31.05.2014 Views

Processus de Lévy en Finance - Laboratoire de Probabilités et ...

Processus de Lévy en Finance - Laboratoire de Probabilités et ...

Processus de Lévy en Finance - Laboratoire de Probabilités et ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Thèse prés<strong>en</strong>tée pour obt<strong>en</strong>ir le titre <strong>de</strong><br />

DOCTEUR DE L’ECOLE POLYTECHNIQUE<br />

specialité: Mathématiques Appliquées<br />

par<br />

P<strong>et</strong>er TANKOV<br />

<strong>Processus</strong> <strong>de</strong> Lévy <strong>en</strong> <strong>Finance</strong>:<br />

Problèmes Inverses <strong>et</strong><br />

Modélisation <strong>de</strong> Dép<strong>en</strong>dance<br />

Lévy Processes in <strong>Finance</strong>: Inverse Problems and Dep<strong>en</strong><strong>de</strong>nce Mo<strong>de</strong>lling<br />

Thèse sout<strong>en</strong>ue le 21 septembre 2004 <strong>de</strong>vant le jury composé <strong>de</strong><br />

Ole E. Barndorff-Niels<strong>en</strong><br />

Rama Cont<br />

Nicole El Karoui<br />

Monique Jeanblanc<br />

Bernard Lapeyre<br />

Agnès Sulem<br />

Rapporteur<br />

Directeur <strong>de</strong> thèse<br />

Rapporteur


Résumé<br />

C<strong>et</strong>te thèse traite <strong>de</strong> la modélisation <strong>de</strong> prix boursiers par les expon<strong>en</strong>tielles <strong>de</strong> processus<br />

<strong>de</strong> Lévy. La première partie développe une métho<strong>de</strong> non-paramétrique stable <strong>de</strong> calibration<br />

<strong>de</strong> modèles expon<strong>en</strong>tielle-Lévy, c’est-à-dire <strong>de</strong> reconstruction <strong>de</strong> ces modèles à partir <strong>de</strong>s prix<br />

d’options cotées sur un marché financier. J’étudie les propriétés <strong>de</strong> converg<strong>en</strong>ce <strong>et</strong> <strong>de</strong> stabilité<br />

<strong>de</strong> c<strong>et</strong>te métho<strong>de</strong> <strong>de</strong> calibration, décris sa réalisation numérique <strong>et</strong> donne <strong>de</strong>s exemples <strong>de</strong> son<br />

utilisation. L’approche adoptée ici consiste à reformuler le problème <strong>de</strong> calibration comme<br />

celui <strong>de</strong> trouver un modèle expon<strong>en</strong>tielle-Lévy risque-neutre qui reproduit les prix d’options<br />

cotées avec la plus gran<strong>de</strong> précision possible <strong>et</strong> qui a l’<strong>en</strong>tropie relative minimale par rapport<br />

à un processus “a priori” donné. Ce problème est alors résolu <strong>en</strong> utilisant la métho<strong>de</strong> <strong>de</strong><br />

régularisation, prov<strong>en</strong>ant <strong>de</strong> la théorie <strong>de</strong> problèmes inverses mal posés. L’application <strong>de</strong> ma<br />

métho<strong>de</strong> <strong>de</strong> calibration aux données empiriques <strong>de</strong> prix d’options sur indice perm<strong>et</strong> d’étudier<br />

certaines propriétés <strong>de</strong>s mesures <strong>de</strong> Lévy implicites qui correspon<strong>de</strong>nt aux prix <strong>de</strong> marché<br />

La <strong>de</strong>uxième partie est consacrée au développem<strong>en</strong>t d’une métho<strong>de</strong> perm<strong>et</strong>tant <strong>de</strong> caractériser<br />

les structures <strong>de</strong> dép<strong>en</strong>dance <strong>en</strong>tre les composantes d’un processus <strong>de</strong> Lévy multidim<strong>en</strong>sionnel<br />

<strong>et</strong> <strong>de</strong> construire <strong>de</strong>s modèles expon<strong>en</strong>tielle-Lévy multidim<strong>en</strong>sionnels. C<strong>et</strong> objectif<br />

est atteint grâce à l’introduction <strong>de</strong> la notion <strong>de</strong> copule <strong>de</strong> Lévy, qui peut être considérée<br />

comme l’analogue pour les processus <strong>de</strong> Lévy <strong>de</strong> la notion <strong>de</strong> copule, utilisée <strong>en</strong> statistique<br />

pour modéliser la dép<strong>en</strong>dance <strong>en</strong>tre les variables aléatoires réelles. Les exemples <strong>de</strong> familles<br />

paramétriques <strong>de</strong> copules <strong>de</strong> Lévy sont donnés <strong>et</strong> une métho<strong>de</strong> <strong>de</strong> simulation <strong>de</strong> processus <strong>de</strong><br />

Lévy multidim<strong>en</strong>sionnels, dont la structure <strong>de</strong> dép<strong>en</strong>dance est décrite par une copule <strong>de</strong> Lévy,<br />

est proposée.<br />

Mots clefs: processus <strong>de</strong> Lévy, produits dérivés, calibration, problèmes inverses, régularisation,<br />

problèmes mal posés, <strong>en</strong>tropie relative, copules, dép<strong>en</strong>dance.


Abstract<br />

This thesis <strong>de</strong>als with the mo<strong>de</strong>lling of stock prices by the expon<strong>en</strong>tials of Lévy processes.<br />

In the first part we <strong>de</strong>velop a non-param<strong>et</strong>ric m<strong>et</strong>hod allowing to calibrate expon<strong>en</strong>tial Lévy<br />

mo<strong>de</strong>ls, that is, to reconstruct such mo<strong>de</strong>ls from the prices of mark<strong>et</strong>-quoted options. We<br />

study the stability and converg<strong>en</strong>ce properties of this calibration m<strong>et</strong>hod, <strong>de</strong>scribe its numerical<br />

implem<strong>en</strong>tation and give examples of its use. Our approach is first to reformulate the calibration<br />

problem as that of finding a risk-neutral expon<strong>en</strong>tial Lévy mo<strong>de</strong>l that reproduces the option<br />

prices with the best possible precision and has the smallest relative <strong>en</strong>tropy with respect to a<br />

giv<strong>en</strong> prior process, and th<strong>en</strong> to solve this problem via the regularization m<strong>et</strong>hodology, used in<br />

the theory of ill-posed inverse problems. Applying this calibration m<strong>et</strong>hod to the empirical data<br />

s<strong>et</strong>s of in<strong>de</strong>x options allows us to study some properties of Lévy measures, implied by mark<strong>et</strong><br />

prices.<br />

The second part of this thesis proposes a m<strong>et</strong>hod allowing to characterize the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

structures among the compon<strong>en</strong>ts of a multidim<strong>en</strong>sional Lévy process and to construct multidim<strong>en</strong>sional<br />

expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls. This is done by introducing the notion of Lévy copula,<br />

which can be se<strong>en</strong> as an analog for Lévy processes of the notion of copula, used in statistics<br />

to mo<strong>de</strong>l <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> real-valued random variables. We give examples of param<strong>et</strong>ric<br />

families of Lévy copulas and <strong>de</strong>velop a m<strong>et</strong>hod for simulating multidim<strong>en</strong>sional Lévy processes<br />

with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce giv<strong>en</strong> by a Lévy copula.<br />

Key words: Lévy processes, option pricing, calibration, inverse problems, regularization,<br />

ill-posed problems, relative <strong>en</strong>tropy, copulas, <strong>de</strong>p<strong>en</strong><strong>de</strong>nce


Remerciem<strong>en</strong>ts<br />

Je voudrais, <strong>en</strong> premier lieu, exprimer toute ma gratitu<strong>de</strong> à Rama Cont, mon directeur <strong>de</strong><br />

thèse, non seulem<strong>en</strong>t pour sa disponibilité <strong>et</strong> son souti<strong>en</strong> au cours <strong>de</strong> ces trois années, mais<br />

aussi pour avoir fait preuve, à plusieurs reprises, <strong>de</strong> sa confiance <strong>en</strong> moi, <strong>et</strong> pour m’avoir aidé,<br />

<strong>en</strong> m’<strong>en</strong>voyant à <strong>de</strong>s nombreuses confér<strong>en</strong>ces <strong>et</strong> séminaires, à faire connaître mon travail <strong>de</strong><br />

recherche.<br />

Je n’aurais jamais p<strong>en</strong>sé à travailler dans le domaine <strong>de</strong> la finance mathématique si je n’avais<br />

pas suivi d’excell<strong>en</strong>ts cours <strong>de</strong> Nicole El Karoui, d’abord à l’Ecole Polytechnique <strong>et</strong> <strong>en</strong>suite au<br />

DEA <strong>de</strong> Probabilités <strong>et</strong> Applications <strong>de</strong> l’Université Paris VI. Je la remercie pour cela mais aussi<br />

pour les nombreuses interactions qu’on a eu au cours <strong>de</strong> ma thèse, notamm<strong>en</strong>t à l’occasion du<br />

groupe <strong>de</strong> travail du CMAP.<br />

Jan Kalls<strong>en</strong> m’a expliqué comm<strong>en</strong>t généraliser la notion <strong>de</strong> copule <strong>de</strong> Lévy, que j’ai introduite<br />

pour les processus <strong>de</strong> Lévy aux sauts positifs, au cas <strong>de</strong>s processus <strong>de</strong> Lévy généraux. Je<br />

le remercie égalem<strong>en</strong>t pour avoir bi<strong>en</strong> voulu écrire un article joint sur ce suj<strong>et</strong> <strong>et</strong> pour <strong>de</strong>s<br />

nombreuses discussions qu’on a eu p<strong>en</strong>dant ce travail.<br />

Je remercie Ole Barndorff-Niels<strong>en</strong>, Thomas Mikosch <strong>et</strong> Elisa Nicolato pour avoir fait <strong>de</strong> ma<br />

visite à MaPhySto <strong>en</strong> août-septembre 2002 un mom<strong>en</strong>t inoubliable <strong>et</strong> pour les discussions qu’on<br />

a eu p<strong>en</strong>dant <strong>et</strong> après mon séjour au Danemark, qui ont été toutes très utiles.<br />

Je voudrais égalem<strong>en</strong>t exprimer ma reconnaissance à Douglas N. Arnold and Scot Adams qui<br />

m’ont accueilli p<strong>en</strong>dant ma visite à l’Institut <strong>de</strong> Mathématiques <strong>et</strong> Applications <strong>de</strong> l’Université<br />

<strong>de</strong> Minnesota, un séjour auquel c<strong>et</strong>te thèse doit beaucoup.<br />

Mes remerciem<strong>en</strong>ts vont aussi à Philippe Durand <strong>et</strong> Emmanuel Bacry, qui ont été responsables<br />

<strong>de</strong> <strong>de</strong>ux stages que j’ai effectués avant <strong>de</strong> comm<strong>en</strong>cer c<strong>et</strong>te thèse, l’un au Crédit Lyonnais<br />

<strong>et</strong> l’autre au CMAP, <strong>et</strong> qui ont beaucoup fait pour ma formation sci<strong>en</strong>tifique.<br />

Je suis reconnaissant à Christoph Schwab <strong>et</strong> Ana-Maria Matache d’avoir partagé avec moi<br />

leurs idées sur les métho<strong>de</strong>s numériques pour les processus <strong>de</strong> Lévy, au cours <strong>de</strong> leurs visites à<br />

Paris <strong>et</strong> <strong>de</strong> ma visite à Zurich.<br />

Je remercie <strong>en</strong>fin tous les participants du groupe <strong>de</strong> travail du CMAP <strong>de</strong> toutes nos interactions<br />

sci<strong>en</strong>tifiques, mais aussi d’avoir contribué à une ambiance très agréable <strong>et</strong> chaleureuse<br />

dans notre équipe <strong>de</strong> modélisation financière.


Cont<strong>en</strong>ts<br />

Remarks on notation 11<br />

Introduction <strong>et</strong> principaux résultats 13<br />

Principaux résultats . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

Calibration <strong>de</strong> modèles expon<strong>en</strong>tielle-Lévy . . . . . . . . . . . . . . . . . . . . . . 17<br />

Modélisation multidim<strong>en</strong>sionnelle avec les processus <strong>de</strong> Lévy . . . . . . . . . . . . 22<br />

Introduction 27<br />

I Calibration of one-dim<strong>en</strong>sional exp-Lévy mo<strong>de</strong>ls 31<br />

1 Lévy processes and exp-Lévy mo<strong>de</strong>ls 33<br />

1.1 Lévy processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

1.1.1 Stochastic expon<strong>en</strong>tial of Lévy processes . . . . . . . . . . . . . . . . . . . 36<br />

1.1.2 Change of measure and absolute continuity of Lévy processes . . . . . . . 38<br />

1.1.3 Tightness and converg<strong>en</strong>ce of sequ<strong>en</strong>ces of Lévy processes . . . . . . . . . 40<br />

1.2 Expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls: <strong>de</strong>finition and main properties . . . . . . . . . . . . . 41<br />

1.3 Expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls in finance . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

1.4 Pricing European options in exp-Lévy mo<strong>de</strong>ls via Fourier transform . . . . . . . 47<br />

2 The calibration problem and its regularization 57<br />

2.1 Least squares calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

2.1.1 Lack of i<strong>de</strong>ntification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

2.1.2 Exist<strong>en</strong>ce of solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

7


8 CONTENTS<br />

2.1.3 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

2.1.4 Numerical difficulties of least squares calibration . . . . . . . . . . . . . . 68<br />

2.2 Selection of solutions using relative <strong>en</strong>tropy . . . . . . . . . . . . . . . . . . . . . 69<br />

2.3 The use of relative <strong>en</strong>tropy for pricing and calibration: review of literature . . . . 72<br />

2.3.1 Pricing with minimal <strong>en</strong>tropy martingale measure . . . . . . . . . . . . . 72<br />

2.3.2 Calibration via relative <strong>en</strong>tropy minimization . . . . . . . . . . . . . . . . 75<br />

2.3.3 The pres<strong>en</strong>t study in the context of previous work on this subject . . . . 79<br />

2.4 Relative <strong>en</strong>tropy of Lévy processes . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

2.4.1 Properties of the relative <strong>en</strong>tropy functional . . . . . . . . . . . . . . . . . 82<br />

2.5 Regularizing the calibration problem . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

2.5.1 Properties of regularized solutions . . . . . . . . . . . . . . . . . . . . . . 87<br />

2.5.2 Converg<strong>en</strong>ce of regularized solutions . . . . . . . . . . . . . . . . . . . . . 92<br />

3 Numerical implem<strong>en</strong>tation of the calibration algorithm 95<br />

3.1 Discr<strong>et</strong>izing the calibration problem . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

3.2 Choice of the prior Lévy process . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

3.3 Choice of the regularization param<strong>et</strong>er . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

3.3.1 A posteriori param<strong>et</strong>er choice rules for attainable calibration problems . . 103<br />

3.3.2 A posteriori param<strong>et</strong>er choice rule for non-attainable calibration problems 110<br />

3.3.3 Computing the noise level . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

3.4 Choice of the weights of option prices . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

3.5 Numerical solution of calibration problem . . . . . . . . . . . . . . . . . . . . . . 113<br />

3.5.1 Computing the calibration functional . . . . . . . . . . . . . . . . . . . . . 114<br />

3.5.2 Computing the gradi<strong>en</strong>t of the calibration functional . . . . . . . . . . . . 117<br />

3.5.3 Overview of the algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

3.6 Numerical and empirical tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

3.6.1 Tests on simulated data . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

3.6.2 Empirical properties of implied Lévy measures . . . . . . . . . . . . . . . 123<br />

3.6.3 Testing time homog<strong>en</strong>eity . . . . . . . . . . . . . . . . . . . . . . . . . . . 126


CONTENTS 9<br />

II Multidim<strong>en</strong>sional mo<strong>de</strong>lling with Lévy processes 131<br />

4 Characterization of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of multidim<strong>en</strong>sional Lévy processes 133<br />

4.1 Introduction to <strong>de</strong>p<strong>en</strong><strong>de</strong>nce mo<strong>de</strong>lling . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

4.2 Dep<strong>en</strong><strong>de</strong>nce concepts for multidim<strong>en</strong>sional Lévy processes . . . . . . . . . . . . . 138<br />

4.3 Increasing functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

4.4 Lévy copulas for spectrally positive Lévy processes . . . . . . . . . . . . . . . . . 146<br />

4.5 Lévy copulas for g<strong>en</strong>eral Lévy processes . . . . . . . . . . . . . . . . . . . . . . . 152<br />

4.6 Examples of Lévy copulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158<br />

5 Applications of Lévy copulas 165<br />

5.1 Param<strong>et</strong>ric families of Lévy copulas . . . . . . . . . . . . . . . . . . . . . . . . . . 167<br />

5.2 Simulation of multidim<strong>en</strong>sional <strong>de</strong>p<strong>en</strong><strong>de</strong>nt Lévy processes . . . . . . . . . . . . . 174<br />

5.3 A two-dim<strong>en</strong>sional variance gamma mo<strong>de</strong>l for option pricing . . . . . . . . . . . . 182<br />

Conclusions and perspectives 187<br />

Bibliography 189<br />

Subject in<strong>de</strong>x 199


Remarks on notation<br />

Basic notation<br />

the range of F .<br />

For a function F , Dom F and Ran F <strong>de</strong>note, respectively, the domain and<br />

δ x is a measure giv<strong>en</strong> by δ x (A) = 1 x∈A for every Borel s<strong>et</strong> A and λ| A is the restriction of<br />

the Lebesgue measure onto a Borel s<strong>et</strong> A.<br />

R d + <strong>de</strong>notes [0, ∞) d \ {0} and ¯R <strong>de</strong>notes the ext<strong>en</strong><strong>de</strong>d real line: ¯R := R ∪ {−∞} ∪ {∞}<br />

For two vectors a, b ∈ ¯R d , inequalities like a ≤ b should be interpr<strong>et</strong>ed as a i ≤ b i for<br />

i = 1, . . . , d. For any such vectors, [a, b) <strong>de</strong>notes a right-op<strong>en</strong> left-closed interval of ¯R d : [a, b) :=<br />

[a 1 , b 1 ) × · · · × [a d , b d ). Other types of intervals: (a, b], [a, b] and (a, b) are <strong>de</strong>fined similarly.<br />

Wh<strong>en</strong> we want to consi<strong>de</strong>r an interval as a s<strong>et</strong> of its vertices, without specifying wh<strong>et</strong>her the<br />

boundary is inclu<strong>de</strong>d, we will use the notation |a, b|. Finally, 〈a, b〉 <strong>de</strong>notes the scalar product<br />

of a and b.<br />

Measures and processes In this thesis, we fix a time horizon T ∞ < ∞ and Ω <strong>de</strong>notes the<br />

space of R d -valued càdlàg functions on [0, T ∞ ], equipped with the Skorokhod topology (see [54]).<br />

Unless otherwise m<strong>en</strong>tioned, X is the coordinate process: for every ω ∈ Ω, X t (ω) := ω(t). F is<br />

the smallest σ-field, for which the mappings ω ∈ Ω ↦→ ω(s) are measurable for all s ∈ [0, T ∞ ].<br />

For any t ∈ [0, T ∞ ], F t is the smallest σ-field, for which the mappings ω ∈ Ω ↦→ ω(s) are<br />

measurable for all s ∈ [0, t].<br />

For a semimartingale X, [X] <strong>de</strong>notes is quadratic variation and [X] c is the continuous part<br />

of [X].<br />

P(Ω) <strong>de</strong>notes the s<strong>et</strong> of all probability measures (stochastic processes) on (Ω, F). For two<br />

probability measures P, Q ∈ P(Ω), Q ≪ P means that Q is absolutely continuous with respect<br />

to P , that is, for every B ∈ F, P (B) = 0 implies Q(B) = 0.<br />

11


12<br />

For {P n } n≥1 , P ∈ P(Ω) the notation P n ⇒ P means that the sequ<strong>en</strong>ce {P n } n≥1 converges<br />

weakly to P , that is, for every continuous boun<strong>de</strong>d function f : Ω → R, P n (f) → P (f).<br />

The s<strong>et</strong> L consists of all probability measures P ∈ P(Ω), un<strong>de</strong>r which (X, P ) is a Lévy<br />

process.<br />

For a constant B > 0, L B and L + B<br />

<strong>de</strong>note the s<strong>et</strong>s of all probabilities P ∈ L such that (X, P )<br />

satisfies, P [|∆X t | ≤ B ∀t : 0 ≤ t ≤ T ∞ ] = 1 for L B (jumps are boun<strong>de</strong>d by B in absolute value)<br />

and P [∆X t ≤ B ∀t : 0 ≤ t ≤ T ∞ ] = 1 for L + B<br />

(jumps are boun<strong>de</strong>d by B from above).<br />

L NA is the s<strong>et</strong> of all probability measures P ∈ L corresponding to Lévy processes <strong>de</strong>scribing<br />

mark<strong>et</strong>s with no arbitrage opportunity, as <strong>de</strong>fined in Proposition 1.8.<br />

Finally, M stands for the s<strong>et</strong> of all probability measures P ∈ P(Ω), un<strong>de</strong>r which e Xt is a<br />

martingale.


Introduction <strong>et</strong> principaux résultats<br />

Un processus <strong>de</strong> Lévy est un processus stochastique aux accroissem<strong>en</strong>ts indép<strong>en</strong>dants <strong>et</strong> stationnaires:<br />

si {X t } t≥0 est un processus <strong>de</strong> Lévy, X t −X s avec t > s est indép<strong>en</strong>dant <strong>de</strong> l’histoire<br />

du processus avant le temps s <strong>et</strong> sa loi ne dép<strong>en</strong>d pas <strong>de</strong> t ou s séparém<strong>en</strong>t mais seulem<strong>en</strong>t<br />

<strong>de</strong> t − s. C<strong>et</strong>te propriété <strong>de</strong>s accroissem<strong>en</strong>ts évoque une analogie avec <strong>de</strong>s fonctions linéaires:<br />

on peut dire que les processus <strong>de</strong> Lévy sont, dans un certain s<strong>en</strong>s, <strong>de</strong>s “processus linéaires” ou<br />

additifs.<br />

Malgré c<strong>et</strong>te simplicité appar<strong>en</strong>te, les processus <strong>de</strong> Lévy ont <strong>de</strong>s nombreuses propriétés<br />

intéressantes <strong>et</strong> constitu<strong>en</strong>t un domaine d’étu<strong>de</strong> <strong>en</strong> plein développem<strong>en</strong>t: plusieurs ouvrages ont<br />

été publiés récemm<strong>en</strong>t [17, 87] <strong>et</strong> une série <strong>de</strong> confér<strong>en</strong>ces internationales dédiées aux processus<br />

<strong>de</strong> Lévy <strong>et</strong> applications a r<strong>en</strong>contré un grand succès [6].<br />

Sur le plan <strong>de</strong> modélisation financière, les processus <strong>de</strong> Lévy fourniss<strong>en</strong>t une classe <strong>de</strong><br />

modèles avec sauts qui est à la fois suffisamm<strong>en</strong>t riche pour bi<strong>en</strong> décrire les données empiriques<br />

<strong>et</strong> assez simple pour faire beaucoup <strong>de</strong> calculs analytiquem<strong>en</strong>t. L’intérêt <strong>de</strong> tels modèles <strong>en</strong><br />

finance est principalem<strong>en</strong>t dû aux trois facteurs suivants.<br />

Premièrem<strong>en</strong>t, dans un modèle aux trajectoires continues comme un modèle <strong>de</strong> diffusion,<br />

le processus <strong>de</strong> prix se comporte localem<strong>en</strong>t comme un mouvem<strong>en</strong>t browni<strong>en</strong> <strong>et</strong> la probabilité<br />

que le prix bouge beaucoup p<strong>en</strong>dant un temps court est très p<strong>et</strong>ite si la valeur <strong>de</strong> volatilité <strong>de</strong><br />

volatilité n’est pas déraisonnablem<strong>en</strong>t gran<strong>de</strong>. Cela implique que dans <strong>de</strong> tels modèles les prix<br />

<strong>de</strong>s options “hors <strong>de</strong> la monnaie” doiv<strong>en</strong>t être beaucoup plus p<strong>et</strong>its que ce que l’on observe<br />

<strong>en</strong> réalité. Par contre, si le processus <strong>de</strong> prix peut sauter, même pour une maturité courte on<br />

ne peut pas négliger la probabilité d’un mouvem<strong>en</strong>t inatt<strong>en</strong>du du prix qui déplacerait l’option<br />

dans la monnaie.<br />

Deuxièmem<strong>en</strong>t, du point <strong>de</strong> vue <strong>de</strong> la couverture, les modèles aux trajectoires continues<br />

13


14 INTRODUCTION EN FRANCAIS<br />

correspon<strong>de</strong>nt <strong>en</strong> général aux marchés compl<strong>et</strong>s ou aux marchés qui peuv<strong>en</strong>t être complétés<br />

<strong>en</strong> ajoutant un ou <strong>de</strong>ux actifs supplém<strong>en</strong>taires comme dans les modèles à volatilité stochastique.<br />

Ceci implique que le flux terminal <strong>de</strong> n’importe quel actif conting<strong>en</strong>t est parfaitem<strong>en</strong>t<br />

repliquable, <strong>et</strong> les options cotées <strong>de</strong>vi<strong>en</strong>n<strong>en</strong>t donc redondantes. Le “mystère” est facilem<strong>en</strong>t<br />

expliquée par la prés<strong>en</strong>ce <strong>de</strong> sauts dans les prix: <strong>en</strong> prés<strong>en</strong>ce <strong>de</strong> sauts la couverture parfaite est<br />

impossible <strong>et</strong> les options perm<strong>et</strong>t<strong>en</strong>t aux participants du marché <strong>de</strong> couvrir les risques qui ne<br />

sont pas repliquables par une stratégie <strong>de</strong> trading ne faisant interv<strong>en</strong>ir que le sous-jac<strong>en</strong>t.<br />

Le troisième argum<strong>en</strong>t <strong>en</strong> faveur <strong>de</strong> l’utilisation <strong>de</strong> modèles discontinus, qui est peut-être le<br />

plus fort, est la prés<strong>en</strong>ce-même <strong>de</strong> sauts dans les prix. Figure 1 <strong>de</strong> page 28 trace l’évolution<br />

du taux <strong>de</strong> change DM/USD p<strong>en</strong>dant une pério<strong>de</strong> <strong>de</strong> <strong>de</strong>ux semaines <strong>en</strong> 1992. Sur ce graphique<br />

il y a au moins trois <strong>en</strong>droits ou le taux a bougé <strong>de</strong> plus <strong>de</strong> 100 points <strong>de</strong> base <strong>en</strong> moins <strong>de</strong> 5<br />

minutes. Il est claire que <strong>de</strong>s variations <strong>de</strong> prix comme celles-là ne peuv<strong>en</strong>t pas être expliquées<br />

par un modèle aux trajectoires continues, mais elles doiv<strong>en</strong>t être prises <strong>en</strong> compte pour une<br />

gestion <strong>de</strong> risques fiable.<br />

Quand c<strong>et</strong> étu<strong>de</strong> a été comm<strong>en</strong>cé, plusieurs thèses <strong>de</strong> doctorat avai<strong>en</strong>t déjà été sout<strong>en</strong>ues [78,<br />

79, 81, 86, 96] <strong>et</strong> quelques c<strong>en</strong>taines d’articles avai<strong>en</strong>t été publiés dans le domaine <strong>de</strong> modélisation<br />

financière avec <strong>de</strong>s processus <strong>de</strong> Lévy. Cep<strong>en</strong>dant, <strong>de</strong>ux questions majeures, qui apparaiss<strong>en</strong>t<br />

dans le titre <strong>de</strong> c<strong>et</strong>te thèse étai<strong>en</strong>t restées ouvertes.<br />

Premièrem<strong>en</strong>t, alors que la piste <strong>de</strong> recherche, privilégiée dans la litterature, était le développem<strong>en</strong>t<br />

<strong>de</strong> métho<strong>de</strong>s efficaces <strong>de</strong> valorisation <strong>de</strong> produits dérivés dans les modèles expon<strong>en</strong>tielle-Lévy,<br />

une étape ess<strong>en</strong>tielle d’utilisation d’un tel modèle est <strong>de</strong> trouver les valuers <strong>de</strong>s<br />

paramètres, où, plus généralem<strong>en</strong>t, le tripl<strong>et</strong> caractéristique du processus <strong>de</strong> Lévy sous-jac<strong>en</strong>t,<br />

qui soit compatible avec les prix <strong>de</strong>s options cotées sur le marché financier. Ce problème,<br />

connu sous le nom du problème <strong>de</strong> calibration, est un problème inverse à celui <strong>de</strong> valorisation<br />

<strong>de</strong>s options europé<strong>en</strong>nes dans un modèle expon<strong>en</strong>tielle-Lévy <strong>et</strong> il est n<strong>et</strong>tem<strong>en</strong>t plus difficile à<br />

résoudre que ce <strong>de</strong>rnier. Le problème <strong>de</strong> calibration a été traité par plusieurs auteurs (cf. par<br />

exemple [4, 35, 53, 85]) dans le cadre d’un modèle <strong>de</strong> diffusion markovi<strong>en</strong>ne, où le paramètre<br />

inconnu est la fonction <strong>de</strong> volatilité locale σ(S t , t). Cep<strong>en</strong>dant, dans le contexte <strong>de</strong> processus<br />

avec sauts, bi<strong>en</strong> que plusieurs articles propos<strong>en</strong>t <strong>de</strong>s modèles paramétriques à base <strong>de</strong> processus<br />

<strong>de</strong> Lévy [5, 22, 36, 62, 67, 71], il n’existait pas, avant c<strong>et</strong> étu<strong>de</strong>, <strong>de</strong> métho<strong>de</strong> perm<strong>et</strong>tant d’<strong>en</strong><br />

choisir un parmi c<strong>et</strong>te multitu<strong>de</strong> <strong>de</strong> modèles ni d’algorithme stable <strong>de</strong> calibration <strong>de</strong> paramètres


INTRODUCTION EN FRANCAIS 15<br />

<strong>de</strong> tels modèles à partir <strong>de</strong> prix d’options europé<strong>en</strong>nes. Dans la première partie <strong>de</strong> la thèse je<br />

développe donc une métho<strong>de</strong> non-paramétrique robuste <strong>de</strong> calibration <strong>de</strong> modèles expon<strong>en</strong>tielle-<br />

Lévy, étudie ses propriétés <strong>de</strong> converg<strong>en</strong>ce <strong>et</strong> <strong>de</strong> stabilité, décris sa réalisation numérique <strong>et</strong><br />

donne <strong>de</strong>s exemples <strong>de</strong> son utilisation. Mon approche consiste à reformuler le problème <strong>de</strong><br />

calibration comme celui <strong>de</strong> trouver un processus <strong>de</strong> Lévy risque neutre qui reproduit les prix<br />

d’options cotées avec la plus gran<strong>de</strong> précision possible <strong>et</strong> qui a l’<strong>en</strong>tropie relative minimale par<br />

rapport à un processus a priori donné. Ce problème est alors résolu <strong>en</strong> utilisant la métho<strong>de</strong> <strong>de</strong><br />

régularisation, prov<strong>en</strong>ant <strong>de</strong> la théorie <strong>de</strong> problèmes inverses mal posés [40].<br />

Le <strong>de</strong>uxième problème, traité dans c<strong>et</strong>te thèse est celui <strong>de</strong> la modélisation multidim<strong>en</strong>sionnelle<br />

avec <strong>de</strong>s processus <strong>de</strong> Lévy. Alors que la plupart d’applications financières nécessit<strong>en</strong>t<br />

un modèle multidim<strong>en</strong>sionnel perm<strong>et</strong>tant <strong>de</strong> pr<strong>en</strong>dre <strong>en</strong> compte la dép<strong>en</strong>dance <strong>en</strong>tre les actifs,<br />

la quasi-totalité <strong>de</strong> modèles paramétriques disponibles dans la littérature ne s’appliqu<strong>en</strong>t<br />

qu’au cas d’un seul actif. Les <strong>de</strong>ux métho<strong>de</strong>s perm<strong>et</strong>tant <strong>de</strong> construire <strong>de</strong>s modèles multidim<strong>en</strong>sionnels<br />

à base <strong>de</strong> processus <strong>de</strong> Lévy que l’on trouve dans la littérature concern<strong>en</strong>t soit le<br />

cas d’un mouvem<strong>en</strong>t browni<strong>en</strong> multivarié changé <strong>de</strong> temps par un processus <strong>de</strong> Lévy croissant<br />

unidim<strong>en</strong>sionnel [79] soit le cas où toutes les composantes ont une int<strong>en</strong>sité finie <strong>de</strong> sauts [65].<br />

Dans les <strong>de</strong>ux cas, la gamme <strong>de</strong> structures <strong>de</strong> dép<strong>en</strong>dance que l’on peut obt<strong>en</strong>ir est très limitée.<br />

La <strong>de</strong>uxième partie <strong>de</strong> c<strong>et</strong>te thèse est alors consacrée au développem<strong>en</strong>t d’une méthodologie<br />

générale perm<strong>et</strong>tant <strong>de</strong> caractériser les structures <strong>de</strong> dép<strong>en</strong>dance <strong>en</strong>tre les composantes d’un<br />

processus <strong>de</strong> Lévy multidim<strong>en</strong>sionnel <strong>et</strong> <strong>de</strong> construire <strong>de</strong>s modèles multidim<strong>en</strong>sionnels à base<br />

<strong>de</strong> processus <strong>de</strong> Lévy. C<strong>et</strong> objectif est atteint grâce à l’introduction <strong>de</strong> la notion <strong>de</strong> copule <strong>de</strong><br />

Lévy, qui peut être considérée comme l’analogue pour les processus <strong>de</strong> Lévy <strong>de</strong> la notion <strong>de</strong><br />

copule, utilisée <strong>en</strong> statistique pour modéliser la dép<strong>en</strong>dance <strong>en</strong>tre les variables aléatoires réelles<br />

[56, 76].<br />

Principaux résultats<br />

Le premier chapitre <strong>de</strong> la thèse comm<strong>en</strong>ce par une brève prés<strong>en</strong>tation <strong>de</strong>s processus <strong>de</strong> Lévy,<br />

complétée d’un recueil <strong>de</strong> leurs principales propriétés, utilisées dans la suite <strong>de</strong> la thèse, <strong>et</strong> <strong>de</strong>s<br />

modèles exp-Lévy, c’est-à-dire, les modèles où le prix d’une action est décrit par l’expon<strong>en</strong>tielle<br />

d’un processus <strong>de</strong> Lévy. Dans la section suivante je passe <strong>en</strong> revue les différ<strong>en</strong>ts modèles exp-


16 INTRODUCTION EN FRANCAIS<br />

Lévy paramétriques, disponibles dans la littérature financière. La <strong>de</strong>rnière section <strong>de</strong> ce chapitre<br />

décrit une métho<strong>de</strong>, due à Carr <strong>et</strong> Madan [23], perm<strong>et</strong>tant <strong>de</strong> valoriser les options europé<strong>en</strong>nes<br />

dans les modèles exp-Lévy à l’ai<strong>de</strong> <strong>de</strong> la transformée <strong>de</strong> Fourier. La métho<strong>de</strong> est fondée sur<br />

l’observation suivante: si on soustrait du prix <strong>de</strong> l’option call sa valeur intrinsèque:<br />

z T (k) = e −rT E[(e rT +X T<br />

− e k ) + ] − (1 − e k−rT ) + ,<br />

alors la quantité qui reste est, sous certaines conditions, intégrable <strong>et</strong> on peut évaluer sa transformée<br />

<strong>de</strong> Fourier:<br />

ζ T (v) :=<br />

∫ +∞<br />

−∞<br />

e ivk z T (k)dk = e ivrT Φ T (v − i) − 1<br />

,<br />

iv(1 + iv)<br />

où Φ T est la fonction caractéristique <strong>de</strong> X T . Les prix d’options peuv<strong>en</strong>t donc être évalués <strong>en</strong><br />

calculant la transformée <strong>de</strong> Fourier inverse <strong>de</strong> ζ T .<br />

C<strong>et</strong>te section est la seule du premier chapitre à cont<strong>en</strong>ir <strong>de</strong>s résultats originaux. Premièrem<strong>en</strong>t,<br />

j’ai démontré qu’on peut diminuer considérablem<strong>en</strong>t l’erreur <strong>de</strong> troncature dans le calcul<br />

<strong>de</strong> la transformée <strong>de</strong> Fourier inverse <strong>en</strong> remplaçant la valeur intrinsèque <strong>de</strong> l’option par son prix<br />

dans le modèle <strong>de</strong> Black <strong>et</strong> Scholes: si on dénote<br />

˜z T (k) = e −rT E[(e rT +X T<br />

− e k ) + ] − C Σ BS(k),<br />

où CBS Σ (k) est le prix Black <strong>et</strong> Scholes d’une option call avec volatilité Σ <strong>et</strong> log-strike k, <strong>et</strong><br />

Φ Σ T (v) = exp(− Σ2 T<br />

2 (v2 + iv)), alors la transformée <strong>de</strong> Fourier <strong>de</strong> ˜z T (k) est donnée par<br />

˜ζ T (v) = e ivrT Φ T (v − i) − Φ Σ T<br />

(v − i)<br />

.<br />

iv(1 + iv)<br />

Pour presque tous les modèles paramétriques, discutés dans la littérature c<strong>et</strong>te quantité converge<br />

vers zéro plus vite que toute puissance <strong>de</strong> |v| lorsque |v| → ∞ (notons que ζ T (v) converge<br />

seulem<strong>en</strong>t comme |v| −2 ).<br />

Le <strong>de</strong>uxième apport <strong>de</strong> c<strong>et</strong>te section est le développem<strong>en</strong>t d’une métho<strong>de</strong> <strong>de</strong> contrôle d’erreur<br />

pour l’algorithme <strong>de</strong> Carr <strong>et</strong> Madan. Supposons que la transformée <strong>de</strong> Fourier inverse <strong>de</strong> ˜ζ T<br />

est approchée comme suit:<br />

∫<br />

1 ∞<br />

e −ivk ˜ζT (v)dv =<br />

2π −∞<br />

N−1<br />

L ∑<br />

2π(N − 1)<br />

m=0<br />

w m ˜ζT (v m )e −ikvm + ε T + ε D , (1)


INTRODUCTION EN FRANCAIS 17<br />

où ε T est l’erreur <strong>de</strong> troncature, ε D est l’erreur <strong>de</strong> discrétisation, v m = −L/2 + m∆, ∆ =<br />

L/(N − 1) est le pas <strong>de</strong> discrétisation <strong>et</strong> w m sont les poids qui correspon<strong>de</strong>nt à la métho<strong>de</strong><br />

d’intégration choisie. Alors, notant le tripl<strong>et</strong> caractéristique <strong>de</strong> X par (A, ν, γ), l’erreur <strong>de</strong><br />

troncature peut être évaluée comme suit:<br />

1. Si on suppose A > 0 alors l’erreur <strong>de</strong> troncature satisfait:<br />

|ε T | ≤<br />

8<br />

πT Σ 2 L 3 e− T L2 Σ 2 8<br />

8 +<br />

πT AL 3 e− T L2 A<br />

8<br />

2. Si on suppose que ν a une <strong>de</strong>nsité <strong>de</strong> la forme ν(x) = e−x<br />

|x|<br />

f(x), où f est une fonction<br />

croissante sur (−∞, 0) <strong>et</strong> décroissante sur (0, ∞) alors l’erreur <strong>de</strong> troncature satisfait:<br />

8<br />

|ε T | ≤<br />

πT Σ 2 L 3 e− T L2 Σ 2<br />

8 + |Φ T (L/2 − i)| + |Φ T (−L/2 − i)|<br />

.<br />

πL<br />

Si l’intégrale dans l’équation (1) est approchée <strong>en</strong> utilisant la métho<strong>de</strong> <strong>de</strong> trapèzes, l’erreur<br />

<strong>de</strong> discrétisation ε D satisfait<br />

|ε D | ≤ ∆2<br />

6π<br />

2∑<br />

l=0<br />

C 3−l + C Σ 3−l<br />

(3 − l)!<br />

×<br />

{ (<br />

∆ + π 2<br />

( √ )<br />

)<br />

e |k−rT | L L<br />

+ log<br />

2 + 2<br />

4 + 1<br />

}<br />

|k − rT | l<br />

,<br />

l!<br />

où<br />

⎧<br />

⎪⎨<br />

C k =<br />

⎪⎩<br />

|Φ (k)<br />

T<br />

(−i)|, k pair<br />

|Φ (k+1)<br />

T<br />

(−i)| k<br />

k+1 , k impair<br />

<strong>et</strong> C Σ k sont calculés <strong>en</strong> utilisant la même formule à partir <strong>de</strong> la fonction caractéristique ΦΣ T .<br />

Calibration <strong>de</strong> modèles expon<strong>en</strong>tielle-Lévy<br />

Le <strong>de</strong>uxième chapitre est consacré au traitem<strong>en</strong>t théorique du problème <strong>de</strong> calibration. Je<br />

comm<strong>en</strong>ce par discuter la calibration au s<strong>en</strong>s <strong>de</strong> moindres carrés linéaires, qui est la métho<strong>de</strong><br />

couramm<strong>en</strong>t utilisée par les pratici<strong>en</strong>s <strong>et</strong> chercheurs. Dans c<strong>et</strong>te approche, pour trouver une<br />

solution, on minimise la somme pondérée <strong>de</strong>s carrés <strong>de</strong>s écarts <strong>en</strong>tre les prix <strong>de</strong> marché <strong>et</strong> les<br />

prix du modèle:


18 INTRODUCTION EN FRANCAIS<br />

Calibration au s<strong>en</strong>s <strong>de</strong> moindres carrés non-linéaires.<br />

Etant donnés les prix C M <strong>de</strong>s<br />

options cotées sur le marché financier, trouver un modèle expon<strong>en</strong>tielle-Lévy risque neutre Q ∗ ,<br />

tel que<br />

‖C M − C Q∗ ‖ 2 w = inf<br />

Q ‖C M − C Q ‖ 2 w,<br />

où le inf est calculé sur tous les modèles expon<strong>en</strong>tielle-Lévy risque neutres, C Q dénote les prix<br />

calculés dans le modèle Q <strong>et</strong><br />

‖C M − C Q ‖ 2 w :=<br />

N∑<br />

w i (C M (T i , K i ) − C Q (T i , K i )) 2 .<br />

i=1<br />

L’<strong>en</strong>semble <strong>de</strong>s solutions du problème <strong>de</strong> calibration au s<strong>en</strong>s <strong>de</strong> moindres carrés sera noté<br />

Q LS . Je comm<strong>en</strong>ce par démontrer que dans le contexte <strong>de</strong> calibration non-paramétrique <strong>de</strong><br />

modèles exp-Lévy c<strong>et</strong>te métho<strong>de</strong> ne perm<strong>et</strong> pas toujours <strong>de</strong> trouver une solution <strong>et</strong> est, dans<br />

certaines situations, instable par rapport aux perturbations <strong>de</strong>s données d’<strong>en</strong>trée C M . Ensuite<br />

je donne <strong>de</strong>s conditions suffisantes (assez restrictives), sous lesquelles le problème <strong>de</strong> calibration<br />

adm<strong>et</strong> une solution continue par rapport aux données d’<strong>en</strong>trée (Théorème 2.1 <strong>et</strong> Proposition<br />

2.5). Comme ces conditions sont rarem<strong>en</strong>t vérifiées <strong>en</strong> pratique, il est nécessaire <strong>de</strong> trouver<br />

une métho<strong>de</strong> plus fiable <strong>de</strong> résolution du problème <strong>de</strong> calibration. Le premier pas dans c<strong>et</strong>te<br />

direction est <strong>de</strong> le reformuler comme le problème <strong>de</strong> trouver un processus <strong>de</strong> Lévy risque neutre<br />

qui reproduit les prix d’options cotées avec la plus gran<strong>de</strong> précision possible <strong>et</strong> qui a l’<strong>en</strong>tropie<br />

relative minimale par rapport à un processus a priori à donné parmi toutes les solutions du<br />

problème <strong>de</strong> calibration au s<strong>en</strong>s <strong>de</strong> moindres carrés:<br />

Calibration au s<strong>en</strong>s <strong>de</strong> moindres carrés avec minimisation d’<strong>en</strong>tropie<br />

Etant donnés<br />

les prix C M <strong>de</strong>s options cotées sur le marché financier, <strong>et</strong> un processus <strong>de</strong> Lévy a priori P ,<br />

trouver un modèle expon<strong>en</strong>tielle-Lévy risque neutre Q ∗ ∈ Q LS , tel que<br />

I(Q ∗ |P ) =<br />

inf I(Q|P ).<br />

Q∈QLS Les solutions du problème ci-<strong>de</strong>ssus seront appelées solutions au s<strong>en</strong>s <strong>de</strong> moindres carrés


INTRODUCTION EN FRANCAIS 19<br />

d’<strong>en</strong>tropie minimale. Le li<strong>en</strong> <strong>de</strong> c<strong>et</strong>te approche avec d’autres métho<strong>de</strong>s <strong>de</strong> calibration <strong>et</strong> <strong>de</strong><br />

valorisation d’options à l’ai<strong>de</strong> <strong>de</strong> l’<strong>en</strong>tropie relative, disponibles dans la littérature, est discuté<br />

dans section 2.3.<br />

C<strong>et</strong>te formulation résout le problème <strong>de</strong> non-i<strong>de</strong>ntifiabilité, mais souffre toujours <strong>de</strong> manque<br />

<strong>de</strong> stabilité. Cep<strong>en</strong>dant, elle peut être transformée <strong>en</strong> un problème <strong>de</strong> calibration stable <strong>et</strong><br />

adm<strong>et</strong>tant toujours une solution, <strong>en</strong> utilisant la métho<strong>de</strong> <strong>de</strong> régularisation, prov<strong>en</strong>ant <strong>de</strong> la<br />

théorie <strong>de</strong> problèmes inverses mal-posés. Pour régulariser le problème <strong>de</strong> calibration au s<strong>en</strong>s <strong>de</strong><br />

moindres carrés, je propose <strong>de</strong> minimiser la somme <strong>de</strong> l’erreur <strong>de</strong> valorisation <strong>et</strong> d’un terme <strong>de</strong><br />

pénalisation donné par l’<strong>en</strong>tropie relative du modèle par rapport à P :<br />

Problème <strong>de</strong> calibration régularisé Etant donnés les prix C M <strong>de</strong>s options cotées sur<br />

le marché <strong>et</strong> un processus <strong>de</strong> Lévy a priori P , trouver un modèle expon<strong>en</strong>tielle-Lévy risque<br />

neutre Q ∗ , tel que<br />

J α (Q ∗ ) = inf<br />

Q J α(Q),<br />

où le inf est calculé sur tous les modèles expon<strong>en</strong>tielle-Lévy risque neutres,<br />

J α (Q) = ‖C M − C Q ‖ 2 w + αI(Q|P )<br />

<strong>et</strong> α est le paramètre <strong>de</strong> régularisation qui détermine l’int<strong>en</strong>sité <strong>de</strong> pénalisation.<br />

Dans le cadre <strong>de</strong> processus <strong>de</strong> Lévy, l’<strong>en</strong>tropie relative peut être calculée explicitem<strong>en</strong>t<br />

(Théorème 2.9):<br />

I(Q|P ) = I(Q| FT∞ |P | FT∞ ) = T { ∫ 1<br />

2<br />

∞<br />

γ Q − γ P − x(ν Q − ν )(dx)} P 1 A≠0 +<br />

2A<br />

−1<br />

∫ ∞<br />

( )<br />

dν<br />

Q<br />

dνQ<br />

T ∞ log<br />

dνP dν P + 1 − dνQ<br />

dν P ν P (dx).<br />

C<strong>et</strong>te formule perm<strong>et</strong> d’étudier les propriétés <strong>de</strong> solutions du problème <strong>de</strong> calibration régularisé<br />

<strong>et</strong> démontrer les résultats suivants, sous condition que les sauts du processus a priori P sont<br />

bornés supérieurem<strong>en</strong>t par une constante B <strong>et</strong> que P correspond à un modèle expon<strong>en</strong>tielle-Lévy<br />

sans opportunité d’arbitrage<br />

−∞<br />

• Le problème <strong>de</strong> calibration régularisé adm<strong>et</strong> au moins une solution.


20 INTRODUCTION EN FRANCAIS<br />

• Si P satisfait une condition <strong>de</strong> régularité supplém<strong>en</strong>taire (2.28), alors la solution Q est une<br />

mesure équival<strong>en</strong>te à P (<strong>en</strong> général, Q est une mesure martingale, absolum<strong>en</strong>t continue<br />

par rapport à P ).<br />

• Les solutions du problème <strong>de</strong> calibration régularisé sont continues par rapport aux données<br />

<strong>de</strong> marché: si {C n M } n≥1 est une suite <strong>de</strong> données telle que ‖C n M − C M‖ w → 0, <strong>et</strong> pour<br />

chaque n, Q n est la solution du problème <strong>de</strong> calibration régularisé avec donnée C n M , loi<br />

a priori P <strong>et</strong> paramètre <strong>de</strong> régularisation α, alors {Q n } n≥1 a une sous-suite faiblem<strong>en</strong>t<br />

converg<strong>en</strong>te, <strong>et</strong> la limite <strong>de</strong> chaque sous-suite faiblem<strong>en</strong>t converg<strong>en</strong>te <strong>de</strong> {Q n } n≥1 est la<br />

solution du problème <strong>de</strong> calibration régularisé avec donnée C M , loi a priori P <strong>et</strong> paramètre<br />

<strong>de</strong> régularisation α.<br />

• Lorsque le niveau d’erreur dans les données <strong>de</strong> marché t<strong>en</strong>d vers 0, si le paramètre <strong>de</strong><br />

régularisation α est choisi <strong>de</strong> façon appropriée, alors les solutions du problème régularisé<br />

converg<strong>en</strong>t vers les solutions au s<strong>en</strong>s <strong>de</strong> moindres carrés d’<strong>en</strong>tropie minimale (lorsqu’il<br />

existe plusieurs solutions au s<strong>en</strong>s <strong>de</strong> moindres carrés d’<strong>en</strong>tropie minimale, la converg<strong>en</strong>ce<br />

est <strong>en</strong>t<strong>en</strong>due dans le même s<strong>en</strong>s que ci-<strong>de</strong>ssus).<br />

Le troisième chapitre traite la résolution numérique du problème <strong>de</strong> calibration régularisé,<br />

construit au chapitre précé<strong>de</strong>nt. Tout d’abord, pour se ram<strong>en</strong>er dans un espace fini-dim<strong>en</strong>sionnel,<br />

la mesure <strong>de</strong> Lévy du processus a priori P est discrétisée sur une grille:<br />

ν P =<br />

M−1<br />

∑<br />

k=0<br />

p k δ {xk }(dx), (2)<br />

ce qui implique que la mesure <strong>de</strong> Lévy <strong>de</strong> la solution Q a la même forme:<br />

ν Q =<br />

M−1<br />

∑<br />

k=0<br />

q k δ {xk }(dx).<br />

Ensuite, je démontre que la solution du problème <strong>de</strong> calibration avec une loi a priori quelconque<br />

peut être approchée avec une précision arbitraire par une suite <strong>de</strong> solutions avec lois a priori <strong>de</strong><br />

la forme (2) (théorème 3.2 <strong>et</strong> lemme 3.1). Section 3.2 discute le choix du processus a priori <strong>et</strong><br />

estime son influ<strong>en</strong>ce sur la solution, <strong>en</strong> effectuant <strong>de</strong>s tests numériques. La conclusion <strong>de</strong> c<strong>et</strong>te<br />

section est que la solution est peu s<strong>en</strong>sible aux p<strong>et</strong>ites variations du processus a priori, mais<br />

qu’il est important <strong>de</strong> spécifier sa forme qualitative correctem<strong>en</strong>t.


INTRODUCTION EN FRANCAIS 21<br />

Section 3.3 prés<strong>en</strong>te un algorithme <strong>de</strong> choix du paramètre <strong>de</strong> régularisation α “a posteriori” à<br />

partir <strong>de</strong>s données. Ma métho<strong>de</strong>, qui est une adaptation du principe <strong>de</strong> discrépance, développé<br />

par Morozov [74] pour la régularisation <strong>de</strong> Tikhonov dans les espaces <strong>de</strong> Banach, consiste à<br />

choisir <strong>de</strong>ux constantes c 1 <strong>et</strong> c 2 proches <strong>de</strong> 1, telles que 1 < c 1 < c 2 <strong>et</strong> trouver la valeur <strong>de</strong><br />

paramètre α pour laquelle c 1 δ 2 ≤ ε δ (α) ≤ c 2 δ 2 , où δ est le niveau <strong>de</strong> bruit dans les données <strong>et</strong><br />

la discrépance ε δ (α) est définie par<br />

ε δ (α) := ‖C Qδ α<br />

− CM‖ δ 2 ,<br />

avec Q δ α la solution du problème régularisé avec paramètre <strong>de</strong> régularisation α <strong>et</strong> niveau <strong>de</strong><br />

bruit δ.<br />

Section 3.5 est consacrée à l’algorithme numérique perm<strong>et</strong>tant <strong>de</strong> résoudre le problème <strong>de</strong><br />

calibration discrétisé, une fois que le processus a priori <strong>et</strong> le paramètre <strong>de</strong> régularisation ont<br />

été choisis. Le problème est résolu <strong>en</strong> minimisant la fonctionnelle <strong>de</strong> calibration approchée<br />

Ĵ α (Q) par la métho<strong>de</strong> BFGS, faisant interv<strong>en</strong>ir le gradi<strong>en</strong>t <strong>de</strong> la fonctionnelle à minimiser. La<br />

fonctionnelle approchée est donnée par<br />

Ĵ α (Q) =<br />

N∑<br />

w i (ĈQ (T i , K i ) − C M (T i , K i )) 2<br />

i=1<br />

⎛<br />

⎞<br />

+ α ⎝ A M−1<br />

2A 2 + ∑<br />

bP + (e x j<br />

− 1)q j<br />

⎠<br />

j=0<br />

2<br />

M−1<br />

∑<br />

+ α (q j log(q j /p j ) + 1 − q j ) ,<br />

j=0<br />

où ĈQ (T i , K i ) est l’approximation du prix d’une option call avec maturité T i <strong>et</strong> prix d’exercice<br />

K i , calculée dans le modèle Q <strong>en</strong> utilisant la métho<strong>de</strong> <strong>de</strong> transformée <strong>de</strong> Fourier. Le gradi<strong>en</strong>t<br />

<strong>de</strong> la fonctionnelle <strong>de</strong> calibration est évalué analytiquem<strong>en</strong>t <strong>en</strong> utilisant l’équation (3.32).<br />

Dans section 3.6 l’algorithme est appliqué d’abord aux données simulées (à partir d’un<br />

modèle exp-Lévy avec une mesure <strong>de</strong> Lévy connue) <strong>et</strong> <strong>en</strong>suite aux données réelles. Les tests<br />

sur <strong>de</strong>s données artificielles m<strong>et</strong>t<strong>en</strong>t <strong>en</strong> évi<strong>de</strong>nce la stabilité <strong>de</strong> l’algorithme <strong>et</strong> sa capacité <strong>de</strong><br />

reconstruire la vraie mesure <strong>de</strong> Lévy partout sauf dans un p<strong>et</strong>it voisinage <strong>de</strong> 0 (parce que les<br />

sauts <strong>de</strong> taille 0 n’ont pas d’influ<strong>en</strong>ce sur les prix d’options). Les tests sur les données réelles<br />

perm<strong>et</strong>t<strong>en</strong>t <strong>de</strong> tirer un nombre <strong>de</strong> conclusions importantes.<br />

- Premièrem<strong>en</strong>t, un modèle exp-Lévy perm<strong>et</strong> <strong>de</strong> calibrer <strong>de</strong>s prix d’options d’une seule<br />

maturité avec une gran<strong>de</strong> précision. C<strong>et</strong>te conclusion contredit les résultats <strong>de</strong> Medve<strong>de</strong>v


22 INTRODUCTION EN FRANCAIS<br />

<strong>et</strong> Scaill<strong>et</strong> [70] qui ont observé que “les sauts dans les prix ne perm<strong>et</strong>t<strong>en</strong>t pas d’expliquer<br />

la p<strong>en</strong>te <strong>de</strong> volatilité implicite à la monnaie”. Toutefois, il est important <strong>de</strong> remarquer<br />

que l’étu<strong>de</strong> <strong>de</strong> Medve<strong>de</strong>v <strong>et</strong> Scaill<strong>et</strong> porte sur les options sur l’indice S&P 500 <strong>de</strong> courte<br />

maturité, tandis que les tests <strong>de</strong> c<strong>et</strong>te thèse ont été effectués <strong>en</strong> utilisant les options sur<br />

l’indice DAX.<br />

- De plus, la qualité <strong>de</strong> calibration est déjà excell<strong>en</strong>te <strong>en</strong> utilisant seulem<strong>en</strong>t <strong>de</strong>s modèles à int<strong>en</strong>sité<br />

finie <strong>de</strong> sauts, ce qui rem<strong>et</strong> <strong>en</strong> question la nécessité, du point <strong>de</strong> vue <strong>de</strong> modélisation<br />

du smile <strong>de</strong> volatilité, <strong>de</strong>s modèles plus compliqués à int<strong>en</strong>sité infinie.<br />

- La troisième conclusion est que même dans le cadre non-paramétrique, il est impossible,<br />

<strong>en</strong> utilisant un modèle exp-Lévy, <strong>de</strong> calibrer <strong>de</strong>s prix d’options <strong>de</strong> plusieurs maturités<br />

avec une précision satisfaisante. Les tests ont donc confirmé l’observation déjà faite par<br />

certains auteurs [13, 68] que le cadre <strong>de</strong> modèles exp-Lévy est trop rigi<strong>de</strong> pour pouvoir<br />

décrire correctem<strong>en</strong>t la structure par terme <strong>de</strong> volatilités implicites.<br />

Modélisation multidim<strong>en</strong>sionnelle avec les processus <strong>de</strong> Lévy<br />

Le quatrième chapitre comm<strong>en</strong>ce par une revue <strong>de</strong> <strong>de</strong>ux métho<strong>de</strong>s, disponibles dans la littérature,<br />

perm<strong>et</strong>tant <strong>de</strong> construire <strong>de</strong>s modèles multidim<strong>en</strong>sionnels à base <strong>de</strong> processus <strong>de</strong> Lévy. Dans la<br />

première approche il s’agit <strong>de</strong> changer l’échelle <strong>de</strong> temps d’un mouvem<strong>en</strong>t browni<strong>en</strong> multivarié<br />

par un processus <strong>de</strong> Lévy croissant, le même pour toutes les composantes [79], <strong>et</strong> la <strong>de</strong>uxième<br />

métho<strong>de</strong>, qui ne concerne que le cas <strong>de</strong> processus <strong>de</strong> Lévy à int<strong>en</strong>sité finie <strong>de</strong> sauts, consiste à<br />

modéliser directem<strong>en</strong>t la dép<strong>en</strong>dance <strong>en</strong>tre les tailles <strong>de</strong>s sauts dans les différ<strong>en</strong>tes composantes,<br />

<strong>en</strong> utilisant les copules. Malheureusem<strong>en</strong>t, dans ces métho<strong>de</strong>s la gamme <strong>de</strong> types <strong>de</strong> dép<strong>en</strong>dance<br />

possibles est très limitée <strong>et</strong> <strong>de</strong>s contraintes sont imposées sur le choix <strong>de</strong> modèles paramétriques<br />

pour les composantes (on ne peut pas coupler <strong>de</strong>s composantes quelconques). La discussion<br />

<strong>de</strong>s défauts <strong>de</strong> ces métho<strong>de</strong>s perm<strong>et</strong> d’établir une liste <strong>de</strong> conditions qu’une bonne métho<strong>de</strong> <strong>de</strong><br />

modélisation multidim<strong>en</strong>sionnelle doit satisfaire:<br />

• On doit pouvoir choisir librem<strong>en</strong>t un processus <strong>de</strong> Lévy unidim<strong>en</strong>sionnel pour chacune <strong>de</strong><br />

composantes.


INTRODUCTION EN FRANCAIS 23<br />

• La gamme <strong>de</strong> structures <strong>de</strong> dép<strong>en</strong>dance possibles doit inclure l’indép<strong>en</strong>dance <strong>et</strong> la dép<strong>en</strong>dance<br />

complète avec une transition continue <strong>en</strong>tre ces <strong>de</strong>ux cas extrêmes.<br />

• La métho<strong>de</strong> doit perm<strong>et</strong>tre <strong>de</strong> construire <strong>de</strong>s modèles <strong>de</strong> dép<strong>en</strong>dance paramétriques.<br />

Pour atteindre ces conditions, je propose <strong>de</strong> séparer la structure <strong>de</strong> dép<strong>en</strong>dance d’un processus<br />

<strong>de</strong> Lévy <strong>de</strong>s lois marginales <strong>de</strong> ces composantes. Dans le cadre plus restreint <strong>de</strong> variables<br />

aléatoires à valeurs dans R d , c<strong>et</strong>te séparation est réalisée par la notion <strong>de</strong> copule. Je développe<br />

une notion analogue, applicable au cas <strong>de</strong>s processus <strong>de</strong> Lévy.<br />

Section 4.4 introduit les copules <strong>de</strong> Lévy dans le cas plus simple <strong>de</strong> processus <strong>de</strong> Lévy<br />

n’adm<strong>et</strong>tant que <strong>de</strong>s sauts positifs dans chacune <strong>de</strong> composantes. Dans ce cas la copule <strong>de</strong><br />

Lévy est définie comme une fonction F : [0, ∞] d → [0, ∞] avec les propriétés suivantes:<br />

1. F (u 1 , . . . , u d ) = 0 dès que u i = 0 pour au moins un i ∈ {1, . . . , d},<br />

2. F est une fonction croissante <strong>en</strong> d dim<strong>en</strong>sions, c.-à-d., pour tout a, b ∈ Dom F ,<br />

∑<br />

(−1) N(c) F (c) ≥ 0,<br />

où la somme est calculé sur tous les somm<strong>et</strong>s c <strong>de</strong> [a, b] <strong>et</strong> N(c) := #{k : c k = a k }.<br />

3. F a les fonctions marginales uniformes: F (u 1 , . . . , u d )| ui =∞,i≠k = u k pour tout k ∈<br />

{1, . . . , d}, u k ∈ [0, ∞].<br />

En représ<strong>en</strong>tant une mesure <strong>de</strong> Lévy ν sur R d + par son intégrale <strong>de</strong> queue:<br />

U(x 1 , . . . , x d ) = ν([x 1 , ∞) × · · · × [x d , ∞)),<br />

je démontre que<br />

• Pour toute mesure <strong>de</strong> Lévy ν sur R d + ayant l’intégrale <strong>de</strong> queue U <strong>et</strong> les mesures <strong>de</strong> Lévy<br />

marginales ν 1 , . . . , ν d , il existe une copule <strong>de</strong> Lévy F sur [0, ∞] d , telle que<br />

U(x 1 , . . . , x d ) = F (U 1 (x 1 ), . . . , U d (x d )), (x 1 , . . . , x d ) ∈ [0, ∞) d , (3)<br />

où U 1 , . . . , U d sont <strong>de</strong>s intégrales <strong>de</strong> queue <strong>de</strong> ν 1 , . . . , ν d .<br />

• Si F est une copule <strong>de</strong> Lévy sur [0, ∞] d <strong>et</strong> ν 1 , . . . , ν d sont <strong>de</strong>s mesures <strong>de</strong> Lévy sur (0, ∞)<br />

avec les intégrales <strong>de</strong> queue U 1 , . . . , U d , alors équation (3) définit une intégrale <strong>de</strong> queue<br />

d’une mesure <strong>de</strong> Lévy sur R d + ayant les mesures <strong>de</strong> Lévy marginales ν 1 , . . . , ν d .


24 INTRODUCTION EN FRANCAIS<br />

Ce résultat montre que pour chaque processus <strong>de</strong> Lévy il existe une copule <strong>de</strong> Lévy qui décrit<br />

sa structure <strong>de</strong> dép<strong>en</strong>dance, <strong>et</strong> que pour chaque copule <strong>de</strong> Lévy <strong>et</strong> chaque combinaison <strong>de</strong> lois<br />

unidim<strong>en</strong>sionnelles il existe un processus <strong>de</strong> Lévy avec <strong>de</strong>s lois <strong>de</strong> composantes données <strong>et</strong> dont<br />

la dép<strong>en</strong>dance est décrite par c<strong>et</strong>te copule <strong>de</strong> Lévy.<br />

Section 4.5 ét<strong>en</strong>d la notion <strong>de</strong> copule <strong>de</strong> Lévy <strong>et</strong> les résultats associés aux processus <strong>de</strong> Lévy<br />

généraux. Les résultats <strong>de</strong> c<strong>et</strong>te section ont été obt<strong>en</strong>us dans un travail joint avec Jan Kalls<strong>en</strong><br />

[59]. En particulier, la structure <strong>de</strong> dép<strong>en</strong>dance d’un processus <strong>de</strong> Lévy général est caractérisé<br />

par une copule <strong>de</strong> Lévy sur (−∞, ∞] d , c.-à-d., une fonction F : (−∞, ∞] d → (−∞, ∞] avec les<br />

propriétés suivantes:<br />

1. F (u 1 , . . . , u d ) < ∞ si (u 1 , . . . , u d ) ≠ (∞, . . . , ∞),<br />

2. F (u 1 , . . . , u d ) = 0 si u i = 0 pour au moins un i ∈ {1, . . . , d},<br />

3. F est une fonction croissante <strong>en</strong> d dim<strong>en</strong>sions,<br />

4. F a les fonctions marginales uniformes: pour tout k ∈ {1, . . . , d} <strong>et</strong> pour tout x k ∈<br />

(−∞, ∞],<br />

lim<br />

c→∞<br />

∑<br />

sgn x j = x k .<br />

(x j ) j≠k ∈{−c,∞} d−1 F (x 1 , . . . , x d ) ∏ j≠k<br />

Dans la <strong>de</strong>rnière section <strong>de</strong> ce chapitre je calcule les copules <strong>de</strong> Lévy qui correspon<strong>de</strong>nt aux<br />

types <strong>de</strong> dép<strong>en</strong>dance particuliers. Les composantes d’un processus <strong>de</strong> Lévy multidim<strong>en</strong>sionnel<br />

sans partie martingale continue sont indép<strong>en</strong>dantes si <strong>et</strong> seulem<strong>en</strong>t s’il a une copule <strong>de</strong> Lévy<br />

suivante:<br />

d∑ ∏<br />

F ⊥ (x 1 , . . . , x d ) := x i 1 {∞} (x j ).<br />

i=1 j≠i<br />

Les sauts d’un processus <strong>de</strong> Lévy d-dim<strong>en</strong>sionnel sont dits complètem<strong>en</strong>t dép<strong>en</strong>dants s’il existe<br />

un sous-<strong>en</strong>semble strictem<strong>en</strong>t ordonné S <strong>de</strong> K := {x ∈ R d : sgn x 1 = · · · = sgn x d }, tel que<br />

∆X t := X t − X t− ∈ S, t ≥ 0. La copule <strong>de</strong> Lévy <strong>de</strong> la dép<strong>en</strong>dance complète est<br />

d∏<br />

F ‖ (x 1 , . . . , x d ) := min(|x 1 |, . . . , |x d |)1 K (x 1 , . . . , x d ) sgn x i .<br />

Le <strong>de</strong>rnier chapitre <strong>de</strong> la thèse est dédié aux applications <strong>de</strong> copules <strong>de</strong> Lévy <strong>en</strong> finance.<br />

Dans les applications on n’a souv<strong>en</strong>t pas assez d’information sur la dép<strong>en</strong>dance pour employer<br />

i=1


INTRODUCTION EN FRANCAIS 25<br />

<strong>de</strong>s métho<strong>de</strong>s non-paramétriques <strong>et</strong> je comm<strong>en</strong>ce donc par donner <strong>de</strong>s outils perm<strong>et</strong>tant <strong>de</strong><br />

construire <strong>de</strong>s familles paramétriques <strong>de</strong> copules <strong>de</strong> Lévy. Un exemple <strong>de</strong> famille <strong>de</strong> copules <strong>de</strong><br />

Lévy <strong>en</strong> <strong>de</strong>ux dim<strong>en</strong>sions, dép<strong>en</strong>dant <strong>de</strong> <strong>de</strong>ux paramètres est donné par<br />

F (u, v) = (|u| −θ + |v| −θ ) −1/θ (η1 uv≥0 − (1 − η)1 uv


26 INTRODUCTION EN FRANCAIS<br />

dans un tel modèle <strong>en</strong> utilisant la métho<strong>de</strong> <strong>de</strong> Monte Carlo. Pour estimer l’importance <strong>de</strong> la<br />

dép<strong>en</strong>dance dans les queues <strong>de</strong> distribution pour la valorisation <strong>de</strong>s options, je pr<strong>en</strong>ds <strong>de</strong>ux jeux<br />

<strong>de</strong> paramètres η <strong>et</strong> θ, qui correspon<strong>de</strong>nt tous les <strong>de</strong>ux à la corrélation <strong>de</strong> r<strong>en</strong><strong>de</strong>m<strong>en</strong>ts <strong>de</strong> 50%<br />

mais ont les structures <strong>de</strong> dép<strong>en</strong>dance différ<strong>en</strong>tes. Les prix <strong>de</strong>s options sur panier pour les <strong>de</strong>ux<br />

jeux <strong>de</strong> paramètres diffèr<strong>en</strong>t <strong>de</strong> 10% à la monnaie, ce qui montre que la dép<strong>en</strong>dance dans les<br />

queues <strong>de</strong> distribution est importante à pr<strong>en</strong>dre <strong>en</strong> compte pour la valorisation d’options sur<br />

panier, <strong>et</strong> que les copules <strong>de</strong> Lévy sont un outil adapté pour décrire c<strong>et</strong>te dép<strong>en</strong>dance.


