• Keine Ergebnisse gefunden

Examination of the closedness of spaces of stochastic integrals

N/A
N/A
Protected

Academic year: 2021

Aktie "Examination of the closedness of spaces of stochastic integrals"

Copied!
135
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Examination of the Closedness of Spaces of

Stochastic Integrals

Von der Fakultät für Mathematik und Naturwissenschaften der Carl von Ossietzky Universität Oldenburg zur Erlangung des Grades und Titels eines

Doktors der Naturwissenschaften (Dr. rer. nat.)

angenommene Dissertation

von Herrn Jörg Thomas Best geboren am 06.08.1988 in Aurich.

(2)

Betreuende Gutachterin: Prof. Dr. Angelika May Zweitgutachter: Prof. Dr. Marcus Christiansen Tag der Disputation: 29.06.2020

(3)
(4)

Abstract

We study stochastic hedging concepts for a general semimartingale model in incomplete markets. To ensure the existence of solutions, one wants the spaces of terminal values of stochastic integrals R ϑdX (representing the cumulative gains from trade) to lie in a closed subset of L2. In addition, we consider constrained

hedging strategies with the notion of correspondences (i.e. the trading strategies are restricted to lie in a predened set) and ensure the according closedness as well. This is a main contribution of this work. Using the closedness and a change of the according probability measure (using a so-called variance-minimizing martingale measure ˜P) to transfer the problem to a (local) martingale setting then allows us, to determine explicit hedging strategies for the respective spaces.

(5)
(6)

Zusammenfassung

In der vorliegenden Dissertation beschäftigen wir uns mit stochastischen

Hedgingkonzepten in einem allgemeinen Semimartingal Modell. Um die Existenz einer Lösung zu gewährleisten, ist es naheliegend, zunächst die L2-Abgeschlossenheit des

Raumes der Endwerte stochastischer Integrale R ϑdX (welche die kummulierten Gewinne einer Handelsstrategie repräsentieren) sicherzustellen. Zusätzlich betrachten wir Han-delsbeschränkungen mit Hilfe von Korrespondenzen (d.h., dass die Handelsstrategien in soweit eingeschränkt sind, dass sie in eine vorher festgelegte Menge abbilden müssen) und untersuchen auch in diesem Kontext die entsprechende L2-Abgeschlossenheit. Dies

ist ein wesentlicher Beitrag dieser Arbeit.

Unter Verwendung der L2-Abgeschlossenheit und eines entsprechenden Maÿwechsels

hin zu dem sogenannten varianzminimierenden lokalen Martingalmaÿ ˜P, mit dem wir unser Hedgingproblem in ein (lokales) Martingal-Setting überführen, bestimmen wir abschlieÿend explizite Hedgingstrategien für die entsprechenden Räume.

(7)
(8)

Contents

1 Introduction 1

1.1 Motivation . . . 1

1.2 Theoretical Background and Objective . . . 2

1.3 Outline . . . 3

2 Preliminaries 4 2.1 Stochastic Processes . . . 4

2.2 Martingales . . . 6

2.3 The Quadratic Variation of a Semimartingale . . . 13

2.4 Compensators . . . 14

2.5 The Bichteler-Dellacherie Theorem . . . 15

2.6 Local Martingales . . . 16

3 Financial Market Model 19 4 Examination of the Closedness of Spaces of Stochastic Integrals in L2 22 4.1 The Space GT(Θeqp ) . . . 22

4.2 The Space GT(Θv) . . . 36

4.3 The Space GT(Θ) . . . 43

5 Examination of the Closedness of Spaces of Stochastic Integrals in L2 under Convex Trading Constraints 56 5.1 The Space GT(ΘS(C)) . . . 56

5.2 The Space GT(Θeq2 (C)) . . . 66

5.3 The Space GT(Θv(C)) . . . 72

6 Calculation of Explicit Hedging Strategies 77 6.1 The space GT(ΘS) . . . 79 6.2 The Space GT(Θeq2 ) . . . 92 6.3 The Space GT(Θv) . . . 103 7 Conclusion 115 8 Outlook 117 9 List of Notation 119

(9)
(10)

1 Introduction

1.1 Motivation

Due to recent crises at global nancial markets and thereupon strengthened regulatory requirements for investments, the question of how reassuring oneself against potential losses is of utmost importance. Having a profound look on current hedging strategies, one recognizes that although much work on this eld has already been done, developing hedging strategies for a huge variety of investments is continuously one of the main issues of research in Financial Mathematics. Impressively, the answers to very fundamental questions on how to protect oneself from nancial losses have not been given completely yet. For example, the following and very intuitive question stays to be of concern:

Having sold a European contingent claim, how does one secure oneself against the risk that may arise from the sale in the future?

One of the reasons why this question is quite dicult to answer is probably that one wants to work in incomplete markets, this means that we have to pay attention to claims that are not replicable. This way, it is not always possible to create a portfolio that replicates the payo of the sold claim exactly. In this context our question turns into an optimization problem, where we have to solve the problem of minimizing the incurring hedging error. In addition, the answer highly depends on the chosen nancial model, i.e. dierent conditions on price processes and trading strategies may as well very likely lead to dierent results for according hedging strategies.

Facing the above question in a mathematical way, the claim can be modelled as a square integrable random variable H describing a certain payo that is generated at time T . In addition, each terminal portfolio value achieved by a self-nancing trading strategy θ can be represented by a stochastic integral R ϑdX, whereas X describes the price process of the underlying stock. For a non-replicable claim, hedging can now be interpreted as nding the terminal portfolio value, that is closest (i.e. minimizing the L2-norm) to the random variable H, which models the claim. In the Hilbert space L2, this means to orthogonally project the claim onto the subset of L2 of terminal portfolio

values. Clearly, the respective subspaces of stochastic integrals should be closed in L2.

In this thesis, for subspaces modelling terminal portfolio values we consider the three dierent spaces GT(ΘS), GT(Θv) and GT(Θeqp ).

(11)

sub-spaces, as well for unrestricted as for restricted trading strategies, and to nd risk-minimizing strategies for all of these subspaces. Furthermore, to weaken the depen-dency of the chosen nancial market model, we provide conditions under which as well the respective hedging strategies as the respective subspaces coincide.

1.2 Theoretical Background and Objective

Subsequently, we want to give a brief overview about what has already been established in the literature and what this work contributes to the current research. We will specify this in the respective sections.

The space GT(ΘS)has been established by Delbaen/Monat/Stricker/Schweizer/

Schachermayer (1991) who prove the L2-closedness in the case that the price process

X is a continuous semimartingale, where it turns out that their results do not hold in the discontinuous case. Delbaen et al. (1997) then found necessary and sucient conditions for the closedness. Further results have been found by Grandits/Krawczyk (1998) and by Choulli/Krawczyk/Stricker (1999). The problem of nding explicit hedg-ing strategies has been worked on by Schweizer (1996), Rheinlaender/Schweizer (1997), Hou/Karatzas (2005) and Arai (2005). In addition, GT(ΘS)is the only space where the

closedness under trading constraints already has been established. This can be found in Czichowsky/Schweizer (2012).

For GT(ΘS), the contribution of this work will be providing a counterexample for the

closedness in the discontinuous case and subsequently providing conditions to ensure the closedness in the discontinuous case, to deepen the results on the closedness of GT(ΘS(C)), i.e the closedness of GT(ΘS) under trading constraints and to specify the

shape of the hedging strategy.

GT(Θv) has been established in Xia/Yan (2006). They provide conditions for the

closedness and develop a duality between the closedness of GT(Θv)and signed martingale

measures. Concerning the explicit hedging strategy, they consider the problem of nding mean-variance ecient portfolios. We expand those results by solving the more general mean-variance hedging problem.

GT(Θeqp ) is implicitly developed in Delbaen/Schachermayer (1996) whereas the

cor-responding hedging strategy for solving the mean-variance hedging problem is worked on in Gouriéroux/Laurent/Pham (1998). We will expand the results on that hedging strategy and in addition solve another type of hedging problem similar to the problem of nding mean-variance ecient portfolios. We also provide conditions under which GT(Θeqp ) and GT(Θv) coincide and solve the closedness question under constraints for

(12)

these spaces.

In conclusion, under special conditions we will establish the closedness with and with-out constraints, solve the mean-variance hedging problem and the problem of nding mean-variance ecient portfolios for all three spaces and draw a variety of relationships between them.

1.3 Outline

The thesis is divided into ve main sections. In section 2 and 3, we establish the necessary mathematical framework, where we provide fundamental results from stochastic analysis and dene the nancial market model which we will work on. Section 4 is dedicated to establish the L2-closedness of all of our three subspaces which will enable us in section 6

to nd hedging strategies by projecting the non-replicable claim onto our closed subsets. In addition, in section 5 we analyze the L2-closedness under trading constraints

im-posed on our strategies ϑ ∈ Θ and the respective closedness of the occurring spaces GT(Θ) of stochastic integrals. In section 6, we use those results from section 4 and 5

to nd the solutions to dierent hedging problems by using Hilbert space methods. In sections 4-6, we will not only state results for each space separately, but also develop connections between them and transfer conclusions from one space to another.

