• Ei tuloksia

About mean-variance hedging with basis risk

N/A
N/A
Info
Lataa
Protected

Academic year: 2022

Jaa "About mean-variance hedging with basis risk"

Copied!
56
0
0

Kokoteksti

(1)

About mean-variance hedging with basis risk

Sami L¨ ahdem¨ aki

Master’s thesis of mathematics

University of Jyv¨askyl¨a

Deparment of Mathematics and Statistics Autumn 2021

(2)

i

Abstract: Sami L¨ahdem¨aki, About mean-variance hedging with basis risk, Master’s Thesis of mathematics, 52p., University of Jyv¨askyl¨a, Deparment of Mathematics and Statistics, autumn 2021.

In this thesis we introduce a mean-variance hedging problem in an incomplete market. As a main source we follow X. Xue, J. Zhang and C. Weng article Mean- variance Hedging with Basis Risk. We assume a time interval [0, T] for some T >0, an arbitrage free financial market, and consider one risk-free asset and (m+ 1) risky assets. The dynamics of the assets are given by stochastic differential equations with deterministic and Borel-measurable coefficients. One risky asset is connected to the pay-off function which we want to hedge. We assume that this connected asset can not be used in hedging and this makes the market incomplete. Because of incompleteness perfect hedging is not possible.

We define a profit-and-loss random variable by using the difference between the value of the hedging portfolio and the pay-off function. A mean-variance criterion is used to this random variable and by that the solution is a hedging strategy which maximizes the difference between the expected value and variance of the profit-and- loss random variable.

To find a solution we start by recalling some important results from probability theory and stochastic analysis. We introduce shortly multiple stochastic integrals and properties of them. These integrals are used to define the Malliavin derivative.

The mean-variance hedging problem is solved by using Linear-Quadratic theory. We consider an auxiliary problem and show that by solving the auxiliary problem we are able to solve the original problem. The solving method with Linear-Quadratic theory is connected to the backward stochastic differential equations (BSDE) and in the thesis we see also the connection of the BSDEs to the Malliavin derivative. We compute an explicit formula for the Malliavin derivative of a forward contract and an European put and call option.

The pay-off function in this thesis is assumed to be Malliavin differentiable and hence we are able to give an explicit solution for the problem. As a main theorem we formulate an explicit hedging strategy which solves the mean-variance hedging problem in the incomplete market.

(3)

ii

Tiivistelm¨a:Sami L¨ahdem¨aki,About mean-variance hedging with basis risk, matema- tiikan pro gradu -tutkielma, 52 s., Jyv¨askyl¨an yliopisto, Matematiikan ja tilastotieteen laitos, syksy 2021.

T¨ass¨a tutkielmassa perehdyt¨a¨an odotusarvo-varianssi -suojausongelmaan (engl.

mean-variance hedging problem) ep¨at¨aydellisill¨a sijoitusmarkkinoilla. P¨a¨al¨ahteen¨a seu- raamme X. Xuen, J. Zhanging ja C. Wengin artikkelia Mean-variance Hedging with Basis risk. Oletamme aikav¨alin [0, T] jollekinT >0, arbitraasivapaan sijoitusmarkki- nan, yhden riskitt¨om¨an sijoituskohteen ja (m+ 1) riskillist¨a sijoituskohdetta. N¨aiden kohteiden arvon oletetaan noudattavan stokastisia differentiaaliyht¨al¨oit¨a, joissa ker- toimet ovat deterministisi¨a ja Borel-mitallisia. Yksi n¨aist¨a riskillisist¨a sijoituskohteista oletetaan liittyv¨an vaateeseen, jolle haluamme rakentaa suojaussalkun. T¨at¨a kyseis- t¨a sijoituskohdetta ei voida k¨aytt¨a¨a suojaussalkun rakentamisessa, mik¨a aiheuttaa sijoitusmarkkinan ep¨at¨aydellisyyden. T¨am¨an vuoksi my¨os t¨aydellisen suojaussalkun rakentaminen ei ole mahdollista.

M¨a¨arittelemme voittoa/tappiota kuvaavan satunnaismuuttujan k¨aytt¨am¨all¨a suo- jaussalkun arvon ja vaateen erotusta. Odotusarvo-varianssi -kriteeri¨a k¨aytet¨a¨an t¨ah¨an satunnaismuuttujaan ja t¨am¨an johdosta ratkaisu on suojaussalkku, joka maksimoi erotuksen voittoa/tappiota kuvaavan satunnaismuuttujan odotusarvon ja varianssin v¨alill¨a.

Ratkaisun l¨oyt¨amiseksi aloitamme kertaamalla t¨arkeit¨a ja tarpeellisia tuloksia to- denn¨ak¨oisyysteoriasta ja stokastisesta analyysist¨a. T¨am¨an j¨alkeen esittelemme lyhyes- ti moninkertaiset stokastiset integraalit ja niiden ominaisuuksia sek¨a k¨ayt¨amme n¨ai- t¨a Malliavin derivaatan m¨a¨arittelyyn. Odotusarvo-varianssi -ongelman ratkaisun l¨oy- t¨amiseksi k¨ayt¨amme ”Linear-Quadratic” -teoriaa. Oletamme apuongelman ja osoi- tamme, ett¨a ratkaisemalla apuongelman on mahdollista ratkaista my¨os alkuper¨ainen ongelma. K¨aytt¨am¨amme ”Linear-Quadratic” -teoria on yhteydess¨a takaperoisiin sto- kastisiin differentiaaliyht¨al¨oihin ja tutkielmassa n¨aemme n¨aiden yhteyden Malliavin derivaattaan. Johdamme my¨os eksplisiittiset ratkaisut suoran sopimuksen ja Euroop- palaisen myynti- ja osto-option Malliavin derivaatalle.

