15.01.2013 Views

Tail Dependence - ETH - Entrepreneurial Risks - ETH Zürich

Tail Dependence - ETH - Entrepreneurial Risks - ETH Zürich

Tail Dependence - ETH - Entrepreneurial Risks - ETH Zürich

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.

Eidgenössische Technische Hochschule <strong>Zürich</strong><br />

Swiss Federal Institute of Technology Zurich<br />

Master Thesis<br />

<strong>Tail</strong> <strong>Dependence</strong>:<br />

Implementation, Analysis, and Study<br />

of the most recent concepts<br />

Sebastian N. Schmuki<br />

Department of Management, Technology, and Economics - DMTEC<br />

Chair of <strong>Entrepreneurial</strong> <strong>Risks</strong> - ER<br />

Supervisors:<br />

Prof. Dr. Didier Sornette<br />

Prof. Dr. Yannick Malevergne<br />

October 2008


Acknowledgements<br />

I would like to thank Prof. Dr. Didier Sornette (Chair of <strong>Entrepreneurial</strong> <strong>Risks</strong>, <strong>ETH</strong><br />

<strong>Zürich</strong>) and Prof. Dr. Yannick Malevergne (EM Lyon Business School, Dept. Economics,<br />

Finance & Control) for their motivating way of supervising this thesis. I’am very<br />

grateful for their support, suggestions, and guidance, which gave me insights into many<br />

new fields and made the completion of this thesis possible.<br />

2


Abstract<br />

Recent events i.e. severe market crashes have shown that there is a need for action in<br />

financial risk management to quantify dependences between extreme events. So far,<br />

there is no generally applied measure for extreme financial risks.<br />

The concept of tail dependence represents a powerful means to describe the amount of<br />

extreme co-movements between asset prices. Several approaches for the estimation of<br />

the tail dependence coefficient were developed.<br />

It is the purpose of this work to review and implement the most recent concepts for<br />

the estimation of tail dependence and to study their relevance for practical applications<br />

providing a complete analysis of the different approaches on historical and synthetic<br />

time series.<br />

Programs for the implementation of the discussed concepts are enclosed to the appendix<br />

as Matlab m-codes allowing the non-expert reader to apply them to his own purposes.


Contents<br />

1 Introduction 2<br />

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

2 Important Concepts used in Multivariate Extreme Value Theory 4<br />

2.1 Multivariate Extreme Value Distributions . . . . . . . . . . . . . . . . . 5<br />

2.2 The Survival Function . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.3 Multivariate <strong>Dependence</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.3.1 Positive Orthant <strong>Dependence</strong> . . . . . . . . . . . . . . . . . . . 6<br />

2.3.2 Quadrant <strong>Dependence</strong> . . . . . . . . . . . . . . . . . . . . . . . 7<br />

2.3.3 Associated Variables . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

2.4 Copulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.4.1 Copula and Survival Copula . . . . . . . . . . . . . . . . . . . . 9<br />

2.4.2 Empirical Copula . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.5 <strong>Tail</strong> <strong>Dependence</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

3 Concepts for the Estimation of <strong>Tail</strong> <strong>Dependence</strong> 11<br />

3.1 Approaches according to Poon, Rockinger, and Tawn . . . . . . . . . . 12<br />

3.1.1 Theoretical Background . . . . . . . . . . . . . . . . . . . . . . 12<br />

3.1.2 Implementation of Non-Parametric χ and χ . . . . . . . . . . . 15<br />

3.2 Approaches according to Sornette and Malevergne . . . . . . . . . . . . 21<br />

3.2.1 Theoretical Background . . . . . . . . . . . . . . . . . . . . . . 21<br />

3.2.2 Implementation of the Non-Parametric Approach . . . . . . . . 31<br />

3.2.3 Analysis of TDC estimated by the Non-Parametric Approach . . 34<br />

3.2.4 Implementation of the Parametric Approach . . . . . . . . . . . 57<br />

3.2.5 Analysis of TDC estimated by the Parametric Approach . . . . 66<br />

3.3 β-Smile Improvement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

3.4 Non-Parametric Approaches according to<br />

Schmidt & Stadtmüller . . . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

3.4.1 Theoretical Background . . . . . . . . . . . . . . . . . . . . . . 104<br />

3.4.2 Implementation of Non-Parametric Approaches . . . . . . . . . 105<br />

3.4.3 Analysis of Coefficients . . . . . . . . . . . . . . . . . . . . . . . 107<br />

3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

4 Application of Concepts to various Data and Purposes 127<br />

4.1 Application to major Financial Centers in the World . . . . . . . . . . 127<br />

4.2 Application to Exchange Rates . . . . . . . . . . . . . . . . . . . . . . 142<br />

4.3 Application to Synthetic Time Series . . . . . . . . . . . . . . . . . . . 147<br />

5 Concluding Remarks 156<br />

4


Appendix 160<br />

M-Files . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

Descriptive Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171


List of Figures<br />

3.1 <strong>Tail</strong> index estimators ˆν (Hill’s estimator) for the upper tails of index<br />

