elementary stochastic calculus with finance in view_thomas mikosch (world 1998 212p)

112

Click here to load reader

Upload: juan-camilo-achury-urrego

Post on 05-Mar-2015

1.349 views

Category:

Documents


124 download

TRANSCRIPT

Page 1: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

ELEMENTARY

STOCHASTIC

CACULUS with Finance in View

Page 2: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

ADVANCED SERIES ON STATISTICAL SCIENCE & APPLIED PROBABILITY

Editor: Ole E. Barndorff-Nielsen

Published Vol. 1: Random Walks of Infinitely Many Particles

by P. Revesz

Vol. 4: Principles of Statistical Inference from a Neo-Fisherian Perspective by L. Pace and A. Salvan

Vol. 5: Local Stereology by Eva 6. Vedel Jensen

Vol. 6: Elementary Stochastic Calculus - With Finance in View by T. Mikosch

Vol. 7: Stochastic Methods in Hydrology: Rain, Landforms and Floods eds. 0. E. Barndofl-Nielsen et a/.

Forthcoming

Vol. 2: Ruin Probability by S. Asmussen

Vol. 3: Essentials of Stochastic Finance by A. Shiryaev

ELEMENTARY

STOCHASTIC

CALCULUS with Finance in View

Thomas Mikosch Department of Mathematics

University of Groningen The Netherlands

World Scientific b S~ngapore New Jersey* London Hong Kong

Page 3: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Published by

World Scientific Publishing Co. Pte. Ltd.

P 0 Box 128, Farrer Road, Singapore 9 12805

USA ofice: Suite IB, 1060 Main Street, River Mge, NJ 07661

UK ofice: 57 Shelton Street, Covent Garden, London WC2H 9HE

Library of Congress Cataloging-in-Publication Data Mikosch, Thomas.

Elementary stochastic calculus with finance in view /Thomas Mikosch.

p. cm. -- (Advanced series on statistical science & applied probability ; v. 6 )

Includes bibliographical references and index. ISBN 9810235437 (alk. paper) 1. Stochastic analysis. I. Title. 11. Series.

QA274.2.M54 1998 519.2--dc21 98-2635 1

CIP

British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library

First published 1998 Reprinted 1999

Copyright O 1998 by World Scientific Publishing Co. Pte. Ltd.

All rights reserved. This book, or parts thereof; muy nut be reproduced in unyfurnl or by any nheans. electronic or mechanical, including photocopying, recordrng or any information storage and retrieval system now known or to be invented, without written permission from the Publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center, lnc., 222 Rosewood Drive, Uanvers, MA 01923, USA. In this case permission to photocopy is not required from the publisher

Printed in Singapore.

Preface

Ten years ago I would not have dared to write a book like this: a non-rigorous treatment of a mathematical theory. I adrr~it that I would have been ashamed, and I am afraid that most of my colleagues in mathematics still think like this. However, my experience with students and practitioners convinced me that there is a strong demand for popular mathematics.

I started writing this book as lecture notes in 1992 when I prepared a course on stochastic calculus for the students of the Commerce Faculty at Victoria University Wellington (New Zealand). Since I had failed in giving tutorials on portfolio theory and investment analysis, I was expected to teach something I knew better. At that time, staff members of economics and mathematics departments already discussed the usc of the Black and Scholes option pricing formula; courses on stochastic fir~ar~ce were offered at leading institutions such as ETH Ziirich, Columbia and Stanford; and there was a general agreement that not only students and staff members of economics and mathematics de- partments, but also practitioners in financial institutions should know more ahout this new topic.

Soon I realized that there was not very much literature which could be used for teaching stochastic calculus at a rather elementary level. I am fully aware of the fact that a combination of "elementary" and LLstochastic calculus" is a contradiction in itself. Stochastic calculus requires advanced mathematical techniques; this theory cannot be fully understood if one does not know about the basics of measure theory, functional analysis and the theory of stochastic processes. However, I strongly believe that an interested person who knows about elementary probability theory and who can handle the rules of inte- gration and differentiation is able to understand the main ideas of stochastic calculus. This is supported by my experience which I gained in courses for economics, statistics and mathematics students at VUW Wellington and the Department of Mathematics in Groningen. I got the same impression as a lecturer of crash courses on stochastic calculus at the Summer School of the

Page 4: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Swiss Association of Actuaries in Lausanne 1994, the Workshop on Financial Mathematics in Groningen 1997 and at the University of Leuven in May 1998.

Various colleagues, friends and students had read my 1ect11re notes a.nd suggested that I extend them to a small book. Among those are Claudia Kliippelberg and Paul Embrechts, my coauthors from a book about extremal events, and David Vere-Jones, my former colleague at the Institute of Statis- tics and Operations Research in Wellington. Claudia also proposed to get in contact with Ole Barndorff-Nielsen who is the editor of the probability series of World Scicntific. I am indebted to him for encouraging me throughout the long process of writing this book.

Many colleagues and students helped in proofreading parts of the book at various stages. In particular, I would like to thank Leigh Roberts from Wellington, Bojan Basrak and Diemer Salome from Groningen. Their criticism was very helpful. I am most grateful to Carole Proctor from Sussex University. She was a constant source of inspiration, both on stylistic and mathematical issues. I also take pleasure in thanking the Department of Mathematics a t the University of Groningen, my colleagues and students for their much appreciated support.

Thomas Mikosch Groningen, June 1, 1998

Contents

Reader Guidelines 1

1 Preliminaries 5 1.1 Basic Concepts from Probability Theory . . . . . . . . . . . . . 6

1.1.1 Random Variables . . . . . . . . . . . . . . . . . . . . . 6 1.1.2 Random Vectors . . . . . . . . . . . . . . . . . . . . . . 14 1.1.3 Independence and Dependence . . . . . . . . . . . . . . 19

1.2 Stochastic Processes . . . . . . . . . . . . . . . . . . . . . . . . 23 1.3 Brownian Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 33

1.3.1 Defining Properties . . . . . . . . . . . . . . . . . . . . . 33 1.3.2 Processes Derived from Brownian blotion . . . . . . . . 40 1.3.3 Simulation of Brownian Sample Paths . . . . . . . . . . 44

1.4 Conditional Expectation . . . . . . . . . . . . . . . . . . . . . . 56 1.4.1 Conditional Espectatio~l under Discrete Condit,ion . . . 56 1.4.2 About a-Fields . . . . . . . . . . . . . . . . . . . . . . . 62 1.4.3 The General Conditional Expectation . . . . . . . . . . 67 1.4.4 Rules for the Calculation of Conditional Expectations . 70 1.4.5 The Projection Property of Conditional Expectations . 74

1.5 Martingales . . . . . . . . . . . . . . . . . . . . . . .... . . . . 77 1.5.1 Defir~ir~g Properties . . . . . . . . . . . . . . . . . . . . . 77 1.5.2 Exarrlples . . . . . . . . . . . . . . . . . . . .... . . . . 81 1.5.3 The Interpretation of a Martingale as a Fair Game . . . 84

2 The Stochastic Integral 8 7 2.1 The Riemann and Riernarlll-Stieltjes Integrals . . . . . . . . . . 88

2.1.1 The Ordinary Riemann Integral . . . . . . . . . . . . . . 88 2.1.2 The Riemann-Stieltjes Integral . . . . . . . . . . . . . . 92

2.2 The It6 Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 2.2.1 A Motivating Example . . . . . . . . . . . . . . . . . . . 96

Page 5: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

... v11l CONTENTS

2.2.2 The It6 Stochastic Integral for Simple Processes . . . . 101 2.2.3 The General It8 Stochastic Integral . . . . . . . . . . . . 107

2.3 The Tt. ri Ternma . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 . . . . . . . 2.3.1 The Classical Chain Rule of Differentiation 113

2.3.2 A Simple Version of the l to Lemma . . . . . . . . . . . 114 2.3.3 Extended Versions of the I t6 Lemma . . . . . . . . . . . 117

2.4 The Stratonovich and Other Integrals . . . . . . . . . . . . . . 123

3 Stochastic Differential Equations 131 3.1 Deterministic Differential Equations . . . . . . . . . . . . . . . 132 3.2 It6 Stochastic Differential Equations . . . . . . . . . . . . . . . 134

3.2.1 What is a Stochastic Differential Equation? . . . . . . . 134 3.2.2 Solving It6 Stochastic Differential Equations by the It6

Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 3.2.3 Solving It6 Differential Equations via Stratonovich Cal-

culus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 3.3 The General Linear Differential Equation . . . . . . . . . . . . 150

3.3.1 Linear Equations with Additive Noise . . . . . . . . . . 150 3.3.2 Homogeneolls Equations with Multiplicative Noise . . . 153 3.3.3 The General Case . . . . . . . . . . . . . . . . . . . . . 155 3.3.4 The Expectation and Variance Functions of the Solution 156

3.4 Numerical Solution . . . . . . . . . . . . . . . . . . . . . . . . . 157 3.4.1 The Euler Approximation . . . . . . . . . . . . . . . . . 156 3.4.2 The Milstein m appro xi ma ti or^ . . . . . . . . . . . . . . . 162

4 Applications of Stochastic Calculus in Finance 167 4.1 The Black-Scholes Option Pricing Formula . . . . . . . . . . . 168

4.1.1 A Short Excursion into Finance . . . . . . . . . . . . . . 168 4.1.2 What is an Option? . . . . . . . . . . . . . . . . . . . . 170 4.1.3 A Mathematical Formulation of the Option Pricing Pro-

blem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 . . . . . . . . . . . . . . 4.1.4 The Black and Scholes Formula 174

. . . . . . . . . . . . . 4.2 A Useful Technique: Change of Measure 176 . . . . . 4.2.1 What is a Change of the Underlying Measure? 176

4.2.2 An Interpretation of the Black-Scholes Formula by Chan- . . . . . . . . . . . . . . . . . . . . . . . . ge of Measure 180

Appendix 185 A1 Modes of Convergence . . . . . . . . . . . . . . . . . . . . . . . 185 A2 Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

CONTENTS ix

A3 Non-Differentiability and Unbounded Variation of Brownian Sanl- . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ple Paths 188

A4 Proof of the Existence of the General it6 Stochastic Integral . . 190 . . . . . . . . . . . . . . . . . . A5 The Radon-Nikodym Theorem 193

A6 Proof of the Existence and Uniqueness of the Conditional Ex- . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . pectation 194

Bibliography 195

Index 199

List of Abbreviations and Symbols 209

Page 6: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Reader Guidelines

This book grew out of lecture notes for a course on stochastic calculus for eco- nomics studcnts. When I prepared the first lectures I realized that there was no adequate textbook treatment for non-mathematicians. On the other hand, there was and indeed is an increasing demand to learn about stochastic cal- culus, in particular in economics, insurance, finance, econometrics. The main reason for this interest originates from the fact that this mathematical theory is the basis for pricing financial derivatives such as options and futures. The fundamental idea of Black, Scholes and Merton from 1973 to use It6 stochas- tic calculus for pricing and hedging of derivative instruments has conquered the real world of finance; the Black-Scholes formula has been known to many people in mathematics and economics long before Merton and Scholes were awarded the Nobel prize for economics in 1997.

For whom is this book written? ! I In contrast to the increasing popularity of financial mathematics, its theoretical

basis is by no means trivial. Who ever tried to read the first few pagcs of a book on stochastic calculus will certainly agree. Tools from measure theory and functional analysis are usually required.

In this book I have tried to keep the mathematical level low. The reader 1 will not be burdened with measure theory, but it cannot be avoided altogether.

Then we will have to rely on heuristic arguments, stressing the underlying ideas rather than technical details. Notions such as measurable function and measurable set are not introduced, and therefore the formulation and proof of various statements and results are necessarily incomplete or non-rigorous. This may sometimes discourage the mathematically oriented reader, but for those, excellent mathematical textbooks on stochastic calculus exist.

In discussions with economists and practitioners from banks and insur- ance companies I frequently listened to the argument: "It6 calculus can be

Page 7: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

2 READER G UTT>ELIIC%S

understood 07111/ by niathernaticians." It is the main objective of this book to overcome this superstition.

Every mathematical theory has its routs i11 real life. Therefore t,he notions of It8 integral, It.6 lemma and stochastic differential equation can be explained to anybody who ever attended courses on elementary calculus and probability theory: physicists, chemists, biologists, actuaries, engineers, economists, . . . . In the course of this book the reader will learn about the basic rules of stochastic calculus. Finally, you will be able to solve some simple stochastic differential equations, to sirnulate these solutions on a computer and to understand the mat,henlatical ideology behind the modern theory of option pricing.

What are the prerequisites for this book?

You should be familiar with the rules of integration and differentiation. Ideally, you also know about differential equations, but it is not essential. You must know about elementary probability theory. Chapter 1 will help you to recall some facts about probability, expectation, distribution, etc., but this will not be a proper basis for the rest of the book. You would be advised to read one of the recommended books on probability theory, if this is new to you, before you attempt to read this book.

How should you read this book?

It depends on your knowledge of probability theory. I reco~rimend tha.t you browse through the "boxes" of Chapter 1. If you know everything that is written there, you can start with Chapter 2 on It6 stochastic calculus and colitinue with Chapter 3 on stochastic differential equations.

You cannot proceed in this way if you are not familiar with the following basic notions: stochastic process, Brownian motion, conditional expectation and martingale. There is no doubt that you will struggle with the noJion of conditional expectation, unless you have some background on measure the- ory. Conditional expectation is one of the key notions underlying stochastic integration.

The ideal reader can handle simulations on a comput,er. Computer graphs of Brownian motion and solutions to stochastic differential equations will help you to experience the theory. The theoretical tools for these simulations will be provided in Sections 1.3.3 and 3.4.

I have not included lists of exercises, but I will ask you various questiorls in the course of this book. Try to answer them. They are riot difficult, but they aim at testing the level of your understanding.

Besides Sections 1.3-1.5, the core material is contained in Chapter 2 and the first sections of Chapter 3. Chapter 2 provides the construction of the It6 integral and a heuristic derivation of the It6 lemma, the chain rule of stochastic calculus. In Chapter 3 you will learn how to solve some simple stochastic differential equations. Section 3.3 on linear stochastic differential equation is mainly included in order to exercise the use of the It6 lemma. Section 3.4 will be interesting to those who want to visualize solutions to stochastic differential equations.

Chapter 4 is for those readers who want to scc how stochastic calculus enters financial applications. Prior knowledge of economic theory is not re- quired, but we will introduce a minimum of economic terminology which can be understood by everybody. If you can read through Section 4.1 on option pricing without major difficulties as regards stochastic calculus, you will have passed the examination on this course on elementary stochastic calculus.

At the end of this book you may want to know more about stochastic calculus and its applications. References to more advanced literature are given in the Notes and Comments a t the end of each section. These references are not exhaustive; they do not include the theoretically most advanced textbook treatments, but they can be useful for the continuation of your studies.

You are now ready to start. Good luck!

T.M.

R.EA DER. GUIDELINES 3

Page 8: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Preliminaries

In this chapter we collect some basic facts needed for defining stochastic inte- grals. At a first reading, most parts of this chapter can be skipped, provided you have some basic knowledge of probability thcory and stochastic processes. Yo11 ma.y then want to start with Chapter 2 on It6 stochastic calculus and recall some facts from this chapter if necessary.

I n Section 1.1 we recall elementary notions from probability theory such as random variable, random vector, distribution, distribution function, den- situ, expectation, m o m e n t , variance and covariance. This small review cannot replace a whole course on probability, and so you are well recommended to consult your old leclure riotes or a standard textbook. Section 1.2 is about stochastic processes. A stochastic process is a natural model for describing the evolution of real-life processes, objects and systems in time and space. One particular stochastic process plays a central role in this book: Brownian motion. We introduce it in Section 1.3 and discuss some of its elementary properties, in particular the non-differentiability and the unbounded variation of its sample paths. These properties indicate that Brownian sample paths are very irregular, and therefore a new, stochastic calculus has to be introduced for integrals with respect to Brownian motion.

In Section 1.4 we shortly review conditional expectations. Their precise definition is based or) a deep rriathematical theory, and therefore we only give some intuition on this concept. The same remark applies to Section 1.5, where we introduce an important class of stochastic processes: the martingales. It includes Brownian motion and indefinite Ito integrals as particular examples.

Page 9: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

6 CHAPTER 1. 1.1. BASIC CONCEPTS FROM PROBABILITY THEORY 7

1.1 Basic Concepts from Probability Theory

1.1.1 Random Variables

The outcome of a.n experiment or game is random. A simple example is coin tossing: the possible outcomes "head" or "tail" are not predictable in the sense that they appear according to a random mechanism which is determined by the physical properties of the coin. A more complicated experiment is the stock market. There the random outcomes of the brokers' activities (which actually represent economic tendencies, political interests and their own instincts) are for example share prices and exchange rates. Another game is called "com- petition" and can be watched where products are on sale: the price of 1 kg bananas, say, is the outcome of a game between the shop owners, on the one hand, and betwee11 1he shop owners and the customers, on the other hand.

The scientific treatment of an experiment requires that we assign a number to each random outcome. When tossing a coin, we can write "1" for "head" and "0" for "tail". Thus we get a random variable X = X(w) E {O,1), where w belongs to the outcome space R = {head,tail). The value of a share price of stock is already a random number, and so is the banana price in a greengrocers. These numbers X(w) provide us with information about the experiment, even if we do not know who plays the game or what drives it.

Mathematicians make a clear cut between reality and a mathematical model: they dcfine an abstract space 0 collecting all possible outcomes w of the underlying exper i~~ler~t . It is an abstract space, i.e. it does not really matter what the ws are. In mathematical language, the random variable X = X(w) is nothing but a real-valued function defined on R.

The next step in the process of abstraction from reality is the probabilistic description of the random variahle X:

Which are the mos t likelg values X(w), what are they concentrated around, what i s their spread?

To approach these problems, one first collects "good" subsets of R, the events, in a class 3 , say. In advanced textbooks 3 is called a u-field or u-algebra; see p. 62 for a precise definition. Such a class is supposed to contain all interesting events. What could F be for coin tossing? Certainly, {w : X ( w ) = 0) = {tail) and {u : X(w) = 1) = {head) must belong to F, but also the union, difference, intersection of any events in 3, the set R = {head,tail) and its complement, the empty set 0. This is a trivial example, but it shows what F should be like: if A F F , so is its complement AC, and if A, B E F, so are A n B , A u B, A u BC, B u Ac, il n Bc, B n 4" , etc.

If we consider a share price 9, not only the events {w : x(w) = c) should belong to F, but also

{ w : a < X(w) < b) , {w : b < Xjw) ) , {w: X(w) <a),

and many more event,s which can be relevant in this or that situation. As in the case of coin tossing, we would like that element.ary operations such as n,u,' on the events of 3 should not lead outside the class 3. This is the intuitive meaning of a a-field F.

Probability, Distribution and Distribution Function

Now, where do the probabilities come in? When flipping a coin, "head" or "tail" occurs. Probabilities measure the likelihood that these events happen. If the coin is "fair" we assign the probability 0.5 to both events, i.e. P({w : X(w) = 0)) = P({w : X(w) = 1)) = 0.5. This mathematical definition is based on empirical evidence: if we flip a fair coin a large number of times, we expect that about 50% of the outcomes are heads and about 50% are tails. In probability theory, the law of large numbers gives the theoretical justification for this empirical observation.

This elementary example explains what a probability measure on the class 3 of the events is: t o each event A E 3 it assigns a number P ( A ) E [0, 11. This number is the expected fraction of occurrences of the event A in a long series of experiments where A or AC are observed.

Some elementary properties of probability measures are easily summarized:

For events A, B E 3-,

P ( A U B ) = P(A) + P ( B ) - P ( A n B) ,

and, if A and B are disjoint,

P ( A U B) = P(A) + P ( B ) .

Moreover,

P (Ac) = 1 - P(A) , P ( R ) = 1 and P(L4) = 0 .

The relationship between random variables and probability can be character- I ized by certain numerical quantities. In what follows, we consider some of

them. I

Page 10: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

CHAPTER 1. I . I . R.4STC CONCEPTS FR.OA/I PR.OBAB1LZTY THE0R.Y

The collection of the probabilities

Fx(x) = P ( X 5 x) = P({w : X(w) 5 x)) , x E R = ( - m , m ) ,

is the di.strihirtion firnction F,Y of X .

It yields the probability that X belongs to the interval (a, b]. Indeed,

Moreover, we also obtain the probability that X is equal to a number:

P ( X = x)

= P({w : X(w) = 2)) = P({w : X 5 x) ) - P({w : X < 2))

= P({w : X(w) < x)) - lim P({w : X(w) 5 a: - h)) hi0

= Fx (2) - lim Fx (X - h) . h&O

With these probabilities one can approximate the probability of the event {w : X(w) E B) for very complicated subsets B of R.

The collection of the probabilities

P,y(B) = P ( X E B ) = P({w : X(w) E B))

for suitable subsets B C R is the distribution of X .

"Suitable" subsets of R are the so-called Borel sets. They are obtained by a countable number of operations n, U or acting on the intervals; see p. 64 for a precise definition.

The distribution P.Y. and the distribution function Fx are equivalent no- tions in the sense that both of them can be used to calculate the probability of any event { X E B).

A distribution function is continuous or it has jumps. We first consider the special case when the distribution function Fx is a pure jump function:

F x ( x ) = x Pk, X E R , (I,1) k : x k s x

where 00

O 5 p c 5 1 forall k a n d xPh=l. k= 1

The distribution function (1.1) and the corresponding distribution are sa.id t,o he discrete; a. ra.ndom va.ria.hle wit,h distrihi~tion function (1.1) is a discrete random variable.

A discrete random variable only assumes a finite or countably infinite number of values X I , X ~ , . . ., and pk = P(,Y = xk). In particular, the distribution function Fx has an upward jump of size pk a t x = xh. For example, the random variable X related to coin tossing is discrete: it assumes the values 0 and 1. The price for any product on sale in a supermarket is a discrete random variable: it may assume the values 0 ,0.01,0.02, . . . $, say.

Example 1.1.1 (Two important discrete distributions) Important discrete distributions are the binomial distribution Bin(n,p) with parameters n E N = {0,1,2 , . . .) and p E (0 , l ) :

and the Poisson distribution PoiiX) with parameter X > 0:

See Figure 1.1.2 for an illustration.

In contrast to discrete distributions and random variables, the distribution function of a continuous random variable does not have jumps, hence P ( X = .x) = 0 for all x, or equivalently,

lim Fx (x + h) = F x (x) for all x, h+O

(1.2)

i.e. such a random variable assumes any particular value with probability 0. A continuous random variable gains its name from the continuity property (1.2) of the distribution function Fx.

Page 11: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

10 CHAPTER 1 1.1. BASIC CONCEPTS FROM PROBABILITY THEORY 11

Figure 1.1.2 Left: the probabilities P ( X = k ) , k = 0,1,2,. . ., of the Poisson distri- bution with parameter A = 10. Right: the corresponding distribution function.

Figure 1.1.3 Left: the density of the standard normal distribution ( m e a n 0, variance 1 ) . Right: the cor r . c spo~rd i~~~ distl-ib.utiorr function.

Figure 1.1.4 Left: the denszty of the log-returns Xt = In Yt - In Yt-I of the daily closing prices l't of the S&P index. Thc S&P is one of the basic US industrial in- dzces. Right: the cor~esponding distribution junction. A comparison with Figure 1 . I .3 zndicates that the latter distribution as certainly not normal.

Most continuous distributions of interest have a dens i t y f x :

I where

f x ( x ) > 0 for every x E R and

Example 1.1.5 (The normal and uniform distributions) An important continuous distribution is the n o r m a l or Gauss ian d is t r ibut ion N ( p , a" with parameters p E R, aa > 0. I t has density

If X is N ( 0 , l ) (standard normal) we write p for the density f x and cP for the distribut,ion function F,y. For a.n ill~ist,rat,ion of t,he st,a.nda.rd normal densihy and the corresponding distribution function, see Figure 1.1.3.

The u n i f o r m distribution U ( a , b ) on ( a , b) has density

( 0 , otherwise.

The value of an exchange rate or share price can, a t least theoretically, assume any positive real number. Clearly, there are technical limitations: a computer or pocket calculator is not capable of saving the value of an exchange rate with infinitely many digits, fi say; every number in the computer's memory is rounded off. Therefore any random variable of practical interest is actually discrete. . . . However, it is often convenient to think of such a variable as a continuous one. There may be theoretical reasorls. For example, the normal distribution appears as a limit distribution via the central limit theorem; see p. 45. Many functions of a sample are therefore approximately normal, hence their limit distribution is continuous. But there are also practical reasons: it is often less tedious to work with a well-studied continuous distribution (such as t h ~ normal, exponential, gamma, uniform) because we can use standard knowl- edge (in particular, standard software packages) about its density, moments, quantiles, etc., and we can possibly o b t a n some nice expl~clt expressions for these quantities.

Page 12: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

12 CHAPTER 1. 1 .l. BASIC CONCEPTS FROM PROBABILITY THEORY 13

Expectation, Variance and Moments

Interesting characteristics of a random variable X are the expectation E X , the variance var(X) and the moments E ( X L ) .

The expectation or mean value of a random variable X with density fx is given by

03

p x = E X = Lrn x.~x(x) d x .

The variance of X is defined as 00

u i = var(X) = (z - f x ( x ) d x . 1, The l th moment of X for 1 E M is defined as

03

E ( x L ) = xL f,y(x) d x . J_, For a real-valued function g the expectation of g(X) is given by

03

The expectation or mean value of a discrete random variable X with probabilities pk = P ( X = xk) is given by

M

p,y = E X = ~ X ~ ~ ~ . k=l

The variance of X is defined as

M

u i = var(X) = x ( x k - px) \~k . k= 1

The lth moment of X for 1 E M is defined as

00

E(X" = C X : ~ ~ . $= 1

We can regard the expectation p x as the "center of gravity" of the random variable X , i.e. the random values X(w) are concentrated around the non- random number p x . The expectation is quite often taken as a surrogate for the size of the random variable. For example, it is a simple means for predicting the future values in a time series.

The spread or dispersion of the random values X(w) around the expecta- tion p x is described by the variance

and the standard deviataon u x . Recall the normal density from (1.3). The parameter p is the expectation

p x and the parameter u2 is the variance u$ of a random variable X with density (1.3). I t is a well-known f ~ c t (and easy to verify, for example with the computer) that for an N ( p , u2) random variable X ,

So there is a 95% chance that the normal random variable X assumes values in [p - 1.96~7, p + 1.96 u]. Analogously to (1.4) one can formulate a heuristic 2u-rule (it is nothing but this: one can construct counterexamples) that for a "nice" random variable A' the probability

I is close to 1. This rule is also supported by the Chebyshev inequality

I

which provides us with a sufficiently accurate bound of the probability that the absolute deviation of the random variable X from its expectation exceeds some threshold x.

Page 13: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

14 CHAPTER 1.

1.1.2 Random Vectors

In what follows we frequently make use of finite-dimensional and infinite-di- merlsional random structures. We commence with finite-dimensional random vectors as a first step toward the definition of a stochastic process.

X = (XI , . . . , X,,) is an n-dimensional random vector if its components X I , . . . , X,, are one-dimensional real-valued random variables.

If we interpret t = 1,. . . ,n as equidistant instants of time, ,Yt can stand for the outcome of an experiment a t time t . Such a time series may, for example, consist of BMW share prices Xt a t n succeeding days. Clearly, t is "mathematical time", thus nothing but an index or a counting variable. For example, a random vector can describe the state of the weather in Wellington (NZ) a t a given time: X1 could be the temperature, X2 the air pressure and X 3 the windspeed (the latter is usually close to infinity!).

Analogously to one-dimensional random variables one can introduce the distribution function, the expectation, moments and the covariance matrix of a random vector in order to describe its distribution and its dependence structure. The latter aspect is a new one; dependence does not q8ke sense when we talk just about one random variable.

Probability, Distribution and Distribution Function

Toss a fair coin twice. We consider the four pairs (H, H ) , (T, T ) , (H, T ) and (T, H ) (H=head, T=tail) as outcomes of the experiment "flipping a coin twice". These four pairs constitute the outcome space R. As before, we assign 1 t,o H and O to T. In this way we obtain two randorn variables X1 and Xz, and X = ( X I , X2) is a two-dimensional random vector. Notice that

If the coin is indeed fair, we can assign the probability 0.25 to each of the four outcomes, i.e.

P({w : X(w) = (k, 1))) = 0.25, k , 1 € {O, 1 ) .

As before, we consider a collection F of subsets of R and define a probability Irleasule on it, i.e. we assign a number P(A) E [O,1] to each .4 E F.

1.1. BASIC CONCEPTS FROM PROBABILITY THEORY

b2

a 1 bl Figure 1.1.6

The collection of the probabilities

FX (x) = P(X1 < x1 , . . . , Xrr 5 xn) (1.5)

= P({w : >YI(w) < X l , . . . , X,,(w) 5 x n ) ) ,

X = ( x l , . . . , x n ) E R n ,

is the distribution function Fx of X.

It provides us with the probability of the event that X assumes values in the rectangle

(a, b] = {x : ai < xi 5 bi , i = 1, . . . , n} .

Fol- example, if X is two-dimensional,

P ( X E (a, b]) = Fx(bl,bz) + Fx.(al ;az) - Fx(a1, b2) - Fx(b1,az).

Check that this formula is correct; see also Figure 1.1.6. As in the case of one-dimensional random variables, these probabilities approximate P ( X E B) for very general sets B.

Page 14: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

16 CHAPTER 1.

fi ( B ) = P(X E B) = P({w : X ( w ) E B ) )

"Suitable" subsets of Rn are the Bore1 sets which are obtained by a countable number of ope~ations U, n or acting on the intervals in Rn; see p. 64 for a precise definition. For example, point sets, balls and rcctanglcs arc Borcl scts. In a mathematical sense, the distribution and the distribution function of a random vector X are equivalent notions. Both, Fx and Px, can be used to calculate the probability of any event {X E B).

Notice that the distribution of X = (XI , . . . , X,) contains the whole in- formation about the distribution of the components Xi , pairs (Xi, Xj) , triples (Xi ,Xj , X k ) , etc. This is easily seen from (1.5): you get the distribution func- tion of X1 by formally setting xa = . . . = xn = cm, the distribution function of (XI , X2) by setting x3 = . . . = xn = cm, etc.

Ar~alogously to random variables one can introduce discrete and continu- ous random vectors and distributions. For our purposes, continuous random vectors with a density will be relevant, and so we mostly restrict our attention to them.

If the distribution of a random vector X has density fx , one can represent the distribution function Fx of X as

I where the density is a function satisfying I f x (x) 2 0 for every x E Rn

and

~ x ( Y I , . . . , Y , , ) ~ Y I . . . dun = 1

If a vector X has density fx , all its components X,, the vectors of the pairs (Xi, X j ) , triples (Xi, X j , Xk) , etc., have a density. They are called marginal densities.

1.1. B.4SIC CONCEPTS FROM PROBABILITI' THEORY 17

E x a m p l e 1.1.7 (Marginal densities: the case n = 3) We consider the case n = 3. Then the marginal densities are obtained as follows:

f x ( x ) dxadx3 , f . x , , . ~ ~ (x i , x2) =

f . ~ , (x2) is obtained by integrating f x ( x ) with respect to X I and 23, fx,,x, by integrating f x ( x ) with respect to 22, etc.

One case is particularly simple: if the density f x ( x ) can be written as a product of non-negative functions gi:

In this case, Srrn gi(xi) dxi = 1 for i = 1 , . . . , n , i.e. the functions gi(xi) are one-dimensional probability densities, and then necessarily

fx, (xi) = gi(xi) ~x ; ,x , (xi, z j ) = gi(xi) g j (x j ) , etc.

Verify this!

E x a m p l e 1.1.8 (Gaussian random vector) A Gaussian or normal random vector has a Gaussian or normal distribution. The n-dimensional normal or Gaussian distrih~rtion is given by its density

1 with parameters p E Rn and C. (Here and in what follows, y' denotes the 1 I

transpose of the vector y; see p. 211.) The quantity C is a symmetric positive-

i definite n x n matrix, C-I is its inverse and det C its determinant. See Fig-

l ure 1.1.10 for an illustration in the case n = 2.

Expec ta t ion , Var iance a n d Covar iance I

The expectation of a random vector has a similar function as the mean value of a random variable. The values X(w) are concentrated around it.

The expectation or mean value of a random vector X is given by

p x = E X = ( E X l , . . . ,EX, , ) .

Page 15: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

18 CHAPTER 1.

The cova~~iance matr ix of X is defined as

where

cov(Xi, X j ) = E[(Xi - p x i ) ( X j - p,yj ) I = E(XiXj ) - p x i PX, 1

is the covariance of Xi and ,Xi. Notice that cov(Xi, Xi) = u i i - -

Example 1.1.9 (Continuation of Example 1.1.8) Recall from (1.6) the density of a multivariate Gaussian rar~dorr~ vector X. The parameter p is the expert,a.t,ion /LX of X, and C is its covariance matrix Ex . Thus the density of a Gaussian vector (hence its dist,ribution) is completely determined via its expectation and covariance matrix. In particular, if p = 0 and C is the n-dimensional identity matrix I,, we have det I,, = 1 and C-' = I,. The density f x is then simply the product of n standard normal densities:

We write N ( p , C) for the distribution of an n-dimensional Gaussian vector X with expectation p and covariance matrix C. Such a vector has the appealing property that it remains Gaussian under linear transformations (recall that p' and A' denotc thc transposes of p and .4, respectively):

Let X = ( X I , . . . , X,,) have an N ( p , C) distribution and A be an 7n x 71 matrix. Then AX' has an N(A p', A C A') distribution.

It is convenient to standardize covariances by dividing the corresponding ran- dom variables by their standard deviations. The resulting quantity

- - E[(xi - p x , )(JLIa - PX, )I

is the correlation of XI and X2. As a result of this standardization, the corre- lation of two random variables is always between -1 and +l. Check this fact by an application of the Cauchy-Schwarz inequality; see p. 188.

1.1. BASIC CONCEPTS FR,OAi PR,OBABlLITY THEORY

Figure 1.1.10 Densi ty o f the 2-dimensional standard normal densi ty ( p = 0 and C = I2 is the identi ty m a t r i x ) .

1.1.3 Independence and Dependence

Flip a fair coin twice and let the random numbers X L (w), X2(w) E {O,l) be the corresponding outcomes of the first and second experiment. It is easy to verify that

This property is called independence of the random variables XI and X g . In- tuitively, independence means that the first experiment does not influence the second one, and vice versa. For example, knowledge of X1 does not allow one to predict the value of X 2 , and vice versa.

Below we recall some of the essential definitions and properties of inde- I pendent events and independent random variables.

Two events .41 and A2 are independent if

P ( A l n A,) = P(A1) P (A2) .

Page 16: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

CHAPTER 1. 1.1. BASIC CONCEPTS FR0A.I PROBABILITY THEORY 2 1

Two random variables X1 and X2 are independent if

P ( X 1 E B1, X 2 E B2) = P ( X l E B1) P ( X 2 E B2)

for all suitable subsets B1 and B2 of EX. This means that the events {XI E B1) and {X2 E B2) are independent.

Alternatively, one can define independence via distribution functions and den- sities. The random variables X1 and X2 are independent if and only if

Assume that (XI, Xz) has density fx,,x, with marginal densities fxl and fx,: see p. 16. Then the random variables X1 and X 2 are independent if and only if

fx1,xz(" l ,x2) = fx1(x1) fxz (x2) , Xl, x2 E R .

The definition of independence can be extended to an arbitrary finite num- ber of events and random vectors. Notice that independence of the components of a random vector implies the independence of each pair of its components, but the converse is in general not true.

The events A1,. . . ,A, are independent if, for every choice of indices 1 I i l < . . . < i k 5 n and integers 1 5 k 5 n ,

P(Ail n . . . n Ai,) = P ( A i l ) - . . P(Aik) . 1 The random variables X I , . . . , X , are independent if, for every choice of indices 1 5 il < . . . < ik 5 n, integers 1 5 k 5 n and all suitable subsets B1, . . . , B, of R,

P(Xi l E Bil , . . . , X,, E B,,) = P ( X z l E Bi l ) ... P(Xik E B i k ) .

This means that the events {XI E B1), . . . , {X,, E B,,) are independent.

The random variables X I , . . . , X, are independent if and only if their joint distribution function can be written as follows:

If the random vector X = ( X I , . . . ,X,,) has density fx, then X I , . . . , S, are independent if and only if

Example 1.1.1 1 (Continuation of Example 1.1.9) Recall from (1.6) the density of an n-dimensional Gaussian vector X. It is easily seen from the form of the density that its components are independent if and only if the covariance matrix C is diagonal. This mea.ns t,ha.t, cnrr(Xi, A>) = cov(Xi, X,) = O for i # j , and so we can write the density of X in the form (1.7). Thus, i n the Gaussian case, uncorrelatedness and independence are equivalent notions. This statement is wrong for non-Gaussian random vectors; sec Examplc 1.1.12.

An important consequence of the independence of random variables is the following property:

If X I , . . . , X, are independent, then for any real-valued functions Sl , . . . , !?n ,

E[s~ (Xi ) . . . g,(Xn)I = E s ~ (XI ) . . . Eg,,(Xn) ,

provided the considered expectations are well defined.