Introduction<br />

Lévy processes are <strong>de</strong>fined as stochastic processes with stationary and in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt increm<strong>en</strong>ts:<br />

if {X t } t≥0 is a Lévy process, th<strong>en</strong> X t −X s with t > s is in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt of the history of the process<br />

up to time s, and its law only <strong>de</strong>p<strong>en</strong>ds on t − s but not on t or s separately. This property of<br />

increm<strong>en</strong>ts suggests an analogy with linear functions: one can say that Lévy processes are, in<br />

some s<strong>en</strong>se, “linear processes”.<br />

Despite this appar<strong>en</strong>t simplicity, Lévy processes have many interesting properties and constitute<br />

an exciting field of study: this is shown by the rec<strong>en</strong>t publication of several monographs<br />

(see for example [17, 87]) and by the success of the series of international confer<strong>en</strong>ces on Lévy<br />

processes and applications (see [6]).<br />

From the point of view of financial mo<strong>de</strong>lling, Lévy processes provi<strong>de</strong> a class of mo<strong>de</strong>ls<br />

with jumps that is both suffici<strong>en</strong>tly rich to reproduce empirical data and simple <strong>en</strong>ough to do<br />

many computations analytically. The interest of jump mo<strong>de</strong>ls in finance is mainly due to three<br />

reasons.<br />

First, in a mo<strong>de</strong>l with continuous paths like a diffusion mo<strong>de</strong>l, the price process behaves<br />

locally like a Brownian motion and the probability that the stock moves by a large amount<br />

over a short period of time is very small, unless one fixes an unrealistically high value of the<br />

volatility of volatility. Therefore, in such mo<strong>de</strong>ls the prices of short term out of the money<br />

options should be much lower than what one observes in real mark<strong>et</strong>s. On the other hand,<br />

if stock prices are allowed to jump, ev<strong>en</strong> wh<strong>en</strong> the time to maturity is very short, there is a<br />

non-negligible probability that after a sud<strong>de</strong>n change in the stock price the option will move in<br />

the money.<br />

Second, from the hedging point of view, continuous mo<strong>de</strong>ls of stock price behavior g<strong>en</strong>erally<br />

lead to a compl<strong>et</strong>e mark<strong>et</strong> or to a mark<strong>et</strong> that can be ma<strong>de</strong> compl<strong>et</strong>e by adding one or two<br />

27


28 INTRODUCTION<br />

7500<br />

Taux <strong>de</strong> change DM/USD<br />

7400<br />

7300<br />

7200<br />

7100<br />

7000<br />

17/09/1992 2/10/1992<br />

Figure 1: Jumps in the trajectory of DM/USD exchange rate, sampled at 5-minute intervals.<br />

additional instrum<strong>en</strong>ts, like in stochastic volatility mo<strong>de</strong>ls. Since in such a mark<strong>et</strong> every terminal<br />

payoff can be exactly replicated, the very exist<strong>en</strong>ce of tra<strong>de</strong>d options becomes a puzzle. The<br />

mystery is easily solved by allowing for discontinuities: in real mark<strong>et</strong>s, due to the pres<strong>en</strong>ce of<br />

jumps in the prices, perfect hedging is impossible and options <strong>en</strong>able the mark<strong>et</strong> participants<br />

to hedge risks that cannot be hedged by using the un<strong>de</strong>rlying only.<br />

The third and the strongest argum<strong>en</strong>t for using discontinuous mo<strong>de</strong>ls is simply the pres<strong>en</strong>ce<br />

of jumps in observed prices. Figure 1 <strong>de</strong>picts the evolution of the DM/USD exchange rate over<br />

a two-week period in 1992, and one can see at least three points where the rate moved by over<br />

100 bp within a 5-minute period. Price moves like these ones clearly cannot be accounted for<br />

in the framework of a diffusion mo<strong>de</strong>l with continuous paths, but they must be <strong>de</strong>alt with if<br />

the mark<strong>et</strong> risk is to be measured and managed correctly.<br />

Although wh<strong>en</strong> this thesis was started, the field of financial mo<strong>de</strong>lling with Lévy processes<br />

was already a well-<strong>de</strong>veloped one, with several Ph. D. theses [78, 79, 81, 86, 96] and a few hundred<br />

research papers (see refer<strong>en</strong>ces in [27]) already writt<strong>en</strong> on the subject, two major issues, that<br />

appear in the title of the pres<strong>en</strong>t study, remained unresolved.<br />

First, while the main concern in the literature has be<strong>en</strong> to find effici<strong>en</strong>t analytical and<br />

numerical procedures for computing option prices in expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls, an ess<strong>en</strong>tial<br />

step in using such mo<strong>de</strong>ls is to obtain the param<strong>et</strong>ers — here the characteristic tripl<strong>et</strong> of the


INTRODUCTION 29<br />

un<strong>de</strong>rlying Lévy process — consist<strong>en</strong>t with the mark<strong>et</strong>-quoted prices of tra<strong>de</strong>d options. This<br />

problem — called mo<strong>de</strong>l calibration — is an inverse problem to that of pricing a European option<br />

in an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l and is much more difficult to solve than the latter. The calibration<br />

problem was addressed by several authors (see for example [4, 35, 53, 85]) in the framework of a<br />

markovian diffusion mo<strong>de</strong>l, the unknown param<strong>et</strong>er being the local volatility function σ(S t , t).<br />

However, in the s<strong>et</strong>ting of processes with jumps, although many papers proposed param<strong>et</strong>ric<br />

Lévy-based mo<strong>de</strong>ls [5, 22, 36, 62, 67, 71], prior to this study there existed no systematic approach<br />

to mo<strong>de</strong>l selection, and no stable calibration m<strong>et</strong>hod, that could be implem<strong>en</strong>ted in practice<br />

was available. In the first part of the thesis we <strong>de</strong>velop a non-param<strong>et</strong>ric m<strong>et</strong>hod allowing to<br />

calibrate expon<strong>en</strong>tial-Lévy mo<strong>de</strong>ls, study its stability and converg<strong>en</strong>ce properties, <strong>de</strong>scribe its<br />

numerical implem<strong>en</strong>tation and give examples of its use. Our approach is first to reformulate the<br />

calibration problem as that of finding a risk-neutral expon<strong>en</strong>tial Lévy mo<strong>de</strong>l that reproduces<br />

the observed option prices with the best possible precision and has the smallest relative <strong>en</strong>tropy<br />

with respect to a giv<strong>en</strong> prior, and th<strong>en</strong> to solve this problem via the regularization approach,<br />

used in the theory of ill-posed inverse problems [40].<br />

The second issue, addressed in this thesis, is that of multidim<strong>en</strong>sional mo<strong>de</strong>lling with Lévy<br />

process. Although most financial applications (bask<strong>et</strong> option pricing, risk managem<strong>en</strong>t, portfolio<br />

optimization <strong>et</strong>c.) require a multidim<strong>en</strong>sional mo<strong>de</strong>l taking into account the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

b<strong>et</strong>we<strong>en</strong> various ass<strong>et</strong>s, the vast majority of param<strong>et</strong>ric mo<strong>de</strong>ls, available in the literature cover<br />

only the case of a single ass<strong>et</strong>. The only attempts to construct multidim<strong>en</strong>sional Lévy-based<br />

mo<strong>de</strong>ls that were ma<strong>de</strong> prior to this study concern either the case of a multivariate Brownian<br />

motion, time-changed with a one-dim<strong>en</strong>sional increasing Lévy process [79] or the case where all<br />

compon<strong>en</strong>ts have finite jump int<strong>en</strong>sity [65]. In both cases the scope of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures<br />

that one can obtain is quite limited. In the second part of this thesis we therefore propose a<br />

g<strong>en</strong>eral m<strong>et</strong>hodology to characterize <strong>de</strong>p<strong>en</strong><strong>de</strong>nce among the compon<strong>en</strong>ts of multidim<strong>en</strong>sional<br />

Lévy processes and to construct multidim<strong>en</strong>sional exp-Lévy mo<strong>de</strong>ls. This is done by introducing<br />

the notion of Lévy copula, which can be se<strong>en</strong> as an analog for Lévy processes of the notion<br />

of copula, used in statistics to mo<strong>de</strong>l <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> real random variables [56, 76].<br />

This thesis is structured as follows. The first chapter contains a brief review of the properties<br />

of Lévy processes and expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls, that are used in the sequel. The last section<br />

<strong>de</strong>scribes a m<strong>et</strong>hod, due to Carr and Madan [23], for pricing European options in expon<strong>en</strong>tial


30 INTRODUCTION<br />

Lévy mo<strong>de</strong>ls by Fourier transform. This section is the only one of the first chapter to contain<br />

new results: we propose original improvem<strong>en</strong>ts to this m<strong>et</strong>hod and obtain estimates for the<br />

truncation and discr<strong>et</strong>ization errors, that are not giv<strong>en</strong> in the original refer<strong>en</strong>ce [23].<br />

The second chapter is <strong>de</strong>dicated to the theor<strong>et</strong>ical treatm<strong>en</strong>t of the calibration problem.<br />

We start by discussing the least squares calibration m<strong>et</strong>hod, commonly used by aca<strong>de</strong>mics<br />

and practitioners. We show that in the context of non-param<strong>et</strong>ric calibration of expon<strong>en</strong>tial<br />

Lévy mo<strong>de</strong>ls, least squares calibration does not always allow to find a solution and ev<strong>en</strong> if it<br />

does, the solution is typically very s<strong>en</strong>sitive to small perturbations of input data. To solve the<br />

calibration problem in a stable manner, we first reformulate it as the problem of finding an<br />

expon<strong>en</strong>tial Lévy mo<strong>de</strong>l that has the smallest relative <strong>en</strong>tropy with respect to the prior among<br />

all solutions of the least squares calibration problem. This problem, called minimal <strong>en</strong>tropy<br />

least squares calibration, is still ill-posed, so we regularize it using the technique of minimal<br />

<strong>en</strong>tropy regularization, from the theory of ill-posed inverse problems.<br />

The third chapter discusses the numerical implem<strong>en</strong>tation of our calibration algorithm. To<br />

solve the calibration problem numerically, it is expressed in terms of the characteristic tripl<strong>et</strong>s<br />

of the prior and the solution and the Lévy measure of the prior is discr<strong>et</strong>ized on a uniform grid<br />

so that the calibration problem becomes finite-dim<strong>en</strong>sional.<br />

In the fourth chapter we first review the two available m<strong>et</strong>hods to mo<strong>de</strong>l <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong><br />

compon<strong>en</strong>ts of Lévy processes. Un<strong>de</strong>rstanding the drawbacks of these m<strong>et</strong>hods allows to<br />

formulate the <strong>de</strong>sirable properties of a multidim<strong>en</strong>sional mo<strong>de</strong>lling approach. These properties<br />

lead us to <strong>de</strong>fining the notion of Lévy copula, first in the case of Lévy processes with only<br />

positive jumps in every compon<strong>en</strong>t and th<strong>en</strong> in the g<strong>en</strong>eral case. After proving a repres<strong>en</strong>tation<br />

theorem, which shows that Lévy copulas compl<strong>et</strong>ely characterize <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures of Lévy<br />

processes, we compute Lévy copulas that correspond to various special types of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce.<br />

The fifth and the last chapter provi<strong>de</strong>s the tools necessary to apply Lévy copulas to finance.<br />

We first give m<strong>et</strong>hods to construct param<strong>et</strong>ric families of Lévy copulas and th<strong>en</strong> <strong>de</strong>velop an effici<strong>en</strong>t<br />

algorithm to simulate multidim<strong>en</strong>sional Lévy processes with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures giv<strong>en</strong><br />

by Lévy copulas. The last section of this chapter contains an example of a multidim<strong>en</strong>sional<br />

expon<strong>en</strong>tial Lévy mo<strong>de</strong>l for option pricing, constructed using Lévy copulas.


Part I<br />

Calibration of one-dim<strong>en</strong>sional<br />

exp-Lévy mo<strong>de</strong>ls<br />

31


Chapter 1<br />

Lévy processes and exp-Lévy mo<strong>de</strong>ls<br />

This introductory chapter serves ess<strong>en</strong>tially two purposes. First, in Section 1.1, we give an<br />

overview of the probabilistic properties of Lévy processes that will be used in the sequel. Second,<br />

we <strong>de</strong>fine expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls (Section 1.2), review various param<strong>et</strong>rizations of the Lévy<br />

measure, proposed by other authors (Section 1.3) and discuss a m<strong>et</strong>hod for option pricing in<br />

these mo<strong>de</strong>ls, based on Fourier transform (Section 1.4). This option pricing m<strong>et</strong>hod is later used<br />

for the numerical solution of the calibration problem. While the results of the first three sections<br />

can be found in the literature, Section 1.4 contains new material: we improve the m<strong>et</strong>hod due<br />

to Carr and Madan [23] in a number of ways and provi<strong>de</strong> estimates of the truncation and<br />

discr<strong>et</strong>ization error, not found in the original paper.<br />

1.1 Lévy processes<br />

Proofs of the the results of this section can be found, unless otherwise m<strong>en</strong>tioned in [87]. For<br />

additional <strong>de</strong>tails on Lévy processes the rea<strong>de</strong>r may consult [17] or [54]. The latter book treats<br />

a far more g<strong>en</strong>eral class of semimartingales but properties of Lévy processes are oft<strong>en</strong> discussed<br />

as examples or corollaries of the g<strong>en</strong>eral results.<br />

Definition 1.1 (Lévy process). A stochastic process {X t } t≥0 on (Ω, F, P ) such that X 0 = 0<br />

is called a Lévy process if it possesses the following properties:<br />

1. In<strong>de</strong>p<strong>en</strong><strong>de</strong>nt increm<strong>en</strong>ts: for every increasing sequ<strong>en</strong>ce of times t 0 . . . t n , the random variables<br />

X t0 , X t1 − X t0 , . . . , X tn − X tn−1<br />

are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt.<br />

33


34 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

2. Stationary increm<strong>en</strong>ts: for every h > 0, the law of X t+h − X t does not <strong>de</strong>p<strong>en</strong>d on t.<br />

The law of a Lévy process is compl<strong>et</strong>ely i<strong>de</strong>ntified by its characteristic tripl<strong>et</strong> (A, ν, γ), where<br />

A is a symm<strong>et</strong>ric nonnegative-<strong>de</strong>finite d × d matrix, γ ∈ R d and ν is a Lévy measure, that is, a<br />

positive measure on R d \ {0}, satisfying<br />

∫<br />

(|x| 2 ∧ 1)ν(dx) < ∞.<br />

R d \{0}<br />

In particular, the characteristic function of X t can be computed from this tripl<strong>et</strong> as follows.<br />

Theorem 1.1 (Lévy-Khintchine repres<strong>en</strong>tation). L<strong>et</strong> {X t } t≥0 be a Lévy process on R d<br />

with characteristic tripl<strong>et</strong> (A, ν, γ). Th<strong>en</strong><br />

E[e i〈z,Xt〉 ] = e tψ(z) , z ∈ R d (1.1)<br />

with ψ(z) = − 1 2 〈z, Az〉 + i〈γ, z〉 + ∫R d (e i〈z,x〉 − 1 − i〈z, x〉1 |x|≤1 )ν(dx).<br />

Remark 1.1. L<strong>et</strong> h : R d → R d be a measurable function, such that for every z, e i〈z,x〉 − 1 −<br />

i〈z, h(x)〉 is integrable with respect to ν. Th<strong>en</strong> the second equation in the Lévy-Khintchine<br />

formula (1.1) may be rewritt<strong>en</strong> as<br />

ψ(z) = − 1 ∫<br />

2 〈z, Az〉 + i〈γ h, z〉 + (e i〈z,x〉 − 1 − i〈z, h(x)〉)ν(dx),<br />

R d<br />

where γ h = γ +<br />

∫ ∞<br />

−∞<br />

(h(x) − x1 |x|≤1 )ν(dx).<br />

The tripl<strong>et</strong> (A, ν, γ h ) is called the characteristic tripl<strong>et</strong> of {X t } t≥0 with respect to the truncation<br />

function h. For instance, if ∫ R d (|x|∧1)ν(dx) < ∞, one may take h ≡ 0 and the Lévy-Khintchine<br />

repres<strong>en</strong>tation becomes<br />

ψ(z) = − 1 ∫<br />

2 〈z, Az〉 + i〈γ 0, z〉 + (e i〈z,x〉 − 1)ν(dx).<br />

R d<br />

The vector γ 0 is in this case called drift of the process X.<br />

The tail behavior and mom<strong>en</strong>ts of the distribution of a Lévy process at a giv<strong>en</strong> time are<br />

<strong>de</strong>termined by the Lévy measure, as shown by the following proposition (see Proposition 2.5<br />

and Theorems 25.3 and 25.17 in [87]).


1.1. LEVY PROCESSES 35<br />

Proposition 1.2 (Mom<strong>en</strong>ts and cumulants of a Lévy process).<br />

1. L<strong>et</strong> {X t } t≥0 be a Lévy process on R with characteristic tripl<strong>et</strong> (A, ν, γ) and l<strong>et</strong> n ≥ 1.<br />

E[|X t | n ] < ∞ for some t > 0 or equival<strong>en</strong>tly for every t if and only if ∫ |x|≥1 |x|n ν(dx) < ∞.<br />

In this case Φ t (z), the characteristic function of X t , is of class C n and the first n mom<strong>en</strong>ts<br />

of X t can be computed by differ<strong>en</strong>tiation:<br />

E[Xt k ] = 1 ∂ k<br />

i k ∂z k Φ t(z)| z=0 , k = 1, . . . , n.<br />

The cumulants of X t , <strong>de</strong>fined by<br />

c k (X t ) := 1 ∂ k<br />

i k ∂z k log Φ t(z)| z=0 ,<br />

have a particularly simple structure:<br />

∫<br />

c 1 (X t ) ≡ E[X t ] = t(γ + xν(dx)),<br />

c 2 (X t ) ≡ Var X t = t(A +<br />

c k (X t ) = t<br />

∫ ∞<br />

−∞<br />

|x|≥1<br />

∫ ∞<br />

−∞<br />

x 2 ν(dx)),<br />

x k ν(dx) for 3 ≤ k ≤ n.<br />

2. L<strong>et</strong> {X t } t≥0 be a Lévy process on R with characteristic tripl<strong>et</strong> (A, ν, γ) and l<strong>et</strong> u ∈ R.<br />

E[e uXt ] < ∞ for some t or, equival<strong>en</strong>tly, for all t > 0 if and only if ∫ |x|≥1 eux ν(dx) < ∞.<br />

In this case<br />

E[e uXt ] = e tψ(−iu) .<br />

where ψ is the characteristic expon<strong>en</strong>t of the Lévy process <strong>de</strong>fined by (1.1).<br />

Corollary 1.1. L<strong>et</strong> {X t } t≥0 be a Lévy process on R with characteristic tripl<strong>et</strong> (A, ν, γ).<br />

1. {X t } t≥0 is a martingale if and only if ∫ |x|≥1<br />

|x|ν(dx) < ∞ and<br />

∫<br />

γ + xν(dx) = 0.<br />

|x|≥1<br />

2. {exp(X t )} t≥0 is a martingale if and only if ∫ |x|≥1 ex ν(dx) < ∞ and<br />

∫<br />

A<br />

∞<br />

2 + γ + (e x − 1 − x1 |x|≤1 )ν(dx) = 0. (1.2)<br />

−∞


36 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

1.1.1 Stochastic expon<strong>en</strong>tial of Lévy processes<br />

The stochastic (Doléans-Da<strong>de</strong>) expon<strong>en</strong>tial of a semimartingale {X t } t≥0 (see [54, Theorem<br />

I.4.61]) is <strong>de</strong>fined as the unique càdlàg process {Z t } t≥0 such that<br />

dZ t = Z t− dX t , Z 0 = 1. (1.3)<br />

Z is giv<strong>en</strong> by:<br />

∏<br />

Z t = e Xt−X 0− 1 2 [X]c t (1 + ∆X s )e −∆Xs . (1.4)<br />

0≤s≤t<br />

The stochastic expon<strong>en</strong>tial of X is <strong>de</strong>noted by E(X). The following result, due to Goll and<br />

Kalls<strong>en</strong> [47] clarifies the relation b<strong>et</strong>we<strong>en</strong> the stochastic and the ordinary expon<strong>en</strong>tial wh<strong>en</strong> X<br />

is a Lévy process.<br />

Proposition 1.3 (Relation b<strong>et</strong>we<strong>en</strong> ordinary and stochastic expon<strong>en</strong>tials).<br />

1. L<strong>et</strong> {X t } t≥0 be a real-valued Lévy process with characteristic tripl<strong>et</strong> (A, ν, γ) and Z = E(X)<br />

its stochastic expon<strong>en</strong>tial. If Z > 0 a.s. th<strong>en</strong> there exists another Lévy process {L t } t≥0<br />

such that for all t, Z t = e Lt , where<br />

L t = log Z t = X t − At<br />

2 + ∑<br />

0≤s≤t<br />

Its characteristic tripl<strong>et</strong> (A L , ν L , γ L ) is giv<strong>en</strong> by:<br />

{<br />

log(1 + ∆Xs ) − ∆X s<br />

}<br />

. (1.5)<br />

A L = A,<br />

ν L (B) = ν({x : log(1 + x) ∈ B}) =<br />

∫ ∞<br />

−∞<br />

1 B (log(1 + x))ν(dx), B ∈ B(R \ {0}), (1.6)<br />

γ L = γ − A ∫ ∞<br />

2 + ν(dx) { log(1 + x)1 [−1,1] (log(1 + x)) − x1 [−1,1] (x) } .<br />

−∞<br />

2. L<strong>et</strong> {L t } t≥0 be a real-valued Lévy process with characteristic tripl<strong>et</strong> (A L , ν L , γ L ) and S t =<br />

exp L t its expon<strong>en</strong>tial. Th<strong>en</strong> there exists a Lévy process {X t } t≥0 such that S t is the<br />

stochastic expon<strong>en</strong>tial of X: S = E(X) where<br />

X t = L t + At<br />

2 + ∑<br />

0≤s≤t<br />

{<br />

e<br />

∆L s<br />

− 1 − ∆L s<br />

}<br />

. (1.7)


1.1. LEVY PROCESSES 37<br />

The characteristic tripl<strong>et</strong> (A, ν, γ) of X is giv<strong>en</strong> by:<br />

A = A L ,<br />

ν(B) = ν L ({x : e x − 1 ∈ B}) =<br />

γ = γ L + A L<br />

2 + ∫ ∞<br />

−∞<br />

∫ ∞<br />

−∞<br />

1 B (e x − 1)ν L (dx), B ∈ B(R \ {0}), (1.8)<br />

ν L (dx) { (e x − 1)1 [−1,1] (e x − 1) − x1 [−1,1] (x) } .<br />

Equation (1.3) shows that the stochastic expon<strong>en</strong>tial of a martingale is necessarily a local<br />

martingale.<br />

For Lévy processes, however, a stronger result is true, namely, the stochastic<br />

expon<strong>en</strong>tial of a martingale Lévy process is a (true) martingale.<br />

This fact seems to belong<br />

to the folklore of the Lévy process theory but we have be<strong>en</strong> unable to find its proof in the<br />

literature.<br />

Proposition 1.4 (Martingale preserving property). If {X t } t≥0 is a Lévy process and a<br />

martingale, th<strong>en</strong> its stochastic expon<strong>en</strong>tial Z = E(X) is also a martingale.<br />

Proof. L<strong>et</strong> {X t } t≥0 be a Lévy process on R with tripl<strong>et</strong> (A, ν, γ) such that γ + ∫ |x|≥1 xν(dx) = 0<br />

(this implies that X is a martingale). First, suppose that |∆X s | ≤ ε < 1 a.s. Th<strong>en</strong><br />

and Equation (1.4) implies that<br />

⎛<br />

log(1 + ∆X s ) − ∆X s ≥ −<br />

∆X2 s<br />

2(1 − ε) 2<br />

Z t ≥ exp ⎝X t − X 0 − 1 2 [X]c t −<br />

1 ∑<br />

2(1 − ε) 2<br />

0≤s≤t<br />

∆X 2 s<br />

⎞<br />

⎠ > 0<br />

a.s.<br />

Therefore by Proposition 1.3 there exists a Lévy process L such that e Lt = Z t for all t. Moreover,<br />

this process has boun<strong>de</strong>d jumps and therefore admits all expon<strong>en</strong>tial mom<strong>en</strong>ts.<br />

Proposition 1.3, we can write:<br />

Again, by<br />

γ L + A ∫ ∞<br />

L<br />

2 + (e z − 1 − z1 |z|≤1 )ν L (dz) = γ +<br />

−∞<br />

+<br />

∫ ∞<br />

−∞<br />

∫ 1<br />

−1<br />

(e z − 1 − z1 |z|≤1 )ν L (dz) =<br />

{zν L (dz) − zν(dz)}<br />

∫ ∞<br />

−∞<br />

{(e z − 1)ν L (dz) − zν(dz)} = 0<br />

because ∆X s = e ∆Ls − 1 for all s. Therefore, by Corollary 1.1, Z t = e Lt is a martingale.<br />

The second step is to prove the proposition wh<strong>en</strong> X is a comp<strong>en</strong>sated compound Poisson<br />

process. In this case, the stochastic expon<strong>en</strong>tial has a very simple form:<br />

∏<br />

Z t = e bt (1 + ∆X s ),<br />

0≤s≤t


38 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

where b = − ∫ ∞<br />

−∞<br />

xν(dx). D<strong>en</strong>oting the jump int<strong>en</strong>sity of X by λ, and the g<strong>en</strong>eric jump of X<br />

by ∆X, we obtain, conditioning on the number of jumps of X in the interval [0, t]:<br />

E[Z t ] = e −λt+bt<br />

∞ ∑<br />

n=0<br />

(λt) n<br />

(1 + E[∆X]) n = 1.<br />

n!<br />

Because the increm<strong>en</strong>ts of X are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and stationary, this proves that Z is a martingale.<br />

Now l<strong>et</strong> X be an arbitrary martingale Lévy process.<br />

It can be <strong>de</strong>composed into a sum<br />

of a comp<strong>en</strong>sated compound Poisson process X ′ and an in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt martingale Lévy process<br />

with jumps smaller than ε, <strong>de</strong>noted by X ′′ . Since these two processes never jump tog<strong>et</strong>her,<br />

E(X ′ + X ′′ ) = E(X ′ )E(X ′′ ) (cf. Equation II.8.19 in [54]). Moreover, each of the factors is a<br />

martingale and they are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt, h<strong>en</strong>ce E(X ′ + X ′′ ) is a martingale.<br />

1.1.2 Change of measure and absolute continuity of Lévy processes<br />

The following proposition (see [54], Theorem IV.4.39) provi<strong>de</strong>s a criterion of absolute continuity<br />

of one Lévy process with respect to another in terms of their characteristic tripl<strong>et</strong>s.<br />

Proposition 1.5 (Absolute continuity of Lévy processes). L<strong>et</strong> {X t } t≥0 be a real-valued<br />

Lévy process on (Ω, F, Q) and on (Ω, F, P ) with respective characteristic tripl<strong>et</strong>s (A Q , ν Q , γ Q )<br />

and (A P , ν P , γ P ). Th<strong>en</strong> Q| Ft ≪ P | Ft for all t ∈ [0, T ∞ ] or, equival<strong>en</strong>tly, Q| Ft ≪ P | Ft for some<br />

t ∈ (0, T ∞ ] if and only if the following conditions are satisfied:<br />

1. A Q = A P := A.<br />

2. ν Q ≪ ν P .<br />

3. ∫ ∞<br />

−∞ (√ φ(x) − 1) 2 ν P (dx) < ∞, where φ := dν Q<br />

dν P<br />

.<br />

4. ∫ |x|≤1 |x(1 − φ(x))|ν P (dx) < ∞.<br />

5. If A = 0 th<strong>en</strong> γ Q − γ P = ∫ |x|≤1 x(φ(x) − 1)ν P (dx).<br />

Remark 1.2. In fact, condition 4 above is a consequ<strong>en</strong>ce of condition 3. In<strong>de</strong>ed, on the s<strong>et</strong><br />

{x : φ(x) > 4}, we have ( √ φ(x) − 1) 2 > 1 and (√ φ(x)−1) 2<br />

∫<br />

{φ(x)>4}<br />

and<br />

∫<br />

∫<br />

ν P (dx) ≤<br />

{φ(x)>4}<br />

{φ(x)>4}<br />

∫<br />

φ(x)ν P (dx) < 4<br />

φ(x)<br />

( √ φ(x) − 1) 2 ν P (dx) < ∞<br />

{φ(x)>4}<br />

> 1 4<br />

. Therefore, condition 3 implies<br />

( √ φ(x) − 1) 2 ν P (dx) < ∞.


1.1. LEVY PROCESSES 39<br />

On the other hand,<br />

∫<br />

∫<br />

|x(1 − φ(x))|ν P (dx) =<br />

|x|≤1<br />

|x|≤1;φ(x)>4<br />

The first term in the right-hand si<strong>de</strong> satisfies:<br />

∫<br />

∫<br />

|x(1 − φ(x))|ν P (dx) ≤<br />

|x|≤1;φ(x)>4<br />

and for the second term one can write:<br />

∫<br />

|x|≤1;φ(x)≤4<br />

|x(1 − φ(x))|ν P (dx)<br />

∫<br />

≤<br />

|x|≤1;φ(x)≤4<br />

∫<br />

|x(1 − φ(x))|ν P (dx) +<br />

|x(1 − φ(x))|ν P (dx).<br />

|x|≤1;φ(x)≤4<br />

|x|≤1;φ(x)>4<br />

∫<br />

ν P (dx) +<br />

φ(x)ν P (dx) < ∞,<br />

|x|≤1;φ(x)>4<br />

∫<br />

|x| 2 ν P (dx) +<br />

( √ φ(x) − 1) 2 ( √ φ(x) + 1) 2 ν P (dx) < ∞.<br />

|x|≤1;φ(x)≤4<br />

A real-valued Lévy process with characteristic tripl<strong>et</strong> (A, ν, γ) is said to be of jump-diffusion<br />

type if A > 0 and ν(R \ {0}) < ∞.<br />

Corollary 1.2. L<strong>et</strong> {X t } t≥0 be a Lévy process of jump-diffusion type un<strong>de</strong>r P . Th<strong>en</strong> Q| Ft ≪<br />

P | Ft for all t ∈ [0, T ∞ ] or, equival<strong>en</strong>tly, Q| Ft ≪ P | Ft for some t ∈ (0, T ∞ ] if and only if X is of<br />

jump-diffusion type un<strong>de</strong>r Q and conditions 1 and 2 of Proposition 1.5 are satisfied.<br />

Proof. Observe that for all u, v and r > 0,<br />

(1 − r)u 2 + (1 − 1/r)v 2 ≤ (u + v) 2 ≤ (1 + r)u 2 + (1 + 1/r)v 2 .<br />

Taking u = √ φ(x), v = −1 and r = 1/2, we obtain:<br />

Suppose that Q| Ft ≪ P | Ft<br />

1<br />

2 φ(x) − 1 ≤ (√ φ(x) − 1) 2 ≤ 3 φ(x) + 3. (1.9)<br />

2<br />

for all t. Th<strong>en</strong> (1.9) implies that<br />

ν Q (R \ {0}) ≤ 2ν P (R \ {0}) + 2<br />

∫ ∞<br />

−∞<br />

and therefore {X t } t≥0 is of jump-diffusion type un<strong>de</strong>r Q.<br />

( √ φ(x) − 1) 2 ν P (dx) < ∞<br />

Suppose that {X t } t≥0 is of jump-diffusion type un<strong>de</strong>r Q and conditions 1 and 2 are satisfied.<br />

On account of Remark 1.2, we only need to verify condition 3 of the proposition. Equation<br />

(1.9) implies that<br />

∫ ∞<br />

−∞<br />

which compl<strong>et</strong>es the proof.<br />

( √ φ(x) − 1) 2 ν P (dx) ≤ 3ν P (R \ {0}) + 3 2 ν Q(R \ {0}) < ∞,


40 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

1.1.3 Tightness and converg<strong>en</strong>ce of sequ<strong>en</strong>ces of Lévy processes<br />

L<strong>et</strong> E be a Polish space with Borel σ-field F. The space of all probability measures on (E, F) is<br />

<strong>de</strong>noted by P(E). The weak topology on P(E) is the coarsest topology for which the mappings<br />

µ ∈ P(E) ↦→ µ(f) are continuous for all boun<strong>de</strong>d continuous functions f on E. We recall that<br />

a subs<strong>et</strong> B ⊂ P(E) is called tight if for every ε > 0 there exists a compact subs<strong>et</strong> K ⊂ E such<br />

that µ(E \K) ≤ ε for all µ ∈ B. Prohorov’s theorem states that a subs<strong>et</strong> B ⊂ P(E) is relatively<br />

compact for the weak topology if and only if it is tight.<br />

The space of all càdlàg functions on [0, T ∞ ], equipped with Skorokhod topology (see Chapter<br />

VI in [54]) is a Polish space and all of the above remains valid. The following result is a corollary<br />

(an adaptation for the case Lévy processes) of Theorem VI.4.18 in [54].<br />

Proposition 1.6. L<strong>et</strong> {X n } n≥1 be a sequ<strong>en</strong>ce of real-valued Lévy processes with characteristic<br />

tripl<strong>et</strong>s (A n , ν n , γ n ). For {X n } to be tight it suffices that<br />

1. The sequ<strong>en</strong>ce of Lévy measures {ν n } satisfies<br />

2. The sequ<strong>en</strong>ce {γ n } is boun<strong>de</strong>d.<br />

lim sup ν n ({x : |x| > a}) = 0.<br />

a→∞<br />

3. The sequ<strong>en</strong>ce {C n } is boun<strong>de</strong>d, where C n = A n + ∫ ∞<br />

−∞ (x2 ∧ 1)ν n (dx).<br />

n<br />

The next result characterizes the weak converg<strong>en</strong>ce of sequ<strong>en</strong>ces of Lévy processes (cf.<br />

Corollary VII.3.6 in [54]).<br />

Proposition 1.7. L<strong>et</strong> {X n } n≥1 and X be real-valued Lévy processes with characteristic tripl<strong>et</strong>s<br />

(A n , ν n , γ h n) n≥1 and (A, ν, γ h ) with respect to a truncation function h (cf. Remark 1.1), that<br />

is continuous, boun<strong>de</strong>d and satisfies h(x) ≡ x on a neighborhood of 0. There is equival<strong>en</strong>ce<br />

b<strong>et</strong>we<strong>en</strong><br />

1. X n d → X.<br />

2. X n 1<br />

d<br />

→ X 1 .<br />

3. The following conditions are satisfied:<br />

(a) γ h n → γ h .


1.2. EXPONENTIAL LEVY MODELS 41<br />

(b) C n → C, where C ∗ = A ∗ + ∫ ∞<br />

−∞ h2 (x)ν ∗ (dx).<br />

(c) ∫ ∞<br />

−∞ f(x)ν n(dx) → ∫ ∞<br />

−∞<br />

f(x)ν(dx) for every continuous boun<strong>de</strong>d function f such that<br />

f(x) ≡ 0 on a neighborhood of zero, or, equival<strong>en</strong>tly, for every continuous boun<strong>de</strong>d<br />

function f satisfying f(x) = o(|x| 2 ) wh<strong>en</strong> x → 0.<br />

1.2 Expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls: <strong>de</strong>finition and main properties<br />

Expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls are obtained by replacing the Brownian motion with drift in the<br />

classical Black-Scholes-Samuelson mo<strong>de</strong>l of ass<strong>et</strong> price, by a Lévy process:<br />

S t = S 0 e rt+Xt , (1.10)<br />

where X is a Lévy process on (Ω, F, P ), and the interest rate term rt is introduced to simplify<br />

the notation below. Wh<strong>en</strong> P is the probability that <strong>de</strong>scribes the evolution of stock prices<br />

in the real world, also called the historical probability, the mo<strong>de</strong>l (1.10) is called a historical<br />

expon<strong>en</strong>tial Lévy mo<strong>de</strong>l.<br />

By the first fundam<strong>en</strong>tal theorem of ass<strong>et</strong> pricing (see [34]), a financial mark<strong>et</strong> does not<br />

allow for arbitrage opportunity (more precisely, satisfies the No Free Lunch with Vanishing Risk<br />

condition) if there exists a probability measure Q, equival<strong>en</strong>t to P , such that the discounted<br />

prices e −rt V t of all ass<strong>et</strong>s are Q-local martingales. Q is called a risk-neutral probability. The<br />

abs<strong>en</strong>ce of arbitrage in the mo<strong>de</strong>l (1.10) is therefore equival<strong>en</strong>t to the exist<strong>en</strong>ce of a probability<br />

Q ∼ P , such that e X is a Q-local martingale. The following result shows that if X is a Lévy<br />

process un<strong>de</strong>r P , one can almost always find a probability Q ∼ P , un<strong>de</strong>r which X is still a Lévy<br />

process and e X is a martingale.<br />

Proposition 1.8 (Abs<strong>en</strong>ce of arbitrage in exp-Lévy mo<strong>de</strong>ls). L<strong>et</strong> {X t } t≥0 be a Lévy<br />

process on (Ω, F, P ) with characteristic tripl<strong>et</strong> (A, ν, γ). If the trajectories of X are neither<br />

almost surely increasing nor almost surely <strong>de</strong>creasing, th<strong>en</strong> there exists a probability measure Q<br />

equival<strong>en</strong>t to P such that un<strong>de</strong>r Q, {X t } t≥0 is a Lévy process and {e Xt } t≥0 is a martingale.<br />

Q may be chos<strong>en</strong> in such way that (X, Q) will have the characteristic tripl<strong>et</strong> (A Q , ν Q , γ Q ),


42 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

where<br />

for some θ ∈ R.<br />

A Q = A, ν Q = e −x2 +θx ν, (1.11)<br />

∫<br />

γ Q = γ + Aθ + x(e −x2 +θx − 1)ν(dx) (1.12)<br />

|x|≤1<br />

In other words, this proposition shows that expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls are arbitrage-free in<br />

the following (not mutually exclusive) cases:<br />

• X has infinite variation ∫ 1<br />

−1<br />

|x|ν(dx) = ∞ and/or A > 0.<br />

• X has both positive and negative jumps.<br />

• X has positive jumps and negative drift or negative jumps and positive drift.<br />

In the sequel, we will <strong>de</strong>note the s<strong>et</strong> of Lévy processes satisfying at least one of these conditions<br />

by L NA and the corresponding expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls will be called arbitrage-free expon<strong>en</strong>tial<br />

Lévy mo<strong>de</strong>ls.<br />

The first part of this proposition was proved in [55], see also [25, Theorem 4.6]. Since the<br />

second part will also be used below, we give a proof here.<br />

Proof of Proposition 1.8. For any θ ∈ R we <strong>de</strong>note the probability measure un<strong>de</strong>r which X has<br />

the characteristic tripl<strong>et</strong> (1.11)–(1.12) by Q θ . From Proposition 1.5 it follows easily that for<br />

every θ ∈ R, both Q θ ≪ P and P ≪ Q θ . Therefore, we only need to prove that for some θ, the<br />

tripl<strong>et</strong> (A, ν Qθ , γ Qθ ) satisfies the martingale condition (1.2), or equival<strong>en</strong>tly, that the equation<br />

has a solution, where<br />

f(θ) := A 2 + Aθ + ∫<br />

|x|≤1<br />

f(θ) + γ = 0 (1.13)<br />

x(e −x2 +θx − 1)ν(dx) +<br />

∫ ∞<br />

−∞<br />

(e x − 1 − x1 |x|≤1 )e −x2 +θx ν(dx)<br />

The dominated converg<strong>en</strong>ce theorem implies that f is continuous and differ<strong>en</strong>tiable and that<br />

f ′ (θ) = A + ∫ ∞<br />

−∞ x(ex − 1)e −x2 +θx ν(dx) ≥ 0, which means that f(θ) is an increasing function.<br />

Moreover, if A > 0 or if ν((0, ∞)) > 0 and ν((−∞, 0)) > 0 th<strong>en</strong> f ′ is everywhere boun<strong>de</strong>d from<br />

below by a positive number. Therefore in these cases lim θ→+∞ f(θ) = +∞, lim θ→−∞ f(θ) =<br />

−∞ and Equation (1.13) admits a solution.


1.3. EXP-LEVY MODELS IN FINANCE 43<br />

It remains to treat the case wh<strong>en</strong> A = 0 and ν is supported by one of the half-axes. Suppose<br />

that ν((−∞, 0)) = 0 and ν((0, ∞)) > 0. In this case lim θ→+∞ f(θ) = +∞ and<br />

lim f(θ) = − xν(dx).<br />

θ→−∞<br />

∫0


44 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

where {N t } t≥0 is the Poisson process counting the jumps of X and Y i are jump sizes (i.i.d.<br />

variables).<br />

In the Merton mo<strong>de</strong>l [71], which is the first mo<strong>de</strong>l of this type, suggested in the literature,<br />

jumps in the log-price X are assumed to have a Gaussian distribution: Y i ∼ N(µ, δ 2 ). In the<br />

risk-neutral version the characteristic expon<strong>en</strong>t of the log stock price takes the following form:<br />

(<br />

ψ(u) = − Au2<br />

2 + u 2 A /2+iµu λ{e−δ2 − 1} − iu<br />

2 + /2+µ λ(eδ2 − 1))<br />

. (1.16)<br />

In the Kou mo<strong>de</strong>l [62], jump sizes are distributed according to an asymm<strong>et</strong>ric Laplace law<br />

with a <strong>de</strong>nsity of the form<br />

ν 0 (dx) = [pλ + e −λ +x 1 x>0 + (1 − p)λ − e −λ −|x| 1 x 0, λ − > 0 governing the <strong>de</strong>cay of the tails for the distribution of positive and negative<br />

jump sizes and p ∈ [0, 1] repres<strong>en</strong>ting the probability of an upward jump. The probability<br />

distribution of r<strong>et</strong>urns in this mo<strong>de</strong>l has semi-heavy (expon<strong>en</strong>tial) tails.<br />

In the above two mo<strong>de</strong>ls, the dynamical structure of the process is easy to un<strong>de</strong>rstand and<br />

<strong>de</strong>scribe, since the distribution of jump sizes is known. They are easy to simulate and effici<strong>en</strong>t<br />

Monte Carlo m<strong>et</strong>hods for pricing path-<strong>de</strong>p<strong>en</strong><strong>de</strong>nt options have be<strong>en</strong> <strong>de</strong>veloped. Mo<strong>de</strong>ls of this<br />

type also perform quite well for the purposes of implied volatility smile interpolation. However,<br />

they do not lead to closed-form <strong>de</strong>nsities: statistical estimation and computation of mom<strong>en</strong>ts<br />

or quantiles may be quite difficult.<br />

The second category consists of mo<strong>de</strong>ls with infinite number of jumps in every interval,<br />

which we will call infinite activity or infinite int<strong>en</strong>sity mo<strong>de</strong>ls. In these mo<strong>de</strong>ls, one does not<br />

need to introduce a Brownian compon<strong>en</strong>t since the dynamics of jumps is already rich <strong>en</strong>ough<br />

to g<strong>en</strong>erate nontrivial small time behavior [21] and it has be<strong>en</strong> argued [21, 43, 66] that such<br />

mo<strong>de</strong>ls give a more realistic <strong>de</strong>scription of the price process at various time scales. In addition,<br />

many mo<strong>de</strong>ls from this class can be constructed via Brownian subordination, which gives them<br />

additional tractability compared to jump-diffusion mo<strong>de</strong>ls.<br />

Table 1.1 compares the advantages and drawbacks of these two categories. It should be kept<br />

in mind that since the price process is observed on a discr<strong>et</strong>e grid, it is difficult if not impossible<br />

to see empirically to which category the price process belongs. The choice is more a question<br />

of mo<strong>de</strong>lling conv<strong>en</strong>i<strong>en</strong>ce than an empirical one.


1.3. EXP-LEVY MODELS IN FINANCE 45<br />

Table 1.1: Compound Poisson or infinite int<strong>en</strong>sity: a comparison of two mo<strong>de</strong>lling approaches<br />

Jump-diffusion mo<strong>de</strong>ls<br />

Must contain a Brownian compon<strong>en</strong>t.<br />

Jumps are rare ev<strong>en</strong>ts.<br />

Distribution of jump sizes is known.<br />

Perform well for implied volatility smile interpolation.<br />

D<strong>en</strong>sities not known in closed form.<br />

Easy to simulate.<br />

Infinite int<strong>en</strong>sity mo<strong>de</strong>ls<br />

Brownian compon<strong>en</strong>t is not nee<strong>de</strong>d.<br />

The process moves only by jumps and <strong>de</strong>terministic<br />

linear drift.<br />

“Distribution of jump sizes” does not exist:<br />

jumps arrive infinitely oft<strong>en</strong>.<br />

Give a realistic <strong>de</strong>scription of the historical<br />

price process.<br />

Closed form <strong>de</strong>nsities available in some<br />

cases.<br />

In some cases can be repres<strong>en</strong>ted via Brownian<br />

subordination, which gives additional<br />

tractability.


46 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

There are three standard ways to <strong>de</strong>fine a param<strong>et</strong>ric Lévy process with infinite jump int<strong>en</strong>sity,<br />

summarized in Table 1.2.<br />

The first approach is to obtain a Lévy process by subordinating a Brownian motion with an<br />

in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt increasing Lévy process (such a process is called a subordinator). Here the characteristic<br />

function of the resulting process can be obtained immediately, but an explicit formula<br />

for the Lévy measure is not always available. Due to the conditionally Gaussian structure of<br />

the process, simulation and some computations can be consi<strong>de</strong>rably simplified (for instance, call<br />

option price can be expressed as an integral involving Black-Scholes prices). The interpr<strong>et</strong>ation<br />

of the subordinator as “business time” [44] makes mo<strong>de</strong>ls of this type easier to un<strong>de</strong>rstand and<br />

interpr<strong>et</strong>. Multidim<strong>en</strong>sional ext<strong>en</strong>sions are also possible: one can take a multivariate Brownian<br />

motion and change the time scale of all compon<strong>en</strong>ts with the same subordinator. Two examples<br />

of mo<strong>de</strong>ls from this class are the variance gamma process and the normal inverse Gaussian<br />

process. The variance gamma process [23, 67] is obtained by time-changing a Brownian motion<br />

with a gamma subordinator and has the characteristic expon<strong>en</strong>t of the form:<br />

ψ(u) = − 1 κ log(1 + u2 σ 2 κ<br />

2<br />

− iθκu). (1.18)<br />

The <strong>de</strong>nsity of the Lévy measure of the variance gamma process is giv<strong>en</strong> by<br />

where c = 1/κ, λ + =<br />

ν(x) = c<br />

|x| e−λ −|x| 1 x0 , (1.19)<br />

√<br />

θ 2 +2σ 2 /κ<br />

σ 2<br />

− θ<br />

σ 2 and λ − =<br />

√<br />

θ 2 +2σ 2 /κ<br />

σ 2<br />

+ θ<br />

σ 2 .<br />

The normal inverse Gaussian process [5, 7, 86] is the result of time-changing a Brownian<br />

motion with the inverse Gaussian subordinator and has the characteristic expon<strong>en</strong>t:<br />

ψ(u) = 1 κ − 1 κ√<br />

1 + u 2 σ 2 κ − 2iθuκ.<br />

The second approach is to specify the Lévy measure directly. The main example of this<br />

category is giv<strong>en</strong> by the tempered stable process, introduced by Kopon<strong>en</strong> [61] and used for<br />

financial mo<strong>de</strong>lling in [18, 26, 27], as well as in [22] (un<strong>de</strong>r the name of CGMY mo<strong>de</strong>l) and in<br />

[19] (un<strong>de</strong>r the name of KoBoL process). The tempered stable process has a Lévy measure with<br />

<strong>de</strong>nsity of the form:<br />

ν(x) =<br />

c −<br />

|x| 1+α − e−λ −|x| 1 x0 (1.20)


1.4. PRICING EUROPEAN OPTIONS 47<br />

with α + < 2 and α − < 2. This rich param<strong>et</strong>ric form of the Lévy measure is probably suffici<strong>en</strong>t<br />

for most applications.<br />

The third approach is to specify the <strong>de</strong>nsity of increm<strong>en</strong>ts of the process at a giv<strong>en</strong> time<br />

scale, say ∆, by taking an arbitrary infinitely divisible distribution. G<strong>en</strong>eralized hyperbolic<br />

processes (see [36–38]) can be constructed in this way. In this approach it is easy to simulate<br />

the increm<strong>en</strong>ts of the process at the same time scale and to estimate param<strong>et</strong>ers of the distribution<br />

if data are sampled with the same period ∆, but, unless this distribution belongs to<br />

some param<strong>et</strong>ric class closed un<strong>de</strong>r convolution, we do not know the law of the increm<strong>en</strong>ts at<br />

other time scales. Also, giv<strong>en</strong> an infinitely divisible distribution, one may not know its Lévy-<br />

Khintchine repres<strong>en</strong>tation, so it may not be easy to see wh<strong>et</strong>her the corresponding Lévy process<br />

has a Gaussian compon<strong>en</strong>t, finite or infinite jump int<strong>en</strong>sity, <strong>et</strong>c.<br />

1.4 Pricing European options in exp-Lévy mo<strong>de</strong>ls via Fourier<br />

transform<br />

In this section we pres<strong>en</strong>t a m<strong>et</strong>hod, adapted from [23], for pricing European call options in exp-<br />

Lévy mo<strong>de</strong>ls using Fourier transform and, in particular, the Fast Fourier transform algorithm<br />

[29]. We suggest several improvem<strong>en</strong>ts to the original procedure and give a rigorous analysis of<br />

truncation and discr<strong>et</strong>ization errors. These results are of in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt interest and will also be<br />

used for the numerical solution of the calibration problem in Chapter 3. Similar analysis has<br />

rec<strong>en</strong>tly appeared in [63].<br />

L<strong>et</strong> {X t } t≥0 be a Lévy process satisfying the martingale condition (1.2). To compute the<br />

price of a call option<br />

C(k) = e −rT E[(e rT +X T<br />

− e k ) + ], (1.21)<br />

we would like to express its Fourier transform in log strike in terms of the characteristic function<br />

Φ T (v) of X T and th<strong>en</strong> find the prices for a range of strikes by Fourier inversion. However we<br />

cannot do this directly because C(k) is not integrable (it t<strong>en</strong>ds to 1 as k goes to −∞). The<br />

key i<strong>de</strong>a is to instead compute the Fourier transform of the (modified) time value of the option,<br />

that is, the function<br />

z T (k) = e −rT E[(e rT +X T<br />

− e k ) + ] − (1 − e k−rT ) + . (1.22)


48 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

Table 1.2: Three approaches to building param<strong>et</strong>ric exp-Lévy mo<strong>de</strong>ls<br />

Brownian subordination Specifying the Lévy<br />

measure<br />

Specifying<br />

<strong>de</strong>nsity for t = ∆<br />

probability<br />

Interpr<strong>et</strong>ation as “Brownian<br />

motion in business time”.<br />

Clear vision of the pathwise<br />

properties.<br />

Structure of jumps is not<br />

known.<br />

Simulation is easy if we know<br />

how to simulate the subordinator.<br />

Simulation is quite involved.<br />

Simulation is easy on a grid<br />

of size ∆.<br />

Estimation via maximum<br />

likelihood may be difficult.<br />

Estimation can be done by<br />

approximating the transition<br />

<strong>de</strong>nsity.<br />

Estimation is easy for data<br />

with sampling interval ∆.<br />

Multivariate g<strong>en</strong>eralizations<br />

possible using multidim<strong>en</strong>sional<br />

Brownian motion.<br />

Rich vari<strong>et</strong>y of mo<strong>de</strong>ls.<br />

The infinite divisibility of a<br />

giv<strong>en</strong> mo<strong>de</strong>l may be difficult<br />

to prove.


1.4. PRICING EUROPEAN OPTIONS 49<br />

Proposition 1.9 (Carr and Madan [23]). L<strong>et</strong> {X t } t≥0 be a real-valued Lévy process satisfying<br />

the martingale condition (1.2), such that<br />

E[e (1+α)Xt ] < ∞ (1.23)<br />

for some α > 0. Th<strong>en</strong> the Fourier transform in log-strike k of the time value of a call option is<br />

giv<strong>en</strong> by:<br />

ζ T (v) :=<br />

∫ +∞<br />

−∞<br />

e ivk z T (k)dk = e ivrT Φ T (v − i) − 1<br />

iv(1 + iv)<br />

(1.24)<br />

Remark 1.3. Since typically Φ T (z) → 0 as Rz → ∞, ζ T (v) will behave like |v| −2 at infinity which<br />

means that the truncation error in the numerical evaluation of the inverse Fourier transform<br />

will be large. The reason of such a slow converg<strong>en</strong>ce is that the time value (1.22) is not smooth;<br />

therefore its Fourier transform does not <strong>de</strong>cay suffici<strong>en</strong>tly fast at infinity. For most mo<strong>de</strong>ls<br />

the converg<strong>en</strong>ce can be improved by replacing the time value with a smooth function of strike.<br />

Instead of subtracting the (non-differ<strong>en</strong>tiable) intrinsic value of the option from its price, we<br />

suggest to subtract the Black-Scholes call price with a non-zero volatility (which is a smooth<br />

function). The resulting function will be both integrable and smooth. Suppose that hypothesis<br />

(1.23) is satisfied and <strong>de</strong>note<br />

˜z T (k) = e −rT E[(e rT +X T<br />

− e k ) + ] − C Σ BS(k),<br />

where CBS Σ (k) is the Black-Scholes price of a call option with volatility Σ and log-strike k for<br />

the same un<strong>de</strong>rlying value and the same interest rate. Proposition 1.9 th<strong>en</strong> implies that the<br />

Fourier transform of ˜z T (k), <strong>de</strong>noted by ˜ζ T (v), satisfies<br />

˜ζ T (v) = e ivrT Φ T (v − i) − Φ Σ T<br />

(v − i)<br />

, (1.25)<br />

iv(1 + iv)<br />

where Φ Σ T (v) = exp(− Σ2 T<br />

2 (v2 + iv)). Since for most exp-Lévy mo<strong>de</strong>ls found in the literature<br />

(except variance gamma) the characteristic function <strong>de</strong>cays faster than every power of its argum<strong>en</strong>t<br />

at infinity, this means that the expression (1.25) will also <strong>de</strong>cay faster than every power<br />

of v as Rv → ∞, and the truncation error in the numerical evaluation of the inverse Fourier<br />

transform will be very small for every Σ > 0.<br />

Figure 1.1 shows the behavior of |˜ζ T | for differ<strong>en</strong>t values of Σ compared to the behavior of<br />

|ζ T | in Merton’s jump-diffusion mo<strong>de</strong>l with volatility 0.2, jump int<strong>en</strong>sity equal to 5 and jump


50 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

0.035<br />

0.03<br />

No smoothing<br />

Σ=0.2<br />

Σ=0.3575<br />

10 −2<br />

10 0 No smoothing<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

10 −4<br />

10 −6<br />

10 −8<br />

10 −10<br />

10 −12<br />

Σ=0.2<br />

Σ=0.3575<br />

0.005<br />

10 −14<br />

0<br />

−50 0 50<br />

10 −16<br />

−50 0 50<br />

Figure 1.1: Converg<strong>en</strong>ce of Fourier transform of option’s time value to zero in Merton mo<strong>de</strong>l.<br />

Left graph: linear scale; right graph: logarithmic scale.<br />

param<strong>et</strong>ers µ = −0.1 and δ = 0.1 for the time horizon T = 0.5. The converg<strong>en</strong>ce of ˜ζ T to zero<br />

is faster than expon<strong>en</strong>tial for all values of Σ and it is particularly good for Σ = 0.3575, the<br />

value of Σ for which ˜ζ(0) = 0.<br />

Proof of Proposition 1.9. Since the discounted price process is a martingale,<br />

∫ ∞<br />

z T (k) = e −rT µ T (dx)(e rT +x − e k )(1 k≤x+rT − 1 k≤rT ),<br />

−∞<br />

where µ T is the probability distribution of X T . Condition (1.23) <strong>en</strong>ables us to compute ζ T (v)<br />

by interchanging integrals:<br />

ζ T (v) = e −rT ∫ ∞<br />

−∞<br />

−∞<br />

∫ ∞<br />

dk<br />

µ T (dx)e ivk (e rT +x − e k )(1 k≤x+rT − 1 k≤rT )<br />

−∞<br />

∫ rT<br />

∫ ∞<br />

= e −rT µ T (dx) e ivk (e k − e rT +x )dk<br />

−∞<br />

x+rT<br />

∫ {<br />

}<br />

∞<br />

e ivrT (1 − e x )<br />

= µ T (dx)<br />

−<br />

ex+ivrT<br />

iv + 1 iv(iv + 1) + e(iv+1)x+ivrT<br />

iv(iv + 1)<br />

The first term in braces disappears due to the martingale condition and the other two, after<br />

computing the integrals, yield (1.24).


1.4. PRICING EUROPEAN OPTIONS 51<br />

Numerical Fourier inversion.<br />

the inverse Fourier transform of ˜ζ T :<br />

Option prices can be computed by evaluating numerically<br />

˜z T (k) = 1<br />

2π<br />

∫ +∞<br />

−∞<br />

e −ivk ˜ζT (v)dv (1.26)<br />

This integral can be effici<strong>en</strong>tly computed for a range of strikes using the Fast Fourier transform,<br />

an algorithm due to Cooley and Tukey [29] which allows to compute the discr<strong>et</strong>e Fourier<br />

transform DFT[f] n=0 N−1 , <strong>de</strong>fined by,<br />

DFT[f] n :=<br />

using only O(N log N) operations.<br />

N−1<br />

∑<br />

k=0<br />

f k e −2πink/N , n = 0 . . . N − 1, (1.27)<br />

To approximate option prices, we truncate and discr<strong>et</strong>ize the integral (1.26) as follows:<br />

∫<br />

1 ∞<br />

e −ivk ˜ζT (v)dv = 1 ∫ L/2<br />

e −ivk ˜ζT (v)dv + ε T<br />

2π −∞<br />

2π −L/2<br />

=<br />

N−1<br />

L ∑<br />

w m ˜ζT (v m )e −ikvm + ε T + ε D , (1.28)<br />

2π(N − 1)<br />

where ε T is the truncation error, ε D is the discr<strong>et</strong>ization error, v m = −L/2+m∆, ∆ = L/(N −1)<br />

is the discr<strong>et</strong>ization step and w m are weights, corresponding to the chos<strong>en</strong> integration rule (for<br />

instance, for the trapezoidal rule w 0 = w N−1 = 1/2 and all other weights are equal to 1).<br />

Now, choosing k n = k 0 + 2πn<br />

N∆<br />

transform:<br />

N−1<br />

L<br />

∑<br />

2π(N − 1) eiknL/2 w m ˜ζT (k m )e −ik0m∆ e −2πinm/N<br />

m=0<br />

m=0<br />

we see that the sum in the last term becomes a discr<strong>et</strong>e Fourier<br />

=<br />

L<br />

2π(N − 1) eiknL/2 DFT n [w m ˜ζT (k m )e −ik 0m∆ ]<br />

Therefore, the FFT algorithm allows to compute ˜z T and option prices for the log strikes k n =<br />

k 0 + 2πn<br />

N∆<br />

. The log strikes are thus equidistant with the step d satisfying<br />

d∆ = 2π<br />

N .<br />

Error control.<br />

We start with the truncation error.<br />

Lemma 1.10. L<strong>et</strong> {X t } t≥0 be a Lévy process with characteristic tripl<strong>et</strong> (A, ν, γ) and characteristic<br />

function Φ T . Th<strong>en</strong>


52 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

1. ∀ v ∈ R, |Φ T (v − i)| ≤ exp(− T Av2<br />

2<br />

).<br />

2. Suppose that ν has a <strong>de</strong>nsity of the form ν(x) = e−x<br />

|x|<br />

f(x), where f is increasing on (−∞, 0)<br />

and <strong>de</strong>creasing on (0, ∞). Th<strong>en</strong> |Φ T (v − i)| is increasing on v ∈ (−∞, 0) and <strong>de</strong>creasing<br />

on v ∈ (0, ∞).<br />

Proof. The martingale condition implies that ∀ v ∈ R,<br />

|Φ T (v − i)| = exp<br />

{− T ∫ ∞<br />

}<br />

Av2 − e x (1 − cos(vx))ν(dx) . (1.29)<br />

2<br />

This immediately <strong>en</strong>tails the first part of the lemma. For the second part, l<strong>et</strong> 0 > v 1 > v 2 (the<br />

case v 1 < v 2 < 0 can be shown in the same way). Th<strong>en</strong>, for all t ∈ R,<br />

Therefore,<br />

∫ ∞<br />

−∞<br />

e x (1 − cos v 1 x)ν(dx) =<br />

∫ ∞<br />

−∞<br />

e t/v 1<br />

ν(t/v 1 )<br />

v 1<br />

≤ <strong>et</strong>/v 2<br />

ν(t/v 2 )<br />

v 2<br />

.<br />

(1 − cos t) <strong>et</strong>/v 1<br />

ν(t/v 1 )<br />

dt<br />

−∞<br />

∫ ∞<br />

≤<br />

−∞<br />

v 1<br />

(1 − cos t) <strong>et</strong>/v 2<br />

ν(t/v 2 )<br />

v 2<br />

dt =<br />

and it follows from Equation (1.29) that |Φ T (v 1 − i)| ≥ |Φ T (v 2 − i)|.<br />

∫ ∞<br />

−∞<br />

e x (1 − cos v 2 x)ν(dx),<br />

Proposition 1.11. L<strong>et</strong> {X t } t≥0 be a real-valued Lévy process with tripl<strong>et</strong> (A, ν, γ) and characteristic<br />

function Φ T and l<strong>et</strong> Σ > 0.<br />

1. Suppose A > 0. Th<strong>en</strong> the truncation error in Equation (1.28) satisfies:<br />

|ε T | ≤<br />

8<br />

πT Σ 2 L 3 e− T L2 Σ 2 8<br />

8 +<br />

πT AL 3 e− T L2 A<br />

8 . (1.30)<br />

2. Suppose that ν has a <strong>de</strong>nsity of the form ν(x) = e−x<br />

|x|<br />

f(x), where f is increasing on (−∞, 0)<br />

and <strong>de</strong>creasing on (0, ∞). Th<strong>en</strong> the truncation error in Equation (1.28) satisfies:<br />

Proof. From Equation, (1.28),<br />

|ε T | ≤ 1 ∫ ∞<br />

|˜ζ T (v)|dv + 1<br />

2π L/2<br />

2π<br />

≤ 1 ∫<br />

2π<br />

8<br />

|ε T | ≤<br />

πT Σ 2 L 3 e− T L2 Σ 2<br />

8 + |Φ T (L/2 − i)| + |Φ T (−L/2 − i)|<br />

.<br />

πL<br />

∫ −L/2<br />

−∞<br />

(−∞,−L/2)∪(L/2,∞)<br />

|˜ζ T (v)|dv<br />

|Φ T (v − i)|dv<br />

v 2 + 1 ∫<br />

|Φ Σ T<br />

(v − i)|dv<br />

2π (−∞,−L/2)∪(L/2,∞) v 2 .


