0
H
s
dA
s
.
For the process H · A to be adapted, we need a condition.
Definition
(Previsible process)
.
A process
H
: Ω
×
[0
,
)
R
is previsible if
it is measurable with respect to the previsible
σ
-algebra
P
generated by the sets
E × (s, t], where E F
s
and s < t. We call the generating set Π.
Very roughly, the idea is that a previsible event is one where whenever it
happens, you know it a finite (though possibly arbitrarily small) before.
Definition
(Simple process)
.
A process
H
: Ω
×
[0
,
)
R
is simple, written
H E, if
H(ω, t) =
n
X
i=1
H
i1
(ω)1
(t
i1
,t
i
]
(t)
for random variables H
i1
F
i1
and 0 = t
0
< ··· < t
n
.
Fact. Simple processes and their limits are previsible.
Fact.
Let
X
be a c`adl`ag adapted process. Then
H
t
=
X
t
defines a left-
continuous process and is previsible.
In particular, continuous processes are previsible.
Proof.
Since
X
is c`adl`ag adapted, it is clear that
H
is left-continuous and
adapted. Since
H
is left-continuous, it is approximated by simple processes.
Indeed, let
H
n
t
=
2
n
X
i=1
H
(i1)2
n
1
((i1)2
n
,i2
n
]
(t) n E.
Then H
n
t
H for all t by left continuity, and previsibility follows.
Exercise. Let H be previsible. Then
H
t
F
t
= σ(F
s
: s < t).
Example. Brownian motion is previsible (since it is continuous).
Example. A Poisson process (N
t
) is not previsible since N
t
6∈ F
t
.
Proposition. Let A be a finite variation process, and H previsible such that
Z
t
0
|H(ω, s)| |dA(ω, s)| < for all (ω, t) × [0, ).
Then H · A is a finite variation process.
Proof.
The finite variation and c`adl`ag parts follow directly from the deterministic
versions. We only have to check that
H · A
is adapted, i.e. (
H · A
)(
·, t
)
F
t
for
all t 0.
First,
H ·A
is adapted if
H
(
ω, s
) = 1
(u,v]
(
s
)1
E
(
ω
) for some
u < v
and
E F
u
,
since
(H · A)(ω, t) = 1
E
(ω)(A(ω, t v) A(ω, t u)) F
t
.
Thus,
H ·A
is adapted for
H
=
1
F
when
F
Π. Clearly, Π is a
π
system, i.e. it
is closed under intersections and non-empty, and by definition it generates the
previsible
σ
-algebra
P
. So to extend the adaptedness of
H · A
to all previsible
H, we use the monotone class theorem.
We let
V = {H : Ω × [0, ) R : H · A is adapted}.
Then
(i) 1 V
(ii) 1
F
V for all F Π.
(iii) V is closed under monotone limits.
So V contains all bounded P-measurable functions.
So the conclusion is that if
A
is a finite variation process, then as long as
reasonable finiteness conditions are satisfied, we can integrate functions against
d
A
. Moreover, this integral was easy to define, and it obeys all expected
properties such as dominated convergence, since ultimately, it is just an integral
in the usual measure-theoretic sense. This crucially depends on the fact that
A
is a finite variation process.
However, in our motivating example, we wanted to take
A
to be Brownian
motion, which is not of finite variation. The work we will do in this chapter
and the next is to come up with a stochastic integral where we let
A
be a
martingale instead. The heuristic idea is that while martingales can vary wildly,
the martingale property implies there will be some large cancellation between
the up and down movements, which leads to the possibility of a well-defined
stochastic integral.
2.2 Local martingale
From now on, we assume that (Ω
, F,
(
F
t
)
t
, P
) satisfies the usual conditions,
namely that
(i) F
0
contains all P-null sets
(ii) (F
t
)
t
is right-continuous, i.e. F
t
= (F
t+
=
T
s>t
F
s
for all t 0.
We recall some of the properties of continuous martingales.
Theorem
(Optional stopping theorem)
.
Let
X
be a c`adl`ag adapted integrable
process. Then the following are equivalent:
(i) X is a martingale, i.e. X
t
L
1
for every t, and
E(X
t
| F
s
) = X
s
for all t > s.
(ii)
The stopped process
X
T
= (
X
T
t
) = (
X
T t
) is a martingale for all stopping
times T .
(iii)
For all stopping times
T, S
with
T
bounded,
X
T
L
1
and
E
(
X
T
| F
S
) =
X
T S
almost surely.
(iv) For all bounded stopping times T , X
T
L
1
and E(X
T
) = E(X
0
).
For X uniformly integrable, (iii) and (iv) hold for all stopping times.
In practice, most of our results will be first proven for bounded martingales,
or perhaps square integrable ones. The point is that the square-integrable
martingales form a Hilbert space, and Hilbert space techniques can help us say
something useful about these martingales. To get something about a general
martingale
M
, we can apply a cutoff
T
n
=
inf{t >
0 :
M
t
n}
, and then
M
T
n
will be a martingale for all
n
. We can then take the limit
n
to recover
something about the martingale itself.
But if we are doing this, we might as well weaken the martingale condition
a bit we only need the
M
T
n
to be martingales. Of course, we aren’t doing
this just for fun. In general, martingales will not always be closed under the
operations we are interested in, but local (or maybe semi-) martingales will be.
In general, we define
Definition
(Local martingale)
.
A c`adl`ag adapted process
X
is a local martingale
if there exists a sequence of stopping times
T
n
such that
T
n
almost surely,
and X
T
n
is a martingale for every n. We say the sequence T
n
reduces X.
Example.
(i)
Every martingale is a local martingale, since by the optional stopping
theorem, we can take T
n
= n.
(ii) Let (B
t
) to be a standard 3d Brownian motion on R
3
. Then
(X
t
)
t1
=
1
|B
t
|
t1
is a local martingale but not a martingale.
To see this, first note that
sup
t1
EX
2
t
< , EX
t
0.
Since
EX
t
0 and
X
t
0, we know
X
cannot be a martingale. However,
we can check that it is a local martingale. Recall that for any f C
2
b
,
M
f
= f(B
t
) f(B
1
)
1
2
Z
t
0
f(B
s
) ds
is a martingale. Moreover,
1
|x|
= 0 for all
x 6
= 0. Thus, if
1
|x|
didn’t have
a singularity at 0, this would have told us
X
t
is a martingale. Thus, we
are safe if we try to bound |B
s
| away from zero.
Let
T
n
= inf
t 1 : |B
t
| <
1
n
,
and pick
f
n
C
2
b
such that
f
n
(
x
) =
1
|x|
for
|x|
1
n
. Then
X
T
t
X
T
n
1
=
M
f
n
tT
n
. So X
T
n
is a martingale.
It remains to show that
T
n
, and this follows from the fact that
EX
t
0.
Proposition.
Let
X
be a local martingale and
X
t
0 for all
t
. Then
X
is a
supermartingale.
Proof. Let (T
n
) be a reducing sequence for X. Then
E(X
t
| F
s
) = E
lim inf
n→∞
X
tT
n
| F
s
lim
n→∞
E(X
tT
n
| F
s
)
= lim inf
T
n
→∞
X
sT
n
= X
s
.
Recall the following result from Advanced Probability:
Proposition. Let X L
1
(Ω, F, P). Then the set
χ = {E(X | G) : G F a sub-σ-algebra}
is uniformly integrable, i.e.
sup
Y χ
E(|Y |1
|Y |
) 0 as λ .
Recall also the following important result about uniformly integrable random
variables:
Theorem
(Vitali theorem)
. X
n
X
in
L
1
iff (
X
n
) is uniformly integrable and
X
n
X in probability.
With these, we can state the following characterization of martingales in
terms of local martingales:
Proposition. The following are equivalent:
(i) X is a martingale.
(ii) X is a local martingale, and for all t 0, the set
χ
t
= {X
T
: T is a stopping time with T t}
is uniformly integrable.
Proof.
(a)
(b): Let
X
be a martingale. Then by the optional stopping theorem,
X
T
=
E
(
X
t
| F
T
) for any bounded stopping time
T t
. So
χ
t
is uniformly
integrable.
(b)
(a): Let
X
be a local martingale with reducing sequence (
T
n
), and
assume that the sets
χ
t
are uniformly integrable for all
t
0. By the
optional stopping theorem, it suffices to show that
E
(
X
T
) =
E
(
X
0
) for any
bounded stopping time T .
So let T be a bounded stopping time, say T t. Then
E(X
0
) = E(X
T
n
0
) = E(X
T
n
T
) = E(X
T T
n
)
for all
n
. Now
T T
n
is a stopping time
t
, so
{X
T T
n
}
is uniformly
integrable by assumption. Moreover,
T
n
T T
almost surely as
n
,
hence
X
T T
n
X
T
in probability. Hence by Vitali, this converges in
L
1
.
So
E(X
T
) = E(X
0
).
Corollary.
If
Z L
1
is such that
|X
t
| Z
for all
t
, then
X
is a martingale. In
particular, every bounded local martingale is a martingale.
The definition of a local martingale does not give us control over what the
reducing sequence
{T
n
}
is. In particular, it is not necessarily true that
X
T
n
will be bounded, which is a helpful property to have. Fortunately, we have the
following proposition:
Proposition. Let X be a continuous local martingale with X
0
= 0. Define
S
n
= inf{t 0 : |X
t
| = n}.
Then
S
n
is a stopping time,
S
n
and
X
S
n
is a bounded martingale. In
particular, (S
n
) reduces X.
Proof. It is clear that S
n
is a stopping time, since (if it is not clear)
{S
n
t} =
\
kN
sup
st
|X
s
| > n
1
k
=
\
kN
[
s<t,sQ
|X
s
| > n
1
k
F
t
.
It is also clear that S
n
, since
sup
st
|X
s
| n S
n
t,
and by continuity and compactness, sup
st
|X
s
| is finite for every (ω, t).
Finally, we show that
X
S
n
is a martingale. By the optional stopping theorem,
X
T
n
S
n
is a martingale, so
X
S
n
is a local martingale. But it is also bounded by
n. So it is a martingale.
An important and useful theorem is the following:
Theorem.
Let
X
be a continuous local martingale with
X
0
= 0. If
X
is also a
finite variation process, then X
t
= 0 for all t.
This would rule out interpreting
R
H
s
d
X
s
as a Lebesgue–Stieltjes integral
for
X
a non-zero continuous local martingale. In particular, we cannot take
X
to be Brownian motion. Instead, we have to develop a new theory of integration
for continuous local martingales, namely the Itˆo integral.
On the other hand, this theorem is very useful. We will later want to define
the stochastic integral with respect to the sum of a continuous local martingale
and a finite variation process, which is the appropriate generality for our theorems
to make good sense. This theorem tells us there is a unique way to decompose a
process as a sum of a finite variation process and a continuous local martingale
(if it can be done). So we can simply define this stochastic integral by using the
Lebesgue–Stieltjes integral on the finite variation part and the Itˆo integral on
the continuous local martingale part.
Proof.
