Stochastic Processes II/ Wahrscheinlichkeitstheorie III. Lecture Notes

Similar documents
1. Stochastic Processes and filtrations

Stochastic integration. P.J.C. Spreij

Stochastic Processes. Winter Term Paolo Di Tella Technische Universität Dresden Institut für Stochastik

Fundamental Inequalities, Convergence and the Optional Stopping Theorem for Continuous-Time Martingales

Lecture 21 Representations of Martingales

Lecture 19 L 2 -Stochastic integration

Part III Stochastic Calculus and Applications

Risk-Minimality and Orthogonality of Martingales

Brownian Motion. 1 Definition Brownian Motion Wiener measure... 3

Dynkin (λ-) and π-systems; monotone classes of sets, and of functions with some examples of application (mainly of a probabilistic flavor)

STOCHASTIC CALCULUS JASON MILLER AND VITTORIA SILVESTRI

Brownian motion. Samy Tindel. Purdue University. Probability Theory 2 - MA 539

Lecture 12. F o s, (1.1) F t := s>t

An essay on the general theory of stochastic processes

n E(X t T n = lim X s Tn = X s

Doléans measures. Appendix C. C.1 Introduction

16.1. Signal Process Observation Process The Filtering Problem Change of Measure

Lecture 22 Girsanov s Theorem

Chapter 1. Poisson processes. 1.1 Definitions

Some SDEs with distributional drift Part I : General calculus. Flandoli, Franco; Russo, Francesco; Wolf, Jochen

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

Exercises Measure Theoretic Probability

Jump Processes. Richard F. Bass

A Concise Course on Stochastic Partial Differential Equations

Solutions to the Exercises in Stochastic Analysis

The Lebesgue Integral

Applications of Ito s Formula

Lecture 17 Brownian motion as a Markov process

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Filtrations, Markov Processes and Martingales. Lectures on Lévy Processes and Stochastic Calculus, Braunschweig, Lecture 3: The Lévy-Itô Decomposition

Stochastic Calculus (Lecture #3)

Verona Course April Lecture 1. Review of probability

3 Integration and Expectation

4th Preparation Sheet - Solutions

Probability Theory II. Spring 2016 Peter Orbanz

GENERALIZED COVARIATION FOR BANACH SPACE VALUED PROCESSES, ITÔ FORMULA AND APPLICATIONS

On pathwise stochastic integration

(B(t i+1 ) B(t i )) 2

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Notes 1 : Measure-theoretic foundations I

Solution for Problem 7.1. We argue by contradiction. If the limit were not infinite, then since τ M (ω) is nondecreasing we would have

Martingales, standard filtrations, and stopping times

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

A D VA N C E D P R O B A B I L - I T Y

Stochastic Calculus. Alan Bain

A NOTE ON STOCHASTIC INTEGRALS AS L 2 -CURVES

1 Stochastic Dynamic Programming

Measure and integration

Additive Summable Processes and their Stochastic Integral

MATHS 730 FC Lecture Notes March 5, Introduction

THEOREMS, ETC., FOR MATH 515

The Skorokhod problem in a time-dependent interval

SUMMARY OF RESULTS ON PATH SPACES AND CONVERGENCE IN DISTRIBUTION FOR STOCHASTIC PROCESSES

Introduction and Preliminaries

Random Process Lecture 1. Fundamentals of Probability

Stochastic Processes III/ Wahrscheinlichkeitstheorie IV. Lecture Notes

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

µ X (A) = P ( X 1 (A) )

Spring 2014 Advanced Probability Overview. Lecture Notes Set 1: Course Overview, σ-fields, and Measures

Part II Probability and Measure

PROBABILITY: LIMIT THEOREMS II, SPRING HOMEWORK PROBLEMS

Part III. 10 Topological Space Basics. Topological Spaces

PROBABILITY: LIMIT THEOREMS II, SPRING HOMEWORK PROBLEMS

CONVERGENCE OF RANDOM SERIES AND MARTINGALES

Topology, Math 581, Fall 2017 last updated: November 24, Topology 1, Math 581, Fall 2017: Notes and homework Krzysztof Chris Ciesielski

(A n + B n + 1) A n + B n

Measure Theory on Topological Spaces. Course: Prof. Tony Dorlas 2010 Typset: Cathal Ormond

GAUSSIAN PROCESSES; KOLMOGOROV-CHENTSOV THEOREM

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

Stochastic Analysis I S.Kotani April 2006

STAT 331. Martingale Central Limit Theorem and Related Results

Some Background Material

Lecture 2. We now introduce some fundamental tools in martingale theory, which are useful in controlling the fluctuation of martingales.

Metric Spaces and Topology

Summable processes which are not semimartingales

Math212a1413 The Lebesgue integral.

FOUNDATIONS OF MARTINGALE THEORY AND STOCHASTIC CALCULUS FROM A FINANCE PERSPECTIVE

Integration on Measure Spaces

Lecture 5. If we interpret the index n 0 as time, then a Markov chain simply requires that the future depends only on the present and not on the past.

7 Complete metric spaces and function spaces

Optimal stopping for non-linear expectations Part I

Probability Theory. Richard F. Bass

Inference for Stochastic Processes

PROBABILITY THEORY II

The Uniform Integrability of Martingales. On a Question by Alexander Cherny

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

Exercises Measure Theoretic Probability

CHAPTER 1. Martingales

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define

Tools from Lebesgue integration

Introduction to Topology

4 Expectation & the Lebesgue Theorems

Real Analysis Notes. Thomas Goller

STAT 7032 Probability Spring Wlodek Bryc

An Introduction to Stochastic Processes in Continuous Time

Probability Theory II. Spring 2014 Peter Orbanz

1 Math 241A-B Homework Problem List for F2015 and W2016

Selected Exercises on Expectations and Some Probability Inequalities

(c) For each α R \ {0}, the mapping x αx is a homeomorphism of X.

Transcription:

BMS Basic Course Stochastic Processes II/ Wahrscheinlichkeitstheorie III Michael Scheutzow Lecture Notes Technische Universität Berlin Sommersemester 218 preliminary version October 12th 218

Contents 1 Processes, Filtrations, Martingales 1 1.1 Foundations....................................... 1 1.2 Filtrations and stopping times............................. 2 1.3 Martingales....................................... 7 1.4 Semimartingales, Quadratic Variation........................ 11 2 Stochastic Integrals and Stochastic Differential Equations 19 2.1 The stochastic integral................................. 19 2.2 Itô s formula...................................... 24 2.3 Representation of local martingales.......................... 27 2.4 Stochastic differential equations............................ 28 3 Lévy Processes 39 4 Markov Processes 43 4.1 Markov transition functions and Markov processes................. 43 4.2 Strong Markov property................................ 49 4.3 Hille-Yosida Theory.................................. 54 4.4 Invariant measures: existence............................. 61 4.5 Invariant measures: uniqueness............................ 65 4.6 Coupling, optimal transportation and Wasserstein distance............ 69 5 Appendix 71 Bibliography 73 i