T¨ass¨a tutkielmassa vaateen oletetaan olevan Malliavin derivoituva ja t¨am¨a mah- dollistaa eksplisiittisen ratkaisun l¨oyt¨amisen. P¨a¨ateoreemana muotoilemme eksplisiit- tisen suojaussalkun, joka ratkaisee odotusarvo-varianssi -ongelman tilanteessa, jossa sijoitusmarkkina on ep¨at¨aydellinen.

(4)

Contents

Introduction 1

Chapter 1. Probability theory and stochastic analysis 3

1.1. Probability space and random variables 3

1.2. Lebesgue integral 4

1.3. Stochastic processes 5

1.4. Conditional expectation and martingales 6

1.5. Itˆo integral and Itˆo’s formula 8

1.6. Stochastic differential equations 12

Chapter 2. Malliavin calculus 14

2.1. A multiple stochastic integral 14

2.2. The Malliavin derivative 21

Chapter 3. Mean-variance hedging with basis risk:

Formulation of the problem 24

3.1. Continuous financial market 24

3.2. Formulation of the problem 25

Chapter 4. The hedging strategy as stochastic linear-quadratic problem 27

Chapter 5. The optimal hedging strategy 42

Appendix A. 47

Appendix B. Notations 49

Bibliography 51

iii

(5)

Introduction

In this thesis we are interested in optimal hedging strategies for a given pay-off, that means we try to determine a trading strategy which replicates the pay-off. We assume that the financial market is arbitrage free and consists of one risk-free asset earning by constant rate and (m+ 1) risky assets with dynamics given by stochastic differential equations. In a complete market it is always possible to find a hedging strategy which replicates a given pay-off, and this possibility is actually given as a definition of completeness in [11]. However, in the setting of this thesis the asset which is connected to our hedging objective is not allowed to be used in hedging. So we only can use other assets which are stochastically dependent on the one which we can not trade with. This causes the market to be incomplete. So our aim here is to determine trading strategies such that the outcome is close to the given pay-off in some sense.

We use a mean-variance criterion to measure the closeness. The portfolio selection using this criteria has been proposed by Markowitz [17], where the variance is assumed to be a measure for risk. We define a profit-and-loss random variable at terminal time T >0 by setting

Vθ(T) =Xθ(T)−G(S0;T),

where Xθ is the value of the hedging portfolio using the hedging strategy θ, and G is the pay-off function. By the mean-variance criterion the aim is to find a strategy which solves the problem

maxθ∈Θ

n

E[Vθ(T)]− γ

2Var[Vθ(T)]

o ,

where γ >0 can be interpreted by [16] as the weight which the investor puts on the variance.

To solve this problem we use stochastic Linear-Quadratic theory as a tool. This method is used in [26] and [16] for example. Both papers assume a complete market where all assets are possible to use for hedging. So we can only use the idea. The main source for us which we follow is [24].

In Linear-Quadratic theory solving the mean-variance problem is connected to solving two different equations. In our case these are a backward differential equation and a backward stochastic differential equation. It is shown that if these two equations are solvable, then also our problem can be solved.

We will see that solving the backward stochastic differential equation has a con- nection to Malliavin derivatives so we also shortly introduce Malliavin calculus by using [19]. Also the Malliavin derivative of the pay-off function is needed in the opti- mal hedging strategy so we compute the Malliavin derivatives of a forward contract, a European put and call option and an Asian option.

1

(6)

INTRODUCTION 2

As a main result we are able to prove an explicit formula for the hedging strategy which solves the mean-variance hedging problem in an incomplete market where the basis risk follows from the setting.

The thesis is organized as follows: In Chapter 1 we recall some important results from Probability theory and Stochastic analysis which are needed later. In Chapter 2 the basics of Malliavin calculus is given and at the end of the chapter the reader gets some useful tools for calculating Malliavin derivatives. Chapter 3 is used to formulate our hedging problem. In Chapter 4 we give the basic idea of stochastic Linear- Quadratic problems and derive important results concerning our problem. Chapter 5 concludes the thesis and contains the main result, the optimal solution to our mean- variance hedging problem. In Appendix A is a collection of results needed in this thesis, and Appendix B contains a list of notations.

(7)

CHAPTER 1

Probability theory and stochastic analysis

Here we will give some basic definitions and tools from probability theory and stochastic analysis. In this chapter we will use basically [7], [10], [13] and [15].

1.1. Probability space and random variables

Definition 1.1.1. Let Ω be a non-empty set. Then a systemF of subsetsA⊂Ω is called σ-algebra on Ω if the following holds

(1) ∅,Ω∈F

(2) if A∈F then Ac∈F

(3) if A1, A2, ...∈F then ∪i=1Ai ∈F.

In this situation the pair (Ω,F) is called a measurable space.

For later use we give a definition of the Borel σ-algebra.

Definition 1.1.2 ([25]Definition 1.4(ii)). Let B(Rd) be the system of all open subsets in Rd for all d∈N. Then it is called Borel σ-algebra onRd.

Definition 1.1.3. Let (Ω,F) be a measurable space. Then a mapµ:F→[0,∞]

is called measure if the following two properties holds (1) µ(∅) = 0

(2) for all A1, A2, ...∈F with Ai∩Aj =∅ for i6=j it holds µ(∪i=1Ai) =

X

i=1

µ(Ai).

Then (Ω,F, µ) is called measure space.

Definition 1.1.4. Let (Ω,F) be a measurable space. If for a measure µ : F → [0,∞] it holds that µ(Ω) = 1 then we denote µ=P and the measure space (Ω,F,P) is called probability space.

A special probability space, called a complete probability space, is used in many cases.

Definition 1.1.5. Let (Ω,F,P) be a probability space. Let A ∈ F such that P(A) = 0. IfB ⊆A implies thatB ∈F, then probability space is called complete.

To define random variables we start with simple functions.