1.4. PRICING EUROPEAN OPTIONS 53<br />

By the first part of Lemma 1.10,<br />

∫<br />

1 ∞<br />

|Φ T (v − i)|<br />

2π L/2 v 2 dv ≤ 1 ∫ ∞<br />

dv<br />

2π L/2 v 2 e−T Av2 /2 = 1 ∫ ∞<br />

4π L 2 /4<br />

≤ 1<br />

4π<br />

dz Az/2<br />

e−T<br />

z3/2 ∫<br />

8 ∞<br />

L 3 dze −T Az/2 =<br />

L 2 /4<br />

4<br />

πT AL 3 e− L2 AT<br />

8 ,<br />

and since a similar bound can be obtained for Φ Σ T<br />

, this proves the first part of the proposition.<br />

To prove the second statem<strong>en</strong>t, observe that from the second part of Lemma 1.10,<br />

∫<br />

1 ∞<br />

2π L/2<br />

|Φ T (v − i)|<br />

v 2 dv ≤ |Φ ∫<br />

T (L/2 − i)| ∞<br />

dv<br />

2π L/2<br />

v 2 = |Φ T (L/2 − i)|<br />

.<br />

πL<br />

To compute a bound for the discr<strong>et</strong>ization (sampling) error ε D in Equation (1.28), we <strong>de</strong>fine<br />

constants C k , k ≥ 1 by<br />

⎧<br />

⎪⎨<br />

C k =<br />

⎪⎩<br />

|Φ (k)<br />

T<br />

(−i)|, k is ev<strong>en</strong>,<br />

|Φ (k+1)<br />

T<br />

(−i)| k<br />

k+1 , k is odd,<br />

(1.31)<br />

and start by proving a technical lemma.<br />

Lemma 1.12. L<strong>et</strong> {X t } t≥0 be a real-valued Lévy process with tripl<strong>et</strong> (A, ν, γ) and characteristic<br />

function Φ T , such that the condition (1.23) is satisfied. L<strong>et</strong> Γ T (v) := Φ T (v−i)−1<br />

v<br />

. Th<strong>en</strong> for all<br />

n ≥ 0,<br />

|Γ (n)<br />

T (v)| ≤ C n+1<br />

n + 1<br />

∀ v ∈ R<br />

Proof. Un<strong>de</strong>r the condition (1.23), for all n ≥ 1, E[|X T | n e X T<br />

] < ∞. Therefore, by Lebesgue’s<br />

dominated converg<strong>en</strong>ce theorem, for all n ≥ 1,<br />

Φ (n)<br />

T<br />

∫ ∞<br />

(v − i) = (ix) n e i(v−i)x µ T (dx),<br />

where µ T is the probability distribution of X T . For ev<strong>en</strong> n > 1 this implies:<br />

−∞<br />

∫ ∞<br />

|Φ (n)<br />

T<br />

(v − i)| ≤ x n e x µ T (dx) = C n ,<br />

−∞<br />

and for odd n ≥ 1, using J<strong>en</strong>s<strong>en</strong>’s inequality for the probability measure ˜µ T = e x µ T yields:<br />

∫ ∞<br />

(∫ ∞<br />

) n<br />

|Φ (n)<br />

T<br />

(v − i)| ≤ |x| n e x µ T (dx) ≤ |x| n+1 e x n+1<br />

µ T (dx) = Cn ,<br />

−∞<br />

−∞


54 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS<br />

From the martingale condition,<br />

Γ T (v) = Φ T (v − i) − Φ T (−i)<br />

v<br />

=<br />

∫ 1<br />

0<br />

Φ ′ T (vt − i)dt.<br />

Applying the dominated converg<strong>en</strong>ce theorem once again, we conclu<strong>de</strong> that<br />

and therefore |Γ (n)<br />

T (v)| ≤ C n+1<br />

n+1<br />

Γ (n)<br />

T (v) = ∫ 1<br />

0<br />

for all n ≥ 0.<br />

t n Φ (n+1)<br />

T<br />

(vt − i)dt,<br />

Proposition 1.13. Suppose that the integral in (1.28) is approximated using the trapezoidal<br />

rule. Th<strong>en</strong> the discr<strong>et</strong>ization error ε D satisfies<br />

{<br />

2∑<br />

|ε D | ≤ ∆2 C 3−l + C3−l<br />

Σ (<br />

∆ + π 6π (3 − l)! 2<br />

l=0<br />

)<br />

e |k−rT | + log<br />

(<br />

L<br />

2 + √<br />

L 2<br />

4 + 1 )<br />

where C Σ k is computed as in (1.31) for the characteristic function ΦΣ T .<br />

}<br />

|k − rT | l<br />

, (1.32)<br />

l!<br />

If this integral is approximated using the Simpson rule, ε D satisfies<br />

{ (<br />

4∑<br />

|ε D | ≤ ∆4 C 5−l + C5−l<br />

Σ (<br />

2∆ + π √ ) }<br />

)<br />

e |k−rT | L L<br />

+ log<br />

5π (5 − l)!<br />

2<br />

2 + 2<br />

4 + 1 |k − rT | l<br />

, (1.33)<br />

l!<br />

l=0<br />

Proof. The trapezoidal rule (cf. [32]) is <strong>de</strong>fined by<br />

∫ x+h<br />

x<br />

f(ξ)dξ = 1 2 h(f(x) + f(x + h)) + R with R = − 1 12 h3 f ′′ (x ∗ )<br />

for some x ∗ ∈ [x, x + h]. The sampling error in (1.28) therefore satisfies:<br />

|ε D | ≤ ∆3 N−2<br />

∑<br />

24π<br />

= ∆3<br />

24π<br />

m=0 v∈[−L/2+m∆,−L/2+(m+1)∆]<br />

N−2<br />

∑<br />

m=0<br />

sup<br />

∂ 2<br />

∣∂v 2 e−ivk ˜ζT (v)<br />

∣<br />

{<br />

sup<br />

∂ 2<br />

∣ ∣∣∣<br />

∣∂v 2 e−ivk ζ T (v)<br />

∣ + sup ∂ 2<br />

}<br />

∂v 2 e−ivk ζT Σ (v)<br />

∣ , (1.34)<br />

where ζT Σ (v) := e ivrT ΦΣ T<br />

(v − i) − 1<br />

.<br />

iv(1 + iv)<br />

Observe that e −ivk ζ T (v) = e−iv(k−rT )<br />

i(1+iv)<br />

Γ T (v), and therefore<br />

∂ n<br />

∣∂v n (eivk ζ T (v))<br />

∣<br />

≤<br />

n∑<br />

l=0<br />

≤ n!<br />

|Γ (n−l)<br />

T<br />

n∑<br />

l=0<br />

n!<br />

(v)|<br />

(n − l)!<br />

C n−l+1<br />

(n − l + 1)!<br />

l∑ |k − rT | j (<br />

1<br />

√<br />

j! 1 + v 2<br />

j=0<br />

) l−j+1<br />

l∑ |k − rT | j ( )<br />

1 l−j+1<br />

√ := ˜g n (v).<br />

j! 1 + v 2<br />

j=0


1.4. PRICING EUROPEAN OPTIONS 55<br />

Since the bound ˜g n (v) is increasing on (−∞, 0) and <strong>de</strong>creasing on (0, ∞),<br />

∆ 3 N−2<br />

∑<br />

sup<br />

∂ 2<br />

∫ L/2<br />

24π ∣∂v 2 e−ivk ζ T (v)<br />

∣ ≤ ∆3<br />

12π ˜g 2(0) + ∆2<br />

24π<br />

m=0<br />

−L/2<br />

⎧<br />

2∑<br />

≤ ∆2 C<br />

⎨<br />

3−l<br />

l∑<br />

12π (3 − l)! ⎩ 2∆ |k − rT | j ∑l−1<br />

|k − rT | j<br />

+<br />

j!<br />

j!<br />

≤ ∆2<br />

6π<br />

l=0<br />

2∑<br />

l=0<br />

j=0<br />

j=0<br />

{<br />

C (<br />

3−l<br />

∆ + π )<br />

e |k−rT | + log<br />

(3 − l)! 2<br />

˜g 2 (v)dv<br />

∫ L/2<br />

−L/2<br />

(<br />

L<br />

2 + √<br />

L 2<br />

4 + 1 )<br />

dv |k − rT |l<br />

+<br />

1 + v2 l!<br />

}<br />

|k − rT | l<br />

l!<br />

.<br />

∫ L/2<br />

−L/2<br />

⎫<br />

dv<br />

⎬<br />

√<br />

1 + v 2 ⎭<br />

To compl<strong>et</strong>e the proof of (1.32) it remains to substitute this bound and a similar bound for<br />

∆<br />

∑<br />

∣ ∣ 3 N−2 ∣∣<br />

24π m=0 sup ∂ 2<br />

e −ivk ζ Σ ∣∣<br />

∂v 2 T (v) into (1.34).<br />

The Simpson rule (cf. [32]) is <strong>de</strong>fined by:<br />

∫ x+2h<br />

x<br />

f(ξ)dξ = 1 h(f(x) + 4f(x + h) + f(x + 2h)) + R,<br />

3<br />

where R = 1<br />

90 h4 f (4) (x ∗ )<br />

for some x ∗ ∈ [x, x + 2h]. The proof of (1.33) can be carried out in the same way as for the<br />

trapezoidal rule.<br />

Example 1.1. L<strong>et</strong> us compute the truncation and discr<strong>et</strong>ization errors in the Merton mo<strong>de</strong>l<br />

(1.16) with param<strong>et</strong>ers σ = 0.1, λ = 2, δ = 0.1, µ = 0 for options with maturity T = 0.25.<br />

Taking Σ = 0.1, we obtain the following values of the coeffici<strong>en</strong>ts C k and C Σ k :<br />

k C k Ck<br />

Σ<br />

1 0.0871 0.05<br />

2 0.0076 0.0025<br />

3 0.0024 2.85 · 10 −4<br />

4 3.29 · 10 −4 1.88 · 10 −5<br />

5 1.82 · 10 −4 2.99 · 10 −6<br />

With 4048 points and the log strike step d equal to 0.01, the truncation error is extremely small:<br />

ε T = 2 · 10 −59 , and the discr<strong>et</strong>ization error for at the money options is giv<strong>en</strong> by ε D = 0.0013<br />

for the trapezoidal rule and by ε D = 3.8 · 10 −5 for the Simpson rule. Since the call option price<br />

in this s<strong>et</strong>ting (S = K = 1, r = q = 0) is C = 0.0313, we conclu<strong>de</strong> that the Simpson rule gives<br />

an acceptable pricing error for this choice of param<strong>et</strong>ers.


56 CHAPTER 1. LEVY PROCESSES AND EXP-LEVY MODELS


Chapter 2<br />

The calibration problem and its<br />

regularization<br />

This chapter lays the theor<strong>et</strong>ical foundations of our calibration m<strong>et</strong>hod. Discr<strong>et</strong>ization of the<br />

regularized calibration problem and numerical implem<strong>en</strong>tation of the calibration algorithm are<br />

addressed in the next chapter.<br />

The calibration problem consists, roughly speaking, of finding a risk-neutral expon<strong>en</strong>tial<br />

Lévy mo<strong>de</strong>l consist<strong>en</strong>t with mark<strong>et</strong> prices of tra<strong>de</strong>d options {C M (T i , K i )} i∈I for some in<strong>de</strong>x s<strong>et</strong><br />

I. In other words, the solution of the calibration problem is a probability Q on the path space<br />

(Ω, F), such that (X, Q) is a Lévy process, satisfying the martingale condition (1.2) and such<br />

that the option prices, computed using (1.14) are in some s<strong>en</strong>se close to mark<strong>et</strong> prices C M .<br />

Suppose first that the mark<strong>et</strong> data C M are consist<strong>en</strong>t with the class of expon<strong>en</strong>tial Lévy<br />

mo<strong>de</strong>ls. This is for example the case wh<strong>en</strong> the true mo<strong>de</strong>l un<strong>de</strong>rlying mark<strong>et</strong> data is an expon<strong>en</strong>tial<br />

Lévy mo<strong>de</strong>l, but this is not the only situation where the above is true: many mo<strong>de</strong>ls<br />

may give the same prices for a giv<strong>en</strong> s<strong>et</strong> of European options. For instance, it is easy to construct,<br />

using Dupire’s formula, a local volatility mo<strong>de</strong>l that gives the same prices, for a s<strong>et</strong> of<br />

European options, as a giv<strong>en</strong> exp-Lévy mo<strong>de</strong>l. D<strong>en</strong>oting by L the s<strong>et</strong> of all probabilities Q<br />

such that (X, Q) is a Lévy process and by M the s<strong>et</strong> of probabilities Q such that {e Xt } t≥0 is a<br />

Q-martingale, the calibration problem assumes the following form:<br />

57


58 CHAPTER 2. THE CALIBRATION PROBLEM<br />

Calibration problem for data, consist<strong>en</strong>t with exp-Lévy mo<strong>de</strong>l.<br />

of call options {C M (T i , K i )} i∈I , find Q ∗ ∈ M ∩ L, such that<br />

Giv<strong>en</strong> mark<strong>et</strong> prices<br />

∀i ∈ I, C Q∗ (T i , K i ) = C M (T i , K i ). (2.1)<br />

In most cases, however, Equations (2.1) cannot be solved exactly, either because the observed<br />

option prices contain errors or because the mark<strong>et</strong> in question cannot be <strong>de</strong>scribed by an<br />

expon<strong>en</strong>tial Lévy mo<strong>de</strong>l. In this case a common practice is to replace the exact constraints<br />

(2.1) by a nonlinear least squares calibration problem. Section 2.1 takes a critical look at<br />

this approach and establishes some limits of its applicability to non-param<strong>et</strong>ric calibration of<br />

expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls.<br />

An important difficulty of least squares calibration is its lack of i<strong>de</strong>ntification, meaning<br />

that a finite number of observed option prices does not allow to reconstruct the law of a Lévy<br />

process in a unique fashion. To address this issue, we suggest, in Section 2.2 to reformulate<br />

the calibration problem as that of finding the risk-neutral expon<strong>en</strong>tial Lévy mo<strong>de</strong>l that has the<br />

smallest relative <strong>en</strong>tropy with respect to a giv<strong>en</strong> prior probability measure among all solutions<br />

of the least squares calibration problem. Section 2.3 reviews the literature on the use of relative<br />

<strong>en</strong>tropy for pricing and calibration and places the pres<strong>en</strong>t study into the context of previous<br />

work on this subject.<br />

The use of relative <strong>en</strong>tropy for selection of solutions removes to some ext<strong>en</strong>t the i<strong>de</strong>ntification<br />

problem but the resulting calibration problem is still ill-posed: small errors in mark<strong>et</strong> data may<br />

lead to large changes of its solution. The last section of this chapter uses the m<strong>et</strong>hod of<br />

regularization to approximate the solutions of this problem in a stable manner in pres<strong>en</strong>ce of<br />

data errors.<br />

2.1 Least squares calibration<br />

Wh<strong>en</strong> the mark<strong>et</strong> data is not consist<strong>en</strong>t with the class of expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls, the exact<br />

calibration problem may not have a solution. In this case one may consi<strong>de</strong>r an approximate<br />

solution: instead of reproducing the mark<strong>et</strong> option prices exactly, one may look for a Lévy tripl<strong>et</strong><br />

which reproduces them in the best possible way in the least squares s<strong>en</strong>se. L<strong>et</strong> w be a probability


2.1. LEAST SQUARES CALIBRATION 59<br />

measure on [0, T ∞ ] × [0, ∞) (the weighting measure, <strong>de</strong>termining the relative importance of<br />

differ<strong>en</strong>t options). An option data s<strong>et</strong> is <strong>de</strong>fined as a mapping C : [0, T ∞ ] × [0, ∞) → [0, ∞) and<br />

the data s<strong>et</strong>s that coinci<strong>de</strong> w-almost everywhere are consi<strong>de</strong>red i<strong>de</strong>ntical. One can introduce a<br />

norm on option data s<strong>et</strong>s via<br />

∫<br />

‖C‖ 2 w :=<br />

[0,T ∞]×[0,∞)<br />

C(T, K) 2 w(dT × dK). (2.2)<br />

The quadratic pricing error in mo<strong>de</strong>l Q is th<strong>en</strong> giv<strong>en</strong> by ‖C M − C Q ‖ 2 w. If the number of constraints<br />

is finite th<strong>en</strong> w = ∑ N<br />

i=1 w iδ (Ti ,K i )(dT × dK) (we suppose that there are N constraints),<br />

where {w i } 1≤i≤N are positive weights that sum up to one. Therefore, in this case<br />

‖C M − C Q ‖ 2 w =<br />

N∑<br />

w i (C M (T i , K i ) − C Q (T i , K i )) 2 . (2.3)<br />

i=1<br />

The calibration problem now takes the following form:<br />

Least squares calibration problem. Giv<strong>en</strong> prices C M of call options, find Q ∗ ∈ M ∩ L,<br />

such that<br />

‖C M − C Q∗ ‖ 2 w =<br />

inf ‖C M − C Q ‖ 2 w. (2.4)<br />

Q∈M∩L<br />

In the sequel, any such Q ∗ will be called a least squares solution and the s<strong>et</strong> of all least<br />

squares solutions will be <strong>de</strong>noted by Q LS .<br />

Several authors (see for example [2, 10]) have used the form (2.4) without taking into account<br />

that the least squares calibration problem is ill-posed in several ways. The principal difficulties<br />

of theor<strong>et</strong>ical nature are the lack of i<strong>de</strong>ntification (knowing the prices of a finite number of<br />

options is not suffici<strong>en</strong>t to reconstruct the Lévy process), abs<strong>en</strong>ce of solution (in some cases<br />

ev<strong>en</strong> the least squares problem may not admit a solution) and abs<strong>en</strong>ce of continuity of solution<br />

with respect to mark<strong>et</strong> data. On the other hand, ev<strong>en</strong> if a solution exists, it is very difficult to<br />

find numerically, because the functional ‖C M −C Q ‖ 2 is typically non-convex and has many local<br />

minima, prev<strong>en</strong>ting a gradi<strong>en</strong>t-based minimization algorithm from finding the true solution. In<br />

the rest of this section we discuss these difficulties in <strong>de</strong>tail.


60 CHAPTER 2. THE CALIBRATION PROBLEM<br />

2.1.1 Lack of i<strong>de</strong>ntification<br />

If the data are consist<strong>en</strong>t with an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l and call option prices are known for<br />

one maturity and all strikes, the characteristic tripl<strong>et</strong> of the un<strong>de</strong>rlying Lévy process could be<br />

<strong>de</strong>duced in the following way:<br />

• Compute the characteristic function Φ T<br />

Equation (1.24).<br />

of log stock price by Fourier transform as in<br />

• Deduce the unit variance of the Gaussian compon<strong>en</strong>t A and the Lévy measure ν from the<br />

characteristic function Φ T . First, A can be found as follows (see [87, page 40]):<br />

Now, <strong>de</strong>noting ψ(u) ≡ log Φ T (u)<br />

T<br />

∫ 1<br />

−1<br />

A = lim<br />

u→∞ −2 log Φ T (u)<br />

T u 2 (2.5)<br />

+ Au2<br />

2<br />

, it can be shown (see Equation (8.10) in [87]) that<br />

∫ ∞<br />

(ψ(u) − ψ(u + z))dz = 2<br />

−∞<br />

e iux (1 − sin x )ν(dx) (2.6)<br />

x<br />

Therefore, the left-hand si<strong>de</strong> of (2.6) is the Fourier transform of the positive finite measure<br />

2(1 − sin x<br />

x<br />

)ν(dx). This means that this measure, and, consequ<strong>en</strong>tly, the Lévy measure ν<br />

is uniquely <strong>de</strong>termined by ψ.<br />

However, call prices are only available for a finite number of strikes. This number may be quite<br />

small (b<strong>et</strong>we<strong>en</strong> 10 and 40 in real examples). Therefore, the above procedure cannot be applied<br />

and ν and A must be computed by minimizing the pricing error ‖C M − C Q ‖ 2 w. Giv<strong>en</strong> that<br />

the number of calibration constraints (option prices) is finite and not very large, there may be<br />

many Lévy tripl<strong>et</strong>s which reproduce call prices with equal precision. This means that the error<br />

landscape may have flat regions in which the error has a low s<strong>en</strong>sitivity to variations in A and<br />

ν.<br />

One may think that in a param<strong>et</strong>ric mo<strong>de</strong>l with few param<strong>et</strong>ers one will not <strong>en</strong>counter this<br />

problem since there are (many) more options than param<strong>et</strong>ers. This is not true, as illustrated by<br />

the following empirical example. Figure 2.1 repres<strong>en</strong>ts the magnitu<strong>de</strong> of the quadratic pricing<br />

error for the Merton mo<strong>de</strong>l (1.16) on a data s<strong>et</strong> of DAX in<strong>de</strong>x options, as a function of the<br />

diffusion coeffici<strong>en</strong>t σ and the jump int<strong>en</strong>sity λ, other param<strong>et</strong>ers remaining fixed. It can be<br />

observed that if one increases the diffusion volatility while simultaneously <strong>de</strong>creasing the jump


2.1. LEAST SQUARES CALIBRATION 61<br />

int<strong>en</strong>sity in a suitable manner, the calibration error changes very little: there is a long “valley”<br />

in the error landscape (highlighted by the dashed white line in Figure 2.1). A gradi<strong>en</strong>t <strong>de</strong>sc<strong>en</strong>t<br />

m<strong>et</strong>hod will typically succeed in locating the valley but will stop at a more or less random point<br />

in it. At first glance this does not seem to be too much of a problem: since the algorithm finds<br />

the valley’s bottom, the best calibration quality will be achieved anyway. However, after a small<br />

change in option prices, the outcome of this calibration algorithm may shift a long way along<br />

the valley. This means that if the calibration is performed every day, one may come up with<br />

wildly oscillating param<strong>et</strong>ers of the Lévy process, leading to important changes in the quantities<br />

computed from these param<strong>et</strong>ers, like prices of exotic options, ev<strong>en</strong> if the mark<strong>et</strong> option prices<br />

change very little.<br />

3<br />

x 10 5<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

2<br />

1.5<br />

1<br />

λ<br />

0.5<br />

0<br />

0.1<br />

0.12<br />

0.14<br />

σ<br />

0.16<br />

0.18<br />

0.2<br />

Figure 2.1: Sum of squared differ<strong>en</strong>ces b<strong>et</strong>we<strong>en</strong> mark<strong>et</strong> prices (DAX options maturing in 10<br />

weeks) and mo<strong>de</strong>l prices in Merton mo<strong>de</strong>l, as a function of param<strong>et</strong>ers σ and λ the other ones<br />

being fixed. Dashed white line shows the “valley” along which the error function changes very<br />

little.<br />

Figure 2.2 illustrates the same problem in the non-param<strong>et</strong>ric s<strong>et</strong>ting. The two graphs<br />

repres<strong>en</strong>t the result of a non-linear least squares minimization where the variable is the vector<br />

of discr<strong>et</strong>ized values of ν on a grid (see Section 3.1). In both cases the same option prices are<br />

used, the only differ<strong>en</strong>ce being the starting points of the optimization routines. In the first case<br />

(solid line) a Merton mo<strong>de</strong>l with int<strong>en</strong>sity λ = 1 is used and in the second case (dashed line)


62 CHAPTER 2. THE CALIBRATION PROBLEM<br />

we used a Merton mo<strong>de</strong>l with int<strong>en</strong>sity λ = 5. As can be se<strong>en</strong> in Figure 2.2 (left graph), the<br />

results of the minimization are totally differ<strong>en</strong>t! However, although the calibrated measures are<br />

differ<strong>en</strong>t, the prices that they correspond to are almost the same (see Figure 2.2, right graph),<br />

and the final values of the calibration functional for the two curves differ very little.<br />

The i<strong>de</strong>ntification problem will be addressed in the next section by adding information in<br />

the form of a prior Lévy process to the calibration problem.<br />

20<br />

18<br />

λ 0<br />

= 1, ε = 0.154<br />

λ 0<br />

= 5, ε = 0.153<br />

1.1<br />

1<br />

mark<strong>et</strong><br />

λ 0<br />

= 1<br />

λ 0<br />

=5<br />

16<br />

0.9<br />

14<br />

0.8<br />

12<br />

0.7<br />

10<br />

0.6<br />

8<br />

0.5<br />

6<br />

0.4<br />

4<br />

0.3<br />

2<br />

0.2<br />

0<br />

−1.2 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6<br />

0.1<br />

3500 4000 4500 5000 5500 6000 6500 7000 7500<br />

Figure 2.2: Left: Lévy measures calibrated to DAX option prices, maturity 3 months via nonlinear<br />

least squares m<strong>et</strong>hod. The starting jump measure for both graphs is Gaussian; the jump<br />

int<strong>en</strong>sity λ 0 is initialized to 1 for the solid curve and to 5 for the dashed one. ε <strong>de</strong>notes the<br />

value of the calibration functional wh<strong>en</strong> the gradi<strong>en</strong>t <strong>de</strong>sc<strong>en</strong>t algorithm stops. Right: implied<br />

volatility smiles corresponding to these two measures.<br />

2.1.2 Exist<strong>en</strong>ce of solutions<br />

In this section, we will first give an (artificial) example which shows that the least squares<br />

formulation (2.4) does not in g<strong>en</strong>eral guarantee the exist<strong>en</strong>ce of solution. However, un<strong>de</strong>r<br />

additional conditions we are able to prove the exist<strong>en</strong>ce of a solution of the least squares<br />

calibration problem.<br />

Example 2.1. Suppose that S 0 = 1, there are no interest rates or divi<strong>de</strong>nds and the (equally<br />

weighted) mark<strong>et</strong> data consist of the following two observations:<br />

C M (T = 1, K = 1) = 1 − e −λ and C M (T = 1, K = e λ ) = 0, (2.7)


2.1. LEAST SQUARES CALIBRATION 63<br />

with some λ > 0. It is easy to see that these prices are, for example, compatible with the<br />

(martingale) ass<strong>et</strong> price process S t = e λt 1 t≤τ1 , where τ 1 is the time of the first jump of a<br />

Poisson process with int<strong>en</strong>sity λ. We will show that if the mark<strong>et</strong> data are giv<strong>en</strong> by (2.7), the<br />

calibration problem (2.4) does not admit a solution.<br />

Abs<strong>en</strong>ce of arbitrage implies that in every risk-neutral mo<strong>de</strong>l Q, for fixed T , C Q (T, K) is a<br />

convex function of K and that C Q (T, K = 0) = 1. The only convex function which satisfies this<br />

equality and passes through the mark<strong>et</strong> data points (2.7) is giv<strong>en</strong> by C(T = 1, K) = (1−Ke −λ ) + .<br />

Therefore, in every arbitrage-free mo<strong>de</strong>l that is an exact solution of the calibration problem with<br />

mark<strong>et</strong> data (2.7), for every K ≥ 0, P [S 1 ≤ K] = e −λ 1 K≤e λ. Since in an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l<br />

P [S 1 > 0] = 1, there is no risk-neutral expon<strong>en</strong>tial Lévy mo<strong>de</strong>l for which ‖C M − C Q ‖ w = 0.<br />

On the other hand, inf Q∈M∩L ‖C M − C Q ‖ 2 w = 0. In<strong>de</strong>ed, l<strong>et</strong> {N t } t≥0 be a Poisson process<br />

with int<strong>en</strong>sity λ. Th<strong>en</strong> for every n, the process<br />

X n t := −nN t + λt(1 − e −n ) (2.8)<br />

belongs to M ∩ L and<br />

lim<br />

n→∞ E[(eXn t<br />

− K) + ] = lim<br />

∞∑<br />

n→∞<br />

k=0<br />

(<br />

−λt (λt)k<br />

+<br />

e e −nk+λt(1−e−n) − K)<br />

= (1 − Ke −λt ) + .<br />

k!<br />

We have shown that inf Q∈M∩L ‖C M − C Q ‖ 2 = 0 and that for no Lévy process Q ∈ M ∩ L,<br />

‖C M − C Q ‖ 2 = 0. Tog<strong>et</strong>her this <strong>en</strong>tails that the calibration problem (2.4) does not admit a<br />

solution.<br />

This example makes clear that to solve the calibration problem (2.4), we must at least<br />

impose a bound on the jumps of the solution. The following theorem provi<strong>de</strong>s an exist<strong>en</strong>ce<br />

result un<strong>de</strong>r this and another important condition (informally speaking, to control the variance<br />

of the solution we need to find a Lévy process for which the pricing error is already suffici<strong>en</strong>tly<br />

low). In the theorem and below L B <strong>de</strong>notes the s<strong>et</strong>s of all probabilities P ∈ L such that<br />

P [|∆X t | ≤ B ∀t : 0 ≤ t ≤ T ∞ ] = 1.<br />

Theorem 2.1. Suppose that for some Q 0 ∈ M ∩ L B and some couple (T 0 , K 0 ) with nonzero<br />

weight w 0 := w({T 0 , K 0 }) > 0,<br />

‖C Q 0<br />

− C M ‖ w < √ w 0 (S 0 − C M (T 0 , K 0 )). (2.9)


64 CHAPTER 2. THE CALIBRATION PROBLEM<br />

Th<strong>en</strong> the problem (2.4) has a solution in M ∩ L B : there exists Q ∗ ∈ M ∩ L B , such that<br />

‖C M − C Q∗ ‖ 2 w =<br />

The proof is based on the following lemmas.<br />

inf ‖C M − C Q ‖ 2 w.<br />

Q∈M∩L B<br />

Lemma 2.2. The pricing error functional Q ↦→ ‖C M − C Q ‖ 2 w, <strong>de</strong>fined by (2.2), is uniformly<br />

boun<strong>de</strong>d and weakly continuous on M ∩ L.<br />

Proof. From Equation (1.14), C Q (T, K) ≤ S 0 . Abs<strong>en</strong>ce of arbitrage in the mark<strong>et</strong> implies that<br />

the mark<strong>et</strong> option prices satisfy the same condition. Therefore, (C M (T, K) − C Q (T, K)) 2 ≤ S0<br />

2<br />

and since w is a probability measure, ‖C M − C Q ‖ 2 w ≤ S0 2.<br />

L<strong>et</strong> {Q n } n≥1 ⊂ M ∩ L and Q ∈ M ∩ L be such that Q n ⇒ Q. For all T, K,<br />

lim<br />

n<br />

C Qn (T, K) = e −rT lim<br />

n<br />

E Qn [(S 0 e rT +X T<br />

− K) + ]<br />

= e −rT lim<br />

n<br />

E Qn [S 0 e rT +X T<br />

− K] + e −rT lim<br />

n<br />

E Qn [(K − S 0 e rT +X T<br />

) + ]<br />

= S 0 − Ke −rT + e −rT E Q [(K − S 0 e rT +X T<br />

) + ] = C Q (T, K).<br />

Therefore, by the dominated converg<strong>en</strong>ce theorem, ‖C M − C Qn ‖ 2 w → ‖C M − C Q ‖ 2 w.<br />

Lemma 2.3. For all B > 0, T > 0, K > 0 and all C with 0 < C < S 0 , the s<strong>et</strong> of Lévy processes<br />

Q ∈ M ∩ L B satisfying C Q (T, K) ≤ C is relatively weakly compact.<br />

Proof. By Prohorov’s theorem, weak relative compactness is implied by the tightness of this<br />

family of probability measures. Since the jumps are boun<strong>de</strong>d, to prove the tightness, by Proposition<br />

1.6 it is <strong>en</strong>ough to show that there exist constants C 1 and C 2 such that for every Lévy<br />

process Q ∈ M∩L B such that C Q (T, K) ≤ C, its characteristic tripl<strong>et</strong> (A, ν, γ) satisfies |γ| ≤ C 1<br />

and A + ∫ 1<br />

−1 x2 ν(dx) ≤ C 2 .<br />

By the risk-neutrality condition (1.2),<br />

γ = − A ∫ B<br />

2 − (e x − 1 − x1 |x|≤1 )ν(dx)<br />

−B<br />

It is easy to see that |e x − 1 − x1 |x|≤1 | ≤ e B∧1 x 2<br />

e B∧1 (A + ∫ ∞<br />

−∞ x2 ν(dx)).<br />

for x ∈ (−∞, B] and therefore |γ| ≤<br />

This means that to prove the lemma, it is suffici<strong>en</strong>t to show that


2.1. LEAST SQUARES CALIBRATION 65<br />

A + ∫ ∞<br />

−∞ x2 ν(dx) < C 3 for some constant C 3 , in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt on the choice of the Lévy process Q.<br />

Since for all α ∈ [0, 1], e x ∧ 1 ≤ e αx , we have:<br />

S 0 − C Q (T, K) = Ke −rT E[e X T −m ∧ 1] ≤ Ke −rT E[e α(X T −m) ] = S α 0 (Ke −rT ) 1−α E[e αX T<br />

],<br />

where m = log(Ke −rT /S 0 ). The right-hand si<strong>de</strong> can be computed using the Lévy-Khintchine<br />

formula:<br />

E[e αX T<br />

] = exp T<br />

{<br />

− A 2 (α − α2 ) −<br />

Choosing α = 1/2, the above reduces to<br />

E[e X T /2 ] = exp T {− A 8 − 1 2<br />

On the other hand, it is easy to check that<br />

and therefore<br />

A<br />

8 + 1 2<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

}<br />

(αe αx − e αx − α + 1)ν(dx) .<br />

−∞<br />

(e x/2 − 1) 2 ν(dx)}.<br />

(e x/2 − 1) 2 ν(dx) ≥ e−B<br />

8 (A + ∫ ∞<br />

−∞<br />

x 2 ν(dx))<br />

∫ √ √ ∞<br />

A + x 2 ν(dx) ≤ 8eB<br />

−∞<br />

T log S0 Ke −rT<br />

S 0 − C Q (T, K) ≤ 8eB<br />

T<br />

log S0 Ke −rT<br />

S 0 − C ,<br />

which finishes the proof of the lemma.<br />

In the following lemma, L + B <strong>de</strong>notes the s<strong>et</strong>s of all probabilities P ∈ L such that P [∆X t ≤<br />

B ∀t : 0 ≤ t ≤ T ∞ ] = 1.<br />

Lemma 2.4. Both M ∩ L B and M ∩ L + B<br />

are weakly closed for every B > 0.<br />

Proof. We give the proof for M ∩ L B ; the proof for M ∩ L + B<br />

can be done in a similar fashion.<br />

L<strong>et</strong> {Q n } ∞ n=1 ⊂ M ∩ L B with characteristic tripl<strong>et</strong>s (A n , ν n , γ h n) with respect to a continuous<br />

boun<strong>de</strong>d truncation function h, satisfying h(x) = x in a neighborhood of 0, and l<strong>et</strong> Q be a Lévy<br />

process with characteristic tripl<strong>et</strong> (A, ν, γ h ) with respect to h, such that Q n ⇒ Q. Note that<br />

a sequ<strong>en</strong>ce of Lévy processes cannot converge to anything other than a Lévy process because<br />

due to converg<strong>en</strong>ce of characteristic functions, the limiting process must have stationary and<br />

in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt increm<strong>en</strong>ts. Define a function f by<br />

⎧<br />

0, |x| ≤ B,<br />

⎪⎨<br />

f(x) := 1, |x| ≥ 2B,<br />

⎪⎩ |x|−B<br />

B<br />

B < |x| < 2B.


66 CHAPTER 2. THE CALIBRATION PROBLEM<br />

By Proposition 1.7, ∫ ∞<br />

−∞ f(x)ν(dx) = lim n→∞<br />

∫ ∞<br />

−∞ f(x)ν n(dx) = 0, which implies that the<br />

jumps of Q are boun<strong>de</strong>d by B.<br />

Define a function g by<br />

⎧<br />

⎨ e x − 1 − h(x) − 1 2<br />

g(x) :=<br />

h2 (x), x ≤ B,<br />

⎩<br />

e B − 1 − h(B) − 1 2 h2 (B), x > B.<br />

Th<strong>en</strong>, by Proposition 1.7 and because Q n satisfies the martingale condition (1.2) for every n,<br />

γ h + A 2 + ∫ ∞<br />

−∞<br />

(e x − 1 − h(x))ν(dx) = γ h + A + ∫ ∞<br />

−∞ h2 (x)ν(dx)<br />

2<br />

= lim<br />

n→∞<br />

which shows that Q also satisfies the condition (1.2).<br />

{γn h + A n + ∫ ∞<br />

−∞ h2 (x)ν n (dx)<br />

+<br />

2<br />

Proof of Theorem 2.1. L<strong>et</strong> {Q n } n≥1 ⊂ M ∩ L B be such that<br />

Condition (2.9) implies that for every n,<br />

∫ ∞<br />

+<br />

lim ‖C M − C Qn ‖ 2<br />

n→∞<br />

w = inf ‖C M − C Q ‖ 2 w<br />

Q∈M∩L B<br />

−∞<br />

∫ ∞<br />

−∞<br />

and ‖C M − C Qn ‖ 2 w ≤ ‖C M − C Q 0<br />

‖ 2 w for all n.<br />

g(x)ν(dx)<br />

g(x)ν n (dx)<br />

}<br />

= 0,<br />

|S 0 − C Qn (T 0 , K 0 )| ≥ |S 0 − C M (T 0 , K 0 )| − |C M (T 0 , K 0 ) − C Qn (T 0 , K 0 )|<br />

≥ |S 0 − C M (T 0 , K 0 )| − ‖C M − C Qn ‖ w<br />

√<br />

w0<br />

≥ |S 0 − C M (T 0 , K 0 )| − ‖C M − C Q 0<br />

‖ w<br />

√<br />

w0<br />

> 0.<br />

Therefore, by Lemmas 2.3 and 2.4, there exists a subsequ<strong>en</strong>ce {Q nm } m≥1 of {Q n } n≥1 and<br />

Q ∗ ∈ M ∩ L B such that Q nm ⇒ Q. By Lemma 2.2,<br />

‖C M − C Q∗ ‖ 2 w = lim ‖C M − C Qnm ‖ 2<br />

m→∞<br />

w = inf ‖C M − C Q ‖ 2 w,<br />

Q∈M∩L B<br />

which shows that Q ∗ is a solution of the least squares calibration problem (2.4).<br />

2.1.3 Continuity<br />

Mark<strong>et</strong> option prices are typically <strong>de</strong>fined up to a bid-ask spread and the close prices used for<br />

calibration may therefore contain numerical errors. If the solution of the calibration problem


2.1. LEAST SQUARES CALIBRATION 67<br />

is not continuous with respect to mark<strong>et</strong> data, these small errors may dramatically alter the<br />

result of calibration, r<strong>en</strong><strong>de</strong>ring it compl<strong>et</strong>ely useless. On the other hand, ev<strong>en</strong> if mark<strong>et</strong> data<br />

did not contain errors, in abs<strong>en</strong>ce of continuity, small daily changes in prices could lead to large<br />

variations of calibrated param<strong>et</strong>ers and of other quantities computed using these param<strong>et</strong>ers<br />

(like prices of exotic options).<br />

Wh<strong>en</strong> the calibration problem has more than one solution, care should be tak<strong>en</strong> in <strong>de</strong>fining<br />

what is meant by continuity. In the sequel, we will use the following <strong>de</strong>finition (see, e.g. [40]),<br />

that applies to all calibration problems, discussed in this chapter.<br />

Definition 2.1. The solutions of a calibration problem are said to <strong>de</strong>p<strong>en</strong>d continuously on input<br />

data at the point C M if for every sequ<strong>en</strong>ce of data s<strong>et</strong>s {C n M } n≥0 such that ‖C n M −C M‖ w −−−→<br />

n→∞ 0,<br />

if Q n is a solution of the calibration problem with data C n M th<strong>en</strong><br />

1. {Q n } n≥1 has a weakly converg<strong>en</strong>t subsequ<strong>en</strong>ce {Q nm } m≥1 .<br />

2. The limit Q of every weakly converg<strong>en</strong>t subsequ<strong>en</strong>ce of {Q n } n≥1 is a solution of the<br />

calibration problem with data C M .<br />

Note that if the solution of the calibration problem with the limiting data C M is unique, this<br />

<strong>de</strong>finition reduces to the standard <strong>de</strong>finition of continuity, because in this case every subsequ<strong>en</strong>ce<br />

of {Q n } has a further subsequ<strong>en</strong>ce converging towards Q, which implies that Q n ⇒ Q.<br />

It is easy to construct an example of mark<strong>et</strong> data leading to a least squares calibration<br />

problem (2.4) that does not satisfy the above <strong>de</strong>finition.<br />

Example 2.2. Suppose that S 0 = 1, there are no interest rates or divi<strong>de</strong>nds and the mark<strong>et</strong> data<br />

for each n are giv<strong>en</strong> by a single observation:<br />

C n M(T = 1, K = 1) = E[(e Xn 1 − 1) + ] for n ≥ 1 and C M (T = 1, K = 1) = 1 − e −λ ,<br />

where Xt<br />

n is <strong>de</strong>fined by Equation (2.8) and λ > 0. Th<strong>en</strong> ‖CM n −C M‖ w −−−→ 0 and<br />

n→∞ Xn t is clearly<br />

a solution for data CM n , but the sequ<strong>en</strong>ce {Xn t } has no converg<strong>en</strong>t subsequ<strong>en</strong>ce (cf. Proposition<br />

1.7).<br />

Un<strong>de</strong>r conditions, similar to those of Theorem 2.1, Lemmas 2.2–2.4 allow to prove a continuity<br />

result for the least squares calibration problem (2.4).


68 CHAPTER 2. THE CALIBRATION PROBLEM<br />

Proposition 2.5. Suppose that there exists a couple (T 0 , K 0 ) with w 0 := w({T 0 , K 0 }) > 0 and<br />

l<strong>et</strong> C M be such that the condition (2.9) is satisfied for some Q 0 ∈ M ∩ L B . Th<strong>en</strong> the solutions<br />

of the least squares calibration problem (2.4) on M ∩ L B <strong>de</strong>p<strong>en</strong>d continuously on the mark<strong>et</strong><br />

data at the point C M .<br />

Proof. L<strong>et</strong> {CM n } n≥0 be a sequ<strong>en</strong>ce of data such that ‖CM n − C M‖ w −−−→<br />

n→∞<br />

n l<strong>et</strong> Q n be a solution of the calibration problem (2.4) on M ∩ L B with data CM<br />

n<br />

0, and for every<br />

(we can<br />

suppose without loss of g<strong>en</strong>erality that a solution exists for all n because it exists starting with<br />

a suffici<strong>en</strong>tly large n by Theorem 2.1). Th<strong>en</strong>, using the triangle inequality several times, we<br />

obtain:<br />

|S 0 − C Qn (T 0 , K 0 )|<br />

≥ |S 0 − C M (T 0 , K 0 )| − |C n M(T 0 , K 0 ) − C Qn (T 0 , K 0 )| − |C n M(T 0 , K 0 ) − C M (T 0 , K 0 )|<br />

≥ |S 0 − C M (T 0 , K 0 )| − ‖Cn M − CQ 0<br />

‖ w<br />

√<br />

w0<br />

− |C n M(T 0 , K 0 ) − C M (T 0 , K 0 )|<br />

≥ |S 0 − C M (T 0 , K 0 )| − ‖C M − C Q 0<br />

‖ w<br />

√<br />

w0<br />

− 2 ‖Cn M − C M‖ w<br />

√<br />

w0<br />

> C ′ > 0<br />

for some C ′ , starting from a suffici<strong>en</strong>tly large n. Therefore, by Lemmas 2.3 and 2.4, {Q n } has<br />

a subsequ<strong>en</strong>ce that converges weakly towards some Q ∗ ∈ M ∩ L B .<br />

L<strong>et</strong> {Q nm } ⊆ {Q n } with Q nm ⇒ Q ∗ ∈ M ∩ L B and l<strong>et</strong> Q ∈ M ∩ L B . Using Lemma 2.2 and<br />

the triangle inequality, we obtain:<br />

‖C Q∗ − C M ‖ w = lim ‖C Qnm − C M ‖ w ≤ lim inf<br />

m m<br />

{‖CQnm − C nm<br />

M ‖ w + ‖C nm<br />

M − C M‖ w }<br />

≤ lim inf<br />

m<br />

‖CQnm − C nm<br />

M ‖ w ≤ lim inf<br />

m<br />

‖CQ − C nm<br />

M ‖ w ≤ ‖C Q − C M ‖ w ,<br />

which shows that Q ∗ is in<strong>de</strong>ed a solution of the calibration problem (2.4) with data C M .<br />

2.1.4 Numerical difficulties of least squares calibration<br />

A major obstacle for the numerical implem<strong>en</strong>tation of the least squares calibration is the nonconvexity<br />

of the optimization problem (2.4), which is due to the non-convexity of the domain<br />

(M ∩ L), where the pricing error functional Q ↦→ ‖C Q − C M ‖ 2 is to be optimized. Due to this<br />

difficulty, the pricing error functional may have several local minima, and the gradi<strong>en</strong>t <strong>de</strong>sc<strong>en</strong>t


2.2. SELECTION USING RELATIVE ENTROPY 69<br />

algorithm used for numerical optimization may stop in one of these local minima, leading to a<br />

much worse calibration quality than that of the true solution.<br />

Figure 2.3 illustrates this effect in the (param<strong>et</strong>ric) framework of the variance gamma mo<strong>de</strong>l<br />

(1.18). The left graph shows the behavior of the calibration functional in a small region around<br />

the global minimum. Since in this mo<strong>de</strong>l there are only three param<strong>et</strong>ers, the i<strong>de</strong>ntification<br />

problem is not pres<strong>en</strong>t, and a nice profile appearing to be convex can be observed. However,<br />

wh<strong>en</strong> we look at the calibration functional on a larger scale (κ changes b<strong>et</strong>we<strong>en</strong> 1 and 8),<br />

the convexity disappears and we observe a ridge (highlighted with a dashed black line), which<br />

separates two regions: if the minimization is initiated in the region (A), the algorithm will<br />

ev<strong>en</strong>tually locate the minimum, but if we start in the region (B), the gradi<strong>en</strong>t <strong>de</strong>sc<strong>en</strong>t m<strong>et</strong>hod<br />

will lead us away from the global minimum and the required calibration quality will never be<br />

achieved.<br />

B<br />

15000<br />

x 10 5<br />

2<br />

1.8<br />

10000<br />

5000<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

A<br />

0.8<br />

0.25<br />

0<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

κ<br />

0<br />

min<br />

0.14<br />

0.16<br />

0.18<br />

σ<br />

0.2<br />

0.22<br />

0.24<br />

0.6<br />

8<br />

7<br />

6<br />

5<br />

4<br />

κ<br />

3<br />

2<br />

1<br />

0<br />

0.1<br />

0.15<br />

σ<br />

0.2<br />

Figure 2.3: Sum of squared differ<strong>en</strong>ces b<strong>et</strong>we<strong>en</strong> mark<strong>et</strong> prices (DAX options maturing in 10<br />

weeks) and mo<strong>de</strong>l prices in the variance gamma mo<strong>de</strong>l (1.18) as a function of σ and κ, the third<br />

param<strong>et</strong>er being fixed. Left: small region around the global minimum. Right: error surface on<br />

a larger scale.<br />

2.2 Selection of solutions using relative <strong>en</strong>tropy<br />

Wh<strong>en</strong> the option pricing constraints do not <strong>de</strong>termine the expon<strong>en</strong>tial Lévy mo<strong>de</strong>l compl<strong>et</strong>ely<br />

(this is for example the case if the number of constraints is finite), additional information may


70 CHAPTER 2. THE CALIBRATION PROBLEM<br />

be introduced into the problem by specifying a prior mo<strong>de</strong>l: we suppose giv<strong>en</strong> a Lévy process<br />

P and look for the solution of the problem (2.4) that has the smallest relative <strong>en</strong>tropy with<br />

respect to P . For two probabilities P and Q on the same measurable space (Ω, F), the relative<br />

<strong>en</strong>tropy of Q with respect to P is <strong>de</strong>fined by<br />

⎧ [ ]<br />

⎨ E P dQ dQ<br />

dP<br />

log<br />

dP<br />

, if Q ≪ P,<br />

I(Q|P ) =<br />

⎩<br />

∞, otherwise,<br />

(2.10)<br />

where by conv<strong>en</strong>tion x log x = 0 wh<strong>en</strong> x = 0. For a time horizon T ≤ T ∞ we <strong>de</strong>fine I T (Q|P ) :=<br />

I(Q| FT |P | FT ).<br />

Minimum <strong>en</strong>tropy least squares calibration problem<br />

Giv<strong>en</strong> prices C M of call options<br />

and a prior Lévy process P , find a least squares solution Q ∗ ∈ Q LS , such that<br />

I(Q ∗ |P ) =<br />

inf I(Q|P ). (2.11)<br />

Q∈QLS In the sequel, any such Q ∗ will be called a minimum <strong>en</strong>tropy least squares solution (MELSS)<br />

and the s<strong>et</strong> of all such solutions will be <strong>de</strong>noted by Q MELS .<br />

The prior probability P must reflect our a priori knowledge about the risk-neutral distribution<br />

of the un<strong>de</strong>rlying. A natural choice of prior, <strong>en</strong>suring abs<strong>en</strong>ce of arbitrage in the calibrated<br />

mo<strong>de</strong>l, is an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l, estimated from the time series of r<strong>et</strong>urns. The effect of<br />

the choice of prior on the solution of the calibration problem and the possible ways to choose<br />

it in practice are discussed in Section 3.2.<br />

Using relative <strong>en</strong>tropy for selection of solutions removes, to some ext<strong>en</strong>t, the i<strong>de</strong>ntification<br />

problem of least-squares calibration. Whereas in the least squares case, this was an important<br />

nuisance, now, if two measures reproduce mark<strong>et</strong> option prices with the same precision and have<br />

the same <strong>en</strong>tropic distance to the prior, this means that both measures are compatible with all<br />

the available information. Knowledge of many such probability measures instead of one may be<br />

se<strong>en</strong> as an advantage, because it allows to estimate mo<strong>de</strong>l risk and provi<strong>de</strong> confi<strong>de</strong>nce intervals<br />

for the prices of exotic options. However, the calibration problem (2.11) remains ill-posed: since<br />

the minimization of <strong>en</strong>tropy is done over the results of least squares calibration, problem (2.11)<br />

may only admit a solution if problem (2.4) does. Also, Q LS is not necessarily a compact s<strong>et</strong>,


2.2. SELECTION USING RELATIVE ENTROPY 71<br />

so ev<strong>en</strong> if it is nonempty, the least squares solution minimizing the relative <strong>en</strong>tropy may not<br />

exist. Other un<strong>de</strong>sirable properties like abs<strong>en</strong>ce of continuity and numerical instability are also<br />

inherited from the least squares approach. In Section 2.5 we will propose a regularized version<br />

of problem (2.11) that does not suffer from these difficulties.<br />

The choice of relative <strong>en</strong>tropy as a m<strong>et</strong>hod for selection of solutions of the calibration<br />

problem is driv<strong>en</strong> by the following consi<strong>de</strong>rations:<br />

• The relative <strong>en</strong>tropy is a conv<strong>en</strong>i<strong>en</strong>t notion of distance for probability measures. In<strong>de</strong>ed,<br />

it is convex, nonnegative functional of Q for fixed P , equal to zero if and only if dQ<br />

dP = 1<br />

P -a.s. To see this, observe that<br />

E P [ dQ<br />

dP<br />

log<br />

dQ<br />

dP<br />

]<br />

= E P [ dQ<br />

dP<br />

log<br />

dQ<br />

dP − dQ<br />

dP + 1 ]<br />

,<br />

and that z log z − z + 1 is a convex nonnegative function of z, equal to zero if and only if<br />

z = 1.<br />

• The relative <strong>en</strong>tropy of two Lévy processes is easily expressed in terms of their characteristic<br />

tripl<strong>et</strong>s (see Theorem 2.9).<br />

• Relative <strong>en</strong>tropy, also called Kullback-Leibler distance, is a well-studied object. It appears<br />

in many domains including the theory of large <strong>de</strong>viations and the information theory. It<br />

has also already be<strong>en</strong> used in finance for pricing and calibration (see Section 2.3). We can<br />

therefore use the known properties of this functional (see e.g. [93]) as a starting point of<br />

our study.<br />

The minimum <strong>en</strong>tropy least squares solution need not always exist, but if the prior is chos<strong>en</strong><br />

correctly, that is, if there exists at least one solution of problem (2.4) with finite relative <strong>en</strong>tropy<br />

with respect to the prior, th<strong>en</strong> the MELSS will also exist, as shown by the following lemma.<br />

We recall that L NA stands for the s<strong>et</strong> of Lévy processes, satisfying the no arbitrage conditions<br />

of Proposition 1.8.<br />

Lemma 2.6. L<strong>et</strong> P ∈ L NA ∩ L + B<br />

for some B > 0 and suppose that the problem (2.4) admits a<br />

solution Q + with I(Q + |P ) = C < ∞. Th<strong>en</strong> the problem (2.11) admits a solution.<br />

Proof. The proof of this lemma uses the properties of the relative <strong>en</strong>tropy functional (2.10)<br />

that will be prov<strong>en</strong> in Section 2.4 below. Un<strong>de</strong>r the condition of the lemma, it is clear that the


72 CHAPTER 2. THE CALIBRATION PROBLEM<br />

solution Q ∗ of problem (2.11), if it exists, satisfies I(Q ∗ |P ) ≤ C. This <strong>en</strong>tails that Q ∗ ≪ P ,<br />

which means by Proposition 1.5 that Q ∗ ∈ L + B . Therefore, Q∗ belongs to the s<strong>et</strong><br />

L + B ∩ {Q ∈ M ∩ L : ‖CQ − C M ‖ = ‖C Q+ − C M ‖} ∩ {Q ∈ L : I(Q|P ) ≤ C}. (2.12)<br />

Lemma 2.10 and the Prohorov’s theorem <strong>en</strong>tail that the level s<strong>et</strong> {Q ∈ L : I(Q|P ) ≤ C}<br />

is relatively weakly compact. On the other hand, by Corollary 2.1, I(Q|P ) is weakly lower<br />

semicontinuous with respect to Q for fixed P . Therefore, the s<strong>et</strong> {Q ∈ P(Ω) : I(Q|P ) ≤ C} is<br />

weakly closed and since by Lemma 2.4, M ∩ L + B is also weakly closed, the s<strong>et</strong> M ∩ L+ B ∩ {Q ∈<br />

L : I(Q|P ) ≤ C} is weakly compact. Lemma 2.2 th<strong>en</strong> implies that the s<strong>et</strong> (2.12) is also weakly<br />

compact. Since I(Q|P ) is weakly lower semicontinuous, it reaches its minimum on this s<strong>et</strong>.<br />

2.3 The use of relative <strong>en</strong>tropy for pricing and calibration: review<br />

of literature<br />

2.3.1 Pricing with minimal <strong>en</strong>tropy martingale measure<br />

It is known (see [25, Theorem 4.7] or [27, Remark 10.3]) that all expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls<br />

with the exception of the Black-Scholes mo<strong>de</strong>l and the mo<strong>de</strong>l driv<strong>en</strong> by the comp<strong>en</strong>sated Poisson<br />

process correspond to incompl<strong>et</strong>e mark<strong>et</strong>s meaning that there exists no unique equival<strong>en</strong>t<br />

martingale measure.<br />

In the incompl<strong>et</strong>e mark<strong>et</strong> s<strong>et</strong>ting many authors including [24, 41, 42, 49, 72, 73, 89] have investigated<br />

the minimal <strong>en</strong>tropy martingale measure, that is, the pricing measure that minimizes<br />

the relative <strong>en</strong>tropy with respect to the historical probability P . A probability measure Q ∗ ∈ M<br />

is called minimal <strong>en</strong>tropy martingale measure if<br />

I(Q ∗ |P ) = min I(Q|P ). (2.13)<br />

Q∈M<br />

Frittelli [42] proves the exist<strong>en</strong>ce of a minimal <strong>en</strong>tropy martingale measure provi<strong>de</strong>d that the<br />

stock price process is boun<strong>de</strong>d and that there exists a martingale measure with finite relative<br />

<strong>en</strong>tropy with respect to P .<br />

He further shows that if there exists an equival<strong>en</strong>t martingale<br />

measure with finite relative <strong>en</strong>tropy with respect to P , the MEMM is also equival<strong>en</strong>t to P .<br />

Several authors including [24, 49, 73] discuss minimal <strong>en</strong>tropy martingale measures for expon<strong>en</strong>tial<br />

Lévy mo<strong>de</strong>ls. The following result is prov<strong>en</strong> in Miyahara and Fujiwara [73].


2.3. RELATIVE ENTROPY IN THE LITERATURE 73<br />

Theorem 2.7. L<strong>et</strong> P be a Lévy process with characteristic tripl<strong>et</strong> (A, ν, γ). If there exists a<br />

constant β ∈ R such that<br />

∫<br />

{x>1}<br />

γ +<br />

e x e β(ex −1) ν(dx) < ∞, (2.14)<br />

( 1<br />

2 + β )<br />

A +<br />

∫<br />

|x|≤1<br />

{<br />

} ∫<br />

(e x − 1)e β(ex−1) − x ν(dx) + (e x − 1)e β(ex−1) ν(dx) = 0,<br />

|x|>1<br />

th<strong>en</strong> there exists a minimal <strong>en</strong>tropy martingale measure Q ∗ with the following properties:<br />

1. The measure Q ∗ corresponds to a Lévy process: Q ∗ ∈ L with characteristic tripl<strong>et</strong><br />

A ∗ = A,<br />

ν ∗ (dx) = e β(ex−1) ν(dx),<br />

∫<br />

γ ∗ = γ + βA + x(e β(ex−1) − 1)ν(dx).<br />

|x|≤1<br />

2. The measure Q ∗ is an equival<strong>en</strong>t martingale measure: Q ∗ ∼ P .<br />

3. The minimal relative <strong>en</strong>tropy is giv<strong>en</strong> by<br />

{ ∫ β ∞<br />

}<br />

I(Q ∗ |P ) = −T<br />

2 (1 + β)A + βγ + {e β(ex−1) − 1 − βx1 |x|≤1 }ν(dx) . (2.15)<br />

−∞<br />

Remark 2.1. It is easy to show, along the lines of the proof of Proposition 1.8, that condition<br />

(2.14) is satisfied, in particular, if P ∈ L NA ∩ L + B<br />

for some B > 0, which corresponds to a stock<br />

price process with jumps boun<strong>de</strong>d from above in a mark<strong>et</strong> without arbitrage opportunity.<br />

Since large positive jumps do not happ<strong>en</strong> very oft<strong>en</strong> in real mark<strong>et</strong>s, (2.14) turns out to be<br />

much less restrictive (and easier to check) than the g<strong>en</strong>eral hypotheses in [42]. This shows that<br />

the notion of MEMM is especially useful and conv<strong>en</strong>i<strong>en</strong>t in the context of expon<strong>en</strong>tial Lévy<br />

mo<strong>de</strong>ls.<br />

In addition to its computational tractability, the interest of the minimal <strong>en</strong>tropy martingale<br />

measure is due to its economic interpr<strong>et</strong>ation as the pricing measure that corresponds to the<br />

limit of utility indiffer<strong>en</strong>ce price for the expon<strong>en</strong>tial utility function wh<strong>en</strong> the risk aversion<br />

coeffici<strong>en</strong>t t<strong>en</strong>ds to zero. Consi<strong>de</strong>r an investor with initial <strong>en</strong>dowm<strong>en</strong>t c, whose utility function<br />

is giv<strong>en</strong> by<br />

U α (x) := 1 − e −αx , (2.16)


74 CHAPTER 2. THE CALIBRATION PROBLEM<br />

where α is a risk-aversion coeffici<strong>en</strong>t. Giv<strong>en</strong> some s<strong>et</strong> of admissible trading strategies Θ, the<br />

utility indiffer<strong>en</strong>ce price p α (c, H) of a claim H for this investor is <strong>de</strong>fined as the solution of the<br />

following equation:<br />

∫<br />

sup E[U α (c + p α (c, H) +<br />

θ∈Θ<br />

(0,T ]<br />

∫<br />

θ u dS u − H)] = sup E[U α (c + θ u dS u )].<br />

θ∈Θ<br />

(0,T ]<br />

Due to the special form (2.16) of the utility function, the initial <strong>en</strong>dowm<strong>en</strong>t c cancels out of the<br />

above equation and we see that<br />

p α (c, H) = p α (0, H) := p α (H).<br />

Using the results in [33], Miyahara and Fujiwara [73] established the following properties of<br />

utility indiffer<strong>en</strong>ce price in expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls. Similar results have be<strong>en</strong> obtained by El<br />

Karoui and Rouge [39] in the s<strong>et</strong>ting of continuous processes.<br />

Proposition 2.8. L<strong>et</strong> (X, P ) be a Lévy process such that P ∈ L + B ∩ L NA, and l<strong>et</strong> Q ∗ be the<br />

MEMM <strong>de</strong>fined by (2.13). L<strong>et</strong> S t := e Xt and l<strong>et</strong> Θ inclu<strong>de</strong> all predictable S-integrable processes<br />

θ such that ∫ (0,t] θ udS u is a martingale for each local martingale measure Q, with I(Q|P ) < ∞.<br />

Th<strong>en</strong> the corresponding utility indiffer<strong>en</strong>ce price p α (H) of a boun<strong>de</strong>d claim H has the following<br />

properties:<br />

1. p α (H) ≥ E Q∗ [H] for any α > 0.<br />

2. If 0 < α < β th<strong>en</strong> p α (H) ≤ p β (H).<br />

3. lim α↓0 p α (H) = E Q∗ [H].<br />

The price of a claim H computed un<strong>de</strong>r the MEMM thus turns out to be the highest price<br />

at which all investors with expon<strong>en</strong>tial utility function will be willing to buy this claim.<br />

Kalls<strong>en</strong> [58] <strong>de</strong>fines neutral <strong>de</strong>rivative prices for which the optimal trading strategy consists<br />

in having no conting<strong>en</strong>t claim in one’s portfolio. This approach to valuation in incompl<strong>et</strong>e<br />

mark<strong>et</strong>s corresponds to the notion of fair price introduced by Davis [31]. Kalls<strong>en</strong> further shows<br />

that such prices are unique and correspond to a linear arbitrage-free pricing system, <strong>de</strong>fined<br />

by an equival<strong>en</strong>t martingale measure (neutral pricing measure). If the utility function of an<br />

investor has the form (2.16), the neutral pricing measure coinci<strong>de</strong>s with the minimal <strong>en</strong>tropy


2.3. RELATIVE ENTROPY IN THE LITERATURE 75<br />

martingale measure. The neutral pricing measure Q ∗ also corresponds to the least favorable<br />

mark<strong>et</strong> compl<strong>et</strong>ion from the point of view of the investor in the following s<strong>en</strong>se. L<strong>et</strong><br />

V (c, Q) := sup{E[U(c + X − E Q [X])] : X is F T -measurable}<br />

be the maximum possible expected utility that an investor with initial <strong>en</strong>dowm<strong>en</strong>t c and utility<br />

function U can g<strong>et</strong> by trading in a mark<strong>et</strong> where the price of every conting<strong>en</strong>t claim X is equal<br />

to E Q [X]. Th<strong>en</strong><br />

V (c, Q ∗ ) =<br />

inf V (c, Q),<br />

Q∈EMM(P )<br />

that is, the neutral pricing measure minimizes the maximum possible expected utility that an<br />

investor can g<strong>et</strong> in a compl<strong>et</strong>ed mark<strong>et</strong>, over all possible arbitrage-free compl<strong>et</strong>ions. The notion<br />

of neutral pricing measure thus coinci<strong>de</strong>s with the minimax measures studied in [16], [49] and<br />

other papers.<br />

2.3.2 Calibration via relative <strong>en</strong>tropy minimization<br />

Despite its analytic tractability and interesting economic interpr<strong>et</strong>ation, the MEMM has an important<br />

drawback which makes it impossible to use for pricing in real mark<strong>et</strong>s: it does not take<br />

into account the information obtained from prices of tra<strong>de</strong>d options. To tackle this problem, Goll<br />

and Rüsch<strong>en</strong>dorf [48] introduced the notion of minimal distance martingale measure consist<strong>en</strong>t<br />

with observed mark<strong>et</strong> prices. In particular, giv<strong>en</strong> prices of call options {C M (T i , K i )} N i=1 , a probability<br />

measure Q ∗ ∈ M is called consist<strong>en</strong>t minimal <strong>en</strong>tropy martingale measure (CMEMM)<br />

if<br />

I(Q ∗ |P ) = min I(Q|P ),<br />

Q<br />

where the minimum is tak<strong>en</strong> over all martingale measures Q such that C M (T i , K i ) = C Q (T i , K i )<br />

for i = 1, . . . , N.<br />

Kalls<strong>en</strong> [58] shows that consist<strong>en</strong>t minimal distance martingale measures correspond to<br />

constant <strong>de</strong>mand <strong>de</strong>rivative pricing (wh<strong>en</strong> the optimal portfolio must contain not zero but<br />

a fixed amount of each tra<strong>de</strong>d <strong>de</strong>rivative), thus providing an economic rationale for distance<br />

minimization un<strong>de</strong>r constraints, and, in particular, for <strong>en</strong>tropy minimization.<br />

A drawback of CMEMM is that it may be very difficult to compute numerically. In particular,<br />

if X is a Lévy process un<strong>de</strong>r the historical measure P , it will in g<strong>en</strong>eral no longer be


76 CHAPTER 2. THE CALIBRATION PROBLEM<br />

a Lévy process un<strong>de</strong>r a consist<strong>en</strong>t minimal <strong>en</strong>tropy martingale measure and no analytic results<br />

similar to Theorem 2.7 are available for the constrained case. For example, l<strong>et</strong> {X t } t≥0 be a<br />

real-valued Lévy process on (A, ν, γ) such that for every t,<br />

X t = N ′ t − N ′′<br />

t , (2.17)<br />

where N ′ and N ′′ are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt Poisson processes with int<strong>en</strong>sity 1 un<strong>de</strong>r P . It follows from<br />

Proposition 1.5 that the s<strong>et</strong> of Lévy processes equival<strong>en</strong>t to P and satisfying the martingale<br />

condition (1.2) contains all processes un<strong>de</strong>r which X has the form (2.17) where N ′ and N ′′ are<br />

still in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and have respective int<strong>en</strong>sities λ ′ = λ and λ ′′ = eλ for some λ > 0. The s<strong>et</strong> of<br />

all equival<strong>en</strong>t martingale measures un<strong>de</strong>r which X remains a Lévy process is thus param<strong>et</strong>rized<br />

by one param<strong>et</strong>er λ.<br />

If one allows X to be an additive process (process with in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt<br />

increm<strong>en</strong>ts, continuous in probability) un<strong>de</strong>r the new measure, the class of equival<strong>en</strong>t martingale<br />

measures is much larger: it follows from Theorem IV.4.32 in [54] that N ′ and N ′′ can now<br />

have variable (but <strong>de</strong>terministic) int<strong>en</strong>sities λ ′ (t) = λ(t) and λ ′′ (t) = eλ(t) for some function<br />

λ : [0, T ∞ ] → (0, ∞). It is clear that many mark<strong>et</strong> data s<strong>et</strong>s (involving several maturities) can<br />

be reproduced by an equival<strong>en</strong>t martingale measure, un<strong>de</strong>r which X is an additive process but<br />

not by a martingale measure un<strong>de</strong>r which X is a Lévy process, which implies that un<strong>de</strong>r the<br />

consist<strong>en</strong>t minimal <strong>en</strong>tropy martingale measure X will not be a Lévy process.<br />