Chapter 1 Processes, Filtrations, Martingales 1.1 Foundations Most of this chapter can be found in the first chapter of the monograph [KS91] or in [RY94]. Definition 1.1. Let Ω, F and E, E be measurable spaces and I a non-empty index set. A collection X = X t t I of measurable maps from Ω, F to E, E is called E-valued process. If in addition a probability measure is specified on Ω, F, then we call X a stochastic process. Notation 1.2. If E, d is a metric space or, more generally, E is a topological space, then we denote its Borel-σ-algebra i.e. the smallest σ-algebra containing all open subsets of E by BE. Unless we say something different, we always assume that subsets of R d are equipped with the Euclidean metric. Sometimes we write B[a, b] instead of B[a, b] etc. Notation 1.3. Let E, E be a measurable space. E I denotes the smallest σ-algebra on E I such that for each i I the projection map π i : E I E defined as π i x := x i, x E I is measurable. E I is called product σ-algebra. The following proposition is easy to prove. Proposition 1.4. Let X t t I be an E, E-valued process. Let X : Ω, F E I, E I be defined by Xωt := X t ω. Then X is measurable. Remark 1.5. If I is uncountable then the σ-algebra E I is rather crude. Subsets of E I which are not determined by countably many i I are never in E I. For example {f R [,1] : sup s [,1] fs 1} / B [,1], no matter which σ-algebra B on R we choose. Definition 1.6. Let X be an E, E-valued stochastic process on Ω, F, P. Then P X := PX 1 is called the distribution or the law of X. We often write LX instead of P X. Remark 1.7. Note that P X is a probability measure on E I, E I. Definition 1.8. Let X and Y be two stochastic processes on Ω, F, P with the same index set I and the same target space E, E. a X and Y are called modifications if for all t I we have P{ω : X t ω = Y t ω} = 1. b X and Y are called indistinguishable or equivalent if P{ω : X t ω = Y t ω for all t I} = 1. 1

2 Wahrscheinlichkeitstheorie III Remark 1.9. a Strictly speaking, the previous definition should be written more carefully. It is possible that for processes X and Y the set {ω : X t ω = Y t ω} is not in F when the space E, E is a bit pathological. Therefore the phrase we have... in part a should be replaced by the set {ω : X t ω = Y t ω} contains a set F F such that PF = 1. A similar remark applies to part b. As an example, take I = {}, Ω, F = E, E := {a, b}, {, E}, Xa = a, Xb = b, Y a = Y b = a and let P be the unique probability measure on Ω, F. Then {ω : Xω = Y ω} = {a} which is not in F. Note that X and Y are not modifications. b If X and Y are modifications then P X = P Y. c If X and Y are indistinguishable then they are modifications. d If X and Y are modifications and I is countable i.e. finite or countably infinite then they are indistinguishable. Example 1.1. Ω = [, 1], F = B[, 1], P = λ [,1], I = [, 1] where λ [,1] is Lebesgue measure on the Borel σ-algebra B[, 1] on [, 1]. X t ω := { 1, t = ω, t ω Y t ω. Then X and Y are modifications but not indistinguishable. Definition 1.11. If Ω, F, P is a probability space, then a set A F is called P-negligible if PA =. 1.2 Filtrations and stopping times In this section we consider stochastic processes and filtrations indexed by the interval [,. We could formulate these concepts for more general totally or even partially ordered index sets but we prefer not to be too general. All stochastic processes are assumed to have index set I = [, unless we explicitly say something different. Note that we will usually write X instead of X and similarly for other letters denoting stochastic processes. Definition 1.12. An E, E-valued process X on Ω, F is called measurable if, for every A E, we have X 1 A B[, F where X 1 A := {t, ω [, Ω : X t ω A}. Definition 1.13. Let Ω, F be a measurable space. F = F t t is called a filtration on Ω, F if for each t [,, F t is a sub-σ algebra of F such that F s F t whenever s t <. In this case, Ω, F, F is called a filtered measurable space FMS. If, moreover, Ω, F, P is a probability space, then Ω, F, F, P is called a filtered probability space FPS. We define F as the smallest σ-algebra which contains all F t. Definition 1.14. An E, E-valued process X on the FMS Ω, F, F is called adapted or F- adapted if for each t, X t is F t, E-measurable. The filtration F t := σx s, s t is called the natural filtration of X or the filtration generated by X. Definition 1.15. If Ω, F, F is a FMS, then we define F + s := F s + := t>s F t, s and F + := F s + s.

Version October 12th 218 3 It is easy to check that F + is also a filtration and that F + + = F +. Definition 1.16. If Ω, F, F is a FMS such that F = F +, then F is called right continuous. The following assumption about a filtration is common in stochastic analysis but not so common in the theory of Markov processes. Definition 1.17. A filtered probability space Ω, F, F, P is said to satisfy the usual conditions or to be normal if the following hold: i F contains all P-negligible sets in F, ii F = F +, i.e. F is right continuous. Remark 1.18. Some authors require in addition that the space Ω, F, P is complete, i.e. that each subset of a P-negligible set belongs to F. Here we follow [KS91] who do not require completeness. Definition 1.19. Let Ω, F, F be a FMS. A map τ : Ω [, ] is called a weak F-stopping time or optional time, if for all t we have {ω Ω : τω t} F t +. τ is called a strict F-stopping time or just F-stopping time if for all t we have {ω Ω : τω t} F t. The specification F is often omitted when the choice of the filtration is clear from the context. Remark 1.2. A strict F-stopping time is a weak F-stopping time and a weak F-stopping time is the same as an F + -stopping time. If the filtration F is right continuous, then a weak F-stopping time is automatically an F-stopping time. Proposition 1.21. Let Ω, F, F be a FMS and τ : Ω [, ] a map. equivalent: The following are i τ is a weak stopping time ii {ω : τω < t} F t for all t. Proof. i ii: Let τ be a weak stopping time and t. Then {τ < t} = n N ii i: Let τ satisfy ii and t. Then {τ t} = n N { 1 } τ t Ft. n { 1 } τ < t + F + n t. Example 1.22. Let Ω = {a, b}, F = 2 Ω. Define X t a := t and X t b = t for t. Let F be the filtration generated by X, i.e. F = {, Ω}, F t = 2 Ω for t >. Then τω := inf{t : X t ω > } is clearly an optional time but not a stopping time. Note that F is not right continuous.