In particular, we may conclude that the independent random variables X1 and X 2 are uncorrelated, i.e. corr(X1, X2) = cov(X1, .Yz) = 0. The converse is in general not true.

Exanlple 1.1.12 (U~lcurrelated random variables are not necessarily indepen- dent .) Let X be a standard nornlal random variable. Since X is symmetric (i.e. X and - X have the same distribution), so is X3, and therefore both X and X3 have expectation zero. Thus

but X and X 2 are clearly dependent: since { X E [-I, 11) = {X2 E [O,l]), we obtain

P ( X E [-I, 11, x2 E [0, 11) = P ( X E [-I, 11)

> P(X E [-I, l])P(X" [0, 11) = [ P ( X E [-I, 1])12.

Page 17: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

22 CHAPTER 1. 1.2. STOCH.4STIC PROCESSES 2 3

Example 1 .I . 13 (Autocorrelations of a time series) For a time series &, X1,X2, . . . the autocorrelation a t lag h is defined by COI.I.(X~, X h ) , h = 0,1, . . .. A claim which can frequently be found in the literature is that financial time series (derived from stock indices, share prices, exchange rates, etc.) are nearly uncorrelated. This is supported by the sample autocorrelations of the daily log-returns Xt of the S&P index; see Figure 1.1.14. In contrast to this observation, the estimated autocorrelations of the absolute values lXtl are different from zero even for large lags h. This indicates that there is dependence in this time series.

elementary level and Gut (1995) for an intermediate course. Also rrlany text- books on statistics often begin with an introduction to probability theory; see for example Mendenhall, Wackerly and Scheaffer (1990).

1.2 . Stochastic Processes

We suppose that the exchange rate NZ$/US$ a t every fixed instant t between 9 a.m. and 10 a.m. this morning is random. Therefore we can interpret it as a realization Xt(w) of the random variable Xi, and so we observe Xt(w), 9 < t 5 10. In order to make a guess a t 10 a.m. about the exchange rate XI (w) a t 11 a.m. it is reasonable to look a t the whole evolution of Xt(w) between 9 a.m. and 10 a.m. This is also a demand of the high standard technical devices which provide us with al~rlost c o ~ ~ t i ~ l u v u s information about the process considered. A mathematical model for describing such a phenomenon is called a stochastic process.

( X t , t E T ) = (Xt(w) , t E T , w E R ) ,

Figure 1.1.14 T l ~ e eslirnuled u~ulocorrelaliur~s oJ the S&P index (left) and of its absolute values (right); see Example 1.1.13; cf. the comments i n Figure 1.1.4.

In what follows, we will often deal with infinite collections (Xt, t E T ) of random variables X t , i.e. T is an infinite index set. In this set-up, we may also introduce independence:

choice of distinct indices t l , . . . , t , E T and n > 1 the random variables St , , . . . , St,, are independent. This collection is i ndependen t a n d iden - t ically distributed ( i i d ) if it is independent and all random variables X t have the same distribution.

Notes and Comments

In this section we recalled some elerrle~ltar-y probability theory which can be found in every textbook on the topic; see for instance Pitman (1993) for an

For our purposes, T is often an interval, for example T = [a, b ] , [a, b) or [a, co) for a < b. Then we call X a con t inuous - t ime process in contrast t o discrete- t i m e processes. In the latter case, T is a finite or countably infinite set. For obvious reasons, the index t of the random variable X t is frequently referred to as tirne, and we will follow this convention.

A stochastic process X is a function of two variables.

For a fixed instant of time t , it is a random variable:

For a fixed random outcome w E R , it is a function of time: I This function is called a realization, a trajectory or a samp le path of the process X .

These two aspects of a stochastic process are illustrated in Figure 1.2.1.

Page 18: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

24 CHAPTER 1

Figure 1.2.1 5 sample paths of a stochastic process ( X t , t E [0, I]). Top: every path corresponds to a different w E Q. Middle and bottom: the values on the verticnl 1in.e~ at t = 0 . 1 , . . . ,0 .9 visualize the random variables X ' o . 1 , . . . , X o . g ; they occur as the projections of the sample paths on the vertzcal lznes.

1.2. STOCHASTIC PROCESSES

Figure 1.2.2 The (scaled) daily values of the S&P index over a period of 7,422 days. The graph suggests that we consider the S&P t ime series as the sample path of a continuous-time process. If there are many values in a t ime series such that the instants of t ime t E T are "dense" i n an interval, then one may want t o interpret this discrete-time process as a continuous-time process. The sample paths of a real-life continuous-time process are always reported at discrete instants of t ime. Depending o n the situation, one has to make a decision which model (discrete- or continuous- t ime) is more appropriate.

E x a m p l e 1.2.3 A t ime series

is a discrete-time process with T = Z = (0, f 1, f 2, . . .). Time series constitute an important class of stochastic processes. They are relevant models in many applications, where one is interested in the evolution of a real-life process. Such series represent, for example, the daily body temperature of a patient in a hospital, the daily returns of a price or the monthly number of air traffic passengers in the US. The most popular theoret,ical t,ime series models are the ARMA (AutoRegressive Moving Average) processes. They are given by certain difference equativns in which an iid sequence (Z t ) (see p. 22), the so- called noise, is involved. For example, a moving average of order q > l is defined as

X t = Z t + O I Z t - l +. . .+ OqZtPq, t E Z ,

Page 19: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

and an autoregressive process of order 1 is given by

CHAPTER 1.

Here &,. . . ,8, and 4 are given real parameters. Time series models can be understood as discretizations of stochastic differential equations. We will see this for the autoregressive process on p. 141.

Figure 1.2.4 shows two examples.

I Figure 1.2.4 Two time scries X t , t = 1 , . . . ,100. Lcft: 100 successive daily log-

I returns of the S&P index; see Figure 1.1.4. Right: a simulated sample path of the I autoregressive process Xt = 0.5Xt-1 +Z t , where Zt are iid N ( 0 , l ) random variables;

see Example 1.2.3.

',

We see that the concepts of a random variable X and of a stochastic process (Xi, t e T ) are not so much different. Both have random realizations, but the realization X ( w ) of a random variable is a number, whereas the realization X t ( w ) , t E T , of a stochastic process is a function on T . So we are completely correct if we understand a stochastic process to be a "random element" tak- ing functions as values. Moreover, we can interpret a random variable and a rand0111 vector as special stochastic processes with a finite index set T .

~ Distribution

In analogy to random variables and random vectors we want to introduce non-random characteristics of a stochastic process such as its distribution, expectation, etc, and describe its dependence structure. This is a task much

1.2. STOCHASTIC PROCESSES 2 7

more complicated than the description of a random vector. Indeed, a non- trivial stochastic process X = (Xt , t e T ) with infinite index set T is an infinite-dimensional cbject; it can be understood as the infinite collection of the random variables X t , t E T. Since the \dues of X are functions on T , the dzstribution of X should be defined on subsets of a certain "function space", i.e.

P ( X E A ) , A E F , (1.8)

where .F is a collection of suitable subsets of this space of functions. This approach is possible, but requires advanced mathematics, and so we try to find some simpler means.

The key observation is that a stochastic process can be interpreted as a collection of random vectors.

The f inite-dimensional distributions (f idis) of the stochastic process X are the distributions of the finite-dimensional vectors

t t , t l , . . . , t , E T ,

for all possible choices of times tl , . . . , t , E T and every n 2 1.

We can imagine the fidis much easier than the complicated distribution (1.8) of a stochastic process. I t can be shown that the fidis determine the distribution of X . In this sense, we refer to the collection of the fidis as the distrabution of the stochastic process.

Stochastic proccsscs can be classified according to different criteria. One of them is the kind of fidis.

E x a m p l e 1.2.5 (Gaussian process) %call from (1.6) the definition of an n-dimensional Gaussian density. A stochastic process is called Gaussian if all its fidis are multivariate Gaussiar~. We learnt in Example 1.1.9 that the parameters p and C of a Gaussian vector are its expectation and covariance matrix, respectively. Hence the distribution of a Gaussian stochastic process is dctcrmincd only by the collection of the expectations and covariance matrices of the fidis.

A simple Gaussian process on T = [O, 11 consists of iid N ( 0 , l ) random vari- ables. In this case the fidis are characterized by the distribution functions

Page 20: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

28 CHAPTER 1

The sample paths of this process are very irregular. See Figure 1.2.6 for an illustration.

Figure 1.2.6 A sample path of the Gaussian process ( X t , t E [0, l]), where /he Xts are aid N(0, l ) ; see Example 1.2.5. The expectation function is ,ux(t) = 0 and the dashed lines indicate the curves f 2ox(t) = f 2; see Example 1.2.7.

Expectation and Covariance Functior~

For a random vector X = ( X I , . . . ,X,) we defined the expectation p x = (EX1,. . . , EX,,) and the covariance mat,rix Ex = (rov(Xi, Xj) , i, j = 1,. . . , n). A stochastic process X = ( X t , t E T ) can be considered as the collection of the random vectors ( X t l , . . . , X t n ) for t l , . . . , t, E T and n 2 1. Ebr each of them we can determine the expectation and covariance matrix. Alternatively, we can consider these quantities as functions of t E T:

The ezpectution firzctioiz of X is given by

p x ( t ) = p x c = E X t , t E T .

The couariance function of X is given by

1.2. STOCH.ASTIC PROCESSES 29

r The variance function of X is given by

I &(t) = cx (t , t ) = var(Xt) , t E T . I We learnt in Example 1.2.5 that Gaussian processes are determined only via their expectation and covariance functions. This is not correct for a non- Gaussian process.

As ful- a ~ a r l d o i i ~ vector, the expectation function p x (t) is a deterministic quantity around which the sample paths of X are concentrated. The covariance function cx ( t , s) is a measure of dependence in the process X . The variance function a$(t) can be considered as a measure of spread of the sample paths of X around p x ( t ) . Tn ront,rast t,o the one-dimensional case, a statement like "95% of all sample paths lie between the graphs of px(t) - 2ux( t ) and px( t ) + 2 ~ 7 ~ (t)" is very difficult to show (even for Gaussian processes), and is in general not correct. We will sometimes consider computer graphs with paths of certain stochastic processes and also indicate the curves p x ( t ) and p x (t) d~ 2ux (t), t E T. The latter have to be interpreted for every fixed t , i.e. fur every irlcliviclual random varisble X t . Only in a heuristic sense, do they give bounds for the paths of the process X. See Figure 1.2.6 for an illustration.

Example 1.2.7 (Continuatlion of Example 1.2.5) Consider the Gaussian process ( X t , t E [O, 11) uf iid N(O, l ) random variables Xt . Its expectation and covariance functions are given by

Dependence Structure

We have already introduced Gaussian processes by specifying their fidis as multivariate Gaussian. Another way of classifying stochastic processes consists of imposing a special dependence structure.

The process X = (Xi , t E T ) , T c R, is strictly stationmry if t,he fidis are invariant under shifts of the index t:

for all possible choices of indices t l , . . . ,t,, E T , n 2 1 and h such t,ha.t, t l + d

h , . . . , t , + h E T. Here = stands for the identity of the distributions; see p. 211 for the definition. For the random vectors in (1.9) this means that their distribution functions are identical.

Page 21: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Example 1.2.8 (Stationary Gaussian processes) Consider a process S = ( X t , t E T ) with T = [ O , c o ) or T = Z. A trivial example of a strictly stationary process is a sequence of iid random variables St , t E Z. Since a Gaussiarl process X is determined by its expectation and covariance functions, condition (1.9) reduces to

p , ~ (t + h ) = p . ~ (t) and c x (t, s ) = c x (t + h , s + h)

for all s , t E T such that s + h , t + h E T . But this rrleans that p x ( t ) = p x (0 ) for all t , whereas cx( t , s) = Ex (It - sl) for some function Esy of one variable. Hence, for a Gaussian process, strict stationarity means that the expectation function is constant and the covariance function only depends on the distance It -s/. More generally, if a (possibly non-Gaussian) process X has the two aforementioned properties, it is called a stationary ( in the wide sense) or (second-order) stationary process.

If we describe a real-life process by a (strictly or in the wide sense) stationary stochastic process, then we believe that the characteristic properties of this process do not change when time goes by. The dependence structure described by the fidis or the covariance function is invariant under shifts of time. This is a relatively strong restriction on the underlying process. However, it is a

-- -

standard assumption in Inany probability related fields such as statistics and t h e series analvsis.

Stationarity can also bc imposed on the increments of a process. The process itself is then not necessarily stationary.

Let X = (Xt, t E T) be a stochastic process and T c R be an interval

S is said to havc stutionary increrncnts if

X is said to have independent increments if for every choice of t ; E T with t l < . . . < t , and n 2 1,

Xt, - Xtl , . . . , Xl,, - Xt,-,

are independent random variables.

One of the prime examples of processes with independent, stationary incre- ments is the homogeneous Poisson process. Honiogeneity is here another word- ing for stationarity of the increments.

- 1.2. STOCHASTIC PROCESSES

0 .

I -

* t go.

n

0 -

Figure 1.2.9 Sample paths of a homogeneous Poisson process (Xt , t E [0, w)) with in tens i ty X = 1 ; see Example 1.2.10. T h e straight solid l ine s tands for the expectation funct ion p~ (t) = t .

Page 22: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

32 C H A P T E R 1

E x a m p l e 1.2.10 (Homogeneous Poisson process) A stochastic process (Xt , t E [O, 00)) is called an homogeneous Poisson pro- cess or simply a Poisson process with intensity or rate X > 0 if the following conditions are satisfied:

It starts a t zero: ;"=o = 0.

It has stationary, independent increments.

For every t > 0 , Xi has a Poisson Poi(Xt ) distribution; see Example 1.1.1 for the definition of the Poisson distribution.

Figure 1.2.9 shows several Poissonian sample paths.

Notice that , by stationarity of the increments, X t - X, with t > s has the same distribution as Xtp , - Xo = Xt-,, i.e. a Poi(X( t - s)) distribution.

An alternative definition of the Poisson process is given by

where #A denotes the number of elements of any particular set A, T,-= Yl +. . . + Yn and (Y,) is a sequence of iid exponential E x p ( X ) random variables with common distribution function

This definition shows nicely what kind of sample path a Poisson process has. I t is a pure jump function: it is constant on [T,, Tn+1) and has upward jumps of size 1 a t the random times T,.

The r6le of the Poisson process and its modifications and ramifications is com- parable with the role of Brownian motion. The Poisson process is a counting process; see (1.10). I t has a large variety of applications in the most different fields. To name a few: for a given time interval [0, t], X t is a model for the number of

tclcphonc calls to bc handled by an operator,

customers waiting for service in a queue,

claims arriving in an insurance portfolio.

1.3. BROWNIAN MOTION 33

Notes and Comments a Introductions to the theory of stochastic processes are based on non-elementary

facts from rveasure theory and functional analysis. Standard texts are Ash and Gardner (1975), Gilchman and Skorokhod (1975), Karlin and Taylor (1975,1981) and many others. An entertaining introduction to the theory of applied stochas- tic processes is Resnick (1992). Grimett and Stirzaker (1994) is an introduction "without burdening the reader with a great deal of measure theory".

1.3 Brownian Motion

1.3.1 Defining Properties

Brownian motion plays a central r6le in probability theory, the theory of stochastic processes, physics, finance,. . ., and also in this book. We start with the definition of this important process. Then we continue with some of its elementary properties.

A stochastic process B = (Bt, t E [0, co)) is called (standard) Brownian motion or a Wiener process if the following conditions are satisfied:

I I t starts a t zero: Bo = 0 . I I t has stationary, independent increment.^; see p. 30 for the defini- tion.

For every t > 0 , Bt has a normal N(O, t ) distribution.

I I t has continuous sample paths: "no jumps". I I-

I See Figure 1.3.1 for a visualization of Brownian sample paths. Brownian motion is named after the biologist Robert Brown whose research

dates to the 1820s. Early in this century, Louis Bachelier (1900), Albert Ein- stein (1905) and Norbert Wiener (1923) began developing the mathematical theory of Brownian motion. The construction of Bachelier (1900) was erro- neous but i t captured many of the essential properties of the process. Wier~er (1923) was the first t o put Brownian motion on a firm mathematical basis.

D i s t r ibu t ion , E x p e c t a t i o n a n d Covar iance Func t ions

The fidis of Brownian motion are multivariate Gaussian, hence B is a Gaus- sian process. Check this statement by observing that Brownian motion has

Page 23: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

1.3. BROWNIAN MOTION 35

independent Gaussian increments and by using the formula for linear transfor- mations of a Gaussian random vector; see p. 18.

The random variables Bt - B, and Bt-, have an N(0, t - s ) distribution for s < t.

This follows from the stationarity of the increments. Indeed, Bt - B, has the same distribution as Bt-, - Bo = Bt-, which is normal with mean zero and variance t - s. Thus the variance is proportional t o the length of the interval [s, t]. This means intuitively: the larger the interval, the larger the fluctuations of Brownian motion on this interval. This observation is also supported by simulated Brownian sample paths; see for example Figure 1.3.2.

Notice: d

The distributional identity Bt - B, = Bt-, does not imply pathwise identity: in general,

Bt(w) - Bs(w) # Bt-s(w).

I t is worthwhile to compare Brownian motion with the Poisson process; scc Example 1.2.10. Their definitions coincide insofar that they are processes with stationary, independent increments. The crucial difference is the kind of distribution of the increments. The requirement of the Poisson distribution makes the sample paths pure jump functions, whereas the Gaussian assumption makes the sample paths continuous.

I t is immediate from the definition that Brownian motion has expectation function

pg( t ) = E B t = 0 , t 2 0 ,

and, since the increments B, - Bo = B, and Bt - B, are independent for t > s , it has covariance function (recall its definition from p. 28)

~ g ( t 1 ~ ) = E[[(Bt - Bs) + B,] B,] = E[(Bt - B,) B,] + EB:

Figure 1.3.1 Sample paths of Brownian motion on [ O , l ] .

Since a Gaussian process is characterized by its expectation and covariance functions (see Example 1.2.5), we can give an alternative definition:

Brownian motion is a Gaussian process with

/rg(t) = O and cg(t , s) = min(s, t) . (1.11)

Page 24: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

36 CHAPTER 1 . 1.3. BROIVNIAN MOTION 3 7

P a t h P roper t i e s : Non-Differentiabil i ty and U n b o u n d e d Var ia t ion

In what follows, we fix one sample path B t ( w ) , t > 0, and consider its proper- ties. We already know from the definition of Brownian motion that its sample paths arc continuous. However, a glance a t simulated Brownian paths imme- diately convinces us that these functions of t are extremely irregular: they oscillate wildly. The main reason is that the increments of B are independent. In particular, increments of Brownian motion on adjacent intervals are inde- pendent whatever the length of the intervals. Since we can imagine the sample path as constructed from its independent increments on adjacent intervals, it is rather surprising that continuity of the path results.

Thus:

ZZow irregular is a Brownian sample path?

Before we answer this questiur~ we make a short excursion to a class of stochas- tic processes which contains Brownian motion as a special case. All members of this class have irregular sample paths.

A stochastic process ( X t , t E [0, m)) is H-self-similar for some H > 0 if its fidis satisfy the condition

d ( T H B t l , . . .. T H B t , ) = ( B T t l , . . . , BT~ , ) (1.12)

for every T > 0, any choice of ti > 0, i = I , . . . , n, and n 2 1.

Notice:

Self-similarity is a distributional, not a pathwise property. In (1.12), one must d

not replace = wzth =.

Ruugllly speakir~g, self-si~riilarity means that the properly scaled patterns of a sample path in any small or large time interval have a similar shape, but they are not identical. See Figure 1.3.2 for an illustration.

The sample paths of a self-similar process are nowhere differentiable; see Proposition A3.1 on p. 188. And here it comes:

Brownian motion is 0.5-self-similar, i.e.

(T'/'B,, , . . . , T ' / ~ B ~ , ) d ( B T l l , . . . . BTLn ) (1.13)

for cvcry T > 0, any choicc of ti > 0, i = 1 , . . . , n, and n > 1.

Hence its sample paths are nowhere differentiable.

Figure 1.3.2 Self-similarity: the snme Rrown.ian sample path on, different .scales. The shapes of the curves on different intervals loolc similar, but they are not simply scaled copies of each other.

Page 25: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

38 CHAPTER 1.

Figure 1.3.3 Left: a dzfferentiable function. A t every point its graph can be approx- imated by a linear function which i s the unique tangcnt at this point. Right: this function is not differentiable at z = 1. There are infinitely many tangents t o the curve of the function at this point.

One can easily check the distributional identity (1.13). Indeed, the left- and right-hand sides of (1.13) are Gaussian random vectors, and therefore it suffices to verify that they have the same expectation and covariance matrix. Check these properties hy ~ising (1.11).

Differentiability of a function f means that its graph is smooth. Indeed, if the limit

. f (xo + Ax) - f (xo) f f ( x 0 ) = lim AX+O Ax

exists and is finite for some xo E (0, t ) , say, then we may write for small Ax

f (20 + Ax) = f (20) + fl(xO)Ax + h(x0, Ax) A x ,

where h(x0, Ax) + 0 as Ax + 0. Hence, in a small neighborhood of xo, the function f is roughly linear (as a function of Ax). This explains its smoothness. Alternatively, differentiability of f a t xo implies that we have a unique tangent to the curve of the function f a t this p o i ~ ~ t ; see Figure 1.3.3 for an illustration. In this figure you can also see a function which is not differentiable a t one point.

Now try to imagine a nowhere differentiable function: the graph of this function changes its shape in the neighborhood of any point in a completely non-predictable way. You will admit that you cannot really imagine such a

1.3. BROWNIAN AfOTION 39

function: it is physically impossible. Nevertheless, Brownian motion is consid- ered as a very good approximation to many real-life phenomena. We will see in Section 1.3.3 that Brownian motion is a limit process of certain sum processes.

The self-similarity property of Brownian motion has a nice consequence for the si~rlulatioil of its sample paths. In order to simulate a path on [0, TI it suffices to simulate one path on [0, 11, then scale the time interval by the factor T and the sample path by the factor ~ ' 1 ~ . Then we are done.

In some books one can find the claim that the limit

lim var a t - t o

does not exist and therefore the sample paths of Brownian motion are non-dif- ferentiable. It is easy to check (do it!) that the limit (1.14) does not exist, but without further theory it would be wrong to conclude from this distributional result that the paths of the process are non-differentiable.

The existence of a nowhere differentiable continuous function was discov- ered in the 19th century. Such a function was constructed by Weierstrass. It was considered as a curiosity, far away from any practical application. Brown- ian motion is a process with nowhere differentiable sample paths. Currently it is considered as one of those processes which have a multitude of applications in very different fields. One of them is stochastic calculus; see Chapters 2 and 3.

A further indication of the irregularity of Brownian sample paths is given by the following fact:

Brownian sample paths do not have bounded variation on any finite interval [0, TI. This means that

71.

SUP C IBt,(w) - Bti-, (w)I = m, i=l

where the supremum (see p. 211 for its definition) is taken over all pos- sible partitions T : 0 = to < - . . < tn = T of [0, TI.

A proof of this fact is provided by Proposition A3.2 on p. 189. We mention a t this point that the unbounded variation and non-differentiability of Brownian sample paths are major reasons for the failure of classical integration methods, when applied to these paths, and for the introduction of stochastic calculus.

Page 26: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

40 CHAPTER 1.

1.3.2 Processes Derived from Brownian Motion

The purpose of this section is to get some feeling for the distributional and pathwise properties of Brownian motion. If you want to start with Chapter 2 on stocha.st,ic calculus as soon as possible, you can easily skip this section and return to it whenever you need a reference to a property or definition.

Various Gaussian and non-Gaussian stochastic processes of practical rel- evance can be derived from Brownian motion. Below we introduce some of those processes which will find further applications in the course of this book. As before, B = (Bt, t E [O,co)) denotes Brownian motion.

p.0 0:2 0.4 0:6 1 .O t

Figure 1.3.4 A sample path of the Brownzan bridge.

Example 1.3.5 (Brownian bridge) Consider the process

X t = B t - t B 1 , O I t i 1 .

Obviously,

X o = B o - O B 1 = O and X 1 = B 1 - l B 1 = O .

For this simple reason, the process X bears the name (standard) Brownian bridge or tied down Brownian motion. A glance a t the sarnple paths of this "bridge" (see Figure 1.3.4) may or may not convince you that this name is justified.

I..?. BROWNIAN MOTION 41

Using the formula for linear transformations of Gaussian random vectors (see p. 18), one can show that the fidis of X are Gaussian. Verify this! Hence X is a Gaussian process. You can easily calculate the expectation and covariance functions of the Brownian bridge:

p x ( t ) = O and c x ( t , s ) = m i n ( t , s ) - t s , s , t E [ O , l ] .

Since X is Gaussian, the Brownian bridge is characterized by these two func- tions.

The Brownian bridge appears as the limit process of the normalized empirical distribution function of a sample of iid uniform U(0, l ) random variables. This is a fundamental result from non-parametric statistics; it is the basis for nu- merous goodness-of-fit tests in statistics. See for example Shorack and Wellner (1986).

Figure 1.3.6 A sample path of Brownian motion with drift Xt = 20 Bt + 10 t on [0,100]. The dashed line stands for the drift function yx(t) = l o t . !

Example 1.3.7 (Brownian motion with drift) Consider the process

X t = p t + u B t , t .0,

for constants a > 0 and p E R. Clearly, it is a Gaussian process (why?) with expectation and covariance functions

p x ( t ) = p t and c x ( t , s ) = a % i n ( t , s ) , s , t 2 0 .

I

Page 27: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

42 CHAPTER I .

The expectation function p x ( t ) = p t (the deterministic "drift" of the pro- cess) essentially determines the characteristic shape of the sample paths; see Figure 1.3.6 for an illustration. Therefore X is called Broiunian motion with (linear) drif t .

With the fundamental discovery of Bachelier in 1900 that prices of risky assets (stock indices, exchange rates, share prices, etc.) can be well described by Brownian motion, a new area of applications of stochastic processes was born. However, Brownian motion, as a Gaussian proccss, may assume negative val- ues, which is not a very desirable property of a price. In their celebrated papers from 1973, Black, Scholes and Merton suggested another stochastic process as a model for speculative prices. In Section 4.1 we consider their approach to the pricing of European call options in more detail. I t is one of the promising and motivating examples for the use of stochastic calculus.

Example 1.3.8 (Geometric Brownian motion) The process suggested by Black, Scholes and Merton is given by

i.e. i t is the exponential of Brownian motion with drift; see Example 1.3.7. Clearly, X is not a Gaussian process (why?).

For the purpose of later use, we calculate the expectation and covariance func- tions of geometric Brownian motion. For readers, familiar with probability theory, you rrlay recall that for an N ( 0 , l ) random variable 2,

E ~ X Z = e ~ 2 / 2 , X E B . (1.15)

I t is easily derived as shown below:

Here we used the fact that ( 2 ~ ) - ' / % x ~ { - ( a - X)2/2) is the density of an N (A, 1) random variable.

From (1.15) and the self-similarity of Brownian motion it follows immediately that

p X ( t) = e f i t ~ e u B i = ef i tEeu t112~1 - - e(~+0.5 u2) t . (1.16)

1.3. BROWNIAN MOTION 43 I

Figure 1.3.9 Sample paths of geometric Brownian motion X t = exp{O.Olt+O.OIBt} I on [0, 101, the expectation function p x ( t ) (dashed line) and the graphs of the functions px ( t ) f 2ax ( t ) (solid lines). The latter curves have to be interpreted with care sznce the distributions of the X t s are not normal.

d For s < t, Bt - B, and B, are independent, and Bt - B, = Bt-,. Hcncc

In particular, geometric Brownian motion has variance function I See Figure 1.3.9 for an illustration of various sample paths of geometric Brow- nian motion.

Page 28: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

CHAPTER 1.

E x a m p l e 1.3.10 (Gaussian white and colored noise) In statistics and time series analysis one often uses the name "white noise" for a sequence of iid or uncorrelated random variables. This is in contrast to physics, where white noise is understood as a certain derivative of Brownian sample paths. This does not contradict our previous remarks since this derivative is not obtained by ordinary differentiation. Since white noise is "physically impossible", one considers an approximation to it, called colored noise. I t is a Gaussian process defined as

where h > 0 is some fixed constant. Its expectation and covariance functions are give11 by

p x ( t ) = O and c x ( t , s ) = h - 2 [ ( s + h ) - m i n ( s + h , t ) ] , s s t

Notice that c x ( t , s ) = 0 if t - s 2 h, hence Xt and X , are independent, but if t - s < h, cx( t , s ) = hK2[h - (t - s)]. Since X is Gaussian and cx(t , s) is a function only of t - s, i t is stationary (see Example 1.2.8).

Clearly, if B was differentiable, we could let h in (1.19) go to zero, and in the limit we would obtain the ordinary derivative of B a t t . But, as we know,

lves an this argument is not applicable. The variance function a$ (t) = h-' g' indication t,hat the fllict,liat,ions of colored noise become larger as h decreases. Simulated paths of colored noise look very much like the sample paths in Figure 1.2.6.

1.3.3 Simulation of Brownian Sample Paths

This section is not necessary for the understanding of stochastic calculus. How- ever, it will characterize Brownian motion as a distributional limit of partial sum processes (so-called functional central limit theorem). This observation will help you to understand the Brownian path properties (non-differentiability, unbounded variation) much better. A second objective of this section is to show that Brownian sample paths can easily be simulated by using standard software.

Using the almost unlimited power of modern computers, you can visualize the paths of almost every stochastic process. This is desirable because we like to see sample paths in order to understand the stochastic process better. On the other hand, simulations of the paths of stochastic processes are sometimes unavoidable if you want to say something about the distributional properties of such a process. In most cases, we cannot determine the exact distribution

1.3. BROWNIAN MOTION

of a stochastic process and its functionals (such as its maximum or minimum on a given interval). Then simulations and numerical techniques offer some alternative to calculate these distributions.

S imula t ion v ia t h e Funct ional C e n t r a l L imi t T h e o r e m

From an elementary course in probability theory we know about the central limit theorem (CLT). I t is a fundamental result: it explains why the normal distribution plays such an important r61e in probability theory and statistics. The CLT says that the properly normalized and centered partial sums of an iid finite variance sequence converge in distribution to a normal distribution. To be precise: let Yl , Y2, . . . be iid non-degenerate (i.e. non-constant) random variables with mean p y = EYl and variance 0; = var(Yl). Define the partial sums

& = O , R n = Y l + . . . + Y n , n L 1 .

Recall that @ denotes the distribution function of a standard normal random variable.

If Yl has finite variance, then the sequence (Y,) obeys the CLT, i.e.

[var(Rn)I1I2 -

Thus, for large 12, the distribution of (R, - p y IZ)/(U$II) ' /~ is approxin~ately standard normal. This is an amazing fact, since the CLT holds independently of the distribution of the Y,s; all one needs is a finite variance a$.

T l ~ e CLT has an a ~ ~ a l o g u e for stocl~aslic processes. Consider the process with continuous sample paths on [0, 11:

( u ; ~ ) - ' / ~ ( R ~ - py i) , if t = i / n , i = 0 , . . . , n , s n ( t ) = (1.20)

linearly interpolated, elsewhere.

In Figures 1.3.11 and 1.3.12 you will find realizations of S, for various n. Assume for the moment that the k;s are iid N ( 0 , l ) and consider the

restriction of the process Sn to the points i l n . We immediately see that the following properties hold: . S, starts a t ?zero: Sn(0) = 0.

Page 29: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Figure 1.3.11 Sample paths of the process S, for one sequence of realizations Y l ( w ) , . . . , Y ~ ( w ) and n = 2 , . . . ,9.

Figure 1.3.12 Sample paths of the process S, for dzfferent n and the same sequence of realizations Y l ( w ) , . . . , Y~oo,ooo (a).

1.3. BROWNI14N MOTION 47

S, has independent increments, i.e. for all integers 0 < i l I . . . 5 im 5 n, the random variables

are independent.

For every O 5 i I n, S,(i/n) has a normal N(O,i/n) distribution.

Thus, S, and Brownian motion B on [O,l], when restricted to the points .i/n, have very much the same properties; cf. the definition of Brownian motion on p. 33. Naturally, the third property above is not valid if we drop the assumption that the Y,s are iid Gaussian. However, in an asymptotic sense the stochastic process S, is close to Brownian motion:

If Yl has finite variance, then the sequence (Y,) obeys the functional CLT, also called Donsker's invariance principle, i.e. the processes S, converge in distribution to Brownian motion B on [0, 11.

Convergence in distribution of S, has a two-fold meaning. The first one is quite intuitive:

The fidis of S, converge to the corresponding fidis of B , i.e.

P (Sn(t1) I 21,. . . , Sn(tm) I xm) + f.' (Btl < 21,. . . r Bt, I xm) (1.21)

for all possible choices of ti E [0, 11, xi E B, i = 1, . . . , m, and all integers m 2 1.

But convergence of the fidis is not sufficient for the convergence in distribu- tion of stochastic processes. Fidi convergence determines the Gaussian limit distribution for every choice of finitely many fixed instants of time ti, but stochastic processes are infinite-dimensional objects, and therefore unexpected events may happen. For example, the sample paths of the converging processes may fluctuate vcry wildly with increasing n, in contrast t o the limiting process of Brownian motion which has continuous sample paths. In order to avoid such irregular behavior,

a so-called tightness or stochastic compactness condition must be satis- fied.

Fortunately, the partial sum processes S, are tight, but it is beyond the scope of this book to show it.

Page 30: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

48 CHAPTER 1.

- Figure 1.3.13 One sclmple p t h of the processes SI) (dottcd line) and S 9 (solid line) for the same sequence of realizations Y l ( w ) , . . . ,Yg(w). See (1.20) and (1.22) for the definitions of Sn and S,.

What was said about the convergence of the processes S, remains valid for the processes

% ( t ) = - p y [nt]), o 5 t 5 1 , (1.22)

where [nt] denotes the integer part of the real number - n t ; see Figure 1.3.13 for

an illustration. In contrast t o S,, the process S, is constan: on the intervals [(i - l ) / n , i ln) and has jumps a t the points iln. But S , and S , coincide a t the points iln, and the differences between these two processes are asymptotically negligible: the normalization n1/2 makes the jumps of & arbitrarily small for large n. As for Sn, we can formulate the following functional CLT:

If Yl has finite variance, then the sequence (Y,) obeys the functional CLT, i.e. the processes ?, converge in distribution to Brownian motion B on [0, 11.

- Since S, is a jump process, the notion of convergence in distribution becomes even more complicated than for S,. We refrain from discussing details.

1.3. BROWNIAN MOTION 49

Thus we have a first simple tool for simulating Brownian sample paths in our hands:

Plot the paths of the processes S,, or S",, for suficiently large n, and you get a reasonable approximation to Brownian sample paths.

Also notice:

Since Brownian motion appears as a distributional limit you will see completely different graphs for different values of n, for the same sequence of realizations K ( w ) . From the self-similarity property we also know how to obtain an approximation to the sample paths of Brownian motion on any interval [0, TI:

~ -

Simulate one path of S,, or 5, o n [0, 11, then scale the t ime interval by the factor T and the sample path by the factor T I / ' .

Standard software (such as Splus, Mathematics, Matlah, etc.) provides you with quick and reliable algorithms for generating random numbers of standard distributions. Random number generators are frequently based on natural processes, for example radiation, or on algebraic methods. The generated "random" numbers can be considered as "pseudo" realizations Y,(w) of iid random variables Yi.

For practical purposes, you may want to choose the realizations Y , ( w ) (or, as you like, the random numbers Y,(w)) from an appropriate distribution. If you are interested in "good" approximations to the Gaussian distribution of Brownian n~otion, you would generate the k;(w)s from a Gaussian distribution. - If you are for-ced to simulate many sample paths of S,, or S,, in a short period of time, you would perhaps choose the Y,(w)s as realizations of iid Bernoulli random variables, i.e. P(Y, = f 1) = 0.5, or of iid uniform random variables.

Simulation via Series Representations

Recall from a course on calculus that every continuous 2~-periodic function f on R (i.e. f (x + 2x) = f ( x ) for d l x E R) has a Fourier series representation, i.e. it can be written as an infinite series of trigonometric functions.

Since Brownian sample paths arc continuous functions, we can try to ex- pand them in a Fourier series. However, the paths are random functions: for different w we obtain different functions. This means that the coefficients of this Fourier series are random variables, and since the process is Gaussian, they must be Gaussian as well.

The followirlg representation of Brownian motion on the interval [0, 2 ~ ] is called Paley-Wiener representation: let (Z,,n 2 O) be a sequence of iid

Page 31: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

50 CHAPTER, I .

i I -i I

Figure 1.3.14 Simulation of one Brownian sample path from the discretization (1.24) of the Paley-Wiener representation with N = 1,000. Top left: all paths for M = 2 , . . . ,40. Top right: the path only for M = 40. Bottom left: M = 100. Bottom right: M = 800.