(13)

2 Preliminaries

Before we can start working on our actual topic, namely the examination of the closed-ness of spaces consisting of stochastic integrals representing terminal portfolio values, we will rst provide some basic denitions used in the context of stochastic analysis. The aim of this section is to clarify the setting in terms of the probability space that we work on and to give the reader exact denitions of the dierent stochastic processes that we will use during this work. Here, the latter case is of utmost importance since those denitions are not uniformly given in the literature and all our results may only be valid given this exact set of denitions and framework. Particularly, the dierent types of martingales and how they are connected shall be pointed out in this section, as they will be used to model price processes of stocks and therefore will play a very important role throughout all of this work.

The section will only give a brief overview to establish all of the necessary framework for our thesis. For further information the reader may have a look at Protter (2005) which we clearly orientate us on. Especially for the construction and properties of the stochastic integral, we refer to it.

2.1 Stochastic Processes

We assume as given a complete probability space (Ω, F, P ) and a ltration (Ft)0≤t≤∞.

By a ltration we mean a family of σ-algebras (Ft)0≤t≤∞ that is increasing, i.e. Fs⊂ Ft

if s ≤ t. For convenience, we will usually write F for the ltration (Ft)0≤t≤∞.

Denition 1. A ltered complete probability space (Ω, F, F, P ) is said to satisfy the usual hypothesis if

F0 contains all the P -null sets of F;

Ft=Tu>tFu, all t, 0 ≤ t ≤ ∞; that is, the ltration F is right continuous.

We always assume that the usual hypothesis holds.

Denition 2. A random variable T : Ω → [0, ∞] is a stopping time if the event {T ≤ t} ∈ Ft, for every t, 0 ≤ t ≤ ∞.

One important consequence of the right continuity of the ltration is the following theorem.

(14)

A stochastic process X on (Ω, F, P ) is a collection of R-valued or Rd-valued random

variables (Xt)0≤t≤∞. The process X is said to be adapted if Xt ∈ Ft (that is, is Ft

measurable) for each t. We must take care to be precise about the concept of equality of two stochastic processes.

Denition 3. Two stochastic processes X and Y are modications if Xt= Yt a.s., each

t. Two processes X and Y are indistinguishable if a.s., for all t, Xt= Yt.

Remark 1. If X and Y are modications there exists a null set, Nt, such that if ω /∈ Nt,

then Xt(ω) = Yt(ω). The null set Nt depends on t.

Denition 4. A stochastic process X is said to be càdlàg if it a.s. has sample paths which are right continuous, with left limits. Similarly, a stochastic process X is said to be càglàd if it a.s. has sample paths which are left continuous, with right limits.

(The nonsensical words càdlàg and càglàd are acronyms from the French for continu à droite, limité à gauche and continu à gauche, limité à droite, respectively.)

Theorem 2. Let X and Y be two stochastic processes, with X a modication of Y . If X and Y have right continuous paths a.s., then X and Y are indistinguishable.

Corollary 1. Let X and Y be two stochastic processes which are càdlàg. If X is a modication of Y , then X and Y are indistinguishable.

Càdlàg processes provide natural examples of stopping times.

Denition 5. Let X be a stochastic process and let Λ be Borel set in R. Dene

T (ω) = inf{t > 0 : Xt∈ Λ}.

Then T is called the hitting time of Λ for X.

Theorem 3. Let X be an adapted càdlàg stochastic process, and let Λ be an open set. Then the hitting time of Λ is a stopping time.

Theorem 4. Let X be an adapted càdlàg stochastic process, and let Λ be a closed set. Then the random variable

T (ω) = inf{t > 0 : Xt(ω) ∈ Λ or Xt−(ω) ∈ Λ}

(15)

Theorem 5. Let S, T be stopping times. Then the following are stopping times: (i) S ∧ T = min(S, T );

(ii) S ∨ T = max(S, T ); (iii) S + T ;

(iv) αS, where α > 1.

The σ-algebra Ftcan be thought of as representing all (theoretically) observable events

up to and including time t. We would like to have an analogous notion of events that are observable before a random time.

Denition 6. Let T be a stopping time. The stopping time σ-algebra FT is dened to

be

{Λ ∈ F : Λ ∩ {T ≤ t} ∈ Ft, all t ≥ 0}.

The previous denition is not especially intuitive. However it does well represent 'knowl-edge' up to time T , as the next theorem illustrates.

Theorem 6. Let T be a nite stopping time. Then FT is the smallest σ-algebra

con-taining all càdlàg adapted processes sampled at T . That is,

FT = σ{XT; X all adapted càdlàg processes}.

Theorem 7. Let X be adapted and càdlàg. If ∆XT1{T <∞} = 0 a.s. for each stopping

time T , then ∆X is indistinguishable from the zero process.

2.2 Martingales

In this section we give, mostly without proofs, only the essential results from the theory of continuous time martingales.

Also, recall that we will always assume as given a ltered, complete probability space (Ω, F , F, P ), where the ltration F = (Ft)0≤t≤∞ is assumed to be right continuous.

Denition 7. A real-valued, adapted process X = (Xt)0≤t<∞ is called a martingale

(resp. supermartingale, submartingale) with respect to the ltration F if (i) Xt∈ L1(dP ); that is, E{|Xt|} < ∞;

(16)

(ii) if s ≤ t, then E{Xt|Fs} = Xs, a.s. (resp. E{Xt|Fs} ≤ Xs, resp. ≥ Xs).

Note that martingales are only dened on [0, ∞); that is, for nite t and not t = ∞. It is often possible to extend the denition to t = ∞.

Denition 8. A martingale X is said to be closed by a random variable Y if E{|Y |} < ∞ and Xt = E{Y |Ft}, 0 ≤ t < ∞.

Theorem 8. Let X be a supermartingale. The function t 7→ E{Xt} is right continuous

if and only if there exists a modication Y of X which is càdlàg. Such a modication is unique.

Corollary 2. If X = (Xt)0≤t<∞ is a martingale then there exists a unique modication

Y of X which is càdlàg.

Denition 9. A family of random variables (Uα)α∈A is uniformly integrable if

lim

n→∞supα

Z

{|Uα|≥n}

|Uα|dP = 0.

Theorem 9. Let (Uα)α∈A be a subset of L1. The following are equivalent:

(i) (Uα)α∈A is uniformly integrable.

(ii) supα∈AE{|Uα|} < ∞, and for every  > 0 there exists δ > 0 such that Λ ∈ F,

P (Λ) ≤ δ, imply E{|Uα1Λ|} < .

(iii) There exists a positive, increasing, convex function G(x) dened on [0, ∞) such that limx→∞G(x)x = +∞.

Theorem 10. Let X be a right continuous martingale which is uniformly integrable. Then Y = limt→∞Xt a.s. exists, E{|Y |} < ∞, and Y closes X as a martingale.

Theorem 11. Let X be a (right continuous) martingale. Then (Xt)t≥0 is closed if and

only if (Xt)t≥0 is uniformly integrable, and if and only if Y = limt→∞Xt exists a.s.,

E{|Y |} < ∞, and (Xt)0≤t≤∞ is a martingale, where X∞ = Y.

Remark 2. If X is a uniformly integrable martingale, then Xt converges to X∞ = Y in

L1 as well as almost surely.

Denition 10. Let X be a stochastic process and let T be a random time. XT is said

to be the process stopped at T if XT

(17)

Note that if X is adapted and càdlàg and if T is a stopping time, then XtT = Xt∧T = Xt1{t<T }+ XT1{t≥T }

is also adapted. A martingale stopped at a stopping time is still a martingale, as the next theorem shows; see its corollary.

Theorem 12. Let X be a uniformly integrable right continuous martingale, and let T be a stopping time. Then XT = (X

t∧T)0≤t≤∞ is also a uniformly integrable right continuous

martingale.

Corollary 3. Let M be a martingale, and T a nite valued stopping time. Then MT,

the martingale stopped at T , is still a martingale.

Corollary 4. Let Y be an integrable random variable and let S, T be stopping times. Then

E{E{Y |FS}|FT} = E{E{Y |FT}|FS}

= E{Y |FS∧T}.

The next inequality is elementary, but indispensable. Theorem 13. (Jensen`s Inequality)

Let φ : R → R be convex, and let X and φ(X) be integrable random variables. For any σ-algebra G,

φ ◦ E{X|G} ≤ E{φ(X)|G}.

Corollary 5. Let X be a martingale, and let φ be convex such that φ(Xt) is integrable,

0 ≤ t < ∞. Then φ(X) is a submartingale. In particular, if M is a martingale, then |M | is a submartingale.

Corollary 6. Let X be a submartingale and let φ be convex, non-decreasing, and such that φ(Xt)0≤t<∞ is integrable. Then φ(X) is also a submartingale.