4 Wahrscheinlichkeitstheorie III Warning 1.23. For each stopping time τ on a FMS Ω, F, F we have for all t {τ = t} = {τ t} \ {τ < t} F t 1.2.1 by the previous proposition, but the converse is not true, i.e. a map τ : Ω [, ] satisfying 1.2.1 for every t is not necessarily a stopping time not even a weak one. As an example, take Ω, F, P := [, 1], B[, 1], λ [,1] and define X t ω = 1 if t = ω and X t ω = otherwise. Let F be the filtration generated by X, i.e. F t := σ{ω}, ω t. Then τω := inf{s : X s ω = 1} is not a stopping time not even a weak one since {τ < t} is not in F t for any t, 1 but {τ = t} F t. Definition 1.24. Let Ω, F, F be a FMS and τ : Ω [, ] a map. F τ := {A F : A {τ t} F t for all t }, F + τ := {A F : A {τ t} F + t for all t }. The proof of the following lemma is very easy and is left to the reader. Lemma 1.25. Let Ω, F, F be a FMS and τ : Ω [, ] an F-stopping time. a F τ is a σ-algebra. b If τ s [, ], then F τ = F s. c τ is F τ -measurable. d If σ τ, then F σ F τ even if σ is not a stopping time. e If σ n n N is a sequence of stopping times, then so is σ := sup n σ n and σ := inf n σ n is a weak stopping time. f If σ is another stopping time then F σ τ = F σ F τ. Remark 1.26. If τ is a weak stopping time then F τ need not be a σ-algebra since it does not necessarily contain the set Ω. In fact F τ is a σ-algebra if and only if τ is a strict stopping time. Next, we want to investigate under which conditions first entrance times of a process X are stopping times. Definition 1.27. Let X be an E, E-valued process on a FMS Ω, F, F and G E. Then the maps S G resp. T G from Ω to [, ] defined as S G ω : = inf{t > : X t ω G} T G ω : = inf{t : X t ω G} are called hitting time resp. debut time of X in G. Proposition 1.28. Let E, d be a metric space with Borel σ-algebra E and let X be an E-valued F-adapted process. a If G E is open and X has right or left continuous paths, then T G and S G are weak stopping times.

Version October 12th 218 5 b If G E is closed and X has continuous paths, then T G is a strict stopping time. Proof. a Let t >. Then {T G < t} = {X s G} = {X s G} F t, s<t s [,t Q and analogously for S G, where we used the facts that G is open and the continuity assumption in the second equality. b For n N, let G n := {x E : dx, G < 1 n }. Then G n is open for each n and T Gn < T G on {T G, }. To see this, observe that the fact that G is closed and X is right continuous implies X T G G on the set {T G < } and left continuity of X then shows that T Gn < T G on {T G, }. On the set {T G, }, the sequence T Gn is clearly increasing and therefore converges to some T ω T G ω for each ω Ω. Using again left-continuity of X we see that T ω = T G ω for all ω Ω. From this it follows that {T G } = {T G = } = {X G} F, where we used the fact that G is closed and X is right-continuous. Further, for t >, we have {T G t} = {T G, t]} {T G = } = n=1 {T Gn < t} F t by part a since G n is open. We will see later in Theorem 1.37 that under slight additional assumptions on the filtration we get a much better result. Definition 1.29. Let Ω, F, F be a FMS. a The family {A [, Ω : A [, t] Ω B[, t] F t for all t } is called the progressive σ-algebra. b An E, E-valued process X is called progressive or progressively measurable if X : [, Ω E is measurable with respect to the progressive σ-algebra. Remark 1.3. Note that the progressive σ-algebra is a σ-algebra. Further note that a progressive set A satisfies A B[, F and that a progressive process X is adapted and measurable. On the other hand an adapted and measurable process X is not necessarily progressive. We will provide an example at the end of this section Example 1.38. Proposition 1.31. Let X be an E, E-valued progressive process on a FMS Ω, F, F and let T be a stopping time. Then a X T is F T -measurable, i.e. for B E we have {ω : T ω < and X T ω ω B} F T. b The process X T t t is progressive. Proof. a Fix t [,. Then the map ϕ : Ω [, t] Ω defined as ϕω := T ω t, ω is F t, B[, t] F t -measurable since for a t and A F t we have ϕ 1 [a, t] A = A {ω : T ω a} F t

6 Wahrscheinlichkeitstheorie III and the sets [a, t] A of this form generate B[, t] F t. Therefore, for B E, {T <, X T B} {T t} = {X T t B} {T t} = {X ϕ B} {T t} F t and {T <, X T B} = n=1{t n, X T B} F. b The proof of this part is similar. In the proof of the following proposition we will need the following lemma. Lemma 1.32. If E, d is a metric space with Borel-σ-algebra E and X n : Ω, F E, E, n N are measurable and lim n X n ω = Xω for all ω Ω, then X is also measurable. Proof. Let A E be closed and A n := {x E : dx, A < 1 n }. Then A n is open and X 1 A = n=1 k=1 m=k X 1 m A n F. Remark 1.33. The previous lemma does not hold on arbitrary topological spaces E even if E is compact Hausdorff! For example, if I = [, 1] carries the usual topology and E := I I is equipped with the product topology which is compact and Hausdorff, then there exists a sequence of continuous and hence measurable X n : I E which converge pointwise to a map X which is not measurable! The reader is invited to check this and/or to construct a different counterexample. Proposition 1.34. If X is a right or left continuous adapted process on a FMS Ω, F, F taking values in a metric space E, d with Borel σ-algebra E, then X is progressive. Proof. Assume that X is right continuous and adapted. Fix t. For n N and s [, t] define { [ Xs n X k+1 := 2 n t, s k 2 t, k+1 n 2 t, k < 2 n n X t s = t. For any B E we have {s, ω [, t] Ω : Xs n ω B} = [ k 2 n t, k + 1 { } 2 n t X k+1 2 n t ω B {t} {X t ω B} B[, t] F t, k<2 n so X n is B[, t] F t -measurable. Further, lim n Xs n ω = X s ω for every ω Ω and s [, t] by right continuity. The result follows using Lemma 1.32. The result for left continuous processes follows similarly. In this case we define { ] Xs n X k := 2 n t, s k 2 t, k+1 n 2 t, k < 2 n n X s =. The rest of the proof is as in the right continuous case.

Version October 12th 218 7 Definition 1.35. Let Ω, F be a measurable space and µ M 1 Ω, F, the space of probability measures on Ω, F. We denote by F µ the completion of F with respect to µ, i.e. the smallest σ-algebra which contains F and all subsets of µ-negligible sets in F. Then F u := µ M1 Ω,FF µ is called the universal completion of F. Note that F u u = F u. We state the following important and deep projection theorem. For a proof, see e.g. [He79] or [DM78]. Theorem 1.36. Let Ω, F be a measurable space and let R be the Borel-σ-algebra of a Polish i.e. complete separable metric space R. If A R F, then pr Ω A F u, where pr Ω A := {ω Ω : s R : s, ω A} is the projection of A onto Ω. Theorem 1.37. Let Ω, F, F be a FMS such that F t = Ft u for each t. Let X be E, E- valued progressive and let A E. Then both S A and T A are weak stopping times. Proof. For each t > we have to show that {S A < t} and {T A < t} are in F t. To show this for T A define A t := {s, ω : s < t, X s ω A} = n N which is in B[, t] F t. Therefore, {T A < t} = pr Ω A t F t by Theorem 1.36. The proof for S A is similar: just consider which is also in B[, t] F t. { s, ω : s t 1 n, X sω A }, Ã t := {s, ω : < s < t, X s ω A} = A t {, ω : ω Ω} Now we provide an example of an adapted and measurable process which is not progressive. Example 1.38. Let Ω, F = [, 1], L, where L denotes the σ-algebra of Lebesgue sets in [, 1] which by definition is the completion of B[, 1] with respect to Lebesgue measure. Let L be the σ-algebra on [, 1] containing all sets in L which have Lebesgue measure or 1. Define F t := L for all t. Define the set A [, Ω by A = {x, x : x [, 1/2]}. Then A B[, F but for each t >, A [, t] Ω / B[, t] F t otherwise the projection of the intersection onto Ω would be in F t by Theorem 1.36 which is however not the case. Therefore the indicator of A is measurable and check! adapted but not progressive. 1.3 Martingales We will assume that the reader is familiar with discrete time martingales. Definition 1.39. Let Ω, F, F, P be a FPS. A real-valued stochastic process M = M t t is called an F-martingale resp. F-submartingale resp. F-supermartingale if