Definition 1.1.6. Let (Ω,F) be a measurable space. A function f : Ω → R is called simple function if there exists α1, . . . , αn∈R and A1, . . . , An ∈F such that

f(ω) =

n

X

i=1

αi1Ai(ω),

3

(8)

1.2. LEBESGUE INTEGRAL 4

where 1A(ω) = 1 if ω ∈A and 0 otherwise.

Definition 1.1.7. Let (Ω,F) be a measurable space and f : Ω → R. If there is a sequence of simple functions (fn)n=1 such that

f(ω) = lim

n→∞fn(ω)

for all ω ∈ Ω, then function f is called measurable. If the measurable space has a probability measure P, then the measurable function is called random variable.

We do not always have to show that a function is a random variable by using simple functions. For a measurable function there exists an equivalent condition which is often more useful.

Proposition 1.1.8 ([7] Proposition 3.1.3). Let (Ω,F) be a measurable space and let f : Ω→R be a function. Then f is measurable if and only if

f−1((a, b)) :={ω ∈Ω :a < f(ω)< b∈F}, for all −∞< a < b <∞.

An important concept in probability is the independence of random variables. It will be needed later.

Definition1.1.9. Let (Ω,F,P) be a probability space andfi : Ω→Rbe random variables for all i= 1,2, . . . , n. If for all B1, . . . , Bn∈B(R) one has

P(f1 ∈B1, . . . , fn∈Bn) =P(f1 ∈B1). . .P(fn ∈Bn), then the random variables f1, . . . , fn are called independent.

1.2. Lebesgue integral

We assume that the reader is familiar with the Lebesgue integral and recall only some important tools. As a reference we recommend [7]. Our first tool is Dominated convergence.

Proposition1.2.1 (Dominated convergence, ([7] Proposition 5.4.5)).Let(Ω,F, µ) be a measure space andg, f1, f2,· · ·: Ω→Rbe measurable functions such that|fn| ≤g for all n ∈ N. Assume that g is integrable and limn→∞fn = f. Then f is integrable and

n→∞lim Z

fndµ= Z

f dµ.

Another tool that gives a possibility to calculate integrals and especially expecta- tions explicitly is the Change of variable formula which we formulate for a probability space.

Proposition 1.2.2 (Change of variable, ([7] Proposition 5.6.1)). Let (Ω,F,P) be a probability space, (R,B(R)) measurable space, f : Ω → R a random variable and ϕ : R → R Borel-measurable function. Assume that Pf is the distribution of f, meaning

Pf(B) =P({ω ∈Ω :f(ω)∈B}) =P(f−1(B)), for all B ∈R. Then

Z

B

ϕdPf = Z

f−1(B)

ϕ(f)dP,

(9)

1.3. STOCHASTIC PROCESSES 5

for all B ∈R.

1.3. Stochastic processes

Families of random variables play an important role in stochastic analysis and in our case we will give a definition for a continuous time interval [0, T], where T >0.

Definition 1.3.1. Let T > 0 and [0, T]. Then a family of random variables X = (Xt)t∈[0,T] with Xt : Ω → R is called stochastic process with a continuous interval [0, T].

We can think a σ-algebra as all information what we have. Next we introduce a definition which tells about the information what we have at some time point.

Definition 1.3.2. Let (Ω,F,P) be a probability space. Then the family of σ- algebras (Ft)t∈[0,T] is called filtration if Fs ⊆ Ft ⊆ F for all 0 ≤ s ≤ t ≤ T and (Ω,F,P,(Ft)t∈[0,T]) is called stochastic basis.

Using a filtration one can say something about types of measurability of stochastic processes.

Definition1.3.3. Let (Ω,F,P,(Ft)t∈[0,T]) be a stochastic basis andX = (Xt)t∈[0,T], Xt: Ω→R a stochastic process. Then

(1) The process X is called measurable if the function (ω, t)7→ Xt(ω) seen as a map between Ω×[0, T] and R is measurable with respect to F⊗B([0, T]) and B(R).

(2) The process X is called progressively measurable with respect to (Ft)t∈[0,T] if for alls∈[0, T] the function (ω, t)7→Xt(ω) seen as a map between Ω×[0, s]

and Ris measurable with respect to Fs⊗B([0, s]) andB(R).

(3) The process X is called adapted with respect to (Ft)t∈[0,T] if for all t∈[0, T] the random variable Xt isFt-measurable.

Between these three different kinds of measurability one has the following connec- tions.

Proposition 1.3.4 ([10] Propositions 2.1.10. and 2.1.11.). The following holds (1) A progressively measurable process is measurable and adapted.

(2) An adapted process with left- or right-continuous paths is progressively mea- surable.

Next we look at one famous stochastic process. It was first observed by Robert Brown when he was looking at pollen in water by a microscope. He realized that particles move incessant and irregular and published papers 1928 and 1929 about this movement. After this observation 1905 Albert Einstein gave a correct explanation for this phenomenon. From the perspective of mathematics in 1900 Louis Bachelier gave a first, but not rigorous, definition for Brownian motion when he studied fluctuation of stock prices. This was without a connection to Brownian motion in physics. The first rigorous mathematical construction was given by Norbert Wiener in 1923.

Definition 1.3.5 (Brownian motion, ([20] Definition 1.2.1)). Let (Ω,F,P) be a probability space.

(10)

1.4. CONDITIONAL EXPECTATION AND MARTINGALES 6

(1) A process W = (Wt)t∈[0,T] with W0 = 0 is called standard Brownian motion if

(a) (Wt)t∈[0,T] is continuous.

(b) For all n ∈ N and 0 ≤ t1 ≤ t2 ≤ · · · ≤ tn = T the increments Wtn − Wtn−1, . . . , Wt2 −Wt1 are independent.

(c) For all 0 ≤s≤t≤T it holdsWt−s ∼N(0, t−s).

