<<

Stochastics II

Lecture notes

Prof. Dr. Evgeny Spodarev

Ulm 2015 Contents

1 General theory of random functions1 1.1 Random functions...... 1 1.2 Elementary examples...... 5 1.3 Regularity properties of trajectories...... 6 1.4 Differentiability of trajectories...... 11 1.5 Moments und covariance...... 12 1.6 Stationarity and Independence...... 14 1.7 Processes with independent increments...... 15 1.8 Additional exercises...... 16

2 Counting processes 19 2.1 Renewal processes...... 19 2.2 Poisson processes...... 28 2.2.1 Poisson processes...... 28 2.2.2 Compound Poisson process...... 33 2.2.3 Cox process...... 35 2.3 Additional exercises...... 35

3 Wiener process 38 3.1 Elementary properties...... 38 3.2 Explicit construction of the Wiener process...... 39 3.2.1 Haar- and Schauder-functions...... 39 3.2.2 Wiener process with a.s. continuous paths...... 42 3.3 Distribution and path properties of Wiener processes...... 45 3.3.1 Distribution of the maximum...... 45 3.3.2 Invariance properties...... 47 3.4 Additional exercises...... 50

4 Lévy Processes 52 4.1 Lévy Processes...... 52 4.1.1 Infinitely Divisibility...... 52 4.1.2 Lévy-Khintchine Representation...... 55 4.1.3 Examples...... 59 4.1.4 Subordinators...... 61 4.2 Additional Exercises...... 64

5 Martingales 67 5.1 Basic Ideas...... 67 5.2 (Sub-, Super-)Martingales...... 69 5.3 Uniform Integrability...... 72

i ii Contents

5.4 Stopped Martingales...... 74 5.5 Lévy processes and Martingales...... 78 5.6 Martingales and Wiener Process...... 80 5.7 Additional Exercises...... 83

6 Stationary of Random Variables 86 6.1 Sequences of Independent Random Variables...... 86 6.2 Stationarity in the Narrow Sense and Ergodic Theory...... 87 6.2.1 Basic Ideas...... 87 6.2.2 Mixing Properties and Ergodicity...... 90 6.2.3 Ergodic Theorem...... 94 6.3 Stationarity in the Wide Sense...... 96 6.3.1 Correlation Theory...... 96 6.3.2 Orthogonal Random Measures...... 98 6.3.3 Integral regarding an Orthogonal Random ...... 99 6.3.4 Spectral Representation...... 100 6.4 Additional Exercises...... 101

Bibliography 101 1 General theory of random functions

1.1 Random functions

Let ˆΩ, A, P be a space and ˆS, B a measurable space, Ω, S x g. Definition 1.1.1 A random element X ¢ Ω S is a ASB-measurable mapping (Notation: X > ASB), i.e.,  X 1ˆB ˜ω > Ω ¢ Xˆω > B > A,B > B. If X is a random element, then Xˆω is a realization of X for arbitrary ω > Ω. We say that the σ-algebra B of subsets of S is induced by the system M (Elements of M are also subsets of S), if B  F FaM F-σ-algebra on S (Notation: B σˆM). If S is a toplological or , then M is often chosen as a class of all open sets of S and σˆM is called the Borel σ-algebra (Notation: B BˆS). Example 1.1.1 1. If S R, B BˆR, then a random element X is called a . m m 2. If S R , B BˆR , m A 1, then X is called a random vector. Random variables and random vectors are considered in the lectures „Elementare Wahrscheinlichkeitsrechnung und Statistik“ and „Stochastik I“. m 3. Let S be the class of all closed sets of R . Let m M ˜˜A > S ¢ A 9 B x g ,B – arbitrary compactum of R  . Then X ¢ Ω S is a random closed set. m As an example we consider n independent uniformly distributed points Y1,...,Yn > 0, 1¥ and R1,...,Rn A 0 (almost surely) independent random variables, which are defined on the ˆ  8n ˆ  ˆ  ˜ > same Ω, A, P as Y1,...,Yn. Consider X i 1BRi Yi , where Br x y m R ¢ SSyxSS B r. Obviously, this is a random closed set. An example of a realization is provided in Figure 1.1. Exercise 1.1.1 Let ˆΩ, A and ˆS, B be measurable spaces, B σˆM, where M is a class of subsets of S.  Prove that X ¢ Ω S is ASB-measurable if and only if X 1ˆC > A, C > M. Definition 1.1.2 Let T be an arbitrary index set and ˆSt, Btt>T a family of measurable spaces. A family X ˜Xˆt, t > T  of random elements Xˆt ¢ Ω St defined on ˆΩ, A, P and ASBt-measurable for all t > T is called a random (associated with ˆSt, Btt>T ).

1 2 1 General theory of random functions

86 ˆ  Abb. 1.1: Example of a random set X i 1BRi Yi

Therefore it holds X ¢ Ω  T ˆSt, t > T , i.e. Xˆω, t > St for all ω > Ω, t > T and Xˆ , t > ASBt, t > T . We often omit ω in the notation and write Xˆt instead of Xˆω, t. Sometimes ˆSt, Bt does not depend on t > T as well: ˆSt, Bt ˆS, B for all t > T .

Special cases of random functions:

1. T b Z ¢ X is called a random or in discrete time. Example: T Z, N.

2. T b R ¢ X is called a stochastic process in continuous time. Example: T R, a, b¥, ª @ a @ b @ ª, R. d 3. T b R , d C 2 ¢ X is called a random field. d d d d Example: T Z , R, R , a, b¥ . d 4. T b BˆR  ¢ X is called set-indexed process. If Xˆ  is almost surely non-negative and σ-additive on the σ-algebra T , then X is called a .

The tradition of denoting the index set with T comes from the interpretation of t > T for the cases 1 and 2 as time parameter. For every ω > Ω, ˜Xˆω, t, t > T  is called a trajectory or path of the random function X. We would like to prove that the random function X ˜Xˆt, t > T  is a random element within the corresponding function space, which is equipped with a σ-algebra that now is spec- ified. Let ST Lt>T St be the cartesian product of St, t > T , i.e., x > ST if xˆt > St, t > T . The elementary cylindric set in ST is defined as

CT ˆB, t ˜x > ST ¢ xˆt > B , where t > T is a selected point from T and B > Bt a subset of St. CT ˆB, t therefore contains all trajectories x, which go through the „gate“ B, see Figure 1.2. Definition 1.1.3 The cylindric σ-algebra BT is introduced as a σ-algebra induced in ST by the family of all 1 General theory of random functions 3

Abb. 1.2: Trajectories which pass a „gate“ Bt.

T elementary cylinders. It is denoted by BT at>T Bt. If Bt B for all t > T , then B is written instead of BT . Lemma 1.1.1 The family X ˜Xˆt, t > T  is a random function on ˆΩ, A, P with phase spaces ˆSt, Btt>T if and only if for every ω > Ω the mapping ω ( Xˆω,  is ASBT -measurable. Exercise 1.1.2 Proof Lemma 1.1.1. Definition 1.1.4 Let X be a random element X ¢ Ω S, i.e. X be ASB-measurable. The distribution of X is 1 the PX on ˆS, B such that PX ˆB PˆX ˆB, B > B. Lemma 1.1.2 An arbitrary probability measure µ on ˆS, B can be considered as the distribution of a random element X.

Proof Take Ω S, A B, P µ and Xˆω ω, ω > Ω.

When does a random function with given properties exist? A random function, which consists of independent random elements always exists. This assertion is known as Theorem 1.1.1 (Lomnicki, Ulam): Let ˆSt, Bt, µtt>T be a sequence of probability spaces. It exists a random sequence X ˜Xˆt, t > T  on a probability space ˆΩ, A, P (associated with ˆSt, Btt>T ) such that

1. Xˆt, t > T are independent random elements.

2. PXˆt µt on ˆSt, Bt, t > T .

A lot of important classes of random processes is built on the basis of independent random elements; cf. examples in Section 1.2. Definition 1.1.5 Let X ˜Xˆt, t > T  be a random function on ˆΩ, A, P with phase space ˆSt, Btt>T . The ˆ ˆ  ˆ T finite-dimensional distributions of X are defined as the distribution law Pt1,...,tn of X t1 ,...,X tn ˆ  > >   on St1,...,tn , Bt1,...,tn , for arbitrary n N, t1, . . . , tn T , where St1,...,tn St1 ... Stn and 4 1 General theory of random functions

a a   Bt1,...,tn Bt1 ... Btn is the σ-algebra in St1,...,tn , which is induced by all sets Bt1 ... Btn , > ˆ  ˆˆ ˆ  ˆ T >  > Bti Bti , i 1, . . . , n, i.e., Pt1,...,tn C P X t1 ,...,X tn C , C Bt1,...,tn . In particular,   > for C B1 ... Bn, Bk Btk one has ˆ    ˆ ˆ  > ˆ  >  Pt1,...,tn B1 ... Bn P X t1 B1,...,X tn Bn .

Exercise 1.1.3 ˆ ˆ  ˆ T S Prove that Xt1,...,tn X t1 ,...,X tn is a A Bt1,...,tn -measurable random element. Definition 1.1.6 Let St R for all t > T . The random function X ˜Xˆt, t > T  is called symmetric, if all ˆ  of its finite-dimensional distributions are symmetric probability measures, i.e., Pt1,...,tn A ˆ  > > > Pt1,...,tn A for A Bt1,...,tn and all n N, t1, . . . , tn T , whereby

ˆ  ˆˆ ˆ   ˆ T >  Pt1,...,tn A P X t1 ,..., X tn A . Exercise 1.1.4 Prove that the finite-dimensional distributions of a random function X have the following > C ˜  ` > properties: for arbitrary n N, n 2, t1, . . . , tn T , Bk Stk , k 1, . . . , n and an arbitrary permutation ˆi1, . . . , in of ˆ1, . . . , n it holds:

Symmetry P ˆB  ...  B  P ˆB  ...  B  1. : t1,...,tn 1 n ti1 ,...,tin i1 in ˆ     ˆ    2. Consistency: Pt1,...,tn B1 ... Bn1 Stn Pt1,...,tn1 B1 ... Bn1

The following theorem evidences that these properties are sufficient to prove the existence of a random function X with given finite-dimensional distributions. Theorem 1.1.2 (Kolmogorov): ˜ > ˜  `  Let Pt1,...,tn , n N, t1, . . . , tn T be a family of probability measures on

m m m m ˆR  ...  R , BˆR  a ... a BˆR , which fulfill conditions 1 and 2 of Exercise 1.1.4. Then there exists a random function X ˜Xˆt, t > T  defined on a probability space ˆΩ, A, P with finite-dimensional distributions

Pt1,...,tn .

Proof See [13], Section II.9.

m This theorem also holds on more general (however not arbitrary!) spaces than R , on so- called Borel spaces, which are in a sense isomorphic to ˆ 0, 1¥ , B 0, 1¥ or a subspace of that. Definition 1.1.7 Let X ˜Xˆt, t > T  be a random function with values in ˆS, B, i.e., Xˆt > S almost surely for arbitrary t > T . Let (T,C) be itself a measurable space. X is called measurable if the mapping X ¢ ˆω, t ( Xˆω, t > S, ˆω, t > Ω  T , is A a CSB-measurable. Thus, Definition 1.1.7 not only provides the measurability of X with respect to ω > Ω: Xˆ , t > ASB for all t > T , but Xˆ ,  > A a CSB as a function of ˆω, t. The measurability of X is of significance if Xˆω, t is considered at random moments τ ¢ Ω T , i.e., Xˆω, τˆω. This is in particular the case in the theory of martingales if τ is a so-called stopping time for X. The distribution of Xˆω, τˆω might differ considerably from the distribution of Xˆω, t, t > T . 1 General theory of random functions 5

1.2 Elementary examples

The theorem of Kolmogorov can be used directly for the explicit construction of random pro- cesses only in few cases, since for a lot of random functions their finite-dimensional distributions are not given explicitly. In these cases a new random function X ˜Xˆt, t > T  is built as Xˆt gˆt, Y1,Y2,..., t > T , where g is a and ˜Yn a sequence of random elements (also random functions), whose existence has already been ensured. For that we give several examples. Let X ˜Xˆt, t > T  be a real-valued random function on a probability space ˆΩ, A, P. 1. : Definition 1.2.1 The random function X ˜Xˆt, t > T  is called white noise, if all Xˆt, t > T , are independent and identically distributed (i.i.d.) random variables.

White noise exists according to the Theorem 1.1.1. It is used to model the noise in (electromagnetic or acoustical) signals. If Xˆt ¢ Berˆp, p > ˆ0, 1, t > T , one means Salt-and-pepper noise, the binary noise, which occurs at the transfer of binary data in computer-networks. If Xˆt ¢ N ˆ0, σ2, σ2 A 0, t > T , then X is called Gaussian white noise. It occures e.g. in acoustical signals.

2. Gaussian random function: Definition 1.2.2 The random function X ˜Xˆt, t > T  is called Gaussian, if all of its finite-dimensional distributions are Gaussian, i.e. for all n > N, t1, . . . , tn ` T it holds ˆˆ ˆ  ˆ — ¢ ˆ Q  Xt1,...,tn X t1 ,...,X tn N µt1,...,tn , , t1,...,tn ˆ ˆ  ˆ — where the mean is given by µt1,...,tn EX t1 ,..., EX tn and the covariance P ˆˆ ˆ ˆ  ˆ n is given by t1,...,tn cov X ti ,X tj i,j 1. Exercise 1.2.1 Proof that the distribution of an Gaussian random function X is uniquely determined by its mean value function µˆt EXˆt, t > T , and covariance function Cˆs, t E XˆsXˆt¥, s, t > T , respectively.

An example for a Gaussian process is the so-called Wiener process (or Brownian motion) X ˜Xˆt, t C 0, which has the expected value zero (µˆt ¡ 0, t C 0) and the covariance function Cˆs, t min ˜s, t, s, t C 0. Usually it is addionally required that the paths of X are continuous functions. We shall investigate the regularity properties of the paths of random functions in more detail in Section 1.3. Now we can say that such a process exists with probability one (with almost surely continuous trajectories). Exercise 1.2.2 Prove that the Gaussian white noise is a Gaussian random function.

3. Lognormal- and χ2-functions: ˆ  The random function X ˜Xˆt, t > T  is called lognormal, if Xˆt eY t , where Y 6 1 General theory of random functions

˜Y ˆt, t > T  is a Gaussian random function. X is called χ2-function, if Xˆt YY ˆtY2, n where Y ˜Y ˆt, t > T  is a Gaussian random function with values in R , for which ˆ  ¢ ˆ  > ˆ   ˆ  ¢ 2 Y t N 0, I , t T ; here I is the n n -unit matrix. Then it holds that X t χn, t > T .

4. Cosine wave: º X ˜Xˆt, t > R is defined by Xˆt 2 cosˆ2πY  tZ, where Y ¢ Uˆ 0, 1¥ and Z is a random variable, which is independent of Y . Exercise 1.2.3 Let X1,X2,... be i.i.d. cosine waves. Determine the weak limit of the finite-dimensional š º1 Pn X ˆt, t > Ÿ n ª distributions of the random function n k 1 k R for .

5. Poisson process: ˜Y  Y ¢ ˆλ, λ A Let n n>N be a sequence of i.i.d. random variables n Exp 0. The stochastic ˜ ˆ  C  ˆ  ˜ > ¢ n B  process X X t , t 0 defined as X t max n N Pk 1 Yk t is called Poisson process with intensity λ A 0. Xˆt counts the number of certain events until the time t A 0, where the typical interval between two of these events is Expˆλ-distributed. These events can be claim arrivals of an insurance portfolio the records of elementary particles in the Geiger counter, etc. Then Xˆt represents the number of claims or particles within the time interval 0, t¥.

1.3 Regularity properties of trajectories

The theorem of Kolmogorov provides the existence of the distribution of a random function with given finite-dimensional distributions. However, it does not provide a statement about the properties of the paths of X. This is understandable since all random objects are defined in the almost surely sense (a.s.) in , with the exception of a set A ` Ω with PˆA 0. Example 1.3.1 Let ˆΩ, A, P ˆ 0, 1¥ , Bˆ 0, 1¥, ν1, where ν1 is the Lebesgue measure on 0, 1¥. We define X ˜Xˆt, t > 0, 1¥ by Xˆt ¡ 0, t > 0, 1¥ and Y ˜Y ˆt, t > 0, 1¥ by

1, t U, Y ˆt œ 0, sonst, where Uˆω ω, ω > 0, 1¥, is a Uˆ 0, 1¥-distributed random variable defined on ˆΩ, A, P. d Since PˆY ˆt 0 1, t > T , because of PˆU t 0, t > T , it is clear that X Y . Nevertheless, X and Y have different path properties since X has continuous and Y has discontinuous trajectories, and PˆXˆt 0, ¦t > T  1, where PˆY ˆt 0, ¦t > T  0. It may well be that the „set of exceptions“ A (see above) is very different for Xˆt for every t > T . Therefore, we require that all Xˆt, t > T , are defined simultaneously on a subset Ω0 b Ω with PˆΩ0 1. The so defined random function X˜ ¢ Ω0  T R is called modification of X ¢ ΩT R. X and X˜ differ on a set Ω~Ω0 with probability zero. Therefore we indicate later when stating that „random function X possesses a property C“ that it exists a modification of X with this property C. Let us hold it in the following definition: 1 General theory of random functions 7

Definition 1.3.1 The random functions X ˜Xˆt, t > T  and Y ˜Y ˆt, t > T  defined on the same prob- ability space ˆΩ, A, P associated with ˆSt, Btt>T have equivalent trajectories (or are called stochastically indistinguishable) if

A ˜ω > Ω ¢ Xˆω, t x Y ˆω, t for a t > T  > A and PˆA 0. This term implies that X and Y have paths, which coincide with probability one. Definition 1.3.2 The random functions X ˜Xˆt, t > T  and Y ˜Y ˆt, t > T  defined on the same probability space ˆΩ, A, P are called (stochastically) equivalent, if

Bt ˜ω > Ω ¢ Xˆω, t x Y ˆω, t > A, t > T, and PˆBt 0, t > T . We also can say that X and Y are versions or modifications of one and the same random function. If the space ˆΩ, A, P is complete (i.e. the implication of A > A ¢ PˆA 0 is for all B ` A: B > A (and then PˆB 0)), then the indistinguishable processes are stochastically equivalent, but vice versa is not always true (it is true for so-called separable processes. This is the case if T is countable). Exercise 1.3.1 Prove that the random functions X and Y in Example 1.3.1 are stochastically equivalent. Definition 1.3.3 The random functions X ˜Xˆt, t > T  and Y ˜Y ˆt, t > T  (not necessarily defined on the same probability space) are called equivalent in distribution, if PX PY on ˆSt, Btt>T . d Notation: X Y . According to Theorem 1.1.2 it is sufficient for the equivalence in distribution of X and Y that they possess the same finite-dimensional distributions. It is clear that stochastic equivalence implies equivalence in distribution, but not the other way around. Now, let T and S be Banach spaces with norms S ST and S SS , respectively. The random function X ˜Xˆt, t > T  is now defined on ˆΩ, A, P with values in ˆS, B. Definition 1.3.4 The random function X ˜Xˆt, t > T  is called

P a) stochastically continuous on T , if Xˆs ÐÐ Xˆt, for arbitrary t > T , i.e. s t

PˆSXˆs  XˆtSS A ε ÐÐ 0, for all ε A 0. s t

Lp b) Lp-continuous on T , p C 1, if Xˆs ÐÐ Xˆt, t > T , i.e. ESXˆs  XˆtSp ÐÐ 0. For p 2 s t s t the specific notation „continuity in the square mean “is used.

f.s. c) a.s. continuous on T , if Xˆs ÐÐ Xˆt, t > T , i.e., PˆXˆs ÐÐ Xˆt 1, t > T . s t s t d) continuous, if all trajectories of X are continuous functions. 8 1 General theory of random functions

In applications one is interested in the cases c) and d), although the weakest form of continuity is the stochastic continuity.

Lp-continuity Ô¨ stochastic continuity Ô a.s. continuity Ô continuity of all paths

Why are cases c) and d) important? Let us consider an example. Example 1.3.2 ¥ ˆ  0,1¥ Let T 0, 1 and Ω, A, P be the canonical probability space with Ω R , i.e. Ω Lt> 0,1¥ R. Let X ˜Xˆt, t > 0, 1¥ be a stochastic process on ˆΩ, A, P. Not all events are elements of A, like e.g. A ˜ω > Ω ¢ Xˆω, t 0 for all t > 0, 1¥ 9t> 0,1¥ ˜Xˆω, t 0, since this is an intersection of measurable events from A in uncountable number. If however X is continuous, then all of its paths are continuous functions and one can write A 9t>D ˜Xˆω, t 0, where D is a dense countable subset of 0, 1¥, e.g., D Q 9 0, 1¥. Then it holds that A > A. However, in many applications (like e.g. in financial mathematics) it is not realistic to consider stochastic processes with continuous paths as models for real phenomena. Therefore, a bigger class of possible trajectories of X is allowed: the so-called càdlàg-class (càdlàg = continue à droite, limitée à gauche (fr.)). Definition 1.3.5 A stochastic process X ˜Xˆt, t > R is called càdlàg, if all of its trajectories are right-side continuous functions, which have left-side limits. Now, we would like to consider the properties of the notion of continuity (introduced above) in more detail. One can note e.g. that the stochastic continuity is a property of the two- dimensional distribution Ps,t of X. This is shown by the following lemma. Lemma 1.3.1 Let X ˜Xˆt, t > T  be a random function associated with ˆS, B, where S and T are Banach spaces. The following statements are equivalent:

P a) Xˆs ÐÐ Y , s t0

d b) Ps,t ÐÐÐ PˆY,Y , s,t t0 where t0 > T and Y is a S-valued ASB-random element. For the stochastic continuity of X, one should choose t0 > T arbitrarily and Y Xˆt0.

Proof a ¨ b P — P — Xˆs ÐÐ Y means ˆXˆs,Xˆt ÐÐÐ ˆY,Y  . s t0 s,t t0

PˆSˆXˆs,Xˆt  ˆY,Y S A ε D PˆSXˆs  Y S A ε~   PˆSXˆt  Y S A ε~  ÐÐÐ 2 S 2 S 2 0 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ s,t t0 ˆS ˆ  S2 S ˆ  S2 1~2 X s Y S X t Y S

d P d This results in Ps,t PˆY,Y , since -convergence is stronger than -convergence. b ¨ a For arbitrary ε A 0 we consider a gε ¢ R 0, 1¥ with gεˆ0 0, gεˆx 1, 1 General theory of random functions 9

x ¶ Bεˆ0. It holds for all s, t > T that

EgεˆSXˆs  XˆtSS  PˆSXˆs  XˆtSS A ε  EˆgεˆSXˆs  XˆtSS 1ˆSXˆs  XˆtSS B 卍, ˆS ˆ   ˆ S A  B ˆS ˆ   ˆ S  ˆS  S  ˆ ˆ  ÐÐ hence P X s X t S ε Egε X s X t S RS RS gε x y S Ps,t d x, y s t0 t t0 ˆS  S  ˆ ˆ  ™ˆ  > 2 ¢ ž RS RS gε x y S PˆY,Y  d x, y 0, since PˆY,Y  is concentrated on x, y S x y and P g ˆ  ˜Xˆs Xˆs ÐÐ Y. ε 0 0. Thus s t0 is a fundamental sequence (in probability), therefore s t0

It may be that X is stochastically continuous, although all of the paths of X have jumps, i.e. X cannot possess any a.s. continuous modification. The descriptive explanation for that is that such X may have a jump at concrete t > T with probability zero. Therefore jumps of the paths of X always occur at different locations. Exercise 1.3.2 Prove that the Poisson process is stochastically continuous, although it does not possess any a.s. continuous modification. Exercise 1.3.3 Let T be compact. Prove that if X is stochastically continuous on T , then it also is uniformly stochastically continuous, i.e., for all ε, η A 0 §δ A 0, such that for all s, t > T with Ss  tST @ δ it holds that PˆSXˆs  XˆtSS A ε @ η. 2 Now let S R, EX ˆt @ ª, t > T , EXˆt 0, t > T . Let Cˆs, t E XˆsXˆt¥ be the covariance function of X. Lemma 1.3.2 2 For all t0 > T and a random variable Y with EY @ ª the following assertions are equivalent: L2 a) Xˆs ÐÐ Y s t0 b) Cˆs, t ÐÐÐ EY 2 s,t t0 Proof a ¨ b The assertion results from the Cauchy-Schwarz inequality: SCˆs, t  EY 2S SEˆXˆsXˆt  EY 2S SE ˆXˆs  Y  Y ˆXˆt  Y  Y ¥  EY 2S B ESˆXˆs  Y ˆXˆt  Y S  ESˆXˆs  Y Y S  ESˆXˆt  Y Y S ¿ Á 2 2 B ÁEˆXˆs  Y  EˆXˆt  Y  ÀÁ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ SSXˆsY SS2 SSXˆtY SS2 ¿ L2 ¿L2 ÁEY 2 EˆXˆs  Y 2  ÁEY 2 EˆXˆt  Y 2 ÐÐÐ Á Á 0 ÀÁ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ÀÁ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ s,t t0 SSXˆsY SS2 SSXˆtY SS2 L2 L2 with assumption a). b ¨ a

EˆXˆs  Xˆt2 EˆXˆs2  2E XˆsXˆt¥  EˆXˆt2 Cˆs, s  Cˆt, t  2Cˆs, t ÐÐÐ 2EY 2  2EY 2 0. s,t t0 10 1 General theory of random functions

2 2 L Thus, ˜Xˆs, s t0 is a fundamental sequence in the L -sense, and we get Xˆs ÐÐ Y . s t0 A random function X, which is continuous in the mean-square sense, may still have uncon- tinuous trajectories. In most of the cases which are practically relevant, X however has an a.s. continuous modification. Later on this will become more precise by stating a corresponding theorem. Corollary 1.3.1 The random function X, which satisfies the conditions of Lemma 1.3.2, is continuous on T in 2 the mean-square sense if and only if its covariance function C ¢ T R is continuous on the 2 ™ˆ  > 2 ¢ ž ˆ  ˆ  ˆ  > diagonal diag T s, t T s t , i.e., lims,t t0 C s, t C t0, t0 Var X t0 for all t0 T.

Proof Choose Y Xˆt0 in Lemma 1.3.2.

Remark 1.3.1 If X is not centered, then the continuity of µˆ  together with the continuity of C on diag T 2 is required to ensure the L2-continuity of X on T . Exercise 1.3.4 Give an example of a stochastic process with a.s. discontinuous trajectories, which is L2- continuous. Now we consider the property of (a.s.) continuity in more detail. As mentioned before, we can merely talk about continuous modification or version of a process. The possibility to possess such a version also depends on the properties of the two-dimensional distributions of the process. This is proven by the following theorem (originally proven by A. Kolmogorov). Theorem 1.3.1 Let X ˜Xˆt, t > a, b¥, ª @ a @ b B ª be a real-valued stochastic process. X has a continuous version, if there exist constants α, c, δ A 0 such that

 ESXˆt  h  XˆtSα @ cShS1 δ, t > ˆa, b, (1.3.1) for sufficiently small ShS.

Proof See, e.g. [7], Theorem 2.23.

Now we turn to processes with càdlàg-trajectories. Let ˆΩ, A, P be a complete probability space. Theorem 1.3.2 Let X ˜Xˆt, t C 0 be a real-valued stochastic process and D a countable dense subset of 0, ª. If

P a) X is stochastically right-handside continuous, i.e., Xˆt  h ÐÐÐ Xˆt, t > 0, ª, h 0 b) the trajectories of X at every t > D have finite right- and left-handside limits, i.e., S limh 0 Xˆt  hS @ ª, t > D a.s., then X has a version with a.s. càdlàg-paths. Without proof. 1 General theory of random functions 11

Lemma 1.3.3 Let X ˜Xˆt, t C 0 and ˜Y Y ˆt, t C 0 be two versions of a random function, both defined on the probability space ˆΩ, A, P, with property that X and Y have a.s. right-handside continuous trajectories. Then X and Y are indistinguishable.

Proof Let ΩX , ΩY be „sets of exception“, for which the trajectories of X and Y , respec- tively are not right-sided continuous. It holds that PˆΩX  PˆΩY  0. Consider At ˜ > ¢ ˆ  x ˆ  > ª 8 9 ª ω Ω X ω, t Y ω, t , t 0, and A t>Q At, where Q Q 0, . Since X and Y are stochastically equivalent, it holds that PˆA 0 and therefore

P ˆA˜ B PˆA  PˆΩX   PˆΩY  0, where A˜ A 8 ΩX 8 ΩY . Therefore Xˆω, t Y ˆω, t holds for t > Q and ω > Ω  A˜. Now, we prove this for all t C 0. For arbitrary t C 0 a sequence ˜tn ` Q exists, such that tn £ t. Since Xˆω, tn Y ˆω, tn for all n > N and ω > Ω  A˜, it holds that Xˆω, t limn ª Xˆω, tn limn ª Y ˆω, tn Y ˆω, t for t C 0 and ω > Ω  A˜. Therefore X and Y are indistinguishable. Corollary 1.3.2 If càdlàg-processes X ˜Xˆt, t C 0 and Y ˜Y ˆt, t C 0 are versions of the same random function then they are indistinguishable.

1.4 Differentiability of trajectories

Let T be a linear normed space. Definition 1.4.1 A real-valued random function X ˜Xˆt, t > T  is differentiable on T in direction h > T stochastically, in the Lp-sense, p C 1, or a.s., if

ˆ    ˆ  œ X t hl X t ˆ  > lim Xh t , t T l 0 l exists in the corresponding sense, namely stochastically, in the Lp-space or a.s.. The Lemmas 1.3.1- 1.3.2 show that the stochastic differentiability is a property that is deter- XˆthlXˆt mined by three-dimensional distributions of X (because the common distribution of œ l Xˆthl Xˆt œ and l should converge weakly), whereas the differentiability in the mean-square sense is determined by the smoothness of the covariance function Cˆs, t. Exercise 1.4.1 Show that

1. the Wiener process is not stochastically differentiable on 0, ª.

2. the Poisson process is stochastically differentiable on 0, ª, however not in the Lp-mean, p A 1. Lemma 1.4.1 A centered random function X ˜Xˆt, t > T  (i.e., EXˆt ¡ 0, t > T ) with E X2ˆt¥ @ ª, t > T is L2-differentiable in t > T in direction h > T if its covariance function C is differentiable œœ ∂2Cˆs,t œ twice in ˆt, t in direction h, i.e., if § C ˆt, t U .X ˆt is L2-continuous in t > T hh ∂sh∂th s t h 12 1 General theory of random functions

œœ ∂2Cˆs,t œœ if C ˆs, t is continuous in s t. Moreover, C ˆs, t is the covariance function of hh ∂sh∂th hh œ ˜ œ ˆ  >  Xh Xh t , t T .

Proof According to Lemma 1.3.2 it is enough to show that

ˆ    ˆ  ˆ  œ   ˆ  ŒX t lh X t X s l h X s ‘ I limœ E œ l,l 0 l l exists for s t. Indeed we get

1 œ œ I ŠCˆt  lh, s  l h  Cˆt  lh, s  Cˆt, s  l h  Cˆt, s llœ œ œ ˆ     ˆ   ˆ    ˆ  œœ 1 ŒC t lh, s l h C t lh, s  C t, s l h C t, s ‘ ÐÐÐ ˆ  œ œ œ Chh s, t . l l l l,l 0

All other statements of the lemma result from this relation.

Remark 1.4.1 The properties of the L2-differentiability and a.s. differentiability of random functions are disjoint in the following sense: there are stochastic processes that have L2-differentiable paths, although they are a.s. discontinuous, and vice versa, processes with a.s. differentiable paths are not always L2-differentiable, since e.g. the first derivative of their covariance function is not continuous. Exercise 1.4.2 Give appropriate examples!

1.5 Moments und covariance

Let X ˜Xˆt, t > T  be a random function that is real-valued, and let T be an arbitrary index space. Definition 1.5.1 ˆj1,...,jn n The mixed moment µ ˆt1, . . . , tn of X of order ˆj1, . . . , jn > N , t1, . . . , tn > T is given ˆj1,...,jn j1 jn by µ ˆt1, . . . , tn E ¡X ˆt1 ... X ˆtn¦, where it is required that the expected value exists and is finite. Then it is sufficient to assume that ESXˆtSj @ ª for all t > T and j j1  ...  jn. Important special cases:

ˆ  1. µ ˆt µ 1 ˆt EXˆt, t > T – mean value function of X.

ˆ  2. µ 1,1 ˆs, t E XˆsXˆt¥ Cˆs, t – (non-centered) covariance function of X. Whereas ˆ  the centered covariance function is: Kˆs, t covˆˆXˆs,Xˆt µ 1,1 ˆs, t  µˆsµˆt, s, t > T .