Stutzer [91] suggests a three-step algorithm for numerical evaluation of <strong>de</strong>rivative prices<br />

un<strong>de</strong>r the CMEMM in a mo<strong>de</strong>l with a single time horizon T . This m<strong>et</strong>hod allows to compute<br />

prices of European options with maturity T , consist<strong>en</strong>t with prices of mark<strong>et</strong>-quoted European<br />

options with the same maturity. First, possible ass<strong>et</strong> price values at time T , {P h } N h=1<br />

and the<br />

corresponding probabilities ˆπ(h) are estimated nonparam<strong>et</strong>rically using a histogram estimator<br />

from the historical r<strong>et</strong>urn values. Second, one needs to find the probabilities π ∗ (h), that satisfy<br />

the martingale constraint and the pricing constraints and have the smallest relative <strong>en</strong>tropy with<br />

respect to ˆπ. In this simplified one period mo<strong>de</strong>l the martingale condition reduces to a single<br />

constraint that the discounted expectation of stock price un<strong>de</strong>r π ∗ be equal to its pres<strong>en</strong>t value.<br />

The European options expiring at T can th<strong>en</strong> be priced using these martingale probabilities π ∗ .<br />

In this paper, Stutzer suggests an information theor<strong>et</strong>ic rationale for using the relative<br />

<strong>en</strong>tropy (also called Kullback-Leibler information criterion) for pricing and calibration. From<br />

a Bayesian point of view, the historical prices constitute a prior information about the future


2.3. RELATIVE ENTROPY IN THE LITERATURE 77<br />

risk-neutral distribution of ass<strong>et</strong> prices.<br />

This prior must be updated to take into account<br />

the martingale constraint and the observed option prices, and it is natural to <strong>de</strong>mand that the<br />

updated distribution incorporate no additional information other than the martingale constraint<br />

and the pricing constraints. One must therefore minimize some quantitative measure of relative<br />

information of the two distributions, and it has be<strong>en</strong> shown (see [52]) that every measure of<br />

information satisfying a s<strong>et</strong> of natural axioms must be proportional to Kullback-Leibler relative<br />

<strong>en</strong>tropy.<br />

Because it only allows to reconstruct the ass<strong>et</strong> price distribution for one specified time<br />

horizon, and only takes into account the prices of options that expire at this horizon, Stutzer’s<br />

m<strong>et</strong>hod does not provi<strong>de</strong> any information about the risk-neutral process and thus cannot be<br />

used to price any <strong>de</strong>rivative that <strong>de</strong>p<strong>en</strong>ds on the stock price at times other than T . Avellaneda<br />

<strong>et</strong> al. [3] pursue the same logic further and propose a m<strong>et</strong>hod allowing to construct a discr<strong>et</strong>e<br />

approximation of the law of the process un<strong>de</strong>rlying the observed option prices. They consi<strong>de</strong>r<br />

N < ∞ fixed trajectories {X 1 , . . . , X N } that the price process can take, simulated beforehand<br />

from a prior mo<strong>de</strong>l. The new state space Ω ′ thus contains a finite number of elem<strong>en</strong>ts: Ω ′ =<br />

{X 1 , . . . , X N }. The new prior P ′ is the uniform law on Ω ′ : P ′ (X i ) = 1 N<br />

and the paper<br />

suggests to calibrate the weights (probabilities) of these trajectories q i := Q ′ (X i ) to reproduce<br />

mark<strong>et</strong>-quoted option prices correctly. Minimizing relative <strong>en</strong>tropy I(Q ′ |P ′ ) is th<strong>en</strong> equival<strong>en</strong>t<br />

to maximizing the <strong>en</strong>tropy of Q ′ and the calibration problem becomes:<br />

maximize −<br />

N∑<br />

q i log q i un<strong>de</strong>r constraints E Q′ [H j ] = C j , j = 1, . . . , M,<br />

i=1<br />

where H j are terminal payoffs and C j the observed mark<strong>et</strong> prices of M tra<strong>de</strong>d options. D<strong>en</strong>oting<br />

by g ij the payoff of j-th option on the i-th trajectory, and introducing Lagrange multipliers<br />

λ 1 , . . . , λ M , Avellaneda <strong>et</strong> al. reformulate the calibration problem as a minimax problem:<br />

⎧<br />

⎨ N (<br />

min max<br />

λ q ⎩ − ∑<br />

M∑ N<br />

) ⎫ ∑<br />

⎬<br />

q i log q i + λ j q i g ij − C j<br />

⎭ .<br />

i=1<br />

The inner maximum can be computed analytically and, since it is tak<strong>en</strong> over linear functions<br />

of λ 1 , . . . , λ M , yields a convex function of Lagrange multipliers. The outer minimum can th<strong>en</strong><br />

j=1<br />

be evaluated numerically using a gradi<strong>en</strong>t <strong>de</strong>sc<strong>en</strong>t m<strong>et</strong>hod.<br />

This technique (weighted Monte Carlo) is attractive for numerical computations but has<br />

a number of drawbacks from the theor<strong>et</strong>ical viewpoint.<br />

i=1<br />

First, the result of calibration is a


78 CHAPTER 2. THE CALIBRATION PROBLEM<br />

probability measure on a finite s<strong>et</strong> of paths chos<strong>en</strong> beforehand; it does not allow to reconstruct<br />

a process on the original space Ω. Second, the martingale condition is not imposed in this<br />

approach (because this would correspond to an infinite number of constraints). As a result,<br />

<strong>de</strong>rivative prices computed with the weighted Monte Carlo algorithm may contain arbitrage<br />

opportunities, especially wh<strong>en</strong> applied to forward start contracts.<br />

Nguy<strong>en</strong> [77] studies the converg<strong>en</strong>ce of the above m<strong>et</strong>hod by Avellaneda <strong>et</strong> al. wh<strong>en</strong> the<br />

number of paths t<strong>en</strong>ds to infinity as well as the stability of the calibration procedure, minimization<br />

criteria other than relative <strong>en</strong>tropy, possible ways to impose the (approximate) martingale<br />

condition and many other issues related to this calibration m<strong>et</strong>hodology. Statistical properties<br />

of weighted Monte Carlo estimators are also studied in [46].<br />

Chapter 8 of [77] introduces an interesting theor<strong>et</strong>ical approach to minimal <strong>en</strong>tropy calibration,<br />

that is related to the pres<strong>en</strong>t work and is worth being discussed in more <strong>de</strong>tail. Starting<br />

with a very g<strong>en</strong>eral prior jump-diffusion mo<strong>de</strong>l P for the stock price S t of the form<br />

∫<br />

dS t = S t− [b(t, S t− )dt + σ(t, S t− )dW t + Φ(t, S t− , z)(Π(dt, dz) − π(dt, dz))], (2.18)<br />

z∈R<br />

where Π is a homog<strong>en</strong>eous Poisson random measure with int<strong>en</strong>sity measure π(dt, dz) = dt ×<br />

ρ(dz), and the coeffici<strong>en</strong>ts satisfy some regularity hypotheses not listed here, Nguy<strong>en</strong> suggests<br />

to find a martingale measure Q ∗ which reproduces the observed option prices correctly and has<br />

the smallest relative <strong>en</strong>tropy with respect to P in a subclass M ′ of all martingale measures on<br />

(Ω, F). This subclass M ′ contains all martingale measures Q K un<strong>de</strong>r which S t has the same<br />

volatility σ and the comp<strong>en</strong>sator of Π is giv<strong>en</strong> by K(t, X t− , z)π(dt, dz) with<br />

K ∈ K H := {K(t, x, z) : [0, T ] × R × R → (0, ∞) Borel with | log K| ≤ log(H)φ 0 }<br />

for some φ 0 ∈ L 2 (ρ) positive with ‖φ 0 ‖ ∞ ≤ 1. In other words, the object of calibration here is<br />

the int<strong>en</strong>sity of jumps of a Markovian jump diffusion.<br />

The fact that the s<strong>et</strong><br />

M ′ := { Q K : K ∈ K H}<br />

is convex <strong>en</strong>ables Nguy<strong>en</strong> to use classical convex analysis m<strong>et</strong>hods to study the calibration<br />

problem. In particular, he proves the exist<strong>en</strong>ce and uniqu<strong>en</strong>ess of the solution of the (p<strong>en</strong>alized)<br />

calibration problem un<strong>de</strong>r the condition that the coeffici<strong>en</strong>ts appearing in (2.18) are suffici<strong>en</strong>tly<br />

regular and the drift b and option price constraint C belong to some neighborhood of (r, C r ),


2.3. RELATIVE ENTROPY IN THE LITERATURE 79<br />

where r is the interest rate and C r <strong>de</strong>notes the option prices computed for S t giv<strong>en</strong> by (2.18)<br />

with b = r. Unfortunately, no insight is giv<strong>en</strong> as to the size of this neighborhood.<br />

An important drawback of the calibration m<strong>et</strong>hodology advocated by Nguy<strong>en</strong> [77] is the<br />

difficulty of numerical implem<strong>en</strong>tation. The solution is obtained by minimizing a certain function<br />

of Lagrange multipliers (the value function), which is itself a solution of nonlinear partial<br />

integro-differ<strong>en</strong>tial equation of HJB type, and the author gives no indication about the numerical<br />

m<strong>et</strong>hods that can be used. In addition, since a finite s<strong>et</strong> of option prices does not typically<br />

contain <strong>en</strong>ough information to i<strong>de</strong>ntify a three-dim<strong>en</strong>sional int<strong>en</strong>sity function K(t, x, z), the<br />

choice of the prior process, not discussed in [77], will play a crucial role.<br />

2.3.3 The pres<strong>en</strong>t study in the context of previous work on this subject<br />

In this thesis, we suppose that the prior probability P is a Lévy process and restrict the class<br />

of martingale measures Q that we consi<strong>de</strong>r as possible solutions of the calibration problem to<br />

those un<strong>de</strong>r which (X, Q) remains a Lévy process. Since this class is rather narrow, it may not<br />

contain a measure that reproduces the observed option prices exactly. Therefore, in problem<br />

(2.11) we only require that these prices be reproduced in the least squares s<strong>en</strong>se.<br />

Restricting the calibration to the class of Lévy processes <strong>en</strong>ables us to compute the solution<br />

of the calibration problem numerically using a relatively simple algorithm (see Chapter 3). Our<br />

m<strong>et</strong>hod can thus be se<strong>en</strong> as a computable approximation of the consist<strong>en</strong>t minimal <strong>en</strong>tropy<br />

martingale measure, discussed by Kalls<strong>en</strong> [58] and Goll and Rüsch<strong>en</strong>dorf [48]. However, contrary<br />

to the approach by Avellaneda <strong>et</strong> al. [3], we do not discr<strong>et</strong>ize the space of sample paths:<br />

our m<strong>et</strong>hod yields a Lévy process on the initial state space Ω and not on the finite space of<br />

trajectories simulated beforehand. The solution of calibration problem (2.11) and its regularized<br />

version (2.27) can therefore be used for all types of computations with arbitrary precision.<br />

Another advantage of restricting the class of martingale measures is that using a finite number<br />

of option prices it is easier to calibrate a one-dim<strong>en</strong>sional object (the Lévy measure) than<br />

a three-dim<strong>en</strong>sional object (the int<strong>en</strong>sity function as in [77, Chapter 8]). Therefore, in our<br />

approach the influ<strong>en</strong>ce of the prior on the solution is less important (see Chapter 3).<br />

Unlike the class of all Markov processes and unlike the class of probabilities consi<strong>de</strong>red in<br />

[77, Chapter 8], the class of all probabilities Q, un<strong>de</strong>r which (X, Q) remains a Lévy process is


80 CHAPTER 2. THE CALIBRATION PROBLEM<br />

not convex. We are therefore unable to apply the m<strong>et</strong>hods of convex analysis to our formulation<br />

of the calibration problem and are limited to standard functional analysis tools. In particular,<br />

in our s<strong>et</strong>ting we cannot hope to obtain a uniqu<strong>en</strong>ess result for the solution of the calibration<br />

problem, but, as m<strong>en</strong>tioned in Section 2.2, uniqu<strong>en</strong>ess is not always a <strong>de</strong>sirable property from<br />

the financial viewpoint.<br />

2.4 Relative <strong>en</strong>tropy of Lévy processes<br />

In this subsection we explicitly compute the relative <strong>en</strong>tropy of two Lévy processes in terms of<br />

their characteristic tripl<strong>et</strong>s. Un<strong>de</strong>r additional assumptions this result was shown in [24] (where<br />

it is supposed that Q is equival<strong>en</strong>t to P and the Lévy process has finite expon<strong>en</strong>tial mom<strong>en</strong>ts<br />

un<strong>de</strong>r P ) and in [77] (where log dνQ<br />

dν P<br />

is supposed boun<strong>de</strong>d from above and below). We consi<strong>de</strong>r<br />

it important to give an elem<strong>en</strong>tary proof valid for all Lévy processes.<br />

Theorem 2.9 (Relative <strong>en</strong>tropy of Lévy processes). L<strong>et</strong> {X t } t≥0 be a real-valued Lévy<br />

process on (Ω, F, Q) and on (Ω, F, P ) with respective characteristic tripl<strong>et</strong>s (A Q , ν Q , γ Q ) and<br />

(A P , ν P , γ P ). Suppose that the conditions 1–5 of Proposition 1.5 are satisfied and <strong>de</strong>note A :=<br />

A Q = A P . Th<strong>en</strong> for every time horizon T ≤ T ∞ the relative <strong>en</strong>tropy of Q| FT<br />

P | FT<br />

can be computed as follows:<br />

with respect to<br />

I T (Q|P ) = I(Q| FT |P | FT ) = T { ∫ 1<br />

γ Q − γ P −<br />

2A<br />

T<br />

−1<br />

∫ ∞<br />

−∞<br />

x(ν Q − ν P )(dx)} 2<br />

1 A≠0 +<br />

( dν<br />

Q<br />

)<br />

dνQ<br />

log<br />

dνP dν P + 1 − dνQ<br />

dν P ν P (dx). (2.19)<br />

Proof. L<strong>et</strong> {X c t } t≥0 be the continuous martingale part of X un<strong>de</strong>r P (a Brownian motion), µ<br />

be the jump measure of X and φ := dνQ<br />

dν P . From [54, Theorem III.5.19], the <strong>de</strong>nsity process<br />

Z t := dQ| F t<br />

dP | Ft<br />

is the Doléans-Da<strong>de</strong> expon<strong>en</strong>tial of the Lévy process {N t } t≥0 <strong>de</strong>fined by<br />

∫<br />

N t := βXt c + (φ(x) − 1){µ(ds × dx) − ds ν P (dx)},<br />

[0,t]×R<br />

where β is giv<strong>en</strong> by<br />

⎧<br />

⎨ 1<br />

A<br />

β =<br />

{γQ − γ P − ∫ |x|≤1 x(φ(x) − 1)νP (dx)} if A > 0,<br />

⎩<br />

0 otherwise.


2.4. RELATIVE ENTROPY OF LEVY PROCESSES 81<br />

Choose 0 < ε < 1 and l<strong>et</strong> I := {x : ε ≤ φ(x) ≤ ε −1 }.<br />

martingales:<br />

∫<br />

N t ′ := βXt c +<br />

[0,t]×I<br />

(φ(x) − 1){µ(ds × dx) − ds ν P (dx)}<br />

∫<br />

N t ′′ := (φ(x) − 1){µ(ds × dx) − ds ν P (dx)}.<br />

[0,t]×(R\I)<br />

We split N t into two in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt<br />

Since N ′ and N ′′ never jump tog<strong>et</strong>her, [N ′ , N ′′ ] t = 0 and E(N ′ + N ′′ ) t = E(N 1 ) t E(N 2 ) t (cf.<br />

Equation II.8.19 in [54]). Moreover, since N ′ and N ′′ are Lévy processes and martingales, their<br />

stochastic expon<strong>en</strong>tials are also martingales (Proposition 1.4). Therefore,<br />

I T (Q|P ) = E P [Z T log Z T ]<br />

if these expectations exist.<br />

and<br />

= E P [E(N ′ ) T E(N ′′ ) T log E(N ′ ) T ] + E P [E(N ′ ) T E(N ′′ ) T log E(N ′′ ) T ]<br />

= E P [E(N ′ ) T log E(N ′ ) T ] + E P [E(N ′′ ) T log E(N ′′ ) T ] (2.20)<br />

Since ∆N ′ t > −1 a.s., E(N ′ ) t is almost surely positive. Therefore, from Proposition 1.3,<br />

U t := log E(N ′ ) t is a Lévy process with characteristic tripl<strong>et</strong>:<br />

This implies that e Ut<br />

A U = β 2 A,<br />

ν U (B) = ν P (I ∩ {x : log φ(x) ∈ B}) ∀B ∈ B(R),<br />

∫<br />

γ U = − β2 A ∞<br />

2 − (e x − 1 − x1 |x|≤1 )ν U (dx).<br />

−∞<br />

is a martingale and that U t has boun<strong>de</strong>d jumps and all expon<strong>en</strong>tial<br />

mom<strong>en</strong>ts. Therefore, E[U T e U T<br />

] < ∞ and can be computed as follows:<br />

E P [U T e U T<br />

] = −i d dz EP [e izU T<br />

]| z=−i = −iT ψ ′ (−i)E P [e U T<br />

] = −iT ψ ′ (−i)<br />

= T (A U + γ U +<br />

= β2 AT<br />

2<br />

I<br />

∫ ∞<br />

−∞<br />

(xe x − x1 |x|≤1 )ν U (dx))<br />

∫<br />

+ T (φ(x) log φ(x) + 1 − φ(x))ν P (dx) (2.21)<br />

It remains to compute E P [E(N ′′ ) T log E(N ′′ ) T ]. Since N ′′ is a compound Poisson process,<br />

E(N ′′ ) t = e bt ∏ s≤t<br />

(1 + ∆N<br />

′′<br />

s ), where b = ∫ R\I (1 − φ(x))νP (dx). L<strong>et</strong> ν ′′ be the Lévy measure of<br />

N ′′ and λ its jump int<strong>en</strong>sity. Th<strong>en</strong><br />

E(N ′′ ) T log E(N ′′ ) T = bT E(N ′′ ) T + e bT ∏ s≤T<br />

(1 + ∆N s ′′ ) ∑ log(1 + ∆N s ′′ )<br />

s≤T


82 CHAPTER 2. THE CALIBRATION PROBLEM<br />

and<br />

E P [E(N ′′ ) T log E(N ′′ ) T ] = bT + e bT<br />

∞ ∑<br />

k=0<br />

e<br />

−λT (λT )k<br />

k!<br />

E[ ∏ s≤T<br />

(1 + ∆N s ′′ ) ∑ log(1 + ∆N s ′′ )|k jumps]<br />

s≤T<br />

Since, un<strong>de</strong>r the condition that N ′′ jumps exactly k times in the interval [0, T ], the jump sizes<br />

are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and i<strong>de</strong>ntically distributed, we find, <strong>de</strong>noting the g<strong>en</strong>eric jump size by ∆N ′′ :<br />

E P [E(N ′′ ) T log E(N ′′ ) T ]<br />

∑<br />

∞<br />

= bT + e bT e<br />

k=0<br />

−λT (λT )k<br />

= bT + λT E[(1 + ∆N ′′ ) log(1 + ∆N ′′ )]<br />

∫ ∞<br />

= bT + T (1 + x) log(1 + x)ν ′′ (dx)<br />

−∞<br />

∫<br />

= T (φ(x) log φ(x) + 1 − φ(x))ν P (dx).<br />

R\I<br />

k!<br />

kE[1 + ∆N ′′ ] k−1 E[(1 + ∆N ′′ ) log(1 + ∆N ′′ )]<br />

In particular, E P [E(N ′′ ) T log E(N ′′ ) T ] is finite if and only if the integral in the last line is finite.<br />

Combining the above expression with (2.21) and (2.20) finishes the proof.<br />

2.4.1 Properties of the relative <strong>en</strong>tropy functional<br />

Lemma 2.10. L<strong>et</strong> P, {P n } n≥1 ⊂ L + B for some B > 0, such that P n ⇒ P . Th<strong>en</strong> for every r > 0,<br />

the level s<strong>et</strong> L r := {Q ∈ L : I(Q|P n ) ≤ r for some n} is tight.<br />

Proof. For any Q ∈ L r , P Q <strong>de</strong>notes any elem<strong>en</strong>t of {P n } n≥1 , for which I(Q|P Q ) ≤ r.<br />

characteristic tripl<strong>et</strong> of Q is <strong>de</strong>noted by (A Q , ν Q , γ Q ) and that of P Q by (A P Q, ν P Q, γ P Q). In<br />

addition, we <strong>de</strong>fine φ Q := dνQ . From Theorem 2.9,<br />

dν P Q<br />

∫ ∞<br />

−∞<br />

(φ Q (x) log φ Q (x) + 1 − φ Q (x))ν P Q<br />

(dx) ≤ r/T ∞ .<br />

The<br />

Therefore, for u suffici<strong>en</strong>tly large,<br />

∫<br />

{φ Q >u}<br />

∫<br />

φ Q ν P Q<br />

(dx) ≤<br />

{φ Q >u}<br />

2φ Q [φ Q log φ Q + 1 − φ Q ]ν P Q(dx)<br />

φ Q log φ Q<br />

≤<br />

2r<br />

T ∞ log u ,<br />

which <strong>en</strong>tails that for u suffici<strong>en</strong>tly large,<br />

∫<br />

{φ Q >u}<br />

ν Q (dx) ≤<br />

2r<br />

T ∞ log u


2.4. RELATIVE ENTROPY OF LEVY PROCESSES 83<br />

uniformly with respect to Q ∈ L r . L<strong>et</strong> ε > 0 and choose u such that ∫ {φ Q >u} νQ (dx) ≤ ε/2 for<br />

every Q ∈ L r . By Proposition 1.7,<br />

∫ ∞<br />

−∞<br />

f(x)ν Pn (dx) →<br />

∫ ∞<br />

−∞<br />

f(x)ν P (dx)<br />

for every continuous boun<strong>de</strong>d function f that is i<strong>de</strong>ntically zero on a neighborhood of zero.<br />

Since the measures ν P and ν Pn<br />

for all n ≥ 1 are finite outsi<strong>de</strong> a neighborhood of zero, we can<br />

choose a compact K such that ν Pn (R \ K) ≤ ε/2u for every n. Th<strong>en</strong><br />

∫<br />

ν Q (R \ K) =<br />

(R\K)∩{φ Q ≤u}<br />

which proves property 1 of Proposition 1.6.<br />

∫<br />

φ Q ν P Q<br />

(dx) +<br />

ν Q (dx) ≤ ε,<br />

(R\K)∩{φ Q >u}<br />

It is easy to check by computing <strong>de</strong>rivatives that for every u > 0, on the s<strong>et</strong> {x : φ Q (x) ≤ u},<br />

(φ Q − 1) 2 ≤ 2u(φ Q log φ Q + 1 − φ Q ).<br />

Therefore, for u suffici<strong>en</strong>tly large and for all Q ∈ L r ,<br />

∫<br />

∣<br />

|x|≤1<br />

x(φ Q − 1)ν P Q (dx) ∣<br />

∫<br />

∣ ∣∣∣∣ ≤<br />

x(φ Q − 1)ν P Q (dx)<br />

∣ |x|≤1, φ Q ≤u<br />

∣<br />

∫|x|≤1, + x(φ Q − 1)ν P Q (dx)<br />

φ Q >u<br />

∣<br />

∫<br />

∫<br />

∫<br />

≤ x 2 ν P Q<br />

(dx) +<br />

(φ Q − 1) 2 ν P Q<br />

(dx) + 2 φ Q ν P Q<br />

(dx)<br />

∫<br />

≤<br />

∫<br />

≤<br />

|x|≤1<br />

|x|≤1<br />

|x|≤1<br />

x 2 ν P Q<br />

(dx) + 2u<br />

|x|≤1, φ Q ≤u<br />

∫ ∞<br />

−∞<br />

φ Q >u<br />

(φ Q log φ Q + 1 − φ Q )ν P Q<br />

(dx) +<br />

4r<br />

T ∞ log u<br />

x 2 ν P Q<br />

(dx) + 3ru<br />

T ∞<br />

. (2.22)<br />

By Proposition 1.6, applied to the sequ<strong>en</strong>ce {P n } n≥1 ,<br />

∫<br />

A Pn + x 2 ν Pn (dx) (2.23)<br />

|x|≤1<br />

is boun<strong>de</strong>d uniformly on n, which implies that the right hand si<strong>de</strong> of (2.22) is boun<strong>de</strong>d uniformly<br />

with respect to Q ∈ L r . From Proposition 1.5, A Q = A P Q<br />

for all Q ∈ L r because for the relative<br />

<strong>en</strong>tropy to be finite, necessarily Q ≪ P Q . From Theorem 2.9 and Proposition 1.5 it follows that<br />

{ ∫ 1<br />

} 2<br />

γ Q − γ P − x(ν Q − ν P )(dx) ≤ 2AP Qr<br />

.<br />

−1<br />

T ∞


84 CHAPTER 2. THE CALIBRATION PROBLEM<br />

From (2.23), A Pn<br />

is boun<strong>de</strong>d uniformly on n. Therefore, inequality (2.22) shows that |γ Q | is<br />

boun<strong>de</strong>d uniformly with respect to Q, which proves property 2 of Proposition 1.6.<br />

Once again, for u suffici<strong>en</strong>tly large,<br />

A Q +<br />

∫ ∞<br />

−∞<br />

∫<br />

(x 2 ∧ 1)φ Q ν P Q<br />

(dx) ≤ A Q + u (x 2 ∧ 1)ν P Q<br />

(dx)<br />

∫<br />

φ Q ≤u<br />

∫ ∞<br />

+ φ Q ν P Q<br />

(dx) ≤ A P Q<br />

+ u<br />

φ Q >u<br />

−∞<br />

(x 2 ∧ 1)ν P Q<br />

(dx) +<br />

2r<br />

T ∞ log u<br />

and (2.23) implies that the right hand si<strong>de</strong> is boun<strong>de</strong>d uniformly with respect to Q ∈ L r .<br />

Therefore, property 3 of Proposition 1.6 also holds and the proof is compl<strong>et</strong>ed.<br />

Lemma 2.11. L<strong>et</strong> Q and P be two probability measures on (Ω, F). Th<strong>en</strong><br />

I(Q|P ) =<br />

where C b (Ω) is space of boun<strong>de</strong>d continuous functions on Ω.<br />

{∫ ∫<br />

}<br />

sup fdQ − (e f − 1)dP , (2.24)<br />

f∈C b (Ω) Ω<br />

Ω<br />

Proof. Observe that<br />

⎧<br />

⎨ x log x + 1 − x, x > 0,<br />

φ(x) =<br />

⎩<br />

∞, x ≤ 0<br />

and φ ∗ (y) = e y − 1 are proper convex functions on R, conjugate to each other and apply<br />

Corollary 2 to Theorem 4 in [83].<br />

Corollary 2.1. The relative <strong>en</strong>tropy functional I(Q|P ) is weakly lower semicontinuous with<br />

respect to Q for fixed P .<br />

Lemma 2.12. L<strong>et</strong> P, {P n } n≥1 ⊂ L NA ∩ L + B for some B > 0 such that P n ⇒ P . There exists a<br />

sequ<strong>en</strong>ce {Q n } n≥1 ⊂ M ∩ L + B and a constant C < ∞ such that I(Q n|P n ) ≤ C for every n.<br />

Proof. For every n ≥ 1, by Remark 2.1, Theorem 2.7 can be applied to P n . L<strong>et</strong> Q n be the<br />

minimal <strong>en</strong>tropy martingale measure and β n be the corresponding constant, <strong>de</strong>fined in (2.14).<br />

We must show that the minimal relative <strong>en</strong>tropy, giv<strong>en</strong> by Equation (2.15), is boun<strong>de</strong>d uniformly<br />

on n.<br />

First, l<strong>et</strong> us show that the sequ<strong>en</strong>ce {β n } n≥1 is boun<strong>de</strong>d. L<strong>et</strong> h be a continuous boun<strong>de</strong>d<br />

truncation function, satisfying h(x) = x in a neighborhood of zero and for any Lévy process Q


2.5. REGULARIZING THE CALIBRATION PROBLEM 85<br />

with characteristic tripl<strong>et</strong> (A, ν, γ h ) with respect to the truncation function h, <strong>de</strong>fine<br />

( ) ∫ 1 ∞<br />

f(β, Q) := γ h +<br />

2 + β A +<br />

−∞<br />

( ) ( 1<br />

= γ h +<br />

2 + β A +<br />

∫ ∞<br />

+<br />

−∞<br />

∫ ∞<br />

−∞<br />

{<br />

(e x − 1)e β(ex −1) − h(x) −<br />

{<br />

}<br />

(e x − 1)e β(ex−1) − h(x) ν(dx)<br />

)<br />

h 2 (x)ν(dx)<br />

( ) }<br />

1<br />

2 + β h 2 (x) ν(dx). (2.25)<br />

Since (e x − 1)e β(ex −1) − x − ( 1<br />

2 + β) x 2 = o(|x| 2 ) and the integrand in the last term of (2.25) is<br />

boun<strong>de</strong>d on (−∞, B], by Proposition 1.7, for every β, lim n f(β, P n ) = f(β, P ).<br />

The support of ν Pn<br />

is boun<strong>de</strong>d from above by B, and the dominated converg<strong>en</strong>ce theorem<br />

allows to compute the <strong>de</strong>rivative of f(β, P n ) by interchanging the <strong>de</strong>rivative and the integral:<br />

f ′ β (β, P n) = A Pn +<br />

∫ ∞<br />

−∞<br />

(e x − 1) 2 e β(ex −1) ν Pn (dx) > 0.<br />

Therefore, β n is the unique solution of f(β, P n ) = 0. L<strong>et</strong> β ∗ be the solution of f(β, P ) = 0. The<br />

support of ν P is also boun<strong>de</strong>d from above by B and f ′ β (β∗ , P ) > 0. This means that there exist<br />

ε > 0 and finite constants β − < β ∗ and β + > β ∗ such that f(β − , P ) < −ε and f(β + , P ) > ε.<br />

One can th<strong>en</strong> find N such that for all n ≥ N, f(β − , P n ) < −ε/2 and f(β + , P n ) > ε/2, which<br />

means that β n ∈ [β − , β + ] and the sequ<strong>en</strong>ce {β n } is boun<strong>de</strong>d.<br />

To show that the sequ<strong>en</strong>ce of relative <strong>en</strong>tropies is boun<strong>de</strong>d, observe that for |x| ≤ 1,<br />

∣<br />

∣e β(ex−1) − 1 − βx∣ ≤ βe β(e−1)+1 (1 + βe)|x| 2<br />

and that for x ≤ B,<br />

∣<br />

∣e β(ex −1) − 1 − βx1 |x|≤1<br />

∣ ∣∣ ≤ βe<br />

β(e B +1) + 1 + βB.<br />

The uniform boun<strong>de</strong>dness of the sequ<strong>en</strong>ce of relative <strong>en</strong>tropies I(Q n |P n ) now follows from<br />

Proposition 1.6 and Equation (2.15).<br />

2.5 Regularizing the calibration problem<br />

As observed in Section 2.2, problem (2.11) is ill-posed and hard to solve numerically. In particular,<br />

its solutions, wh<strong>en</strong> they exist, may not be stable with respect to perturbations of mark<strong>et</strong><br />

data. If we do not know the prices C M exactly but only know the perturbed prices C δ M that


86 CHAPTER 2. THE CALIBRATION PROBLEM<br />

are within an error δ of C M , and want to construct an approximation to MELSS(C M ), the<br />

solution of problem (2.11) with the true data, it is not a good i<strong>de</strong>a to solve problem (2.11) with<br />

the noisy data C δ M<br />

because MELSS(Cδ M ) may be very far from MELSS(C M). We therefore<br />

need to regularize the problem (2.11), that is, construct a family of continuous “regularization<br />

operators” {R α } α>0 , where α is the param<strong>et</strong>er which <strong>de</strong>termines the int<strong>en</strong>sity of regularization,<br />

such that R α (CM δ ) converges to MELSS of the calibration problem as the noise level δ t<strong>en</strong>ds to<br />

zero if, for each δ, the regularization param<strong>et</strong>er α is chos<strong>en</strong> appropriately. The approximation<br />

to MELSS(C M ) using the noisy data CM δ is th<strong>en</strong> giv<strong>en</strong> by R α(CM δ ) with an appropriate choice<br />

of α.<br />

Following classical results on regularization of ill-posed problems (see [40]), we suggest to<br />

construct a regularized version of (2.11) by using the relative <strong>en</strong>tropy for p<strong>en</strong>alization rather<br />

than for selection, that is, to <strong>de</strong>fine<br />

J α (Q) = ‖C δ M − C Q ‖ 2 w + αI(Q|P ), (2.26)<br />

where α is the regularization param<strong>et</strong>er, and solve the following regularized calibration problem:<br />

Regularized calibration problem<br />

Giv<strong>en</strong> prices C M of call options, a prior Lévy process<br />

P and a regularization param<strong>et</strong>er α > 0, find Q ∗ ∈ M ∩ L, such that<br />

J α (Q ∗ ) =<br />

inf J α(Q). (2.27)<br />

Q∈M∩L<br />

Problem (2.27) can be thought of in two ways:<br />

• If the minimum <strong>en</strong>tropy least squares solution with the true data C M exists, (2.27) allows<br />

to construct a stable approximation of this solution using the noisy data.<br />

• If the MELSS with the true data does not exist, either because the s<strong>et</strong> of least squares<br />

solutions is empty or because the least squares solutions are incompatible with the prior,<br />

the regularized problem (2.27) allows to find a “compromise solution”, achieving a tra<strong>de</strong>off<br />

b<strong>et</strong>we<strong>en</strong> the pricing constraints and the prior information.<br />

In the rest of this section we study the regularized calibration problem. Un<strong>de</strong>r our standing<br />

hypothesis that the prior Lévy process has jumps boun<strong>de</strong>d from above and corresponds to


2.5. REGULARIZING THE CALIBRATION PROBLEM 87<br />

an arbitrage free mark<strong>et</strong> (P ∈ L NA ∩ L + B<br />

), we show that the regularized calibration problem<br />

always admits a solution that <strong>de</strong>p<strong>en</strong>ds continuously on the mark<strong>et</strong> data. In addition, we give the<br />

condition that the prior P must satisfy in or<strong>de</strong>r for the solution to be an equival<strong>en</strong>t martingale<br />

measure, and show how the regularization param<strong>et</strong>er α must be chos<strong>en</strong> <strong>de</strong>p<strong>en</strong>ding on the noise<br />

level δ if the regularized solutions are to converge to the solutions of the minimum <strong>en</strong>tropy least<br />

squares calibration problem (2.11).<br />

2.5.1 Properties of regularized solutions<br />

We start with an exist<strong>en</strong>ce theorem showing that if the prior Lévy process has jumps boun<strong>de</strong>d<br />

from above and corresponds to an arbitrage-free mark<strong>et</strong>, the regularized calibration problem<br />

admits a solution.<br />

Theorem 2.13. L<strong>et</strong> P ∈ L NA ∩ L + B<br />

a solution Q ∗ ∈ M ∩ L + B .<br />

for some B > 0. Th<strong>en</strong> the calibration problem (2.27) has<br />

Proof. By Lemma 2.12, there exists Q 0 ∈ M ∩ L with I(Q 0 |P ) < ∞. The solution, if it<br />

exists, must belong to the level s<strong>et</strong> L Jα(Q 0 ) := {Q ∈ L : I(Q|P ) ≤ J α (Q 0 )}. Since J α (Q 0 ) =<br />

‖C M − C Q0 ‖ 2 w + I(Q 0 |P ) < ∞, by Lemma 2.10, L Jα(Q 0 ) is tight and, by Prohorov’s theorem,<br />

weakly relatively compact. Corollary 2.1 <strong>en</strong>tails that I(Q|P ) is weakly lower semicontinuous<br />

with respect to Q. Therefore {Q ∈ P(Ω) : I(Q|P ) ≤ J α (Q 0 )} is weakly closed and since by<br />

Lemma 2.4, M ∩ L + B is weakly closed, M ∩ L+ B ∩ L J α(Q 0 ) is weakly compact. Moreover, by<br />

Lemma 2.2, the squared pricing error is weakly continuous, which <strong>en</strong>tails that J α (Q) is weakly<br />

lower semicontinuous. Therefore, J α (Q) achieves its minimum value on M ∩ L + B ∩ L J α(Q 0 ),<br />

which proves the theorem.<br />

Since every solution Q ∗ of the regularized calibration problem (2.27) has finite relative<br />

<strong>en</strong>tropy with respect to the prior P , necessarily Q ∗ ≪ P . However, Q ∗ need not in g<strong>en</strong>eral be<br />

equival<strong>en</strong>t to the prior. Wh<strong>en</strong> the prior corresponds to the real world (historical) probability,<br />

abs<strong>en</strong>ce of arbitrage is guaranteed if options are priced using an equival<strong>en</strong>t martingale measure.<br />

The next theorem gives a suffici<strong>en</strong>t condition for this equival<strong>en</strong>ce.<br />

Theorem 2.14. L<strong>et</strong> P ∈ L NA ∩ L + B and suppose that the characteristic function ΦP T of P


88 CHAPTER 2. THE CALIBRATION PROBLEM<br />

satisfies<br />

∫ ∞<br />

−∞<br />

|Φ P T (u)|du < ∞ (2.28)<br />

for some T < T 0 , where T 0 is the shortest maturity, pres<strong>en</strong>t in the mark<strong>et</strong> data. Th<strong>en</strong> every<br />

solution Q ∗ of the calibration problem (2.27) satisfies Q ∗ ∼ P .<br />

Remark 2.2. Condition (2.28) implies that the prior Lévy process has a continuous <strong>de</strong>nsity at<br />

time T and all subsequ<strong>en</strong>t times. Two important examples of processes satisfying the condition<br />

(2.28) for all T are<br />

• Processes with non-trivial Gaussian compon<strong>en</strong>t (A > 0). This follows directly from the<br />

Lévy-Khintchine formula (1.1).<br />

• Processes with stable-like behavior of small jumps, that is, processes whose Lévy measure<br />

satisfies<br />

∃β ∈ (0, 2),<br />

lim inf<br />

ε↓0<br />

For proof, see Proposition 28.3 in [87].<br />

(1.20) with α + > 0 and/or α − > 0.<br />

Theorem 2.14 will be prov<strong>en</strong> after the following technical lemma.<br />

∫ ε<br />

ε −β |x| 2 ν(dx) > 0. (2.29)<br />

−ε<br />

This class inclu<strong>de</strong>s tempered stable processes<br />

Lemma 2.15. L<strong>et</strong> P ∈ M ∩ L + B<br />

with characteristic tripl<strong>et</strong> (A, ν, γ) and characteristic expon<strong>en</strong>t<br />

ψ. There exists C < ∞ such that<br />

Proof. From (1.1) and (1.2),<br />

Observe first that<br />

ψ(v − i)<br />

∣ (v − i)v ∣ ≤ C ∀v ∈ R.<br />

ψ(v − i) = − 1 ∫ ∞<br />

2 Av(v − i) + (e i(v−i)x + iv − e x − ive x )ν(dx). (2.30)<br />

−∞<br />

e i(v−i)x + iv − e x − ive x = iv(xe x + 1 − e x ) + θv2 x 2 e x<br />

2<br />

for some θ with |θ| ≤ 1.<br />

Therefore, for all v with |v| ≥ 2,<br />

e i(v−i)x + iv − e x − ive x<br />

∣ (v − i)v ∣ ≤ xex + 1 − e x + x 2 e x . (2.31)


2.5. REGULARIZING THE CALIBRATION PROBLEM 89<br />

On the other hand,<br />

e i(v−i)x + iv − e x − ive x<br />

(v − i)v<br />

= iex (e ivx − 1)<br />

v<br />

− i(ei(v−i)x − 1)<br />

v − i<br />

= −xe x − ivx2<br />

2 eθ 1ivx + x +<br />

i(v − i)x2<br />

e θ 2i(v−i)x<br />

2<br />

with some θ 1 , θ 2 ∈ [0, 1]. Therefore, for all v with |v| ≤ 2,<br />

e i(v−i)x + iv − e x − ive x<br />

∣ (v − i)v ∣ ≤ x(1 − ex ) + x2<br />

2 (v + √ 1 + v 2 e x ) ≤ x(1 − e x ) + x 2 (1 + 2e x ). (2.32)<br />

Since the support of ν is boun<strong>de</strong>d from above, the right-hand si<strong>de</strong>s of (2.31) and (2.32) are<br />

ν-integrable, and the proof of the lemma is compl<strong>et</strong>ed.<br />

Proof of Theorem 2.14. L<strong>et</strong> Q ∗ be a solution of (2.27) with prior P . By Theorem 2.7, there<br />

exists Q 0 ∈ M ∩ L such that Q 0 ∼ P . D<strong>en</strong>ote the characteristic tripl<strong>et</strong> of Q ∗ by (A, ν ∗ , γ ∗ ) and<br />

that of Q 0 by (A, ν 0 , γ 0 ).<br />

L<strong>et</strong> Q x be a Lévy process with characteristic tripl<strong>et</strong><br />

(A, xν 0 + (1 − x)ν ∗ , xγ 0 + (1 − x)γ ∗ ).<br />

From the linearity of the martingale condition (1.2), it follows that for all x ∈ [0, 1], Q x ∈ M∩L.<br />

Since Q ∗ realizes the minimum of J α (Q), necessarily J α (Q x ) − J α (Q ∗ ) ≥ 0 for all x ∈ [0, 1].<br />

Our strategy for proving the theorem is first to show that ‖C M −C Qx ‖ 2 −‖C M −C Q∗ ‖ 2<br />

x<br />

is boun<strong>de</strong>d<br />

as x → 0 and th<strong>en</strong> to show that if I(Qx|P )−I(Q∗ |P )<br />

x<br />

is boun<strong>de</strong>d from below as x → 0, necessarily<br />

Q ∗ ∼ P .<br />

The first step is to prove that the characteristic function Φ ∗ of Q ∗ satisfies the condition<br />

(2.28) for some T < T 0 . If A > 0, this is trivial; suppose therefore that A = 0. In this case,<br />

|Φ ∗ T (u)| = exp(T ∫ ∞<br />

−∞ (cos(ux) − 1)ν∗ (dx)). D<strong>en</strong>ote dν∗<br />

dν P<br />

:= φ ∗ . Since Q ∗ ≪ P , by Theorem<br />

1.5, ∫ ∞<br />

−∞ (√ φ ∗ (x) − 1) 2 ν P (dx) ≤ K < ∞ for some constant K. Therefore, there exists another<br />

constant C > 0 such that<br />

∫<br />

{φ ∗ (x)>C}<br />

(1 − cos(ux))|φ ∗ − 1|ν P (dx) < C


90 CHAPTER 2. THE CALIBRATION PROBLEM<br />

uniformly on u. For all r > 0,<br />

∫ ∞<br />

−∞<br />

∫<br />

(1 − cos(ux))|φ ∗ − 1|ν P (dx) ≤ C +<br />

≤ C + r 2<br />

This implies that<br />

∫ ∞<br />

−∞<br />

∫<br />

{φ ∗ (x)≤C}<br />

{φ ∗ (x)≤C}(1 − cos(ux)) 2 ν P (dx) + 1 2r<br />

(cos(ux) − 1)ν ∗ (dx) ≤ (1 + r)<br />

≤ C + r<br />

∫ ∞<br />

−∞<br />

(1 − cos(ux))|φ ∗ − 1|ν P (dx)<br />

∫ ∞<br />

−∞<br />

∫<br />

{φ ∗ (x)≤C}<br />

(φ ∗ − 1) 2 ν P (dx)<br />

(1 − cos(ux))ν P (dx) + K(√ C + 1) 2<br />

.<br />

2r<br />

(cos(ux) − 1)ν P (dx) + C + K(√ C + 1) 2<br />

for all r > 0. Therefore, if the characteristic function of P satisfies the condition (2.28) for some<br />

T , the characteristic function of Q ∗ will satisfy it for every T ′ > T .<br />

Since P ∈ L NA ∩ L + B , Q x ∈ M ∩ L + B<br />

for all x ∈ [0, 1]. Therefore, condition (1.23) is satisfied<br />

and option prices can be computed by inverting the Fourier transform (1.24):<br />

C Qx (T, K) = (1−Ke −rT ) + + 1 ∫ ∞<br />

2π −∞<br />

e −iv log K+ivrT exp(T (1 − x)ψ∗ (v − i) + T xψ 0 (v − i)) − 1<br />

dv,<br />

iv(1 + iv)<br />

where ψ 0 and ψ ∗ <strong>de</strong>note the characteristic expon<strong>en</strong>ts of Q 0 and Q ∗ . It follows that<br />

2r<br />

C Qx (T, K) − C Q∗ (T, K)<br />

x<br />

= 1 ∫ ∞<br />

2π −∞<br />

e −iv log K+ivrT eT (1−x)ψ∗ (v−i)+T xψ 0 (v−i) − e T ψ∗ (v−i)<br />

dv<br />

iv(1 + iv)x<br />

The fact that Rψ 0 (v − i) ≤ 0 and Rψ ∗ (v − i) ≤ 0 for all v ∈ R tog<strong>et</strong>her with Lemma 2.15<br />

implies that<br />

∣ e−iv log K+ivrT eT (1−x)ψ∗ (v−i)+T xψ 0 (v−i) − e T ψ∗ (v−i)<br />

iv(1 + iv)x<br />

∣<br />

≤ T |eT (1−x)ψ∗ (v−i) ||ψ 0 (v − i) − ψ ∗ (v − i)|<br />

|v(1 + iv)|<br />

≤ T |e T (1−x)ψ∗ (v−i) |C ′<br />

for some constant C ′ . From the dominated converg<strong>en</strong>ce theorem and since Q ∗ satisfies (2.28),<br />

∂C Qx (T,K)<br />

∂x<br />

| x=0 exists and is boun<strong>de</strong>d uniformly on T and K in the mark<strong>et</strong> data s<strong>et</strong>. This in<br />

turn means that ‖C M −C Qx ‖ 2 −‖C M −C Q∗ ‖ 2<br />

x<br />

is boun<strong>de</strong>d as x → 0.<br />

To compl<strong>et</strong>e the proof, it remains to show that if I(Qx|P )−I(Q∗ |P )<br />

x<br />

is boun<strong>de</strong>d from below as<br />

x → 0, necessarily Q ∗ ∼ P . Using the convexity (with respect to ν Q and γ Q ) of the two terms


2.5. REGULARIZING THE CALIBRATION PROBLEM 91<br />

in the expression (2.19) for relative <strong>en</strong>tropy, we have:<br />

I(Q x |P ) − I(Q ∗ |P )<br />

x<br />

{<br />

= T ∫<br />

2<br />

∞<br />

xγ 0 + (1 − x)γ ∗ − γ P − z(xν 0 + (1 − x)ν ∗ − ν P )ν (dz)} P 1 A≠0<br />

2Ax<br />

|z|≤1<br />

{<br />

− T ∫<br />

2<br />

∞<br />

γ ∗ − γ P − (ν ∗ − ν P )ν (dz)} P 1 A≠0<br />

2Ax<br />

≤ T ∞<br />

2A<br />

+ T ∞<br />

x<br />

− T ∞<br />

x<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

−∞<br />

{ ∫<br />

γ 0 − γ P −<br />

|z|≤1<br />

{(xφ 0 + (1 − x)φ ∗ ) log(xφ 0 + (1 − x)φ ∗ ) − xφ 0 − (1 − x)φ ∗ + 1}ν P (dz)<br />

{φ ∗ log(φ ∗ ) − φ ∗ + 1}ν P (dz)<br />

|z|≤1<br />

(ν 0 − ν P )ν P (dz)} 2<br />

1 A≠0<br />

{<br />

− T ∫<br />

2<br />

∞<br />

γ ∗ − γ P − (ν ∗ − ν P )ν (dz)} P 1 A≠0<br />

2A<br />

|z|≤1<br />

∫<br />

∫<br />

+ T ∞ {φ 0 log(φ 0 ) − φ 0 + 1}ν P (dz) − T ∞<br />

+ T ∞<br />

∫<br />

{φ ∗ >0}<br />

{φ ∗ =0}<br />

{φ ∗ >0}<br />

{φ 0 log(xφ 0 ) − φ 0 }ν P (dz) ≤ I(Q 0 |P ) + T ∞<br />

∫<br />

{φ ∗ log(φ ∗ ) − φ ∗ + 1}ν P (dz)<br />

{φ ∗ =0}<br />

(φ 0 log x − 1)ν P (dx)<br />

Since J α (Q x ) − J α (Q ∗ ) ≥ 0, this expression must be boun<strong>de</strong>d from below. Therefore, ν P ({φ ∗ =<br />

0}) = 0, and Proposition 1.5 <strong>en</strong>tails that P ≪ Q ∗ .<br />

Continuity of solutions with respect to data<br />

Theorem 2.16. L<strong>et</strong> {C n M } n≥1 and C M be data s<strong>et</strong>s of option prices such that<br />

lim<br />

n<br />

‖C n M − C M ‖ w → 0.<br />

L<strong>et</strong> P ∈ L NA ∩ L + B , α > 0, and for each n, l<strong>et</strong> Q n be a solution of the calibration problem<br />

(2.27) with data C n M , prior Lévy process P and regularization param<strong>et</strong>er α. Th<strong>en</strong> {Q n} n≥1 has<br />

a subsequ<strong>en</strong>ce, converging weakly to a process Q ∗ ∈ M ∩ L + B<br />

, and the limit of every converging<br />

subsequ<strong>en</strong>ce of {Q n } n≥1 is a solution of calibration problem (2.27) with data C M , prior P and<br />

regularization param<strong>et</strong>er α.<br />

Proof. By Lemma 2.12, there exists Q 0 ∈ M ∩ L with I(Q 0 |P ) < ∞. Since, by Lemma 2.2,<br />

‖C Q0 − C n M ‖2 ≤ S 2 0 for all n, αI(Q n|P ) ≤ S 2 0 + αI(Q0 |P ) for all n. Therefore, by Lemmas 2.4


92 CHAPTER 2. THE CALIBRATION PROBLEM<br />

and 2.10 and Prohorov’s theorem, {Q n } n≥1 is weakly relatively compact, which proves the first<br />

part of the theorem.<br />

Choose any subsequ<strong>en</strong>ce of {Q n } n≥1 , converging weakly to a process Q ∗ ∈ M ∩ L + B . , To<br />

simplify notation, this subsequ<strong>en</strong>ce is <strong>de</strong>noted again by {Q n } n≥1 . The triangle inequality and<br />

Lemma 2.2 imply that<br />

‖C Qn − C n M‖ 2 −−−→<br />

n→∞ ‖CQ∗ − C M ‖ 2 (2.33)<br />

Since, by Lemma 2.11, the relative <strong>en</strong>tropy functional is weakly lower semicontinuous in Q,<br />

for every Q ∈ M ∩ L + B ,<br />

‖C Q∗ − C M ‖ + αI(Q|P ) ≤ lim inf{‖C Qn − C n<br />

n<br />

M‖ 2 + αI(Q n |P )}<br />

≤ lim inf{‖C Q − C n<br />

n<br />

M‖ 2 + αI(Q|P )}<br />

= lim<br />

n<br />

‖C Q − C n M‖ 2 + αI(Q|P )<br />

= ‖C Q − C M ‖ 2 + αI(Q|P ),<br />

where the second inequality follows from the fact that Q m is the solution of the calibration<br />

problem with data C m M<br />

and the last line follows from the triangle inequality.<br />

2.5.2 Converg<strong>en</strong>ce of regularized solutions<br />

In this section we study the converg<strong>en</strong>ce of solutions of the regularized calibration problem<br />

(2.27) to the solutions of the minimum <strong>en</strong>tropy least squares calibration problem (2.11) wh<strong>en</strong><br />

the noise level in the data t<strong>en</strong>ds to zero.<br />

Theorem 2.17. L<strong>et</strong> {C δ M } be a family of data s<strong>et</strong>s of option prices such that ‖C M − C δ M ‖ ≤ δ,<br />

l<strong>et</strong> P ∈ L NA ∩ L + B and suppose that there exist a solution Q of problem (2.4) with data C M (a<br />

least squares solution) such that I(Q|P ) < ∞.<br />

and<br />

If ‖C Q − C M ‖ = 0 (the constraints are reproduced exactly), l<strong>et</strong> α(δ) be such that α(δ) → 0<br />

δ2<br />

δ<br />

α(δ)<br />

→ 0 as δ → 0. Otherwise, l<strong>et</strong> α(δ) be such that α(δ) → 0 and<br />

Th<strong>en</strong> every sequ<strong>en</strong>ce {Q δ k}, where δ k → 0 and Q δ k<br />

α(δ)<br />

→ 0 as δ → 0.<br />

is a solution of problem (2.27) with data<br />

C δ k<br />

M , prior P and regularization param<strong>et</strong>er α(δ k), has a weakly converg<strong>en</strong>t subsequ<strong>en</strong>ce. The<br />

limit of every converg<strong>en</strong>t subsequ<strong>en</strong>ce is a solution of problem (2.11) (MELSS) with data C M<br />

and prior P . If such a MELSS Q + is unique th<strong>en</strong> lim δ→0 Q δ = Q + .


2.5. REGULARIZING THE CALIBRATION PROBLEM 93<br />

Proof. By Lemma 2.6, there exists at least one MELSS with data C M and prior P , that has<br />

finite relative <strong>en</strong>tropy with respect to the prior. L<strong>et</strong> Q + be any such MELSS. Since Q δ k<br />

solution of the regularized problem, for every k,<br />

is the<br />

‖C Qδk − C δ k<br />

M ‖2 + α(δ k )I(Q δ k<br />

|P ) ≤ ‖C Q+ − C δ k<br />

M ‖2 + α(δ k )I(Q + |P ).<br />

Using the fact that for every r > 0 and for every x, y ∈ R,<br />

(1 − r)x 2 + (1 − 1/r)y 2 ≤ (x + y) 2 ≤ (1 + r)x 2 + (1 + 1/r)y 2 ,<br />

we obtain that<br />

(1 − r)‖C Qδk − C M ‖ 2 + α(δ k )I(Q δ k<br />

|P )<br />

≤ (1 + r)‖C Q+ − C M ‖ 2 + 2δ2 k<br />

r + α(δ k)I(Q + |P ), (2.34)<br />

and since Q + is a least squares solution with data C M , this implies for all r ∈ (0, 1) that<br />

α(δ k )I(Q δ k<br />

|P ) ≤ 2r‖C Q+ − C M ‖ 2 + 2δ2 k<br />

r + α(δ k)I(Q + |P ). (2.35)<br />

If the constraints are reproduced exactly, th<strong>en</strong> ‖C Q+ −C M ‖ = 0 and with the choice r = 1/2,<br />

the above expression yields:<br />

I(Q δ k<br />

|P ) ≤<br />

4δ2 k<br />

α(δ k ) + I(Q+ |P ).<br />

Since, by the theorem’s statem<strong>en</strong>t, in the case of exact constraints<br />

δk<br />

2<br />

α(δ k )<br />

→ 0, this implies that<br />

lim sup{I(Q δ k<br />

|P )} ≤ I(Q + |P ). (2.36)<br />

k<br />

If ‖C Q+ − C M ‖ > 0 (misspecified mo<strong>de</strong>l) th<strong>en</strong> the right-hand si<strong>de</strong> of (2.35) achieves its<br />

maximum wh<strong>en</strong> r = δ k ‖C Q+ − C M ‖ −1 , in which case we obtain<br />

I(Q δ k<br />

|P ) ≤<br />

Since in the case of approximate constraints,<br />

4δ k<br />

α(δ k ) ‖CQ+ − C M ‖ + I(Q + |P ).<br />

δ k<br />

α(δ k )<br />

→ 0, we obtain (2.36) once again.<br />

Inequality (2.36) implies in particular that I(Q δ k|P ) is uniformly boun<strong>de</strong>d, which proves,<br />

by Lemmas 2.10 and 2.4, that {Q δ k} is relatively weakly compact in M ∩ L + B .


94 CHAPTER 2. THE CALIBRATION PROBLEM<br />

Choose a subsequ<strong>en</strong>ce of {Q δ k}, converging weakly to Q ∗ ∈ M ∩ L + B<br />

. To simplify notation,<br />

this subsequ<strong>en</strong>ce is <strong>de</strong>noted again by {Q δ k} k≥1 . Substituting r = δ into Equation (2.34) and<br />

making k t<strong>en</strong>d to infinity shows that<br />

Tog<strong>et</strong>her with Lemma 2.2 this implies that<br />

lim sup ‖C Qδk − C M ‖ 2 ≤ ‖C Q+ − C M ‖ 2 .<br />

k<br />

‖C Q∗ − C M ‖ 2 ≤ ‖C Q+ − C M ‖ 2 ,<br />

h<strong>en</strong>ce Q ∗ is a least squares solution. By weak lower semicontinuity of I (cf. Lemma 2.11) and<br />

using (2.36),<br />

I(Q ∗ |P ) ≤ lim inf<br />

k<br />

I(Q δ k<br />

|P ) ≤ lim sup I(Q δ k<br />

|P ) ≤ I(Q + |P ),<br />

k<br />

which means that Q ∗ is a MELSS. The last assertion of the theorem follows from the fact that<br />

in this case every subsequ<strong>en</strong>ce of {Q δ k} has a further subsequ<strong>en</strong>ce converging toward Q + .


Chapter 3<br />

Numerical implem<strong>en</strong>tation of the<br />

calibration algorithm<br />

Before solving the calibration problem (2.27) numerically, we reformulate it as follows:<br />

• The calibration problem (2.27) is expressed in terms of the characteristic tripl<strong>et</strong>s (A, ν Q , γ Q )<br />

and (A, ν P , γ P ) of the prior Lévy process and the solution. This can be done using Equations<br />

(1.1), (1.25) and (1.26) (option pricing by Fourier transform) and Equation (2.19)<br />

(expression of the relative <strong>en</strong>tropy in terms of characteristic tripl<strong>et</strong>s).<br />

• The Lévy measure ν Q is discr<strong>et</strong>ized (approximated by a finite-dim<strong>en</strong>sional object). This<br />

discussed in <strong>de</strong>tail in Section 3.1.<br />

The prior Lévy process P is a crucial ingredi<strong>en</strong>t that must be specified by the user. Section<br />

3.2 suggests differ<strong>en</strong>t ways to do this and studies the effect of a misspecification of the prior<br />

on the solutions of the calibration problem. Section 3.3 discusses the m<strong>et</strong>hods to choose the<br />

regularization param<strong>et</strong>er α based on the data and Section 3.4 treats the choice of weights w i of<br />

differ<strong>en</strong>t options. Section 3.5 <strong>de</strong>tails the numerical algorithm that we use to solve the calibration<br />

problem once the prior, the regularization param<strong>et</strong>er and the weights have be<strong>en</strong> fixed.<br />

The calibration algorithm, including the automatic choice of the regularization param<strong>et</strong>er,<br />

has be<strong>en</strong> implem<strong>en</strong>ted in a computer program levycalibration and various tests have be<strong>en</strong><br />

carried out, both on simulated option prices (computed in a known expon<strong>en</strong>tial Lévy mo<strong>de</strong>l)<br />

and using real mark<strong>et</strong> data. Section 3.6 discusses the results of these tests.<br />

95


96 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

3.1 Discr<strong>et</strong>izing the calibration problem<br />

A conv<strong>en</strong>i<strong>en</strong>t way to discr<strong>et</strong>ize the calibration problem is to take a prior Lévy process P with<br />

Lévy measure supported by a finite number of points:<br />

ν P =<br />

M−1<br />

∑<br />

k=0<br />

p k δ {xk }(dx). (3.1)<br />

In this case, by Proposition 1.5, the Lévy measure of the solution necessarily satisfies ν Q ≪ ν P ,<br />

therefore<br />

ν Q =<br />

M−1<br />

∑<br />

k=0<br />

q k δ {xk }(dx), (3.2)<br />

that is, the solution belongs to a finite-dim<strong>en</strong>sional space and can be computed using a numerical<br />

optimization algorithm. The advantage of this discr<strong>et</strong>ization approach is that we are solving the<br />

same problem (2.27), only with a differ<strong>en</strong>t prior measure, so all results of Section 2.5 (exist<strong>en</strong>ce<br />

of solution, continuity <strong>et</strong>c.) hold in the finite-dim<strong>en</strong>sional case.<br />

Taking Lévy measures of the form (3.1) we implicitly restrict the class of possible solutions<br />

to Lévy processes with boun<strong>de</strong>d jumps and finite jump int<strong>en</strong>sity. However, in this section we<br />

will see that this restriction is not as important as it seems: the solution of a calibration problem<br />

with any prior can be approximated (in the weak s<strong>en</strong>se) by a sequ<strong>en</strong>ce of solutions of calibration<br />

problems with priors having Lévy measures of the form (3.1). Moreover, in Section 3.6.1 we<br />

will observe empirically that smiles produced by infinite int<strong>en</strong>sity mo<strong>de</strong>ls can be calibrated with<br />

arbitrary precision by such jump-diffusion mo<strong>de</strong>ls.<br />

We start with a lemma showing that every Lévy process can be approximated by Lévy<br />

processes with atomic Lévy measures.<br />

Lemma 3.1. L<strong>et</strong> P be a Lévy process with characteristic tripl<strong>et</strong> (A, ν, γ) with respect to a<br />

continuous boun<strong>de</strong>d truncation function h, satisfying h(x) = x in a neighborhood of 0, and for<br />

every n, l<strong>et</strong> P n be a Lévy process with characteristic tripl<strong>et</strong> (A, ν n , γ) (with respect to the same<br />

truncation function) where<br />

ν n :=<br />

2n∑<br />

k=1<br />

δ {xk }(dx) µ([x k − 1/ √ n, x k + 1/ √ n))<br />

1 ∧ x 2 ,<br />

k<br />

x k := (2(k − n) − 1)/ √ n and µ is a finite measure on R, <strong>de</strong>fined by µ(B) := ∫ B (1 ∧ x2 )ν(dx)<br />

for all B ∈ B(R). Th<strong>en</strong> P n ⇒ P .


3.1. DISCRETIZING THE CALIBRATION PROBLEM 97<br />

Proof. For a function f ∈ C b (R), <strong>de</strong>fine<br />

⎧<br />

0, x ≥ 2 √ n,<br />

⎪⎨<br />

f n (x) := 0, x < −2 √ n,<br />

⎪⎩<br />

f(x i ), x ∈ [x i − 1/ √ n, x i + 1/ √ n) with 1 ≤ i ≤ 2n,<br />

Th<strong>en</strong> clearly<br />

∫<br />

∫<br />

(1 ∧ x 2 )f(x)ν n (dx) =<br />

f n (x)µ(dx).<br />

Since f(x) is continuous, f n (x) → f(x) for all x and since f is boun<strong>de</strong>d, the dominated converg<strong>en</strong>ce<br />

theorem implies that<br />

lim<br />

n<br />

∫<br />

(1 ∧ x 2 )f(x)ν n (dx) = lim<br />

n<br />

∫<br />

∫<br />

f n (x)µ(dx) =<br />

∫<br />

f(x)µ(dx) =<br />

(1 ∧ x 2 )f(x)ν(dx). (3.3)<br />

With f(x) ≡ h2 (x)<br />

1∧x 2<br />

the above yields:<br />

∫<br />

∫<br />

lim h 2 (x)ν n (dx) =<br />

n<br />

h 2 (x)ν(dx).<br />

On the other hand, for every g ∈ C b (R) such that g(x) ≡ 0 on a neighborhood of 0, f(x) := g(x)<br />

1∧x 2<br />

belongs to C b (R). Therefore, from Equation (3.3), lim<br />

n<br />

∫<br />

g(x)νn (dx) = ∫ g(x)ν(dx), and by<br />

Proposition (1.7), P n ⇒ P .<br />

Theorem 3.2. L<strong>et</strong> P, {P n } n≥1 ⊂ L NA ∩ L + B such that P n ⇒ P , l<strong>et</strong> α > 0, l<strong>et</strong> C M be a data<br />

s<strong>et</strong> of option prices and for each n l<strong>et</strong> Q n be a solution of the calibration problem (2.27) with<br />

prior P n , regularization param<strong>et</strong>er α and data C M . Th<strong>en</strong> the sequ<strong>en</strong>ce {Q n } n≥1 has a weakly<br />

converg<strong>en</strong>t subsequ<strong>en</strong>ce and the limit of every weakly converg<strong>en</strong>t subsequ<strong>en</strong>ce of {Q n } n≥1 is a<br />

solution of the calibration problem (2.27) with prior P .<br />

Proof. By Lemma 2.12, there exists C < ∞ such that for every n, one can find ˜Q n ∈ M∩L with<br />

I( ˜Q n |P n ) ≤ C. Since, by Lemma 2.2, ‖C ˜Q n<br />

− C M ‖ 2 w ≤ S 2 0 for every n and Q n is the solution<br />

of the calibration problem, I(Q n |P n ) ≤ S0 2 /α + C < ∞ for every n. Therefore, by Lemma<br />

2.10, {Q n } is tight and, by Prohorov’s theorem and Lemma 2.4, weakly relatively compact in<br />

M ∩ L + B . Choose a subsequ<strong>en</strong>ce of {Q n}, converging weakly to Q ∈ M ∩ L + B<br />

. To simplify<br />

notation, this subsequ<strong>en</strong>ce is also <strong>de</strong>noted by {Q n } n≥1 . It remains to show that Q is in<strong>de</strong>ed a<br />

solution of the calibration problem with prior P .