(2) AnRd-valued stochastic processW = (Wt)t∈[0,T],Wt= (Wt1, . . . , Wtd), where W1, . . . , Wd are Brownian motions independent from each other, is called a d-dimensional Brownian motion.

Information from Brownian motion can be collected to a special σ-algebra.

Definition 1.3.6. Let W = (Wt)t∈[0,T] be a Brownian motion which generates σ-algebra for allt∈[0, T]

FtW =σ(Ws : 0≤s≤t).

If

N=

A⊆Ω : there exist aB ∈FTW such that A⊆B and P(B) = 0 , then (Ft)t∈[0,T] with Ft=FtW ∨N is called augmentation of (FWt )t∈[0,T].

1.4. Conditional expectation and martingales

Assume a probability space (Ω,F,P), a random variable f : Ω → R and another σ-algebra Gwhich is included in F. Then if the random variable f is F-measurable it is not always G-measurable. The following proposition and definition introduce the concept of conditional expectation.

Proposition 1.4.1 ([7] Proposition 7.3.1). Let (Ω,F,P) be a probability space, G⊆F be a sub-σ-algebra and f a random variable such that f ∈L(Ω,F,P). Then

(1) There exists a random variable g ∈L(Ω,G,P) such that Z

B

f dP= Z

B

gdP, for all B ∈G. (2) If there is g and g’ such that for both above hold, then

P(g 6=g0) = 0.

Definition 1.4.2 ([7] Definition 7.3.2). The random variable g ∈ L(Ω,G,P) in Proposition 1.4.1 is called conditional expectation of f given G. It is denoted by

g =E[f|G].

It is good to notice that the conditional expectation can be changed on sets with probability zero so it is unique only almost surely. The conditional expectation has many properties which are listed below. For later use especially the property called tower property is useful for us.

Proposition 1.4.3 ([7] Proposition 7.3.3 (1)-(9)). Let (Ω,F,P) be a probability space and H ⊆G⊆F sub-σ-algebras of F. Assume f, g ∈L(Ω,F,P). Then we have the following properties almost surely:

(11)

1.4. CONDITIONAL EXPECTATION AND MARTINGALES 7

(1) Linearity: Let λ, µ∈R. Then

E[λf +µf|G] =λE[f|G] +µE[g|G].

(2) Monotonicity: Let g ≤f almost surely. Then E[g|G]≤E[f|G].

(3) Positivity: Let f ≥0 almost surely. Then E[f|G]≥0.

(4) Convexity: |E[f|G]| ≤E[|f||G].

(5) Projection property: Let f be G-measurable. Then E[f|G] =f. (6) Tower property: E[E[f|G]|H] =E[E[f|H]|G] =E[f|H].

(7) Let h: Ω→R be G-measurable and f h∈L(Ω,F,P). Then E[f h|G] =hE[f|G].

(8) Let G={∅,Ω}. Then E[f|G] =E[f].

(9) Let f be independent from G. Then E[f|G] =E[f].

For the conditional expectation we have a similar property as Proposition 1.2.1 states for the Lebesgue integral. It is called Dominated convergence for conditional expectation.

Proposition 1.4.4 ([1] Equation (15.14)). Let (Ω,F,P) be a probability space, G ⊆ F be a sub-σ-algebra and (Xn)n∈N a sequence of random variables such that

|Xn| ≤Y for all n ∈N and for some integrable random variable Y. Assume also that Xn→X almost surely when n → ∞. Then

n→∞lim E[Xn|G] =E[X|G].

The conditional expectation is used in the definition of martingales, which are important processes in the field of stochastics.

Definition 1.4.5. Let (Ω,F,P,(Ft)t∈[0,T]) be a stochastic basis. A stochastic process M = (Mt)t∈[0,T] is called martingale with respect to a filtration (Ft)t∈[0,T] if

(1) Mt isFt-measurable for all t∈[0, T] (2) E|Mt|<∞ for all t ∈[0, T]

(3) E[Mt|Fs] =Ms for all 0≤s≤t≤T.

Moreover, a martingale M = (Mt)t∈[0,T] belongs to Mc2 if (1) E|Mt|2 <∞for all t ∈[0, T]

(2) the paths t→Mt(ω) are continuous for allω ∈Ω.

The space Mc,02 consists of allM ∈Mc2 with M0 = 0.

Sometimes the setMc2 is not large enough, so we need also a definition for a larger set of processes. For that a map called stopping time is needed.

Definition 1.4.6. Let (Ω,F) be a measurable space with filtration (Ft)t∈[0,T]. Then the map τ : Ω → [0, T] is called stopping time with respect to the filtration (Ft)t∈[0,T] if

{τ ≤t} ∈Ft, for all t∈[0, T].

Definition 1.4.7. Let M = (Mt)t∈[0,T] be a continuous and adapted process with M0 = 0. If there exists an increasing sequence (τn)n=0 of stopping times with limn→∞τn(ω) = ∞ for all ω ∈ Ω such that Mτn = (Mt∧τn)t∈[0,T] is martingale for all n ∈ N, then the process M is called local martingale. Moreover, the set of local martingales is denoted by Mc,0loc.

(12)

1.5. IT ˆO INTEGRAL AND IT ˆO’S FORMULA 8

We have a sufficient condition for a local martingale to be a martingale.

Lemma 1.4.8. Let (Mt)t∈[0,T] be a (continuous) local martingale and G such that sup

t∈[0,T]

|Mt|< G and EG <∞.

Then (Mt)t∈[0,T] is martingale.

Proof. Let (τN)N≥1 be a localizing sequence. Now we have for 0 ≤ s < t ≤ T that

E[Mt∧τN|Fs] =Ms∧τN.

By dominated convergence for conditional expectation (Proposition 1.4.4) we get

N→∞lim E[Mt∧τN|Fs] = E h

N→∞lim Mt∧τN|Fs

i

= E[Mt|Fs]. On the other hand we have

N→∞lim Ms∧τN =Ms. So we conclude

E[Mt|Fs] =Ms.