Exercise 1.5.1 Show that the centered covariance function of a real-valued random function X

1. is symmetric, i.e., Kˆs, t Kˆt, s, s, t > T . 1 General theory of random functions 13

2. is positive semidefinite, i.e., for n > N, t1, . . . , tn > T , z1, . . . , zn > R it holds that n Q Kˆti, tjzizj C 0. i,j 1

3. satisfies Kˆt, t Var Xˆt, t > T . Property 2) also holds for the non-centered covariance function Cˆs, t. The mean value function µˆt shows a (non random) trend. If µˆt is known, the random function X can be centered by considering a random function Y ˜Y ˆt, t > T  with Y ˆt Xˆt  µˆt, t > T . The covariance function Kˆs, t (Cˆs, t, respectively) contains information about the depen- Kˆs,t dence structure of X. Sometimes the correlation function Rˆs, t » for all s, t > T : Kˆs,sKˆt,t Kˆs, s Var Xˆs A 0, Kˆt, t Var Xˆt A 0 is used instead of K and C, respectively. Because of the Cauchy-Schwarz inequality it holds that SRˆs, tS B 1, s, t > T . The set of all mixed moments in general does not (uniquely) determine the distribution of a random function. Exercise 1.5.2 Give an example of different random functions X ˜Xˆt, t > T  und Y ˜Y ˆt, t > T , for which it holds that EXˆt EY ˆt, t > T and EˆXˆsXˆt EˆY ˆsY ˆt, s, t > T . Exercise 1.5.3 Let µ ¢ T R be a measurable function and K ¢ T  T R be a positive semidefinite sym- metric function. Prove that a random function X ˜Xˆt, t > T  exists with EXˆt µˆt, covˆXˆs,Xˆt Kˆs, t, s, t > T . Let now X ˜Xˆt, t > T  be a real-valued random function with E SXˆtSk @ ª, t > T , for a k > N. Definition 1.5.2 k The mean increment of order k of X is given by γkˆs, t EˆXˆs  Xˆt , s, t > T . ˆ  1 ˆ  1 ˆ ˆ  ˆ 2 > Special attention is paid to the function γ s, t 2 γ2 s, t 2 E X s X t , s, t T , which is called variogram of X. In geostatistics the variogram is often used instead of the covariance function. A lot of times we discard the condition EX2ˆt @ ª, t > T , instead we assume that γˆs, t @ ª for all s, t > T . Exercise 1.5.4 Prove that there exist random functions without finite second moments with γˆs, t @ ª, s, t > T . Exercise 1.5.5 Show that for a random function X ˜Xˆt, t > T  with mean value function µ and covariance function K it holds that: Kˆs, s  Kˆt, t 1 γˆs, t  Kˆs, t  ˆµˆs  µˆt2, s, t > T. 2 2 2 If the random function X is complex-valued, i.e., X ¢ Ω  T C, with E SXˆtS @ ª, t > T , then the covariance function of X is introduced as Kˆs, t EˆXˆs  EXˆsˆXˆt  EXˆt, s, t > T , where z is the complex conjugate of z > C. Then it holds that Kˆs, t Kˆt, s, s, t > T , and K is positive semidefinite, i.e, for all n > N, t1, . . . , tn > T , z1, . . . , zn > C it holds that n ˆ  C Pi,j 1 K ti, tj zizj 0. 14 1 General theory of random functions

1.6 Stationarity and Independence

T be a subset of the linear with operations ,  over space R. Definition 1.6.1 The random function X ˜Xˆt, t > T  is called stationary (strict sense stationary) if for all n > N, h, t1, . . . , tn > T with t1  h, . . . , tn  h > T it holds that:

PˆXˆt1,...,Xˆtn PˆXˆt1h,...,Xˆtnh, i.e., all finite-dimensional distributions of X are invariant with repsect to translations in T . Definition 1.6.2 A (complex-valued) random function X ˜Xˆt, t > T  is called second-order stationary (or wide sense stationary) if ESXˆtS2 @ ª, t > T , and µˆt ¡ EXˆt ¡ µ, t > T , Kˆs, t covˆXˆs,Xˆt Kˆs  h, t  h for all h, s, t > T ¢ s  h, t  h > T . If X is second-order stationary, it is convenient to introduce a function Kˆt ¢ Kˆ0, t, t > T whereby 0 > T . Strict sense stationarity and wide sense stationarity do not result from each other. However it is clear that if a complex-valued random function is strict sense stationary and possesses finite second-order moments, then the function is also second-order stationary. Definition 1.6.3 A real-valued random function X ˜Xˆt, t > T  is intrinsic second-order stationary if γkˆs, t, s, t > T exist for k B 2, and for all s, t, h > T , s  h, t  h > T it holds that γ1ˆs, t 0, γ2ˆs, t γ2ˆs  h, t  h. For real-valued random functions, intrinsic second-order stationarity is more general than second-order stationarity since the existence of ESXˆtS2, t > T is not required. The analogue of the stationarity of increments of X also exists in strict sense. Definition 1.6.4 Let X ˜Xˆt, t > T  be a real-valued stochastic process, T ` R. It is said that X

1. possesses stationary increments if for all n > N, h, t0, t1, t2, . . . , tn > T , with t0 @ t1 @ t2 @ ... @ tn, ti  h > T , i 0, . . . , n the distribution of — ˆXˆt1  h  Xˆt0  h,...,Xˆtn  h  Xˆtn1  h does not depend on h.

2. possesses independent increments if for all n > N, t0, t1, . . . , tn > T with t0 @ t1 @ ... @ tn the random variables Xˆt0,Xˆt1Xˆt0,...,XˆtnXˆtn1 are pairwise independent.

Let ˆS1, B1 and ˆS2, B2 be measurable spaces. In general it is said that two random elements X ¢ Ω S1 and X ¢ Ω S2 are independent on the same probability space ˆΩ, A, P if PˆX > A1,Y > A2 PˆX > A1PˆY > A2 for all A1 > B1, A2 > B2. This definition can be applied to the independence of random functions X and Y with phase space ˆST , BT , since they can be considered as random elements with S1 S2 ST , B1 B2 BT (cf. Lemma 1.1.1). The same holds for the independence of a random element (or a random function) X and of a sub-σ-algebra G > A: this is the case if Pˆ˜X > A 9 G PˆX > APˆG, for all A > B1, G > G (or A > BT , G > G). 1 General theory of random functions 15

1.7 Processes with independent increments

In this section we concentrate on the properties and existence of processes with independent increments. Let ˜ϕs,t, s, t C 0 be a family of characteristic functions of probability measures Qs,t, s, t C Bˆ  z > s, t C ϕ ˆz R eizxQ ˆdx 0 on R , i.e., for R, 0 it holds that s,t R s,t . Theorem 1.7.1 There exists a stochastic process X ˜Xˆt, t C 0 with independent increments with the property that for all s, t C 0 the characteristic function of Xˆt  Xˆs is equal to ϕs,t if and only if

ϕs,t ϕs,uϕu,t (1.7.1) for all 0 B s @ u @ t @ ª. Thereby the distribution of Xˆ0 can be chosen arbitrarily.

Proof The necessity of the condition (1.7.1) is clear since for all s > ˆ0, ª ¢ s @ u @ t it holds XˆtXˆs Xˆt  Xˆu  Xˆu  Xˆs, and XˆtXˆu and XˆuXˆs are independent. ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ Y1 Y2

Then it holds ϕs,t ϕY1Y2 ϕY1 ϕY2 ϕs,uϕu,t. Now we prove the sufficiency. If the existence of a process X with independent increments and property ϕXˆtXˆs ϕs,t on a probability space ˆΩ, A, P had already been proven, one could define the characteristic functions of all of its finite-dimensional distributions with the help of ˜ϕs,t as follows. — Let n > N, 0 t0 @ t1 @ ... @ tn @ ª and Y ˆXˆt0,Xˆt1  Xˆt0,...,Xˆtn  Xˆtn1 . The independence of increments results in ˆ  i`z,Y e ˆ  ˆ  ˆ  > n1 ϕY z0, z1, . . . , zn Ee ϕXˆt0 z0 ϕt0,t1 z1 . . . ϕtn1,tn zn , z R , ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ z ˆ  ˆ  where the distribution of X t0 is an arbitrary probability measure Q0 on B R . For Xt0,...,tn ˆ ˆ  ˆ  ˆ — X t0 ,X t1 ,...,X tn however it holds that Xt0,...,tn AY , where

’ 1 0 0 ... 0 “ – — – 1 1 0 ... 0 — – — A – ... — . – 1 1 1 0 — – ...... — ” 1 1 1 ... 1 •

ˆ  ˆ  i`z,AY e i`A—z,Y e ˆ —  Then ϕXt ,...,tn z ϕAY z Ee Ee ϕY A z holds. Therefore the distribu- 0 ˆ  ˆ  ˆ  ˆ  tion of Xt0,...,tn possesses the characteristic function ϕXt ,...,t z ϕQ0 l0 ϕt0,t1 l1 . . . ϕtn1,tn ln , — — 0 n where l ˆl1, l1, . . . , ln A z, thus ¢   ¨ l0 z0 ... zn ¨   ¦ l1 z1 ... zn ¨ ¦ ¨ ¤ ln zn

ϕ ˆ  ϕ ϕ ˆz , . . . , z  ϕ ˆ , z , . . . , z  z > Thereby X t0 Q0 and Xt1,...,tn 1 n Xt0,...,tn 0 1 n holds for all i R. Now we prove the existence of such a process X. 16 1 General theory of random functions

For that we construct the family of characteristic functions ˜ @ @ @ @ ª >  ϕt0 , ϕt0,t1,...,tn , ϕt1,...,tn , 0 t0 t1 ... tn , n N ˜ B @  from ϕQ0 and ϕs,t, 0 s t as above, thus ˆ  ˆ  > ϕt0 ϕQ0 , ϕt1,...,tn 0, z1, . . . , zn ϕt0,t1,...,tn 0, z1, . . . , zn , zi R, ˆ  ˆ    ˆ    ˆ  ϕt0,...,tn z ϕt0 z1 ... zn ϕt0,t1 z1 ... zn . . . ϕtn1,tn zn . Now we have to check whether the corresponding probability measures of these characteristic functions fulfill the conditions of Theorem 1.1.2. We will do that in equivalent form since by Exercise 1.8.1 the conditions of symmetry and consistency in Theorem 1.1.2 are equivalent to: ϕ ˆz , . . . , z  ϕ ˆz , . . . , z  ˆ , , . . . , n ( a) ti0 ,...,tin i0 in t0,...,tn 0 n for an arbitrary permutation 0 1 ˆi0, i1, . . . , in, ˆ  ˆ  b) ϕt0,...,tm1,tm1,...,tn z0, . . . , zm1, zm1, . . . , zn ϕt0,...,tn z0,..., 0, . . . , zn , for all z0, . . . , zn > R, m > ˜1, . . . , n. Condition a) is obvious. Conditon b) holds since ˆ     ˆ    ˆ  ϕtm1,tm 0 zm1 ... zn ϕtm,tm1 zm1 ... zn ϕtm1,tm1 zm1, . . . , zn for all m > ˜1, . . . , n. Thus, the existence of X is proven.

Example 1.7.1 1. If T N0 N8˜0, then X ˜Xˆt, t > N0 has independent increments ˆ  d n ˜  d if and only if X n Pi 0 Yi, where Yi are independent random variables and Yn Xˆn  Xˆn  1, n > N. Such a process X is called random walk. It also may be defined m for Yi with values in R . 2. The Poisson process with intensity λ has independent increments (we will show that later).

3. The Wiener process possesses independent increments. Exercise 1.7.1 Proof that! Exercise 1.7.2 Let X ˜Xˆt, t C 0 be a process with independent increments and g ¢ 0, ª R an arbitrary (deterministic) function. Show that the process Y ˜Y ˆt, t C 0 with Y ˆt Xˆtgˆt, t C 0, also possesses independent increments.

1.8 Additional exercises

Exercise 1.8.1 ˆ n ˆ n Prove the following assertion: The family of probability measures Pt1,...,tn on R , B R , — n n C 1, t ˆt1, . . . , tn > T fulfills the conditions of the theorem of Kolmogorov if and only if — n for n C 2 and for all s ˆs1, . . . , sn > R the following conditions are fulfilled: — — a) ϕP ˆˆs1, . . . , sn  ϕP ˆˆs ˆ , . . . , s ˆ   for all π > Sn. t1,...,tn tπˆ1,...,tπˆn π 1 π n 1 General theory of random functions 17

— — ϕ ˆˆs , . . . , s    ϕ ˆˆs , . . . , s  ,   b) Pt1,...,tn1 1 n 1 Pt1,...,tn 1 n 1 0 .

Remark: ϕˆ  denotes the characteristic function of the corresponding measure. Sn denotes the group of all permutations π ¢ ˜1, . . . , n ˜1, . . . , n. Exercise 1.8.2 Show the existence of a random function whose finite-dimensional distributions are multivariate- ˆ  normally distributed and explicitly give the measurable spaces Et1,...,tn , Et1,...,tn . Exercise 1.8.3

Give an example of a family of probability measures Pt1,...,tn , which do not fulfill the conditions of the theorem of Kolmogorov. Exercise 1.8.4 Let X ˜Xˆt, t > T  and Y ˜Y ˆt, t > T  be two stochastic processes which are defined on the same complete probability space ˆΩ, F, P and which take values in the measurable space ˆS, B.

a) Proof that: X and Y are stochastically equivalent Ô¨ PX PY .

b) Give an example of two processes X and Y for which holds: PX PY , but X and Y are not stochastically equivalent.

c) Proof that: X and Y are stochastically indistinguishable Ô¨ X and Y are stochastically equivalent.

d) Proof in the case of countability of T : X and Y are stochastically equivalent Ô¨ X and Y are stochastically indistinguishable.

e) Give in the case of uncountability of T an example of two processes X and Y for which holds: X and Y are stochastically equivalent but not stochastically indistinguishable.

Exercise 1.8.5 Let W ˜W ˆt, t > R be a Wiener Process. Which of the following processes are Wiener processes as well?

a) W1 ˜W1ˆt ¢ W ˆt, t > R, º b) W2 ˜W2ˆt ¢ tW ˆ1, t > R,

c) W3 ˜W3ˆt ¢ W ˆ2t  W ˆt, t > R. Exercise 1.8.6 Given a stochastic process X ˜Xˆt, t > 0, 1¥ which consists of idependent and identically distributed random variables with density fˆx, x > R. Show that such a process can not be continuous in t > 0, 1¥. Exercise 1.8.7 Give an example of a stochastic process X ˜Xˆt, t > T  which is stochastically continuous on T , and prove why this is the case. Exercise 1.8.8 In connection with the continuity of stochastic processes the so-called criterion of Kolmogorov 18 1 General theory of random functions plays a central role. (see also theorem 1.3.1 in the lecture notes): Let X ˜Xˆt, t > a, b¥ be a real-valued stochastic process. If constants α, ε A 0 and C ¢ Cˆα, ε A 0 exist such that  ESXˆt  h  XˆtSα B CShS1 ε (1.8.1) for sufficient small h, then the process X possesses a continuous modification. Show that:

a) If you fix the variable ε 0 in condition (1.8.1), then in general the condition is not sufficient for the existence of a continuous modification. Hint: Consider the Poisson process.

b) The Wiener process W ˜W ˆt, t > 0, ª possesses a continuous modification. Hint: Consider the case α 4. Exercise 1.8.9 Show that the Wiener process W is not stochastically differentiable at any point t > 0, ª. Exercise 1.8.10 Show that the covariance function Cˆs, t of a complex-valued stochastic process X ˜Xˆt, t > T 

a) is symmetric, i.e. Cˆs, t Cˆt, s, s, t > T ,

b) fulfills the identity Cˆt, t Var Xˆt, t > T ,

c) is positive semidefinite, i.e. for all n > N, t1, . . . , tn > T , z1, . . . , zn > C it holds that: n n Q Q Cˆti, tjziz¯j C 0. i 1 j 1

Exercise 1.8.11 Show that it exists a random function X ˜Xˆt, t > T  which simultaneously fulfills the conditions:

• The second moment EX2 does not exist.

• The variogram γˆs, t is finite for all s, t > T . Exercise 1.8.12 Give an example of a stochastic process X ˜Xˆt, t > T  whose paths are simultaneously L2-differentiable but not almost surely differentiable, and prove why this is the case. Exercise 1.8.13 Give an example of a stochastic process X ˜Xˆt, t > T  whose paths are simultaneously almost surely differentiable but not L1-differentiable, and prove why this is the case. Exercise 1.8.14 Proof that the Wiener process possesses independent increments. Exercise 1.8.15 Proof: A (real-valued) stochastic process X ˜Xˆt, t > 0, ª with independent increments already has stationary increments if the distibution of the random variable Xˆt  h  Xˆh is independent of h. 2 Counting processes

In this chapter we consider several examples of stochastic processes which model the counting of events and thus possess piecewise constant paths. ˆ , A, P ˜S  Let Ω be a probability space and n n>N a non-decreasing sequence of a.s. non- negative random variables, i.e. 0 B S1 B S2 B ... B Sn B .... Definition 2.0.1 The stochastic process N ˜Nˆt, t C 0 is called counting process if

ª Nˆt Q 1ˆSn B t, n 1 where 1ˆA is the indicator function of the event A > A.

Nˆt counts the events which occur at Sn until time t. Sn e.g. may be the time of occurence of

1. the n-th elementary particle in the Geiger counter, or

2. a damage in the insurance of material damage, or

3. a data paket at a server within a computer network, etc.

A special case of the counting processes are the so-called renewal processes.

2.1 Renewal processes

Definition 2.1.1 ˜  ˆ A  A Let Tn n>N be a sequence of i.i.d. non-negative random variables with P T1 0 0.A ˜ ˆ  C  ˆ  n > counting process N N t , t 0 with N 0 0 a.s., Sn Pk 1 Tk, n N, is called renewal process. Thereby Sn is called the time of the n-th renewal, n > N. The name „renewal process“ is given by the following interpretation. The „interarrival times“ Tn are interpreted as the lifetime of a technical spare part or mechanism within a system, thus Sn is the time of the n-th break down of the system. The defective part is immediately replaced by a new part (comparable with the exchange of a lightbulb). Thus, Nˆt is the number of repairs (the so-called „renewals“) of the system until time t.

Remark 2.1.1 1. It is Nˆt ª if Sn B t for all n > N.

2. Often it is assumed that only T2,T3,... are identically distributed with ETn @ ª. The distribution of T1 is freely selectable. Such a process N ˜Nˆt, t C 0 is called delayed renewal process (with delay T1).

3. Sometimes the requirement Tn C 0 is omitted.

19 20 2 Counting processes

Abb. 2.1: Konstruktion und Trajektorien eines Erneuerungsprozesses

˜  Pn > 4. It is clear that Sn n>N0 with S0 0 a.s., Sn k 1 Tk, n N is a random walk.

œ 5. If one requires that the n-th exchange of a defective part in the system takes a time Tn, ˜  œ > then by Tn Tn Tn, n N a different renewal process is given. Its stochastic property does not differ from the process which is given in definition 2.1.1.

In the following we assume that µ ETn > ˆ0, ª, n > N. Theorem 2.1.1 (Individual ergodic theorem): Let N ˜Nˆt, t C 0 be a renewal process. Then it holds that:

Nˆt 1 lim a.s.. t ª t µ

C > ˜ ˆ   ˜ B @  B @ Proof For all t 0 and n N it holds that N t n Sn t Sn1 , therefore SNˆt t SNˆt1 and

SNˆt t SNˆt1 Nˆt  1 B B . Nˆt Nˆt Nˆt  1 Nˆt

SNˆt a.s a.s. t a.s If we can show that ˆ  ÐÐ µ and Nˆt ÐÐ ª, then ˆ  ÐÐ µ holds and therefore the N t t ª t ª N t t ª assertion of the theorem. According to the strong law of large numbers of Kolmogorov (cf. lecture notes „Wahrschein- Sn a.s. a.s. lichkeitsrechnung“ (WR), theorem 7.4) it holds that ÐÐÐ µ, thus Sn ÐÐÐ ª and therefore n n ª n ª PˆNˆt @ ª 1 since PˆNˆt ª Pˆ Sn B t, ¦n 1  Pˆ§n ¢ ¦m > N0 Snm A t 1  1 0. ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ a.s 1, if SnÐÐÐ ª n ª Then Nˆt, t C 0, is a real random variable. a.s. We show that Nˆt ÐÐ ª. All trajectories of Nˆt are monotonously non-decreasing in t C 0, t ª 2 Counting processes 21

thus § limt ª Nˆω, t for all ω > Ω. Moreover it holds that

ˆ‡ Pˆ lim Nˆt @ ª lim Pˆ lim Nˆt @ n lim lim PˆNˆt @ n t ª n ª t ª n ª t ª n lim lim PˆSn A t lim lim PˆQ Tk A t n ª t ª n ª t ª k 1 n t B lim lim Q PˆTk A  0. n ª t ª n k 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ÐÐ 0 t ª ‡ ˜ ˆ  @  ˜§ > ¢ ¦ C ˆ  @  8 9 The equality ( ) holds since limt ª N t n t0 Q t t0 N t n t0>Q t>Q tCt0 ˜ ˆ  @  ˜ ˆ  @  N t n lim inf t>Q N t n , then the continuity of the probability measure is used, t ª S 9 ˜ > ¢ C  > Sn Nˆt where Q Q R q Q q 0 . Since for every ω Ω it holds that limn ª n limt ª Nˆt ˆ  SNˆt a.s (the of a realization of N is a subsequence of N), it holds that limt ª Nˆt µ.

Remark 2.1.2 One can generalize the ergodic theorem to the case of non-identically distributed Tn. Thereby 1 n we require that µn ETn, ˜Tn  µn > are uniformly integrable and P µk ÐÐÐ µ A 0. n N n k 1 n ª Nˆt P Then we can prove that ÐÐ 1 (cf. [2], page 276). t t ª µ Theorem 2.1.2 (Central limit theorem): 2 If µ > ˆ0, ª, σ Var T1 > ˆ0, ª, it holds that ˆ   t 3 N t µ d µ 2 º ÐÐ Y, σ t t ª where Y ¢ N ˆ0, 1.

Proof According to the central limit theorem for sums of i.i.d. random variables (cf. theorem 7.5, WR) it holds that Sn  nµ d º ÐÐÐ Y. (2.1.1) nσ2 n ª ¥ > σ2 Let x be the whole part of x R. It holds for a µ3 that ˆ   t º ’N t µ “ t P º B x P ‹Nˆt B x at   P ‰Smˆt A tŽ , ” at • µ º ˆ  ¢  t §  C ˆ  ª where m t x at µ 1, t 0, and limt ª m t . Therefore we get that

R t R R ’Nˆt  “ R RP º µ B x  ϕˆxR TP ‰S A tŽ  ϕˆxT R ” • R mˆt R at R R R R ’Smˆt  µmˆt t  µmˆt“ R RP » A »  ϕˆxR ¢ I ˆx R ” • R t R σ mˆt σ mˆt R 22 2 Counting processes for arbitrary t C 0 and x > R, where ϕ is the distribution function of the N ˆ0, 1-distribution.  ˆ  t »µm t For fixed x > we introduce Zt   x, t C 0. Then it holds that R σ mˆt R R R ’Smˆt  µmˆt “ R I ˆx RP »  Z A x  ϕˆxR . t R ” t • R R σ mˆt R

If we can prove that Zt ÐÐ 0, then applying (2.1.1) and the theorem of Slutsky (theorem t ª S µmˆt d mˆ»t 6.4.1, WR) would result in  Zt ÐÐ Y ¢ N ˆ0, 1 since Zt ÐÐ 0 a.s. results σ mˆt t ª t ª d in Zt ÐÐ 0. Therefore we could write Itˆx ÐÐ Sϕ¯ˆx  ϕˆxS Sϕˆx  ϕˆxS 0, where t ª t ª ϕ¯ˆx 1  ϕˆx is the tail function of the N ˆ0, 1-distribution, and the property of symmetry of N ˆ0, 1 ¢ ϕ¯ˆx ϕˆx, x > R was used.  ˆ  º t »µm t t Now we show that Zt ÐÐ 0, thus ÐÐ x. It holds that mˆt x at   εˆt, t ª σ mˆt t ª µ where εˆt > 0, 1. Then it holds that º º t  µmˆt t  µx at  t  µεˆt at  µ µεˆt » » x ¼ º  » σ mˆt σ mˆt  t  ˆ  σ mˆt σ x at µ ε t xµ µ  εˆt  ½  » ˆ  ˆ  σ ºx  1  ε t σ m t at µa at x µ µεˆt ½ σ  » ÐÐ x. 2 ˆ  ˆ  t ª µ  ºx  ε t σ m t σ2 at at ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ÐÐ 0 ª ÐÐ x t t ª

Remark 2.1.3 In Lineberg form, the central limit theorem can also be proven for non-identically distributed Tn, cf. [2], pages 276 - 277. Definition 2.1.2 The function Hˆt ENˆt, t C 0 is called renewal function of the process N (or of the sequence ˜S  n n>N). Let FT ˆx PˆT1 B x, x > R be the distribution function of T1. For arbitrary distribution ¢ ¥ ‡ ‡ ˆ  x ˆ   ˆ  functions F,G R 0, 1 the convolution F G is defined as F G x Rª F x y dG y . ‡k The k-fold convolution F of the distribution F with itself, k > N0, is defined inductive: ‡0 F ˆx 1ˆx > 0, ª, x > R, ‡1 F ˆx F ˆx, x > R, ‡ˆk1 ‡k F ˆx F ‡ F ˆx, x > R. Lemma 2.1.1 The renewal function H of a renewal process N is monotonously non-decreasing and right-sided continuous on R. Moreover it holds that ª ª ˆ  ˆ B  ‡nˆ  C H t Q P Sn t Q FT t , t 0. (2.1.2) n 1 n 1 2 Counting processes 23

Proof The monotony and right-sided continuity of H are consequences from the almost surely monotony and right-sided continuity of the trajectories of N. Now we prove (2.1.2):

ª ª ª ª ˆ‡ ˆ  ˆ  ˆ B  ˆ B  ˆ B  ‡nˆ  H t EN t E Q 1 Sn t Q E1 Sn t Q P Sn t Q FT t , n 1 n 1 n 1 n 1 ˆ B  ˆ   B  ‡nˆ  C ˆ‡ since P Sn t P T1 ... Tn t FT t , t 0. The equality holds for all partial sums on both sides, therefore in the limit as well.

Except for exceptional cases it is impossible to calculate the renewal function H by the formula (2.1.2) analytically. Therefore the Laplace transform of H is often used in calulations. For a monotonously (e.g. monotonously non-decreasing) right-sided continuous function G ¢ ª ˆ ˆ  ª sx ˆ  C 0, R the Laplace transform is defined as lG s R0 e dG x , s 0. Here the integral is to be understood as the Lebesgue-Stieltjes integral, thus as a Lebesgue integral with respect to ˆˆ ¥ ˆ  ˆ  B @ @ ª the measure µG on BR defined by µG x, y G y G x , 0 x y , if G is monotonously non-decreasing. Just to remind you: the Laplace transform ˆlX of a random variable X C 0 is defined by ˆ ˆ  ª sx ˆ  C lX s R0 e dFX x , s 0. Lemma 2.1.2 For s A 0 it holds that: ˆl ˆs ˆl ˆs T1 . H  ˆ ˆ  1 lT1 s Proof It holds that:

ª ª ª ª ª ˆ  sx ˆ  (2.1.2 ) sx Œ ‡nˆ ‘ sx ‡nˆ  ˆlH s S e dH x S e d Q FT x Q S e dF x 0 0 n 1 n 1 0 ª ª ˆ ˆ  n lT1 s Q ˆl   ˆs Q ‰ˆl ˆsŽ , T1 ... Tn T1  ˆ ˆ  n 1 n 1 1 lT1 s

A ˆ ˆ  @ Pª ‰ˆ ˆ Žn where for s 0 it holds that lT1 s 1 and thus the geometric n 1 lT1 s converges.

Remark 2.1.4 If N ˜Nˆt, t C 0 is a delayed renewal process (with delay T1), the statements of lemmas 2.1.1- 2.1.2 hold in the following form:

1. ª ˆ  Qˆ ‡ ‡nˆ  C H t FT1 FT2 t , t 0, n 0 C where FT1 and FT2 , respectively are the distribution functions of T1 and Tn, n 2, respectively.

2. ˆlT ˆs ˆl ˆs 1 , s C 0, (2.1.3) H  ˆ ˆ  1 lT2 s ˆ ˆ C where lT1 and lT2 are the Laplace transforms of the distribution of T1 and Tn, n 2. 24 2 Counting processes

For further observations we need a theorem (of Wald) about the expected value of a sum (with random number) of independent random variables. Definition 2.1.3 ν ˜X  Let be a N-valued random variable and be n n>N a sequence of random variables defined on the same probability space. ν is called independent of the future, if for all n > N the event ˜ν B n does not depend on the σ-algebra σˆ˜Xk, k A n. Theorem 2.1.3 (Wald’s identity): ˜X  ESX S @ ª EX a n > ν Let n n>N be a sequence of random variables with sup n , n , N and be a N-valued random variable which is independent of the future, with Eν @ ª. Then it holds that

ν EˆQ Xn a Eν. n 1 n > ª ˆ C  Proof Calculate Sn Pk 1 Xk, n N. Since Eν Pn 1 P ν n , the theorem follows from Lemma 2.1.3.

Lemma 2.1.3 (Kolmogorov-Prokhorov): Let ν be a N-valued random variable which is independent of the future and it holds that ª Q Pˆν C nESXnS @ ª. (2.1.4) n 1 ª ˆ C  C Then ESν Pn 1 P ν n EXn holds. If Xn 0 a.s., then condition (2.1.4) is not required. ν ª ˆ C  Proof It holds that Sν Pn 1 Xn Pn 1 Xn1 ν n . We introduce the notation Sν,n n ˆ C  > C > ¡ Pk 1 Xk1 ν k , n N. First, we prove the lemma for Xn 0 f.s., n N. It holds Sν,n Sν, n ª for every ω > Ω, and thus according to the monotone convergence theorem it holds that: n ˆ ˆ C  ˜ C  ˜ B  c ESν limn ª ESν,n lim Pk 1 E Xk1 ν k . Since ν k ν k 1 does not depend on σˆXk ` σˆ˜Xn, n C k it holds that EˆXk1ˆν C k EXkPˆν C k, k > N, and thus ª ˆ C  ESν Pn 1 P ν n EXn. S S n n ˆ C  > Now, let Xn be arbitrary. Take Yn Xn , Zn Pn 1 Yn, Zν,n Pk 1 Yk1 ν k , n N. C > ª ˆ S ˆ C  @ ª Since Yn 0, n N, it holds that EZν Pn 1 E Xn P ν k from (2.1.4). Since SSν,nS B Zν,n B Zν, n > N, according to the dominated convergence theorem of Lebesgue it holds ª ˆ C  that ESν limn ª ESν,n Pn 1 EXnP ν n , where this series converges absolutely.

Conclusion 2.1.1 1. Hˆt @ ª, t C 0.

2. For an arbitrary Borel measurable function g ¢ R R and the renewal process N ˜Nˆt, t C 0 with interarrival times ˜Tn, Tn i.i.d., µ ETn > ˆ0, ª it holds that

’Nˆt1 “ E Q gˆT  ˆ  HˆtEgˆT , t C . ” n • 1 1 0 k 1

Proof 1. For every t C 0 it is obvious that ν 1  Hˆt does not depend on the future of ˜  ˆ  > Tn n>N, the rest follows from theorem 2.1.3 with Xn g Tn , n N. ˆs 2. For s A 0 consider Tn min˜Tn, s, n > N. Choose s A 0 such that for freely selected A ¢ ˆs ˆs C  A ˆs (but fixed) ε 0 µ ET1 µ ε 0. Let N be the renewal process which is based 2 Counting processes 25

˜ ˆs ˆsˆ  Pª ˆ ˆs B  C on the sequence Tn n>N of interarrival times: N t n 1 1 Tn t , t 0. It holds ˆ  Nˆt B N s ˆt, t C 0, a.s., according to conclusion 2.1.1:

ˆ  ˆ  ˆ  ˆµ  午EN ˆsˆt   B µˆsˆEN ˆsˆt   ES s EˆS s  T s  B t  s, 1 1 N ˆsˆt1 N ˆsˆt N ˆsˆt1 ´¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ Bt Bs

C ˆs ˆs   ˆs > ˆ  ˆ  B ˆsˆ  B ts C t 0, where Sn T1 ... Tn , n N. Thus H t EN t EN t µε , t 0. A Hˆt B 1 ˆ  @ ª Since ε 0 is arbitrary, it holds that lim supt ª t µ , and also our assertion H t , t C 0.