Let
X
be a finite-variation continuous local martingale with
X
0
= 0. Since
X
is finite variation, we can define the total variation process (
V
t
) corresponding
to X, and let
S
n
= inf{t 0 : V
t
n} = inf
t 0 :
Z
1
0
|dX
s
| n
.
Then
S
n
is a stopping time, and
S
n
since
X
is assumed to be finite
variation. Moreover, by optional stopping,
X
S
n
is a local martingale, and is also
bounded, since
X
S
n
t
Z
tS
n
0
|dX
s
| n.
So X
S
n
is in fact a martingale.
We claim its
L
2
-norm vanishes. Let 0 =
t
0
< t
1
< ··· < t
n
=
t
be a
subdivision of [0
, t
]. Using the fact that
X
S
n
is a martingale and has orthogonal
increments, we can write
E((X
S
n
t
)
2
) =
k
X
i=1
E((X
S
n
t
i
X
S
n
t
i1
)
2
).
Observe that
X
S
n
is finite variation, but the right-hand side is summing the
square of the variation, which ought to vanish when we take the limit
max |t
i
t
i1
| 0. Indeed, we can compute
E((X
S
n
t
)
2
) =
k
X
i=1
E((X
S
n
t
i
X
S
n
t
i1
)
2
)
E
max
1ik
|X
S
n
t
i
X
S
n
t
i1
|
k
X
i=1
|X
S
n
t
i
X
S
n
t
i1
|
!
E
max
1ik
|X
S
n
t
i
X
S
n
t
i1
|V
tS
n
E
max
1ik
|X
S
n
t
i
X
S
n
t
i1
|n
.
Of course, the first term is also bounded by the total variation. Moreover, we
can make further subdivisions so that the mesh size tends to zero, and then the
first term vanishes in the limit by continuity. So by dominated convergence, we
must have
E
((
X
S
n
t
)
2
) = 0. So
X
S
n
t
= 0 almost surely for all
n
. So
X
t
= 0 for all
t almost surely.
2.3 Square integrable martingales
As previously discussed, we will want to use Hilbert space machinery to construct
the Itˆo integral. The rough idea is to define the Itˆo integral with respect to a
fixed martingale on simple processes via a (finite) Riemann sum, and then by
calculating appropriate bounds on how this affects the norm, we can extend this
to all processes by continuity, and this requires our space to be Hilbert. The
interesting spaces are defined as follows:
Definition (M
2
). Let
M
2
=
X : Ω × [0, ) R : X is adl´ag martingale with sup
t0
E(X
2
t
) <
.
M
2
c
=
X M
2
: X(ω, ·) is continuous for every ω
We define an inner product on M
2
by
(X, Y )
M
2
= E(X
Y
),
which in aprticular induces a norm
kXk
M
2
=
E(X
2
)
1/2
.
We will prove this is indeed an inner product soon. Here recall that for
X M
2
,
the martingale convergence theorem implies
X
t
X
almost surely and in
L
2
.
Our goal will be to prove that these spaces are indeed Hilbert spaces. First
observe that if
X M
2
, then (
X
2
t
)
t0
is a submartingale by Jensen, so
t 7→ EX
2
t
is increasing, and
EX
2
= sup
t0
EX
2
t
.
All the magic that lets us prove they are Hilbert spaces is Doob’s inequality.
Theorem (Doob’s inequality). Let X M
2
. Then
E
sup
t0
X
2
t
4E(X
2
).
So once we control the limit
X
, we control the whole path. This is why
the definition of the norm makes sense, and in particular we know
kXk
M
2
= 0
implies that X = 0.
Theorem. M
2
is a Hilbert space and M
2
c
is a closed subspace.
Proof.
We need to check that
M
2
is complete. Thus let (
X
n
)
M
2
be a Cauchy
sequence, i.e.
E((X
n
X
m
)
2
) 0 as n, m .
By passing to a subsequence, we may assume that
E((X
n
X
n1
)
2
) 2
n
.
First note that
E
X
n
sup
t0
|X
n
t
X
n1
t
|
!
X
n
E
sup
t0
|X
n
t
X
n1
t
|
2
1/2
(CS)
X
n
2E
|X
n
X
n1
|
2
1/2
(Doob’s)
2
X
n
2
n/2
< .
So
X
n=1
sup
t0
|X
n
t
X
n1
t
| < a.s. ()
So on this event, (
X
n
) is a Cauchy sequence in the space (
D
[0
,
)
, k · k
) of
adl´ag sequences. So there is some X(ω, ·) D[0, ) such that
kX
n
(ω, ·) X(ω, ·)k
0 for almost all ω.
and we set X = 0 outside this almost sure event (). We now claim that
E
sup
t0
|X
n
X|
2
0 as n .
We can just compute
E
sup
t
|X
n
X|
2
= E
lim
m→∞
sup
t
|X
n
X
m
|
2
lim inf
m→∞
E
sup
t
|X
n
X
m
|
2
(Fatou)
lim inf
m→∞
4E(X
n
X
m
)
2
(Doob’s)
and this goes to 0 in the limit n as well.
We finally have to check that
X
is indeed a martingale. We use the triangle
inequality to write
kE(X
t
| F
s
) X
s
k
L
2
kE(X
t
X
n
t
| F
s
)k
L
2
+ kX
n
s
X
s
k
L
2
E(E((X
t
X
n
t
)
2
| F
s
))
1/2
+ kX
n
s
X
s
k
L
2
= kX
t
X
n
t
k
L
2
+ kX
n
s
X
s
k
L
2
2E
sup
t
|X
t
X
n
t
|
2
1/2
0
as
n
. But the left-hand side does not depend on
n
. So it must vanish. So
X M
2
.
We could have done exactly the same with continuous martingales, so the
second part follows.
2.4 Quadratic variation
Physicists are used to dropping all terms above first-order. It turns out that
Brownian motion, and continuous local martingales in general oscillate so wildly
that second order terms become important. We first make the following definition:
Definition
(Uniformly on compact sets in probability)
.
For a sequence of
processes (X
n
) and a process X, we say that X
n
X u.c.p. iff
P
sup
s[0,t]
|X
n
s
X
s
| > ε
!
0 as n for all t > 0, ε > 0.
Theorem.
Let
M
be a continuous local martingale with
M
0
= 0. Then there
exists a unique (up to indistinguishability) continuous adapted increasing process
(
hMi
t
)
t0
such that
hMi
0
= 0 and
M
2
t
hMi
t
is a continuous local martingale.
Moreover,
hMi
t
= lim
n→∞
hMi
(n)
t
, hMi
(n)
t
=
d2
n
te
X
i=1
(M
t2
n
M
(i1)2
n
)
2
,
where the limit u.c.p.
Definition (Quadratic variation). hMi is called the quadratic variation of M .
It is probably more useful to understand
hMi
t
in terms of the explicit formula,
and the fact that
M
2
t
hM i
t
is a continuous local martingale is a convenient
property.
Example.
Let
B
be a standard Brownian motion. Then
B
2
t
t
is a martingale.
Thus, hBi
t
= t.
The proof is long and mechanical, but not hard. All the magic happened
when we used the magical Doob’s inequality to show that
M
2
c
and
M
2
are
Hilbert spaces.
Proof.
To show uniqueness, we use that finite variation and local martingale
are incompatible. Suppose (
A
t
) and (
˜
A
t
) obey the conditions for
hMi
. Then
A
t
˜
A
t
= (
M
2
t
˜
A
t
)
(
M
2
t
A
t
) is a continuous adapted local martingale starting
at 0. Moreover, both
A
t
and
˜
A
t
are increasing, hence have finite variation. So
A
˜
A = 0 almost surely.
To show existence, we need to show that the limit exists and has the right
property. We do this in steps.
Claim. The result holds if M is in fact bounded.
Suppose
|M
(
ω, t
)
| C
for all (
ω, t
). Then
M M
2
c
. Fix
T >
0 deterministic.
Let
X
n
t
=
d2
n
T e
X
i=1
M
(i1)2
n
(M
i2
n
t
M
(i1)2
n
t
).
This is defined so that
hMi
(n)
k2
n
= M
2
k2
n
2X
n
k2
n
.
This reduces the study of hM i
(n)
to that of X
n
k2
n
.
We check that (
X
n
t
) is a Cauchy sequence in
M
2
c
. The fact that it is a
martingale is an immediate computation. To show it is Cauchy, for
n m
, we
calculate
X
n
X
m
=
d2
n
T e
X
i=1
(M
(i1)2
n
M
b(i1)2
mn
c2
m
)(M
i2
n
M
(i1)2
n
).
We now take the expectation of the square to get
E(X
n
X
m
)
2
= E
d2
n
T e
X
i=1
(M
(i1)2
n
M
b(i1)2
mn
c2
m
)
2
(M
i2
n
M
(i1)2
n
)
2
E
sup
|st|≤2
m
|M
t
M
s
|
2
d2
n
T e
X
i=1
(M
i2
n
M
(i1)2
n
)
2
= E
sup
|st|≤2
m
|M
t
M
s
|
2
hMi
(n)
T
!
E
sup
|st|≤2
m
|M
t
M
s
|
4
!
1/2
E
(hMi
(n)
T
)
2
1/2
(Cauchy–Schwarz)
We shall show that the second factor is bounded, while the first factor tends to
zero as
m
. These are both not surprising the first term vanishing in
the limit corresponds to
M
being continuous, and the second term is bounded
since M itself is bounded.
To show that the first term tends to zero, we note that we have
|M
t
M
s
|
4
16C
4
,
and moreover
sup
|st|≤2
m
|M
t
M
s
| 0 as m by uniform continuity.
So we are done by the dominated convergence theorem.
To show the second term is bounded, we do (writing N = d2
n
T e)
E
(hMi
(n)
T
)
2
= E
N
X
i=1
(M
i2
n
M
(i1)2
n
)
2
!
2
=
N
X
i=1
E
(M
i2
n
M
(i1)2
n
)
4
+ 2
N
X
i=1
E
(M
i2
n
M
(i1)2
n
)
2
N
X
k=i+1
(M
k2
n
M
(k1)2
n
)
2
!
We use the martingale property and orthogonal increments the rearrange the
off-diagonal term as
E
(M
i2
n
M
(i1)2
n
)(M
N2
n
M
i2
n
)
2
.
Taking some sups, we get
E
(hMi
(n)
T
)
2
12C
2
E
N
X
i=1
(M
i2
n
M
(i1)2
n
)
2
!
= 12C
2
E
(M
N2
n
M
0
)
2
12C
2
· 4C
2
.
So done.
So we now have X
n
X in M
2
c
for some X M
2
c
. In particular, we have
sup
t
|X
n
t
X
t
|
L
2
0
So we know that
sup
t
|X
n
t
X
t
| 0
almost surely along a subsequence Λ.
Let N be the events on which this convergence fails. We define
A
(T )
t
=
(
M
2
t
2X
t
ω \ N
0 ω N
.
Then
A
(T )
is continuous, adapted since
M
and
X
are, and (
M
2
tT
A
(T )
tT
)
t
is a
martingale since
X
is. Finally,
A
(T )
is increasing since
M
2
t
X
n
t
is increasing
on 2
n
Z
[0
, T
] and the limit is uniform. So this
A
(T )
basically satisfies all the
properties we want hM i
t
to satisfy, except we have the stopping time T .