98 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

Lemma 2.11 <strong>en</strong>tails that<br />

I(Q, P ) ≤ lim inf<br />

n I(Q n, P n ), (3.4)<br />

and since, by Lemma 2.2, the pricing error is weakly continuous, we also have<br />

‖C Q − C M ‖ 2 w + αI(Q, P ) ≤ lim inf<br />

n {‖CQn − C M ‖ 2 w + αI(Q n , P n )}. (3.5)<br />

L<strong>et</strong> φ ∈ C b (Ω) with φ ≥ 0 and E P [φ] = 1. Without loss of g<strong>en</strong>erality we can suppose that for<br />

every n, E Pn [φ] > 0 and therefore Q ′ n, <strong>de</strong>fined by Q ′ n(B) := EPn [φ1 B ]<br />

, is a probability measure<br />

E Pn [φ]<br />

on Ω. Clearly, Q ′ n converges weakly to Q ′ <strong>de</strong>fined by Q ′ (B) := E P [φ1 B ]. Therefore, by Lemma<br />

2.2,<br />

Moreover,<br />

lim<br />

n<br />

I(Q ′ n|P n ) = lim<br />

n<br />

∫<br />

Ω<br />

lim<br />

n<br />

‖C Q′ n<br />

− C M ‖ 2 w = ‖C Q′ − C M ‖ 2 w. (3.6)<br />

φ<br />

E Pn [φ] log<br />

= lim<br />

n<br />

1<br />

E Pn [φ]<br />

φ<br />

E Pn [φ] dP n<br />

∫<br />

φ log φdP n − lim<br />

n<br />

Ω<br />

∫ ∫<br />

log φdP n = φ log φdP. (3.7)<br />

Ω<br />

Ω<br />

For the rest of this proof, for every φ ∈ L 1 (P ) with φ ≥ 0 and E P [φ] = 1 l<strong>et</strong> Q φ <strong>de</strong>note the<br />

probability measure on Ω, <strong>de</strong>fined by Q φ (B) := E P [φ1 B ] for every B ∈ F. Using (3.5–3.7) and<br />

the optimality of Q n , we obtain that for every φ ∈ C b (Ω) with φ ≥ 0 and E P [φ] = 1,<br />

‖C Q − C M ‖ 2 w + I(Q, P ) ≤ ‖C Q φ<br />

− C M ‖ 2 w + I(Q φ |P ) (3.8)<br />

To compl<strong>et</strong>e the proof of the theorem, we must g<strong>en</strong>eralize this inequality to all φ ∈ L 1 (P ) with<br />

φ ≥ 0 and E P [φ] = 1.<br />

First, l<strong>et</strong> φ ∈ L 1 (P ) ∩ L ∞ (P ) with φ ≥ 0 and E P [φ] = 1. Th<strong>en</strong> there exists a sequ<strong>en</strong>ce<br />

{φ n } ⊂ C b (Ω) such that φ n → φ in L 1 (P ), φ n ≥ 0 for all n and φ n are boun<strong>de</strong>d in L ∞ norm<br />

uniformly on n. Moreover, φ ′ n := φ n /E P [φ n ] also belongs to L 1 (P ), is positive and φ ′ n<br />

because by the triangle inequality,<br />

‖φ ′ n − φ‖ L 1 ≤<br />

1 (<br />

‖φn<br />

E P − φ‖<br />

[φ n ]<br />

L 1 + ‖φ − φE P )<br />

[φ n ]‖ L 1 −−−→ 0.<br />

n→∞<br />

In addition, it is easy to see that Q φ ′ n<br />

⇒ Q φ . Therefore,<br />

L 1 (P )<br />

−−−→ φ<br />

lim<br />

n<br />

‖C Q φ ′ n − C M ‖ 2 w = ‖C Q φ<br />

− C M ‖ 2 w


3.2. CHOICE OF THE PRIOR 99<br />

Since φ ′ n are boun<strong>de</strong>d in L ∞ norm uniformly on n, φ ′ n log φ ′ n is also boun<strong>de</strong>d and the dominated<br />

converg<strong>en</strong>ce theorem implies that lim n I(Q φ ′ n<br />

|P ) = I(Q φ |P ). Passing to the limit in (3.8), we<br />

obtain that this inequality holds for every φ ∈ L 1 (P ) ∩ L ∞ (P ) with φ ≥ 0 and E P [φ] = 1.<br />

L<strong>et</strong> us now choose a nonnegative φ ∈ L 1 (P ) with E P [φ] = 1. If I(Q φ |P ) = ∞ th<strong>en</strong> surely<br />

(3.8) holds, therefore we can suppose I(Q φ |P ) < ∞. L<strong>et</strong> φ n = φ ∧ n. Th<strong>en</strong> φ n → φ in L 1 (P )<br />

because<br />

∫<br />

‖φ n − φ‖ L 1 ≤<br />

φ≥n<br />

∫<br />

φdP =<br />

φ≥n<br />

D<strong>en</strong>oting φ ′ n := φ n /E P [φ n ] as above, we obtain that<br />

φ log φ<br />

log φ dP ≤ I(Q φ|P )<br />

log n → 0.<br />

lim<br />

n<br />

‖C Q φ ′ n − C M ‖ 2 w = ‖C Q φ<br />

− C M ‖ 2 w<br />

Since, for a suffici<strong>en</strong>tly large n, |φ n (x) log φ n (x)| ≤ |φ(x) log φ(x)|, we can once again apply the<br />

dominated converg<strong>en</strong>ce theorem:<br />

∫<br />

lim φ ′<br />

n<br />

n log φ ′ 1<br />

ndP =<br />

lim n E P [φ n ] lim n<br />

∫<br />

∫<br />

φ n log φ n dP − lim log E P [φ n ] =<br />

n<br />

φ log φdP<br />

Therefore, by passage to the limit, (3.8) holds for all φ ∈ L 1 (P ) with φ ≥ 0 and E P [φ] = 1,<br />

which compl<strong>et</strong>es the proof of the theorem.<br />

To approximate numerically the solution of the calibration problem (2.27) with a giv<strong>en</strong> prior<br />

P , we need to construct, using Lemma 3.1, an approximating sequ<strong>en</strong>ce {P n } of Lévy processes<br />

with atomic measures such that P n ⇒ P . The sequ<strong>en</strong>ce {Q n } of solutions corresponding to this<br />

sequ<strong>en</strong>ce of priors will converge (in the s<strong>en</strong>se of Theorem 3.2) to a solution of the calibration<br />

problem with prior P .<br />

3.2 Choice of the prior Lévy process<br />

The prior Lévy process must, g<strong>en</strong>erally speaking, reflect the user’s view of the mo<strong>de</strong>l. It is<br />

one of the most important ingredi<strong>en</strong>ts of the calibration m<strong>et</strong>hod and cannot be <strong>de</strong>termined<br />

compl<strong>et</strong>ely automatically because the choice of the prior has a strong influ<strong>en</strong>ce on the outcome<br />

of the calibration. The user should therefore specify a characteristic tripl<strong>et</strong> (A P , ν P , γ P ) of<br />

the prior P . A natural solution, justified by the economic consi<strong>de</strong>rations of Section 2.3 is<br />

to take the historical probability, resulting from statistical estimation of an expon<strong>en</strong>tial Lévy


100 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

mo<strong>de</strong>l from the time series of ass<strong>et</strong> r<strong>et</strong>urns. Since, un<strong>de</strong>r the conditions of Theorem 2.14,<br />

the calibration procedure yields a martingale probability equival<strong>en</strong>t to the prior, the choice<br />

of historical probability as prior <strong>en</strong>sures the abs<strong>en</strong>ce of arbitrage opportunity involving stock<br />

and options in the calibrated mo<strong>de</strong>l. Estimation of expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls is not discussed<br />

here and interested rea<strong>de</strong>r is referred to [27, Chapter 7].<br />

Specific expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls<br />

are discussed in [57] (Merton mo<strong>de</strong>l), [79] (g<strong>en</strong>eralized hyperbolic mo<strong>de</strong>l) and [21] (CGMY or<br />

tempered stable mo<strong>de</strong>l). [9] and [12] treat more g<strong>en</strong>eral jump-diffusion type mo<strong>de</strong>ls.<br />

To <strong>en</strong>sure the stability of calibrated Lévy measures over time, one can systematically choose<br />

the calibrated expon<strong>en</strong>tial Lévy mo<strong>de</strong>l of the previous day as today’s prior. This choice guarantees<br />

that the curr<strong>en</strong>t calibrated Lévy measure will be altered in the least possible way, to<br />

accommodate the arrival of new option pricing information.<br />

Alternatively, the prior can simply correspond to a mo<strong>de</strong>l that seems “reasonable” to the<br />

user (e.g. an analyst). In this case our calibration m<strong>et</strong>hod may be se<strong>en</strong> as a way to make the<br />

smallest possible change in this mo<strong>de</strong>l to take into account the observed option prices.<br />

Wh<strong>en</strong> the user does not have such <strong>de</strong>tailed views, it should be possible to g<strong>en</strong>erate a refer<strong>en</strong>ce<br />

measure P automatically from options data. To do this we choose an auxiliary mo<strong>de</strong>l Q(θ) (e.g.<br />

Merton mo<strong>de</strong>l) <strong>de</strong>scribed by a finite-dim<strong>en</strong>sional param<strong>et</strong>er vector θ and calibrate it to data<br />

using the least squares procedure:<br />

θ ∗ = arg min<br />

θ<br />

‖C Q(θ) − C M ‖ 2 w (3.9)<br />

It is g<strong>en</strong>erally not a good i<strong>de</strong>a to recalibrate this param<strong>et</strong>ric mo<strong>de</strong>l every day, because in this<br />

case the prior will compl<strong>et</strong>ely lose its stabilizing role. On the contrary, one should try to find<br />

typical param<strong>et</strong>er values for a particular mark<strong>et</strong> (e.g.<br />

averages over a long period) and fix<br />

them once and for all. Since the objective function in (3.9) is usually not convex, a simple<br />

gradi<strong>en</strong>t <strong>de</strong>sc<strong>en</strong>t procedure may not give the global minimum. However, the solution Q(θ ∗ ) will<br />

be corrected at later stages and should only be viewed as a way to regularize the optimization<br />

problem (2.4) so the minimization procedure at this stage need not be precise.<br />

Theorem 3.2 shows that the calibrated solutions are continuous with respect to the prior,<br />

that is, small changes in the prior process induce small changes in the solutions.<br />

To assess<br />

empirically the influ<strong>en</strong>ce of finite changes in the prior on the result of calibration, we have<br />

carried out two series of numerical tests. In the first series of tests, Lévy measure was calibrated


3.2. CHOICE OF THE PRIOR 101<br />

twice to the same s<strong>et</strong> of option prices using prior mo<strong>de</strong>ls that were differ<strong>en</strong>t but close to each<br />

other (see Section 3.5.3 for the <strong>de</strong>scription of the calibration algorithm). Namely, in the test A<br />

we used Merton mo<strong>de</strong>l with diffusion volatility σ = 0.2, zero mean jump size, jump standard<br />

<strong>de</strong>viation of 0.1 and int<strong>en</strong>sity λ = 3, whereas in the test B all the param<strong>et</strong>ers except int<strong>en</strong>sity<br />

had the same values and the int<strong>en</strong>sity was equal to 2. The result of the test is shown in<br />

Figure 3.1. The solid curves correspond to calibrated measures and the dotted ones <strong>de</strong>pict<br />

the prior measures. Notice that there is very little differ<strong>en</strong>ce b<strong>et</strong>we<strong>en</strong> the calibrated measures,<br />

which means that, in harmony with Theorem 3.2, the result of calibration is robust to minor<br />

variations of the param<strong>et</strong>ers of prior measure, as long as its qualitative shape remains the same.<br />

12<br />

10<br />

Test A<br />

8<br />

6<br />

4<br />

Test B<br />

2<br />

0<br />

−1.2 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6<br />

Figure 3.1: Lévy measures calibrated to the same data s<strong>et</strong> using two prior measures similar to<br />

each other. Solid curves correspond to calibrated measures and dotted ones <strong>de</strong>pict the priors.<br />

In the second series of tests we have again calibrated the Lévy measure twice to the same<br />

s<strong>et</strong> of option prices, this time taking two radically differ<strong>en</strong>t priors. Namely, in test A we used<br />

Merton mo<strong>de</strong>l with diffusion volatility σ = 0.2, zero mean jump size, jump standard <strong>de</strong>viation<br />

of 0.1 and int<strong>en</strong>sity λ = 2, whereas in test B we took a uniform Lévy measure on the interval<br />

[−1, 0.5] with int<strong>en</strong>sity λ = 2. The calibrated measures (solid lines in Figure 3.2) are still similar<br />

but exhibit much more differ<strong>en</strong>ces than in the first series of tests. Not only they are differ<strong>en</strong>t


102 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

in the middle, but also the behavior of tails of the calibrated Lévy measure with uniform prior<br />

is more erratic than in the case where Merton mo<strong>de</strong>l was used as prior (see Figure 3.2, right<br />

graph).<br />

Comparison of Figures 3.1 and 3.2 shows that the exact values of param<strong>et</strong>ers of the prior<br />

mo<strong>de</strong>l are not very important, but it is crucial to find the right shape of the prior.<br />

8<br />

10 2 Test A<br />

7<br />

6<br />

5<br />

Test A<br />

10 1<br />

10 0<br />

4<br />

10 −1<br />

Test B<br />

3<br />

2<br />

1<br />

Test B<br />

10 −2<br />

10 −3<br />

0<br />

−1.2 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6<br />

10 −4<br />

−0.5 −0.4 −0.3 −0.2 −0.1 0<br />

Figure 3.2: Lévy measures calibrated to the same data s<strong>et</strong> using two qualitatively differ<strong>en</strong>t<br />

priors. Solid curves correspond to calibrated measures and dotted ones <strong>de</strong>pict the priors. Right<br />

graph: zoom of the left tail of Lévy <strong>de</strong>nsities on log scale.<br />

3.3 Choice of the regularization param<strong>et</strong>er<br />

Theorem 2.17 of the preceding chapter suggests that for the regularization m<strong>et</strong>hod to converge,<br />

the param<strong>et</strong>er choice strategy α(δ) must satisfy the following limiting relations, wh<strong>en</strong> the data<br />

error δ goes to zero:<br />

• α(δ) → 0,<br />

• If the exact data C M is attainable by the mo<strong>de</strong>l, one must have<br />

δ 2<br />

α(δ)<br />

δ<br />

α(δ) → 0.<br />

→ 0 and if not,


3.3. CHOICE OF THE REGULARIZATION PARAMETER 103<br />

A typical example of a param<strong>et</strong>er choice rule that works in both attainable and unattainable<br />

cases would be<br />

α(δ) = Cδ µ (3.10)<br />

with any C > 0 and any µ ∈ (0, 1).<br />

However, in real calibration problems the error level δ is fixed and finite and rules of type<br />

(3.10) do not tell us how we should choose α in this case. Good performance and error control<br />

for finite values of δ is achieved by using the so called a posteriori param<strong>et</strong>er choice rules<br />

(α <strong>de</strong>p<strong>en</strong>ds not only on δ but also on the data CM δ ), the most wi<strong>de</strong>ly used of which is the<br />

discrepancy principle, originally <strong>de</strong>veloped by Morozov for Tikhonov regularization in Banach<br />

spaces [74, 75], see also [40]. In the rest of this section we apply this principle and its variants<br />

that is b<strong>et</strong>ter suited for nonlinear problems to our <strong>en</strong>tropic regularization m<strong>et</strong>hod.<br />

3.3.1 A posteriori param<strong>et</strong>er choice rules for attainable calibration problems<br />

In this subsection we make the following standing assumptions:<br />

1. The prior Lévy process corresponds to an arbitrage-free mo<strong>de</strong>l with jumps boun<strong>de</strong>d from<br />

above by B: P ∈ L NA ∩ L + B<br />

measure Q ∗ ).<br />

(this implies the exist<strong>en</strong>ce of a minimal <strong>en</strong>tropy martingale<br />

2. There exists a solution Q + of problem (2.11) with data C M (minimum <strong>en</strong>tropy least<br />

squares solution) such that<br />

I(Q + |P ) < ∞<br />

and<br />

‖C Q+ − C M ‖ w = 0<br />

(the data is attainable by an exp-Lévy mo<strong>de</strong>l).<br />

3. There exists δ 0 such that<br />

ε max := inf<br />

δ≤δ 0<br />

‖C Q∗ − C δ M‖ 2 w > δ 2 0. (3.11)<br />

Remark 3.1. In the condition (3.11), δ 0 can be se<strong>en</strong> as the highest possible noise level of all data<br />

s<strong>et</strong>s that we consi<strong>de</strong>r. If, for some δ, ‖C Q∗ − C δ M ‖ w < δ 0 th<strong>en</strong> either Q ∗ is already suffici<strong>en</strong>tly<br />

close to the solution or the noise level is so high that all the useful information in the data is


104 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

compl<strong>et</strong>ely blurred and the data does not allow to construct a b<strong>et</strong>ter approximation of the true<br />

solution than Q ∗ .<br />

L<strong>et</strong> Q δ α <strong>de</strong>note a solution of the regularized problem (2.27) with data C δ M<br />

param<strong>et</strong>er α. The function<br />

and regularization<br />

ε δ (α) := ‖C Qδ α<br />

− C δ M‖ 2 w<br />

is called the discrepancy function of the calibration problem (2.27). Note that since this problem<br />

can have many solutions, ε δ (α) is a priori a multivalued function. Giv<strong>en</strong> two constants c 1 and<br />

c 2 satisfying<br />

the discrepancy principle can be stated as follows:<br />

1 < c 1 ≤ c 2 < εmax<br />

δ0<br />

2 , (3.12)<br />

Discrepancy principle<br />

For a giv<strong>en</strong> noise level δ, choose α > 0 that satisfies<br />

δ 2 < c 1 δ 2 ≤ ε δ (α) ≤ c 2 δ 2 , (3.13)<br />

If, for a giv<strong>en</strong> α, the discrepancy function has several possible values, the above inequalities<br />

must be satisfied by each one of them.<br />

The intuition behind this principle is as follows. We would like to find a solution Q of<br />

the equation C Q = C M . Since the error level in the data is of or<strong>de</strong>r δ, it is the best possible<br />

precision that we can ask for in this context, so it does not make s<strong>en</strong>se to calibrate the noisy<br />

data CM δ with a precision higher than δ. Therefore, we try to solve ‖C Qδ α − C<br />

δ<br />

M<br />

‖ 2 w ≤ δ 2 . In<br />

or<strong>de</strong>r to gain stability we must sacrifice some precision compared to δ, therefore, we choose a<br />

constant c with 1 c, for example, c = 1.1 and look for Q δ α in the level s<strong>et</strong><br />

‖C Qδ α<br />

− CM‖ δ 2 w ≤ cδ 2 . (3.14)<br />

Since, on the other hand, by increasing precision, we <strong>de</strong>crease the stability, the highest stability<br />

is obtained wh<strong>en</strong> the inequality in (3.14) is replaced by equality and we obtain<br />

ε δ (α) ≡ ‖C Qδ α<br />

− CM‖ δ 2 w = cδ 2 .<br />

To make the numerical solution of the equation easier, we do not impose a strict value of the<br />

discrepancy function but allow it to lie b<strong>et</strong>we<strong>en</strong> two bounds, obtaining (3.13).


3.3. CHOICE OF THE REGULARIZATION PARAMETER 105<br />

Supposing that an α solving (3.13) exists, it is easy to prove the converg<strong>en</strong>ce of the regularized<br />

solutions to the minimal <strong>en</strong>tropy least squares solution, wh<strong>en</strong> the regularization param<strong>et</strong>er<br />

is chos<strong>en</strong> using the discrepancy principle.<br />

Proposition 3.3. Suppose that the hypotheses 1–3 of page 103 are satisfied and l<strong>et</strong> c 1 and c 2 be<br />

as in (3.12). L<strong>et</strong> {C δ k<br />

M } k≥1 be a sequ<strong>en</strong>ce of data s<strong>et</strong>s such that ‖C M − C δ k<br />

M ‖ w < δ k and δ k → 0.<br />

Th<strong>en</strong> the sequ<strong>en</strong>ce {Q δ k<br />

αk<br />

} k≥1 where Q δ k<br />

αk<br />

is a solution of problem (2.27) with data C δ k<br />

M , prior<br />

P and regularization param<strong>et</strong>er α k chos<strong>en</strong> according to the discrepancy principle, has a weakly<br />

converg<strong>en</strong>t subsequ<strong>en</strong>ce. The limit of every such subsequ<strong>en</strong>ce of {Q δ k<br />

αk<br />

} k≥1 is a MELSS with<br />

data C M and prior P .<br />

Proof. Using the optimality of Q δk<br />

α k<br />

, we can write:<br />

ε δk (α k ) + α k I(Q δk<br />

α k<br />

|P ) ≤ ‖C Q+ − C δ k<br />

M ‖2 w + α k I(Q + |P ) ≤ δ 2 k + α kI(Q + |P ).<br />

The discrepancy principle (3.13) th<strong>en</strong> implies that<br />

I(Q δk<br />

α k<br />

|P ) ≤ I(Q + |P ), (3.15)<br />

By Lemma 2.10, the sequ<strong>en</strong>ce {Q δ k<br />

αk<br />

} k≥1 is tight and therefore, by Prohorov’s theorem and<br />

Lemma 2.4, relatively weakly compact in M ∩ L + B .<br />

Choose a subsequ<strong>en</strong>ce of {Q δ k<br />

αk<br />

} k≥1 , converging weakly to a limit Q and <strong>de</strong>noted, to simplify<br />

notation, again by {Q δ k<br />

αk<br />

} k≥1 . Inequality (3.13) and the triangle inequality yield:<br />

‖C Qδ k αk<br />

− C M ‖ w ≤ ‖C Qδ k αk<br />

− C δ k<br />

M ‖ w + δ k ≤ δ k (1 + √ c 2 ) −−−→<br />

k→∞ 0.<br />

By Lemma 2.2 this means that ‖C Q − C M ‖ w = 0 and therefore Q is a solution. By weak lower<br />

semicontinuity of I (cf. Corollary 2.1) and using (3.15),<br />

I(Q|P ) ≤ lim inf<br />

k<br />

which means that Q is a MELSS.<br />

I(Q δ k<br />

αk<br />

|P ) ≤ lim sup<br />

k<br />

I(Q δ k<br />

αk<br />

|P ) ≤ I(Q + |P ),<br />

The discrepancy principle performs well for regularizing linear operators but may fail in<br />

nonlinear problems like (2.27), because Equation (3.13) may have no solution due to discontinuity<br />

of the discrepancy function ε δ (α). Examples of nonlinear ill-posed problems to which the


106 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

discrepancy principle cannot be applied are giv<strong>en</strong> in [95] and [50]. However, in our numerical<br />

tests (see Section 3.6) we have always be<strong>en</strong> able to find a solution to (3.13)<br />

We will now give a simple suffici<strong>en</strong>t condition, adapted from [94], un<strong>de</strong>r which (3.13) admits<br />

a solution.<br />

Proposition 3.4. Suppose that the hypotheses 1–3 of page 103 are satisfied and l<strong>et</strong> c 1 and c 2<br />

satisfy (3.12). If ε δ (α) is a single-valued function th<strong>en</strong> there exists an α satisfying (3.13).<br />

This proposition is a direct consequ<strong>en</strong>ce of the following lemma.<br />

Lemma 3.5. The function ε δ (α) is non-<strong>de</strong>creasing and satisfies the following limit relations:<br />

lim ε δ (α) ≤ δ 2 ,<br />

α↓0<br />

lim ε δ(α) = ‖C Q∗ − C δ<br />

α→∞<br />

M‖ 2 w.<br />

If, at some point α > 0, ε δ (α) is single-valued, th<strong>en</strong> it is continuous at this point.<br />

The function<br />

J δ (α) := ‖C Qδ α<br />

− CM‖ δ 2 w + αI(Q δ α|P ).<br />

is non-<strong>de</strong>creasing, continuous, and satisfies the following limit relations:<br />

lim J δ (α) ≤ δ 2 ,<br />

α↓0<br />

lim J δ(α) ≥ ‖C Q∗ − C δ<br />

α→∞<br />

M‖ 2 w.<br />

Proof. L<strong>et</strong> γ δ (α) := I(Q δ α|P ) and l<strong>et</strong> 0 < α 1 < α 2 . By the optimality of Q δ α 1<br />

and Q δ α 2<br />

we have:<br />

and therefore<br />

ε δ (α 1 ) + α 1 γ δ (α 1 ) ≤ ε δ (α 2 ) + α 1 γ δ (α 2 ),<br />

ε δ (α 2 ) + α 2 γ δ (α 2 ) ≤ ε δ (α 1 ) + α 2 γ δ (α 1 )<br />

ε δ (α 2 ) − ε δ (α 1 ) ≥ α 1 (γ δ (α 1 ) − γ δ (α 2 ))<br />

ε δ (α 2 ) − ε δ (α 1 ) ≤ α 2 (γ δ (α 1 ) − γ δ (α 2 )),<br />

which implies that ε δ (α 2 ) ≥ ε δ (α 1 ) and γ δ (α 1 ) ≥ γ δ (α 2 ). To prove the first limit relation for<br />

ε δ (α), observe that for all α > 0,<br />

ε δ (α) ≤ ‖C Q+ − C δ M‖ 2 w + αI(Q + |P ) ≤ δ 2 + αI(Q + |P ) −−−→<br />

α→0 δ2 .<br />

To prove the second limit relation for ε δ (α), one can write, using the optimality of Q δ α:<br />

ε δ (α) + αγ δ (α) ≤ ‖C Q∗ − C δ M‖ 2 w + αI(Q ∗ |P ).


3.3. CHOICE OF THE REGULARIZATION PARAMETER 107<br />

From [30, Theorem 2.2],<br />

γ δ (α) ≥ I(Q δ α|Q ∗ ) + I(Q ∗ |P ). (3.16)<br />

Therefore<br />

Using the inequality<br />

I(Q δ α|Q ∗ ) ≤ ‖CQ∗ − C δ M ‖2 w<br />

α<br />

−−−→<br />

α→∞ 0.<br />

|P − Q| ≤ √ 2I(P |Q), (3.17)<br />

where |P − Q| <strong>de</strong>notes the total variation distance (see [30, Equation (2.3)]), this implies that<br />

Q δ α converges to Q ∗ in total variation distance (and therefore also weakly) as α goes to infinity.<br />

The limit relation now follows from Lemma 2.2.<br />

To prove the continuity of ε δ (α), l<strong>et</strong> {α n } be a sequ<strong>en</strong>ce of positive numbers, converging to<br />

α > 0. By the optimality of Q δ α n<br />

, I(Q δ α n<br />

|P ) is boun<strong>de</strong>d and one can choose a subsequ<strong>en</strong>ce of<br />

{Q δ α n<br />

}, converging weakly toward some measure Q ′ , and <strong>de</strong>noted, to simplify notation, again<br />

by {Q δ α n<br />

} n≥1 . We now need to prove that Q ′ is the solution of the calibration problem with<br />

regularization param<strong>et</strong>er α. By weak continuity of the pricing error (Lemma 2.2) and weak<br />

lower semicontinuity of the relative <strong>en</strong>tropy (Lemma 2.11), we have for any other measure Q:<br />

‖C Q′ − C δ M‖ 2 w + αI(Q ′ |P ) ≤ lim inf<br />

n {‖CQδ αn − C<br />

δ<br />

M ‖ 2 w + αI(Q δ α n<br />

|P )}<br />

= lim inf<br />

n {‖CQδ αn − C<br />

δ<br />

M ‖ 2 w + α n I(Q δ α n<br />

|P ) + (α − α n )I(Q δ α n<br />

|P )}<br />

≤ lim inf<br />

n {‖CQ − C δ M‖ 2 w + α n I(Q|P ) + (α − α n )I(Q δ α n<br />

|P )}<br />

= ‖C Q − C δ M‖ 2 w + αI(Q|P ).<br />

Therefore, the sequ<strong>en</strong>ce {ε δ (α n )} converges to one of the possible values of ε δ (α). If this function<br />

is single-valued in α, this means that every subsequ<strong>en</strong>ce of the original sequ<strong>en</strong>ce {ε δ (α n )} has<br />

a further subsequ<strong>en</strong>ce that converges toward ε δ (α), and therefore, the original sequ<strong>en</strong>ce also<br />

converges toward ε δ (α).<br />

The fact that J δ (α) is non<strong>de</strong>creasing as a function of α is trivial. To show the continuity,<br />

observe that for 0 < α 1 < α 2 ,<br />

J δ (α 2 ) − J δ (α 1 ) ≤ ε δ (α 1 ) + α 2 γ δ (α 1 ) − ε δ (α 1 ) − α 1 γ δ (α 1 )<br />

= (α 2 − α 1 )γ δ (α 1 ) ≤ (α 2 − α 1 )I(Q + |P ).


108 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

The limit relations for J δ (α) follow from the relations for ε δ (α).<br />

We will now pres<strong>en</strong>t an alternative a posteriori param<strong>et</strong>er choice rule, which reduces to the<br />

discrepancy principle wh<strong>en</strong> inequality (3.13) has a solution but also works wh<strong>en</strong> this is not<br />

the case. However, if the param<strong>et</strong>er is chos<strong>en</strong> according to the alternative rule, the sequ<strong>en</strong>ce<br />

of regularized solutions does not necessarily converge to a minimum <strong>en</strong>tropy solution as in<br />

Proposition 3.3 but to a solution with boun<strong>de</strong>d <strong>en</strong>tropy (see Proposition 3.7). Our treatm<strong>en</strong>t<br />

partly follows [50] where this param<strong>et</strong>er choice rule is applied to Tikhonov regularization.<br />

Alternative principle<br />

For a giv<strong>en</strong> noise level δ, if there exists α > 0 that satisfies<br />

c 1 δ 2 ≤ ε δ (α) ≤ c 2 δ 2 , (3.18)<br />

choose one such α; otherwise, choose an α > 0 that satisfies<br />

ε δ (α) ≤ c 1 δ 2 , J δ (α) ≥ c 2 δ 2 . (3.19)<br />

Proposition 3.6. Suppose that the hypotheses 1–3 of page 103 are satisfied and l<strong>et</strong> c 1 and c 2<br />

be as in (3.12). Th<strong>en</strong> there exists α > 0 satisfying either (3.18) or (3.19).<br />

Proof. Suppose that (3.18) does not admit a solution. We need to prove that there exists α > 0<br />

satisfying (3.19). L<strong>et</strong><br />

B := {α > 0 : ε δ (α) ≤ c 1 δ 2 } and U := {α > 0 : ε δ (α) > c 2 δ 2 }.<br />

The limit relations of Lemma 3.5 imply that both s<strong>et</strong>s are nonempty. Moreover, since we have<br />

assumed that (3.18) does not admit a solution, necessarily sup B = inf U. L<strong>et</strong> α ∗ := sup B ≡<br />

inf U. Now we need to show that<br />

By continuity of J δ (α),<br />

J δ (α ∗ ) > c 2 δ 2 . (3.20)<br />

J δ (α ∗ ) ≥ c 2 δ 2 + lim<br />

α↓α ∗ γ δ(α).<br />

If lim α↓α ∗ γ δ (α) > 0 th<strong>en</strong> (3.20) holds. Otherwise from (3.16), P is the minimal <strong>en</strong>tropy martingale<br />

measure and (3.17) implies that Q δ α ⇒ P as α ↓ α ∗ . Therefore, J δ (α ∗ ) = lim α↓α ∗ ε δ (α) =<br />

‖C Q∗ − C δ M ‖2 w > c 2 δ 2 and (3.20) is also satisfied.


3.3. CHOICE OF THE REGULARIZATION PARAMETER 109<br />

By continuity of J δ (α), there exists ∆ > 0 such that J δ (α ∗ − ∆) > c 2 δ 2 . However, since<br />

α ∗ = sup B and ε δ (α) is non<strong>de</strong>creasing, necessarily ε δ (α ∗ − ∆) ≤ c 1 δ 2 . Therefore, α − ∆ is a<br />

solution of (3.19).<br />

Remark 3.2. If c 1 < c 2 , one can show, along the lines of the above proof, that there exists not<br />

a single α that satisfies either (3.18) or (3.19) but an interval of nonzero l<strong>en</strong>gth (α 1 , α 2 ), such<br />

that each point insi<strong>de</strong> this interval satisfies one of the two conditions. From the computational<br />

viewpoint this means that a feasible α can be found by bisection with a finite number of<br />

iterations.<br />

Proposition 3.7. Suppose that the hypotheses 1–3 of page 103 are satisfied and that c 1 and c 2<br />

are chos<strong>en</strong> according to (3.12). L<strong>et</strong> {C δ k<br />

M } k≥1 be a sequ<strong>en</strong>ce of data s<strong>et</strong>s such that ‖C M −C δ k<br />

M ‖ w <<br />

δ k and δ k → 0.<br />

Th<strong>en</strong> the sequ<strong>en</strong>ce {Q δ k<br />

αk<br />

} k≥1 where Q δ k<br />

αk<br />

is a solution of problem (2.27) with data C δ k<br />

M , prior<br />

P and regularization param<strong>et</strong>er α k chos<strong>en</strong> according to the alternative principle, has a weakly<br />

converg<strong>en</strong>t subsequ<strong>en</strong>ce. The limit Q ′ of every such subsequ<strong>en</strong>ce of {Q δ k<br />

αk<br />

} k≥1 satisfies<br />

‖C Q′ − C M ‖ w = 0<br />

I(Q ′ |P ) ≤ c 2<br />

c 2 − 1 I(Q+ |P ).<br />

Proof. Using the optimality of Q δk<br />

α k<br />

, we can write:<br />

ε δk (α k ) + α k I(Q δk<br />

α k<br />

|P ) ≤ ‖C Q+ − C δ k<br />

M ‖2 w + α k I(Q + |P ) ≤ δ 2 + α k I(Q + |P ). (3.21)<br />

If α k satisfies (3.18), the above implies that<br />

otherwise (3.19) <strong>en</strong>tails that<br />

and tog<strong>et</strong>her with (3.21) this gives<br />

I(Q δk<br />

α k<br />

|P ) ≤ I(Q + |P ),<br />

ε δk (α k ) + α k I(Q δk<br />

α k<br />

|P ) ≥ c 2 δ 2 ,<br />

I(Q δk<br />

α k<br />

|P ) ≤ c 2<br />

c 2 − 1 I(Q+ |P ).<br />

In both cases I(Q δk<br />

α k<br />

|P ) is uniformly boun<strong>de</strong>d, which means, by Lemma 2.10, that the sequ<strong>en</strong>ce<br />

{Q δ k<br />

αk<br />

} k≥1 is tight and therefore, by Prohorov’s theorem, relatively weakly compact. The rest<br />

of the proof follows the lines of the proof of Proposition 3.3.


110 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

3.3.2 A posteriori param<strong>et</strong>er choice rule for non-attainable calibration problems<br />

In this subsection we suppose that hypothesis 1 of page 103 is satisfied and that hypotheses 2<br />

and 3 are replaced with the following ones:<br />

2a. There exists a solution Q + of problem (2.11) with data C M such that<br />

I(Q + |P ) < ∞<br />

and<br />

‖C Q+ − C M ‖ w > 0<br />

(the data is not attainable by an exp-Lévy mo<strong>de</strong>l).<br />

3a. There exists δ 0 such that<br />

Giv<strong>en</strong> two constants c 1 and c 2 satisfying<br />

ε max := inf<br />

δ≤δ 0<br />

{‖C Q∗ − C δ M‖ 2 w − ‖C Q+ − C δ M‖ 2 w} > 0.<br />

the discrepancy principle can be stated as follows:<br />

0 < c 1 ≤ c 2 < εmax<br />

δ0<br />

2 , (3.22)<br />

Discrepancy principle for non-attainable data For a giv<strong>en</strong> noise level δ, choose α > 0<br />

that satisfies<br />

c 1 δ 2 ≤ ε δ (α) − ε δ (0) ≤ c 2 δ 2 , (3.23)<br />

Proofs of the following two results are similar to the proofs of Propositions 3.4 and 3.3 of<br />

the preceding subsection and are therefore omitted.<br />

Proposition 3.8. Suppose that hypothesis 1 of page 103 and hypotheses 2a and 3a of page 110<br />

are satisfied and l<strong>et</strong> c 1 and c 2 satisfy (3.22). If ε δ (α) is a single-valued function th<strong>en</strong> there exists<br />

an α satisfying (3.23).<br />

Proposition 3.9. Suppose that hypothesis 1 of page 103 and hypotheses 2a and 3a of page 110<br />

are satisfied and l<strong>et</strong> c 1 and c 2 satisfy (3.22) for all δ. L<strong>et</strong> {C δ k<br />

M } k≥1 be a sequ<strong>en</strong>ce of data s<strong>et</strong>s<br />

such that ‖C M − C δ k<br />

M ‖ w ≤ δ k and δ k → 0.


3.4. CHOICE OF THE WEIGHTS 111<br />

Th<strong>en</strong> the sequ<strong>en</strong>ce {Q δ k<br />

αk<br />

} k≥1 where Q δ k<br />

αk<br />

is a solution of problem (2.27) with data C δ k<br />

M , prior<br />

P and regularization param<strong>et</strong>er α k chos<strong>en</strong> according to the discrepancy principle, has a weakly<br />

converg<strong>en</strong>t subsequ<strong>en</strong>ce. The limit of every such subsequ<strong>en</strong>ce of {Q δ k<br />

αk<br />

} k≥1 is a MELSS with<br />

data C M and prior P .<br />

The alternative principle of the preceding section does not carry over to non-attainable<br />

problems as easily and is not discussed here.<br />

3.3.3 Computing the noise level<br />

If the bid and ask prices are known, the noise level δ and an estimate of option prices C δ M can<br />

be computed directly using<br />

CM(T, δ K) := Cbid M (T, K) + Cask (T, K)<br />

, ∀ T, K,<br />

2<br />

δ := ‖Cbid M<br />

− Cask M<br />

Since for all i, the true option prices satisfy C M (T i , K i ) ∈ (C bid<br />

C M ‖ w ≤ δ.<br />

2<br />

‖ w<br />

.<br />

M , Cask M<br />

), we clearly have ‖Cδ M −<br />

If the bid and ask prices are unavailable, one can assess the or<strong>de</strong>r of magnitu<strong>de</strong> of δ directly<br />

from C δ M , using the following heuristic argum<strong>en</strong>t. Suppose that the exact data C M is attainable<br />

and that the errors in differ<strong>en</strong>t option prices are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt. Th<strong>en</strong> it is reasonable to assume<br />

that the main part of error, pres<strong>en</strong>t in C δ M<br />

will lead to violations of arbitrage constraints on<br />

option prices (e.g. convexity). Since the least squares solution Q + is an arbitrage-free mo<strong>de</strong>l,<br />

these violations of no-arbitrage constraints will contribute to the discrepancy ‖C Q+ − C δ M ‖2 w ≡<br />

ε δ (0), as shown in Figure 3.3. Therefore, ε δ (0) will have the same or<strong>de</strong>r of magnitu<strong>de</strong> as δ 2<br />

and one can approximately take δ ∼ √ ε δ (0). Since the noise level δ is only used to choose<br />

the regularization param<strong>et</strong>er α, we do not need to know it with high precision and this rough<br />

estimate is suffici<strong>en</strong>t.<br />

3.4 Choice of the weights of option prices<br />

The relative weights w i of option prices in the pricing error term (2.3) should reflect our confi<strong>de</strong>nce<br />

in individual data points, which is <strong>de</strong>termined by the liquidity of a giv<strong>en</strong> option. This


112 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

C<br />

Least squares calibration<br />

Noisy mark<strong>et</strong> data<br />

K<br />

Figure 3.3: Estimating the noise level in the data.<br />

can be assessed from the bid-ask spreads by choosing<br />

w i :=<br />

w<br />

(CM bid(T<br />

i, K i ) − CM ask(T<br />

i, K i )) 2 ,<br />

where w is the normalization constant, <strong>de</strong>termined from ∑ i w i = 1. However, the bid and<br />

ask quotes are not always available from option price data bases. On the other hand, it is<br />

known that at least for the options that are not too far from the money, the bid-ask spread<br />

in implied volatility units is of or<strong>de</strong>r of 1%. This means that to have errors proportional to<br />

bid-ask spreads, one must minimize the differ<strong>en</strong>ces of implied volatilities and not those of the<br />

option prices. However, this would prohibitively increase the computational bur<strong>de</strong>n since one<br />

would have to invert numerically the Black-Scholes formula once for each data point at each<br />

minimization step. each A feasible solution is to minimize the squared differ<strong>en</strong>ces of option<br />

prices weighted by the Black Scholes “vegas” evaluated at the points corresponding to mark<strong>et</strong><br />

option prices. D<strong>en</strong>oting the implied volatility computed in the mo<strong>de</strong>l Q for strike K and<br />

maturity T by Σ Q (T, K) and the corresponding mark<strong>et</strong> implied volatility by Σ M (T, K), we


3.5. NUMERICAL SOLUTION 113<br />

have the following approximation:<br />

N∑<br />

(Σ Q (T i , K i ) − Σ M (T i , K i )) 2<br />

i=1<br />

≈<br />

=<br />

N∑<br />

i=1<br />

( ) ∂Σ<br />

2<br />

∂C (C M(T i , K i ))|C Q (T i , K i ) − C M (T i , K i )|<br />

N∑ (C Q (T i , K i ) − C M (T i , K i )) 2<br />

Vega 2 , (3.24)<br />

(Σ M (T i , K i ))<br />

i=1<br />

where Vega <strong>de</strong>notes the <strong>de</strong>rivative of the Black-Scholes option price with respect to volatility:<br />

Vega(σ) = S √ ( ( )<br />

1 S<br />

T n<br />

σ √ T log Ke −rT + 1 )<br />

2 σ√ T ,<br />

with n <strong>de</strong>noting the CDF of the standard normal distribution. Therefore, by using w i =<br />

w<br />

Vega 2 i<br />

one can achieve the correct weighting of option prices without increasing the computational<br />

bur<strong>de</strong>n 1 (because w i can be computed in advance.)<br />

3.5 Numerical solution of calibration problem<br />

Once the (discr<strong>et</strong>e) prior P = P (A, ν P , γ P ), the regularization param<strong>et</strong>er α and the weights w i<br />

have be<strong>en</strong> fixed, it remains to find the numerical solution of the calibration problem (2.27).<br />

We construct an approximate solution of the calibration problem (2.27) by minimizing an<br />

approximation of the calibration functional J α (Q), <strong>de</strong>noted by<br />

Ĵ α (Q) := ‖C M − ĈQ ‖ 2 w + αI(Q|P ), (3.25)<br />

where ĈQ is the approximate option price <strong>de</strong>fined by Equation (3.31) below. The minimization<br />

is done over all Lévy processes Q ∈ M ∩ L + B<br />

with Lévy measures of the form (3.2). The<br />

calibration functional therefore becomes a function of a finite number of argum<strong>en</strong>ts:<br />

Ĵ α (Q) = Ĵα(q 0 , . . . , q M−1 ), (3.26)<br />

To minimize (3.26) we use a variant of the popular Broy<strong>de</strong>n-Fl<strong>et</strong>cher-Goldfarb-Shanno (BFGS)<br />

variable m<strong>et</strong>ric algorithm. For a <strong>de</strong>scription of the algorithm see [80]. This algorithm requires<br />

1 Since vegas can be very small for out of the money options, to avoid giving them too much weight one should<br />

impose a lower bound on the vegas used for weighting, i.e., use weights of the form w i =<br />

constant C ∼ 10 −2 ÷ 10 −1 .<br />

w<br />

(Vega i ∧C) 2<br />

with some


114 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

the user to supply the function value and its gradi<strong>en</strong>t and tries, iteratively, to build up a good<br />

approximation of the inverse Hessian matrix of the functional being optimized. In our numerical<br />

examples we used the LBFGS implem<strong>en</strong>tation by Jorge Nocedal <strong>et</strong> al. [20]. The algorithm is<br />

typically initialized with the prior Lévy measure.<br />

Since gradi<strong>en</strong>t-based m<strong>et</strong>hods only allow to find one local minimum of the objective function<br />

and the calibration functional (3.26) is not convex, there is no guarantee that the BFGS algorithm<br />

will converge to its true global minimum. However, starting the optimization procedure<br />

from differ<strong>en</strong>t initializers, we have empirically observed (see Section 3.6.1) that in pres<strong>en</strong>ce of<br />

ev<strong>en</strong> a small regularization, the minimizers do not <strong>de</strong>p<strong>en</strong>d on the starting point of the minimization<br />

algorithm and that using a gradi<strong>en</strong>t-based optimization procedure produces an acceptable<br />

calibration quality.<br />

In the rest of this section we show how to compute numerically the functional (3.26) and<br />

its gradi<strong>en</strong>t. For simplicity, from now on we will suppose that the prior Lévy process has a<br />

non-zero diffusion part (A > 0).<br />

3.5.1 Computing the calibration functional<br />

Substituting the discr<strong>et</strong>e expressions (3.1) and (3.2) into formula (3.25), we obtain:<br />

Ĵ α (Q) =<br />

N∑<br />

w i (ĈQ (T i , K i ) − C M (T i , K i )) 2<br />

i=1<br />

⎛<br />

⎞<br />

+ α ⎝ A M−1<br />

2A 2 + ∑<br />

bP + (e x j<br />

− 1)q j<br />

⎠<br />

j=0<br />

2<br />

M−1<br />

∑<br />

+ α (q j log(q j /p j ) + 1 − q j ) , (3.27)<br />

j=0<br />

where b P = γ P − ∫ |x|≤1 xνP (dx) is the drift of the prior process. The last two terms of the<br />

above equation can be evaluated directly using a finite number of computer operations, we will<br />

therefore conc<strong>en</strong>trate on the first term.<br />

The approximate option prices ĈQ (T i , K i ) are computed using the Fourier transform algorithm<br />

of Section 1.4 as follows: for each maturity date, pres<strong>en</strong>t in the data, prices are first<br />

computed on a fixed grid of strikes and th<strong>en</strong> interpolated to obtain prices at strikes K i corresponding<br />

to this maturity date. The number of Fourier transforms is thus <strong>de</strong>termined by the<br />

number of maturity dates pres<strong>en</strong>t in the data, which is typically smaller than 10.<br />

To use the Fourier transform m<strong>et</strong>hod, the first step is to compute the characteristic expon<strong>en</strong>t


3.5. NUMERICAL SOLUTION 115<br />

of X t un<strong>de</strong>r Q at the points u − i for real u:<br />

ψ Q (u − i) = − A M−1<br />

2 u(u − i) − (1 + iu) ∑<br />

Introducing uniform grids<br />

j=0<br />

M−1<br />

∑<br />

(e x j<br />

− 1)q j −<br />

j=0<br />

M−1<br />

∑<br />

q j +<br />

j=0<br />

e iux j<br />

e x j<br />

q j .<br />

x j = x 0 + jd and u k = u 0 − k∆,<br />

the last term becomes:<br />

M−1<br />

∑<br />

j=0<br />

e iu kx j<br />

e x j<br />

q j = e −ik∆x 0<br />

M−1<br />

∑<br />

j=0<br />

e −ikjd∆ e (iu 0+1)x j<br />

q j<br />

and comparing this to Equation (1.27), we see that if d∆ = 2π M , the last term of {ψQ (u k −i)} M−1<br />

k=0<br />

becomes a discr<strong>et</strong>e Fourier transform (cf. Equation (1.27)):<br />

ψ k := ψ Q (u k − i) = − A M−1<br />

2 u ∑<br />

k(u k − i) − (1 + iu k )<br />

j=0<br />

(e x j<br />

− 1)q j −<br />

M−1<br />

∑<br />

j=0<br />

q j<br />

+ e −ik∆x 0<br />

DFT k [e (iu 0+1)x j<br />

q j ]. (3.28)<br />

This expression can therefore be computed using the fast Fourier transform algorithm. Note<br />

that at this stage all computations are exact: there are no truncation or discr<strong>et</strong>ization errors.<br />

For a giv<strong>en</strong> maturity date T we now need to compute the Fourier transform ˜ζ T<br />

modified time value of options (see Equation (1.25)) at the points {u k } M−1<br />

k=0 :<br />

of the<br />

˜ζ T (u k ) = e iu krT eT ψ k<br />

− e − AT<br />

2 (u k(u k −i))<br />

iu k (1 + iu k )<br />

. (3.29)<br />

Option time values are th<strong>en</strong> approximated using Equation (1.28) on the same grid of log-strikes<br />

{x j } M−1<br />

j=0<br />

that was used to discr<strong>et</strong>ize the Lévy measure. 2 In the following equation and below<br />

we <strong>de</strong>note approximated quantities by putting a hat over the corresponding variables.<br />

ˆ˜z T (x j ) = ∆ M−1<br />

∑<br />

2π e−ix ju M−1<br />

k=0<br />

w k ˜ζT (u M−1−k )e −ix 0∆k e −2πijk/M<br />

= ∆ 2π e−ix ju M−1<br />

DFT j [w k ˜ζT (u M−1−k )e −ix 0∆k ], (3.30)<br />

2 Actually, the grid of log-strikes may be shifted but we do not do this here to simplify the formulas.


116 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

where w k are the weights corresponding to the chos<strong>en</strong> integration rule. For a giv<strong>en</strong> strike K<br />

such that x 0 ≤ log K ≤ x M−1 , the option price can now be computed by linear interpolation<br />

(other interpolation m<strong>et</strong>hods can also be used):<br />

Ĉ Q √<br />

(T, K) = C A<br />

BS (T, K) + log K − x n K ˆ˜z T (x nK +1) + x n K +1 − log K<br />

ˆ˜z T (x nK ), (3.31)<br />

x nK +1 − x nK x nK +1 − x nK<br />

√<br />

where n K := sup{n : x n ≤ log K} and C A<br />

BS<br />

(T, K) <strong>de</strong>notes the Black Scholes price of a call<br />

option with time to maturity T , strike K and volatility √ A.<br />

Error control The above m<strong>et</strong>hod of evaluating option prices contains three types of numerical<br />

errors: truncation and discr<strong>et</strong>ization errors appear wh<strong>en</strong> the integral is replaced by a finite sum<br />

using Equation (1.28), and interpolation error appears in Equation (3.31). Supposing that the<br />

grid in Fourier space is c<strong>en</strong>tered, that is, u 0 = L/2 and ∆ = L/(M − 1) for some constant L,<br />

Section 1.4 allows to obtain the following bounds for the first two types of error.<br />

Since we have supposed that A > 0, Equation (1.30) provi<strong>de</strong>s a uniform (with respect to Q)<br />

bound for the truncation error:<br />

|ε T | ≤<br />

16e−T<br />

AL2<br />

πT AL 3 .<br />

Proposition 1.13 does not give uniform bounds for the discr<strong>et</strong>ization (sampling) error, however,<br />

since the calibrated Lévy measures are usually similar to the Lévy measures of the prior<br />

process (see Section 3.6), the or<strong>de</strong>r of magnitu<strong>de</strong> of the discr<strong>et</strong>ization error can be estimated<br />

by computing the bounds of Proposition 1.13 for the prior process. The exact form of the error<br />

<strong>de</strong>p<strong>en</strong>ds on the integration rule used; for example for Simpson’s rule one has<br />

( L 4 )<br />

log L<br />

|ε D | = O<br />

M 4 .<br />

A uniform bound for the interpolation error <strong>de</strong>p<strong>en</strong>ds on the interpolation m<strong>et</strong>hod that one<br />

is using. For linear interpolation using Equation (3.31) one easily obtains<br />

|ε I | ≤ d2<br />

8<br />

max ˜z′′ T ,<br />

and the second <strong>de</strong>rivative of the time value is uniformly boun<strong>de</strong>d un<strong>de</strong>r the hypothesis A > 0


3.5. NUMERICAL SOLUTION 117<br />

since<br />

∣<br />

∣˜z ′′<br />

T (k) ∣ ∣ = 1<br />

2π<br />

∣<br />

∫ ∞<br />

−∞<br />

v 2 e −ivk ˜ζT (v)dv<br />

∣<br />

≤ 1 ∫ ∞<br />

|Φ T (v − i)|dv + 1<br />

2π −∞<br />

2π<br />

∫ ∞<br />

−∞<br />

√<br />

√<br />

|Φ A<br />

2<br />

T (v − i)|dv ≤ πAT ,<br />

where we have used Equation (1.25) and Lemma 1.10. Combining the above two expressions,<br />

one obtains:<br />

|ε I | ≤ 1 L 2 √<br />

π 3<br />

2AT .<br />

Taking L and M suffici<strong>en</strong>tly large so that L/M becomes small, one can make the total error<br />

ε = ε T + ε D + ε I arbitrarily small. In practice, the param<strong>et</strong>ers M and L of the discr<strong>et</strong>ization<br />

scheme should be chos<strong>en</strong> such that the total numerical error is of the same or<strong>de</strong>r as the noise<br />

level in the data. The continuity result (Theorem 2.16) th<strong>en</strong> guarantees that the numerical<br />

approximation of the calibrated measure will be close to the true solution.<br />

3.5.2 Computing the gradi<strong>en</strong>t of the calibration functional<br />

We emphasize that for the optimization algorithm to work correctly, we must compute the exact<br />

gradi<strong>en</strong>t of the approximate functional (3.26), rather than an approximation of the gradi<strong>en</strong>t of<br />

the exact functional. The gradi<strong>en</strong>t of the approximate calibration functional is computed as<br />

follows:<br />

∂Ĵα(q 0 , . . . , q M−1 )<br />

∂q k<br />

=<br />

N∑<br />

w i (ĈQ (T i , K i ) − C M (T i , K i )) ∂ĈQ (T i , K i )<br />

∂q k<br />

⎛<br />

⎞<br />

+ α(ex k<br />

− 1)<br />

⎝ A M−1<br />

2A 2 + ∑<br />

bP + (e x j<br />

− 1)q j<br />

⎠ + α log(q k /p k ).<br />

i=1<br />

The nontrivial part is therefore to compute the gradi<strong>en</strong>t of the approximate option price<br />

Ĉ Q (T i , K i ) and for this, due to the linear structure of the interpolating formula (3.31) it suffices<br />

to know the gradi<strong>en</strong>t of ˆ˜z T at the points {x i }<br />

i=0 M−1 . From Equation (3.30),<br />

[<br />

∂ˆ˜z T (x j )<br />

= ∆ ∂q m 2π e−ix ju M−1<br />

∂<br />

DFT j w ˜ζ<br />

]<br />

T (u M−1−k )<br />

k e −ix 0∆k<br />

,<br />

∂q m<br />

j=0


118 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

and formulas (3.28–3.29) yield:<br />

where ψ A k<br />

∂ ˜ζ T (u k )<br />

∂q m<br />

= T eiu krT +T ψ k {<br />

e<br />

x m+iu k x m<br />

− 1 − (1 + iu k )(e xm − 1) }<br />

iu k (1 + iu k )<br />

= T (e xm − 1) eiu krT +T ψ k<br />

+ T e xm eiukrT (e T ψ k<br />

− e T ψA k )<br />

1 + iu k iu k (1 + iu k )<br />

− T e xm eiu krT (e T ψ k<br />

− e T ψA k )<br />

iu k (1 + iu k )<br />

e iu kx m<br />

+ T e xm eiu krT +T ψk A (e iu kx m<br />

− 1)<br />

,<br />

iu k (1 + iu k )<br />

= − A<br />

2 u k(u k − i). Suppose that the grid in strike space is such that x 0 = m 0 d for<br />

some m 0 ∈ Z. Th<strong>en</strong> for every j ∈ {0, . . . , M − 1},<br />

M−1<br />

∑<br />

DFT j [f k e iu M−1−kx m<br />

] = e iu M−1x m<br />

Introducing additional notation:<br />

k=0<br />

e −2πik(j−m−m 0)/M f k<br />

H k = eiu krT +ψ k T<br />

1 + iu k<br />

G k = eiu krT +ψ A k T<br />

iu k (1 + iu k ) ,<br />

we obtain the final formula for computing the gradi<strong>en</strong>t:<br />

∂ˆ˜z T (x j )<br />

∂q m<br />

= ∆ 2π e−ix ju M−1<br />

T (e xm − 1)DFT j<br />

[<br />

w k e −ix 0∆k H M−1−k<br />

]<br />

= e iu M−1x m<br />

DFT (j−m0 −m) mod M [f k ].<br />

+ T e xm (e iu M−1x m ˆ˜z T (x (j−m0 −m) mod M ) − ˆ˜z T (x j ))<br />

+ ∆ ]<br />

2π ei(xm−x j)u M−1<br />

T e xm DFT (j−m0 −m) mod M<br />

[w k e −ix0∆k G M−1−k<br />

− ∆ [<br />

]<br />

2π e−ix ju M−1<br />

T e xm DFT j w k e −ix0∆k G M−1−k . (3.32)<br />

The time values ˆ˜z T have already be<strong>en</strong> computed in all points of the grid wh<strong>en</strong> evaluating option<br />

prices, and the Fourier transforms of G k do not need to be reevaluated at each step of the<br />

algorithm because G k do not <strong>de</strong>p<strong>en</strong>d on q i . Therefore, compared to evaluating the functional<br />

alone, to compute the gradi<strong>en</strong>t of Ĵ α one only needs one additional fast Fourier transform per<br />

maturity date. The complexity of evaluating the gradi<strong>en</strong>t using the above formula is thus only<br />

about 1.5 times higher than that of evaluating the functional itself (because evaluating option<br />

prices requires two Fourier transforms per maturity date). If the gradi<strong>en</strong>t was to be evaluated<br />

numerically, the complexity would typically be M times higher. The analytic formula (3.32)<br />

therefore allows to reduce the overall running time of the calibration algorithm from several<br />

hours to less than a minute on a standard PC.


3.6. NUMERICAL AND EMPIRICAL TESTS 119<br />

3.5.3 Overview of the algorithm<br />

Here is the final numerical algorithm, as implem<strong>en</strong>ted in the computer program levycalibration,<br />

used to run the tests of Section 3.6.<br />

• Fix the prior using one of the m<strong>et</strong>hods <strong>de</strong>scribed in Section 3.2. In the tests below, a<br />

user-specified prior was tak<strong>en</strong>.<br />

• Compute the weights of mark<strong>et</strong> option prices (Section 3.4) and estimate the noise level<br />

(Section 3.3.3).<br />

• Use one of the a posteriori m<strong>et</strong>hods of Section 3.3.1 to compute an optimal regularization<br />

param<strong>et</strong>er α ∗ achieving tra<strong>de</strong>-off b<strong>et</strong>we<strong>en</strong> precision and stability. The optimal α ∗ is computed<br />

by bisection, minimizing ˜J α several times for differ<strong>en</strong>t values of α with low precision.<br />

In the tests below we have always be<strong>en</strong> able to choose α ∗ using the discrepancy principle<br />

(3.13) with c 1 = 1.1 and c 2 = 1.3, so there was no need to resort to the alternative scheme<br />

(3.19).<br />

• Minimize ˜J α ∗ with high precision to find the regularized solution Q ∗ .<br />

3.6 Numerical and empirical tests<br />

Our numerical tests, performed using the levycalibration program, fall into two categories.<br />

First, to assess the accuracy and numerical stability of our m<strong>et</strong>hod, we tested it on option prices<br />

produced by a known expon<strong>en</strong>tial-Lévy mo<strong>de</strong>l (Section 3.6.1). We th<strong>en</strong> applied our algorithm<br />

to real options data, using the prices of European options on differ<strong>en</strong>t European stock mark<strong>et</strong><br />

indices, provi<strong>de</strong>d by Thomson Financial R and studied the properties of Lévy measures, implied<br />

by mark<strong>et</strong> data.<br />

3.6.1 Tests on simulated data<br />

A compound Poisson example: Kou’s mo<strong>de</strong>l In the first series of tests, option prices<br />

were g<strong>en</strong>erated using Kou’s jump diffusion mo<strong>de</strong>l [62] with a diffusion part with volatility<br />

σ 0 = 10% and a Lévy <strong>de</strong>nsity giv<strong>en</strong> by Equation (1.17).


120 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

0.5<br />

0.45<br />

simulated<br />

calibrated<br />

0.5<br />

0.45<br />

simulated<br />

calibrated<br />

0.4<br />

0.4<br />

0.35<br />

0.35<br />

0.3<br />

0.3<br />

0.25<br />

0.25<br />

0.2<br />

0.2<br />

0.15<br />

0.15<br />

0.1<br />

6 7 8 9 10 11 12 13 14<br />

0.1<br />

6 7 8 9 10 11 12 13 14<br />

Figure 3.4: Calibration quality (implied volatilities as a function of strike prices) for Kou’s jump<br />

diffusion mo<strong>de</strong>l. Left: no perturbations. Right: 1% perturbations were ad<strong>de</strong>d to the data<br />

The Lévy <strong>de</strong>nsity used in the tests was asymm<strong>et</strong>ric with the left tail heavier than the<br />

right one (α 1 = 1/0.07 and α 2 = 1/0.13). The jump int<strong>en</strong>sity λ was equal to 1 and the last<br />

param<strong>et</strong>er p was chos<strong>en</strong> such that the Lévy <strong>de</strong>nsity is continuous at x = 0. The option prices<br />

were computed using the Fourier transform m<strong>et</strong>hod of Section 1.4. The maturity of the options<br />

was 5 weeks and we used 21 equidistant strikes ranging from 6 to 14 (the spot being at 10). To<br />

capture tail behavior it is important to have strikes quite far in and out of the money. Merton’s<br />

jump diffusion mo<strong>de</strong>l [71] was used as prior.<br />

After g<strong>en</strong>erating s<strong>et</strong>s of call option prices from Kou’s mo<strong>de</strong>l, in the first test, the algorithm<br />

<strong>de</strong>scribed in Section 3.5.3 was applied to these prices directly, and in the second test, to mo<strong>de</strong>l<br />

the pres<strong>en</strong>ce of data errors, random perturbations of or<strong>de</strong>r of 1% of implied volatility were<br />

ad<strong>de</strong>d to the simulated prices. As can be observed from Figure 3.4, in both cases the accuracy<br />

of calibration at the level of implied volatilities is satisfying with only 21 options.<br />

Figure 3.5 compares the non-param<strong>et</strong>ric reconstruction of the Lévy <strong>de</strong>nsity to the true Lévy<br />

<strong>de</strong>nsity which, in this case, is known to be of the form (1.17). It can be observed that ev<strong>en</strong> in<br />

pres<strong>en</strong>ce of data errors we r<strong>et</strong>rieve successfully the main features of the true <strong>de</strong>nsity with our<br />

non-param<strong>et</strong>ric approach. The only region in which we observe a <strong>de</strong>tectable error is near zero:<br />

very small jumps have a small impact on option prices. The gradi<strong>en</strong>t of option’s time value<br />

giv<strong>en</strong> by Equation (3.32) vanishes at zero (i.e. for m = m 0 ) which means that the Lévy <strong>de</strong>nsity


3.6. NUMERICAL AND EMPIRICAL TESTS 121<br />

5<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Prior<br />

True<br />

Calibrated<br />

10 1 Prior<br />

True<br />

10 0<br />

Calibrated<br />

10 −1<br />

10 −2<br />

10 −3<br />

10 −4<br />

0<br />

−0.6 −0.4 −0.2 0 0.2 0.4 0.6<br />

5<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Prior<br />

True<br />

Calibrated<br />

0<br />

−0.6 −0.4 −0.2 0 0.2 0.4 0.6<br />

10 −5<br />

−0.6 −0.4 −0.2 0 0.2 0.4 0.6<br />

10 1 Prior<br />

True<br />

10 0<br />

Calibrated<br />

10 −1<br />

10 −2<br />

10 −3<br />

10 −4<br />

10 −5<br />

−0.6 −0.4 −0.2 0 0.2 0.4 0.6<br />

Figure 3.5: Lévy measure calibrated to option prices simulated from Kou’s jump diffusion mo<strong>de</strong>l<br />

with σ 0 = 10%, on linear and logarithmic scale. Top: No perturbations were ad<strong>de</strong>d. Bottom:<br />

1% perturbations were ad<strong>de</strong>d to the data.<br />

at or near zero is <strong>de</strong>termined exclusively by the prior: the int<strong>en</strong>sity of small jumps cannot be<br />

r<strong>et</strong>rieved accurately.<br />

Figure 3.6 illustrates the redundancy of small jumps and diffusion: the two graphs were<br />

calibrated to the same prices and only differ in the diffusion coeffici<strong>en</strong>ts. Comparing the two<br />

graphs shows that the algorithm comp<strong>en</strong>sates the error in the diffusion coeffici<strong>en</strong>t by adding<br />

jumps to the Lévy <strong>de</strong>nsity so that, overall, the accuracy of calibration is maintained (the<br />

standard <strong>de</strong>viation in both graphs is 0.2%). The redundancy of small jumps and diffusion<br />

compon<strong>en</strong>t has be<strong>en</strong> pointed out by other authors in the context of statistical estimation on time<br />

series [15, 69]. Here we r<strong>et</strong>rieve another version of this redundancy in a context of calibration<br />

to a cross sectional data s<strong>et</strong> of options.


122 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

5<br />

4.5<br />

prior<br />

true<br />

calibrated<br />

5<br />

4.5<br />

prior<br />

true<br />

calibrated<br />

4<br />

4<br />

3.5<br />

3.5<br />

3<br />

3<br />

2.5<br />

2.5<br />

2<br />

2<br />

1.5<br />

1.5<br />

1<br />

1<br />

0.5<br />

0.5<br />

0<br />

−0.5 0 0.5<br />

0<br />

−0.5 0 0.5<br />

Figure 3.6: Lévy measure calibrated to option prices simulated from Kou’s jump diffusion mo<strong>de</strong>l<br />

with σ 0 = 10%. Left: σ = 10.5% > σ 0 . Right: σ = 9.5% < σ 0 .<br />

The stability of the algorithm with respect to initial conditions can be tested by altering<br />

the starting point of the optimization routine and examining the effect on the output. As<br />

illustrated in Figure 3.7, the <strong>en</strong>tropy p<strong>en</strong>alty removes the s<strong>en</strong>sitivity observed in the non-linear<br />

least squares algorithm (see Figure 2.2 and Section 2.1). The only minor differ<strong>en</strong>ce b<strong>et</strong>we<strong>en</strong><br />

the two calibrated measures is observed in the neighborhood of zero i.e. the region which, as<br />

remarked above, has little influ<strong>en</strong>ce on option prices.<br />

Variance gamma mo<strong>de</strong>l In a second series of tests we examine how our m<strong>et</strong>hod performs<br />

wh<strong>en</strong> applied to option prices g<strong>en</strong>erated by an infinite int<strong>en</strong>sity process such as the variance<br />

gamma mo<strong>de</strong>l. We assume that the user, ignoring that the data g<strong>en</strong>erating process has infinite<br />

jump int<strong>en</strong>sity, chooses a (misspecified) prior which has a finite jump int<strong>en</strong>sity (e.g. the Merton<br />

mo<strong>de</strong>l).<br />

We g<strong>en</strong>erated option prices for 21 equidistant strikes b<strong>et</strong>we<strong>en</strong> 0.6 and 1.4 (the spot being at<br />

1) using the variance gamma mo<strong>de</strong>l [67] with no diffusion compon<strong>en</strong>t and applied the calibration<br />

algorithm using a Merton jump diffusion mo<strong>de</strong>l as prior to these prices. The param<strong>et</strong>ers of the<br />

variance gamma process (cf Equation (1.18)) were σ = 0.3, θ = −0.2 and k = 0.04. All the<br />

options were maturing in five weeks. The left graph in Figure 3.8 shows that ev<strong>en</strong> though the<br />

prior is misspecified, the calibrated mo<strong>de</strong>l reproduced the simulated implied volatilities with


3.6. NUMERICAL AND EMPIRICAL TESTS 123<br />

5<br />

4.5<br />

4<br />

true<br />

calibrated with λ 0<br />

=2<br />

calibrated with λ 0<br />

=1<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5 0 0.5<br />

Figure 3.7: Levy <strong>de</strong>nsities calibrated to option prices g<strong>en</strong>erated from Kou mo<strong>de</strong>l, using two<br />

differ<strong>en</strong>t initial measures with int<strong>en</strong>sities λ = 1 and λ = 2.<br />

good precision (for both values of σ 0 that we took, the standard <strong>de</strong>viation is less than 0.5% in<br />

implied volatility units).<br />

The calibrated Lévy <strong>de</strong>nsities for two differ<strong>en</strong>t values of prior diffusion volatility σ 0 are<br />

shown in the right graph of Figure 3.8: a smaller value of the volatility param<strong>et</strong>er leads to a<br />

greater int<strong>en</strong>sity of small jumps.<br />

Here we observe once again the redundancy of volatility and small jumps, this time in an<br />

infinite int<strong>en</strong>sity context. More precisely this example shows that call option prices g<strong>en</strong>erated<br />

from an infinite int<strong>en</strong>sity expon<strong>en</strong>tial Lévy mo<strong>de</strong>l can be reproduced with arbitrary precision<br />

using a jump-diffusion with finite jump int<strong>en</strong>sity. This leads us to conclu<strong>de</strong> that giv<strong>en</strong> a finite<br />

(but realistic) number of option prices, the shape of implied volatility skews/smiles does not<br />

allow to distinguish infinite activity mo<strong>de</strong>ls like variance gamma from jump diffusions.<br />

3.6.2 Empirical properties of implied Lévy measures<br />

To illustrate our calibration m<strong>et</strong>hod we have applied it to a data s<strong>et</strong> spanning the period from<br />

1999 to 2001 and containing daily prices of options on the DAX (German stock mark<strong>et</strong> in<strong>de</strong>x)<br />

for a range of strikes and maturities.<br />

Figure 3.10 shows the calibration quality for three differ<strong>en</strong>t maturities. Note that here<br />

each maturity has be<strong>en</strong> calibrated separately (three differ<strong>en</strong>t expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls). These


124 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

0.4<br />

0.38<br />

Variance gamma<br />

Calibrated with σ=0.25<br />

Calibrated with σ=0.2<br />

15<br />

True VG Levy measure<br />

Calibrated with σ=0.25<br />

Calibrated with σ=0.2<br />

0.36<br />

10<br />

0.34<br />

0.32<br />

5<br />

0.3<br />

0.28<br />

0.7 0.8 0.9 1 1.1 1.2 1.3 1.4<br />

0<br />

−0.5 0 0.5<br />

Figure 3.8: Calibration of options simulated from a variance gamma mo<strong>de</strong>l using a finite int<strong>en</strong>sity<br />

prior with differ<strong>en</strong>t values of diffusion volatility σ 0 . Left: calibration quality (implied<br />

volatilities as a function of strike prices). Right: Calibrated Lévy measures. Increasing the<br />

diffusion coeffici<strong>en</strong>t <strong>de</strong>creases the int<strong>en</strong>sity of small jumps in the calibrated measure.<br />

graphs show that using an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l one can calibrate with high precision the<br />

prices of a s<strong>et</strong> of options with common maturity. This conclusion contradicts the findings of<br />

Medve<strong>de</strong>v and Scaill<strong>et</strong> [70] who observe that “jumps in r<strong>et</strong>urns do not help explaining the slope<br />

of implied volatilities at the money”. It is important to note, however, that Medve<strong>de</strong>v and<br />

Scaill<strong>et</strong> used short term options on S&P 500 in<strong>de</strong>x whereas our analysis is based on DAX in<strong>de</strong>x<br />

options.<br />

Figure 3.9, left graph, illustrates the typical shape of risk neutral Lévy <strong>de</strong>nsities obtained<br />

from our data s<strong>et</strong>. It is readily se<strong>en</strong> that the Lévy measures obtained are far from being<br />

symm<strong>et</strong>ric: the distribution of jump sizes is highly skewed toward negative values. Figure 3.11<br />

shows the same result across cal<strong>en</strong>dar time, showing that this asymm<strong>et</strong>ry persists across time.<br />

This effect also <strong>de</strong>p<strong>en</strong>ds on the maturity of options in question: for longer maturities (see Figure<br />

3.11, right graph) the support of the Lévy measure ext<strong>en</strong>ds further to the left. Most of the<br />

<strong>de</strong>nsities we obtained are bimodal with one mo<strong>de</strong> corresponding to small jumps that are hard<br />

to separate from a diffusion compon<strong>en</strong>t and another clearly distinguishable mo<strong>de</strong> corresponding<br />

to a large negative jump that can reflect the mark<strong>et</strong> participants’ “fear of crash” [11].<br />

The logarithmic scale in the right graph of Figure 3.9 allows the tails to be se<strong>en</strong> more clearly.