1.3. BROWNIAN MOTION

N(0 , l ) random variables, then

t 2 O0 sin(nt/2) , t E [0, 27r] . (1.23)

I n= 1

This series converges for every fixed t, but also uniformly for t E [O, 2 ~ 1 , i.e. the

1 rate of convergence is comparable for all t For an application of this formula, 1 one has to decide about the number M of sine functions and the number N

of discretization points at which the sine functio~ls will be evaluated. This amounts to calculating the values

The problem of choosing the "right" values for M and N is similar to the choice of the sample size n in the functional CLT; it is diffic~llt, to give a simple rule of thumb for the choices of M and N.

In Figure 1.3.14 you can see that the shape of the sample paths does not change very much if one switches from M = 100 to M = 800 sine functions. A visual inspection in the case M = 100 gives the impression that the sample path is still too smooth. This is not completely surprising since a sum of M sine functions is differentiable; only in the limit (as M + oo) do we gair~ a non-differentiable sample path.

The Paley-Wiener representation is just one of infinitely many possible series representations of Brownian motion. Another well-known such repre- sentation is due to L6vy. In the LCvg rep~esen~tation, the sine functions are replaced by certain polygonal functions (the Schauder functions).

To be precise, first define the Haar functions Hn on [O,1] as follows:

HZm+l (t) = 1 -2mlZ, if t E [ I - g . 1 1 9

( 0 , elsewhere,

Page 32: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

CHAPTER 1

( 0 , elsewhere,

From these fi~nct~ions define the system of the Schauder functions on [O,1] by integrating the Haar functions: I

t Bn(t) = 1 H,(s) d s , n = 1 , 2 , . . .

Figures 1.3.15 and 1.3.16 show the graphs of Hn and H, for the first n. A series representation for a Brownian sample path or1 [O,l] is then given by

1 00

= x .G(w)fin(t) , t E [O, 11, (1.25) n=l

where the convergence of this series is uniform for t E [O,l] and the Zn(w)s are realizations of an iid N(0 , l ) sequence (2,). As for simulations of Brow- nian motion via sine functions, one has to choose a truncation point M of the infinite series (1.25). In Figure 1.3.17 we show how a Brownian sample path is approximated by the superposition of the first M terms in the series representation (1.25). In contrast to Figure 1.3.14, the polygonal shape of the Schauder functions already anticipates the irregular behavior of a Brownian path (its non-differentiability) for relatively small M.

The Paley-Wiener and Levy representations are just two of infinitely many possible series representations of Brownian motion. They are special cases of the so-called Le'vy-Ciesielski representation. Ciesielski showed that Brownian motion on [O,1] can be represented in the form

00 , t

where 2, are iid N(0 , l ) random variables and (4,) is a complete orthonormal function system on [O,l].

i Notes and Comments I Brownian motion is the best studied stochastic process. Various books are I devoted to it, for example Borodin and Salminen (1996), Hida (1980), Karatzas 1

Figure 1.3.15 The Haar functions H I , . . . , Hs

Page 33: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

54 CHAPTER 1. 1.3. BROWNZANMOTZON 55

- Figure 1.3.16 The Schauder functions 61,. . . , Hg.

Figure 1.3.17 The first steps i n the constructzon of one Brownian sample path from the Lc'vy representation (1.25) via hf Schauder functions, M = 1 , . . . ,8 .

Page 34: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

56 CHAPTER 1. I 1.4 . CONDITIONAL EXPECTATION

and Shreve (1988) and Revuz and \'or (1991). The reader of these books must be familiar with the theory of stochastic processes, functional analysis, special functions and measure theory. Every textbook on stochastic processes also contains a t least one chapter about Brownian motion; see the references on p. 33.

In addition to the non-differentiability and unbounded variation, Brownian motion has many more exciting path and distributional properties. Hida (1980) is a good reference to read about them.

Thc functional CLT is to bc found in advanced tcxtboolcs on stochastic pro-

! cesses and the convergence of probability measures; see for example Billingsley (1968) or Pollard (1984). The series representations of Brownian motion can be found in Hida (1980); see also Ciesielski (1965).

1.4 Conditional Expectation

You cannot avoid this section; it contains material which is essential for the understanding of martingales, and more generally, It6 stochastic integrals.

If you are not interested in details you may try to read from one box to another. .4t the end of Section 1.4 you should know:

the u-field generated by a random variable, a random vector or a stochas- tic process; see Section 1.4.2,

the conditional expectation of a random variable given a u-field; see Section 1.4.3,

the most common rules for calculating conditional expectations; sek Sec- tion 1.4.4.

You should start with Section 1.4.1, whcrc an cxamplc of a conditional cxpcc- tation is given. It will give you some motivation for the abstract notion of conditional expectation given a u-field, and every time when you get lost in this section, you should return t,o Section 1.4.1 and try t,o figure out what the geueral theory says in this concrete case.

1.4.1 Conditional Expectation under Discrete Condition

From an elementary course on probability theory we know the conditional probabzlzty of A given B, i.e.

Figure 1.4.1 The classical conditional probability: zf we know that B occurred, we assign the new probability 1 t o it. Events outside B cannot occur, hence they have the new probability 0 . I

Clearly,

P ( A I B) = P(A) if and only if A and B are independent. I I I

For the definition of P(AI B) it is crucial that P ( B ) is positive. It is the objective of Section 1.4.3 to relax this condition.

The probability P ( A I B ) can be interpreted as follows. Assume the event B occurred. This is additional information which substantially changes the underlying probability measure. In particular, we assign the new probabilities 0 to Bc (we know that Bc will not happen) and 1 to B. The event B becomes our new probability space a', say. All events of interest are now. subsets of R': A n B c R'. In order to get a new probability measure on R' we have to normalize the old probabilities P ( A n B ) by k'(B). In sum, the occurrence of B makes our original space R shrink to 61, and thk original probabilities P(A) have to be replaced with P ( A I B). . . .

Given that P ( B ) > 0, we can define the conditional distribution function of a random variable X given B

Page 35: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

58 CHAPTER 1 . I and also the conditional expectation of X given B I

where 1 if W E B ,

IB(w) = 0 if w e B ,

dcnotcs the indicator function of the event B . In order to discuss the definition (1.26), we assume for the moment that R = R. If X is a discrete random variable with values x l , 2 2 , . . ., then (1.26) becomes

00 P({W : X ( W ) = x k ) n B ) E(X ( B ) = xzt

k=l P ( B )

If X has density f x , then (1.26) beco~r~es

It is common usage to write JB g(x)dx for J-2 g ( x ) I ~ ( x ) d x .

Example 1.4.2 (The conditional expectation of a uniform random variable) We consider the random variable X ( w ) = w on the space R = (0 ,1] , endowed with the probability measure P such that

P((a , b]) = b - a , (a, b] C (O, l ] .

Clearly, X has a uniform distribution on (0,1]: its distribution function is given by

Fx(x ) = P({w : X(w) = w 5 x ) )

P(0) = O i f x < O ,

P ( (0 , X I ) = x if x E (O,1] , P((O, l ] ) = 1 if x > 1 .

1.4. CONDlTIONAL EXPECTATION

Figure 1.4.3 Left: a uniform random variable X on (O,1] (dotted line) and its ex- pectation (solid line); see Example 1.4.2. Right: the random variable X (dotted line) and the conditional ezpectations E ( X I A i ) (solid lines), where A; = ( ( i - 1 ) / 5 , i / 5 ] , i = 1, . . . , 5 . These wnditional expectations can be interpreted as the values of a discrete random variable E ( X 1 Y ) with distinct constant values on the sets A,; see Example 1.4.4.

Both, the random variable X and its expectation EX = 0.5, are represented in Figure 1.4.3.

Now assume that one of the events

Ai = ((i - l)/n,i/n] , i = 1,. . . , n ,

occurred. Recall that f x ( x ) = 1 on (O,1] and notice that P(A i ) = l / n . Then

The conditional expectations E(X I Ai) are illustrated in Figure 1.4.3. The value E(X I Ai) is the updated expectation on the new space Ai, given the information that Ai occurred.

distinct values yi on the sets Ai, i.e.

Ai = { w : Y ( w ) = y i ) , i = 1,2 , . . . .

Page 36: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

60 CHAPTER 1. / 1.4. CONDITIOIV.~L E-WECT.4TION

Clearly, ( A i ) is a disjoint partition of 0, i.e.

05

Ai n A j = 0 for i # j and U Ai = 0. (1.28) Here we also used (1.28) so that i= 1

We also assume for convenience that P ( A i ) > 0 for all i .

For a random variable X on R with ElXl < m we define the conditional expectation of X given Y as the discrete random variablc

E ( X I Y ) ( w ) = E ( X I A i ) = E ( X I Y = y i ) for w E A i ,

i = l , 2 , ... . (1.29)

If we know that Ai occurred, we may restrict ourselves to ws in Ai. For those ws, E ( X I Y ) (w) coincides with the classical conditional expectation E ( X I A i ) ; see (1.26).

Example 1.4.4 (C~ntinuat~ion of Example 1.4.2) We interpret the E ( X I Ai)s in (1.27) as the values of a discrete random variable E ( X I Y ) , where Y is constant on the sets Ai = ( ( i - l ) / n , i l n ] . See Figure 1.4.3 for an illustration of this random variable. In this sense, E ( X I Y ) is nothing but a coarser version of the original random variable X , i.e. an approximation to X , given the information that any of the Ais occurred.

In what follows we give some elementary properties of the random variable E ( X I Y ) . The first property can be easily checked. (Do it!)

constants cl , c2,

E ( [ C I X I + C Z X Z ] I Y ) = c I E ( X I I Y ) + c z E ( X 2 I Y ) .

The expectations of X and E ( X I Y ) are the same: EX = E [ E ( X I k')].

This follows by a direct application of the defining properties (1.29), (1.26) and by the observation that E ( X I Y) is a discrete random variable:

If S and 1' are independent, then E ( X I Y ) = E X . (1.30)

Recall from p. 20 that independence of X and I' implies

P ( X E A , Y = y i ) = P ( X E A ) P ( Y = y i ) = P ( X E A ) P ( A i ) . (1.31)

Consider the random variable IAi and not,ice that,

{ w : I A i ( w ) = 1 ) = A i = { w : Y(w) = y i ) .

Thus we can rewrite (1.31) as follows

P ( S E -4, IA, = 1) - P ( X E A) P( IA , - 1 ) .

The analogous relation, where { I A , = 1 ) is replaced with { I A , = O), also holds. Hence t,he random variables X and IA, are independent and for w E .4,,

1 where we used that

E I A , = 0 P ( A t ) + 1 P ( A i ) = P ( A i )

This proves (1.30).

Up to this point, we have learnt: I

The conditional expectation E ( X I 1') of X given a discrete random variable Y is a discrete random variable.

It coincides with the classical conditional expectation E ( X 12' = y i ) or1 the sets Ai = { w ; Z'(w) = yi).

In this sense, it is a coarser version of X; see again Figure 1.4.3.

Page 37: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

The fewer values Y has, the coarser the random variable E ( X I Y) . In particular, if I' = c o ~ ~ s t , thcn E ( S 11') = E X ; if I' assulncs two distinct values, so does E(X I Y ) , etc.

The conditional expectation E ( X I Y ) is not a function of X , hut rrlerely a function of Y. The random variable X only determines the kind of function. Indeed. we can write

DC)

E ( X 11') = , where g(y) = E ( X I Y = yi) I{,,) (y) . z=1

1.4.2 About a-Fields

In the previous section we introduced the conditional expectation E(X I Y) of a random variable X under the discrete condition (i.e. discrete random variable) Y. Recall from (1.29) that the values of 1' did not really matter for the definition of E ( X I Y), but it was crucial that Y assumed the distinct values y, on the w-sets Ai. Thus the conditional expectation E(X 11') can actually he understood as a random variable constructed from a collection a(1'), say, of subsets of R. So we may, in a syrnbolic way, write

E ( X I k') = E ( X I a ( Y ) )

Obviously, the collection a ( Y ) provides us with the information about the struct,ur.e of the randorr~ variable I'(w) as a function uf w E 0 .

In what follows, we want to make precise what "collection ~ ( 1 , ~ ) of subsets of 0 " means. We will call it a u-field or a u-algebra. Its definition follows:

-4 u-field 3 (on R) is a collection of subsets of R satisfying the following conditions:

It is not empty: 0 E 3 and R E 3 .

If A E 3, then A' E 3 .

/ 1 . 4 CONDITIONAL EXPECT4TION

11 If ill, Az, . . . E 3 , then

DC) 00

U A i e 3 and n A i ~ 3 . z = 1 2= 1

Example 1.4.5 (Some elementary a-fields) Check that the following collections of subsets of R are a-fields:

FZ = {0, 0, A, AC) for some A # 0 and A # R ,

Fl is the smallest a-field on R, and F3 , the power set of R, is the biggest one, as it contains all possible subsets of R.

Now suppose that C is a collection of subsets of R, but not necessarily a a-field. By adding more sets to C, one can always obtain a a-field, for example the power set P (R) . However, there are mathematical reasons showing that P ( R ) is in general too big. But one can also prove that ,

for a given collection C of subsets of 0 , there exists a smallest a-field a(C) on R containing C.

1 We call a(C) the a-field generated by C.

Example 1.4.6 (Generating a-fields from collections of subsets of 0) Recall the u-fields from Example 1.4.5. Prove that Fi = a(Ci), where

C1 = (01, C2 = {A) , C3 = F3

Using the definition of a u-field, you have to check which sets necessarily belong to the u-field a(Ci).

Now consider C4 = {A, B) and Cg = {A, B , C) ,

where A, B , C c R and determine a(C4) and a(C5). Before you start: first think about the structure of the elements in u(C4) and a(C5). Notice that you can get every element of a(C4) by taking all possible unions of the sets

I For u(C5) you can proceed in a similar way.

Page 38: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

64 CHAPTER I .

In general, it is difficult, if not impossible, to give a constructive description of the elements in u(C). The u-field u(l') generated by a discrete random variable 1' is ari exception.

E x a m p l e 1.4.7 (The a-field generated by a discrete random variablc) ~ e d a l l the set-up of Section 1.4.1. We considered a discrete random variable 1' with distinct values yi and defined the subsets Ai = {w : Y(w) = yi), which constitute a disjoint partition of 0. Choose

Since a(C) is a a-field it must contain all sets of the form

where I is any subset of N = {1,2,. . .), including I = 0 (giving A = 0) and I = N (giving A = 0). Verify that the sets (1.32) constitute a a-field a ( Y ) . Since the sets (1.32) necessarily belong to a(C), the smallest a-field containing C, we also have a (Y) = a(C). Later we will call a ( IT) the a-field generated by Y.

Notice that a (Y) contains all sets of the form

Indeed, the set I = {i : a < yi < b) is a subset of N, hence

E x a m p l e 1.4.8 (The Borel sets) Take R = R and

c(') = {(a, b] : -co < a < b < oo}

The a-field Bl = CT(C(~)) contains very general suhset ,~ of R. It is called the Borel a-field, its elements are the Borel sets. A normal human being cannot imagine the large variety of Borel sets. For example, it is a true fact,, but not easy to verify, that Bl is a genuine subset of the power set 'P(R). One can also int,roduce the a-field of the n-dimensional Borel sets B,, = u ( C ( ' ~ ) ) , where R = B7' and

c('" = {(a, b] : - m < ai < bi < oo, i = 1 , . . . , n ) .

I 1.4. CONDITIONAL EXPECTATION 6 5

The sets (a, b] are called rectangles. Any "reasonable" subset of R"" is a Borel set. For example, balls, spheres, smooth curves, surfaces, etc., are Borel sets, and so are the open and the closed sets. In order to show that a give11 subset C C Br is a Borel set, it is necessary to obtain C by a countable number of

1, oparatiurrs n, U,' actirrg ou the rectarrgles.

1/ For example, show for 72 = 1 that every point set { a ) , a E R, is a Bore1 set. Also check that the intervals (a, b), [a, b) (-co,a), (b, co) are Borel sets.

Now recall Example 1.4.7, where we generated the u-field o ( Y ) from the sets Ai = { w : Y(w) = yi) for a discrete random variable Y with values yi. If Y is a random variable assuming a continuum of values, the a-field generated from the sets {w : Y(w) = y), y E B, is, from a mathematical point of view, not rich enough. For example, sets of the form {w : a < Y ( w ) 5 b) do not belong to such a a-field, but it is desirable to have them in a ( Y ) .

We saw in Example 1.4.7 that , for a discrete random variable Y, the sets {w : a < Y(w) 5 b) belong to a (Y) . Therefore we will require as a minimal assumption that these sets belong to u(1') when Y is any random variable. Since we are also intcrcstcd in random vectors Y , we immediately define the a-fields a ( Y ) for the multivariate case:

Let Y = (Yl,. . . , Y,) be an n-dimensional random vector, i.e. Yl, . . . , Y, are random variables. The a-field a ( Y ) is the smallest a-field containing all sets of the form

{ Y E (a, b]) = {w : ai < I>(w) 5 hi, i = 1 , . . . , 7 ~ ) ,

- o o < a j < b j < o o , j = l , . . . , n .

We call u (Y) the a-field generated by the random vector Y .

An element of a ( Y ) tells us for which w E R the random vector Y assumes values in a rectangle (a, b] or in a more general Borel set.

The a-field u(Y) generated by Y contains the essential information about the structure of the random vector Y as a function of w E R . I t contains all sets of the form {w : Y(w) E C ) for all Borel sets C c Rn.

You will admit that it is difficult t o ilnagine lhe a-field a ( Y ) , and it is even more complicated for a stochastic process.

I

Page 39: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

For a stochastic process Y = (k;, t E T, w E R), the a-.field a ( l r ) is the smallest a-field containing all sets of the form

{w : the sample path (l;(w), t E T ) belongs to C )

for all suitable sets C of functions on T . Then a (Y) is called the a-field generated b y Y.

66 CH.APTER, 1.

In view of our restricted tools, we cannot make this definition more precise, but we will support our imagination about the a-field a (Y) by considering an example:

1.4. CONDITIONAL EXPECTATION 6 7

Example 1.4.9 Let B = (B,,s 5 t) be Brownian motion on [O,t], see Sec- tion 1.3.1, and

Ft = a ( B ) = a(B,, s 5 t) . This is the smallest a-field containing the essential information about the struc- ture of the process B on [0, t]. One can show that this a-field is generated by all sets of the form

for any n-dimensional Bore1 set C , any choices of ti E [0, t], n L 1.

We learnt that the a-field a ( Y ) is a non-trivial object. In what follows, we want tan iise the following r~ile of thumh in order to be able to imagine a(Y):

For a random variable, a random vector or a stochastic process Y on R, the a-field a(Y) generated by Y contains the essential information about the structure of Y as a function of w E R. It consists of all subsets {w : Y (w) E C ) for suitable sets C.

Because Y ger~erates a a-field, we also say that Y contains information represented b y a(Y) or Y carries the information a(Y).

We conclude with a useful remark. Let f be a funct,ion acting on Y, and consider the set

{w : f (Y(w)) E C I

for suitable sets C. For "nice" functions f , this set belongs to a (Y) , i.e.

1 Example 1.4.10 As before, let B be Brownian motion and define the a-fields

This means that a function f acting on Y does not provide new information about the structure of Y. We say that the information carried by f ( Y ) is contained in t,he information a(Y). We give a simple example:

Consider the function f ( B ) = Bt for a fixed t . Given that we know the structure of the whole process (B,, s 5 t) on R, then we also know the structure of the random variable Bt, thus a (Bt ) c Ft . The converse is clearly not true. If we know the random variable Bt we cannot reconstruct the whole process (B,, s 5 t) from it.

1.4.3 The General Conditional Expectation

In Section 1.4.1 we defined the conditional expect,at,ion E(X I Y) of a random variable X given the discrete random variable Y. This definition does not make explicit use of the values yi of Y, hut, it depends on the subsets Ai = {w : Y(w) = y i ) of R. We learnt in Example 1.4.7 that the collection of the sets Ai generates the a-field a(Y). This is the starting point for the definition of the general conditional expectation E ( X I F) given a a-field 3 on R . In applications we will always choose F = a(Y) for appropriate random variables, random vectors or stochastic processes Y. In the previous section we tried to convince you that the essential information about the structure of Y is contained in the a-field a (Y) , generated by Y. In this sense, we say that Y carries the information o(Y).

Let Y, Yl, Y2 be random variables, random vectors or stochastic processes on R and 3 be a a-field on R . We say that

the information of Y is contained in 3 or Y does not contain more information than that contained in 3 if a (Y) C 3, and

Y2 contains more information than Yl if a(Y1) c a(Y2).

Now we are prepared to give a rigorous definition of the conditional expectation E ( X 1 3 ) under an abstract a-field F :

Page 40: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

68 CH-4P TER 1.

A random variable Z is called the conditional expectation of X given the a-field F (we write Z = E ( X IF ) ) if

Z does not contain more information than that contained in F : a (Z ) c F .

Z satisfies the relation

E ( X I A ) = E ( Z I A ) for all A E 3. (1.33)

In Appendix A6 we show the existence and uniqueness of E ( X 1 3 ) by measure- theoretic means.

Property (1.33) shows that the random variables X and E ( X I .F) are "close" to each other, not in the sense that they coincide for any w, but aver- ages (expectations) of X and E ( X 1 3 ) on suitable sets A are the same. This supports our imagination about the random variable E ( X 1 3 ) which we' gained in Section 1.4.1:

The conditional expectation E ( X 1 3 ) is a coarser version of the original random variable X .

We mention that the defining property (1.33) leaves us some freedom for the coils~ructioil of llle random variable Z = E ( X 1 F ) . This means that there can exist versions Z' of E ( X I F ) which can differ from Z on'sets of probability 0. For this reason, all relations involving E ( X I F ) should actually be interpreted in an almost sure sense, and we should also indicate this uncertainty in all relevant formulae with the label "as.". However, once having pointed o i ~ t this fact, we will find it convenient to suppress this label everywhere.

Example 1.4.11 (Conditional expectation under discrete condition) We want to show that the definition of E ( X I Y) in Section 1.4.1 yields a special case of the above definition of E ( X 1 3 ) when F = a(Y). Rccall from Examplc 1.4.7 that every element A of a(Y) is of the form

In (1.29) we defined E (X I Y ) as the random variable Z such that

Z(w) = E ( X I Ai) for w E Ai .

We learnt on p. 62 that Z is merely a function of Y, not of S, hence u(Z) c ~ ( 1 ' ) ; see the discussion at the end of Section 1.4.2. Moreover, for A given by (1.341,

1 On the other hand, we see that ZIA is a discrete random variable with expec- tation

Thus Z satisfies the defining relation (1.33) and it does not contain more information than Y (it is a function of Y). Therefore it is indeed the conditional cxpcctation of X given the a-field u(Y) .

In the case of a discrete random variable Y we have just learnt that E ( X I Y) and E ( X I a(Y)) represent the same random variable. This suggests the fol- lowing definition:

Let Y be a raildbm variable, a random vector or a stochastic process on R and a[1') theta-field generated by Y.

The codi t ional 'expectat ion of a random variable X given Y is defined by

E (X 1 1') = E (.Y 1 o (1')) .

1 Example 1.4.12 (The classical conditional probability and conditional ex- pectation) The classical conditional probability and conditional expectation are special I

f cases of the gerleral no ti or^ of cor~ditional expectation defined on p. 68. Indeed, let B be such that P ( B ) > 0, P (BC) > 0 and define FB = a({B)) . We know from Example 1.4.6 that FB = {O,R, B, Bc). An appeal to Example 1.4.11 yields that

E (XIFB) (W) = E(XI B) for w E B.

This is the classical notion of conditional expectation. If wc spccify X = IA (1.34) for some event A, we obtain for w E B,

1 The right-hand side is the classical conditional probability of A given B.

Page 41: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

70 CH-APTER 1. I 1.4. CONDITIONAL EXPECTATION 71 !

1.4.4 Rules for the Calculation of Conditional Expecta- 1 tions I

The defining property (1.33) of E ( X IT) is not a constructive one. Therefore it is in general difficult, if not impossible, to calculate E ( X IT) . The case 7 = g(Y) for a discrete random variable Y, which was discussed in Exam- ple 1.4.11, is an exception. For this reason, it is important to be able to work with conditional expectations without knowing their particular forms. This means we have to know a few rules in order to deal with conditional expecta- tions in standard situations.

In what follows, we collect the most common rules. In some cases we will give reasons for these rules, in the other situations we have to rely on intuitive arguments. We start with an existence and uniqueness result.

If EIXI < oo, the conditional expectation E ( X I T ) exists and is unique in the sense as discussed in Appendix A6.

In Section 1.4.1 we encountered rules for the calculation of conditional expec tations under a discrete condition. They remain valid in the general case.

Rule 1

The conditional expectation is linear : for random variables X1, X2 and constants cl, c2,

E( [c lX i + ~2x21 I T ) = ci E ( X I I T ) + c2 E(X2 I T ) . (1.35)

Now take A = in property (1.33). You immediately obtain:

Rule 3

If X and the 0-field T are independent, then E ( X ( T ) = E X .

In particular, if X and Y are independent, then E ( X I Y) = E X .

This statement needs some discussion: what does independence between X and 3 mean? It means that we do not gain any information about X , if we know 3 , and vice versa. More formally: the random variables X and IA are independent for all A E F. Nnw, by independence,

A comparison with (1.33) yields that the constant random variable Z = E X is the conditional expectativri E ( X IT). This proves Rule 3.

This follows by an application of the defining property (1.33) to the right- and left-hand sides of (1.35) (try to prove it!). I

Rule 2

The expeclatioris of X and E ( X I T ) are the same:

E X = E [ E ( X I T)] . 1

Rule 4

If the 0-field (T(X), generated by the random variable X , is contained in T , then

E ( X I 3 ) = X .

In particular, if X is a function of Y, o ( X ) C o(Y), thus E ( X I Y) = X .

This means that the information contained in T provides us with the whole

j : information about thc random variablc X. If we know everything about the structure of X, we can deal with it as if it was non-random and write the value

ti X(w) in front of the conditional expectation E ( 1 I T ) = 1:

This rule can be extended to more general situations:

Also the third claim (see (1.30)) carries over from the discrete case: I

I

I I I

Rule 5

If the g-field a ( X ) , generated by the random variable X , is contained in T , then for ariy raridvrri variable G,

E ( X G 1 T ) = X E ( G / T ) .

In particular, if X is a function of Y, a ( X ) C o ( Y ) , thus E ( X G I Y ) = XE(G1Y).

Page 42: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

72 CHAPTER 1.

Indeed, given 3 , we can deal with X as if it was a constant, hence we can pull X ( w ) out of the updated expectation and write it in front of E ( G IT) .

Rule 6

If T and T' are two 0-fields with 3 C TI, then

E ( X 13) = E ( E ( X 13') I F ) , (1.36)

E ( X I 3 ) = E ( E ( X I T ) I F r ) . (1.37)

For obvious reasons, Rule 6 is sometimes called the tower property of condi- tional expectations.

Rule (1.37) can be justified by Rule 4: since F C TI, E ( X I F ) does not contain more information than F', i.e. given T', we can deal with E ( X I T) like a constant:

Rule (1.36) can formally be derived from the defining property (1.33) which says that for A E T and Z = E ( X I F),

On the other hand, by Rule 5 and since A E 3 c T r ,

Now take expectations and apply Rule 2 to the right-hand side:

Thus Z' = E ( E ( X 13') I T ) also satisfies (1.38), but since E ( X 1 3 ) is unique, we must have Z = Z', which proves (1.36).

We finish with a generalization of Rule 3.

Rule 7

If X is independent of 3, and the information carried by the random variable, the random vector or the stochastic process G is contained in 3, then for any function h(x, y) ,

E(h(X, G) I T ) = E ( E x [h(X, GI1 I T ) ,

where Ex[h(X,G)] means that we fix G and take the expectation with respect to X.

1.4. CONDITIONAL EXPECTATION 73

We illustrate Rule 7 by an example.

Example 1.4.13 Let X' and 1' be independent random variables. Then Rules 7 and 5 give

E ( X Y Y ) = E(E*y(X'Y) I Y) = E ( Y E X 1 Y) = Y E X ,

In the following two examples we want to exercise the rules for calculating conditional expectations.

Example 1.4.14 (Brownian motion) Recall the definition of~Brownian niotion B = (Bt, t > 0) from p. 33. We associate with B an increasing strea~n of information about the structure of the prvcess which is represented by the 0-fields 3, = o(BX, z 5 s). We want to calculate

E ( B t I F s ) = E ( B t I B z , x 5 s ) for s 2 0 .

Clearly, if s > t , 3, 3 F t , and so Rule 4 gives

Now assume s < t. Then, by linearity of the conditional expectation (Rule l ) ,

Since Bt - B, is independent of F s , Rule 3 applies:

E(Bt - B, I 3 ,) = E(Bt - B,) = 0

Moreover, g(B,) C g(Bx, x 5 s) = FS, hence E(B, I 3 , ) = B,. Finally,

Example 1.4.15 (Squared Brownian motion) As in Example 1.4.14, B denotes Brownian niotion and F8 = n(B,,x 5 s). Lie consider the stochastic process St = B; - t, t > 0. The same arguments as in Example 1.4.14 yield

E(St 1 FS) = Xt for s 2 t

Page 43: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

74 CHAPTER I . 1.4. CONDITIONAL EXPECT4TION

For s < t , we write

Taking conditional expectations, we obtain

Notice that Bt - B, and (Bt - B,)' are independent of F s , and that u(Bz) C

a ( B s ) c Fs. Applying Rules 3-5, we obtain

= ( t - s ) + ~ : + O - t = x , .

Finally,

E ( X t I Fs) = Xmin(s,t) . Later we will take this relation as the defining property for a martingale; see (1.41).

1.4.5 The Projection Property of Conditional Expecta- tions

In what follows, 3 is a 0-field and L 2 ( 3 ) is the collection of random variahles Z on R, satisfying the following conditions:

Figure 1.4.16 A n illustration of the projection property of the conditional expecta- t ion E ( X I F). W e mention that < Z, 1' >= E(Z1') for Z , Y with EZ' < w and EI" < w defines an inner product and llZ - YII = J< Z - Y , Z - Y > a distance between Z and Y . As in Euclidean space, we say that Z and Y are orthogonal if < Z, Y >= 0. I n this sense, E (X 13) i s the orthogonal projection of X on L Z ( 3 ) : < X - E ( X IF) , Z >= 0 for all Z E L 2 ( 3 ) , and < X - 2, X - Z > is minimal for Z = E ( X I F ) .

Z has a finite second moment: E Z 2 < w, l i In this sense, E ( X I 3) is the projection of the random tianable X on the space The information carried by Z is contained in 7: a ( Z ) c F. If F = a ( Y ) L 2 ( F ) of the random variables Z carrying part of the information f i also see

this means that Z is a function of Y. the comments to Figure 1.4.16.

The Projection Property

Let X be a random variable with E X 2 < w. The conditional expecta- tion E ( X I T ) is that random variable in L 2 ( T ) which is closest to X in the mean square sense. This means that

E [ X - E ( X I T ) I 2 = min E ( X - 2 ) ' . Z E L Z ( F )

(1.39)

If 7 = a ( Y ) , E ( X I Y ) is that function of Y which has a finite second moment and which is closest to X in the mean square sense.

The random variable E ( X I T ) can be understood as an updated version of the expectation of X , given the information 3. The conditional expectation has a certain optimality property in the class L 2 ( 3 ) :

In what follows, we sometimes refer to E ( X 17) as the best prediction of X given 3. This means that relation (1.39) holds. To give some meaning to the word "prediction", we recorisider Examples 1.4.14 and 1.4.15. We proved that

E ( B t l B , , z ~ s ) = B s and E ( B ~ ~ - ~ ~ B , , z < s ) = B ~ - s .

Thus the best predictions of the future values Bt and B? - t , given the in- formation about Brownian motion until the present time s, are the present

Page 44: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

76 CHAPTER 1 . 1 .,5. MA RTINGALES I 77

values B, and Bz - S, respectively. This property characterizes the whole class of martingales with a finite second moment: the best prediction of the future values of the stochastic process is the present value; see Section 1.5.1.

In what follows, we indicate the steps for the proof of the projection prop- erty (1.39). We only make use of the rules for calculatillg conditio~lal expecta- tions. First notice that the random variable Z' 2 E ( X I T ) belongs to L2(T): it carries information only about F and has a finite second moment. The latter property follows by an application of Jensen's inequality (see (A.2) on p. 188) in combination with RIIIF! 2:

E [ ( E ( X IT))2] 5 E[E(X~ I T)] = EX^.

Now we are going to check the projection property (1.39). Let Z be any random variable in L2 ( T ) . Then

E ( X - 2 ) ' = E ( ( X - Z ' ) + ( Z ' - Z ) ) ~

= E ( X - Z ' ) ~ + E ( Z 1 - 2 ) ' + 2 E [ ( X - Z r ) ( Z ' - Z ) ] .

We treat the terms on the right-hand side separately. First consider E [ ( X - Z')(Z1 - Z)]. Since both, Z and Z', belong to L2(T) , so does Z - 2'. In particular, by Rule 5, -

E [ ( X - Z r ) ( Z ' - Z ) [ F ] = ( 2 - Z 1 ) E ( X - Z ' I F ) .

But by Rules 1 and 4,

Thus we proved

Hence,

E ( X - Z)2 2 E ( X - z ' ) ~ for all random variables Z in L2(T) . (1.40)

We also see that equality in (1.40) is achieved if Z = Z', so Z' = E ( X I T ) really represents that element of L2(T) , for which the minim4m of E ( X - 2)' is attained. This proves (1.39).

Notes add Comments

The notion.of conditional expectation is one of the most difficult ones in prob- ability theory, but it is also one of the most powerful tools. Its definition as a random variable given a a-field goes back to Kolmogorov. For its complctc thc- oretical understanding, measure-theoretic probability theory is unavoidable.

The conditional expectation is treated in every advanced textbook on prob- ability theory, see for example Billingsley (1995) or Williams (1991).

1.5 . Martingales

1.5.1 Defining Properties

The notion of a martingale is crucial for the understanding of the It6 stochas- tic integral. Indefinite It6 stochastic integrals are constructed in such a way that they constitute martingales. The idea underlying a martingale is a fair game where the net winnings are evaluated via conditional expectations. For- tunately, we now have this powerful instrument in our tool box; see Section 1.4.

Assume that (FL, t 2 0) is a collection of a-fields on the same spacc 0 and that all F t s are subsets of a larger c~-field T on n, say.

The collection (F t , t 2 0) of a-fields on n is called a jiltration if

Fs C Ft for all 0 5 s 5 t .

Thus a filtration is an increasing stream of information.

If (Fn , n = 0,1,. . .) is a sequence of a-fields on 0 and Fn c Fn+l for all n, we call (&) a filtration as well.

For our applications, a filtration is usually linked up with a stochastic process:

The stochastic process Y = (&, t 2 0) is said to be adapted to the filtra- tion (Tt , t >_ 0) if

a ( Y t ) c F t f o r a l l t > O .

The stochastic process Y is always adapted to the natural filtration gen- erated b y Y :

Tt = o(Y,, s 5 t ) .

Thus adaptedness of a stochastic process Y means that the &s do not carry more information than Ft .

Page 45: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

78 CHAPTER 1. ( 1.5. MARTINGALES 70

If Y = (Y,, n = 0,1, . . .) is a discrete-time process we define adaptedness in an analogous way: for a filtration ( F n , n = 0,1, . . .) we require that o(Yn) c Fn.

Example 1.5.1 (Examples of adapted processes) Let (B t , t > 0) be Brownian motion and (Fk, t > 0) be the corresponding natural filtration. Stochastic processes of the form

where f is a function of two variables, are adapted to (Ft , t > 0). This includes the processes

But also processes, which may depend on the whole past of Brownian motion, can be adapted. For example,

(6) - max B, or (7) - min ~ , 2 . X t - o<s<t Xt - osr<t

If the stochastic process Y is adapted to the natural Brownian filtration (F t , t 2 O), we will say that Y is adapted to B ~ o w n i a n motion. This means that is a function of B,, s 5 t .

The following processes are not adapted to Brownian motion:

where T > O is a fixed number. Indeed, these processes require the knowledge (9) of Brownian motion a t future instants of time. For example, consider Xt .

For its definition you have to know BT a t times t < T. u

Clearly, one can enlarge the natural filtration and obtain another filtration to which the stochastic process is adapted.

Example 1.5.2 (Enlarging a filtration) Consider Brownian motion B = ( B t , t 2 0) and the corresponding natural filtration Ft = o(Bs, s 5 t), t 2 0. The stochastic process Xt = Bf generates the natural filtration

which is smaller than (Ft). Indeed, for every t , F{ C Ft, since we can only reconstruct the whole information about JBtJ from B f , but not about Bt: we

I car1 say nothing about the sign of Bt. Then ( F L ) is also a filtration for (Bz) .

Thus we can work with different filtrations for the same process. To illus- trate this aspect from an applied point of view, we consider an example from stochastic finance. In this field, it is believed that share prices, exchange rates, interest rates, etc., can be modelled by solutions of stochastic differential equa- tions which are driven by Brownian motion; see Chapter 4. These solutions are then functions of Brownian motion. This process models the fluctuations of the financial market (independent movements up and down on disjoint time intervals). These floctuations actually represent the information about the market. This relevant knowledge is contained in the natural filtration. I t does not take information from outside into account. However, in f i~ance there are always people who know more than the others. For example, they might know that an essential political decision will be taken in the very near future which will completely change the financial landscape. This enables the informed per- sons to act with more competence than the others. Thus they have their own filtrations which can be bigger than the natural filtration.

