stochastic calculus1. introduction 1 2. preliminaries 4 3. local martingales 10 4. the stochastic...

65
STOCHASTIC CALCULUS JASON MILLER AND VITTORIA SILVESTRI Contents Preface 1 1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic differential equations 49 8. Diffusion processes 59 Preface These lecture notes are for the University of Cambridge Part III course Stochastic Calculus, given Lent 2017. The contents are very closely based on a set of lecture notes for this course due to Christina Goldschmidt. Please notify [email protected] for comments and corrections. 1. Introduction In ordinary calculus, one learns how to integrate, differentiate, and solve ordinary differential equations. In this course, we will develop the theory for the stochastic analogs of these constructions: the Itˆ o integral, the Itˆ o derivative, and stochastic differential equations (SDEs). Let us start with an example. Suppose a gambler is repeatedly tossing a fair coin, while betting £1 on head at each coin toss. Let ξ 1 2 3 ... be independent random variables denoting the consecutive outcomes, i.e. ξ k = +1 if head at k th coin toss, -1 otherwise, (ξ k ) k1 i.i.d. 1

Upload: others

Post on 19-Jun-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS

JASON MILLER AND VITTORIA SILVESTRI

Contents

Preface 1

1. Introduction 1

2. Preliminaries 4

3. Local martingales 10

4. The stochastic integral 16

5. Stochastic calculus 36

6. Applications 44

7. Stochastic differential equations 49

8. Diffusion processes 59

Preface

These lecture notes are for the University of Cambridge Part III course Stochastic Calculus,

given Lent 2017. The contents are very closely based on a set of lecture notes for this course

due to Christina Goldschmidt. Please notify [email protected] for comments and corrections.

1. Introduction

In ordinary calculus, one learns how to integrate, differentiate, and solve ordinary differential

equations. In this course, we will develop the theory for the stochastic analogs of these

constructions: the Ito integral, the Ito derivative, and stochastic differential equations (SDEs).

Let us start with an example. Suppose a gambler is repeatedly tossing a fair coin, while betting

£1 on head at each coin toss. Let ξ1, ξ2, ξ3 . . . be independent random variables denoting the

consecutive outcomes, i.e.

ξk =

+1 if head at kth coin toss,

−1 otherwise,(ξk)k≥1 i.i.d.

1

Page 2: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

2 JASON MILLER AND VITTORIA SILVESTRI

Then the gambler’s net winnings after n coin tosses are given by Xn := ξ1 + ξ2 + . . . + ξn.

Note that (Xn)n≥0 is a simple random walk on Z starting at X0 = 0, and it is therefore a

discrete time martingale with respect to the filtration (Fn)n≥0 generated by the outcomes, i.e.

Fn = σ(ξ1, . . . , ξn). Now suppose that, at the mth coin toss, the gambler bets an amount Hm

on head (we can also allow Hm to be negative by interpreting it as a bet on tail). Take, for

now, (Hm)m≥1 to be deterministic (for example, the gambler could have decided in advance

the sequence of bets). Then it is easy to see that

(1.1) (H ·X)n :=n∑k=1

Hk(Xk −Xk−1)

gives the net winnings after n coin tosses. We claim that the stochastic process H · Xis a martingale with respect to (Fn)n≥0. Indeed, integrability and adaptedness are clear.

Furthermore,

(1.2) E((H ·X)n+1 − (H ·X)n

∣∣Fn) = Hn+1E(Xn+1 −Xn|Fn) = 0

for all n ≥ 1, which shows that H · X is a martingale. In fact, this property is much more

general. Indeed, instead of deciding his bets in advance, the gambler might want to allow its

(n+1)th bet to depend on the first n outcomes. That is, we can take Hn+1 to be random, as long

as it is Fn-measurable. Such processes are called previsible, and they will play a fundamental

role in this course. In this more general setting, (H ·X)n still represents the net amount of

winnings after n bets, and (1.2) still holds, so that H ·X is again a discrete time martingale.

For this reason, the process H ·X is usually referred to as discrete martingale transform.

Remark 1.1. This teaches us that we cannot make money by gambling on a fair game!

Our goal for the first part of the course is to generalise the above reasoning to the continuous

time setting, i.e. to define the process ∫ t

0HsdXs

for (Ht)t≥0 suitable previsible1 process and (Xt)t≥0 continuous martingale (think, for example,

of X as being standard one-dimensional Brownian Motion, the analogue to a simple random

walk in the continuum). To achieve this, as a first attempt one could try to generalise the

Lebesgue-Stieltjes theory of integration to more general integrators: this will lead us to talk

about finite variation processes. Unfortunately, as we will see, the only continuous martingales

which have finite variation are the constant ones, so a new theory is needed in order to integrate,

for example, with respect to Brownian Motion. This is what is commonly called Ito’s theory

of integration. To see how one could go to build this new theory, note that definition (1.1) is

1At this point we have not yet defined the notion of previsible process in the continuous time setting: this

will be clarified later in the course.

Page 3: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 3

perfectly well-posed in continuous time whenever the integrand H is piecewise constant taking

only finitely many values. In order to accommodate more general integrands, one could then

think to take limits, setting, for example,

(1.3) (H ·X)t ” := ” limε→0

bt/εc∑k=0

Hkε(X(k+1)ε −Xkε).

On the other hand, one quickly realises that the r.h.s. above in general does not converge in the

usual sense, due to the roughness of X (think, for example, to Brownian Motion). What is the

right notion of convergence that makes the above limit finite? As we will see in the construction

of Ito integral, to get convergence one has to exploit the cancellations in the above sum. As an

example, take X to be Brownian Motion, and observe that

E[( bt/εc∑

k=0

Hkε(X(k+1)ε −Xkε)

)2]=

= E[ bt/εc∑k=0

H2kε

(X(k+1)ε −Xkε

)2+

bt/εc∑j 6=k

HjεHkε(X(k+1)ε −Xkε)(X(j+1)ε −Xjε)

]

= E[ bt/εc∑k=0

H2kε

(X(k+1)ε −Xkε

)2 ]= ε

bt/εc∑k=0

H2kε →

∫ t

0H2sds as ε→ 0,

where the cancellations are due to orthogonality of the martingale increments. These type

of considerations will allow us to provide a rigorous construction of the Ito integral for X

continuous martingale (such as Brownian Motion), and in fact even for slightly more general

integrands, namely local martingales and semimartingales.

Once the stochastic integral has been defined, we will discuss its properties and applications. We

will look, for example, at iterated integrals, the stochastic chain rule and stochastic integration

by part. Writing, to shorten the notation,

dYt = HtdXt

to mean Yt =∫ t

0 HsdXs, we will learn how to express df(Yt) in terms of dYt by mean of the

celebrated Ito’s formula. This is an extremely useful tool, of which we will present several

applications. For example, we will be able to show that any continuous martingale is a

time-changed Brownian Motion (this is the so-called Dubins-Schwarz theorem).

In the second part of the course we will then focus on the study of Stochastic Differential

Equations (SDEs in short), that is equations of the form

(1.4) dXt = b(t,Xt)dt+ σ(t,Xt)dBt, X0 = x0,

Page 4: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

4 JASON MILLER AND VITTORIA SILVESTRI

for suitably nice functions a, σ. These can be thought of as arising from randomly perturbed

ODEs. To clarify this point, suppose we are given the ODE

(1.5)

dx(t)

dt= b(t, x(t)),

x(0) = x0,

which writes equivalently as x(t) = x0 +∫ t

0 b(s, x(s))ds. In many applications we may wish to

perturb (1.5) by adding some noise, say Brownian noise. This gives a new (stochastic) equation

Xt = x0 +

∫ t

0b(s,Xs)ds+ σBt,

where the real parameter σ controls the intensity of the noise. In fact, we may also want to

allow such intensity to depend on the current time and state of the system, to get

Xt = x0 +

∫ t

0b(s,Xs)ds+

∫ t

0σ(s,Xs)dBs,

where the last term is an Ito integral. This, written in differential form, gives (1.4). Of such

SDEs we will present several notions of solutions, and discuss under which conditions on the

functions b and σ such solutions exist and are unique.

Finally, in the last part of the course we will talk about diffusion processes, which are Markov

processes characterised by martingales properties. We will explain how such processes can be

constructed from SDEs, and how they can be used to build solutions to certain (deterministic)

PDEs involving second order elliptic operators.

2. Preliminaries

2.1. Finite variation integrals. Recall that a function a : [0,∞)→ R is said to be cadlag if

it is right-continuous and has left limits. In other words, for all x ∈ (0,∞) we have both

limy→x−

a(y) exists and limy→x+

a(y) = a(x).

We write a(x−) for the left limit at x, and let ∆a(x) := a(x) − a(x−). Assume that a is

non-decreasing and cadlag with a(0) = 0. Then there exists a unique Borel measure da on

[0,∞) such that for all s < t, we have that

da((s, t]) = a(t)− a(s).

For any f : R → R measurable and integrable function we can define the Lebesgue-Stieltjes

integral f · a by setting

(f · a)(t) =

∫(0,t]

f(s)da(s)

Page 5: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 5

for all t ≥ 0. It is easy to see that f · a is right-continuous. Moreover, if a is continuous then

f · a is itself continuous. In this case, we can write∫(0,t]

f(s)da(s) =

∫ t

0f(s)da(s)

unambiguously. We are now interested in enlarging the class of functions a against which we can

integrate. A first, trivial observation is that we can use linearity to integrate against functions

a which write as a = a+ − a−, for some a+, a− satisfying the conditions above. Then we set

(f · a)(t) := (f · a+)(t)− (f · a−)(t)

for all f measurable functions and t ≥ 0, provided both terms in the r.h.s. are finite. How large

is the class of such functions? It turns out to exactly coincide with the class of finite variation

functions.

Definition 2.1. Let a : [0,∞)→ R be a cadlag function. For any n ∈ N and t ≥ 0, let

(2.1) vn(t) =

d2nte−1∑k=0

|a((k + 1)2−n)− a(k2−n)|.

Then v(t) := limn→∞ vn(t) exists for all t, and it is called total variation of a on (0, t]. If

v(t) <∞ then a is said to have finite variation on (0, t]. We say that a is of finite variation if

it is of finite variation on (0, t] for all t ≥ 0.

To show that the above definition makes sense we have to show that limn→∞ vn(t) exists at all

t. Indeed, let t+n = 2−nd2nte, t−n = 2−n(d2nte− 1), and note that t+n ≥ t ≥ t−n for all n. We have

vn(t) =

2nt−n−1∑k=0

|a((k + 1)2−n)− a(k2−n)|+ |a(t+n )− a(t−n )|.

Now, the first term is increasing in n by the triangle inequality, and hence it has a limit as

n→∞. Moreover, by the cadlag property the second term converges to |∆a(t)| as n→∞. In

all, vn(t) has a limit as n→∞.

Lemma 2.2. Let a : [0,∞)→ R be a cadlag function of finite variation, that is v(t) <∞ for

all t ≥ 0. Then v is also with ∆v(t) = |∆a(t)|, and non-decreasing in t. In particular, if a is

continuous then also v is continuous.

Proof. See Example Sheet 1.

We can now state the aforementioned characterisation result.

Proposition 2.3. A cadlag function a : [0,∞) → R can be written as difference of two

right-continuous, non-decreasing functions if and only if a is of finite variation.

Page 6: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

6 JASON MILLER AND VITTORIA SILVESTRI

Proof. Assume that a = a+ − a−, for a+, a− non-decreasing, right-continuous functions. We

now show that a is of finite variation. Indeed, for any s ≤ t we have

|a(t)− a(s)| ≤ (a+(t)− a+(s)) + (a−(t)− a−(s)),

where both terms are non-negative since a+, a− are non-decreasing. Plugging this into (2.1),

and observing that the summations defining the total variation of a+, a− are telescopic, we

gather that

vn(t) ≤ a+(t+n )− a+(0) + a−(t+n )− a−(0).

Being a+, a− cadlag , the r.h.s. converges to a+(t) − a+(0) + a−(t) − a−(0), which is finite.

Since t is arbitrary, this shows that v(t) <∞ for all t ≥ 0, i.e. a is of finite variation.

We now prove the reverse implication by constructing a+ and a− explicitly. Assume that

v(t) <∞ for all t ≥ 0. Set

a+ =1

2(v + a) , a− =

1

2(v − a) .

Then clearly, a = a+ − a−, and a+, a− are cadlag since v and a are. It remains to show that

they are non-decreasing. For s < t, define t±n as above, and s±n analogously. Then we have

a+(t)− a+(s) = limn→∞

1

2

[(vn(t)− vn(s)) + (a(t)− a(s))

]= lim

n→∞

1

2

[ 2nt−n−1∑k=2ns+n

(|a((k + 1)2−n)− a(k2−n)|+

(a((k + 1)2−n)− a(k2−n)

)︸ ︷︷ ︸≥0

)

+ |a(t+n )− a(t−n )|+ (a(t+n )− a(t−n ))︸ ︷︷ ︸≥0

]≥ 0.

The same argument can be used to show that a− is non-decreasing.

2.2. Random integrators. We now discuss integration with respect to random finite variation

functions. Let (Ω,F , (Ft)t≥0,P) denote a filtered probability space. Recall that a process

X : Ω× [0,∞)→ R is adapted to the filtration (Ft)t≥0 if Xt = X(·, t) is Ft-measurable for all

t ≥ 0. Moreover, X is cadlag if X(ω, ·) is cadlag for all ω ∈ Ω (that is, if all realizations of X

are cadlag ).

Definition 2.4. Given a cadlag adapted process A : Ω× [0,∞)→ R, its total variation process

V : Ω× [0,∞)→ R is defined pathwise by letting V (ω, ·) be the total variation of A(ω, ·), for

all ω ∈ Ω.

We say that A is of finite variation if A(ω, ·) is of finite variation for all ω ∈ Ω (that is, if all

realizations of A are of finite variation).

Page 7: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 7

Lemma 2.5. Let A be a cadlag , adapted process of finite variation V . Then V is cadlag ,

adapted and pathwise non-decreasing.

Proof. By Lemma 2.2, for all ω ∈ Ω the path V (ω, ·) is cadlag and non-decreasing, so the

process V is cadlag and pathwise non-decreasing. To see that it is adapted, recall that for any

t ≥ 0 we define t−n = 2−n(d2nte − 1), and introduce for all n the process

V nt =

2nt−n−1∑k=0

|A(k+1)2−n −Ak2−n |.

Then V n is adapted for all n, since t−n ≤ t. Moreover,

Vt = limn→∞

V nt + |∆At|,

which shows that Vt is Ft-measurable since both terms in the r.h.s. are Ft-measurable.

Our next goal is to be able to integrate random functions against A. Clearly, we can define the

integral pathwise, but doing so does not guarantee that the resulting process is both cadlag

and adapted. To have this, we need to restrict our class of integrands to so called previsible

processes.

2.2.1. Previsible processes. Recall from the introduction that a discrete time process H =

(Hn)n≥1 is previsible with respect to a filtration (Fn)n≥0 if Hn+1 is Fn-measurable for all n ≥ 0.

Here is the analogous definition in continuous time.

Definition 2.6. The previsible σ-algebra P on Ω× (0,∞) is the σ-algebra generated by sets of

the form E × (s, t], for E ∈ Fs and s < t. A process H : Ω× (0,∞) is said to be previsible if it

is measurable with respect to P.

It is easy to see directly from the definition that, for example, any process of the form

H(ω, t) = Z(ω)1(t1,t2](t), Z Ft1-measurable,

is previsible. Thus clearly if we let

(2.2) H(ω, t) =

n∑k=0

Zk(ω)1(tk,tk+1](t)

for some n ∈ N, 0 = t0 < t1 · · · < tn <∞ and Zk Ftk -measurable for all k, then H is previsible.

Processes of the form (2.2) are called simple, and they will play an important role in the

construction of the Ito integral later in the course.

Remark 2.7. Note that all simple processes are left-continuous and adapted. In general, it is

possible to show that P is the smallest σ-algebra on Ω× (0,∞) such that all left-continuous,

adapted processes are measurable.

Page 8: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

8 JASON MILLER AND VITTORIA SILVESTRI

Although not all cadlag adapted processes are previsible, the next result shows that their

left-continuous modification is.

Proposition 2.8. Let X be a cadlag adapted process, and define for all t the process Ht := Xt− .

Then H is previsible.

Proof. Since X is cadlag and adapted, H is left-continuous and adapted. Now to see that H is

previsible we express it as a limit of previsible processes. For n ≥ 1, define the left-continuous,

piecewise constant process

Hnt :=

∞∑k=0

Hk2−n1(k2−n,(k+1)2−n](t).

Then the processes Hn are clearly previsible for all n ≥ 1. Moreover, by left-continuity of H

we have that limn→∞Hnt = Ht for all t. Thus H is a limit of previsible processes, and it is

therefore previsible.

Remark 2.9. By the above proposition, all continuous processes (such as Brownian Motion)

are previsible.

The following result provides a useful tool to determine that a process is not previsible.

Proposition 2.10. If a process H is previsible, then Ht is measurable with respect to Ft− =

σ(Fs : s < t), for all t ≥ 0.

Proof. See Example sheet 1.

This shows, for example, that a Poisson process (Nt)t≥0 is not previsible since Nt is not

Ft−-measurable.

We can now show that integrating previsible processes against cadlag , adapted processes of

finite variation results in cadlag , adapted processes of finite variation.

Theorem 2.11. Let A : Ω × [0,∞) → R be a cadlag , adapted process of finite variation V .

Let H be a previsible process, and assume that for all ω ∈ Ω it holds

(2.3)

∫(0,t]|H(ω, s)|dV (ω, s) <∞, ∀t > 0.

Then the process (H ·A) : Ω× [0,∞)→ R defined by

(2.4) (H ·A)(ω, t) :=

∫(0,t]

H(ω, s)dA(ω, s), (H ·A)(ω, 0) ≡ 0

is cadlag , adapted and of finite variation.

Page 9: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 9

Proof. Note that the integral in (2.4) is well defined thanks to the integrability condition

(2.3). Indeed, write H± := max±H, 0 and A± = 12(V ± A), so that H = H+ − H− and

A = A+ −A−. Then

H ·A = (H+ −H−) · (A+ −A−) = H+ ·A+ −H− ·A+ −H+ ·A− +H− ·A−,

and all terms in the r.h.s. are finite by (2.3).

Step 1: Cadlag. Note that 1(0,s] → 1(0,t] as s t and 1(0,s] → 1(0,t) as s t. Recall that

(H · A)t =∫Hs1(s ∈ (0, t])dAs. By the dominated convergence theorem (with ω fixed), we

have that

(H ·A)t =

∫Hs lim

rt1(s ∈ (0, r])dAs

= limrt

∫Hs1(s ∈ (0, r])dAs

= limrt

(H ·A)r.

Therefore H ·A is right-continuous. A similar argument implies that H ·A has left limits.

Moreover, we have that

∆(H ·A)t =

∫Hs1(s = t)dAs = Ht∆At.