S&P 500 and 9 major assets included plotted in dependence of threshold<br />

k on a time interval ranging from July 1985 to April 2008. . . . . . . . 25<br />

3.2 <strong>Tail</strong> index estimators ˆ b γ n<br />

(Gabaix’s estimator) for the upper tails of index<br />

S&P 500 and 9 major assets included plotted in dependence of threshold<br />

k on a time interval ranging from July 1985 to April 2008. . . . . . . . 26<br />

3.3 <strong>Tail</strong> index estimators ˆν (Hill’s estimator) for the lower tails of index S&P<br />

500 and 9 major assets included plotted in dependence of threshold k on<br />

a time interval ranging from July 1985 to April 2008. . . . . . . . . . . 26<br />

3.4 <strong>Tail</strong> index estimators ˆ b γ n (Gabaix’s estimator) for the lower tails of index<br />

S&P 500 and 9 major assets included plotted in dependence of threshold<br />

k on a time interval ranging from July 1985 to April 2008. . . . . . . . 27<br />

3.5 ˆ l(k) = Xk,N/Yk,N with k = 1 . . .150 (k/N = 0% . . .6%) for the lower<br />

tail of all assets with the index S&P 500 on the smaller data set ranging<br />

3.6<br />

from January 1991 up to December 2000. . . . . . . . . . . . . . . . . .<br />

ˆ<br />

� 1 k Xj,N<br />

l(k) = k j=1 with k = 1 . . .150 (or k/N = 0% . . .6%) for the lower<br />

Yj,N<br />

tail of all assets with the index S&P 500 on the smaller data set ranging<br />

32<br />

3.7<br />

from January 1991 up to December 2000. . . . . . . . . . . . . . . . . .<br />

ˆl(k)c =<br />

32<br />

1 �k Xj,N<br />

k−c+1 j=Y with k = 13 . . .150 (or k/N = 0.5% . . .6%),<br />

Yj,N<br />

and c = [0.05 · N] ([·] denotes integer numbers) applied to the lower tail<br />

of all assets with the index S&P 500 on the smaller data set ranging from<br />

3.8<br />

January 1991 up to December 2000. . . . . . . . . . . . . . . . . . . . .<br />

ˆ 1<br />

λ(k) = � �ˆν with k = 1 . . .343 (or k/N = 0% . . .6%), applied to<br />

33<br />

max<br />

1, l<br />

ˆβ<br />

the lower tail of all assets with the index S&P 500 on the bigger data set<br />

ranging from July 1985 to April 2008. . . . . . . . . . . . . . . . . . . . 37<br />

3.9 λ(k) = 1<br />

� �ˆν with k = 1 . . .343 (or k/N = 0% . . .6%), applied to<br />

max<br />

1, ˆ l<br />

ˆβ<br />

the lower tail of all assets with the index S&P 500 on the bigger data set<br />

ranging from July 1985 to April 2008. . . . . . . . . . . . . . . . . . . . 37<br />

3.10 λ(k) = 1<br />

� �ˆν with k = 29 . . .343 (or k/N = 0.5% . . .6%) and c =<br />

max<br />

1, ˆ lĉ<br />

β<br />

[0.005 · N] ([·] denotes integer numbers), applied to the lower tail of all<br />

assets with the index S&P 500 on the bigger data set ranging from July<br />

1985 to April 2008. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

3.11 β(N) estimated for 9 assets X and index S&P 500 Y using the linear<br />

single factor model: X = β·Y +ε for rolling time horizon windows of S =<br />

2500 considered data points from N = (1 . . .S), (2 . . .S +1), . . .,(5736−<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 43<br />

II


3.12 β(N) estimated for 9 assets X and index S&P 500 Y using the linear<br />

single factor model: X = β·Y +ε for rolling time horizon windows of S =<br />

1600 considered data points from N = (1 . . .S), (2 . . .S +1), . . .,(5736−<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 43<br />

3.13 β(N) estimated for 9 assets X and index S&P 500 Y using the linear<br />

single factor model: X = β·Y +ε for rolling time horizon windows of S =<br />

800 considered data points from N = (1 . . .S), (2 . . .S + 1), . . .,(5736 −<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 44<br />