8 Wahrscheinlichkeitstheorie III i M t t is adapted. ii E M t < for each t. iii For each s t < we have EM t F s = resp., resp. M s a.s. Example 1.4. Let Ω, F, F, P be a FPS. An F-adapted real-valued process B is called an F-Brownian motion if i PB = = 1 ii For every s < t <, B t B s is independent of F s. iii For every s < t <, LB t B s = N, t s. iv B has continuous paths. Note that a Brownian motion as defined in Wahrscheinlichkeitstheorie II hereafter abbreviated WT2 is an F-Brownian motion with respect to the filtration generated by B. Any F-Brownian motion is an F-martingale even if it only satisfies i-iii and not iv: for s < t < we have almost surely EB t F s = EB t B s + B s F s = EB t B s F s + EB s F s = + B s = B s. Further, the process M t := B 2 t t, t is a martingale, since for s t <, we have EM t F s = E B t B s + B s 2 F s t = EBt B s 2 + B 2 s + 2B s EB t B s t = M s. Another example, the compensated Poisson process N, will be introduced in class. The following theorem which is essentially the same as Theorem 3.8 in [KS91] states some fundamental results for right continuous submartingales which were shown for discrete time submartingales in WT2. We will answer the question whether and if so in which sense every submartingale admits a right continuous modification afterwards. For a function f : [, R and a finite subset F [, and < a < b < we let U F a, b; f be the number of upcrossings of f restricted to F of the interval [a, b] as defined in WT2. For a general subset I [,, we define U I a, b; f := sup{u F a, b; f; F I, F finite}. Theorem 1.41. Let X be a submartingale on Ω, F, F, P with right continuous paths and let [s, u] [,. For real numbers a < b, λ > and p > 1 we have i Maximal inequality: ii Upcrossing inequality: λp { sup X t λ } EX u +. s t u EU [s,u] a, b; Xω EX+ u + a b a

Version October 12th 218 9 iii Doob s L p -inequality: if in addition X is nonnegative, then p p pex E X t p p 1 u. sup s t u Proof. i Let F be a finite subset of [s, u] Q {s, u} which contains s and u. From WT2 we know that for any µ > we have µp { sup X t > µ } EX u +. 1.3.1 t F Take an increasing sequence F n of such sets whose union equals G := [s, u] Q {s, u}. The continuity of P implies that 1.3.1 holds true if F is replaced by G. Further, by right continuity of X the same estimate holds if G is replaced by [s, u]. The claim follows if we apply this to µ := λ 1 n and let n. ii For any finite subset F [s, u] containing u, the inequality EU F a, b; Xω EX+ u + a b a was proved in WT2. Defining F n and G as in part i and using right continuity of X the claim follows. iii Again this follows from the corresponding discrete time result as in parts i and ii. Obviously the statements of the previous theorem remain true if only almost all paths of X are right continuous. Next, we state and prove the submartingale convergence theorem. Theorem 1.42. Let X be a right continuous F-submartingale and assume that sup t EX + t <. Then there exists an integrable and F -measurable random variable X such that X ω = lim t X t ω for almost all ω Ω. Further, E X lim inf t E X t. Proof. The proof is as in the discrete time case using the upcrossing inequality in Theorem 1.41 ii and the right continuity of the paths of X. Remark 1.43. If we drop the right continuity assumption in the previous theorem but keep the other assumption, then it is still true that there exists an integrable and F -measurable random variable X such that for every fixed deterministic sequence t < t 1 <... such that lim n t n = we have X ω = lim n X tn ω for almost all ω Ω but the exceptional set of measure will depend on the sequence. Further, X t converges to X in probability but not necessarily almost surely [Exercise: show this!]. For ease of exposition, we now look at martingales only and refer the interested reader to the literature for the submartingale case. The following statement follows easily from the corresponding result for discrete time martingales the proof of iii i was actually provided in WT2 in the general case.

1 Wahrscheinlichkeitstheorie III Theorem 1.44. For an F-martingale X not necessarily with right continuous paths the following are equivalent: i The family X t, t is uniformly integrable. ii There exists an F -measurable integrable random variable X such that lim t E X t X =. iii There exists an F -measurable integrable random variable X such that for every t, we have X t = EX F t a.s. Now we state the optional sampling theorem. Theorem 1.45 Optional Sampling. Let X be a right continuous F-martingale for which one of the equivalent conditions in the previous theorem hold. Let S and T be F-stopping times. Then EX T F S = X S T and EX T F + S = X S T a.s. Proof. Consider the random times { [ k S n ω := 2, if Sω k 1 n 2, k n 2, k {1,..., n2 n }, n if Sω n, and analogously for T n. Both S n and T n are decreasing sequences of stopping times and converge to S resp. T for each ω Ω. The discrete time optional sampling theorem tells us that for each n N EX Tn F Sn = X Sn T n, i.e. for each A F Sn and hence for each A F + S since S n > S on the set {S < } implies F + S F S n X Tn dp = X Sn Tn dp. 1.3.2 A By right continuity of X the integrands converge to X T resp. X S. If we were allowed to interchange the limit n and the integral on both sides of 1.3.2, then the result would follow. A sufficient condition for the interchange to be legal is that the families X Tn resp. X Sn Tn are uniformly integrable. The following proof which shows that this is true is inspired by the proof of Lemma 1.2 in [Kl6]. Since X t is uniformly integrable there exists a nonnegative nondecreasing convex function f : [, [, such that L := sup t [, ] Ef X t < see WT2. For n, N N we obtain, using Jensen s inequality, E f X Tn 1l {Tn N} = E f XTn N 1l {Tn N} E E f XN F Tn N 1l{Tn N} = E f X N 1l {Tn N} Ef XN L, where we used the discrete time stopping theorem to show the first, so and A E f X Tn 1l {Tn< } L Ef X Tn = E f X Tn 1l {Tn< } + E f XTn 1l {Tn= } 2L. Uniform integrability the family X Tn follows. Uniform integrability of XSn T n follows in the same way and the proof is complete.