Step 2: Adapted. We prove adaptedness by a monotone class argument. Suppose that

H = 1B×(s,u] where B ∈ Fs and u > s. Then we have that

(H ·A)t = 1B (At∧u −At∧s) ,

which is clearly Ft-measurable. Let

A = C ∈ P : 1C ·A is adapted to (Ft)t≥0,

we aim to show that A = P. To this end, let

Π = B × (s, u] : B ∈ Fs, s < u.

Then Π is a π-system and we have shown above that Π ⊆ A. To prove the reverse inclusion,

observe that A is a d-system, so by Dynkin’s lemma we have that P = σ(Π) ⊆ A ⊆ P.

Therefore A = P, i.e. H ·A is adapted for all H = 1C , C ∈ P.

Now to generalise, suppose that H ≥ 0 is previsible. Set

Hn = (2−nb2nHc) ∧ n

=2nn−1∑k=1

2−nk1(H ∈ [2−nk, 2−n(k + 1))︸ ︷︷ ︸∈P

) + n1(H > n︸ ︷︷ ︸∈P

),

Page 10: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

10 JASON MILLER AND VITTORIA SILVESTRI

and note that Hn is a finite linear combination of functions of the form 1C for C ∈ P. It

therefore follows by linearity and that (Hn ·A)t is Ft-measurable for all t. Moreover Hn H

almost surely as n→∞, and so by the Monotone Convergence Theorem (MCT) we have that

(Hn ·A)t → (H ·A)t as n→∞. It follows that (H ·A)t is Ft-measurable for all t ≥ 0, as limit

of Ft-measurable functions. This extends in the usual way to the case of general H by writing

H = H+ −H− and observing that H+ ·A and H− ·A are both adapted, so such is H ·A.

Step 3: Finite variation. Simply observe that

H ·A = (H+ −H−) · (A+ −A−) = (H+ ·A+ +H− ·A−)− (H− ·A+ +H+ ·A−)

is a difference of non-decreasing processes, hence of finite variation.

Our next goal is to integrate against continuous martingales, such as Brownian Motion. By the

previous theorem, were we able to identify a class of continuous martingales with finite variations

we would be done, as then we would be able to apply the definition in (2.4). Unfortunately – but

also interestingly – we will see in the next section that the only continuous martingales which

have finite variation are the constant ones. This shows that in order to be able to integrate with

respect to general (non-constant) continuous martingales one need to develop a new theory of

integration, which we discuss in Section 4.

3. Local martingales

Let (Ω,F , (Ft)t≥0,P) be a filtered probability space.

Definition 3.1. A filtration (Ft)t≥0 is said to satisfy the usual condition if

• F0 contains all P-null sets,

• (Ft)t≥0 is right-continuous, that is for all t ≥ 0 it holds Ft = Ft+ :=⋂s>tFs.

We assume that (Ft)t≥0 satisfies the usual condition throughout the course.

Recall that an integrable, adapted process X = (Xt)t≥0 is a martingale with respect to (Ft)t≥0

if for all 0 ≤ s ≤ t it holds E(Xt|Fs) = Xs almost surely. Similarly, X is a supermartingale

(resp. submartingale) if for all 0 ≤ s ≤ t it holds E(Xt|Fs) ≤ Xs (resp. E(Xt|Fs) ≥ Xs) almost

surely. Moreover, a random variable T : Ω→ [0,∞] is called a stopping time if T ≤ t ∈ Ftfor all t ≥ 0. Note that if X is continuous (or cadlag ) and we let

FT := E ∈ F : E ∩ T ≤ t ∈ Ft ∀t ≥ 0

then XT is FT -measurable. Crucially, if X is a martingale then the stopped process XT :=

(Xt∧T )t≥0 is again a martingale. In fact, this property can be used to characterise martingales,

as shown in the next theorem.

Page 11: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 11

Theorem 3.2 (Optional Stopping Theorem, OST). Let X be an adapted integrable process.

Then the following are equivalent:

(1) X is a martingale,

(2) XT = (Xt∧T )t≥0 is a martingale for all stopping times T ,

(3) for all bounded stopping times S ≤ T it holds E(XT |FS) = XS a.s.,

(4) for all bounded stopping times T it holds E(XT ) = E(X0).

In other words, the class of cadlag martingales is stable under stopping. This leads us to define

a slightly more general class of such processes, called local martingales.

Definition 3.3 (Local martingale). A cadlag adapted process is called a local martingale if

there exists a sequence (Tn)n≥1 of stopping times with Tn +∞ almost surely, such that the

stopped process XTn is a martingale for all n ≥ 1. In this case we say that the sequence (Tn)n≥1

reduces X.

Note that any martingale is a local martingale, since any deterministic sequence (Tn)n≥1 will

reduce it. The following important example shows that, in fact, the class of local martingales is

strictly larger.

Example 3.4. Let B be a standard Brownian motion in R3. Let Mt = 1/|Bt|. Recall from

Example Sheet 3, Exercise 7.7 of Advanced Probability that:

(i) (Mt)t≥1 is bounded in L2, that is supt≥1

E(M2t ) <∞.

(ii) E(Mt)→ 0 as t→∞.

(iii) M is a supermartingale.

Note that M cannot be a martingale, as if it were it should have constant expectation, which

would have to vanish by (ii). We now show that it is a local martingale by exhibiting a reducing

sequence of stopping times.

For each n, let Tn = inft ≥ 1 : |Bt| ≤ 1/n = inft ≥ 1 : |Mt| ≥ n. We aim to show that

(MTnt )t≥1 is a martingale for all n and Tn ∞ almost surely.

We begin by noting that if n ≤ M1(ω) then Tn(ω) = 1 and if n > M1(ω) then Tn(ω) > 1.

Moreover, since |Bt| cannot hit 1/(n+ 1) without first having hit 1/n, it follows that Tn(ω) is

non-decreasing in n.

Recall from Advanced Probability that if f ∈ C2b (R3) then

f(Bt)− f(B0)− 1

2

∫ t

0∆f(Bs)ds, t ≥ 0,

is a martingale. Note that f(x) = 1/|x| is harmonic in R3 \ 0. To deal with the singularity

at the origin, let (fn)n≥1 be a sequence of bounded C2(R3) functions such that fn ≡ f on

Page 12: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

12 JASON MILLER AND VITTORIA SILVESTRI

|x| ≥ 1/n, for all n ≥ 1. Note that if 0 < |B1| ≤ 1/n, then Tn = 1 and MTn ≡M1 is trivially

a martingale. Moreover, since B1 6= 0 almost surely, we will have that |B1| > 1/n for n large

enough, almost surely. Suppose therefore without loss of generality that |B1| > 1/n for all n.

Then

f(Bt∧Tn) = fn(Bt∧Tn) ∀t ≥ 1.

It follows that

Mt∧Tn = f(Bt∧Tn)− f(B1) + f(B1)

=

[f(Bt∧Tn)− f(B1)− 1

2

∫ t∧Tn

1∆f(Bs)ds

]+ f(B1)

=

[fn(Bt∧Tn)− fn(B1)− 1

2

∫ t∧Tn

1∆fn(Bs)ds

]+ fn(B1),

and so MTn = (Mt∧Tn)t≥1 is a martingale for all n ≥ 1.

It only remains to show that Tn ∞ almost surely as n→∞. Note that limn→∞ Tn exists

since the Tn’s are non-decreasing, so we only need to show that such limit is infinite. For each

R > 0 we let SR = inft ≥ 1 : |Bt| > R = inft ≥ 1 : Mt < 1/R, and note that SR +∞almost surely as R→∞. Then clearly

P(

limn→∞

Tn <∞)≤ P

(∃R : Tn < SR for all n

)= lim

R→∞limn→∞

P(Tn < SR).

To compute the latter probability note that by the OST we have

E(MTn∧SR) = E(M1) =: µ ∈ (0,∞).

On the other hand, we can also rewrite the left hand side as

nP(Tn < SR) +1

RP(SR ≤ Tn).

Using that P(SR ≤ Tn) = 1−P(Tn < SR), we can solve for P(Tn < SR) in the above to get that

P(Tn < SR) =µ− 1/R

n− 1/R→ 0 as n→∞.

We thus conclude that (Mt)t≥1 is:

• non-negative,

• a local martingale but not a martingale,

• a supermartingale, and

• L2 bounded.

In fact, we could have concluded that M is a supermartingale from the first 2 properties, as

the following result shows.

Page 13: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 13

Proposition 3.5. If X is a local martingale, and Xt ≥ 0 for all t ≥ 0, then X is a super-

martingale.

Proof. This follows from the conditional Fatou’s lemma. Indeed, if (Tn)n≥1 is a reducing

sequence for X, then for any s ≤ t we have

E(Xt|Fs) = E(

lim infn→∞

Xt∧Tn |Fs)≤ lim inf

n→∞E(Xt∧Tn |Fs) = lim inf

n→∞Xs∧Tn = Xs

almost surely.

Recall that our goal is to define integrals with respect to continuous martingales, such as

Brownian Motion. In fact, we will be able to integrate with respect to continuous local

martingales, with almost no additional effort. But why bother working with local martingales?

One good reason is that, for example, this frees us from worrying about integrability, which can

be difficult to check. Indeed, in Definition 3.3 the process X is only required to be integrable at

stopping times. Moreover, we will often encounter stochastic processes which are only defined

up to a random time τ . In this case it is not clear how to adapt the definition of martingale,

since the process might fail to be defined on the whole interval [0,∞). On the other hand, it’s

clear what the definition of local martingale on [0, τ) should be: simply ask that Tn τ as

n→∞ in Definition 3.3.

With this in mind, we ask how big is the class of continuous local martingales with respect to

the one of continuous martingales. We will then prove the main result of this section, stating

that the only continuous local martingales with finite variation are the constant ones.

Definition 3.6 (Uniform integrability, UI). A set X of random variables is said to be uniformly

integrable if

supX∈X

E(|X|1(|X| > λ))→ 0 as λ→∞.

Remark 3.7. Any family of uniformly bounded random variable is trivially UI. More generally,

if |X| ≤ Y for all X ∈ X and some integrable random variable Y , then X is UI.

Lemma 3.8. If X is in L1(Ω,F ,P), then the family

X = E(X|G) : G is a sub σ-algebra is UI.

Proof. See Example sheet 1.

We use the above result to prove a useful criterion to determine whether a local martingale is

in fact a martingale.

Proposition 3.9. The following statements are equivalent:

(i) X is a martingale,

Page 14: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

14 JASON MILLER AND VITTORIA SILVESTRI

(ii) X is a local martingale, and for all t ≥ 0 the family

Xt = XT : T is a stopping time with T ≤ t is UI.

Proof. Suppose that X is a martingale. By the OST, if T is a stopping time with T ≤ t, then

we have that XT = E(Xt | FT ). Lemma 3.8 then implies that Xt is UI.

Suppose, on the other hand, that X is a local martingale and the family Xt is UI for all t ≥ 0.

We prove that X is a true martingale by showing that for all bounded stopping times T we

have E(XT ) = E(X0). Indeed, let (Tn)n≥0 be a reducing sequence for X. Then, for all bounded

stopping times T ≤ t,

E(X0) = E(XTn0 ) = E(XTn

T ) = E(XT∧Tn).

As XT∧Tn : n ≥ 0 is UI, we know from Advanced Probability that XT∧Tn converges almost

surely and in L1 as n→∞, the limit being XT since Tn +∞. It follows that E(XT∧Tn)→E(XT ) as n → ∞. This shows that E(X0) = E(XT ) and, since T was arbitrary, that X is a

martingale.

Combining the above proposition with Remark 3.7 we obtain the following useful corollary.

Corollary 3.10. A bounded local martingale is a true martingale. More generally, if X is a

local martingale such that |Xt| ≤ Y for all t ≥ 0 and some integrable random variable Y , then

X is a true martingale.

We can finally turn to the main result of this section.

Theorem 3.11. Let X be a continuous local martingale with X0 = 0, and assume that X has

finite variation. Then X ≡ 0 almost surely.

Proof. Let V = (Vt)t≥0 denote the total variation process of X. Then V0 = 0 and V is

continuous, adapted and pathwise non-decreasing by Lemmas 2.2-2.5. Define the sequence of

stopping times

(3.1) Tn = inft ≥ 0 : Vt = n, n ≥ 1.

Then Tn ∞ since V is non-decreasing and finite by assumption. Moreover,

|XTnt | = |Xt∧Tn | ≤ |Vt∧Tn | ≤ n,

which shows that XTn is bounded, and hence a true martingale for all n ≥ 1. Thus (Tn)n≥1

reduces X. To show that X ≡ 0, it clearly suffices to show that XTn ≡ 0 for all n ≥ 1. Fix

any n ≥ 1 and write Y = XTn to shorten the notation. Then Y is a continuous, bounded

Page 15: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 15

martingale with Y0 = 0. By continuity, to see that Y ≡ 0 it suffices to show that E(Y 2t ) = 0 for

all t ≥ 0. Fix t > 0 and for N ≥ 1 write tk = kt/N for 0 ≤ k ≤ N . Then we have

E(Y 2t ) = E

(N−1∑k=0

(Y 2tk+1− Y 2

tk)

)= E

(N−1∑k=0

(Ytk+1− Ytk)2

)

≤ E(

sup0≤k<N

|Ytk+1− Ytk |︸ ︷︷ ︸

≤Vt≤n

·N−1∑k=0

|Ytk+1− Ytk |︸ ︷︷ ︸

≤Vt≤n

),

where the second equality follows from the orthogonality of increments, and the last from the

definition of the stopping time Tn. Now, by continuity of Y we have that

limN→∞

sup0≤k<N

|Ytk+1− Ytk | = 0

almost surely. We can therefore apply the Bounded Convergence Theorem to conclude that

E(

sup0≤k<N

|Ytk+1− Ytk | ·

N−1∑k=0

|Ytk+1− Ytk |

)→ 0 as N →∞,

and thus E(Y 2t ) = 0 for all t ≥ 0, which is what we wanted to show.

Remark 3.12. Note that:

(1) The continuity assumption cannot be removed, as we will see in the next section.

(2) The theorem shows that Brownian Motion has infinite variation, and hence the Lebesgue-

Stieltjes theory of integration cannot be used to define integrals with respect to it.

We conclude this section by observing that, although we were able to introduce an explicit

reducing sequence (3.1) for X in the proof of Theorem 3.11, this was based on the total

variation of X being finite, which fails to hold when dealing with non-constant continuous local

martingales. The next result provides an explicit reducing sequence for this class of processes.

Proposition 3.13. Let X be a continuous local martingale with X0 = 0. For n ≥ 1 define the

stopping times

Tn = inft ≥ 0 : |Xt| = n.Then the sequence (Tn)n≥1 reduces X.

Note that this is the same sequence of stopping times we used to show that Mt = 1/|Bt| is a

local martingale in Example 3.4.

Proof. To see that (Tn)n≥1 is a sequence of stopping times, note that for all t ≥ 0 we can write

Tn ≤ t =

sups∈[0,t]

|Xs| ≥ n

=⋂k≥1

⋃s≤ts∈Q

|Xs| > n− 1

k

∈ Ft,

Page 16: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

16 JASON MILLER AND VITTORIA SILVESTRI

which shows that Tn is a stopping time for al n. We now prove that Tn(ω)∞ for all ω ∈ Ω.

Indeed, fix any ω ∈ Ω. Since sups∈[0,t] |Xs(ω)| < ∞ by continuity of X, for any t ≥ 0 there

exists n(ω, t) finite such that

n(ω, t) > sups∈[0,t]

|Xs(ω)|,

which implies that Tn(ω) > t for all n ≥ n(ω, t), and hence that Tn(ω)→ +∞ as n→∞.

It only remains to show that (Tn)n≥1 reduces X. Let (T ∗n)n≥1 be a reducing sequence for X

(such sequence exists by assumption, since X is a local martingale). If Tn = T ∗n for all n there

is nothing to prove. Otherwise, we know that XT ∗n is a martingale for all n, and we need to

deduce that the same holds for XTn . Indeed, pick any n ≥ 1. Then by the OST the process

XT ∗n∧Tm is a martingale for all m, and hence XTn is a local martingale with reducing sequence

(T ∗m)m≥1. But XTn is also bounded by definition of Tn, and so it is a true martingale. Since n

was arbitrary, this shows that (Tn)n≥1 is a reducing sequence for X.

4. The stochastic integral

In this section we develop a new theory of integration, which will allow us to integrate against

continuous local martingales such as Brownian Motion or the process defined in Example 3.4.

Recall that a real-valued process X = (Xt)t≥0 is L2-bounded if supt≥0 ‖Xt‖L2 < ∞, or

equivalently if supt≥0 E(X2t ) <∞. The following two results are known to the reader from the

Advanced Probability course, and they will be repeatedly used in this section.

Theorem 4.1. Let X be a cadlag and L2-bounded martingale. Then there exists a random

variable X∞ such that

Xt → X∞ almost surely and in L2 as t→∞.

Moreover, Xt = E(X∞|Ft) a.s. for all t ≥ 0.

Proposition 4.2 (Doob’s L2 inequality). Let X be a cadlag and L2-bounded martingale. Then

E(

supt≥0|Xt|2

)≤ 4E(X2

∞).

Define the following sets:

M2 =L2 bounded, cadlag martingales

,

M2c =

L2 bounded, continuous martingales

,

M2c,loc =

L2 bounded, continuous local martingales

.

(4.1)

Page 17: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 17

Definition 4.3 (Simple process). A process H : Ω× (0,∞)→ R is said to be a simple process

if it writes as

H(ω, t) =

n−1∑k=0

Zk(ω)1(tk,tk+1](t)

for some n ∈ N, 0 = t0 < t1 · · · < tn <∞, and Zk bounded Ftk-measurable random variables

for all k. We let S denote the set of simple processes.

We will start by defining integrals of simple processes against L2 bounded, cadlag martingales

(H ∈ S, M ∈ M2). To extend this definition to more general integrands we will restrict to

continuous integrators (M ∈M2c), and equip M2

c with a Hilbert space structure. We will then

introduce the concept of quadratic variation, and use it to define a Hilbert space structure on

a larger space of integrands (depending on the integrator), which we call L2(M). Finally, we

will note that the stochastic integral defines a Hilbert space isometry between M2c and L2(M)

when restricted to simple integrands, and show that asking for this property to hold for all

integrands in L2(M) uniquely defines the stochastic integral on this larger space.

4.1. Simple integrands. In this section we define integrals of simple processes against L2

bounded cadlag martingales M ∈M2. Throughout, our integrals will be over the interval (0, t]

and will take the value 0 at 0.

Definition 4.4. For Ht =∑n−1

k=0 Zk1(tk,tk+1](t) simple process, and M ∈M2, set

(H ·M)t :=n−1∑k=0

Zk(Mtk+1∧t −Mtk∧t).

Note that this is a continuous time version of the discrete martingale transform defined in (1.1).

The next result show that the stochastic integral of a simple process against M ∈M2 is inM2.

Proposition 4.5. If H ∈ S and M ∈M2, then also H ·M ∈M2. Moreover,

(4.2) E[(H ·M)2∞] =

n−1∑k=0

E[Z2k(Mtk+1

−Mtk)2] ≤ 4‖H‖∞E[(M∞ −M0)2].