3.6. NUMERICAL AND EMPIRICAL TESTS 125<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Maturity 8 days<br />

Maturity 36 days<br />

Maturity 71 days<br />

Prior<br />

10 1 Maturity 8 days<br />

Maturity 36 days<br />

10 0 Maturity 71 days<br />

Prior<br />

10 −1<br />

10 −2<br />

10 −3<br />

10 −4<br />

10 −5<br />

0<br />

−1 −0.5 0 0.5<br />

10 −6<br />

−1 −0.5 0 0.5<br />

Figure 3.9: Lévy measures calibrated to DAX options on the same cal<strong>en</strong>dar date for three<br />

differ<strong>en</strong>t maturities, linear and logarithmic scale.<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

Maturity 8 days, mark<strong>et</strong><br />

Maturity 8 days, mo<strong>de</strong>l<br />

Maturity 36 days, mark<strong>et</strong><br />

Maturity 36 days, mo<strong>de</strong>l<br />

Maturity 71 days, mark<strong>et</strong><br />

Maturity 71 days, mo<strong>de</strong>l<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

3500 4000 4500 5000 5500 6000 6500 7000 7500<br />

Figure 3.10: Calibration quality for differ<strong>en</strong>t maturities: mark<strong>et</strong> implied volatilities for DAX<br />

options against mo<strong>de</strong>l implied volatilities. Each maturity has be<strong>en</strong> calibrated separately.


126 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

5<br />

4.5<br />

4<br />

3.5<br />

11 May 2001, 8 days<br />

11 June 2001, 4 days<br />

11 July 2001, 9 days<br />

Prior<br />

4<br />

3.5<br />

3<br />

11 May 2001, 36 days<br />

11 June 2001, 39 days<br />

11 July 2001, 37 days<br />

Prior<br />

3<br />

2.5<br />

2.5<br />

2<br />

2<br />

1.5<br />

1.5<br />

1<br />

1<br />

0.5<br />

0.5<br />

0<br />

−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4<br />

0<br />

−1 −0.8 −0.6 −0.4 −0.2 0 0.2<br />

Figure 3.11: Results of calibration at differ<strong>en</strong>t dates for shortest (left) and second shortest<br />

(right) maturity. DAX in<strong>de</strong>x options.<br />

Recall that the prior <strong>de</strong>nsity is Gaussian, which shows up as a symm<strong>et</strong>ric parabola on log scales.<br />

The area un<strong>de</strong>r the curves shown here is to be interpr<strong>et</strong>ed as the (risk neutral) jump int<strong>en</strong>sity.<br />

While the shape of the curve does vary slightly across cal<strong>en</strong>dar time, the int<strong>en</strong>sity stays<br />

surprisingly stable: its numerical value is empirically found to be λ ≃ 1, which means around<br />

one jump a year. Of course note that this is the risk neutral int<strong>en</strong>sity: jump int<strong>en</strong>sities are not<br />

invariant un<strong>de</strong>r equival<strong>en</strong>t change of measures. Moreover this illustrates that a small int<strong>en</strong>sity<br />

of jumps λ can be suffici<strong>en</strong>t for explaining the shape of the implied volatility skew for small<br />

maturities. Therefore the motivation of infinite activity processes from the point of view of<br />

option pricing does not seem clear to us. On the other hand, from the mo<strong>de</strong>lling perspective<br />

infinite int<strong>en</strong>sity processes do pres<strong>en</strong>t a particular interest since they do not have a diffusion<br />

compon<strong>en</strong>t and therefore do not suffer from the problem of redundancy of small jumps and<br />

diffusion, <strong>de</strong>scribed above, leading to more parsimonious and i<strong>de</strong>ntifiable mo<strong>de</strong>ls.<br />

3.6.3 Testing time homog<strong>en</strong>eity<br />

While the literature on jump processes in finance has focused on time homog<strong>en</strong>eous (Lévy)<br />

mo<strong>de</strong>ls, practitioners have t<strong>en</strong><strong>de</strong>d to use time <strong>de</strong>p<strong>en</strong><strong>de</strong>nt jump or volatility param<strong>et</strong>ers. Despite<br />

the fact that, as our non-param<strong>et</strong>ric analysis and several empirical studies by other authors<br />

[21, 67] have shown, Lévy processes reproduce the implied volatility smile for a single maturity


3.6. NUMERICAL AND EMPIRICAL TESTS 127<br />

0.28<br />

Implied volatility<br />

0.26<br />

0.24<br />

0.22<br />

0.2<br />

0.18<br />

0.16<br />

0<br />

Maturity<br />

0.5<br />

1<br />

7000<br />

6500<br />

6000<br />

Strike<br />

5500<br />

5000<br />

0.3<br />

Implied volatility<br />

0.28<br />

0.26<br />

0.24<br />

0.22<br />

0.2<br />

0.18<br />

0.16<br />

0<br />

Maturity<br />

0.5<br />

1<br />

7000<br />

6500<br />

6000<br />

Strike<br />

5500<br />

Implied volatility<br />

0.28<br />

0.26<br />

0.24<br />

0.22<br />

0.2<br />

0.18<br />

0.16<br />

0<br />

5000<br />

0.5<br />

Maturity<br />

1<br />

7000<br />

6500<br />

6000<br />

Strike<br />

5500<br />

5000<br />

Figure 3.12: Top: Mark<strong>et</strong> implied volatility surface. Bottom left: implied volatility surface in<br />

an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l, calibrated to mark<strong>et</strong> prices of the first maturity. Bottom right:<br />

implied volatility surface in an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l, calibrated to mark<strong>et</strong> prices of the last<br />

maturity.


128 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

quite well, wh<strong>en</strong> it comes to calibrating several maturities at the same time, the calibration by<br />

Lévy processes becomes much less precise. This is clearly se<strong>en</strong> from the three graphs of Figure<br />

3.12. The top graph shows the mark<strong>et</strong> implied volatilities for four maturities and differ<strong>en</strong>t<br />

strikes. The bottom left graphs <strong>de</strong>picts implied volatilities, computed in an expon<strong>en</strong>tial Lévy<br />

mo<strong>de</strong>l calibrated using our nonparam<strong>et</strong>ric algorithm to the first maturity pres<strong>en</strong>t in the mark<strong>et</strong><br />

data. One can see that while the calibration quality is acceptable for the first maturity, it<br />

quickly <strong>de</strong>teriorates as the time to maturity increases: the smile in an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l<br />

flatt<strong>en</strong>s too fast. The same effect can be observed in the bottom right graph: here, the mo<strong>de</strong>l<br />

was calibrated to the last maturity, pres<strong>en</strong>t in the data. As a result, the calibration quality is<br />

poor for the first maturity: the smile in an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l is more pronounced and its<br />

shape does not resemble that of the mark<strong>et</strong>.<br />

It is difficult to calibrate an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l to options of several maturities because<br />

due to in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce and stationarity of their increm<strong>en</strong>ts, Lévy processes have a very rigid term<br />

structure of cumulants. In particular, the skewness of a Lévy process is proportional to the<br />

inverse square root of time and the excess kurtosis is inversely proportional to time [68]. A<br />

number of empirical studies have compared the term structure of skewness and kurtosis implied<br />

in mark<strong>et</strong> option prices to the skewness and kurtosis of Lévy processes. Bates [13], after an<br />

empirical study of implicit kurtosis in $/DM exchange rate options conclu<strong>de</strong>s that “while<br />

implicit excess kurtosis does t<strong>en</strong>d to increase as option maturity shrinks, . . . , the magnitu<strong>de</strong><br />

of maturity effects is not as large as predicted [by a Lévy mo<strong>de</strong>l]”. For stock in<strong>de</strong>x options,<br />

Madan and Konikov [68] report ev<strong>en</strong> more surprising results: both implied skewness and kurtosis<br />

actually <strong>de</strong>crease as the l<strong>en</strong>gth of the holding period becomes smaller. It should be m<strong>en</strong>tioned,<br />

however, that implied skewness/kurtosis cannot be computed from a finite number of option<br />

prices with high precision.<br />

Our non-param<strong>et</strong>ric approach allows to investigate time homog<strong>en</strong>eity by calibrating the Lévy<br />

measure separately to various option maturities. Figure 3.9 shows Lévy measures obtained by<br />

running the algorithm for options of differ<strong>en</strong>t maturity. The hypothesis of time homog<strong>en</strong>eity<br />

would imply that all the curves are the same, which is appar<strong>en</strong>tly not the case here. However,<br />

computing the areas un<strong>de</strong>r the curves yields similar jump int<strong>en</strong>sities across maturities: this<br />

result can be interpr<strong>et</strong>ed by saying that the risk neutral jump int<strong>en</strong>sity is relatively stable<br />

through time while the shape of the (normalized) jump size <strong>de</strong>nsity can actually change. The


3.6. NUMERICAL AND EMPIRICAL TESTS 129<br />

50<br />

30−day ATM options<br />

450−day ATM options<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

2/01/1996 2/01/1997 2/01/1998 30/12/1998<br />

Figure 3.13: Implied volatility of at the money European options on CAC40 in<strong>de</strong>x.<br />

type of time <strong>de</strong>p<strong>en</strong><strong>de</strong>nce is therefore more complicated than simply a time <strong>de</strong>p<strong>en</strong><strong>de</strong>nt int<strong>en</strong>sity.<br />

A second major difficulty arising while trying to calibrate an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l is the<br />

time evolution of the smile. Expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls belong to the class of so called “sticky<br />

moneyness” mo<strong>de</strong>ls, meaning that in an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l, the implied volatility of an<br />

option with giv<strong>en</strong> moneyness (strike price to spot ratio) does not <strong>de</strong>p<strong>en</strong>d on time. This can<br />

be se<strong>en</strong> from the following simple argum<strong>en</strong>t.<br />

In an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l Q, the implied<br />

volatility σ of a call option with moneyness m, expiring in τ years, satisfies:<br />

e −rτ E Q [(S t e rτ+Xτ − mS t ) + |F t ] = e −rτ σ2<br />

rτ+σWτ −<br />

E[(S t e 2 τ − mS t ) + |F t ]<br />

Due to the in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt increm<strong>en</strong>ts property, S t cancels out and we obtain an equation for<br />

the implied volatility σ which does not contain t or S t . Therefore, in an exp-Lévy mo<strong>de</strong>l this<br />

implied volatility does not <strong>de</strong>p<strong>en</strong>d on date t or stock price S t . This means that once the smile<br />

has be<strong>en</strong> calibrated for a giv<strong>en</strong> date t, its shape is fixed for all future dates. Wh<strong>et</strong>her or not this<br />

is true in real mark<strong>et</strong>s can be tested in a mo<strong>de</strong>l-free way by looking at the implied volatility<br />

of at the money options with the same maturity for differ<strong>en</strong>t dates. Figure 3.13 <strong>de</strong>picts the<br />

behavior of implied volatility of two at the money options on the CAC40 in<strong>de</strong>x, expiring in<br />

30 and 450 days. Since the maturities of available options are differ<strong>en</strong>t for differ<strong>en</strong>t dates, to<br />

obtain the implied volatility of an option with fixed maturity T for each date, we have tak<strong>en</strong><br />

two maturities, pres<strong>en</strong>t in the data, closest to T from above and below: T 1 ≤ T and T 2 > T .


130 CHAPTER 3. NUMERICAL IMPLEMENTATION<br />

The implied volatility Σ(T ) of the hypoth<strong>et</strong>ical option with maturity T was th<strong>en</strong> interpolated<br />

using the following formula:<br />

Σ 2 (T ) = Σ 2 (T 1 ) T 2 − T<br />

T 1 − T + Σ2 (T 2 ) T − T 1<br />

T 2 − T 1<br />

.<br />

As we have se<strong>en</strong>, in an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l the implied volatility of an option which is at<br />

the money and has fixed maturity must not <strong>de</strong>p<strong>en</strong>d on time or stock price. Figure 3.13 shows<br />

that in reality this is not so: both graphs are rapidly varying random functions.<br />

This simple test shows that real mark<strong>et</strong>s do not have the “sticky moneyness” property:<br />

arrival of new information can alter the form of the smile. The expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls are<br />

therefore “not random <strong>en</strong>ough” to account for the time evolution of the smile.<br />

Moreover,<br />

mo<strong>de</strong>ls based on additive processes, that is, time-inhomog<strong>en</strong>eous processes with in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt<br />

increm<strong>en</strong>ts, although they perform well in calibrating the term structure of implied volatilities<br />

for a giv<strong>en</strong> date [27], are not likely to <strong>de</strong>scribe the time evolution of the smile correctly since in<br />

these mo<strong>de</strong>ls the future form of the smile is still a <strong>de</strong>terministic function of its pres<strong>en</strong>t shape<br />

[27]. To <strong>de</strong>scribe the time evolution of the smile in a consist<strong>en</strong>t manner, one may need to<br />

introduce additional stochastic factors (e.g. stochastic volatility) [9, 10, 22].<br />

One of the important conclusions of this chapter is that at least in the domain of stock in<strong>de</strong>x<br />

options, to which we have applied our tests, expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls, although they provi<strong>de</strong> a<br />

consi<strong>de</strong>rable improvem<strong>en</strong>t compared to Black-Scholes mo<strong>de</strong>l, do not allow to calibrate the term<br />

structure of implied volatilities with suffici<strong>en</strong>t precision. Our calibration algorithm therefore<br />

should not be se<strong>en</strong> as a m<strong>et</strong>hod to find an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l consist<strong>en</strong>t with all available<br />

mark<strong>et</strong> data but rather as a way to construct a Lévy measure, implied by option prices of a<br />

particular maturity. It can therefore be ext<strong>en</strong><strong>de</strong>d to other mo<strong>de</strong>ls where jumps are repres<strong>en</strong>ted,<br />

in a nonparam<strong>et</strong>ric way, by a Lévy measure. Its many possible applications inclu<strong>de</strong>, among<br />

others,<br />

• Interpolation of option prices for a single maturity;<br />

• Calibration, using short-maturity data, of the jump part of hybrid mo<strong>de</strong>ls, including both<br />

jumps and stochastic volatility;<br />

• Separate calibration of Lévy <strong>de</strong>nsities for differ<strong>en</strong>t maturities in or<strong>de</strong>r to <strong>de</strong>termine the<br />

correct pattern of time <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of Lévy measure for additive processes.


Part II<br />

Multidim<strong>en</strong>sional mo<strong>de</strong>lling with<br />

Lévy processes<br />

131


Chapter 4<br />

Characterization of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of<br />

multidim<strong>en</strong>sional Lévy processes<br />

4.1 Introduction to <strong>de</strong>p<strong>en</strong><strong>de</strong>nce mo<strong>de</strong>lling<br />

Many financial applications require a multidim<strong>en</strong>sional mo<strong>de</strong>l with jumps, taking into account<br />

the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> compon<strong>en</strong>ts. However, such mo<strong>de</strong>ls are more difficult to construct than<br />

one-dim<strong>en</strong>sional ones and the applications continue to be dominated by (geom<strong>et</strong>ric) Brownian<br />

motion.<br />

A simple m<strong>et</strong>hod to introduce jumps into a multidim<strong>en</strong>sional mo<strong>de</strong>l is to take a multivariate<br />

Brownian motion and time change it with a one-dim<strong>en</strong>sional increasing Lévy process.<br />

This approach, advocated in [36, 79], allows to construct multidim<strong>en</strong>sional versions of many<br />

popular one-dim<strong>en</strong>sional mo<strong>de</strong>ls, including variance gamma, normal inverse Gaussian and g<strong>en</strong>eralized<br />

hyperbolic process. The principal advantage of this m<strong>et</strong>hod is its simplicity and analytic<br />

tractability; in particular, processes of this type are easy to simulate. However, the range of<br />

<strong>de</strong>p<strong>en</strong><strong>de</strong>nce patterns that one can obtain using this approach is quite limited (for instance, in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

is not inclu<strong>de</strong>d), and all compon<strong>en</strong>ts must follow the same param<strong>et</strong>ric mo<strong>de</strong>l (e.g.,<br />

either all of the compon<strong>en</strong>ts are variance gamma or all of the compon<strong>en</strong>ts are normal inverse<br />

Gaussian <strong>et</strong>c.)<br />

To be more specific, suppose that two stock price processes {St 1 } t≥0 and {St 2 } t≥0 are mod-<br />

133


134 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

elled as follows:<br />

S 1 t = exp(X 1 t ), X 1 t = B 1 (Z t ) + µ 1 Z t ,<br />

S 2 t = exp(X 2 t ), X 2 t = B 2 (Z t ) + µ 2 Z t ,<br />

where B 1 and B 2 are two compon<strong>en</strong>ts of a planar Brownian motion, with variances σ 2 1 and<br />

σ 2 2 and correlation coeffici<strong>en</strong>t ρ, and {Z t} t≥0 is the stochastic time change (an increasing Lévy<br />

process). The correlation of r<strong>et</strong>urns, ρ(X 1 t , X 2 t ), can be computed by conditioning with respect<br />

to Z t :<br />

ρ(X 1 t , X 2 t ) =<br />

σ 1 σ 2 ρE[Z t ] + µ 1 µ 2 Var Z t<br />

(σ 2 1 E[Z t] + µ 2 1 Var Z t) 1/2 (σ 2 2 E[Z t] + µ 2 2 Var Z t) 1/2 .<br />

In the compl<strong>et</strong>ely symm<strong>et</strong>ric case (µ 1 = µ 2 = 0) and in this case only ρ(X 1 t , X 2 t ) = ρ: the<br />

correlation of r<strong>et</strong>urns equals the correlation of Brownian motions that are being subordinated.<br />

However, the distributions of real stocks are skewed and in the skewed case the correlation of<br />

r<strong>et</strong>urns will be differ<strong>en</strong>t from the correlation of Brownian motions that we put into the mo<strong>de</strong>l.<br />

Ev<strong>en</strong> if the Brownian motions are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt, the covariance of r<strong>et</strong>urns is equal to µ 1 µ 2 Var Z t<br />

and if the distributions of stocks are not symm<strong>et</strong>ric, they are correlated.<br />

In the symm<strong>et</strong>ric case, if Brownian motions are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt, the two stocks are <strong>de</strong>correlated<br />

but not in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt. Since the compon<strong>en</strong>ts of the Brownian motion are time changed with<br />

the same subordinator, large jumps in the two stocks (that correspond to large jumps of the<br />

subordinator) will t<strong>en</strong>d to arrive tog<strong>et</strong>her, which means that absolute values of r<strong>et</strong>urns will be<br />

correlated. If µ 1 = µ 2 = 0 and ρ = 0 th<strong>en</strong> the covariance of squares of r<strong>et</strong>urns is<br />

Cov((X 1 t ) 2 , (X 2 t ) 2 ) = σ 1 σ 2 Cov((B 1 (Z t )) 2 , (B 2 (Z t )) 2 ) = σ 1 σ 2 Var Z t ,<br />

which means that squares of r<strong>et</strong>urns are correlated unless the subordinator Z t is <strong>de</strong>terministic.<br />

This ph<strong>en</strong>om<strong>en</strong>on can lead to mispricing and errors in evaluation of risk measures.<br />

In finite activity mo<strong>de</strong>ls, a more accurate mo<strong>de</strong>lling of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce may be achieved by<br />

specifying directly the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of individual jumps in one-dim<strong>en</strong>sional compound Poisson<br />

processes (see [65]). This approach is useful in pres<strong>en</strong>ce of few sources of jump risk (e.g., wh<strong>en</strong> all<br />

compon<strong>en</strong>ts jump at the same time) because in this case it allows to achieve a precise <strong>de</strong>scription<br />

of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce within a simple mo<strong>de</strong>l. Suppose that we want to improve a d-dim<strong>en</strong>sional Black-<br />

Scholes mo<strong>de</strong>l by allowing for “mark<strong>et</strong> crashes.” The dates of mark<strong>et</strong> crashes can be mo<strong>de</strong>lled


4.1. INTRODUCTION 135<br />

as jump times of a standard Poisson process {N t } t≥0 . This leads us to the following mo<strong>de</strong>l for<br />

the log-price processes of d ass<strong>et</strong>s:<br />

∑N t<br />

Xt i = µ i t + Bt i + Yj i , i = 1 . . . d,<br />

j=1<br />

where (B t ) is a d-dim<strong>en</strong>sional Brownian motion with covariance matrix Σ, and {Y j } ∞ j=1 are<br />

i.i.d.<br />

d-dim<strong>en</strong>sional random vectors which <strong>de</strong>termine the sizes of jumps in individual ass<strong>et</strong>s<br />

during a mark<strong>et</strong> crash. This mo<strong>de</strong>l contains only one driving Poisson shock because we only<br />

account for jump risk of one type (global mark<strong>et</strong> crash affecting all ass<strong>et</strong>s). To construct a<br />

param<strong>et</strong>ric mo<strong>de</strong>l, we need to specify the distribution of jumps in individual ass<strong>et</strong>s during a<br />

crash (distribution of Y i<br />

∗<br />

a simplifying assumption that {Y i<br />

j }d i=1<br />

for all i) and the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> jumps in ass<strong>et</strong>s. If we make<br />

are Gaussian random vectors, th<strong>en</strong> we need to specify<br />

their covariance matrix Σ ′ and the mean vector m, thus obtaining a multivariate version of<br />

Merton’s mo<strong>de</strong>l [71].<br />

If the jumps are not Gaussian, we must specify the distribution of jumps in each compon<strong>en</strong>t<br />

and the copula 1 <strong>de</strong>scribing their <strong>de</strong>p<strong>en</strong><strong>de</strong>nce.<br />

The mo<strong>de</strong>l is thus compl<strong>et</strong>ely specified by a<br />

covariance matrix Σ, d jump size distributions, a d-dim<strong>en</strong>sional copula C and a jump int<strong>en</strong>sity<br />

param<strong>et</strong>er λ. However, som<strong>et</strong>imes it is necessary to have several in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt shocks to account<br />

for ev<strong>en</strong>ts that affect individual companies or individual sectors rather than the <strong>en</strong>tire mark<strong>et</strong>.<br />

In this case we need to introduce several driving Poisson processes into the mo<strong>de</strong>l, which now<br />

takes the following form:<br />

where N 1 t , . . . , N M t<br />

X i t = µ i t + B i t +<br />

N M∑ ∑t<br />

k Yj,k i , i = 1 . . . d,<br />

k=1 j=1<br />

are Poisson processes driving M in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt shocks and Y i<br />

j,k<br />

is the size of<br />

jump in i-th compon<strong>en</strong>t after j-th shock of type k. The vectors {Yj,k i }d i=1 for differ<strong>en</strong>t j and/or k<br />

are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt. To <strong>de</strong>fine a param<strong>et</strong>ric mo<strong>de</strong>l compl<strong>et</strong>ely, one must specify a one-dim<strong>en</strong>sional<br />

distribution for each compon<strong>en</strong>t for each shock type — because differ<strong>en</strong>t shocks influ<strong>en</strong>ce the<br />

same stock in differ<strong>en</strong>t ways — and one d-dim<strong>en</strong>sional copula for each shock type. This adds up<br />

to M × d one-dim<strong>en</strong>sional distributions and M one-dim<strong>en</strong>sional copulas. How many differ<strong>en</strong>t<br />

1 For an introduction to copulas see [76] — this monograph treats mostly the bivariate case — and [56] for<br />

the multivariate case.


136 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

shocks do we need to <strong>de</strong>scribe suffici<strong>en</strong>tly rich <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures? The answer <strong>de</strong>p<strong>en</strong>ds<br />

on the particular problem, but to <strong>de</strong>scribe all possible <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures, such that the<br />

d-dim<strong>en</strong>sional process remains a Lévy process of compound Poisson type, one needs a total of<br />

2 M − 1 shocks (M shocks that affect only one stock, M(M−1)<br />

2<br />

shocks that affect two stocks <strong>et</strong>c.,<br />

adding up to 2 M − 1). It is clear that as the dim<strong>en</strong>sion of the problem grows, this kind of<br />

mo<strong>de</strong>lling quickly becomes infeasible. Not only the number of param<strong>et</strong>ers grows expon<strong>en</strong>tially,<br />

but also, wh<strong>en</strong> the number of shocks is greater than one, one cannot specify directly the laws<br />

of compon<strong>en</strong>ts because the laws of jumps must be giv<strong>en</strong> separately for each shock.<br />

Comparison of advantages and drawbacks of these two m<strong>et</strong>hods leads to an un<strong>de</strong>rstanding<br />

of the required properties of a multidim<strong>en</strong>sional mo<strong>de</strong>lling approach for Lévy processes.<br />

g<strong>en</strong>eral, such an approach must satisfy the following conditions:<br />

• One should be able to choose any one-dim<strong>en</strong>sional Lévy process for each of the compon<strong>en</strong>ts.<br />

In particular, it should be possible to couple a compound Poisson process with<br />

a process which has infinite jump int<strong>en</strong>sity. This is particularly important for financial<br />

applications because information about margins and about <strong>de</strong>p<strong>en</strong><strong>de</strong>nce may come from<br />

differ<strong>en</strong>t sources. For instance, for pricing bask<strong>et</strong> options the mark<strong>et</strong> practice is to estimate<br />

the correlations (<strong>de</strong>p<strong>en</strong><strong>de</strong>nce) from historical data while using implied volatilities,<br />

computed from the prices of tra<strong>de</strong>d options.<br />

In<br />

• The range of possible <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures should inclu<strong>de</strong> compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce and<br />

compl<strong>et</strong>e in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce with a “smooth” transition b<strong>et</strong>we<strong>en</strong> these two extremes.<br />

• It should be possible to mo<strong>de</strong>l <strong>de</strong>p<strong>en</strong><strong>de</strong>nce in a param<strong>et</strong>ric fashion (e.g., not through<br />

the <strong>en</strong>tire multidim<strong>en</strong>sional Lévy measure). A parsimonious <strong>de</strong>scription of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

is especially important because one typically does not have <strong>en</strong>ough information about<br />

the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure to estimate many param<strong>et</strong>ers or proceed with a nonparam<strong>et</strong>ric<br />

approach.<br />

To implem<strong>en</strong>t this program, we suggest to mo<strong>de</strong>l separately the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure of<br />

a Lévy process and the behavior of its compon<strong>en</strong>ts (margins). It has long be<strong>en</strong> known that<br />

the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure of a random vector on R d can be dis<strong>en</strong>tangled from its margins<br />

via the notion of copula.<br />

Since the law of a Lévy process is compl<strong>et</strong>ely <strong>de</strong>termined by its


4.1. INTRODUCTION 137<br />

distribution at time t for any fixed t > 0, the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure of a multidim<strong>en</strong>sional<br />

Lévy process X ≡ {X i t} i=1...d<br />

t≥0<br />

can be param<strong>et</strong>rized by the copula C t of the random vector<br />

{X i t} i=1...d for some t > 0. However, this approach has a number of drawbacks. First, for giv<strong>en</strong><br />

infinitely divisible one-dim<strong>en</strong>sional laws µ 1 t , . . . , µ d t , it is unclear, which copulas C t will yield a<br />

d-dim<strong>en</strong>sional infinitely divisible law. Second, the copula C t may <strong>de</strong>p<strong>en</strong>d on t and C s for some<br />

s ≠ t cannot in g<strong>en</strong>eral be computed from C t alone; to compute it one also needs to know the<br />

marginal distributions at time t and at time s. We will now construct an explicit example of a<br />

Lévy process with nonconstant copula.<br />

Example 4.1. L<strong>et</strong> Z := {Z t } t≥0 be a 2-dim<strong>en</strong>sional Cauchy process, that is, a Lévy process with<br />

characteristic tripl<strong>et</strong> (0, ν Z , 0), where ν Z has a <strong>de</strong>nsity, also <strong>de</strong>noted by ν Z , giv<strong>en</strong> by<br />

ν Z (x, y) =<br />

1<br />

(x 2 + y 2 ) 3/2 .<br />

The probability distribution of Z t for every t > 0 has a <strong>de</strong>nsity<br />

p Z t (x, y) = 1 t<br />

2π {(x 2 + y 2 ) 2 + t 2 } 3/2 .<br />

The copula of Z t does not <strong>de</strong>p<strong>en</strong>d on time and can be computed explicitly:<br />

⎧ ⎫<br />

C Z (u, v) = − 1 4 + u 2 + v 2 + 1 ⎨<br />

2π arctan tan π(u − 1 2 ) tan π(v − 1 2 ) ⎬<br />

⎩<br />

√1 + tan 2 π(u − 1 2 ) + tan2 π(v − 1 2 ) ⎭ ,<br />

which is clearly differ<strong>en</strong>t from the in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce copula C ⊥ (u, v) = uv.<br />

L<strong>et</strong> W := {W t } t≥0 be a standard planar Brownian motion, in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt from Z. Since the<br />

compon<strong>en</strong>ts of W are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt, the copula of W t for each t > 0 is the in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce copula<br />

C ⊥ (u, v) = uv. For every t, l<strong>et</strong> X t = Z t + W t . {X t } t≥0 is a Lévy process with characteristic<br />

tripl<strong>et</strong> (Id 2 , ν Z , 0).<br />

X t √t<br />

Since Z is a 1-stable process and W is 1 2 -stable, X t<br />

t<br />

d = Z √<br />

t<br />

+ W 1 . Therefore, the random variable Xt<br />

t<br />

d<br />

= Z 1 + W 1/t and<br />

is infinitely divisible with characteristic<br />

tripl<strong>et</strong> ( 1 t Id 2, 0, ν Z ) and the random variable Xt √<br />

t<br />

is infinitely divisible with characteristic tripl<strong>et</strong><br />

(Id 2 , 0, √ tν Z ). From Proposition 1.7 it follows that Xt d<br />

t<br />

−−−→ Z 1 and Xt d<br />

√<br />

t→∞<br />

t<br />

−−→ W 1 . Since the<br />

t→0<br />

copula is invariant with respect to transformations of margins by strictly increasing functions,<br />

X t , Xt √<br />

t<br />

and Xt<br />

t<br />

have the same copula. Therefore, Theorem 2.1 in [64] implies that the copula<br />

C t of X t has the following properties<br />

∀u, v, C t (u, v) → C Z (u, v) as t → ∞,<br />

∀u, v, C t (u, v) → C ⊥ (u, v) as t → 0.


138 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

Since C Z (u, v) is differ<strong>en</strong>t from C ⊥ (u, v), we conclu<strong>de</strong> that C t is not constant over time.<br />

Every Lévy process X := {X t } t≥0 is <strong>de</strong>scribed in a time-<strong>de</strong>p<strong>en</strong><strong>de</strong>nt fashion by its characteristic<br />

tripl<strong>et</strong> (A, ν, γ). It seems therefore natural to <strong>de</strong>scribe the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> compon<strong>en</strong>ts<br />

of X also in terms of its characteristic tripl<strong>et</strong>. Since the continuous martingale compon<strong>en</strong>t of<br />

X is compl<strong>et</strong>ely <strong>de</strong>scribed by the covariance matrix A and is in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt from the jump part,<br />

we will focus on the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of the jump part of X, that is, we only consi<strong>de</strong>r Lévy processes<br />

with A = 0. For such a process, separate mo<strong>de</strong>lling of margins and <strong>de</strong>p<strong>en</strong><strong>de</strong>nce will be<br />

achieved by introducing Lévy copulas, which play the same role for Lévy measures as copulas<br />

for probability measures. After showing how basic <strong>de</strong>p<strong>en</strong><strong>de</strong>nce patterns for Lévy processes are<br />

expressed in terms of their characteristic tripl<strong>et</strong>s in Section 4.2 and introducing the necessary<br />

notation and <strong>de</strong>finitions in Section 4.3, first, in Section 4.4 we treat the conceptually simpler<br />

case of Lévy processes, admitting only positive jumps in every compon<strong>en</strong>t or, equival<strong>en</strong>tly, having<br />

Lévy measures supported by [0, ∞) d . Lévy copulas for g<strong>en</strong>eral Lévy processes, introduced<br />

in a joint work of the pres<strong>en</strong>t author with Jan Kalls<strong>en</strong> [59], are discussed in Sections 4.5 and<br />

4.6 of this chapter.<br />

4.2 Dep<strong>en</strong><strong>de</strong>nce concepts for multidim<strong>en</strong>sional Lévy processes<br />

In this section, X := {X i t} i=1,...,d<br />

t≥0<br />

(A, ν, γ).<br />

<strong>de</strong>notes a Lévy process on R d with characteristic tripl<strong>et</strong><br />

For I ⊂ {1, . . . , d} we <strong>de</strong>fine I c := {1, . . . , d} \ I and |I| <strong>de</strong>notes the number of<br />

elem<strong>en</strong>ts in I. We start by <strong>de</strong>fining the margins of a Lévy process.<br />

Definition 4.1. L<strong>et</strong> I ⊂ {1, . . . , d} nonempty. The I-margin of X is the Lévy process X I :=<br />

{X i t} i∈I<br />

t≥0 .<br />

The following lemma explains that the Lévy measure of X I<br />

measure of X and shows how it can be computed.<br />

only <strong>de</strong>p<strong>en</strong>ds on the Lévy<br />

Lemma 4.1 (Marginal Lévy measures). L<strong>et</strong> I ⊂ {1, . . . , d} nonempty.<br />

process X I has Lévy measure ν I giv<strong>en</strong> by<br />

Th<strong>en</strong> the Lévy<br />

ν I (B) = ν({x ∈ R d : (x i ) i∈I ∈ B}), ∀B ∈ B(R |I| \ {0}). (4.1)<br />

Proof. This lemma is a direct consequ<strong>en</strong>ce of Proposition 11.10 in [87].


4.2. DEPENDENCE CONCEPTS FOR LEVY PROCESSES 139<br />

In view of the above lemma, for a giv<strong>en</strong> Lévy measure ν we will refer to the Lévy measure<br />

ν I <strong>de</strong>fined by Equation (4.5) as the I-margin of ν. To simplify notation, wh<strong>en</strong> I = {k} for some<br />

k, the I-margin of ν will be <strong>de</strong>noted by ν k and called simply k-th margin of ν.<br />

Next we would like to characterize the in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce of Lévy processes in terms of their<br />

characteristic tripl<strong>et</strong>s.<br />

Lemma 4.2. The compon<strong>en</strong>ts X 1 , . . . , X d of an R d -valued Lévy process X are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt<br />

if and only if their continuous martingale parts are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and the Lévy measure ν is<br />

supported by the coordinate axes. ν is th<strong>en</strong> giv<strong>en</strong> by<br />

ν(B) =<br />

d∑<br />

ν i (B i ) ∀B ∈ B(R d \ {0}), (4.2)<br />

i=1<br />

where for every i, ν i <strong>de</strong>notes the i-th margin of ν and<br />

B i = {x ∈ R : ( 0, . . . , 0<br />

} {{ }<br />

, x, 0, . . . , 0) ∈ B}.<br />

i − 1 times<br />

Proof. Since the continuous martingale part and the jump part of X are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt, we<br />

can assume without loss of g<strong>en</strong>erality that X has no continuous martingale part, that is, its<br />

characteristic tripl<strong>et</strong> is giv<strong>en</strong> by (0, ν, γ).<br />

The “if” part. Suppose ν is supported by the coordinate axes. Th<strong>en</strong> necessarily for every<br />

B ∈ B(R d \ {0}), ν(B) = ∑ d<br />

i=1 ˜ν i(B i ) with some measures ˜ν i , and Lemma 4.1 show that these<br />

measures coinci<strong>de</strong> with the margins of ν: ˜ν i = ν i ∀i. Using the Lévy-Khintchine formula for<br />

the process X, we obtain:<br />

∫<br />

E[e i〈u,Xt〉 ] = exp t{i〈γ, u〉 + (e i〈u,x〉 − 1 − i〈u, x〉1 |x|≤1 )ν(dx)}<br />

R d \{0}<br />

d∑<br />

∫<br />

= exp t {iγ k u k + (e iu kx k<br />

− 1 − iu k x k 1 |xk |≤1)ν k (dx k )} =<br />

k=1<br />

R\{0}<br />

which shows that the compon<strong>en</strong>ts of X are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt Lévy processes.<br />

d∏<br />

E[e iu kXt k ],<br />

The “only if” part. Define a measure ˜ν on R d \ {0} by ˜ν(B) = ∑ d<br />

i=1 ν i(B i ), where ν i is the<br />

i-th marginal Lévy measure of X and B i is as above. It is straightforward to check that ˜ν is<br />

a Lévy measure. Since the compon<strong>en</strong>ts of X are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt, applying the Lévy-Khintchine<br />

formula to each compon<strong>en</strong>t of X, we find:<br />

∫<br />

E[e i〈u,Xt〉 ] = exp t{i〈γ, u〉 +<br />

R d \{0}<br />

k=1<br />

(e i〈u,x〉 − 1 − i〈u, x〉1 |x|≤1 )˜ν(dx)}.


140 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

Now from the uniqu<strong>en</strong>ess of Lévy-Khintchine repres<strong>en</strong>tation we conclu<strong>de</strong> that ˜ν is the Lévy<br />

measure of X.<br />

The compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of Lévy processes is a new notion that is worth being discussed<br />

in <strong>de</strong>tail. First, the following <strong>de</strong>finition is in or<strong>de</strong>r.<br />

Definition 4.2. A subs<strong>et</strong> S of R d is called or<strong>de</strong>red if, for any two vectors v, u ∈ S, either<br />

v k ≤ u k , k = 1, . . . , d or v k ≥ u k , k = 1, . . . , d. S is called strictly or<strong>de</strong>red if, for any two<br />

differ<strong>en</strong>t vectors v, u ∈ S, either v k < u k , k = 1, . . . , d or v k > u k , k = 1, . . . , d.<br />

We recall that random variables Y 1 , . . . , Y d are said to be compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt or comonotonic<br />

if there exists a strictly or<strong>de</strong>red s<strong>et</strong> S ⊂ R d such that (Y 1 , . . . , Y d ) ∈ S with probability 1.<br />

However, saying that the compon<strong>en</strong>ts of a Lévy process are compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt only if they<br />

are compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt for every fixed time is too restrictive; the compon<strong>en</strong>ts of a Lévy process<br />

can be compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt as processes without being compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt as random<br />

variables for every fixed time. The following example clarifies this point.<br />

Example 4.2 (Dynamic compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce for Lévy processes). L<strong>et</strong> {X t } t≥0 be a Lévy process<br />

with characteristic tripl<strong>et</strong> (A, ν, γ) such that A = 0 and γ = 0 and l<strong>et</strong> {Y t } t≥0 be a Lévy process,<br />

constructed from the jumps of X: Y t = ∑ s≤t ∆X3 s . From the dynamic point of view X and Y<br />

are compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt, because the trajectory of any one of them can be reconstructed from<br />

the trajectory of the other. However, the copula of X t and Y t is not that of compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

because Y t is not a <strong>de</strong>terministic function of X t . In<strong>de</strong>ed, if X is a compound Poisson process<br />

having jumps of size 1 and 2 and X t = 3 for some t, this may either mean that X has three<br />

jumps of size 1 in the interval [0, t], and th<strong>en</strong> Y t = 3, or that X has one jump of size 1 and one<br />

jump of size 2, and th<strong>en</strong> Y t = 9.<br />

This example motivates the following <strong>de</strong>finition. In this <strong>de</strong>finition and below,<br />

K := {x ∈ R d : sgn x 1 = · · · = sgn x d }. (4.3)<br />

Definition 4.3. L<strong>et</strong> X be a R d -valued Lévy process. Its jumps are said to be compl<strong>et</strong>ely<br />

<strong>de</strong>p<strong>en</strong><strong>de</strong>nt or comonotonic if there exists a strictly or<strong>de</strong>red subs<strong>et</strong> S ⊂ K such that ∆X t :=<br />

X t − X t− ∈ S, t ≥ 0 (except for a s<strong>et</strong> of paths having zero probability).


4.3. INCREASING FUNCTIONS 141<br />

Clearly, an elem<strong>en</strong>t of a strictly or<strong>de</strong>red s<strong>et</strong> is compl<strong>et</strong>ely <strong>de</strong>termined by one coordinate<br />

only. Therefore, if the jumps of a Lévy process are compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt, the jumps of all<br />

compon<strong>en</strong>ts can be <strong>de</strong>termined from the jumps of any single compon<strong>en</strong>t. If the Lévy process<br />

has no continuous martingale part, th<strong>en</strong> the trajectories of all compon<strong>en</strong>ts can be <strong>de</strong>termined<br />

from the trajectory of any one compon<strong>en</strong>t, which indicates that Definition 4.3 is a reasonable<br />

dynamic notion of compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce for Lévy processes. The condition ∆X t ∈ K means that<br />

if the compon<strong>en</strong>ts of a Lévy process are comonotonic, they always jump in the same direction.<br />

For any R d -valued Lévy process X with Lévy measure ν and for any B ∈ B(R d \ {0}) the<br />

number of jumps in the time interval [0, t] with sizes in B is a Poisson random variable with<br />

param<strong>et</strong>er tν(B). Therefore, Definition 4.3 can be equival<strong>en</strong>tly restated in terms of the Lévy<br />

measure ν of X as follows:<br />

Definition 4.4. L<strong>et</strong> X be a R d -valued Lévy process with Lévy measure ν. Its jumps are said<br />

to be compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt or comonotonic if there exists a strictly or<strong>de</strong>red subs<strong>et</strong> S of K such<br />

that ν(R d \ S) = 0.<br />

4.3 Increasing functions<br />

In this section we sum up some aspects of the theory of increasing functions of several variables.<br />

Most <strong>de</strong>finitions are well known (see, e.g. [88]) but some are introduced either here for the first<br />

time or in [59].<br />

L<strong>et</strong> ¯R := R ∪ {−∞} ∪ {∞} <strong>de</strong>note the ext<strong>en</strong><strong>de</strong>d real line. In the sequel, for a, b ∈ ¯R d such<br />

that a ≤ b (inequalities like this one are to be interpr<strong>et</strong>ed compon<strong>en</strong>twise), l<strong>et</strong> [a, b) <strong>de</strong>note a<br />

right-op<strong>en</strong> left-closed interval of ¯R d (d-box):<br />

[a, b) := [a 1 , b 1 ) × · · · × [a d , b d ),<br />

In the same way we <strong>de</strong>fine other types of intervals: (a, b], [a, b] and (a, b). To simplify notation it<br />

is conv<strong>en</strong>i<strong>en</strong>t to introduce a “g<strong>en</strong>eral” type of interval: wh<strong>en</strong> we want to consi<strong>de</strong>r a d-dim<strong>en</strong>sional<br />

interval as a s<strong>et</strong> of its vertices, without specifying wh<strong>et</strong>her the boundary is inclu<strong>de</strong>d or not, we<br />

will write |a, b|.


142 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

Definition 4.5 (F-volume). L<strong>et</strong> F : S ⊂ ¯R d → ¯R. For a, b ∈ S, with a ≤ b, the F -volume of<br />

|a, b| is <strong>de</strong>fined by<br />

V F (|a, b|) := ∑ (−1) N(c) F (c), (4.4)<br />

where the sum is tak<strong>en</strong> over all vertices c of |a, b|, and N(c) := #{k : c k = a k }.<br />

The notion of F-volume should only be se<strong>en</strong> as a conv<strong>en</strong>i<strong>en</strong>t notation for the sum in the<br />

right-hand si<strong>de</strong> of (4.4). It does not in g<strong>en</strong>eral correspond to any measure because the measure<br />

of |a, b| will <strong>de</strong>p<strong>en</strong>d on wh<strong>et</strong>her the boundary is inclu<strong>de</strong>d or not.<br />

In particular, in two dim<strong>en</strong>sions (d = 2) the F -volume of a rectangle B = |x 1 , x 2 | × |y 1 , y 2 |<br />

satisfies<br />

V F (B) = F (x 2 , y 2 ) − F (x 2 , y 1 ) − F (x 1 , y 2 ) + F (x 1 , y 1 ).<br />

If F (u) = ∏ d<br />

i=1 u i, the F -volume of any interval is equal to its Lebesgue measure.<br />

Definition 4.6 (d-increasing function). A function F : S ⊂ ¯R d → ¯R is called d-increasing<br />

if for all a, b ∈ S with a ≤ b, V F (|a, b|) ≥ 0.<br />

Definition 4.7 (Groun<strong>de</strong>d function). For each k, l<strong>et</strong> S k ⊂ ¯R be such that inf S k ∈ S k . A<br />

function F : S 1 × · · · × S d → ¯R is groun<strong>de</strong>d if F (x 1 , . . . , x d ) = 0 wh<strong>en</strong>ever x k = inf S k for at<br />

least one k.<br />

Definition 4.8 (margins of a d-increasing groun<strong>de</strong>d function). For each k, l<strong>et</strong> S k ⊂ ¯R<br />

be such that inf S k ∈ S k and sup S k ∈ S k and l<strong>et</strong> F : S 1 × · · · × S d → ¯R be d-increasing and<br />

groun<strong>de</strong>d. Th<strong>en</strong> for I ⊂ {1, . . . , d} nonempty, the I-margin of F is a function F I : ∏ i∈I S i → ¯R<br />

<strong>de</strong>fined by<br />

F I ((x i ) i∈I ) := F (x 1 , . . . , x d )| xi =sup S i , i∈Ic. (4.5)<br />

Wh<strong>en</strong> I = {k}, to simplify notation, the (one-dim<strong>en</strong>sional) I-margin of F is <strong>de</strong>noted by F k .<br />

Example 4.3. The distribution function F of a random vector X ∈ R d is usually <strong>de</strong>fined by<br />

F (x 1 , . . . , x d ) := P [X 1 ≤ x 1 , . . . , X d ≤ x d ]


4.3. INCREASING FUNCTIONS 143<br />

for x 1 , . . . , x d ∈ ¯R. F is th<strong>en</strong> clearly increasing because for every a, b ∈ ¯R d with a ≤ b,<br />

V F (|a, b|) = P [X ∈ (a, b]] (4.6)<br />

It is groun<strong>de</strong>d because F (x 1 , . . . , x d ) = 0 if x k = −∞ for some k, and the margins of F are the<br />

distribution functions of the margins of X: for example, F 1 (x) = F (x, ∞, . . . , ∞) = P [X 1 ≤ x].<br />

The following technical lemma will be useful in the sequel.<br />

Lemma 4.3. For each k, l<strong>et</strong> S k ⊂ ¯R be such that inf S k ∈ S k and l<strong>et</strong> F, H : S 1 × · · · × S d → ¯R<br />

be d-increasing and groun<strong>de</strong>d. Th<strong>en</strong> their product F H is d-increasing and groun<strong>de</strong>d.<br />

Proof. We will prove this lemma by induction on d. The product of two increasing groun<strong>de</strong>d<br />

functions on ¯R is clearly increasing and groun<strong>de</strong>d. Suppose d ≥ 2, and for each k l<strong>et</strong> a k , b k ∈ S k<br />

with a k ≤ b k . Consi<strong>de</strong>r the function<br />

˜F (u 2 , . . . , u d ) := F (b 1 , u 2 , . . . , u d )H(b 1 , u 2 , . . . , u d ) − F (a 1 , u 2 , . . . , u d )H(a 1 , u 2 , . . . , u d )<br />

= H(a 1 , u 2 , . . . , u d )[F (b 1 , u 2 , . . . , u d ) − F (a 1 , u 2 , . . . , u d )]<br />

+ F (b 1 , u 2 , . . . , u d )[H(b 1 , u 2 , . . . , u d ) − H(a 1 , u 2 , . . . , u d )]<br />

Since H is groun<strong>de</strong>d,<br />

V H(a1 ,∗)(|a 2 , b 2 | × · · · × |a d , b d |) = V H (|0, a 1 | × |a 2 , b 2 | × · · · × |a d , b d |) ≥ 0,<br />

V H(b1 ,∗)−H(a 1 ,∗)(|a 2 , b 2 | × · · · × |a d , b d |) = V H (|a 1 , b 1 | × |a 2 , b 2 | × · · · × |a d , b d |) ≥ 0,<br />

and the same is true for F . Therefore ˜F is increasing by the induction hypothesis. Since<br />

V F H (|a 1 , b 1 | × · · · × |a d , b d |) = V ˜F<br />

(|a 2 , b 2 | × · · · × |a d , b d |),<br />

this finishes the proof of the lemma.<br />

To <strong>de</strong>velop the theory of Lévy copulas for g<strong>en</strong>eral Lévy processes, we will need to <strong>de</strong>fine the<br />

notion of margins for a function that is not groun<strong>de</strong>d. The following example gives an intuition<br />

of how this can be done.<br />

Example 4.4. Consi<strong>de</strong>r the following “alternative” <strong>de</strong>finition of a distribution function of a<br />

random vector X:<br />

d∏<br />

˜F (x 1 , . . . , x d ) := P [X 1 ∈ (x 1 ∧ 0, x 1 ∨ 0]; . . . ; X d ∈ (x d ∧ 0, x d ∨ 0]] sgn x i (4.7)<br />

i=1


144 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

for x 1 , . . . x d ∈ ¯R. This function satisfies Equation (4.6) and can therefore play the role of a<br />

distribution function. However, the margins of ˜F (e.g. the distribution functions computed<br />

using Equation (4.7) for the compon<strong>en</strong>ts of X) are no longer giv<strong>en</strong> by (4.5). It can be shown<br />

that<br />

˜F I ((x i ) i∈I ) := P [X i ∈ (x i ∧ 0, x i ∨ 0], i ∈ I] ∏ i∈I<br />

sgn x i<br />

=<br />

∑<br />

(x i ) i∈I c ∈{−∞,∞} |Ic |<br />

˜F (x 1 , . . . , x d ) ∏ i∈I c sgn x i .<br />

The above example motivates the following <strong>de</strong>finitions.<br />

notion of “increasing and groun<strong>de</strong>d” function.<br />

First, we need to g<strong>en</strong>eralize the<br />

Definition 4.9 (Volume function). L<strong>et</strong> S k ⊂ ¯R for k = 1, . . . , d. A function F : S 1 × · · ·×S d<br />

is called volume function if it is d-increasing and there exists x ∗ ∈ S 1 × · · · × S d such that<br />

F (x 1 , . . . , x d ) = 0 wh<strong>en</strong>ever x k = x ∗ k<br />

for some k.<br />

The term “volume function” is due to the fact that an increasing function is a volume<br />

function if and only if there exists x ∗ ∈ S 1 × · · · × S d such that<br />

d∏<br />

F (x 1 , . . . , x d ) = V F (|x 1 ∧ x ∗ 1, x 1 ∨ x ∗ 1| × · · · × |x d ∧ x ∗ d , x d ∨ x ∗ d |) sgn(x i − x ∗ i )<br />

for all x ∈ S 1 ×· · ·×S d . Every increasing groun<strong>de</strong>d function is a volume function. The function<br />

˜F , <strong>de</strong>fined in Example 4.4 is a volume function but is not groun<strong>de</strong>d.<br />

Definition 4.10 (Margins of a volume function). L<strong>et</strong> S k ⊂ ¯R for k = 1, . . . , d and l<strong>et</strong><br />

F : S 1 × · · · × S d → ¯R be a volume function. Th<strong>en</strong> for I ⊂ {1, . . . , d} nonempty, the I-margin<br />

of F is a function F I : ∏ i∈I S i → ¯R <strong>de</strong>fined by<br />

( )<br />

∏<br />

F I ((x i ) i∈I ) := sgn(x i − x ∗ i )<br />

i∈I<br />

∑<br />

× sup<br />

(−1) N((x i) i∈I c ) F (x 1 , . . . , x d ), (4.8)<br />

a i ,b i ∈S i :i∈I c<br />

(x i ) i∈I c ∈ ∏<br />

j∈I c{a j,b j }<br />

where N((x i ) i∈I c) = #{i ∈ I c : x i = a i }.<br />

In particular, for d = 2 we have F 1 (x) = sgn(x − x ∗ ) sup y1 ,y 2 ∈S 2<br />

{F (x, y 2 ) − F (x, y 1 )}.<br />

Equation (4.8) looks so complicated because it applies without modification to all cases that<br />

i=1


4.3. INCREASING FUNCTIONS 145<br />

are of interest to us: distribution functions, copulas, Lévy copulas with closed or op<strong>en</strong> domains.<br />

In particular cases we will obtain simplifications that are easier to work with. Using the F-<br />

volume notation, Equation (4.8) can be rewritt<strong>en</strong> as follows:<br />

⎛ ⎧<br />

⎞<br />

( )<br />

∏<br />

F I ((x i ) i∈I ) = sgn(x i − x ∗ ⎜<br />

d∏<br />

⎪⎨ |x i ∧ x ∗ i , x i ∨ x ∗ i |, i ∈ I ⎟<br />

i ) sup V F ⎝<br />

⎠ .<br />

i∈I<br />

a i ,b i ∈S i :a i ≤b i ,i∈I c i=1<br />

⎪⎩ |a i , b i |, i ∈ I c<br />

(4.9)<br />

Wh<strong>en</strong> F satisfies the conditions of Definition 4.8, since F is d-increasing, the sup is achieved<br />

wh<strong>en</strong> for every i ∈ I c , a i = inf S i and b i = sup S i . Therefore, in this case the above formula<br />

yields F I ((x i ) i∈I ) = F (x 1 , . . . , x d )| xi =sup S i , i∈I c<br />

and the two <strong>de</strong>finitions coinci<strong>de</strong>.<br />

The following important property of increasing functions will be useful in the sequel.<br />

Lemma 4.4. L<strong>et</strong> S k ⊂ ¯R for k = 1, . . . , d and l<strong>et</strong> F : S 1 × · · · × S d → ¯R be a volume function.<br />

L<strong>et</strong> (x 1 , . . . , x d ) and (y 1 , . . . , y d ) be any points in Dom F . Th<strong>en</strong><br />

|F (x 1 , . . . , x d ) − F (y 1 , . . . , y d )| ≤<br />

Proof. From the triangle inequality,<br />

d∑<br />

|F k (x k ) − F k (y k )|. (4.10)<br />

k=1<br />

|F (x 1 , . . . , x d ) − F (y 1 , . . . , y d )| ≤ |F (x 1 , . . . , x d ) − F (y 1 , x 2 . . . , x d )| + . . .<br />

+ |F (y 1 , . . . , y d−1 , x d ) − F (y 1 , . . . , y d )|,<br />

h<strong>en</strong>ce it suffices to prove that<br />

|F (x 1 , . . . , x d ) − F (y 1 , x 2 . . . , x d )| ≤ |F 1 (x 1 ) − F 1 (y 1 )|. (4.11)<br />

Without loss of g<strong>en</strong>erality suppose that x 1 ≥ y 1 ≥ x ∗ 1 and that x k ≥ x ∗ k<br />

B := |x ∗ 2 , x 2| × · · · × |x ∗ d , x d|. Th<strong>en</strong><br />

for k = 2, . . . , d and l<strong>et</strong><br />

F (x 1 , . . . , x d ) − F (y 1 , x 2 . . . , x d ) = V F (|y 1 , x 1 | × B). (4.12)<br />

Moreover, for all x ≥ x ∗ 1 ,<br />

F 1 (x) =<br />

sup V F (|x ∗ 1, x| × |a 2 , b 2 | × · · · × |a d , b d |).<br />

a i ,b i ∈S i :a i ≤b i ,i=2,...,d


146 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

Fix ε > 0 and choose a i , b i ∈ S i : a i ≤ b i , i = 2, . . . , d such that<br />

F 1 (y 1 ) ≤ ε + V F (|x ∗ 1, y 1 | × |a 2 , b 2 | × · · · × |a d , b d |).<br />

For i = 2, . . . , d l<strong>et</strong> ã i := a i ∧ x ∗ i and ˜b i := b i ∨ x i and <strong>de</strong>note ˜B := |ã 2 , ˜b 2 | × · · · × |ã d , ˜b d |. Th<strong>en</strong>,<br />

since F is d-increasing, V F (|y 1 , x 1 | × B) ≤ V F (|y 1 , x 1 | × ˜B) and V F (|x ∗ 1 , y 1| × |a 2 , b 2 | × · · · ×<br />

|a d , b d |) ≤ V F (|x ∗ 1 , y 1| × ˜B). Therefore<br />

V F (|y 1 , x 1 | × B) ≤ V F (|x ∗ 1, x 1 | × ˜B) − F 1 (y 1 ) + ε ≤ F 1 (x 1 ) − F 1 (y 1 ) + ε.<br />

Since the above is true for all ε > 0, in view of (4.12), the proof of (4.11) is compl<strong>et</strong>e.<br />

We close this section with the <strong>de</strong>finition of (ordinary) copula and the Sklar’s theorem, which<br />

relates copulas to distribution functions. The proof of Sklar’s theorem can be found in [90].<br />

Definition 4.11 (Copula). A d-dim<strong>en</strong>sional copula (a d-copula) is a function C : [0, 1] d →<br />

[0, 1] such that<br />

1. C is groun<strong>de</strong>d and d-increasing.<br />

2. C has margins C k , k = 1, 2, . . . , d, which satisfy C k (u) = u for all u in [0, 1].<br />

Theorem 4.5 (Sklar). L<strong>et</strong> F be a d-dim<strong>en</strong>sional distribution function with margins F 1 , . . . , F d .<br />

Th<strong>en</strong> there exists a d-dim<strong>en</strong>sional copula C such that for all x ∈ ¯R d ,<br />

F (x 1 , x 2 , . . . , x d ) = C(F 1 (x 1 ), F 2 (x 2 ), . . . , F d (x d )). (4.13)<br />

If F 1 , . . . , F d are continuous th<strong>en</strong> C is unique; otherwise, C is uniquely <strong>de</strong>termined on Ran F 1 ×<br />

· · · × Ran F d . Conversely, if C is a d-copula and F 1 , . . . , F d are distribution functions, th<strong>en</strong> the<br />

function F <strong>de</strong>fined by (4.13) is a d-dim<strong>en</strong>sional distribution function with margins F 1 , . . . , F d .<br />

4.4 Lévy copulas for spectrally positive Lévy processes<br />

This section discusses the notion of Lévy copula for Lévy processes with only positive jumps in<br />

each compon<strong>en</strong>t. This notion was introduced by the pres<strong>en</strong>t author in [92]. Examples of Lévy<br />

copulas for spectrally positive Lévy processes and m<strong>et</strong>hods to construct them will be giv<strong>en</strong> in<br />

Sections 4.6 and 5.1 tog<strong>et</strong>her with the examples of g<strong>en</strong>eral Lévy copulas. Further properties of


4.4. LEVY COPULAS: SPECTRALLY POSITIVE CASE 147<br />

Lévy copulas in the spectrally positive case, including a characterization of the converg<strong>en</strong>ce of<br />

Lévy measures in terms of Lévy copulas can be found in a rec<strong>en</strong>t paper by Barndorff-Niels<strong>en</strong><br />

and Lindner [8].<br />

As the laws of random variables are repres<strong>en</strong>ted by their distribution functions, Lévy measures<br />

can be repres<strong>en</strong>ted by their tail integrals.<br />

Definition 4.12. L<strong>et</strong> ν be a Lévy measure on R d + := [0, ∞) d \ {0}. The tail integral U of ν is<br />

a function [0, ∞) d → [0, ∞] such that<br />

1. U(0, . . . , 0) = ∞.<br />

2. For (x 1 , . . . , x d ) ∈ R d +,<br />

U(x 1 , . . . , x d ) = ν([x 1 , ∞) × · · · × [x d , ∞)).<br />

The Lévy measure is uniquely <strong>de</strong>termined by its tail integral, because the above <strong>de</strong>finition<br />

implies that for every x, y ∈ R d + with x ≤ y,<br />

V U (|x, y|) = (−1) d ν([x 1 , y 1 ) × · · · × [x d , y d )). (4.14)<br />

Definition 4.13. L<strong>et</strong> X be a R d -valued Lévy process and l<strong>et</strong> I ⊂ {1, . . . , d} non-empty. The<br />

I-marginal tail integral U I of X is the tail integral of the process X I := (X i ) i∈I . The onedim<strong>en</strong>sional<br />

margins are, as usual, <strong>de</strong>noted by U i := U {i} .<br />

Lemma 4.1 <strong>en</strong>tails that for I ⊂ {1, . . . , d} nonempty, the I-marginal tail integral of a Lévy<br />

measure ν on R d + can be computed from the tail integral U of ν by substituting 0 instead of<br />

argum<strong>en</strong>ts with indices not in I:<br />

U I ((x i ) i∈I ) = U(x 1 , . . . , x d )| (xi ) i∈I c =0. (4.15)<br />

A Lévy copula is <strong>de</strong>fined similarly to ordinary copula but on a differ<strong>en</strong>t domain.<br />

Definition 4.14. A function F : [0, ∞] d → [0, ∞] is a Lévy copula if<br />

1. F (u 1 , . . . , u d ) < ∞ for (u 1 , . . . , u d ) ≠ (∞, . . . , ∞),<br />

2. F is groun<strong>de</strong>d: F (u 1 , . . . , u d ) = 0 wh<strong>en</strong>ever u i = 0 for at least one i ∈ {1, . . . , d},


148 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

3. F is d-increasing,<br />

4. F i (u) = u for any i ∈ {1, . . . , d}, u ∈ [0, ∞].<br />

The following theorem is an analog of the well-known Sklar’s theorem for copulas. The<br />

proof is done using multilinear interpolation and is inspired by Sklar’s proof of his theorem in<br />

[90].<br />

Theorem 4.6. L<strong>et</strong> ν be a Lévy measure on R d + with tail integral U and marginal Lévy measures<br />

ν 1 , . . . , ν d . There exists a Lévy copula F on [0, ∞] d such that<br />

U(x 1 , . . . , x d ) = F (U 1 (x 1 ), . . . , U d (x d )), (x 1 , . . . , x d ) ∈ [0, ∞) d , (4.16)<br />

where U 1 , . . . , U d are tail integrals of ν 1 , . . . , ν d . This Lévy copula is unique on ∏ d<br />

i=1 Ran U i.<br />

Conversely, if F is a Lévy copula on [0, ∞] d and ν 1 , . . . , ν d are Lévy measures on (0, ∞)<br />

with tail integrals U 1 , . . . , U d th<strong>en</strong> Equation (4.16) <strong>de</strong>fines a tail integral of a Lévy measure on<br />

R d + with marginal Lévy measures ν 1 , . . . , ν d .<br />

Remark 4.1. In particular, the Lévy copula F is unique if the marginal Lévy measures ν 1 , . . . , ν d<br />

are infinite and have no atoms, because in this case Ran U k = [0, ∞] for every k.<br />

The first part of this theorem states that all types of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of Lévy processes (with<br />

only positive jumps), including compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce and in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce, can be repres<strong>en</strong>ted<br />

with Lévy copulas. The second part shows that one can construct multivariate Lévy process<br />

mo<strong>de</strong>ls by specifying separately jump <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure and one-dim<strong>en</strong>sional laws for the<br />

compon<strong>en</strong>ts. The laws of compon<strong>en</strong>ts can have very differ<strong>en</strong>t structure, in particular, it is possible<br />

to construct examples of Lévy processes with some compon<strong>en</strong>ts being compound Poisson<br />

and others having an infinite jump int<strong>en</strong>sity.<br />

Proof of theorem 4.6. For the purposes of this proof, we introduce some auxiliary functions and<br />

measures. First, for every k = 1, . . . , d and every x ∈ [0, ∞] we <strong>de</strong>fine<br />

⎧<br />

⎪⎨ U k (x), x ≠ ∞<br />

Ũ k (x) :=<br />

⎪⎩ 0, x = ∞.<br />

Ũ (−1)<br />

k<br />

(t) := sup{x ≥ 0 : Ũk(x) ≥ t}.