Now consider a stochastic process X = (Xt, t > 0) on a and suppose you have the information Fa a t the present time s.

How does this in formation influence our knowledge about the behavior of the process X in the future?

If Fs and X are dependent, we can expect that our information reduces the uncertainty about the values of Xt a t a future instant of time t . If we know that certain events happened in the past, we may include this knowledge in our calculations. Thus Xt can be better predicted with the information Fs than without it. A mathematical tool to describe this gain of information is the conditional expectation of Xt given Fs:

E(Xt I ;F,) for 0 5 s < t

We learnt in Section 1.4.5 that E(Xt I Fs) is the best prediction of Art given the information F s . Also recall from Examples 1.4.14 and 1.4.15 that the stochastic processes Xt = Bt and Xt = Bz - t ( B is Brownian motion) satisfy the condition E(Xt I FS) = X, for s < t . For these processes X, the best prediction of the future value Xt given Fs is the present value X,. Clearly, this may change if we consider arlother filtration, a11d therefore we must always say which filtration we consider.

Page 46: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

80 CHAPTER. 1. I 1.5. MARTINGALES 8 1

The stochastic process X = (Xt , t 2 0) is called a continuous-time mar- tingale with respect to the hlti-ation (Ft, t > 0). we write (X, (F t ) ) , if

I EIXtI < co for all t 2 0 . I I X is adapted to (F t ) ; see p. 77 for the definition. I 1 E ( S , IFs) = X, for all 0 5 s < t ,

i.e. X, is the best prediction of Xt given FS.

1 I

It is also possible to define a discrete-time martingale X = (X,, n = 0 ,1 , . . .). In this case, we adapt the defining property (1.41) as follows:

+ 1 n = n 7 2 0 . (1.42)

We show that it suffices to require (1.42) for k = 1. Indeed, recalling Rule 6 from p. 72,

E(Xn+l I F71) = E[E(Xn+2 1 Fn+l) 1 F n ] = E(Xn+2 1 F71)

The sequence (Y,) is then called a martzngale diflerence sequence with respect to the filtratzon (F,).

In what fbllows, we often say that "(Xt, t >. 0), respectively ( X n , n = 0,1, . . .), is a martingale" without pointing out which filtration we lisp. This will be clear from the context.

A martingale has the remarkable property that its expectation function is constant.

t Indeed, using the &;fining property E(Xt 1 F,) = X, for s < t and Rule 2 on p. 70, we obtain

EX, = E[E(Xi 1 F8)] = EXt for all s and t .

This provides an easy way qf proving that a stochastic process is not a mar- tingale. For examplc, if B is: Brownian motion, EBt2 = t fur all t . Hence (BZ) cannot be a martingale. However, we cannot use this means to prove that a stochastic process is a martingale, since a process that has a constant expec- tation function need not be a martingale as the following example shows: we have EB: = 0 for all t , but (B:) is also not a martingale; see Example 1.5.5.

= E[E(Xn+3 1 Fn+2) I F n I = E(Xn+3 I & I ) 1 1.5.2 Examples

Now we define a martingale in the discrete-time case. . .I

The stochastic process X = (X, , ,n = 0,1,. . .) is called a discrete-t'ime martingale with respect to the filtration (Fn , n = 0,1, . . .), we write (X, (Fn)), if

EIX',I < CQ for all n = 0 ,1 , . . ..

X is adapted to (Fn).

E(Xn+l I F,) = X,, for all n = 0,1, . . . , (1.43)

i.e. X, is the best prediction of X,,+1 given F n .

It is not difficult to see that the defining property (1.43) can be rewritten in the forrn

E(EL+l I F,) = 0 , where Yn+i = Xn+l - X,, n = 0,1, . . . . (1.44)

In this section we collect some simple examples of stochastic processes which have the martingale property.

Example 1.5.3 (Partial sums of independent random variables constitute a martingale) Let (2,) be a sequence of independent random variables with finite expecta- tions and Zo = 0. Consider the partial sums I

I I

and the corresponding natural filtration ;F, = o(R0,. . . , R,) for n 2 0. Notice that

-T,=o(Zo ,..., Z,), n > 0 .

Indeed, the random vectors (20, .. . , Z n ) and (Ro, . . . , Rn) contain the same information since Ri = Z1 + . . . + Zi and Zi = Ri - for i = 1 , . . . , n . An application of Rules 1, 3 and 4 in Section 1.4.4 yields

Hence, if EZ, = 0 for all n, then (R,, n = 0,1,. . .) is a martingale with respect i to (F,, n = 0,1, . . .).

Page 47: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

82 CHAPTER 1. 1.5. M-ARTINGALES 83 I Example 1.5.4 (Collectirig information about a random variable) Let Z be a random variable on 0 with El21 < oo and ( F t , t 2 0) be a filtration on R Define the stochastic process X as follows:

Since Ft increases when time goes by, Xt gives us more and more information about the random variable Z. In particular, if o ( Z ) C Ft for some t, then Xt = 2. Wc show that X is a martingale.

An appeal to Jensen's inequality (A.2) on p. 188 and to Rule 2 on p. 70 yields

Moreover, XL is obtained by conditioning on the information F t . Hence it does not contain more information than Ft, so o(Xt) c Ft. It remains to check (1.41). Let s < t. Then an application of Rule 6 on p. 72 yields

E(Xt I A) = E [ E ( Z I F t ) 1 Fs] = E ( Z I F s ) = X y

Thus A' obeys the defining properties of a continuous-time martingale; see p. 80. . . Example 1.5.5 (Brownian motion is a martingale) Let B = (BL, t > 0) be Brownian motion. We conclude from Examples 1.4.14 and 1.4.15 that both, (Bt , t > 0) and (Bt2 - t , t 2 O), are martingales with respect to the natural filtration 3 t = o(Bs, s 5 t) .

In the same way, you can show (do it!) that ((B: -3tBt), (F t ) ) is a martingale.

Try to find a stochastic process (At) such that ((B! +At), (Ft)) is a martingale. Hint: first calculate E[((Bt - B,) + B , ) ~ I Fs] for s < t.

The following example of a martingale transform is a first step toward the defi- nition of the It6 strochastic int#egral. Indeed, such a transform can be considered as a discrete analogue of a stochastic integral.

Example 1.5.6 (Martingale transform) Let Y = (Y,,, n = 0,1, . . .) be a martingale difference sequence with respect to the filtration (3,T,, ,n = 0 , 1 , . . .); see (1.44) for the definition. Consider a stochastic process C = (C,, n = 1 ,2 , . . .) and assume that, for every n, the information carried by C, is contained in Fn-1, i.e.

This means that, given FTLP1, we completely know Cn a t time n - 1.

If the sequence (C,, n = 1,2, . . .) satisfies (1.45), we call it previsible or predictable with respect to (Fn).

Now define the stochastic process

For obvious reasons, the process X is denoted by C Y. It is called the martingale transform of Y by C.

The martingale transform C Y is a martingale if EC: < oo and EY: < oo for all n. We check the three defining properties on p. 80:

where we made use of the Cauchy-Schwarz inequality EICiY,I 5 [EC?EY,']'/~ (see p. 188). Clearly, Xn is adapted to Fn since Yl,. . . , Yn do not carry more information than ;F,, and C1,. . . , C, is predictable. Hence o(Xn) C F n . More- over, applying Rule 5 on p. 71 and recalling that (C,) is predictable, we obtain

In the last step we used the defining property (1.44) of the martingale difference sequence (Y,). Hence (X, - XnP1) is a martingale difference sequence, and (X,) is a martingale with respect to (F,).

Example 1.5.7 (A Brownian martingale transform) Consider Brownian motion B = (By, s < t) and a partition

Using the independent increment property of B , it is not difficult to see (check it!) that the sequence

forms a martingale difference sequence A B with respect to the filtration given by

F 0 = { 0 , 0 ) , = o ( B , l i ) i = 1 , ..., n .

Page 48: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Now consider the transforming sequence

It is predictable with respect to (F,) (why?). The martingale transform

B A B is then a martingale:

The sums on the right-hand side have the typical form of Riemann-Stieltjes sums which would be used for the definition of the Riemann-Stieltjes inte- gral J; B, dB,; see Section 2.1.2. However, this integral does not esist in the Riemann-Stieltjes sense since the sample p_aths of Brownian motion are too irregular. We will see in Section 2.2.1 that B A B is a discrete-time analogue of the It6 stochastic integral S i B,dB,.

1.5.3 The Interpretation of a Martingale as a Fair Game

At the beginning of this section we mentioned that martingales are consid- ered as models for fair games. This interpretation comes froni the defining properties of a martingale; see p. 80.

Suppose you play a game in continuous time; i.e. a t every instant of time t you have a random variable Art as the value of the game. Suppose that (Xt) is adapted to the filtratioll (.Ft). Think of Xt - X, as the net winnings of the game per unit stake in the time frame (s, t ] . Then the best prediction (in the sense of Section 1.4.5) of the net winnings given the information .Fs a t time s < t has value

E(Xt - X , I F s ) = E(Xt I F s ) - Xs .

If (X, (.Ft)) is a martingale, the right-hand side vanishes. This means:

The best prediction of the future net winnings per unit stake in the interval ( s , t] is zero.

This is exactly what you would expect to be a fair game

People in finance do certainly gamble (although they woilld not admit this in public), and they even believe that they play fair. This is why they have become attracted by the world of marti~~gales and stochastic integrals.

\VP know that t h ~ 11lnrt.ingale transfornl C I' of a martingale difference sequence T 7 is a martingale: see Exarliple 1.5.6. It has an int>eresting interpre- tat,iori in tlie context of fair games: think of 1; a.s your net wirlriirigs per unit stake a t the nth game which are adapt,ed to a filtration (Fa) . Your stakes C,, constitute a predictable sequence with respect to (.Fn), i.e. a t the n t h game your stake C,, does not contain more information than does. At time n. - 1 this is the best information we have about the game. The mar- tingale transform C . I' gives you the net winnings per game. In particular, ( C Ir),, = EL1 Ci are the net winnings up to time n , and C,l:, are the net winnings per stake C, a t the nth game. It is fair since the best prediction of the net winnings C,I:, of the nth game, just before the nth game starts, is zero: E(C,I',, I FT,-l) = 0.

In Chapter 2 we will learn about the continuous-time analogue to martin- gale transforms: the It6 stochastic integral.

Notes and Comments

Martingales constitute an in~port~ant class of stochastic processes. The theory of mart,ingales is considered in all modern textbooks on stochastic processes. See Karatzas and Shreve (1988) and Revuz and Yor (1991) for the advanced theory of continuous-time martingales. The book by Williams (1991) contains an introduction to conditional expect,ations arid discrete-time martingales.

Page 49: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

The Stochastic Integral

In this chapter we introduce the important notion of It6 stochastic integral. We know from Section 1.3.1 that the sample paths of Brownian motion are nowhere differentiable and have unbounded variation. This has major consequences for the definition of a stochastic integral with respect to Brownian sample paths.

We discuss the notion of ordinary integral in Section 2.1 and also consider an extension, the Riemann-Stieltjes integral. The latter can be applied to Brownian sample paths if the integrand process is sufficiently smooth. How- ever, we will see in Sectio~i 2.1.2 that B1.uwriia~i sarriple patl~s caririot be iri- tegrated with respect to themselves, which is a major handicap. Therefore we are forced to consider an integral which is not defined path by path. This will be done in Section 2.2. There we define the It6 stochastic integral as a mean square limit of certain Riemann-Stieltjes sums. This has the disadvantage that we lose the intuitive interpretation of such an integral which is naturally provided by a pathwise integral. However, it will be convenient to think of the stochastic integral in terms of an approximating Riemann-Stieltjes sum.

In Section 2.3 we will learn about another important tool: the It6 lemma. It is the stochastic analogue of the ordinary chain rule of differentiation. The It8 lemma will be crucial in Chapter 3, where we intend to solve It8 stochastic differential eqllations.

In Section 2.4 we will learn about one more stochastic integral whose value usually differs frorn the It6 stochastic integral. It is called the Stratonovich stochastic integral and can be useful as an auxiliary tool for the solution of It8 stochastic differential equations.

Page 50: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

88 CHAPTER 2.

2.1 The Riemann and Riemann-Stieltjes Inte- grals

This section is not needed for the understanding of the Itb stochastic int,egral, but i t gives some information about the history of integration and explains why classical integrals may fail when the integrand or the integrator are Brownian sample paths.

We discuss some approaches to pathwise integration. In particular, we are interested in integrals of the form J: f ( s ) dB,(w), where (Bt(w), t 2 0) is a given Brownian sample path and f is a deterministic function or the sample path of a stochastic process. In particular, we recall the Riemann integral J; f (t) dt a s t h e classical integral, and some of its properties. Then we discuss Riemann-Stieltjes integrals which are close in spirit to the pathwise integral J; f (t) dBl(w). We will see that , under some smoothness conditions on f , the latter integral is well defined, and we will also point out that , for example, the

1 Riemann-Stieltjes integral So Bt(w) dBt(w) cannot be defined.

2.1.1 The Ordinary Riemann Integral

We want to recall the notion of ordinary integral, which is also called the Riemann integral. You should be familiar with this integral from a course on elementary calculus.

Suppose, for simplicity only, that f is a real-valued function defined on [O, 11, but instead we could consider any interval [a, b ] .

Consider a partition of the interval [0, 11:

and define & = t i - t i P l , i = l , ..., n

An intermediate partition 0, of T, is given by any values yi satisfying ti-l 5 yi 5 t i for i = I , . . . ,n. For given partitions rn and 0, we can define the Riemann sum

Thus a Riemann sum is nothing but a weighted average of the values f (yi), where the weights are the corresponding lengths Ai of the intervals [ti-1, ti]. We also know that Sn is an approximation to the area between the graph of

I 2.1. THE RIEAlA4NN AND RIEIC1,4NNPSTIELTJES INTEGRALS 80

t t 1 t 2 t 3 t 4 t 5

Y 1 Y'2 Y 3 Y 4 Y 5

Figure 2.1.1 An illustration of a Rie~nann sum with partition (ti) and intermediate partztion (y,). The sum of the rectangular areas approximates the area between the graph of f ( t ) and the t-axis, i.e. Jd5 f ( t ) dt.

the function f and the t-axis, provided f only assumes non-negative values. See Figure 2.1.1 for a n illustration.

Now let the mesh of the partition T, go to zero, i.e.

mesh(-r,) := max A, = , max (ti - ti-1) -t 0 i=l, ..., n z = l , ..., n I 'F.

If we proceed in this way, the points ti = t!") clearly have to depend on n, but we suppress this dependence in our notation.

If the limit n

S = n + w lim Sn = n + w lim f (gi) Ai i = l

exists as mesh(r7,) -+ 0 and S is independent of the choice of the parti- tions rn and their inter~r~ediate partitions a,, the11 S is called the ordi- nary or Riemann integral of f on [O, 11.

Page 51: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

90 CHAPTER 2.

We write

We know that J; f ( t ) dt exists if f is sufficiently smooth, for example, contin- uous or piecewise continuous on [0, I ] .

Figure 2.1.2 Two Riemann sums for the integral t d t ; see Example 2.1.3. The partztion is given by t i = i / 5 , i = 0 , . . . , 5 . The left (left figure) and the right (right figure) end points of the intervals [(i - 1 ) / 5 , i / 5 ] are taken as points yi of the inlerrnediute purtition.

Example 2.1.3 We want to calculate the integral J; t dt as the limit of certain Riemann sums. See Figure 2.1.2 for an illustration. We already know that

1 So t dt = 0.5; it represents the triangular area between the graph of f ( t ) = t and the t-axis. For numerical approximations it is convenient to choose an equidistant partition of [O, 11:

First take the left end points of the intervals [( i - l ) /n , i /n] for the intermediate art it ion:

2.1. THE R,IEAIANN AND RIEhlA N N -STlET,TJES lNTEGR.4LS 91

Using the well-known sum formula

Analogously, take the right end points of the i ~ ~ t e ~ v a l s [(i - l ) / r ~ , i l r~] :

and conclude, again using (2.2) , that

Now choose the yis as the middle points of the intervals [(i - l ) / n , i / n ] and

denote the corresponding Riemann sums by sArn). Since s:) 5 < - we may conclude that s!~~' -t S = 0.5.

The Riemann integral is taken as a model for the definition of any type of integral. The new kind of integral should have as many properties in common with the Riemann integral as possible. Such properties are given below.

For Riemann integrable functions f , fl and f2 on [O, I] the following properties hold:

The Riemann integral is linear, i.e. for any constants cl and ca :

1 1 1 1 [c1 f i ( f ) + c2 f2 ( t , ]d t = c1 1 f i ( t ) dt + c2 1 f d t ) d t .

The Riemann integral is linear on adjacent intervals:

~ l i ( t ) d t = ~ a f ( t ) d t + ~ l f ( t ) d t for O < a < l .

Page 52: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

92 C H A P T E R 2.

One can define the indefinite Riemann integral as a function of the upper limit:

2.1.2 The Riemann-Stieltjes Integral

In probability theory it is usual to denote the expectation of a random variable

where Fx denotes the distribution function of X. This notation refers to the fact that E X is defined as a Riemann-Stieltjes integral, or as a Lebesgue- Stieltjes integral. This means, roughly speaking, that

..

I-, t ~ F x (t) )Pi [Fx (ti) - FX (ti-I)] i

for a partition (ti) of B and a corresponding intermediate partition (yi). Also recall from a course on elementary calculus that you can define the integral

f (t) dg(t) as 5: f (t)gl(t) dt, provided the derivative g' (t) exists.

Both, Srm t dFX (t) and 1; f (t) dg(t), are examples of how one could ap- proach the problem of integrating one function f with respect to another func- tion g. It is the aim of the present chapter to suggest some methods. In

particular, we would like to obtain an integral of type J: f (t) dBt (w), where f is a function or a stochastic process on [O, I] and Bt(w) is a Brownian sample path. We have a major difficulty in defining such an integral since the path Bt(w) does not have a derivative; see Section 1.3.1. However, one can find some kind of pathwisc intcgral which sometimes allows one to evaluate the integral f (t) dBt(w). This kind of integral is called the Riemann-Stieltjes integral. It is a classical tool in many fields of mathematics.

In what follows we give a precise definition of the Riemann-Stieltjes inte- gral. As for t,he R i ~ m a n n int,egral, we consider a partition of the interval [O,l]:

2.1. T H E R I E M A N N A N D RIEILIANN-STIELTJES I N T E G R A L S 93

and an intermediate partition a, of T?,:

a,& : 5 yi 5 ti for 2 = 1 ,..., 7 % .

Let f and g be two real-valued functions on [0, I] and define

The Riemann-Stieltjes sum corresponding to T, and a, is given by

Notice the similarity with a Riemann sum; see (2.1). In that case, g(t) = t . Thus, a Riemann-Stieltjes sum is obtained by weighing the values f (yi) with the increments Aig of g in the intervals [ti-1, ti].

If the limit n

S = lim S,, = lim f (yi) Aig n+m ~E+W

i= l

exists as mesh(r,) -t 0 and S is independent of the choice of the par- titions 7, and their intermediate partitions a,, then S is called the Riemann-Stieltjes integral of f with respect to g on [O, I].

We write

It raises the question:

When does the Riemann-Stieltjes integral f (t) dg(t) exist, and is it possible to take g = B for Brownian motion B on [0, I] ?

This question does not have a simple answer and needs some more sophistica- tion. In some textbooks one can find the following partial answers:

The functions f and g must not have discontinuities at the same point t E [O,l].

Assume that f is corltinuous and g has bounded variation, i.e.

Page 53: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

where the suprenlum (cf. its definition on p. 211) is taken over all possible partitions r of [0, I ] .

Then the Riemann-Stieltjes integral J; f ( t ) d g ( t ) exists.

Notice that the latter result is not applicable to the integral J: f ( t ) d B t ( w ) . Indeed, we learnt in Section 1.3.1 that Brownian sample paths B t ( w ) do not have bounded variation. However,

bounded variation of g is not necessary for the existence of the Riernann-Stieltjes integral J: f ( t ) d g ( t ) ,

although the contrary claim can be found in some books. Weaker conditions for the existence of J: f ( t ) d g ( t ) are not very well known, but they were already found by L.C. Young in 1936; see the recent papers by Dudley and NorvaiSa (1998a,b) for an extensive discussion.

Without going into detail, we give a sufficient condition for Riemann- Stieltjes integrability which is close to necessity. First we give a definition:

The real-valued function h on [0, I ] is said to have bounded p-variation for some p > 0 if

1 where the supremum is taken over all partitions i of [0, 11.

Notice that h has bounded variation if p = 1.

' lhe Riemann-Stieltjes integral J,' f ( t ) d g ( t ) exists if the following con- ditions are satisfied:

The functions f and g do not have discontinuities at the same point t E [0, 11.

The function f has bounded pvariation and the function g has bounded q-variation for some p > 0 and q > O such that p-' + q-' > 1.

In some cases these conditions can be verified. We give an example for the integral J,' f ( t ) d B t (w) .

2.1. T H E RlEhlA N N ANT) RIEhfANN S T I E L T J E S I N T E G R A L S 95

Example 2.1.4 As before, B = ( B t , t > 0 ) denotes Brownian motion. I t is well known that sample paths of Brownian motion have bounded pvariation on ally fixed finit,e interval, provided that p > 2, and unbounded pvariation for p < 2. See Taylor (1972).

Now consider a deterministic function f ( t ) on [0, I ] or a sample path of a stochastic process f ( t , w ) . According to the above theory, we can define the Rieinann-Stieltjes integral J,' f ( t ) d B t ( w ) with respect to the sample path B t ( w ) , provided f has bounded q-variation for some q < 2. This is satis- fied if f has bounded variation, i.e. q = 1.

Assume that f is a differentiable function with bounded derivative f l ( t ) . Then it follows by an application of the mean value theorem that

where K > 0 is a constant depending on f . Then

11 11

sup C I f ( t i ) - f ( tz-l)l I K ( t i - t i - 1 ) = K < 00 l=l P= 1

Hence f has bounded variation, and so the following statement holds:

Assume f is a deterministic function or the sample path of a stochastic process. If f is differentiable with a bounded derivative on [O, I], then the Riemann-Stieltjes integral

i1 f ( t ) d B t ( w )

exists for every Brownian sample path B t ( w ) .

In particular, you can define the following integrals as Riemann-Stieltjes inte- grals:

This does not rnean that you can evaluate these integrals explicitly in terms of Brownian motion.

The above discussion may be slightly misleading since it suggests that one can define the integral J: f ( t ) d B t ( w ) for very general integrands f as a Riemann- Stieltjes integral with respect to a Brownian sample path. However, this is not

Page 54: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

96 CHAPTER 2.

the case. A famous example, which is also one of the motivating examples for the introduction of the It6 stochastic integral, is the integral

Brownian motion has bounded pvariation for p > 2, not for g < 2, and so the sufficient condition 2p-1 > 1 for the existence of I ( B ) (see p. 94) is not satisfied. Furthermore, it can be show11 that this sirrlple integral does not exist as a Riemann-Stieltjes integral.

Moreover, one can show the following: if J,' f ( t) dg(t) exists as a Riemann- Stieltjes integral for all continuous functions f on [0, I], then g necessarily has bounded variation; see Protter (1992), Theorem 52. Since one of our aims is to define the integrals J,' f (t) dBt(@) for all continuous deterministic functions f on 10, I], the Rienla~~n-Slieltjes i~ltegral approach rmust fail since Brownian sample paths do not have bounded variation on any finite interval.

It is the objective of the following sections to find another approach to the stochastic integral. Since pathwise integration with respect to a Brownian sample pa.th, as suggested by the Riemann-Stieltjes integral, does not lead to a sufficiently large class of integrable functions f , we will try to define the integral as a probabilistic average. This approach has the disadvantage in that it is iess intuitive than the Riemann-Stieltjes integral, regarding the form of the integrals.

Notes and Comments The Riemann and Riemann-Stieltjes integrals are treated in textbooks on calculus. However, it is not easy to find a conlprehe~isive trealrr~e~lt of the Riemann-Stieltjes integral. Young (1936) is still one of the best references on the topic. Extensions of the Riemann-Stieltjes integral are treated in the recent work by Dudley and NorvaiSa (1998a,b).

2.2 The It6 Integral

2.2.1 A Motivating Example

At the end of the previous section we learnt that it is impossible to define the integral J,' Bt(w) dBt(w) path by path as a Riemann-Stieltjes integra!. Here and in what follows, B = (Bt, t > 0) denotes Brownian motion. In order to

1 2.2. THE IT^ lNTECRAT,

I understand what is going on, we consider the Riemann-Stieltjes sums

where r,: O = to < tl < . . . < t,z-l < t , = t , (2.5)

is a partition of [0, t] and, for any function f on [0, t],

I! A f : Aif = f ( t i ) f ( t i - i ) , i = l , ..., n ,

I are the corresponding increments of f and

Thus the Riemann-Stieltjes sum Sn corresponds to the partition rn and the intermediate partition (yi) with yi = t iPl , i.e. yi is the left end point of the interval [t,-1, t,]. This choice is typical for the definition of the It6 stochastic integral. We will also see that another choice of intermediate partition (yi)

I

gives rise to the definition of another type of stochastic integral.

L Notice that S, can be written in the following form:

.l 1" (Check it by using the binomial formula

Recall that Brownian motion has independent and stationary increments (see

! p. 33). Hence

I Then it is immediate that I

Page 55: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

and, again by the independent increments of Brownian motion and since the relation var(X) = E(x-') - holds,

It is a well-known fact that for the N ( 0 , l ) random variable B1, EB: = 3. Hence,

4 ~ ( n i ~ ) ~ = EB,$-~,-, = a [(A,)'/z L I ~ ] = 3 A: ,

which implies that

Thus, if mesh(r,) = max Ai -t 0 ,

i=l, ..., n

we obtain that

Since var(Q,,(t)) = E(Q, , ( t) - t ) 2 we showed:

Qn(t) converges to t in mean square, hence in probability; see Ap- pendix A1 for the different modes of convergence.

The limiting function f ( t ) = t is characteristic only for Brownian motion. It is called the quadratic variation of Brownian motion on [0, t].

One can show that Q,,(t) does not converge for a given Brownian sample path and suitable choices of partitions r,,. This fact clearly indicates that we cannot define J: B,(w) dB,(w) as a Riemann-Stieltjes integral. However,

we could define J: B, dB, as a mean square limit.

Indeed, since S, = 0.5 [B; - Qn(t)] converges in mean square to 0.5 (Bz - t ) , we could take this limit as the value of the integral J; B, dB,. Later it. will turn out that this is the value of the It6 stochastic integral:

2.2. THE ITO INTEGRAL 99

Fro~n this evaluation we have learnt various facts which will be useful in what folluws:

Integrals with respect to Brownian sample paths, which cannot be de- fined in the Rienlann-Stieltjes sense, can hopefully be defined in the mean square sense.

The increment A i B = Bt, - Bt,-, on the interval [ti-1, ti] satisfies E A i B = 0 a.nd E(Ai R)2 = Ai = ti - ti-l. The mean square limit of Q,(t) is t . These properties suggest that (AiB)"s of the order Ai.

In terms of differentials, we write

(dB*)' = (Bttdt - ~ t ) ' = d t , ( 2 . 7 )

and in terms of integrals,

it (dB,)-' = J1 ds = t . 0

(2.8)

The right-hand side is the quadratic variation of Brownian motion on [O, t ] .

At the moment,, relations (2.7) and (2.8) are nothing but hcuristic rulcs. Thcy can be made mathematically correct (in the mean square sense). In this book we do not have the mathematical m%ans to prove them. In what follows, we will use (2.7) and (2.8) as given rules; they will help us to understand the results of It6 calculus, in particular the It6 le~rima i11 Sectiori 2.3.

Next we try to understand why we chose the Riemann-Stieltjes sums S, in (2.4) in such a specific way: the values of Brownian motion were evaluated a t the left end points of the intervals [ t ipl , ti]. Assume that we have a partition r, of [0, t] as in (2.5). We learnt in Example 1.5.7 that the martingale transform - B A B given by

is a martingale with rcspcct to thc filtration a ( B t , , i = I , . . . , k ) , k = 1,. . . , n,. As a result, the mean square limit 0.5 (B; - t ) of the approximating Riemann- Stieltjes sums is a martingale with respect to the natural Brownian filtration.

Page 56: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

100 CHAPTER 2.

Another definition of the Riemann-Stieltjes sums makes them lose the martingale property. As a matter of fact we mention that, for partitions r, = (ti) of [O, t] with mesh(rn) -t 0, t,he Riemann-St,ielt,jes slims

n

5, = 1 B,, A i B i=l

with 1

yi = Z(ti-l + t i ) , t = 1 , . . . , n ,

have the mean square limit 0.5 B; (you can check this by the same arguments and tools as given for S,). This quantity can be interpreted as the value of another stochastic integral, J,' B, o dB, say. The Riemann-Stieltjes sums

~ f = ~ B,, AiB, Ic = 1,. . . , n, do not constitute a martingale, and neither does the limit process 0 . 5 ~ ; (check it!). However, the latter process enjoys another nice property. The relation

indicates that the classical chain rule holds. To be precise, let b(t) be a deterministic differentiable function with b(0) =

0. We know t,hat, 1 db2 (s) -- db(s) = b(s)-. 2 ds ds

Integrating both sides, we obtain by the classical rules of calculus,

If we formally replace the function b(t) with Brownian motion Bt, we obtain the same value as for the stochastic integral (2.9). However, it is only a formal replacement, since this chain rule is applicable to differentiable functions, not to Brownian sample paths.

The stochastic integral, which is obtained as the mean square limit of the Riemann-Stieltjes sums, evaluated a t the middle points of the intervals [ti-l, ti], is called a Stratonovich integral. We will consider it in Section 2.4. It will turn out to be a useful tool for solving It8 stochastic differential equatiu~ls.

Thus we have learnt one more useful fact:

2.2. THE ITO INTEGRAL 101

t The formula So B, dB, = 0.5 (B: - t) suggests that the classical chain rule of integration does not hold for It6 stochastic integrals. In Sec- tion 2.3 we will find a chain rule which is well suited f o ~ It6 i11teglat.io11: the It6 lemma.

1 2.2.2 he It6 Stochastic Integral for Simple Processes We start the investigation of the It6 stochastic integral for a class of processes whose paths assume only a finite number of values. As usual, B = (Bt , t > 0) denotes Brownian motion, and

is the corresponding natural filtration. Recall that a stochastic process X = ( X ; , t _> 0) is adapted to Brownian mot,ion if X is adapt,ed to (Ft, t > 0). This means that, for every t , Xt is a function of the past and present of Brownian motion.

I In what follows, we consider all processes on a fixed interval [0, TI.

First we introduce an appropriate class of It6 integrable processes.

The stochastic process C = (Ci, t E [O, TI) is said to be simple if it sat- isfies t,he following properties:

There exists a partition

7,: O = t O < t l < . . . < t n - l < t n = T ,

and a sequence ( Z i , i = 1, . . . , T I ) of ra~ldom variables such that

ct= { z n , if t = T 3

Z i , if tiPl < t < t i , i = l , . . . , n .

The sequence (Zi) is adapted to (Ftz- , , i = 1, . . . , n), i.e. Zi is a function of Brownian motion up to time ti-l, and satisfies EZ: < w for all i.

Page 57: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

102 C H . 4 P T E R 2. 2.2. T H E ITO INTEGRAL 103

'I

I \ I \:

i I '.'

1-

00 0 2 04 06 0.6 10 0 0 0 2 0 4 0 6 06 1 0 ! I

t Thus the value of the It6 stochastic integral lo C, dB, is the Riemann-Stieltjes Figure 2.2.2 Two approximations of a Brownian sample path b y simple processes surn of the path of C , evaluated at the left end points of the intervals [ti-1, t i ] , C (dashed lines), given b y (2.10). with rcspcct to Brownian motion. If t < t,,, the point t can formally be inter-

preted as the last point of the partition of [0 , t ] .

The It6 stochastic integral of a simple process C on [0 , t ] , tk-l 5 t 5 t k , is given by

L G d B , :=

k - 1 iT ~s ~ ~ ~ , t / ( s ) dBa = ~i A ~ B + Z k (Bt - Btk-,) , (2 .11) i= 1

where c:=, Zi AiB = 0.

is a step function, hence it is a simple process.

Now define the process

for a given partition T", of [O, TI. It is a simple process: the paths are piece- wise constant, and Ct is a function of Brownian motion until time t . For an illustration of the process C for two different partitions, see Figure 2.2.2.

I - ~ h e it6 stochastic integral of a simple process C on 10, TI is given by I Figure 2.2.3 The It6 stochastic integral S; C, d B , corresponding to the paths of C and B given in Figure 2.2.2.

Example 2.2.4 ( C o ~ ~ t i ~ ~ u a t i o ~ ~ of Example 2.2.1) R.ecal1 the sirnple processes f,, and C from Example 2.2.1. The corresponding It6 stochastic integrals are then given by

Page 58: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

104 CHAPTER 2. 1 2.2. THE ITC) INTEGRAL

for t E [(k - l ) / n , k /n] , and by I (C) is adapted to (F+) .

for t E [ t k - 1 , t k ] . See Figures 2.2.2 and 2.2.3 for a visualization of the sample paths of B, C and the corresponding It6 stochastic integrals.

We know from Example 2.1.4 that

where the right-hand side is a Riemann-Stieltjes integral wit,h respect t,o a given I3rownian sa~r~p le palli. We also learnt in Section 2.1.2 that the Riemann- Stieltjes sums (2.12) do not in general converge, as mesh(rn) + 0, for a, given sample path of Brownian motion.

The form of the It6 stochastic integral for simple processes very much reminds u_s of a martingale transform. On p. 83 we introduced the martingale transform C Y , wherc

Here Y = (Y,) is a martingale difference sequence with respect to a given filtration and (? = (e,,) is a previsible sequence. In this sense, the sequence of It6 stochastic integrals

of a simple process (C,, s 5 t ) is a martingale transform with respect to the filtration (Ft, , lc = 0, . . . , n).

But even more is true:

The ~t~ochastic process I t(C) = J: C, dB,, t E [0, TI, is a martingale with respect to the natural Brownian filtration (Ft , t E [0, TI).

We check the defining properties of a martingale; scc p. 80. For convenience we recall them in terms of I (C):

E(It (C) 1 F,) = I, ( C ) for s < t

The first property follows, for example, from the isometry property (2.14) given below.

Adaptedness of I (C) to Brownian motion is easily seen: a t time t , the random variables Z1, . . . , Zk and Al B, . . . , Ak-l B , Bt - Bt,-, , occurring in the defining relation (2.11), are all functions of Brownian motion up to time t .

It remains to show the crucial relation (2.13). This is a good exercise in using the rules of conditional expectation, which we suggest you try for yourself. We have not indicated which rules to use as we assume that you k~iow tl1e111 by 11ow.

First assume that s < t and s , t E [ t k - 1 , t k ] . Notice that

I f ( C ) = Itk-1 ( C ) f Zk ( B s - B t k - l ) + Zk ( B t - B S )

= Is(C) +Zk(Bt - Bs) ,

where I,(C) and Zk are functions of Brownian motion up to time s, and Bt -Bs is independent of F,. Hence,

This proves (2.13) in this case. The case, when s E [ t r P l , tr] and t E [tk-l , t k ] for some 1 < k, can be han-

dled analogously. Check (2.13); make use of the decomposition

= Is(C) + R(s, t ) ,

and show that E(R(s , t ) I F,) = 0.

Properties of thc It6 Stochastic Integral for Simple Processes

It is easy to see that EIt(C) = 0 since 2, and h , B in the definition (2.11) arc independent, and E(Zi AiB) = EZi EAiB = 0. Thus:

Page 59: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

CHAPTER 2. 7 2.2. THE ITO INTEGRAL

The It6 stochastic integral has expectation zero. The It6 stochastic integral is linear: For constants cl, cz and simple processes C(' ) and C(2) on [O, TI,

t t t 1 [cl c!') + c2 c :~)] dB, = cl 1 c:') dB, + c2 1 c!~) dB,. 0

The It6 stochastic integral is linear on adjacent intervals: For 0 5 t 5 T ,

T t T 1 ~ d ~ , = l C.dB, + CsdBs.

Alternatively, we can argue as follows: since I (C) is a martingale, it has a constant expectation function. Moreover, by definition, Io(C) = 0. and so EIt(C) = EIo(C) = 0.

Another property of the It6 stochastic integral for simple processes will turn out to be crucial for the definition of the general It6 stochastic integral.

The proof of the linearity is not difficult; try it yourself. Before you start you We show this property. For the ease of presentation, we assume that t = t k for have to define C ( l ) and C(') on the same partition. This is always possible: if some k. Indeed, i f tk -1 < t < t k , observe that O = to < . - . < tk- l < t i := t is T;') is the partition corresponding to C(') and T$?) the partition corresponding a partition of [0, t ] , so you can formally consider t as a point of the partition. to C ( q , you get a joint partition T with at most n + m distinct points by taking We have, for Wi = Zi AiB, the union of T:) and 72 ' . Clearly, this is a refinement of the two original

k k partitions. The values of the I ( C ( ~ ) ) S remain the same with this new partition.