Proof. We first show that H · M is a martingale. For tk ≤ s ≤ t ≤ tk+1, we have that

(H ·M)t − (H ·M)s = Zk(Mt −Ms), so that

E[(H ·M)t − (H ·M)s | Fs] = ZkE[Mt −Ms | Fs] = 0,

Page 18: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

18 JASON MILLER AND VITTORIA SILVESTRI

as M is a martingale. This extends to general s ≤ t by the tower property. Indeed, take

tj ≤ s ≤ tj+1 ≤ tk ≤ t ≤ tk+1. Then we have

E[(H ·M)t − (H ·M)s|Fs] =

= E[ k−1∑i=0

Zi(Mti+1 −Mti) + Zk(Mt −Mtk)−j−1∑i=0

Zi(Mti+1 −Mti)− Zj(Ms −Mtj )∣∣∣Fs]

=k−1∑i=j+1

E[Zi(Mti+1 −Mti)|Fs] + E[Zj(Mtj+1 −Ms)|Fs] + E[Zk(Mt −Mtk)|Fs] = 0,

since

• E[Zi(Mti+1 −Mti)|Fs] = E[ZiE(Mti+1 −Mti |Fti)|Fs] = 0 for all j + 1 ≤ i ≤ k − 1,

• E[Zj(Mtj+1 −Ms)|Fs] = ZjE(Mtj+1 −Ms|Fs) = 0, and

• E[Zk(Mt −Mtk)|Fs] = E[ZkE(Mt −Mtk |Ftk)|Fs] = 0.

We thus conclude that H ·M is a martingale. To see that H ·M is bounded in L2, we observe

that if j < k then

E[Zj(Mtj+1 −Mtj )Zk(Mtk+1−Mtk)] = E[Zj(Mtj+1 −Mtj )ZkE[Mtk+1

−Mtk | Ftk ]] = 0.

Therefore

E[(H ·M)2t ] = E

(n−1∑k=0

Zk(Mtk+1∧t −Mtk∧t)

)2

=n−1∑k=0

E[Z2k(Mtk+1∧t −Mtk∧t)

2]

≤ ‖H‖2L∞n−1∑k=0

E[(Mtk+1∧t −Mtk∧t)2]

= ‖H‖2L∞E[(Mtn∧t −M0)2]

≤ 4‖H‖2L∞E[(M∞ −M0)2],

where the last line follows from Doob’s L2 inequality. Since this bound is uniform in t, we

conclude that H ·M is bounded in L2, and so H ·M ∈ M2. Finally, to see that (4.2) holds,

note that

E[(H ·M)2∞] = lim

t→∞E[(H ·M)2

t ] ≤ supt≥0

E[(H ·M)2t ] ≤ 4‖H‖2L∞E[(M∞ −M0)2]

by the previous inequality.

In what follows it will also be useful to understand how the stochastic integral defined above

behaves under stopping. The following result tells us that stopping the stochastic integral is

equivalent to stopping the martingale M .

Page 19: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 19

Proposition 4.6. Let H ∈ S and M ∈M2. Then for any stopping time T it holds

H ·MT = (H ·M)T .

Proof. We have that

(H ·MT )t =

n−1∑k=0

Zk(MTtk+1∧t −M

Ttk∧t)

=n−1∑k=0

Zk(MT∧tk+1∧t −MT∧tk∧t)

= (H ·M)t∧T = (H ·M)Tt

for any t ≥ 0.

4.2. The space of integrators. In this section we equip the space of integrators M2c with a

Hilbert space structure.

For X cadlag adapted process, define the norm

|||X||| = ‖X∗‖L2 where X∗ = supt≥0|Xt|,

and let

C2 =

cadlag adapted processes X with |||X||| <∞.Recall from (4.1) the definition of the sets M2, M2

c , M2c,loc, and define the following norm on

M2

‖X‖ := ‖X∞‖L2 .

This is clearly a seminorm. To see that it is a norm we have to show that ‖X‖ = 0 implies

X ≡ 0 almost surely. Indeed, if ‖X‖ = ‖X∞‖L2 = 0 then X∞ = 0 almost surely. Moreover, by

Theorem 4.1 we have

Xt = E(X∞|Ft) = 0 almost surely,

for all t ≥ 0. It then follows by continuity of X that X ≡ 0 almost surely.

Define further

M =

cadlag martingales,

Mc =

continuous martingales,

Mc,loc =

continuous local martingales.

(4.3)

Proposition 4.7. The following hold:

(a) (C2, ||| · |||) is complete.

(b) M2 =M∩ C2.

(c) (M2, ‖ · ‖) is a Hilbert space and M2c =Mc ∩M2 is a closed subspace.

Page 20: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

20 JASON MILLER AND VITTORIA SILVESTRI

(d) The map M2 → L2(F∞) given by X 7→ X∞ is an isometry.

Remark 4.8. We can identify an element of M2 with its terminal value and then (M2, ‖ · ‖)inherits the Hilbert space structure of (L2(F∞), ‖ · ‖L2). Moreover, since (M2

c , ‖ · ‖) is a closed

linear subspace of M2 by (c), it is itself a Hilbert space: this will be our space of integrators.

Proof. To prove (a) we have to show that any Cauchy sequence in (C2, ||| · |||) converges to an

element of C2. To this end, we will first show that a cadlag subsequential limit exists for any

such sequence, and then prove that in fact the whole sequence converges to such limit in ||| · |||norm.

Let (Xn)n≥1 be a Cauchy sequence in (C2, ||| · |||). Then we can find a subsequence (Xnk)k≥1

such that∑

k |||Xnk+1 −Xnk ||| <∞. It follows that∥∥∥∥∥∑k

supt≥0|Xnk+1

t −Xnkt |

∥∥∥∥∥L2

≤∑k

|||Xnk+1 −Xnk ||| <∞.

Therefore∑

k supt≥0 |Xnk+1

t −Xnkt | <∞ almost surely, and hence supt≥0 |X

nk+1

t −Xnkt | → 0 as

k →∞ almost surely. This shows that (Xnk(ω))k≥1 converges uniformly on [0,∞) as k →∞to some limit, which must then be cadlag as uniform limit of cadlag functions.

Let X denote such limit: it remains to show that Xn → X in ||| · ||| norm. Indeed,

|||Xn −X|||2 = E(

supt≥0|Xn

t −Xt|2)

≤ E(

limk→∞

supt≥0|Xn

t −Xnkt |2

)≤ lim inf

k→∞E(

supt≥0|Xn

t −Xnkt |2

)(Fatou’s lemma)

= lim infk→∞

|||Xn −Xnk |||2 → 0 as n→∞

as (Xn) is Cauchy. This proves (a).

To see (b) we show that both inclusions hold. Suppose that X ∈ C2 ∩M. Then |||X||| <∞, and

so

supt≥0‖Xt‖L2 ≤

∥∥∥∥ supt≥0|Xt|

∥∥∥∥L2

= |||X|||.

Therefore X ∈M2. On the other hand, if X ∈M2, by Doob’s L2-inequality we find

|||X||| ≤ 2‖X∞‖L2 <∞,

so X ∈ C2 ∩M. Hence M2 =M∩ C2, which proves (b).

We now prove (c).

Page 21: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 21

Note that (X,Y ) 7→ E(X∞Y∞) defines an inner product on M2. For X ∈ M2, we have just

shown that

‖X‖ ≤ |||X||| ≤ 2‖X‖,

so ‖ · ‖ and ||| · ||| are equivalent norms. Therefore, in order to show that (M2, ‖ · ‖) is complete

we can equivalently show that (M2, ||| · |||) is complete. By (a), it suffices to show that M2 is

closed in (C2, ||| · |||). To see this, let (Xn)n≥1 be a sequence in M2 such that |||Xn −X||| → 0 as

n→∞ for some X. Then X is cadlag , adapted and bounded in L2. To see that X is also a

martingale, note that for s < t∥∥E(Xt | Fs)−Xs

∥∥L2 ≤

∥∥E(Xt −Xnt | Fs) +Xn

s −Xs

∥∥L2

≤∥∥E(Xt −Xn

t | Fs)∥∥L2 +

∥∥Xns −Xs

∥∥L2 (Minkowski’s inequality)

≤∥∥Xt −Xn

t

∥∥L2 +

∥∥Xns −Xs

∥∥L2

≤ 2|||Xn −X||| → 0 as n→∞.

Therefore X ∈M2, which is therefore closed. Exactly the same argument applies to show that

M2c is a closed subspace of (M2, ‖ · ‖). This proves (c).

Part (d) follows from the definition.

In this section we have equipped our space of integrators with a Hilbert space structure. In the

next section we do the same for a suitable space of integrands.

4.3. The space of integrands. The crucial tool that will allow us to equip the space of

integrands with a Hilbert space structure is the so-called quadratic variation, that we now

discuss in detail.

Definition 4.9 (Ucp convergence). Let (Xn)n≥1 be a sequence of processes. We say that

Xn → X uniformly on compacts in probability (ucp) if for every ε, t > 0 we have that

P(

sups≤t|Xn

s −Xs| > ε)→ 0 as n→∞.

The following result defines the quadratic variation of a continuous local martingale.

Theorem 4.10 (Quadratic variation). Let M be a continuous local martingale. Then there

exists a unique (up to indistinguishability) continuous, adapted and non-decreasing process [M ]

such that [M ]0 = 0 and M2 − [M ] is a continuous, local martingale. Moreover, if we define

[M ]nt =

d2nte−1∑k=0

(M(k+1)2−n −Mk2−n)2

then [M ]n → [M ] ucp as n→∞. The process [M ] is called the quadratic variation of M .

Page 22: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

22 JASON MILLER AND VITTORIA SILVESTRI

Proof of Theorem 4.10. By replacing M with Mt −M0 if necessary, we may assume without

loss of generality that M0 = 0.

Step 1: Uniqueness. If A and A′ are two increasing processes satisfying the conditions of the

theorem, then we can write

At −A′t = (M2t −A′t)− (M2

t −At).

The left hand side is a continuous process of bounded variation as it is given by the difference of

two non-decreasing functions while the right hand side is a continuous local martingale as it is

given by the difference of two continuous local martingales. Since a continuous local martingale

starting from 0 with bounded variation is almost surely equal to 0, it follows that A−A′ ≡ 0.

That is, A ≡ A′, as desired.

Step 2: Existence, M bounded. We shall first consider the case that M is a bounded, continuous

martingale. Then we have that M ∈ M2c . We first build the quadratic variation process on

compact time sets [0, T ], for T deterministic and finite. Fix such T > 0 and let

Hnt =

d2nT e−1∑k=0

Mk2−n1(k2−n,(k+1)2−n](t).

Note that Hn approximates MT for large n. Then, since Hn is a simple process for all n, we

have that

Xnt = (Hn ·M)t =

d2nT e−1∑k=0

Mk2−n(M(k+1)2−n∧t −Mk2−n∧t)

is an L2-bounded continuous martingale, that is Xn ∈ M2c for all n. We next show that the

sequence (Xn)n≥1 is Cauchy in ‖ · ‖ norm, and therefore it has a limit in M2c . Pick n ≥ m ≥ 1

and write H in place of Hn −Hm to shorten the notation. Then

Xn −Xm = (Hn −Hm) ·M = H ·M,

Page 23: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 23

and so

‖Xn −Xm‖2 = E[(H ·M)2∞] = E[(H ·M)2

T ]

= E[( d2nT e−1∑

k=0

Hk2−n(M(k+1)2−n −Mk2−n)

)2]

= E[ d2nT e−1∑

k=0

H2k2−n(M(k+1)2−n −Mk2−n)2

]

≤ E[

supt∈[0,T ]

|Ht|2d2nT e−1∑k=0

(M(k+1)2−n −Mk2−n)2

]

≤ E[(

supt∈[0,T ]

|Ht|2)2]1/2

E[( d2nT e−1∑

k=0

(M(k+1)2−n −Mk2−n)2

)2]1/2

= E[

supt∈[0,T ]

|Hnt −Hm

t |4]1/2

E[( d2nT e−1∑

k=0

(M(k+1)2−n −Mk2−n)2

)2]1/2

,

where we have used Cauchy-Schwartz in the last inequality. Note that the first term is bounded

by

supt∈[0,T ]

|Hnt −Hm

t |4 ≤ 16‖M‖4L∞

since M is bounded, and

supt∈[0,T ]

|Hnt −Hm

t |4 → 0 almost surely as n,m→∞

by uniform continuity of M on [0, T ]. It therefore follows from the Bounded Convergence

Theorem that

E[

supt∈[0,T ]

|Hnt −Hm

t |4]1/2

→ 0

as n,m→∞. We claim that the second term is bounded since M is.

Lemma 4.11. Let M ∈M be bounded. Then, for any N ≥ 1 and 0 = t0 ≤ t1 ≤ . . . ≤ tN <∞we have that

E[(N−1∑

k=0

(Mtk+1−Mtk)2

)2]≤ 48‖M‖4L∞ .

Proof. Note that

E[(N−1∑

k=0

(Mtk+1−Mtk)2

)2]=

=

N−1∑k=0

E[(Mtk+1−Mtk)4] + 2

N−1∑k=0

E[(Mtk+1

−Mtk)2N−1∑j=k+1

(Mtj+1 −Mtj )2

].

(4.4)

Page 24: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

24 JASON MILLER AND VITTORIA SILVESTRI

For each fixed k, we have

E[(Mtk+1

−Mtk)2N−1∑j=k+1

(Mtj+1 −Mtj )2

]=

= E[(Mtk+1

−Mtk)2 E( N−1∑j=k+1

(Mtj+1 −Mtj )2∣∣∣Ftk+1

)]

= E[(Mtk+1

−Mtk)2 E[( N−1∑

j=k+1

(Mtj+1 −Mtj )

)2∣∣∣Ftk+1

)]= E

[(Mtk+1

−Mtk)2E[(MtN −Mtk+1)2|Ftk+1

]]

= E[(Mtk+1

−Mtk)2(MtN −Mtk+1)2]

Inserting this into (4.4) we then get

E[(N−1∑

k=0

(Mtk+1−Mtk)2

)2]≤

≤ E[(

sup0≤j≤N−1

|Mtj+1 −Mtj |2 + 2 sup0≤j≤N−1

|MtN −Mtj |2)N−1∑

k=0

(Mtk+1−Mtk)2

]

≤ 12‖M‖2L∞ E[N−1∑k=0

(Mtk+1−Mtk)2

]= 12‖M‖2L∞ E[(MtN −Mt0)2]

≤ 48‖M‖4L∞ .

Going back to the proof of the theorem, this implies that

‖Xn −Xm‖2 → 0 as n,m→∞,

and so (Xn)n≥1 is Cauchy in (M2c , ‖ ·‖). It follows that there exists Y ∈M2

c such that Xn → Y

in ‖ · ‖ norm as n→∞.

Now, for any n and 1 ≤ k ≤ d2nT e, we can write

M2k2−n − 2Xn

k2−n =k−1∑j=0

(M(j+1)2−n −Mj2−n)2 = [M ]nk2−n .

Consequently, for each n ≥ 1 the process M2 − 2Xn is non-decreasing when restricted to the

set of times k2−n : 1 ≤ k ≤ d2nT e. To show that the same holds for the limiting process

M2 − 2Y it suffices to prove that Xn → Y almost surely along a subsequence, which is true by

the equivalence of the norms ‖ · ‖ and ||| · |||.

Page 25: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 25

Set

[M ]t = M2t − 2Yt.

Then [M ] = ([M ]t)t≤T is a continuous, adapted non-decreasing process and M2 − [M ] = 2Y is

a martingale on [0, T ].

We can extend this definition to all times in [0,∞) by applying the above for each T = k ∈ N.

By uniqueness, we note that the process obtained with T = k must be the restriction to [0, k]

of the process obtained with T = k + 1.

Finally we show that [M ]n → [M ] ucp as n → ∞. To this end, note that the convergence

Xn → Y in (M2c , ‖ · ‖) implies, by the equivalence of ‖ · ‖ and ||| · |||, that sup0≤t≤T |Xn

t −Yt| → 0

in L2 as n→∞, and so in particular

(4.5) sup0≤t≤T

|Xnt − Yt| → 0 in probability as n→∞.

Now,

[M ]nt = M22−nb2ntc − 2Xn

2−nb2ntc

and so

sup0≤t≤T

∣∣[M ]t − [M ]nt∣∣ ≤ sup

0≤t≤T

∣∣M22−nb2ntc −M

2t

∣∣+ 2 sup

0≤t≤T

∣∣Xn2−nb2ntc − Y2−nb2ntc

∣∣+ 2 sup

0≤t≤T

∣∣Y2−nb2ntc − Yt∣∣.

Each term in the right hand side converges to 0 in probability as n→∞. Indeed, this follows

for the first and third terms by uniform continuity of M2 and Y on [0, T ], and for the second

term by (4.5). Thus [M ]n → [M ] ucp as n→∞.

Step 3: Existence, M ∈Mc,loc. We now obtain existence for the general case M ∈Mc,loc by

mean of a localisation argument. For each n ≥ 1, define

Tn = inft ≥ 0 : |Mt| ≥ n.

Then (Tn)n≥1 reduces M and moreover MTn is a bounded martingale by construction for all n.

Thus we can apply Step 2 to conclude that there exists a unique continuous, adapted and non-

decreasing process [MTn ] on [0,∞) such that [MTn ]0 = 0 and (MTn)2 − [MTn ] ∈Mc,loc. Write

An = [MTn ] for simplicity. By uniqueness, (An+1t∧Tn)t≥0 and (Ant )t≥0 must be indistinguishable.

We thus let A be the non-decreasing process such that At∧Tn = Ant for all n. By construction,

M2t∧Tn −At∧Tn is a martingale for each n. Therefore M2−A is a local martingale with reducing

sequence (Tn)n≥1, and we can take [M ]t = At.

Page 26: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

26 JASON MILLER AND VITTORIA SILVESTRI

Finally, to see that [M ]n → [M ] ucp as n→∞, note that [MTk ]n → [MTk ] as n→∞ ucp by

Step 2, for all k, that is for all ε > 0 it holds

(4.6) P(

supt∈[0,T ]

∣∣[MTk ]nt − [MTk ]t∣∣ > ε

)→ 0 as n→∞.

Observing that on the event Tk > T we have [M ]nt = [MTk ]nt and [M ]t = [MTk ]t for all t ≤ T ,

we have that

P(

supt∈[0,T ]

∣∣[M ]nt − [M ]t∣∣ > ε

)≤ P(Tk ≤ T ) + P

(supt∈[0,T ]

∣∣[MTk ]nt − [MTk ]t∣∣ > ε

)→ 0

as n→∞, since P(Tk ≤ T )→ 0 as k →∞, and the second term in the r.h.s. converges to 0 as

n→∞ for all k by (4.6).

Example 4.12. Let B be a standard Brownian motion. Then (B2t − t)t≥0 is a martingale.