We next observe that for any
T
1,
A
(T )
tT
=
A
(T +1)
tT
for all
t
almost surely.
This essentially follows from the same uniqueness argument as we had at the
beginning of the proof. Thus, there is a process (hM i
t
)
t0
such that
hMi
t
= A
(T )
t
for all
t
[0
, T
] and
T N
, almost surely. Then this is the desired process. So
we have constructed hMi in the case where M is bounded.
Claim. hMi
(n)
hMi u.c.p.
Recall that
hMi
(n)
t
= M
2
2
n
b2
n
tc
2X
n
2
n
b2
n
tc
.
We also know that
sup
tT
|X
n
t
X
t
| 0
in L
2
, hence also in probability. So we have
|hMi
t
hMi
(n)
t
| sup
tT
|M
2
2
n
b2
n
tc
M
2
t
|
+ sup
tT
|X
n
2
n
b2
n
tc
X
2
n
b2
n
tc
| + sup
tT
|X
2
n
b2
n
tc
X
t
|.
The first and last terms
0 in probability since
M
and
X
are uniformly
continuous on [0
, T
]. The second term converges to zero by our previous assertion.
So we are done.
Claim. The theorem holds for M any continuous local martingale.
We let
T
n
=
inf{t
0 :
|M
t
| n}
. Then (
T
n
) reduces
M
and
M
T
n
is a
bounded martingale. So in particular
M
T
n
is a bounded continuous martingale.
We set
A
n
= hM
T
n
i.
Then (
A
n
t
) and (
A
n+1
tT
n
) are indistinguishable for
t < T
n
by the uniqueness argu-
ment. Thus there is a process
hMi
such that
hMi
tT
n
=
A
n
t
are indistinguishable
for all
n
. Clearly,
hMi
is increasing since the
A
n
are, and
M
2
tT
n
hM i
tT
n
is a
martingale for every n, so M
2
t
hMi
t
is a continuous local martingale.
Claim. hMi
(n)
hMi u.c.p.
We have seen
hM
T
k
i
(n)
hM
T
k
i u.c.p.
for every k. So
P
sup
tT
|hMi
(n)
t
hM
t
i| > ε
P(T
k
< T ) + P
sup
tT
|hM
T
k
i
(n)
t
hM
T
k
i
t
> ε
.
So we can fisrt pick
k
large enough such that the first term is small, then pick
n
large enough so that the second is small.
There are a few easy consequence of this theorem.
Fact.
Let
M
be a continuous local martingale, and let
T
be a stopping time.
Then alsmot surely for all t 0,
hM
T
i
t
= hMi
tT
Proof.
Since
M
2
t
hMi
t
is a continuous local martingle, so is
M
2
tT
hMi
tT
=
(M
T
)
2
t
hMi
tT
. So we are done by uniqueness.
Fact.
Let
M
be a continuous local martingale with
M
0
= 0. Then
M
= 0 iff
hMi = 0.
Proof.
If
M
= 0, then
hMi
= 0. Conversely, if
hMi
= 0, then
M
2
is a continuous
local martingale and positive. Thus EM
2
t
EM
2
0
= 0.
Proposition.
Let
M M
2
c
. Then
M
2
hMi
is a uniformly integrable martin-
gale, and
kM M
0
k
M
2
= (EhMi
)
1/2
.
Proof. We will show that hM i
L
1
. This then implies
|M
2
t
hMi
t
| sup
t0
M
2
t
+ hMi
.
Then the right hand side is in
L
1
. Since
M
2
hMi
is a local martingale, this
implies that it is in fact a uniformly integrable martingale.
To show hM i
L
1
, we let
S
n
= inf{t 0 : hM i
t
n}.
Then S
n
, S
n
is a stopping time and moreover hMi
tS
n
n. So we have
M
2
tS
n
hMi
tS
n
n + sup
t0
M
2
t
,
and the second term is in L
1
. So M
2
tS
n
hMi
tS
n
is a true martingale.
So
EM
2
tS
n
EM
2
0
= EhMi
tS
n
.
Taking the limit
t
, we know
EM
2
tS
n
EM
2
S
n
by dominated convergence.
Since
hMi
tS
n
is increasing, we also have
EhMi
tS
n
EhMi
S
n
by monotone
convergence. We can take n , and by the same justification, we have
EhMi EM
2
EM
2
0
= E(M
M
0
)
2
< .
2.5 Covariation
We know
M
2
c
not only has a norm, but also an inner product. This can also be
reflected in the bracket by the polarization identity, and it is natural to define
Definition
(Covariation)
.
Let
M, N
be two continuous local martingales. Define
the covariation (or simply the bracket) between M and N to be process
hM, Ni
t
=
1
4
(hM + N i
t
hM N i
t
).
Then if in fact M, N M
2
c
, then putting t = gives the inner product.
Proposition.
(i) hM, N i
is the unique (up to indistinguishability) finite variation process
such that M
t
N
t
hM, Ni
t
is a continuous local martingale.
(ii) The mapping (M, N) 7→ hM, Ni is bilinear and symmetric.
(iii)
hM, Ni
t
= lim
n→∞
hM, Ni
(n)
t
u.c.p.
hM, Ni
(n)
t
=
d2
n
te
X
i=1
(M
i2
n
M
(i1)2
n
)(N
i2
n
N
(i1)
2
n
).
(iv) For every stopping time T ,
hM
T
, N
T
i
t
= hM
T
, Ni
t
= hM, Ni
tT
.
(v)
If
M, N M
2
c
, then
M
t
N
t
hM, Ni
t
is a uniformly integrable martingale,
and
hM M
0
, N N
0
i
M
2
= EhM, Ni
.
Example.
Let
B, B
0
be two independent Brownian motions (with respect to
the same filtration). Then hB, B
0
i = 0.
Proof.
Assume
B
0
=
B
0
0
= 0. Then
X
±
=
1
2
(
B ± B
0
) are Brownian motions,
and so hX
±
i = t. So their difference vanishes.
An important result about the covariation is the following Cauchy–Schwarz
like inequality:
Proposition
(Kunita–Watanabe)
.
Let
M, N
be continuous local martingales
and let H, K be two (previsible) processes. Then almost surely
Z
0
|H
s
||K
s
||dhM, Ni
s
|
Z
0
H
2
s
dhMi
s
1/2
Z
0
H
2
s
hNi
s
1/2
.
In fact, this is Cauchy–Schwarz. All we have to do is to take approximations
and take limits and make sure everything works out well.
Proof. For convenience, we write
hM, Ni
t
s
= hM, Ni
t
hM, Ni
s
.
Claim. For all 0 s t, we have
|hM, Ni
t
s
|
p
hM, Mi
t
s
p
hN, Ni
t
s
.
By continuity, we can assume that s, t are dyadic rationals. Then
|hM, Ni
t
s
| = lim
n→∞
2
n
t
X
i=2
n
s+1
(M
i2
n
M
(i1)2
n
)(N
i2
n
N
(i1)2
n
)
lim
n→∞
2
n
t
X
i=2
n
s+1
(M
i2
n
M
(i1)2
n
)
2
1/2
×
2
n
t
X
i=2
n
s+1
(N
i2
n
N
(i1)2
n
)
2
1/2
(Cauchy–Schwarz)
=
hM, Mi
t
s
1/2
hN, Ni
t
s
1/2
,
where all equalities are u.c.p.
Claim. For all 0 s < t, we have
Z
t
s
|dhM, Ni
u
|
p
hM, Mi
t
s
p
hN, Ni
t
s
.
Indeed, for any subdivision s = t
0
< t
1
< ···t
n
= t, we have
n
X
i=1
|hM, Ni
t
i
t
i1
|
n
X
i=1
q
hM, Mi
t
i
t
i1
q
hN, Ni
t
i
t
i1
n
X
i=1
hM, Mi
t
i
t
i1
!
1/2
n
X
i=1
hN, Ni
t
i
t
i1
!
1/2
.
(Cauchy–Schwarz)
Taking the supremum over all subdivisions, the claim follows.
Claim. For all bounded Borel sets B [0, ), we have
Z
B
|dhM, Ni
u
|
s
Z
B
dhMi
u
s
Z
B
dhNi
u
.
We already know this is true if
B
is an interval. If
B
is a finite union of
integrals, then we apply Cauchy–Schwarz. By a monotone class argument, we
can extend to all Borel sets.
Claim. The theorem holds for
H =
k
X
`=1
h
`
1
B
`
, K =
n
X
`=1
k
`
1
B
`
for B
`
[0, ) bounded Borel sets with disjoint support.
We have
Z
|H
s
K
s
| |dhM, Ni
s
|
n
X
`=1
|h
`
k
`
|
Z
B
`
|d
¯
M, Ni
s
|
n
X
`=1
|h
`
k
`
|
Z
B
`
dhMi
s
1/2
Z
B
`
dhNi
s
1/2
n
X
`=1
h
2
`
Z
B
`
dhMi
s
!
1/2
n
X
`=1
k
2
`
Z
B
`
dhNi
s
!
1/2
To finish the proof, approximate general
H
and
K
by step functions and take
the limit.
2.6 Semi-martingale
Definition
(Semi-martingale)
.
A (continuous) adapted process
X
is a (contin-
uous) semi-martingale if
X = X
0
+ M + A,
where
X
0
F
0
,
M
is a continuous local martingale with
M
0
= 0, and
A
is a
continuous finite variation process with A
0
= 0.
This decomposition is unique up to indistinguishables.
Definition
(Quadratic variation)
.
Let
X
=
X
0
+
M
+
A
and
X
0
=
X
0
0
+
M
0
+
A
0
be (continuous) semi-martingales. Set
hXi = hMi, hX, X
0
i = hM, M
0
i.
This definition makes sense, because continuous finite variation processes do
not have quadratic variation.
Exercise. We have
hX, Y i
(n)
t
=
d2
n
te
X
i=1
(X
i2
n
X
(i1)2
n
)(Y
i2
n
Y
(i1)2
n
) hX, Y i u.c.p.
3 The stochastic integral
3.1 Simple processes
We now have all the background required to define the stochastic integral, and
we can start constructing it. As in the case of the Lebesgue integral, we first
define it for simple processes, and then extend to general processes by taking a
limit. Recall that we have
Definition
(Simple process)
.
The space of simple processes
E
consists of func-
tions H : Ω × [0, ) R that can be written as
H
t
(ω) =
n
X
i=1
H
i1
(ω)1
(t
i1
,t
i
]
(t)
for some 0 t
0
t
1
··· t
n
and bounded random variables H
i
F
t
i
.
Definition (H · M ). For M M
2
and H E, we set
Z
t
0
H dM = (H · M )
t
=
n
X
i=1
H
i1
(M
t
i
t
M
t
i1
t
).
If
M
were of finite variation, then this is the same as what we have previously
seen.
Recall that for the Lebesgue integral, extending this definition to general
functions required results like monotone convergence. Here we need some similar
results that put bounds on how large the integral can be. In fact, we get
something better than a bound.