Theorem 14. Let X be a positive submartingale. For all p > 1, with q conjugate to p (i.e. 1 p + 1 q = 1), we have || sup t Xt||Lp ≤ q sup t ||Xt||Lp.

(18)

For a real valued process, we let X∗ denote sup

s|Xs|. Note that if M is a martingale

with M∞∈ L2, then |M| is a positive submartingale, and taking p = 2 we have

E{(M∗)2} ≤ 4E{M2 ∞}.

This last inequality is called Doob's maximal quadratic inequality. Denition 11. An adapted counting process N is a Poisson process if (i) for any s, t, 0 ≤ s < t < ∞, Nt− Ns is independent of Fs;

(ii) for any s, t, u, v, 0 ≤ s < t < ∞, 0 ≤ u < v < ∞, t−s = v −u, then the distribution of Nt− Ns is the same as that of Nv − Nu.

Properties (i) and (ii) are known respectively as increments independent of the past, and stationary increments.

Denition 12. An adapted process B = (Bt)0≤t<∞ taking values in Rn is called an

n-dimensional F Brownian motion if

(i) for 0 ≤ s < t < ∞, Bt− Bs is independent of Fs (increments are independent of the

past);

(ii) for 0 < s < t, Bt− Bs is a Gaussian random variable with mean zero and variance

matrix (t − s)C, for a given, non-random matrix C. The Brownian motion starts at x if P (B0 = x) = 1.

Denition 13. An adapted process X = (Xt)t≥0 with X0 = 0 a.s. is a Lévy process if

(i) X has increments independent of the past; that is, Xt − Xs is independent of Fs,

0 ≤ s < t < ∞; and

(ii) X has stationary increments; that is, Xt− Xs has the same distribution as Xt−s,

0 ≤ s < t < ∞; and

(iii) Xt is continuous in probability; that is, limt→sXt = Xs, where the limit is taken in

probability.

Denition 14. A process H is said to be simple predictable if H has a representation Ht = H01{0}(t) +

n

X

i=1

Hi1(Ti,Ti+1](t)

where 0 = T1 ≤ · · · ≤ Tn+1 < ∞ is a nite sequence of stopping times, Hi ∈ FTi with

(19)

Note that we can take T1 = T0 = 0 in the above denition, so there is no 'gap'

be-tween T0 and T1. We can topologize S by uniform convergence in (t, ω), and we denote

S endowed with this topology by Su. We also write L0 for the space of nite-valued

random variables topologized by convergence in probability.

Let X be a stochastic process. An operator IX, induced by X should have two

funda-mental properties to earn the name 'integral'. The operator IX should be linear, and it

should satisfy some version of the Bounded Convergence Theorem. A particularly weak form of the Bounded Convergence Theorem is that the uniform convergence of processes Hn to H implies only the convergence in probability of IX(Hn)to IX(H).

Inspired by the above considerations, for a given process X we dene a linear mapping IX : S → L0 by letting IX(H) = H0X0+ n X i=1 Hi(XTi+1− XTi),

where H ∈ S has the representation

Ht= H01{0}+ n

X

i=1

Hi1(Ti,Ti+1].

Since this denition is a path-by-path denition for the step functions H(ω), it does not depend on the choice of the representation of H in S.

Denition 15. A process X is a total semimartingale if X is càdlàg, adapted, and IX : Su → L0 is continuous.

Recall that for a process X and a stopping time T , the notation XT denotes the

process (Xt∧T)t≥0.

Denition 16. A process X is called a semimartingale if, for each t ∈ [0, ∞), Xt is a

total semimartingale.

Denition 17. An adapted, càdlàg process A is a nite variation process (FV) if almost surely the paths of A are of nite variation on each compact interval of [0, ∞). We write R∞

0 |dAs| or |A|∞ for the random variable which is the total variation of the paths of A.

Theorem 15. Let X be a semimartingale. If X has a decomposition Xt = X0+ Mt+ At

(20)

with M a local martingale and A a predictably measurable FV process, M0 = A0 = 0,

then such a decomposition is unique.

Theorem 16. The set of (total) semimartingales is a vector space.

Theorem 17. If Q is a probability and absolutely continuous with respect to P , then every (total) P semimartingale X is a (total) Q semimartingale.

Theorem 18. Let (Pk)k≥1 be sequence of probabilities such that X is a Pk

semimartin-gale for each k. Let R = P∞

k=1λkPk, where λk ≥ 0, each k, and P ∞

k=1λk = 1. Then X

is a semimartingale under R as well. Theorem 19. (Stricker's Theorem)

Let X be a semimartingale for the ltration F. Let G be a subltration of F, such that X is adapted to the G ltration. Then X is a G semimartingale.

Theorem 20. Each square integrable martingale with càdlàg paths is a semimartingale. Corollary 7. Each càdlàg, locally square integrable local martingale is a semimartingale. Corollary 8. A local martingale with continuous paths is a semimartingale.

Corollary 9. The Wiener process is a semimartingale.

Denition 18. An adapted, càdlàg process X is decomposable if there exist processes N, A such that

Xt= X0+ Nt+ At

with N0 = A0 = 0, N a locally square integrable local martingale, and A an FV process.

Theorem 21. A decomposable process is a semimartingale. Corollary 10. A Lévy process is a semimartingale.

Denition 19. Let A = [a, b] be an interval. The variation of paths of a process B is dened to be VA(ω) = sup π∈P X ti∈π |Bti+1− Bti|

(21)

Denition 20. An adapted, càdlàg process Y is a classical semimartingale if there exist processes N, B with N0 = B0 = 0 such that

Yt= Y0+ Nt+ Bt

where N is a local martingale and B is an FV process.

Theorem 22. Let X be an adapted, càdlàg process. The following are equivalent: (i) X is a semimartingale;

(ii) X is decomposable;

(iii) given β > 0, there exist M, A with M0 = A0 = 0, M a local martingale with jumps

bounded by β, A an FV process, such that Xt = X0+ Mt+ At;

(iv) X is a classical semimartingale

Denition 21. Let X be a semimartingale. If X has a decomposition Xt = X0+ Mt+ At

with M0 = A0 = 0, M a local martingale, A an FV process, and with A predictable, then

X is said to be a special semimartingale.

Denition 22. If X is a special semimartingale, then the unique decomposition X = M + A

with M0 = X0 and A0 = 0 and A predictable is called the canonical decomposition.

Theorem 23. If X is a special semimartingale, then its decomposition X = M + A

with A predictable is unique.

Denition 23. Let X be a stochastic process. A property π is said to hold locally if there exists a sequence of stopping times (Tn)n≥1 increasing to ∞ a.s. such that XTn1{Tn>0}

has property π each n ≥ 1.

Theorem 24. Let X be a semimartingale. X is special if and only if the process Xt∗ = sup

s≤t

(22)

is locally integrable.

Denition 24. An adapted, càdlàg process X is a quasimartingale on [0, ∞] if E[|Xt|] <

∞, for each t, and if V ar[X] < ∞. Theorem 25. (Rao's Theorem)

A quasimartingale X has a unique decomposition X = M + A, where M is a local martingale and A is a predictable process with paths of locally integrable variation and A0 = 0.

2.3 The Quadratic Variation of a Semimartingale

The quadratic variation process of a semimartingale, also known as the bracket process, is a simple object that nevertheless plays a fundamental role.

Denition 25. Let X, Y be semimartingales. The quadratic variation process of X, denoted [X, X] = ([X, X]t)t≥0, is dened by

[X, X] = X2− 2 Z

X−dX

(recall that X0− = 0). The quadratic covariation of X, Y , also called the bracket process

of X, Y , is dened by [X, Y ] = XY − Z X−dY − Z Y−dX.

It is clear that the operation (X, Y ) → [X, Y ] is bilinear and symmetric. We therefore have a polarization identity

[X, Y ] = 1

2([X + Y, X + Y ] − [X, X] − [Y, Y ]).

The next theorem gives some elementary properties of [X, X]. (X is assumed to be a given semimartingale throughout this section).

Theorem 26. The quadratic variation process of X is a càdlàg, increasing, adapted process. Moreover, it satises the following.

(i) [X, X]0 = X02 and ∆[X, X] = (∆X)2.

(ii) If T is a stopping time, then

(23)

Corollary 11. The bracket process [X, Y ] of two semimartingales has paths of nite variation on compacts, and it is also a semimartingale.

Corollary 12. (Integration by parts)

Let X, Y be two semimartingales. Then XY is a semimartingale and XY =

Z

X−dY +

Z

Y−dX + [X, Y ].

In the integration by parts formula above, we have (X−)0 = (Y−)0 = 0. Hence

evaluating at 0 yields

X0Y0 = (X−)0Y0+ (Y−)0X0+ [X, Y ]0.

Since [X, Y ]0 = ∆X0∆Y0 = X0Y0, the formula is valid. Without the convention that

(X−)0 = 0, we could have written the formula

XtYt = Z t 0+ Xs−dYs+ Z t 0+ Ys−dXs+ [X, Y ]t.