Therefore [B]t = t. As we will see later, the fact that [B]t = t singles out the standard Brownian

motion among continuous local martingales. This is the so-called Levy characterization of

Brownian motion.

The following result will be useful later on, and it shows that the a priori local martingale

M2 − [M ] is in fact a true UI martingale if M ∈M2c .

Theorem 4.13. If M ∈M2c , then M2 − [M ] is a uniformly integrable martingale.

Proof. We show that the martingale is dominated by an integrable random variable.

Let Sn = inft ≥ 0 : [M ]t ≥ n for n ≥ 1. Then Sn +∞ and Sn is a stopping time for all

n, with [M ]t∧Sn ≤ n. Therefore the local martingale M2t∧Sn

− [M ]t∧Sn is dominated by the

random variable n+ supt≥0M2t , which is integrable by Doob’s L2 inequality. It then follows

from Corollary 3.10 that M2t∧Sn

− [M ]t∧Sn is a true martingale. The OST then gives

E[[M ]t∧Sn ] = E[M2t∧Sn

].

Letting t→∞ and using the Monotone Convergence Theorem on the l.h.s. and the Dominated

Convergence Theorem on the r.h.s., we conclude that

E[[M ]Sn ] = E[M2Sn

].

Sending n→∞ and using the same theorems to exchange limits and expectation, we find that

E[[M ]∞] = E[M2∞] <∞,

thus showing that [M ]∞ is integrable. Moreover,

|M2t − [M ]t| ≤ sup

t≥0M2t + [M ]∞ ∀t ≥ 0,

and the r.h.s. is integrable, so we conclude that M2t − [M ]t is a true UI martingale.

Page 27: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 27

4.3.1. The space L2(M). We can finally define our space of integrands. Recall that P denotes

the previsible σ-algebra on Ω× (0,∞). Given M ∈M2c , we define a measure µ on P by setting

µ(E × (s, t]) = E(1(E)([M ]t − [M ]s)

)for s ≤ t, E ∈ Fs.

Since P is generated by a π-system of events of the form E × (s, t] for E, s, t as above, this

uniquely specifies µ. Note that this is equivalent to defining

µ(dω ⊗ dt) := d[M ](ω, dt)P(dω),

where, for fixed ω ∈ Ω, the measure d[M ](ω, ·) is the Lebesgue-Stieltjes measure associated to

the non-decreasing function [M ](ω). Then for a previsible process H ≥ 0 we have that∫Ω×(0,∞)

Hdµ = E(∫ ∞

0Hsd[M ]s

).

Definition 4.14. Let L2(M) := L2(Ω× (0,∞),P, µ) and write

‖H‖L2(M) = ‖H‖M :=

[E(∫ ∞

0H2sd[M ]s

)]1/2

.

Then L2(M) is the set of previsible processes H such that ‖H‖M <∞. This is clearly a Hilbert

space, and it will be our space of integrands.

Remark 4.15. Note that the space of integrands (L2(M), ‖ · ‖M ) depends on the integrator

M ∈M2c , since the measure µ depends on it. On the other hand, recalling that S denotes the

set of simple processes, we have that S ⊆ L2(M) for all M ∈M2c .

4.4. Ito integrals. Up to now, we have identified two Hilbert spaces, that of integrators

(M2c , ‖ · ‖) and that of integrands (L2(M), ‖ · ‖M ). For a subset S of L2(M), that of simple

processes of the form Ht =∑n−1

k=0 Zk1(tk,tk+1](t), we have defined the integral H ·M by setting

(H ·M)t =

n−1∑k=0

Zk(Mtk+1∧t −Mtk∧t) ∈M2c .

We now argue that for any M ∈ M2c the map H 7→ H ·M provides an isometry between

(L2(M), ‖ · ‖M ) and (M2c , ‖ · ‖) when restricted to simple processes S ⊆ L2(M). To see this, we

have to show that ‖H ·M‖ = ‖H‖M for all H ∈ S, M ∈M2c . Indeed, recall from (4.2) that

‖H ·M‖2 =∥∥(H ·M)∞

∥∥2

L2 =

n−1∑k=0

E[Z2k(Mtk+1

−Mtk)2].

Page 28: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

28 JASON MILLER AND VITTORIA SILVESTRI

Moreover, by Theorem 4.13 the process M2 − [M ] is a true martingale, so that

E[Z2k(Mtk+1

−Mtk)2]

= E[Z2k E((Mtk+1

−Mtk)2 | Ftk)]

= E[Z2k E(M2tk+1−M2

tk| Ftk

)]= E

[Z2k E([M ]tk+1

− [M ]tk | Ftk)]

= E[Z2k ([M ]tk+1

− [M ]tk)].

Therefore

(4.7)∥∥H ·M∥∥2

= E( n−1∑k=0

Z2k ([M ]tk+1

− [M ]tk)

)= E

(∫ ∞0

H2sd[M ]s

)= ‖H‖2M ,

thus showing that the stochastic integral provides an isometry between L2(M) and M2c when

restricted to simple processes. The following theorem is the main result of this section, and it

tells us that this property uniquely defines the stochastic integral on the whole L2(M).

Theorem 4.16 (Ito’s isometry). There exists a unique isometry I : L2(M)→M2c such that

I(H) = H ·M for all H ∈ S.

Definition 4.17. For any M ∈M2c and H ∈ L2(M), let

H ·M := I(H),

where I is the unique isometry from (L2(M), ‖ · ‖M ) to (M2c , ‖ · ‖) of Theorem 4.16.

We now proceed with the proof of Theorem 4.16 above. To start with, we show that simple

processes are dense in L2(M), so that there is hope to extend the definition of the stochastic

integral from S to the whole L2(M) uniquely.

Lemma 4.18. Let ν be a finite measure on P, the previsible σ-algebra. Then S is a dense

subspace of L2(P, ν). In particular, taking ν = µ we have that S is a dense subspace of L2(M)

for any M ∈M2c .

Proof. It is certainly the case that S ⊆ L2(P, ν). Denote by S the closure of S in L2(P, ν).

We aim to show that S = L2(P, ν). Let A = A ∈ P : 1A ∈ S. Then clearly A ⊆ P. To

see the reverse inclusion, observe that A is a d-system containing the π-system E × (s, t] :

E ∈ Fs, s ≤ t, which generates P. So, by Dynkin’s lemma, we have that A ⊇ P and hence

A = P. This shows that S contains indicator functions of all previsible processes. The result

then follows because finite linear combinations of such indicators are dense in L2(P, ν).

We can now prove Ito’s isometry.

Page 29: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 29

Proof of Theorem 4.16. Take any H ∈ L2(M). Then by Lemma 4.18 there exists a sequence

(Hn)n≥1 in S ⊆ L2(M) such that ‖Hn −H‖M → 0 as n→∞, that is

E(∫ ∞

0(Hn

s −Hs)2d[M ]s

)→ 0 as n→∞.

In particular, (Hn)n≥1 is Cauchy in ‖ · ‖M norm. We argue that this implies that (I(Hn))n≥1

is Cauchy in ‖ · ‖ norm. Indeed, by linearity of the integral for simple processes we have

‖I(Hn)− I(Hm)‖ = ‖Hn ·M −Hm ·M‖ = ‖(Hn −Hm) ·M‖ = ‖Hn −Hm‖M → 0

as n,m→∞, where the last equality follows from (4.7). Thus I(Hn) = Hn ·M converges in

‖ · ‖ norm to an element of M2c , that we define to be I(H). Note that this definition does not

depend on the choice of the approximating sequence. Indeed, if (Kn)n≥1 is another sequence of

simple processes converging to H in ‖ · ‖M norm, then

‖I(Hn)− I(Kn)‖ = ‖Hn ·M −Kn ·M‖ = ‖Hn −Kn‖M ≤ ‖Hn −H‖M + ‖Kn −H‖M → 0

as n→∞, thus showing that the limits of Hn ·M and Kn ·M must be indistinguishable. This

uniquely extends I to the whole L2(M). It only remains to show that the resulting map is an

isometry from L2(M) to M2c . Indeed, since convergence in norm implies convergence of the

norms by the reverse triangle inequality2, we have

‖H ·M‖ = limn→∞

‖Hn ·M‖ = limn→∞

‖Hn‖M = ‖H‖M

for all H ∈ L2(M), as required.

We write

I(H)t = (H ·M)t =

∫ t

0HsdMs.

The process H ·M is Ito’s stochastic integral of H with respect to M .

Remark 4.19. Recall from the introduction that we were seeking to define∫HdM in such a

way that if Hn is a sequence of simple processes converging to H, then∫HndM →

∫HdM .

The above definition matches this requirement by construction, where the convergence Hn → H

is to be intended in ‖ ·‖M norm, and the one of the integrals∫HndM → HdM is to be intended

in ‖ · ‖ norm.

2|‖x‖ − ‖y‖| ≤ ‖x− y‖.

Page 30: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

30 JASON MILLER AND VITTORIA SILVESTRI

4.5. Extensions. We now extend the definition of stochastic integral to a larger class of

integrators and integrands. To this end, we first need to understand how the stochastic integral

behaves under stopping.

Proposition 4.20. Let M ∈M2c and H ∈ L2(M). Let T be a stopping time. Then

(H ·M)T = (H1(0,T ]) ·M = H · (MT ).

Proof. It is easy to check that H ∈ L2(M) implies that also H1(0,T ] ∈ L2(M) and H ∈ L2(MT )

so that all the integrals above are well defined. We split the proof into several steps.

Case 1. Fix H ∈ S, M ∈M2c , T taking only finitely many values. Then one can check directly

from the definition that H1(0,T ] ∈ S and (H ·M)T = (H1(0,T ]) ·M = H · (MT ).

Case 2. Fix H ∈ S, M ∈ M2c , T a general stopping time. Then we have already shown that

(H ·M)T = H · (MT ) in Proposition 4.6. To see that (H ·M)T = (H1(0,T ]) ·M we reduce to

Case 1 by approximating T . For n,m ≥ 1, we let Tn,m = (2−nd2nT e) ∧m. Then Tn,m takes on

only finitely many values and Tn,m T ∧m as n→∞. Thus,

‖H1(0,Tn,m] −H1(0,T∧m]‖2M = E(∫ ∞

0H2t 1(T∧m,Tn,m]d[M ]t

)→ 0 as n→∞

by the Dominated Convergence Theorem. Since convergence of the integrands implies conver-

gence of the integrals, we get that

(H1(0,Tn,m]) ·M → (H · 1(0,T∧m]) ·M in M2c as n→∞.

On the other hand, the l.h.s. equals (H ·M)Tn,m by Case 1, and

(H ·M)Tn,m → (H ·M)T∧m pointwise almost surely as n→∞

by continuity of H ·M . It then follows from uniqueness of the limit3 that

(H · 1(0,T∧m]) ·M = (H ·M)T∧m

for all m ≥ 1. Sending now m→∞ and using the same argument as above we conclude that

(H1(0,T ]) ·M = (H ·M)T .

Case 3. H ∈ L2(M), M ∈ M2c , T a general stopping time. Let (Hn)n≥1 be a sequence in S

such that Hn → H in L2(M). Then

‖(Hn ·M)T − (H ·M)T ‖ = ‖(Hn ·M)T − (H ·M)T ‖L2

≤∥∥∥ supt≥0|(Hn ·M)t − (H ·M)t|

∥∥∥L2

≤ 2‖(Hn ·M)∞ − (H ·M)∞‖L2 (Doob’s L2 inequality)

= ‖(Hn −H) ·M‖ = ‖Hn −H‖M → 0

3Recall from the proof of Proposition 4.7 that convergence in ‖ · ‖ is equivalent to convergence in ||| · |||, which

in turn implies almost sure uniform convergence of a subsequence.

Page 31: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 31

as n→∞, that is

(4.8) (Hn ·M)T → (H ·M)T in M2c .

On the other hand,

‖Hn1(0,T ] −H1(0,T ]‖2M = E(∫ T

0(Hn

t −Ht)2d[M ]t

)≤ ‖Hn −H‖2M → 0 as n→∞.

Thus

(Hn1(0,T ]) ·M → (H1(0,T ]) ·M in M2c .

Since (Hn ·M)T = (Hn1(0,T ]) ·M for all n, this shows that (H ·M)T = (H1(0,T ]) ·M . Finally,

to see that (H ·M)T = H · (MT ) it suffices to show that if Hn → H in L2(M) then also

Hn → H in L2(MT ). Indeed,

‖Hn −H‖2MT = E(∫ ∞

0(Hn

s −Hs)2d[MT ]s

)= E

(∫ T

0(Hn

s −Hs)2d[M ]s

)≤ ‖Hn −H‖2M → 0 as n→∞.

Thus Hn · (MT )→ H · (MT ) in M2c . This together with (4.8) gives (H ·M)T = H · (MT ).

We use Proposition 4.20 to extend the definition of Ito’s integral.

Definition 4.21 (Locally bounded process). We say that a previsible process H is locally

bounded if there exists a sequence (Sn)n≥1 of stopping times with Sn ∞ almost surely as

n→∞ such that H1(0,Sn] is bounded for all n.

Definition 4.22 (Ito’s integral, H locally bounded, M ∈Mc,loc). Let H be a locally bounded

previsible process such that H1(0,Sn] is bounded for all n, for (Sn)n≥1 sequence of stopping

times with Sn ∞ almost surely. Let moreover M be a continuous local martingale with

reducing sequence (S′n)n≥1 given by S′n = inft ≥ 0 : |Mt| ≥ n, so that MS′n ∈ M2c for all n.

Set Tn := Sn ∧ S′n for all n ≥ 1, and define

(4.9) (H ·M)t := ((H1(0,Tn]) ·MTn)t for t ≤ Tn.

Note that the r.h.s. in the above definition is well defined, since if M ∈M2c then any bounded

previsible process is in L2(M), and clearly MTn = (MS′n)Tn ∈ M2c . Moreover, it is easy to

check that the definition of the integral does not depend on the choice of the reducing sequence

(Sn)n≥1 for H, and it is consistent with our previous definitions of Ito’s integral.

Page 32: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

32 JASON MILLER AND VITTORIA SILVESTRI

Proposition 4.23. Let M ∈ Mc,loc and let H be a locally bounded previsible process. Then

H ·M ∈Mc,loc and (Tn)n≥1 defined as in (4.9) is a reducing sequence. Moreover, for any T

stopping time we have

(H ·M)T = (H1(0,T ]) ·M = H · (MT ).

Proof. The fact that H ·M ∈Mc,loc with reducing sequence (Tn)n≥1 follows from the definition,

since the r.h.s. of (4.9) is in M2c for all n. Moreover,

(H ·M)T = limn→∞

((H1(0,Tn]) ·MTn)T

= limn→∞

((H1(0,Tn]1(0,T ]) ·MTn)

= (H1(0,T ]) ·M

where the limits hold pointwise almost surely, and the second equality follows from Proposition

4.20. The same argument shows that (H ·M)T = H · (MT ).

The next result allows us to express the quadratic variation of the integral in terms of the

quadratic variation of the integrator.

Proposition 4.24. Let M ∈Mc,loc and let H be a locally bounded previsible process. Then

[H ·M ] = H2 · [M ].

Proof. Assume first that H,M are uniformly bounded. Then we have that

E[(H ·M)2T ] = E[((H1(0,T ]) ·M)2

∞]

= E[((H21(0,T ]) · [M ])∞] (Ito’s isometry)

= E[(H2 · [M ])T ]

for all bounded stopping times T . By the OST, we thus have that (H ·M)2 −H2 · [M ] is a

continuous martingale. On the other hand, the unique non-decreasing, adapted, continuous

process X such that (H ·M)2 −X is a martingale is [H ·M ]. Therefore it must be [H ·M ] =

H2 · [M ].

We now treat the general case H locally bounded by mean of a localization argument. Let

(Tn)n≥1 be a sequence of stopping times such that H1(0,Tn] is uniformly bounded for all n.

Then we have that

[H ·M ] = limn→∞

[H ·M ]Tn (pointwise, almost surely)

= limn→∞

[(H ·M)Tn ]

= limn→∞

[(H1(0,Tn]) ·M ]

= limn→∞

(H1(0,Tn])2 · [M ] (H1(0,Tn] uniformly bounded)

= H2 · [M ] (pointwise, almost surely)

Page 33: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 33

where the last equality holds by the Monotone Convergence Theorem.

If H is a locally bounded previsible process, and M ∈Mc,loc, then we have seen that H ·M ∈Mc,loc. It follows that we can iterate the integral.

Proposition 4.25 (Iterated integral). Let M ∈Mc,loc and let H,K be locally bounded previsible

processes. Then

(4.10) H · (K ·M) = (HK) ·M.

Proof. The fact that (4.10) holds for H,K simple is an elementary exercise. For general H,K,

assume first that both H and K are uniformly bounded. Let us first show that both integrals

are well defined. We have

‖H‖2L2(K·M) = E[(H2 · [K ·M ])∞] = E[(H2 · (K2 · [M ]))∞]

= E[((HK)2 · [M ])∞] = ‖HK‖2L2(M)

≤ min‖H‖2L∞‖K‖2L2(M); ‖K‖

2L∞‖H‖2L2(M)

,

where the third equality follows from the corresponding property for the Lebesgue-Stieltjes

integral. Thus both H ∈ L2(K ·M) and HK ∈ L2(M), and the integrals are well defined. Note

that incidentally we have shown that

(4.11) ‖H‖L2(K·M) = ‖HK‖L2(M),

a property that we will use later on. Let (Hn)n≥1 and (Kn)n≥1 be sequences of simple processes

converging to H and K respectively in L2(M), and assume that ‖Hn‖L∞ and ‖Kn‖L∞ are

uniformly bounded over all n ≥ 1 without loss of generality. We have already observed that

Hn · (Kn ·M) = (HnKn) ·M . Moreover,

‖Hn · (Kn ·M)−H · (K ·M)‖ ≤ ‖(Hn −H) · (Kn ·M)‖+ ‖H · ((Kn −K) ·M)‖

= ‖Hn −H‖L2(Kn·M) + ‖H‖L2((Kn−K)·M) (Ito’s isometry)

= ‖(Hn −H)Kn‖L2(M) + ‖H(Kn −K)‖L2(M)

≤ ‖Hn −H‖L2(M)‖Kn‖L∞ + ‖H‖L∞‖Kn −K‖L2(M) → 0

as n → ∞, where the second equality follows from (4.11). A similar argument implies that

(HnKn) ·M → (HK) ·M inM2c as n→∞, thus showing that H · (K ·M) = (HK) ·M . Finally,

for H,K locally bounded we again use a localization argument. Let (Tn)n≥1 be a sequence of

stopping times such that both H1(0,Tn] and K1(0,Tn] are bounded for all n (it is always possible

to construct such sequence by setting Tn = Sn ∧ S′n, where (Sn)n≥1 and (S′n)n≥1 are sequences

of stopping times making H and K uniformly bounded respectively). Then by the previous

point we have

(H1(0,Tn]) · ((K1(0,Tn]) ·M) = (HK1(0,Tn]) ·M

Page 34: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

34 JASON MILLER AND VITTORIA SILVESTRI

for all n. Moreover, (K1(0,Tn]) ·M = (K ·M)Tn , from which

(H1(0,Tn]) · ((K1(0,Tn]) ·M) = (H1(0,Tn]) · (K ·M) = (H · (K ·M))Tn → H · (K ·M)

pointwise almost surely as n→∞. Finally,

(HK1(0,Tn]) ·M = ((HK) ·M)Tn → (HK) ·M

pointwise almost surely as n→∞. We thus conclude that H · (K ·M) = (HK) ·M .