Proposition. If M M
2
c
and H E, then H · M M
2
c
and
kH · M k
2
M
2
= E
Z
0
H
2
s
dhMi
s
. ()
Proof.
We first show that
H · M M
2
c
. By linearity, we only have to check it
for
X
i
t
= H
i1
(M
t
i
t
M
t
i1
t
)
We have to check that
E
(
X
i
t
| F
s
) = 0 for all
t > s
, and the only non-trivial case
is when t > t
i1
.
E(X
i
t
| F
s
) = H
i1
E(M
t
i
t
M
t
i1
t
| F
s
) = 0.
We also check that
kX
i
k
M
2
2kHk
kMk
M
2
.
So it is bounded. So H · M M
2
c
.
To prove (), we note that the X
i
are orthogonal and that
hX
i
i
t
= H
2
i1
(hMi
t
i
t
hMi
t
i1
t
).
So we have
hH ·M, H ·M i =
X
hX
i
, X
i
i =
X
H
2
i1
(hMi
t
i
t
hM i
t
i1
t
) =
Z
t
0
H
2
s
dhMi
s
.
In particular,
kH · M k
2
M
2
= EhH · M i
= E
Z
0
H
2
s
dhMi
s
.
Proposition. Let M M
2
c
and H E. Then
hH · M, N i = H · hM, Ni
for all N M
2
.
In other words, the stochastic integral commutes with the bracket.
Proof. Write H · M =
P
X
i
=
P
H
i1
(M
t
i
t
M
t
i1
t
) as before. Then
hX
i
, Ni
t
= H
i1
hM
t
i
t
M
t
i1
t
, Ni = H
i1
(hM, Ni
t
i
t
hM, Ni
t
i1
t
).
3.2 Itˆo isometry
We now try to extend the above definition to something more general than
simple processes.
Definition
(
L
2
(
M
))
.
Let
M M
2
c
. Define
L
2
(
M
) to be the space of (equiva-
lence classes of) previsible H : Ω × [0, ) R such that
kHk
L
2
(M)
= kHk
M
= E
Z
0
H
2
s
dhMi
s
1/2
< .
For H, K L
2
(M), we set
(H, K)
L
2
(M)
= E
Z
0
H
s
K
s
dhMi
s
.
In fact,
L
2
(
M
) is equal to
L
2
(Ω
×
[0
,
)
, P,
d
P
d
hMi
), where
P
is the
previsible σ-algebra, and in particular is a Hilbert space.
Proposition. Let M M
2
c
. Then E is dense in L
2
(M).
Proof.
Since
L
2
(
M
) is a Hilbert space, it suffices to show that if (
K, H
) = 0 for
all H E, then K = 0.
So assume that (K, H) = 0 for all H E and
X
t
=
Z
t
0
K
s
dhMi
s
,
Then
X
is a well-defined finite variation process, and
X
t
L
1
for all
t
. It
suffices to show that
X
t
= 0 for all
t
, and we shall show that
X
t
is a continuous
martingale.
Let 0
s < t
and
F F
s
bounded. We let
H
=
F
1
(s,t]
E
. By assumption,
we know
0 = (K, H) = E
F
Z
t
s
K
u
dhMi
u
= E(F (X
t
X
S
)).
Since this holds for all F
s
measurable F , we have shown that
E(X
t
| F
s
) = X
s
.
So X is a (continuous) martingale, and we are done.
Theorem. Let M M
2
c
. Then
(i)
The map
H E 7→ H ·M M
2
c
extends uniquely to an isometry
L
2
(
M
)
M
2
c
, called the Itˆo isometry.
(ii) For H L
2
(M), H · M is the unique martingale in M
2
c
such that
hH · M, N i = H · hM, Ni
for all
N M
2
c
, where the integral on the LHS is the stochastic integral
(as above) and the RHS is the finite variation integral.
(iii) If T is a stopping time, then (1
[0,T ]
H) · M = (H · M )
T
= H · M
T
.
Definition
(Stochastic integral)
. H · M
is the stochastic integral of
H
with
respect to M and we also write
(H · M )
t
=
Z
t
0
H
s
dM
s
.
It is important that the integral of martingale is still a martingale. After
proving Itˆo’s formula, we will use this fact to show that a lot of things are in
fact martingales in a rather systematic manner. For example, it will be rather
effortless to show that
B
2
t
t
is a martingale when
B
t
is a standard Brownian
motion.
Proof.
(i)
We have already shown that this map is an isometry when restricted to
E
.
So extend by completeness of M
2
c
and denseness of E.
(ii)
Again the equation to show is known for simple
H
, and we want to show
it is preserved under taking limits. Suppose
H
n
H
in
L
2
(
M
) with
H
n
L
2
(M). Then H
n
· M H · M in M
2
c
. We want to show that
hH · M, N i
= lim
n→∞
hH
n
· M, Ni
in L
1
.
H · hM, N i = lim
n→∞
H
n
· hM, Ni in L
1
.
for all N M
2
c
.
To show the first holds, we use the Kunita–Watanabe inequality to get
E|hH · M H
n
· M, Ni
| E (hH · M H
n
· Mi
)
1/2
(EhNi
)
1/2
,
and the first factor is
kH ·M H
n
·Mk
M
2
0, while the second is finite
since N M
2
c
. The second follows from
E |((H H
n
) · hM, Ni)
| kH H
n
k
L
2
(M)
kNk
M
2
0.
So we know that
hH · M, N i
= (
H · hM, N i
)
. We can then replace
N
by the stopped process N
t
to get hH · M, N i
t
= (H · hM, N i)
t
.
To see uniqueness, suppose
X M
2
c
is another such martingale. Then we
have
hX H · M, N i
= 0 for all
N
. Take
N
=
X H · M
, and then we
are done.
(iii) For N M
2
c
, we have
h(H · M )
T
, Ni
t
= hH · M, N i
tT
= H · hM, N i
tT
= (H1
[0,T ]
· hM, Ni)
t
for every N . So we have shown that
(H · M )
T
= (1
[0,T ]
H · M )
by (ii). To prove the second equality, we have
hH ·M
T
, Ni
t
= H ·hM
T
, Ni
t
= H ·hM, N i
tT
= ((H1
[0,T ]
·hM, N i)
t
.
Note that (ii) can be written as
*
Z
()
0
H
s
dM
s
, N
+
t
=
Z
t
0
H
s
dhM, Ni
s
.
Corollary.
hH · M, K · N i = H · (K · hM, Ni) = (HK) · hM, Ni.
In other words,
*
Z
()
0
H
s
dM
s
,
Z
()
0
K
s
dN
s
+
t
=
Z
t
0
H
s
K
s
dhM, Ni
s
.
Corollary.
Since
H · M
and (
H · M
)(
K · N
)
hH · M, K · Ni
are martingales
starting at 0, we have
E
Z
t
0
H dM
s
= 0
E

Z
t
0
H
s
dM
s
Z
t
0
K
s
dN
s

=
Z
t
0
H
s
K
s
dhM, Ni
s
.
Corollary.
Let
H L
2
(
M
), then
HK L
2
(
M
) iff
K L
2
(
H · M
), in which
case
(KH) · M = K · (H · M).
Proof. We have
E
Z
0
K
2
s
H
2
s
dhM
s
i
= E
Z
0
K
2
s
hH · M i
s
,
so kKk
L
2
(H·M)
= kHKk
L
2
(M)
. For N M
2
c
, we have
h(KH)·M, Ni
t
= (KH ·hM, Ni)
t
= (K ·(H ·hM, Ni))
t
= (K ·hH ·M, Ni)
t
.
3.3 Extension to local martingales
We have now defined the stochastic integral for continuous martingales. We next
go through some formalities to extend this to local martingales, and ultimately
to semi-martingales. We are not doing this just for fun. Rather, when we
later prove results like Itˆo’s formula, even when we put in continuous (local)
martingales, we usually end up with some semi-martingales. So it is useful to be
able to deal with semi-martingales in general.
Definition (L
2
bc
(M)). Let L
2
bc
(M) be the space of previsible H such that
Z
t
0
H
2
s
dhMi
s
< a.s.
for all finite t > 0.
Theorem. Let M be a continuous local martingale.
(i)
For every
H L
2
bc
(
M
), there is a unique continuous local martingale
H ·M
with (H · M )
0
= 0 and
hH · M, N i = H · hM, Ni
for all N, M.
(ii) If T is a stopping time, then
(1
[0,T ]
H) · M = (H · M )
T
= H · M
T
.
(iii)
If
H L
2
loc
(
M
),
K
is previsible, then
K L
2
loc
(
H ·M
) iff
HK L
2
loc
(
M
),
and then
K · (H · M) = (KH) · M.
(iv)
Finally, if
M M
2
c
and
H L
2
(
M
), then the definition is the same as
the previous one.
Proof.
Assume
M
0
= 0, and that
R
t
0
H
2
s
d
hMi
s
<
for all
ω
(by setting
H = 0 when this fails). Set
S
n
= inf
t 0 :
Z
t
0
(1 + H
2
s
) dhMi
s
n
.
These S
n
are stopping times that tend to infinity. Then
hM
S
n
, M
S
n
i
t
= hM, Mi
tS
n
n.
So M
S
n
M
2
c
. Also,
Z
0
H
s
dhM
S
n
i
s
=
Z
S
n
0
H
2
s
dhMi
s
n.
So
H L
2
(
M
S
n
), and we have already defined what
H · M
S
n
is. Now notice
that
H · M
S
n
= (H · M
S
m
)
S
n
for m n.
So it makes sense to define
H · M = lim
n→∞
H · M
S
n
.
This is the unique process such that (
H · M
)
S
n
=
H · M
S
n
. We see that
H · M
is a continuous adapted local martingale with reducing sequence S
n
.
Claim. hH · M, Ni = H · hM, Ni.
Indeed, assume that
N
0
= 0. Set
S
0
n
=
inf{t
0 :
|N
t
| n}
. Set
T
n
=
S
n
S
0
n
. Observe that N
S
0
n
M
2
c
. Then
hH · M, N i
T
n
= hH · M
S
n
, N
S
0
n
i = H · hM
S
n
, N
S
0
n
i = H · hM, Ni
T
n
.
Taking the limit n gives the desired result.
The proofs of the other claims are the same as before, since they only use
the characterizing property hH · M, N i = H · hM, Ni.
3.4 Extension to semi-martingales
Definition
(Locally boounded previsible process)
.
A previsible process
H
is
locally bounded if for all t 0, we have
sup
st
|H
s
| < a.s.
Fact.
(i) Any adapted continuous process is locally bounded.
(ii)
If
H
is locally bounded and
A
is a finite variation process, then for all
t 0, we have
Z
t
0
|H
s
| |dA
s
| < a.s.
Now if
X
=
X
0
+
M
+
A
is a semi-martingale, where
X
0
F
0
,
M
is a
continuous local martingale and
A
is a finite variation process, we want to define
R
H
s
d
X
s
. We already know what it means to define integration with respect to
d
M
s
and d
A
s
, using the Itˆo integral and the finite variation integral respectively,
and X
0
doesn’t change, so we can ignore it.