4.4. LEVY COPULAS: SPECTRALLY POSITIVE CASE 149<br />

The following properties of Ũ (−1)<br />

k<br />

follow directly from its <strong>de</strong>finition:<br />

Ũ (−1)<br />

k<br />

(t) is nonincreasing in t,<br />

Ũ (−1)<br />

k<br />

(∞) = 0, (4.17)<br />

Ũ k (Ũ (−1)<br />

k<br />

(t)) = t ∀t ∈ Ran Ũk. (4.18)<br />

For every (x 1 , . . . , x d ) ∈ [0, ∞] d , we <strong>de</strong>fine<br />

⎧<br />

⎪⎨<br />

0, x k = ∞ for some k<br />

Ũ(x 1 , . . . , x d ) =<br />

⎪⎩ U(x 1 , . . . , x d ), otherwise.<br />

Finally, introduce a measure ˜ν on [0, ∞] d \{0} by ˜ν(B) := ν(B ∩R d +) for all B ∈ B([0, ∞] d \{0}).<br />

Clearly, Equation (4.14) still holds for all x, y ∈ [0, ∞] d \{0} with x ≤ y, if U and ν are replaced<br />

by Ũ and ˜ν.<br />

First part. To prove the exist<strong>en</strong>ce of a Lévy copula, we construct the required Lévy copula<br />

in two stages.<br />

1. First consi<strong>de</strong>r the function ˜F : D := Ran Ũ1 × · · · × Ran Ũd → [0, ∞] <strong>de</strong>fined by<br />

˜F (x 1 , . . . , x d ) =<br />

Ũ(Ũ<br />

(−1)<br />

1 (x 1 ), . . . , Ũ (−1)<br />

d<br />

(x d )).<br />

Suppose that x k = 0 for some k. Without loss of g<strong>en</strong>erality we can take k = 1. Th<strong>en</strong><br />

˜F (0, x 2 , . . . , x d ) =<br />

Ũ(z, Ũ<br />

(−1)<br />

2 (x 2 ), . . . , Ũ (−1)<br />

d<br />

(x d )),<br />

where z is such that Ũ1(z) = 0. Since Ũ is nonnegative and nonincreasing in each argum<strong>en</strong>t,<br />

0 ≤<br />

Ũ(z, Ũ<br />

(−1)<br />

2 (x 2 ), . . . , Ũ (−1)<br />

d<br />

(x d )) ≤ Ũ(z, 0, . . . , 0) = Ũ1(z) = 0.<br />

Therefore, ˜F is groun<strong>de</strong>d. L<strong>et</strong> a, b ∈ D with ak ≤ b k , k = 1, . . . , d and <strong>de</strong>note<br />

B := (a 1 , b 1 ] × · · · × (a d , b d ]<br />

and<br />

Since Ũ (−1)<br />

k<br />

˜B := [Ũ (−1)<br />

1 (b 1 ), Ũ (−1)<br />

1 (a 1 )) × · · · × [Ũ (−1)<br />

(b k ) ≤ Ũ (−1)<br />

k<br />

(a k ) for every k, formula (4.14) <strong>en</strong>tails that<br />

d<br />

(b d ), Ũ (−1)<br />

d<br />

(a d )).<br />

V F (B) = (−1) d VŨ( ˜B) = ˜ν( ˜B) ≥ 0,


150 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

which means that ˜F is a d-increasing. Equations (4.17) and (4.18) show that for every k,<br />

˜F (∞, . . . , ∞, x k , ∞, . . . , ∞) = x k , ∀x k ∈ Ran Ũk. Therefore, ˜F has uniform margins.<br />

2. The second stage is to ext<strong>en</strong>d ˜F to the function F <strong>de</strong>fined on [0, ∞] d . The ext<strong>en</strong>sion can<br />

be carried out step by step, constructing a sequ<strong>en</strong>ce of d + 1 groun<strong>de</strong>d d-increasing functions<br />

F 0 , . . . , F d with uniform margins such that F 0 ≡ ˜F , F d ≡ F ,<br />

Dom F k = [0, ∞] k × Ran Ũk+1 × · · · × Ran Ũd,<br />

and for each k = 1, . . . , d − 1, F k ≡ F k+1 on Dom F k . In other words, at each step, we ext<strong>en</strong>d<br />

˜F along one coordinate only. Since all d steps are performed in the same way, we need only to<br />

<strong>de</strong>scribe one of them. To simplify the notation and without loss of g<strong>en</strong>erality, we <strong>de</strong>scribe the<br />

first one, that is, we show how to ext<strong>en</strong>d ˜F to a function <strong>de</strong>fined on [0, ∞]×Ran Ũ2×· · ·×Ran Ũd.<br />

First, l<strong>et</strong> us ext<strong>en</strong>d ˜F by continuity to a function <strong>de</strong>fined on Ran Ũ1 ×Ran Ũ2 ×· · ·×Ran Ũd.<br />

Giv<strong>en</strong> a sequ<strong>en</strong>ce of real numbers {ξ i } ∞ i=1 such that ξ i ∈ Ran Ũ1 for all i and lim i ξ i /∈ Ran Ũ1,<br />

we <strong>de</strong>fine<br />

F (lim<br />

i<br />

ξ i , x 2 , . . . , x d ) := lim<br />

i<br />

F (ξ i , x 2 , . . . , x d ).<br />

Since ∞ ∈ Ran Ũ1, lim i ξ i is finite. Therefore, by Lemma 4.4, the limit in the right-hand si<strong>de</strong> of<br />

the above expression exists and is uniform in other coordinates so F is well-<strong>de</strong>fined. It is also<br />

clear that F is d-increasing, groun<strong>de</strong>d and has uniform margins.<br />

Now we can suppose that Ran Ũ1 is closed and ext<strong>en</strong>d ˜F using linear interpolation to a<br />

function <strong>de</strong>fined on ([0, λ 1 ] ∪ {∞}) × Ran Ũ2 × · · · × Ran Ũd where λ 1 = lim x↓0 Ũ 1 (x).<br />

any x ≤ λ 1 such that x /∈ Ran Ũ1, we introduce x := sup{ξ ∈ Ran Ũ1, ξ ≤ x}, ¯x := inf{ξ ∈<br />

Ran Ũ1, ξ ≥ x}. Since Ran Ũ1 is closed, it contains both x and ¯x. Define<br />

F (x, x 2 , . . . , x d ) := ˜F (x, x 2 , . . . , x d ) ¯x − x<br />

¯x − x + ˜F (¯x, x 2 , . . . , x d ) x − x<br />

¯x − x . (4.19)<br />

The function F is clearly groun<strong>de</strong>d and has uniform margins; we only need to prove that it<br />

is d-increasing. Fix a d-box B = |x 1 , y 1 | × · · · × |x d , y d | with vertices in the domain of F and<br />

<strong>de</strong>note ˜B = |x 2 , y 2 | × · · · × |x d , y d |.<br />

If both x 1 and y 1 belong to Ran Ũ1 th<strong>en</strong> all vertices of B are in Dom ˜F and V F (B) =<br />

V ˜F<br />

(B) ≥ 0.<br />

If x 1 /∈ Ran Ũ1, y 1 /∈ Ran Ũ1 and b<strong>et</strong>we<strong>en</strong> x 1 and y 1 there are no points that belong to<br />

For


4.4. LEVY COPULAS: SPECTRALLY POSITIVE CASE 151<br />

Ran Ũ1, th<strong>en</strong> a straightforward computation using Equation (4.19) shows that<br />

Suppose that x 1<br />

V F (B) = y 1 − x 1<br />

¯x 1 − x 1<br />

V ˜F<br />

(|x 1 , ¯x 1 | × ˜B) ≥ 0.<br />

/∈ Ran Ũ1 but y 1 ∈ Ran Ũ1. L<strong>et</strong> z = inf{ζ ≥ x 1 , ζ ∈ Ran Ũ1}. Because<br />

Ran Ũ1 is closed, z ∈ Ran Ũ1. The F -volume of B can be <strong>de</strong>composed as follows:<br />

V F (B) = V F (|x 1 , z| × ˜B) + V F (|z, y 1 | × ˜B).<br />

The second term is positive because all vertices of the corresponding d-box are in the domain<br />

of ˜F . The first term can again be computed using formula (4.19):<br />

V F (|x 1 , z| × ˜B) = z − x 1<br />

z − x 1<br />

V ˜F<br />

(|x 1 , z| × ˜B) ≥ 0.<br />

The case where x 1 ∈ Ran Ũ1 and y 1 /∈ Ran Ũ1 can be treated in this same way and if x 1 /∈ Ran Ũ1,<br />

y 1 /∈ Ran Ũ1 and b<strong>et</strong>we<strong>en</strong> x 1 and y 1 there are points that belong to Ran Ũ1, the interval |x 1 , y 1 |<br />

can be split onto two intervals of types that we have already discussed.<br />

We have thus ext<strong>en</strong><strong>de</strong>d ˜F to a function <strong>de</strong>fined on ([0, λ 1 ]∪{∞})×Ran Ũ2 ×· · ·×Ran Ũd. If<br />

λ 1 = ∞ th<strong>en</strong> we are done; otherwise we ext<strong>en</strong>d ˜F to [0, ∞] × Ran Ũ2 × · · · × Ran Ũd by <strong>de</strong>fining<br />

F (x 1 , . . . , x d ) := ˜F (g λ1 (x 1 ), x 2 , . . . , x d ) + (x 1 − λ 1 ) + 1 x2 =∞ . . . 1 xd =∞, (4.20)<br />

where<br />

⎧<br />

x, x ≤ λ 1<br />

⎪⎨<br />

g λ1 (x) := λ 1 , λ 1 < x < ∞<br />

⎪⎩<br />

∞, x = ∞.<br />

The function F , <strong>de</strong>fined by (4.20) is an increasing function because it is a sum of two increasing<br />

functions. The groun<strong>de</strong>dness and marginal properties can be verified by direct substitution.<br />

This compl<strong>et</strong>es the construction of the Lévy copula.<br />

To prove the uniqu<strong>en</strong>ess, assume that there exist two functions with required properties,<br />

i.e., we have<br />

F 1 (U 1 (x 1 ), . . . , U d (x d )) = F 2 (U 1 (x 1 ), . . . , U d (x d )), ∀x 1 , . . . , x d .<br />

For each vector (t 1 , . . . , t d ) ∈ Ran U 1 × · · · × Ran U d there exists a vector (x 1 , . . . , x d ) ∈ [0, ∞] d<br />

such that U 1 (x 1 ) = t 1 , . . . , U d (x d ) = t d . This means that for every (t 1 , . . . , t d ) ∈ Ran U 1 × · · · ×


152 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

Ran U d we have F 1 (t 1 , . . . , t d ) = F 2 (t 1 , . . . , t d ). The uniqu<strong>en</strong>ess statem<strong>en</strong>t now follows from<br />

Lemma 4.4.<br />

The converse statem<strong>en</strong>t. Since U 1 , . . . , U d are left-continuous (as tail integrals of Lévy<br />

measures), Lemma 4.4 <strong>en</strong>tails that U is left-continuous in each variable. Therefore, there exists<br />

a unique positive measure ν on B(R d +) such that for every right-op<strong>en</strong> left-closed interval I of<br />

R d +, ν(I) = (−1) d V U (I) (see Section 4.5 in [60]).<br />

It remains to prove that ∫ |x|≤1 |x|2 ν(dx) < ∞. L<strong>et</strong> us fix k ∈ {1 . . . d} and consi<strong>de</strong>r the<br />

measure ν ′ k <strong>de</strong>fined by ν ′ k (A) := ∫<br />

ν(dx)<br />

{x∈R d + :x k∈A}<br />

for A ∈ B((0, ∞)).<br />

By the uniqu<strong>en</strong>ess of the ext<strong>en</strong>sion this measure coinci<strong>de</strong>s with ν k because it is straightforward<br />

to check that the two measures coinci<strong>de</strong> on right-op<strong>en</strong> left-closed intervals of (0, ∞). Therefore<br />

∫<br />

|x| 2 ν(dx) =<br />

[0,1] d<br />

d∑<br />

∫<br />

d∑<br />

∫<br />

x 2 k ν(dx) ≤<br />

[0,1] d<br />

k=1<br />

k=1<br />

[0,1]<br />

x 2 k ν′ k (dx k) =<br />

d∑<br />

∫<br />

k=1<br />

[0,1]<br />

x 2 k ν k(dx k ).<br />

Since ν k for every k is by the theorem’s statem<strong>en</strong>t a Lévy measure, the right-hand si<strong>de</strong> of the<br />

above expression is finite, which shows that ν is in<strong>de</strong>ed a Lévy measure.<br />

4.5 Lévy copulas for g<strong>en</strong>eral Lévy processes<br />

This section discusses the notion of Lévy copula for g<strong>en</strong>eral Lévy processes, introduced in a<br />

joint work of the pres<strong>en</strong>t author with Jan Kalls<strong>en</strong> [59]. In the sequel, we will need a special<br />

interval associated with any x ∈ R:<br />

⎧<br />

⎨ [x, ∞), x ≥ 0,<br />

I(x) :=<br />

⎩<br />

(−∞, x), x < 0.<br />

(4.21)<br />

Similarly to Lévy processes with positive jumps, the Lévy measure of a g<strong>en</strong>eral Lévy process<br />

will be repres<strong>en</strong>ted by its tail integral.<br />

Definition 4.15. L<strong>et</strong> X be a R d -valued Lévy process with Lévy measure ν. The tail integral<br />

of ν is the function U : (R \ {0}) d → R <strong>de</strong>fined by<br />

⎛ ⎞<br />

d∏<br />

d∏<br />

U(x 1 , . . . , x d ) := ν ⎝ I(x j ) ⎠ sgn(x i ) (4.22)<br />

j=1 i=1


4.5. LEVY COPULAS: GENERAL CASE 153<br />

Note that the tail integral is <strong>de</strong>fined so that (−1) d U is d-increasing and left-continuous in<br />

each orthant as was the case for the tail integral of Lévy measure on R d +.<br />

Since the tail integral is only <strong>de</strong>fined on (R \ {0}) d , it does not <strong>de</strong>termine the Lévy measure<br />

uniquely (unless we know that the latter does not charge the coordinate axes). However, we<br />

will see from the following lemma that the Lévy measure is compl<strong>et</strong>ely <strong>de</strong>termined by its tail<br />

integral and all its marginal tail integrals (cf. Definition 4.13).<br />

Lemma 4.7. L<strong>et</strong> X be a R d -valued Lévy process. Its marginal tail integrals {U I : I ⊂<br />

{1, . . . , d} non-empty} are uniquely <strong>de</strong>termined by its Lévy measure ν. Conversely, its Lévy<br />

measure is uniquely <strong>de</strong>termined by the s<strong>et</strong> of its marginal tail integrals.<br />

An important differ<strong>en</strong>ce b<strong>et</strong>we<strong>en</strong> the spectrally positive case and the g<strong>en</strong>eral case is that in<br />

the spectrally positive case the Lévy measure is <strong>de</strong>termined by a single tail integral U, <strong>de</strong>fined<br />

on [0, ∞) d , whereas in the g<strong>en</strong>eral case the Lévy measure is <strong>de</strong>termined jointly by all marginal<br />

tail integrals, each one being <strong>de</strong>fined on (R \ {0}) I . In the spectrally positive case Equation<br />

4.15 shows that specifying the values of U wh<strong>en</strong> some of the argum<strong>en</strong>ts equal 0 is equival<strong>en</strong>t<br />

to specifying the marginal tail integrals. In principle, we could do the same in the g<strong>en</strong>eral case<br />

and <strong>de</strong>fine the tail integral on R d by Equation (4.22). However, due to a more complicated<br />

structure of margins of a Lévy copula in the g<strong>en</strong>eral case, with this <strong>de</strong>finition of the tail integral<br />

the repres<strong>en</strong>tation formula (4.16) does not hold on R d . Therefore, we keep the Definition 4.15<br />

and <strong>de</strong>scribe the Lévy measure using its tail integral, <strong>de</strong>fined on (R \ {0}) d and all its marginal<br />

tail integrals.<br />

Proof of Lemma 4.7. The fact that marginal tail integrals are uniquely <strong>de</strong>termined by the Lévy<br />

measure ν follows from Lemma 4.1.<br />

The converse statem<strong>en</strong>t.<br />

It is suffici<strong>en</strong>t to prove that ν([a, b)) is compl<strong>et</strong>ely <strong>de</strong>termined<br />

by the tail intergals for any a, b ∈ R d with a ≤ b and 0 /∈ [a, b). We prove by induction on<br />

k = 0, . . . , d that ν I ( ∏ i∈I [a i, b i )) is <strong>de</strong>termined by the tail integrals for any a, b ∈ R d such that<br />

a ≤ b and a i b i ≤ 0 for at most k indices and any non-empty I ⊂ {1, . . . , d} with 0 /∈ ∏ i∈I [a i, b i ).<br />

If k = 0, Definitions 4.15 and 4.13 <strong>en</strong>tail that<br />

ν I ( ∏<br />

i∈I[a i , b i )<br />

)<br />

= (−1) |I| V U I<br />

( ∏<br />

[a i , b i )<br />

i∈I<br />

)<br />

.


154 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

L<strong>et</strong> a, b ∈ R d such that a i b i ≤ 0 for at most k indices. For ease of notation we suppose that<br />

a i b i ≤ 0 for i = 1, . . . , k. L<strong>et</strong> I ⊂ {1, . . . , d} non-empty with 0 /∈ ∏ i∈I [a i, b i ). By induction<br />

hypothesis, ν I ( ∏ i∈I [a i, b i )) is uniquely <strong>de</strong>termined if k ∉ I. Suppose that k ∈ I. If b k = 0, th<strong>en</strong><br />

( ) ⎛<br />

⎞<br />

∏<br />

ν I i , b i ) = lim ν<br />

α↑0<br />

i∈I[a I ⎝ ∏<br />

[a i , b i ) × [a k , α) × ∏<br />

[a i , b i ) ⎠<br />

i∈I,ik<br />

and the right-hand si<strong>de</strong> is uniquely <strong>de</strong>termined by the induction hypothesis. If b k ≠ 0, th<strong>en</strong><br />

( )<br />

⎛<br />

⎞<br />

∏<br />

ν I i , b i ) = ν<br />

i∈I[a I\{k} ⎝ ∏<br />

[a i , b i ) ⎠<br />

which is uniquely <strong>de</strong>termined as well.<br />

i∈I\{k}<br />

⎛<br />

⎞<br />

− lim ν I ⎝ ∏<br />

[a i , b i ) × [b k , c) × ∏<br />

[a i , b i ) ⎠<br />

c↑∞<br />

i∈I,ik<br />

⎛<br />

⎞<br />

− lim<br />

α↑a k ;c↓−∞ νI ⎝ ∏<br />

[a i , b i ) × [c, a k ) × ∏<br />

[a i , b i ) ⎠ ,<br />

i∈I,ik<br />

Lévy copulas in the g<strong>en</strong>eral case are <strong>de</strong>fined similarly to the spectrally positive case.<br />

Definition 4.16. A function F : (−∞, ∞] d → (−∞, ∞] is a Lévy copula if<br />

1. F (u 1 , . . . , u d ) < ∞ for (u 1 , . . . , u d ) ≠ (∞, . . . , ∞),<br />

2. F (u 1 , . . . , u d ) = 0 if u i = 0 for at least one i ∈ {1, . . . , d},<br />

3. F is d-increasing,<br />

4. F i (u) = u for any i ∈ {1, . . . , d}, u ∈ (−∞, ∞].<br />

Remark 4.2. Since F is d-increasing, the sup in Equation (4.8) for the margins of F may be<br />

computed by taking b i = ∞ and a i → −∞ for every i ∈ I c . Therefore, Equation (4.8) reduces<br />

to<br />

F I ((x i ) i∈I ) := lim<br />

∑<br />

c→∞<br />

(x j ) j∈I c ∈{−c,∞} |Ic |<br />

F (x 1 , . . . , x d ) ∏<br />

j∈I c sgn x j . (4.23)<br />

Wh<strong>en</strong> F is a Lévy copula on [0, ∞] d (Definition 4.14), it can be ext<strong>en</strong><strong>de</strong>d to a Lévy copula F ext<br />

on (−∞, ∞] d by taking<br />

⎧<br />

⎪⎨ F (x 1 , . . . , x d ), (x 1 , . . . , x d ) ∈ [0, ∞] d<br />

F ext (x 1 , . . . , x d ) :=<br />

⎪⎩<br />

0 otherwise.


4.5. LEVY COPULAS: GENERAL CASE 155<br />

In the following, the term Lévy copula without specifying the domain refers to a Lévy copula<br />

on (−∞, ∞] d .<br />

Example 4.5. It is easy to check that the function<br />

F (u, v) := 1 uv<br />

2 1 + |u| + |v|<br />

<strong>de</strong>fines a Lévy copula on (−∞, ∞] 2 .<br />

In view of Lemma 4.7, the following theorem is analogous to Theorem 4.6 for Lévy measures<br />

on R d + and to Sklar’s theorem for probability measures.<br />

Theorem 4.8. L<strong>et</strong> ν be a Lévy measure on R d \ {0}. Th<strong>en</strong> there exists a Lévy copula F such<br />

that the tail integrals of ν satisfy:<br />

U I ((x i ) i∈I ) = F I ((U i (x i )) i∈I ) (4.24)<br />

for any non-empty I ⊂ {1, . . . , d} and any (x i ) i∈I ∈ (R \ {0}) I . The Lévy copula F is unique<br />

on ∏ d<br />

i=1 Ran U i.<br />

Conversely, if F is a d-dim<strong>en</strong>sional Lévy copula and ν 1 , . . . , ν d are Lévy measures on R\{0}<br />

with tail integrals U i , i = 1, . . . , d th<strong>en</strong> there exists a unique Lévy measure on R d \ {0} with onedim<strong>en</strong>sional<br />

marginal tail integrals U 1 , . . . , U d and whose marginal tail integrals satisfy Equation<br />

(4.24) for any non-empty I ⊂ {1, . . . , d} and any (x i ) i∈I ∈ (R \ {0}) I .<br />

Remark 4.3. In particular, the Lévy copula F is unique if for each k, the marginal Lévy measure<br />

ν k has no atoms and both ν k ((−∞, 0)) = ∞ and ν k ((0, ∞)) = ∞, because in this case Ran U k =<br />

[−∞, ∞] for each k.<br />

Definition 4.17. For a Lévy process X on R d with characteristic tripl<strong>et</strong> (A, ν, γ), any Lévy<br />

copula as in statem<strong>en</strong>t of Theorem 4.8 is called the Lévy copula of X.<br />

Proof of Theorem 4.8. First part. D<strong>en</strong>ote the marginal Lévy measures of ν by ν 1 , . . . , ν d . For<br />

the purposes of this proof we s<strong>et</strong> for x ∈ (−∞, ∞], i = 1, . . . , d,<br />

⎧<br />

U ⎪⎨ i (x) for x ≠ 0 and x ≠ ∞,<br />

Ũ i (x) := 0 for x = ∞,<br />

⎪⎩<br />

∞ for x = 0


156 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

and<br />

⎧<br />

⎨ lim ξ↓x U i (ξ) − U i (x) for x ≠ 0 and x ≠ ∞,<br />

∆U i (x) :=<br />

⎩<br />

0 for x = ∞ or x = 0.<br />

L<strong>et</strong> m be the measure on ((−∞, ∞] d \ {0}) × [0, 1] d × R <strong>de</strong>fined by<br />

m := ˜ν ⊗ λ| (0,1) d ⊗ δ 0 +<br />

d∑<br />

i=1<br />

δ (0,...,0<br />

,∞,0,...,0) ⊗ δ (0,...,0 ) ⊗ λ| (νi ((0,∞)),∞)∪(−∞,−ν<br />

}{{}<br />

}{{}<br />

i ((−∞,0))),<br />

i−1<br />

d<br />

where ˜ν is the ext<strong>en</strong>sion of ν to (−∞, ∞] d \ {0}, i.e. ˜ν(B) := ν(B ∩ R d ). L<strong>et</strong><br />

g i : (−∞, ∞] × [0, 1] × R → (−∞, ∞],<br />

(x, y, z) ↦→ Ũi(x) + y∆U i (x) + z<br />

and <strong>de</strong>fine a measure ˜m on (−∞, ∞] d \ {∞, . . . , ∞} via<br />

˜m(B) := m(˜g −1 (B))<br />

with<br />

˜g(x 1 , . . . , x d , y 1 , . . . , y d , z) := (g 1 (x 1 , y 1 , z), . . . , g d (x d , y d , z)).<br />

Finally, l<strong>et</strong> F be giv<strong>en</strong> by<br />

⎧ ( d∏<br />

) d∏<br />

⎪⎨ ˜m (u i ∧ 0, u i ∨ 0] sgn u i , (u 1 , . . . , u d ) ∈ (−∞, ∞] d \ (∞, . . . , ∞)<br />

F (u 1 , . . . , u d ) := i=1<br />

i=1<br />

⎪⎩ ∞, (u 1 , . . . , u d ) = (∞, . . . , ∞).<br />

Properties 1 and 2 in Definition 4.16 are obvious. From the fact that ˜m is a positive measure<br />

it follows immediately that F is d-increasing. L<strong>et</strong> I ⊂ {1, . . . , d} nonempty and (u i ) i∈I ∈<br />

(−∞, ∞] I . For ease of notation, we consi<strong>de</strong>r only the case of non-negative u i . The g<strong>en</strong>eral case<br />

follows analogously. By <strong>de</strong>finition of F we have<br />

∑<br />

F I ((u i ) i∈I ) = lim<br />

F (u 1 , . . . , u d ) ∏<br />

sgn u j<br />

c→∞<br />

(u j ) j∈I c ∈{−c,∞} Ic j∈I c<br />

( )<br />

∏<br />

u i ] × (−∞, ∞]<br />

i∈I(0, Ic<br />

= ˜m<br />

= m( {<br />

(x 1 , . . . , x d , y 1 , . . . , y d , z) ∈ ((−∞, ∞] d \ {0}) × [0, 1] d × R :<br />

Ũ i (x i ) + y i ∆U i (x i ) + z ∈ (0, u i ] for i ∈ I} ) .


4.5. LEVY COPULAS: GENERAL CASE 157<br />

If I = {i}, th<strong>en</strong> the <strong>de</strong>finition of m implies that this equals<br />

(<br />

νi ⊗ λ| (0,1)<br />

) (<br />

{(x, y) ∈ R × [0, 1] : Ũ i (x) + y∆U i (x) ∈ (0, u i ]} ) + (u i − ν i ((0, ∞)))1 {ui >ν i ((0,∞))}.<br />

Introducing x ∗ := inf{x ≥ 0 : Ũi(x) + ∆U i (x) ≤ u i }, this can be expressed as<br />

ν i ((x ∗ , ∞)) + (u i − Ũi(x ∗ ) − ∆U i (x ∗ ))1 {x∗≠0} + (u i − ν i ((0, ∞)))1 {x ∗ =0} = u i ,<br />

i.e. property 4 in Definition 4.16 is m<strong>et</strong>.<br />

Now, l<strong>et</strong> (x i ) i∈I ∈ (R \ {0}) I . Again, we consi<strong>de</strong>r only the case where all the x i are nonnegative.<br />

Th<strong>en</strong><br />

F I ((U i (x i )) i∈I ) = m( {<br />

(˜x 1 , . . . , ˜x d , y 1 , . . . , y d , z) ∈ (−∞, ∞] d × [0, 1] d × R :<br />

= ν<br />

} )<br />

Ũ i (˜x i ) + y i ∆U i (˜x i ) + z ∈ (0, U i (x i )] for i ∈ I<br />

( ∏<br />

i∈I[x i , ∞) × R Ic )<br />

= ν I ( ∏<br />

i∈I[x i , ∞)<br />

= U I ((x i ) i∈I )<br />

as claimed. The uniqu<strong>en</strong>ess statem<strong>en</strong>t follows from (4.24) and Lemma 4.4.<br />

The converse statem<strong>en</strong>t.<br />

)<br />

Since F is d-increasing and continuous (by Lemma 4.4), there<br />

exists a unique measure µ on (−∞, ∞] d \ {∞, . . . , ∞} such that V F (|a, b|) = µ((a, b]) for any<br />

a, b ∈ (−∞, ∞] d \ {∞, . . . , ∞} with a ≤ b. (see [60], Section 4.5). For a one-dim<strong>en</strong>sional tail<br />

integral U(x), we <strong>de</strong>fine<br />

⎧<br />

⎨<br />

U (−1) sup{x > 0 : U(x) ≥ u} ∨ 0, u ≥ 0<br />

(u) :=<br />

⎩<br />

sup{x < 0 : U(x) ≥ u}, u < 0.<br />

L<strong>et</strong> ˜ν := f(µ) be the image of µ un<strong>de</strong>r<br />

f : (u 1 , . . . , u d ) ↦→ (U (−1)<br />

1 (u 1 ), . . . , U (−1)<br />

d<br />

(u d ))<br />

(4.25)<br />

and l<strong>et</strong> ν be the restriction of ˜ν to R d \ {0}. We need to prove that ν is a Lévy measure and<br />

that its marginal tail integrals U I ν satisfy<br />

U I ν ((x i ) i∈I ) = F I ((U i (x i )) i∈I )


158 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

for any non-empty I ⊂ {1, . . . , d} and any (x i ) i∈I ∈ (R \ {0}) I . Suppose for ease of notation<br />

that x i > 0, i ∈ I. Th<strong>en</strong><br />

U I ν ((x i ) i∈I ) = ν({ξ ∈ R d : ξ i ∈ [x i , ∞), i ∈ I})<br />

= µ({u ∈ (−∞, ∞] d : U (−1)<br />

i<br />

(u i ) ∈ [x i , ∞), i ∈ I})<br />

= µ({u ∈ (−∞, ∞] d : 0 < u i ≤ U(x i ), i ∈ I}).<br />

= F I ((U i (x i )) i∈I ).<br />

This proves in particular that the one-dim<strong>en</strong>sional marginal tail integrals of ν equal U 1 , . . . , U d .<br />

Since the marginals ν i of ν are Lévy measures on R, we have ∫ (x 2 i ∧ 1)ν i(dx i ) < ∞ for<br />

i = 1, . . . , d. This implies<br />

∫<br />

∫<br />

(|x| 2 ∧ 1)ν(dx) ≤<br />

d∑<br />

d∑<br />

∫<br />

(x 2 i ∧ 1)ν(dx)<br />

i=1<br />

i=1<br />

(x 2 i ∧ 1)ν i (dx i ) < ∞<br />

and h<strong>en</strong>ce ν is a Lévy measure on R d . The uniqu<strong>en</strong>ess of ν follows from the fact that it is<br />

uniquely <strong>de</strong>termined by its marginal tail integrals (cf. Lemma 4.7).<br />

4.6 Examples of Lévy copulas<br />

In this section we <strong>de</strong>rive the form of Lévy copulas corresponding to special <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures<br />

of Lévy processes: in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce, compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce and the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of stable<br />

processes. Examples of param<strong>et</strong>ric families of Lévy copulas will be giv<strong>en</strong> in the next chapter.<br />

To characterize in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce of compon<strong>en</strong>ts of a multidim<strong>en</strong>sional Lévy process in terms of its<br />

Lévy copula, we need to restate Lemma 4.2 in terms of tail integrals.<br />

Lemma 4.9. The compon<strong>en</strong>ts X 1 , . . . , X d of an R d -valued Lévy process X are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt if<br />

and only if their continuous martingale parts are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and the tail integrals of the Lévy<br />

measure satisfy U I ((x i ) i∈I ) = 0 for all I ⊂ {1, . . . , d} with card I ≥ 2 and all (x i ) i∈I ∈ (R\{0}) I .<br />

Proof. The “only if” part. L<strong>et</strong> I ⊂ {1, . . . , d} with card I ≥ 2 and (x i ) i∈I ∈ (R \ {0}) I . Th<strong>en</strong><br />

the compon<strong>en</strong>ts of the Lévy process (X i ) i∈I are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt as well. Applying Lemma 4.2 to<br />

this process, we conclu<strong>de</strong>, using Equation (4.2), that U I ((x i ) i∈I ) = 0.<br />

The “if” part. L<strong>et</strong> ν be <strong>de</strong>fined by Equation (4.2), where ν i is the Lévy measure of X i for<br />

i = 1, . . . , d. Th<strong>en</strong> all marginal tail integrals of ν coinci<strong>de</strong> with those of the Lévy measure of


4.6. EXAMPLES OF LEVY COPULAS 159<br />

X. Therefore, ν is the Lévy measure of X (cf. Lemma 4.7), which <strong>en</strong>tails by Lemma 4.2 that<br />

X 1 , . . . , X d are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt.<br />

Theorem 4.10. The compon<strong>en</strong>ts X 1 , . . . , X d of an R d -valued Lévy process X are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt<br />

if and only if their Brownian motion parts are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and X has a Lévy copula of the<br />

form<br />

F ⊥ (x 1 , . . . , x d ) :=<br />

d∑ ∏<br />

x i 1 {∞} (x j ) (4.26)<br />

i=1 j≠i<br />

Proof. It is straightforward to see that Equation (4.26) <strong>de</strong>fines a Lévy copula.<br />

The “if” part. L<strong>et</strong> I ⊂ {1, . . . , d} with card I ≥ 2. Equation (4.23) <strong>en</strong>tails that F⊥ I ((u i) i∈I ) =<br />

0 for all (u i ) i∈I ∈ R I . Therefore, the tail integrals of X satisfy U I ((x i ) i∈I ) = 0 for all (x i ) i∈I ∈<br />

(R \ {0}) I by (4.24). From Lemma 4.9 we conclu<strong>de</strong> that X 1 , . . . , X d are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt.<br />

The “only if” part. By Lemma 4.9, the tail integrals of X satisfy U I ((x i ) i∈I ) = 0 for all<br />

(x i ) i∈I ∈ (R \ {0}) I and all I ⊂ {1, . . . , d} with card I ≥ 2. Since also F⊥ I ((u i) i∈I ) = 0, we<br />

conclu<strong>de</strong> that F ⊥ is a Lévy copula for X.<br />

Observing that F ⊥ | [0,∞] d<br />

following easy corollary.<br />

is a Lévy copula in the s<strong>en</strong>se of Definition 4.14, we obtain the<br />

Corollary 4.1. The compon<strong>en</strong>ts X 1 , . . . , X d of an R d -valued spectrally positive Lévy process X<br />

are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt if and only if their Brownian motion parts are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and X has a Lévy<br />

copula on [0, ∞] d of the form F ⊥ | [0,∞] d, where F ⊥ is <strong>de</strong>fined by (4.26).<br />

The following theorem <strong>de</strong>scribes compl<strong>et</strong>e jump <strong>de</strong>p<strong>en</strong><strong>de</strong>nce in terms of Lévy copulas.<br />

Theorem 4.11. L<strong>et</strong> X be a R d -valued Lévy process whose Lévy measure is supported by an<br />

or<strong>de</strong>red s<strong>et</strong> S ⊂ K, where K is as in Equation (4.3). Th<strong>en</strong> the compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce Lévy<br />

copula giv<strong>en</strong> by<br />

d∏<br />

F ‖ (x 1 , . . . , x d ) := min(|x 1 |, . . . , |x d |)1 K (x 1 , . . . , x d ) sgn x i (4.27)<br />

is a Lévy copula of X.<br />

Conversely, if F ‖ is a Lévy copula of X, th<strong>en</strong> the Lévy measure of X is supported by an<br />

or<strong>de</strong>red subs<strong>et</strong> of K. If, in addition, the tail integrals U i of X i are continuous and satisfy<br />

lim x→0 U i (x) = ∞, i = 1, . . . , d, th<strong>en</strong> the jumps of X are compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt.<br />

i=1


160 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

The proof is based on the following repres<strong>en</strong>tation of an or<strong>de</strong>red s<strong>et</strong> as a union of a strictly<br />

or<strong>de</strong>red s<strong>et</strong> and countable many segm<strong>en</strong>ts that are perp<strong>en</strong>dicular to some coordinate axis.<br />

Lemma 4.12. L<strong>et</strong> S ⊂ R d be an or<strong>de</strong>red s<strong>et</strong>. It can be writt<strong>en</strong> as<br />

∞⋃<br />

S = S ∗ ∪ S n , (4.28)<br />

n=1<br />

where S ∗ ⊂ R d is strictly or<strong>de</strong>red and for every n, S n ⊂ R d and there exist k(n) and ξ(n) such<br />

that x k(n) = ξ(n) for all x ∈ S n .<br />

Proof. For the purposes of this proof we <strong>de</strong>fine the l<strong>en</strong>gth of an or<strong>de</strong>red s<strong>et</strong> S ′ by |S ′ | :=<br />

∑<br />

sup d<br />

a,b∈S ′ i=1 (a i − b i ). L<strong>et</strong><br />

S(ξ, k) = {x ∈ R d : x k = ξ} ∩ S. (4.29)<br />

First, we want to prove that there is at most a countable number of such segm<strong>en</strong>ts with nonzero<br />

l<strong>en</strong>gth. Consi<strong>de</strong>r N differ<strong>en</strong>t segm<strong>en</strong>ts of this type (S i := S(ξ i , k i )) N i=1 with k i = k and<br />

|S i | ≥ ε > 0 for all i. Since S i are differ<strong>en</strong>t, ξ i must all be differ<strong>en</strong>t and we can suppose without<br />

loss of g<strong>en</strong>erality that ξ i < ξ i+1 for all i. Th<strong>en</strong> x i ≤ x i+1 for all i, where x i and x i are the upper<br />

and the lower bounds of S i . Since all S i are subs<strong>et</strong>s of S, which is an or<strong>de</strong>red s<strong>et</strong>, this implies<br />

that | ⋃ N<br />

i=1 S i| ≥ Nε. Therefore, for all A > 0 and for all ε > 0, the s<strong>et</strong> [−A, A] d contains a<br />

finite number of segm<strong>en</strong>ts of type (4.29) with l<strong>en</strong>gth greater or equal to ε. This means that<br />

there is at most a countable number of segm<strong>en</strong>ts of non-zero l<strong>en</strong>gth, and one can <strong>en</strong>umerate<br />

them in a sequ<strong>en</strong>ce {S n } ∞ n=1 with S n := S(ξ(n), k(n)).<br />

Now l<strong>et</strong> S ∗ = S \ ⋃ ∞<br />

n=1 S n. S ∗ is or<strong>de</strong>red because it is a subs<strong>et</strong> of S. L<strong>et</strong> x, y ∈ S ∗ . If<br />

x k = y k for some k th<strong>en</strong> either x and y are the same or they are in some segm<strong>en</strong>t of type (4.29)<br />

h<strong>en</strong>ce not in S ∗ . Therefore, either x k < y k for every k or x k > y k for every k, which <strong>en</strong>tails<br />

that S ∗ is strictly or<strong>de</strong>red and we have obtained the <strong>de</strong>sired repres<strong>en</strong>tation for S.<br />

Proof of Theorem 4.11. We start by proving that F ‖ is in<strong>de</strong>ed a Lévy copula in the s<strong>en</strong>se of<br />

Definition 4.16. Properties 1 and 2 are obvious. To show property 3, introduce a positive<br />

measure µ on ¯R d by<br />

µ(B) = λ({x ∈ R : (x, . . . , x) ∈ B}), B ∈ B(¯R d ),<br />

} {{ }<br />

d times


4.6. EXAMPLES OF LEVY COPULAS 161<br />

where λ <strong>de</strong>notes the Lebesgue measure on R. Th<strong>en</strong> V F‖ ((a, b]) = µ((a, b]) for any a ≤ b, and<br />

therefore F ‖ is d-increasing. The margins of F have the same form as F , namely<br />

F‖ I ((x i) i∈I ) = min |x i|1 {(−1,...,−1),(1,...,1)} ((sgn x i ) i∈I ) ∏ sgn x i . (4.30)<br />

i∈I<br />

i∈I<br />

Therefore, the one-dim<strong>en</strong>sional margins satisfy F {i} (u) = u.<br />

The first part. L<strong>et</strong> x ∈ (0, ∞) d . Clearly, U(x) ≤ U k (x k ) for any k. On the other hand, since<br />

S is an or<strong>de</strong>red s<strong>et</strong>, we have<br />

{y ∈ R d : x k ≤ y k } ∩ S = {y ∈ R d : x ≤ y} ∩ S<br />

for some k. In<strong>de</strong>ed, suppose that this is not so. Th<strong>en</strong> there exist points z 1 , . . . , z d ∈ S and<br />

indices j 1 , . . . , j d such that zk k ≥ x k and zj k k<br />

< x jk for k = 1, . . . , d. Choose the greatest elem<strong>en</strong>t<br />

among z 1 , . . . , z d (this is possible because they all belong to an or<strong>de</strong>red s<strong>et</strong>) and call it z k .<br />

Th<strong>en</strong> z k j k<br />

< x jk . However, by construction of z 1 , . . . , z d we also have z j k<br />

jk<br />

contradiction to the fact that z k is the greatest elem<strong>en</strong>t. Therefore,<br />

U(x) = min(U 1 (x 1 ), . . . , U d (x d )).<br />

≥ x jk , which is a<br />

Similarly, it can be shown that for every x ∈ (−∞, 0) d ,<br />

U(x) = (−1) d min(|U 1 (x 1 )|, . . . , |U d (x d )|).<br />

Since U(x) = 0 for any x /∈ K, we have shown that<br />

U(x) = F ‖ (U 1 (x 1 ), . . . , U d (x d ))<br />

for any x ∈ (R \ {0}) d . Since the marginal Lévy measures of X are also supported by or<strong>de</strong>red<br />

s<strong>et</strong>s and the margins of F ‖ have the same form as F ‖ , we have<br />

U I ((x i ) i∈I ) = F I ‖ ((U i(x i )) i∈I ) (4.31)<br />

for any I ⊂ {1, . . . , d} and any (x i ) i∈I ∈ (R \ {0}) I .<br />

The converse statem<strong>en</strong>t. L<strong>et</strong> S := supp ν. L<strong>et</strong> us first show that S ⊆ K. Suppose that<br />

this is not so. Th<strong>en</strong> there exists x ∈ S such that for some m and n, x m < 0 and x n > 0 and<br />

for every neighborhood N of x, ν(N) > 0. This implies that U {m,n} (x m /2, x n /2) > 0, which<br />

contradicts Equation (4.31).


162 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

Suppose now that S is not an or<strong>de</strong>red s<strong>et</strong>.<br />

Th<strong>en</strong> there exist two points u, v ∈ S such<br />

that u m > v m and u n < v n for some m and n. Moreover, either u i ≥ 0 and v i ≥ 0 for all<br />

i or u i ≤ 0 and v i ≤ 0 for all i. Suppose that u i ≥ 0 and v i ≥ 0, the other case being<br />

analogous. L<strong>et</strong> x = u+v<br />

2 . Since u, v ∈ S, we have ν({z ∈ Rd : z m < x m , z n ≥ x n }) > 0 and<br />

ν({z ∈ R d : z m ≥ x m , z n < x n }) > 0. However<br />

ν({z ∈ R d : z m < x m , z n ≥ x n }) = U n (x n ) − U {m,n} (x m , x n )<br />

= U n (x n ) − min(U m (x m ), U n (x n ))<br />

and<br />

ν({z ∈ R d : z m ≥ x m , z n < x n }) = U m (x m ) − min(U m (x m ), U n (x n )),<br />

which is a contradiction because these expressions cannot be simultaneously positive.<br />

For the last assertion, we assume that the tail integrals U i of X i are continuous and satisfy<br />

lim x→0 U i (x) = ∞, i = 1, . . . , d. It suffices to show that ν(S n ) = 0 for any n in <strong>de</strong>composition<br />

(4.28). If ξ(n) ≠ 0, th<strong>en</strong><br />

ν(S n ) = lim<br />

ε↓0<br />

(U k(n) (ξ(n) − ε) − U k(n) (ξ(n))) = 0,<br />

because U k(n) is continuous. Suppose now that ξ(n) = 0. Since S n does not reduce to a single<br />

point, we must have either x m > 0 or x m < 0 for some x ∈ S n and some m. Suppose that<br />

x m > 0, the other case being analogous. Since S is or<strong>de</strong>red, we have<br />

ν({x ∈ R d : x k(n) ≥ ε} ∩ S) ≤ ν({ξ ∈ R d : ξ m ≥ x m } ∩ S) < ∞<br />

uniformly in ε > 0. This implies lim x↓0 U k(n) (x) < ∞ in contradiction to lim x→0 U k(n) (x) = ∞.<br />

H<strong>en</strong>ce, ξ(n) > 0 for any n. Therefore, ν(R d \ S ∗ ) = 0 and the proof is compl<strong>et</strong>ed.<br />

A characterization of compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of spectrally positive Lévy processes can be<br />

obtained as a corollary of Theorem 4.11.<br />

Corollary 4.2. L<strong>et</strong> X be a R d -valued spectrally positive Lévy process whose Lévy measure is<br />

supported by an or<strong>de</strong>red s<strong>et</strong> S ⊂ R d +. Th<strong>en</strong><br />

F ‖ (x 1 , . . . , x d )| [0,∞] d ≡ min(x 1 , . . . , x d )<br />

is a Lévy copula of X.


4.6. EXAMPLES OF LEVY COPULAS 163<br />

Conversely, if F ‖ | [0,∞] d is a Lévy copula of a spectrally positive Lévy process X, th<strong>en</strong> the<br />

Lévy measure of X is supported by an or<strong>de</strong>red subs<strong>et</strong> of R d +. If, in addition, the tail integrals<br />

U i of X i are continuous and satisfy lim x↓0 U i (x) = ∞, i = 1, . . . , d, th<strong>en</strong> the jumps of X are<br />

compl<strong>et</strong>ely <strong>de</strong>p<strong>en</strong><strong>de</strong>nt.<br />

Lévy copulas provi<strong>de</strong> a simple characterization of possible <strong>de</strong>p<strong>en</strong><strong>de</strong>nce patterns of multidim<strong>en</strong>sional<br />

stable processes.<br />

Theorem 4.13. L<strong>et</strong> X := (X 1 , . . . , X d ) be a Lévy process on R d and l<strong>et</strong> α ∈ (0, 2). X is<br />

α-stable if and only if its compon<strong>en</strong>ts X 1 , . . . , X d are α-stable and it has a Lévy copula F that<br />

is a homog<strong>en</strong>eous function of or<strong>de</strong>r 1:<br />

∀r > 0, ∀u 1 , . . . , u d , F (ru 1 , . . . , ru d ) = rF (u 1 , . . . , u d ). (4.32)<br />

Proof. The “only if” part. L<strong>et</strong> X be α-stable. For each i = 1, . . . , d, three situations are<br />

possible: Ran U i = (−∞, 0] (only negative jumps), Ran U i = [0, ∞) (only positive jumps) or<br />

Ran U i = (−∞, 0) ∪ (0, ∞) (jumps of both signs). We exclu<strong>de</strong> the trivial case of a compon<strong>en</strong>t<br />

having no jumps at all. L<strong>et</strong> I 1 = {i : Ran U i = (−∞, 0]} and I 2 = {i : Ran U i = [0, ∞)} and<br />

for each i, l<strong>et</strong> ¯X i be a copy of X i , in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt from X and from ¯X k for k ≠ i. Define a Lévy<br />

process ˜X on R d by<br />

⎧<br />

⎨ X i , i /∈ I 1 ∪ I 2 ,<br />

˜X i =<br />

⎩<br />

X i − ¯X i , i ∈ I 1 ∪ I 2 .<br />

L<strong>et</strong> ˜ν be the Lévy measure of ˜X, Ũ be its tail integral and ˜F be its Lévy copula (it exists by<br />

Theorem 4.8). The process ˜X is clearly α-stable and each compon<strong>en</strong>t of this process has jumps<br />

of both signs (Ran Ũi = R \ {0}). By Theorem 14.3 in [87], for every B ∈ B(R d \ {0}) and for<br />

every r > 0,<br />

˜ν(B) = r α˜ν(rB). (4.33)<br />

Therefore, for every I ⊂ {1, . . . , d} nonempty and for every (x i ) i∈I ∈ (R \ {0}) |I| ,<br />

Ũ I ((x i ) i∈I ) = r α Ũ I ((rx i ) i∈I ). (4.34)<br />

By Theorem 4.8 this implies that for all (u 1 , . . . , u d ) ∈ (R \ {0}) d ,<br />

˜F I ((u i ) i∈I ) = r −1 ˜F I ((ru i ) i∈I ),


164 CHAPTER 4. DEPENDENCE OF LEVY PROCESSES<br />

and therefore (4.32) holds for ˜F . However, ˜F is also the copula of X. In<strong>de</strong>ed, l<strong>et</strong> I ⊂ {1, . . . , d}<br />

nonempty and (x i ) i∈I ∈ (R \ {0}) |I| . Two situations are possible:<br />

• For every i ∈ I, Ũi(x i ) = U(x i ). Th<strong>en</strong> it is easy to see that Equation (4.24) holds with F<br />

replaced by ˜F .<br />

• For some k ∈ I, U k (x k ) = 0. Th<strong>en</strong> ˜F I ((U i (x i )) i∈I ) = 0, but on the other hand<br />

|U I ((x i ) i∈I )| ≤ |U k (x k )| = 0, and (4.24) also holds.<br />

The “if” part. L<strong>et</strong> X have α-stable margins and a homog<strong>en</strong>eous Lévy copula. Th<strong>en</strong> the<br />

marginal tail integrals of X satisfy (4.34), and since by Lemma 4.7, the Lévy measure of every<br />

s<strong>et</strong> of the form [a, b) can be expressed as a linear combination of tail integrals, the Lévy measure<br />

of X has the property (4.33) and we conclu<strong>de</strong> by Theorem 14.3 in [87] that X is α-stable.


Chapter 5<br />

Applications of Lévy copulas<br />

To apply Lévy copulas to multidim<strong>en</strong>sional financial problems with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> ass<strong>et</strong>s,<br />

three kinds of tools are required:<br />

1. Param<strong>et</strong>ric families of Lévy copulas. Parsimonious mo<strong>de</strong>ls are nee<strong>de</strong>d because one typically<br />

does not have <strong>en</strong>ough information about the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce to estimate many param<strong>et</strong>ers<br />

or proceed with a non-param<strong>et</strong>ric approach.<br />

2. Algorithms allowing to compute various quantities within a Lévy copula mo<strong>de</strong>l (e.g.,<br />

option prices, risk measures <strong>et</strong>c.)<br />

3. Estimation m<strong>et</strong>hods for Lévy copula mo<strong>de</strong>ls.<br />

In this chapter we give an answer to the first two questions. Section 5.1 <strong>de</strong>scribes several<br />

m<strong>et</strong>hods to construct param<strong>et</strong>ric families of Lévy copulas and gives examples of such families. In<br />

Section 5.2 we show how Lévy copulas can be used to simulate multidim<strong>en</strong>sional Lévy processes<br />

with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> compon<strong>en</strong>ts. This simulation algorithm <strong>en</strong>ables us to compute the<br />

quantities of interest using the Monte Carlo m<strong>et</strong>hod. Estimation m<strong>et</strong>hods for Lévy copulas are<br />

the topic of our curr<strong>en</strong>t research and we do not discuss them here.<br />

A fundam<strong>en</strong>tal advantage of the Lévy copula approach compared to ordinary copulas is the<br />

possibility to work with several time scales at the same time. Wh<strong>en</strong> the time scale is fixed,<br />

the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure of r<strong>et</strong>urns can be <strong>de</strong>scribed with an ordinary copula. However, on a<br />

differ<strong>en</strong>t time scale the copula will not be the same (cf. Example 4.1). On the other hand, a<br />

165


166 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

Lévy copula of a Lévy process is not linked to any giv<strong>en</strong> time scale; it <strong>de</strong>scribes the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

structure of the <strong>en</strong>tire process.<br />

One example of a financial mo<strong>de</strong>lling problem where Lévy copulas naturally appear is the<br />

pricing of multi-ass<strong>et</strong> options on un<strong>de</strong>rlyings, <strong>de</strong>scribed by expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls. In this<br />

s<strong>et</strong>ting the param<strong>et</strong>ers of the marginal Lévy processes can be calibrated from the prices of European<br />

options, quoted at the mark<strong>et</strong>, and the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure will typically be estimated<br />

from the historical time series of r<strong>et</strong>urns. Since the sampling rate of r<strong>et</strong>urns is differ<strong>en</strong>t from<br />

the maturity of tra<strong>de</strong>d options as well as from the maturity of the bask<strong>et</strong> option that one wants<br />

to price, ordinary copulas are not well suited for this problem. Pricing of bask<strong>et</strong> options in a<br />

Lévy copula mo<strong>de</strong>l is discussed in more <strong>de</strong>tail in Section 5.3.<br />

Lévy copulas can also be useful outsi<strong>de</strong> the realm of financial mo<strong>de</strong>lling. Other contexts<br />

where mo<strong>de</strong>lling <strong>de</strong>p<strong>en</strong><strong>de</strong>nce in jumps is required are portfolios of insurance claims and mo<strong>de</strong>ls<br />

of operational risk.<br />

Consi<strong>de</strong>r an insurance company with two subsidiaries, in France and in Germany. The<br />

aggregate loss process of the Fr<strong>en</strong>ch subsidiary is mo<strong>de</strong>lled by the subordinator {X t } t≥0 and<br />

the loss process of the German one is {Y t } t≥0 . The nature of processes X and Y may be<br />

differ<strong>en</strong>t because the subsidiaries may not be working in the same sector and many risks that<br />

cause losses are local. However, common risks like floods and pan-European windstorms will<br />

lead to a certain <strong>de</strong>gree of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> the claims. In this s<strong>et</strong>ting it is conv<strong>en</strong>i<strong>en</strong>t<br />

to mo<strong>de</strong>l the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> X and Y using a Lévy copula on [0, ∞] 2 . In this mo<strong>de</strong>lling<br />

approach, the two-dim<strong>en</strong>sional Lévy measure of (X, Y ) is known and the overall loss distribution<br />

and ruin probability can be computed.<br />

Another example where jump processes naturally appear is giv<strong>en</strong> by mo<strong>de</strong>ls of operational<br />

risk. The 2001 Basel agreem<strong>en</strong>t <strong>de</strong>fines the operational risk as “the risk of direct and indirect<br />

loss resulting from ina<strong>de</strong>quate or failed internal processes, people and systems or from external<br />

ev<strong>en</strong>ts” and allows banks to use internal loss data to compute regulatory capital requirem<strong>en</strong>ts.<br />

Taking into account the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> differ<strong>en</strong>t business lines in this computation, due to<br />

a diversification effect, may lead to substantial reduction of regulatory capital [14]. Aggregate<br />

loss processes from differ<strong>en</strong>t business lines can be dynamically mo<strong>de</strong>lled by subordinators and<br />

the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> them can be accounted for using a Lévy copula on [0, ∞] 2 .


5.1. PARAMETRIC FAMILIES 167<br />

The applications of Lévy copulas are not limited to constructing expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls.<br />

Lévy copula mo<strong>de</strong>ls can be time changed to obtain multidim<strong>en</strong>sional analogs of stochastic<br />

volatility mo<strong>de</strong>ls discussed in [22]. More g<strong>en</strong>erally, since a large class of Markov processes or<br />

ev<strong>en</strong> semimartingales behaves locally as a Lévy process in the s<strong>en</strong>se that its dynamics can be<br />

<strong>de</strong>scribed by a drift rate, a covariance matrix, and a Lévy measure, which may all change<br />

randomly through time (cf. e.g. [54], II.2.9, II.4.19), Lévy copulas could be used to <strong>de</strong>scribe<br />

<strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> processes of these types.<br />

5.1 Param<strong>et</strong>ric families of Lévy copulas<br />

Our first result is a m<strong>et</strong>hod to construct Lévy copulas on [0, ∞] d from ordinary copulas.<br />

Theorem 5.1 (Construction of Lévy copulas from ordinary copulas). L<strong>et</strong> C be a copula<br />

on [0, 1] d and φ : [0, 1] → [0, ∞] be a strictly increasing continuous function with φ(1) = ∞,<br />

φ(0) = 0, having nonnegative <strong>de</strong>rivatives of or<strong>de</strong>rs up to d on (0, 1). Th<strong>en</strong><br />

F (u 1 , . . . , u d ) := φ(C(φ −1 (u 1 ), . . . , φ −1 (u d )))<br />

is a Lévy copula on [0, ∞] d .<br />

Proof. First, note that φ −1 is well <strong>de</strong>fined and satisfies φ −1 (0) = 0 and φ −1 (∞) = 1. Therefore,<br />

properties 1 and 2 of Definition 4.14 are clear, and in view of Equation (4.5), property 4 also<br />

holds. It remains to show that F is a d-increasing function, and since φ −1 is strictly increasing,<br />

it suffices to prove that the function φ(C(u 1 , . . . , u d )) is d-increasing on [0, 1] d .<br />

To this <strong>en</strong>d, l<strong>et</strong> us show by induction on d that if H : [0, 1] d → [0, 1] is d-increasing and<br />

groun<strong>de</strong>d and ψ : [0, 1] → [0, ∞] is a continuous increasing function with ψ(0) = 0 and positive<br />

<strong>de</strong>rivatives of or<strong>de</strong>rs up to d on (0, 1) th<strong>en</strong> ψ(H) is also d-increasing and groun<strong>de</strong>d. For d = 1,<br />

the result is clear. Suppose d ≥ 2. For k = 1, . . . , d l<strong>et</strong> a k , b k ∈ [0, 1] with a k ≤ b k . The function<br />

˜H(u 2 , . . . , u d ) := ψ(H(b 1 , u 2 , . . . , u d )) − ψ(H(a 1 , u 2 , . . . , u d )) satisfies<br />

V ψ(H) (|a 1 , b 1 | × · · · × |a d , b d |) = V ˜H(|a 2 , b 2 | × · · · × |a d , b d |),<br />

h<strong>en</strong>ce, it remains to prove that ˜H is d − 1-increasing. It can be repres<strong>en</strong>ted as follows (“∗”


168 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

stands for u 2 , . . . , u d ):<br />

˜H(∗) =<br />

∫ H(b1 ,∗)<br />

H(a 1 ,∗)<br />

ψ ′ (t)dt<br />

= (H(b 1 , ∗) − H(a 1 , ∗))<br />

∫ 1<br />

= (H(b 1 , ∗) − H(a 1 , ∗))ψ ′ +(0)<br />

+ (H(b 1 , ∗) − H(a 1 , ∗))<br />

0<br />

∫ 1<br />

0<br />

ψ ′ (H(a 1 , ∗) + t(H(b 1 , ∗) − H(a 1 , ∗)))dt<br />

{ψ ′ (H(a 1 , ∗) + t(H(b 1 , ∗) − H(a 1 , ∗))) − ψ ′ +(0)}dt.<br />

Note that ψ ′ +(0), the right <strong>de</strong>rivative of ψ at 0, is well <strong>de</strong>fined because we have supposed that<br />

d ≥ 2 and therefore ψ ′′ exists on (0, 1) and ψ ′ is increasing. In the right-hand si<strong>de</strong> of the above<br />

equation:<br />

• The integrand is d−1-increasing by the induction hypothesis, because H(a 1 , ∗)+t(H(b 1 , ∗)−<br />

H(a 1 , ∗)) is d − 1-increasing and groun<strong>de</strong>d for every t and ˜ψ(t) := ψ ′ (t) − ψ ′ +(0) is an<br />

increasing continuous function with ˜ψ(0) = 0 and nonnegative <strong>de</strong>rivatives of or<strong>de</strong>rs up to<br />

d − 1 on (0, 1).<br />

• The second term is d − 1-increasing by Lemma 4.3,<br />

• The first term is d − 1-increasing because H is,<br />

h<strong>en</strong>ce ˜H is d − 1-increasing as sum of d − 1-increasing functions and the proof is compl<strong>et</strong>e.<br />

One example of an absolutely monotonic (with positive <strong>de</strong>rivatives of all or<strong>de</strong>rs) function<br />

that maps [0, 1] into [0, ∞] is φ(x) =<br />

x<br />

1−x .<br />

The following result allows to construct Lévy copulas on (−∞, ∞] d , analogous to the<br />

Archime<strong>de</strong>an copulas (cf. [76]). It can be used to build param<strong>et</strong>ric families of Lévy copulas<br />

in arbitrary dim<strong>en</strong>sion, where the number of param<strong>et</strong>ers does not <strong>de</strong>p<strong>en</strong>d on the dim<strong>en</strong>sion.<br />

Theorem 5.2 (Archime<strong>de</strong>an Lévy copulas). L<strong>et</strong> φ : [−1, 1] → [−∞, ∞] be a strictly increasing<br />

continuous function with φ(1) = ∞, φ(0) = 0, and φ(−1) = −∞, having <strong>de</strong>rivatives of<br />

or<strong>de</strong>rs up to d on (−1, 0) and (0, 1), and satisfying<br />

L<strong>et</strong><br />

d d φ(e x )<br />

dx d ≥ 0,<br />

d d φ(−e x )<br />

dx d ≤ 0, x ∈ (−∞, 0). (5.1)<br />

˜φ(u) := 2 d−2 {φ(u) − φ(−u)}


5.1. PARAMETRIC FAMILIES 169<br />

for u ∈ [−1, 1]. Th<strong>en</strong><br />

<strong>de</strong>fines a Lévy copula.<br />

F (u 1 , . . . , u d ) := φ<br />

( d∏<br />

i=1<br />

˜φ −1 (u i )<br />

)<br />

, (u 1 , . . . , u d ) ∈ (−∞, ∞] d<br />

Proof. Firstly, note that ˜φ is a strictly increasing continuous function from [−1, 1] to [−∞, ∞],<br />

satisfying ˜φ(1) = ∞ and ˜φ(−1) = −∞, which means that ˜φ −1 exists for all u ∈ (−∞, ∞] and<br />

F is well <strong>de</strong>fined. Properties 1 and 2 of Definition 4.16 are clearly satisfied. For k = 1, . . . , d<br />

and u k ∈ R we have<br />

F {k} (u k ) = lim<br />

=<br />

=<br />

c→∞<br />

∑<br />

sgn u i<br />

F (u 1 , . . . , u d ) ∏<br />

(u i ) i≠k ∈{−c,∞} d−1 i≠k<br />

⎛<br />

⎞<br />

∑<br />

⎝ ˜φ−1 (u k ) ∏ sgn u i<br />

⎠ ∏ sgn u i<br />

i≠k<br />

i≠k<br />

(u i ) i≠k ∈{−∞,∞} d−1 φ<br />

∑d−1<br />

( ) d − 1<br />

(<br />

(−1) i φ ˜φ−1 (u k )(−1) i)<br />

i<br />

i=0<br />

= 2 d−2 {φ( ˜φ −1 (u k )) − φ(− ˜φ −1 (u k ))} = u k ,<br />

which proves property 4. It remains to show that F is d-increasing. Since ˜φ −1 is increasing,<br />

we only need to show that (u 1 , . . . , u d ) ↦→ φ( ∏ d<br />

i=1 u i) is d-increasing on (−1, 1] d , and for this,<br />

because φ( ∏ d<br />

i=1 u i) = φ( ∏ d<br />

i=1 |u i| ∏ d<br />

i=1 sgn u i), it suffices to prove that both (u 1 , . . . , u d ) ↦→<br />

φ( ∏ d<br />

i=1 u i) and (u 1 , . . . , u d ) ↦→ −φ(− ∏ d<br />

i=1 u i) are d-increasing on [0, 1] d or, equival<strong>en</strong>tly, on<br />

(0, 1) d (since φ is continuous). The first condition of (5.1) implies that<br />

∂ d ψ(z 1 , . . . , z d )<br />

∂z 1 . . . ∂z d<br />

≥ 0<br />

on (−∞, 0) d for ψ(z 1 , . . . , z d ) := φ(e z 1+···+z d<br />

). From Definition 4.5 it follows easily that<br />

∫<br />

∂ d ψ(z 1 , . . . , z d )<br />

V ψ (B) =<br />

dz 1 . . . dz d .<br />

∂z 1 . . . ∂z d<br />

B<br />

Therefore, ψ is increasing on (−∞, 0) d , which implies that (u 1 , . . . , u d ) ↦→ φ( ∏ d<br />

i=1 u i) is d-<br />

increasing on (0, 1) d .<br />

The second condition of (5.1) <strong>en</strong>tails similarly that (u 1 , . . . , u d ) ↦→<br />

−φ(− ∏ d<br />

i=1 u i) is d-increasing on (0, 1) d as well.<br />

Remark 5.1. Condition (5.1) is satisfied in particular if for any k = 1, . . . , d,<br />

d k φ(u)<br />

du k ≥ 0, u ∈ (0, 1) and (−1) k dk φ(u)<br />

du k ≤ 0, u ∈ (−1, 0).