Conclusion 2.1.2 (Elementary renewal theorem): For a renewal process N as defined in conclusion 2.1.1, 1) it holds:

Hˆt 1 lim . t ª t µ

Hˆt B 1 Proof In conclusion 2.1.1, part 2) we already proved that lim supt ª t µ . If we show Hˆt C 1 lim inft ª t µ , our assertion would be proven. According to theorem 2.1.1 it holds that Nˆt ÐÐ 1 a.s., therefore according to Fatou’s lemma t t ª µ

1 Nˆt ENˆt Hˆt E lim inf B lim inf lim inf . µ t ª t t ª t t ª t

Remark 2.1.5 1. We can prove that in the case of the finite second moment of Tn (µ2 2 @ ª ˆ  ª ET1 ) we can derive a more exact asymptotics for H t , t :

t µ2 Hˆt   oˆ1, t ª. µ 2µ2

2. The elementary renewal theorem also holds for delayed renewal processes, where µ ET2. ˆ  ˆ  Pª ‡nˆ  > ˆ  We define the renewal measure H on B R by H B n 1 RB dFT x , B B R . It holds Hˆˆª, t¥ Hˆt, Hˆˆs, t¥ Hˆt  Hˆs, s, t C 0, if H is the renewal function as well as the renewal measure. Theorem 2.1.4 (Fundamental theorem of the renewal theory): ˜ ˆ  C  ˜  Let N N t , t 0 be a (delayed) renewal process associated with the sequence Tn n>N, where Tn, n > N are independent, ˜Tn, n C 2 identically distributed, and the distribution of T2 is not arithmetic, thus not concentrated on a regular lattice with probability 1. The distribution of T1 is arbitrary. Let ET2 µ > ˆ0, ª. Then it holds that

t 1 ª S gˆt  xdHˆx ÐÐ S gˆxdx, 0 t ª µ 0 ¢ ¥ > ª S ˆ S @ ª where g R R is Riemann integrable 0, n , for all n N, and Pn 0 maxnBxBn1 g x . 26 2 Counting processes

Abb. 2.2:

Without proof.

u In particular Hˆˆt  u, t¥ ÐÐ holds for an arbitrary u > , thus H asymptotically (for t ª µ R t ª) behaves as the Lebesgue measure.

Definition 2.1.4 The random variable χˆt SNˆt1  t is called excess of N at time t C 0.

Obviously χˆ0 T1 holds. We now give an example of a renewal process with stationary increments. Let N ˜Nˆt, t C 0 be a delayed renewal process associated with the sequence of interarrival ˜  C times Tn n>N. Let FT1 and FT2 be the distribution functions of the delays T1 and Tn, n 2. > ˆ ª ˆ  A We assume that µ ET2 0, , FT2 0 0, thus T2 0 a.s. and

x ˆ  1 ¯ ˆ  C FT1 x S FT2 y dy, x 0. (2.1.5) µ 0

In this case FT1 is called the integrated tail distribution function of T2.

Theorem 2.1.5 Under the conditions we mentioned above, N is a process with stationary increments.

Abb. 2.3:

Proof Let n > N, 0 B t0 @ t1 @ ... @ tn @ ª. Because N does not depend on Tn, n > N the — common distribution of ˆNˆt1  t  Nˆt0  t,...,Nˆtn  t  Nˆtn1  t does not depend on d t, if the distribution of χˆt does not depend on t, thus χˆt χˆ0 T1, t C 0, see Figure .... 2 Counting processes 27

C We show that FT1 Fχˆt, t 0.

ª Fχˆtˆx Pˆχˆt B x Q PˆSn B t, t @ Sn1 B t  x n 0 PˆS0 0 B t, t @ S1 T1 B t  x ª  Q EˆEˆ1ˆSn B t, t @ Sn  Tn1 B t  x S Sn n 1 ª t ˆ    ˆ   Q ˆ  @ B    ˆ  FT1 t x FT1 t S P t y Tn1 t x y dFSn y n 1 0 t ª ˆ    ˆ   ˆ  @ B    ˆQ ˆ  FT1 t x FT1 t S P t y T2 t x y d FSn y . 0 n 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ Hˆy ˆ  y C If we can prove that H y µ , y 0, then we would get

0 ˆ  ˆ    ˆ   1 ˆ ˆ      ˆ  ˆ  Fχˆt x FT1 t x FT1 t S FT2 z x 1 1 FT2 z d z µ t t ˆ    ˆ   1 ˆ ¯ ˆ   ¯ ˆ   FT1 t x FT1 t S FT2 z FT2 z x dz µ 0 tx ˆ    ˆ   ˆ   1 ¯ ˆ  FT1 t x FT1 t FT1 t S FT2 y dy µ x ˆ    ˆ    ˆ  ˆ  C FT1 t x FT1 t x FT1 x FT1 x , x 0, according to the form (2.1.5) of the distribution of T1. ˆ  t C Now we like to show that H t µ , t 0. For that we use the formula (2.1.4): it holds that ª ª ª ˆ ˆ  1 stˆ  ˆ  1 st  1 st ˆ  lT1 s S e 1 FT2 t dt S e dt S e FT2 t dt µ 0 µ 0 µ 0 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ 1 s ª ª 1 ‹  ˆ  st 1 ˆ  st ˆ  Tª  st ˆ  1 S FT2 t de 1 e FT2 t S e dFT2 t µs 0 µs 0 0 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ F ˆ0 0 T2 ˆ ˆ  lT2 s 1 ˆ1  ˆl ˆs, s C 0. µs T2

Using the formula (2.1.4) we get ˆ ˆ  ª lT1 s 1 1 st ˆlH ˆs S e dt ˆl t ˆs, s C 0.  ˆ ˆ  µs µ 0 µ 1 lT2 s Since the Laplace transform of a function uniquely determines this function, it holds that ˆ  t C H t µ , t 0. Remark 2.1.6 In the proof of Theorem 2.1.5 we showed that for the renewal process with delay which possesses ˆ  ¢ t ª the distribution (2.1.5), H t µ not only asymptotical for t (as in the elementary renewal 28 2 Counting processes

ˆ  t C theorem) but it holds H t µ , for all t 0. This means, per unit of the time interval we 1 get an average of µ renewals. For that reason such a process N is called homogeneous renewal process. We can prove the following theorem: Theorem 2.1.6 If N ˜Nˆt, t C 0 is a delayed renewal process with arbitrary delay T1 and non-arithmetic distribution of Tn, n C 2, µ ET2 > ˆ0, ª, then it holds that

x ˆ  1 ¯ ˆ  C lim Fχˆt x S FT2 y dy, x 0. t ª µ 0

This means, the limit distribution of excess χˆt, t ª is taken as the distribution of T1 when defining a homogeneous renewal process.

2.2 Poisson processes

2.2.1 Poisson processes

In this section we generalize the definition of a homogeneous Poisson process (see section 1.2, example 5) Definition 2.2.1 The counting process N ˜Nˆt, t C 0 is called Poisson process with intensity measure Λ if

1. Nˆ0 0 a.s.

2. Λ is a locally finite measure R,i.e., Λ ¢ BˆR R possesses the property ˈB @ ª for every bounded set B > BˆR.

3. N possesses independent increments.

4. Nˆt  Nˆs ¢ PoisˆΛˆˆs, t¥ for all 0 B s @ t @ ª.

Sometimes the Poisson process N ˜Nˆt, t C 0 is defined by the corresponding random Poisson counting measure N ˜NˆB,B > BˆR, i.e., N ˆ 0, t¥, t C 0, where a counting measure is a locally finite measure with values in N0. Definition 2.2.2 A random counting measure N ˜NˆB,B > BˆR is called Poissonsh with locally finite intensity measure Λ if

1. For arbitrary n > N and for arbitrary pairwise disjoint bounded sets B1,B2,...,Bn > BˆR the random variables NˆB1,NˆB2,...,NˆBn are independent.

2. NˆB ¢ PoisˆΛˆB, B > BˆR, B-bounded.

It is obvious that properties 3 and 4 of definition 2.2.1 follow from properties 1 and 2 of definition 2.2.2. Property 1 of definition 2.2.1 however is an autonomous assumption. NˆB, B > BˆR is interpreted as the number of points of N within the set B. 2 Counting processes 29

Remark 2.2.1 As stated in definition 2.2.2, a Poisson counting measure can also be defined on an arbitrary d topological space E equipped with the Borel-σ-algebra BˆE. Very often E R , d C 1 is chosen in applications. Lemma 2.2.1 For every locally finite measure Λ on R there exists a Poisson process with intensity measure Λ.

Proof If such a Poisson process had existed, the characteristic function ϕNˆtNˆsˆ  of the increment Nˆt  Nˆs, 0 B s @ t @ ª would have been equal to ϕs,tˆz ϕPoisˆΛˆˆs,t¥ˆz ˈˆs,t¥ˆeiz1 e , z > R according to property 4 of definition 2.2.1. We show that the family of characteristic functions ˜ϕs,t, 0 B s @ t @ ª possesses property 1.7.1: for all n ¢ 0 B s @ u @ t, ˈˆs,u¥ˆeiz1 ˈˆu,t¥ˆeiz1 ˆΛˆˆs,u¥ˈˆu,t¥ˆeiz1 ˈˆs,t¥ˆeiz1 ϕs,uˆzϕu,tˆz e e e e ϕs,tˆz, z > R since the measure Λ is additive. Thus, the existence of the Poisson process N follows from theorem 1.7.1.

Remark 2.2.2 The existence of a Poisson counting measure can be proven with the help of the theorem of Kolmogorov, yet in a more general form than in theorem 1.1.2. From the properties of the Poisson distribution it follows that ENˆB Var NˆB ˈB, B > BˆR. Thus ˈB is interpreted as the mean number of points of N within the set B, B > BˆR. We get an important special case if ˈdx λdx for λ > ˆ0, ª, i.e., Λ is proportional to the Lebesgue measure ν1 on R. Then we call λ ENˆ1 the intensity of N. Soon we will prove that in this case N is a homogeneous Poisson process with intensity λ. To remind you: In section 1.2 the homogeneous Poisson process was defined as a renewal process with interarrival times TN ¢ Expˆλ: Nˆt sup˜n > N Sn B t, Sn T1  ...  Tn, n > N, t C 0. Exercise 2.2.1 d Show that the homogeneous Poisson process is a homogeneous renewal process with T1 T2 ¢ Expˆλ. Hint: you have to show that for an arbitrary exponential distributed random variable X the integrated tail distribution function of X is equal to FX . Theorem 2.2.1 Let N ˜Nˆt, t C 0 be a counting process. The following statements are equivalent.

1. N is a homogeneous Poisson process with intensity λ A 0.

2. a) Nˆt ¢ Poisˆλt, t C 0

b) for an arbitrary n > N, t C 0, it holds that the random vector ˆS1,...,Sn under condition ˜Nˆt n possesses the same distribution as the order statistics of i.i.d. random variables Ui > Uˆ 0, t¥, i 1, . . . , n.

3. a) N has independent increments, b) ENˆ1 λ, and c) property 2b) holds.

4. a) N has stationary and independent increments, and 30 2 Counting processes

b) PˆNˆt 0 1  λt  oˆt, PˆNˆt 1 λt  oˆt, t £ 0 holds.

5. a) N hast stationary and independent increments, b) property 2a) holds.

Remark 2.2.3 1. It is obvious that Definition 2.2.1 with ˈdx λdx, λ > ˆ0, ª is an equivalent definition of the homogeneous Poisson process according to Theorem 2.2.1.

2. The homogeneous Poisson process N was introduced in the beginning of the 20th century from the physicists A. Einstein and M. Smoluchovsky to be able to model the counting process of elementary particle in the Geiger counter.

3. From 4b) it follows PˆNˆt A 1 oˆt, t £ 0.

N λ ENˆ  1 4. The intensity of has the following interpretation: 1 ETn , thus the mean number of renewals of N within a time interval with length 1.

5. The renewal function of the homogeneous Poisson process is Hˆt λt, t C 0. Thereby Hˆt ˈ 0, t¥, t A 0 holds.

Proof Structure of the proof: 1 ¨ 2 ¨ 3 ¨ 4 ¨ 5 ¨ 1 1 ¨ 2: n ¢ ˆ  ¢ ˆ  > ˆ ˆ   ˆ A From 1) follows Sn Pk 1 Tk Erl n, λ since Tk Pois λ , n N, thus P N t 0 P T1 λt t e , t C 0, and for n > N

PˆNˆt n Pˆ˜Nˆt C n  ˜Nˆt C n  1 PˆNˆt C n  PˆNˆt C n  1 t n n1 t n1 n λ x λx λ x λx PˆSn B t  PˆSn1 B t S e dx  S e dx 0 ˆn  1! 0 n! n n t d ˆλx  ˆλt  S Œ e λx‘ dx e λt, t C 0. 0 dx n! n!

Thus 2a) is proven. Now let’s prove 2b). According to the transformation theorem of random variables (cf. theorem 3.6.1, WR), it follows from ¢ ¨ S1 T1 ¨  ¦ S2 T1 T2 ¨ ¦ ¨ ¤ Sn1 T1  ...  Tn1 ˆ — ˆ — that the density fˆS1,...,Sn of S1,...,Sn1 can be expressed by the density of T1,...,Tn1 , Ti ¢ Expˆλ, i.i.d.:

n1 n1  ˆ     ˆ  M ˆ   M λ tk tk1 n 1 λtn1 fˆS1,...,Sn1 t1, . . . , tn1 fTk tk tk1 λe λ e k 1 k 1 for arbitrary 0 B t1 B ... B tn1, t0 0. ˆ  For all other t1, . . . , tn1 it holds fˆS1,...,Sn1 t1, . . . , tn1 0. 2 Counting processes 31

Therefore ˆ S ˆ   ˆ S B B A  fˆS1,...,Sn t1, . . . , tn N t n fˆS1,...,Sn t1, . . . , tn Sk t, k n, Sn1 t ª R fˆ ˆt , . . . , t  dt  t S1,...,Sn1 1 n 1 n 1 t t t ª R R ... R R fˆ ˆt , . . . , t  dt  dt . . . dt 0 t1 tn1 t S1,...,Sn1 1 n 1 n 1 n 1 ª n1 λt  R λ e n 1 dt  t n 1  t t t ª   R R ... R R λn 1e λtn1 dt  dt . . . dt 0 t1 tn1 t n 1 n 1 Iˆ0 B t1 B t2 B ... B tn B t n! Iˆ0 B t B t B ... B t B t. tn 1 2 n This is exactly the density of n i.i.d. Uˆ 0, t¥-random variables. Exercise 2.2.2 Proof this. 2 ¨ 3 From 2a) obviously follows 3b). Now we just have to prove the independence of the increments of N. For an arbitrary n > N, x1, . . . , xn > N, t0 0 @ t1 @ ... @ tn for x x1  ...  xn it holds that n n Pˆ9 ˜Nˆtk  Nˆtk1 xk Pˆ9 ˜Nˆtk  Nˆtk1 xkSNˆtn x  k 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶k 1  x x! n tk tk1 k L Š  according to 2b) x1!...xn! k 1 tn  PˆNˆtn x ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ˆ x λtn λtn e x! according to 2a) n x ˆλˆt  t   k  ˆ   M k k 1 e λ tk tk1 , k 1 xk!

 n since the probability of (‡) belongs to the polynomial distribution with parameters n, š tk tk1 Ÿ . tn k 1 Because the event (‡)is that at the independent uniformly distributed toss of x points on 0, t¥, exactly xk points occur within the basket of length tk  tk1, k 1, . . . , n:

Abb. 2.4:

ˆ9n ˜ ˆ   ˆ   Ln ˆ˜ ˆ   ˆ   Thus 3a) is proven since P k 1 N tk N tk1 xk k 1 P N tk N tk1 xk . 32 2 Counting processes

3 ¨ 4 We prove that N possesses stationary increments. For an arbitrary n > N0, x1, . . . , xn > N, @ @ @ A ˆ  ˆ9n ˜ ˆ    ˆ    t0 0 t1 ... tn and h 0 we consider I h P k 1 N tk h N tk1 h xk and show that Iˆh does not depend on h > R. According to the formula of the total probability it holds that ª ˆ  ˆ9n ˜ ˆ    ˆ    S ˆ    ˆ ˆ    I h Q P k 1 N tk h N tk1 h xk N tn h m P N tn h m m 0 ª n x m m! t  h  t   h k  ˆ   ˆλˆt  h Q M ‹ k n 1  e λ tn h n   m 0 x1! . . . xn! k 1 tn h h m! ª ˆ9n ˜ ˆ   ˆ  S ˆ     ˆ ˆ    ˆ  Q P k 1 N tk N tk1 xk N tn h m P N tn h m I 0 . m 0

We now show property 4b) for h > ˆ0, 1:

ª ª PˆNˆh 0 Q PˆNˆh 0,Nˆ1 k Q PˆNˆh 0,Nˆ1  Nˆh k k 0 k 0 ª Q PˆNˆ1  Nˆh k, Nˆ1 k k 0 ª Q PˆNˆ1 kPˆNˆ1  Nˆh k S Nˆ1 k k 0 ª Q PˆNˆ1 kˆ1  hk. k 0 ˆ ˆ     ˆ  1 ˆ  ˆ ˆ   We have to show that P N h 0 1 λh o h , i.e., limh ª h 1 P N h 0 λ. Indeed it holds that ª ª 1 1 1  ˆ1  hk ˆ1  PˆNˆh 0 Œ1  Q PˆNˆ1 kˆ1  hk‘ Q PˆNˆ1 k h h k 0 k 1 h ª 1  ˆ1  hk ÐÐ Q PˆNˆ1 k lim h 0 h 0 h k 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ k ª Q PˆNˆ1 kk ENˆ1 λ, k 0 ª ˆ ˆ   @ ª since the series uniformly converges in h because it is dominated by Pk 0 P N 1 k k λ k because of the inequality ˆ1  h C 1  kh, h > ˆ0, 1, k > N. PˆNˆh 1 Pª ˆ ˆ   ˆ  k1  ¨  Similarly one can show that limh 0 h limh 0 k 1 P N 1 k k 1 h λ. 4 5 We have to show that for an arbitrary n > N and t C 0 ˆ n λt λt pnˆt PˆNˆt n e (2.2.1) n!

λt holds. We will prove that by induction with respect to n. First we show that p0ˆt e , h 0. For that we consider p0ˆth PˆNˆth 0 PˆNˆt 0,NˆthNˆt 0 p0ˆtp0ˆh p0ˆtˆ1  λh  oˆh, h 0. Similarly one can show that p0ˆt p0ˆt  hˆ1  λh  oˆh, h 0. 2 Counting processes 33

ˆ   ˆ  œ ˆ  p0 t h p0 t  ˆ  A ˆ  ˆ ˆ   Thus p0 t limh 0 h λp0 t , t 0 holds. Since p0 0 P N 0 0 1, it follows from œ ˆ   ˆ  œ p0 t λp0 t p0ˆ0 1, λt that it exists an unique solution p0ˆt e , t C 0. Now for n the formular (2.2.1) be approved. Let’s prove it for n  1.

pn1ˆt  h PˆNˆt  h n  1 PˆNˆt n, Nˆt  h  Nˆt 1  PˆNˆt n  1,Nˆt  h  Nˆt 0

pnˆt p1ˆh  pn1ˆt p0ˆh

pnˆtˆλh  oˆh  pn1ˆtˆ1  λh  oˆh, h 0, h A 0.

Thus œ p  ˆt λp  ˆt  λp ˆt, t A 0 œ n 1 n 1 n (2.2.2) pn1ˆ0 0

 ˆ  λt ˆλtn ˆ  λt ˆλtn 1 ˆ  Since pn t e n! , we obtain pn1 t e ˆn1! as solution of (2.2.2). (Indeed pn1 t λt œ λt λt Cˆte ¨ C ˆte λCˆte ......  λpnˆt œˆ  λn1tn ¨ ˆ  λn1tn1 ˆ  C t n! C t ˆn1! , C 0 0) 5 ¨ 1 Let N be a counting process Nˆt max˜n ¢ Sn B t, t C 0, which fulfills conditions 5a) and n ¢ ˆ  >  5b). We show that Sn Pk 1 Tk, where Tk i.i.d. with Tk Exp λ , k N. Since Tk Sk Sk1, k > N, S0 0, we consider for b0 0 B a1 @ b1 B ... B an @ bn ˆ9n ˜ @ B  P k 1 ak Sk bk ˆ9n1˜ ˆ   ˆ  ˆ   ˆ   P k 1 N ak N bk1 0,N bk N ak 1 9˜Nˆan  Nˆbn1 0,Nˆbn  Nˆan C 1 n1 MˆPˆNˆak  bk1 0 PˆNˆbk  ak 1  k 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ λˆa b   λˆb a  e k k 1 λˆbkake k k PˆNˆan  bn1 0 PˆNˆbn  an C 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ eλˆanbn1 ˆ1eλˆbnan n1 λˆanbn1 λˆbnan λˆbkbk1 e ˆ1  e  M λˆbk  ake k 1  n 1 b1 bn n1 λan λbn n λyn λ ˆe  e  Mˆbk  ak S ... S λ e dyn . . . y1. k 1 a1 an

— n λyn The common density of ˆS1,...,Sn therefore is given by λ e 1ˆy1 B y2 B ... B yn.

2.2.2 Compound Poisson process Definition 2.2.3 Let N ˜Nˆt, t C 0} be a homogeneous Poisson process with intensity λ A 0, build by means ˜  ˜  of the sequence Tn n>N of interarrival times. Let Un n>N be a sequence of i.i.d. random ˜  variables, independent of Tn n>N. Let FU be the distribution function of U1. For an arbitrary 34 2 Counting processes

C ˆ  PNˆt ˜ ˆ  C  t 0 let X t k 1 Uk. The stochastic process X X t , t 0 is called compound Poisson process with parameters λ, FU . The distribution of Xˆt thereby is called compound Poisson distribution with parameters λt, FU .

The compound Poisson process Xˆt, t C 0 can be interpreted as the sum of „marks“ Un of a homogeneous marked Poisson process ˆN,U until time t. In queueing theory Xˆt is interpreted as the overall workload of a server until time t if the n > requests to the service occur at times Sn Pk 1 Tk, n N and represent the amount of work Un, n > N. In actuarial mathematics Xˆt, t C 0 is the total damage in a portfolio until time t C 0 with number of damages Nˆt and amount of loss Un, n > N. Theorem 2.2.2 Let X ˜Xˆt, t C 0 be a compound Poisson process with parameters λ, FU . The following properties hold:

1. X has independent and stationary increments.

sU1 2. If mˆ U ˆs Ee , s > R, is the moment generating function of U1, such that mˆ U ˆs @ ª, s > R, then it holds that

ˆ ˆ   ˆ  λt mˆ U s 1 > C ˆ  ˆ  2 C mˆ Xˆt s e , s R, t 0, EX t λtEU1, Var X t λtEU1 , t 0.

Proof 1. We have to show that for arbitrary n > N, 0 B t0 @ t1 @ ... @ tn and h

ˆ   ˆ   ˆ  ’ N t1 h N tn h “ n ’ N tk “ Q B Q B M Q B P Ui1 x1,..., Uin xn P Uik xk ” • ” • i1 Nˆt0h1 in Nˆtn1h1 k 1 ik Nˆtk11

for arbitrary x1, . . . , xn > R. Indeed it holds that

ˆ   ˆ   ’ N t1 h N tn h “ P Q U B x ,..., Q U B x ” i1 1 in n• i1 Nˆt0h1 in Nˆtn1h1 ª ’ n “ ‡kj n Q M F ˆx  P ˆ9 ˜Nˆt  h  Nˆt   h k  ” n j • m 1 m m 1 m k1,...,kn 0 j 1 ª ’ n “ n ‡kj Q M F ˆx  Œ M PˆNˆt   Nˆt   k ‘ ” n j • m m 1 m k1,...,kn 0 j 1 m 1 n ª ‡ km ˆ  ˆ ˆ   ˆ   M Q Fn xm P N tm N tm1 km m 1 km 0 ˆ  n ’ N tm “ M P Q B xm ” • m 1 km Nˆtm11

2. Exercise 2.2.3 2 Counting processes 35

2.2.3 Cox process A Cox process is a (in general inhomogeneous) Poisson process with intensity measure Λ which as such is a random measure. The intuitive idea is stated in the following definition. Definition 2.2.4 Let Λ ˜ΛˆB,B > BˆR be a random a.s. locally finite measure. The random counting measure N ˜NˆB,B > BˆR is called Cox counting measure (or doubly stochastic Poisson measure) with random intensity measure Λ if for arbitrary n > N, k1, . . . , kn > N0 and 0 B a1 @ b1 B k n n ˈˆa ,b ¥ Λ i ˆˆai,bi¥ a2 @ b2 B ... B an @ bn it holds that Pˆ9 ˜Nˆˆai, bi¥ ki E ŠL e i i . i 1 i 1 ki! The process ˜Nˆt, t C 0 with Nˆt Nˆˆ0, t¥ is called Cox process (or doubly stochastic Poisson process) with random intensity measure Λ. Example 2.2.1 1. If the random measure Λ is a.s. absolutely continuous with respect to ˆ  ˆ  > ˆ  ˜ ˆ  C  the Lebesgue measure, i.e., Λ B RB λ t dt, B - bounded, B B R , where λ t , t 0 is a stochastic process with a.s. Borel-measurable Borel-integrable trajectories, then λˆt C 0 a.s. for all t C 0 is called the intensity process of N.

2. In particular, it can be that λˆt ¡ Y where Y is a non-negative random variable. Then it holds that ˈB Y ν1ˆB, thus N has a random intensity Y . Such Cox processes are called mixed Poisson processes. A Cox process N ˜Nˆt, t C 0 with intensity process ˜λˆt, t C 0 can be build explicitly as the following. Let N˜ ˜N˜ˆt, t C 0 be a homogeneous Poisson process with intensity 1, d which is independent of ˜λˆt, t C 0. Then N N1, where the process N1 ˜N1ˆt, t C 0 ˆ  ˜ˆ t ˆ   C d is given by N1 t N R0 λ y dy , t 0. The assertion N N1 of course has to be proven. However, we shall assume it without proof. It is also the basis for the simulation of the Cox process N.

2.3 Additional exercises

Exercise 2.3.1 Let ˜NˆttC0 be a renewal process with interarrival times Ti, which are exponentially dis- tributed, i.e. Ti ¢ Expˆλ.

a) Prove that: Nˆt is Poisson distributed for every t A 0.

b) Determine the parameter of this Poisson distribution.

c) Determine the renewal function Hˆt E Nˆt.

Exercise 2.3.2 Prove that a (real-valued) stochastic process X ˜Xˆt, t > 0, ª with independent incre- ments already has stationary increments if the distribution of the random variable Xˆt  h  Xˆh does not depend on h. Exercise 2.3.3 Let N ˜Nˆt, t > 0, ª be a Poisson process with intensity λ. Calculate the that within the interval 0, s¥ exactly i events occur under the condition that within the interval 0, t¥ exactly n events occur, i.e. PˆNˆs i S Nˆt n for s @ t, i 0, 1, . . . , n. 36 2 Counting processes

Exercise 2.3.4 ˆ  ˆ  ˆ  ˆ  Let N 1 ˜N 1 ˆt, t > 0, ª and N 2 ˜N 2 ˆt, t > 0, ª be independent Poisson processes with intensities λ1 and λ2. In this case the independence indicates that the sequences ˆ1 ˆ1 ˆ2 ˆ2 ˜ ˆ  ¢ ˆ1ˆ  ˆ2ˆ  > T1 ,T2 ,... and T1 ,T2 ,... are independent. Show that N N t N t N t , t 0, ª is a Poisson process with intensity λ1  λ2. Exercise 2.3.5 (Queuing paradox): Let N ˜Nˆt, t > 0, ª be a renewal process. Then T ˆt SNˆt1  t is called the time of excess, Cˆt t  SNˆt the current lifetime and Dˆt T ˆt  Cˆt the lifetime at time t A 0. Now let N ˜Nˆt, t > 0, ª be a Poisson process with intensity λ.

a) Calculate the distribution of the time of excess T ˆt.

 b) Show that the distribution of the current lifetime is given by PˆCˆt t e λt and the λs density is given by fCˆtSNˆtA0ˆs λe 1˜s B t.

 c) Show that PˆDˆt B x ˆ1  ˆ1  λ min˜t, xe λx1˜x C 0.

d) To determine ET ˆt, one could argue like this: On average t lies in the middle of the ˆ  ˆ  1 ˆ  surrounding interval of interarriving time SNˆt,SNˆt1 , i.e. ET t 2 E SNˆt1  1 1 SNˆt 2 ETNˆt1 2λ . Considering the result from part (a) this reasoning is false. Where is the mistake in the reasoning?

Exercise 2.3.6 ˜ ˆ  ¢ PNˆt C  ˆ  Nˆt Let X X t i 1 Ui, t 0 be a compound Poisson process. Let MNˆt s Es , s > ˆ0, 1, be the generating function of the Poisson processes Nˆt, L˜Uˆs E exp˜sU the Laplace Transform of Ui, i > N, and L˜Xˆtˆs the Laplace Transform of Xˆt. Prove that

L˜Xˆtˆs MNˆtˆL˜Uˆs, s C 0.

Exercise 2.3.7 Let X ˜Xˆt, t > 0, ª be a compound Poisson process with Ui i.i.d., U1 ¢ Expˆγ, where the intensity of Nˆt is given by λ. Show that for the Laplace transform L˜Xˆtˆs of Xˆt it holds: λts L˜Xˆtˆs exp › . γ  s Exercise 2.3.8 Write a function in R (alternatively: Java) to which we pass time t, intensity λ and a value γ as parameters. The return of the function is a random value of the compound Poisson process with characteristics ˆλ, Expˆγ at time t. Exercise 2.3.9 Let the stochastic process N ˜Nˆt, t > 0, ª be a Cox process with intensity function λˆt Z, where Z is a discrete random variable which takes values λ1 and λ2 with probabilities 1~2. Determine the moment generating function as well as the expected value and the variance of Nˆt. Exercise 2.3.10 ˆ  ˆ  ˆ  ˆ  Let N 1 ˜N 1 ˆt, t > 0, ª and N 2 ˜N 2 ˆt, t C 0 be two independent homogeneous Poisson processes with intensities λ1 and λ2. Moreover, let X C 0 be an arbitrary non-negative 2 Counting processes 37

ˆ  ˆ  random variable which is independent of N 1 and N 2 . Show that the process N ˜Nˆt, t C 0 with ¢ ¨N ˆ1ˆt, t B X, Nˆt ¦ ¨ ˆ1 ˆ2 ¤¨N ˆX  N ˆt  X, t A X is a Cox process whose intensity process λ ˜λˆt, t C 0 is given by ¢ ¨λ , t B X, λˆt ¦ 1 ¨ A ¤¨λ2, t X. 3 Wiener process

3.1 Elementary properties

In Example 2) of Section 1.2 we defined the Brownian motion (or Wiener process) W ˜W ˆt, t C 0 as an Gaussian process with EW ˆt 0 and covˆW ˆs,W ˆt min˜s, t, s, t C 0. The Wiener process is called after the mathematician Norbert Wiener (1894 - 1964). Why does the Brownian motion exist? According to theorem of Kolmogorov (Theorem 1.1.2) it exists a real-valued Gaussian process X ˜Xˆt, t C 0 with mean value EXˆt µˆt, t C 0, and covariance function covˆXˆs,Xˆt Cˆs, t, s, t C 0 for every function µ ¢ R R and every positive semidefinite function C ¢ R  R R. We just have to show that Cˆs, t min˜s, t, s, t C 0 is positive semidefinite. Exercise 3.1.1 Prove this! We now give a new (equivalent) definition. Definition 3.1.1 A stochastic process W ˜W ˆt, t C 0 is called Wiener process (or Brownian motion) if

1. W ˆ0 0 a.s.

2. W possesses independent increments

3. W ˆt  W ˆs ¢ N ˆ0, t  s, 0 B s @ t

The existence of W according to Definition 3.1.1 follows from Theorem 1.7.1 since ϕs,tˆz ˆ   2 ˆ   2 ˆ   2 ˆ   2 izˆW ˆtW ˆs  t s z  t u z  u s z  t s z Ee e 2 , z > R, and e 2 e 2 e 2 for 0 B s @ u @ t, thus ϕs,uˆzϕu,tˆz ϕs,tˆz, z > R. From Theorem 1.3.1 the existence of a version with continuous trajectories follows. Exercise 3.1.2 1 Show that Theorem 1.3.1 holds for α 3, σ 2 . Therefore, it is often assumed that the Wiener process possesses continuous paths (just take its corresponding version). Theorem 3.1.1 Both definitions of the Wiener process are equivalent.