Version October 12th 218 11 The following result shows that the right continuity assumption of the filtration is in some sense without loss of generality. Corollary 1.46. A right continuous F-martingale X is also an F + -martingale. Proof. It is clear that X is F + -adapted. To prove the martingale property, let s < t < and apply the previous theorem to S = s and T = t note that the previous theorem can be applied even if X is not uniformly integrable. Finally, we state a result about the existence of a right continuous modification without proof. The proof can for example be found in [KS91]. Theorem 1.47. Let X be an F-submartingale and assume that F is right continuous. If the function t EX t is right continuous, then X has a right continuous modification X which is also an F-submartingale and for which almost all paths have a finite left limit at every point t >. Example 1.48. Here is an example of an F-martingale which does not have a right continuous modification and for which therefore F is not right continuous: let Ω = {a, b} equipped with the σ-algebra F containing all subsets of Ω and let P be the uniform measure on Ω, F. Define X t ω = for ω Ω and t [, 1], and X t a = 1 and X t b = 1 for t > 1. Let F be the filtration generated by X. Clearly, X is a martingale and X does not admit a right continuous modification. Note that F is not right continuous and that X is not an F + -martingale. 1.4 Semimartingales, Quadratic Variation For the rest of this chapter and the next chapter on stochastic integration, we will only consider martingales or semimartingales with continuous paths. Throughout this section, Ω, F, F, P is a given FPS. Definition 1.49. The class of all real-valued adapted stochastic processes with nondecreasing continuous paths is denoted by A +. The class of all real-valued adapted stochastic processes X with continuous paths of locally bounded variation is denoted by A. Recall that X has paths of locally bounded variation if V t ω := sup X ti tω X ti 1 tω < i holds for all t and all ω Ω, where the supremum is taken over all finite sequences = t,..., t n such that t <... < t n t. We call V the variation process of X. Remark 1.5. X A implies V A +. Lemma 1.51. X A iff there exist S, W A + such that X t ω = S t ω W t ω for all ω Ω. Proof. If X A, then S t := 1 2 V t + X t and W t := 1 2 V t X t are both in A + this is not completely obvious but true and X t = S t W t holds for all t. The converse is clear.

12 Wahrscheinlichkeitstheorie III Definition 1.52. An adapted real-valued stochastic process M = M t t is called an F-local martingale, if there exists a sequence of stopping times τ n n N such that τ n almost surely, and such that for every n N N n t := M t τn, t is a martingale. We denote the family of all local martingales with continuous paths by M loc and the family of all M M loc for which M ω = for all ω Ω by M loc. Further, M denotes the family of all martingales with continuous paths. Remark 1.53. Every martingale is a local martingale but the converse is not true. One example of a continuous local martingale which is not a martingale is M t := 1/ x+b t, where x R 3 \{} and B t is 3-dimensional Brownian motion starting at. It is however not so easy to prove these facts now only after we have Itô s formula at our disposal. The following proposition is a simple consequence of the optional sampling theorem. Proposition 1.54. If M is a right-continuous F-local martingale and T is a stopping time, then N t := M t T, t is also a right continuous F-local martingale. The following criterion for a continuous local martingale to be a martingale will be often used in the sequel. Proposition 1.55. If M M loc satisfies Esup s M s <, then M M. Proof. Let τ n, n N and N n be defined as in Definition 1.52. Then for each t, lim n N n t = M t almost surely and in L 1 the latter by dominated convergence since Esup s M s <. Hence, for s t <, and the claim follows. EM t F s = lim n EN n t F s = lim n N n s = M s, Remark 1.56. The assumptions of the previous proposition are in particular satisfied when M M loc is bounded in t, ω. There exist however examples of L p -bounded M M loc for arbitrarily large p which are not in M. Observe however that the condition in the previous proposition can be weakened to Esup t s M s < for each t. Semi- The following concept of a semimartingale is fundamental in stochastic analysis. martingales will turn out to be exactly the right class of stochastic integrators. Definition 1.57. A stochastic process X of the form X t = M t + A t with M M loc and A A is called a continuous semimartingale. We denote the family of all continuous semimartingales by S. The following theorem is very important. Its proof largely follows [Re13]. Theorem 1.58. M loc A = {}. In particular, the decomposition of X S in the previous definition is unique if we require in addition that M =.

Version October 12th 218 13 Proof. Let X M loc A and let V t, t be the associated variation process. Let us first assume that in addition there exist C 1, C 2 > such that X t ω C 1, V t ω C 2 for all t and all ω Ω. Let ε >, T := and T i+1 := inf{t T i : X t X Ti ε}, i N. Fix t > and define S i := T i t these are stopping times. Then, for n N, EX 2 S n = E = E n 1 X 2 X2 Si+1 S i = E i= n 1 2 XSi+1 X Si εe i= n 1 i= n 1 XSi+1 X Si 2 + 2E n 1 X Si XSi+1 X Si i= X Si+1 X Si εev t εc 2, i= where we used the fact that X M by Proposition 1.55 and therefore the second term after the second equality sign is zero by the optional sampling theorem. Therefore, since T n almost surely, we get using dominated convergence EX 2 t C 2 ε for all t > and all ε > which implies X almost surely. For the general case, let τ n be as in Definition 1.52 and define T n : = inf{s : X s n} ˆT n : = inf{s : V s n} Ť n : = inf{τ n, T n, ˆT n }. Then, for each n N, the process t X t Ť n satisfies the assumptions of the first case with C 1 = C 2 = n which implies X t Ť n = for all t almost surely. Since Ťn almost surely, the assertion follows. Remark 1.59. In the previous theorem, the assumption that the involved processes have continuous paths is important. If we replace continuous by right continuous in all definitions of this section, then the uniqueness of the decomposition does no longer hold in general: As an example consider the Poisson process N t which can be decomposed as N t = N t t + t and N t = + N t. Definition 1.6. Let D be the family of all = { = t < t 1 <...} such that t n. For a stochastic process X and D, we define k 1 Tt 2 2, X := Xti+1 X ti + Xt X tk t i= where k is such that t k t < t k+1. We say that X has finite quadratic variation if there exists a process X, X such that for each t, T t converges in probability to X, X t as, where := sup{t n+1 t n, n N }. We also write X t instead of X, X t.