Definition
(Stochastic integral)
.
Let
X
=
X
0
+
M
+
A
be a continuous semi-
martingale, and
H
a locally bounded previsible process. Then the stochastic
integral H · X is the continuous semi-martingale defined by
H · X = H · M + H · A,
and we write
(H · X)
t
=
Z
T
0
H
s
dX
s
.
Proposition.
(i) (H, X) 7→ H · X is bilinear.
(ii) H · (K · X) = (HK) · X if H and K are locally bounded.
(iii) (H · X)
T
= H1
[0,T ]
· X = H · X
T
for every stopping time T .
(iv)
If
X
is a continuous local martingale (resp. a finite variation process), then
so is H · X.
(v)
If
H
=
P
n
i=1
H
i1
1
(t
i1
,t
i
]
and
H
i1
F
t
i1
(not necessarily bounded),
then
(H · X)
t
=
n
X
i=1
H
i1
(X
t
i
t
X
t
i1
t
).
Proof.
(i) to (iv) follow from analogous properties for
H · M
and
H · A
. The
last part is also true by definition if the
H
i
are uniformly bounded. If
H
i
is not
bounded, then the finite variation part is still fine, since for each fixed
ω
Ω,
H
i
(ω) is a fixed number. For the martingale part, set
T
n
= inf{t 0 : |H
t
| n}.
Then T
n
are stopping times, T
n
, and H1
[0,T
n
]
E. Thus
(H · M )
tT
n
=
n
X
i=1
H
i1
T
[0,T
n
]
(X
t
i
t
X
t
i1
t
).
Then take the limit n .
Before we get to Itˆo’s formula, we need a few more useful properties:
Proposition
(Stochastic dominated convergence theorem)
.
Let
X
be a contin-
uous semi-martingale. Let
H, H
s
be previsible and locally bounded, and let
K
be previsible and non-negative. Let t > 0. Suppose
(i) H
n
s
H
s
as n for all s [0, t].
(ii) |H
n
s
| K
s
for all s [0, t] and n N.
(iii)
R
t
0
K
2
s
d
hMi
s
<
and
R
t
0
K
s
|
d
A
s
| <
(note that both conditions are
okay if K is locally bounded).
Then
Z
t
0
H
n
s
dX
s
Z
t
0
H
s
dX
s
in probability.
Proof.
For the finite variation part, the convergence follows from the usual
dominated convergence theorem. For the martingale part, we set
T
m
= inf
t 0 :
Z
t
0
K
2
s
dhMi
s
m
.
So we have
E
Z
T
m
t
0
H
n
s
dM
s
Z
T
n
t
0
H
s
dM
s
!
2
E
Z
T
n
t
0
(H
n
s
H
s
)
2
dhMi
s
!
0.
using the usual dominated convergence theorem, since
R
T
n
t
0
K
2
s
dhMi
s
m.
Since
T
n
t
=
t
eventually as
n
almost surely, hence in probability, we
are done.
Proposition.
Let
X
be a continuous semi-martingale, and let
H
be an adapted
bounded left-continuous process. Then for every subdivision 0
< t
(m)
0
< t
(m)
1
<
··· < t
(m)
n
m
of [0, t] with max
i
|t
(m)
i
t
(m)
i1
| 0, then
Z
t
0
H
s
dX
s
= lim
m→∞
n
m
X
i=1
H
t
(m)
i1
(X
t
(m)
i
X
t
(m)
i1
)
in probability.
Proof.
We have already proved this for the Lebesgue–Stieltjes integral, and all
we used was dominated convergence. So the same proof works using stochastic
dominated convergence theorem.
3.5 Itˆo formula
We now prove the equivalent of the integration by parts and the chain rule, i.e.
Itˆo’s formula. Compared to the world of usual integrals, the difference is that
the quadratic variation, i.e. “second order terms” will crop up quite a lot, since
they are no longer negligible.
Theorem
(Integration by parts)
.
Let
X, Y
be a continuous semi-martingale.
Then almost surely,
X
t
Y
t
X
0
Y
0
=
Z
t
0
X
s
dY
s
+
Z
t
0
Y
s
dX
s
+ hX, Y i
t
The last term is called the Itˆo correction.
Note that if
X, Y
are martingales, then the first two terms on the right are
martingales, but the last is not. So we are forced to think about semi-martingales.
Observe that in the case of finite variation integrals, we don’t have the
correction.
Proof. We have
X
t
Y
t
X
s
Y
s
= X
s
(Y
t
Y
s
) + (X
t
X
s
)Y
s
+ (X
t
X
s
)(Y
t
Y
s
).
When doing usual calculus, we can drop the last term, because it is second order.
However, the quadratic variation of martingales is in general non-zero, and so
we must keep track of this. We have
X
k2
n
Y
k2
n
X
0
Y
0
=
k
X
i=1
(X
i2
n
Y
i2
n
X
(i1)2
n
Y
(i1)2
n
)
=
n
X
i=1
X
(i1)2
n
(Y
i2
n
Y
(i1)2
n
)
+ Y
(i1)2
n
(X
i2
n
X
(i1)2
n
)
+ (X
i2
n
X
(i1)
2
n
)(Y
i2
n
Y
(i1)2
n
)
Taking the limit
n
with
k
2
n
fixed, we see that the formula holds for
t
a
dyadic rational. Then by continuity, it holds for all t.
The really useful formula is the following:
Theorem
(Itˆo’s formula)
.
Let
X
1
, . . . , X
p
be continuous semi-martingales, and
let
f
:
R
p
R
be
C
2
. Then, writing
X
= (
X
1
, . . . , X
p
), we have, almost surely,
f(X
t
) = f(X
0
) +
p
X
i=1
Z
t
0
f
x
i
(X
s
) dX
i
s
+
1
2
p
X
i,j=1
Z
t
0
2
f
x
i
x
j
(X
s
) dhX
i
, X
j
i
s
.
In particular, f (X) is a semi-martingale.
The proof is long but not hard. We first do it for polynomials by explicit
computation, and then use Weierstrass approximation to extend it to more
general functions.
Proof.
Claim. Itˆo’s formula holds when f is a polynomial.
It clearly does when
f
is a constant! We then proceed by induction. Suppose
Itˆo’s formula holds for some f . Then we apply integration by parts to
g(x) = x
k
f(x).
where x
k
denotes the kth component of x. Then we have
g(X
t
) = g(X
0
) +
Z
t
0
X
k
s
df(X
s
) +
Z
t
0
f(X
s
) dX
k
s
+ hX
k
, f(X)i
t
We now apply Itˆo’s formula for f to write
Z
t
0
X
k
s
df(X
s
) =
p
X
i=1
Z
t
0
X
k
s
f
x
i
(X
s
) dX
i
s
+
1
2
p
X
i,j=1
Z
t
0
X
k
s
2
f
x
i
x
j
(X
s
) dhX
i
, X
j
i
s
.
We also have
hX
k
, f(X)i
t
=
p
X
i=1
Z
t
0
f
x
i
(X
s
) dhX
k
, X
i
i
s
.
So we have
g(X
t
) = g(X
0
) +
p
X
i=1
Z
t
0
g
x
i
(X
s
) dX
i
s
+
1
2
p
X
i,j=1
Z
t
0
2
g
x
i
x
j
(X
s
) dhX
i
, X
j
i
s
.
So by induction, Itˆo’s formula holds for all polynomials.
Claim.
Itˆo’s formula holds for all
f C
2
if
|X
t
(
ω
)
| n
and
R
t
0
|
d
A
s
| n
for
all (t, ω).
By the Weierstrass approximation theorem, there are polynomials
p
k
such
that
sup
|x|≤k
|f(x) p
k
(x)| + max
i
f
x
i
p
x
i
+ max
i,j
2
f
x
i
x
j
p
k
x
i
x
j
1
k
.
By taking limits, in probability, we have
f(X
t
) f(X
0
) = lim
k→∞
(p
k
(X
t
) p
k
(X
0
))
Z
t
0
f
x
i
(X
s
) dX
i
s
= lim
k→∞
p
k
x
i
(X
s
) dX
i
s
by stochastic dominated convergence theorem, and by the regular dominated
convergence, we have
Z
t
0
f
x
i
x
j
dhX
i
, X
j
i
s
= lim
k→∞
Z
t
0
2
p
k
x
i
x
j
dhX
i
, X
j
i.
Claim. Itˆo’s formula holds for all X.
Let
T
n
= inf
t 0 : |X
t
| n or
Z
t
0
|dA
s
| n
Then by the previous claim, we have
f(X
T
n
t
) = f(X
0
) +
p
X
i=1
Z
t
0
f
x
i
(X
T
n
s
) d(X
i
)
T
n
s
+
1
2
X
i,j
Z
t
0
2
f
x
i
x
j
(X
T
n
s
) dh(X
i
)
T
n
, (X
j
)
T
n
i
s
= f(X
0
) +
p
X
i=1
Z
tT
n
0
f
x
i
(X
s
) d(X
i
)
s
+
1
2
X
i,j
Z
tT
n
0
2
f
x
i
x
j
(X
s
) dh(X
i
), (X
j
)i
s
.
Then take T
n
.
Example.
Let
B
be a standard Brownian motion,
B
0
= 0 and
f
(
x
) =
x
2
. Then
B
2
t
= 2
Z
t
0
B
S
dB
s
+ t.
In other words,
B
2
t
t = 2
Z
t
0
B
s
dB
s
.
In particular, this is a continuous local martingale.
Example.
Let
B
= (
B
1
, . . . , B
d
) be a
d
-dimensional Brownian motion. Then
we apply Itˆo’s formula to the semi-martingale
X
= (
t, B
1
, . . . , B
d
). Then we
find that
f(t, B
t
) f(0, B
0
)
Z
t
0
s
+
1
2
f(s, B
s
) ds =
d
X
i=1
Z
t
0
x
i
f(s, B
s
) dB
i
s
is a continuous local martingale.
There are some syntactic tricks that make stochastic integrals easier to
manipulate, namely by working in differential form. We can state Itˆo’s formula
in differential form
df(X
t
) =
p
X
i=1
f
x
i
dX
i
+
1
2
p
X
i,j=1
2
f
x
i
x
j
dhX
i
, X
j
i,
which we can think of as the chain rule. For example, in the case case of Brownian
motion, we have
df(B
t
) = f
0
(B
t
) dB
t
+
1
2
f
00
(B
t
) dt.
Formally, one expands
f
using that that “(d
t
)
2
= 0” but “(d
B
)
2
= d
t
”. The
following formal rules hold:
Z
t
Z
0
=
Z
t
0
H
s
dX
s
dZ
t
= H
t
dX
t
Z
t
= hX, Y i
t
=
Z
t
0
dhX, Y i
t
dZ
t
= dX
t
dY
t
.
Then we have rules such as
H
t
(K
t
dX
t
) = (H
t
K
t
) dX
t
H
t
(dX
t
dY
t
) = (H
t
dX
t
) dY
t
d(X
t
Y
t
) = X
t
dY
t
+ Y
t
dX
t
+ dX
t
dY
t
df(X
t
) = f
0
(X
t
) dX
t
+
1
2
f
00
(X
t
) dX
t
dX
t
.