3.14 Constant lU(N) for upper tails plotted in green and constant lL(N) for<br />

lower tails plotted in black estimated for 9 assets and index S&P 500<br />

using relation: ˆl(k) = 1 �k j=1 Xj,S/Yj,S with k = 0.04 · S, Xj,S and Yj,S<br />

k<br />

rank ordered return observations, and rolling time horizon window of S =<br />

2500 considered data points from N = (1 . . .S), (2 . . .S +1), . . .,(5736−<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 44<br />

3.15 Constant lU(N) for upper tails plotted in green and constant lL(N) for<br />

lower tails plotted in black estimated for 9 assets and index S&P 500<br />

using: ˆl(k) = 1 �k j=1 Xj,S/Yj,S with k = 0.04 · S, Xj,S and Yj,S rank<br />

k<br />

ordered return observations, and rolling time horizon window of S =<br />

1600 considered data points from N = (1 . . .S), (2 . . .S +1), . . .,(5736−<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 45<br />

3.16 Constant lU(N) for upper tails plotted in green and constant lL(N) for<br />

lower tails plotted in black estimated for 9 assets and index S&P 500<br />

using: ˆl(k) = 1 �k j=1 Xj,S/Yj,S with k = 0.04 · S, Xj,S and Yj,S rank<br />

k<br />

ordered return observations, and rolling time horizon window of S = 800<br />

considered data points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S +<br />

1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . . . 45<br />

3.17 νk,U(N) for upper tail of index S&P 500 plotted in red and νk,L(N)<br />

for lower tail plotted in black using Hill’s estimator given by equation<br />

� kj=1 ˆν = log Yj,S<br />

�−1 with k = 0.04 · S and Xj,S and Yj,S rank or-<br />

�<br />

1<br />

k<br />

Yk,N<br />

dered return observations for rolling time horizon windows of S = 2500<br />

considered data points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S +<br />

1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . . . 46<br />

3.18 νk,U(N) for upper tail of index S&P 500 plotted in red and νk,L(N)<br />

for lower tail plotted in black using Hill’s estimator given by equation<br />

� kj=1 ˆν = log Yj,S<br />

�−1 with k = 0.04 · S and Xj,S and Yj,S rank or-<br />

�<br />

1<br />

k<br />

Yk,N<br />

dered return observations for rolling time horizon windows of S = 1600<br />

considered data points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S +<br />

1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . . . 46<br />

3.19 νk,U(N) for upper tail of index S&P 500 plotted in red and νk,L(N) for<br />

lower tail plotted in black using Hill’s estimator given by equation ˆν =<br />

� kj=1 log Yj,S<br />

�−1 with k = 0.04 ·S and Xj,S and Yj,S rank ordered re-<br />

�<br />

1<br />

k<br />

Yk,N<br />

turn observations for rolling time horizon windows of S = 800 considered<br />

data points from N = (1 . . .S), (2 . . .S + 1), . . .,(5736 − S + 1 . . .5736)<br />

or a total time interval from July 1985 to Mars 2008. . . . . . . . . . . 47


3.20 λU(N) for the upper tails of index S&P 500 and the nine assets plotted<br />

in blue and λL(N) for the lower tails plotted in red using non-parametric<br />

approach given by equation ˆ λ +,− �<br />

= 1/ max 1, l<br />

�ˆν with coefficients ˆν,<br />

β<br />

l and β for rolling time horizon windows of S = 2500 considered data<br />

points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S + 1 . . .5736) or a<br />

total time interval from July 1985 to Mars 2008. . . . . . . . . . . . . . 48<br />

3.21 λU(N) for the upper tails of index S&P 500 and the nine assets plotted<br />

in blue and λL(N) for the lower tails plotted in red using non-parametric<br />

approach given by equation ˆ λ +,− �<br />

= 1/ max 1, l<br />

�ˆν with coefficients ˆν,<br />

β<br />

l and β for rolling time horizon windows of S = 1600 considered data<br />

points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S + 1 . . .5736) or a<br />

total time interval from July 1985 to Mars 2008. . . . . . . . . . . . . . 49<br />

3.22 λU(N) for the upper tails of index S&P 500 and the nine assets plotted<br />

in blue and λL(N) for the lower tails plotted in red using non-parametric<br />

approach given by equation ˆ λ +,− �<br />

= 1/ max 1, l<br />

�ˆν with coefficients ˆν,<br />

β<br />

l and β for rolling time horizon windows of S = 800 considered data<br />

points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S + 1 . . .5736) or a<br />

total time interval from July 1985 to Mars 2008. . . . . . . . . . . . . . 50<br />