Proof 1. From definition in Section 1.2 follows Definition 3.1.1. W ˆ0 0 a.s. follows from VarˆW ˆ0 min˜0, 0 0. Now we prove that the increments of W are independent. If Y ¢ N ˆµ, K is a n-dimensional Gaussian random vector — and A a ˆn  n-matrix, then AY ¢ N ˆAµ, AKA  holds, this follows from the explicit form of the characteristic function of Y . Now let n > N, 0 t0 B t1 @ ... @ tn, Y

38 3 Wiener process 39

— — ˆW ˆt0,W ˆt1,...,W ˆtn . For Z ˆW ˆt0,W ˆt1  W ˆt0,...,W ˆtn  W ˆtn1 it holds that Z AY , where ’ 1 0 0 ...... 0 “ – — – 1 1 0 ...... 0 — – — A –  ... — . – 0 1 1 0 0 — – ...... — ” 0 0 0 ... 1 1 • Thus Z is also Gaussian with a covariance matrix which is diagonal. Indeed, it holds covˆW ˆti1  W ˆti,W ˆtj1  W ˆtj min˜ti1, tj1  min˜ti1, tj  min˜ti, tj1  min˜ti, tj 0 for i x j. Thus the coordinates of Z are uncorrelated, which means inde- pendence in case of a multivariate Gaussian distribution. Thus the increments of W are independent. Moreover, for arbitrary 0 B s @ t it holds that W ˆt  W ˆs ¢ N ˆ0, t  s. The normal distribution follows since Z AY is Gaussian, obviously it holds that EW ˆtEW ˆs 0 and VarˆW ˆtW ˆs VarˆW ˆt2 covˆW ˆs,W ˆtVarˆW ˆs t  2 min˜s, t  s t  s. 2. From Definition 3.1.1 the definition in Section 1.2 follows. Since W ˆt  W ˆs ¢ N ˆ0, t  s for 0 B s @ t, it holds covˆW ˆs,W ˆt E W ˆsˆW ˆtW ˆsW ˆs¥ EW ˆsEˆW ˆtW ˆsVar W ˆs s,

thus it holds covˆW ˆs,W ˆt min˜s, t. From W ˆtW ˆs ¢ N ˆ0, ts and W ˆ0 0 it also follows that EW ˆt 0, t C 0. The fact that W is a Gaussian process, follows from  point 1) of the proof, relation Y A 1Z.

Definition 3.1.2 — The process ˜W ˆt, t C 0, W ˆt ˆW1ˆt,...,Wdˆt , t C 0, is called d-dimensional Brownian motion if Wi ˜Wiˆt, t C 0 are independent Wiener processes, i 1, . . . , d. The definitions above and Exercise 3.1.2 ensure the existence of a Wiener process with continuous paths. How do we find an explicit way of building these paths? We will show that in the next section.

3.2 Explicit construction of the Wiener process

First we construct the Wiener process on the interval 0, 1¥. The main idea of the construction is to introduce a stochastic process X ˜Xˆt, t > 0, 1¥ which is defined on a probability ˆ  d ˆ  Pª ˆ  > ¥ ˜  subspace of Ω, A, P with X W , where X t n 1 cn t Yn, t 0, 1 , Yn n>N is a sequence ˆ  ˆ  t ˆ  > ¥ > ˜  of i.i.d. N 0, 1 -random variables and cn t R0 Hn s ds, t 0, 1 , n N. Here, Hn n>N is the orthonormed Haar basis in L2ˆ 0, 1¥ which is introduced shortly now.

3.2.1 Haar- and Schauder-functions Definition 3.2.1 The functions Hn ¢ 0, 1¥ R, n > N, are called Haar functions if H1ˆt 1, t > 0, 1¥, n n n1 ˆ  1 ˆ   1 ˆ  ˆ  2 ˆ ˆ   ˆ  > ¥ @ B H2 t 1 ¥ t 1ˆ ¥ t , Hk t 2 1In,k t 1Jn,k t , t 0, 1 , 2 k 2 , where 0, 2 2 ,1 n1 n1 n n n In,k an,k, an,k  2 ¥, Jn,k ˆan,k  2 , an,k  2 ¥, an,k 2 ˆk  2  1, n > N. 40 3 Wiener process

Abb. 3.1: Haar functions

Lemma 3.2.1 ˜  2ˆ ¥ The function system Hn n>N is an orthonormal basis in L 0, 1 with scalar product @ A 1 ˆ  ˆ  > 2ˆ ¥ f, g R0 f t g t dt, f, g L 0, 1 .

Proof The orthonormality of the system `Hk,Hne δkn, k, n > N directly follows from def- ˜  inition 3.2.1. Now we prove the completeness of Hn n>N. It is sufficient to show that for 2 arbitrary function g > L ˆ 0, 1¥ with `g, Hne 0, n > N, it holds g 0 almost everywhere on ¥ n1 0, 1 . In fact, we always can write the indicator function of an interval 1 an,k,an,k2 ¥ as a linear combination of Hn, n > N. ˆ   H1 H2 1 0, 1 ¥ , 2 2 ˆ   H1 H2 1ˆ 1 ,1¥ , 2 2 ˆ  º1  1 0, 1 ¥ H2 2 2 1 0, 1 ¥ , 4 2 ˆ  º1  1 0, 1 ¥ H2 2 2 1ˆ 1 , 1 ¥ , 4 2 2 ¦  n ˆ  n  2  1an,k,an,k 2 2 Hk n n1 1 a ,a 2n1¥ , 2 @ k B 2 . n,k n,k 2

ˆk1 2n ˆ  > n  ˆ  t ˆ  Therefore it holds R k g t dt 0, n N0, k 1,..., 2 1, and thus G t R0 g s ds 0 for 2n k > n  ¥ ˆ  > ¥ t 2n , n N0, k 1,..., 2 1. Since G is continuous on 0, 1 , it follows G t 0, t 0, 1 , and thus gˆs 0 for almost every s > 0, 1¥.

From lemma 3.2.1 it follows that two arbitrary functions f, g > L2ˆ 0, 1¥ have expan- ª ` e ª ` e 2ˆ ¥ sions f Pn 1 f, Hn Hn and g Pn 1 g, Hn Hn (these series converge in L 0, 1 ) and ` e ª ` e` e f, g Pn 1 f, Hn g, Hn (Parseval identity). Definition 3.2.2 ˆ  t ˆ  ` e > ¥ > The functions Sn t R0 Hn s ds 1 0,t¥,Hn , t 0, 1 , n N are called Schauder functions. 3 Wiener process 41

Abb. 3.2: Schauder functions

Lemma 3.2.2 It holds:

1. Snˆt C 0, t > 0, 1¥, n > N  ˜1, 2n 1  n P n ˆ  B 2 > ¥ > 2. k 1 S2 k t 2 2 , t 0, 1 , n N, ˜  ˆ ε @ 1 ª 3. Let an n>N be a sequence of real numbers with an O n , ε 2 , n . Then the series ª ˆ  > ¥ Pn 1 anSn t converges absolutly and uniformly in t 0, 1 and therefore is a continuous function on 0, 1¥. Proof 1. follows directly from definition 3.2.2.

n 2. follows since functions S2nk for k 1,..., 2 have disjoint supports and S2nkˆt B   n  ˆ 2k 1  2 1 > ¥ S2nk 2n1 2 , t 0, 1 .

3. It suffices to show that Rn sup > ¥ PkA2n SakSSkˆt ÐÐÐ 0. For every k > and c A 0 t 0,1 n ª N ε it holds SakS B ck . Therefore it holds for all t > 0, 1¥, n > N

ˆn1ε ˆn1ε  n 1 εnˆ 1 ε Q SakSSkˆt B c 2 Q Skˆt B c 2 2 2 B c 2 2 . 2n@kB2n1 2n@kB2n1

1 ε nˆ 1 ε Since ε @ , it holds Rm B c 2 PnCm 2 2 ÐÐÐ 0. 2 m ª

Lemma 3.2.3 ˜  ˆ  Let Yn n>N be a sequence of (not necessarily independent) random variables defined on Ω, A, P , 1 Yn ¢ N ˆ0, 1, n > N. Then it holds SYnS Oˆˆlog n 2 , n ª, a.s. º Proof We have to show that for c A 2 and almost all ω > Ω it exists a n0 n0ˆω, c > N such 1 that SYnS B cˆlog n 2 for n C n0. If Y ¢ N ˆ0, 1, x A 0, it holds

ª 2 ª 2 1  y 1 1  y PˆY A x º S e 2 dy º S ‹  d ‹e 2  2π x 2π x y 2 ª 2 1 1  y  y 1 1 1  x2 º ‹ e 2  S e 2 dy B º e 2 . 2π x x y2 2π x 42 3 Wiener process

2 º 1 1  x ¯ˆx ¢ º e 2 x ª c A (We also can show that Φ 2π x , .) Thus for 2 it holds º 2 1 2 1 1 2  1  c log n c 2  1  c Q PˆSYnS A cˆlog n 2  B c º Qˆlog n 2 e 2 º Qˆlog n 2 n 2 @ ª. nC2 2π nC2 π nC2

According to the Lemma of Borel-Cantelli (cf. WR, Lemma 2.2.1) it holds Pˆ9n 8kCn Ak 0 1 ˆ  @ ª ˜S S A ˆ  2  > if Pk P Ak with Ak Yk e log k , k N. Thus Ak occurs in infinite number only 1 with probability 0, with SYnS B cˆlog n 2 for n C n0.

3.2.2 Wiener process with a.s. continuous paths

Lemma 3.2.4 ˜  ˆ  ˜  Let Yn n>N be a sequence of independent N 0, 1 -distributed random variables. Let an n>N 2m  m 2m  m ˜  P S m S B 2 P S m S B 2 > and bn n>N be sequences of numbers with k 1 a2 k 2 , k 1 b2 k 2 , m N. Then ª ª ¢ ˆ ª 2  ¢ ˆ ª 2  the limits U Pn 1 anYn and V Pn 1 bnYn, U N 0, Pn 1 an , V N 0, Pn 1 bn exist a.s., ˆ  ª ˆ  where cov U, V Pn 1 anbn. U and V are independent if and only if cov U, V 0.

Proof Lemma 3.2.2 and 3.2.3 reveal the a.s. existence of the limits U and V (replace an by Yn and Sn by e.g. bn in Lemma 3.2.2). From the stability under convolution of the normal distri- ˆm m ˆm m ˆm ¢ ˆ m 2  ˆm ¢ bution it follows for U Pn 1 anYn, V Pn 1 bnYn, that U N 0, Pn 1 an , V ˆ m 2  ˆm Ð d ˆm Ð d ¢ ˆ ª 2  ¢ ˆ ª 2  N 0, Pn 1 bn . Since U U, V V it follows U N 0, Pn 1 an , V N 0, Pn 1 bn . Moreover, it holds

ˆ  ˆ  covˆU, V  lim covˆU m ,V m  m ª m lim Q aibj covˆYi,Yj m ª i,j 1 m ª lim Q aibi Q aibi, m ª i 1 i 1 according to the dominated convergence theorem of Lebesgue, since according to Lemma 3.2.3 1 it holds SYnS B c ˆlog n 2 , for n C N0, and the dominated series converges according to Lemma ´¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¶ B ε @ 1 cn , ε 2 3.2.2:

2m1 2m1 a.s. 2 ε ε 2εˆm1  m  m ˆ12εm Q anbkYnYk B Q anbkc n k B 2 2 2 2 2 2 , 1  2ε A 0. n,k 2m n,k 2m

ª B ª ˆ12εj @ ª For sufficient large m it holds Pn,k m anbkYnYk Pj m 2 , and this series converges a.s. Now we show covˆU, V  0 ¨ U and V are independent

Independence always results in the uncorrelation of random variables. We prove the other 3 Wiener process 43

ˆm ˆm d direction. From ˆU ,V  ÐÐÐ ˆU, V  it follows ϕˆ ˆm ˆm ÐÐÐ ϕˆU,V , thus m ª U ,V m ª

m m ϕˆ ˆm ˆmˆs, t lim E exp˜iˆt Q akYk  s Q bnYn U ,V m ª k 1 n 1 m m lim E exp˜i Qˆtak  sbkYk lim M E exp˜iˆtak  sbkYk m ª m ª k 1 k 1 m 2 ª 2 ˆtak  sbk ˆtak  sbk lim M exp˜  exp˜ Q  m ª k 1 2 k 1 2 ¢ £ ¨ ¨ ¨ ¨ t2 ª ¨ ª ¨ s2 ª œ Q 2¡ ¦ Q § œ Q 2¡ ˆ  ˆ  exp ak exp ¨ts akbk ¨ exp bk ϕU t ϕV s , 2 k 1 ¨ k 1 ¨ 2 k 1 ¨ ´¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¶ ¨ ¨ ¨ ¤ covˆU,V  0¥

s, t > R. Thus, U and V are independent if covˆU, V  0.

Theorem 3.2.1 Let ˜Yn, n > N be a sequence of i.i.d. random variables that are N ˆ0, 1-distributed, defined on a probability space ˆΩ, A, P. Then there exists a probability space ˆΩ0, A0, P of ˆΩ, A, P ˜ ˆ  > ¥ ˆ  ª ˆ  ˆ  and a stochastic process X X t , t 0, 1 on it such that X t, ω Pn 1 Yn ω Sn t , > ¥ > d ˜  t 0, 1 , ω Ω0 and X W . Here, Sn n>N is the family of Schauder functions.

Proof According to Lemma 3.2.2, 2) the coefficients Snˆt fulfill the conditions of Lemma 3.2.4 > ¥ ` for every t 0, 1 . In addition to that it exists according to Lemma 3.2.3 aº subset Ω0 Ω, Ω0 > A with PˆΩ0 1, such that for every ω > Ω0 the relation SYnˆωS Oˆ log n, n ª, 9 ˆ  holds. Let A0 A Ω0. We restrict the probabilityº space to Ω0, A0, P . Then condition ˆ  ˆ ε @ 1 @ ε an Yn ω O n , ε 2 , is fulfilled since log n n for sufficient large n, and according ª ˆ  ˆ  > ¥ to Lemma 3.2.2, 3) the series Pn 1 Yn ω Sn t converges absolutely and uniformly in t 0, 1 to the function Xˆω, t, ω > Ω0, which is a continuous function in t for every ω > Ω0. Xˆ , t is a random variable since in Lemma 3.2.4 the convergence of this series holds almost surely. ˆ  ¢ ˆ ª 2ˆ  > ¥ Moreover it holds X t N 0, Pn 1 Sn t , t 0, 1 . We show that this stochastic process, defined on ˆΩ0, A0, P, is a Wiener process. For that we check the conditions of Definition 3.1.1. We consider arbitrary times 0 B t1 @ t2, t3 @ t4 B 1 and 44 3 Wiener process evaluate ª ª covˆXˆt2  Xˆt1,Xˆt4  Xˆt3 covˆQ YnˆSnˆt2  Snˆt1, Q YnˆSnˆt4  Snˆt3 n 1 n 1 ª QˆSnˆt2  Snˆt1ˆSnˆt4  Snˆt3 n 1 ª Qˆ` e  ` e  Hn, 1 0,t2¥ Hn, 1 0,t1¥ n 1 ˆ` e  ` e Hn, 1 0,t4¥ Hn, 1 0,t3¥ ª Q`  e`  e Hn, 1 0,t2¥ 1 0,t1¥ Hn, 1 0,t4¥ 1 0,t3¥ n 1 `   e 1 0,t2¥ 1 0,t1¥, 1 0,t4¥ 1 0,t3¥ ` e  ` e 1 0,t2¥, 1 0,t4¥ 1 0,t1¥, 1 0,t4¥ ` e  ` e 1 0,t2¥, 1 0,t3¥ 1 0,t1¥, 1 0,t3¥ min˜t2, t4  min˜t1, t4  min˜t2, t3  min˜t1, t3, @ A min˜s,t ˜  > ¥ B by Parseval inequality and since 1 0,s¥, 1 0,t¥ R0 du min s, t , s, t 0, 1 . If 0 t1 @ t2 B t3 @ t4 @ 1, it holds covˆXˆt2  Xˆt1,Xˆt4  Xˆt3 t2  t1  t2  t1 0, thus the increments of X (according to Lemma 3.2.4) are uncorrelated. Moreover it holds Xˆ0 ¢ ˆ ª 2ˆ  ˆ  ˆ  a.s. N 0, Pn 1 Sn 0 N 0, 0 , therefore X 0 0. For t1 0, t2 t, t3 0, t4 t it follows that VarˆXˆt t, t > 0, 1¥, and for t1 t3 s, t2 t4 t, that VarˆXˆtXˆs tsss ts, 0 B s @ t B 1. Thus it holds Xˆt  Xˆs ¢ N ˆ0, t  s, and according to Definition 3.1.1 it holds d X W .

Remark 3.2.1 1. Theorem 3.2.1 is the basis for an approximative simulation of the paths ˆnˆ  n ˆ  > ¥ of a Brownian motion through the partial sums X t Pk 1 YkSk t , t 0, 1 , for sufficient large n > N. 2. The construction in Theorem 3.2.1 can be used to construct the Wiener process with ¥ A ˜ ˆ  > ¥ continuous paths on the interval 0, t0 for arbitrary t0 0. If W º W t , t 0, 1 is a t Wiener process on 0, 1¥ then Y ˜Y ˆt, t > 0, t0¥ with Y ˆt t0W ˆ , t > 0, t0¥, is t0 a Wiener process on 0, t0¥. Exercise 3.2.1 Prove that.

3. The Wiener process W with continuous paths on R can be constructed as follows. Let ˆ  ˆ  W n ˜W n ˆt, t > 0, 1¥ be independent copies of the Wiener process as in Theorem ˆ  ª ˆ >  ¥ n1 ˆkˆ   ˆnˆ  ˆ  ¥ C 3.2.1. Define W t Pn 1 1 t n 1, n Pk 1 W 1 W t n 1 , t 0, thus, ¢ ˆ1ˆ  > ¥ ¨ W t , t 0, 1 , ¨ ˆ  ˆ  ¨ W 1 ˆ1  W 2 ˆt  1, t > 1, 2¥, W ˆt ¦ ˆ  ˆ  ˆ  ¨ W 1 ˆ1  W 2 ˆ1  W 3 ˆt  2, t > 2, 3¥, ¨ ¤¨ etc.

Exercise 3.2.2 Show that the introduced stochastic process W ˜W ˆt, t C 0 is a Wiener process on R. 3 Wiener process 45

Abb. 3.3:

3.3 Distribution and path properties of Wiener processes

3.3.1 Distribution of the maximum

Theorem 3.3.1 Let W ˜W ˆt, t > 0, 1¥ be the Wiener process defined on a probability space ˆΩ, F, P. Then it holds: ¾ ª 2 2  y P Œ max W ˆt A x‘ S e 2 dy (3.3.1) t> 0,1¥ π x for all x C 0.

The mapping maxt> 0,1¥ W ˆt ¢ Ω 0, ª given in relation (3.3.1) is a well-defined random ˆ  ˆ i  > variable since it holds: maxt> 0,1¥ W t, ω limn ª maxi 1,...,k W k , ω for all ω Ω since the trajectories of ˜W ˆt, t > 0, 1¥ are continuous. From 3.3.1 it follows that maxt> 0,1¥ W ˆt has an exponential bounded tail: thus maxt> 0,1¥ W ˆt has finite k-th moments. Useful ideas for the proof of Theorem 3.3.1 Let ˜W ˆt, t > 0, 1¥ be a Wiener process and Z1,Z2,... a sequence of independent random ˆ  ˆ   1 C > ˜ ˜ nˆ  > variables with P Zi 1 P Zi 1 2 for all i 1. For every n N we define W t , t S Z , ¥ W˜ nˆt º nt  ˆnt  nt ºnt1 S Z  ...  Z i C S 0 1 by n n , where i 1 i, 1, 0 0. Lemma 3.3.1 For every k C 1 and arbitrary t1, . . . , tk > 0, 1¥ it holds:

— ˆn ˆn d — ŠW˜ ˆt1,..., W˜ ˆtk ˆW ˆt1,...,W ˆtk .

Proof Consider the special case k 2 (for k A 2 the proof is analogous). Let t1 @ t2. For all 46 3 Wiener process s1, s2 > R it holds:

S nt  ˆS nt   S nt 1 s W˜ ˆnˆt   s W˜ ˆnˆt  ˆs  s  º 1  s 2 º 1 1 1 2 2 1 2 n 2 n

s1 s2 Z  Œˆnt  nt º  º ‘ nt1 1 1 1 n n s2 Z  ˆnt  nt º , nt2 1 2 2 n    since S nt2 S nt1 S nt2 S nt11 S nt11. Now observe that the 4 summands on the right-hand-side of the previous equation are inde- pendent and moreover that the latter two summands converge (a.s. and therefor particularly in distribution) to zero. Consequently, it holds

 ˆ  ˆ  s1º s2 ºs2 iˆs W˜ n ˆt s W˜ n ˆt  i S  i ˆS S   lim Ee 1 1 2 2 lim Ee n nt1 Ee n nt2 nt1 1 n ª n ª ¼ S nt  nt  ˆ   1 » 1 ºs2 i s1 s2 n i S   lim Ee nt1 Ee n nt2 nt1 1 n ª  CLT, CMT  t1 ˆs s 2  t2 t1 s2 e 2 1 2 e 2 2  1 ‰s2t 2s s t s2t Ž e 2 1 1 1 2 1 2 2  1 ‰s2t 2s s min˜t ,t s2t Ž e 2 1 1 1 2 1 2 2 2 ˆ  ϕˆW ˆt1,W ˆt2 s1, s2 , ˆ ˆ  ˆ  where ϕˆW ˆt1,W ˆt2 is the characteristic function of W t1 ,W t2 . Lemma 3.3.2 ˆn ˆn Let W˜ maxt> 0,1¥ W˜ ˆt. Then it holds:

ˆn d º1 W˜ max Sk, for all n > N n k 1,...,n and ¾ x 2 ˆn 2  y lim PˆW˜ B x S e 2 dy, for all x C 0. n ª π 0 Without proof

Proof of Theorem 3.3.1. We shall prove only the upper bound in Theorem 3.3.1. From Lemma 3.3.1 and the continuous mapping theorem it follows for x C 0, k C 1 and t1, . . . , tk > 0, 1¥ that

P Œ W˜ ˆnˆt A x‘ P Œ W ˆt A x‘ , lim ª max max n t>˜t1,...,tk t>˜t1,...,tk since ˆx1, ..., xk ( maxˆx1, ..., xk is continuous. Consequently, it holds

P Œ W˜ ˆnˆt A x‘ C P Œ W ˆt A x‘ , lim infª max max n t> 0,1¥ t>˜t1,...,tk 3 Wiener process 47

ˆ  ˆ  since ˜ max W˜ n ˆt A x b ˜ max W˜ n ˆt A x. t>˜t1,...,tk t> 0,1¥ — — 1 k i With ˆt1, . . . , tk ‰ ,..., Ž and max W ˆt lim max W ‰ Ž a.s. (and therefore partic- k k t> 0,1¥ k ª i 1,...,k k ularly in distribution) it holds

ˆ  i lim inf P Œ max W˜ n ˆt A x‘ C lim P Œ max W ‹  A x‘ P Œ max W ˆt A x‘ . n ª t> 0,1¥ k ª i 1,...,k k t> 0,1¥

Conclusively, the assertion follows from lemma 3.3.2. Corollary 3.3.1 Let ˜W ˆt, t C 0 be a Wiener process. Then it holds:

W ˆt P Œ lim 0‘ 1. t ª t

Proof Exercise.

3.3.2 Invariance properties Specific transformations of the Wiener process again reveal the Wiener process. Theorem 3.3.2 ˆ  Let ˜W ˆt, t C 0 be a Wiener process. Then the stochastic processes ˜Y i ˆt, t C 0, i 1,..., 4, with

ˆ  Y 1 ˆt W ˆt, (Symmetry) ˆ2ˆ  ˆ    ˆ  A Y t Wº t t0 W t0 for a t0 0, Translation of the origin) ˆ3ˆ  ˆ t  A Y t cW c for a c 0, (Scaling) 1 ˆ  tW ˆ , t A 0, Y 4 ˆt œ t (Reflection at t=0) 0, t 0. are Wiener processes as well.

ˆi ˆi ˆi Proof 1. Y , i 1,..., 4, have independent increments with Y ˆt2Y ˆt1 ¢ N ˆ0, t2  t1. ˆ  2. Y i ˆ0 0, i 1,..., 4. ˆ  ˆ  3. Y i , i 1,..., 3, have continuous trajectories. ˜Y i ˆt, t C 0 has continuous trajectories for t A 0. ˆ4ˆ  ˆ 1  4. We have to prove that Y t is continuous at t 0, i.e. that limt 0 tW t 0. ˆ 1  W ˆt a.s. limt 0 tW t limt ª t 0 because of Corollary 3.3.1.

Corollary 3.3.2 Let ˜W ˆt, t C 0 be the Wiener process. Then it holds:

P Œsup W ˆt ª‘ P ‹inf W ˆt ª 1. tC0 tC0 48 3 Wiener process and consequently

P Œsup W ˆt ª, inf W ˆt ª‘ . tC0 tC0 Proof For x, c A 0 it holds: t x x P Œsup W ˆt A x‘ P Œsup W ‹  A º ‘ P Œsup W ˆt A º ‘ tC0 tC0 c c tC0 c

¨ P Œœsup W ˆt 0¡ 8 œsup W ˆt ª¡‘ P Œsup W ˆt 0‘  P Œsup W ˆt ª‘ 1. tC0 tC0 tC0 tC0 Moreover it holds

P Œsup W ˆt 0‘ P Œsup W ˆt B 0‘ B P ŒW ˆt B 0, sup W ˆt B 0‘ tC0 tC0 tC1

P ŒW ˆ1 B 0, supˆW ˆt  W ˆ1 B W ˆ1‘ tC1 0 S P Œsup W ˆt  W ˆ1 B W ˆt S W ˆ1 x‘ P ˆW ˆ1 > dx ª tC1 0 S P ŒsupˆW ˆt  W ˆ1 B x S W ˆ1 x‘ P ˆW ˆ1 > dx ª tC0 0 S P Œsup W ˆt 0‘ P ˆW ˆ1 > dx ª tC0 1 P Œsup W ˆt 0‘ , tC0 2 ˆ ˆ   ˆ ˆ  ª thus P suptC0 W t 0 0 and thus P suptC0 W t 1. Analogously one can show that P ˆinftC0 W ˆt ª 1. The remaining part of the claim follows from P ˆA 9 B 1 for any A, B > F with P ˆA P ˆB 1. Remark 3.3.1 ˆ ˆ  ª ˆ  ª P suptC0 X t , inftC0 X t 1 implies that the trajectories of W oscillate between positive and negative values on 0, ª an infinite number of times. Corollary 3.3.3 Let ˜W ˆt, t C 0 be a Wiener process. Then it holds P ˆω > Ω ¢ W ˆω is nowhere differentiable in 0, ª 1. Proof

˜ω > Ω ¢ W ˆω is nowhere differentiable in 0, ª 9ª ˜ > ¢ ˆ    n 0 ω Ω W ω is nowhere differentiable in n, n 1 .

It is sufficient to show that Pˆω > Ω ¢ W ˆω is differentiable for a t0 t0ˆω > 0, 1¥ 0. Define the set 4 A ›ω > Ω ¢ it exists a t t ˆω > 0, 1¥ with SW ˆt ˆω  h, ω  W ˆt ˆω, ωS B mh, ¦h > £0, ¨ . nm 0 0 0 0 n 3 Wiener process 49

Then it holds

˜ω > Ω ¢ W ˆω differentiable for a t0 t0ˆω Anm. mC1 nC1

We still have to show Pˆ8mC1 8nC1 Anm 0. ˆ  ˜ k C ˆ  > Let k0 ω mink 1,2,... n t0 ω . Then it holds for ω Anm and j 0, 1, 2

k ˆω  j  1 k ˆω  j k ˆω  j  1 WW Œ 0 , ω‘  W Œ 0 , ω‘W B WW Œ 0 , ω‘  W ˆt ˆω, ωW n n n 0 k ˆω  j  WW Œ 0 , ω‘  W ˆt ˆω, ωW n 1 8m B . n ˆ  ˆ k1   ˆ k  Let ∆n k W n W n . Then it holds

’ n 2 8m“ PˆAnm B P  S∆nˆk  jS B ”k 0 j 0 n • 3 n ’ 2 8m “ n 8m B Q P  ›S ˆk  jS B ŒQ P ‹S ˆ S B ‘ ” ∆n • ∆n 0 k 0 j 0 n k 0 n 3 16m B ˆn  1 Œº ‘ 0, n ª, 2πn by the independence of the increments of the Wiener Process. Since Anm ` An1,m, it follows P ˆAnm 0.

Corollary 3.3.4 With probability 1 it holds:

n sup sup Q SW ˆti  W ˆti1S ª, nC1 0Bt0@...@tnB1 i 1 i.e. ˜W ˆt, t > 0, 1¥ possesses a.s. trajectories with unbounded variation.

Proof Since every continuous function g ¢ 0, 1¥ R with bounded variation is differentiable almost everywhere, the assertion follows from Corollary 3.3.3.

Alternative proof P2n U ‰ it Ž  Š ˆi1t U ª It is sufficient to show that limn ª i 1 W 2n W 2n . 2 P2n Š ‰ it Ž  Š ˆi1t   2 2 n1 Let Zn i 1 W 2n W 2n t. Hence EZn 0 and EZn t 2 and with Tchebysh- eff’s inequality

EZ2 t 2 ª P ˆSZ S @ ε B n ‹  n1, Q P ˆSZ S A ε a.s. . n 2 2 i.e. n 0 ε ε i 1 50 3 Wiener process

From lemma of Borel-Cantelli it follows that limn ª Zn 0 almost surely and thus

n 2 2 it ˆi  1t B t B Q ŒW ‹   W Œ ‘‘ 0 n n i 1 2 2 n kt ˆk  1t 2 it ˆi  1t B lim inf max WW ‹   W Œ ‘W Q WW ‹   W Œ ‘W . n ª B B n n n n n 1 k 2 2 2 i 1 2 2

Hence the assertion follows since W has continuous trajectories and therefore

kt ˆk  1t lim max WW ‹   W Œ ‘W 0. n ª 1BkB2n 2n 2n

3.4 Additional exercises

Exercise 3.4.1 Give an intuitive (exact!) method to realize trajectories of a Wiener process W ˜W ˆt, t > 0, 1¥. Thereby use the independence and the distribution of the increments of W . Addition- ally, write a program in R for the simulation of paths of W . Draw three paths t ( W ˆt, ω for t > 0, 1¥ in a common diagram. Exercise 3.4.2 ˜ ˆ  > ¥ ¢ ˆ  Given are the Wiener process W W t , t 0, 1 and L argmaxt> 0,1¥W t . Show that it holds: 2 º PˆL B x arcsin x, x > 0, 1¥. π d Hint: Use relation maxr> 0,t¥ W ˆr SW ˆtS. Exercise 3.4.3 For the simulation of a Wiener process W ˜W ˆt, t > 0, 1¥ we also can use the approximation

n Wnˆt Q Skˆtzk k 1 where Skˆt, t > 0, 1¥, k C 1 are the Schauder functions, and zk ¢ N ˆ0, 1 i.i.d. random variables and the series converges almost surely for all t > 0, 1¥ (n ª).

2 a) Show that for all t > 0, 1¥ the approximation Wnˆt also converges in the L -sense to W ˆt.

b) Write a program in R (alternative: C) for the simulation of a Wiener process W ˜W ˆt, t > 0, 1¥.

c) Simulate three paths t ( W ˆt, ω for t > 0, 1¥ and draw these paths into a common k 8  diagram. Hereby consider the sampling points tk n , k 0, . . . , n with n 2 1. Exercise 3.4.4 For the Wiener process W ˜W ˆt, t C 0 we define the process of the maximum that is given by M ˜Mˆt ¢ maxs> 0,t¥ W ˆs, t C 0. Show that it holds: 3 Wiener process 51

a) The density fMˆt of the maximum Mˆt is given by ¾ 2 x2 fMˆtˆx exp œ ¡ 1˜x C 0. πt 2t

Hint: Use property PˆMˆt A x 2PˆW ˆt A x.

b) Expected value and variance of Mˆt are given by ¾ 2t EMˆt , Var Mˆt tˆ1  2~π. π

ˆ  ¢ ˜ ˆ   Now we define τ x argmin s>R W s x as the first point in time for which the Wiener process takes value x.

c) Determine the density of τˆx and show that: Eτˆx ª. Exercise 3.4.5 Let W ˜W ˆt, t C 0 be a Wiener process. Show that the following processes are Wiener processes as well: ¢ ¨ , t , º ˆ  ¦0 0 ˆ  ˆ ~  A W1 t ¨ W2 t cW t c , c 0. ¤¨tW ˆ1~t, t A 0,

Exercise 3.4.6 The Wiener process W ˜W ˆt, t C 0 is given. Size Qˆa, b denotes the probability that the process exceeds the half line y at  b, t C 0, a, b A 0. Proof that:

a) Qˆa, b Qˆb, a and Qˆa, b1  b2 Qˆa, b1Qˆa, b2,

b) Qˆa, b is given by Qˆa, b exp˜2ab. 4 Lévy Processes

4.1 Lévy Processes

Definition 4.1.1 A stochastic process ˜Xˆt, t C 0 is called Lévy process, if

1. Xˆ0 0,

2. ˜Xˆt has stationary and independent increments,

3. ˜Xˆt is stochastically continuous, i.e for an arbitrary ε A 0, t0 C 0:

lim PˆSXˆt  Xˆt0S A ε 0. t t0

Remark 4.1.1 One can easily see, that compound Poisson processes fulfill the 3 conditions, since for arb. ε A 0 it holds λStt0S P ˆSXˆt  Xˆt0S @ ε C P ˆSXˆt  Xˆt0S A 0 B 1  e ÐÐ 0. t t0

Further holds for the Wiener process for arb. ε A 0 ¾ 2 ª y2 P ˆSXˆt  Xˆt0S A ε S exp Œ ‘ dy πˆt  t0 t 2ˆt  t0 y x » ª 2 tt0 2  x S e 2 dx ÐÐ . t 0 π » t t0 tt0

4.1.1 Infinitely Divisibility

Definition 4.1.2 Let X ¢ Ω R be an arbitrary random variable. Then X is called infinitely divisible, if for > ˆn ˆn d ˆn   ˆn arbitrary n N there exist i.i.d. random variables Y1 ,...,Yn with X Y1 ... Yn . Lemma 4.1.1 The random variable X ¢ Ω R is infinitely divisible if and only if the characteristic function ϕX of X can be expressed for every n C 1 in the form

n ϕX ˆs ˆϕnˆs for all s > R, where ϕn are characteristic functions of random variables.