14 Wahrscheinlichkeitstheorie III Remark 1.61. Recall that for Brownian motion B we showed that Bt 2 t is a martingale. It therefore follows from the next theorem or rather Theorem 1.65 that B, B t = t almost surely. We emphasize however that nevertheless for almost every ω Ω we can find a sequence n ω D for which n ω such that T nω t sequence n n N it may happen that lim sup n T n t Bω. Even for a deterministic Bω = almost surely. Theorem 1.62. Let M M loc be bounded in t, ω and hence M M. Then M has finite quadratic variation and M, M is the unique process in A + vanishing at zero such that t Mt 2 M, M t is a martingale. Proof. The proof is largely taken from [RY94] or [Re13]. Uniqueness follows immediately from Theorem 1.58. Now we show existence of such a process. For D and t i s t i+1 we have Therefore, for s < t <, E M ti+1 M ti 2 F s = E Mti+1 M s 2 F s + Ms M ti 2. E T t M F s = T s M + E M t M s 2 F s = T s M M 2 s + E M 2 t F s, so t Mt 2 Tt M is a continuous martingale. Tt M is not yet our claimed process since it is in general not nondecreasing. We want to show that Tt M converges in probability as. Fix a >. Let, D and denote the union of both subdivisions by. Then X t := Tt M Tt M is in M and bounded on [, t] Ω for each t and therefore t Xt 2 Tt X is in M and vanishes at. In particular, we have E Ta M Ta M 2 = E Xa 2 = E Ta X. We want to show that E X 2 a as +. 1.4.1 Assume for a moment that 1.4.1 has been shown for each a > we refer the reader to [RY94] or [Re13] for the proof. Since L 2 Ω, F, P is complete, there exists a random variable M, M a such that lim T a M = M, M a in L 2. All that remains to be shown is that the process t M, M t has a modification in A + which vanishes at and that t M 2 t M, M t is a martingale. Doob s L 2 -inequality shows that for each fixed a > the processes T s M M, M s converge to zero even uniformly on [, a] in L 2 i.e. the supremum converges to in L 2 and therefore the limit has a continuous modification. It is also easy to see by approximation that we have M, M = almost surely and that the process is nondecreasing. Being the limit of adapted processes it also has a modification which is adapted. Finally, the martingale property follows since L 2 -limits of martingales are martingales. We aim at generalizing the previous theorem to arbitrary M M loc. To do that we need the following lemma. Lemma 1.63. For M as in the previous theorem and a stopping time T define M T t := M t T. We have M T, M T t = M, M t T.

Version October 12th 218 15 Proof. By the optional sampling theorem and Theorem 1.62, the process t M 2 T t M, M t T is a martingale. The claim of the lemma now follows using the uniqueness statement in Theorem 1.62. Definition 1.64. If X n and Y are real-valued stochastic processes indexed by [,, then we write lim n X n = Y ucp, if for each T >, we have that sup s T X n s Y s converges to in probability as n ucp stands for uniformly on compact sets in probability. Theorem 1.65. Let M M loc. There exists a unique process M, M A+ which vanishes at such that M 2 M, M M loc. Further, M, M = lim T M ucp. Proof. Uniqueness follows from Theorem 1.58. Define T n := inf{t : M t = n} and M n := M Tn. Then M n is a bounded continuous martingale. By Theorem 1.62, there exists a process M n, M n such that M n 2 M n, M n is a martingale. For m n, we have M n Tm = M m and therefore, by Lemma 1.63, M n, M n t Tm = M m, M m t. For fixed t and on the set {ω : T m ω t} we define M, M t := M m, M m t. Note that M, M is well-defined and equals lim n M n, M n. Further, since M n 2 M n, M n is a martingale, by taking the limit n, we see that M 2 M, M M loc. To show the second part of the theorem, we fix δ > and t >. Defining T n as above, we pick some k N such that PT k < t δ. On the set {s T k }, we have M, M s = M, M s Tk = M T k, M T k s. Therefore, for ε >, P sup T s M M, M s ε δ + P sup T s M M, M s ε, T k t s t s T k = δ + P sup Ts M T k M T k, M T k s ε, T k t s T k δ + P sup Ts M T k M T k, M T k s ε. s T k Since M T k is a bounded martingale, Theorem 1.62 implies lim sup P sup T s M M, M s ε s t δ. Since δ, ε > are arbitrary, the asserted ucp convergence follows.

16 Wahrscheinlichkeitstheorie III Next, we define the bracket of two local martingales. Theorem 1.66. Let M, N M loc. There exists a unique process M, N A vanishing at such that M N M, N M loc. Further, we have where lim T M, N = M, N ucp, T s M, N := M ti s M ti 1 s N ti s N ti 1 s. Proof. Define Since M, N := 1 4 M + N, M + N M N, M N. M N = 1 4 M + N 2 M N 2, we obtain from Theorem 1.65 the fact that M N M, N M loc. Since M, N is the difference of two processes in A + it belongs to A but not necessarily to A +!. The rest follows as before. Proposition 1.67. If X S has a decomposition X = M + A with M M loc then T X M, M ucp. and A A, Proof. We have T s X = i 2 Xti s X ti 1 s = T s M + 2Ts M, A + Ts A. Now T s M, A = i Mti s M ti 1 s Ati s A ti 1 s sup M ti M ti 1 A ti s A ti 1 s t i,t i s since M has continuous paths in fact the convergence is almost surely uniform on compact sets. The fact that T s A follows in the same way. The proposition now follows from Theorem 1.65. Definition 1.68. Let { } H := M M : supe M t 2 < t be the set of all L 2 -bounded continuous martingales. The meaning of an additional upper index should be clear. Many authors use the symbol H 2 or H 2 to denote this space. Theorem 1.69. 1. If M H, then M converges to some M almost surely and in L 2. 2. H is a real Hilbert space after forming equivalence classes if equipped with the norm M H := lim t E M t 2 1/2 = E M 2 1/2 i

Version October 12th 218 17 3. The following norm on H is equivalent to. H : M := Esup M t 2 1/2 t 4. If M H, then M, M := lim t M, M t exists almost surely and M 2 H = E M, M. Proof. 1. This follows from the submartingale convergence theorem Theorem 1.42 together with results from WT2 L 2 -boundedness of a martingale implies L 2 -convergence. 2. The second equality follows from L 2 -convergence. Since L 2 Ω, F, P is a Hilbert space, it only remains to show that H is closed or complete. To see this pick a Cauchy sequence M n in H. Then M := lim n M n exists in L 2. Further, by Doob s L 2 -inequality Theorem 1.41 iii, we see that E sup Mt n Mt m 2 4E M n M m 2, n, m. t Therefore, there exists a subsequence and a stochastic process M with continuous paths such that lim M n t n = M t uniformly in t for almost all ω Ω. Taking the limit n in E M n F t = M n t, we obtain E M F t = Mt, so M is a continuous martingale with limit M. 3. This follows from Esup t M 2 t 4EM 2 4Esup t M 2 t, where the first inequality is an application of Doob s L 2 -inequality Theorem 1.41 iii. 4. We know from Theorem 1.65 that M 2 t M, M t M loc. Let T n := inf{t : M t M t = n}. Then EM 2 t T n = E M, M t Tn. 1.4.2 Letting n and t, the right hand side converges to E M, M by monotone convergence. Further, lim n M 2 t T n = M 2 t almost surely and, by dominated convergence using 3. above, in L 1. Letting t and using 1. we see that the left side of 1.4.2 converges to EM 2 when we let n and then t, so the claim follows.