3.23 λU(N) for the upper tails of index S&P 500 and the nine assets for rolling<br />

time horizon windows of size S = 2500 plotted in black, S = 1600 plotted<br />

in blue, and S = 800 plotted in red using non-parametric approach given<br />

by equation ˆ λ +,− �<br />

= 1/ max 1, l<br />

�ˆν with coefficients ˆν, l and β from<br />

β<br />

N = (2501 − S . . .S), (2502 . . .S + 1), . . .,(5736 − S + 1 . . .5736). . . . 51<br />

3.24 λL(N) for the lower tails of index S&P 500 and the nine assets for rolling<br />

time horizon windows of size S = 2500 plotted in black, S = 1600 plotted<br />

in blue, and S = 800 plotted in red using non-parametric approach given<br />

by equation ˆ λ +,− �<br />

= 1/ max 1, l<br />

�ˆν with coefficients ˆν, l and β from<br />

β<br />

N = (2501 − S . . .S), (2502 . . .S + 1), . . .,(5736 − S + 1 . . .5736). . . . 52<br />

3.25 Return data for index S&P 500 ranging from January 1950 to April 2008 55<br />

3.26 Autocorrelation for a return series on the left and for a squared return<br />

series on the right of index S&P 500 ranging from January 1950 to April<br />

2008 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

3.27 Scale factors ĈY (k) = k<br />

N · (Yk,N) ˆν of the index S&P 500 and Ĉε(k) =<br />

k<br />

N · (εk,N) ˆν of the nine assets’ residues plotted for the upper tails of the<br />

bigger data set ranging from July 1985 to April 2008 with k=1, 2,...,<br />

458 (k/N = 0% . . .8%). . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

3.28 Scale factors ĈY (k) = k bˆ · (Yk,N)<br />

N γ n of the index S&P 500 and Ĉε(k) =<br />

k bˆ · (εk,N)<br />

N γ n of the nine assets’ residues plotted for the upper tails of the<br />

bigger data set ranging from July 1985 to April 2008 with k=1, 2,...,<br />

458 (k/N = 0% . . .8%). . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

3.29 Fraction of scale factors<br />

� Ĉε(k)<br />

ĈY (k)<br />

� 1/α<br />

= εk,N<br />

Yk,N<br />

of the nine assets’ residues<br />

and the index S&P 500, plotted for the upper tails of the bigger data<br />

set ranging from July 1985 to April 2008 with k=1, 2,..., 458 (k/N =<br />

0% . . .8%). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59


3.30 Fraction of scale factors<br />

� �1/α Ĉε<br />

ĈY<br />

of the nine assets’ residues and the<br />

index S&P 500, plotted for the upper tails of the bigger data set ranging<br />

from July 1985 to April 2008 with k=1, 2,..., 458 (k/N = 0% . . .8%). . 59<br />

3.31 Empirical complementary cumulative distribution functions F ε,empirical =<br />

k<br />

plotted in red and parametric complementary cumulative distribution<br />

N<br />

functions F ε,parametric = Ĉε ·ε−ˆν with Ĉε for (k/N = 4%) plotted in black<br />

for k/N = 4% . . .0% (from left to right) in dependence of returns for the<br />

upper tails of the nine assets. . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.32 Empirical complementary cumulative distribution functions F ε,empirical =<br />

k plotted in red and parametric complementary cumulative distribu-<br />

N<br />

tion functions F ε,parametric = Ĉε,c · ε−ˆν with Ĉε,c = 1 �k k−c j=c Ĉε for<br />

c = 0.005 · N plotted in black for k/N = 4% . . .0% (from left to right)<br />

in dependence of returns for the upper tails of the nine assets. . . . . . 62<br />

3.33 Empirical complementary cumulative distribution functions F εi,empirical =<br />

k plotted in green and parametric complementary cumulative distribu-<br />

N<br />

tion functions F ε,parametric = Ĉε · ε−ˆb γ n with Ĉε for (k/N = 4%) plotted<br />

in black for the upper tails of k/N = 4% . . .0% in dependence of returns<br />

for the nine assets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.34 Empirical complementary cumulative distribution functions F εi,empirical =<br />

k<br />

plotted in green and parametric complementary cumulative distribu-<br />

N<br />

tion functions F ε,parametric = Ĉε,c · ε−ˆb γ n with Ĉε,c = 1 �k k−c j=c Ĉε for<br />

c = 0.005 · N plotted in black for k/N = 4% . . .0% in dependence of<br />

returns for the upper tails of the nine assets. . . . . . . . . . . . . . . . 64<br />