E[lt(C)12 = =C~(J+'iwj). Linearity follows from the linearity of the underlying Riemann-Stieltjes sums.

i=l j=1 The linearity on adjacent intervals follows from linearity since

T T You can check thal, TOI- i > j , the ra~ldo~rl variables Wi and W j are uncor- C,dB, = 1 [c!') + cj2)] dB,, related: notice that W j and Zi are functions of Brownian motion up to time t iwl , hence they are independent of A, B. We conclude that

where c!" = C,I,o,tl ( s ) and cL2' = CsI(t,T] ( s ) are two simple processes.

E(WiWj) = E(WjZi) E(A iB) = 0 . Finally, we state the following property:

Thus the terms E(WiWj) in (2.15) vanish Tor i # j , and so we obtain

k k k This follows from the definition of I (C) : the relation E [ I ~ ( c ) ] ~ = C E ( Z ~ A ~ B ) ~ = C EZ; E ( A ~ B ) ~ = C EZ; (ti - t i P l ) .

i=l i=l i=l I+(C) = I t ; , (C) +Zi(Bt -Bt i ti-I 5 t < t i ,

The right-hand side is nothing but the Riemann integral Sot f ( s ) ds of the step holds and the sample paths of Brownian motion are continuous. function f ( s ) = EC;, which coincides with Ez: for ti-l 5 t < ti (check this!). Thus we have proved the isometry property of the It6 stochastic integral. 2.2.3 The General It6 Stochastic Integral

Thc It8 stochastic intcgral has several properties in common with the Riemann and Riemann-Stieltjes integrals. In the previous section we introduced the Ito stochastic integral for simple

processes C , i.e. for stochastic processes whose sample paths are step functions. t

The It6 stochastic integral So C, dB, is then simply the Riemann-Stieltjes sum

The process I (C) has continuous sample paths.

Page 60: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

CHAPTER 2. 2.2. THE ITO INTEGRAL

of C , evaluated at the left end points of the intervals [ti-l, ti], with respect to Brownian motion. We learnt in Section 2.1.2 that, in general, we cannot define the It6 stochastic integral as a pathwise limit of Riemann-Stieltjes sums. In Section 2.2.1 we suggested that we define the It6 stochastic integral as the mean square limit of suitable Riemann-Stieltjes sums. This idea works under quite general conditions, and it is the aim of this section to advocate this approach. At a certain point we will need some tools from Hilbert space theory, which we do not require as common knowledge. Then we will depend on some heuristic arguments. However, the interested reader can find a proof of the cxistence of the general It6 stochastic integral, using tools from measure theory and functional analysis, in Appendix A4.

In what follows, the process C = (Ct, t E [0, TI) serves as the integrand of the It6 stochastic integral. We suppose that the following conditions are satisfied:

Assumptions on the Integrand Process C:

C is adapted to Brownian motion on [0, TI, i.e. Ct is a function of B,, s 5 t .

The integral JbT EC: ds is finite.

Notice that the Assumptions are trivially satisfied for a simple process; cf. p. 101. (Verify them!) Another class of admissible integrands consists of the

deterministic functions c(t) on [0, T] with J: c2(t) dt < oo. It includes the continuous functions on [O, TI.

In what follows, we give a heuristic approach to the It6 stochastic integral. First recall how we proceeded for the definition of the integral J: B, dB, in Section 2.2.1:

For fixed t and a given partition (ti) of [0, t] we introduced the Riemann- Stieltjes sums C&, Bt;-l (Bt; - Bta-,) .

Notice: in Section 2.2.2 we defined exactly these Riemann-St,ielt,jes slims as the It6 stochastic integrals I ( c ( ~ ) ) = J; tin) dBs of the simple func- tions C(n) which assume the value Bt,-, on the interval [ti-1, t i ) .

Let C be a process satisfying the Assumptions. Then one can find a sequence (C(n) ) of simple processes such that

iT E[Cs - c:")]' ds + 0 .

The roof of this property is beyond the scope of this book; see Appendix A4 for a reference.

Thus the simple processes C(n) converge in a certain mean square sense to the integrand process C. Since dn) is simple we can evaluate the It6 stochastic integrals I t ( O n ) ) = Jt c:") dB, for every n and t.

The next step is to show that the sequence (I(C("))) of It6 stochastic integrals converges in a certain mean square sense t,n a, unique limit process. Indeed, one can show the existence of a process I ( C ) on [0, TI such that

E sup [L(c) - O<t<T

See Appendix A4 for a proof.

The mean square liniit I ( C ) is called the It6 stochastic integral of C . It is denoted by

t

b ( c ) = l CsdB., t € [ O , T ] .

For a simple process C , the It6 stochastic integral has the Riemann- Stieltjes sum representation (2.11).

After this general definition we do not feel very comfortable with the notion of It6 stochastic integral. We have lost the intuition because we are not able to write the integral S: C, dB, in simple terms of Brownian motion. In particular cases we will be able to obtain explicit formulae for It6 stochastic integrals, but this requires knowledge of the It6 lemma; see Section 2.3. For our general intuition, and for practical purposes, thc following rule of thumb is helpful:

Then we considered the mean square limit of the Riemann-Stieltjes sums I(c(")) and obtained the value 0.5 (B; - t).

This approach can be made to work for processes C satisfying the Assumptions. We commence with an interesting observation:

I

Page 61: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

CHAPTER 2.

Figure 2.2.5 1st and 3rd rows: approximations to a Brownian sample path by simple processes c ( ~ ) (dashed lines) assuming n = 4,10,20,40 distinct values. Below every figure yuufind l l ~ e path of the corresponding It6 stochastic integral I ( C ( " ) ) . Thc lattcr processes approximate the process It ( B ) = Ji B, dB, = 0.5 (B: - t ) . Compare with Figure 2.3.1, where the corresponding sample path of I ( B ) is drawn.

! 2.2. T H E l ~ h 1NTEGR.AL

The It8 stochastic integrals I t ( C ) = Ji C, dB,, t E [O,T], constitute a stochastic process. For a given partition

7,: O = t 0 < t 1 <. . .< tn- l < t , = T ,

and t E [tk-1,tk], the random variable It(C) is "close" to the Riemann- Stieltjes sum

k-l

C cti-1 (Bt, - Bt,-1) + Ct,-, (Bt - Btk-, 1 , i=l

and this approximation is the closer (in the mean square sense) to the value of It (C) the n o r e dense the par Lition T~ ill [0, TI.

Properties of the General It6 Stochastic Integral

The general I t8 stochastic integral inherits the properties of the It6 stochastic integral for simple processes, see Section 2.2.2. In general, we will not be able to prove these properties. However, some proofs can be found in Appendix A4.

The stochastic process It(C) = J: C8 dB8, t E [0, TI, is a martingale with respect to the natural Brownian filtration (Ft, t E [0, TI).

/ This results from the particular choice of the approximating Riemann Stieltjes sums of C, which are evaluated a t the left end points of the intervals [ti-*, ti].

The It8 stochastic integral has expectation zero.

For the proof of the existence of the I t8 stochastic integral the isometry prop- erty (2.14) for simple processes is essential. I t remains valid in the general case.

E ([ C, dB,) = EC: ds , t E [O, T ] .

The It8 stochastic integral also has some properties in common with the Rie- mann and Riemann-Stieltjes integrals.

Page 62: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

The Ito stochastic integral is linear:

For constants c l , c.2 and processes C ( l ) and c ( ~ ) on [0, TI, satisfying the L4ssumptions,

I t

[cl c!" + Q CL2) d B s = cl it c:') d B s + c2 c :~) d B s . 0

The It6 stochastic integral is linear on adjacent intervals:

For 0 5 t 5 T ,

T t T /d Cs dBs = 1 Cs d B , + 1 Cs d B s .

Finally, we state the following property:

/ The process I ( C ) has continuous sample paths. 1 Notes and Comments

The definition of the It6 stochastic integral goes back to It6 (1942,1944) who introduced the stochastic integral with a random integrand. Doob (1953) re- alized the connection of I t6 integration and martingale theory. It6 integration is by now part of advanced textbooks on probability theory and stochastic processes. Standard references are Chung and Williams (1990), Ikeda and Watanabe (1989), Karatzas and Shreve (1988), Bksendahl (1985) and Prot- ter (1992). Nowadays, some texts on finance also contain a survival kit on stochastic integration; see for example Lamberton and Lapeyre (1996).

Some of the aforementioned books define t,he stochastic integral with re- spect to processes more general than Brownian motion, including processes

T with jumps. Moreover, the assumption So E C ; ds < oo for the existence of

Jof C s dBS can be substantially relaxed.

2.3 The It6 Lemma

In the last two sections we learnt about the It6 stochastic integral. Now we know how the integral Sof C s dB"s defined, but with the exception of simple processes C we do not have the tools to calculate It8 stochastic integrals and to proceed with some sitnple operations on them. I t is the objective of this section to provide such a tool, the I t8 lemma.

2.3.1 The Classical Chain Rule of Differentiation

The It8 lemma is the stochastic analogue of the classical chain rule of differ- entiation. We mentioned the chain rule on p. 100 in a particular case. There we recalled that , for a differentiable function b ( s ) ,

This relation implies that, with b(0) = 0 ,

We cannot simply replace in (2.16) the deterministic function b(s ) with a sam- ple path B s ( w ) of Brownian motion. Indeed, such a path is nowhere differen- tiable. Using some elementary means, we discovered in Section 2.2.1 t h a t the I t6 stochastic integral J: B, d B , has value 0.5 (Bf2 - t ) . This is in contrast to (2.16). Notice that the integral value 0.5 (B: - t ) is the value 0.5 B:, which is to be expected from the classical chain r~ i l e of differentiat,ion, corrected by -0.5 t . This suggests that we have to find a correction term to the classical chain rule. Another appeal to the discussion in Section 2.2.1 tells us that this term comes from the mean square limit of the quadratic functional C:=l ( B t , - B, , -1 )2 .

In order to understand the stochastic chain rule, we first recall the deter- ministic chain rule. For simplicity, we write h l ( t ) , h l1( t ) , etc., for the ordinary derivatives of the function h a t t .

Let f and g be differentiable functions. Then the classical chain rule of differentiation holds:

[ f ( g ( s ) ) l l = f 1 ( 9 ( s ) ) g l ( s ) . (2.17) I Because we are interested in integrals, we rewrite (2.17) in integral form, i.e. we i ~ ~ t e g r a t e both sides of (2.17) or1 [0, t ] .

The chain rule i n integral form :

Page 63: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

We give a heuristic argument for the validity of the chain rule, which we will t,ake as a motivation for the I t6 lemma. 111 the language of differentials, (2 .17) becomes df ( g ) = f' dg. This differential can he interpreted as the first order term in a Taylor expansion:

Here d g ( t ) = g ( t + d t ) - g ( t ) is the increment of g on [ t , t + d t ] . Under suitable conditions, the second and higher order terms in this expansion are negligible for small d t .

2.3.2 A Simple Version of the It6 Lemma

Now assume that f is a twice differentiable function, but replace g ( t ) in (2.19) with a sample path B t ( w ) of Brownian motion. Write d B t = Bt+dt - Bt for the increment of B on [ t , t + d t ] . Using the same arguments as above, we obtain

On p. 99 we gave some motivation that the squared differential ( d B t ) 2 can be interpreted as dt . Thus,

in contrast t o the determinzstic case, the contribution of the second order t e r m i n the Taylor expansion (2.20) is no t negligible.

This fact is the reason for the deviation from the classical chain rule. Integrating both sides of (2.20) in a formal sense and neglecting terms of

order I~igher Lhan 3 o11 the right-hand side, we o b t a i ~ ~ for s < t ,

The following question naturally arises:

How do we have to interpret the integrals i n (2 .21) and (2.22) ?

2.3. TIIE ITO LEMMA 115

therefore it is natural to assign the value f ( B t ) - f (B,) to it. In what follows, t we always give the value l/t - V, to the symbolic integral S, d V x , whatever the

I stochastic process V.

Let f be twice continuously differentiable.

The formula

t t

f ( B t ) - f ( B s ) = J f 1 ( B , ) d B z +: J f " ( B x ) h , < t , (2.23)

is a simple form of the It6 l e m m a or of the It6 formula.

Using sophisticated methods, which are beyond the scope of this book, one can show that (2.23) is mathematically correct.

Now we want to see the It8 lemma at work.

Figure 2.3.1 One sample path of the process /: B , d B , = 0.5 ( B ~ ~ - t ) . See also Figure 2.2.5.

Example 2.3.2 Choose f ( t ) = t 2 . Then f r ( t ) = 2t and f r ' ( t ) = 2. Hence the It6 lemma yields for s < t ,

The first integral in (2 .22) is the It6 stochastic integral of f l ( B ) , and the second integral has to be interpreted as the Riemann integral o f f "(B). Relation (2.21) defines the quantity J~~ df (B,) which is an analogue of a telescoping sum, and

Page 64: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

116 CHAPTER 2. 2.3. THE ITO LEA4AfA 117

For s = 0 we obtain the formula 2.3.3 Extended Versions of the It6 Lemma

In this section we extend the simple It6 lemma, given in (2.21)-(2.22), in

which is already familiar to us; cf. (2.6) or the discussion in Section 2.3.1.

We try to get a similar formula for Bt3 - B:. In this case, f (t) = t3, f l ( t ) = 3t2 that f ( t ,x) has continuous partial derivatives of a t least second order. A

and f"(t) = G t . The It8 lemma immediately yields modification of the Taylor expansion argument used in Section 2.3.2 is also successful in this more general case.

B:-B: = 3 1 t B : d B , + 3 1 t B , d 3 . Recall that a second order Taylor expansion yields that,

t f (t + dt, Bt+dt) - f ( t , Bt) = (2.25)

We cannot express Js B, dx in simpler terms of Brownian motion. In order t to get an impression of the kind of sample paths of the processes Ss B, dx or f1 ( t ,B t )d t + f2(t ,Bt)dBt

Ji B: dB, one has to rely on simulations.

Now exercise the I t8 lemma yourself: calculate f ( B t ) - f(B,) for a power function f (t) = tn for some integer n > 3.

Example 2.3.3 (The exponential function is not the It6 exponential) Here, and in what follows, we use the following notation for the partial deriva- In classical calculus, the exponential function f (t) = exp{t) has the spectacular property that f r ( t ) = f (t). Equivalently,

t

f ( t ) - f(.) = l f ( x ) d z .

Is there a stochastic process X such that , i , j = 1 , 2 .

As in classical calculus, higher order terms in (2.25) are negligible, and so are

. the terms with factors dt dBt and (dt)2. However, since we interpret in the It6 sense? This would be a candidate for the It6 exponential. Since the as dt, the term with (dBt)2 cannot be neglected. Arguing in the same way as classical rules of integration fail in general when Brownian motion is involved, in Section 2.3.2, i.e. formally integrating the left-hand and right-hand sides in we may also expect that f (Bt) = exp{Bt) i shot the It6 exponential. This can (2.25) and collecting all terms with dt and dBt separately, we end up with the be checked quite easily: since f (t) = f t ( t ) = f"(t) the It6 lemma yields for following formula: s < t ,

t eBi - eBs - 1

- eBz dB, + eBa d x .

Obviously, the second, Riemann, integral yields a positive value, and so (2.24) cannot be satisfied.

At the moment we cannot give an answer to the above question. We will return to this problem when we have a more advanced form of the It6 lemma; see Example 2.3.5..

Extension I of the It6 Lemma: Let f (t, x) be a function whose second order partial derivatives are con- tinuous. Then

f (t, Bt) - f (s, Bs) = (2.25)

1 l t . [ f 1 ( ~ , ~ , ) + I fxz(x,Bx) dx + f ~ ( x , B , ) d B , , s < t . 1 6 '

Page 65: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

118 CHAPTER 2. 2.3. T H E ITO LEhIhL4

where c and u > 0 are constants. Notice that,

f ( t : x ) = e(c-0.50Z) t + o ~ f l ( t , x ) = ( c - O . 5 ~ " f ( t 3 ~ ) ,

5 2 ( t , 5 ) = o f ( t , x ) , f 2 2 ( t r x ) = u 2 f ( t , x ) .

An application of the It6 lemma (2.26) yields that the process X satisfies the linear stochastic differential equation

t t

st - X0 = c 1 E;,ds + CJ 1 *.,dB,. (2 .28)

For later use we will need an even lrlore ge~leral versior~ of the It6 lemma. We will consider processes of the form f ( t , X t ) , where X is given by

Figure 2.3.4 One sample path of the exponential of Brownian motion exp{Bt) (left) and of the It6 exponential exp{Bt - 0 .5 t ) (right) for the same ~ a t h of B ; cf. Exam- ples 2.3.3 and 2.3.5.

and both: .A(') and A ( ~ ) , are adapted to Brownian motion. Here it is assumed

We apply this formula to find the It6 exponential: that tile above iritegrals are well defined in the Riemann and It6 senses, re-

Example 2.3.5 (The It6 exponential) We learnt in Example 2.3.3 that the stochastic process exp{Bt) is not the It6 exponential in the sense of (2 .24) . Now we choose the function

f (t, x ) = .

Then direct calculation shows that

1 I , ) = - ( 2 , ) = f f 2 2 ( t l x ) -- f ( t , x ) . Recall that the g e v ~ ~ ~ e t r i c Brownian motion (2.27) satisfies (2 .28) . Hence it is

an I t8 process with A(') = c X and A(2) = OX. An application of the above It6 lemma gives Now, using a similar argument with a Taylor expansion as above, one car1

show the following formula.

In the sense of (2 .24) , f ( t , Bt) is the It6 exponential. See Figure 2.3.4 for a comparison of the paths of the processes exp{Bt) and exp{Bt - 0.5 t ) .

Example 2.3.6 (Geometric Brownian motion) Consider a particular form of geometric Brownian motion (cf. Example 1.3.8):

Xt = f ( t , B t ) = e (c-0.50Z) t + u Bt 1

Page 66: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

120 CIL4PTER 2. 2.3. THE IT^ LEMMA 121

Extension I1 of the It6 Lemma: ' A Taylor series expansion argument similar to the one above yields the follow-

Let X be an It,o process with representation (2.29) and f (t, r) be a function whose second order partial derivatives are continuous. 'l'heri

f ( t , X , ) - ~ ( S , ~ Y ~ ) =

1 J t [fl(y. s,) + A? f2(yr + i. [.4;)i2n2(Y. x y ) ] dy s

t

+ " : ) f 2 ( ~ , ~ . y ) d ~ y , s < t .

Give a justification of this formula. Use a Taylor expansion for f ( t+dt, X t + d t ) -

f ( t , X t ) as in (2.25), where B is replaced with X: and X has representation (2.29). Neglect higher order terms, starting with terms involving (dt)' and dt dBt , and make use of (dBt)" ddt.

Formula (2.30) is frequently given ~ I I the followir~g for111:

f ( t , X t ) - f(s,-Z=s)

The integrals with respect to x:) have to be interpreted in the same way as in (2.32). Formula (2.34) can be extended in the straightforward way to

where dayy = -4;) dy + A?) dB, functions f ( t , x:'), . . . , xjrn)) , where the X( i ) s are It8 processes with respect

to the same Brownian.motion. Wc mention that one can also consider such a formula for I t6 processes with respect to different Brownian motions. However,

The latter identity is a symbolic way of writing the I t8 representation (2.29). this requires that we define the I t8 stochastic integral for such a setting. The integral with respect to X in (2.31) has to be interpreted as follows: The following example is an application of the I t8 lemma (2.34).

Example 2.3.7 (Integration by parts) Consider the function f (t, X I , z2) = ~ 1 x 2 . Then we obtain (we suppress thc arguments of the functions)

Finally, we consider the It6 lemma for stochastic processes of the form f (t , x;", x j 2 ' ) , where both, x(') and x ( ~ ) , are It6 processes with respect to fi = O r f 2 = ~ 2 , f3 = 2 1 , f22 = f33 = 0 and f23 = f32 = 1.

the same Brownian motion: Now apply formula (2.34) to obtain:

Extension I11 of the It6 Lemma: Let x(') and x(') be two I t8 processes given by (2.33) and f (t , X I , x2 ) be a function whose second order partial derivatives are continuous. Then for s < t,

f ( t , x,('), x,(~)) - f ( s , xi1), ~ 6 ~ ) ) (2.34)

t

= ~ ( Y . X ~ ' ) , X ~ ) ) dy

+ 2 [ fi(y, X;') , x")) Y dX(') Y

,=2 *

3 3

+ ' c c I t f i j(yl x('), Y Y Y Y ~ ( 2 ) ) ~ ( 2 , i ) ~ ( 2 , j ) d y . i=2 j = 2 8

Here fi(t, X I , x2), f i j( tr X I , xz) are the partial derivatives of f ( t , x l , x2 ) with respect to the i th, the i th and j t h variables, respectively; cf. p. 117.

Page 67: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

122 CHAPTER 2.

The Integration by Parts Formula: Let x i 1 ) and x i 2 ) be two It8 processes given by (2.33). Then

Using the It8 representation (2.33), we obtain an alternative expression for the differentials:

(1) (2) d ( x t X, ) = (X,(l)Ai'32) + X,(2)Afl , l ) + Ai2,1)Af2t2) dt

+ ( x , ( ~ ) A ~ ~ ' ~ ) + x ~ ( ~ ) A / ~ ' ~ ) ) dBt .

As particular examples we consider

Obviously,

*jl.') = et Ai2,') = , 0 , = 0 , ~ i ~ , ~ ) = 1

Hence, integration by parts yields

More generally, show that for any continuously differentiable function f on [0, TI the following relation holds:

Notes and Comments

The It6 lemma is the most important tool in It8 calculus. We will use it very often in the following sections. A first version of this fundamental result was proved by It8 (1951). Various versions of the It8 lemma and their proofs can be found in textbooks on stochastic calculus, for exa~r~ple, ~ I I C1lu11g and Williains (1990), Ikeda and Watanabe (1989), Karatzas and Shreve (1988), Bksendahl (1985) or Protter (1992).

2.4. T H E STRATONOVICH AND OTHER INTEGRALS 123

2.4 The Stratonovich and Other Integrals

In this section we discuss some other stochastic integrals and their relation with the It8 stochastic integral. It is certainly useful for you to realize that there is a large variety of other integrals, the It6 integral being just one member of this family. On the other hand, you do not need the information of this section (which is also slightly technical) unless you are interested in solving It8 stochastic differential equations by the so-called Stratonovich calculus.

In the previous sections we studied the It8 stochastic integrals

Here and in what follows, B = (Bt , t 2 O) is Brownian motion and C = (Ct , t E [0, TI) is a process adapted to the natural Brownian filtration Ft = u(B,, s 5 t) , t E [0, TI. The main point in the definition of the It8 stochastic integral was the approximation of I t (C) by Riemann-Stieltjes sums of the form

for partitions T,: O = t o < t i < ... < t n - i < t , = T

such that mesh(r,) = , max (ti - ti-1) 4 0

t = l , ..., n

As usual, we write AiB = Bt. - Bt, , . In the Riemann-Stieltjes sums (2.36) we chose the values of C at the left

end points of the subintervals [ti-', ti]. This choice was made for mathematical convenience. Our gain was that the stochastic process (It(C), t E [0, TI) has a rich mathematical structure. It inherits the martingale property from the approximating Riemann-Stieltjes sums (2.36). Thus a high level of the theory of stochastic processes can bc applied. As subsidiary properties we get that the expectation of the It8 stochastic integral is always zero and its variance can be expressed by an application of the isometry property. The price one has to pay for the nice mathematical structure is that the chain rule of classical calculus is not valid anymore. In It8 calculus, the It8 lemma replaces the classical chain rule.

In this section we consider another type of stochastic integral, the Stra- tonovich stochastic integral. We start with its definition in the case that the

Page 68: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

124 CHAPTER 2.

integrand process C is given by

for a twice differentiable function f on [O,T]. D e h e the Riemann-Stieltjes sums

where

One can show that the mean square limit of the Riemann-Stieltjes sums gn exists if mesh(rn) + 0.

. ..

mean square limit ST(^ (B)) of the Riemann-Stieltjes sums gn exists if J~~ Ef 2(Bt) dt < cm. The limit is called the Stratonovich stochastic integral off (B). I t is denoted by

T

Clearly, as for the l t6 stochastic integral, we can define the stochastic process of the Stratonovich stochastic integrals

as the mean square limit of the corresponding Riemann-Stieltjes sums.

Example 2.4.1 On p. 100 we considered the Riemann-Stieltjes sums

for a given partition rn of [0, TI. We also mentioned that it is not difficult t o verify that the sequence (gn) has the mean square limit 0.5 B;. This is the value of the corresponding Stratonovich stochastic integral:

2.4. THE STRilTONOVICH AND OTHER INTEGRALS

Figure 2.4.2 One sample path of the process Si B: dBs (lower curve) and of the pTOCeSS J,' B.: o dB, (upper curve) for the Sam? Brownzan sample path.

Obviously, the stochastic process (0.5 B:, t E [0, TI) is not a martingale. How- ever, we see that the Stratonovich stochastic integral (2.38) formally satisfies the classical chain rule; see p. 100 for a discussion.

The formal validity of the classical chain rule is the reason for the use of Strato-

I ~iovich st,ochastic integrals. Thus, despite the "poor" mathematical structure of the Stratonovich stochastic' integral, it has this lice" property. \Ve will exploit it when we solve It6 stochastic differential equations.

To get sonie feeling for the Strato~lovich stochastic integral, we consider a transformation formula wliich links the It6 and the corresponding Stratonovich st,ochast,ic irlt,egrals with the sarrle integrand process f ( B ) . We assume

~ ~ E [ f ( B ~ ) ] ~ d t < cm and lT E[f '(Bt)12 dt < cm . (2.38)

The first condition is needed for the definition of the Stratonovich stochastic integral ST(f (B) ) and the It6 stochastic integral 1 ~ ( f (B)) .

First observe that the Taylor expansion

holds, where we neglect higher order terms. Thus an appoxiniat,ilig Rieniann-

Page 69: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Assume the function f satisfies (2.39). Then the transformation formula holds:

T T iT f (4) o dBt = l f ( B t ) dBt + i f l ( B t ) dt . (2.41)

The Stratonovich stochastic integral satisfies the chain rule of classical calculus:

JdT ~ I ( B , ) 0 = g ( a ) - ~ ( B o ) . (2.42)

126 C H A P T E R 2. 2.4. THE STRATONOVICH AND O T H E R INTEGRALS

Stieltjes sum for ST(^ ( B ) ) can be written as follows: = LT f ( B s ) d ~ s + 2 f l ( B s ) d s ,

2 f(ayi a i ~ iT

and the right-hand side is equal to J:~'(B,) o dB,. And so we have: i=l

n n

= C f ( ~ t ; - ~ ) A,B + C f ' ( ~ t ~ - ~ ) (B,, - B t i l ) A i B + . . . i=l i= 1

n n

= C f ( B t ; - , ) a i B + C ~ ' ( B ~ , - ~ ) ( B , , - B t i - I ) 2 i=l i=l

V L Notice: this statement does not mean that the Stratonovich stochastic integral

+ C f l ( B t i - , ) ( B u t - B t i - l ) ( B t i - B y , ) + . . . is a classical (i.e. Riemann) integral. We only clainl that the correspo~ldirlg i=l chain rules have a similar structure.

- - 2;' +gh2) +gA3) + . . . , (2.40) Example 2.4.3 a) Take g( t ) = t2 . Then g l ( t ) = 2t. We obtain from (2.42) that

where we again neglected higher order terms. By definition of the I t6 integral, T 32' has the mean square limit So f (B, ) dB,. An application of condition LT 1 1 1

B~ O ~ B ~ = - B+ - - B~ - - B ~ 2 2 0 - 2 T '

(2.39) shows that 3;) has the mean square limit zero, and some calculations give that This is in agreement with our previous observations about the value of S T ( B ) .

T

SF'+ ; Jd f l ( B t ) d t b) Take g ( t ) = exp{t). Then g l ( t ) = g ( t ) . We obtain from (2.42) that

JdT eBt o iB t = eBT - eBo = e B ~ - 1 . in mean square. Combining the mean square convergences of st', i = 1,2 ,3 , we finally obtain the following result from (2.40):

We conclude from the latter relation that the process Xt = exp{Bt), t E [O, T I , is the Stratonovich exponentzal. Also recall from Example 2.3.5 that the It6 exponential is a totally different process.

In what follows, we want to give a transformation formula for the Stratonovich stochastic integral which is more general than (2.41). There we only addressed integrand processes C = f ( B ) . Now we assume that the integrand is of the fur111

FI-0111 this lurrr~ula it is irr~~rlecliate that ( S t ( j ( B ) ) , t E [ O , TI) dues 11ot curlstitute c t = f ( t , x t ) , t E [ O , T ] , (2.43)

a martingale. Check this by taking expectations in (2.41). where f (t , x) is a function with continuous partial derivatives of order two. The Now take the particular function f ( t ) = g l ( t ) . An application of the I t8 process X is supposed to be an It6 process (see p. 119) given by the stochastic

lemma (2.23) to Yt = g ( B t ) yields differential equation:

T t t

~ ( B T ) - g(Bo) = X t = X o + Jd a ( s , X , ) d s + 1 b ( s , X s ) d B s ,

Page 70: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

1'28 CHAPTER 2

where the continuous functions a( t , x) and b(t, x) satisfy the existence and uniqueness conditions on p. 138. For integrands (2.43) it is not straightforward how to get arl exter~sior~ of the integral with C = f (B). One possible way to define the Stratonovich stochastic integral is via approximating Riemann- Stieltjes su~lls

T The mean square limit So f (t, X t ) o d B t of these Riemann-Stieltjes sums exists if

One can show that this definition is consistent with the previous one (with f (t , x) = f (x), X = B and the approximating Riemann-Stieltjes sums (2.37)).

For later use, we give here another identity:

Under the assumptions on f and given above, the following transfor- mation formula holds:

1 where, as usual, f 2 ( t , r ) is the partial derivative of f with respect to s. 1 From the above discussion it might have become clear that we can define

very different stochastic integrals. For every p E [0, 11, a given partition T, = ( t i ) of [O, T] and a process C adapted to Brownian motion, we can define the Riemann-Stieltjes sums

where

2.4. THE STRATONOVICH AND OTHER lNTEGRALS 129

If mesh(r,) -t 0, the mean square limit of the Riemann-Stieltjes sums zip) exists, provided C satisfies some furt,her assumptions. The limit can be con- sidered as the (p)-stochastic integral of C:

We studied two particular cases: p = 0 (the I t8 case) and p = 0.5 (the Strato- novich case). For non-trivial integrands C the values I ~ ) ( c ) differ for distinct ps. For example, using arguments similar to the I t8 and Stratonovich cases, one can show that

I t is also possible to get a transformation formula such as (2.44) in order to relate the (p)-integrals, p E [O, 11, to the corresponding It6 or Stratonovich integrals. For applications, the It6 and the Stratonovich integrals are most relevant. The reasons were explained above. See also Chapter 3 on stochastic differential equations.

Notes and Comments

The Stratonovich stochastic integral was introduced by Fisk, and indepen- dently by Stratonovich (1966). The mathematical theory of the Stratonovich integral can be found in Stratonovich (1966); see also Arnold (1973).

Both, the It6 and Stratonovich integrals, are defined in a mathematically correct, way. In applications one has to make a derision ahout which stochastic integral is appropriate. This is then a question of modelling; see also the discussion on p. 150.

In Section 3.2.3 we will use the rules of Stratonovich calculus (i.e. the rules of classical calculus) for the solution of Ito stochastic differential equations.

Page 71: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Stochastic Differential Equations

This chapter is devoted to stochastic differential equations and their soli~t,ion. In Section 3.1 we start with a short introduction to ordinary differential equa- tions. Stochastic differential equations can be understood as deterministic differential equations which are perturbed by random noise.

In Section 3.2.1 we introduce It6 stochastic differential equations and ex- plain what a solution is. It will turn out that stochastic differential equa- tions are actually stochastic integral equations which involve ordinary and It6 stochastic integrals. Therefore the notion of "different,ial equation" might be slightly misleading, but since i t is the general custom to use this term, we will follow it. In Section 3.2.1 we formulate conditions for the existence and uniqueness of solilt,ions t,o Tt,6 st,oc,hastic differential equations. In Section 3.2.2 we give a simple method for solving It6 stochastic differential equations. It is based on the It6 lemma. We continue in Section 3.2.3 with the solution of It6 stochastic differential equations which are derived from an equivalent Stratonovich stochastic differential equation.

In Section 3.3 we consider the general linear stochastic differential equa- tion. It is a special class of stochastic differei~tial equations which have an explicit solution in terms of the coefficient functions and of the underlying Brownian motion. Linear stochastic differential equations are relevant in many applications. We also provide a method for calculating the expectation and variance functions in this case.

As for deterministic differential equations, it is very exceptional to obtain an explicit solution to a stochastic differential equation. In general, one has to

Page 72: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

132 CHAPTER 3.

rely on numerical methods; see Section 3.4 for an introduction to this topic.

3.1 Deterministic Differential Equations

The theory of differential equations is the cradle of classical calculus. Such equations were the motivating examples for the creation of the differential and integral calculus. The idea underlying a differential equation is simple: we are given a functional relationship

involving the time t , an unknown function x ( t ) and its derivatives. It is the aim to find a function x ( t ) which satisfies (3.1). I t is called a solution of the differential equation (3.1). Hopefully, this solution is unique for a given initial condition x (0 ) = xo, say.

The simplest differential equations are those of order 1. They involve only t , x ( t ) and the first derivative x l ( t ) . Ideally, such a differential equation is given in the form

for a known function a(t , x ) . It is standard to write (3.2) in the more intuitive form

dx( t ) = a ( t , x ( t ) ) d t , x (0 ) = x o . (3.3)

If you interpret x ( t ) as the location of a one-dimensional particle in space a t time t , (3.3) describes the change of location of the particle in a small time interval [ t , t + dt] , say. Relation (3.3) then tells us that dx( t ) = x ( t + d t ) - x ( t ) is proportional to the time increment dt with factor u(t , z( t ) ) . Alterr~atively, (3.2) tells us that the velocity x f ( t ) of the particle is a given function of time t and location x ( t ) .

Example 3.1.1 (Some simple differential equations) Assume that the velocity is only a function of t :

Integration on both sides yields the solution

This is the simplest form of a differential equation, but it is also a trivial one because the right-hand side does not depend on the unknown function x ( t ) .

Now assume that the velocity x t ( t ) is proportional to the location x ( t ) : \,

for some constant c. Here simple integration does not help. But we know the solution to this differential equation: it is the exponential function x ( t ) = z ( 0 ) expic t ) .

With an elementary trick one can sometimes solve the differential equation (3.2).

Example 3.1.2 (Separation of variables) Suppose that the right-hand side of (3.2) can be separated into a product of two functions:

x t ( t ) = a1 ( t ) aa(x ( t ) ) .

Symbolically, we rewrite this differential equation as

dx -- - a1 ( t ) dt a2 (2)

Now integrate both sides:

011 the left-hand side we obtain a function of x ( t ) , on the right-hand side a function of t . Hopefully, we obtain an explicit 1or111 of the function x ( t ) .

This approach is justified by differentiating both sides of (3.5) as integrals of the upper limits. Then we obtain

x' ( t )

a2 (x ( t ) ) = a l ( t ) ,

which is another form of writing (3.4).

We apply this method to the differential equation x t ( t ) = c x ( t ) for some x (0 ) # 0. Then

giving x ( t ) = x (0 ) exp{ct). This is the solution we guessed in Example 3.1.1.

Page 73: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

134 C H A P T E R 3.

From the above discussion we learnt about some of the important aspects of ordinary differential equations:

I I

Sokuliur~s of ckiflereratial equations are junctions. Th-y describe the evolution or dynamics of a real-life process over a given period of time.

In order to obtain a unique solution, one has to know the initial condition x(0) = xo. If this unique solution x(t) starts from a point xo a t the present t = 0, say, the function x(t) is completely determined in the future, i.e. for t > 0.

Explicit solutions of different,ial eq~iat,ions are t,he exception from the rule. In general, one has to rely on numerical solutions to differential equations.

Integrating both sides of the differential equation (3.2), one obtains an equivalent integral equation:

Although this transformed equation is in general not of great use for finding the solution of (3.4), it gives an idea of how we could define a stochastic differential equation: as a stochastic integral equation.

3.2 It6 Stochastic Differential Equations

3.2.1 What is a Stochastic Differential Equation?

Consider the deterministic differential equation

The easiest way to introduce randomness in this equation is to randomize the initial condition. The solution x(t) then becomes a stochastic process (Xt. t E [O,T]):

d X t = a ( t , X t ) d t , X o ( w ) = Y ( w ) .

i 3.2. ITO STOCHASTIC DIFFERENTI.4L EQUATIONS

Figure 3.2.1 10 solutions Xt i- Xoet of the random differential equation dXt = Xt dt with initial condztion X o = exp{N), where N has an N ( 0 , crZ) distribzltion. Left: a2 = 0.01. Right: a2 = 0.0001.