3.6 The evy characterization
A more major application of the stochastic integral is the following convenient
characterization of Brownian motion:
Theorem
(L´evy’s characterization of Brownian motion)
.
Let (
X
1
, . . . , X
d
) be
continuous local martingales. Suppose that
X
0
= 0 and that
hX
i
, X
j
i
t
=
δ
ij
t
for all
i, j
= 1
, . . . , d
and
t
0. Then (
X
1
, . . . , X
d
) is a standard
d
-dimensional
Brownian motion.
This might seem like a rather artificial condition, but it turns out to be quite
useful in practice (though less so in this course). The point is that we know that
hH ·M i
t
=
H
2
t
·hMi
t
, and in particular if we are integrating things with respect
to Brownian motions of some sort, we know
hB
t
i
t
=
t
, and so we are left with
some explicit, familiar integral to do.
Proof.
Let 0
s < t
. It suffices to check that
X
t
X
s
is independent of
F
s
and
X
t
X
s
N(0, (t s)I).
Claim. E(e
·(X
t
X
s
)
| F
s
) = e
1
2
|θ|
2
(ts)
for all θ R
d
and s < t.
This is sufficient, since the right-hand side is independent of
F
s
, hence so is
the left-hand side, and the Fourier transform characterizes the distribution.
To check this, for θ R
d
, we define
Y
t
= θ · X
t
=
d
X
i=1
θ
i
X
i
t
.
Then Y is a continuous local martingale, and we have
hY i
t
= hY, Y i
t
=
d
X
i,j=1
θ
j
θ
k
hX
j
, X
k
i
t
= |θ|
2
t.
by assumption. Let
Z
t
= e
iY
t
+
1
2
hY i
t
= e
·X
t
+
1
2
|θ|
2
t
.
By Itˆo’s formula, with X = iY +
1
2
hY i
t
and f (x) = e
x
, we get
dZ
t
= Z
t
idY
t
1
2
dhY i
t
+
1
2
dhY i
t
= iZ
t
dY
t
.
So this implies
Z
is a continuous local martingale. Moreover, since
Z
is bounded
on bounded intervals of
t
, we know
Z
is in fact a martingale, and
Z
0
= 1. Then
by definition of a martingale, we have
E(Z
t
| F
s
) = Z
s
,
And unwrapping the definition of Z
t
shows that the result follows.
In general, the quadratic variation of a process doesn’t have to be linear in
t
.
It turns out if the quadratic variation increases to infinity, then the martingale
is still a Brownian motion up to reparametrization.
Theorem
(Dubins–Schwarz)
.
Let
M
be a continuous local martingale with
M
0
= 0 and hM i
= . Let
T
s
= inf{t 0 : hM i
t
> s},
the right-continuous inverse of
hMi
t
. Let
B
s
=
M
T
s
and
G
s
=
F
T
s
. Then
T
s
is a
(F
t
) stopping time, hM i
T
s
= s for all s 0, B is a (G
s
)-Brownian motion, and
M
t
= B
hMi
t
.
Proof.
Since
hMi
is continuous and adapted, and
hMi
=
, we know
T
s
is a
stopping time and T
s
< for all s 0.
Claim. (G
s
) is a filtration obeying the usual conditions, and G
= F
Indeed, if A G
s
and s < t, then
A {T
t
u} = A {T
s
u} {T
t
u} F
u
,
using that
A {T
s
u} F
u
since
A G
s
. Then right-continuity follows from
that of (F
t
) and the right-continuity of s 7→ T
s
.
Claim. B is adapted to (G
s
).
In general, if
X
is adl´ag and
T
is a stopping time, then
X
T
1
{T <∞}
F
T
.
Apply this is with X = M , T = T
s
and F
T
= G
s
. Thus B
s
G
s
.
Claim. B is continuous.
Here this is actually something to verify, because
s 7→ T
s
is only right contin-
uous, not necessarily continuous. Thus, we only know
B
s
is right continuous,
and we have to check it is left continuous.
Now B is left-continuous at s iff B
s
= B
s
, iff M
T
s
= M
T
s
. Now we have
T
s
= inf{t 0 : hM i
t
s}.
If
T
s
=
T
s
, then there is nothing to show. Thus, we may assume
T
s
> T
s
.
Then we have
hMi
T
s
=
hMi
T
s
. Since
hMi
t
is increasing, it means
hMi
T
s
is
constant in [T
s
, T
s
]. We will later prove that
Lemma. M is constant on [a, b] iff hM i being constant on [a, b].
So we know that if T
s
> T
s
, then M
T
s
= M
T
s
. So B is left continuous.
We then have to show that B is a martingale.
Claim. (M
2
hMi)
T
s
is a uniformly integrable martingale.
To see this, observe that
hM
T
s
i
=
hMi
T
s
=
s
, and so
M
T
s
is bounded. So
(M
2
hMi)
T
s
is a uniformly integrable martingale.
We now apply the optional stopping theorem, which tells us
E(B
s
| G
r
) = E(M
T
s
| G
s
) = M
T
t
= B
t
.
So B
t
is a martingale. Moreover,
E(B
2
s
s | G
r
) = E((M
2
hMi)
T
s
| F
T
r
) = M
2
T
r
hMi
T
r
= B
2
r
r.
So
B
2
t
t
is a martingale, so by the characterizing property of the quadratic
variation,
hBi
t
=
t
. So by evy’s criterion, this is a Brownian motion in one
dimension.
The theorem is only true for martingales in one dimension. In two dimen-
sions, this need not be true, because the time change needed for the horizontal
and vertical may not agree. However, in the example sheet, we see that the
holomorphic image of a Brownian motion is still a Brownian motion up to a
time change.
Lemma. M is constant on [a, b] iff hM i being constant on [a, b].
Proof.
It is clear that if
M
is constant, then so is
hMi
. To prove the converse,
by continuity, it suffices to prove that for any fixed a < b,
{M
t
= M
a
for all t [a, b]} {hM i
b
= hMi
a
} almost surely.
We set N
t
= M
t
M
t
a. Then hN i
t
= hMi
t
hMi
ta
. Define
T
ε
= inf{t 0 : hN i
t
ε}.
Then since N
2
hNi is a local martingale, we know that
E(N
2
tT
ε
) = E(hNi
tT
ε
) ε.
Now observe that on the event
{hMi
b
=
hMi
a
}
, we have
hNi
b
= 0. So for
t [a, b], we have
E(1
{hMi
b
=hMi
a
}
N
2
t
) = E(1
{hMi
b
=hMi
a
N
2
tT
ε
) = E(hNi
tT
ε
) = 0.
3.7 Girsanov’s theorem
Girsanov’s theorem tells us what happens to our (semi)-martingales when we
change the measure of our space. We first look at a simple example when we
perform a shift.
Example.
Let
X N
(0
, C
) be an
n
-dimensional centered Gaussian with
positive definite covariance
C
= (
C
ij
)
n
i,j=1
. Put
M
=
C
1
. Then for any
function f , we have
Ef(X) =
det
M
2π
1/2
Z
R
n
f(x)e
1
2
(x,Mx)
dx.
Now fix an a R
n
. The distribution of X + a then satisfies
Ef(X + a) =
det
M
2π
1/2
Z
R
n
f(x)e
1
2
(xa,M(xa))
dx = E[Zf (X)],
where
Z = Z(x) = e
1
2
(a,Ma)+(x,Ma)
.
Thus, if P denotes the distribution of X, then the measure Q with
dQ
dP
= Z
is that of N (a, C) vector.
Example.
We can extend the above example to Brownian motion. Let
B
be a
Brownian motion with
B
0
= 0, and
h
: [0
,
)
R
a deterministic function. We
then want to understand the distribution of B
t
+ h.
Fix a finite sequence of times 0 =
t
0
< t
1
< ··· < t
n
. Then we know that
(
B
t
i
)
n
i=1
is a centered Gaussian random variable. Thus, if
f
(
B
) =
f
(
B
t
1
, . . . , B
t
n
)
is a function, then
E(f(B)) = c ·
Z
R
n
f(x)e
1
2
P
n
i=1
(x
i
x
i1
)
2
t
i
t
i1
dx
1
···dx
n
.
Thus, after a shift, we get
E(f(B + h)) = E(Zf (B)),
Z = exp
1
2
n
X
i=1
(h
t
i
h
t
i1
)
2
t
i
t
i1
+
n
X
i=1
(h
t
i
h
t
i1
)(B
t
i
B
t
i1
)
t
i
t
i1
!
.
In general, we are interested in what happens when we change the measure
by an exponential:
Definition
(Stochastic exponential)
.
Let
M
be a continuous local martingale.
Then the stochastic exponential (or Dol´eans–Dade exponential) of M is
E(M )
t
= e
M
t
1
2
hMi
t
The point of introducing that quadratic variation term is
Proposition.
Let
M
be a continuous local martingale with
M
0
= 0. Then
E(M ) = Z satisfies
dZ
t
= Z
t
dM,
i.e.
Z
t
= 1 +
Z
t
0
Z
s
dM
s
.
In particular,
E
(
M
) is a continuous local martingale. Moreover, if
hMi
is
uniformly bounded, then E(M ) is a uniformly integrable martingale.
There is a more general condition for the final property, namely Novikov’s
condition, but we will not go into that.
Proof. By Itˆo’s formula with X = M
1
2
hMi, we have
dZ
t
= Z
t
d
M
t
1
2
dhMi
t
+
1
2
Z
t
dhMi
t
= Z
t
dM
t
.
Since
M
is a continuous local martingale, so is
R
Z
s
d
M
s
. So
Z
is a continuous
local martingale.
Now suppose hM i
b < . Then
P
sup
t0
M
t
a
= P
sup
t0
M
t
a, hM i
b
e
a
2
/2b
,
where the final equality is an exercise on the third example sheet, which is true
for general continuous local martingales. So we get
E
exp
sup
t
M
t

=
Z
0
P(exp(sup M
t
) λ) dλ
=
Z
0
P(sup M
t
log λ) dλ
1 +
Z
1
e
(log λ)
2
/2b
dλ < .
Since hM i 0, we know that
sup
t0
E(M )
t
exp (sup M
t
) ,
So E(M) is a uniformly integrable martingale.
Theorem
(Girsanov’s theorem)
.
Let
M
be a continuous local martingale with
M
0
= 0. Suppose that
E
(
M
) is a uniformly integrable martingale. Define a new
probability measure
dQ
dP
= E(M )
Let
X
be a continuous local martingale with respect to
P
. Then
X hX, M i
is
a continuous local martingale with respect to Q.
Proof. Define the stopping time
T
n
= inf{t 0 : |X
t
hX, Mi
t
| n},
and
P
(
T
n
) = 1 by continuity. Since
Q
is absolutely continuous with
respect to
P
, we know that
Q
(
T
n
) = 1. Thus it suffices to show that
X
T
n
hX
T
n
, Mi is a continuous martingale for any n. Let
Y = X
T
n
hX
T
n
, Mi, Z = E(M ).