52 4 Lévy Processes 53

Proof „ ¨ “ ˆn ˆn d ˆn ˆn n   ˆ  L ˆ  ˆ  Y1 ,...,Yn i.i.d., X Y1 ... Yn . Hence, it follows that ϕX s i 1 ϕ n s Yi n ˆϕnˆs . “ “ ˆ  ˆ ˆ n ¨ ˆn ˆn ϕX s ϕn s there exist Y1 ,...,Yn i.i.d. with characteristic function ϕn and ˆ  ˆ ˆ n ˆ  ϕY1,...,Yn s ϕn s ϕX s . With the uniqueness theorem for characteristic functions d ˆn   ˆn it follows that X Y1 ... Yn . Theorem 4.1.1 Let ˜Xˆt, t C 0 be a Lévy process. Then the random variable Xˆt is infinitely divisible for every t C 0.

Proof For arbitrary t C 0 and n > N it obviously holds that t 2t t nt ˆn  1t X ˆt X ‹   ‹X ‹   X ‹   ...  ŒX ‹   X Œ ‘‘ . n n n n n

Since ˜Xˆt has independent and stationary increments, the summands are obviously inde- pendent and identically distributed random variables. Lemma 4.1.2 Let X1,X2,... ¢ Ω R be a sequence of random variables. If there exists a function ϕ ¢ R C, ˆ  ˆ  ˆ  > such that ϕ s is continuous in s 0 and limn ª ϕXn s ϕ s for all s R, then ϕ is the d characteristic function of a random variable X and it holds that Xn Ð X. Definition 4.1.3 Let ν be a measure on the measure space ˆR, BˆR. Then ν is called a Lévy measure, if 툘0 0 and

S min ™y2, 1ž νˆdy @ ª. (4.1.1) R

Abb. 4.1: y ( minˆy2, 1

Note • Apparently every Lévy measure is σ-finite and

ν ˆˆε, εc @ ε, for all ε A 0, (4.1.2)

c where ˆε, ε R  ˆε, ε. 54 4 Lévy Processes

• In particular every finite measure ν is a Lévy measure, if 툘0 0. ˆ  ˆ  ˆ  ¢ 1 >  • If ν dy g y dy then g y SySδ for y 0, where δ 0, 3 . • An equivalent condition to (4.1.2) is y2 y2 y2 νˆdy @ ª, B ™y2, ž B . S 2 since 2 min 1 2 2 (4.1.3) R 1  y 1  y 1  y

Theorem 4.1.2 Let a > R, b C 0 be arbitrary and let ν be an arbitrary Lévy measure. Let the characteristic function of a random variable X ¢ Ω R be given through the function ϕ ¢ R C with 2 bs isy ϕˆs exp œias   S ‰e  1  isy1ˆy > ˆ1, 1Ž νˆdy¡ for all s > R. (4.1.4) 2 R Then X is infinitely divisible. Remark 4.1.2 • The formula (4.1.4) is also called Lévy-Khintchine formula.

• The inversion of Theorem 4.1.2 also holds, hence every infinitely divisible random variable has such a representation. Therefore the characteristic triplet ˆa, b, ν is also called Lévy characteristic of an infinitely divisible random variable.

Proof of Theorem 4.1.2 1st step Show that ϕ is a characteristic function. For y > ˆ1, 1 it holds • ª ˆisyk ª ˆisyk ª sk Teisy  1  isyT WQ  1  isyW WQ W B y2 WQ W B y2c k 0 k! k 2 k! k 2 k! ´¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¶ ¢ c Hence it follows from (4.1.1) that the integral in (4.1.4) exists and therefore it is well- defined.

• Let now ˜cn be an arbitrary sequence of numbers with cn A cn1 A ... A 0 and limn ª cn 0. Then the function ϕn ¢ R C with 2 bs isy ϕnˆs ¢ exp œis ‹a  S yνˆdy  ¡ exp ›S ‰e  1Ž νˆdy c c cn,cn¥ 9ˆ1,1 2 cn,cn¥

ˆn ˆn is the characteristic function of the sum of independent random variables Z1 and Z2 , since – the first factor is the characteristic function of the normal distribution with expec-

a  R c yνˆdy b tation cn,cn¥ 9ˆ1,1 and variance . – the second factor is the characteristic function of a compound Poisson process with parameters

c c c λ νˆ cn, cn¥  and PU ˆ  νˆ 9 cn, cn¥ ~νˆ cn, cn¥ 

. 4 Lévy Processes 55

• Furthermore limn ª ϕnˆs ϕˆs for all s > R, where ϕ is obviously continuous in 0, since it holds for the function ψ ¢ R C in the exponent of (4.1.4)

isy ψˆs S ‰e  1  isy1 ˆy > ˆ1, 1Ž νˆdy for all s > R R S ˆ S 2 2 ˆ   T isy  T ˆ  that ψ s cs Rˆ1,1 y ν dy Rˆ1,1c e 1 ν dy . Out of this and from (4.1.3) it follows by Lebesgue’s theorem that lim ψˆs 0. s 0

• Lemma 4.1.2 yields that the function ϕ given in (4.1.4) is the characteristic function of a random variable.

2nd step The infinite divisibility of this random variable follows from Lemma 4.1.1 and out of the fact, > ν that for arbitrary n N n is also a Lévy measure and that ¢ £ ¨ b s2 ¨ ˆ  ¦ a  n  ‰ isy   ˆ > ˆ Ž ‹ ν  ˆ § > ϕ s exp ¨i s S e 1 isy1 y 1, 1 dy ¨ for all s R. ¤¨ n 2 R n ¥¨

Remark 4.1.3 The map η ¢ R C with

bs2 ηˆs ias   S ‰eisy  1  isy1ˆy > ˆ1, 1Ž νˆdy 2 R from (4.1.4) is called Lévy exponent of this infinitely divisible distribution.

4.1.2 Lévy-Khintchine Representation

˜Xˆt, t C 0 – Lévy process. We want to represent the characteristic function of Xˆt, t C 0, through the Lévy-Khintchine formula. Lemma 4.1.3 Let ˜Xˆt, t C 0 be a stochastically continuous process, i.e. for all ε A 0 and t0 C 0 it holds ˆS ˆ   ˆ S A  > z ˆ  that limt t0 P X t X t0 ε 0. Then for every s R, t ϕXˆt s is a continuous map from 0, ª to C.

isy Proof • y z e continuous in 0, i.e. for all ε A 0 there exists a δ1 A 0, such that ε sup Teisy  1T @ . y>ˆδ1,δ1 2

• ˜Xˆt, t C 0 is stochastically continuous, i.e. for all t0 C 0 there exists a δ2 A 0, such that ε sup P ˆSXˆt  Xˆt0S A δ1 @ . tC0, Stt0S@δ2 4 56 4 Lévy Processes

Hence, it follows that for s > R, t C 0 and St  t0S @ δ2 it holds

ˆ  ˆ  ˆ  ˆ ˆ  ˆ  T ˆ   ˆ T U Š isX t  isX t0 U B U isX t0 Š is X t X t0  U ϕXˆt s ϕXˆt0 s E e e E e e 1 ˆ ˆ  ˆ  B U is X t X t0  U T isy  T ˆ  E e 1 S e 1 PXˆtXˆt0 dy R B T isy  T ˆ  S e 1 PXˆtXˆt0 dy ˆδ1,δ1 isy  S Te  1T PXˆtXˆt ˆdy ˆ c 0 δ1,δ1 ´¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¶ B2 isy B sup Te  1T  2P ˆSXˆt  Xˆt0S A δ1 B ε. y>ˆδ1,δ1

Theorem 4.1.3 Let ˜Xˆt, t C 0 be a Lévy process. Then for all t C 0 it holds ˆ  tηˆs > ϕXˆt s e , s R,

, where η ¢ R C is a continuous function. In particular it holds that

t ˆ  tηˆs Š ηˆs ‰ ˆ Žt > C ϕXˆt s e e ϕXˆ1 s , for all s R, t 0.

Proof Due to stationarity and independence of increments we have

ˆ  isXˆttœ Š isXˆt isˆXˆttœXˆt ˆ  ˆ  > ϕXˆttœ s Ee E e e ϕXˆt s ϕXˆtœ s , s R.

¢ ª ˆ  ˆ  > ˆ  œ ˆ  ˆ œ œ C Let gs 0, C be defined by gs t ϕXˆt s , s R, gs t t gs t gs t , t, t 0. Xˆ0 0. ¢ ˆ  œ ˆ  ˆ œ œ C ¨ gs t t gs t gs t , t, t 0, ¦ gsˆ0 1, ¨ ¤ gs ¢ 0, ª C continuous. ¢ ˆ  ηˆst > C ˆ  ηˆs Hence there exists η R C, such that gs t e for all s R, t 0. ϕXˆ1 s e and it follows that η is continuous.

Lemma 4.1.4 (Prokhorov): Let µ1, µ2,... be a sequence of finite measures (on BˆR) with ˆ  @ @ ª 1. supnC1 µn R c, c const (uniformly bounded)

2. for all ε A 0 there exists Bε > BˆR compact, such that fulfills the tightness condition ˆ c B supnC1 µn Bε ε. Hence follows that there exists a subsequence µn1 , µn2 ,... and a finite measure over BˆR, such that for all f ¢ R C, bounded, continous, it holds that

ˆ  ˆ  ˆ  ˆ  lim S f y µnk dy S f y µ dy k ª R R

Proof See [14], page 122 - 123. 4 Lévy Processes 57

Theorem 4.1.4 Let ˜Xˆt, t C 0 be a Lévy process. Then there exist a > R, b C 0 and a Lévy measure ν, such that 2 ias bs isy ˆ  2  ‰   ˆ > ˆ Ž ˆ  > ϕXˆ1 s e S e 1 iy1 y 1, 1 ν dy , for all s R. R

Proof For all sequences ˆtnn> b ˆ0, ª, with lim tn 0, it holds N n ª

ˆ  œ tnη s  ˆ   ˆ  e 1 ϕXˆtn s 1 ηˆs Šetη s  V , lim ª lim ª (4.1.5) t 0 n tn n tn since η ¢ R C is continuous. The latter convergence is even uniform in s > s0, s0¥ for any s0 A 0, since Taylor’s theorem yields

tnηˆs ª k e  1 1 ˆtnηˆs lim Wηˆs  W lim Wηˆs  Q W n ª n ª tn tn k 1 k! ª k 1 ˆtnηˆs lim W Q W n ª tn k 2 k! ª k ˆtnηˆs lim Wηˆs Q W n ª ˆ   k 1 k 1 ! ª ˆ ˆ k1 2 tnη s lim Wη ˆstn Q W n ª ˆ   k 1 k 1 ! ª S Sk1 2 tnM B lim M tn Q , where M ¢ sup SηˆsS @ ª n ª ˆ   k 1 k 1 ! s> s0,s0¥ ª S Sk1 2 tnM 1 lim M tn Q n ª ˆ   ˆ   k 1 k 1 ! k k 1 ª S Sk1 2 tnM B lim M tn Q n ª ˆ   k 1 k 1 ! 2 StnMS lim M tne n ª 0. 1 ˆ 1  Now let tn n and Pn be the distribution of X n . Hence it follows that

ϕ ˆ 1 ˆs  1 isy X n lim n S ˆe  1Pnˆds lim n ηˆs n ª n ª 1 R n s0 s0 S n S ‰eisy  Ž P ˆdyds S ηˆsds lim ª 1 n n R s0 s0 and consequently

sinˆs y 1 s0 1 s0 n S Œ  0 ‘ P ˆdy n S  S ‰eisy  Ž dsP ˆdy  S ηˆsds. lim ª 1 n lim ª 1 n n R s0y n R 2s0 s0 2s0 s0

Since η ¢ R C is continuous with ηˆ0 0 and it follows from the mean value theorem, that 1 s0 sinˆs0y 1 for all ε A 0 it exists δ0 A 0, such that U R ηˆsdsU @ ε. Since 1  C , for Ss0yS C 2, 2s0 s0 s0y 2 58 4 Lévy Processes

it holds: for all ε A 0 there exist s0 A 0, n0 A 0, such that

n sinˆs y S P ˆdy B n S Œ  0 ‘ P ˆdy @ ε. lim sup 2 n lim sup 1 n n ª 2 šy¢SySC Ÿ n ª s0y s0 R

Hence for all ε A 0 there exist s0 A 0, n0 A 0, such that

n S Pnˆdy B 4ε, for all n C n0. šy¢SySC 2 Ÿ s0

Decreasing s0 gives

n S Pnˆdy B 4ε, for all n C 1. šy¢SySC 2 Ÿ s0 y2 sin y B c ‹1   , for all y x 0 and a c A 0. 1  y2 y Hence, it follows that y2 n P ˆdy B cœ cœ @ ª. sup S 2 n for a nC1 R 1  y

Let now µn ¢ BˆR 0, ª be defined as

y2 µnˆB n S Pnˆdy for all B > BˆR. B 1  y2

2 ˜µ  µ ˆ  @ cœ y B It follows that n n>N is uniformly bounded, supnC1 n R . Furthermore holds 1y2 1, 2 sup C µn Ššy ¢ SyS A Ÿ B 4ε and ˜µn > relatively compact. After lemma 4.1.3 it holds: there n 1 s0 n N ˜µ  exists nk k>N, such that

ˆ  ˆ  ˆ  ˆ  lim S f y µnk dy S f y µ dy k ª R R for a measure µ and f continuous and bounded. Let for s > R the function fs ¢ R C be defined as ¢ 2 ¨ ‰ isy   ˆ Ž 1y x ¨ e 1 is sin y 2 , y 0, f ˆy ¦ y s ¨  s2 ¤ 2 , otherwise.

Hence follows that fs is bounded and continuous and

ηˆs n S ‰eisy  Ž P ˆdy lim ª 1 n n R ‹S f ˆyµ ˆdy  isn S yP ˆdy lim ª s n sin n n R R ‹ ˆ  ˆ   ˆ  lim S fs y µnk dy isnk S sin yPnk dy k ª R R ˆ  ˆ   ˆ  S fs y µ dy lim isnk S sin yPnk dy R k ª R 2 œ bs ηˆs ia s   S ‰eisy  1  is sin yŽ νˆdy, 2 R 4 Lévy Processes 59

œ s > a ª isn R yP ˆdy @ ª b µ ˆ˜  ν ¢ Bˆ  , ª for all R with limk k R sin nk , 0 , R 0 , ¢ ¨ 1y2 ˆ  x ˆ  ¦ y2 µ dy , y 0, ν dy ¨ ¤¨ 0 , y 0.

S Sy1ˆy > ˆ1, 1  sin yS νˆdy @ ª. R 2 1  y œœ œœ Sy1ˆy > ˆ1, 1  sin yS @ c , for all y x 0 and a c A 0. y2 Hence follows that

2 bs isy ηˆs ias   S ‰e  1  isy1 ˆy > ˆ1, 1Ž νˆdy, for all s > R. 2 R

œ a a  S ˆy1ˆy > ˆ1, 1  sin y νˆdy. R

4.1.3 Examples

1. Wiener process (it is enough to look at Xˆ1)  s2 Xˆ1 ¢ N ˆ0, 1, ϕXˆ1ˆs e 2 and hence follows

ˆa, b, ν ˆ0, 1, 0.

Let X ˜Xˆt, t C 0 be a Wiener proess with drift µ, i.e. Xˆt µt  σW ˆt, W ˜W ˆt, t C 0 – Brownian motion. It follows

ˆa, b, ν ˆµ, σ2, 0.

2 isXˆ1 ˆµσW ˆ1is µis isµσ2 s ˆ  ˆ  2 > ϕXˆ1 s Ee Ee e ϕW ˆ1 σs e , s R.

2. Compound Poisson process with parameters ˆλ, Pn ˆ  PNˆt ˆ  ¢ ˆ  ¢ X t i 1 Ui, N t Pois λt , Ui i.i.d. PU .

isx ϕXˆ1ˆs exp ›λ S ‰e  1Ž PU ˆdx R isx exp ›λis S x1ˆx > 1, 1¥PU ˆdx  λ S ‰e  1  isx1ˆx > 1, 1¥Ž PU ˆdx R R 1 ‰ isx Ž exp ›λis S xPU ˆdx  λ S e  1  isx1ˆx > 1, 1¥ PU ˆdx , s > R. 1 R Hence follows

1 ˆa, b, ν ‹λ S xPU ˆdx, 0, λPU  , PU – finite on R. 1 60 4 Lévy Processes

3. Process of Gauss-Poisson type X ˜Xˆt, t C 0, Xˆt X1ˆt  X2ˆt, t C 0. X1 ˜X1ˆt, t C 0 and X2 ˜X2ˆt, t C 0 independent. 2 X1 – Wiener process with drift µ and variance σ , X2 – Compound Poisson process with parameters λ, PU . ˆ  ˆ  ˆ  ϕXˆt s ϕX1ˆt s ϕX2ˆt s ’ 2 2 “ σ s isx exp –isµ   λ S e  1PU ˆdx— ” 2 • R 1 σ2s2 exp œis ‹µ  λ S xPU ˆdx  1 2 ‰ isx Ž  S λ e  1  isx1ˆx > 1, 1¥ PU ˆdx , s > R. R Hence follows 1 2 ˆa, b, ν ‹µ  λ S xPU ˆdx, σ , λPU  . 1 4. Stable Lévy process X ˜Xˆt, t C 0 – Lévy process with Xˆt ¢ α stable distribution, α > ˆ0, 2¥. To introduce α-stable laws ν, let us begin with an example. 2 If X W (Wiener process), then Xˆ1 ¢ N ˆ0, 1. Let Y,Y1,...,Yn be i.i.d. N ˆµ, σ - variables. Since the normal distribution is stable w.r.t. convolution it holds º º Y  ...  Y ¢ N ˆnµ, nσ2 d nY  nµ  nµ 1 n º º nY  µ ‰n  nŽ 1 2 1 n 2 Y  µ Šn 2  n 2 

1 1 n α Y  µ Šn  n α  , α 2.

Definition 4.1.4 The distribution of a random variable Y is called α-stable, if for all n > N independent copies Y1,...,Yn of Y exist, such that

d 1 Y1  ...  Yn n α Y  dn, where dn is deterministic. The constant α > ˆ0, 2¥ is called index of stability. Moreover, one can show that

¢ 1 ¨ Š  α  x ¦ µ n n , α 1, dn ¨ ¤¨ µn log n , α 1. Example 4.1.1 • α 2: Normal distribution, with any mean and any variance. • α 1: Cauchy distribution with parameters ˆµ, σ2. The density: σ fY ˆx , x > R. π ‰ˆx  µ2  σ2Ž

It holds EY 2 ª, EY does not exist. 4 Lévy Processes 61

1 ˆ 2 • α 2 : Lévy distribution with parameters µ, σ . The density:

¨¢ 1 ¨ ‰ σ Ž 2 1 š σ Ÿ A 2π 3 exp 2ˆxµ , x µ, f ˆx ¦ ˆxµ 2 Y ¨ ¤¨ 0 , otherwise.

These examples are the only examples of α-stable distribution, where an explicit form of > ˆ  x 1 the density is available. For other α 0, 2 , α 2 , 1, the α-stable distribution is introduced through its characteristic function. In general holds: If Y α-stable, α > ˆ0, 2¥, then ESY Sp @ ª, 0 @ p @ α. Definition 4.1.5 d The distribution of a random variable is called symmetric, if Y Y . If Y has a symmetric α-stable distribution, α > ˆ0, 2¥, then

α ϕY ˆs exp ˜c SsS  , s > R.

Indeed, it follows from the stability of Y that

1 n idns ˆϕY ˆs e ϕY Šn α s , s > R.

idns idns It follows that dn 0, since ϕY ˆs ϕY ˆs varphiY ˆs. It holds: e e , s > R and dn 0. The rest is left as an exercise. Lemma 4.1.5 Lévy-Khintchine representation of the characteristic function of a stable distribution. Any stable law is infinitely divisible with the Lévy triplet ˆa, b, ν, where a > R arbitrary,

σ2, α 2, b œ 0 , α @ 2. and , α , ˆ  œ 0 2 ν dx c1 ˆ C   c2 ˆ @  @ C ¢  A x1α 1 x 0 dx SxS1α 1 x 0 dx, α 2, c1, c2 0 c1 c2 0 Without proof Exercise: Prove that ¢ 2 ¨  x ¢ e 2σ2 , α , P ˆSY S C x ª ¦ 2 x ¨ c @ ¤ xα , α 2. Definition 4.1.6 The Lévy process X ˜Xˆt, t C 0 is called stable, if Xˆ1 has an α-stable distribution, α > ˆ0, 2¥ (α 2: Brownian motion (with drift)).

4.1.4 Subordinators Definition 4.1.7 A Lévy process X ˜Xˆt, t C 0 is called subordinator, if for all 0 @ t1 @ t2, Xˆt1 B Xˆt2 a.s. Since Xˆ0 0 a.s. ¨ Xˆt C 0, t C 0, a.s. 62 4 Lévy Processes

b ˆ  ˆ  This class of Subordinators is important since you can easily introduce Ra g t dX t as a Lebesgue-Stieltjes-integral. Theorem 4.1.5 The Lévy process X Xˆt, t C 0 is a subordinator if and only if the Lévy-Khintchine repre- sentation can be expressed in the form

ˆ  ›  ‰ isx  Ž ˆ  > ϕXˆ1 s exp ias S e 1 ν dx , s R, (4.1.6) R where a > 0, ª and ν is the Lévy measure, with ª ν ˆˆª, 0 0, S min ™1, y2ž νˆdy @ ª. 0 Proof Sufficiency It has to be shown that Xˆt2 C Xˆt1 a.s., if t2 C t1 C 0. First of all we show that Xˆ1 C 0 a.s.. If ν ¡ 0, then Xˆ1 a C 0 a.s., hence ˆ  ‰ ˆ Žt iats > ϕXˆt s ϕXˆ1 s e , s R. Xˆt at a.s. and therefore it follows that Xˆt ¡ and X is a subordinator. ˆ ª A A C @ ‰¡ 1 ªŽŽ @ ª If ν 0, 0, then there exists N 0 such that for all n N it holds 0 ν n , . It follows ª isx ias ϕXˆ1ˆs exp œias  lim S ‰e  1Ž νˆdx¡ e lim ϕnˆs, s > , n ª 1 n ª R n ª isx where ϕnˆs R 1 ‰e  1Ž νˆdx is the characteristic function of a compound Poisson process n ’ ν‰ 9£ 1 ,ªŽ “ ‰¡ 1 ªŽŽ n > distribution with parameters ν n , , ‰¡ 1 ªŽŽ for all n N. Let Zn be the random ” ν n , • PNn ¢ ‰ ‰¡ 1 ªŽŽŽ variable with characteristic function ϕn. It holds: Zn i 1 Ui, Nn Pois ν n , , ν‰ 9¡ 1 ,ªŽŽ d ¢ n C ˆ   C Ui ‰¡ 1 ªŽŽ ; hence follows Zn 0 a.s. and X 1 a lim Zn 0 a.s.. Since X is a Lévy ν n , ® ´¹¹¹¹¹¸¹¹¹¹¹¹¶ C 0 C0 process, it holds 1 2 1 n n  1 X ˆ1 X ‹   ‹X ‹   X ‹   ...  ‹X ‹   X ‹  , n n n n n ‰ k Ž  ‰ k1 Ž a.s.C where, because of stationarity and independence of the increments, X n X n 0 for 1 B k B n for all n. Hence Xˆq2  Xˆq1 C 0 a.s. for all q1, q2 > Q, q2 C q1 C 0. Now let t1, t2 > R B B š ˆn ˆnŸ ˆn B ˆn such that 0 t1 t2. Let q1 , q2 be sequences of numbers from Q with q1 q2 such ˆn £ ˆn ¡ ª A that q1 t1, q2 t2, n . For ε 0 ’ “ – — ˆ ˆ   ˆ  @   – ˆ   Š ˆn  Š ˆn  Š ˆn  Š ˆn  ˆ  @  — P X t2 X t1 ε P –X t2 X q2 X q2 X q1 X q1 X t1 ε— ” ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ • C0 B Š ˆ   Š ˆn  Š ˆn  ˆ  @   P X t2 X q2 X q1 X t1 ε ˆn ε ˆn ε B P ‹Xˆt2  X Šq  @    P ‹X Šq   Xˆt1 B   ÐÐÐ 0, 2 2 1 2 n ª 4 Lévy Processes 63 since X is stochastically continuous. Then

P ˆXˆt2  Xˆt1 @ ε 0 for all ε A 0 and

P ˆXˆt2  Xˆt1 @ 0 lim P ˆXˆt2  Xˆt1 @ ε 0 ε 0 ¨ Xˆt2 C Xˆt1 a.s. Necessity Let X be a Lévy process, which is a subordinator. It has to be shown that ϕXˆ1ˆ  has the form (4.1.6). After the Lévy-Khintchine representation for Xˆ1 it holds that 2 2 ª ˆ  œ  b s  ‰ isx   ˆ >  ¥Ž ˆ ¡ > ϕXˆ1 s exp ias S e 1 isx1 x 1, 1 ν dx , s R. 2 0 a.s. The measure ν is concentrated on 0, ª, since Xˆt C 0 for all t C 0 and from the proof of Theorem 4.1.4 ν ˆˆª, 0 0 can be chosen. 2 2 ª b s isx ϕXˆ1ˆs exp œias  ¡ exp ›S ‰e  1  isx1 ˆx > 1, 1¥Ž νˆdx 2 0 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ¢ ¢ ϕ ϕY ˆs Y1ˆs 2 2 Hence it follows that Xˆ1 Y1  Y2, where Y1 and Y2 are independent, Y1 ¢ N ˆa, b  and therefore b 0. (Otherwise Y1 could attain negative values and consequently Xˆ1 could attain negative values as well.) For all ε > ˆ0, 1 1 ε ª isx isx ϕXˆ1ˆs exp ›is ‹a  S xνˆdx  S ‰e  1  isxŽ νˆdx  S ‰e  1Ž νˆdx . ε 0 0 ª ‰ isx  Ž ˆ  ª ‰ isx  Ž ˆ  @ ª It has to be shown that for ε 0 it holds Rε e 1 ν dx R0 e 1 ν dx with 1 ˜  ˆ  @ ª ˆ  š Š  1 ˆ Ÿ ˆ  ˆ  R0 min x, 1 ν dx . ϕXˆ1 s exp is a Rε xν dx ϕZ1 s ϕZ2 s , where Z1 and Z2 ε ˆ  œ ‰ isx   Ž ˆ ¡ ˆ  ™ ª ‰ isx  Ž ˆ ž are independent, ϕZ1 s exp R e 1 isx ν dx , ϕZ2 s exp Rε e 1 ν dx , 0 d ˆ  EZ2 ˆ  s > Xˆ  a  R 1 xνˆdx  Z  Z ϕ 2 ˆ  1 @ ª ϕ 1 ˆ  iEZ R. Then 1 ε 1 2. There exist Z1 0 2 , Z1 0 0 1 and it therefore follows that EZ1 0 and PˆZ1 B 0 A 0. On the other hand, Z2 has a compound Š ˆ ª νˆ 9 ε,ª¥  > ˆ  Poisson distribution with parameters ν ε, , νˆ ε,ª , ε 0, 1 .

¨ P ˆZ2 B 0 A 0, since PˆZ2 0 A 0.

¨ P ˆZ1  Z2 B 0 C P ˆZ1 B 0,Z2 B 0 P ˆZ1 B 0 P ˆZ2 B 0 A 0 ˆ   1 ˆ  C > ˆ  C For X 1 to be positive it follows that a Rε xν dx 0 for all ε 0, 1 . Hence a 0 and ª S min ˜x, 1 dx @ ª. 0 d Moreover, for ε £ 0 it holds Z1 0 and consequently 1 ª ˆ  › ‹  ˆ   ‰ isx  Ž ˆ  > ϕXˆ1 s exp is a S xν dx S e 1 ν dx , s R. 0 0 64 4 Lévy Processes

Example 4.1.2 (α-stable subordinator): Let X ˜Xˆt, t C 0 be a subordinator, with a 0 and the Lévy measure

α 1 A 1α dx, x 0, νˆdx œ È1፠x 1 B 0 x1α dx 0, x 0.

By Lemma 4.1.5 it follows that X is an α-stable Lévy process. sXˆt tsα We show that ˆlXˆtˆs Ee e for all s, t C 0.

ª ˆ  ‰ ˆ Žt œ ‰ isx  Ž α 1 ¡ > ϕXˆt s ϕXˆ1 s exp t S e 1  dx , s R. 0 È1  ፠x1 α

It has to be shown that ª α α ˆ  ux dx C u S 1 e  , u 0. È1  ፠0 x1 α ˆ  ˜ > ¢ C  ˆ  This is enough since ϕXˆt can be continued analytically to z C Imz 0 , i.e. ϕXˆt iu ˆlXˆtˆu, u C 0. In fact, it holds that

ª ª x ˆ  ux dx uy 1α S 1 e  S u S e dyx dx 0 x1 d 0 0 ª ª Fubini    S S ue uyx 1 αdxdy 0 y ª ª    S S x 1 αdxue uydy 0 y ª u   S e uyy αdy α 0 ª Subst. u z α 1 ‹ z  S e z  d α 0 u α u α ª u  ˆ   S e zz 1 α 1dz α 0 uα È1  ፠α

tsα and hence follows ˆlXˆtˆs e , t, s C 0.

4.2 Additional Exercises

Exercise 4.2.1 Let X eb a random variable with distribution function F and characteristic function ϕ. Show that the following statements hold:

a) If X is infinitely divisible, then it holds ϕˆt x 0 for all t > R. Hint: Show that 2 n 2 limn ª SϕnˆsS 1 for all s > R, if ϕˆs ˆϕnˆs . Note further that SϕnˆsS is 1 again a characteristic function and limn ª x n 1 holds for x A 0.

b) Give an example (with explanation) for a distribution, which is not infinitely divisible. 4 Lévy Processes 65

Exercise 4.2.2 Let X ˜Xˆt, t C 0 be a Lévy process. Show that the random variable Xˆt is then infinitely divisible for every t C 0. Exercise 4.2.3 Show that the sum of two independent Lévy processes is again a Lévy process, and state the corresponding Lévy characteristic. Exercise 4.2.4 Look at the following function ϕ ¢ R C with ª ˆ   ϕˆt eψ t , where ψˆt 2 Q 2 kˆcosˆ2kt  1. k ª Show that ϕˆt is the characteristic function of an infinitely divisible distribution. Hint: Look k k at the Lévy-Khintchine representation with measure 툘2  2 , k > Z. Exercise 4.2.5 Let the Lévy process ˜Xˆt, t C 0 be a Gamma process with parameters b, p A 0, that is, for every t C 0 it holds Xˆt ¢ Èb, pt. Show that ˜Xˆt, t C 0 is a subordinator with the Laplace ˆ  ªˆ  uy ˆ  ˆ  1 by A exponent ξ u R0 1 e ν dy with ν dy py e dy, y 0. (The Laplace exponent  ˆ   ˆ  of ˜Xˆt, t C 0 is the function ξ ¢ 0, ª 0, ª, for which holds that Ee uX t e tξ u for arbitrary t, u C 0) Exercise 4.2.6 Let ˜Xˆt, t C 0 be a Lévy process with charactersistic Lévy exponent η and ˜τˆs, s C 0 a independent subordinator witch characteristic Lévy exponent γ. The stochastic process Y be defined as Y ˜Xˆτˆs, s C 0. (a) Show that iθY ˆs γˆiηˆθs E Še  e , θ > R, where Imz describes the imaginary part of z. ˆ  ˆ  Hint: Since τ is a process with non-negative values, it holds Eeiθτ s eγ θ s for all θ > ˜z > C ¢ Imz C 0 through the analytical continuation of Theorem 4.1.3. (b) Show that Y is a Lévy process with characteristic Lévy exponent γˆiηˆ . Exercise 4.2.7 Let ˜Xˆt, t C 0 be a compound Poisson process with Lévy measure º λ 2  x2 νˆdx º e 2σ2 dx, x > , σ π R where λ, σ A 0. Show that ˜σW ˆNˆt, t C 0 has the same finite-dimensional distributions as X, where ˜Nˆs, s C 0 is a Poisson process with intensity 2λ and W is a standard Wiener process independent from N. Hint to exercise 4.2.6 a) and exercise 4.2.7

• In order to calculate the expectation for the characteristic function, the identity EˆX EˆEˆXSY  R EˆXSY yF ˆdy X Y R Y for two random variables and can be used. In doing so, it should be conditioned on τˆs. 66 4 Lévy Processes

2 º 2 ª  y  as ˆ  2a 2 A > • Rª cos sy e dy 2πa e for a 0 and s R. Exercise 4.2.8 α Let W be a standard Wiener process and τ an independent 2 -stable subordinator, where α > ˆ0, 2. Show that ˜W ˆτˆs, s C 0 is a α-stable Lévy process. Exercise 4.2.9 Show that the subordinator T with marginal density

t  3  t2 fT ˆtˆs º s 2 e 4s 1˜s A 0 2 π 1 ˆ  is a 2 -stable subordinator. (Hint: Differentiate the Laplace transform of T t and solve the differential equation) 5 Martingales

5.1 Basic Ideas

Let ˆΩ, F, P be a complete probability space. Definition 5.1.1 Let ˜Ft, t C 0 be a family of σ-algebras Ft ` F. It is called

1. a filtration if Fs b Ft, 0 B s @ t.