1.5. Itˆo integral and Itˆo’s formula

In this section we recall the Itˆo integral. We assume the usual conditions meaning that we have a complete probability space (Ω,F,P) and a right continuous filtration (Ft)t∈[0,T] which means ∩>0Ft+ =Ft for all t ∈ [0, T]. We also assume that all sets of probability zero are included in F0 and W = (Wt)t∈[0,T] is a standard Brownian motion. We give only the definitions and main properties of the Itˆo integral. For more information for example [10] is recommend. First we need simple stochastic processes and a stochastic integral for them.

Definition 1.5.1. Let L= (Lt)t∈[0,T] be a stochastic process. It is called simple if there exist

(1) a sequence of time points such that 0 =t0 < t1 <· · ·< tn =T and (2) Fti-measurable bounded random variables υi : Ω→R,

such that Lhas a representation Lt(ω) =

n

X

i=1

υi−1(ω)1(ti−1,ti](t).

The set of all simple processes is denoted byL0.

Definition 1.5.2. LetL∈ L0 and t∈[0, T]. Then stochastic integral is defined as

It(L)(ω) =

n

X

i=1

υi−1(ω) Wti∧t(ω)−Wti−1∧t(ω) .

The stochastic integral of a simple process is a continuous, square integrable mar- tingale as the next proposition states.

(13)

1.5. IT ˆO INTEGRAL AND IT ˆO’S FORMULA 9

Proposition 1.5.3. Let L∈L0. Then I(L) = (It(L)t∈[0,T])∈Mc,02 .

Stochastic integrals of simple processes can be extended to a larger set of inte- grands.

Definition 1.5.4. The set of all progressively measurable stochastic processes L= (Lt)t∈[0,T], Lt : Ω→R with

kLkL2,t =

E Z t

0

L2udu 12

<∞ for all t∈[0, T], is denoted by L2.

Proposition 1.5.5 ([10]Proposition 3.1.12 (i)-(v)). The map I :L0 → Mc,02 can be extended to J :L2 →Mc,02 with properties:

(1) Linearity: Let α, β ∈R and L, K ∈L2. Then

Jt(αL+βK) = αJt(L) +βJt(K) a.s. for t ∈[0, T].

(2) Extension property: Let L∈L0. Then It(L) =Jt(L) a.s. for t∈[0, T].

(3) Itˆo isometry: Let L∈L2. Then

E

Jt(L)212

=

E Z t

0

L2udu 12

for t∈[0, T].

(4) Continuity property: Let (K(n))n=1 be a sequence where K(n) ∈ L2 for all n∈N. Let L∈L2. If d(K(n), L) =P

m=12−mmin

1,kK(n)−LkL2,m →0 whenn → ∞, then

E

"

sup

t∈[0,T]

|Jt(L)−Jt(K(n))|2

#

→0 as n→ ∞.

(5) Uniqueness: If Jˆ:L2 →Mc,02 is another mapping for which above properties hold, then

P

Jt(L) = ˆJt(L), t∈[0, T]

= 1, for all L∈L2.

ForL∈L2 we have thatJ(L) is a square integrable martingale. But this property is vanishing in the next extension.

Definition 1.5.6. (1) The set of progressively measurable stochastic process L= (Lt)t∈[0,T] with

P

ω ∈Ω : Z T

0

Lu(ω)2du <∞

= 1 is denoted by Lloc2 .

(2) Let (τn)n=1 be a sequence of stopping times. It is called localizing for L = (Lt)t∈[0,T]∈Lloc2 if

(a) 0≤τ0(ω)≤τ1(ω)≤ · · · ≤T with limn→∞τn=T for all ω∈Ω and (b) Lτn =Lt1t≤τn ∈L2 for all n= 0,1,2, . . .

The next lemma states the existence of a stochastic integral for processes in Lloc2 .

(14)

1.5. IT ˆO INTEGRAL AND IT ˆO’S FORMULA 10

Lemma 1.5.7 ([10] Lemma 3.1.19). Let L ∈ Lloc2 . Then there is a unique and adapted process X= (Xt)t∈[0,T] with X0 = 0 such that

P(Jt(Lτn) =Xt, t∈[0, τn]) = 1, for all localizing sequences (τn)n=0 of L and for all n= 1,2, . . ..

Remark 1.5.8. The uniquenes in Lemma 1.5.7 means that if there is a Y = (Yt)t∈[0,T] with the same properties, then P(Xt=Yt, t∈[0, T]) = 1.

Definition 1.5.9. LetL∈Lloc2 . Then the process X in Lemma 1.5.7 is called Ito integral and it is denoted by

X = Z t

0

LudWu

t∈[0,T]

.

For integrands from Lloc2 the Itˆo integral is a local martingale:

Proposition 1.5.10 ([10]Proposition 3.1.23 (i)). LetL∈Lloc2 . Then one has that Rt

0 LudWu

t∈[0,T]∈Mc,0loc.

In many situations the task is to show that the integrand is inL2implying that the Itˆo integral is a martingale. As a notation we get for the Itˆo isometry in Proposition 1.5.5

E

(

Z t 0

LudWu)2 12

=

E Z t

0

L2udu 12

for t ∈[0, T].

A very useful tool in stochastic analysis is called Itˆo’s formula which allows us to write stochastic processes in different form. To recall this we need first some definitions.

Definition 1.5.11. LetA = (At)t∈[0,T], At: Ω →R be a stochastic process such that

sup

n∈N

sup

0=t0≤t1≤...≤tn=t n

X

k=1

|Atk(ω)−Atk−1(ω)|<∞ a.s. for all t ∈[0, T].

Then the process A is said to be of bounded variation.