170 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

Corollary 5.1 (Archime<strong>de</strong>an Lévy copulas for spectrally positive processes). L<strong>et</strong><br />

ψ : [0, ∞] → [0, ∞] be a strictly <strong>de</strong>creasing continuous function with ψ(0) = ∞, ψ(∞) = 0,<br />

having <strong>de</strong>rivatives of or<strong>de</strong>rs up to d on (0, ∞) and satisfying<br />

for all x ∈ (0, ∞). Th<strong>en</strong><br />

(−1) d dd ψ(x)<br />

dx d ≥ 0<br />

is a Lévy copula on [0, ∞] d .<br />

d∑<br />

F (u 1 , . . . , u d ) := ψ( ψ −1 (u i )) (5.2)<br />

i=1<br />

Proof. L<strong>et</strong><br />

⎧<br />

⎨ ψ(− log u), u ≥ 0<br />

φ(u) =<br />

, u ∈ [−1, 1].<br />

⎩<br />

−ψ(− log(−u)), u < 0<br />

Th<strong>en</strong> φ satisfies the conditions of Theorem 5.2 and therefore<br />

( d∏<br />

)<br />

¯F (u 1 , . . . , u d ) := φ φ −1 (u i /2 d−1 )<br />

i=1<br />

is a Lévy copula on (−∞, ∞] d . This implies that the function ˜F := ¯F | [0,∞] d has properties 1,<br />

2 and 3 of Definition 4.14. However, it is easy to check that ˜F (u 1 , . . . , u d )| ui =∞,i≠k = u k /2 d−1<br />

and therefore<br />

d∑<br />

2 d−1 ˜F (u1 , . . . , u d ) = 2 d−1 ψ( ψ −1 (u i /2 d−1 ))<br />

is a Lévy copula on [0, ∞] d , which means that (5.2) also <strong>de</strong>fines a positive Lévy copula.<br />

i=1<br />

Example 5.1. L<strong>et</strong><br />

with θ > 0 and η ∈ (0, 1). Th<strong>en</strong><br />

φ(x) := η(− log |x|) −1/θ 1 x≥0 − (1 − η)(− log |x|) −1/θ 1 x


5.1. PARAMETRIC FAMILIES 171<br />

2.5<br />

2.5<br />

2<br />

2<br />

1.5<br />

1.5<br />

1<br />

1<br />

0.5<br />

0.5<br />

0<br />

0<br />

−0.5<br />

−0.5<br />

−1<br />

−1<br />

−1.5<br />

−1.5<br />

−2<br />

−2<br />

−2.5<br />

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5<br />

2.5<br />

−2.5<br />

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

−2<br />

−2.5<br />

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5<br />

Figure 5.1: Contour plots of Lévy <strong>de</strong>nsity with 1-stable margins and Lévy copula (5.4) with<br />

θ = 5 and differ<strong>en</strong>t values of η. Top left: η = 0. Top right: η = 1. Bottom: η = 0.5. On each<br />

curve the Lévy <strong>de</strong>nsity ν(x, y) is constant and it increases from outer curves to inner curves.<br />

<strong>de</strong>fines a two-param<strong>et</strong>er family of Lévy copulas (F is in fact a Lévy copula for all θ > 0 and<br />

η ∈ [0, 1]). The role of param<strong>et</strong>ers is most easy to analyze in the case d = 2, wh<strong>en</strong> (5.3) becomes<br />

F (u, v) = (|u| −θ + |v| −θ ) −1/θ (η1 uv≥0 − (1 − η)1 uv


172 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

2.5<br />

2.5<br />

2<br />

2<br />

1.5<br />

1.5<br />

1<br />

1<br />

0.5<br />

0.5<br />

0<br />

0<br />

−0.5<br />

−0.5<br />

−1<br />

−1<br />

−1.5<br />

−1.5<br />

−2<br />

−2<br />

−2.5<br />

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5<br />

2.5<br />

−2.5<br />

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5<br />

2.5<br />

2<br />

2<br />

1.5<br />

1.5<br />

1<br />

1<br />

0.5<br />

0.5<br />

0<br />

0<br />

−0.5<br />

−0.5<br />

−1<br />

−1<br />

−1.5<br />

−1.5<br />

−2<br />

−2<br />

−2.5<br />

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5<br />

−2.5<br />

−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5<br />

Figure 5.2: Contour plots of Lévy <strong>de</strong>nsity with 1-stable margins and Lévy copula (5.4) with<br />

η = 1. Top left: θ = 0.2. Top right: θ = 0.7. Bottom left: θ = 1. Bottom right: θ = 5. On each<br />

curve the Lévy <strong>de</strong>nsity ν(x, y) is constant and it increases from outer curves to inner curves.<br />

values of η, positive jumps in one compon<strong>en</strong>t can correspond to both positive and negative<br />

jumps in the other compon<strong>en</strong>t. The param<strong>et</strong>er θ is responsible for the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of absolute<br />

values of jumps in differ<strong>en</strong>t compon<strong>en</strong>ts: from Figure 5.2 it is se<strong>en</strong> that as θ grows from 0.2 to<br />

5, the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure changes from in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce to almost compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce.<br />

Theorem 5.2 can be used to construct parsimonious mo<strong>de</strong>ls of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce; this is typically<br />

useful wh<strong>en</strong> one has little information about the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure of the problem. If a more<br />

precise vision is necessary, a possible strategy is to mo<strong>de</strong>l the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce in each corner of the<br />

Lévy measure separately. A d-dim<strong>en</strong>sional Lévy copula can be constructed from 2 d positive


5.1. PARAMETRIC FAMILIES 173<br />

Lévy copulas (one for each orthant) as follows:<br />

Theorem 5.3. For each {α 1 , . . . , α d } ∈ {−1, 1} d l<strong>et</strong> g (α 1,...,α d ) (u) : [0, ∞] → [0, 1] be a nonnegative,<br />

increasing function satisfying<br />

∑<br />

α∈{−1,1} d with α k =−1<br />

g (α 1,...,α d ) (u) = 1<br />

and<br />

∑<br />

α∈{−1,1} d with α k =1<br />

g (α 1,...,α d ) (u) = 1<br />

for all u ∈ [0, ∞] and all k ∈ {1, . . . , d}. Moreover, l<strong>et</strong> F (α 1,...,α d ) be a positive Lévy copula<br />

that satisfies the following continuity property at infinity: for all I ⊂ {k : α k = −1}, (u i ) i∈I c ∈<br />

[0, ∞] Ic we have<br />

lim F (α 1,...,α d ) (u 1 , . . . , u d ) = F (α 1,...,α d ) (v 1 , . . . , v d ),<br />

{u i } i∈I →(∞,...,∞)<br />

where v i = u i for i ∈ I c and v i = ∞ otherwise. Th<strong>en</strong><br />

F (u 1 , . . . , u d ) :=<br />

( ) d∏<br />

F (sgn u 1,...,sgn u d )<br />

|u 1 |g (sgn u 1,...,sgn u d ) (|u 1 |), . . . , |u d |g (sgn u 1,...,sgn u d ) (|u d |) sgn u i<br />

<strong>de</strong>fines a Lévy copula.<br />

i=1<br />

Proof. Properties 1 and 2 of Definition 4.16 are obvious. Property 3 follows after observing that<br />

u ↦→ ug (α 1,...,α d ) (u) is increasing on [0, ∞] for any {α 1 , . . . , α d } ∈ {−1, 1} d . To prove property<br />

4, note that<br />

( )<br />

F (α 1,...,α d )<br />

|u 1 |g (α 1,...,α d ) (|u 1 |), . . . , |u d |g (α 1,...,α d ) (|u d |) = |u k |g (α 1,...,α d ) (|u k |)<br />

for any {α 1 , . . . , α d } ∈ {−1, 1} d and any {u 1 , . . . , u d } ∈ ¯R d with u i = ∞ for all i ≠ k. Therefore,<br />

⎧<br />

⎪⎨<br />

F {k} (u) =<br />

⎪⎩<br />

∑<br />

d∏ ∏<br />

ug (α 1,...,α d ) (u) α i α j if u ≥ 0<br />

α∈{−1,1} d with α k =1<br />

∑<br />

i=1<br />

|u|g (α 1,...,α d ) (|u|)<br />

j≠k<br />

α∈{−1,1} d with α k =−1<br />

i=1 j≠k<br />

d∏ ∏<br />

α i α j if u < 0<br />

= u.


174 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

Example 5.2. L<strong>et</strong><br />

⎧<br />

⎨<br />

g (α 1,...,α d ) 1 for α 1 = · · · = α d<br />

(u) =<br />

⎩<br />

0 otherwise.<br />

Th<strong>en</strong> the Lévy copula F in Theorem 5.3 satisfies F (u 1 , . . . , u d ) = 0 if u i u j < 0 for some i, j.<br />

This means that the Lévy measure is supported by the positive and the negative orthant: either<br />

all compon<strong>en</strong>ts of the process jump up or all compon<strong>en</strong>ts jump down.<br />

5.2 Simulation of multidim<strong>en</strong>sional <strong>de</strong>p<strong>en</strong><strong>de</strong>nt Lévy processes<br />

Lévy copulas turn out to be a conv<strong>en</strong>i<strong>en</strong>t tool for simulating multidim<strong>en</strong>sional Lévy processes<br />

with specified <strong>de</strong>p<strong>en</strong><strong>de</strong>nce. In this section we first give the necessary <strong>de</strong>finitions and two auxiliary<br />

lemmas and th<strong>en</strong> prove two theorems which show how multidim<strong>en</strong>sional Lévy processes<br />

with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures giv<strong>en</strong> by Lévy copulas can be simulated in the finite variation case<br />

(Theorem 5.6) and in the infinite variation case (Theorem 5.7).<br />

To simulate a Lévy process {X t } 0≤t≤1 on R d with Lévy measure ν, we will first simulate a<br />

Poisson random measure on [0, 1] × R d with int<strong>en</strong>sity measure λ [0,1] ⊗ ν, where λ <strong>de</strong>notes the<br />

Lebesgue measure. The Lévy process can th<strong>en</strong> be constructed via the Lévy-Itô <strong>de</strong>composition.<br />

L<strong>et</strong> F be a Lévy copula on (−∞, ∞] d such that for every I ∈ {1, . . . , d} nonempty,<br />

lim F (x 1, . . . , x d ) = F (x 1 , . . . , x d )| (xi ) i∈I =∞. (5.5)<br />

(x i ) i∈I →∞<br />

This Lévy copula <strong>de</strong>fines a positive measure µ on R d with Lebesgue margins such that for each<br />

a, b ∈ R d with a ≤ b,<br />

V F (|a, b|) = µ((a, b]). (5.6)<br />

For a one-dim<strong>en</strong>sional tail integral U, the inverse tail integral U (−1) was <strong>de</strong>fined in Equation<br />

(4.25). In the sequel we will need the following technical lemma.<br />

Lemma 5.4. L<strong>et</strong> ν be a Lévy measure on R d with marginal tail integrals U i , i = 1, . . . , d, and<br />

Lévy copula F on (−∞, ∞] d , satisfying (5.5), l<strong>et</strong> µ be <strong>de</strong>fined by (5.6) and l<strong>et</strong><br />

Th<strong>en</strong> ν is the image measure of µ by f.<br />

f : (u 1 , . . . , u d ) ↦→ (U (−1)<br />

1 (u 1 ), . . . , U (−1)<br />

d<br />

(u d )).


5.2. SIMULATION OF DEPENDENT LEVY PROCESSES 175<br />

Proof. We must prove that for each A ∈ B(R d ),<br />

ν(A) = µ({u ∈ R d : f(u) ∈ A}),<br />

but in view of Lemma 4.7, it is suffici<strong>en</strong>t to show that for each I ⊂ {1, . . . , d} nonempty and<br />

for all (x i ) i∈I ∈ (R \ {0}) |I| ,<br />

U I ((x i ) i∈I ) = µ({u ∈ R d : U (−1)<br />

i<br />

(u i ) ∈ I(x i ), i ∈ I}),<br />

where I(x) was <strong>de</strong>fined in (4.21). However, since U i is left-continuous, for every i, U (−1)<br />

i<br />

(u) ∈<br />

I(x) if and only if u ∈ (U i (x) ∧ 0, U i (x) ∨ 0]. Therefore,<br />

µ({u ∈ R d : U (−1)<br />

i<br />

(u i ) ∈ I(x i ), i ∈ I})<br />

= µ({u ∈ R d : u i ∈ (U i (x i ) ∧ 0, U i (x i ) ∨ 0], i ∈ I}) = F I ((U i (x i )) i∈I ),<br />

and an application of Theorem 4.8 compl<strong>et</strong>es the proof.<br />

By Theorem 2.28 in [1], there exists a family, in<strong>de</strong>xed by ξ ∈ R, of positive Radon measures<br />

K(ξ, dx 2 · · · dx d ) on R d−1 , such that<br />

ξ ↦→ K(ξ, dx 2 · · · dx d )<br />

is Borel measurable and<br />

µ(dx 1 . . . dx d ) = λ(dx 1 ) ⊗ K(x 1 , dx 2 · · · dx d ). (5.7)<br />

In addition, K(ξ, R d−1 ) = 1 λ-almost everywhere, that is, K(ξ, ∗) is, almost everywhere, a probability<br />

distribution. In the sequel we will call {K(ξ, ∗)} ξ∈R the family of conditional probability<br />

distributions associated to the Lévy copula F .<br />

L<strong>et</strong> F ξ be the distribution function of the measure K(ξ, ∗):<br />

F ξ (x 2 , . . . , x d ) := K(ξ, (−∞, x 2 ] × · · · × (−∞, x d ]), (x 2 , . . . , x d ) ∈ R d−1 . (5.8)<br />

The following lemma shows that it can be computed in a simple manner from the Lévy copula<br />

F .


176 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

Lemma 5.5. L<strong>et</strong> F be a Lévy copula on (−∞, ∞] d , satisfying (5.5), and F ξ be the corresponding<br />

conditional distribution function, <strong>de</strong>fined by (5.8). Th<strong>en</strong>, there exists N ⊂ R with λ(N) = 0<br />

such that for every fixed ξ ∈ R \ N, F ξ (∗) is a probability distribution function, satisfying<br />

F ξ (x 2 , . . . , x d ) = sgn(ξ) ∂ ∂ξ V F ((ξ ∧ 0, ξ ∨ 0] × (−∞, x 2 ] × · · · × (−∞, x d ]) (5.9)<br />

in every point (x 2 , . . . , x d ), where F ξ is continuous.<br />

Remark 5.2. Since the law of a random variable is compl<strong>et</strong>ely <strong>de</strong>termined by the values of its<br />

distribution function at the continuity points of the latter, being able to compute F ξ at all<br />

points where it is continuous is suffici<strong>en</strong>t for all practical purposes.<br />

Proof. Since it has already be<strong>en</strong> observed that K(ξ, R d−1 ) = 1 λ-almost everywhere, we only<br />

need to prove the second part of the lemma. L<strong>et</strong><br />

G(x 1 , . . . , x d ) := sgn x 1 V F ((x 1 ∧ 0, x 1 ∨ 0] × (−∞, x 2 ] × · · · × (−∞, x d ])<br />

By Theorem 2.28 in [1], for each f ∈ L 1 (R d , µ),<br />

∫<br />

∫ ∞<br />

f(x 1 , . . . , x d )µ(dx 1 · · · dx d ) = dx 1 f(x 1 , . . . , x d )K(x 1 , dx 2 · · · dx d ), (5.10)<br />

R d −∞<br />

∫R d<br />

which implies that<br />

∫<br />

G(x 1 , . . . , x d ) = sgn x 1<br />

(x 1 ∧0,x 1 ∨0]<br />

dξF ξ (x 2 , . . . , x d ),<br />

Therefore, for fixed (x 2 , . . . , x d ), (5.9) holds ξ-almost everywhere. Since a union of countably<br />

many s<strong>et</strong>s of zero measure is again a s<strong>et</strong> of zero measure, there exists a s<strong>et</strong> N ⊂ R with λ(N) = 0<br />

such that for every ξ ∈ R \ N, (5.9) holds for all (x 2 , . . . , x d ) ∈ Q d , where Q <strong>de</strong>notes the s<strong>et</strong> of<br />

rational numbers.<br />

Fix ξ ∈ R \ N and l<strong>et</strong> x ∈ R d−1 and {x + n } and {x − n } be two sequ<strong>en</strong>ces of d − 1-dim<strong>en</strong>sional<br />

vectors with coordinates in Q, converging to x from above and from below (compon<strong>en</strong>twise).<br />

Since F ξ is increasing in each coordinate (as a probability distribution function), the limits<br />

lim n F ξ (x + n ) and lim n F ξ (x − n ) exist. Suppose that<br />

lim F ξ (x +<br />

n<br />

n ) = lim F ξ (x −<br />

n<br />

n ) = F ∗ (5.11)<br />

and observe that for every δ ≠ 0,<br />

G(ξ + δ, x − n ) − G(ξ, x − n )<br />

δ<br />

≤<br />

G(ξ + δ, x) − G(ξ, x)<br />

δ<br />

≤ G(ξ + δ, x+ n ) − G(ξ, x + n )<br />

.<br />

δ


5.2. SIMULATION OF DEPENDENT LEVY PROCESSES 177<br />

For every ε > 0, in view of (5.11), there exists N 0 such that for every n ≥ N 0 , F ξ (x + n )−F ∗ ≤ ε/2<br />

and F ∗ − F ξ (x − n ) ≤ ε/2. Since G is differ<strong>en</strong>tiable with respect to the first variable at points<br />

(ξ, x + n ) and (ξ, x − n ), we can choose δ small <strong>en</strong>ough so that<br />

and<br />

This proves that<br />

G(ξ + δ, x − n ) − G(ξ, x − n )<br />

∣<br />

δ<br />

G(ξ + δ, x + n ) − G(ξ, x + n )<br />

∣<br />

δ<br />

− F ξ (x − n )<br />

∣ ≤ ε/2<br />

− F ξ (x + n )<br />

∣ ≤ ε/2<br />

G(ξ + δ, x) − G(ξ, x)<br />

lim<br />

= F ∗ .<br />

δ→0 δ<br />

We have thus shown that F ξ satisfies Equation (5.9) in all points where (5.11) holds, that is,<br />

where F ξ is continuous.<br />

In the following two theorems we show how Lévy copulas may be used to simulate multidim<strong>en</strong>sional<br />

Lévy processes with specified <strong>de</strong>p<strong>en</strong><strong>de</strong>nce. Our results can be se<strong>en</strong> as an ext<strong>en</strong>sion<br />

to Lévy processes, repres<strong>en</strong>ted by Lévy copulas, of the series repres<strong>en</strong>tation results, <strong>de</strong>veloped<br />

by Rosinski and others (see [84] and refer<strong>en</strong>ces therein). The first result concerns the simpler<br />

case wh<strong>en</strong> the Lévy process has finite variation on compacts.<br />

Theorem 5.6. (Simulation of multidim<strong>en</strong>sional Lévy processes, finite variation case)<br />

L<strong>et</strong> ν be a Lévy measure on R d , satisfying ∫ (|x| ∧ 1)ν(dx) < ∞, with marginal tail integrals U i ,<br />

i = 1, . . . , d and Lévy copula F (x 1 , . . . , x d ), such that the condition (5.5) is satisfied, and l<strong>et</strong><br />

K(x 1 , dx 2 · · · dx d ) be the corresponding conditional probability distributions, <strong>de</strong>fined by (5.8). L<strong>et</strong><br />

{V i } be a sequ<strong>en</strong>ce of in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt random variables, uniformly distributed on [0, 1]. Introduce<br />

d random sequ<strong>en</strong>ces {Γ 1 i }, . . . , {Γd i }, in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt from {V i} such that<br />

• N = ∑ ∞<br />

i=1 δ {Γ 1 i } is a Poisson random measure on R with int<strong>en</strong>sity measure λ.<br />

• Conditionally on Γ 1 i , the random vector (Γ2 i , . . . , Γd i ) is in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt from Γk j<br />

and all k and is distributed on R d−1 with law K(Γ 1 i , dx 2 · · · dx d ).<br />

with j ≠ i<br />

Th<strong>en</strong><br />

{Z t } 0≤t≤1 where Z k t =<br />

∞∑<br />

i=1<br />

U (−1)<br />

i<br />

(Γ k i )1 [0,t] (V i ), k = 1, . . . , d, (5.12)


178 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

is a Lévy process on the time interval [0, 1] with characteristic function<br />

( ∫<br />

)<br />

e i〈u,Zt〉 = exp t (e i〈u,z〉 − 1)ν(dz) . (5.13)<br />

R d<br />

Remark 5.3. The probability distribution function of (Γ 2 i , . . . , Γd i ) conditionally on Γ1 i<br />

from Lemma 5.9.<br />

is known<br />

Remark 5.4. The sequ<strong>en</strong>ce {Γ 1 i } i≥1 can be constructed, for example, as follows. L<strong>et</strong> {X i } i≥1 be<br />

a sequ<strong>en</strong>ce of jump times of a Poisson process with jump int<strong>en</strong>sity equal to 2. Th<strong>en</strong> it is easy<br />

to check that one can <strong>de</strong>fine Γ 1 i by Γ1 i = X i(−1) i .<br />

Proof. First note that {Γ k i } are well <strong>de</strong>fined since by Lemma 5.5, K(x 1, ∗) is a probability<br />

distribution for almost all x 1 . L<strong>et</strong><br />

Z k τ,t =<br />

∑<br />

−τ≤Γ 1 i ≤τ U (−1)<br />

k<br />

(Γ k i )1 Vi ≤t, k = 1, . . . , d.<br />

By Proposition 3.8 in [82],<br />

∫<br />

Zτ,t k =<br />

U (−1)<br />

k<br />

(x k )M(ds × dx 1 · · · dx d ),<br />

[0,t]×[−τ,τ]×R d−1<br />

where M is a Poisson random measure on [0, 1] × R d<br />

µ(dx 1 · · · dx d ), and the measure µ was <strong>de</strong>fined in Equation (5.6).<br />

with int<strong>en</strong>sity measure λ [0,1] (dt) ⊗<br />

By Lemma 5.4 and Proposition 3.7 in [82],<br />

∫<br />

Zτ,t k = x k N τ (ds × dx 1 · · · dx d ), (5.14)<br />

[0,t]×R d<br />

for some Poisson random measure N τ on [0, 1]×R d with int<strong>en</strong>sity measure λ [0,1] (ds)⊗ν τ (dx 1 · · · dx d ),<br />

where<br />

ν τ := 1 (−∞,U<br />

(−1)<br />

1 (−τ)]∪[U (−1)<br />

1 (τ),∞) (x 1)ν(dx 1 · · · dx d ) (5.15)<br />

The Lévy-Itô <strong>de</strong>composition [87, Theorem 19.2] implies that Z τ,t is a Lévy process on the time<br />

interval [0, 1] with characteristic function<br />

( ∫<br />

)<br />

e i〈u,Zτ,t〉 = exp t (e i〈u,z〉 − 1)ν τ (dz)<br />

R d .<br />

L<strong>et</strong> h be a boun<strong>de</strong>d continuous truncation function such that h(x) ≡ x on a neighborhood of<br />

0. Since lim τ→∞ U (−1)<br />

1 (τ) = 0 and lim τ→∞ U (−1)<br />

1 (−τ) = 0, by dominated converg<strong>en</strong>ce,<br />

∫<br />

∫<br />

∫<br />

∫<br />

h 2 (x)ν τ (dx) −−−→ h 2 (x)ν(dx) and h(x)ν τ (dx) −−−→ h(x)ν(dx).<br />

R d τ→∞<br />

R d R d τ→∞<br />

R d


5.2. SIMULATION OF DEPENDENT LEVY PROCESSES 179<br />

Moreover, for every f ∈ C b (R d ) such that f(x) ≡ 0 on a neighborhood of 0,<br />

∫<br />

∫<br />

f(x)ν τ (dx) = f(x)ν(dx)<br />

R d R d<br />

starting from a suffici<strong>en</strong>tly large τ. Therefore, Proposition 1.7 allows to conclu<strong>de</strong> that {Z τ,t } 0≤t≤1<br />

converges in law to a Lévy process with characteristic function giv<strong>en</strong> by (5.13).<br />

If the Lévy process has paths of infinite variation on compacts, it can no longer be repres<strong>en</strong>ted<br />

as the sum of its jumps and we have to introduce a c<strong>en</strong>tering term into the series<br />

(5.12).<br />

Theorem 5.7. (Simulation of multidim<strong>en</strong>sional Lévy processes, infinite variation<br />

case)<br />

L<strong>et</strong> ν be a Lévy measure on R d with marginal tail integrals U i , i = 1, . . . , d and Lévy copula<br />

F (x 1 , . . . , x d ), such that the condition (5.5) is satisfied. L<strong>et</strong> {V i } and {Γ 1 i }, . . . , {Γd i } be as in<br />

Theorem 5.6. L<strong>et</strong><br />

∫<br />

A k (τ) =<br />

|x|≤1<br />

where ν τ is giv<strong>en</strong> by (5.15). Th<strong>en</strong> the process<br />

{Z τ,t } 0≤t≤1 , where Z k τ,t = ∑<br />

x k ν τ (dx 1 · · · dx d ), k = 1 . . . d,<br />

U (−1)<br />

k<br />

−τ≤Γ 1 i ≤τ<br />

(Γ k i )1 Vi ≤t − tA k (τ),<br />

converges in law as τ → ∞ to a Lévy process {Z t } 0≤t≤1 on the time interval [0, 1] with characteristic<br />

function<br />

( ∫<br />

)<br />

e i〈u,Zt〉 = exp t (e i〈u,z〉 − 1 − i〈u, z〉)1 |z|≤1 ν(dz) . (5.16)<br />

R d<br />

Proof. The proof is ess<strong>en</strong>tially the same as in Theorem 5.6. Similarly to Equation (5.14), Z k τ,s<br />

can now be repres<strong>en</strong>ted as<br />

Z k τ,s =<br />

∫<br />

[0,s]×{x∈R d :|x|≤1}<br />

x k {N τ (ds × dx 1 · · · dx d ) − dsν τ (dx 1 · · · dx d )}<br />

∫<br />

+<br />

x k N τ (ds × dx 1 · · · dx d ),<br />

[0,s]×{x∈R d :|x|>1}<br />

where N τ is a Poisson random measure on [0, 1] × R d with int<strong>en</strong>sity measure λ [0,1] (ds) ⊗ ν τ ,<br />

and ν τ is <strong>de</strong>fined by (5.15). This <strong>en</strong>tails that Z τ,s is a Lévy process (compound Poisson) with


180 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

−10 −8 −6 −4 −2 0 2 4 6 8 10<br />

Figure 5.3: Conditional distribution function F ξ , corresponding to the Lévy copula of Example<br />

5.1 with ξ = 1, θ = 1 and η = 0.25.<br />

characteristic function<br />

e i〈u,Zτ,t〉 = exp<br />

( ∫<br />

)<br />

t (e i〈u,z〉 − 1 − i〈u, z〉1 |z|≤1 )ν τ (dz) .<br />

R d<br />

Proposition 1.7 once again allows to conclu<strong>de</strong> that (Z τ,s ) 0≤s≤1 converges in law to a Lévy process<br />

with characteristic function (5.16).<br />

Example 5.3. L<strong>et</strong> d = 2 and F be the Lévy copula of Example 5.1. A straightforward computation<br />

yields:<br />

F ξ =<br />

⎧ (<br />

⎨<br />

⎩ (1 − η) + 1 +<br />

∣ ) ⎫<br />

ξ ∣∣∣<br />

θ<br />

−1−1/θ<br />

⎬<br />

∣<br />

(η − 1 x2


5.2. SIMULATION OF DEPENDENT LEVY PROCESSES 181<br />

8<br />

1<br />

7<br />

0.5<br />

6<br />

0<br />

5<br />

−0.5<br />

4<br />

−1<br />

3<br />

−1.5<br />

2<br />

−2<br />

1<br />

−2.5<br />

0<br />

−3<br />

−1<br />

−3.5<br />

−2<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

−4<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Figure 5.4: Trajectories of two variance gamma processes with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure giv<strong>en</strong> by<br />

the Lévy copula of Example 5.1. In both graphs both variance gamma processes are driftless<br />

and have param<strong>et</strong>ers c = 10, λ − = 1 and λ + = 1 (cf. Equation (1.19)). In the left graph, the<br />

<strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> the two compon<strong>en</strong>ts is strong both in terms of sign and absolute value<br />

(η = 0.9 and θ = 3): the processes jump mostly in the same direction and the sizes of jumps<br />

are similar. In the right graph the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of absolute values is weak (θ = 0.5) and the<br />

<strong>de</strong>p<strong>en</strong><strong>de</strong>nce of jump sign is negative (η = 0.25).<br />

If ν is a Lévy measure on R 2 , satisfying ∫ (|x|∧1)ν(dx) < ∞ with marginal tail integrals U 1 , U 2<br />

and Lévy copula F of Example 5.1, the Lévy process with characteristic function (5.13) can be<br />

simulated as follows. L<strong>et</strong> {V i } and {Γ 1 i } be as in Theorem 5.6 and l<strong>et</strong> {W i} be an in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt<br />

sequ<strong>en</strong>ce of in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt random variables, uniformly distributed on [0, 1].<br />

For each i, l<strong>et</strong><br />

Γ 2 i = F −1 (W<br />

Γ 1 i ). Th<strong>en</strong> the Lévy process that we want to simulate is giv<strong>en</strong> by Equation (5.12).<br />

i<br />

Figure 5.4 shows the simulated trajectories of variance gamma processes with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

structure giv<strong>en</strong> by the Lévy copula of Example 5.1 with differ<strong>en</strong>t values of param<strong>et</strong>ers. The<br />

number of jumps for each trajectory was limited to 2000 and the inverse tail integral of the<br />

variance gamma Lévy measure was computed by inverting numerically the expon<strong>en</strong>tial integral<br />

function (function expint available in MATLAB). Simulating two trajectories with 2000 jumps<br />

each takes about 1 second on a P<strong>en</strong>tium III computer running MATLAB, but this time could be<br />

reduced by several or<strong>de</strong>rs of magnitu<strong>de</strong> if the inverse expon<strong>en</strong>tial integral function is tabulated<br />

and a lower-level programming language (e.g. C++) is used.


182 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

5.3 A two-dim<strong>en</strong>sional variance gamma mo<strong>de</strong>l for option pricing<br />

In this section we pres<strong>en</strong>t a case study showing how one particular mo<strong>de</strong>l, constructed using<br />

Lévy copulas, can be used to price multi-ass<strong>et</strong> options.<br />

The mo<strong>de</strong>l We suppose that un<strong>de</strong>r the risk-neutral probability, the prices {S 1 t } t≥0 and<br />

{S 2 t } t≥0 of two risky ass<strong>et</strong>s satisfy<br />

S 1 t = e rt+X1 t , S 2 t = e rt+X2 t , (5.18)<br />

where (X 1 , X 2 ) is a Lévy process on R d with characteristic tripl<strong>et</strong> (0, ν, b) with respect to<br />

zero truncation function. X 1 and X 2 are supposed to be variance gamma processes, that is,<br />

the margins ν 1 and ν 2 of ν are of the form (1.19) with param<strong>et</strong>ers c 1 , λ 1 +, λ 1 − and c 2 , λ 2 +, λ 2 −.<br />

The Lévy copula F of ν is supposed to be of the form (5.4) with param<strong>et</strong>ers θ and η. The<br />

no-arbitrage condition imposes that for i = 1, 2, λ i + > 1 and the drift coeffici<strong>en</strong>ts satisfy<br />

(<br />

b i = c i log 1 − 1<br />

λ i + 1<br />

+ λ i − 1 )<br />

− λ i .<br />

+ λi −<br />

The problem In the rest of this section, mo<strong>de</strong>l (5.18) will be used to price two differ<strong>en</strong>t kinds<br />

of multi-ass<strong>et</strong> options: the option on weighted average, whose payoff at expiration date T is<br />

giv<strong>en</strong> by<br />

( 2∑ +<br />

H T = w i ST i − K)<br />

with w 1,2 ≥ 0 and w 1 + w 2 = 1,<br />

i=1<br />

and the best-of or alternative option with payoff structure<br />

H T =<br />

( ( S<br />

1<br />

) ) +<br />

N max T<br />

S0<br />

1 , S2 T<br />

S0<br />

2 − K<br />

Option pricing by Monte Carlo Bask<strong>et</strong> options, <strong>de</strong>scribed above can be priced by Monte<br />

Carlo m<strong>et</strong>hod using European options on individual stocks as control variates. D<strong>en</strong>ote the<br />

discounted payoffs of European options by<br />

V i T = e −rT (S i T − K) + for i = 1, 2.


5.3. TWO-DIMENSIONAL VARIANCE GAMMA MODEL 183<br />

0.2<br />

0.2<br />

0.1<br />

0.1<br />

0<br />

0<br />

−0.1<br />

−0.1<br />

−0.2<br />

−0.2<br />

−0.2 −0.1 0 0.1 0.2<br />

−0.2 −0.1 0 0.1 0.2<br />

Figure 5.5: Scatter plots of r<strong>et</strong>urns in a 2-dim<strong>en</strong>sional variance gamma mo<strong>de</strong>l with correlation<br />

ρ = 50% and differ<strong>en</strong>t tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce. Left: strong tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce (η = 0.75 and θ = 10).<br />

Right: weak tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce (η = 0.99 and θ = 0.61).<br />

and the discounted payoff of the bask<strong>et</strong> option by V T = e −rT H T . Th<strong>en</strong> the Monte Carlo<br />

estimate of bask<strong>et</strong> option price is giv<strong>en</strong> by<br />

Ê[V T ] = ¯V T + a 1 (E[V 1 T ] − ¯V 1 T ) + a 2 (E[V 2 T ] − ¯V 2 T ),<br />

where a bar over a random variable <strong>de</strong>notes the sample mean over N i.i.d. realizations of this<br />

variable, that is, ¯V T = 1 ∑ N<br />

N i=1 V (i)<br />

(i)<br />

T<br />

, where V<br />

T<br />

are in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt and have the same law as V T .<br />

The coeffici<strong>en</strong>ts a 1 and a 2 should be chos<strong>en</strong> in or<strong>de</strong>r to minimize the variance of Ê[V T ]. It is easy<br />

to see that this variance is minimal if a = Σa 0 , where Σ ij = Cov(V i T , V j T ) and a0 i = Cov(V T , V i T ).<br />

In practice these covariances are replaced by their in-sample estimates; this may introduce a<br />

bias into the estimator Ê[V T ], but for suffici<strong>en</strong>tly large samples this bias is small compared to<br />

the Monte Carlo error [45].<br />

To illustrate the option pricing procedure, we fixed the following param<strong>et</strong>ers of the marginal<br />

distributions of the two ass<strong>et</strong>s: c 1 = c 2 = 25, λ 1 + = 28.9, λ 1 − = 21.45, λ 2 + = 31.66 and<br />

λ 2 − = 25.26. In the param<strong>et</strong>rization (1.18) this corresponds to θ 1 = θ 2 = −0.2, κ 1 = κ 2 = 0.04,<br />

σ 1 = 0.3 and σ 2 = 0.25. To emphasize the importance of tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce for pricing multi-ass<strong>et</strong><br />

options, we used two s<strong>et</strong>s of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce param<strong>et</strong>ers, which correspond both to a correlation<br />

of 50% (the correlation is computed numerically) but lead to r<strong>et</strong>urns with very differ<strong>en</strong>t tail<br />

<strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures:


184 CHAPTER 5. APPLICATIONS OF LEVY COPULAS<br />

0.04<br />

0.035<br />

Tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

No tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

0.04<br />

0.035<br />

Tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

No tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

0.03<br />

0.03<br />

0.025<br />

0.025<br />

0.02<br />

0.02<br />

0.015<br />

0.015<br />

0.01<br />

0.01<br />

0.005<br />

0.005<br />

0<br />

0.95 1 1.05<br />

0<br />

0.95 1 1.05<br />

Figure 5.6: Prices of options on weighted average (left) and of best-of options (right) for two<br />

differ<strong>en</strong>t <strong>de</strong>p<strong>en</strong><strong>de</strong>nce patterns.<br />

Pattern 1 Strong tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce: θ = 10 and η = 0.75. The scatter plot of r<strong>et</strong>urns is shown<br />

in Figure 5.5, left graph. Although the signs of r<strong>et</strong>urns may be differ<strong>en</strong>t, the probability<br />

that the r<strong>et</strong>urns will be large in absolute value simultaneously in both compon<strong>en</strong>ts is very<br />

high.<br />

Pattern 2 Weak tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce: θ = 0.61 and η = 0.99. The scatter plot of r<strong>et</strong>urns in shown<br />

in Figure 5.5, right graph. With this <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure the r<strong>et</strong>urns typically have the<br />

same sign but their absolute values are not correlated.<br />

In each of the two cases, a sample of 1000 realizations of the couple (XT 1 , X2 T<br />

) with T = 0.02<br />

(one-week options) was simulated using the procedure <strong>de</strong>scribed in Example 5.3. The cutoff<br />

param<strong>et</strong>er τ (see Equation (5.14)) was tak<strong>en</strong> equal to 1000, which lead to limiting the average<br />

number of jumps for each trajectory to about 40. For this value of τ, Ui<br />

−1 (τ) is of or<strong>de</strong>r of<br />

10 −19 for both ass<strong>et</strong>s. Since for the variance gamma mo<strong>de</strong>l the converg<strong>en</strong>ce of U −1 to zero as<br />

τ → ∞ is expon<strong>en</strong>tial, the error resulting from the truncation of small jumps is of the same<br />

or<strong>de</strong>r, h<strong>en</strong>ce, negligible.<br />

Figure 5.6 shows the prices of bask<strong>et</strong> options, computed for differ<strong>en</strong>t strikes with <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

patterns giv<strong>en</strong> above. The initial ass<strong>et</strong> prices were S0 1 = S2 0 = 1, and the interest rate was tak<strong>en</strong><br />

to be r = 0.03. For the option on weighted average, the weights w i were both equal to 0.5<br />

and for the best-of option the coeffici<strong>en</strong>t was N = 1. The prices of European options, used for


5.3. TWO-DIMENSIONAL VARIANCE GAMMA MODEL 185<br />

variance reduction, were computed using the Fourier transform algorithm <strong>de</strong>scribed in Chapter<br />

1. The standard <strong>de</strong>viation of Monte Carlo estimates of option prices was below 2 · 10 −4 at the<br />

money in all cases.<br />

The differ<strong>en</strong>ce b<strong>et</strong>we<strong>en</strong> option prices computed with and without tail <strong>de</strong>p<strong>en</strong><strong>de</strong>ce is clearly<br />

important for both types of options: as se<strong>en</strong> from Figure 5.6, neglecting tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce may<br />

easily lead to a 10% error on the option price at the money. On the other hand, this example<br />

shows that using Lévy copulas allows to take into account the tail <strong>de</strong>p<strong>en</strong><strong>de</strong>nce and discriminate<br />

b<strong>et</strong>we<strong>en</strong> two situations that would be undistinguishable in a log-normal framework.


186 CHAPTER 5. APPLICATIONS OF LEVY COPULAS


Conclusions and perspectives<br />

In the first part of this thesis we have solved, using <strong>en</strong>tropic regularization, the ill-posed problem<br />

of calibrating an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l to options data and proposed a stable numerical<br />

m<strong>et</strong>hod for computing this solution. Applying our m<strong>et</strong>hod to prices of in<strong>de</strong>x options allowed us<br />

to estimate the risk-neutral Lévy measures, implied by mark<strong>et</strong> prices. This object is the analog,<br />

for expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls, of implied volatility, used in the Black-Scholes framework. Our<br />

empirical results allow to make a number of important conclusions. First, using an expon<strong>en</strong>tial<br />

Lévy mo<strong>de</strong>l one can calibrate with high precision the prices of a s<strong>et</strong> of options with common<br />

maturity. Moreover, high quality of calibration is achieved already by using finite-int<strong>en</strong>sity Lévy<br />

processes. Therefore, from the point of view of option pricing the imperative for using more<br />

complex infinite-int<strong>en</strong>sity mo<strong>de</strong>ls is not clear. The third conclusion is that ev<strong>en</strong> in the nonparam<strong>et</strong>ric<br />

s<strong>et</strong>ting it is impossible, using an expon<strong>en</strong>tial Lévy mo<strong>de</strong>l, to calibrate accurately<br />

the prices of stock in<strong>de</strong>x options of several maturities at the same time: options of differ<strong>en</strong>t<br />

maturities produce differ<strong>en</strong>t implied Lévy measures. This confirms the observation already ma<strong>de</strong><br />

by several authors [13, 68] that the framework of expon<strong>en</strong>tial Lévy mo<strong>de</strong>ls is not suffici<strong>en</strong>tly<br />

flexible to reproduce the term structure of implied volatilities correctly.<br />

In view of the above conclusions, we plan to continue the line of research initiated by this<br />

thesis, by ext<strong>en</strong>ding its results to mo<strong>de</strong>ls of stock price behavior that do allow to <strong>de</strong>scribe the<br />

<strong>en</strong>tire term structure of implied volatilities, e.g. mo<strong>de</strong>ls based on additive processes (processes<br />

with in<strong>de</strong>p<strong>en</strong><strong>de</strong>nt but not stationary increm<strong>en</strong>ts) and hybrid mo<strong>de</strong>ls including both jumps and<br />

stochastic volatility. The second important direction of future research is to investigate the<br />

impact of our calibration m<strong>et</strong>hodology on the m<strong>et</strong>hods of hedging in pres<strong>en</strong>ce of jumps in stock<br />

prices.<br />

In the second part of this thesis we introduced the notion of Lévy copula, providing a<br />

187


188 CONCLUSIONS AND PERSPECTIVES<br />

g<strong>en</strong>eral framework in which the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce structures of multidim<strong>en</strong>sional Lévy processes can<br />

be <strong>de</strong>scribed. Lévy copulas compl<strong>et</strong>ely characterize the possible <strong>de</strong>p<strong>en</strong><strong>de</strong>nce patterns of Lévy<br />

processes in the s<strong>en</strong>se that for every Lévy process, there exists a Lévy copula that <strong>de</strong>scribes its<br />

<strong>de</strong>p<strong>en</strong><strong>de</strong>nce structure and for every Lévy copula and every n one-dim<strong>en</strong>sional Lévy processes,<br />

there exists an n-dim<strong>en</strong>sional Lévy process with this Lévy copula and with margins giv<strong>en</strong> by<br />

these one-dim<strong>en</strong>sional processes. Multidim<strong>en</strong>sional Lévy process mo<strong>de</strong>ls for applications can<br />

thus be constructed by taking any n one-dim<strong>en</strong>sional processes and a Lévy copula from a<br />

(possibly param<strong>et</strong>ric) family. The simulation m<strong>et</strong>hods, <strong>de</strong>veloped in the last chapter of this<br />

thesis, allow to compute various quantities of interest in a Lévy copula mo<strong>de</strong>l using the Monte<br />

Carlo m<strong>et</strong>hod.<br />

The scope of pot<strong>en</strong>tial applications of Lévy copula mo<strong>de</strong>ls in finance and other domains is<br />

large. Financial applications inclu<strong>de</strong> bask<strong>et</strong> option pricing, and portfolio managem<strong>en</strong>t. Lévy<br />

copula mo<strong>de</strong>ls are also useful in insurance and in risk managem<strong>en</strong>t, to mo<strong>de</strong>l the <strong>de</strong>p<strong>en</strong><strong>de</strong>nce<br />

b<strong>et</strong>we<strong>en</strong> loss processes of differ<strong>en</strong>t business lines, and more g<strong>en</strong>erally, in all multivariate problems<br />

where <strong>de</strong>p<strong>en</strong><strong>de</strong>nce b<strong>et</strong>we<strong>en</strong> jumps needs to be tak<strong>en</strong> into account.<br />

From the point of view of applications, the next step is to <strong>de</strong>velop the m<strong>et</strong>hods of estimating<br />

Lévy copula mo<strong>de</strong>ls from the data, using, for example, simulation-based techniques of statistical<br />

infer<strong>en</strong>ce [51]. A more theor<strong>et</strong>ical research direction that we curr<strong>en</strong>tly pursue in collaboration<br />

with Jan Kalls<strong>en</strong> is to investigate the relation b<strong>et</strong>we<strong>en</strong> the Lévy copula of a Lévy process and<br />

its probabilistic copula at a giv<strong>en</strong> time.


Bibliography<br />

[1] L. Ambrosio, N. Fusco, and D. Pallara, Functions of Boun<strong>de</strong>d Variation and Free<br />

Discontinuity Problems, Oxford University Press, 2000.<br />

[2] L. An<strong>de</strong>rs<strong>en</strong> and J. Andreas<strong>en</strong>, Jump-diffusion mo<strong>de</strong>ls: Volatility smile fitting and<br />

numerical m<strong>et</strong>hods for pricing, Rev. Derivatives Research, 4 (2000), pp. 231–262.<br />

[3] M. Avellaneda, R. Buff, C. Friedman, N. Grandchamp, L. Kruk, and J. Newman,<br />

Weighted Monte Carlo: a new technique for calibrating ass<strong>et</strong>-pricing mo<strong>de</strong>ls, Int. J.<br />

Theor. Appl. <strong>Finance</strong>, 4 (2001), pp. 91–119.<br />

[4] M. Avellaneda, C. Friedman, R. Holmes, and D. Samperi, Calibrating volatility<br />

surfaces via relative <strong>en</strong>tropy minimization, Applied Mathematical <strong>Finance</strong>, 4 (1997),<br />

pp. 37–64.<br />

[5] O. Barndorff-Niels<strong>en</strong>, Processes of normal inverse Gaussian type, <strong>Finance</strong> Stoch., 2<br />

(1998), pp. 41–68.<br />

[6] O. Barndorff-Niels<strong>en</strong>, T. Mikosch, and S. Resnick, eds., Lévy Processes — Theory<br />

and Applications, Birkhäuser, Boston, 2001.<br />

[7] O. E. Barndorff-Niels<strong>en</strong>, Normal inverse Gaussian distributions and stochastic volatility<br />

mo<strong>de</strong>lling, Scand. J. Statist., 24 (1997), pp. 1–13.<br />

[8] O. E. Barndorff-Niels<strong>en</strong> and A. Lindner, Some aspects of Lévy copulas. Download<br />

from www-m4.ma.tum.<strong>de</strong>/m4/pers/lindner, 2004.<br />

189


190 BIBLIOGRAPHY<br />

[9] O. E. Barndorff-Niels<strong>en</strong> and N. Shephard, Non-Gaussian Ornstein-Uhl<strong>en</strong>beck based<br />

mo<strong>de</strong>ls and some of their uses in financial econom<strong>et</strong>rics, J. R. Statistic. Soc. B, 63 (2001),<br />

pp. 167–241.<br />

[10] D. Bates, Jumps and stochastic volatility: the exchange rate processes implicit in<br />

Deutschemark options, Rev. Fin. Studies, 9 (1996), pp. 69–107.<br />

[11] D. Bates, Post-87 crash fears in the S&P 500 futures option mark<strong>et</strong>, J. Econom<strong>et</strong>rics, 94<br />

(2000), pp. 181–238.<br />

[12] D. Bates, Maximum likelihood estimation of lat<strong>en</strong>t affine processes. Working paper, available<br />

from http://www.biz.uiowa.edu/faculty/dbates/, 2004.<br />

[13] D. S. Bates, Testing option pricing mo<strong>de</strong>ls, in Statistical M<strong>et</strong>hods in <strong>Finance</strong>, vol. 14 of<br />

Handbook of Statistics, North-Holland, Amsterdam, 1996, pp. 567–611.<br />

[14] N. Baud, A. Frachot, and T. Roncalli, How to avoid over-estimating capital charge<br />

in operational risk. Download from gro.creditlyonnais.fr/cont<strong>en</strong>t/fr/home_mc.htm,<br />

2002.<br />

[15] S. Beckers, A note on estimating param<strong>et</strong>ers of a jump-diffusion process of stock r<strong>et</strong>urns,<br />

J. Fin. and Quant. Anal., 16 (1981), pp. 127–140.<br />

[16] F. Bellini and M. Frittelli, On the exist<strong>en</strong>ce of minimax martingale measures, Math.<br />

<strong>Finance</strong>, 12 (2002), pp. 1–21.<br />

[17] J. Bertoin, Lévy Processes, Cambridge University Press, Cambridge, 1996.<br />

[18] J. Bouchaud and M. Potters, Théorie <strong>de</strong>s Risques Financiers, Aléa, Saclay, 1997.<br />

[19] S. Boyarch<strong>en</strong>ko and S. Lev<strong>en</strong>dorskiĭ, Non-Gaussian Merton-Black-Scholes Theory,<br />

World Sci<strong>en</strong>tific, River Edge, NJ, 2002.<br />

[20] R. Byrd, P. Lu, and J. Nocedal, A limited memory algorithm for bound constrained<br />

optimization, SIAM Journal on Sci<strong>en</strong>tific and Statistical Computing, 16 (1995), pp. 1190–<br />

1208.


BIBLIOGRAPHY 191<br />

[21] P. Carr, H. Geman, D. Madan, and M. Yor, The fine structure of ass<strong>et</strong> r<strong>et</strong>urns: An<br />

empirical investigation, Journal of Business, 75 (2002), pp. 305–332.<br />

[22] P. Carr, H. Geman, D. Madan, and M. Yor, Stochastic volatility for Lévy processes,<br />

Math. <strong>Finance</strong>, 13 (2003), pp. 345–382.<br />

[23] P. Carr and D. Madan, Option valuation using the fast Fourier transform, J. Comput.<br />

<strong>Finance</strong>, 2 (1998), pp. 61–73.<br />

[24] T. Chan, Pricing conting<strong>en</strong>t claims on stocks driv<strong>en</strong> by Lévy processes, Ann. Appl.<br />

Probab., 9 (1999), pp. 504–528.<br />

[25] A. S. Cherny and A. N. Shiryaev, Change of time and measure for Lévy processes.<br />

MaPhySto lecture notes series 2002-13, 2002.<br />

[26] R. Cont, J.-P. Bouchaud, and M. Potters, Scaling in financial data: Stable laws<br />

and beyond, in Scale Invariance and Beyond, B. Dubrulle, F. Graner, and D. Sorn<strong>et</strong>te, eds.,<br />

Springer, Berlin, 1997.<br />

[27] R. Cont and P. Tankov, Financial Mo<strong>de</strong>lling with Jump Processes, Chapman & Hall /<br />

CRC Press, 2004.<br />

[28] R. Cont and P. Tankov, Non-param<strong>et</strong>ric calibration of jump-diffusion option pricing<br />

mo<strong>de</strong>ls, J. Comput. <strong>Finance</strong>, 7 (2004).<br />

[29] J. W. Cooley and J. W. Tukey, An algorithm for the machine calculation of complex<br />

Fourier series, Math. Comp., 19 (1965), pp. 297–301.<br />

[30] I. Csiszar, I-diverg<strong>en</strong>ce geom<strong>et</strong>ry of probability distributions and minimization problems,<br />

Annals of Probability, 3 (1975), pp. 146–158.<br />

[31] M. H. A. Davis, Option pricing in incompl<strong>et</strong>e mark<strong>et</strong>s, in Mathematics of Derivative<br />

Securities, M. H. A. Dempster and S. R. Pliska, eds., 1997.<br />

[32] P. J. Davis and P. Rabinowitz, M<strong>et</strong>hods of Numerical Integration, Aca<strong>de</strong>mic Press,<br />

1975.


192 BIBLIOGRAPHY<br />

[33] F. Delba<strong>en</strong>, P. Grandits, T. Rheinlän<strong>de</strong>r, D. Samperi, M. Schweizer, and<br />

C. Stricker, Expon<strong>en</strong>tial hedging and <strong>en</strong>tropic p<strong>en</strong>alties, Math. <strong>Finance</strong>, 12 (2002),<br />

pp. 99–123.<br />

[34] F. Delba<strong>en</strong> and W. Schachermayer, The fundam<strong>en</strong>tal theorem of ass<strong>et</strong> pricing for<br />

unboun<strong>de</strong>d stochastic processes, Math. Ann., 312 (1998), pp. 215–250.<br />

[35] B. Dupire, Pricing with a smile, RISK, 7 (1994), pp. 18–20.<br />

[36] E. Eberlein, Applications of g<strong>en</strong>eralized hyperbolic Lévy motion to <strong>Finance</strong>, in Lévy<br />

Processes — Theory and Applications, O. Barndorff-Niels<strong>en</strong>, T. Mikosch, and S. Resnick,<br />

eds., Birkhäuser, Boston, 2001, pp. 319–336.<br />

[37] E. Eberlein, U. Keller, and K. Prause, New insights into smile, mispricing and<br />

Value at Risk: The hyperbolic mo<strong>de</strong>l, Journal of Business, 71 (1998), pp. 371–405.<br />

[38] E. Eberlein and S. Raible, Some analytic facts on the g<strong>en</strong>eralized hyperbolic mo<strong>de</strong>l, in<br />

European Congress of Mathematics, Vol. II (Barcelona, 2000), vol. 202 of Progr. Math.,<br />

Birkhäuser, Basel, 2001, pp. 367–378.<br />

[39] N. El Karoui and R. Rouge, Pricing via utility maximization and <strong>en</strong>tropy, Math.<br />

<strong>Finance</strong>, 10 (2000), pp. 259–276.<br />

[40] H. W. Engl, M. Hanke, and A. Neubauer, Regularization of Inverse Problems,<br />

vol. 375, Kluwer Aca<strong>de</strong>mic Publishers Group, Dordrecht, 1996.<br />

[41] H. Föllmer and M. Schweizer, Hedging of conting<strong>en</strong>t claims un<strong>de</strong>r incompl<strong>et</strong>e information,<br />

in Applied Stochastic Analysis, M. Davis and R. Elliott, eds., vol. 5, Gordon and<br />

Breach, London, 1991, pp. 389–414.<br />

[42] M. Frittelli, The minimal <strong>en</strong>tropy martingale measure and the valuation problem in<br />

incompl<strong>et</strong>e mark<strong>et</strong>s, Math. <strong>Finance</strong>, 10 (2000), pp. 39–52.<br />

[43] H. Geman, Pure jump Lévy processes for ass<strong>et</strong> price mo<strong>de</strong>ling, Journal of Banking and<br />

<strong>Finance</strong>, 26 (2002), pp. 1297–1316.<br />

[44] H. Geman, D. Madan, and M. Yor, Time changes for Lévy processes, Math. <strong>Finance</strong>,<br />

11 (2001), pp. 79–96.


BIBLIOGRAPHY 193<br />

[45] P. Glasserman, Monte Carlo M<strong>et</strong>hods in Financial Engineering, Springer, New York,<br />

2003.<br />

[46] P. Glasserman and B. Yu, Large sample properties of weighted Monte Carlo estimators.<br />

Preprint, available from www.gsb.columbia.edu/faculty/pglasserman/Other/, 2002.<br />

[47] T. Goll and J. Kalls<strong>en</strong>, Optimal portfolios for logarithmic utility, Stochastic Process.<br />

Appl., 89 (2000), pp. 31–48.<br />

[48] T. Goll and L. Rüsch<strong>en</strong>dorf, Minimal distance martingale measures and optimal<br />

portfolios consist<strong>en</strong>t with observed mark<strong>et</strong> prices. Preprint, available from<br />

www.stochastik.uni-freiburg.<strong>de</strong>/~goll/start.html, 2000.<br />

[49] , Minimax and minimal distance martingale measures and their relationship to portfolio<br />

optimization, <strong>Finance</strong> Stoch., 5 (2001), pp. 557–581.<br />

[50] A. V. Goncharsky, A. S. Leonov, and A. G. Yagola, Applicability of the disparity<br />

principle in the case of non-linear incorrectly posed problems, and a new regularizing<br />

algorithm for solving them, USSR Comp. Math. Math. Phys, 15 (1975), pp. 8–16.<br />

[51] C. Gouriéroux and A. Monfort, Simulation-Based Econom<strong>et</strong>ric M<strong>et</strong>hods, Oxford University<br />

Press, 1996.<br />

[52] A. Hobson, Concepts in Statistical Mechanics, Gordon and Breach, Langhorne, P<strong>en</strong>nsylvania,<br />

1971.<br />

[53] N. Jackson, E. Süli, and S. Howison, Computation of <strong>de</strong>terministic volatility surfaces,<br />

Journal of Computational <strong>Finance</strong>, 2 (1999), pp. 5–32.<br />

[54] J. Jacod and A. N. Shiryaev, Limit Theorems for Stochastic Processes, Springer, Berlin,<br />

2nd ed., 2003.<br />

[55] P. Jakub<strong>en</strong>as, On option pricing in certain incompl<strong>et</strong>e mark<strong>et</strong>s, Proceedings of the<br />

Steklov Mathematical Institute, 237 (2002).<br />

[56] H. Joe, Multivariate Mo<strong>de</strong>ls and Dep<strong>en</strong><strong>de</strong>nce Concepts, Chapman & Hall, London, 1997.


194 BIBLIOGRAPHY<br />

[57] P. Jorion, On jump processes in the foreign exchange and stock mark<strong>et</strong>s, Rev. Fin. Studies,<br />

1 (1988), pp. 427–445.<br />

[58] J. Kalls<strong>en</strong>, Utility-based <strong>de</strong>rivative pricing, in Mathematical <strong>Finance</strong> — Bachelier<br />

Congress 2000, Springer, Berlin, 2001.<br />

[59] J. Kalls<strong>en</strong> and P. Tankov, Characterization of <strong>de</strong>p<strong>en</strong><strong>de</strong>nce of multidim<strong>en</strong>sional<br />

Lévy processes using Lévy copulas. Preprint, available from<br />

www.cmap.polytechnique.fr/~tankov, 2004.<br />

[60] J. F. C. Kingman and S. J. Taylor, Introduction to Measure and Probability, Cambridge<br />

University Press, Cambridge, 1966.<br />

[61] I. Kopon<strong>en</strong>, Analytic approach to the problem of converg<strong>en</strong>ce of truncated Lévy flights<br />

towards the Gaussian stochastic process., Physical Review E, 52 (1995), pp. 1197–1199.<br />

[62] S. Kou, A jump-diffusion mo<strong>de</strong>l for option pricing, Managem<strong>en</strong>t Sci<strong>en</strong>ce, 48 (2002),<br />

pp. 1086–1101.<br />

[63] R. W. Lee, Option pricing by transform m<strong>et</strong>hods: ext<strong>en</strong>sions, unification and error control,<br />

J. Comput. <strong>Finance</strong>, 7 (2004).<br />

[64] A. Lindner and A. Szimayer, A limit theorem for copulas. Download from<br />

www-m4.ma.tum.<strong>de</strong>/m4/pers/lindner, 2003.<br />

[65] F. Lindskog and A. McNeil, Common Poisson shock mo<strong>de</strong>ls: Applications to insurance<br />

and credit risk mo<strong>de</strong>lling. Available from www.risklab.ch, September 2001.<br />

[66] D. Madan, Financial mo<strong>de</strong>ling with discontinuous price processes, in Lévy Processes<br />

— Theory and Applications, O. Barndorff-Niels<strong>en</strong>, T. Mikosch, and S. Resnick, eds.,<br />

Birkhäuser, Boston, 2001.<br />

[67] D. Madan, P. Carr, and E. Chang, The variance gamma process and option pricing,<br />

European <strong>Finance</strong> Review, 2 (1998), pp. 79–105.<br />

[68] D. Madan and M. Konikov, Option pricing using variance gamma Markov chains, Rev.<br />

Derivatives Research, 5 (2002), pp. 81–115.


BIBLIOGRAPHY 195<br />

[69] C. Mancini, Dis<strong>en</strong>tangling the jumps from the diffusion in a geom<strong>et</strong>ric jumping Brownian<br />

motion, Giornale <strong>de</strong>ll’Istituto Italiano <strong>de</strong>gli Attuari, LXIV (2001), pp. 19–47.<br />

[70] A. Medve<strong>de</strong>v and O. Scaill<strong>et</strong>, A simple calibration procedure of stochastic<br />

volatility mo<strong>de</strong>ls with jumps by short term asymptotics. Download from<br />

www.hec.unige.ch/professeurs/SCAILLET_Olivier/pages_web/Home_Page.htm, 2004.<br />

[71] R. Merton, Option pricing wh<strong>en</strong> un<strong>de</strong>rlying stock r<strong>et</strong>urns are discontinuous, J. Financial<br />

Economics, 3 (1976), pp. 125–144.<br />

[72] Y. Miyahara, Canonical martingale measures in incompl<strong>et</strong>e ass<strong>et</strong>s mark<strong>et</strong>s, in Proceedings<br />

of the sev<strong>en</strong>th Japan-Russian symposium (Probability theory and mathematical statistics,<br />

Tokyo 1995), World Sci<strong>en</strong>tific, 1996, pp. 343–352.<br />

[73] Y. Miyahara and T. Fujiwara, The minimal <strong>en</strong>tropy martingale measures for geom<strong>et</strong>ric<br />

Lévy processes, <strong>Finance</strong> Stoch., 7 (2003), pp. 509–531.<br />

[74] V. Morozov, On the solution of functional equations by the m<strong>et</strong>hod of regularization,<br />

Sovi<strong>et</strong> Math. Doklady, 7 (1966), pp. 414–417.<br />

[75] V. A. Morozov, The error principle in the solution of operational equations by the regularization<br />

m<strong>et</strong>hod, USSR Comp. Math. Math. Phys., 8 (1968), pp. 63–87.<br />

[76] R. Nels<strong>en</strong>, An Introduction to Copulas, Springer, New York, 1999.<br />

[77] L. Nguy<strong>en</strong>, Calibration <strong>de</strong> Modèles Financiers par Minimisation d’Entropie Relative <strong>et</strong><br />

Modèles aves Sauts, PhD thesis, Ecole Nationale <strong>de</strong> Ponts <strong>et</strong> Chaussées, 2003.<br />

[78] E. Nicolato, Stochastic Volatility Mo<strong>de</strong>ls of Ornstein-Uhl<strong>en</strong>beck Type, PhD thesis, Aarhus<br />

University, Aarhus, 2000.<br />

[79] K. Prause, The G<strong>en</strong>eralized Hyperbolic Mo<strong>de</strong>l: Estimation, Financial Derivatives, and<br />

Risk Measures, PhD thesis, Universität Freiburg i. Br., 1999.<br />

[80] W. H. Press, S. A. Teukolsky, W. T. V<strong>et</strong>terling, and B. P. Flannery, Numerical<br />

Recipes in C: the Art of Sci<strong>en</strong>tific Computing, Cambridge University Press, Cambridge,<br />

1992.


196 BIBLIOGRAPHY<br />

[81] S. Raible, Lévy Processes in <strong>Finance</strong>: Theory, Numerics and Empirical Facts, PhD thesis,<br />

Freiburg University, 1998.<br />

[82] S. I. Resnick, Extreme Values, Regular Variation and Point Processes, Springer-Verlag,<br />

1987.<br />

[83] R. T. Rockafellar, Integrals which are convex functionals, Pacific journal of mathematics,<br />

24 (1968), pp. 525–539.<br />

[84] J. Rosiński, Series repres<strong>en</strong>tations of Lévy processes from the perspective of point processes,<br />

in Lévy Processes — Theory and Applications, O. Barndorff-Niels<strong>en</strong>, T. Mikosch,<br />

and S. Resnick, eds., Birkhäuser, Boston, 2001.<br />

[85] M. Rubinstein, Implied binomial trees, Journal of <strong>Finance</strong>, 49 (1994), pp. 771–819.<br />

[86] T. H. Rydberg, The normal inverse Gaussian Lévy process: simulation and approximation,<br />

Comm. Statist. Stochastic Mo<strong>de</strong>ls, 13 (1997), pp. 887–910.<br />

[87] K. Sato, Lévy Processes and Infinitely Divisible Distributions, Cambridge University<br />

Press, Cambridge, UK, 1999.<br />

[88] B. Schweizer and A. Sklar, Probabilistic M<strong>et</strong>ric Spaces, North Holland, 1983.<br />

[89] M. Schweizer, On the minimal martingale measure and the Föllmer-Schweizer <strong>de</strong>composition,<br />

Stochastic Anal. Appl., 13 (1995), pp. 573–599.<br />

[90] A. Sklar, Random variables, distribution functions, and copulas — a personal look<br />

backward and forward, in Distributions with Fixed Marginals and Related Topics,<br />

L. Rüsch<strong>en</strong>dorf, B. Schweizer, and M. D. Taylor, eds., Institute of Mathematical Statistics,<br />

Hayward, CA, 1996.<br />

[91] M. Stutzer, A simple nonparam<strong>et</strong>ric approach to <strong>de</strong>rivative security valuation, Journal<br />

of <strong>Finance</strong>, 51 (1996), pp. 1633–1652.<br />

[92] P. Tankov, Dep<strong>en</strong><strong>de</strong>nce structure of spectrally positive multidim<strong>en</strong>sional Lévy processes.<br />

Download from www.cmap.polytechnique.fr/~tankov, 2003.


BIBLIOGRAPHY 197<br />

[93] M. Teboulle and I. Vajda, Converg<strong>en</strong>ce of best φ-<strong>en</strong>tropy estimates, IEEE transactions<br />

on information theory, 39 (1993), pp. 297–301.<br />

[94] A. N. Tikhonov and V. Ars<strong>en</strong>in, Solutions of Ill-Posed Problems, Wiley, New York,<br />

1977.<br />

[95] A. N. Tikhonov, A. V. Goncharsky, V. V. Stepanov, and A. G. Yagola, Numerical<br />

M<strong>et</strong>hods for the Solution of Ill-Posed Problems, Kluwer Aca<strong>de</strong>mic Publishers, 1995.<br />

[96] X. Zhang, Analyse Numérique <strong>de</strong>s Options Américaines dans un Modèle <strong>de</strong> Diffusion avec<br />

Sauts, PhD thesis, Ecole Nationale <strong>de</strong>s Ponts <strong>et</strong> Chaussées, 1994.


198 BIBLIOGRAPHY


In<strong>de</strong>x<br />

a posteriori param<strong>et</strong>er choice, 103<br />

alternative principle, 108<br />

bask<strong>et</strong> option, 182<br />

best-of option, 182<br />

calibration problem<br />

continuity of solutions, 67<br />

discr<strong>et</strong>ized, 96<br />

least squares, 59<br />

minimum <strong>en</strong>tropy least squares, 70<br />

numerical solution, 113<br />

regularized, 86<br />

with exact constraints, 58<br />

Cauchy process, 137<br />

characteristic tripl<strong>et</strong>, 34<br />

compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce, 140<br />

compound Poisson mo<strong>de</strong>l, 45<br />

control variate, 182<br />

copula, 136, 146<br />

discrepancy principle, 104<br />

distribution function, 142<br />

drift, 34<br />

expon<strong>en</strong>tial Lévy mo<strong>de</strong>l, 41<br />

abs<strong>en</strong>ce of arbitrage, 41<br />

risk-neutral, 43<br />

F-volume, 141<br />

Fast Fourier transform, 51<br />

g<strong>en</strong>eralized hyperbolic mo<strong>de</strong>l, 47<br />

groun<strong>de</strong>d function, 142<br />

implied volatility, 129<br />

increasing function, 142<br />

jump-diffusion, 43<br />

Kou mo<strong>de</strong>l, 44, 119<br />

Lévy copula, 147, 154<br />

archime<strong>de</strong>an, 168<br />

of a stable process, 163<br />

of compl<strong>et</strong>e <strong>de</strong>p<strong>en</strong><strong>de</strong>nce, 159<br />

of in<strong>de</strong>p<strong>en</strong><strong>de</strong>nce, 159<br />

Lévy-Khintchine repres<strong>en</strong>tation, 34<br />

Lévy measure, 34<br />

Lévy process, 33<br />

absolute continuity, 38<br />

converg<strong>en</strong>ce, 40<br />

cumulants, 35<br />

expon<strong>en</strong>tial mom<strong>en</strong>ts, 35<br />

jump-diffusion, 39<br />

mom<strong>en</strong>ts, 35<br />

simulation, 174<br />

199


200 INDEX<br />

tightness, 40<br />

levycalibration (computer program), 95, 119<br />

margins<br />

of a groun<strong>de</strong>d function, 142<br />

of a Lévy measure, 139<br />

of a Lévy process, 138<br />

of a tail integral, 147<br />

of a volume function, 144<br />

Merton mo<strong>de</strong>l, 44, 60<br />

minimal <strong>en</strong>tropy martingale measure, 72<br />

consist<strong>en</strong>t, 75<br />

Monte-Carlo m<strong>et</strong>hod, 182<br />

vega, 113<br />

volume function, 144<br />

weighted Monte Carlo, 77<br />

neutral <strong>de</strong>rivative pricing, 74<br />

normal inverse Gaussian process, 46<br />

or<strong>de</strong>red s<strong>et</strong>, 140<br />

prior Lévy process, 99<br />

regularization param<strong>et</strong>er, 102<br />

relative <strong>en</strong>tropy, 70<br />

stochastic expon<strong>en</strong>tial, 36<br />

martingale preserving property, 37<br />

strictly or<strong>de</strong>red s<strong>et</strong>, 140<br />

subordinated Brownian motion, 48<br />

tail integral, 147, 152<br />

tempered stable process, 46<br />

truncation function, 34<br />

utility indiffer<strong>en</strong>ce price, 73<br />

variance gamma process, 46, 69, 122, 182

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!