2.4 Compensators

Let A be a process of locally integrable variation, hence a FV process. A is then locally a quasimartingale, and hence by Rao's Theorem, there exists a unique decomposition

A = M + ˜A

where ˜A is a predictable FV process. In other words, there exists a unique, predictable FV process ˜A such that A − ˜A is a local martingale.

Denition 26. Let A be an FV process with A0 = 0, with locally integrable total

varia-tion. The unique FV predictable process ˜A such that A− ˜A is a local martingale is called the compensator of A.

Theorem 27. Let A be an increasing process of integrable variation, and let M be a bounded martingale. The compensator of Rt

0 MsdAs is R t

0 Ms−dAs.

Denition 27. Let X be a semimartingale such that its quadratic variation process [X, X] is locally integrable, that is, there exist stopping times Tn increasing to ∞ a.s.

(24)

denoted < X, X >= (< X, X >t)t≥0, exists and it is dened to be the compensator of

[X, X]. That is < X, X >= ^[X, X].

If X is a continuous semimartingale then [X, X] is also continuous and hence already predictable; thus

[X, X] =< X, X >

when X is continuous. In particular for a standard Brownian motion B, [B, B]t=< B, B >t= t,

all t ≥ 0. The conditional quadratic variation is also known in the literature by its notation. It is sometimes called the sharp bracket, the angle bracket, or the oblique bracket. It has properties analogous to that of the quadratic variation processes. For example, if X and Y are two semimartingales such that < X, X >, < Y, Y >, and < X + Y, X + Y >all exist, then < X, Y > exists and can be dened by polarization

< X, Y >= 1

2(< X + Y, X + Y > − < X, X > − < Y, Y >).

However, < X, Y > can be dened independently as the compensator of [X, Y ] which provides of course that [X, Y ] is locally of integrable variation. In other words, there exist stopping times (Tn)

n≥1 increasing to ∞ a.s. such that E{R Tn

0 |d[X, Y ]s|} < ∞

for each n. Also, < X, X > is a non-decreasing process by the preceding discussion, since [X, X] is non-decreasing. The conditional quadratic variation is inconvenient since unlike the quadratic variation it doesn't always exist. Moreover, while [X, X], [X, Y ] and [Y, Y ] all remain invariant with a change to an equivalent probability measure, the sharp brackets in general change with a change to an equivalent probability measure and may even no longer exist. Although the angle bracket is ubiquitous in the literature it is sometimes unnecessary as one can often use the quadratic variation instead, and indeed whenever possible we use the quadratic variation rather than the conditional quadratic variation. < X, X > occurs naturally in extensions of Girsanov's theorem for example, and it has become indispensable in many areas of advanced analysis in the theory of stochastic processes.

2.5 The Bichteler-Dellacherie Theorem

The next theorem shows us that a process X is a classical semimartingale if and only if it is a semimartingale.

(25)

Theorem 28. (Bichteler-Dellacherie Theorem)

An adapted, càdlàg process X is a semimartingale if and only if it is a classical semi-martingale. That is, X is a semimartingale if and only if it can be written X = M + A, where M is a local martingale and A is an FV process.

Theorem 29. Let X be an adapted, càdlàg process. The following are equivalent: (i) X is a semimartingale;

(ii) X is decomposable;

(iii) given β > 0, there exist M, A with M0 = A0 = 0, M a local martingale with jumps

bounded by β, A an FV process, such that Xt = X0+ Mt+ At;

(iv) X is a classical semimartingale.

2.6 Local Martingales

Recall that for a process X and a stopping time T we further recall that XT denotes

the stopped process

XtT = Xt∧T = Xt1{t<T }+ XT1{t≥T }.

Denition 28. An adapted, càdlàg process X is a local martingale if there exists a se-quence of increasing stopping times, Tn, with limn→∞Tn= ∞a.s. such that Xt∧Tn1{Tn>0}

is a uniformly integrable martingale for each n. Such a sequence (Tn) of stopping times

is called a fundamental sequence.

Denition 29. A stopping time T reduces a process M if MT is a uniformly integrable

martingale.

Theorem 30. Let M, N be local martingales and let S and T be stopping times. (a) If T reduces M and S ≤ T a.s., then S reduces M.

(b) The sum M + N is also a local martingale.

(c) If S, T both reduce M, then S ∨ T also reduces M. (d) The processes MT, MT1

{T >0} are local martingales.

(e) Let X be a càdlàg process and let Tn be a sequence of stopping times increasing to ∞

a.s. such that XTn1

{Tn>0} is a local martingale for each n. Then X is a local martingale.

Theorem 31. Let X be a process which is locally a square integrable martingale. Then X is a local martingale.

Theorem 32. Let M be adapted, càdlàg and let (Tn)n≥1 be a sequence of stopping times

increasing to ∞ a.s.. If MTn1

{Tn>0} is a martingale for each n, then M is a local

(26)

It is often of interest to determine when a local martingale is actually a martingale. A simple condition involves the maximal function. Recall that X∗

t = sups≤t|Xs| and

X∗ = sups|Xs|.

Theorem 33. Let X be a local martingale such that E{X∗

t} < ∞for every t ≥ 0. Then

X is a martingale. If E{X∗} < ∞, then X is a uniformly integrable martingale.

Remark 3. Note that in particular a bounded local martingale is a uniformly integrable martingale.

Denition 30. Two probability laws P , Q on (Ω, F) are said to be equivalent if P << Q and Q << P . (Recall that P << Q denotes that P is absolutely continuous with respect to Q.) We write Q ≈ P to denote equivalence.

If Q << P , then there exists a random variable Z in L1(dP ) such that dQ

dP = Z and

EP{Z} = 1, where EP denotes expectation with respect to the law P . We let

Zt= EP{

dQ dP|Ft}

be the right continuous version. Then Z is a uniformly integrable martingale and hence a semimartingale. Note that if Q is equivalent to P then dP

dQ ∈ L 1(dQ)and dP dQ = ( dQ dP) −1.

Lemma 1. Let Q ≈ P , and Zt = EP{dQdP|Ft}. An adapted, càdlàg process M is a Q

local martingale if and only if MZ is a P local martingale. Theorem 34. (Girsanov-Meyer Theorem)

Let P and Q be equivalent. Let X be a semimartingale under P with decomposition X = M +A. Then X is also a classical semimartingale under Q and has a decomposition X = L + C, where Lt = Mt− Z t 0 1 Zs d[Z, M ]s

is a Q local martingale, and C = X − L is a Q FV process. Theorem 35. (Girsanov-Meyer Theorem: Predictable Version)

Let X be a P local martingale with X0 = 0. Let Q be another probability equivalent to

P and let Zt = E{dQdP|Ft}. If < X, Z > exists for the probability P , then the canonical

Q decomposition of X is Xt = (Xt− Z t 0 1 Zs− d < X, Z >s) + Z t 0 1 Zs− d < X, Z >s.

(27)

Theorem 36. Let X be a P local martingale with X0 = 0. Let Q be another probability

absolutely continuous with respect to P , and let Zt = E{dQdP|Ft}. Assume that < X, Z >

exists for P . Then At =

Rt

0 1

Zs−d < X, Z >s exists a.s. for the probability Q, and

Xt−

Rt

0 1

Zs−d < X, Z >s is a Q local martingale.

Theorem 37. (Lenglart-Girsanov Theorem)

Let X be a P local martingale with X0 = 0. Let Q be a probability absolutely continuous

with respect to P , and let Zt = EP{dQdP|Ft}, R = inf{t > 0 : Zt = 0, Zt− > 0}, and

Ut= ∆XR1{t≥R}. Then Xt− Z t 0 1 Zs d[X, Z]s+ ˜Ut is a Q local martingale.

(28)

3 Financial Market Model

To describe a nancial market operating in continuous time, we begin with a probability space (Ω, F, P ), a time horizon T ∈ (0, ∞) and a ltration F = (Ft)0≤t≤T satisfying the

usual conditions of right-continuity and completeness. Additionally, we assume that F0

is trivial and we set FT = F. We have d+1 basic primary assets available for trade with

price processes Si = (Si

t)0≤t≤T for i = 0, 1, . . . , d. To simplify the notation, we assume

that S0 has a strictly positive price. Then we use S0 as a numéraire and describe all

quantities in this paper discounted with S0. This way asset 0 has (discounted) price

1 at all times and the other assets (discounted) prices are Xi = Si

S0 for i = 1, . . . , d.

Furthermore, we assume that time t prices Xt are Ft-measurable, i.e. observable at time

t and that X is a càdlàg semimartingale.

In this setting a contingent claim is an FT-measurable random variable H describing a

random payo to be made at time T of some nancial instrument. This way our claims are of European type because the maturity is xed.

As far as hedging is now concerned, the following question arises: Having sold H, how can one insure against the random loss at time T ?