2. a complete filtration if it is a filtration such that F0 (and therefore all Fs, s A 0) contains all sets of zero probability. Later on we will always assume, that we have a complete filtration.

3. a right-continuous filtration if for all t C 0 Ft 9sAtFs.

4. a natural filtration for a stochastic process ˜Xˆt, t C 0, if it is generated by the past of the process until time t C 0, i.e. for all t C 0 Ft is the smallest σ-algebra which contains — n the sets ˜ω > Ω ¢ ˆXˆt1,...,Xˆtn ` B, for all n > N, 0 B t1, . . . , tn B t, B > BˆR .

A random variable τ ¢ Ω R is called stopping time (w.r.t. the filtration ˜Ft, t C 0), if for all t C 0 ˜ω > Ω ¢ τˆω B t > Ft. If ˜Ft, t C 0 is the natural filtration of a stochastic process ˜Xˆt, t C 0, then τ being a stopping time means that by looking at the past of the process X you can tell, whether the moment τ occurred. Lemma 5.1.1 Let ˜Ft, t C 0 be a right-continuous filtration. τ is a stopping time w.r.t. ˜Ft, t C 0 if and only if ˜τ @ t > Ft for all t C 0. ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ˜ω>Ω¢τˆω@t>Ft

Proof „ “ Let ˜τ @ t > Ft, t C 0. To show: ˜τ B t > Ft. ˜τ B t 9s>ˆt,t単τ @ s for all ε A 0 ¨ ˜τ B t > 9sAtFs Ft

„ ¨ “ To show: ˜τ B t > Ft, t C 0 ¨ ˜τ @ t > Ft, t C 0. ˜τ @ t 8s>ˆ0,t˜τ B t  s > 8s>ˆ0,tFts ` Ft

Definition 5.1.2 Let ˆΩ, F, P be a probability space, ˜Ft, t C 0 a filtration (Ft ` F, t C 0) and X ˜Xˆt, t C 0 a stochastic process on ˆΩ, F, P. X is adapted w.r.t. the filtration ˜Ft, t C 0, if Xˆt is Ft-measurable for all t C 0, i.e. for all B > BˆR ˜Xˆt > B > Ft.

67 68 5 Martingales

Definition 5.1.3 The time τBˆω inf˜t C 0 ¢ Xˆt > B, ω > Ω, is called first hitting time of the set B > BˆR by the stochastic process X ˜Xˆt, t C 0 (also called: first passage time, first entrance time). Theorem 5.1.1 Let ˜Ft, t C 0 be a right-continuous filtration and X ˜Xˆt, t C 0 an adapted (w.r.t. ˜Ft, t C 0) càdlàg process. For open B ` R, τB is a stopping time. If B is closed, then τ˜Bˆω inf˜t C 0 ¢ Xˆt > B or Xˆt > B is a stopping time, where Xˆt lims¡t Xˆs.

Proof 1. Let B > BˆR be open. Because of Lemma 5.1.1 it is enough to show that ˜τB @ t > Ft, t C 0. Because of right-continuity of the trajectories of X it holds: ˜ @  8 ˜ ˆ  >  > 8 b b @ τB t s>Q9 0,t X s B s>Q9 0,tFs Ft, since Fs Ft, s t.

2. Let B > BˆR be closed. For all ε A 0 let Bε ˜x > R ¢ dˆx, B @ ε be a parallel set of B, where dˆx, B infy>B Sx  yS. Bε is open for all ε A 0. ’ “ ’ “ ˜τ˜B B t ˜Xˆs > B 8  ˜Xˆs > B 1  > Ft, ” • ” C n • s>Q9ˆ0,t¥ n 1 s>Q9ˆ0,t

since X is adapted w.r.t. ˜Ft, t C 0.

Lemma 5.1.2 Let τ1, τ2 be stopping times w.r.t. the filtration ˜Ft, t C 0. Then min˜τ1, τ2, max˜τ1, τ2, τ1  τ2 and ατ1, α C 1, are stopping times (w.r.t. ˜Ft, t C 0).

Proof For all t C 0 holds: ˜min˜τ1, τ2 B t ˜τ1 B t 8 ˜τ2 B t > Ft, ´¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¶ >Ft >Ft ˜max˜τ1, τ2 B t ˜τ1 B t 9 ˜τ2 B t > Ft, t t ˜ατ1 B t ˜τ1 B  > F t ` Ft, since B t, α α α c ˜τ1  τ2 B t ˜τ1  τ2 A t ˜τ1 A t 8 ˜τ2 A t 8 ˜τ1 C t, τ2 A 0 8 ˜τ2 C t, τ1 A 0 8 ´¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ >Ft >Ft >Ft >Ft ˜0 @ τ2 @ t, τ1  τ2 A t 8 ˜0 @ τ1 @ t, τ2  τ1 A t, To show: ˜0 @ τ2 @ t, τ1  τ2 A t > Ft. (That also ˜0 @ τ1 @ t, τ2  τ1 A t > Ft works analogously.) ˜0 @ τ2 @ t, τ1  τ2 A t  ˜s @ τ1 @ t, τ2 A t  s > Ft s>Q9ˆ0,t Theorem 5.1.2 Let τ be an a.s. finite stopping time w.r.t. the filtration ˜Ft, t C 0 on the probability space ˆ  ˆ @ ª ˜  Ω, F, P , i.e. P τ 1. Then there exists a sequence of discrete stopping times τn n>N, τ1 C τ2 C τ3 C ..., such that τn £ τ, n ª a.s.

Proof For all n > N let ˆ  œ 0, if τ ω 0 τn k1 k @ ˆ  B k1 > 2n , if 2n τ ω 2n , for a k N0 C > § > ¢ k B B k1 ˜ B  ˜ B k  For all t 0 and for all n N k N0 2n t 2n , i.e. it holds τn t τn 2n ˜ B k  > ` ¨ £ ª τ n F k Ft τn is a stopping time. Obviously τn τ, n a.s. 2 2n 5 Martingales 69

Corollary 5.1.1 Let τ be an a.s. finite stopping time w.r.t. the filtration ˜Ft, t C 0 and X ˜Xˆt, t C 0 a càdlàg process on ˆΩ, F, P, Ft ` F for all t C 0. Then Xˆω, τˆω, ω > Ω is a random variable on ˆΩ, F, P.

Proof To show: Xˆτ ¢ Ω R measurable, i.e. for all B > BˆR ˜Xˆτ > B > F. Let τn £ τ, n ª be as in Theorem 5.1.2. Since X is càdlàg, it holds that Xˆτn ÐÐÐ Xˆτ a.s.. Then n ª Xˆτ is F-measurable as the limit of Xˆτn, which are themselves F-measurable. Indeed, for all B > BˆR it holds

’ “ ª – k k — ˜Xˆτ  > B –›τ 9 ›Xˆ  > B — > F n – n n n — k 0 – 2 2 — ”´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶• >F >F

5.2 (Sub-, Super-)Martingales

Definition 5.2.1 Let X ˜Xˆt, t C 0 be a stochastic process adapted w.r.t. to a filtration ˜Ft, t C 0, Ft ` F, t C 0, on the probability space ˆΩ, F, P, E SXˆtS @ ª, t C 0. X is called martingale, resp. sub- or supermartingale, if EˆXˆt S Fs Xˆs a.s., resp. EˆXˆt S Fs C Xˆs a.s. or EˆXˆt S Fs B Xˆs a.s. for all s, t C 0 with t C s: ¨ EˆXˆt EˆXˆs const for all s, t. Examples

Very often martingales are constructed on the basis of a stochastic process Y ˜Y ˆt, t C 0 as follows: Xˆt gˆY ˆt  EgˆY ˆt for some measurable function g ¢ R R, or by iuY ˆt Xˆt e , for any fixed u > . ϕY ˆtˆu R

1. Poisson process Let Y ˜Y ˆt, t C 0 be the homogeneous Poisson process with intensity λ A 0. EY ˆt Var Y ˆt λt since Y ˆt ¢ Poisˆλt, t C 0. a) Xˆt Y ˆt  λt, t C 0 ¨ Xˆt is a martingale w.r.t. the natural filtration ˜Fs, s C 0.

EˆXˆt S FssBt EˆY ˆt  λt  ˆY ˆs  λs  ˆY ˆs  λs S Fs

Y ˆs  λs  EˆY ˆt  Y ˆs  λˆt  s S Fs ind.incr. Y ˆs  λs  EˆY ˆt  Y ˆs  λˆt  s Y ˆs  λs  EˆY ˆt  s λˆt  s ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ λˆts Y ˆs  λs a.s. Xˆs 70 5 Martingales

2 b) X˜ˆt X ˆt  λt, t C 0 ¨ X˜ˆt is a martingale w.r.t. ˜Fs, s C 0.

2 2 EˆX˜ˆt S Fs EˆX ˆt  λt S Fs EˆˆXˆt  Xˆs  Xˆs  λt S Fs 2 2 EˆˆXˆt  Xˆs  2ˆXˆt  XˆsXˆs  X ˆs  λs  λˆt  s S Fs X˜ˆs  EˆˆXˆt  Xˆs2 2Xˆs EˆXˆt  Xˆs λˆt  s ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ VarˆY ˆtY ˆs λˆts 0 a.s. Xœˆs, s B t.

2. Compound Poisson process ˆ  PNˆt C A Y t i 1 Ui, t 0, N – homogeneous Poisson process with intensity λ 0, Ui – independent identically distributed random variables, ESUiS @ ª, ˜Ui independent of N. Let Xˆt Y ˆt  EY ˆt Y ˆt  λtEU1, t C 0. Exercise 5.2.1 Show that X ˜Xˆt, t C 0 is a martingale w.r.t. its natural filtration.

3. Wiener process Let W ˜W ˆt, t C 0 be a Wiener process, ˜Fs, s C 0 be the natural filtration. a) Y ˜Y ˆt, t C 0, where Y ˆt ¢ W 2ˆt  EW 2ˆt W 2ˆt  t, t C 0, is a martingale w.r.t. ˜Fs, s C 0. Indeed, it holds

EˆY ˆt S Fs 2 EˆˆW ˆt  W ˆs  W ˆs  s  ˆt  s S Fs 2 2 EˆˆW ˆt  W ˆs  2W ˆsˆW ˆt  W ˆs  W ˆs S Fs  s  ˆt  s 2 2 EˆˆW ˆt  W ˆs   2W ˆsEˆW ˆt  W ˆs  EˆW ˆs S Fs  s  ˆt  s t  s  W 2ˆs  s  ˆt  s W 2ˆs  s a.s. Y ˆs, s B t.

uW ˆtu2 t b) Y˜ ˆt ¢ e 2 , t C 0 and a fixed u > R. u2 t uW ˆt u2 t u2 t ESY˜ ˆtS e 2 Ee e 2 e 2 1 @ ª. We show that Y˜ ˜Y˜ ˆt, t C 0 is a martingale w.r.t. ˜Fs, s C 0.

ˆ   uˆW ˆtW ˆsW ˆsu2 s u2 t s EˆY˜ ˆt S Fs Eˆe 2 2 S Fs ˆ   u2 s uW ˆs u2 t s uˆW ˆtW ˆs e 2 e e 2 Eˆe S Fs ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ œˆ  Y s 2 ˆts EˆeuW ˆts eu 2 ˆ   ˆ   u2 t s u2 t s Y˜ ˆse 2 e 2 Y˜ ˆs, s B t.

4. Regular martingale Let X be a random variable (on ˆΩ, F, P) with ESXS @ ª. Let ˜Fs, s C 0 be a filtration on ˆΩ, F, P. Construct Y ˆt EˆX S Ft, t C 0. Y ˜Y ˆt, t C 0 is a martingale. Indeed, ESY ˆtS ESEˆX S FtS B EˆEˆSXSS Ft ESXS @ ª, t C 0. 5 Martingales 71

a.s. EˆY ˆt S Fs EˆEˆX S Ft S Fs EˆX S Fs Y ˆs, s B t since Fs b Ft. If ˜Y ˆnS n 1, ..., N, where N > N, is a martingale, then it is regular since Y ˆn EˆY ˆN S Fn for all n 1, ..., N, i.e. X ¢ Y ˆN. However, the latter is not possible for processes of the form ˜Y ˆnS n > N or ˜Y ˆtS t C 0.

5. Lévy processes Let X ˜Xˆt, t C 0 be a Lévy process with Lévy exponent η and natural filtration ˜Fs, s C 0.

a) If ESXˆ1S @ ª, define Y ˆt Xˆt  tEXˆ1, t C 0. As in the previous cases it can ´¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¶ EXˆt be shown that Y ˜Y ˆt, t C 0 is martingale w.r.t. the filtration ˜Fs, s C 0. (Compare Example1 and2) ˆ  b) In the general case one can use the combination from example 3b – normalize eiuX t eiuXˆt eiuXˆt iuXˆttηˆu by the characteristic function of Xˆt, i.e. let Y ˆt ˆ  ˆ  e , ϕ u etη u Xˆt t C 0, u > R. To show: Y ˜Y ˆt, t C 0 is a complex-valued martingale. tηˆu ESY ˆtS Se S @ ª, since η ¢ R C. EY ˆt 1, t C 0. Furthermore, it holds

iuˆXˆtXˆsˆtsηˆu iuXˆssηˆu EˆY ˆt S Fs Eˆe e S Fs eiuXˆssηˆueˆtsηˆuEˆeiuˆXˆtXˆs Y ˆseˆtsηˆueˆtsηˆu a.s. Y ˆs

6. Monotone Submartingales/Supermartingales Every integrable stochastich process X ˜Xˆt, t C 0, which is adapted w.r.t. to a filtration ˜Fs, s C 0 and has a.s. monotone nondecreasing (resp. non-increasing) trajectories, is a sub- (resp. a super-)martingale. a.s. a.s. a.s. In fact, it holds e.g. Xˆt C Xˆs, t C s ¨ EˆXˆt S Fs C EˆXˆs S Fs Xˆs. In particular, every subordinator is a submartingale. Lemma 5.2.1 Let X ˜Xˆt, t C 0 be a stochastic process, which is adapted w.r.t. a filtration ˜Ft, t C 0 and let f ¢ R R be convex such that ESfˆXˆtS @ ª, t C 0. Then Y ˜fˆXˆt, t C 0 is a sub-martingale, if

a) X is a martingale, or

b) X is a sub-martingale and f is monotone nondecreasing.

Proof Use Jensen’s inequality for conditional expectations.

a.s. a.s. EˆfˆXˆt S Fs C fˆEˆXˆt S Fs C fˆXˆs. ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ b) a) C Xˆs, Xˆs 72 5 Martingales

5.3 Uniform Integrability

a.s. L1 It is known, that in general Xn ÐÐÐ X does not yield Xn ÐÐÐ X. Here X,X1,X2,... n ª n ª a.s. are random variables, defined on the probability space ˆΩ, F, P. When does „Xn ÐÐÐ X“ n ª L1 ¨ „Xn ÐÐÐ X“hold? The answer for this provides the notion of uniformly integrability of n ª ˜Xn, n > N. Definition 5.3.1 The sequence ˜Xn, n > N of random variables is called uniformly integrable, if ESXnS @ ª, n > , and sup EˆSXnS1ˆSXnS A 卍 ÐÐÐ 0. N ε ª n>N Lemma 5.3.1 The sequence ˜Xn, n > N of random variables is uniformly integrable if and only if

1. sup ESXnS @ ª (uniformly bounded) and n>N

2. for every ε A 0 there is a δ A 0 such that EˆSXnS1ˆA @ ε for all n > N and all A > F with PˆA @ δ.

Proof Let ˜Xn be a sequence of random variables. It has to be shown that

1 sup ESXnS @ ª sup EˆSXnS1ˆSXnS A x ÐÐÐ 0 ¨ n>N x ª n>N 2 ¦ε A 0 §δ A 0 ¢ EˆSXnS1ˆA @ ε ¦A > F ¢ PˆA @ δ „ “ ˜S S A  > A ˆ  B 1 S S Set An Xn x for all n N and x 0. It holds P An x E Xn by virtue of Markov’s in- 1 c equality and consequently sup PˆAn B sup ESXnS B ÐÐÐ 0 ¨ §N A 0 ¢ ¦x A N PˆAn @ δ n x n x x ª  ¨2 ˆS S ˆ  B supn E Xn 1 An ε. Since ε A 0 can be chosen arbitrarily small ¨ sup EˆSXnS1ˆSXnS A x ÐÐÐ 0. n x ª

„¨“

1.

sup ESXnS B supˆEˆSXnS1ˆSXnS A x  EˆSXnS1ˆSXnS B x n n B supˆEˆSXnS1ˆSXnS A x  x PˆSXnS B x n ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ B1 B ε  x @ ª A § A ˆS S ˆS S A  @ ε 2. For all ε 0 x 0 such that E Xn 1 Xn x 2 because of uniform integrability. A @ ε Choose δ 0 such that xδ 2 . Then it holds

EˆSXnS1ˆA Eˆ SXnS 1ˆSXnS B x 1ˆA  EˆSXnS1ˆSXnS A x 1ˆA  ± ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ² Bx B1 B1 B xPˆA  EˆSXnS1ˆSXnS A x, ´¹¹¹¹¹¹¸¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ B ε B ε 2 2 5 Martingales 73

Lemma 5.3.2 P Let ˜Xnn> be a sequence of random variables with ESXnS @ ª, n > , Xn ÐÐÐ X. Then N N n ª L1 Xn ÐÐÐ X if and only if ˜Xnn> is uniformly integrable. n ª N P Particularly, Xn ÐÐÐ X implies EXn ÐÐÐ EX. n ª n ª

Proof 1) Let ˜Xnn> be uniformly integrable. It has to be shown that ESXn  XS ÐÐÐ 0. N n ª P Since Xn ÐÐÐ X one obtains that ˆXnn> has a subsequence ˆXn k> converging almost n ª N k N surely to X. Consequently, Fatou’s Lemma yields

ESXS B lim inf SXn S B sup ESXnS, ª E k k n>N and therefore ESXS @ ª by Lemma 5.3.1.1). Moreover, one infers from lim P ˆSXn  XS A ε 0 and Lemma 5.3.1.2) that n ª

lim EˆSXn  XS1 ˆSXn  XS A 卍 0 n ª

A ˜  for all ε 0. (It is an easy exercise to verify that the uniform integrability of Xn n>N implies ˜   the uniform integrability of Xn X n>N.) Conclusively it follows

lim EˆSXn  XS lim EˆSXn  XS1 ˆSXn  XS A 卍  EˆSXn  XS1 ˆSXn  XS B 卍 B ε n ª n ª

L1 for all ε A 0, i.e. Xn ÐÐÐ X. n ª

2) Now let ESXn  XS ÐÐÐ 0. The properties 1 and 2 of Lemma 5.3.1 have to be shown. n ª

L1 1. sup ESXnS B sup ESXn  XS  ESXS @ ª, since Xn Ð X. n n

2. For all A ` F, PˆA B δ:

ε EˆSXnS1ˆA B EˆSXn  XS 1ˆA   EˆSXS1ˆA B ESXn  XS  ε ² ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ 2 B1 @ ε 2

with an appropriate choice of δ, since ESXS @ ª and since for all ε A 0 §N, such that for A S  S @ ε all n N E Xn X 2 . 74 5 Martingales

5.4 Stopped Martingales

Notation: x ˆx maxˆx, 0, x > R. Theorem 5.4.1 (Doob’s inequality): Let X ˜Xˆt, t C 0 be a càdlàg process, adapted w.r.t. the filtration ˜F, t C 0. Let X be a submartingale. Then for arbitary t A 0 and arbitrary x A 0 it holds:

EˆXˆt P Œ sup Xˆx A x‘ B 0BsBt x Proof W.l.o.g. assume Xˆt C 0, t C 0 a.s.. ˆ ˆ  A  ˆ ˆˆ ˆ  A  C A ˜ ˆ  A  P sup0BsBt X s x P sup0BsBt X s x , for all t 0, x 0. A supt1,...,tn X s x , B @ @ @ B 8n 0 t1 t2 ... tn t – arbitrary times. A k 1Ak,

A1 ˜Xˆt1 A x A2 ˜Xˆt2 B x, Xˆt2 A x ¦

Ak ˜Xˆt1 B x, Xˆt1 B x, . . . , Xˆtk1 B x, Xˆtk A x, k 2, . . . , n, Ai 9 Aj g, i x j. ˆ ˆ  ˆ  B E X tn It has to be shown that P A x . ˆ ˆ  C ˆ ˆ  ˆ  n ˆ ˆ  ˆ  C n ˆ  ˆ   E X tn E X tn 1 A Pk 1 E X tn 1 Ak x Pk 1 P Ak xP A , k 1, . . . , n 1, since X is a martingale and thus follows that EˆXˆtn1ˆAk C EˆXˆtk1ˆAk C Eˆx1ˆAk xPˆAk, k 1, . . . , n  1, tn A tk. Let B ` 0, t¥ be a finite subset, 0 > B, t > B ¨ it is proven similarly that Pˆmaxs>B Xˆs A  B EXˆt x x . ¨  9 8 ˜  8ª `  9 8 ˜  ` @ Q is dense in R 0, t Q t k 1Bk, Bk 0, t Q t finite, Bk Bn, k n. By the monotonicity of the probability measure it holds: EXˆt lim P ‹max Xˆs C x P ‹8n˜max Xˆs A x P Œ sup Xˆs A x‘ B n ª > > s B s Bn s>8nBn x ˆ ˆ  A  B EXˆt By the right-continuity of the paths of X it holds P sup0BsBt X s x x . Corollary 5.4.1 For the Wiener process W ˜W ˆt, t C 0 we are looking at the Wiener process with negative drift: Y ˆt W ˆt  µt, µ A 0, t C 0. From example nr.3 of section 5.3 Xˆt exp˜uˆY ˆt    u2t  C tµ 2 , t 0 is a martingale w.r.t. the natural filtration of W . For u 2µ it holds Xˆt exp˜2µY ˆt, t C 0. ˜ ˆ  A  ™ 2µY ˆs A 2µxž B Ee2µY ˆt 2µx A P sup0BsBt Y s x P sup0BsBt e e e2µx e , x 0 and consequently  P ˆsup Y ˆs A x lim P ˆ sup Y ˆs A x B e 2µx. sC0 t ª 0BsBt Theorem 5.4.2 Let X ˜Xˆt, t C 0 be a martingale w.r.t. the filtration ˜Ft, t C 0 with càdlàg paths. If τ ¢ Ω 0, ª is a finite stopping time w.r.t. the filtration ˜Ft, t C 0, then, the stochastic process ˜Xτ,tˆt C 0 is called a stopped martingale. Here a , b min˜a, b. 5 Martingales 75

Lemma 5.4.1 Let X ˜Xˆt, t C 0 be a martingale with càdlàg-trajectories w.r.t. the filtration ˜Ft, t C 0. ˜  Let τ be a finite stopping time and let τn n>N be the sequence of discrete stopping times out £ ª ˜ ˆ ,  of Theorem 5.1.2, for which τn τ, n , holds. Then X τn t n>N is uniformly integrable for every t C 0.

Proof

œ 0 , if τ 0 τn k1 k @ B k1 > 2n , if 2n τ 2n , for a k N0

1. It is to be shown: ESXˆτn , tS @ ª for all n. S ˆ , S B S ˆ k S S ˆ S @ ª E X τn t P ¢ k @ E X n E X t , since X is a martingale, therefore integrable. k 2n t 2

2. It is to be shown: sup EˆSXˆτn , tS1ˆSXˆτn , tS A x ÐÐÐ 0. n ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ x ª An

sup EˆSXˆτn , tS1ˆAn n ’ “ k k – Q E ‹VX ‹ V 1 ‹›τ 9 A   E ˆSXˆtS 1 ˆτ A t 1 ˆA — sup n n n n n n n ” ¢ k @ 2 2 • k 2n t ’ “ subm. k B – Q E ‹SXˆtS 1 ‹›τ 9 A   E ˆSXˆtS 1 ˆ˜τ A t 9 A — sup n n n n n n ” ¢ k @ 2 • k 2n t ’ ’ “ “ k –E –SXˆtS Q 1 ‹›τ 9 A —  E ˆSXˆtS 1 ˆ˜τ A t 9 A — sup n n n n n n ” ” ¢ k @ 2 • • k 2n t sup E ˆSXˆtS 1 ˆAn B sup E ˆSXˆtS 1 ˆY A x n n E ˆSXˆtS 1 ˆY A x ,

where 1ˆAn B 1ˆsup SXˆτn , tS A x. It remains to prove that PˆY A x ÐÐÐ 0. (The ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶n x ª Y latter obviously implies lim E ˆSXˆtS 1 ˆY A x 0, since ESXˆtS @ ª for all t C 0.) x ª Doob’s inequality yields

ESXˆtS PˆY A x B Pˆ SXˆsS A x B ÐÐÐ . sup ª 0 0BsBt x x

Proof of Theorem 5.4.2 It is to be shown that ˜Xˆτ , t, t C 0 is a martingale.

a.s. 1. ESXˆτ ,tS @ ª for all t C 0. As in conclusion 5.1.1 τn £ τ, n ª ¨ Xˆτn,t ÐÐÐ Xˆτ ,t n ª is approximated, but since ESXˆτn , tS @ ª for all n it follows ESXˆτ , tS @ ª because of Lemma 5.4.1, since uniform integrability gives L1-convergence. 76 5 Martingales

2. Martingale property It is to be shown: a.s. EˆXˆτ , t S Fs Xˆτ , s, s B t  a.s. EˆXˆτ , t1ˆA EˆXˆτ , s1ˆA,A > Fs

First of all, we show that EˆSXˆτn , tS1ˆA EˆSXˆτn , sS1ˆA, A > Fs, n > N. Let t1, . . . , tk > ˆs, t be discrete values, which τn takes with positive probability in ˆs, t. ˆ ˆ ,  S  ˆ ˆ ˆ ,  S  S  E X τn t Fs E E X τn t Ftk Fs EˆEˆXˆτn , t 1ˆτn B tk S Ft  S Fs ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ k Xˆtk

EˆEˆXˆτn , t 1ˆτn A tk S Ft  S Fs ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ k Xˆt ˆ ˆ  ˆ B  S   ˆ ˆ A  ˆ ˆ  S  S  E X tk 1 τn tk Fs E 1 τn tk E X t Ftk Fs EˆXˆtk , τn S Fs ... EˆXˆtk1 , τn S Fs ...

EˆXˆt1 , τn S Fs ... EˆXˆτn , s S Fs a.s. Xˆτn , s

a.s. Since X is càdlàg and τn £ τ, n ª, it holds Xˆτn , t ÐÐÐ Xˆτn , t. Furthermore n ª ˜ ˆ ,  1 X τn t n>N is uniformly integrable because of L -convergence. Therefore follows that

EˆXˆτn , t1ˆA EˆXˆτn , s1ˆA for all A > Fs £ £ EˆXˆτ , t1ˆA EˆXˆτ , s1ˆA

¨ ˜Xˆτ , t, t C 0 is a martingale.

Definition 5.4.1 Let τ ¢ Ω R be a stopping time w.r.t. the filtration ˜Ft, t C 0, Ft ` F, t C 0. The „stopped“ σ-algebra Fτ is defined by A > Fτ ¨ A 9 ˜τ B t > Ft for all t C 0. a.s. Lemma 5.4.2 1. Let η, τ be stopping times w.r.t. the filtration ˜Ft, t C 0, η B τ. Then it holds Fη ` Fτ .

2. Let X ˜Xˆt, t C 0 be a martingale with càdlàg-trajectories w.r.t. the filtration ˜Ft, t C 0 and let τ be a stopping time w.r.t. ˜Ft, t C 0. Then Xˆτ is Fτ -measurable.

Proof 1. A > Fs ¨ A 9 ˜η B t > Ft, t C 0. A 9 ˜τ B t A 9 ˜η B t 9 ˜τ B t > Ft for all t C 0 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¶ >Ft >Ft ¨ A > Fτ .

2. Xˆτ g X f, f ¢ Ω Ω  R, fˆω ˆω, τˆω, g ¢ Ω  R R, gˆω, s Xˆs, ω. S   S ¨ X S It has to be shown: f-F F BR -measurable, g-F BR Fτ -measurable g f-F Fτ - measurable. 5 Martingales 77

S  f-F F BR -measurable is obvious, since τ is a random variable. If we’re looking at the restriction of X ˜Xˆs, s C 0 on s > 0, t¥, t C 0. It has to be shown: ˜Xˆτ > B 9 ˜τ B t > Ft for all t C 0, B > BˆR. ¨ ˆ  ˆ  ˆ   P2n ˆ k  ˆ k1 @ B k  ¨ X – càdlàg X s, ω X 0, ω 1 s 0 limn ª k 1 X t 2n , ω 1 2n t s 2n t Xˆs, ω is B 0,t¥  Ft-measurable ¨ Xˆτ is F S Fτ -measurable.

Theorem 5.4.3 (Optional sampling theorem): Let X ˜Xˆt, t C 0 be a martingale with càdlàg-trajectories w.r.t. a filtration ˜Ft, t C 0, a.s. and let τ be a finite stopping time w.r.t. ˜Ft, t C 0. Then EˆXˆt S Fτ  Xˆτ , t, t A 0.

ˆ ˆ  S  a.s. ˆ ,  C > £ ª Proof First of all we show that E X t Fτn X τn t , t 0, n N, where τn τ, n , is the discrete approximation of τ, cf. Theorem 5.1.2. Let t1 B t2 B ... B tk t be the values, , > attained by τn t with positive probability. It is to be shown that for all A Fτn it holds: EˆXˆt1ˆA EˆXˆτn , t1ˆA.

k1 tk t ˆXˆt  Xˆτn , t1ˆA QˆXˆtk  Xˆti1ˆ˜τn , t ti 9 A i 1 k QˆXˆtk  Xˆti11ˆ˜τn , t ti1 9 A i 2 k k Q QˆXˆtj  Xˆtj11ˆ˜τn , t ti1 9 A i 2 j i k j Q QˆXˆtj  Xˆtj11ˆ˜τn , t ti1 9 A j 2 i 2 k QˆXˆtj  Xˆtj11ˆ˜τn , t B tj1 9 A j 2 k E ˆXˆt  Xˆτn , t1ˆA¥ Q E ˆXˆtj  Xˆtj11ˆ˜τn , t B tj1 9 A¥ j 2 k Q ˆ ˆ   ˆ  ˆ˜ , B  9 ¥ S ¥ E E X tj X tj1 1 τn t tj1 A Ftj1 j 2 ´¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¶ ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ > > Ftj1 Ftj1 k Q ˆ˜ , B  9 ˆ ˆ ˆ  S ¥  ˆ ¥ E 1 τn t tj1 A E X tj Ftj1 X tj1 j 2 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ Xˆtj1 0

ˆ ˆ  S  a.s. ˆ ,  by Definition 5.4.1 and martingale property. Hence it holds E X t Fτn X τn t , since ˆ  B ¦ b X τn is Fτn -measurable. Since τ τn, n, by Lemma 5.4.2, Fτ Fτn . Then

ˆ ˆ  S  a.s. ˆ ˆ  S  ˆ ,  a.s. ˆ ,  ˆ ª E X t Fτ E X t Fτn X τn t X τ t , n , since X is càdlàg. 78 5 Martingales

Corollary 5.4.2 Let X ˜Xˆt, t C 0 be a càdlàg-martingale and let η, τ be finite stopping times, such a.s. that Pˆη B τ 1. Then it holds EˆXˆτ , t S Fη EˆXˆη , t, t C 0. In particular EˆXˆτ , t EˆXˆ0 holds.

Proof Since X is a martingale, by Theorem 5.4.2 ˜Xˆτ ,t, t A 0 is also a martingale. Apply Theorem 5.4.3 to it:

a.s. a.s. EˆXˆτ , t S Fη Xˆτ , η , t Xˆη , t,

a.s. since η B τ. Set η 0, then EˆEˆXˆτ , t S F0 EXˆ0 , t EXˆ0.

5.5 Lévy processes and Martingales

Theorem 5.5.1 Let X ˜Xˆt, t C 0 be a Lévy process with characteristics ˆa, b, ν.