3.35 ˆ λ = with k = 1 . . .458 (or k/N = 0% . . .8%), applied to the<br />

1<br />

1+ ˆ β −ˆν · Cε<br />

Cy<br />

lower tails of all assets with the index S&P 500 on the bigger data set<br />

ranging from July 1985 to April 2008. . . . . . . . . . . . . . . . . . . . 66<br />

3.36 λ =<br />

1<br />

with k = 1 . . .458 (or k/N = 0% . . .8%), applied to the<br />

1+ ˆ β −ˆν · Ĉε<br />

Ĉy<br />

lower tails of all assets with the index S&P 500 on the bigger data set<br />

ranging from July 1985 to April 2008. . . . . . . . . . . . . . . . . . . . 67<br />

3.37 λ = 1 with k = 29 . . .458 (or k/N = 0.5% . . .8%), and c =<br />

1+ ˆ β −ˆν · Ĉε,c<br />

Ĉy,c<br />

[0.005 · N] ([·] denotes integer numbers) applied to the lower tails of all<br />

assets with the index S&P 500 on the bigger data set ranging from July<br />

1985 to April 2008. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

3.38 Empirical complementary cumulative distribution functions F εi,empirical =<br />

k plotted in red and parametric complementary cumulative distribution<br />

N<br />

functions F ε,parametric = Ĉε · ε−ˆν with Ĉε = 1 �k k j=1 Ĉε plotted in black<br />

for k/N = 4% . . .0% (from left to right) in dependence of returns for the<br />

lower tails of the nine assets for a time interval ranging from January<br />

1991 to December 2000. . . . . . . . . . . . . . . . . . . . . . . . . . . 71


3.39 Empirical complementary cumulative distribution functions F εi,empirical =<br />

k plotted in red and parametric complementary cumulative distribu-<br />

N<br />

tion functions F εi,parametric = Ĉε, c · ε−ˆν with Ĉε,c = 1 �k k−c j=c Ĉε for<br />

c = 0.005 · N plotted in black for k/N = 4% . . .0% (from left to right)<br />

in dependence of returns for the lower tails of the nine assets for a time<br />

interval ranging from January 1991 to December 2000. . . . . . . . . . 72<br />

3.40 Empirical complementary cumulative distribution functions F εi,empirical =<br />

k plotted in red and parametric complementary cumulative distribution<br />

N<br />

functions F ε,parametric = Ĉε · ε−ˆν with Ĉε = 1 �k k j=1 Ĉε plotted in black<br />

for k/N = 4% . . .0% in dependence of returns for the lower tails of the<br />

nine assets for a time interval ranging from January 1991 to December<br />

2000. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

3.41 Empirical complementary cumulative distribution functions F εi,empirical =<br />

k plotted in red and parametric complementary cumulative distribu-<br />

N<br />

tion functions F εi,parametric = Ĉε, c · ε−ˆν with Ĉε,c = 1 �k k−c j=c Ĉε for<br />

c = 0.005 · N plotted in black for k/N = 4% . . .0% in dependence of<br />

returns for the lower tails of the nine assets for a time interval ranging<br />

from January 1991 to December 2000. . . . . . . . . . . . . . . . . . . . 74<br />

3.42 ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in<br />

black calculated by least square method applied on linear additive single<br />

factor model: X = β·Y +ε and first SI condition: Y ≥ Y (k)∩X ≥ X(k)<br />

in dependence of threshold k. Reference ˆ β calculated for all data is given<br />

in green. Y denotes the index return vector of S&P 500 and Y denotes<br />

asset return vector of the nine assets for a time interval ranging from<br />

January 1991 to December 2000. . . . . . . . . . . . . . . . . . . . . . . 80<br />

3.43 ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in<br />

black calculated by least square method applied on linear additive single<br />

factor model: X = β · Y + ε and second SI condition: Y ≥ Y (k) in<br />

dependence of threshold k. Reference ˆ β calculated for all data is given<br />

in green. Y denotes the index return vector of S&P 500 and Y denotes<br />

asset return vector of the nine assets for a time interval ranging from<br />

January 1991 to December 2000. . . . . . . . . . . . . . . . . . . . . . . 81<br />

3.44 ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in<br />

black calculated by least square method applied on linear additive single<br />

factor model: X = β · Y + ε and third SI condition: X ≥ X(k) in<br />

dependence of threshold k. Reference ˆ β calculated for all data is given<br />

in green. Y denotes the index return vector of S&P 500 and Y denotes<br />

asset return vector of the nine assets for a time interval ranging from<br />

January 1991 to December 2000. . . . . . . . . . . . . . . . . . . . . . . 82<br />