Remark 4.26. In both the proof of Proposition 4.24 and the one of Proposition 4.25 we have

used a so-called localization argument to reduce the case of locally bounded integrands to the

one of uniformly bounded integrands. This is a standard procedure that we have chosen to spell

out in the two proofs above, but will be left for the reader to check in the rest of these notes.

4.5.1. Semimartingales. As a last extension of the Ito integral we enlarge the class of integrators

to the so-called semimartingales.

Definition 4.27 (Semimartingale). A continuous adapted process X is called a semimartingale

if it can be written in the form

X = X0 +M +A

where M is a continuous local martingale, A is a process of finite variation and M0 = A0 = 0.

This is called the Doob-Meyer decomposition of X.

For X = X0 +M +A continuous semimartingale define its quadratic variation to be that of its

martingale part, i.e. [X] := [M ]. This definition is justified by the fact that

d2nte−1∑k=0

(X(k+1)2−n −Xk2−n)2 → [X]t ucp as n→∞,

which is easy to check. Then we can easily extend the definition of Ito’s integral by linearity,

since we know how to integrate both with respect to M (Ito’s theory) and with respect to A

(Lebesgue-Stieltjes theory).

Definition 4.28. For H locally bounded previsible process, and X = X0 +M +A continuous

semimartingale, define

H ·X := H ·M +H ·A.

Note that H ·X defined above is again a continuous semimartingale, since H ·M ∈Mc,loc and

H ·A is of finite variation.

The next result shows that, under the additional assumption that H is left-continuous, the

integral can be obtained as the limit of its Riemann sum approximations.

Page 35: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 35

Proposition 4.29. Let X be a continuous semimartingale and H be a locally bounded left-

continuous process which is adapted. Then

(4.12)

b2ntc−1∑k=0

Hk2−n(X(k+1)2−n −Xk2−n)→ (H ·X)t ucp as n→∞.

Proof. To start with, note that the integral H ·X is well defined, as H is previsible since left

continuous and adapted. We can treat the local martingale and finite variation parts separately.

The latter is Problem 2 on Example Sheet 1, so this leaves just the martingale part. By applying

a localization argument, we can reduce to the case M ∈M2c and H bounded. Moreover, since

convergence in ‖ · ‖ norm implies ucp convergence, it suffices to show that the convergence in

(4.12) holds in ‖ · ‖ norm. To this end, let Hnt = H2−nb2ntc. Then Hn is simple for all n, and

by the left-continuity of H we have that Hnt → Ht almost surely as n→∞. Moreover,

(Hn ·M)t =

b2ntc−1∑k=0

Hk2−n

(M(k+1)2−n −Mk2−n

)+H2−nb2ntc(Mt −M2−nb2ntc).

Since H is bounded and |Mt −M2−nb2ntc| → 0 almost surely as n → ∞ by the continuity of

M , the Dominated Convergence Theorem implies that the second term in the right hand side

above converges to 0 in ‖ · ‖ norm as n→∞. It thus suffices to show that the remaining term

converges to the right limit, that is to show that Hn ·M → H ·M in ‖ · ‖ norm as n→∞. By

Ito’s isometry, this is equivalent to showing that Hn → H in ‖ · ‖M norm. Indeed, we have

‖Hn −H‖2M = E(∫ ∞

0(Hn

t −Ht)2d[M ]t

)→ 0 as n→∞

by the Bounded Convergence Theorem.

This concludes our construction of the Ito’s integral, which we now briefly summarize.

Summary of the stochastic integral

Step 1. H ∈ S, M ∈M2c . If Ht =

∑n−1k=0 Zk1(tk,tk+1](t) where Zk is a bounded, Ftk -measurable

random variable for each k, define

(H ·M)t =

n−1∑k=0

Zk(Mtk+1∧t −Mtk∧t).

Then H ·M ∈M2c .

Step 2. Equip M2c with a Hilbert space structure with norm ‖X‖ := ‖X∞‖L2 .

Step 3. Establish the existence of [M ] for all M ∈ Mc,loc. This is the unique, adapted,

continuous, non-decreasing process [M ] such that M2 − [M ] ∈Mc,loc.

Page 36: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

36 JASON MILLER AND VITTORIA SILVESTRI

Step 4. For M ∈M2c , use [M ] to define the Hilbert space (L2(M), ‖ · ‖M ) with

‖H‖M =

(E(∫ ∞

0H2sd[M ]s

))1/2

.

Step 5. Extend the integral to H ∈ L2(M), M ∈M2c by requiring that

‖H ·M‖ = ‖H‖M (Ito’s isometry)

for all H ∈ L2(M). Then H ·M ∈M2c .

Step 4. Extend to H locally bounded previsible and M ∈Mc,loc by setting

(H ·M)t = ((H1(0,Tn]) ·MTn)t for t ≤ Tn.

Then H ·M ∈Mc,loc.

Step 5. Extend to H locally bounded previsible and X = X0 +M+A continuous semimartingale

by setting

H ·X = H ·M +H ·A.

Then H ·X is again a continuous semimartingale.

5. Stochastic calculus

In this section we develop the usual tools of calculus for the Ito’s integral. We start with

some preliminary results on the so called covariation of two local martingales. This will then

be used to introduce and prove the celebrated Ito’s formula, which can be thought of as the

generalization of the chain rule to the Ito setting. Finally, we discuss a different notion of

stochastic integral called Stratonovich integral, for which the classical rules of calculus apply.

5.1. Covariation. It is easy to see that the quadratic variation defines a quadratic form. We

now define the associated bilinear form, called covariation.

Definition 5.1 (Covariation). For M,N ∈Mc,loc, define the covariation of M and N , denoted

[M,N ], by setting

(5.1) [M,N ] :=1

4([M +N ]− [M −N ]) (Polarization identity).

Note that [M,M ] = [M ].

Theorem 5.2. Let M,N ∈Mc,loc. Then the following hold:

(a) [M,N ] is the unique (up to indistinguishability) continuous, adapted, finite variation process

such that [M,N ]0 = 0 and MN − [M,N ] ∈Mc,loc.

Page 37: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 37

(b) For n ≥ 1, define

[M,N ]nt :=

d2nte−1∑k=0

(M(k+1)2−n −Mk2−n)(N(k+1)2−n −Nk2−n).

Then [M,N ]nt → [M,N ]t ucp as n→∞.

(c) If M,N ∈M2c , then MN − [M,N ] is a UI martingale.

(d) For H locally bounded and previsible, it holds

[H ·M,N ] + [M,H ·N ] = 2H · [M,N ].

Note that [M,N ] is a symmetric bilinear form, as one can easily see from (a).

Proof of Theorem 5.2. Note that MN = 14((M +N)2 − (M −N)2), so

MN − [M,N ] =1

4

((M +N)2 − [M +N ]︸ ︷︷ ︸

∈Mc,loc

−(M −N)2 + [M −N ]︸ ︷︷ ︸∈Mc,loc

),

which shows that MN−[M,N ] ∈Mc,loc. The covariation process [M,N ] is continuous, adapted

and of finite variation since difference of continuous, adapted non-decreasing processes. Finally,

uniqueness follows from the same argument used to show uniqueness of the quadratic variation.

This proves (a).

Note further that

[M,N ]nt =1

4([M +N ]nt − [M −N ]nt ) .

Thus (b)– (c) follow from the equivalent properties for quadratic variation.

For (d), we note that

[H · (M ±N)] = H2 · [M ±N ].

Plugging this into (5.1), we have that

[H ·M,H ·N ] = H2 · [M,N ].

Moreover,

(H + 1)2 · [M,N ] = [(H + 1) ·M, (H + 1) ·N ]

= [H ·M +M,H ·N +N ]

= [H ·M,H ·N ] + [M,H ·N ] + [H ·M,N ] + [M,N ],

and

(H + 1)2 · [M,N ] = (H2 + 2H + 1) · [M,N ]

= H2 · [M,N ] + 2H · [M,N ] + [M,N ].

By matching up terms, we see that

2H · [M,N ] = [M,H ·N ] + [H ·M,N ],

Page 38: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

38 JASON MILLER AND VITTORIA SILVESTRI

as required.

The next result tells us how the covariation behaves under integration, and it will be very useful

later on.

Proposition 5.3 (Kunita-Watanabe identity). Let M,N ∈ Mc,loc and H a locally bounded,

previsible process. Then

[H ·M,N ] = H · [M,N ].

Proof. It suffices to show that [H ·M,N ] = [M,H ·N ], since then the result follows from (d).

We have that (H ·M)N − [H ·M,N ] ∈Mc,loc and M(H ·N)− [M,H ·N ] ∈Mc,loc. Suppose

we could show that (H ·M)N −M(H ·N) ∈ Mc,loc. Then [H ·M,N ]− [M,H ·N ] ∈ Mc,loc.

But this last process can be expressed as sums and differences of quadratic variations and, as

quadratic variations are non-decreasing, this implies that [H ·M,N ]− [M,H ·N ] is of finite

variation. Hence [H ·M,N ]− [M,H ·N ] is a continuous local martingale starting from 0 and

of finite variation, which implies that it must be indistinguishable from the 0 process. Thus we

conclude that [H ·M,N ] = [M,H ·N ].

It only remains to show that (H ·M)N −M(H ·N) ∈Mc,loc. By localization, we may assume

that M,N ∈M2c and that H is bounded. By the OST, it suffices to show that

E((H ·M)TNT ) = E(MT (H ·N)T )

for all bounded stopping times T . But the l.h.s. above equals E((H ·MT )∞NT∞), and the r.h.s.

equals E(MT∞(H ·NT )∞), and MT , NT ∈M2

c . It therefore suffices to show that

(5.2) E((H ·M)∞N∞) = E(M∞(H ·N)∞)

for all M,N ∈ M2c and H bounded. Consider first the case that H = Z1(s,t] where Z is

bounded and Fs-measurable. Then

E((H ·M)∞N∞) = E(Z(Mt −Ms)N∞)

= E(ZMtE(N∞ | Ft)− ZMsE(N∞ | Fs)) (tower property)

= E(Z(MtNt −MsNs)).

The same argument implies that

E(M∞(H ·N)∞) = E(Z(MtNt −MsNs))

which proves (5.2) for H = Z1(s,t].

By linearity, this extends to all H ∈ S. If H is bounded, then we can find4 a sequence (Hn)n≥1

of simple processes such that Hn → H in both L2(M) and L2(N). Therefore Hn ·M → H ·M4This can be always insured by taking, for example, µ = d[M ] + d[N ] as measure to define the space of

integrands.

Page 39: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 39

and Hn ·N → H ·N in ‖ · ‖ norm, i.e. (Hn ·M)∞ → (H ·M)∞ and (Hn ·N)∞ → (H ·N)∞ in

L2 as n→∞. It follows that

‖((Hn ·M)∞ − (H ·M)∞)N∞‖2L1 ≤ ‖(Hn ·M)∞ − (H ·M)∞‖2L2‖N∞‖2L2 → 0,

which implies convergence of the expectation, i.e. E((Hn · M)∞N∞) → E((H · M)∞N∞).

Similarly one shows that E(M∞(Hn ·N)∞)→ E(M∞(H ·N)∞), which concludes the proof.

Definition 5.4. For X,Y continuous semimartingales we define their covariation [X,Y ] to be

the covariation of their respective martingale parts in the Doob-Meyer decomposition.

This definition is justified by the fact that

(5.3) [X,Y ]nt :=

d2nte−1∑k=0

(X(k+1)2−n −Xk2−n)(Y(k+1)2−n −Yk2−n)→ [X,Y ]t ucp as n→∞.

Note that the Kunita-Watanabe identity also holds for continuous semimartingales.

An important property of the covariation is that two independent semimartingales have zero

covariation. However, having zero covariation is not a sufficient condition for independence, in

the same way two uncorrelated random variables are not necessarily independent.

Proposition 5.5. Let X,Y be independent5 continuous semimartingales. Then [X,Y ] = 0.

Proof. See Example sheet 2.

5.2. Ito’s formula. We now move to the celebrated Ito’s formula. To start with, we prove a

particular case of it, which can be thought of as the generalization of integration by parts to

stochastic integrals.

Theorem 5.6 (Integration by parts). Let X,Y be continuous semimartingales. Then

XtYt −X0Y0 =

∫ t

0XsdYs +

∫ t

0YsdXs + [X,Y ]t.

Proof. Note that the above integrals are well defined, since any continuous adapted process is

locally bounded and previsible. Since both sides are continuous, it suffices to prove the result

for t of the form t = m2−n for m,n ∈ N. Fix t of this form. Note that for s ≤ t we have that

XtYt −XsYs = Xs(Yt − Ys) + Ys(Xt −Xs) + (Xt −Xs)(Yt − Ys).

5Recall that X,Y are independent if σ(Xt, t ≥ 0) and σ(Yt, t ≥ 0) are independent.

Page 40: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

40 JASON MILLER AND VITTORIA SILVESTRI

Applying this for j ≥ n, we find

XtYt −X0Y0 =m2j−n−1∑k=0

(X(k+1)2−jY(k+1)2−j −Xk2−jYk2−j )

=m2j−n−1∑k=0

(Xk2−j (Y(k+1)2−j − Yk2−j ) + Yk2−j (X(k+1)2−j −Xk2−j )

+ (X(k+1)2−j −Xk2−j )(Y(k+1)2−j − Yk2−j )

)→ (X · Y )t + (Y ·X)t + [X,Y ]t ucp as n→∞,

where convergence follows from (4.12) and (5.3).

Note that the covariation term does not appear in the usual integration by parts formula for

Lebesgue-Stieltjes integrals. This term, sometimes referred to as Ito’s correction term, will

vanish if either X or Y is of finite variation, or if X and Y are independent (cf. Proposition

5.5).

Theorem 5.7 (Ito’s formula). Let X1, . . . , Xd be continuous semimartingales and let X =

(X1, . . . , Xd). Let f : Rd → R be C2. Then

f(Xt) = f(X0) +

d∑i=1

∫ t

0

∂f

∂xi(Xs)dX

is +

1

2

d∑i,j=1

∫ t

0

∂2f

∂xi∂xj(Xs)d[Xi, Xj ]s.

Remark 5.8. Note that Theorem 5.7 implies that f(Xt) is a semimartingale since each of the

summands in Ito’s formula is a semimartingale.

We will often apply Ito’s formula for d = 1, that is

f(Xt) = f(X0) +

∫ t

0f ′(Xs)dXs +

1

2

∫ t

0f ′′(Xs)d[X]s.

One can derive this formula using Taylor expansion. Indeed, we have that

f(Xt) = f(X0) +

b2ntc−1∑k=0

(f(X(k+1)2−n)− f(Xk2−n)

)+(f(Xt)− f(X2−nb2ntc)

)= f(X0) +

b2ntc−1∑k=0

f ′(Xk2−n)(X(k+1)2−n −Xk2−n) +1

2

b2ntc−1∑k=0

f ′′(Xk2−n)(X(k+1)2−n −Xk2−n)2 + error

→ f(X0) +

∫ t

0f ′(Xs)dXs +

1

2

∫ t

0f ′′(Xs)d[X]s ucp as n→∞.

This is Ito’s formula for d = 1. We will not take this route to prove Ito’s formula, since the

error term turns out to be difficult to deal with.

Page 41: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 41

Proof of Theorem 5.7. We are going to give the proof of Ito’s formula for d = 1. We leave

the proof for d > 1 as an exercise. Write X = X0 +M +A where A has corresponding total

variation process V . Let

Tr := inft ≥ 0 : |Xt|+ Vt + [M ]t > r.

Then (Tr)r≥1 is a sequence of stopping times with Tr +∞ almost surely as r → ∞. It

suffices to prove the formula for arbitrary f ∈ C2(R) in the time-interval [0, Tr]. Let A denote

the subset of C2(R) where the formula holds: we aim to show that A = C2(R). To this end,

we will show that:

(a) A contains the functions f(x) ≡ 1 and f(x) = x.

(b) A is a vector space.

(c) A is in fact an algebra. That is, if f, g ∈ A then fg ∈ A.

(d) If (fn)n≥1 is a sequence in A with fn → f in C2(Br) for each r > 0, then f ∈ A. Here,

Br = x : |x| < r and fn → f in C2(Br) means that with

∆n,r = max

supx∈Br

|fn(x)− f(x)|, supx∈Br

|f ′n(x)− f ′(x)|, supx∈Br

|f ′′n(x)− f ′′(x)|

we have that ∆n,r → 0 as n→∞ with r fixed.

Note that (a), (b), (c) together imply that all polynomials are in A. By Weierstrass’ ap-

proximation theorem, polynomials are dense in C2(Br) for each r > 0, so (d) implies that

A = C2(R).

The fact that (a) and (b) hold is easy to check.

Proof of (c) Suppose that f, g ∈ A. Set Ft = f(Xt) and Gt = g(Xt). Since Ito’s formula holds

for f and g, F and G must be continuous semimartingales. By integration by parts, we find

(5.4) FtGt − F0G0 =

∫ t

0FsdGs +

∫ t

0GsdFs + [F,G]t.

Using that H · (K ·M) = (HK) ·M and Ito’s formula for g, we also have that∫ t

0FsdGs =

∫ t

0f(Xs)d

(∫ s

0g′(Xu)dXu +

1

2

∫ u

0g′′(Xu)d[X]u

)=

∫ t

0f(Xs)g

′(Xs)dXs +1

2

∫ t

0f(Xs)g

′′(Xs)d[X]s.

(5.5)

Similarly one shows that

(5.6)

∫ t

0GsdFs =

∫ t

0f ′(Xs)g(Xs)dXs +

1

2

∫ t

0f ′′(Xs)g(Xs)d[X]s.

Finally, by the Kunita-Watanabe identity, we have that [H ·M,N ] = H · [M,N ], hence

(5.7) [F,G]t = [f(X), g(X)]t = [f ′(X) ·X, g′(X) ·X]t =

∫ t

0f ′(Xs)g

′(Xs)d[X]s.

Page 42: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

42 JASON MILLER AND VITTORIA SILVESTRI

Inserting (5.5)-(5.6)-(5.7) into (5.4) implies that Ito’s formula holds for fg. That is, fg ∈ A.

Proof of (d) Suppose that (fn)n≥1 is a sequence in A and we have that fn → f in C2(Br) for

all r > 0. We want to show that Ito’s formula holds for f , that is f ∈ A. Since fn ∈ A for all

n, we have that

fn(Xt) = fn(X0) +

∫ t

0f ′n(Xs)dXs +

1

2

∫ t

0f ′′n(Xs)d[X]s

= fn(X0) +

∫ t

0f ′n(Xs)dAs +

∫ t

0f ′n(Xs)dMs +

1

2

∫ t

0f ′′n(Xs)d[M ]s.