Claim. ZY is a continuous local martingale with respect to P.
We use the product rule to compute
d(ZY ) = Y
t
dZ
t
+ Z
t
dY
t
+ dhY, Zi
t
= Y Z
t
dM
t
+ Z
t
(dX
T
n
dhX
T
n
, Mi
t
) + Z
t
dhM, X
T
n
i
= Y Z
t
dM
t
+ Z
t
dX
T
n
So we see that
ZY
is a stochastic integral with respect to a continuous local
martingale. Thus ZY is a continuous local martingale.
Claim. ZY is uniformly integrable.
Since
Z
is a uniformly integrable martingale,
{Z
T
:
T is a stopping time}
is
uniformly integrable. Since
Y
is bounded,
{Z
T
Y
T
:
T is a stopping time}
is also
uniformly integrable. So ZY is a true martingale (with respect to P).
Claim. Y is a martingale with respect to Q.
We have
E
Q
(Y
t
Y
s
| F
s
) = E
P
(Z
Y
t
Z
Y
s
| F
s
)
= E
P
(Z
t
Y
t
Z
s
Y
s
| F
s
) = 0.
Note that the quadratic variation does not change since
hX hX, M ii = hXi
t
= lim
n→∞
b2
n
tc
X
i=1
(X
i2
n
X
(i1)2
n
)
2
a.s.
along a subsequence.
4 Stochastic differential equations
4.1 Existence and uniqueness of solutions
After all this work, we can return to the problem we described in the introduction.
We wanted to make sense of equations of the form
˙x(t) = F (x(t)) + η(t),
where
η
(
t
) is Gaussian white noise. We can now interpret this equation as saying
dX
t
= F (X
t
) dt + dB
t
,
or equivalently, in integral form,
X
t
X
0
=
Z
T
0
F (X
s
) ds + B
t
.
In general, we can make the following definition:
Definition
(Stochastic differential equation)
.
Let
d, m N
,
b
:
R
+
× R
d
R
d
,
σ
:
R
+
× R
d
R
d×m
be locally bounded (and measurable). A solution to the
stochastic differential equation E(σ, b) given by
dX
t
= b(t, X
t
) dt + σ(t, X
t
) dB
t
consists of
(i) a filtered probability space (Ω, F, (F
t
), P) obeying the usual conditions;
(ii) an m-dimensional Brownian motion B with B
0
= 0; and
(iii) an (F
t
)-adapted continuous process X with values in R
d
such that
X
t
= X
0
+
Z
t
0
σ(s, X
s
) dB
s
+
Z
t
0
b(s, X
s
) ds.
If
X
0
=
x R
d
, then we say
X
is a (weak) solution to
E
x
(
σ, b
). It is a strong
solution if it is adapted with respect to the canonical filtration of B.
Our goal is to prove existence and uniqueness of solutions to a general class
of SDEs. We already know what it means for solutions to be unique, and in
general there can be multiple notions of uniqueness:
Definition
(Uniqueness of solutions)
.
For the stochastic differential equation
E(σ, b), we say there is
uniqueness in law if for every
x R
d
, all solutions to
E
x
(
σ, b
) have the
same distribution.
pathwise uniqueness if when (Ω
, F,
(
F
t
)
, P
) and
B
are fixed, any two
solutions X, X
0
with X
0
= X
0
0
are indistinguishable.
These two notions are not equivalent, as the following example shows:
Example (Tanaka). Consider the stochastic differential equation
dX
t
= sgn(X
t
) dB
t
, X
0
= x,
where
sgn(x) =
(
+1 x > 0
1 x 0
.
This has a weak solution which is unique in law, but pathwise uniqueness fails.
To see the existence of solutions, let
X
be a one-dimensional Brownian motion
with X
0
= x, and set
B
t
=
Z
t
0
sgn(X
s
) dX
s
,
which is well-defined because
sgn
(
X
s
) is previsible and left-continuous. Then we
have
x +
Z
t
0
sgn(X
s
) dB
s
= x +
Z
t
0
sgn(X
s
)
2
dX
s
= x + X
t
X
0
= X
t
.
So it remains to show that
B
is a Brownian motion. We already know that
B
is
a continuous local martingale, so by evy’s characterization, it suffices to show
its quadratic variation is t. We simply compute
hB, Bi
t
=
Z
t
0
dhX
s
, X
s
i = t.
So there is weak existence. The same argument shows that any solution is a
Brownian motion, so we have uniqueness in law.
Finally, observe that if
x
= 0 and
X
is a solution, then
X
is also a solution
with the same Brownian motion. Indeed,
X
t
=
Z
t
0
sgn(X
s
) dB
s
=
Z
t
0
sgn(X
s
) dB
s
+ 2
Z
t
0
1
X
s
=0
dB
s
,
where the second term vanishes, since it is a continuous local martingale with
quadratic variation
R
t
0
1
X
s
=0
ds = 0. So pathwise uniqueness does not hold.
In the other direction, however, it turns out pathwise uniqueness implies
uniqueness in law.
Theorem
(Yamada–Watanabe)
.
Assume weak existence and pathwise unique-
ness holds. Then
(i) Uniqueness in law holds.
(ii)
For every (Ω
, F,
(
F
t
)
, P
) and
B
and any
x R
d
, there is a unique strong
solution to E
x
(a, b).
We will not prove this, since we will not actually need it.
The key, important theorem we are now heading for is the existence and
uniqueness of solutions to SDEs, assuming reasonable conditions. As in the case
of ODEs, we need the following Lipschitz conditions:
Definition
(Lipschitz coefficients)
.
The coefficients
b
:
R
+
× R
d
R
d
,
σ
:
R
+
× R
d
R
d×m
are Lipschitz in
x
if there exists a constant
K >
0 such that
for all t 0 and x, y R
d
, we have
|b(t, x) b(t, y)| K|x y|
|σ(t, x) σ(t, y)| |x y|
Theorem.
Assume
b, σ
are Lipschitz in
x
. Then there is pathwise uniqueness
for the
E
(
σ, b
) and for every (Ω
, F,
(
F
t
)
, P
) satisfying the usual conditions and
every (
F
t
)-Brownian motion
B
, for every
x R
d
, there exists a unique strong
solution to E
x
(σ, b).
Proof. To simplify notation, we assume m = d = 1.
We first prove pathwise uniqueness. Suppose
X, X
0
are two solutions with
X
0
=
X
0
0
. We will show that
E
[(
X
t
X
0
t
)
2
] = 0. We will actually put some
bounds to control our variables. Define the stopping time
S = inf{t 0 : |X
t
| n or |X
0
t
| n}.
By continuity,
S
as
n
. We also fix a deterministic time
T >
0. Then
whenever t [0, T ], we can bound, using the identity (a + b)
2
2a
2
+ 2b
2
,
E((X
tS
X
0
tS
)
2
) 2E
Z
tS
0
(σ(s, X
s
) σ(s, X
0
s
)) dB
s
!
2
+ 2E
Z
tS
0
(b(s, X
s
) b(s, X
0
s
)) ds
!
2
.
We can apply the Lipschitz bound to the second term immediately, while we can
simplify the first term using the (corollary of the) Itˆo isometry
E
Z
tS
0
(σ(s, X
s
) σ(s, X
0
s
)) dB
s
!
2
= E
Z
tS
0
(σ(s, X
s
) σ(s, X
0
s
))
2
ds
!
.
So using the Lipschitz bound, we have
E((X
tS
X
0
tS
)
2
) 2K
2
(1 + T )E
Z
tS
0
|X
s
X
0
s
|
2
ds
!
2K
2
(1 + T )
Z
t
0
E(|X
sS
X
0
sS
|
2
) ds.
We now use Gr¨onwall’s lemma:
Lemma. Let h(t) be a function such that
h(t) c
Z
t
0
h(s) ds
for some constant c. Then
h(t) h(0)e
ct
.
Applying this to
h(t) = E((X
tS
X
0
tS
)
2
),
we deduce that h(t) h(0)e
ct
= 0. So we know that
E(|X
tS
X
0
tS
|
2
) = 0
for every t [0, T ]. Taking n and T gives pathwise uniqueness.
We next prove existence of solutions. We fix (Ω
, F,
(
F
t
)
t
) and
B
, and define
F (X)
t
= X
0
+
Z
t
0
σ(s, X
s
) dB
s
+
Z
t
0
b(s, X
s
) ds.
Then
X
is a solution to
E
x
(
a, b
) iff
F
(
X
) =
X
and
X
0
=
x
. To find a fixed point,
we use Picard iteration. We fix
T >
0, and define the
T
-norm of a continuous
adapted process X as
kXk
T
= E
sup
tT
|X
t
|
2
1/2
.
In particular, if
X
is a martingale, then this is the same as the norm on the
space of L
2
-bounded martingales by Doob’s inequality. Then
B = {X : Ω × [0, T ] R : kXk
T
< ∞}
is a Banach space.
Claim. kF (0)k
T
< , and
kF (X) F (Y )k
2
T
(2T + 8)K
2
Z
T
0
kX Y k
2
t
dt.
We first see how this claim implies the theorem. First observe that the claim
implies
F
indeed maps
B
into itself. We can then define a sequence of processes
X
i
t
by
X
0
t
= x, X
i+1
= F (X
i
).
Then we have
kX
i+1
X
i
k
2
T
CT
Z
T
0
kX
i
X
i1
k
2
t
dt ··· kX
1
X
0
k
2
T
CT
i
i!
.
So we find that
X
i=1
kX
i
X
i1
k
2
T
<
for all
T
. So
X
i
converges to
X
almost surely and uniformly on [0
, T
], and
F (X) = X. We then take T and we are done.
To prove the claim, we write
kF (0)k
T
|X
0
| +
Z
t
0
b(s, 0) ds
+
Z
t
0
σ(s, 0) dB
s
T
.
The first two terms are constant, and we can bound the last by Doob’s inequality
and the Itˆo isometry:
Z
t
0
σ(s, 0) dB
s
T
2E
Z
T
0
σ(s, 0) dB
s
2
= 2
Z
T
0
σ(s, 0)
2
ds.
To prove the second part, we use
kF (X) F (Y )k
2
2E
sup
tT
Z
t
0
b(s, X s) b(s, Y
s
) ds
2
!
+ 2E
sup
tT
Z
t
0
(σ(s, X
s
) σ(s, Y
s
)) dB
s
2
!
.
We can bound the first term with Cauchy–Schwartz by
T E
Z
T
0
|b(s, X
s
) b(s, Y
s
)|
2
!
T K
2
Z
T
0
kX Y k
2
t
dt,
and the second term with Doob’s inequality by
E
Z
T
0
|σ(s, X
s
) σ(s, Y
s
)|
2
ds
!
4K
2
Z
T
0
kX Y k
2
t
dt.
4.2 Examples of stochastic differential equations
Example
(The Ornstein–Uhlenbeck process)
.
Let
λ >
0. Then the Ornstein–
Uhlenbeck process is the solution to
dX
t
= λX
t
dt + dB
t
.