Definition 1.5.12. Let X = (Xt)t∈[0,T] be a continuous and adapted stochastic process. If there exist x0 ∈ R, L ∈ Lloc2 and a progressively measurable process a= (at)t∈[0,T] with

Z t 0

|au(ω)|du <∞

for all t∈[0, T] andω ∈Ω such that X can be represented as Xt=x0+

Z t 0

LudWu

(ω) + Z t

0

audu a.s. for all t∈[0, T], then X is called Itˆo process.

(15)

1.5. IT ˆO INTEGRAL AND IT ˆO’S FORMULA 11

Definition 1.5.13. LetX = (Xt)t∈[0,T] be a continuous and adapted process. If there existx0 ∈R,M ∈Mloc2 and a process A of bounded variation withA0 = 0 such that

Xt=x0+Mt+At, then X is called a continuous semi-martingale.

Remark 1.5.14. Especially Itˆo processes are continuous semi-martingales since sup

0=t0≤t1≤...≤tn=t n

X

k=1

| Z tk

0

audu− Z tk−1

0

audu| = sup

0=t0≤t1≤...≤tn=t n

X

k=1

| Z tk

tk−1

audu|

≤ sup

0=t0≤t1≤...≤tn=t n

X

k=1

Z tk

tk−1

|au|du

= Z t

0

|au|du <∞.

We give a version of Itˆo’s formula which is for continuous semi-martingales.

Proposition1.5.15 ([10]Proposition 3.4.3). Letf ∈C2(Rd)andXt= (Xt1, . . . , Xtd) be a vector of continuous semi-martingales. Then almost surely

f(Xt) = f(X0) +

d

X

i=1

Z t 0

∂f

∂xi(Xu)dXui +1 2

d

X

i,j=1

Z t 0

2f

∂xi∂xj(Xu)d

Mi, Mj

u,

where dXui = dMui +dAiu and hMi, Mjiu = 14 [hMi+Mjiu− hMi−Mjiu] is called cross-variation.

To make it possible to use explicitly above Itˆo’s formula, we need also the following proposition.

Proposition 1.5.16 ([10]Proposition 4.4.3.). Let L∈Lloc2 . Then Z ·

0

LudWu

t

= Z t

0

L2udu,

for all t∈[0, T] almost surely.

We had before the Lemma 1.4.8 which provided a condition when a local martin- gale is a martingale. We are able to give a similar condition called Novikov’s condition for the exponential martingale.

Proposition 1.5.17 ([10]Proposition 4.4.8). Let L∈Lloc2 and t∈[0, T]. Then eR0tLudWu12R0tL2udu

is a martingale if

E h

e12R0TL2udui

<∞.

(16)

1.6. STOCHASTIC DIFFERENTIAL EQUATIONS 12

1.6. Stochastic differential equations

In this section we focus on stochastic differential equations, also called SDEs, and state a proposition about existence and uniqueness of solutions. For more informa- tion about solving SDEs we recommend for example [13]. We assume again that the usual conditions hold as in the previous section but now we assume (Wt)t∈[0,T], Wt= (Wt1, . . . , Wtd) to be a d-dimensional standard Brownian motion adapted to the filtra- tion (Ft)t∈[0,T].

Letb be a d-dimensional vector of functions andσ be a d×k-dimensional matrix of functions. For bj and σij,j ∈ {1,2, . . . , d}, i∈ {1,2, . . . , k} we assume that

(1) bj, σij : [0, T]×Rd×Ω→R.

(2) bj and σij are B([0, T])×B(Rd)×F-measurable.

(3) For all t ∈ [0, T], bj(t,·,·) and σij(t,·,·) are measurable with respect to B(Rd)×Ft.

Definition 1.6.1. Letx0 ∈Rd. Then (1.1)

(dXt=b(t, Xt)dt+σ(t, Xt)dWt X0 =x0

is called stochastic differential equation. The d-dimensional stochastic process X = (Xt)t∈[0,T] is called a strong solution if the following holds:

(1) X is Ft -adapted and has continuous sample paths.

(2) RT

0 (|b(t, Xt)|+|σ(t, Xt)|2)dt <∞a.s. where|·|denotes both thed-dimensional norm and the norm of a matrix.

(3) For allt ∈[0, T] Xt=x0+

Z t 0

b(s, Xs)ds+ Z t

0

σ(s, Xs)dWs a.s.

Does an SDE always have a solution? We provide a property when there exist a unique strong solution.

Proposition 1.6.2 ([13] Theorem 6.2.1). Assume SDE (1.1). If for b and σ the following holds

(1) |b(t, x)|2+|σ(t, x)|2 ≤K(1 +|x|2) a.s.

(2) |b(t, x)−b(t, y)|2+|σ(t, x)−σ(t, y)|2 ≤K|x−y|2 a.s.

for all t ∈ [0, T], x, y ∈ Rd and some constant K > 0, then the SDE has a unique strong solution X = (Xt)t∈[0,T].

Remark 1.6.3. Uniqueness in the above proposition means that if there exists another strong solution Y = (Yt)t∈[0,T] then

P(Xt =Yt, t∈[0, T]) = 1.

Remark 1.6.4. From [13] (proof of Theorem 6.2.1) we get that E

"

sup

t∈[0,T]

|Xt|2

#

<∞.

We conclude this section with some useful inequalities called Burkholder-Davis- Gundy inequality and H¨older inequality.

(17)

1.6. STOCHASTIC DIFFERENTIAL EQUATIONS 13

Proposition 1.6.5 (Burkholder-Davis-Gundy, ([10]Theorem 4.3.1)). Let L ∈ Lloc2 . Then for any 0< p <∞ there exist constants αp, βp >0 such that

βp

s Z T

0

L2tdt p

sup

t∈[0,T]

Z t 0

LsdWs p

≤αp

s Z T

0

L2tdt p

.

Moreover, αp ≤c√

p for 2≤p <∞ and for some constant c >0.

The H¨older inequality we state for a probability space.