Such an equation is called a random differential equation. Its solution does not require stochastic calculus; we can use the classical methods and adjust the solution to the corresponding outcome of the initial condition. Random differential equations can be considered as deterministic differential equations with a perturbed initial condition. Their investigation can be of interest if one wants to study the robustness of the solution to a differential equation under a small change of the initial condition. For example, Figure 3.2.1 shows that the solution of a differential equation can change quite drastically, even if the change of the initial condition is small.

For our purposes, the randomness in the differential equation is introduced via an additional random noise term:

Here, as usual, B = (Bt , t > 0) denotes Brownian motion, and a( t ,x) and b(t,x) are deterministic functions. The solution XI if it exists, is then a stochastic process. The randomness of X = (Xt, t E [0, TI) results, on the one hand, from the initial condition, and on the other hand, from the noise generated by Brownian motion.

A naive interpretation of (3.6) tells us that the change dXt = Xt+dt - xt is caused by a change dt of time, with factor a(t, X t ) , in combination with a change dBt = Bt+dt - Bt of Brownian motion, with factor b(t,Xt). Since Brownian motion does not have differentiable sample paths (see Sectiorl 1.3. I ) , the following question naturally arises:

Page 74: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

136 CHAPTER 3.

In which sense can we interpret (3.6) ?

Clearly, there is no unique answer to this question. But since we know about, It6 calculus, we car1 propose the followi~lg:

Interpret (3.6) as the stochastic integral equation

t

X, = Xo + /OL a(s, XI) ds + b(s, Xs) dBs , 0 < t _< T , (3.7)

where the first integral on the right-hand side is a Riemann integral, and the second one is an It6 stochastic integral.

Equation (3.7) is called an It6 stochastic differential equation.

To call equation (3.7) a differential eqqation is counterintuitive. Its name originates from the symbolic equation (3.6), but the latter is meaningless unless one says what the differentials are. We follow the general custom and call (3.7) an It6 stochastic differential equation.

Brownian motion B is called the driving process of the It6 stochastic differential equation (3.7).

It is possible to replace Brownian motion by other driving processes, but this requires lo define Illore general stochastic integrals. This is not a topic of this book. We refer to the monographs by Chung and Williams (1990) or Prot,t,er (1992) who introduce the stochastic integral with respect to semimartingales. The latter class of processes contains Brownian motion, but also a large variety of jump processes. They are useful tools when one is interested in modelling the jump character of real-life processes, for example, the strong oscillations of foreign exchange rates or crashes of the stock market.

It is a priori not clear whether the integrals in (3.7) are well defined. For example, can we ensure that b(s, X,) is adapted to Brownian motion and is J: E[b(s, X,)I2 ds finite? These are the crucial conditions for the definition of an It6 stochastic integral; see p. 108. But can one check these conditions if we do not know the solution X? Soon we will see that there exist simple condi- tions for the existence and uniqueness of solutions to It6 stochastic differential equatiorls.

We first attempt the question:

3.2. ITO STOCHASTIC DIFFERENTL4L EQU.4TIONS 137

Surprisingly, there is no unique answer. We will find two kinds of solutions to a stochastic differential equation. These are called strong and weak solutions.

A strong solution to the It6 stochastic differential equation (3.7) is a stochastic process X = (Xt , t E [0, TI) which satisfies the following con- ditions:

X is adapted to Brownian motion, i.e. a t time t i t is a function of B,, s 5 t. '

The integrals occurring in (3.7) are well defined as Riemann or It6 stochastic integrals, respectively.

X is a function of the underlying Brownian sample path and of the coefficient functions a(t , x) and b(t, x).

Thus a strong solution to (3.7) is based on the path of the underlying Brownian motion. If we were to change the Brownian motion by another Brownian motion we would get another strong solution which would be given by the same functional relationship, but with the new Brownian motion in it.

But what is now a weak solution?

For these solutions the path behavior is not essent,ial, we are only interested in the distribution of X. The initial condition Xo and the coefficient functions a(t , x) and b(t, x) are given, and we have to find a Brownian motion such that (3.7) holds. We mention that there exist It6 stochastic differential equations which have only weak solutions; see Chung and Williams (1990), p. 248, for an example.

Weak solutions X are sufficient in order to determine the distributional characteristics of X , such as the expectation, variance and covariance functions of the process. In this case, we do not have to know the sample paths of X.

A strong or weak solution X of the It6 stochastic differential equation (3.7) is called a diffusion. Tn part,ir.iilar, taking a(t, x) = 0 and b ( t , x) = 1 in (3.7), we see that Brownian motion is a diffusion process.

In what follows, we only consider strong solutions of It6 stochastic differential equations.

First we give sufficient conditions for the existence and uniqueness of such What is a solution of the It6 stochastic differential equation (3.7) ? solutions. A proof of the following result can be found, for example, in Kloeden

and Platen (1992), Section 4.5, or in Bksendahl (1985), Theorem 5.5.

Page 75: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Assume the initial condition Xo has a finite second moment: EX: < co, and is independent of (Bt, t 2 0).

Assume that, for all t E [0, TI and x, y E R, the coefficient functions a( t , x) and b(t, x) satisfy the following conditions:

They are continuous.

They satisfy a Lipschitz condition with respect to the second vari- able:

la(t,x) - a(t,y)I + Ib(t,x) - b(t,Y)I I K lx - Y I .

Then the It6 stochastic differential equation (3.7) has a unique strong solution X on [0, TI.

for constants ci and a i , i = 1,2.

138 CHA PTER 3. 3.2. ITO STOCHA4STIC DIFFERENTI.4L EQUATIONS

Figure 3.2.3 Sample paths of geometric Brownian mot ion e(C-0.5 "') t+" " with c = 0.01. Left: a 0.01. Right: a = 0.1.

Example 3.2.2 (Linear stochastic differential equation) Consider the It6 stochastic differential equation Example 3.2.4 (Geometric Brownian motion as the solution of a linear It6

st,ochast,ic differential equation with multiplicative noise) Consider the linear It6 stochastic differential equation

The above conditions are satisfied (check them!) for In the second, It6, integral, Brownian motion and the process X are linked in

a ( t ,x ) = C I X + C ~ and b(t,x) = u ~ x + u ~ . a multiplicative way. Therefore the It6 stochastic differential equation (3.10) is

An It6 stochastic differential equation (3.8) with linear (in x) coefficient func- also referred to as a linear It6 stochastic differential equation with multiplicative

tions a( t , x) and b(t, x) is called a linear It6 stochastic differential equation. 111 Section 3.3 we will give the solution of the general linear stochastic differential From (2.28) we know: equation. By virtue of the above theory, linear stochastic differential equations have a unique strong solution on every interval [O,T], whatever the choice of the constants ci and ai.

3.2.2 SolvingIt6 StochasticDifferentialEquations bythe It6 Lemma

In this section we solve some elementary It6 stochastic differential equations by using the It8 lemma. Clearly, the first candidates are linear stochastic differential equations. They are "simple" and have unique strong solutions on Kecall how we verified that X satisfies (3.1U): we applied the It6 lemma. Now

every finite interval [0, TI; see Example 3.2.2. we are faced with the inverse problem:

t t

X t = x o + c l x S d s + u 1 XsdBs , t e [ O , T ] , (3.10)

for given constants c and a > 0.

The particular geometric Brownian motion

xt = xO e(~-0.5 u 2 ) t + 0 Bt , t E [(),TI, (3.11)

solves (3.10), and from Example 3.2.2 we conclude that X is the unique solution of (3.10).

Figure 3.2.3 shows two paths of this process.

Page 76: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

t t

xt = X O + ~ 1 X s d s + n 1 dBs, t E [O,T]. (3.1 6)

Equation (3.16) is usually referred to as Langevin equation.

140 CHAPTER 3. 3.2. IT^) STOCHASTIC DIFFERENTIAL EQUATIONS 141

use the It6 lemma to find the solution (3.11) from (3.10). which agrees with (3.11).

Suppose Xt = f (t, Bt) for some smooth function f (t , x) and recall the It6 .' Thus the solution to an It6 stochastic differential equation can sometimes be lemma from (2.26): derived as the solution of a (deterministic) partial differential equation.

1 1 1 " Example 3.2.5 (The Ornstein-Uhlenbeck process) Xt = xo + lt [fl(s, B,) + 2 f 2 2 ( ~ , B.) ds + h ( ~ , Bs) dBs . (3.12) We consider another linear stochastic differential equation:

The process X is an It6 process, and therefore we may identify the integrands in the Riemann and It6 integrals, respectively, of (3.1 0) and (3.12) (see p. 119 for the necessary argument). This together with the continuity of the sample paths of Brownian motion yields the following system of two partial differential equations for f (because the partial derivatives of f (t , x) are involved):

1 c f (t, X) = f l (t. X) + 5 f22(t1 x) , (3.13)

a f ( t l x ) = f2 ( t lx ) . (3.14)

From (3.14) we obtain f f2 f (t, x) = f22(t, x)

Thus the two differential equations (3.13) and (3.14) can be simplified:

( c -0 .5 f f2 ) f ( t , x )= f l ( t , x ) , n f ( t , x ) = f 2 ( t , x ) . (3.15) 1

Try to write f (t, x) as a product of two functions:

f (tl x) = 9(t) h(x) .

Then (3.15) becomes

(c - 0.5 a2) g(t) = gl(t) , a h(x) = hl(x) .

Both of them can be solved by separation of variables (see Example 3.1.2):

g(t) = g(0) e(~-'3.5"2)t , h(x) = h(0) eux .

Thus we obtain f (t, x) = g(0) h(O) e(c-0~5u2)t+ux .

Now recall that xo = f (0, Bo) = f (0,O) = g(0) h(0).

This finally gives

Xt = f ( t , Bt) = Xo e ( ~ - 0 . 5 U 2 ) t+UBt 1 t E [ O , T l ,

Langevin (1908) studied this kind of stochastic differential equation to model the velocity of a Brownian particle. This was long before a general theory for stocl~aslic dilferer~lial equaliorls existed.

In contrast to (3.10), Brownian motion and the process X are not directly linked in the It6 integral part of (3.16). In the physics literature, the random forcing in (3.16) is called additive noise which is an adequate description of this phenomenon.

Model (3.16) is related to the world of time series analysis. Write (3.16) in the intuitive form

dXt = c X t d t + a d B t ,

and formally set dt = 1. Then we have

Xt+l - Xt = cXt + a ( B t + l - Bt)

or &+I = 4xt + zt ,

where 4 = c + 1 is a constant and the random variables Zt = a (&+I - Bt) constitute an iid sequence of N(0,a2) random variables. This is an autore- gressive process of order 1; cf. Example 1.2.3. This time series model can be considered as a discrete analogue of the solution to the Langevin equation (3.16).

To solve (3.16) the following transformation of X is convenient:

= ePct Xt .

Note that both processes X and Y satisfy the same initial condition:

xo = Yo.

Page 77: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

142 CHAPTER 3

Figure 3.2.6 Five sample paths of the Omstein-Uhlenbeck process (3.17). Left: X o = l , c = O . l , a = l . Right: X o = l O , c = - l , a = l .

Applying the I t6 lemma (2.30) with

and A(') = c X and A ( ~ ) = a ,

we obtain

= lo [o epCs] dB, .

We conclude: I I 1 The process

rt

X t = eCt XO + a eCt e-', dB, . 1 I solves the Langevin stochastic differential equation (3.16). I

3.2. IT^) STOCHASTIC DIFFERENTIAL EQUATIONS 143

For a constarit initial condition X o , this process is called an Ornstein- Ul~lenbeck process.

Figure 3.2.6 shows several paths of this process.

I11 order to verify that the process X, given by (3.17), is actually a solution to (3.16), apply the I t6 lemma (2.30) to the process X t = u(t , Z t ) , where

$ = it e-"' dBs and u(t, z ) = ect Xo + o ect z

Moreover, since the Langevin equat,ion is a linear It6 stochastic different,ial equation, WP may conclude from Example 3.2.2 that X is the unique strong solutior~ to (3.16).

We verify that the Ornstein-Uhlenbcck process is a Gaussian process. Assume for simplicity that Xo = 0. Recall from the definition of the I t6 stochastic integral that

lt e-es dB,

is the mean square limit of approximating Riemann-Stieltjes sums

/ for partitions r,, = ( t , ) of [O, t] with mesh(^,) + 0. The latter sum has a normal distribution with mean zero and variance

(Check this!) Notice that (3.18) is a Riemann sum approximation to the inte- gral

r f 1

Since mean square convergence inlplies convergence in distribution (see Ap- pendix A l ) we may corlclude that the mean square limit St of the normally distributed Rienlann-Stieltjes sums S,, is normally distributed. This follows from the argu~nent given in Example Al.1 on p. 186. We have for Xo = 0:

EXt = 0 and var(Xt) = 2c

Page 78: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

t t t

St = ;Yo + c X. ds + ol X. d ~ ! ' ) + a 2 1 X. ,

for constants c and ai.

1.44 CH-4PTER 3. 3.2. ITO STOCHASTIC DIFFERENTIAL EQUATIONS 145

4 I

2: f ": We interpret the latter equation as

Xt - SO = c X. ds + l X, d [u, B!') + o2 B!~)]

t 0 0 0 5 1 0 ( 5 2 0 2 5 4 6 0 0 0 s 1 I) l l

6.1 6.1 = c l X, ds + (a: + u;)'/' l x s d $ , (3.20)

Figure 3.2.7 Two paths of the two-dimensional process (B!'), B!')), t E [0 ,T] , w_hich is an It6 stochastic differential equation with driving Brownian motion where B!') and B(') are two independent Brownian motions. See Ezample 3.2.8. B. From Example 3.2.4 we can read off the solution:

xt = x [c-0.5 (u:+u,~)] t+(~ :+a ; ) ' /~8~ 0 e

Using the same Riemann-Stieltjes sum approach, you can also calculate the - x [c-0.5 (a:+az)] ~ + [ U ~ B , ( ' ) + U ~ B ~ ~ ) ] - oe covariance function of an Ornstein-Uhlenbeck process with Xo = 0:

In the above definition (3.20) of the stochastic integral we were quite lucky

c o v ( ~ , , xt = d ( e ~ ( t+s) - ec(t-s) ) s < t . (3.19) because X appears as a multiplier in both integrals. Following similar patterns 2c

I as for the definition of the It6 stochastic integral, it is also possible to introduce

Since X is a mean-zero Gaussian process, this covariance function is charac- the stochastic integral

teristic for the Ornstein-Uhlenheck process. t 1' Ail) dB!') + /d A?) dBj2)

Example 3.2.8 (A stochastic differential equation with two independent driv- ing Brownian motions) for more general processes A(i). Moreover, the Brownian motions B ( ~ ) can be

Let B(%) = (B!" , t >_ 0) bc two independent Brownian motions and ai , i = 1,2, dependent, and it is also possible to consider more than two driving Brownian motions.

real numbers. See Figure 3.2.7 for an illustration of the two-dimensional pro- cess (B,('), B,(").

Define the process 3.2.3 Solving It6 Differential Equations via Stratonovich

Calculus $ = (a: + (olBjl) + o2Bj2)) . In section 2.4 we introduced the notion of Stratonovich stochastic integral.

We also explained that some of the basic rules of classical calculus formal ly Using the independence of B(') and B(*), it is rlot dificult to see that I remain valid, in particular one can use the classical chain rule. We will exploit

E E ~ = 0 and cov(Bt, B,) = min(s, t ) , this property to solve some It6 stochastic differential equations. Analogously to an It6 stochastic differential equation, a Stratonovich sto- -

i.e. B has exactly the same expectaeon and covariance functions as standard chas t ic dif ferential equat ion is a stochastic integral equation Brownian motion; see p. 35. Hence B is a Brownian motion. t

Now consider the integral equation xt = xo + Z(S, X,) ds + &t z ( ~ , Xs) 0 dBs , t E [O,T] , (3.21)

I

Page 79: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

146 CHAPTER 3. 3.2. ITO STOCHASTIC DIFFERENTIAL EQUATIONS 147

for given coefficient functions G(t, x ) and z ( t , x ) . The first integral is a Riemann Consider the stochastic process & = u ( t , Xi), where X is the solution to (3.22) integral, and the second one is a Stratonovich stochastic integral. As usual, (equivalently, to (3.25)) and u ( t , x ) is a smooth function. An application of B = (Bt , t 2 0) denotes Brownian motion. A stochastic process X is called the It6 lemma (2.30) with A:') = a(t , X L ) and Aj2) = b(t, X t ) yields a solution if it obeys (3.21). It is common to write (3.21) in the language of differentials:

dXt = Z(t , X t ) dt + %(t, ~ t ) 0 dBt .

Now assume that X is the solution of the It6 stochastic differential equation The functions a, b, u and their derivatives are clearly functions of s and X,, t t but we suppress this dependence, in the notat,ion, in order to make the above

Xt = Xo + 1 o(s ,X , )ds + 1 b ( s , X s ) d B s , t c [ O , T ] , formula more compact. An application of the transformation formula (3.23) for f = bu2 and f 2 = b2u2 + bu22 yields

where the coefficient functions a ( t , x ) and b( t , x ) satisfy the existence and uniqueness conditions on p. 138.

Recall the transformation formula (2.44) for Stratonovich integrals in terms Combining this relation with (3.26), we finally obtain: of It6 and Riemann integrals:

t t 1 f ( s , X,) 0 dBs = 1 f (s, X.1 dB. + 1' b(s, x,) f2(s, xS) d s .

For f = b we obtain

Plugging this relation into (3.22), we arrive a t the Stratonovich st,ochast,ic Indeed, assunie for the nlonlent that ~ ( t ) satisfies the deterministic differential

differential equation: dx( t ) = E(t, x ( t ) ) dt + b(t, x ( t ) ) dc(t) ,

where c ( t ) is a differentiable function. Then the classical rules of differentiation give the following formula:

u ( t + d t , x + &) - u ( t , x )

= ~ l ( t , ~ ) d t + u a ( t , x ) d ~

= [ u l ( t , x ) + i i ( t , x ) u 2 ( t , x ) ] dt + b ( t , x ) u 2 ( t , x ) d c .

t t

xt = xo + /d z ( s , x , ) d s + 1 b ( s , x , ) o d ~ ~ , (3.25)

where

- 1 a ( t , x ) = a ( t , x ) - - b(t , x ) bz(t, x ) .

2

The It6 stochastic differential equation (3.22) is equivalent to the Strato- novich stochastic differential equation (3.25) in the sense that both have the same strong solution provided it exists for one of them.

Yt = Yu+ lt [ ~ 1 + (0 - 0.5bb~) U X ] d~ + bu2 0 dBs

t

/o' = YO + 1 [u, + Zuz] ds + bu2 o dB, . /o' (3.27)

This formula is the exact analogue of the classical chain rule for a twice differentiable function u ( t , x ) .

This formal analogy with (3.27) allows one to solve Stratonovich stochas- tic differential equations by using the rules of classical calculus.

Page 80: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

xt - xo (3.29)

1 t = Jt [q f ( x S ) + l f ( ~ s ) f t ( ~ ~ ) ] ds + 1 f ( x S )

0

for constant q .

148 CHAPTER 3. 3.2. 1 ~ 6 STOCH.4STIC DIFFERENTIAL EQUATIONS 149

In what follows, we give some examples. for a positive integer n > 1.

Example 3.2.9 (Solving an It6 stochastic differential equation by solving an B) A slightly more complicated It6 stochastic differential equation is given by

equivalent Stratonovich stochastic differential equation) Let f be a differentiable function.

A) Consider the It6 stochastic differential equation

Thus 1

a(t , x) = f (x) f '(x) and b(t, x) = f (x) . 1

a( t ,x) = q f (x), + 5 f (x) f l (x) and b(t, x) = f (x) .

The corresponding Stratonovich stochastic differential equation (see (3.25)) with coefficient functions Z(t, x) = 0 and b(t, x) is then given by

This corresponds to the deterministic differential equation

It corresponds to the deter~rlirlistic dilrerential equation dx(t) = q f (x(t)) dt + f (x(t)) dc(t) , (3.31)

dx(t) = f (x(t)) dc(t), where c(t) is a differentiable function. You can use the principle of separation

where c(t) is a differentiable function. Using classical separation of variables of variables to solve this equation:

(see Example 3.1.2), we obtain = g(x(t)) - g(z(0)) =

I for some function g(x). You can check the validity of this approach by dif-

for some function g(x). Then replace x(t) with Xt and c(t) with Bt: ferentiating the integrals on the right- and left-hand sides: you obtain (3.31). I

An appeal to the Stratonovich chain rule (3.27) gives us the solution to (3.30), g ( x t ) - d x o ) = Bt . hence to (3.29), by replacing x(t) with Xt and c(t) with Bt:

Hopefully, you can solve this equation for X to get an explicit solution of .9(X't) - .9(Xo) = qt + Bt . (3.28). NOW solve the It6 stochastic differential equation Now consider the following example f (x) = x + 1 for x > 0 and Xo = 0. Then

t g(x) = ln(1 + x) and ln(1 + Xt) = qt + Bt, i.e. l t

X, = Xo + - n 1 x:.-' ds + X: dB. 2 xt = -1 + eqt+B' .

I

Page 81: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

T h e Genera l Linear Stochastic Differential Equation:

X t = X o + [ ~ ~ ( s ) X , + c ~ ( s ) ] d s + [ U ~ ( S ) X , + ~ J ~ ( S ) ] ( ~ B , , t E [ O , T ] . l l (3.32)

The (deterministic) coefficient functions ci and ui are continuous, hence bounded on [O,T], and so it is not difficult to see that the existence and uniqueness conditions on p. 138 guarantee that (3.32) has a unique strong solution.

T h e Linear Equat ion wi th Additive Noise:

X t = X o + [c , (s )X, + c ~ ( s ) ] ds + O.L(S ) dB , , t E [0, TI . (3.33) l l In terms of differentials it can be written as

dXt = [cl ( t ) X t + c2( t ) ] dt + 02 ( t ) dBt , t E [0, TI .

150 CH-4PTER 3. 3.3. T H E G E N E R A L LINEAR DIFFERENTI.4L EQUATION 151

Notes and Comments

The solution of stochastic differential equations is treated in all textbooks which are related to It6 calculus; see for example the references on p. 112.

In applications of stochastic differential equations one has to make a de- cision as to which kind of integral, It6 or Stratonovich, is more appropriate. There is no clear answer as to which type of differential equation should be used. Both kinds of equations are mathematically meaningful. The answer depends on how precisely we intend the noise process in the differential equa- tion to model the real nolse. Classical calculus is based on classes of smooth functions, in particular o,p differentiable functions. Real processes are often 1 The process X is not directly involved in the stochastic integral, and therefore smooth with a t least a shall degree of correlation. This means that the noise (3.33) gained its name. The solution of (3.33) is particularly simple. is LLsufficiently regular". "1n this case it seems reasonable to model the process One way to solve a differential equation is to guess its form from some by a Stratonovich stochastic differential equation because the rules of classical particular examples. This is perhaps not the most satisfactory approach since calculus remain valid. one certainly needs a lot of cxpcricncc for such a guess. Anyway, let us assume

that

3.3 The General Linear Differential Equation Xt = [Y ( t )]- '~/t , where

Consider ~ ( t ) = exp {- s ( s ) d s } and B = f ( t , X i ) .

for some smooth function f ( t , x). An application of the It6 lemma (2.30) on p. 120 yields

I1 d K = d ( y ( t ) X t ) = cz ( t ) y ( t ) dt + 02( t ) y ( t ) dBt .

Integrating both sides and noticing that y(0) = 1, hence X o = Yo, we obtain

This equation is important for many applications. It is particularly attractive because it has an explicit solution in terms of the coefficient functions and of the underlying Brownian sample path. In what follows, we derive this solution by multiple use of different variants of the It6 lemma.

If -Yo is a constant, this defines a Gaussian process. (You can check this 3.3.1 Linear Equations with Additive Noise by observing that the stochastic integral is the mean square limit, hence the

limit in distribution, of R.iemann-Stieltjes sums which are Gaussian random Setting ul ( t ) = 0 in (3.32), we obtain variables since the integrand P ~ ( s ) ~ ( s ) is deterministic; cf. Example A l . l on

p. 186.)

a

T h e Solution of (3.33):

t

xt = [Y (t)l-l (xo + lt cz ( S ) Y ( s ) d~ + Jld gZ(S)Y(S) d B a ) . (3.34)

Page 82: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

E*mple 3.3.1 (The Langevin equation)

111 Exa~nple 3.2.5 we already considered the Langevin equation. This is equa- tion (3.33) with cl(t) = c, c2(t) = O and a;?(t) = a for constailts c aiid a. The solution is given by

X t = eCt+Xo + uect epCS dB,, t E [0, TI . l We learnt in Exarrlple 3.2.5 that A' is called an Ornstein-Uhlenbeck process provided X o is constant. The covariance function of this Gaussian process is given in (3.19).

Figure 3.3.2 Two sample paths o f t h e Vaszcek znterest rate model drf = c [ p - r t ] dt+ u dBt wzth u = 0.5, ro = 5.05, 11 = 5. Left: c = 0.2. Right: c = 1. Both graphs show th,at rt reverts from the znitial value r o = 5.05 to the mean valur p = 5. The speed at whzch this happens i s determined b y c.

Example 3.3.3 (The Vasicek interest rate model) .

This is one of the standard niodels for describing the interest rate for borrowing and lending money when this rate is not assumed to be a constant,, but a random function rr of time t . Since r. lnay change at every instant, of time, r.t

is also called the insta~~ta~aeous interest rate. In the Vasicek model, it is given by the linear stochastic differential equation

or ill the language of different,ials:

drt = c [ p - rt] dt + (T dBt , t E [0, T ] ,

3.3. T H E G E N E R A L LINEAR DIFFERENTIAL EQUATION 153

where c, p and u are positive constants. The rationale of this model is that r t fluctuates around the value p. When rt deviates from p it will immediately be drawn back to p; one says that r t reverts to the mean p. The speed a t which

i this happens is adjusted by the parameter c. This point is illustrated well in Figure 3.3.2. This can also be seen from the form of the expectation and variance functions; see (3.36) and (3.37). The third parameter u is a measure for the order of the magnitude of the fluctuations of rt around p; see (3.37). It is called volatility.

From the general solution (3.34) we can read off the solution to (3.35):

We assume that r o is a constant. Then r is a Gaussian process. For p = O we obtair~ a11 Or1lstei11-Uhle~lbeck process. It is not difficult to calculate

and u2

var(rt) = - (1 - e-2Ct) . 2c

(3.37)

You can check these formulae by using the results for the Ornstein-Uhlenbeck process; see Example 3.2.5. Alternatively, you can use the results i n Sec- tion 3.3.4.

] As t + oo, the random variable rt converges in distribution to an N(p, u2/(2c)) variable. Since rt is Gaussian, it assumes negative values with positive prob- ability. This property is not very desirable for an interest rate. Nevertheless, if t is large and u2/(2c) is small compared with p, rt is unlikely to be pega- tive; see Figure 3.3.2 for an illustration of sample paths of the Vasicek process. There are various other models for interest rates which overcome this patholog- ical property of the Vasicek process; see for example Lamberton and Lapeyre (1996).

3.3.2 Homogeneous Equations with Multiplicative Noise

Another special case of the general linear stochastic differential equation (3.32) is given by

Page 83: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

154 CHAPTER 3

The Homogeneous Linear Equation:

or in the language of differentials, I I dXt = cl (t) Xt dt + ~1 (t)Xt dBt , t E [0, TI. I

Since Xt appears as a factor of the increments of Brownian motion, (3.38) is called a stochastic differential equation with multiplicative noise. It is called homogeneous because ca(t) = u2(t) = 0 for all t.

Sincc wc may divide both sides of (3.38) by Xo then we may assume that Xo = ,1. (The case Xo = 0 corresponds to the trivial case Xt = 0 for all t which is not of interest.) Since we expect an exponential form of the solution, we assume Xt > 0 for all t. This allows one to consider 1; = In Xt = f (Xt) and to apply the It6 lemma (2.30) on p. 120 with

2 f (t , x) = 1n x , f l ( t , x) = 0 , f2(t, x ) = x-I , f22(t, x) = -x- .

We obtain d l< = [cl ( 1 ) - 0.5 o:(l)] (it + ul (t) dBt . (3.39)

First integrate both sides of this equation, then take exponentials and finally correct for the initial value if & # 1:

The Solution of the Homogeneous Equation (3.38):

St = Xo eexp {l [cl (s) - 0.5 of (s)] ds + ol (s) dB, , 1 ' 1

You can check that X is a solution (do it!) by applying the It8 lemma to f (K) = Xo exp{K), whcrc d K is givcn by (3.39).

Example 3.3.4 (Geometric Brownian motion) The most prominent example of a homogeneous stochastic differential equation

3.3. T H E GENER,AL LINEAR DIFFERENTIAL EQUATION 155

with multiplicative noise was treated in Example 3.2.4: cl (t) = c , ul (t) = u for constants c and u. We discovered that the geometric Brownian motion

is the unique strong solution in this case. This agrees with the solution (3.40) of the general homogeneoils stochastic differential equation (3.38). Geomet- ric Brownian motion is of major importance for applications in finance; see Cha.pt,er 4.

3.3.3 The General Case Now we are well prepared to solve the general linear stochastic differential equation (3.32). We achieve this by embedding (3.32) in a system of two stochastic differential equations.

Recall the solution Y of the homogeneous stochastic differential equation (3.38) with Yo = 1. It is given by formula (3.40)' (with X replaced by Y). Consider the two processes

Apply the It6 lemma (2.30) on p. 120 to Y,-' (with f (t , x) = xP1). After some calculations we obtain

An appeal to the integration by parts formula (2.35) yields

Now integrate both sides and recall that Yo = 1. Having in mind the particular forms of Xi1) and Xi2) , we finally arrive a t

The Solution of the General Linear Equation (3.32):

t

Xt = YL ( ~ ~ + ~ [ c ~ ( t ) - u 1 ( t ) u 2 ( t ) ] l ~ ~ d s + ~ UZ(S)Y;'~B.) ,

t E [O,T] (3.41)

Page 84: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

156 CHAPTER 3.

Here the process k7 is the solutio~l (3.40) of the homogeneous equation (3.38) (where S has to be replaced with 17) with the initial condition Yo = 1.

As a matter of fact, we also derived the solution of the corresponding deter- ministic differential equation

dx( t ) = [cl ( t ) x ( t ) + cz( t )] dt , t E [0, T ] .

Simply set u l ( t ) = uz(t) = 0 for all t . We will use this fact for calculating the expectation and second moment functions of X in Section 3.3.4.

3.3.4 The Expectation and Variance Functions of the So- lut ion

If we have the explicit solution of a stochastic differential equation we can in principle determine its expectation, covariance and variance functions. We were able to proceed in this way for the Ornstein-Uhlenbeck process (see (3.19) on p. 144) and for geometric BI-ownia11 m~otion (see (1.16) and (1.17) on p. 42).

In what follows we go another way to calculate

p x ( t ) = E X , and q x ( t ) = E X y t ) ), t E [O, TI .

Clearly, then we can also calculate u $ ( t ) = var(Xt). Recall the general linear stochastic differential equation:

6 = XO + [ C I ( s ) X s + C Z ( S ) ] ds + [UI ( s ) X s + 0 2 ( s ) ] dBs , t E 10, TI . lo -

Take expectations on both sides and notice that the stochastic integral has expectation zero; see p. 11 1. Hence

This corresponds to the general linear differential equation for p x ( t )

Bearing in mind the remark a t the end of the previous section, we can write down the solution px of this differential equation in terms of cl and c2. Derive

3.4. NUMERICAL SOLUTION 157

p-y for the Ornstein-Uhlenbeck process and for geometric Brownian motion and compare with the known formulae.

Similarly o6e'can proceed for q x ( t ) . To ohta.in a, st,acha,st,ic different,ial equation for X f , first apply the I t8 lemma (2.30) on p. 120:

I then take expectations:

I q x ( t ) = qx ( 0 ) + [[2c1(s) + 0:: ( s ) ] P X ( s ) + 2 [ ~ 2 ( ~ ) + 0 1 ( ~ 1 ~ 2 ( ~ 1 1 PX ( s ) 6' + u; ( s ) ] d~ .

This is a general linear differential equation for q x ( t ) (notice that p x ( t ) is known):

q i ( t ) = ( t ) + u : ( t ) ] q . ~ ( t ) + 2 [ ~ 2 ( t ) f g l ( t )gz ( t ) ] P X ( ~ ) f ,

I whose solution can be derived as a special case of (3.41); see the remark a t the end of the previous section. No doubt: t,he soliit8ians I L ~ and qx look in general quite messy, but in the case of constant coefficient functions ci and ui

11 you might try to obtain the expectation and variance functions of X .

1 3.4 Numerical Solution

i Stochastic differential equations which admit an explicit solution are the ex- ception from the rule. Therefore numerical techniques for the approximation i of the solution to a stochastic differential equation are called for. In what

! follows, such an approximation is called a numerical solution. Numerical solutions are needed for different aims. One purpose is to vi-

sualize a variety of sample paths of the solution. A collcction of such paths is sometimes called a scenario. I t gives an impression of the possible sample path behavior. In this sense, we can get some kind of "prediction" of the stochastic process a t future instants of time. But a scenario has to be interpreted with care. In real life we never know the Brownian sample path driving the stochas- tic differential equation, and the simulation of a couple of such paths is not representative for the general picture.

Page 85: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

158 CHAPTER 3.

4 second objective (perhaps the most important one) is to achieve reason- able approximations to the distributional characteristics of the solution to a stochastic different,ial erl~ia,tion. They inclnde expectations, va.riances, covari- ances and higher-order moments. This is indeed an important matter since only in a few cases one is able to give explicit formulae for these quantities, and even then they frequently involve special functions which have to be approximated numerically. Numerical solutions allow us to simulate as many sample paths as we want; they constitute the basis for Monte-Carlo techniques to obtain the distributional characteristics.

For the purpose of illustration we restrict ourselves to the numerical solu- tion of the stochastic differential equation

dXt = ~ ( X L ) dt + b(Xt) dBt , t E [0, TI, (3.42)

where, as usual, B denotes Brownian motion, and the "differential equation" is actually a stochastic integral equation. We also assume that the coefficient functions a(x) and b(z) are Lipschitz continuous. If in addition EX: < oo, these assumptions guarantee the existence and uriiqueness of a strong solution; see p. 138.

3.4.1 The Euler Approximation

A numerical solution Xtn) = ( .~ , ( ' " , t E [O,T]) of the stochastic differential equation (3.42) is a stochastic process that approximates the solution S =

(Sf, t E [0, TI) of (3.42) in a sense to be made precise later. Such a solution is cliaracterized by a partition T,, of [0, TI:

6n = ~nesli(r,,) = , max (ti - z=l, ..., n

t iPl) = max Ai i=l, ..., T I

The process x ( ~ ) is calculated only a t the points ti of the partition T,,, and so one has some freedom to choose on the intervals [ti-l, ti]. Since we are interested in solutions X with continuous sample paths we will assume t,hat xin) on ( t iPl , t,) is obtained by sirnple linear interpolation of the points

(n) (t,-1,Xtz_1) and (t,,x,':"'). A numerical approximation scheme dctcrmincs x ( ~ ) a t the points ti. A

naive way to obtain a numerical solution is to replace the differentials in the stochastic differential equation (3.42) with differences. This leads to the fol- lowing iterative scheme:

The Euler Approximation Scheme: Denote, as usual,

Ai = ti - ti-, and AiB = Bt i - Bt,-, , i = 1,. . . , n .

Then

xin' = X o ,

x i ) = XAn) + a(.X$)) A, + b(XAn') A,B,

xi:' = Xi:' + a(xl:') A, + b(Xj:)) A 2 B ,

'"' = xt(nn12 + a(X::!,) a n - 1 + b(xt(nn12) ~ ~ - 1 ~ 7 xt"-l

xF) = x:rll + U(X,':!~)A, + ~(X~:!~)A,B.

3.4. NIIMER.lCA L SOLUTION 159

In practice one usually chooses equidistant points ti such that

6, = mesh(rn) = T l n ,

- (n) (n) xj;tjn - X(i-l)Tin + a(X(i-l)T/n) 6n + b ( ~ { r ? ~ ) ~ / ~ ) A i B 7

i = 1, ..., n . (3.43)

Figures 3.4.1 and 3.4.2 show how the Euler approximation works in practice.

Several questions naturally arise:

(a) I n which sense is the numerical solution x ( ~ ) close to the solution X ?

(b) How can we measure the quality of the approzimation of X by x ( ~ ) ?

(c) Is there an approximation to X which is better than the Euler scheme?

The third question has a positive answer; see Section 3.4.2. The first and second questions have several answers. If we are interested in the pathwise approximation of X , we would like that X(")(w) is "close" to the sample path X(w) for a given Brownian path B(w). In practice we never know the latter

Page 86: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

160 CH.4PTER 3. 3.4. N7JMERTC.4T1 SOTJUTTON 161

path, b u t we can easily simulate Brownian paths on the computer; see Sec- tion 1;3.3 for a brief introduction to this topic. As a measure of the quality of pa.t,hwise approximation one can choose the quantity

e,(6,) = EIXT(W) - XP'(w)l .