Chapter 2 Stochastic Integrals and Stochastic Differential Equations 2.1 The stochastic integral Throughout this section Ω, F, F, P will be a FPS. In addition we assume that the usual conditions introduced earlier hold. We aim at defining the stochastic integral process t f sω ds s for S S and a reasonably large class of integrand processes f. It turns out that this cannot be done in a pathwise manner. Since each S S can be uniquely decomposed as S = M + A with M M loc and A A, we define f s ω ds s := f s ω dm s + f s ω da s. So far neither of the two integrals on the right hand side has been defined. The second one a so called Lebesgue-Stieltjes integral turns out to be much easier than the first, so we start by defining the second integral. This can be done pathwise, i.e. for each ω Ω separately only afterwards will we care about measurability properties with respect to ω. Let A A, i.e. A A and A =. Think of ω being fixed, so A is just a continuous function from [, to R of locally bounded variation. A can be written as A t = A + t A t, where both A + and A are continuous, nondecreasing and vanishing at, so they correspond to σ-finite measures µ + and µ on, via A + t = µ +, t] and A = µ, t]. This allows us to define f s da s := f s dµ + s f s dµ s, whenever f is measurable with respect to s and the right hand side is well-defined i.e. is not of the form or. We will in fact assume that f is chosen such that both integrals are finite. Note that the decomposition of A = A + A is not unique and even the integrability of f may depend on the choice of the decomposition, but the value of the left hand side does not depend on the particular decomposition and there is always a unique minimal decomposition, i.e. one for which V t = A + t + A t is the variation process cf. the proof of Lemma 1.51. The integral is then automatically continuous in t since the measures µ + and µ are atomless. We want to ensure that the integral is also adapted. This property will hold when f is progressive and for example locally bounded for almost every ω Ω. This follows essentially like Lemma 4.2 in the following chapter we will discuss this in class. 19

2 Wahrscheinlichkeitstheorie III In the following we will also write f A instead of. f s da s. We will now proceed to define the stochastic integral f sω dm s. This will be done in several steps. We first assume that M H and f is simple in the sense of the following definition. Then we extend to more general f and to general M M loc. Definition 2.1. The stochastic process f : [, Ω R is called a simple process, in short f L s, if there exist = { = t < t 1 <...} such that t i and C < and F ti -measurable ξ i, i =, 1,... such that sup n ξ n ω C for all ω Ω and f t ω = ξ ω1l {} t + ξ i ω1l ti,t i+1 ]t; t <, ω Ω. 2.1.1 i= For M H and f L s as in 2.1.1, we define the stochastic integral in an obvious way: I t f := f s dm s := ξ i Mt ti+1 M t ti, t <. i= We will also write f M instead of I. f. Note that the value of f has no relevance for the stochastic integral. Proposition 2.2. For M H and f L s we have i f M H. ii f M t = f 2 s d M s. iii f M 2 H = E f 2 s d M s. Proof. i f M is clearly adapted, has continuous paths, vanishes at and is L 2 -bounded. To show the martingale property, take s < t < and define k, n such that s [t k 1, t k and t [t n, t n+1. If k n, then n 1 f M t f M s = ξ i Mti+1 M ti + ξn Mt M tn + ξk 1 Mtk M s. i=k The conditional expectation of the right hand side given F s is. If k = n + 1, then E f M t f M s F s = ξn E M t M s F s = and therefore the martingale property follows. ii Since the right hand side of the claimed equality is in A + and vanishes at, it suffices to show that t f M 2 t f 2 s d M s

Version October 12th 218 21 is a martingale. Both summands are integrable. For s < t < we have, assuming without loss of generality that s and t are in the partition, say s = t k and t = t n+1, so the claim follows. E f M 2 t f M 2 s F s = E f Mt f M s 2 Fs = E n ξ i M ti+1 M ti 2 Fs i=k = E n ξi 2 M ti+1 M ti 2 F s i=k = E fr 2 d M r F s, iii This follows from part ii and part 4. of Theorem 1.69. s Property iii of the previous proposition states that I is an isometry from L s to H if L s is equipped with the corresponding norm. To formulate this precisely, we introduce the Doleans measure and various function spaces. Definition 2.3. Let M H. The Doleans measure µ M on [, Ω, B[, F is defined as µ M A := E 1l A s, ω d M s. Note that µ M is a finite measure with total mass E M. Definition 2.4. Let LM := L 2 [, Ω, B[, F, µ M. Note that L s is a subspace of LM. We will work with several subspaces of LM below. As usual, we denote the corresponding spaces of equivalence classes by LM etc. We will always assume that M H is given. Observe that LM is a Hilbert space. We will denote the corresponding norm by. M. The following corollary is a reformulation of part iii of Proposition 2.2. Corollary 2.5. I : L s H defined above is an isometry if L s is equipped with the norm. M. An isometry from some metric space to a complete metric space can always be extended uniquely to an isometry on the completion of the first space. Denote the completion of L s in the Hilbert space LM by L M. We can and will simply define the stochastic integral I t f for f L M as the image of the unique extension of I to a linear isometry from L M to H which we again denote by I. It is of course of interest to have a more explicit description of the space L M. We will deal with this question soon. Observe however that it is definitely not true in general that L M = LM: if L M denotes the closed subspace of LM consisting of all progressive processes in LM more precisely: consisting of all equivalence classes containing some progressive process then L M contains L s since left continuous adapted processes are progressive and therefore also its completion L M. We point out that it may well happen

22 Wahrscheinlichkeitstheorie III that an equivalence class of LM contains both progressive and non-progressive processes: just think of the extreme case in which M and therefore LM = {} then µ M = and all processes in LM are equivalent. One approach to identifying the space L M is to consider the sub-σ-algebra of B[, F which is generated by all processes in L s. This sub-σ-algebra is called the predictable σ- algebra. It is contained in the progressive σ-algebra since each left continuous adapted process is progressive. It is then not hard to show that any predictable process i.e. a process which is measurable with respect to the predictable σ-algebra in LM can be approximated by a sequence of processes from L s with respect to the norm on LM or semi-norm on LM we will show this below and therefore L M is precisely the space of equivalence classes which contain at least one predictable process which is in LM. Even though by far not every predictable process is progressive, Lemma 3.2.7 in [KS91] the proof of which is neither hard nor short shows that each equivalence class in LM which contains a progressive process also contains a predictable process i.e. L M = L M and in this sense we have defined the stochastic integral also called Itô s integral for all progressive integrands in LM. We point out that if one likes to introduce stochastic integrals with respect to general right continuous but not necessarily continuous semimartingales, then one has to restrict the class of integrands to predictable ones since then the corresponding equality L M = L M does not hold in general. Remark 2.6. The predictable σ-algebra on, Ω coincides with the σ-algebra on, Ω which is generated by all sets of the form s, t] A where s t < and A F s. Proposition 2.7. If for 1 i k we have M i H and f is predictable and in each of the spaces LM i, then there exists a sequence of simple processes f n such that f f n M i for all i. Proof. Define R := {f predictable, bounded uniformly in t, ω : there exist f n as claimed}. Clearly, R is a linear space which contains the constant process 1l. The family of indicators of s, t] A where s t < and A F s belongs to R, is closed under multiplication, and generates the predictable σ-algebra. The monotone class theorem see appendix states that R is equal to the set of all bounded predictable functions provided we can show that R is closed with respect to limits of increasing, nonnegative and uniformly bounded elements of R. To show this, take a sequence of non-negative g n R such that g n g and g is bounded in t, ω. Being the pointwise limit of predictable processes g is also predictable. Further, we have g n g M i by dominated convergence and it follows from the usual triangle trick that g R. To complete the proof, let f be as in the proposition and define f N := f1l f <N. Then f N is bounded and predictable and hence in R and f f N M i for each i by dominated convergence. Again the triangle trick shows that the claim in the proposition holds for f. Before proceeding to the next step of the construction of the Itô integral with respect to a continuous semimartingale we state a few properties of the integral defined so far. First note that for general M, N M loc, we have for each s < t almost surely M, N t M, N s M t M s N t N s 2.1.2