To approach this question we consider self-nancing portfolio strategies (V0, ϑ) =

(V0, ϑt)0≤t≤T, where V0 is an F0-measurable random variable that can be interpreted

as the initial outlay to start the strategy and ϑ is a d-dimensional predictable process. In such a strategy, ϑi

t describes the number of units of asset i held at time t and

pre-dictability of ϑ is a mathematical formulation of the informational constraint that ϑ is not allowed to anticipate the movement of X.

Throughout this work we refer to the theory of stochastic integration established in Protter (2005).

At any time t, the cumulative gains from trade up to time t are given by Gt(ϑ) =

Z t

0

ϑsdXs

and the value process V is given by Vt= V0+

Z t

0

(29)

The strategy being self-nancing means that the cost process which is given by Ct= Vt−

Z t

0

ϑsdXs = Vt− Gt(ϑ)

is constant so that it is not allowed to put external money into the portfolio.

Since X is a semimartingale it makes sense to speak of stochastic integrals with respect to X and we denote by L(X) the linear space of all Rd-valued predictable X-integrable

processes ϑ.

A claim H is called attainable if there exists a self-nancing strategy with VT = H P-a.s.

By equation 1 this means that H can be written as H = H0+

Z T

0

ϑHs dXs P −f.s.,

i.e. as the sum of a constant H0 and a stochastic integral with respect to X. We speak of

a complete market if every claim is attainable; otherwise the market is called incomplete. We dene Me q(P ) =  Q ≈ P : dQ dP ∈ L q(P ), X is a local Q − martingale  and Mq(P ) =  Q << P : dQ dP ∈ L q(P ), X is a local Q − martingale 

with 1 ≤ q ≤ ∞ and we assume that Mq(P ) 6= ∅. Note that since we do not assume

that Me

q(P ) 6= ∅ we do not necessarily have an arbitrage-free market.

Since we want to examine dierent spaces of stochastic integrals and see if they are closed or not, we will vary the subspaces Θ ⊆ L(X) of integrands as well as the condi-tions on our semimartingale X that describes the basic stock prices.

First we need to establish some notation that will be used throughout this paper. Suppose we have chosen a linear subspace Θ of L(X). Then a Θ-strategy is a pair (V0, ϑ) ∈ R × Θ with value process V0+ G(ϑ) and the linear space

G := GT(Θ) =

Z T

0

ϑudXu|ϑ ∈ Θ

(30)

describes all outcomes of self-nancing Θ-strategies with initial wealth V0 = 0 and A := R + G =  V0+ Z T 0 ϑudXu|(V0, ϑ) ∈ R × Θ 

is the space of contingent claims replicable by self-nancing Θ-strategies.

Throughout the paper, p ∈ [1, ∞] will be arbitrary (but xed) and q will denote the conjugate exponent, i.e. 1

p + 1

q = 1. On L

p(P ) we shall consider the norm topology, if

(31)

4 Examination of the Closedness of Spaces of

Stochastic Integrals in L

2

To be able to nd suitable hedging strategies for the sale of an option in an incomplete nancial market, we will try to nd the orthogonal projection of the claim onto dierent spaces of terminal portfolio values. Therefore we will analyze the closedness of those subspaces of the Hilbert space L2 in this section.

For trading strategies we will consider the sets Θeq

p , Θv and ΘS, i.e. we will

ex-amine the L2-closedness of the spaces G

T(Θeq2 ), GT(Θv) and GT(ΘS). We will give

explicit information on when and by whom the respective spaces have been intro-duced and what this work contributes to mathematical research in each subsection, but we want to give a brief overview in advance. The notion of GT(Θv) has rst

been mentioned by Xia/Yan (2006) and GT(ΘS) was rst described and worked on

by Delbaen/Monat/Schachermayer/Schweizer/Stricker (1991). GT(Θeq2 ) has only been

described implicitly by Theorem 1.2 in Delbaen/Schachermayer (1996). This work rstly provides a summary of the existing results on the dierent spaces in a uniform setting and gives an explicit denition of the space GT(Θeq2 )and provides a proof for the closedness

in L2. In addition, we develop sucient conditions under which G

T(Θeq2 ) and GT(Θv)

coincide, so that we can transfer all the results from one space to another and vice versa. Lastly, we give a counterexample for the closedness of GT(ΘS)in the discontinuous case.

4.1 The Space G

T

eqp

)

The rst space that we will take a look at is GT(Θeqp ), which is a very intuitive

con-struction when we take a look at the space Dp in Delbaen/Schachermayer (1996). It

is based on the idea of taking linear combinations of stochastic integrals with so-called simple integrands and then taking a look at the corresponding L2-closure. We will give

an explicit denition of the elements of the space (Denition 32) and with the help of Theorem 1.2 in Delbaen/Schachermayer (1996) we will show that it is closed in L2

(Corollary 13). Furthermore, we state an identity between GT(Θeqp ) and another space

in the case that X is continuous (Theorem 39).

We consider X to be an Rd-valued càdlàg semimartingale based on (Ω, F, (F

t)t∈R+, P )

(32)

stopping times increasing to T such that, for each s ∈ N, the family {XU : U stopping time, U ≤ Us}

is bounded in Lp(P ) (and uniformly integrable in the case p = 1).

A predictable Rd-valued process will be called a simple p-admissible integrand for X

if it has a representation Ht = H0χ{0}(t) + n X i=1 Hiχ]]Ti,Ti+1]](t)

where 0 = T1 ≤ . . . ≤ Tn+1 < ∞ is a nite sequence of stopping times dominated

by some corresponding Us and where Hi is in L∞(Ω, FTi, P ). The space of all those

processes is denoted by Θs p.

Using the concepts of Protter (2005) for those processes we then may form the stochas-tic integral Z t 0 HdX = H0X0+ n X i=1 Hi(Xt∧Ti+1 − Xt∧Ti)

and the random variable Z T 0 HdX = H0X0+ n X i=1 Hi(XTi+1− XTi).

For every summand the interpretation here is that Hi denes the trading strategy of

buying at time Ti the amount of Hi(ω) = (Hi1(ω), . . . , Hid(ω)) units of the (discounted)

stocks X = (X1, . . . , Xd)and selling it at time T

i+1. At time t ∈ [0, ∞] the random

vari-able (Rt

0 HdX)(ω)describes the cumulated gain (or loss) up to time t and (R T

0 HdX)(ω)

the nal result, if an agent follows the trading strategy H. We call the linear space con-sisting of the above random variables RT

0 HdX the space of claims attainable by simple

integrands or by simple trading strategies for obvious reasons.

(33)

Denition 31. In the above setting let Kp = GT(Θsp) and Dp = GT(Θsp) − L p +∩ GT(Θsp) + L p +,

where the bar denotes the closure with respect to the norm topology of Lp(P ), for 1 ≤

p < ∞, and with respect to the σ∗-topology, for p = ∞.

The interpretation of the elements in Kp is obvious: a random variable f ∈ Lp is

in Kp if it can be approximated by random variables GT(Θsp). The space Dp (which

clearly contains Kp) has a more intriguing interpretation: a random variable f ∈ Lp

is in Dp if it may be approximately sandwiched between elements of GT(Θsp), i.e. if

there are simple p-admissible strategies H+ and Hsuch that (RT

0 H +dX) − f ) − as well as (RT 0 H −dX) − f )

+ both are small with respect to the topology on Lp(P ). An

eco-nomic agent will wish to approximate f by either RT

0 H

+dX or RT

0 H

dX depending on

whether she wants to buy or sell the contingent claim modelled by the random variable f. Before we state the next theorem we need a lemma that will be needed for the proof of the theorem.

Lemma 2. If a stochastic process Xn converges to X in Lp, p an integer, then E[Xnp]

converges to E[Xp].

Proof 1. Note that by Hölder`s inequality, for each 1 ≤ j ≤ p, we have lim

n→+∞E[|Xn− X|

j|X|p−j] = 0.

Now we can conclude the result using the expansion Xnp = (Xn− X + X)p = p X j=0 p j  (Xn− X)jXp−j = Xp+ p X j=1 p j  (Xn− X)jXp−j.

The following theorem is a partial statement of Theorem 1.2 in Delbaen/Schachermayer (1996). Since we focus on measures Q ∈ Mq(P ), we do not state all their equivalences

and therefore give a slightly dierent proof.

Theorem 38. Let 1 ≤ p ≤ ∞, q its conjugate exponent, X a semimartingale locally in Lp(P ) such that Meq(P ) 6= ∅, and f ∈ Lp(P ). The following assertions are equivalent: (i) f ∈ Dp.

(34)

(ii) There is an X-integrable predictable process H such that, for each Q ∈ Mq(P ), the

process Rt

0 HdX is a uniformly integrable Q-martingale converging to f in the norm of

L1(Q).

(iii) EQ[f ] = 0 for each Q ∈ Mq(P ).

Proof 2. (ii) ⇒ (iii) is obvious (iii) ⇒ (i)

If f /∈ GT(Θsp) − L p

+(P ), then the Hahn-Banach theorem provides us with an element

g of Lq+(P ) vanishing on GT(Θsp) - so that, after normalization, it is the density of a

non-negative probability measure R in Mq(P ) - and such that EP[f g] = ER[f ] > 0.