We show that each term above converges to the corresponding one with f in place of fn, thus

proving Ito’s formula for f .

Let us start from the finite variation part. We have∫ t∧Tr

0|f ′n(Xs)− f ′(Xs)|dVs +

1

2

∫ t∧Tr

0|f ′′n(Xs)− f ′′(Xs)|d[M ]s

≤ ∆n,rVt∧Tr +1

2∆n,r[M ]t∧Tr ≤ r∆n,r → 0 as n→∞.

Consequently,∫ t∧Tr

0f ′n(Xs)dAs +

1

2

∫ t∧Tr

0f ′′n(Xs)d[M ]s →

∫ t∧Tr

0f ′(Xs)dAs +

1

2

∫ t∧Tr

0f ′′(Xs)d[M ]s

uniformly in t, almost surely. Moreover, for the martingale part we have that MTr ∈M2c , from

which

‖(f ′n(X) ·M)Tr − (f ′(X) ·M)Tr‖2 = E(∫ Tr

0(f ′n(Xs)− f ′(Xs))

2d[M ]s

)≤ ∆2

n,r E([M ]Tr) ≤ r∆2n,r → 0 as n→∞.

Therefore (f ′n(X) ·M)Tr → (f ′(X) ·M)Tr in M2c as n→∞.

Hence, for any r, we can pass to the limit in Ito’s formula to obtain

f(Xt∧Tr) = f(X0) +

∫ t∧Tr

0f ′(Xs)dXs +

1

2

∫ t∧Tr

0f ′′(Xs)d[X]s.

Example 5.9. Let X = B where B is a standard Brownian motion in R and let f(x) = x2.

Applying Ito’s formula to f(B), we then have that

B2t = 2

∫ t

0BsdBs + t.

Rearranging, we obtain

B2t − t = 2

∫ t

0BsdBs ∈Mc,loc.

Page 43: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 43

Example 5.10. Let f : R+ × Rd → R be C1,2 and let Xt = (t, B1t , . . . , B

dt ) where B1, . . . , Bd

are independent standard Brownian motions. Then, by Ito’s formula, we have that

f(t, Bt)− f(0, B0)−∫ t

0

(∂

∂t+

1

2∆

)f(s,Bs)ds =

d∑i=1

∫ t

0

∂xif(s,Bs)dB

is ∈Mc,loc.

In particular, if f does not depend on t and f is harmonic in x, then f(Bt) ∈ Mc,loc. If

furthermore f is bounded, then f(Bt) is a true bounded martingale.

5.3. Stratonovich integral. LetX,Y be continuous semimartingales. We define the Stratonovich

integral as ∫ t

0Xs∂Ys =

∫ t

0XsdYs +

1

2[X,Y ]t.

Here, the integral on the right hand side is an Ito integral. Using the Riemann sum approxima-

tions for the terms on the right hand side, we have that

b2ntc−1∑k=0

(X(k+1)2−n +Xk2−n

2

)(Y(k+1)2−n − Yk2−n)→

∫ t

0Xs∂Ys ucp as n→∞.

Proposition 5.11. Let X1, . . . , Xd be continuous semimartingales and let f : Rd → R be C3.

Then

(5.8) f(Xt) = f(X0) +d∑i=1

∫ t

0

∂f

∂xi(Xs)∂X

is.

In particular, the integration by parts formula is given by

(5.9) XtYt −X0Y0 =

∫ t

0Xs∂Ys +

∫ t

0Ys∂Xs.

Note that Proposition 5.11 implies that the Stratonovich integral satisfies the usual rules of

calculus. On the other hand, the price to pay for this property is that the Stratonovich integral

with respect to a local martingale is not necessarily a local martingale. Indeed, for example,∫ t

0Bs∂Bs =

∫ t

0BsdBs +

1

2t =

1

2B2t /∈Mc,loc.

Proof of Proposition 5.11. We will treat the case that d = 1 and leave the case d ≥ 2 as an

exercise. By Ito’s formula, we have that

f(Xt) = f(X0) +

∫ t

0f ′(Xs)dXs +

1

2

∫ t

0f ′′(Xs)d[X]s(5.10)

f ′(Xt) = f ′(X0) +

∫ t

0f ′′(Xs)dXs +

1

2

∫ t

0f ′′′(Xs)d[X]s.(5.11)

Page 44: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

44 JASON MILLER AND VITTORIA SILVESTRI

By the Kunita-Watanabe identity and (5.11), we have that [f ′(X), X] = [∫ t

0 f′′(Xs)dXs, X] =

f ′′(X) · [X]. Combining this with (5.10) and the definition of the Stratonovich integral, we

then have that

f(Xt) = f(X0) +

∫ t

0f ′(Xs)∂Xs.

Remark 5.12. Note that in Proposition 5.11, we require more regularity of the function f

in order to obtain (5.8) in comparison to Ito’s formula. We also emphasize that for the

Stratonovich integral, the integrand must be a semimartingale. This is also in contrast to the

case of Ito’s formula, in which case the integrand need only be locally bounded and previsible.

5.4. Shorthand notation. We now record some shorthand notation which is common for

stochastic calculus.

• If Zt = Z0 +

∫ t

0HsdXs write dZt = HtdXt.

• If Zt = Z0 +

∫ t

0Hs∂Xs write ∂Zt = Ht∂Xt.

• If Zt = [X,Y ]t =

∫ t

0d[X,Y ]s we write dZt = dXtdYt.

We then have the following computational rules:

• Htd(KtdXt) = (HtKt)dXt (iterated integral),

• Ht(dXtdYt) = d(HtdXt)dYt (Kunita-Watanabe),

• d(XtYt) = XtdYt + YtdXt + dXtdYt (integration by parts),

• df(Xt) =

d∑i=1

∂f

∂xi(Xt)dX

it +

1

2

d∑i,j=1

∂2f

∂xi∂xj(Xt)dX

itdX

jt (Ito’s formula).

The relationship between the Ito and Stratonovich integrals is that ∂Zt = Yt∂Xt if and only if

dZt = YtdXt + 12dXtdYt.

6. Applications

6.1. Brownian motion. Throughout, we will work on a filtered probability space (Ω,F , (Ft)t≥0,P)

where (Ft)t≥0 satisfies the usual conditions.

Theorem 6.1 (Levy’s characterization of Brownian motion). Let X1, . . . , Xd be continuous

local martingales and set X = (X1, . . . , Xd). Suppose that X0 = 0 and that [Xi, Xj ]t = δijt for

all i, j and t ≥ 0. Then X is a standard Brownian motion on Rd.

Page 45: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 45

Proof. It suffices to show that, for all 0 ≤ s ≤ t < ∞, we have that Xt −Xs is independent

of Fs and it is distributed as a N(0, (t − s)Id) random variable, where Id denotes the d × didentity matrix. This is equivalent to show that

E(ei(θ,Xt−Xs) | Fs) = exp

(−1

2|θ|2(t− s)

), for all θ ∈ Rd,

where (·, ·) denotes the scalar product in Rd, and |θ|2 = (θ, θ).

Fix θ ∈ Rd and set Yt = (θ,Xt) =∑d

j=1 θjXj

t . Then Y ∈ Mc,loc as Mc,loc is a vector space.

By assumption, we have that

[Y ]t = [Y, Y ]t =

[ d∑j=1

θjXj ,

d∑k=1

θkXk

]t

=

d∑j,k=1

θjθk[Xj , Xk]t = |θ|2t.

Consider

Zt := exp(iYt + 12 [Y ]t) = exp(i(θ,Xt) + 1

2 |θ|2t).

By Ito’s formula with X = iY + 12 [Y ] and f(x) = ex, we have that

dZt = Zt(idYt + 12d[Y ]t)− 1

2Ztd[Y ]t = iZtdYt.

Consequently, Z ∈Mc,loc. Moreover, Z is bounded on [0, t] for each t ≥ 0, hence Z ∈Mc by

Proposition 3.9. Therefore E(Zt | Fs) = Zs, which implies that

E(exp(i(θ,Xt −Xs)) | Fs) = exp(−12 |θ|

2(t− s)).

The following result is a powerful consequence of Levy’s characterization of Brownian motion.

Theorem 6.2 (Dubins-Schwartz theorem). Let M ∈Mc,loc with M0 = 0 and [M ]∞ =∞. Set

τs := inft ≥ 0 : [M ]t > s, Bs := Mτs, and Gs := Fτs. Then τs is an (Ft)-stopping time and

[M ]τs = s for all s ≥ 0. Moreover, B is a (Gs)-Brownian motion and

Mt = B[M ]t .

Remark 6.3. Theorem 6.2 states that any continuous local martingale starting from 0 can be

represented as a stochastic time-change of Brownian motion. Thus, in some sense, Brownian

motion is the “most general” continuous local martingale.

Proof of Theorem 6.2. Since [M ] is continuous and adapted, τs is a stopping time for each

s ≥ 0. The hypothesis [M ]∞ = ∞ implies that τs < ∞ for all s ≥ 0. Note that (Gs) is a

filtration, since if S, T are two stopping times with S ≤ T almost surely, and A ∈ FS , then

A ∩ T ≤ t = (A ∩ S ≤ t) ∩ T ≤ t ∈ Ft,

so A ∈ FT and hence FS ⊆ FT . Applying this with S = τr and T = τs gives that Gr ⊆ Gs for

all r ≤ s.

Page 46: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

46 JASON MILLER AND VITTORIA SILVESTRI

To see that B is adapted to (Gs)s≥0 we have to show that Bτs is Fτs-measurable for all s ≥ 0.

Indeed, this follows from the general fact that if X is a cadlag adapted process and T is a

stopping time, then XT1T<∞ is FT -measurable (see the Advanced Probability course).

To finish the proof, we will show that B is continuous and that it is a local martingale with

[B]t = t. The result will then follow from Levy’s characterization of Brownian motion.

Step 1: B is continuous. We first note that s 7→ τs is non-decreasing and cadlag . Therefore B

is cadlag . To prove the continuity of B, we must therefore show that Bs− = Bs for all s ≥ 0.

Equivalently, Mτs−= Mτs for all s ≥ 0, with τs− = inft ≥ 0 : [M ]t = s. If τs = τs− there

is nothing to show. If, on the other hand, τs > τs− then [M ] must be constant on the time

interval [τs− , τs]. It thus suffices to prove that if [M ] is constant on a given interval then M is

also constant on the same interval.

By applying localization, we may assume that M ∈M2c . It suffices to show that for all q ∈ Q,

q > 0, with Sq = inft > q : [M ]t > [M ]q we have that M is constant on [q, Sq]. We know that

M2 − [M ] is a UI martingale as M ∈M2c . By the OST, we have that

E(M2Sq− [M ]Sq | Fq) = M2

q − [M ]q.

Since [M ]Sq = [M ]q and M is a martingale, we also have that

E((MSq −Mq)2 | Fq) = E(M2

Sq−M2

q |Fq) = E[[M ]Sq − [M ]q | Fq] = 0.

Therefore M is almost surely constant on [q, Sq], hence B is continuous.

Step 2: B is a Brownian motion. Fix s > 0. Then we know that [M τs ]∞ = [M ]τs = s.

Consequently, M τs ∈ M2c since E([M τs ]∞) <∞. Therefore (M2 − [M ])τs is a UI martingale.

By the OST, we have for 0 ≤ r < s <∞ that

E(Bs | Gr) = E(M τs∞ | Fτr) = Mτr = Br

E(B2s − s | Gr) = E((M2 − [M ])τs | Fτr) = M2

τr − [M ]τr = B2r − r.

Combining, we have that B ∈Mc with [B]t = t for all t ≥ 0. Thus by Theorem 6.1, we have

that B is a (Gt)t≥0-Brownian motion.

Here is an extension to Dubins-Schwartz theorem that we state without proof.

Theorem 6.4. Let M ∈ Mc,loc with M0 = 0. Then define a new process B on the same

probability space by setting

Bs =

Mτs , s ≤ [M ]∞,

M∞ + βs−[M ]∞ , s ≥ [M ]∞

where β is a standard Brownian motion defined on the same probability space, independent of

everything else. Then B is a standard Brownian motion, and Mt = B[M ]t for all t ≥ 0.

Page 47: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 47

Example 6.5. Let B be a Brownian motion, and h be a deterministic measurable function in

L2([0,∞)). Let Mt :=∫ t

0 h(s)dBs. Then M0 = 0, M ∈Mc,loc and [M ]t =∫ t

0 h2(s)ds. Then

M∞(d)= B∫∞

0 h2(s)ds ∼ N(0, ‖h‖2L2).

Example 6.6. Let M ∈Mc,loc. then

[M ]∞ <∞ =

limt→∞

Mt exists finite.

Conversely,

[M ]∞ =∞ =

lim inft→∞

Mt = −∞, lim supt→∞

Mt = +∞.

6.2. Exponential martingales. Let M ∈ Mc,loc with M0 = 0. Set Zt = exp(Mt − 12 [M ]t).

By Ito’s formula applied to X = M − 12 [M ] and f(x) = ex, we have that

dZt = Zt(dMt − 12d[M ]t) + 1

2Ztd[M ]t = ZtdMt.

Consequently, Z ∈ Mc,loc as it is the integral of Z with respect to an element of Mc,loc. We

call Z the stochastic exponential of M .

Definition 6.7 (Exponential martingale). Let M ∈ Mc,loc with M0 = 0. Define the process

E(M)t by setting

E(M)t = exp(Mt − 12 [M ]t).

Then E(M) is called the stochastic exponential of (or exponential martingale associated with)

M . Note that E(M) ∈Mc,loc, and it satisfies dE(M)t = E(M)tdMt.

The next result shows that under some additional assumptions on M the stochastic exponential

is a true UI martingale.

Proposition 6.8. Let M ∈Mc,loc with M0 = 0. Suppose that [M ] is uniformly bounded. Then

E(M) is a UI martingale.

To prove the above result we will use a powerful inequality which tells us that if the quadratic

variation of a continuous local martingale is small then the martingale cannot be too large.

Proposition 6.9. Let M ∈Mc,loc with M0 = 0. Then for all ε, δ > 0 we have that

P(

supt≥0

Mt ≥ ε, [M ]∞ ≤ δ)≤ exp

(− ε

2

).

Proof. Fix ε > 0 and set T = inft ≥ 0 : Mt ≥ ε. Fix θ ∈ R and set Zt = exp(θMTt − 1

2θ2[M ]Tt ).

Then Zt ∈Mc,loc since Z = E(θMT ) and θMT ∈Mc,loc. Moreover, |Zt| ≤ eθε for all t ≥ 0 by

Page 48: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

48 JASON MILLER AND VITTORIA SILVESTRI

construction, so Z is in fact a bounded martingale. Therefore E(Z∞) = Z0 = 1. For each δ > 0,

we have that

P(

supt≥0

Mt ≥ ε, [M ]∞ ≤ δ)

= P(

supt≥0

eθMTt ≥ eθε, [M ]∞ ≤ δ

)≤ P

(supt≥0

Zt ≥ eθε−θ2δ/2)

≤ exp(−θε+ θ2δ/2) (Doob’s maximal inequality).

The result follows by optimizing this bound over θ.

We can now prove Proposition 6.8.

Proof of Proposition 6.8. We show that E(M) is uniformly bounded by an integrable random

variable, which implies the result. Note that

supt≥0E(M)t ≤ exp

(supt≥0

Mt

)as [M ]t ≥ 0 for all t ≥ 0. It suffices to show that the r.h.s. is integrable. Let c be such that

[M ]∞ ≤ c. By Proposition 6.9, we have that

P(

supt≥0

Mt ≥ ε)

= P(

supt≥0

Mt ≥ ε, [M ]∞ ≤ c)≤ exp

(− ε

2

2c

).

It follows that

E(

exp(

supt≥0

Mt

))=

∫ ∞0

P(

exp(

supt≥0

Mt

)≥ λ

)dλ

=

∫ ∞0

P(

supt≥0

Mt ≥ log λ)dλ

≤ 1 +

∫ ∞1

exp

(−(log λ)2

2c

)dλ <∞.

Therefore E(M) is UI, as desired.

Proposition 6.8 provides one circumstance in which the hypotheses of the following theorem

are satisfied.

Suppose that P and Q are two probability measures on (Ω,F). Then we say that Q is absolutely

continuous with respect to P and write Q P if for any A ∈ F , we have that P(A) = 0 implies

Q(A) = 0.

Theorem 6.10 (Girsanov’s theorem). Let M ∈ Mc,loc be such that M0 = 0. Suppose that

Z = E(M) is a UI martingale. Then we can define a new probability measure P P on (Ω,F)

by setting P(A) = E(Z∞1(A)) for A ∈ F . If X ∈Mc,loc(P), then X − [X,M ] ∈Mc,loc(P).

Page 49: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 49

Proof. Since Z is UI, we know that Z∞ exists with Z∞ ≥ 0 and E(Z∞) = E(Z0) = 1. It

is not difficult to see that P is a probability measure with P P. Let X ∈ Mc,loc(P) and

set Tn = inft ≥ 0 : |Xt − [X,M ]t] ≥ n. As X − [X,M ] is continuous, we have that

P(Tn ∞) = 1. By absolute continuity, we therefore have that P(Tn ∞) = 1. So, to show

that Y = X − [X,M ] ∈Mc,loc(P), it suffices to show that Y Tn = XTn − [XTn ,M ] ∈Mc(P) for

all n. Write X and Y in place of XTn and Y Tn in what follows.

By integration by parts, we have that

d(ZtYt) = YtdZt + ZtdYt + dYtdZt

= YtdZt + ZtdXt − ZtdXtdMt + dXtdZt

= YtdZt + ZtdXt,

where we have used that dZt = ZtdMt. Consequently, ZY ∈ Mc,loc(P). Also, ZT : T ≤t is a stopping time is UI for all t. Since Y is bounded, we also have that ZTYT : T ≤t is a stopping time is UI for all t as well. Hence, ZY ∈Mc(P). But, for s ≤ t, we have that

E(Yt − Ys | Fs) = E(Z∞Yt − Z∞Ys | Fs) = E(ZtYt − ZsYs | Fs) = 0,

as ZY ∈Mc(P). Therefore Y ∈Mc(P), as required.

Remark 6.11. Note that the quadratic variation does not change when performing a change

of measures from P to P (cf. Example Sheet 4).

Corollary 6.12. Let B be a standard Brownian motion under P and let M ∈ Mc,loc with

M0 = 0. Suppose that Z = E(M) is a UI martingale and P(A) = E(Z∞1A) for A ∈ F . Then

B = B − [B,M ] is a P-Brownian motion.

Proof. Since B is a continuous P-local martingale by Girsanov’s theorem and has [B]t = [B]t = t,

it follows from the Levy characterization that B must be a P-Brownian motion.

7. Stochastic differential equations

In ths section we turn to the study of Stochastic Differential Equations, SDEs in short, of the

form

dXt = b(Xt)dt+ σ(Xt)dBt.