Notice that e, indeed describes the w-wise (i.e. pathwise) comparison of and Xln) a t t = T. The index "s" in e, stands for "strong"; a pathwise approx- imation of X is usually called a strong ntrmerical soltrtion; see below. There are certainly more appropriate criteria to describe the pathwise closeness of X and x("); one of them could be Es~p,,~, , , ] IXt(w) - ~T, (n (w) ( . However, the latter quantity is more difficult t o deal with theoretically. So let us believe that the paths X(")(w) and X(w) are close whenever they are close a t the end of the interval [0, TI.

In contrast to a strong numerical solution, a weak numerical solution aims a t the approximation of the moments of the solution X . In this case it is not

I essential how close Xt(w) and ~ ~ ( ~ ' ( w ) really are. What matters is that the /) difference of moments

4 6 , ) = l ~ f ( x T ) - ~f (x/")) / is small. Here f is chosen from a class of smooth functions, for exa,mple cert.ain polynomials or functions with a specific polynomial growth. We refrain frorn discussing these classes of functions because their definitions are technical, and they also differ in the literature. The "d' in e, stands for "weak". AS before, we may wonder why we compare the moments Ef (XT) and ~f (x?)) only for t = T , but again, this is for theoretical simplicity.

I

Figure 3.4.1 The equidistant Euler scheme (3.43) at work: ~~~urrscr~kc l l solutions (dashed lines) and exact solution to the stochastic differential equation dXt = 0.OlXt dt + O.01Xt dBt, t E [O,l] , with Xo = 1.

We say that A'(,) is a strong numerical solution of the stochastic differ- ential equation (3.42) if

e,(6,) + 0 as 6, = mesh(rn) + 0 . I

Page 87: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

The numerical solution x(,) converges strongly to X with order y > 0 if there exists a constant c > 0 such that

eS(6,) I c 6 : for 6, < h O .

The numerical solution x(") converges weakly to X with order y > 0 if there exists a constant c > 0 such that

,e, (6,) 5 c 6: for 6, 5 do

- .

The equidistant Euler approximation (see (3.43)) converges strongly with order 0.5 and weakly with order 1.0 (for a class of functions with appro- priate polynomial growth).

CHAPTER 3. 3.4. NUAIERICAL SOLUTION 163

In order to measure the quality of the approxinlation of X by ~ ( ~ 1 one intro- The Milstein approximation exploits a so-called Taylor-It6 expansion of duces the order of convergence: relation (3.44). The idea consists of applyi~lg the It6 lemma on p. 120 to

the integrands a(X , ) and b(X,) in (3.44). To make the formulae below more co~npact we write a, b,at, . . . instead of a(X,) , b ( X y ) , a f ( X y ) , . . ..

= It' t , - 1 [ a ( X t ) + r t,-1 (aaf + :b2atf) d y + I s batdBy] ds t i - 1

+ f i t t - 1 ( X ~ , - ~ ) + J ~ t i - 1 (abt+!b2btf) 2 d y + J s t i - 1 bbtd~.] . i ~ ~

= ~ ( x t , - ~ ) Ai + b(Xt ,_ , ) A i B + Ri , (3.45)

where the remainder term is given by

3.4.2 The Milstein Approximation The double stochastic integral R:') is then approximated by

The Euler approximation can be further improved. We illustrate this by in- troducing the Milstein approximation. As before, we consider the stochastic differential equation R!') z b ( X t i - l ) b f ( X t , - l )

t 1

xt = xo + llu a ( X s ) ds + 1 b ( X s ) dBs , t E [OIT] . Denote the double integral on the right-hand side by I. The evaluation of I is quite delicate. Recall from p. 99 that = ds, and therefore

For the points ti of the partition T, we can consider the difference Xt, - Xt,-l and obtain:

Xti = Xti-l + J t i a ( X s ) ds + Jt i b ( X s ) dBs , i = 1,. . . , n (3.44) t i -1 ti-1 Consider the double integral

The Euler approximation is based on a discretization of the integrals in (3.44). (3.48)

To see this, first consider the approximations

ti t ; - - ( ( i t d B s ) ( ( i t d B y ) = ( i t ( ( i t dB.) d B s . / a s d t i 1 i (i b(Xs) dBs b(Xt i - , ) A i B . t,-1 t i - 1 t i - I t,-1 ti-1 t.-1

Alternatively, we interpret the latter integral in a heuristic sense as and then replace Xti with Xi"':

XtI., = Xtinl ,-I + a ( X j ~ ' ~ ) + b ( ~ j 2 , ) A ~ B , i = 1,. . . , n .

Page 88: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Combining (3.47) with (3.48) and (3.49), we have

Under mild assumptions on the coefficient functions a ( x ) and b(x) one can show

that Rj2) is small in comparison with Thus the latter term determines the order of magnitude of Ri. Relations (3.45) and (3.50) give the motivation for the definition of the Milstein approximation scheme. Notice that this scheme is the Euler approximation with an additional correction term containing the squared increrner~ts o l Brow~lia~l 111otio11.

The Milstein Approximation Scheme:

Xin) = Xo and for i = 1,. . . , n,

x!:' = Xi:)), + +a(X,':_', ) Ai + b ( X F l ) Ai B

1 +- 2 b ( ~ , ( , : ) ~ ) b ' ( ~ , ( , ? ~ ) [ ( A i ~ ) 2 - A,] .

I The Milstein approximation leads to a substantial improvement of the quality i I of the numerical solution:

The equidistant Milstein approximation converges strongly with or- der 1.0.

This improvement is illustrated well in Figure 3.4.2.

Notes and Comments

The Milstein approximation can be further improved by applying the It6 lemma to the integrands in t,he remainder term R,; see (3.46). The form of t,he nu- merical solutions then becomes more and more complicated. However, in view of the power of modern computers, this is not an issue one has to worry about.

Taylor-It6 expansions such as (3.45) involve multiple stochastic integrals. Their rigorous treatment requires a more advanced level of the theory. This also concerns the derivation of (3.50); the latter relation was obtained by some heuristic arguments.

Figure 3.4.2 A comparison of the equidistant Euler (left column) bnd Milstein (right column) schemes. In every figure a numerical solution (dashed lines) and the exact solution to the stochastic differen,ti;aL eqt~ation dXt = 0.01 X t dt + 2Xt dBt, t E [ O , l ] , with Xo = 1 are given. Notice the significant improvement of'the approxi- mation by the Milstein scheme for large n , in particular at the end of the interval.

Page 89: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

166 CHAPTER 3.

The numerical solution of stochastic differential equations is a relatively new area of applied probability theory. An overview of the existing numerical techlliques is given in Kloeden and Platen (1992) and the companion book by Kloeden, Platen and Schurz (1994). The latter book aims a t the introduction to stochastic differential equations and their numerical solution via computer experiments.

I

4

Applications of Stochastic Calculus in Finance

Since the celebrated papers by Black and Scholes (1973) and Merton (1973) the idea of using stochastic calculus for modelling prices of risky assets (share prices of stock, stock indices such as the Dow Jones, Nikkei or DAX, foreign exchange rates, interest rates, etc.) has been generally accepted. This led to a new branch of applied probability theory, the field of mathematical finance. It is a symbiosis of stochastic modelling, economic reasoning and practical financial engineering. 1 In this chapter we consider the Black-Scholes model for pricing a Euro- pean call option. Do not worry if you do not have much prior knowledge about economic and financial processes. As a minimum, you will need some words of economic language, and as a maximum, the It6 lemma. In Section 4.1 we will explain the basic terminology of finance: bond, stock, option, portfolio, volatil- i ty , trading strategy, hedging, maturi ty of a contract, self-financing, arbitrage will be the catchy words. The Black-Scholes formula for pricing a European call option will be obtained as the solution of a particular partial differential equation.

Since not,ions like equivalent martingale measure and change of measure have become predominant in the financial literature, we give in Section 4.2 a short introduction to this area. This requires some knowledge of measure- theoretic tools, but, as usual, we will keep the theory on a low level. We hope that the key ideas will become transparent, and we give a second derivation of the Black-Scholes formula by using the change of measure ideology.

Page 90: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

168 CHAPTER 4.

4.1 The Black-Scholes Option Pricing Formula

4.1.1 A Short Excursion into Finance

We assume that the price X t of a risky asset (called stock) a t time t is given by geometric Brownian motion of the form

where, as usual, B = (Bt, t > 0) is Brownian motion, and Xo is assunled to be independent of B. The motivation for this assumption on X comes from the fact that X is the unique strong solution of the linear stochastic differential equation

r t rt

which we can formally write as

This was proved in Exar~~ple 3.2.4. If we interpret this equation in a naive way, we have on [t , t + dt]:

$ . I . THE BLACK-SCHOLES OPTION PRICING FORMULA 169

Now assume that you have a non-risky asset such as a bank account. In financial theory, it is called a bond. We assume that an investment of Po in bond yields an arnou~lt of

Pt = Po ePt

a t time t. Thus your initial capital Po has been continuously compounded with a constant interest rate r > 0. This is an idealization since the interest rate changes with time as well. Note that 0 satisfies the deterministic integral equation

t

8 = Po + 7- Jo A d s . (4.3)

In general, you want to hold certain amounts of shares: at in stock and bt in bond. They constitute your portfolio. We assume that at and bt are stochastic processes adapted to Brownian motion and call the pair

(at, bt) , t E 10, TI ,

a trading strategy. Clearly, you want to choose a strategy, where you do not lose. How to choose (at, b t ) in a reasonable way, will be discussed below. Notice t<h.zt, your rlicnlth I/t (or the ~inluc of gour portfolio) at, t,ime t is now given by

Vt = at Xt + bt pt .

We allow both, at and b t , to assume any positive or negative values. A negative value of at means short sale of stock, i.e. you sell the stock a t time t. A negative value of bt means that you borrow money a t the bond's riskless interest rate r . In reality, you would have to pay transaction costs for operations on stock and sale, but we neglect them here for simplicity. Moreover, we do not assume that at and bt are bounded. So, in principle, you should have a potentially infinite

I amount of capital, and you should allow for unbounded debts as well. Clearly, this is a simplification, which makes our mathematical problems easier. And finally, we assume that you spend no money on other purposes, i.e. you do not make your portfolio smaller by consumption.

We assume that your trading strategy (at, bt ) is self-financing. This means that the increments of your wealth & result only from changes of the prices Xt and pt of your assets. We formulate the self-financing condition in terms of differentials:

dK = d(atXt + btPt) = at dXt + bt dDt ,

which we interpret in the It6 sense as the relation

t t

I4 - VO = l d ( a . X. + b.0,) = /d a,d-Y. + 1 bad@, .

The quantity on the left-hand side is the relative return from the asset in the period of time [t, t + dt]. It tells us that there is a linear trend cdt which is disturbed by a stochastic noise term u dBt. The constant c > 0 is the so-called mean rate of return, and o > 0 is the volatility. A glance a t formula (4.1) tells us that, the larger u, the larger the fluctuations of Xt . You can also check this with the formula for the variance function of geometric Brownian motion, which is provided in (1.18). Thus u is a measure of the riskiness of the a.sset.

I t is believed that the model (4.2) is a reasonable, though crude, first approximation to a real price process. If you forget for the moment the term with a , i.e. assume u = 0, then (4.2) is a deterministic differential equation which has the well-known solution Xt = Xo exp{ct). Thus, if u > 0, we should expect to obtain a randomly perturbed exponential function, and this is the gcomctric Brownian motion (4.1). People in economics believe in exponential growth, and therefore they are quite satisfied with this model.

Page 91: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

170 CHAPTER 4.

The integrals on the right-hand side clearly make sense if you replace d X , with c X , ds + a x , dB,, see (4.2), and dp, with rp, ds, see (4.3). Hence the value & of your portfolio a t time t is precisely equal to the initial investment Vo plus capital gains from stock and bond up to time t .

4.1.2 What is an Option?

Now suppose you purchase a ticket, called an option, a t time t = 0 which entitles you to buy one share of stock until or at time T, the time of maturity or time of expinztion of the option. If you can exercise this option a t a fixed price K , called the exercise price or strike price of the option, only a t time of maturity T , this is called a European call option. If you can exercise i t until or a t time T , it is called an American call option. Note that there are many more different kinds of options in the real world of finance but we will not be able to i~lclude therrl i11 this book.

The holder of a call option is not obliged to exercise it. Thus, if a t time T the price XT is less than K , the holder of the ticket would be silly to exercise it (you could buy one share for XT$ on the market!), and so the ticket expires as a worthless contract. If the price XT exceeds K , it is worthwhile t o exercise the call, i.e. one buys the share a t the price K , then turns around and sells it a t the price XT for a net profit XT - K.

In sum, the purchaser of a European call option is entitled to a payment of

See Figure 4.1.1 for an illustration. A put is an option to sell stock a t a give11 price If on or uiltil a parlicular

date of maturity T. A European put option is exercised only a t time of maturity, an American put can be exercised until or a t time T. The purchaser of a European put makes profit

In our theoretical considerations we restrict ourselves to European calls. This has a simple reason: in this case we can derive explicit solutions and compact formulae for our pricing problerr~s. Thus,

from now on, an option is a European call option.

4.1. THE BLACK-SCHOLES 0PTIONPR.ICING FOR,MllLA

i 1 X - h') = 0 1

Figure 4.1.1 The value of a European call option with exercise price K at time of maturity T .

As an aside, it is interesting to note that the situation can be imagined col- orfully as a game where the reward is the payoff of the option and the option holder pays a fee (the option price) for playing the game.

Since you do not know the price XT a t time t = 0, when you purchase the call, a natural question arises:

How much would you be willing to pay for such a ticket, i.e. what is a rational price for this option a t time t = 0 7

Black, Scholes and Merton defined such .a value as follows:

An individual, after investing this rational value of money in stock and bond a t time t = 0, can manage his/her portfolio according to a self-financing strategy (see p. 169) so as to yield the same payoff ( X T - K)+ as if the option had been purchased.

If the option were offered a t any price other than this rational value, there would be an'opportunity of arbitrage, i.e. for unbounded profits without an acconipanying I-isk of loss.

Page 92: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

172 CHAPTER 4.

4.1.3 A Mathematical Formulation of the Option Pricing Problem

Now suppose we want to find a self-financing strategy (a t , b t ) and an associated valuc process b; such that

for some smooth deterministic function u( t , x). Clearly, this is a restriction: you assume that the value & of your portfolio depends in a smooth way on t and Xt. I t is our aim to find this function u(t ,x). Since the value VT of the portfolio a t time of maturity T shall be (XT - K)+, we get the terminal condition

VT = ~ ( 0 , XT) = (XT - K)+ . (4.4) In the financial literature, the process of building a self-financing strategy such that (4.4) holds is called hedging against the contingent claim (XT - K)+.

We intend t o apply the It6 lemma to the value process Vt = u ( T - t , Xt) . Write f (t , x) = u(T - t , x) and notice that

f l ( t , x) = -u1 ( T - t , 2) , f2(t, x) = u2(T - t , I ) , f22(t, x) = ~ 2 2 ( T - t, 2).

Also recall that X satisfies the It6 integral equation

Now an application of the It6 lemma (2.30) with A( ' ) = c X and A(') = a X yields that

+ 0.5 0' X: ~ 2 2 ( T - S, X,)] ds

rt

+ l o [a X, u2 ( T - S, X,)] dB, . (4.5)

1 4.1. THE BLACK-SCHOLES OPTION PRICING FORMULA

On the other hand, (at, bt) is self-financing:

Since pt = DoePt, dDt = r h e P t d t = r p t d t .

Moreover, Vt = at Xt + bt pt, thus

Combining (4.6)-(4.8), we obtain another expression for

[(c - T) a, X , + TV,] ds + (4.9)

j Now compare formulae (4.5) and (4.9). We learnt on p. 119 that coefficient functions of It6 processes coincide. Thus we may formally identify the inte- grands of the Riemann and It6 integrals, respectively, in (4.5) and (4.9):

= -u l (T - t , Xt) + c X t u2(T - t , X t )

Since Xt may assume any positive value, we can write the last identity as a partial differential equation ("partial" refers to the use of the partial derivatives of u):

Page 93: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

174 CHAPTER 4.

Recalling the terminal condition (4.4), we also require that

VT = u(O,XT) = (XT - K ) + .

This yields the deterministic terminal condition

~ ( 0 , x ) = ( x - K ) + , x > 0 . (4.12)

4.1.4 The Black and Scholes Formula

In general, it is hard to solve a partial differential equation explicitly, and so one has to rely on numerical solutions. So it is somewhat surprising that the partial differential equation (4.11) has an explicit solution. (This is perhaps one of the leasoris for the popularity of the Black-Scholes-Merton approach.) The partial differential equation (4.11) with terminal condition (4.12) has been well-studied; see for example Zauderer (1989). It has the explicit solution

u(t, x) = x Q(g(t, x)) - K eWTt @(h(t, x)) ,

where

and

is the standard normal distribution function. After all these calculations,

what did we actually gain?

Recalling our starting point on p. 171, we see that

VO = u(T, Xo) = XO @(g(T, XO)) - K e C T Q ( ~ ( T , x ~ ) ) (4.13)

is a rational price a t time t = 0 for a European call option with exercise price I<.

The stochastic process Vt = u(T -t, X t ) is the value of your self-financing portfolio a t time t E [O, TI.

' 4 1. THE BLACK-SCHOLES OPTION PRICING FORMULA 175

The self-financing strategy (a t , b t ) is given by

u(T - t , X t ) - a t X t at = u ~ ( T - t , X t ) and bt =

Pt , (4.14)

see (4.10) and (4.8).

.kt time of maturity T , the formula (4.13) yields the net portfolio value of (XT - K)+. Moreover, one can show that at > 0 for all t E [O, TI, but bt < O is not excluded. Thus short sales of stock do not occur, but borrowing money a t the bond's constant interest rate r > 0 may become necessary.

Equation (4.13) is the celebrated Black-Scholes option pricing formula. We see that it is independent of the mean rate of return c for the price S t , but it depends on the volatility c.

If we want to understand q = u(T, Xo) as a rational value in terms of arbi- trage, suppose that the initial option price p # q. If p > q, apply the following strategy: a t time t = 0

sell the option to someone else a t the price p, and

invest q in stock and bond according to the self-financing strategy (4.14).

Thus you gain an initial net profit of p - q > 0. At time of maturity T , the portfolio has value a~ XT + b ~ / 3 ~ = (XT - K ) + , and you have the obligation

I to pay the value ( S T - I()+ to the purchaser of the option. This means: if .YT > K , you must buy the stock for XT, and sell it to the option holder a t the exercise price K, for a net loss of ST - K . If JiT < K , you do not have to pay a~ ly t l~ i~ lg , since the option will not be exercised. Thus the total terminal profit is zero, and the net profit is p - q.

The scale of this game can be increased arbitrarily, by selling n options for np a t time zero and by investing n q in stock and bond according to the self-financing strategy (nut, nbt). The net profit will be n (p - q). Thus the opportunity for arbitrarily large profits exists without an accompanying risk of loss. This means arbitrage. Similar arguments apply if q > p; now the pur- chaser of the option will make arbitrarily big net profits without accompanying risks.

Notes and Comments The idea of using Brownian motion in finance goes back to Bachelier (1900), but only after 1973, when Black, Scholes and Merton published their papers,

Page 94: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

176 CHAPTER 4.

did the theory reach a more advanced level. Since then, options, futures and many other financial derivatives have conquered the international world of finance. This led to a new, applied, dimension of an advanced mathematical theory: stochastic calculus. As we have learnt in this book, this theory requires some non-trivial mathematical tools.

In 1997, Merton and Scholes were awarded the Nobel prize for economics. Most books, which are devoted to mathematical finance, require the knowl-

edge of measure theory and functional analysis. For this reason they can be read only after several years of university education! Here are a few references: Duffie (1996), Musiela and Rutkowski (1997) and Karatzas and Shreve (1998), and also the chapter about finance in Karatzas and Shreve (1988).

By now, there also exist a few texts on mathematical finance which address an elenlentary or intermediate level. Baxter and Rennie (1996) is an casy introduction with a minimum of mathematics, but still precise and with a good explanation of the economic background. The book by Willmot, Howison and Dewynne (1995) focuses 011 partial differential equations and avoids stochastic calculus whenever possible. A course on finance and stochastic calculus is given by Lamberton and Lapeyre (1996) who address an intermediate level based on some knowledge of measure theory. Pliska (1997) is an introduction to finance using only discrete-time models.

4.2 A Useful Technique: Change of Measixre

In this section we consider a very powerful technique of stochastic calculus: the change of the underlying probability measure. In the literature it often appears under the synonym Girsanov's theorem or Cameron-Martin formula.

In what follows, we cannot completely avoid measure-theoretic arguments. If you do not have the necessary background on measure theory, you should a t least try to understand the main idea by considering the applications in Section 4.2.2.

4.2.1 What is a Change of the Underlying Measure?

The main idea of the change of measure technique consists of introducing a new probability measure via a so-called density function which is in general not a probability density function.

We start with a simple example of two distributions on the real line. Recdl

4.2. A USEFUL TECHNIQUE: CHANGE OF AiIEASURE

the probability density of a normal N(p , u" random variable:

and write

@ ~ , u ~ ( x ) = J _ ' _ P , , ~ 2 ( y ) d y , x E R .

for the corresponding distribution function. Consider two pairs (p l , a;) and (p2, ~22) of parameters and define

cp,, ,u: (2) cp,,,,,. (x) fi = and f i ( x ) =

cP,,,,; (.I cp,, ,u:

and

The function f l is called the density function of a,,,,: with respect to +,,,,;, and f2 is the density function of @,2,u22 with respect to @,,,,;. Clearly, fl and fi are not probability densities; they are positive functions, but the integrals

fi(x) dx, i = 1,2, are not equal to 1. Now we consider a more general context:

Let P and Q be two probability measures on the a-field F. If there exists a non-negative function fi such that

we say that fl i s the density of Q with respect t o P and we also say that Q is absolutely continuous with respect to P.

--

The integrals in (4.17) have to be interpreted in the measure-theoretic sense. In a similar way, changing the roles of P and Q, we can introduce the

density f2 of P with respect t o Q , given such a non-negative function exists.

If P is absolutely continuous with respect to Q, and Q is absolutely con- tinuous with respect to P, we say that P and Q are equivalent probability measures. I

Page 95: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

From (4.15) and (4.16) we may conclude that two Gaussian probability mea- sures on the real line are equivalent.

For a characterization of absolute continuity via the Radon-Nikodym the- orem, see Appendix A5.

As usual, B = (Bt, t E [0, TI) denotes standard Brownian motion. The definition of Brownian motion on p. 33 heavily depends on the underlying probability measure P. Indeed, for the definition of the independent, sta- tionary increments of B and for its Gaussian distribution we must know the probability measure P on the a-field 3. Usually, we do riot pay attention to this fact, P i s simply given, but in what follows, it will be a crucial aspect. For example, consider the one-dimensional distribution function P(Bt 5 x), x E R It is the distribution function of an N(O, t) random variable, but if we were to change P for another probability measure Q, this function cnuld differ from a normal distribution' function. At the moment, this remark may sound a little abstract, but we will see soon that the change of measure is a very useful technique.

We are interested in processes of the form

for some constant q. With the only exception when q = 0, B is n o t standard Brownian motion. However, if we change the-underlying probability measure P for an appropriate probability measure Q, B can be shown to be a standard Browr~iar~ ~rlotior~ u ~ ~ d e r . lhe new probability measure Q. This simply means that B satisfies the defining properties of a standard Brownian motion on p. 33, when you replace P with Q. This is the content of the following famous result; see for example Karatzas and Shreve (1988) for a proof. As usual,

is the Brownian filtration.

Girsanov's Theorem:

The following statements hold:

The stochastic process

M t = e x p (4.20)

is a martingale with respect to the natural Brownian filtration (4.19) under the probability measure P.

4.2. A USEFUL TECHNIQUE: CHANGE OF MEASURE 179

1 The relation I I I defines a probability measure Q on 3 which is equivalent to P. I 1

I The process B is adapted to the filtration (4.19). I I

The probability mcasurc Q is called an equivalent martingale measure.

Under the probability measure Q, the process B defined by (4.18) is a standard Brownian motion.

The change of measure serves the purpose of eliminating the drift tcrm in a stochastic differential equation. This will be seen in the applications below. We illustrate this fact by the following simple example.

Example 4.2.1 (Elimination of the drift in a linear stochastic differential equation)

Consider the linear stochastic differential equation

with constant coefficients c and u > 0. We know from Example 3.2.4 that this equation has solution

Introduce

Bt = Bt + (C/O) t , t E [0, TI , and rewrite (4.22) as follows:

d x t = aXt d[(c/a) t + a Bt] = aXt dB,, t E [0, T] . (4.24)

By Girsanov's theorem, B is a standard Brownian motion under the equivalent martingale measure given by (4.21) with q = c/o. The linear stochastic differ- ential equation (4.24) does not have a drift term; its solution X is a martingale under Q, but it is not a martingale under P. You can read off the solution X of (4.24) from (4.23) (choose c = O and replace B with B) :

Page 96: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

This is the solutio~l to the original stochastic differential equation (4.22) driven by B. It scems that we did not gain much. However, if we had known the solution (4.23) only for c = 0, we could have derived the solution of (4.23) for general c from the solution in the case c = 0. Moreover, since X is a martingale under Q , one can make use of the martingale property for proving various results about X . We will use this trick for pricing a European call option in Section 4.2.2.

4.2.2 An Interpretation of the Black-Scholes Formula by Change of Measure

In this section we review the Black-Scholes option pricing formula. We will show that it can be interpreted as the conditional expectation of the discounted overshoot (XT - K ) + a t maturity.

First recall:

The Black-Scholes Model

The price of one share of the risky asset (stock) is described by the stochastic differential equation

where c is the nleurl rnte of r e t u ~ n , u the uulutility, B is starldard Brow- nian motion (under P) and T is the time of maturity of the option.

The price of the riskless asset (bond) is described by the deterministic differential equation

where r > 0 is the interest rate of the bond

Your portfolio a t time t consists of at shares of stock and bt shares of bond. Thus its value a t time t is given by

The portfolio is self-financing, i.e.

dK = at dXt + bt dpt , t E [0, TI

4.2. .4 USEFUL TECHNIQUE: CHANGE OF AfEASURE 181

At time of maturity, I j . is equal to the contingent claim h ( X T ) for a given function l b . For a European call opt,ion, h(x) = (x - K)', where K is thc steke price of the option, and for a European put option, h(;c) = (A- - x)+.

In Section 4.1.4 we derived a rational price for a European call option and showed that there exists a self-financing strategy (at, bt) such that VT = (,YT - K)+. In what follows, we give an intuitive interpretation of this pricing formula by applying the Girsanov Lheor-ern.

If we were naive we could argue as follows. Your gain from the option a t time of maturity is ( X T - K ) + . In order to determine the value of this amount of money a t tirne zero, you have to discount it with given interest rate r:

We do not know XT in advance, so let's assume that X satisfies the linear stochastic differential equation (4.25), and simply take the expect,ation of (4.26) as t,he price for the option a t time zero.

This sounds convi~lcing, although we did not use a,ny theory. And because we did not apply any theory we will see that we are slightly wrong with our arguments, i.e. we do not get the Black-Scholes price in this way. I t will turn out that tk.e rational Black-Scholes price is the expected value of (4.26), but we will have to adjust our expectation by changing the under ly i~~g probabilit,~ measure; see (4.29) for the correct, formula.

This change of the probability nieasure I' will be provided in such a way that the ~ I S C O I L T L ~ P ~ ~ T Z C P of one share of stork

- St = e - ' ' ~ t , t E [O, T I ,

will become a martingale under the new probability measure Q. Write

f (t , x) = e-Ttx

and apply t , h ~ Tt,b l~nlrrla (2.31) on p 120 (notice that f . ~ ~ ( t , x) - 0) to obtain

-re-"lit dt + e C t X t [cdt + u dBt] -

= X t [(c - r ) dt + a dBf]

Page 97: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

182 CHAPTER 4

where Bt = B t + [ ( c - r ) / a ] t , t E [O,T].

From Girsanov's theorem-we know that there exists an equivalent martingale measure Q which turns B into sta.ndard Brownian motion. The solution of (4.28), given by

Zt = Zo e - ~ . ~ ~ Z t + ~ E t , t E [O, T] , turns under Q into a martingale with respect to the natural Brownian filtration.

The existence of the equivalent martingale mcasurc allows for an intuitive interpretation of the Black-Scholes formula:

Assume in the Black-Scholes model that there exists a self-financing strategy (at , bt) such that the value Vt of your portfolio at timc t is given by

Vt = at& + btPt , t E [(),TI,

and that VT is equal to the contingent claim ~ ( X T ) .

Then the value of the portfolio a t time t is given by

Vt = EQ [e-r(l'-t) h ( x T ) I F ~ ] , t E [O,T], (4.29)

where EQ(A I F t ) denotes the conditional expectation of the random vari- able A, given Ft = a(B,, s < t ) , ur~der the new probability measure Q.

IJntil now, the conditional expectatiori E(A 1 Ft) = Ep(A IFt) was always defined under the original probability measure P, and so we did not indicate this dependence in our notation.

First we want to show that (4.29) is correct, and then we will use (4.29) to evaluate the price of a European call option.

Consider the discounted value process - V -e-rtb' -

t - - ePrt (at-& + btP t ) .

The Ito lemma (2.31) yields

Now we use the fact that (a t , bt ) is self-financing together with (4.27):

1 Notice that VO = Vo. R o m (4.30) and (4.28) we have

- Under the equivalent_ martingale measure Q, B is standard Brownian motion and the process (at Xt , t E [0, TI) is adapted to (Ft, t E [0, '1'1). Hence (4.31) constitutes a martingale with respect to (Ft). This is one of the basic properties of the It6 integral; see p. 111. In particular, the martingale property implies that

R = E Q ( ~ T I F ~ T , ) , t E [(),TI,

but - VT = eKTT VT = eCrT h(XT) .

Hence e-Tt vt = EQ [ e - T T h ( ~ T ) I F~] ,

or, equivalently, the value of the portfolio is given by (4.29). Now we want to calculate the value Vt of our portfolio and the Black-

Scholes price in the case of a European option.

I Example 4.2.2 (The value of a European option) -- Write

1 8 = T - t for t E [ O , T ]

I By (4.29), the value K of the portfolio at time t corresponding to the contingent clairn VT = h(XT) is given by

At time t , Xt is a. fiinction of Bt, hence a ( X t ) C F t , and so we can treat it under 5 as if it was a constant. Moreover, under Q, ET - Et is independent of 3t and has an N(O,8) distribution. An application of Rule 7 on p. 72 yields that

v, = f ( 4 Xt) , where

w f( t , z) = e-rQ lm h (z e(*-0-5u2)8+uv81'z

Page 98: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

184 CHAPTER 4.

and p(y) denot.es the standard normal density. For a European call option,

h(x) = (x - K)+ = max(0,x - K ) ,

and so

where @(x) is the standard normal distribution function,

ln(x/K) + ( r + 0.5c2)8 z1 = and zn = zl - ( ~ 8 " ~ .

at1' /2

This is exactly the formula (for t = 0) that we derived on p. 174 for the rational price of a European call option.

The price of the European put option (with h(x) = ( K - x ) + ) can be calculated in the same way with

f (t, x) = K eKrB@(-22) - x @(-21) .

Verify this formula!

Notes and Comments

The derivation of the Black-Scholes price via a change of the underlying mea- sure (also called change of numeraire) was initiated by the fundamental paper by Harrison and Pliska (1981). Since then this techriique belongs to the stan- dard repertoir of mathematical finance; see for example Karatzas and Shreve (1998), Lamberton and Lapeyre (1996) or Musiela and Rutkowski (1997). The interpretation of the Black-Scholes formula as a conditional expectation and related results can be found in any of these books.

A proof of Girsanov's theorem can be found in advanced textbooks on probability theory and stochastic processes; see for example Kallenberg (1997) or Icaratzas and Shreve (1988).

Appendix

The following theory can be found for instance in Feller (1 968), Ka.rr (1993) or L o k e (1978).

We introduce the main modes of convergence for a sequence of random variables A, Al , A2, . . ..

Convergence in Distribution

The sequence (A, , ) converges in distribution or converges weakly to the

random variable -4 (A,, 3 A) if for all bounded, cont,inunus filnctions f the relation

f ( A , ) f ( A ) 7 L + c o ,

holds.

d Notice: A, ---+ A holds if and only if for all continuity points x of thc distri- bution function FA the relation

is satisfied. If b~ is continuous then (A.l) can even be strenghtened to uniform convergence:

sup - IFA,, (2) - FA(x)I + 0, n + co.

Page 99: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

186 APPENDIX

It is also well known that convergence in distribution is equivalent to pointwise convergence of the corresponding characteristic functions:

A,, A if and only if EeitAn + EeitA for all t

Example A l . l (Convergence in distribution of Gaussian random variables) Assume that (A,) is a sequence of normal N(p,, a:) random variables.

First suppose that p, + p and a: + a2, where p and a2 are finite numbers. Then the corresponding characteristic functions converge for every t E R:

The right-hand side is the characteristic function of an N ( p , u2) random vari-

able A. Hence A, J' A. d Also the converse i% true. If we know that A, -+ A, then the characteristic

functions eitfin-0.5u=t necessarily converge for every t . From this fact we con- clude that there exist real numbers p and a2 such that p, -+ p and a: + a2. This implies that A is necessarily a normal N(p, a 2 ) random variable.

Convergence in Probability

The sequence (A,) converges i n probability to the random variable A

(A, 3 A) if for all positive E the relation

P ( I A , - A ~ > E ) + O , n + m ,

holds.

Convergence in probability implies convergence in distribution. The converse is true if and only if A = a for some constant a.

Almost Sure Convergence

The sequence (A,) converges almost surely ( a s . ) or with probability 1 to the random variable A (A, 3 A) if the set of ws with

An(w) + A ( w ) , n + m ,

has probability 1.

APPENDIX 187

This means that

P (A, + A) = P ({w : A,(uJ) + A(w))) = 1 .

Convergence with probability 1 implies convergence in probability, hence con- vergence in distribution. Convergence in probability does not imply conver-

P gence a.s. However, A, + A implies that A,, 3 A for a suitable subse- quence (nk).

LP- Convergence

(A, -+ A) if E[IA,lp + IAlp] < m for all n and

EIA,-AIP+O, n + m .

By Markov's inequality, P(IA, -AJ > E) 5 E-PEIA, - Alp for positive p and E.

Thus A, LP' A implies that A, 5 A. The converse is in general not true. For p = 2, we say that (A,) converges i n mean square to A. This notion

/' can be extended to stochastic processes; see for example Appendix A4. Mean square convergence is convergence in the Hilbert space

endowed with the inner product < X , Y >= E ( X Y ) and the norm J J X J J = d m . The symbol X stands for the equivalence class of random variables

d Y satisfying X = Y.

A2 Inequalities

In this section we give some standard inequalities which are frequently used in this book.

The Chebyshev inequality:

P ( l X - EX1 > x) < C2 var(X) , z > 0 .

Page 100: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

188 APPENDIX

The Cauchy-Schwarz inequality:

ElXYl 5 EX^)'/^ (EY' ) ' /~ .

Let f be a convex function on R. If ElXl and Elf (X)I are finite, then

f(EW I E f ( X ) .

In particular,

(EIX1q)l/q I (E(x~P)''P for 0 < q < p .

This is Lyapunov's inequality. Jensen's inequality remains valid for conditional expectations: let F be a

a-field on R. Then

~ ( E ( x I 3)) 5 E ( ~ ( x ) I a. (A.2) In particular,

A3 Non-Differentiability and Unbounded Va- riation of Brownian Sample Paths

Let B = (Bt, t 2 0) be Brownian motion. Recall the definitions of an H-self- similar process from (1.12) on p. 36 and of a process with stationary increments from p. 30. We also know that BI-ownian rioti ion is a 0.5-self-similar process with stationary, independent increments; see Section 1.3.1. We show the non- differentiability of Brownian sample paths in the more general context of self- similar processes.

Proposition A3.1 (Non-differentiability of self-similar processes) Suppose (Xi) is H-self-simailar with stationary increments for some H E (0 , l ) . Then for every fixed to,

lim sup IXt - xt, I = c o , tit,, t - to

i.e. sample paths of H-self-similar processes are nowhere differentiable with probability 1.

APPENDIX 189

Proof. Without loss of generality we choose to = 0. Let (tn) be a sequence such that t, 4 0. Then, by H-self-similarity, Xo = 0 as . , and hence

P ( n+co lim o s ~ s t , , sup l $ l > x ) = lim P ( ~ ~ ~ i $ l > r ) + O<s<t,

= lim sup P (tf-lIX1 1 > X) n-+m

Hence, with probability 1, limsup,,, ( X t , /tnl = co for any sequence t, 4 0.