The case f /∈ GT(Θsp) + L p

+(P ) is similar.

Since we need the equivalence from (i) and (iii) for the proof of (i) ⇒ (ii) we also show (i) ⇒ (iii)

For Q ∈ Mq(P ), we have that Q takes values ≤ 0 on GT(Θsp) − L p +(P ):

For u ∈ GT(Θsp) and v ∈ L p

+(P ), we have

E[u − v] = E[u] − E[v]

and since E[u] is 0 and E[v] is ≥ 0, we can conclude with lemma 2 that Q takes values ≤ 0 on GT(Θsp) − L

p

+(P ). That Q takes values ≥ 0 on GT(Θsp) + L p

+(P ) can be shown

in a similar way. Hence Q vanishes on Dp.

(i) ⇒ (ii)

Fix f ∈ Dp and Q ∈ Meq(P ); rst note that the identity mapping considered as an

operator from Lp(P ) to L1(Q) is well dened and continuous. Hence we have that,

f ∈ GT(Θsp) + L1+(Q) L1(Q)

which means that there is a sequence (fn)∞n=1 = (

RT 0 HndX) ∞ n=1 in GT(Θsp) such that lim n→∞EQ[f − fn]+ = 0.

But from the martingale property and (iii) - which we have already proven to being equivalent to (i) - we get for each n ∈ N

(35)

which implies that

lim

n→∞EQ[|f − fn|] = 0,

i.e. f is in the L1(Q)-closure of G T(Θsp).

The rest of the proof follows an argument of C. Stricker (1990): We may identify f - as well as each fn - with a uniformly integrable martingale (ft)t∈R+ by letting

ft= EQ[f |Ft].

We are now in a position to apply the theorem of M. Yor (1978, cor. 2.5.2) to exhibit a predictable integrand H with the desired properties w.r.t. Q.

We still have to show that H also has the desired properties w.r.t. each R ∈ Mq(P ). We

have to show that Rt

0 HdX = ER[f |Ft] for each t ∈ R+, which will readily show that the

R-almost surely dened stochastic integral R HdX is indeed a R-uniformly integrable martingale.

As R ∈ Mq(P ) we have that each R HndX is an R-uniformly integrable martingale

so that Rt

0 HndX = E[

RT

0 HndX|Ft]. By the same argument as above, (R T

0 HndX) ∞ n=1

converges to f in L1(R) and therefore (Rt

0HndX) ∞

n=1 is a Cauchy sequence in L1(R)

that converges to Rt

0 HdX in L

1(Q) and therefore also to Rt

0HdX in L

1(R). This shows

that R HdX is indeed a R-uniformly integrable martingale converging to f in L1(R),

thus nishing the proof of the implication (i) ⇒ (ii). Denition 32. Θeq

p consists of all ϑ ∈ L(X) such that GT(ϑ)is in Lp(P ) and the process

G(ϑ) =R ϑdX is a uniformly Q-integrable Q-martingale for every Q ∈ Me q(P ).

Corollary 13. Let X be a Rd-valued semimartingale which is locally in L2(P ) and

Me

2(P ) 6= ∅. Then the space GT(Θeq2 ) is closed in L2(P ).

Proof 3. This is due to Theorem 38 and follows from the equivalence of (i) and (ii). Note that the equivalence also holds true if we assume Q ∈ Me

q(P ) (see

Del-baen/Schachermayer (1996)).

Remark 4. As one might have wondered that we do not explicitly require a no-arbitrage condition in our market model, we will see in section 5.2 that the condition Me

2(P ) 6= ∅

can be interpreted as one.

Remark 5. Due to Theorem 38 we also have the following, more general assertion: Let 1 ≤ p ≤ ∞, q its conjugate exponent, X a Rd-valued semimartingale which is locally

in Lp(P ) and Me

(36)

But since we want to nd hedging strategies in Hilbert spaces, we focus on the case of p = 2.

Of course, the question arises, whether the concept of 'sandwichable' contingent claims is vacuous in the sense that we always have GT(Θsp) = GT(Θeqp ). The next example which

has been established in Delbaen/Schachermayer (1996) (for the case p = 2), shows that this is not always the case, i.e. there are situations where GT(Θeqp ) strictly contains

GT(Θsp). But after that, we supply a sucient condition for X so that we always have

GT(Θsp) = GT(Θeqp ).

Example 1. We construct a uniformly bounded discrete adapted stochastic process X = (Xt)∞t=0 dened on (Ω, F, (F)

t=0, P ) with the following properties.

(i) There exists an equivalent martingale measure Q for X with density function dQ dP ∈

L∞(Ω, F , P ).

(ii) GT(Θs2) is strictly contained in D2.

We work on Ω = N. Denote, for t ∈ N0,

At1 = {3t + 1} , At2 = {3t + 2} , At3 = {3t + 3} , Bt = {3t + 1, 3t + 2, . . .} =[ s≥t (As1 ∪ As 2∪ As3), Ct = {3t + 2, 3t + 3, . . .} = Bt\ At 1.

In order to keep track of the right order of magnitude of the sequences constructed below we shall use the following notation: For sequences (at)∞t=0 and (bt)∞t=0 of positive numbers

we write at ≈ bt if there are constants c, C > 0 such that cat < bt < Cat for all t

suciently big.

F will denote the sigma-algebra of all subsets of Ω and we shall dene measures P and Q on F: let (formally) P (A−11 ) = −1 and dene recursively, for t ≥ 0,

P (At1) = 2−1(1 − 2−(2t−1))P (At−11 ) = 2−(t+1) t Y s=1 (1 − 2−(2s−1)) ≈ 2−t, P (At2) = P (At3) = 2−(2t+2)P (At1) = 2−(3t+3) t Y s=1 (1 − 2−(2s−1)) ≈ 2−3t,

(37)

and (formally) Q(A−1 1 ) = 4 and, for t ≥ 0, Q(At 1) = 2 −3Q(At−1 1 ) = 2 −(3t+1)≈ 2−3t , Q(At 2) = 2 −(t+2)Q(At 1) = 2 −(4t+3)≈ 2−4t , Q(At3) = (1 − 2−2− 2−(t+2))Q(A1t) = (1 − 2−2− 2−(t+2))2−(3t+1)≈ 2−3t. Let us try to explain the idea behind this denition: we start with letting P (A0

1) =

Q(A0

1) = 2−1 so that P (C0) = Q(C0) = 2−1. For each t ∈ N0 the set Ct is broken into

Ct = At2∪ At

3∪ At+11 ∪ Ct+1.

The mass of Ct is divided among these 4 sets such that

P (At+11 ) = P (Ct+1) and Q(At+11 ) = Q(Ct+1). (2) In the case of P the mass of Ct is distributed among the 4 sets above with the weights

2−(2t+2), 2−(2t+2), 2−1(1 − 2−(2t+1)), 2−1(1 − 2−(2t+1))

and in the case of Q with the weights 2−(t+2), (1 − 2−2− 2−(t+2)), 2−3, 2−3

. Clearly the measures P and Q are equivalent and dQ

dP is uniformly bounded.

Now we dene a sequence (ft)∞t=0 of functions on Ω by

ft =          1, on At1, −1, on Ct, 0, elsewhere. In view of 2 we have EP[ft] = EQ[ft] = 0. (3) Let, for t ∈ N0, at = 2t ∞ Y s=t+1 (1 − 2−(2s−1)) ≈ 2t, and dene the function f on Ω by

f =

X

t=0

atft.

(38)

every-where on Ω. It is elementary to calculate explicitly the values of f: f =    ct on At1, −bt on At2∪ At3, where bt = t X s=0 as = t X s=0 (2s ∞ Y r=s+1 (1 − 2−(2r−1))) ≈ 2t and, more precisely, limt→∞ b

t 2t = 2, ct= at− bt−1= 2t ∞ Y s=t+1 (1 − 2−(2s−1)) − t−1 X s=0 2s ∞ Y r=s+1 (1 − 2−(2r−1)) ≈ 1.

Note that f ∈ L2(P ) and that, for all t ∈ N 0,

(ft, f )P = 0, (4)

where (·, ·)P denotes the inner product in L2(P ). Indeed, letting Fn denote the n-th

partial sum of f Fn = n X t=0 atft

and noting the biorthogonality of (ft)∞t=0, we have, for n ≥ t,

(ft, Fn)P = at(ft, ft)P = at2P (At1) = 2t ∞ Y s=t+1 (1 − 2−(2s−1))2(2−(t+1) t Y s=1 (1 − 2−(2s−1))) = ∞ Y s=1 (1 − 2−(2s−1)).