Recall from the introduction that these arise from Ordinary Differential equations by perturba-

tion.

Page 50: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

50 JASON MILLER AND VITTORIA SILVESTRI

7.1. General definitions. Let M d×m(R) denote the set of d×m matrices with real entries.

Suppose that σ : Rd →M d×m(R) and b : Rd → Rd are measurable functions which are bounded

on compact sets. We write σ(x) = (σij(x)) and b(x) = (bi(x)). Consider the equation

(7.1) dXt = σ(Xt)dBt + b(Xt)dt.

We can write this equation component-wise as

dXit =

m∑j=1

σij(Xt)dBjt + bi(Xt)dt.

A solution to (7.1) consists of:

• A filtered probability space (Ω,F , (Ft)t≥0,P) where (Ft)t≥0 satisfies the usual conditions.

• An (Ft)-Brownian motion B = (B1, . . . , Bm) in Rm.

• An (Ft)-adapted continuous process X = (X1, . . . , Xd) in Rd such that

Xt = X0 +

∫ t

0σ(Xs)dBs +

∫ t

0b(Xs)ds.

When, in addition, X0 = x ∈ Rd, we say that X is a solution started from x.

There are several different notions of existence and uniqueness for solutions to SDEs.

We say that an SDE has a weak solution if for all x ∈ Rd, there exists a solution to the

equation started from x. There is uniqueness in law if all solutions to the SDE started from

x have the same distribution. There is pathwise uniqueness if, when we fix (Ω,F , (Ft)t≥0,P)

and B, then any two solutions X and X ′ such that X0 = X ′0 are indistinguishable, that is

P(Xt = X ′t for all t ≥ 0) = 1. We say that a solution X of a SDE started from x is a strong

solution if X is adapted to the filtration generated by B.

In the case of a weak solution, the filtration (Ft)t≥0 is not necessarily the filtration generated by

B. In other words, it is not necessarily true that Xt(ω) is a measurable function of (Bs(ω) : s ≤ t)and x.

Example 7.1. It is possible to have the existence of a weak solution and uniqueness in law

without pathwise uniqueness.

Suppose that β is a Brownian motion in R with β0 = x. Set Bt =∫ t

0 sgn(βs)dβs. We take

the convention that sgn(x) = 1(0,∞)(x)− 1(−∞,0](x). Note that (sgn(βt))t≥0 is previsible since

left-continuous, so that the integral is well-defined. Note that

x+

∫ t

0sgn(βs)dBs = x+

∫ t

0(sgn(βs))

2dβs = x+

∫ t

0dβs = βt.

Page 51: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 51

Consequently, β solves the SDE

(7.2)

dXt = sgn(Xt)dBt

X0 = x.

Therefore (7.2) has a weak solution. By the Levy characterization of Brownian motion, any

solution to this SDE is a Brownian motion. Therefore we have uniqueness in law.

On the other hand, there is not pathwise uniqueness. To see this take x = 0: we show that

both β and −β are solutions to (7.2). Indeed,

−βt = −∫ t

0sgn(βs)dBs =

∫ t

0sgn(−βs)dBs + 2

∫ t

01(βs = 0)dBs.

The second term in the r.h.s. is a continuous local martingale with quadratic variation 4∫ t

0 1(βs =

0)ds = 0, since the zero set of Brownian motion has zero Lebesgue measure by Fubini. It

follows that∫ t

0 1(βs = 0)dBs is indistinguishable from the zero martingale. We thus conclude

that both β and −β are solutions (with the same probability space and Brownian motion) to

(7.2), thus contradicting pathwise uniqueness.

7.2. Lipschitz coefficients. We now turn to the question of sufficient conditions for existence

and uniqueness of solutions to (7.1).

Recall that for U ⊆ Rd and f : U → Rn, we say that f is Lipschitz if there exists K <∞ such

that |f(x)− f(y)| ≤ K|x− y| for all x, y ∈ U . For d,m ≥ 1 integers, let M d×m(R) denote the

space of d×m matrices with real entries. We equip this space with the following Frobenius

norm: if A ∈M d×m(R) with A = (aij), define

|A| :=

d∑i=1

m∑j=1

a2ij

1/2

.

Now, if f : U → M d×m(R), then we say that f Lipschitz if there exists K < ∞ such that

|f(x)− f(y)| ≤ K|x− y| for all x, y ∈ U .

We establish the existence and uniqueness of solutions to SDEs when the coefficients σ and b

are Lipschitz.

Theorem 7.2 (Existence and uniqueness of solutions). Suppose that σ : Rd →M d×m(R) and

b : Rd → Rd are Lipschitz. Then there is pathwise uniqueness for the SDE

(7.3) dXt = σ(Xt)dBt + b(Xt)dt.

Moreover, for each filtered probability space (Ω,F , (F)t≥0,P) satisfying the usual conditions and

each (Ft)-Brownian motion B, there exists for each x ∈ Rd a strong solution starting from x.

Page 52: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

52 JASON MILLER AND VITTORIA SILVESTRI

In our proof the existence of solutions to SDEs with Lipschitz coefficients, we will make use of

the following theorem which we state without proof.

Theorem 7.3 (Contraction mapping theorem). Let (X, d) be a complete metric space.

(a) Suppose that F : X → X is a contraction, i.e. there exists r < 1 such that d(F (x), F (y)) ≤r d(x, y) for all x, y ∈ X. Then F has a unique fixed point.

(b) Suppose that F : X → X is such that there exists r < 1 and n ∈ N with d(Fn(x), Fn(y)) <

r d(x, y) for all x, y ∈ X. Then F has a unique fixed point.

Uniqueness will on the other hand follow as an application of Gronwall’s inequality, that we

now recall.

Lemma 7.4 (Gronwall’s lemma). Let T > 0 and let f be a non-negative, bounded, measurable

function on [0, T ]. Suppose that there exists a, b ≥ 0 such that

f(t) ≤ a+ b

∫ t

0f(s)ds for all t ∈ [0, T ].

Then

f(t) ≤ aebt for all t ∈ [0, T ].

Proof. See Example sheet 3.

We can now prove the existence and uniqueness result.

Proof of Theorem 7.2. We will give the proof in the case that d,m = 1 and leave the general

case as an exercise. Let K > 0 be such that

|σ(x)− σ(y)| ≤ K|x− y| and |b(x)− b(y)| ≤ K|x− y| for all x, y ∈ R.

Pathwise uniqueness.

Suppose that X,X ′ are solutions to the SDE with the same probability space (Ω,F , (Ft)t≥0,P)

and Brownian motion B, and with X0 = X ′0. We will show that they must be indistinguishable,

that is Xt = X ′t for all t ≥ 0 almost surely. Fix M > 0 and let

τ = inft ≥ 0 : |Xt| ∨ |X ′t| ≥M.

Then we have that

Xt∧τ = X0 +

∫ t∧τ

0σ(Xs)dBs +

∫ t∧τ

0b(Xs)ds

X ′t∧τ = X0 +

∫ t∧τ

0σ(X ′s)dBs +

∫ t∧τ

0b(X ′s)ds.

Page 53: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 53

Fix T > 0. If t ∈ [0, T ], then

E((Xt∧τ −X ′t∧τ )2) ≤ 2E

((∫ t∧τ

0(σ(Xs)− σ(X ′s))dBs

)2)

+ 2E

((∫ t∧τ

0(b(Xs)− b(X ′s))ds

)2)

≤ 2E(∫ t∧τ

0(σ(Xs)− σ(X ′s))

2ds

)+ 2TE

(∫ t∧τ

0(b(Xs)− b(X ′s))2ds

)(by Ito’s isometry and Cauchy-Schwarz)

≤ 2K2(1 + T )E(∫ t∧τ

0(Xs −X ′s)2ds

)≤ 2K2(1 + T )

∫ t

0E((Xs∧τ −X ′s∧τ )2)ds.

Let f(t) = E((Xt∧τ − X ′t∧τ )2). Then f is non-negative and bounded by 4M2, and we have

shown that

f(t) ≤ 2K2(1 + T )

∫ t

0f(s)ds for all t ∈ [0, T ].

This is the right setting for applying Gronwall’s lemma, which gives f(t) = 0 for all t ∈ [0, T ].

This implies that Xt∧τ = X ′t∧τ for all t ∈ [0, T ], almost surely. Since M and T were arbitrary,

we conclude that Xt = X ′t for all t ≥ 0, almost surely.

Existence of a strong solution.

Fix a filtered probability space (Ω,F , (Ft)t≥0,P) and a standard (Ft)-Brownian motion B. Let

(FBt ) be the filtration generated by B, and observe that FBt ⊆ Ft for all t ≥ 0.

For any T > 0, we write CT for the set of continuous, adapted processes X : [0, T ]→ R such

that |||X|||T := ‖ supt≤T |Xt|‖2 <∞. Recall that CT is complete. Write further C for the set of

continuous, adapted processes X : [0, T ]→ R such that |||X|||T <∞ for all T > 0.

For each T > 0, we construct a strong solution (on the space (Ω,F , (Ft)t≥0,P) with (Ft)-Brownian motion B) as the fixed point of a contraction map F from CT to itself. We will then

”glue” these solutions to obtain a solution defined on the whole [0,∞).

Fix x ∈ R. Using the Lipschitz property of σ and b, we note that for all y ∈ R we have that

|σ(y)| = |σ(y)− σ(0) + σ(0)| ≤ |σ(0)|+ |σ(y)− σ(0)| ≤ |σ(0)|+K|y|.(7.4)

We similarly have that

|b(y)| ≤ |b(0)|+K|y|.(7.5)

Fix T > 0 and X ∈ CT . Let Mt =∫ t

0 σ(Xs)dBs, 0 ≤ t ≤ T . Then [M ]T =∫ T

0 σ2(Xs)ds.

Consequently, by (7.4) we have that

E([M ]T ) ≤ 2T (|σ(0)|2 +K2|||X|||2T ) <∞.

Page 54: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

54 JASON MILLER AND VITTORIA SILVESTRI

It thus follows that (Mt)t∈[0,T ] is an L2-bounded martingale. Doob’s L2 inequality then implies

that

E

(sup

0≤t≤T

∣∣∣∣∫ t

0σ(Xs)dBs

∣∣∣∣2)≤ 8T

(|σ(0)|2 +K2|||X|||2T

)<∞.

Using (7.5) and the Cauchy-Schwarz inequality, we also have that

E

(sup

0≤t≤T

∣∣∣∣∫ t

0b(Xs)ds

∣∣∣∣2)≤ TE

(∫ T

0b2(Xs)ds

)≤ 2T 2

(|b(0)|2 +K|||X|||2T

)<∞.

Combining, we see that the map F defined on CT by

F (X)t = x+

∫ t

0σ(Xs)dBs +

∫ t

0b(Xs)ds, t ≤ T

takes values in CT . Note that for X,Y ∈ CT , similar arguments to those above imply that

|||F (X)− F (Y )|||2t ≤ 2K2(4 + T )

∫ t

0|||X − Y |||2sds

=: CT

∫ t

0|||X − Y |||2sds.

Thus by iterating, we have that

|||F (n)(X)− F (n)(Y )|||2T ≤ CnT∫ T

0

∫ t1

0

∫ t2

0· · ·∫ tn−1

0|||X − Y |||2tndt1dt2 . . . dtn

≤CnTT

n

n!|||X − Y |||2T .

(7.6)

Note that we can make n sufficiently large so that CnTTn/n! < 1. The contraction mapping

theorem therefore implies that there exists a unique fixed point X(T ) ∈ CT of F .

By pathwise uniqueness, we have that X(T )t = X

(T ′)t for all t ≤ T ∧ T ′, almost surely. We thus

define X by setting Xt = X(N)t whenever t ≤ N and N ∈ N. Then X is the pathwise unique

solution to the SDE started from x. To see that X is a strong solution, it remains to prove

that it is (FBt )-adapted. To this end, we show that for all T ≥ 0 the process (Xt)t≤T is a limit

of (FBt )-adapted processes.

Define a sequence (Y n)n≥0 in CT by setting Y 0 = x and Y n = F (Y n−1) for each n ≥ 1. Then

Y n is (FBt )t≥0 adapted for all n. Since X = F (n)(X) for all n ≥ 0, it follows from (7.6) that

|||X − Y n|||2T = |||F (n)(X)− F (n)(x)|||2T ≤CnTT

n

n!|||X − x|||2T .

Thus Y n → X in CT , and hence there exists a subsequence (Y nk)k≥1 such that Y nk → X

uniformly on [0, T ], almost surely. In particular, (Xt)t≤T is the almost sure limit of (FBt )-

adapted processes, and it is therefore (FBt )-adapted. Since T was arbitrary, this shows that

X = (Xt)t≥0 is (FBt )-adapted, and thus a strong solution.

Page 55: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 55

Remark 7.5. Note that the above proof also shows that the pathwise unique strong solution to

the SDE (7.3) belongs to CT for all T > 0.

Proposition 7.6. Under the hypotheses of Theorem 7.2, there is uniqueness in law for the

SDE dXt = σ(Xt)dBt + b(Xt)dt.

Proof. See Example sheet 3.

Example 7.7 (Ornstein-Uhlenbeck process). Fix λ ∈ R and consider the SDE in R2

dVt = dBt − λVtdt, V0 = v0

dXt = Vtdt, X0 = x0.

When λ > 0, this SDE models the movement of a pollen grain on the surface of a liquid. X

represents the x-coordinate of the position and V is the velocity in the x-direction. The term

−λV is the “damping” due to the viscosity of the liquid. Whenever |V | gets large, the system

acts to reduce |V |.

Theorem 7.2 implies that there exists a unique strong solution to this SDE started at each

x0, v0. This is a rare example of an SDE that we can solve explicitly. By Ito’s formula, we have

that

d(eλtVt) = eλtdVt + λeλtVtdt = eλtdBt

and so

eλtVt = v0 +

∫ t

0eλsdBs.

Rearranging, we find

Vt = e−λtv0 +

∫ t

0eλ(s−t)dBs.

Note that Vt has the normal distribution with mean e−λtv0 and variance (1−e−2λt)/2λ. If λ > 0,

then Vt(d)→ N (0, (2λ)−1) as t→∞. In particular, N (0, (2λ)−1) is the stationary distribution

for V in the sense that if V0 ∼ N (0, (2λ)−1) then Vt ∼ N (0, (2λ)−1) for all t ≥ 0.

7.3. Local solutions. A locally defined process (X, T ) is a stopping time T together with a

map

X : (ω, t) ∈ Ω× [0,∞) : t < T (ω) → R.

It is cadlag if the map t 7→ Xt(ω) from [0, T (ω)) to R is cadlag for all ω ∈ Ω. Let

Ωt = ω ∈ Ω : t < T (ω).

Then (X, T ) is adapted if Xt : Ωt → R is Ft-measurable for all t. We say that (X, T ) is a locally

defined local martingale if there exist stopping times Tn T such that XTn is a martingale

for all n. We say that (H, η) is a locally defined locally bounded previsible process if there exist

Page 56: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

56 JASON MILLER AND VITTORIA SILVESTRI

stopping times Sn η such that H1(0,Sn] is bounded and previsible for all n. Then we can

define (H ·X, T ∧ η) by setting

(H ·X)Sn∧Tnt :=

((H1(0,Sn∧Tn]) ·XSn∧Tn)

t, for all n.

Proposition 7.8 (Local Ito’s formula). Let X1, . . . , Xd be continuous semimartingales and let

X = (X1, . . . , Xd). Let U ⊆ Rd be open and let f : U → Rd be C2. Set T = inft ≥ 0 : Xt /∈ U.Then for all t < T , we have that

f(Xt) = f(X0) +d∑i=1

∫ t

0

∂f

∂xi(Xs)dX

is +

1

2

d∑i,j=1

∫ t

0

∂2f

∂xi∂xj(Xs)d[Xi, Xj ]s.

Proof. This follows by applying Ito’s formula toXTn where Tn = inft ≥ 0 : dist(Xt, Uc) ≤ 1/n,

and observing that Tn T as n→∞.

Example 7.9. Let X = B where B is a standard Brownian motion starting from 1 with d = 1,

U = (0,∞), and f(x) =√x. Then Proposition 7.8 implies that√Bt = 1 +

1

2

∫ t

0B−1/2s ds− 1

8

∫ t

0B−3/2s ds

for t < T = inft ≥ 0 : Bt = 0.

Suppose that U ⊆ Rd is open and σ : U →M d×m(R) and b : U → Rd are measurable functions

which are bounded on compact sets. A local solution to the SDE consists of:

• A filtered probability space (Ω,F , (Ft)t≥0,P) satisfying the usual conditions.

• An (Ft)-Brownian motion B in Rm.

• A continuous (Ft)-adapted locally defined process (X, T ) with X ∈ Rd such that

Xt = X0 +∫ t

0 σ(Xs)dBs +∫ t

0 b(Xs)ds for all t < T .

We say that (X, T ) is maximal if for any other local solution (Y, η) on the same space such

that Xt = Yt for all t < T ∧ η, we have that η ≤ T .

7.4. Locally Lipschitz coefficients. Suppose that U ⊆ Rd is open. Then a function f : U →Rd is called locally Lipschitz if for each compact set C ⊆ U we have that f |C is Lipschitz.

Theorem 7.10. Suppose that U ⊆ Rd is open and σ : U → M d×m(R) and b : U → Rd are

locally Lipschitz. Then for all x ∈ U the SDE dXt = σ(Xt)dBt + b(Xt)dt has a pathwise unique

maximal local solution (X, T ) started from x. Moreover, for all compact sets C ⊆ U , on the

event T <∞ we have that supt < T : Xt ∈ C < T .

Note that the last assertion tells us that if the solution blows up in finite time, then at least it

will not return to any compact set infinitely often as it approaches that time.

Page 57: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 57

Lemma 7.11. Let U ⊆ Rd be open, C ⊆ U compact. Then:

(1) there exists a C∞ function ϕ : Rd → [0, 1] such that ϕ|C ≡ 1 and ϕ|Uc ≡ 0.

(2) given a locally Lipschitz function f : U → R, there exists a (globally) Lipschitz function

g : Rd → R such that f |C = g|C .

Proof. We leave the first part as an exercise. For the second part, we take ϕ as in the first part

and then set g = fϕ.

Proof of Theorem 7.10. We will assume for simplicity that d = m = 1. Fix C ⊆ U compact.

By Lemma 7.11 we can find (globally) Lipschitz functions σ, b on R such that σ|C = σ|C and

b|C = b|C . By Theorem 7.2, there exists a pathwise unique strong solution X to the SDEdXt = σ(Xt)dBt + b(Xt)dt,

X0 = x.

Set T = inft ≥ 0 : Xt /∈ C and denote by X the restriction of X to [0, T ). Then (X,T ) is a

local solution in C and if T <∞ then XT− exists in U \ C. Suppose that (X,T ) and (Y, S)

are both local solutions in C. Consider

f(t) = E(

sups≤S∧T∧t

|Xs − Ys|2).