Proposition A3.2 (Unbounded variation of Brownian sample paths) For almost all Brownian sample paths,

n

v(B(w)) = sup I B ~ , (w) - Bti-, (w) 1 = co a . ~ . , i=l

where the supremum is taken over all possible partitions r : 0 = to < . . . < tn = T of [O,T]. , Proof. For convenience, assume T = 1. Sllpposc! that v(B(w)) < oo for a given w. Let (r,) be a sequence of partitions r, : 0 = to < tl < . . . < tnpl < t, = 1 , such that mesh(r7,) i 0. Recall that A i B = Bt, - Bt,-,. The following chain of inequalities holds:

n

Qn(w) = C ( A ~ B ( W ) ) ~ i= 1

5 max lAiB(w)l C l A i ~ ( w ) l i=l, ..., n i=l

Since B has continuous sample paths with probability 1, we may assume that Bt(w) is a continuous function of t. It is also uniformly continuous on [O, 11 which, in combination with mesh(rn) i 0, implies that maxi,^, ...,, lAiB(w)l -t 0. Hence the right-hand side of (A.3) converges to zero, implying tha.t,

Qn(w) + 0 . (A.4)

Page 101: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

P On the other hand, we know from p. 98 that Q, --+ 1, hence Q,,(w) % 1 for a suitable subsequence (nk); see p. 187. Thus (A.4) is only possible on a null-set, and so

P({w : v(B(w)) = oo)) = 1 .

A4 Proof of the Existence of the General It6 Stochastic Integral

In this section we give a proof of the existence of the It6 stochastic integral. This requires some knowledge of the spaces L2 of square integrable functions and of measure-theoretic arguments. If you think this is too much for you, you should avoid t,his section, blit, a short glance a t the proof would be useful. We do not give all the details; sometimes we refer to some other books.

As usual, B = (Bt, t >. 0) is Brownian motion, and (3i , t > 0) is the cor- responding natural filtration. Let C = (Ct, t E [0, TI) be a stochastic process satisfying the Assumptions on p. 108, which we recall here for convenience:

Ct is a function of B,, s 5 t.

All stochastic processes are supposed to live on thc samc probability space [n, F, PI ( P is a probability measure on the c-field 3 ) .

Lemma A4.1 If C satisfies the Assumptions, then there exists a sequence ( ~ ( ~ 1 ) of simple processes (see p. 101 for the definition) such that

A proof can be found, for example, in Kloeden and Platen (1992), Lemma 3.2.1. Thus simple processes c ( ~ ) are dense in the space L2[R x [0, TI, d P x

dt]. Relat,ion (A.5) means exactly this: in the norm of this space, C can be approximated arbitrarily well by simple processes d n ) .

As usual, denote by

t

4(C'")) = l c!") dB., t E [O, TI ,

the It6 stochastic integral of the simple process C(n) . It is well defined as a Riemann-Stieltjes sum; cf. (2.1 1).

APPENDIX 191

Lemma A4.2 Under the Assumptions, the following relation holds for all I

n, k E N:

Proof. Both, I(C(")) and I(c("+~)) are martingales with respect to Brownian motion, hence I(c("+~)) - I(c(")) = I ( c ( ~ + ~ ) - c(,)) is a martingale with respect to (Ft) .

For a general martingale (M, (3t ) ) on [0, TI, Doob's squared maximal in- equality (see for example Dellacherie and Meyer (1982), Theorem 99 on p. 164) tells us that

E sup M; 5 4 EM$. OStST

In particular,

By the isometry property (2.14), applied to the simple process I (C(n+k)-C(n)) , the right-hand side becomes

which, together with (A.7), proves (A.6).

Since (A.5) holds, ( ~ ( " 1 ) is a Cauchy sequence in L2[R x [0, TI, d P x dt] . This implies, in particular, that the right-hand side in (A.6) converges to zero. Thus we can find a suhseq11enc.e (k,), say, such that

This and the Chebyshev inequalily or1 p. 187 imply that, for every E > 0, I 1 p ( sup ~ I t ( c ( ~ n + l ) ) - ~ t ( c ( ~ " ) ) l > E )

O S l T

Page 102: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

192 APPENDIX

.4n appeal to the Borel-Cantelli lenima (see for example Karr (1993)) yields that the processes I (C(")) converge uniformly on [0, TI to a stochastic process I ( C ) , say, with probability 1. Since the (uniformly) converging processes have coritinuous sample paths, so has the limiting process I ( C ) .

Now, letting k in (A.6) go to infinity, we conclude that

E sup [I t (C) - It (Ccn')]' 5 4 E O < t < T

By Lemma A4.1, the right-hand side converges to zero as n -t oo, and so we have proved that the limit I ( C ) does not depend on the particular choice of the approximating sequence ( c ( " ) ) of simple processes.

The limit process I (C) is the desired It6 stochastic integral process, denoted by

t

4(c) =l CsdBs, t E [().TI.

Its sample paths are continuous with probability 1.

L e m m a A4.3 The pair ( I ( C ) , ( F t ) ) constitutes a martingale.

Proof . We know that ( I ( C ( ~ ) ) , ( F A ) ) is a martingale (see p. 104) for all n. In particular, it is adapted to (F t ) . This does not change in the limit. Moreover, for all n,

E ( I ~ ( c ( " ) ) I F,) = I , (dn)) for s < t .

Using Lebesgue dominated convergence and, if necessary, passing to a sub- sequence (k,) such that I ~ ( C ( ~ * ) ) converges to It(C) uniformly for t , with probability 1, we obtain

lim ~ ( 1 , (c(")) )I Fs) = E( lim It(C("-') 1 Fs) = E(I t (C) I Fs) n-&w n+m

Hence I (C) is a martingale.

L e m m a A4.4 The It6 stochastic integral I ( C ) satisfies the isometry property

APPENDIX 193

Proof. We know that I(c(")) satisfies this property; see (2.14). By (A.6), (A.5) and since the norm is continuous,

EIIt (C'"')]' = E [ I ~ (c(".' - C ) + 1, (C)]" E[1t(C)12 + oil) ,

~ [ C i " ' ] ~ d s = ~ b f E[(C;"' - CS) + Cs]%ds = J: EC,? ds + o(1).

Since the two left-hand sides coincide, by the isometry property for I ( C ( ~ ) ) , we obtain as n -t rn the desired isometry relation (A.8).

L e m m a A4.5 For a simple process C , the particular Riemann-Stieltjes sums in the definition (2.1 1 ) coincide with the It6 stochastic integral I ( C ) .

The proof makes use of the fact that simple processes C(n) axe dense in the underlying L"space.

A5 The Radon-Nikodym Theorem

Let [ f l ,Fr] be a measurable space, i.e. 3' is a u-field on 0. Consider two measures p and v on 3'. We say that p is absolutely continuous with respect to v ( p << v ) if

for all A E .Fr: v (A) = 0 implies that p(A) = 0.

We say that p and v are equivalent measures if p << v and v << p.

T h e Radon-Nikodym Theorem:

Assume p and I/ are two u-finite measures. (This is in particular satisfied if they are probability measures.) Then p << v holds if and only if there exists a non-negative measurable function f such that

p ( ~ ) = 1 f ( ~ ) ~ u ( u ) , A E 3'. A

If this relation holds with f replaced by another non-negative function g , then

v ( f # 9) = 0 .

The v-almost everywhere unique fiinction f is called the density of p with respect to v.

For a proof of this result, see Billingsley (1995) or Dudley (1989), Section 5.5.

Page 103: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

194 APPENDIX

A6 Proof of the Existence and Uniqueness of the Conditional Expectation

Let [ R , 3 , P] be a probability space, i.e. 3 is a u-field on R and P is a probability measure on F. Consider the u-field 3' c 3 and denote by P' the restriction of P to F'. Let X be a random variable on R.

Theorem A6.1 If ElXl < cm, then there exists a random variable Z such that

(a) u(Z) c 3' and

( b ) for all A E F 1 ,

If (a) and ( b ) hold with Z replaced by another random variable Z' then

W e call Z the conditional expectation of X given the u-field 3', and we write Z = E ( X I 3 ) .

Proof. A) Assume that X 2 0 P-a.s. Consider the (smaller) probability space [R, F ' , P'] and define the measure

v(A) = X(w) dP(w) , A E 3'. L If P(A) = 0, then X I A = 0 P-a.s., and so v(A) = E ( X I A ) = 0. Hence v << P', where both measures, P' and v, are defined on 3'. By virtue of the Radon- Nikodym theorem on p. 193, there exists a non-negative random variable Z on [fl, F'] such that

v(A) = Z(W) dPi(w) , A 6 3' , /A

and the random variable Z is unique in the sense of (A.9). Moreover, recalling the definition of the integral and using that u(Z) c 3', one can show that S, Z(w) dP1(w) = SA Z(w) dP(w).

B) Assume that ElXl < cm. Write X = X + - X-, where X + = max(0, X ) and X - = - min(0, X) . Apply the first part of the theorem to Xf and X - separately. It yields two nun-negative random variables Z+ and Z- such that (a) and (b) hold if we replace Z by them. Moreover, EX* = EZ* < cm since ElXl = E X f + E X - < cm. Hence Z* < cm P-a.s., and so we can define Z = Z+ - Z- which is the desired random variable.

I [I] Arnold, L. (1973) Stochastische Dzfferentialgleichungen. Theorie und An- wendungen. R. Oldenbourg Verlag, Miinchen, Wien. [I291

I [2] Ash, R.B. and Gardner, M.F. (1975) Topics i n Stochastic Processes. Aca- demic Press, New York. [33]

I , [3] Bachelier, L. (1900) ThCorie de la spkculation. Ann. Sci. ~ c o l e Norm. Sup. 111-17, 21-86. Translated in: Cootner, P.H. (Ed.) (1964) The Ran- dom Character of Stock Market Prices, pp. 17-78. MIT Press, Cambridge, Mass. [35',42,175]

I [4] Baxter, M. and Rennie, A. (1996) Financial Calculus. A n Introduction to Derivative Pricing. Cambridge University Press, Cambridge. [I '761

I [5] Billingsley, P. (1968) Convergence of Probability Measures. Wiley, New York. [56]

I [6] Billingsley, P. (1995) Probability and Measure, 3rd edition. Wiley, New York. [V, 1931

I [7] Black, F. and Scholes, M. (1973) The pricing of options and corporate liabilities. J. Political Economy 81, 635-654. [I ,42,167,1'75]

I [8] Borodin, A.N. and Salminen, P. (1996) Handbook of Brownian Motion - Facts and Formulae. Birkhauser, Basel. [52]

I [9] Chung, K.L. and Williams, R.J. (1990) Introduction to Stochastic Inte- gration, 2nd edition. Birkhauser, Boston. [112,122,136,137]

[lo] Ciesielski, Z. (1965) Lectures on Brownian Motion. Heat Conduction and Potential Theory. Aarhus University Lecture Notes Series, Aarhus. [56]

Page 104: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

[ll] Dellacherie, C. arid Meyer, P.-A. (1982) Probabilities and Potential B. North-Holland, Amsterdam. [I 911

[12] Doob, J.L. (1953) Stochastic Processes. Wiley, New York. [I121

[13] Dudley, R.M. (1989) Real Analysis and Probability. Wadsworth, Brooks and Cole, Pacific Grove. [I931

[14] Dudley, R.M. and NorvaiSa. R. (1998a) Product integrals, Young integrals and yvariation. Lecture Notes in Matherr~atics. Sprirlger, New York. Tu appear. [94,96/

[15] Dudley, R.M. and NorvaiSa, R. (1998b) A survey on differentiability of six operators in relation to probability and statistics. Lecture Notes in Mathematics. Springer, New York. To appear. [94,96]

[16] Duffie, D. (1996) Dynamic Asset Pricing Theory. Princeton University Press. [I 761

[17] Einstein, A. (1905) On the movement of small particles suspended in a stationary liquid de~narlded by the molecular-kinetic theory of heat. Ann. Phys. 17. [33]

[18] Feller, W. (1968) An Introduction to Probability Theory arid Its Applica- tions I,ZI, 3rd edition. Wiley, New York. [I851

[19] Gikhman, 1.1. and Skorokhod, A.V. (1975) The Theory of Stochastic Pro- cesses. Springer, Berlin. 1331

[20] Grimett, G.R. and Stirzaker, D.R. (1994) Probability and Random Pro- cesses. Clarerldor~ Press, Oxford. [33]

[21] Gut, A. (1995) An Intermediate Course in Probability. Springer, Berlin. [23]

[22] Harrison, M.J. and Pliska, S.R. (1981) Martingales and stochastic inte- grals in the theory of continuous trading. Stoch. Proc. Appl. 11, 215- 260. [I841

[23] Hida, T. (1980) Brownian Motion. Springer, New York. [52]

[24] Ikeda, N. and Watanabe, S. (1989) Stochastic Differential Equations and Diffusion Processes, 2nd edition. North-Holland, Amsterdam. [112,122]

[2G] It6, K. (1944) Stochastic integral. Proc. Imp. Acad. Tokyo 20, 519- 524. [I121

[27] It6, K. (1951) On a formula concerning stochastic differentials. Nagoya Math. J. 3, 55-65. [I221

[28] Kallenberg, 0. (1997) Foundations of Modern Probability. Springer, New York. [I 841

[29] Karatzas, I. and Shreve, S.E. (1988) Brownian Motior~ and Sloc1~asl.i~ Calculus. Springer, Berlin. [56,85,112,122,176,178,184]

I [30] Karatzas, I. and Shreve, S.E. (1998) Methods of Mathematical Finance. Springer, Heidelberg. To appear. [I 76,1841

I [31] Karlin, S. and Ta,ylnr, H.M. (1975) A First Course in Stochastic Processes. Academic Press, New York. [33]

I [32] Karlin, S. and Taylor, H.M. (1981) A Second Course in Stochastic Pro- cesses. Academic Press, New York. [33]

1 1331 Karr, A.F. (1993) Probability. Springer, New York. [185,192]

I [34] Kloeden, P.E. and Platen, E. (1992) Numerical Solution of Stochastic Differential Equations. Springer, Berlin. [I 37,166,190]

I [35] Kloeden, P.E., Platen, E. and Schurz, H. (1994) Numerical Solution of SDE throhgh Computer Experiments. Springer, Berlin. [I661

I [36] Lamberton, D. and Lapeyre, B. (1996) Introduction to Stochastic Calculus Applied to Fznance. Chapman and Hall, London. [112,153,176,184]

I 1371 Langevin, P. (1908) Sur la thborie du mouvement brownien. C.R. Acad. Sci. Paris 146, 530-533. [I411

[38] Logve, M. (1978) Probability Theory I,ZI, 4th edition. Springer, Ber- lin. [I 851

[39] Mendenhall, W., Wackerly, D.D. and SchealTer-, R.L. (1990) Mathematical Statistics with Applications, 4th edition. PWS-Kent Publishing Company, Boston. [23]

[25] It6, K. (1942) Differential equations determining Markov processes (in Japanese). Zenkoku Shzjo Sugaku Danwakai 244, 1352-1400. (1 121

Page 105: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

198 BIBLIOGRAPHY I [41] Musiela, M. and Rutkowski, M. (1997) Martingale Methods i n Financial

Modelling. Springer, Berlin. [I 76,1841

[42] Bksendahl, B. (1985) Stochastic Differential Equations, an Introduction with Applications. Springer, Berlin. [112,122,137]

[43] Pitman, J . (1993) Probability. Springer, Berlin. [22]

[44] Pliska, S.R. (1997) Introduction to Mathematical Finance. Discrete Time Models. Blackwell, Maldcn (MA). [I 761

[45] Pollard, D. (1984) Convergence of Stochastic Processes. Springer, Ber- lin. [56]

[46] Protter, P. (1992) Stochastic Integration and Differential Equations. Springer, Berlin. [96,112,122,136]

[47] Resnick, S.I. (1992) Adventures i n Stochastic Processes. Birkhauser, Boston. [33]

[48] Revuz, D. and Yor, M. (1991) Continuous Martingales and Brownian Motion. Springer, Berlin. [56,85]

[49] Shorack, G.R. and Wellner, J.A. (1986) Empirical Processes with Appli- cations to Statistics. Wiley, New York. [dl]

[50] Stratonovich, R.L. (1966) A new representation for stochastic integrals and equations. S I A M J. Control 4 , 362-371. [I291

[51] Taylor, S.J. (1972) Exact asymptotic estimates of Brownian path varia- tion. Duke Math. J. 39, 219-241. [95]

[52] Wiener, N. (1923) Differential space. J. Math. Phys. 2 , 131-174. [33]

[53] Williams, D. (1991) Probability with Martingales. Cambridge University Press, Cambridge, UK. [77,85]

[54] Willmot, P., Howison, S. and Dewynne, J. (1995) The Mathematics of Financial Derivatives. Cambridge University Press. [l 761

[55] Young, L.C. (1936) An inequality of the Holder type, connected with Stieltjes integration. Acta Math. 67, 251-282. [94,96]

[56] Zauderer, E. (1989) Partial Dafferential Equations of Applied Mathemat- ics, 2nd edition. Wiley, Singapore. [I741

Index

Absolute continuity of two mcasurcs 177, 193 - density function 177, 193 - equivalent measures 177,193 - Radon-Nikodym theorem

193 Adapted stochastic process 77

-with respect to the Brownian filtration 78

Adapted to Brownian motion 78 Additive noise

in a stochastic differential equation 141, 150

Almost sure (a.s.) convergence 186 American call

- see option Arbitrage 171, 175 ARMA process

(autoregressive moving avera- ge)j25

a.s. (dmosl sure, almost surely) Autocorrelation 22 Autoregressive process 26

Binomial distribution 9 Black and Scholes model 168, 180

- arbitrage 171, 175 - bond 169 - consumption 169 - contingent claim 172, 181 - and geometric Brownian mo-

tion 168 - hedging 172 - mean rate of return 168 - option

- - American call 170 - - European call 170

- - European put 170 - - exercise (strike) price

170 - - rational price of an op-

tion 171 - - time of maturity (expi-

ration) 170 - portfolio 169 - trading strategy 169 - self-financing strategy 169 - short sale 169 - stock 168 - transaction costs 169 - volatility 168

Black and Scholes option pricing formula 174 - as conditional expectation of

the discounted contingent claim 182,183

- rational price of an option 171

Bond 169 Borel set

- see Borel a-field Borel a-field 64 Borel Cantelli lemma 192 Bounded variation 39

- and Brownian motion 39 - pvariation 94 - and Riemann-Stieltjes inte-

gral 92 Brownian bridge 40 I3rownian filtration 78 Brownian motion 33

- Brownian bridge 40 - covariance function 35 - distribution 35

Page 106: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

- and Donsker invariance prin- ciple (FCLT) 47

- with drift 41 - expectation function 35 - fiinctional central limit the-

orem 47 - geometric Brownian motion

42 - as a martingale 82 - non-differentiable sample

paths 36, 188 - path properties 36

- - non-differentiability 36 - - unbounded variation 39

- and pvariation 95 - quadratic variation 98 - self-simila,rit,y 36 - simulation 44 - unbounded variation 39, 189

Brownian motion with drift 41

Call option - see option

Cameron-Martin formula - see Girsanov's theorem

Cauchy-Schwarz's inequality 188 Chain rule

- classical 113 - It6 lemma 114, 117 - transformation formula for

Stratonovich integral 126, 128

Change of measure 170; - absolute continuity of mea-

sures 177, 193 - and the Black-Scholes for-

mula 182 - density function 177, 193 - equivalent martingale mea-

sure 179 - equivalent measures 177, 193 - Girsanov's theorem 178 - Radon-Nikodym theorem

193 Central limit theorem (CLT) 45

- FCLT 47 Chebyshev's inequality 13, 187 CLT (central limit theorem) 45

INDEX

Colored noise 44 Conditional distribution function

5 7 Conditional expectation 56, 67

- classical 58 - conditional distribution

function 57 - conditional probability 56 - under discrete condition

56, 68 - existence and uniqueness 194 - and martingales 80 - projection property 74 - rules for calculation 70 - given a u-field 68

Conditional probability 56 Consumption 169 Contingent claim 172, 181 Continuous distribution 9

- density 11 Continuous random variablc 9 Continuous-time stochastic process

23 Convergence modes 185

- almost sure (a.s.) convergen- ce 186

- convergence in distribution 185

- convergence inJp 187 - - mean square convergence

187 convergence in probability

186 - convergence with probability

one 186 - weak convergence 185

Correlation 18 - and dependence 21

- - and Gaussianity 21 Covariance 18 Covariance function

of a stochastic process 28 Covariance matrix

of a random vector 18

Density (probability density) - Gaussian (normal)

- - multivariate 17

INDEX

- - one-dimensional 11 - of a rando~n variable 11 - of a random vector 16

- - marginal density 16 Density function

- and absolute continuity 177, 193

- equivalent measures 177, 193 - Radon-Nikodym theorem

193 Differential equation

- deterministic 132 - initial condition 134 - It6 stochastic differential

equation 134 - random differential equation

135 - solution 132

- - strong 137 - - weak 137

- Stratonovich stochastic dif- ferential equation 145

Diffusion 137 Discounted price 181 Discrete distribution 9 Discrete random variable 9

- conditional expectation given a discrete random variablc 56

- u-field generated by a dis- crete random variable 64

Discrete-time stochastic process 23 - time series 25

Distribution - binomial 9 - continuous 9

- - density 11, 16 - discrete 9 - exponential 32 - Gaussian, normal - - multivariate 17 - - one-dimensional 11

- Poisson 9 - of a random variable 8 - of a random vector 16 - of a stochastic process 27

- - fidis 27 - symmetric 21

- uniform 11 Distribution function

- of a random variable 8 - of a random vector 15

Donsker invariance principlc 47 Doob's inequality 191 Drift

- Brownian motion with drift 4 2

Driving process of a SDE 136 - SDE with two driving pro-

cesses 144

Equivalent measures 177, 193 - and the Black-Scholes for-

mula 182 - equivalent martingale mea-

sures 179 - - Girsanov's theorem 178

Euler approximation 158 European calllput option

- see option, Black-Scholes Exercise price

- see option Expectation

- of a random variable 12 - of a random vector 17

Expectation function of a stochastic process 28

Exponential distribution 32

Fair game - and martingales 84

FCLT (functional central limit the- orem) 47

Fidis (finite-dimensional distributi- ons) 27

Filtration 77 - adapted stochastic process

77 Brownian 78

- natural 77 Finance

- see Black and Scholes Finite-dimensional distributions

(fidis) of a stochast,ic. process 27

- of a Gaussian process 27

Page 107: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

Functional central limit theorem (FCLT) 47 - Donsker invariance principle

4 7

Gaussian (normal) distribution - multivariate

- - density 17 - - with independent com-

ponents 21 - - under linear transforma-

tions 18 - one-dimensional

- - density 11 - standard normal 11

Gaussian process 27 - Brownian bridge 40 - Brownian motion 33 - Brownian motion with drift

4 1 - colored noise 44 - fidis 27 - Ornstein-Uhlenbeck process

142, 152 - self-similar 36 - stationary 30 - Vasicek interest model 152 - white noise 44

Gaussian random variable, random vector

- see Gaussian distribution General linear stochastic differen-

tial equation 150 - with additive noisc 151 - expectation and covariance

functions 156 - homogeneous 153 - with multiplicative noise 153 - solution 155

Generation of a-fields from a collecliori of sets 63

Geometric Brownian motion 42 - and the Black-Scholes for-

mula 168 - covariance function 43

expectation function 42 - and It6 stochastic differential

equation with multiplica-

INDEX

tive noise 118, 139 Girsanov's theorem 178

- equivalent martingale mea- sure 179

Haar functions 51 - Lkvy representation of Brow-

nian motion 51 Hedging 172 Homogcncous lincar stochastic dif-

ferential equation 153 Homogeneous Poisson process

- see Poisson process

iid (independent and identically distributed) 22

Increment of a stochastic process - process with independent in-

crements 30 - process with stationary in-

crements 30 - proccss with stationary, in-

dependent increments - - Brownian motion 33 - - Poisson process 32

Independence - of events 20 - of Gaussian variables 21 - - and ur~correlatedriess 21

- of random variables 20 - of a sequence of random vari-

ables 22 - - iid 22

Inequalities 187 - Cauchy-Schwarz 188 - Chebyshev 187 - Doob 191 - Jensen 188

Information - carried by a random variable,

a random vector or a sto- chastic process 66

Initial condition - of a differential equation 134 - of a random differential equa-

tion or a SDE 135 Integral equation 134 Integrals

- ordinary (Riemann) 88 - Riemann-Stieltjes 92 - stochastic 87 - - It6 96 - - Stratnnovich 123

Integration by parts formula 122 - extended It6 lemma 121

Intensity of a Poisson process 32 Interest rate

- Vasicek model 152 Intermediate partition 88 Isometry property

of the It6 integral - for general processes 11 1 - for simple processes 106

It6 exponential 116, 118 It6 formula

- see It6 lemma It6 integral

- see It6 stochastic integral It6 lemma

(It6 chain rule, It6 formula) - and classical chain rule 113 - extended versions

117, 120, 121 - and integration by parts 122 -and It6 stochastic differential

equations 138 - simple version 114

It6 process 119 It6 stochastic differential equation

136 - driving process 136 - equivalent Stratonovich dif-

ferential equation 146 - initial condition 135 - and It6 integral equation 136 - linear 138

- - with additive noise 141, 150

- - geometric Brownian mo- tion 139

- - homogeneous 153 - - with multiplicative noise

139, 153 - - Ornstein-Uhlenbeck pro-

cess 142, 152 - solution

- - existence and uniqueness 138

- - strong 137 - - weak 137

- with two independent driv- ing processes 144

It6 stochastic integral 96 - existence 190 - for general processes 107

- - assumptions on the inte- grand process 108

- - properties 111, 190 - isometry 106, 111 - It6 lemma 114, 117 - as a martingale 104, 11 1 - relation to Stratonovich sto-

chastic integral - - transformation formula

126, 128 - for simple processes 101

- - properties 105

Jensen's inequality 188

Langevin equation 141, 152 Law of large numbers 7 Lkvy representation

of Brownian motion 51 - Haar functions 51 - Schauder functions 52

Lkvy-Ciesielski representation of Brownian motion 52

Linear stochastic differential equation 138 - with additive noise 141, 150 - general 150

- - with additive noise 150 - - expectation and covari-

ance functions 156 - - homogeneous 153 - - with multiplicative noise

153 - - solution 155

- geometric Brownian motion 139

- Langevin equation 141 - with multiplicative noise

139. 153

Page 108: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

- Ornstein-Uhlenbeck process 142, 152

- Vasicek model 152 Lipscliitz cor~clition 138 Log-return 10

Marginal density 16 Martingale 77

-adapted stochastic process 77 - continuous-time 80 - discrete-time 80 - equivalent martingale measu-

re 179 - - Girsanov's theorem 178

- examples - - partial sums of indepen-

dent random variables 81 - - Brownian motion 82 - - It6 stochastic integral

104, 111 - - martingale transform 82

- as a fair game 84 - filtration 77

- - Brownian 78 - - natural 77

Martingale difference sequence 80 Martingale transform 82

- Brownian 83 - predictable (previsible)

sequence 83 Maturity

- time of maturity (expiration) of an option 170

Mean, mean value - see expectation

Mean rate of return 168 Mean reversion

- see Vasicek interest model Mean square convergence 187 Measure

- see probability measure Mesh of a partition 89 Milstein approximation 162

- Taylor-It,o expansion 163 Modes of convergence

- see convergence modes Moment

of a random variable 12

INDEX

Moving average process 25 Multiplicative noise

of a linear stochastic differen- tial equation 139, 153

Natural filtration 77 Normal distribution

- see Gaussian distribution Numerical solution of a stochastic

differential equation 157 - Euler approximation 158 - Milstein approximation 162 - order of convergence 162 - strong 161 - Taylor-It8 expansion 163 - weak 161

Option - American call 170 - Black and Scholes formula

174, 182, 183 - contingent claim 172, 181 - European call 170 - European put 170 - exercise (strike) price 170 - time of maturity (expiration)

170 Order of convergence

of a numerical solution to a SDE

- strong solution 162 - weak solution 162

Ornstein-Uhlenbeck process 142, 152

- covariance function 144 - as a Gaussian process 143

Outcome space 6

Paley-Wiener representation of Brownian motion 51

Partition of an interval 88 - intermediate partition 88 - mesh of a partition 89

Poisson distribution 9 Poisson process 32

- intensity (rate) 32 Portfolio 169 Power set 63

Predictable (previsible) sequence 83 - martingale transform 82

Previsible - see predictable

Probability, probability measure 7 - change of measure 176

Projection property of the conditional expectation

74 - best prediction 74

Put option - see option

pvariation of a function 94 - of Brownian motion 95

Quadratic variation of Brownian motion 98

Radon-Nikodym theorem 193 - absolute continuity of two

measures 177, 193 - density function 193 - equivalent measures 177, 193

Random differential equation 135 Random variable 6

- continuous 9 - density 11 - discrete 9 - distribution 8 - distribution function 8 - expectation 12 - nioment 12 - standard deviation 13 - variance 12

Randorn vector 14 - correlation 18 - covariance matrix 18 - density 16 - distribution 16 - distribution function 15 - expectation 17 - Gaussian 17

Rate of return 168 Rational price of an optiorl

- see Black-Scholes Realization of a stochastic process

- see sample path Rectangle in Rn 65

Return - log-return 10 - mean rate of return 168 - relative 168 - volatility 168

Riemann (ordinary) integral 88 - properties 91 - Riemanri sum 88

Riemann-Stieltjes integral 92 - and bounded variation 93 - Riemann-Stieltjes sum 93

Riemann-Stieltjes sum 93 Riemann sum 88 Rules for the calculation

of conditional expectations 70

Sample path of a stochastic process

(trajectory) 23 Schauder functions 52

- Levy representation of Brow- nian motion 51

SDE (stochastic differential equati- on) 131

Self-financing strategy 169 Self-similar process 36

- Brownian motion 36 - non-differentiable sample

paths 188 Separation of variables 133 Series representations

of Brownian motion - Levy 51 - LCvy-Ciesielski 52 - Paley-Wiener 51

Short sale 169 u-algebra

- see u-field u-field 6, 62

- Bore1 u-field 64 - conditional expectation

given a u-field 67, 68 - filtration 77 - generated from a collection

of sets 63 - generated by a

- - discrete variable Y 64 - - random vector Y 65

Page 109: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

- - stochastic process Y 66 - information contained in a 0-

field 66 Simple process 101 Simulation

of Brownian motion 33 - via the FCLT 47

via the LCvy representation 51

- via the LCvy-Ciesielski rep- resentation 52

- via the Paley-Wiener repre- sentation 51

Solution of a differential equation 132

Solution of a stochastic differential

equation - existence and uniqueness 138 - numerical 157

- - strong 161 - - weak 161

- - strong 137 - - weak 137

S&P index 10, 22, 26, 25 Standard Brownian motion 33 Standard deviation 13 Standard normal distribution 11 Stationary increment process 30

- process with independent in- crements 30

- - Brownian motion 33 - - Poisson process 32

Stationary stochastic process - Gaussian 30 - strictly stationary 29 - in the wide sense

(second order stationary) 30

Stochastic chain rule - see It6 lemma

Stochastic differential equation

- see It6 SDE - see Stratonovich SDE

Stochastic integral 87 - It6 96 - Stratonovich 123

INDEX

Stochastic integral equation - see stochastic differential

equation Stochastic process 23

- contin~lous-time 23 - covariance function 28 - discrete-time 23 . - distribution 27 - expectation function 28 - finite-dimensional distributi-

ons (fidis) 27 - Gaussian 27

- - Brownian bridge 40 - - Brownian motion 33 - - Brownian motion with

drift 41 - - Ornstein-Uhlenbeck

process 142, 152 - geometric Brownian motion

42 - with independent increments

30 - Poisson process 32 - realization 23 - sample path 23 - self-similar 36 - stationary

- - Gaussian 30 - - strictly stationary 29 - - in the wide sense

(second order stationary) 30

- with stationary increments 30

- with stationary, independent increments

- - Brownian motion 33 - - Poisson process 32

- time series 25 - - ARMA process 25 - - autoregressive process 26 - - moving average process

25 - trajectory 23 - variance function 29

Stock 168 Stratonovich exponential 127 Stratonovich integral 123

INDEX

- and chain rule 126 - relation to It8 stochastic in-

tegral - - transformation formula

126, 128 Stratonovich stochastic differential

equation 145 - equivalent It6 stochastic dif-

ferential equation 146 Strictly stationary stochastic pro-

cess 29 Strike price

- see option Strong numerical solution

of a stochastic differential equation 161

- order of convergence 162 Strong solution

of a stochastic differential equation 137

Symmetric distribution 21

Taylor-Itb expansion 163 Tied down Brownian motion

- see Brownian bridge Time of maturity

- see option Time series 25

- ARMA process 25 - autoregressive process 26 - moving average process 25

Trading strategy 169 - self-financing strategy 169

Trajectory of a stochastic process - see sample path

Transaction costs 169 Transformation formula

for It6 and Stratonovich inte- grals 126, 128

Uniform distribution 11

Variance of a random variable 12

Variance function of a stochastic process 29

Variation

- see bounded variation, pva- riation, quadratic variati- on

Vasicek interest rate model 152 - expectation and covariance

functions 153 Volatility 168

Weak convergence 185 Weak solution

of a stochastic differential equation 137

Weak numerical solution of a stochastic differential

equation 161 - order of convergence 162

Wealth - see portfolio

White noise 44 Wiener process

- see Brownian motion

Page 110: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

-

<

List of Abbreviations and Symbols

I have tried as much as possible to use uniquely defined abbreviations and .' symbols. In various cases, however, symbols can have different meanings in

different sections. The list below gives the most typical usage. Commonly used mathematical symbols are not explained.

Symbol Explanation P.

B , Bt Brownian motion 33 Bin(n,p) binomial distribution with parameters n and p 9 C LT central limit theorem 4 5 corr(X, Y ) correlation of the random variables S and Y 18 cov(,Y, Y) covariance of the random variables X and 5' 18 cx ( t , s ) covariance function of the process X 28 A i length ti - ti-, of the interval [ti-1, t i ]

Ai f increment of the function f on [tipl , ti]: Aif = f (ti) - f ( t i - 1 )

EX expectation of the random variable X 12 EX expectation of the random vector X 17 E(X I Y ) conditional expectation of the random variable X, given

the random variable, the random vector or the stochastic process Y 69

E ( S I F ) conditional expectation of 1he rarldorri variable X given the 0-field 3 68

E x p ( X ) exponential distribution with parameter X 32 FCLT functional central limit theorem 47 fidis finite-dimensional distributions of a stochastic process 27 f i , f i j l f i jk partial derivatives of f with respect to the ith, i, j th,

i, j , kth variables

Page 111: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

210 SYMBOLS

F, F ' , 3* f x f x F ;Y

Fx I A I (C) iid inf

a-fields density of the random variable X density of the random vector X distribution function of a random variable distribution function of the random variable X distribution function of the random vector X indicator function of the set (event) A It6 stochastic integral of the process C indcpcndcnt, idcntically distributed for a sequence of real numbers a,, inf a, = - sup,(-a,). The supremum is defined below. For a finite index set I, infnEI a, = minIEn a,. logarithm with basis e space of (equivalence classes of) random variables Z with ElZlP < oo space of (equivalence classes of) random variables Z with E Z 2 < oo and ( ~ ( 2 ) c 3 for a given u-field F on R. mesh of the partition r expectation of the random variable X expectation function of the stochastic process X expectation of the random vector X set of the positive integers set of the non-negative integers Gaussian (normal) distribution with mean p and variance a2 multivariate Gaussian (normal) distribution with mean p and covariance matrix C standard normal distribution w E R outcome of the space R outcome space empty set probability space probability measure probability of the event A distribution of the random variable X or the stochastic process X distribution of the random vector X power set of R density of the standard normal distribution standard riormal distribution function

SYMBOLS

I Poi (A) lit 1W" s (C) 0:

I ffi (t) Ex

I

a(Y)

-' U(a, b) var (X ) X X Z - -

Poisson distribution with parameter X real line n-dimensional Euclidean space Stratonovich integral of the process C variance of the random variable X variance function of the process X covariance matrix of the random vector X a-field generated by the collection of sets C a-field generated by a random variable, a random vector or a stochastic process Y a = sup, a,: the supremum of a sequence of real numbers a,. If a E R, a > a, for all n and for every E > 0 there exists a k such that a - E < ak. If a = oo, then for every M > 0 there exists a k such that ak > M. For a finite index set 1, slipnEI a, = m a x , ~ ~ a,. uniform random variable on (a, b) variance of the random variable X random variable or stochastic process random vector set of integers a(x) NN b(x) means that a(x) is approximately (roughly) of the same order as b(x). It is only used in a heuristic sense. integer part of s fractional part of x x+ = max(0, x) x- = - min(0, x) transpose of the vector x complement of the set A A, 3 A: a.s. convergence

d A, .--, A: convergence in distribution - L P A, + A: convergence in LP, pth mean convergence L2 A, -+ A: convergence in L2, mean square convergence P A, .--, A: convergence in probability

d A = B: the random elements (random variables, random vectors, stochastic processes) A and B have the same distribution, i.e. P ( A E C ) = P ( B E C) for all suitable sets C. For random variables and random vectors A, B this means that their distribution functions are the s a c .

Page 112: Elementary Stochastic Calculus With Finance in View_thomas Mikosch (World 1998 212p)

# A the number of elements in the set A C B: martingale transform