1. There exists a càdlàg-modification of X˜ ˜X˜ˆt, t C 0 of X with the same characteristics ˆa, b, ν.

2. The natural filtration of càdlàg-Lévy processes is right-continuous.

Without proof Theorem 5.5.2 (Regeneration theorem for Lévy processes): ˜ ˆ  A  ˜ X C  Let X X t , t 0 be a càdlàg-Lévy process with natural filtration Ft , t 0 and let ˜ X C  ˜ ˆ  C  τ be a finite stopping time w.r.t. Ft , t 0 . The process Y Y t , t 0 , given by ˆ  ˆ    ˆ  C ˜ X C  Y t X τ t X τ , t 0, is also a Lévy process, adapted w.r.t. the filtration Fτt, t 0 , X which is independent of Fτ and has the same characteristics as X. τ is called regeneration time.

Abb. 5.1: 5 Martingales 79

Proof For any n > N, take arbitrary 0 B t0 @ @ tn and u1, . . . , un > R. We state that all assertions of the theorem follow from the relation ¢ £ ¢ £ ’ ¨ n ¨“ ’ ¨ n ¨“ E 1ˆA ¦Q iu ˆY ˆt   Y ˆt  § PˆAE ¦Q iu ˆXˆt   Xˆt  § ” exp ¨ j j j 1 ¨• ”exp ¨ j j j 1 ¨• (5.5.1) ¤¨j 1 ¥¨ ¤¨j 1 ¥¨ > X for all A Fτ . 1. By Lemma 5.4.2, since τ and τ  t, t C 0 are stopping times with τ B τ  t a.s., we have X b X ˜ X  ˆ  ˆ    ˆ  ˜ X  Fτs Fτt, i.e. Fτt tC0 is a filtration, and Y t X τ t X τ is Fτt -adapted: ˆ  ˜ X  ˆ   ˜ X  X b X X τ is Fτ -measurable, X τ t is Fτt -measurable, Fτ Fτt. d 2. It follows from (5.5.1) for A Ω that X Y , i.e., Y is a Lévy process with the same Lévy exponent η as X. X 3. It also follows from (5.5.1) that Y and Fτ are independent, since arbitrary increments X of Y do not depend on Fτ . 4. Now let us prove (5.5.1). We begin with the case of

a.) § c A 0: Pˆτ B c 1. By example 5, b) of section 5.2., Y˜j ˜Y˜jˆttC0, j 1, . . . , n with Y˜jˆt exp ˜iujXˆt  tηˆuj, t C 0 are complex-valued martingales. Furthermore, it holds n Eˆ1ˆA exp˜Q iujˆY ˆtj  Y ˆtj1 j 1 n Eˆ1ˆA exp˜Q iujˆXˆτ  tj  Xˆτ  Xˆτ  tj1  Xˆτ j 1 ’ n Y˜ ˆτ  t  exp˜ηˆu ˆτ  t  “ E 1ˆA M j j j j ” ˜ ˆ ˆ  • j 1 Y˜jˆτ  tj1 exp η uj τ tj1 ’ ’ n ˜ ˆ   ““ Yj τ tj X E E 1ˆA M exp˜ˆt  t  ηˆu  S F  ” ” j j 1 j τ tj1 •• j 1 Y˜jˆτ  tj1  ˆ   ˆ  ’ n 1 Y˜ ˆτ  t  tn tn1 η un “ j j ˆtj tj1ηˆuj  e X E 1ˆA M e EˆY˜ ˆτ  t  S F   ” n n τ tn1 • j 1 Y˜jˆτ  tj1 Y˜nˆτ  tn1 < = ’ @n1 ˜ ˆ   A “ @ Yj τ tj ˆt t  ηˆu A ˆt t  ηˆu  E 1ˆA @M e j j 1 j A e n n 1 n ” @ ˜ ˆ   A • >j 1 Yj τ tj1 ? n n ˆ   ˆ  ˆ   ˆ  ... Eˆ1ˆA M e tj tj1 η uj  PˆA M e tj tj1 η uj j 1 j 1 n PˆAEˆexp˜i QˆujˆXˆtj  Xˆtj1 j 1

b.) Prove (5.5.1) for arbitrary finite stopping times τ: Pˆτ @ ª 1. For any k > N it 9 ˜ B  > X > X holds Ak A τ k Fτ,k, if A Fτ . Then it follows from 4. a.) and (5.5.1) that ¢ £ ¢ £ ’ ¨ n ¨“ ’ ¨ n ¨“ ˆ  ¦ ˆ  § ˆ  ¦ ˆ ˆ   ˆ § E 1 A exp Q iu Y Y  P A E exp Q iu X t X t  ” k ¨ j k,tj k,tj 1 ¨• k ” ¨ j j j 1 ¨• ¤¨j 1 ¥¨ ¤¨j 1 ¥¨ (5.5.2) 80 5 Martingales

for Yk,t Xˆˆτ , k  t  Xˆτ , k. Now let n ª on both sides of (5.5.2). By Lebesgue’s theorem on majorized convergence, (5.5.1) follows for any a.s. finite τ, a.s. since τ , k τ, ˆn ª.

5.6 Martingales and Wiener Process

Our goal: We would like to show that if W ˜W ˆt, t C 0 is a Wiener process, then it holds ¾ ª 2 2  y Pˆ max W ˆs A x S e 2t dy, x C 0. s> 0,t¥ πt x

Theorem 5.6.1 (Reflection principle): ˜ W C  Let τ be an arbitrary finite stopping time w.r.t. the natural filtration Ft , t 0 . Let X ˜Xˆt, t C 0 be the reflected Wiener process at time τ, i.e. Xˆt W ˆτ ,tˆW ˆtW ˆτ ,t, d t C 0. Then X W holds.

Abb. 5.2: Reflection principle.

Proof It holds ¢ ¨W ˆt , t @ τ ˆ  ˆ ,   ˆ ˆ   ˆ ,  ¦ X t W τ t W t W τ t ¨ ¤¨2W ˆτ  W ˆt , t C τ

Let X1ˆt W ˆτ , t, X2ˆt W ˆτ  t  W ˆτ, t C 0. From Theorem 5.5.2 it follows that X2 is independent of ˆτ, X1 (W – Lévy process and τ – regeneration time). It holds W ˆt X1ˆt  X2ˆˆt  τ, Xˆt X1ˆt  X2ˆˆt  τ, t C 0. Indeed, ¢ ¨W ˆt  X ˆ  W ˆt , t @ τ ˆ   ˆˆ    ¦ 2 0 X1 t X2 t τ  ¨ ¤¨W ˆτ  W ˆτ  t  τ  W ˆτ W ˆt , t C τ 5 Martingales 81

¨ W ˆt X1ˆt  X2ˆˆt  τ, t C 0. Furthermore ¢ ¨W ˆt  X ˆ  W ˆt , t @ τ ˆ   ˆˆ    ¦ 2 0 X1 t X2 t τ  ¨ ¤¨W ˆτ  W ˆt  W ˆτ 2W ˆτ  W ˆt , t C τ

d ¨ Xˆt X1ˆt  X2ˆˆt  τ, t C 0. From theorems 5.5.2 and 3.3.2 it follows that X2 W , d W W and hence d ˆτ1,X1,X2 ˆτ, X1, X2 + + W d X

˜ ˆ  C  ˆ  ˜ W C  Let W W t , t 0 be the Wiener process on Ω, F, P , let Ft , t 0 be the natural > W ˜ C ¢ ˆ   W ¢ W filtration w.r.t. W . For z R let τ˜z inf t 0 W t z . τ˜z τz is an a.s. finite ˜ W C  A ˜ W B  > W stopping time w.r.t. Ft , t 0 , z 0, since it obviously holds Fz t Ft . Since W has ˜ W C  continuous paths (a.s.), Ft , t 0 is right-continuous. Corollary 5.6.1 Let Mt maxs> 0,t¥ W ˆs, t C 0. Then it follows for all z A 0, y C 0, that PˆMt C z, W ˆt B z  y PˆW ˆt A y  z. ¢ W Proof Mt is a random variable, since W has continuous paths. Let τ τz . By Theorem 5.6.1, ˆ  ˆ , ˆ ˆ  ˆ ,  C d ˜ W  d ˜ Y  it holds for Y t W τ t W t W τ t , t 0, that Y W . Hence, τz ,W τz ,Y , ˆ  W Y since W τ z, τz τz . Therefore ˆ B ˆ  @   ˆ Y B ˆ  @   P τ t, W t z y P τz t, Y t z y , ˜ Y B  9 ˜ ˆ  @   ˜ Y B  9 ˜  ˆ  @   Y B whereas τz t Y t z y τz t 2z W t z y . If τ τz t then Y ˆt 2W ˆτ  W ˆt 2z  W ˆt, and hence

Pˆτ B t, W ˆt @ z  y Pˆτ B t, 2z  W ˆt @ z  y Pˆτ B t, W ˆt A z  y PˆW ˆt A z  y.

W Per definition of τ τz it holds:

Pˆτ B t, W ˆt @ z  y PˆMt C z, W ˆt @ z  y PˆW ˆt A y  z,

˜ W B  ¨ ˜ ˆ  C  since τz t maxs> 0,t¥ W s z . Theorem 5.6.2 (Distribution of the maximum of W ): For t A 0 and x C 0 it holds ¾ ª 2 2  y PˆMt A x S e 2t dy πt x

Proof In Corollary 5.6.1, set y 0 ¨ PˆMt C z, W ˆt @ z PˆW ˆt A z. It holds PˆW ˆt A z PˆW ˆt C z for all t and all z, since W ˆt ¢ N ˆ0, t, thus continuously distributed ¨ PˆMt C z, W ˆt @ z  PˆW ˆt C z PˆW ˆt A z  PˆW ˆt A z ¨ PˆMt C z, W ˆt @ z  PˆMt C z, W ˆt C z PˆMt C z 2PˆW ˆt A z 2 ¼ 2 1 ª  y 2 ª  y ¨ PˆM A z PˆW ˆt A z º R e 2z dy R e 2t dy t 2 2 2πt z πt z 82 5 Martingales

Let Xˆt W ˆt  tµ, t C 0, µ A 0, be the Wiener process with negative drift. Consider ˆ ˆ  A  C P suptC0 X t x , x 0.

Motivation Calculation of the following ruin probability in risk theory.

Assumptions Initial capital x C 0. Let µ be the volume of premiums per time unit.¨ µt – earned premiums at time t C 0. Let W ˆt be the loss process (price development). ¨ Y ˆt x  tµ  W ˆt – remaining capital at time t. The ruin probability is

Pˆinf Y ˆt @ 0 Pˆx  sup Xˆt @ 0 Pˆsup Xˆt A x tC0 tC0 tC0

Theorem 5.6.3 It holds  Pˆsup Xˆt A x e 2µx, x C 0, µ A 0. tC0

X ˜ C ¢ ˆ   Proof Let τ τx inf t 0 X t x . It holds

Pˆsup Xˆt A x Pˆτ @ ª lim Pˆτ @ t. tC0 t ª

Compute this limit. For that, introduce the process Y ˜Y ˆt, t C 0,

u2 Y ˆt exp œuXˆt  t Œ  µu‘¡ , t C 0, 2

œ u C 0 fixed. It can be easily shown that Y is a martingale. τ τ , t is a finite stopping time ˜ X C  ˆ œ ˆ  0 w.r.t. filtration Ft , t 0 . By Corollary 5.4.1, EY τ EY 0 e 1. On other hand,

œ œ œ 1 EY ˆτ  EˆY ˆτ 1ˆτ @ t  EˆY ˆτ 1ˆτ C t EˆY ˆτ1ˆτ @ t  EˆY ˆτ œ1ˆτ C t.

If we can show that œ lim EˆY ˆτ 1ˆτ C t 0, (5.6.1) t ª ˆ ˆ  ˆ @  ˆ  š  Š u2  Ÿ then limt ª E Y τ 1 τ t 1. Since Y τ exp ux τ 2 µu , it follows

2 u  lim E ¤exp œτ Œ  uµ‘¡ 1ˆτ @ t© e ux, t ª 2 and for u 2µ it holds

 lim E ¡e01ˆτ @ t¦ lim Pˆτ @ t e 2µx t ª t ª

 ¨ Pˆsup Xˆt A x e 2µx. Now let us prove (5.6.1). By Corollary 3.3.2, it is known that tC0

W ˆt a.s. 0 ˆt ª. t 5 Martingales 83

Hence, Xˆt W ˆt lim lim Œ  µ‘ µ t ª t t ª t a.s. ¨ Xˆt ª, ˆt ª. Then,

œ a.s. Y ˆτ 1ˆτ C t exp ˜2µXˆt 1ˆτ C t 0 ˆt ª

œ By Lebesgue’s theorem, it holds E Y ˆτ 1ˆτ C t¥ 0, ˆt ª. Theorem 5.6.4 Let X ˜Xˆt, t C 0 be a Lévy process and let τ ˜τˆt, t C 0 be a subordinator, which are both defined on a probability space ˆΩ, F, P. Let X and τ be independent. Then Y ˜Y ˆt, t C 0 definded by Y ˆt Xˆτˆt, t C 0, is a Lévy process. Without proof

5.7 Additional Exercises

Exercise 5.7.1 Let X,Y ¢ Ω R be arbitrary random variables on ˆΩ, F, P with ESXS @ ª, ESY S @ ª, ESXY S @ ª, and let G ` F be an arbitrary sub-σ-Algebra of F. Then it holds a.s. that

(a) EˆXS˜g, ٝ EX, EˆXSF X,

(b) EˆaX  bY SG aEˆXSG  bEˆY SG for arbitrary a, b > R, (c) EˆXSG B EˆY SG, if X B Y ,

(d) EˆXY SG Y EˆXSG, if Y is a ˆG, BˆR-measurable random variable,

(e) EˆEˆXSG2SG1 EˆXSG1, if G1 and G2 are sub-σ-algebras of F with G1 ` G2,

1 (f) EˆXSG EX, if the σ-algebra G and σˆX X ˆBˆR are independent, i.e., œ œ œ if PˆA 9 A  PˆAPˆA  for arbitrary A > G and A > σˆX.

(g) EˆfˆXSG C fˆEˆXSG, if f ¢ R R is a convex function such that ESfˆXS @ ª. Exercise 5.7.2 ˆ  ¥ ˆ  ¥ 1  Look at the two random variables X and Y on the probability space 1, 1 , B 1, 1 , 2 ν with ESXS @ ª, where ν is the Lebesgue measure on 1, 1¥. Determine σˆY  and a version of the conditional expectation EˆXSY  for the following random variables.

(a) Y ˆω ω5 (Hint: Show first that σˆY  Bˆ 1, 1¥) ˆ  ˆ k > ¡ k3 k2 Ž ˆ  (b) Y ω 1 for ω 2 , 2 , k 1,..., 4 and Y 1 1 ˆ  ˆ S  E X1B > ˆ  ˆ  A (Hint: It holds E X B PˆB for B σ Y with P B 0) (c) Calculate the distribution of EˆXSY  in (a) and (b), if X ¢ U 1, 1¥. 84 5 Martingales

Exercise 5.7.3 Let X and Y be random variables on a probability space ˆΩ, F, P. The conditional variance VarˆY SX is defined by VarˆY SX EˆˆY  EˆY SX2SX.

Show that Var Y EˆVarˆY SX  VarˆEˆY SX. Exercise 5.7.4 Let S and τ be stopping times w.r.t. the filtration ˜Ft, t C 0. Show:

(a) A 9 ˜S B τ > Fτ , A > FS

(b) Fmin˜S,τ FS 9 Fτ

Exercise 5.7.5 (a) Let ˜Xˆt, t C 0 be a martingale. Show that EXˆt EXˆ0 holds for all t C 0.

(b) Let ˜Xˆt, t C 0 be a sub- resp. supermartingale. Show that EXˆt C EXˆ0 (resp., EXˆt B EXˆ0) holds for all t C 0. Exercise 5.7.6 The stochastic process X ˜Xˆt, t C 0 is adapted and càdlàg. Show that

EXˆt2 Pˆ Xˆv A x B sup 2 2 0BvBt x  EXˆt holds for arbitrary x A 0 and t C 0, if X is a submartingale with EXˆt 0 and EXˆt2 @ ª. Exercise 5.7.7 (a) Let g ¢ 0, ª 0, ª be a monotone increasing function with

gˆx ª, x ª. x

Show that the sequence X1,X2,... of random variables is uniformly integrable, if ˆS S @ ª supn>N Eg Xn .

(b) Let X ˜Xˆn, n > N be a martingale. Show that the sequence of random variables Xˆτ , 1,Xˆτ , 2,... is uniformly integrable for every finite stopping time τ, if ESXˆτS @ ª and EˆSXˆnS1˜τAn 0 for n ª. Exercise 5.7.8 ˜  n >  A ˆ  Let S Sn a Pi 1 Xi, n N be a symmetric random walk with a 0 and P Xi 1 PˆXi 1 1~2 for i > N. The random walk is stopped at the time τ, when it exceeds or falls below one of the two values 0 and K A a for the first time, i.e.

τ min˜Sk B 0 or Sk C K. kC0

Pn  1 3 ˆPτ  1 ˆ 2  2  Show that Mn i 0 Si 3 Sn is a martingale and E i 0 Si 3 K a a a holds. ˆ S M  A ˆ l 3 B B Hint: To calculate E Mn Fm , n m, you can use E Pi k Xi 0, 1 k l, Mn Pm  Pn  1 3   r 0 Sr r m1 Sr 3 Sn and Sn Sn Sm Sm. 5 Martingales 85

˜  ˜  A discrete martingale w.r.t. a filtration Fn n>N is a sequence of random variables Xn n>N on ˜  ˜  ˆ S  a probability space Ω, F, P , such that Xn is measurable w.r.t. Fn n>N and E Xn1 Xn Xn > ˜  a.s. for all n N. A discrete stopping time w.r.t. Fn n>N is a random variable ¢ 8 ˜ª ˜ B  > > 8 ˜ª ˜ ª  τ Ω N such that τ n Fn for all n N , where Fª σ n 1 Fn . Exercise 5.7.9 ˜  ˜  Let Xn n>N be a discrete martingale and τ a discrete stopping time w.r.t. Fn n>N. Show ˜  ˜  that Xτ,n n>N is also a martingale w.r.t. Fn n>N. Exercise 5.7.10 ˜  Pn Let Sn n>N be a symmetric random walk with Sn i 1 Xi for a sequence of independent and ˆ  ˆ   1 identically distributedº random variables X1,X2,..., such that P X1 1 P X1 1 2 . Let τ inf˜n ¢ SSnS A n and Fn σ˜X1,...,Xn, n > N. ˜  (a) Show that τ is a stopping time w.r.t. Fn n>N. ˜  2  , ˜  (b) Show that Gn n>N with Gn Sτ,n τ n is a martingale w.r.t. Fn n>N. (Hint: Use Exercise 5.7.9)

(c) Show that SGnS B 4τ holds for all n > N. S S B S 2 S  S , S B 2  (Hint: It holds Gn Sτ,n τ n Sτ,n τ) Exercise 5.7.11 Let X1,X2,... be a sequence of independent and identically distributed random variables with S S @ ª ˜  > ˜  E X1 . Let Fn σ X1,...,Xn , n N, and let τ be a stopping time w.r.t. Fn n>N with Eτ @ ª.

(a) Let τ be independent of X1,X2,.... Derive a formula for the characteristic function of τ Sτ Pi 1 Xi and verify Wald’s identity ESτ EτEX1.

(b) Let additionally EX1 0 and τ inf˜n ¢ Sn @ 0. Use Theorem 2.1.3, to show that Eτ ª. (Hint: Proof by contradiction) 6 Stationary Sequences of Random Variables

6.1 Sequences of Independent Random Variables

It is known that the series ª 1 @ ª ¨ A Pn 1 α α 1, n n Pª ˆ1 @ ª ¨ A n 1 nα α 0, 1 since the drift of neighboring terms in the second series has order n1α , i.e. ª ª ˆ1n 1 1 Q   Q Œ  ‘ , α 1 ˆ α ˆ  α n 1 n k 1 2k 2k 1 whereas 1 α α 1 ˆ  α  ˆ α ‰1  Ž  1 ª 1   1  o ‰ Ž 1  1 2k 1 2k 2k k 2k 2k ˆ2kα ˆ2k  1α ˆ2kαˆ2k  1α ˆ2k  1α ˆ2k  1α α  oˆ1 1 1 O Œ ‘ O ‹  , n 2k. 2kˆ2k  1α ˆ2kα1 nα1 ª A P δn For which α 0 does the series n 1 nα converge, where δn are i.i.d. random variables with ˆ   1 Eδn 0, e.g. P δn 1 2 ? ª @ ª More general problem: Under which conditions holds Pn 1 Xn a.s., where Xn are inde- pendent random variables? a.s. It is known that for a sequence of random variables ˜Yn with Yn ÐÐÐ Y it holds that n ª P Yn ÐÐÐ Y . The opposite is in general not true. n ª Theorem 6.1.1 n P a.s. Let Xn, n > be independent random variables. If Sn P Xi ÐÐÐ S then Sn ÐÐÐ S. N i 1 n ª n ª Without proof Corollary 6.1.1 If the Xn, n > N are independent random variables with Var Xn @ ª, EXn 0, n > N and ª @ ª ª Pn 1 Var Xn , then Pn 1 Xn converges a.s. Pn > ˜  2ˆ  Proof Let Sn i 1 Xi for n N. Prove that Sn n>N is a Cauchy sequence in L Ω, F, P . Namely, for n A m it holds n 2 2 EˆSn  Sm YSn  SmY 2 Q Var Xi ÐÐÐÐ 0, L n,m ª i m1 Pª @ ª ˜  2ˆ  since i 1 Var Xi . Hence, Sn n>N is a Cauchy sequence in L Ω, F, P . Then ª §S lim Sn Q Xi n ª i 1

86 6 Stationary Sequences of Random Variables 87

P in the mean square sense and thus Sn ÐÐÐ S. The statement of the corollary then follows n ª from Theorem 6.1.1.

Corollary 6.1.2 Pª 2 @ ª ˜  ˜  If n 1 an , where an n>N is a deterministic sequence, and δn is a sequence of i.i.d. 2 @ ª > ª random variables with Eδn 0, Var δn σ , n N, then the series Pn 1 anδn converges a.s. Exercise 6.1.1 Derive Corollary 6.1.2 from Corollary 6.1.1. 2 A ˆ   1 1 In our motivating example δn i.i.d., Eδn 0, Var δn σ 0 (e.g. P δn 1 2 ), an nα , > Pª 1 @ ª Pª 1 @ ª @ 1 n N. n 1 nα , if n 1 n2α , i.e. for α 2 . Corollary 6.1.3 (Three-Series-Theorem): ˜  Pª Pª @ ª Let Xn n>N be a sequence of independent random variables with n 1 EXn, n 1 Var Xn . ª a.s.@ ª Then Pn 1 Xn .

ª Proof Let Y X EX , thus X EX Y , n > , and EY 0, P a @ ª by our assump- n n n n ±n n N n n 1 n an ª a.s.@ ª > ª @ ª ¨ tions. Then Pn 1 Yn by Corollary 6.1.1, since Var Xn Var Yn, n N, Pn 1 Var Xn ª ª  ª a.s.@ ª Pn 1 Xn Pn 1 an Pn 1 Yn .

6.2 Stationarity in the Narrow Sense and Ergodic Theory

6.2.1 Basic Ideas ˜  Let Xn n>N be a stationary in the narrow sense sequence of random variables, i.e. for all — n, k > N the distribution of ˆXn,...,Xnk is independent of n > N. In particular, this means that all Xn are identically distributed. In the language of Kolmogorov’s theorem:

PˆˆXn,Xn1,... > B PˆˆX1,X2,... > B,

ª ª for all n > N and B > BˆR  where R R  R  ...  .... ˜  Example 6.2.1 (Stationary sequences of random variables): 1. Let Xn n>N be a se- ˜  quence of i.i.d. random variables, then Xn n>N is stationary.   ˜  > 2. Let Yn a0Xn ... akXnk, k – fixed natural number, Xn n>N from 1), a0, . . . , ak R (fixed), n > N. Yn are not independent anymore, but identically distributed. The sequence ˜  Yn n>N is stationary. Pª > ˜  3. Let Yn j 0 ajXnj for arbitrary n N. The sequence aj j>N is a sequence of real ˜ ª ª @ ª numbers and Xn n 1 are independent random variables with EXn 0, Pn 1 Var Xn (compare Corollary 6.1.3). ˜  ˜ ª It is obvious that Yn n>N is a stationary sequence if Xn n 1 is i.i.d. (This construction is important for autoregressive time series (AR-processes), e.g. in econometrics). ˆ  > ¢ ª ˜  ˜  4. Let Yn g Xn,Xn1,... , n N, g R R measurable, Xn n>N from 1). Then Yn n>N is stationary. 88 6 Stationary Sequences of Random Variables

˜  Remark 6.2.1 1. An arbitrary stationary sequence of random variables X Xn n>N can ¯ ˜  be extended to a stationary sequence X Xn n>Z. In fact, the finite dimensional distri- bution of X¯ can be defined as:

d ˆXn,...,Xnk ˆX1,...,Xk1, n > Z, k > N.

Therefore, by Kolmogorov’s theorem, there exists a probability space and a sequence ˜  ¯ ˜  ˜  d Yn n>Z with the above distribution. We set X Yn n>Z and hence follows that Yn n>N ˜  Xn n>N. ª 2. We define a shift of coordinates. Let x > Rª, x ˆxk, k > Z. Define the mapping ª ª θ ¢ Rª Rª, ˆθxk xk1 (shift of the coordinates by 1), k > N, k > Z. If θ is ª 1 considered on Rª, then it is bijective and the inverse mapping would be ˆθ xk xk1, k > Z. Let now X ˜Xn, n > Z be a stationary sequence of random variables. Let X¯ θX,  d d X˜ θ 1X. It is obvious that X¯ and X˜ are again stationary and X¯ X X˜ i.e.,

1 ª Pˆθ X > B PˆθX > B PˆX > B,B > BˆRª.

θ is called a measure preserving map. There are also other maps which have a measure preserving effect. Definition 6.2.1 Let ˆΩ, F, P be an arbitary probability space. A map T ¢ Ω Ω is called measure preserving, if  1. T is measurable, i.e. T 1A > F for all A > F,  2. PˆT 1A PˆA, A > F. Lemma 6.2.1 Let T be a measure preserving mapping and let X0 be a random variable. Define Xnˆω n X0ˆT ˆω, ω > Ω, n > N. Then the sequence of random variables X ˜X0,X1,X2,... is stationary.

Proof Let 2 Xˆω ˆX0ˆω,X0ˆT ˆω,X0ˆT ˆω,..., 2 θXˆω ˆX0ˆT ˆω,X0ˆT ˆω,..., ª B > BˆR , A ˜ω > Ω ¢ Xˆω > B, A1 ˜ω > Ω ¢ θXˆω > B. Therefore ω > A1  T ˆω > A. 1 n Since PˆT A PˆA, it holds PˆA1 PˆA. For An ˜ω > Ω ¢ θ Xˆω > B the same holds, PˆAn PˆA, n > N (Induction).Hence the sequence X is stationary. The sequence X in Lemma 6.2.1 is called the sequence generated by T . Definition 6.2.2 A map T ¢ Ω Ω is called measure preserving in both directions, if

1. T is bijective and T ˆΩ Ω,  2. T and T 1 are measurable,  3. PˆT 1A PˆA, A > F, and therefore PˆTA PˆA. 6 Stationary Sequences of Random Variables 89

Thus, exactly as in Lemma 6.2.1, we can construct stationary sequences of random variables with time parameter n > Z: ˆ  ˜ ˆ nˆ  > X ω X0 T ω n>N, ω Ω, 0 0 where T is a measure preserving map (in both directions), X0ˆT ˆω X0ˆω,(T Id). Lemma 6.2.2 For an arbitrary stationary sequence of random variables X ˆX0,X1,... there exists a mea- ˆ  ˜ ˆ nˆ  sure preserving map T and a random variable Y0, such that Y ω Y0 T ω n>N has the d same distribution as X: X Y . The same statement holds for sequences with time parameter n > Z. ª ª Proof Consider the canonical probability space ˆR , BˆR , PX , Y ˆω ω, ω > R, T θ. ª Thus, Y is constructed since PX ˆA PY ˆA PX ˆY > A, A > BˆR . Ω Example 6.2.2 (Measure preserving maps): 1. Let Ω ˜ω1, . . . , ωk, k C 2, F 2 , ˆ  1  P ωi k , i 1, . . . , k, be a Laplace probability space. T ωi ωi1 for all i 1, . . . , k 1, T ωk ω1.

2. Let Ω 0, 1, F Bˆ 0, 1, P ν1 – Lebesgue-measure on 0, 1. T ω ˆω  s mod 1, s C 0. T is measure preserving in both directions. Sequences of random variables, which in these examples can be generated by the map T , are mostly deterministic resp. cyclic. In example 1) we can consider a random variable X0 ¢ Ω R, n such that Xˆωi xi are all pairwise distinct. Therefore Xnˆω X0ˆT ˆω uniquely defines n1 the value of Xn1ˆω X0ˆT ˆω, for all n > N. Remark 6.2.2 Measure preserving maps play an important role in physics. There, T is interpreted as the change of state of a physical system and the measure P can e.g. be the volume. (Example: T – Change of temperature, measure P – volume of the gas.) Therefore the ergodic theory to be developed can be immediately transfered to some physical processes. Theorem 6.2.1 (Poincaré): If T is a measure preserving map on ˆΩ, F, P, A > F, then for almost all ω > A the relation n ˜T ˆω > A holds for infinitely many n > N. n That means, the trajectory ˜T ˆω, n > N returns to A infinitely often, if ω > A, PˆA A 0. Proof Let N ˜ω > A ¢ T nˆω ¶ A, ¦n C 1. It is obvious that N > F, since ˜ω > Ω ¢ T nˆω ¶   A > F for all n C 1. N 9 T nN g for all n C 1. In fact, if ω > N 9 T nN, ¦ n, then ω > A, n n n T ˆω ¶ A for all n C 1, ω1 T ˆω, ω1 > N. Hence it follows that ω1 > A and T ˆω > A. That is a contradiction with ω > N. It holds n n T N ˜ω > Ω ¢ T ˆω > N. For arbitrary m > N it holds T mN 9 T ˆnmN T mˆN 9 T nN T mˆg g.

n n Hence follows that the sets T N, n > N, are pairwise disjoint, belong to F and PˆT N PˆN a C 0 holds. Then ª ª n n 1 C Pˆ8nT N Q PˆT N Q a n 0 n 0 90 6 Stationary Sequences of Random Variables which is possible only if a 0, i.e., PˆN 0. Hence it follows that for almost all ω > A n1 k ˆω > AƒN there exists a n1 n1ˆω such that T ˆω > A. Now use T instead of T , k > N in the above reasoning. It holds PˆNk 0 and for all ω > AƒNk there exists nk nkˆω such k n n that ˆT  k ˆω > A. Since knk C k it follows for almost all ω > A, that T ˆω > A for infinitely many n. Corollary 6.2.1 Let X C 0 be a random variable, A ˜ω > Ω ¢ Xˆω A 0. Then it holds for almost all ω > A ª ˆ nˆ  ª that Pn 0 X T ω , where T is a measure preserving map. Exercise 6.2.1 Prove it. Remark 6.2.3 The proof of Theorem 6.2.1 holds for the sets A > F ¢ PˆA C 0. If however PˆA 0, it is possible that AƒN g and thus the statement of the theorem is trivial. As an example we consider Ω 0, 1, F B 0,1, P ν1 – Lebesgue-measure, T ˆω ω  s ˆ n mod 1, s ¶ Q. Set A ˜ω0, ω0 > Ω. Then T ˆω0 x ω0 holds for all n, because otherwise >   m > there exist k, m N such that ω0 ks m ω0 and hence follows s k Q. Thus we get a contradiction.

6.2.2 Mixing Properties and Ergodicity Here we study the dependence structure in a stationary sequence of random variables, which is generated by a measure preserving map T . X ˜X  Let n n>N be a stationary sequence (in the narrow sense) of random variables. Then d n d there exists a measure preserving map T ¢ Ω Ω, such that Xnˆω X0ˆT ˆω and Xn X0, and thus X0 gives the marginal distribution of the sequence X. In turn, the map T is responsible for the dependence within X (it indicates the properties of multidimensional distributions). We shall therefore now examine the depedence properties of X generated by T . Definition 6.2.3 1. Event A > F is called invariant w.r.t. (a measure preserving map)  T ¢ Ω Ω, if T 1A A.  2. Event A > F is called almost invariant w.r.t. T , if PˆT 1A Q A 0. Here Q is the symmetric difference of sets. Exercise 6.2.2 ‡ Show that the set of all (almost) invariant events T is a σ-algebra JˆJ . Lemma 6.2.3 ‡ Let A > J . Then there exists B > J such that PˆA Q B 0 n 9ª 8ª k > ˆ Q  Proof Let B lim supn ª T A n 1 k n T A. It is to be shown that B J, P A B 0. 1ˆ  ˆn1 > It is obvious that T B lim supn ª T A B and hence B J. Q ` 8ª ˆ k Q ˆk1  ˆ k Q ˆk1  It is easy to see that A B k 0 T A T A . Since P T A T A 0 for all ‡ k C 1 due to A > J , it follows that PˆA Q B 0.