The solution exists by the previous theorem, but we can also explicitly find one.
By Itˆo’s formula applied to e
λt
X
t
, we get
d(e
λt
X
t
) = e
λt
dX
t
+ λe
λt
X
t
dt = dB
t
.
So we find that
X
t
= e
λt
X
0
+
Z
t
0
e
λ(ts)
dB
s
.
Observe that the integrand is deterministic. So we can in fact interpret this as
an Wiener integral.
Fact.
If
X
0
=
x R
is fixed, then (
X
t
) is a Gaussian process, i.e. (
X
t
i
)
n
i=1
is
jointly Gaussian for all
t
1
< ··· < t
n
. Any Gaussian process is determined by
the mean and covariance, and in this case, we have
EX
t
= e
λt
x, cov(X
t
, X
s
) =
1
2λ
e
λ|ts|
e
λ|t+s|
Proof. We only have to compute the covariance. By the Itˆo isometry, we have
E((X
t
EX
t
)(X
s
EX
s
)) = E
Z
t
0
e
λ(tu)
dB
u
Z
s
0
e
λ(su)
dB
u
= e
λ(t+s)
Z
ts
0
e
λu
du.
In particular,
X
t
N
e
λt
x,
1 e
2λt
2λ
N
0,
1
2λ
.
Fact.
If
X
0
N
(0
,
1
2λ
), then (
X
t
) is a centered Gaussian process with stationary
covariance, i.e. the covariance depends only on time differences:
cov(X
t
, X
s
) =
1
2λ
e
λ|ts|
.
The difference is that in the deterministic case, the
EX
t
cancels the first
e
λt
X
0
term, while in the non-deterministic case, it doesn’t.
This is a very nice example where we can explicitly understand the long-time
behaviour of the SDE. In general, this is non-trivial.
Dyson Brownian motion
Let
H
N
be an inner product space of real symmetric
N ×N
matrices with inner
product
N Tr
(
HK
) for
H, K H
N
. Let
H
1
, . . . , H
dim(H
N
)
be an orthonormal
basis for H
N
.
Definition
(Gaussian orthogonal ensemble)
.
The Gaussian Orthogonal Ensem-
ble GOE
N
is the standard Gaussian measure on H
N
, i.e. H GOE
N
if
H =
dim H
n
X
r=1
H
i
X
i
where each X
i
are iid standard normals.
We now replace each
X
i
by a Ornstein–Uhlenbeck process with
λ
=
1
2
. Then
GOE
N
is invariant under the process.
Theorem. The eigenvalues λ
1
(t) ··· λ
N
(t) satisfies
dλ
i
t
=
λ
i
2
+
1
N
X
j6=i
1
λ
i
λ
j
dt +
r
2
Nβ
dB
i
.
Here
β
= 1, but if we replace symmetric matrices by Hermitian ones, we get
β = 2; if we replace symmetric matrices by symplectic ones, we get β = 4.
This follows from Itˆo’s formula and formulas for derivatives of eigenvalues.
Example
(Geometric Brownian motion)
.
Fix
σ >
0 and
t R
. Then geometric
Brownian motion is given by
dX
t
= σX
t
dB
t
+ rX
t
dt.
We apply Itˆo’s formula to log X
t
to find that
X
t
= X
0
exp
σB
t
+
r
σ
2
2
t
.
Example
(Bessel process)
.
Let
B
= (
B
1
, . . . , B
d
) be a
d
-dimensional Brownian
motion. Then
X
t
= |B
t
|
satisfies the stochastic differential equation
dX
t
=
d 1
2X
t
dt + dB
t
if t < inf{t 0 : X
t
= 0}.
4.3 Representations of solutions to PDEs
Recall that in Advanced Probability, we learnt that we can represent the solution
to Laplace’s equation via Brownian motion, namely if
D
is a suitably nice domain
and
g
:
D R
is a function, then the solution to the Laplace’s equation on
D
with boundary conditions g is given by
u(x) = E
x
[g(B
T
)],
where T is the first hitting time of the boundary D.
A similar statement we can make is that if we want to solve the heat equation
u
t
=
2
u
with initial conditions u(x, 0) = u
0
(x), then we can write the solution as
u(x, t) = E
x
[u
0
(
2B
t
)]
This is just a fancy way to say that the Green’s function for the heat equation is
a Gaussian, but is a good way to think about it nevertheless.
In general, we would like to associate PDEs to certain stochastic processes.
Recall that a stochastic PDE is generally of the form
dX
t
= b(X
t
) dt + σ(X
t
) dB
t
for some
b
:
R
d
R
and
σ
:
R
d
R
d×m
which are measurable and locally
bounded. Here we assume these functions do not have time dependence. We
can then associate to this a differential operator L defined by
L =
1
2
X
i,j
a
ij
i
j
+
X
i
b
i
i
.
where a = σσ
T
.
Example. If b = 0 and σ =
2I, then L = ∆ is the standard Laplacian.
The basic computation is the following result, which is a standard application
of the Itˆo formula:
Proposition.
Let
x R
d
, and
X
a solution to
E
x
(
σ, b
). Then for every
f : R
+
× R
d
R that is C
1
in R
+
and C
2
in R
d
, the process
M
f
t
= f(t, X
t
) f(0, X
0
)
Z
t
0
s
+ L
f(s, X
s
) ds
is a continuous local martingale.
We first apply this to the Dirichlet–Poisson problem, which is essentially to
solve
Lu
=
f
. To be precise, let
U R
d
be non-empty, bounded and open;
f C
b
(
U
) and
g C
b
(
U
). We then want to find a
u C
2
(
¯
U
) =
C
2
(
U
)
C
(
¯
U
)
such that
Lu(x) = f(x) for x U
u(x) = g(x) for x U.
If
f
= 0, this is called the Dirichlet problem; if
g
= 0, this is called the Poisson
problem.
We will have to impose the following technical condition on a:
Definition
(Uniformly elliptic)
.
We say
a
:
¯
U R
d×d
is uniformly elliptic if
there is a constant c > 0 such that for all ξ R
d
and x
¯
U, we have
ξ
T
a(x)ξ c|ξ|
2
.
If
a
is symmetric (which it is in our case), this is the same as asking for the
smallest eigenvalue of a to be bounded away from 0.
It would be very nice if we can write down a solution to the Dirichlet–Poisson
problem using a solution to
E
x
(
σ, b
), and then simply check that it works. We
can indeed do that, but it takes a bit more time than we have. Instead, we shall
prove a slightly weaker result that if we happen to have a solution, it must be
given by our formula involving the SDE. So we first note the following theorem
without proof:
Theorem.
Assume
U
has a smooth boundary (or satisfies the exterior cone
condition),
a, b
are older continuous and
a
is uniformly elliptic. Then for
every older continuous
f
:
¯
U R
and any continuous
g
:
U R
, the
Dirichlet–Poisson process has a solution.
The main theorem is the following:
Theorem.
Let
σ
and
b
be bounded measurable and
σσ
T
uniformly elliptic,
U R
d
as above. Let
u
be a solution to the Dirichlet–Poisson problem and
X
a
solution to E
x
(σ, b) for some x R
d
. Define the stopping time
T
U
= inf{t 0 : X
t
6∈ U}.
Then ET
U
< and
u(x) = E
x
g(X
T
U
) +
Z
T
U
0
f(X
s
) ds
!
.
In particular, the solution to the PDE is unique.
Proof.
Our previous proposition applies to functions defined on all of
R
n
, while
u is just defined on U . So we set
U
n
=
x U : dist(x, U ) >
1
n
, T
n
= inf{t 0 : X
t
6∈ U
n
},
and pick
u
n
C
2
b
(
R
d
) such that
u|
U
n
=
u
n
|
U
n
. Recalling our previous notation,
let
M
n
t
= (M
u
n
)
T
n
t
= u
n
(X
tT
n
) u
n
(X
0
)
Z
tT
n
0
Lu
n
(X
s
) ds.
Then this is a continuous local martingale that is bounded by the proposition,
and is bounded, hence a true martingale. Thus for
x U
and
n
large enough,
the martingale property implies
u(x) = u
n
(x) = E
u(X
tT
n
)
Z
tT
n
0
Lu(X
s
) ds
!
= E
u(X
tT
n
) +
Z
tT
n
0
f(X
s
) ds
!
.
We would be done if we can take
n
. To do so, we first show that
E
[
T
U
]
<
.
Note that this does not depend on
f
and
g
. So we can take
f
= 1 and
g
= 0,
and let v be a solution. Then we have
E(t T
n
) = E
Z
tT
n
0
Lv(X
s
) ds
!
= v(x) E(v(X
tT
n
)).
Since
v
is bounded, by dominated/monotone convergence, we can take the limit
to get
E(T
U
) < .
Thus, we know that t T
n
T
U
as t and n . Since
E
Z
T
U
0
|f(X
s
)| ds
!
kfk
E[T
U
] < ,
the dominated convergence theorem tells us
E
Z
tT
n
0
f(X
s
) ds
!
E
Z
T
U
0
f(X
s
) ds
!
.
Since u is continuous on
¯
U, we also have
E(u(X
tT
n
)) E(u(T
u
)) = E(g(T
u
)).
We can use SDE’s to solve the Cauchy problem for parabolic equations as
well, just like the heat equation. The problem is as follows: for
f C
2
b
(
R
d
), we
want to find u : R
+
× R
d
R that is C
1
in R
+
and C
2
in R
d
such that
u
t
= Lu on R
+
× R
d
u(0, ·) = f on R
d
Again we will need the following theorem:
Theorem.
For every
f C
2
b
(
R
d
), there exists a solution to the Cauchy problem.
Theorem.
Let
u
be a solution to the Cauchy problem. Let
X
be a solution to
E
x
(σ, b) for x R
d
and 0 s t. Then
E
x
(f(X
t
) | F
s
) = u(t s, X
s
).
In particular,
u(t, x) = E
x
(f(X
t
)).
In particular, this implies X
t
is a continuous Markov process.
Proof.
The martingale has
t
+
L
, but the heat equation has
t
L
. So we set
g(s, x) = u(t s, x). Then
s
+ L
g(s, x) =
t
u(t s, x) + Lu(t s, x) = 0.
So g(s, X
s
) g(0, X
0
) is a martingale (boundedness is an exercise), and hence
u(t s, X
s
) = g(s, X
s
) = E(g(t, X
t
) | F
s
) = E(u(0, X
t
) | F
s
) = E(f(X
t
) | X
s
).
There is a generalization to the Feynman–Kac formula.
Theorem
(Feynman–Kac formula)
.
Let
f C
2
b
(
R
d
) and
V C
b
(
R
d
) and
suppose that u : R
+
× R
d
R satisfies
u
t
= Lu + V u on R
+
× R
d
u(0, ·) = f on R
d
,
where V u = V (x)u(x) is given by multiplication.
Then for all t > 0 and x R
d
, and X a solution to E
x
(σ, b). Then
u(t, x) = E
x
f(X
t
) exp
Z
t
0
V (X
s
) ds

.
If
L
is the Laplacian, then this is Schr¨odinger equation, which is why Feynman
was thinking about this.