Version October 12th 218 23 which is just the usual Cauchy-Schwarz inequality which holds true because.,. is almost surely bilinear, symmetric and positive semi-definite. This implies for M, N H using part 4. of Theorem 1.69. E sup t< M, N t M H N H. 2.1.3 Note that item b in the following proposition follows directly from the isometry property of the stochastic integral in case M = N no matter if f and g agree or not since an isometry between Hilbert spaces automatically preserves the inner product. Proposition 2.8 Kunita-Watanabe identity and inequality. Let M, N H and f L M, g L N. Then a fg M, N t 2 f 2 M t g2 N t a.s. b f M, g N = fg M, N c E f M, g N t E f 2 M E g t 2 N t. Proof. a This is easy to show for f, g L s see e.g. [Ku9] and then for general f, g by approximation see e.g. [KS91], Proposition 3.2.14. b For f, g L s the claim in part b follows just like part ii of Proposition 2.2. For general f L M, g L N approximate by f n, g n L s in L M resp. L N and use 2.1.3 and part a exercise. c This is just a combination of parts a and b using the Cauchy Schwarz inequality. We are now ready for the final step in the construction. For M M loc let L loc M := {f progressive : f 2 s d M s < for all t a.s.}. Definition 2.9. For M M loc and f L locm let S n be a sequence of stopping times such that S n and t M t Sn is in H. Further define R n := n inf{t : f s 2 d M s n}. Define T n := R n S n. Then T n almost surely and the T n are stopping times such that Mt n := M t Tn is in H and ft n := f t 1l Tn t is in L M n. We define f M t := f s dm s := f n M n t on {T n ω t}. We have to make sure that the integral is well-defined. Note that once this has been clarified the stochastic integral process is again in M loc. The fact that the stochastic integral is welldefined follows from the following proposition see [KS91], Corollary 3.2.21 for the easy proof. Proposition 2.1. Let M, N H and f L M, g L N and assume that there exists a stopping time T such that almost surely Then, for almost all ω Ω f t T ω ω = g t T ω ω; M t T ω ω = N t T ω ω for all t. f M t T ω = g N t T ω for all t.

24 Wahrscheinlichkeitstheorie III Roughly speaking, all properties derived for stochastic integrals in case M H and f L M which do not involve expected values carry over to the general case and are easily proved. Those properties which do involve expected values will generally not carry over. Our next aim is to formulate and prove Itô s formula, the chain rule of stochastic calculus. To do this, we require a few further properties. The following is a kind of Riemann approximation result. Theorem 2.11. For each n N, let n D such that n. Let M M loc and f adapted and continuous. Then t f t n Mt n i i+1 t M t n i t = f s dm s ucp. lim n i= Proof. We point out that a continuous adapted process is automatically predictable because it can be pointwise approximated by simple processes. First assume that M H and f is in addition bounded. For a given n D we define the approximation f n by f n t := f ti in case t t i, t i+1 ] and say f n := f. Then f n L s and f f n M and therefore the corresponding stochastic integrals converge, so the result follows in this case. The general case follows by stopping. We start by proving the following integration by parts formula or product rule. So far we have defined the bracket M, N only for M, N M loc but of course it also makes sense for M, N S and coincides with the bracket of the local martingale parts. Note that M, N = in case either M or N is in A. If M, N S but not in S, then we define both the bracket and the stochastic integral as the bracket and stochastic integral of M M resp. N N. Proposition 2.12. Let X, Y S. Then X t Y t X Y = X s dy s + Proof. For n D, we have X t Y t X Y = X t n k+1 t X t n k t Y t n k+1 t Y t n k t k= + k= Y t n k t X t n k+1 t X t n k t + Y s dx s + X, Y t. k= X t n k t Y t n k+1 t Y t n k t. Theorem 2.11 shows that as n the second and last sums converge to the corresponding stochastic integrals and by results in the first chapter the first sum converges ucp to the bracket. 2.2 Itô s formula The following Itô s formula is the chain rule of stochastic integration. Theorem 2.13 Itô s formula. Let F C 2 R d and X = X 1,..., X d such that X i S, i = 1,..., d. Then F X 1,..., X d S and F X t = F X + d k=1 F x k X s dx k s + 1 2 d k,l=1 2 F x k x l X s d X k, X l s.

Version October 12th 218 25 In the proof of Itô s formula we will need two properties of stochastic integrals which are also of interest otherwise. Lemma 2.14. Let M H, g L M and f L g M. Then fg L M and fg M = f g M. Proof. Let M := g M. Then M H and M = g 2 M by Proposition 2.8, part ii. Hence, E f s g s 2 d M s = E f 2 s d M s <, so fg L M. Further, for N H, we have using Proposition 2.8, part ii once more, which implies fg M, N = fg M, 1 N = fg M, N = f g M, N = f g M, N = f g M, N, fg M f g M, N =. Since this holds for each N H we can choose N = fg M f g M which implies fg M f g M =. Now Theorem 1.69, part 4. shows that fg M t f g M t = almost surely for all t so the claim follows. Remark 2.15. The previous proposition extends to the case of general M M loc and even M S and integrands f, g for which the stochastic integrals are defined by applying suitable stopping times. The following continuity property is good to know. Lemma 2.16. Let M M loc and f, f n, n N be in L loc M such that f n f ucp. Then, f n M f M ucp. Proof. Without loss of generality we assume that f =. Define T m := inf{s : M s m}; M m s := M s Tm ; τ m,n := inf{s : Then, for fixed t, δ >, s P fu n dm u δ sup s t sup s t P τ m,n T m t + P sup s t s s τm,n T m f n u 2 d M m u 1 m }. fu n dm u δ. Using in this order Chebychev s inequality, Doob s L 2 -inequality and Proposition 2.2 iii, the last summand can be estimated from above by s P fu n 1l [,τm,n]udmu m δ 4 δ 2 E fu n 1l [,τm,n]u 2 d M m u 4 δ 2 m. For given ε, δ >, we first choose m such that PT m t ε/3 and 4 n such that for all n n, Pτ m,n t ε/3. Then we obtain s P fu n dm u δ ε. sup s t The result follows since δ, ε, t > were arbitrary. δ 2 m ε/3. Then choose