3.45 ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in<br />

black calculated by least square method applied on linear additive single<br />

factor model: X = β · Y + ε and fourth SI condition: Y ≥ Y (k) ∪ X ≥<br />

X(k) in dependence of threshold k. Reference ˆ β calculated for all data<br />

is given in green. Y denotes the index return vector of S&P 500 and Y<br />

denotes asset return vector of the nine assets for a time interval ranging<br />

from January 1991 to December 2000. . . . . . . . . . . . . . . . . . . . 83


3.46 ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted<br />

in blue and by the second condition: Y ≥ Y (k) plotted in black for the<br />

upper tail calculated by least squares method applied on linear additive<br />

single factor model: X = β · Y + ε in dependence of threshold k for<br />

k/N = 3% . . .6%. Reference ˆ β calculated for all data is given in green.<br />

Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from January<br />

1991 to December 2000. . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

3.47 ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted<br />

in blue and by the second condition: Y ≥ Y (k) plotted in black for the<br />

lower tail calculated by least squares method applied on linear additive<br />

single factor model: X = β · Y + ε in dependence of threshold k for<br />

k/N = 3% . . .6%. Reference ˆ β calculated for all data is given in green.<br />

Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from January<br />

1991 to December 2000. . . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

3.48 ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted<br />

in blue and by the second condition: Y ≥ Y (k) plotted in black for the<br />

upper tail calculated by least squares method applied on linear additive<br />

single factor model: X = β · Y + ε in dependence of threshold k for<br />

k/N = 3% . . .6%. Reference ˆ β calculated for all data is given in green.<br />

Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from July<br />

1985 to Mars 2008. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

3.49 ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted<br />

in blue and by the second condition: Y ≥ Y (k) plotted in black for the<br />

lower tail calculated by least squares method applied on linear additive<br />

single factor model: X = β · Y + ε in dependence of threshold k for<br />

k/N = 3% . . .6%. Reference ˆ β calculated for all data is given in green.<br />

Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from July<br />

1985 to Mars 2008. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

3.50 ˆ λ + (k) for the upper tail of the nine assets and the index S&P 500 is<br />

plotted in black for ˆ β(k) calculated by least squares method applied to<br />

linear additive single factor model: X = β·Y +ε and adapted to the first<br />

condition: Y ≥ Y (k) ∩ X ≥ X(k), ˆν(k = 4%) and ˆ l(k) in dependence<br />

of threshold k for k/N = 3% . . .10%, in blue for ˆ β(k) and ˆ l(k) adapted<br />

to the first SI condition for comparison, in green for Reference ˆ β and<br />

ˆl(k), and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index<br />

return vector of S&P 500 and X denotes asset return vector of the nine<br />

assets for a time interval ranging from January 1991 to December 2000. 90


3.51 ˆ λ + (k) for the lower tail of the nine assets and the index S&P 500 is<br />

plotted in black for ˆ β(k) calculated by least squares method applied to<br />

linear additive single factor model: X = β·Y +ε and adapted to the first<br />

condition: Y ≥ Y (k) ∩ X ≥ X(k), ˆν(k = 4%) and ˆ l(k) in dependence<br />

of threshold k for k/N = 3% . . .10%, in blue for ˆ β(k) and ˆ l(k) adapted<br />

to the first SI condition for comparison, in green for Reference ˆ β and<br />

ˆl(k), and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index<br />

return vector of S&P 500 and X denotes asset return vector of the nine<br />

assets for a time interval ranging from January 1991 to December 2000. 91<br />

3.52 ˆ λ + (k) for the upper tail of the nine assets and the index S&P 500 is<br />

plotted in black for ˆ β(k) calculated by least squares method applied to<br />

linear additive single factor model: X = β · Y + ε and adapted to the<br />

second condition: Y ≥ Y (k), ˆν(k = 4%) and ˆ l(k) in dependence of<br />

threshold k for k/N = 3% . . .10%, in blue for ˆ β(k) and ˆ l(k) adapted<br />

to the first SI condition for comparison, in green for Reference ˆ β and<br />

ˆl(k), and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index<br />

return vector of S&P 500 and X denotes asset return vector of the nine<br />

assets for a time interval ranging from January 1991 to December 2000. 92<br />

3.53 ˆ λ + (k) for the lower tail of the nine assets and the index S&P 500 is<br />

plotted in black for ˆ β(k) calculated by least squares method applied to<br />

linear additive single factor model: X = β · Y + ε and adapted to the<br />

second condition: Y ≥ Y (k), ˆν(k = 4%) and ˆ l(k) in dependence of<br />

threshold k for k/N = 3% . . .10%, in blue for ˆ β(k) and ˆ l(k) adapted<br />

to the first SI condition for comparison, in green for Reference ˆ β and<br />

ˆl(k), and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index<br />

return vector of S&P 500 and X denotes asset return vector of the nine<br />

assets for a time interval ranging from January 1991 to December 2000. 93<br />

3.54 ˆ βSI(N) plotted in black for the upper tails and plotted in green for the<br />

lower tails for 9 assets given by X and index S&P 500 given by Y using<br />

the linear single factor model: X = β · Y + ε and the first SI condition:<br />

Y ≥ Y (k) ∩ X ≥ X(k) for rolling time horizon windows of S = 2500<br />

considered data points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S +<br />