(39)

On the other hand lim n→∞(ft, f − Fn)P = limn→∞EP[(Fn− f )χC n] = lim n→∞( ∞ X m=n+1 EP[Fn− f ]χAm 1 + ∞ X m=n EP[Fn− f ]χAm 2 ∪Am3 ) = lim n→∞( ∞ X m=n+1 (−bn− cm)P (Am1 ) + ∞ X m=n (−bn+ bm)P (Am2 ∪ A m 3 )) = − lim n→∞bn ∞ X m=n+1 P (Am1 ) = − lim n→∞2 n+1 ∞ X m=n+1 (2−(m+1) m Y s=1 (1 − 2−(2s−1))) = − ∞ Y s=1 (1 − 2−(2s−1)).

Combining these two equalities we obtain 4. For later use we observe that

EP(f χBt) = ∞ X s=t ctP [At1] − ∞ X s=t btP [At2∪ A t 3] ≈ 2 −t . (5)

Also note that lim n→∞||(Fn− f )+|| 2 L2(P )= lim n→∞ ∞ X m=n+1 (bm− bn)2P (Am2 ∪ Am3 ) (6) = lim n→∞ ∞ X m=n+1 22m· 2−3m = 0.

Now dene, for t ∈ N0,

gt=          Mt on At2, −mt on At3, 1 on Bt+1,

where the real numbers Mt and mt will be chosen such that the relations

(f, gt)P = 0 and EQ[gt] = 0 (7)

(40)

equa-tions

Mt· P (At2)(−bt) + mt· P (At3)bt= −EP[f χBt+1]

Mt· Q(At2) − mt· Q(At3) = −Q(B t+1).

We can rearrange these equations to get

Mt− mt = EP(f χBt+1) P (At2)bt ≈ 2t Mt·  −Q(A t 2) Q(At 3)  + mt = Q(Bt+1) Q(At 3) ≈ 1

which yields, in view of Q(At 2)

Q(At 3) ≈ 2

−t,

Mt≈ 2t and mt≈ 1.

Now we are ready to dene the process X: Let X0 = 0 and, for t ≥ 0,

X2t+1− X2t = 2−tft,

X2t+2− X2t+1 = 2−2tgt.

Clearly, (Xt)∞t=0 is a uniformly bounded process and it follows from 3 and 7 that X is a

Q-martingale with respect to its natural ltration (Ft)∞t=0. Note that each Fn is a simple

integral on the process X, hence Fn ∈ GT(Θs2) and we obtain from 6 that

f ∈ GT(Θs2) − L2+(P ). (8)

On the other hand, we claim that for

Gn = Fn+ bngn

we have that

lim

n→∞||(Gn− f )−||L

2(P )= 0 (9)

which will readily imply that

(41)

and therefore, combining 8 and 10,

f ∈ D2.

To prove 9, note that

||(Gn− f )−||2L2(P )= ||(Gn− f ) − χAn 3|| 2 L2(P )≈ (2bn)2P (At3) ≈ 22n2−3n .

Finally, we shall show that f is orthogonal to GT(Θs2), whence in particular

f ∈ D2\ K2. (11)

Indeed, as for each t ≥ 0 the support of Xt+1− Xt is contained in an atom of Ft, the

space GT(Θs2) of simple integrals consists of the linear span of (ft)∞t=0 and (gt)∞t=0 and

therefore we obtain the assertion from 4 and 7. The construction of the example is now completed.

So we have seen that GT(Θsp) = GT(Θeqp ) does not hold in general. But the next

theorem shows that we have GT(Θsp) = GT(Θeqp )for the case of continuous processes. The

content of the theorem has been established in Theorem 2.2 of Delbaen/Schachermayer (1996), but here we will adjust it to our setting and apply it to our newly constructed space GT(Θeqp ).

Theorem 39. Suppose that X = (Xt)t∈R+ is a continuous R

d-valued semimartingale

which is locally in Lp, let 1 ≤ p, q ≤ ∞, 1 p +

1

q = 1 and suppose that M e

q(P ) 6= ∅.

Then we have GT(Θsp) = GT(Θeqp ).

Proof 4. Suppose rst that 1 ≤ p < ∞. Let f ∈ GT(Θeqp ); by theorem 38 there is an

X-integrable predictable process H such that Rt

0 HdX is a uniformly integrable martingale

with respect to each Q ∈ Me

q(P ) and such that almost surely

f = lim

t→∞

Z t

0

HdX.

The denition of GT(Θeqp ) respectively Dp implies the existence of sequences (H+,n)∞n=1

and (H−,n)

n=1 of simple p-admissible integrands such that, for

f+,n = Z T 0 H+,ndX and f−,n = Z T 0 H−,ndX

(42)

we have that ((f − f+,n)

+)∞n=1 and ((f − f−,n)−)∞n=1 tend to zero in Lp(P ). By the

argument used in the proof of theorem 38 we deduce that, in fact, (f − f+,n)

n=1 and

(f −f+,n)∞n=1 tend to zero in L1(Q), for each Q ∈ Meq(P ). It follows from the martingale property that, more generally, for each Q ∈ Me

q(P ) and each stopping time S, we have

lim n→∞|| Z S 0 (H − H+,n)dX||L1(Q) = 0 and lim n→∞|| Z S 0 (H − H−,n)dX||L1(Q) = 0. (12)

We have to show that, for  > 0, there is a simple p-admissible integrand H such that

||f − Z T 0 HdX||Lp(P ) < . For C ∈ R+ let SC = inf  t : | Z t 0 HdX| ≥ C  , so that lim C→∞P ({SC < ∞}) = 0.

Let δ > 0 to be specied below and nd C = C(δ) > 1 such that P ({SC < ∞}) < δ. Let h = Z SC 0 HdX and dene, for n ∈ N

h+,n = Z SC 0 H+,ndX and h−,n = Z SC 0 H−,ndX. We deduce from 12 that (h+,n)

n=1 and (h−,n)∞n=1, converge to h in L1(Q) and therefore

in measure. For η > 0, again to be specied below, we therefore have that, for n = n(η) suciently big,

P (|h − h+,n| > 1 ) < η and P (|h − h−,n| > 1 ) < η.

Fix n such that the above inequalities hold true and dene the FSC-measurable sets A

(43)

and A− by

A+ = {SC < ∞} ∩ {h = −C} ∩|h − h+,n| ≤ 1

and A− = {SC < ∞} ∩ {h = +C} ∩|h − h−,n| ≤ 1 ,

so that A+ and Aare disjoint subsets of {S

C < ∞} covering this set up to a set B =

{SC < ∞} \ (A+∪ A−) of P -measure at most P (B) < 2η.

Dene the predictable integrand ˜H ˜ H = Hχ]0,SC]+ H +,nχ ]SC,∞[χA++ H −,n χ]SC,∞[χA−,

and dene the stopping time S as the rst moment after SC when R t

0 HdX = 0˜ . We

want to show that

||f − ˜f ||Lp(P )<  (13)

if δ = δ() > 0 and η = η(C(δ), ) > 0 are suciently small, where ˜ f = Z S 0 ˜ HdX. As ||f − ˜f ||Lp(P ) ≤ ||(f − ˜f )χ A+||Lp(P )+ ||(f − ˜f )χA−||Lp(P )+ ||(f − ˜f )χB||Lp(P ), (14)

it will suce to show that each of the three terms on the right hand side is less than  3.

As regards the last one note that ||fχB||Lp(P )<  6 and || ˜fχB||Lp(P )≤ C(2η) 1 p <  6 if η = η(C(δ), ) is small enough.

As regards the rst two terms in 14 we only estimate the rst one (the second being analogous): we split the set A+ into A+∩ {S < ∞} and A+∩ {S = ∞}. For the former

set we may estimate

||(f − ˜f )χA+∩{S<∞}||Lp(P )= ||f χA+∩{S<∞}||Lp(P )≤ ||f χA+||Lp(P ),

which is smaller than 

6 if δ = δ() > 0 is suciently small as f ∈ L p(P ).

(44)

For the second set we may estimate ||(f − ˜f )χA+∩{S=∞}||Lp(P ) ≤ ||(1 + |f − f+,n|)χ A+∩{S=∞}||Lp(P ) ≤ ||(1 + (f − f+,n)+)χA+∩{S=∞}||L p(P )+ ||(1 + (f − f +,n )−)χA+∩{S=∞}||Lp(P ) ≤ ||(1 + (f − f+,n) +)χA+∩{S=∞}||Lp(P )+ ||(2 + fA+∩{S=∞}||Lp(P ).

In the last line we have used the fact f+,n is less than or equal to 1 on {T = ∞}. If we

choose δ = δ() > 0 small enough and n = n(, η) big enough the above expression is smaller than 

6.

Summing up, we have shown 13: given  > 0 choose δ = δ() > 0, then C = C(δ) > 0, η = η(C, ) > 0 and nally n = n(, η) ∈ N0. However, we are not yet nished, as ˜H

is a simple p-admissible integrand only after the stopping time SC. But it is standard

to approximate RSC

0 HdX by the stochastic integral of a simple p-admissible integrand ˆH

supported by [0, SC] such that | R t 0 HdX|ˆ is bounded by C and || Z SC 0 ˆ HdX − Z SC 0 HdX||Lp(P ) < .