Definition 6.2.4 1. The measure preserving map T ¢ Ω Ω is called ergodic if for every A > J 0 PˆA œ . 1 6 Stationary Sequences of Random Variables 91

X ˜X  ergodic 2. The stationary sequence of random variables n n>N is called if the measure preserving map T ¢ Ω Ω, which generates X, is ergodic. Lemma 6.2.4 The measure preserving map T is ergodic if and only if the probability of almost invariant sets

0 ‡ PˆA œ for all A > J . 1 Proof „ “ ‡ Obvious, since arbitrary invariant sets are also alsmost invariant, i.e. J ` J „ ¨ “ ‡ Let T be ergodic and A > J . According to Lemma 6.2.3 it follows that there exists B > J such that PˆAQB 0. Therefore P ˆA P ˆA9B P ˆB. Since T is ergodic and B > J it follows 0 PˆA PˆB œ . 1

Definition 6.2.5 A random variable Y ¢ Ω R is called (almost) invariant w.r.t. a measure preserving map T ¢ Ω Ω if Y ˆω Y ˆT ˆω for (almost) all ω > Ω. Theorem 6.2.2 Let T ¢ Ω R be a measure preserving map. The following statements are equivalent: 1. T – ergodic 2. If Y is almost invariant w.r.t. T then Y const a.s. 3. If Y is invariant w.r.t. T then Y const a.s. Proof 1 ¨ 2 ¨ 3 ¨ 1 1 ¨ 2 Let T – ergodic, Y – almost invariant. It is to be shown that Y ˆω const for almost all ω > Ω. Y ˆT ˆω Y ˆω almost surely. Let Av ˜ω > Ω ¢ Y ˆω B v, v > R. Hence it follows that ‡ Av > J for all v > R and by Lemma 6.2.4 0 PˆA  œ for all v. v 1

Let c sup ˜v ¢ PˆAv 0. Show that PˆY c 1. It holds Av ¡ Ω, v ª, Av £ g, v ª ¨ ScS @ ª. Thus ª ˆ @  ‹8ª › B  1  B Q Š  P Y c P n 1 Y c P Ac 1 0, n n 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶n 0 by definition of c. Analogously one proves that PˆY A c 0 and hence PˆY c 1. 2 ¨ 3 is obvious. 0 3 ¨ 1 It is to be shown that T is ergodic, i.e. PˆA œ for all A > J. 1 0 0 Let Y 1 . It is invariant w.r.t. T , hence it follows that 1 const œ and PˆA œ . A A 1 1 92 6 Stationary Sequences of Random Variables

Remark 6.2.4 1. The statements of Theorem 6.2.2 stay true if you demand 3) for a.s. bounded random variables Y .

2. If Y is invariant w.r.t. T then Yn min ˜Y, n, n > N, is also invariant w.r.t. T . ˜  Ω ˆ˜  1 ˆ  Example 6.2.3 1. Let Ω ω1, . . . , ωd , F 2 , P ωi d , i 1, . . . , d. Let T ωi ωi1 T i 1, . . . , d1, ωd z ω1. T is obviously ergodic and hence any invariant random variable is constant.

2. Let Ω 0, 1, F B 0,1, P ν1, T ˆω ˆω  s mod 1. Show that T is ergodic ¨ s ¶ Q.

Proof „ “ 2 Let s ¶ Q, Y be an arbitrary invariant random variable. Let Y be bounded a.s. so that EY @ ª (compare Remark 6.2.4, 1)). We decompose the random variable Y into a Fourier-series. The ˆ  ª 2πinw A Fourier series of Y is Y ω Pn 0 ane . We want to show that an 0, n 0, and hence follows that Y ˆω a0 a.s.. Then T is ergodic by Theorem 6.2.2. Indeed, by definition of T , T ˆω ω  s  k, k > N. Since T is measure preserving and since Y is invariant w.r.t. T it holds

2πinw 2πinw 2πinw 2πins 2πins an @ Y ˆω, e AL2 EˆY ˆωe  EˆY ˆT ˆωe e e an.

Therefore if s ¶ Q then an 0. „ ¨“ m > > @ ˆ  @ If s n Q then T is not ergodic, i.e., there exists A J such that 0 P A 1. Indeed, set 8n1 ™ > ¢ 2k B @ 2k1 ž ˆ  1 ˆ  ‰  2m Ž A k 0 ω Ω 2n ω 2n . It is clear that P A 2 . A is invariant, since T A A 2n mod 1 A.

Definition 6.2.6 1. The measure preserving map T ¢ Ω Ω is called mixing (on average), if n 1 n k for all A1,A2 > F it holds: PˆA1 9T A2 ÐÐÐ PˆA1PˆA2 ( P PˆA1 9T A2 ÐÐÐ n ª n k 1 n ª PˆA1PˆA2, respectively), i.e., by repeated application of T to A2, A1 and A2 are getting asymptotically independent.

X ˜X  2. Let n n>N0 be a stationary sequence of random variables which are generated by a random variable X0 and a measure preserving map T . X is called weakly dependent (on average) if the random variables Xk and Xkn are getting asymptotically independent for ª > > n , i.e. for all B1,B2 BR, and k N0

PˆXk > B1,Xkn > B2 ÐÐÐ PˆX0 > B1PˆX0 > B2 n ª 1 n Œ Q PˆX0 > B1,Xk > B2 ÐÐÐ PˆX0 > B1PˆX0 > B2, respectively n ª n k 1

Theorem 6.2.3 X ˜X  A stationary sequence of random variables n n>N0 , generated by the measure preserving map T , is weakly dependent (on average) if and only if T is mixing (on average).

Proof We prove the equivalence of mixing and weak dependence. The proof of the equivalence of mixing and weak dependence on average is analogous and left as an exercise to the reader. ˜  ˆ  ˆ nˆ  > „ “ If T is mixing we have to show that X Xn n>N0 with Xn ω X0 T ω , n N0 is 6 Stationary Sequences of Random Variables 93 weakly dependent. > Let B1,B2 BR. W.l.o.g., choose k 0. Then n 1 n 1 PˆX0 > B1,X0 X T > B2 PˆX ˆB1 9 T ˆX ˆB2 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶0 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶0 n A1 T A2 ÐÐÐ PˆA1PˆA2 PˆX0 > B1PˆX0 > B2. n ª ¨ ˜  ˆ  ˆ nˆ  > „ “ Let any X Xn n>N0 with Xn ω X0 T ω , n N0 be weakly dependent. For any A1,A2 > F construct the random variable ¢ ¨0, ω ¶ A 8 A ¨ 1 2 ¨ > 9 c ¨1, ω A1 A2 X0ˆω ¦ . ¨2, ω > A 9 A ¨ 1 2 ¨ > c 9 ¤3, ω A1 A2

n Since Xnˆω X0ˆT ˆω yields a weakly dependent sequence, it holds

n PˆA1 9 T A2 Pˆ˜1 B X0 B 2 9 ˜Xn C 2 ÐÐÐ PˆX0 > 1, 2¥PˆX0 C 2 PˆA1PˆA2. n ª Hence, T is mixing. Theorem 6.2.4 Let T be a measure preserving map. T is ergodic iff it is mixing on average.

Proof „ “ It is to be shown that if T is mixing on average then T is ergodic, i.e. for all A > J it holds 0 PˆA œ . Let A > F, A A > J. Then 1 1 2 1 n n P PˆA1 9T ˆA2 PˆA1 9A2 ÐÐÐ PˆA1PˆA2, which is possible only if PˆA1 9A2 n k 1 ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ n ª A2 0 PˆA PˆA . For A A, we get PˆA P2ˆA and hence PˆA œ 1 2 1 1 „ ¨ “ Later.

Now we give the motivation for the term „mixing “. Theorem 6.2.5 Let A > F, PˆA A 0. The measure preserving map T ¢ Ω Ω is ergodic (i.e. mixing on average) iff ˆ8ª n  P n 0T A 1, n i.e. the preimages T A, n > N0, cover almost the whole Ω. Proof „ “ 8ª n 1 8ª n ` Let B n 0T A. Obviously, T B n 1T A B. Since T is measure preserving, i.e.     PˆT 1B PˆB, it follows that PˆT 1B Q B PˆBƒT 1B PˆB  PˆT 1B 0. Hence, ‡ 0 B > J (B – almost invariant w.r.t. T ) and PˆB œ . PˆB C PˆA A 0 ¨ PˆB 1. 1 94 6 Stationary Sequences of Random Variables

„ ¨ “ Let T be non-ergodic. It is to be shown, that PˆB @ 1. T A > J @ PˆA @ B 8ª T nA A If is not ergodic, there exists such that 0 1. n 0 ² and hence A PˆB @ 1.

Remark 6.2.5 So far, the fact that the random variables X are real-valued was never explicitly used. Therefore the above observations can be transfered without modifications to sequences of random elements with values in an arbitrary measurable space M.

6.2.3 Ergodic Theorem ˜ ª ˆ  Let X Xn n 0 be a sequence of random variables on the probability space Ω, F, P . If Xn are i.i.d., then by the strong law of large numbers

n1 1 a.s. Q Xk ÐÐÐ EX0, if ESX0S @ ª. n ª n k 0

We want to prove a similar statement about stationary sequences. Theorem 6.2.6 (Ergodic theorem, Birkhoff-Khinchin): X ˜X  Let n n>N0 be a stationary sequence of random variables, generated by the random variable X0 and a measure preserving map T ¢ Ω Ω. Let J be the σ-algebra of the invariant sets from T and ESX0S @ ª. Then

n1 1 a.s. Q Xk ÐÐÐ EˆX0 S J. n ª n k 0

If X is weakly dependent on average (i.e. T – ergodic) then EˆX0 S J EˆX0. Lemma 6.2.5 ˜  ˆ  n1 ˆ kˆ  ˆ  ˜ ˆ  ˆ  Let Xn , T be as above. Let Sn ω Pk 0 X0 T ω , Mk ω max 0,S1 ω ,...,Sk ω . Under the condition of Theorem 6.2.6 it holds

EˆX01ˆMn A 0 C 0, n > N.

Proof Let ω > ˜ω ¢ Mnˆω A 0. For all k B n it holds Skˆω0 B Mnˆω0, ω0 > Ω. Take ω0 T ˆω. We can add X0 and get

X0ˆω  MnˆT ˆω C X0ˆω  SkˆT ˆω Sk1ˆω.

For k 0 it holds X0ˆω C Sk1ˆω  MnˆT ˆω, k 0, . . . , n  1. Hence it follows that X0ˆω C max ˜S1ˆω,...,Snˆω MnˆT ˆω. Since Mnˆω A 0, then Mn max ˜S1,...,Sn. ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ Mnˆω It follows that

EˆX01ˆMn A 0 C EˆˆMn  MnˆT 1ˆMn A 0 C EˆMn  MnˆT  0, since T is measure preserving. 6 Stationary Sequences of Random Variables 95

Proof of the Theorem 6.2.6 W.l.o.g. let EˆX0 S J 0, otherwise replace X0 by X0  EˆX S J. a.s.  Sn Pn 1 It has to be shown: limn ª n 0, Sn k 0 Xk. It is enough to show that

Sn Sn 0 B lim inf B lim sup B 0. n ª n n ª n

S Sn B Pˆ S A ε  ε A First we show that lim supn ª n 0. It is enough to show that ² 0 for all 0. Aε ‡ ˆ   ‡ Pk1 ‡ˆ jˆ  ‡ ˜ ‡ ‡ Let X0 X0 ε 1Aε , Sk j 0 X0 T ω , Mk max 0,S1 ,...,Sk . By Lemma 6.2.5, it ˆ ‡ ˆ ‡ A  C C follows E X0 1 Mn 0 0 for all n 1. But S‡ ˜ ‡ A  › ‡ A ¡ œ ‡ A ¡ œ k A ¡ œ Sk A ¡ 9 Mn 0 max Sk 0 n ª sup Sk 0 sup 0 sup ε Aε Aε, 1BkBn kC1 kC1 k kC1 k

™ Sk A ž a ™ A ž since supkC1 ε S ε Aε. By Lebesgue’s theorem ‡ ‡k ‡ ‡ ‡ 0 B EˆX 1ˆM A 0 ÐÐÐ EˆX 1A , since ESX S B ESX0S  ε. Hence 0 B EˆX 1A  EˆˆX0  0 n n ª 0 ε 0 0 ε e ε1A  EˆX01A εPˆAε EˆEˆX01A S JεPˆAε Eˆ1A EˆX0 S JεPˆAε εPˆAε ε ε ε ε ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ 0 which means that PˆAε B 0, i.e. PˆAε 0 for all ε A 0. B Sn  In oder to show 0 lim infn ª n S it is enough to look at X0 instead of X0, since ˆ Sn  ˆ Sn  ˆ B  ˆ C  lim supn ª n lim infn ª n . Since P S 0 1 it holds P S 0 1. Consider now the case if T is ergodic. Since Y EˆX0SJ is an invariant random variable by definition of J, it follows from Theorem 6.2.2, 3) that Y const. a.s., i.e., Y EY EˆEˆX0SJ EX0. e ¯ Sn ¯ˆ  ¯ ˜ ¯ A  > : Since S lim sup n is invariant w.r.t. T (S T S) then Aε S ε J, and hence 1Aε n ª is J-measurable. Then ˆ S  ˆ S  E X01Aε J 1Aε E X0 J .

Remark 6.2.6 The peculiarity of the Ergodic Theorem in comparison with the strong law of large numbers lies in the fact that the limiting value EˆX0 S J is random. Example 6.2.4 We consider the probability space from Example 6.2.3 a): Ω ˜ω1, . . . , ωd, d 2l > N. T ¢ Ω Ω be defined by ¢ ˆ   ¨ T ωi ωi2 , i 1, . . . , d 2, ¦ T ˆωd1 ω1 , ¨ ¤ T ˆωd ω2 .

Let A1 ˜ω1, ω3, . . . , ω2l1, A2 ˜ω2, ω4, . . . , ω2l. Since ˆΩ, F, P is a Laplace probability ˆ˜  1 ˆ  1 > space (P ωi d , for all i) it follows that P Ai 2 , i 1, 2. On the other hand, A1,A2 J w.r.t. T and therefore T is not ergodic. For an arbitrary random variable X0 ¢ Ω R it holds n1 2 Pl1 ˆ  1 > 1 n d j 0 X0 ω2j1 , with probability 2 , if ω A1, Q X0 ˆT ˆω ÐÐÐ œ n ª 2 Pl ˆ  1 > n k 0 d j 1 X0 ω2j , with probability 2 , if ω A2. 96 6 Stationary Sequences of Random Variables

Proof of theorem 6.2.4 It has to be shown: If T ¢ Ω Ω is ergodic, then T is mixing on average, i.e. for all A1,A2 > F n1 1 k Q PˆA1 9 T A2 ÐÐÐ PˆA1PˆA2. n ª n k 0

1 n1 k T heorem 6.2.6 k Let Yn P 1ˆT A2 ÐÐÐÐÐÐÐÐ PˆA2, since T is ergodic, thus the sequence ™1ˆT A2ž n k 0 n ª k>N a.s. is weakly dependent on average. By Lebesgue’s theorem it follows from 1ˆA1Yn ÐÐÐ 1ˆA1PˆA2 n ª that n1 1 k Eˆ1ˆA1Yn Q PˆA1 9 T A2 ÐÐÐ PˆA1PˆA2. n ª n k 0

Lemma 6.2.6 ˜X  p C If n n>N is a uniformly integrable sequence of random variables and n,i 0, such that n > n S S > Pi 1 pn,i 1 for all n N, then the sequence of random variables Yn Pi 1 pn,i Xi , n N, uniformly integrable as well. Without proof Conclusion 6.2.1 Under the conditions of theorem 6.2.6 it holds n1 1 L2 Q Xk ÐÐÐ EˆX0 S J n ª n k 0 resp. n1 1 L2 Q Xk ÐÐÐ EˆX0 n ª n k 0 in the ergodic case. ˜  ˆS S ˆS S A  ˆS S ˆS S A  ÐÐ Proof If Xn n> 0 is stationary, it then holds supn E Xn 1 Xn ε E X0 1 X0 ε N ε 0 S S @ ª 1 Pn1 Pn 1 ˜ 1 Pn1 Pn S S 0, since E X0 . Let Sn n k 0 Xk i 1 pn,iXi1, pn,i n , Sn n k 0 Xk i 1 pn,i Xi1 . From lemma 6.2.6, ™S˜nž is also uniformly integrable and after Lemma 5.3.2 it follows from n>Z ÐÐÐ a.s. S S B 1 n1 S S ÐÐÐ Sk 0 that E Sn Pk 0 E Xk 0. k ª n n ª

6.3 Stationarity in the Wide Sense

˜X  ESX S2 @ Let n n>Z be a sequence of random variables, which is stationary in the wide sense: n ª, n > N. ESXnS const, n > N, covˆXn,Xm Cˆn  m, n, m > Z.

6.3.1 Correlation Theory Theorem 6.3.1 (Herglotz): Let C ¢ Z R be a positive semi-definite function. Then there exists a finite measure µ on ˆπ, π, such that π inx Cˆn S e µˆdx, n > Z. π µ is called spectral measure of C. 6 Stationary Sequences of Random Variables 97

Remark 6.3.1 Since covariance function of a stationary sequence is positive semi-definit, the the above repre- sentation holds for an arbitrary covariance function C. Definition 6.3.1 A family ˜Qλ, λ > ˝ of probability measures is called weakly relatively compact, if an arbitrary ˜Q  ˜Q  sequence of measures λn n>N has a subsequence λnk n>N, which converges weakly. Definition 6.3.2 A family of probability measures Q ˜Qλ, λ > ˝ on ˆS, B, B – Borel σ-algebra on a metric space S is called tight, if for all ε A 0 there exists a compactum, such that Kε > B and QλˆKε A 1  ε for all λ > Λ. Theorem 6.3.2 (Prokhorov): If the family of probability measures Q ˜Qλ, λ > ˝ on the metric measurable space ˆS, B is tight, then it is weakly relatively compact. If S is a , then every weakly relatively compact familiy Q ˜Qλ, λ > ˝ of measures is also tight. Without proof

The theorem of Prokhorov is used to prove the weak convergence of a sequence of probability measures, by checking the tightness among other things. In particular, if S is compact, then every family of probability measures on ˆS, B is tight, since Kε S for all ε A 0.

Proof of theorem 6.3.2 „ “ ˆ  π inx ˆ  > > > > If C n Rπ e µ dx , n Z, then for all n N, for all z1, . . . , zn C and t1, . . . , tn Z

n π n 2 izix Q zjz¯jCˆti  tj S WQ zie W µˆdx C 0.  i,j 1 π i 1 Hence follows that C is positive semi-definit. „ ¨ “ C >  ¥ ˆ  1 PN ˆ   ikx ijx C For all N 1, x π, π , define the function gN x 2πN k,j 1 C k j e e 0, which is continuous in x, since C is positive semi-definit. It holds S S 1 n inx gN ˆx Q Œ1  ‘ Cˆne , 2π SnS@N N

2 sine there are N  SnS pairs ˆk, j > ˜1,...,N , such that k  j n. Define the measure µN on ˆ π, π¥, B π,𥍠by µN ˆB R gN ˆxdx, B > Bˆ π, π¥. ¢ S S π π ¨ Š  n  Cˆn, SnS @ N, inx ˆ  inx ˆ  ¦ 1 N S e QN dx S e gN x dx ¨ π 𠤨 0, otherwise, inx 2 since ™e ž is a orthogonale system in L π, π¥. For n 0 it holds QN ˆ π, 𥍠Cˆ0 @ ª, n>Z QN hence š ˆ  Ÿ is a family of probability measures, which is tight. After theorem 6.3.2 there C 0 n> N ˜  exists a subsequence Nk k> , ω N such that QN ÐÐÐ µ. µ – finite measure on π, π¥ and hence follows k ª π S S inx n lim S e gN ˆxdx lim Œ1  ‘ Cˆn Cˆn, for all n > Z. k ª π k ª Nk 98 6 Stationary Sequences of Random Variables

X ˜X  Let n n>Z be a stationary in the wide sense sequence of random variables. Then the following spectral representation holds: π d inx Xn S e Zˆdx, n > Z, π where Z is an orthogonal random measure on ˆ π, π¥, Bˆ π, 𥍍. Therefore both Z and ˆ  π ˆ  ˆ  ¢  ¥ I f Rπ f x Z dx are to be introduced for deterministic functions f π, π C.

6.3.2 Orthogonal Random Measures Construction scheme of Z resp. Iˆ : 1. Z is defined on a semiring K (the sumbset of Λ). 2. Z is defined on the algebra A, which is generated by K. 3. Define the integralI w.r.t. Z for a simple function on σˆA, if the measure µˆΛ @ ª, µ – given measure.

4. Define I as limn ª Iˆfn for arbitrary random functions f, f limn ª fn, fn simple, µˆΛ @ ª.

5. Define I on a σ-finite space Λ 8nΛn, µˆΛn @ ª, Λn 9Λm g, n x m, as Iˆf Pn Iˆf S Λn, In – integralw.r.t. Z on Λn. Hence Z is extended on ˜A > σˆA ¢ µˆA @ ª as ZˆA Iˆ1ˆA. Step 1 Let K be a semiring of the subsets of Λ (Λ – arbitrary space), i.e. for all A, B > K it holds 9 > ` > 9 g x 88n A B K; if A B, then there exist A1,...,An K, Ai Aj , i j, such that B A i 1Ai. Definition 6.3.3 1. A complexvalued random measure Z ˜ZˆB,B > K, given on the probability space ˆΩ, F, P, is called orthogonal, if a) all ZˆB > L2ˆΩ, F, P, B > K, > 9 g ¨ ` ˆ  ˆ e ˆ ˆ  ˆ  b) A, B K, A B Z A ,Z B L2ˆΩ,F,P E Z A , Z B 0,

c) as a random measure the σ-additivity of Z holds: If B,B1,...,Bn,... > K, B 8nBn, a.s. Bi 9 Bj g, i x j, ZˆB Pn ZˆBn, where the convergence of this series is intepreted in L2ˆΩ, F, P terms. ˜ ˆ  >  ˆ  S ˆ S2 ` ˆ  ˆ e 2. The term µ µ B ,B K defined by µ B E Z B Z B ,Z B L2ˆΩ,F,P, B > K, is called stucture measure of Z. It is easy to see that µ is in fact a measure on K. If Λ > K, then µ is finite, otherwise σ-finite, Λ 8nΛn, Λn > K, Λn 9 Λm g, such that µˆΛn @ ª. 3. The orthogonal random measure Z is called centered, if EZˆB 0, B > K. Example 6.3.1 Let Λ 0, ª, K ˜ a, b, 0 B a @ b @ ª, Zˆ a, b W ˆb  W ˆa, 0 B a @ b @ ª, where W ˜W ˆt, t C 0 is the Wiener process. Z is an orthogonal random measure on K, since W has independent increments. Analog, this definition can be transfered to an arbitrary quadratic integrable stochastic process X with independent increments instead of W . 6 Stationary Sequences of Random Variables 99

Step 2 Theorem 6.3.3 Let µ be a σ-finite measure on the algebra A, which is generated by K (after the theorem of Caratheodon µ is uniquely continued on σˆA). Then there exists a probability space ˆΩ, F, P and a centered orthogonal random measure Z on ˆΩ, F, P, defined on ˜B > A ¢ µˆB @ ª, with structure measure (or control measure) µ. Without proof > 8n > 9 g x To the definition of Z on A: for B A, B i 1Bi, Bi K, Bi Bj , i j, we set ˆ  n ˆ  Z B Pi 1 Z Bi .

6.3.3 Integral regarding an Orthogonal Random Measure Step 3 ¢ ˆ  n ˆ >  > > Let f Λ C be a simple function, i.e. f x Pi 1 ci1 x Bi , for ci C and Bi E, 8n 9 g x ˆ  i 1, . . . , n, such that i 1Bi Λ, Bi Bj , i j, and Λ, E, µ be a measurable space with µˆΛ @ ª. Definition 6.3.4 The integral of f w.r.t. an orthogonal random measure Z defined on ˆΩ, F, P is given by ˆ  ¢ ˆ  ˆ  Pn ˆ  I f RΛ f x Z dx i 1 ciZ Bi . Exercise 6.3.1 Show that the definition is correct, i.e. Iˆf does not depend on the representation of f as a simple function. Lemma 6.3.1 (Properties of I): Let Iˆ  be the integral w.r.t. the orthogonal random measure, defined on a simple function Λ C as abovev. The following properties hold: ` ˆ  ˆ e ` e 1. Isometry: I t ,I g L2ˆΩ f, g L2ˆΩ, where f and g are simple functions Λ C, ` e ˆ  ˆ  ˆ  f, g L2ˆΩ RΛ f x g x Λ dx .

a.s. 2. Linearity: For every simple function f, g ¢ Λ C holds Iˆf  g Iˆf  Iˆg. Exercise 6.3.2 Proof it. Step 4 2 Let now f > L ˆΩ, E, µ. Then there exists a sequence of simple functions fn ¢ Λ C, such 2ˆ  L Λ 2 that fn ÐÐÐ f (simple functions are tight in L ˆΛ). Then define Iˆf limn ª Iˆfn, n ª whereas this limit is to be understood in the L2ˆΩ, F, P sense. You can show, that the definition of Iˆf is independent of the choice of the sequence ˜fn. Lemma 6.3.2 The statements of lemma 6.3.1 hold for the general case.

Proof Use the continuity ` , e. Remark 6.3.2 If Z is centered, then EIˆf 0 holds for arbitrary functions f > L2ˆΛ, E, µ. 100 6 Stationary Sequences of Random Variables

Step 5 Let now Λ be σ-finite, i.e. Λ 8nΛn, µˆΛn @ ª, Λn 9 Λm g, n x m. Then for all > 2ˆ  P S 2ˆ 9  f L Λ, E, µ holds f n f Λn . On L Λn, E Λn, µ the integral In w.r.t. Z is defined as in ˆ  ¢ P ˆ S  1)- 4). Now set I f n In f Λn . Theorem 6.3.4 The map g ¢ L2ˆΛ, E, µ L2ˆΩ, F, P is an isometry. In particular, as a result, the random measure Z on ˜B > ε ¢ µˆB @ E can be continued as ZˆB ¢ Iˆ1B, B > E ¢ µˆB @ ª.

6.3.4 Spectral Representation

Let X ˜Xˆt, t > T  be an arbitrary complexvalued stochastic process on ˆΩ, F, P, T – an arbitrary index set, ESXˆtS2 @ ª, t > T , EXˆt 0, t > T (w.l.o.g., otherwise consider X˜ˆt Xˆt  EXˆt), t > T , with Cˆs, t EˆXˆs, Xˆt, s, t > T ). Theorem 6.3.5 (Karhunen): ˆ  ˆ  ˆ  > X has the spectral representation X t RΛ f t, x Z dx , t T (i.e., there exists a centered orthogonal random measure on ˜B > E ¢ µˆB @ ª, where L2ˆΛ, E, µ is an as above defined space), if and only if there exists a system of the functionsfˆt,  > L2ˆΛ, E, µ, t > T , such that ˆ  ˆ  ˆ  ˆ  > 2ˆ  C s, t RΛ f s, x f t, x µ dx , s, t T , and this system F is completely in L Λ, E, µ (i.e. ` ˆ  e > 2ˆ  > ¡ f t, , ψ L2ˆΩ 0, ψ L Ω, E, µ , for all t T and ψ 0, µ almost everywhere). Without proof Theorem 6.3.6 Let ˜Xn, n > Z be a centered complexvalued stationary in the wide sense sequence of ran- dom variables on ˆΩ, F, P. Then there exists an orthogonal centered random measure on ˆ  ¥ ˆ  ¥ ˆ  a.s. π inx ˆ  > π, π , B π, π (defined on Ω, F, P ), such that Xn Rπ e Z dx , n Z.

inx 2 Proof Let F ˜e , x > π, π¥, n > Z. This system in complete on L ˆ π, 𥍠(comp. the theory of the Fourier-series). From the theorem of Herglotz follows that

π inx imx Cˆn, m EˆXnXm S e e µˆdx, π where µ is the spectral measure of X, thus a finite measure on ˆ π, π¥, Bˆ π, 𥍍. Af- a.s. ter theorem 6.3.5 there exists an orthogonal random measure on ˆΩ, F, P, such that Xn π inx ˆ  > Rπ e Z dx , n Z.

Theorem 6.3.7 (Ergodic theorem for stationary (in the wide sense) sequences of random variables): Unter the conditions of theorem 6.3.6 it holds

 1 n 1 L2ˆΩ Q Xk ÐÐÐ Zˆ˜0. n k 0

L2ˆΩ > 1 Pn1 ÐÐÐ In particular if X is not centered, i.e. EXn a, n Z, then n k 0 Xk a converges, if ESZˆ˜0S2 0, thus Z and therefore µ has no atom in zero. ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ µˆ˜0 6 Stationary Sequences of Random Variables 101

1 1einx   x 1 n 1 1 n 1 ikx ˆ  ˆ  œ n 1eix , x 0 >  Proof Sn Pk 0 Xk R Pk 0 e Z dx . ψn x , for all n N. Sn n ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶n 1, x 0 ψnˆx π π 2 2 Zˆ˜0 R ˆψnˆx  1ˆx 0 Zˆdx R ϕnˆxZˆdx. YSn  Zˆ˜0Y 2ˆ  YϕnˆxY 2ˆ  ¥  π ´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶ π L Ω L π,π ,µ ϕnˆx π S ˆ S2 ˆ  ÐÐÐ S ˆ S B 2 ÐÐÐ R ϕn x µ dx 0 after the theorem of Lebesgue, since ϕn x ix 0 for all π n ª nS1e S n ª x > π, π¥.

6.4 Additional Exercises

Exercise 6.4.1 ª Let Z1,Z2,... be a sequence of random variables, such that the series Pi 1 Zi converges almost surely. Let a1, a2,... be a monotone increasing sequence of positive (deterministic) numbers with an ª, n ª. Show that

n 1 a.s. Q akZk 0, n ª. an k 1 Exercise 6.4.2 Let X be a non-negative variable on a probability space ˆΩ, F, P an T ¢ Ω Ω a measure preserving map. Show that ª Q XˆT kˆω ª a.s. k 0 for almost all ω > Ω with Xˆw A 0. Exercise 6.4.3 Let X be a random variable on a probability space ˆΩ, F, P and T ¢ Ω Ω a measure preserving map. Show that EX EˆX X T , i.e.

S XˆT ˆωPˆdω S XˆωPˆdω. Ω Ω (Hint: algebraic induction) Exercise 6.4.4 Let ˆΩ, F, P be a probability space, where Ω 0, 1, F Bˆ 0, 1 and P is the Lebesgue measure. Let λ > ˆ0, 1.

(a) Show that T ˆx ˆx  λ mod 1 is a measure preserving map, where  a  > >  a mod m a b m for a R and b Z and is the Gauss bracket. (b) Show that T ˆx λx and T ˆx x2 are no measure preserving maps. Bibliography

[1] D. Applebaum. Lévy processes and stochastic calculus. Cambridge University Press, Cam- bridge, 2004.

[2] A. A. Borovkov. Wahrscheinlichkeitstheorie. Eine Einfürung. Birkhäuser, Basel, 1976.

[3] W. Feller. An introduction to probability theory and its applications. Vol. I/II. John Wiley & Sons Inc., New York, 1970/71.

[4] P. Gänssler and W. Stute. Wahrscheinlichkeitstheorie. Springer, Berlin, 1977.

[5] G. Grimmett and D. Stirzaker. Probability and random processes. Oxford University Press, New York, 2001.

[6] C. Hesse. Angewandte Wahrscheinlichkeitstheorie. Vieweg, Braunschweig, 2003.

[7] O. Kallenberg. Foundations of modern probability. Springer, New York, 2002.

[8] J. F. C. Kingman. Poisson processes. Oxford University Press, New York, 1993.

[9] N. Krylov. Introduction to the theory of random processes. American Mathematical Society, Providence, RI, 2002.

[10] S. Resnick. Adventures in stochastic processes. Birkhäuser, Boston, MA, 1992.

[11] G. Samorodnitsky and M. Taqqu. Stable non-Gaussian random processes. Chapman & Hall, New York, 1994.

[12] K.-I. Sato. Lévy Processes and Infinite Divisibility. Cambridge University Press, Cam- bridge, 1999.

[13] A. N. Shiryaev. Probability. Springer, New York, 1996.

[14] A. V. Skorokhod. Basic principles and applications of probability theory. Springer, Berlin, 2005.

[15] J. Stoyanov. Counterexamples in probability. John Wiley & Sons Ltd., Chichester, 1987.

[16] A. D. Wentzel. Theorie zufälliger Prozesse. Birkhäuser, Basel, 1979.

102