Then using that b and σ are Lipschitz in C and applying Gronwall’s lemma, as in the proof of

Theorem 7.2, one shows that f ≡ 0, and thus Xt = Yt for all t ≤ S ∧ T , almost surely.

Now to build a local solution to the original SDE we approximate U by growing compact sets

Cn and ”glue” together the corresponding local solutions.

Take a sequence of compact sets (Cn)n≥1 increasing to U and construct for each n a local

solution (Xn, Tn) in Cn by the above procedure. Then if Tn <∞, we have that XnT−n∈ U \ Cn.

Since the times Tn are increasing, we have Tn T := supn Tn. Then we can define a local

solution (X, T ) to the original SDE by setting Xt = Xnt whenever t < Tn, where the definition

is well posed since Xnt = Xn+1

t for all t < Tn and all n ≥ 1.

We will now show that (X, T ) is a maximal solution. Suppose that (Y, η) is another local

solution on the same probability space and, for each n, set6 Sn = inft ≤ η : Yt /∈ Cn∧η. Then

by the uniqueness of the solution on each Cn, we have that Xt = Yt for all t < Tn ∧ Sn, and

therefore Sn ≤ Tn. As n→∞, we have that Sn η and Tn T , hence η ≤ T and Xt = Yt

for all t ≤ η.

6Here we are using the convention that inf ∅ = +∞.

Page 58: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

58 JASON MILLER AND VITTORIA SILVESTRI

We now prove the final assertion of the theorem. Suppose that C1, C2 are compact sets in U

with C1 ⊆ C2 ⊆ C2 ⊆ U and let ϕ : U → R be a C∞ function with ϕ|C1 ≡ 1 and ϕ|(C2 )c ≡ 0.

We then set

R1 = inft ≥ 0 : Xt /∈ C2

Sn = inft ∈ [Rn, T ) : Xt ∈ C1 ∧ T

Rn = inft ∈ [Sn−1, T ) : Xt /∈ C2 ∧ T .

Let N be the number of crossings that X makes from C2 to C1. On T ≤ t,N ≥ n we have

that

n =n∑k=1

(ϕ(XSk

)− ϕ(XRk))

=

∫ t

0

n∑k=1

1(Rk,Sk](s)(ϕ′(Xs)dXs +

1

2ϕ′′(Xs)d[X]s

)=

∫ t

0

(Hns dBs +Kn

s ds)

:= Znt

where Hn and Kn are previsible and bounded uniformly in n. This implies that

n21T ≤t,N≥n ≤ (Znt )2

and hence

P(T ≤ t,N ≥ n

)≤ 1

n2E((Znt )2

).

Note that E((Znt )2) ≤ C < ∞ where C is independent of n as Hn and Kn are bounded

uniformly in n and Znt is defined by integrating Hn,Kn over a time-interval which does not

depend on n. Thus P(T ≤ t,N ≥ n

)≤ C/n2, which is summable, so by Borel-Cantelli we see

that P(T ≤ t,N =∞) = 0. Therefore, on the event T <∞, X eventually stops coming back

to C1, which shows that supt < T : Xt ∈ C1 < T .

Example 7.12 (Bessel processes). Fix ν ∈ R. Consider the SDE in U = (0,∞) given by

dXt = dBt +

(ν − 1

2Xt

)dt, X0 = x0 ∈ U.

Theorem 7.10 implies that there exists a unique maximal local solution (X, T ) in U with

T = inft ≥ 0 : Xt = 0. We call (X, T ) a Bessel process of dimension ν.

Suppose that ν ∈ N and let β be a Brownian motion in Rν with |β0| = x0 > 0. Set Yt = |βt|and η = inft ≥ 0 : βt = 0. By the local Ito formula, we have that

dYt =(βt, dβt)

|βt|+

(ν − 1

2|βt|

)dt, t < η.

Page 59: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 59

Consider Wt =∫ t

0(βs,dβs)|β|s ds for t ≥ 0. Then W ∈Mc,loc and

d[W ]t =1

|βt|2ν∑

i,j=1

βitβjt d[βi, βj ]t = dt.

Therefore, by Levy’s characterization of Brownian motion, we have that W is a standard

Brownian motion in R. It follows that Y solves the SDE

dYt = dWt +

(ν − 1

2Yt

)dt, t < η.

Remark 7.13. This shows that a Bessel process of dimension ν ∈ N+ describes the evolution

of the Euclidean norm of a ν-dimensional Brownian motion, up to the first time it hits 0.

8. Diffusion processes

Suppose that we are given bounded, measurable functions a : Rd →M d×d(R) and b : Rd → Rd

with a symmetric (i.e., each for each x ∈ Rd the matrix a(x) satisfies a(x) = (a(x))T ). For

f ∈ C2b (Rd) (i.e., the space of C2(Rd) bounded functions with bounded derivatives), we set

(8.1) Lf(x) =1

2

d∑i,j=1

aij(x)∂2f

∂xi∂xj(x) +

d∑i=1

bi(x)∂f

∂xi(x).

Let X be a continuous, adapted process in Rd. We say that X is an L-diffusion if for all

f ∈ C2b (Rd), we have that

Mft := f(Xt)− f(X0)−

∫ t

0Lf(Xs)ds

is a martingale. We call a the diffusivity and b the drift. If a and b are clear from the context,

we will sometimes refer to an L-diffusion as an (a, b)-diffusion.

Example 8.1. Suppose that σ and b are constant and a = σσT . Let B be a standard Brownian

motion in Rd. Then the process Xt = σBt + bt is an (a, b)-diffusion.

One special case of this is when σ = I and b = 0. Then Xt = Bt is an L-diffusion with L = 12∆.

The next result tells us how to construct L-diffusions from SDEs.

Proposition 8.2. Suppose that X is a solution to the SDE

(8.2) dXt = σ(Xt)dBt + b(Xt)dt.

Let f ∈ C1,2b (R+ × Rd) (i.e., the space of bounded functions which are C1 in the first variable

and C2 in the remaining d-variables, with bounded derivatives). Then the process

Mft = f(t,Xt)− f(0, X0)−

∫ t

0

(∂

∂s+ L

)f(s,Xs)ds

Page 60: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

60 JASON MILLER AND VITTORIA SILVESTRI

is a continuous local martingale, where a = σσT and L is as in (8.1). In particular, if σ, b are

bounded, then X is an L-diffusion.

Proof. See Example Sheet 3.

Proposition 8.2 tells us that, in particular, if X solves (8.2) with bounded and Lipschitz b, σ,

then X is a (σσT , b)-diffusion. This raises the question of which a can be written as a = σσT

for σ as above.

We suppose that a, b are Lipschitz and bounded and that a is uniformly positive definite (UPD),

that is there exists ε > 0 such that

(8.3) (ξ, a(x)ξ) ≥ ε2|ξ|2 for all x, ξ ∈ Rd.

Then there exists a Lipschitz map σ : Rd → Rd with σσT = a. Indeed, if d = 1 then we can take

σ =√a. For d ≥ 2, we can write a(x) = u(x)λ(x)(u(x))T where λ(x) is the diagonal matrix of

eigenvalues of a(x) and u(x) is an orthogonal matrix having the corresponding eigenvectors

as columns. So, in this case, we can take σ(x) = u(x)√λ(x)(u(x))T . The fact that such σ is

again Lipschitz follows from the differentiability of the square root function on the set of UPD

matrices, and we take it for granted. For such b and σ, Theorem 7.2 implies that the SDE

dXt = σ(Xt)dBt + b(Xt)dt has a unique strong solution, which is then an (a, b)-diffusion. Thus

we can use SDEs to build (a, b) diffusions for all a, b Lipschitz, bounded and such that a is

UPD.

The next result tells us that restarting an L-diffusion at a finite stopping time gives again an

L-diffusion.

Proposition 8.3. Let X be an L-diffusion and T a finite stopping time. Set Xt = XT+t and

Ft = FT+t. Then X is an L-diffusion with respect to (Ft)t≥0.

Proof. Consider for f ∈ C2b (Rd) the process

Mft = f(Xt)− f(X0)−

∫ t

0Lf(Xs)ds.

Then Mft is Ft-adapted and integrable. For A ∈ Fs and n ≥ 0, we note that

E((Mf

t − Mfs )1A∩T≤n

)= E

((Mf

T+t −MfT+s)1A∩T≤n

)= E

((Mf

(T∧n)+t −Mf(T∧n)+s)1A

)= 0

by the OST. Letting n→∞ and using the Dominated Convergence Theorem, we see that Mf

is a martingale.

Page 61: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 61

Lemma 8.4. Let X be an L-diffusion. Then for all f ∈ C1,2b (R+ × Rd), the process

Mft = f(t,Xt)− f(0, X0)−

∫ t

0

(∂

∂s+ L

)f(s,Xs)ds

is a martingale.

Proof. We fix T > 0 and consider

Zn = sup0≤s≤t≤Tt−s≤1/n

|f(s,Xt)− f(s,Xs)|+ sup0≤s≤t≤Tt−s≤1/n

|Lf(s,Xt)− Lf(t,Xt)|.

Then Zn is bounded and Zn → 0 as n→∞ by uniform continuity of continuous functions on

compact sets. Consequently, E(Zn)→ 0 as n→∞ by the Bounded Convergence Theorem. We

note that:

Mft −Mf

s =

(f(t,Xt)− f(s,Xt)−

∫ t

sf(r,Xt)dr

)+(

f(s,Xt)− f(s,Xs)−∫ t

sLf(s,Xr)dr

)+∫ t

s

(f(r,Xt)− f(r,Xr)

)dr +

∫ t

s

(Lf(s,Xr)− Lf(r,Xr)

)dr.

Choose s0 ≤ s1 ≤ · · · ≤ sm with s0 = s and sm = t and sk+1 − sk ≤ 1/n. As the first line of

our formula for Mft −M

fs is zero and the second line has zero conditional expectation given Fs

(as X is an L-diffusion), it follows that

E(∣∣∣E(Mf

sk+1−Mf

sk| Fs)∣∣∣) ≤ (sk+1 − sk)E(Zn).

Combining this with the triangle inequality, we thus see that

E(∣∣∣E(Mf

t −Mfs | Fs

)∣∣∣) ≤ (t− s)E(Zn).

As the right hand side goes to zero as n→∞, it follows that E(Mft | Fs) = Mf

s , as desired.

8.1. Dirichlet and Cauchy problems. Assume that a, b are Lipschitz and that a is UPD,

i.e. (ξ, a(x)ξ) ≥ ε|ξ|2 for all x, ξ ∈ Rd. Let D be an bounded, open subset of Rd with smooth

boundary ∂D.

8.1.1. Dirichlet problems. We shall assume the following result from the theory of PDEs.

Theorem 8.5 (Dirichlet problem). For all f ∈ C(∂D), ϕ ∈ C(D), there exists a unique

function u ∈ C(D) ∩ C2(D) such thatLu+ ϕ = 0 on D

u = f on ∂D.

Page 62: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

62 JASON MILLER AND VITTORIA SILVESTRI

Moreover, there exist continuous functions m : D × ∂D → (0,∞) and g : (x, y) ∈ D ×D : x 6=y → (0,∞) such that for all f and ϕ as above we have that

(8.4) u(x) =

∫Dg(x, y)ϕ(y)dy +

∫∂D

f(y)m(x, y)λ(dy).

We call g the Green kernel and m(x, y)λ(dy) the harmonic measure on ∂D starting from x.

Theorem 8.6. Suppose that u ∈ C(D) ∩ C2(D) satisfiesLu+ ϕ = 0 on D

u = f on ∂D,

with f ∈ C(∂D), ϕ ∈ C(D). Then for any L-diffusion X starting from x ∈ D, we have that

u(x) = Ex(∫ T

0ϕ(Xs)ds+ f(XT )

)where T = inft ≥ 0 : Xt /∈ D. In particular, for all Borel sets A ⊆ D and B ⊆ ∂D, we have

that

Ex(∫ T

01(Xs ∈ A)ds

)=

∫Ag(x, y)dy(8.5)

Px(XT ∈ B) =

∫Bm(x, y)λ(dy).(8.6)

Proof. Fix n ≥ 1 and set Tn = inft ≥ 0 : Xt /∈ Dn where Dn = x ∈ D : dist(x,Dc) > 1/nand consider the process

Mt = u(Xt∧Tn)− u(X0) +

∫ t∧Tn

0ϕ(Xs)ds.

There exists u ∈ C2b (Rd) with u = u on Dn. Then M = (M)Tn where

Mt = u(Xt)− u(X0)−∫ t

0Lu(Xs)ds.

Now, since X is an L-diffusion we have that M u is a martingale, so by the OST we conclude

that M is a martingale. Hence,

(8.7) u(x) = Ex(u(Xt∧Tn) +

∫ t∧Tn

0ϕ(Xs)ds

).

In order to send n→∞ in both terms, we first show that Ex(T ) <∞. To this end, take ϕ ≡ 1

and f ≡ 0, and let u1,0 denote the solution to the associated Dirichlet problem. Then (8.7)

holds for u1,0, and so

Ex(Tn ∧ t) = u1,0(x)− Ex(u1,0(Xt∧Tn)).

Since u1,0 is bounded and Tn T , we deduce by the Monotone Convergence Theorem (with

t = n→∞) that Ex(T ) <∞.

Page 63: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 63

Returning to the general case, we can let t→∞ and n→∞ in (8.7). Then t ∧ Tn T almost

surely. Since u is continuous on D, we have that u(Xt∧Tn)→ f(XT ). Finally, u is also bounded

on D since continuous function on a compact set, so by the Bounded Convergence Theorem we

find

Ex(u(Xt∧Tn))→ Ex(f(XT ))

as t, n→∞. Moreover,

Ex(∫ T

0|ϕ(Xs)|ds

)≤ ‖ϕ‖∞Ex(T ) <∞.

Thus by the Dominated Convergence Theorem we find

Ex(∫ t∧Tn

0ϕ(Xs)ds

)→ Ex

(∫ T

0ϕ(Xs)ds

).

Consequently,

u(x) = Ex(f(XT ) +

∫ T

0ϕ(Xs)ds

).

Finally, combining this with (8.4) we have shown that for all ϕ ∈ C(D) and f ∈ C(∂D) it holds

Ex(f(XT ) +

∫ T

0ϕ(Xs)ds

)=

∫∂D

m(x, y)f(y)λ(dy) +

∫Dg(x, y)ϕ(y)dy.

Taking f ≡ 0 and ϕn → 1A with ϕn ∈ C(D) we get (8.5). Similarly, taking ϕ ≡ 0 and fn → 1B

with fn ∈ C(∂D) yields (8.6).

8.1.2. Cauchy problems. We shall assume the following result from the theory of PDEs.

Theorem 8.7 (Cauchy problem). For each f ∈ C2b (Rd), there exists a unique u ∈ C1,2

b (R+×Rd)such that

∂u

∂t= Lu on R+ × Rd

u(0, ·) = f on Rd.

Moreover, there exists a continuous function p : (0,∞)× Rd × Rd → (0,∞) such that

u(t, x) =

∫Rd

p(t, x, y)f(y)dy

for all (t, x) ∈ R+ × Rd.

In the setting of Theorem 8.7, the function p is called the heat kernel.

Page 64: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

64 JASON MILLER AND VITTORIA SILVESTRI

Theorem 8.8. Assume that f ∈ C2b (Rd). Let u ∈ C1,2

b (R+ × Rd) satisfy∂u

∂t= Lu on R+ × Rd

u(0, ·) = f on Rd.

Then, for any L-diffusion X, for all t ∈ R+, x ∈ Rd, s ≤ t, we have that E(f(Xt) | Fs) =

u(t− s,Xs) almost surely. In particular,

Ex(f(Xt)) = u(t, x) =

∫Rd

p(t, x, y)f(y)dy.

Moreover, under Px the finite dimensional distributions of X are given by

Px(Xt1 ∈ dx1, . . . , Xtn ∈ dxn) = p(t1, x0, x1) · · · p(tn − tn−1, Xtn−1 , Xtn)dx1 · · · dxn

for 0 < t1 < · · · < tn and x1, . . . , xn ∈ Rd.

Remark 8.9. In the language of Markov processes, the above theorem tells us that X is a

Markov process with transition density function p.

Proof. Fix t ∈ (0,∞) and consider the function g(s, x) = u(t− s, x) for s ≤ t. Note that(∂

∂s+ L

)g(s, x) = −u(t− s, x) + Lu(t− s, x) = 0.

Consequently, Mgs = g(s,Xs)− g(0, X0) is a martingale. This implies that E(Mg

t |Fs) = E(Mgs )

almost surely, i.e. E(f(Xt) | Fs) = u(t − s,Xs). Taking s = 0 and X0 = x implies that

E(f(Xt)) = u(t, x). This proves the first part of the theorem.

For the second part of the theorem, we set

Ptf(x) =

∫Rd

p(t, x, y)f(y)dy = u(t, x).

By the uniqueness of solutions to the Cauchy problem, we have that Ps(Ptf) = Ps+tf . We

claim by induction on n that

Ex0

(n∏i=1

fi(Xti)

)=

∫(Rd)n

p(t1, x0, x1)f1(x1) · · · p(tn − tn−1, xn−1, xn)fn(xn)dx1 · · · dxn.

For the induction step, we use that

Ex0

(n∏i=1

fi(Xti)∣∣∣Ftn−1

)=

(n−1∏i=1

fi(Xti)

)E(fn(Xtn) | Ftn−1)

=

(n−1∏i=1

fi(Xti)

)Ptn−tn−1fn(Xtn−1),

and then apply the n− 1 case.

Page 65: STOCHASTIC CALCULUS1. Introduction 1 2. Preliminaries 4 3. Local martingales 10 4. The stochastic integral 16 5. Stochastic calculus 36 6. Applications 44 7. Stochastic di erential

STOCHASTIC CALCULUS 65

Theorem 8.10 (Feynman-Kac). Let f ∈ C2b (Rd) and V ∈ Cb(Rd). Suppose that u ∈ C1,2

b (R+×Rd) satisfies the PDE

∂u

∂t=

1

2∆u+ V u on R+ × Rd and u(0, ·) = f on Rd.

Then for all t ∈ R+, x ∈ Rd we have that

u(t, x) = Ex(f(Bt) exp

(∫ t

0V (Bs)ds

))where B is a standard Brownian motion.

Proof. Let

Et = exp

(∫ t

0V (Bs)ds

).

Fix T ∈ (0,∞) and set Mt = u(T − t, Bt)Et. By Ito’s formula, we have that

dMt =d∑i=1

∂u

∂xi(T − t, Bt)dBi

t +

(1

2∆u(T − t, Bt)− u(T − t, Bt) + u(T − t, Bt)V (Bt)

)︸ ︷︷ ︸

=0

Etdt.

So, M is a local martingale which is uniformly bounded on [0, T ] as u ∈ C1,2b (R+ × Rd) and V

is bounded on [0, T ]. Therefore M is a martingale. Hence,

u(T, x) = M0 = Ex(MT ) = Ex(f(BT )ET ),

as desired.

Statslab, Center of Mathematical Sciences, University of Cambridge, Wilberforce Road, Cam-

bridge CB3 0WB, UK