Modifying ˜H on [0, SC] in the indicated way we obtain the desired simple integrand H

for which

||f − Z T

0

HdX||Lp(P )< 2,

thus nishing the proof.

The case p = ∞ is easy: simply note that, for Q ∈ Me

1(P ), (L

(P ), σ(L(P ), L1(P ))),

(45)

4.2 The Space G

T

v

)

The next set of trading strategies which we will take a profound look at is Θv which was

rst mentioned in Xia/Yan (2006). They used signed martingale measures to establish the L2-closedness of the space. We will expand their ideas in the sense that we will

develop the identity GT(Θeq2 ) = GT(Θv)under certain requirements in our Corollary 14.

Therefore we get an explicit description of the elements in GT(Θv) as well. The reader

may notice that this is crucial for nding hedging strategies later on, since all (existing) results on the strategies for the respective space can be used for the other space and vice versa. We want to clearly indicate that Lemma 3, Lemma 4 and Theorem 41 were established in Xia/Yan (2006). Our contribution to this section is using their results to provide an assertion about relating GT(Θv)to our previously constructed space GT(Θeqp )

in Corollary 14 and Corollary 15 which will be crucial for nding explicit hedging strate-gies later on.

For a process X we establish a duality relation between Kp, the Lp(P )-closure of the

space of claims in Lp(P ), which are attainable by `simple` strategies, and Mq,s, all signed

martingale measures Q with dQ dP ∈ L

q(P ), where p ≥ 1, q ≥ 1 and 1 p +

1 q = 1.

First of all we want to make the following hypothesis (under which X is not necessarily a semimartingale):

Hypothesis 1. X is locally in Lp(P ).

As in the previous chapter we consider GT(Θsp), i.e. the subspace of Lp(P ) spanned

by the simple stochastic integrals of the form

f = h(XT2 − XT1)

where T1 ≤ T2 are stopping times. We dene

Kp = GT(Θsp),

where the bar denotes the closure in Lp(P ).

It is clear that a probability measure Q on (Ω, F) makes X a local martingale under Qif dQ dP ∈ L q(P ) and EQ[f ] = E[ dQ dPf ] = 0

(46)

for all f ∈ GT(Θsp). We set

Mq,s:=g ∈ Lq(P ) : E[gf ] = 0 for all f ∈ G

T(Θsp) and E[g] = 1 ,

Mq,e:= {g ∈ Mq,s: g > 0 a.s.} .

Any element in Mq,s (resp. Mq,e) is called a signed (resp. equivalent) martingale

measure for X.

Before we can prove our rst lemma we need the following corollary from the Hahn-Banach theorem:

Theorem 40. If X is a normed space, U a subspace and if x ∈ X is not contained in the closure of U, then there exists a linear functional f with norm 1 that vanishes on U and for which we have f(x) = ||x||.

Lemma 3. Under Hypothesis 1, we have (a) Mq,s6= ∅ ⇔ 1 /∈ K

p;

(b) For g ∈ Lq(P ), g ∈ Mq,s ⇔ E[g] = 1 and E[gf] = 0 for all f ∈ K p.

Proof 5. (a) First of all we dene the space ˆ

K = span(G(Θs p), 1).

Now according to Theorem 40 the linear functional ϕ on ˆK that satises ϕ|Kp = 0 and

ϕ(1) = 1 is well dened and continuous on the closed space ˆK i 1 /∈ Kp.

(b) This is an immediate consequence of the very denition of Mq,s.

Lemma 3 (b) and the following lemma give a duality relation between Kp and Mq,s.

Lemma 4. Under Hypothesis 1 and that Mq,s6= ∅, we have

f ∈ Kp ⇔ f ∈ Lp(P ) and E[fg] = 0 for all g ∈ Mq,s. (15)

To prove this lemma we rst need the following proposition:

Proposition 1. Let ϕ be a continuous linear functional on Lp(P ). Then there exists a

g ∈ Lq(P ) such that

ϕ(k) = E[gk] for all k ∈ Lp(P ).

(47)

Now we have all the tools to prove the lemma. Proof 6. (proof of Lemma 4)

⇒ is clear. It is left to show ⇐. Assume f ∈ Lp(P ) and E[fg] = 0 for all g ∈ Mq,s.

We want to show f ∈ Kp.

Therefore rstly we will show

1 /∈ span(GT(Θsp), f ).

Actually, we suppose that 1 ∈ span(GT(Θsp), f ). Then there exist a sequence (fn, n ≥

1) ⊂ GT(Θsp) and a sequence (δn, n ≥ 1) ⊂ R such that fn+ δnf → 1 in Lp(P ). Since

we have Mq,s 6= ∅, for any g ∈ Mq,s, we have

0 = E[(fn+ δnf )g] → E[g] = 1,

a contradiction. Because of that, 1 ∈ span(GT(Θsp), f ) is impossible and we have

1 /∈ span(GT(Θsp), f ).

By Theorem 40, there exists a continuous linear functional ϕ1 on Lp(P ) such that

ϕ1(1) 6= 0 and ϕ1(k) = 0 for all k ∈ span(GT(Θsp), f ).

Now we want to show that f must be in

span(GT(Θsp), 1).

Therefore we assume again that f /∈ span(GT(Θsp), 1).

Then we also have by the Hahn-Banach Theorem, that there exists a continuous linear functional ϕ2 on Lp(P ) such that ϕ2(f ) 6= 0and ϕ2(k) = 0for all k ∈ span(GT(Θsp), 1).

Let

ϕ := ϕ1+ ϕ2,

then ϕ is a continuous linear functional on Lp(P ) and hence according to the previous

proposition there exists a g ∈ Lq(P ) such that

(48)

for all k ∈ Lp(P ). It is clear that

E[gk] = ϕ(k) = 0 for all k ∈ Kp, and

E[g] = ϕ1(1) 6= 0,

and

E[gf ] = ϕ2(f ) 6= 0.

So after a normalization, there exists a g ∈ Mq,s such that

E[gf ] 6= 0.

This is a contradiction to the assumption on the right hand side of 15. So f /∈ span(GT(Θsp), 1) is impossible and we have f ∈ span(GT(Θsp), 1).

Now we are able to prove ⇐:

Since f ∈ span(GT(Θsp), 1), there exists a sequence (kn) ⊂ Kps and a sequence (δn) ⊂ R

such that

kn+ δn → f in Lp(P ). For any g ∈ Mq,s, we have

δn= E[(kn+ δn)g] → E[f g] = 0.

Here the last equality obviously follows from the assumption on the right-hand side of 15. Since we know that δn converges to 0 we can conclude that

kn→ f in Lp(P ), which implies

f ∈ Kp.

Throughout the rest of this section, we always assume X to be a semimartingale satisfying Hypothesis 1. Denoted by Θv all Rm-valued predictable X-integrable processes

ϑ such that

GT(ϑ) :=

Z T

0

(49)

and

E[GT(ϑ)g] = 0

for all g ∈ Mq,s. As usual we have

GT(Θv) := {GT(ϑ) : ϑ ∈ Θv} .

If we assume Mq,s6= ∅, then by Lemma 4, we have

GT(Θv) ⊂ Kp. (16)

Furthermore, we make another assumption that Hypothesis 2.

Mq,e 6= ∅,

and get the following theorem:

Theorem 41. Under Hypotheses 1 and 2 we have Kp = GT(Θv).

In order to proof this theorem we will use the following proposition which is a result of Corollary 2.2 in Yor (1978).

Proposition 2. Let M be a local martingale under P . Then Z T

0

HdM : H is predictable and M-integrable and Z

HdMis uniformly integrable 

is closed in L1(Ω, F , P ; R).

Now we are able to prove the theorem. Proof 7. (proof of Theorem 41)

By 16, we only need to show that

Kp ⊂ GT(Θv).

For any w ∈ GT(Θsp), we have

w =

n

X

i=1

Referenzen

ÄHNLICHE DOKUMENTE

These are quantified at the country level using four indicators: life expectancy at birth, mean and expected years of schooling, and the logarithm of Gross National Income per

Chapter 1 contains an introduction to outer measure spaces and an outline of the proof of the bounds of the Carleson Operator (0.1) in the simplified Walsh case.. In Section 1.1

Because we relied on the systems available to LSPs (and provided data in a few cases where they were not able to gain permission to use cus- tomer data), we were not able to

The notion of cyclomatic flow complexity has been derived by means of theoretical considerations. This feature contradicts empirical evidence that the complexity

Here we can recognize one of the motives for Leonardos insistence on proportion and the measure of sight: the correct measure of sight and thus the rationality of perspective

Two models are presented: the case of one lobbyist acting unopposed, and the case of two opposing lobbyists; each is shown t o lead t o a certain concept of

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any

Our tasks directed visual input either to the left or right hemisphere in early visual cortex and required left or right hemispheric motor responses, presumably changing