1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . . . 100<br />

3.55 ˆ βSI(N) plotted in black for the upper tails and plotted in green for the<br />

lower tails for 9 assets given by X and index S&P 500 given by Y using<br />

the linear single factor model: X = β·Y +ε and the second SI condition:<br />

Y ≥ Y (k) for rolling time horizon windows of S = 2500 considered data<br />

points from N = (1 . . .S), (2 . . .S + 1), . . ., (5736 − S + 1 . . .5736) or a<br />

total time interval from July 1985 to Mars 2008. . . . . . . . . . . . . . 101<br />

3.56 λ + (N) for upper tails of index S&P 500 and the nine assets plotted<br />

in blue and λ − (N) for lower tails plotted in red using non-parametric<br />

approach given by equation ˆ λ +,− = 1/ max<br />

�<br />

1,<br />

l<br />

ˆβ +,−<br />

SI<br />

� ˆν<br />

with coefficients<br />

ˆν, l, and ˆ β +,−<br />

SI for rolling time horizon windows of S = 2500 considered<br />

data points from N = (1 . . .S), (2 . . .S+1), . . ., (5736−S+1 . . .5736) or<br />

a total time interval from July 1985 to Mars 2008. ˆ β +,−<br />

SI were calculated<br />

using the linear single factor model: X = β · Y + ε and the first SI<br />

condition: Y ≥ Y (k) ∩ X ≥ X(k). . . . . . . . . . . . . . . . . . . . . . 102


3.57 λ + (N) for upper tail of index S&P 500 and the nine assets plotted in blue<br />

and λ− (N) for lower tail plotted in red using non-parametric approach<br />

given by equation ˆ λ +,− �<br />

l<br />

= 1/ max 1, ˆβ +,−<br />

�ˆν with coefficients ˆν, l, and<br />

SI<br />

ˆβ +,−<br />

SI for rolling time horizon windows of S = 2500 considered data points<br />

from N = (1 . . .S), (2 . . .S+1), . . ., (5736−S+1 . . .5736) or a total time<br />

interval from July 1985 to Mars 2008. ˆ β +,−<br />

SI were calculated using the<br />

linear single factor model: X = β · Y + ε and the second SI condition:<br />

Y ≥ Y (k). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

3.58 ˆ λU,m for upper tails of index S&P 500 and the nine assets plotted in black<br />

and ˆEV T λU,m for upper tails plotted in red, both in dependence of threshold<br />

k on an interval from k/m = 0% . . .6% for m = 2507 daily return values<br />

during a time interval ranging from January 1991 to December 2000. . 109<br />

3.59 ˆ λL,m for lower tails of index S&P 500 and the nine assets plotted in black<br />

and ˆEV T λL,m for lower tails plotted in red, both in dependence of threshold<br />

k on an interval from k/m = 0% . . .6% for m = 2507 daily return values<br />

during a time interval ranging from January 1991 to December 2000. . 110<br />

3.60 Smoothed ˆ λL,m plotted in black and smoothed ˆ λU,m plotted in green for<br />

the nine assets with the index S&P 500 given for the most extreme 8%<br />

during a time interval from January 1991 to December 2000. Smoothing<br />

is performed by locally weighted scatter plot smooth using least squares<br />

quadratic polynomial fitting filters with 25 considered data points by<br />

step or [0.125 · 0.08 · m] values with a total of m = 2507 data points and<br />

[·] denoting integer numbers. . . . . . . . . . . . . . . . . . . . . . . . . 112<br />

3.61 Smoothed ˆ λL,m plotted in black and smoothed ˆ λU,m plotted in green<br />

for the nine assets with the index S&P 500 given for the most extreme<br />

8% during a time interval from July 1985 to April 2008. Smoothing is<br />

performed by locally weighted scatter plot smooth using least squares<br />

quadratic polynomial fitting filters with 57 considered data points by<br />

step or [0.125 · 0.08 · m] values with a total of m = 5736 data points and<br />

[·] denoting integer numbers. . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