Proposition 1.6.6 (H¨older, ([7] Proposition 5.10.5)). Let (Ω,F,P) be a proba- bility space and X, Y : Ω → R random variables. If 1 < p, q < ∞ and 1p + 1q = 1, then

E[|XY|]≤(E[Xp])1p(E[Yq])1q .

(18)

CHAPTER 2

Malliavin calculus

2.1. A multiple stochastic integral

Next we will follow Nualart [18] and give the basics of Malliavin calculus.

Definition 2.1.1. Let (Ω,F,P,(Ft)t∈[0,T]) be a filtered probability space and H a real and separable Hilbert space with scalar product h·,·iH and norm k · kH. We consider a stochastic process indexed by the elements of H:

W ={W(h);h∈H}.

We say that this process is an isonormal Gaussian process if W is a Gaussian family of random variables such that for all h, g ∈H

E[W(h)W(g)] =hh, giH and

E[W(h)] = 0.

In this thesis we realize such an isonormal Gaussian process as follows. We assume a probability space (Ω,F,P) carrying a Brownian motionW = (Wt)t∈[0,T]and a special Hilbert space with

L2 [0, T],B([0, T]), λ

= (

f : [0, T]→R; Z T

0

f(t)2dλ(t) 12

<∞ )

.

For allf, g∈L2([0, T]), we letW(f) =RT

0 f(s)dWsandhf, giL2([0,T]) =E[W(f)W(g)].

Moreover, for all A ∈ B([0, T]), we let W(A) = W(1A). In this way, we have an isonormal Gaussian process as we defined above.

Next we define a set of special elementary functions, which will vanish on diago- nals.

Definition 2.1.2. LetA1, . . . , An∈B([0, T]) such thatAk∩Al=∅for allk 6=l, where k, l∈ {1, . . . , n}. Then we define

Em =

f : [0, T]m →R;f(t1, . . . , tm) = Pn

i1,...,im=1ai1···im1Ai1×···×Aim(t1, . . . , tm) where ai ∈R and ai1···im = 0, if ik =ij for some k6=j.

We define a multiple stochastic integral first for these elementary functions and after that for all functions in L2([0, T]m).

Definition 2.1.3. For f ∈ Em given as above a multiple stochastic integral is defined as

Im(f) =

n

X

i1,...,im=1

ai1···imW(Ai1)· · ·W(Aim).

14

(19)

2.1. A MULTIPLE STOCHASTIC INTEGRAL 15

Remark 2.1.4. One can easily check that this definition is independent from the representation of f. Hence Im is well-defined.

This defined multiple stochastic integral has three important properties. Before we state the properties, we introduce the symmetrization ˜f of a function f. This is

f(t˜ 1, . . . , tm) = 1 m!

X

σ∈Sm

f(tσ(1), . . . , tσ(m)), where Sm is the set of all permutations of {1, . . . , m}.

Proposition 2.1.5. A multiple stochastic integral with integrands from Em has the following properties

(i) Let f, g∈Em and α, β ∈R. Then

Im(αf +βg) = αIm(f) +βIm(g).

(ii) If f˜is symmetrization of f, then

Im( ˜f) =Im(f).

(iii) If f ∈ Em and g ∈ En, then for the product of multiple stochastic integrals it holds

E[Im(f)In(g)] =

( 0 if m6=n m!hf ,˜ ˜giL2

[0,T]m if m=n.

Proof. (i) We can assume that f and g have the same partition A1, . . . , An, because if not, we can make it to be the same by using intersections of sets.

Now

αIm(f) +βIm(g) = α

n

X

i1,...,im=1

ai1···imW(Ai1). . . W(Aim)

n

X

i1,...,im=1

bi1···imW(Ai1). . . W(Aim)

=

n

X

i1,...,im=1

[αai1···im+βbi1···im]W(Ai1). . . W(Aim)

= Im(αf +βg) (ii) Because of (i) we can assume a function

f(t1, . . . , tm) =1Ai1×···×Aim(t1, . . . , tm).

For this we have by definition

Im(f) = W(Ai1). . . W(Aim), and for the symmetrization we have

Im( ˜f) = 1 m!

X

σ∈Sm

W(Aσ(1)). . . W(Aσ(m)).

(20)

2.1. A MULTIPLE STOCHASTIC INTEGRAL 16

Now for all permutations σ we can change order in the product such that we always get

W(Aσ(1)). . . W(Aσ(m)) =W(Ai1). . . W(Aim).

By that we have

Im( ˜f) = 1

m!m!W(Ai1). . . W(Aim) = Im(f).

(iii) Let f ∈ Em and g ∈ Ek. Because of (i) and (ii), we can assume that both functions are symmetric and have the same partition A1, . . . , An. We have now

f(t1, . . . , tm) =

n

X

i1,...,im=1

ai1···im1Ai1×···×Aim(t1, . . . , tm) and

g(t1, . . . , tk) =

n

X

j1,...,jk=1

bj1···jk1Aj1×···×Ajk(t1, . . . , tk).

Because the functions are symmetric, we have for all permutations ai1···im =aσ(i1)···σ(im) and bj1···jk =bσ(j1)···σ(jk),

and because we have m! permutations for the functionf and k! forg, we get Im(f) =m! X

i1<···<im

ai1···imW(Ai1). . . W(Aim) and

Ik(g) = k! X

j1<···<jk

bj1···jkW(Aj1). . . W(Ajk).

With these we get

E[Im(f)Ik(g)] = E

"

m! X

i1<···<im

ai1···imW(Ai1). . . W(Aim)

×k! X

j1<···<jk

bj1···jkW(Aj1). . . W(Ajk)

#

= m!k! X

i1<···<im

X

j1<···<jk

ai1···imbj1···jk

×E[W(Ai1). . . W(Aim)W(Aj1). . . W(Ajk)].

By Definition 2.1.2 we have Ai∩Aj =∅ for all i6=j so this implies thatW(Ai) and W(Aj) are independet for all i6=j. We have now two possibilities:

E[W(Ai1)...W(Aim)W(Aj1)...W(Ajk)] =





E[W(Ai1)2]...E[W(Aim)2], if m =k and iq =jq for all q ∈ {1,2, ..., m}, 0, all other cases.

So we conclude ifm 6=k

E[Im(f)Ik(g)] = 0.

(21)

2.1. A MULTIPLE STOCHASTIC INTEGRAL 17

And for m =k we get by Itˆo’s isometry and the fact that E[W(Ai1). . . W(Aim)W(Aj1). . . W(Ajm)] = 0 unless iq =jq for all q ∈ {1,2, . . . , m} the following:

E[Im(f)Im(g)]

= E

"

m! X

i1<···<im

ai1···imW(Ai1). . . W(Aim)m! X

j1<···<jm

bj1···jmW(Aj1). . . W(Ajm)

#

= E

"

(m!)2 X

i1<···<im

ai1···imbi1···imW(Ai1)2. . . W(A1m)2

#

= (m!)2 X

i1<···<im

ai1···imbi1···imE

W(Ai1)2 . . .E

W(Aim)2

= (m!)2 X

i1<···<im

ai1···imbi1···imλ(Ai1). . . λ(Aim)

= m!

n

X

i1,...,im=1

ai1···imbi1···imλ(Ai1). . . λ(Aim)

= m!

Z T 0

· · · Z T

0

f(t1, . . . , tm)g(t1, . . . , tm)dλ(t1). . . dλ(tm)

= m!hf, giL2([0,T]m).

And because we assumed the functions f and g to be symmetric, we have E[Im(f)Im(g)] =E

h

Im( ˜f)Im(˜g)i

=m!hf ,˜ ˜giL2([0,T]m).

We also notice that if we have a symmetric function f ∈ Em, it holds for all permutations σ ∈Sm

Z

[0,T]m

|f(t1, . . . , tm)|2m = Z

[0,T]m

|f(tσ(1), . . . , tσ(m))|2m, and thanks to the triangle inequality, this gives us

kfk˜ L2([0,T]m) = k 1 m!

X

σ∈Sm

fkL2([0,Tm])

≤ 1 m!

X

σ∈Sm

kfkL2([0,T]m)

= kfkL2([0,T]m). By that we conclude the inequality

(2.1) kf˜kL2([0,T]m) ≤ kfkL2([0,T]m).

Our next step is to extend the multiple stochastic integral to all functions in L2([0, T]m).

Proposition 2.1.6. The set Em is dense in L2([0, T]m,B([0, T]m), λm).

(22)

2.1. A MULTIPLE STOCHASTIC INTEGRAL 18

Proof. We do the proof in three steps.

Step 1: Let > 0. First we show that we can approximate every 1A, where A =A1× · · · ×Am and Ai ∈B([0, T]) for all i∈ {1, . . . , m}, by functions from Em. Because the Lebesgue measure λ has no atoms, we can find for all A∈ B([0, T]m) a measurable set B ⊂A such that

0< λ(B)< λ(A).

Now let ˜B ={B1, . . . , Bn} ⊂B([0, T]), whereBj∩Bk =∅ ifj 6=k andλ(Bj)<

for all i = 1, . . . , n. We choose ˜B such that we can express every Ai as union of Bj ∈B. Since˜ A∈[0, T]m, we letλm(A) = Πmi=1λ(Ai) =α and put i1,···,im be 0 or 1.

In this case we can write 1A=

n

X

i1,...,im=1

i1,···,im1Bi1×···×Bim.

If we define a setI which includes all (i1, . . . , im) wherei1, . . . , im are all different and put Ic =J, we get

1B= X

(i1,...,im)∈I

i1,···,im1Bi1×···×Bim

and also 1B ∈ Em. Because in the set J are at least two of the i1, . . . , im equal, we get

k1A−1Bk2L2([0,T]m) = k X

(i1,...,im)∈J

i1,···,im1Bi1×···×Bimk2L2([0,T]m)

= X

(i1,...,im)∈J

i1,···,imλ(Bi1)...λ(Bim)

≤ X

(i1,...,im)∈J

λ(Bi1)...λ(Bim)

= m

2 n

X

j=1

(λ(Bj))2

n

X

i=1

λ(Bi)m−2

≤ m

2

n

X

i=1

λ(Bi)m−1

≤ m

2

αm−1

→ 0, if →0.

Step 2: Next we show that every bounded function f ∈ L2([0, T]m) can be approximated by simple functions which are defined by using the sets of the form A =A1× · · · ×Am. We use the Monotone class theorem for functions (Proposition 1.0.3). First let H be the set which includes all bounded and measurable functions f

Viittaukset

LIITTYVÄT TIEDOSTOT

We have discussed risk as the basis for a cost sharing scheme. When stakeholder participation in financing the risk is risk-based, it is possible to control for the

When the forward price is an unbiased estimate for the price at harvest, the speculative compo- nent drops out and the optimal hedge under yield uncertainty depends on the

In the theoretical part of the work, we (i) derived the within-cross-section bias, approxi- mative variance and true variance for the area estimators based on the circle area formula

Essay 1: Dividends, costly external capital, and firm value: the case of constant scale This essay studies optimal dividend and capital raising policies for a constant scale firm

The number of risk days for heavy snow loading on the basis of daily mean values of different weather variables proved to predict well the number of days with modelled heavy snow

Parameters in the yield equation (17).. prices at harvest. The Pearson’s correlation test suggests that the correlation coefficient is statistically signifi- cant at

This test is motivated to find differences between hedging decision when a firm is exposed to foreign exchange risk and has decided not to hedge and with firms that do not

This paper employs the outbreak of the pandemic as a quasi-experimental design to explore the hedging abilities of Bitcoin – which has been referred to as ‘virtual