3.62 Lower tail dependence λL(θ) = 2−1/θ and corresponding asymptotic vari-<br />

ance σ2 L (θ) = 2−1/θ − 3<br />

24−1/θ + 1<br />

28−1/θ for the Pareto copula. . . . . . . . 114<br />

3.63 λU,m(m) for upper tails of index S&P 500 and the nine assets plotted in<br />

green and λL,m(m) for lower tails plotted in black using a non-parametric<br />

approach given by equation ˆ λL,m = 1 �m k j=1 1R (j)<br />

m1≤k and R(j)<br />

m2≤k and ˆ λU,m =<br />

� 1 m<br />

k j=1 1R (j)<br />

for rolling time horizon windows of S =<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

2500 considered data points from N = (1 . . .S), (2 . . .S +1), . . .,(5736−<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 119<br />

3.64 λU,m(m) for upper tails of index S&P 500 and the nine assets plotted in<br />

green and λL,m(m) for lower tails plotted in black using a non-parametric<br />

approach given by equation ˆ λL,m = 1 �m k j=1 1R (j)<br />

m1≤k and R(j)<br />

m2≤k and ˆ λU,m =<br />

� 1 m<br />

k j=1 1R (j)<br />

for rolling time horizon windows of S =<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

1600 considered data points from N = (1 . . .S), (2 . . .S +1), . . .,(5736−<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 120


3.65 λU,m(m) for upper tails of index S&P 500 and the nine assets plotted in<br />

green and λL,m(m) for lower tails plotted in black using a non-parametric<br />

approach given by equation ˆ λL,m = 1 �m k j=1 1R (j)<br />

m1≤k and R(j)<br />

m2≤k and ˆ λU,m =<br />

� 1 m<br />

k j=1 1R (j)<br />

for rolling time horizon windows of S =<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

800 considered data points from N = (1 . . .S), (2 . . .S + 1), . . .,(5736 −<br />

S + 1 . . .5736) or a total time interval from July 1985 to Mars 2008. . . 121<br />

3.66 λL,m(N) for lower tails of index S&P 500 and the nine assets for rolling<br />

time horizon windows of size S = 2500 plotted in blue, S = 1600<br />

plotted in green, and S = 800 plotted in red using a non-parametric<br />

approach given by equation ˆ λU,m = 1 �m k j=1 1R (j)<br />

from<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

N = (2501 − S . . .S), (2502 . . .S + 1), . . .,(5736 − S + 1 . . .5736). . . . 122<br />

3.67 λU,m(N) for lower tails of index S&P 500 and the nine assets for rolling<br />

time horizon windows of size S = 2500 plotted in blue, S = 1600 plotted<br />

in green, and S = 800 plotted in red a using non-parametric approach<br />

given by equation ˆ λL,m = 1 �m k j=1 1R (j) from N = (2501 −<br />

m1≤k and R(j)<br />

m2≤k S . . .S), (2502 . . .S + 1), . . ., (5736 − S + 1 . . .5736). . . . . . . . . . . . 123<br />

4.1 Non-parametric probability density distribution functions of index S&P<br />

500, assets CVX (Chevron Corp.), and a synthetic sample estimated<br />

using a Gaussian box kernel for a time interval of N = 2507 data points<br />

ranging from January 1991 to December 2000. . . . . . . . . . . . . . . 149<br />

4.2 Non-parametric probability density distribution functions of index S&P<br />

500, assets CVX (Chevron Corp.), and a synthetic sample plotted on<br />

a semi-log scale and estimated using a Gaussian box kernel for a time<br />

interval of N = 2507 data points ranging from January 1991 to December<br />

2000. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149


List of Tables<br />

3.1 Estimated values of upper and lower tail dependence for S&P 500 index<br />

with a set of nine major assets traded on the New York stock Exchange<br />

calculated by a non-parametric approach: ˆχ +,− = Zk,N ·k<br />

and correspond-<br />

N<br />

� Zk,N 2 k(N−k)<br />

ing standard deviation: ˆσ (ˆχ) = N3 . The tail represents the<br />

most extreme 4% of the return values during a time interval from January<br />

1991 to December 2000. ’tail’ shows the calculation interval of L<br />

with c = 13 and k = 100, and ∗ denotes ˆχ �= 1 and therefore ˆχ = 0. . . 18<br />

3.2 Estimated values of upper and lower tail dependence for S&P 500 index<br />

with a set of nine major assets traded on the New York stock Exchange<br />

calculated by a non-parametric approach: ˆχ +,− = Zk,N ·k<br />

and correspond-<br />

N<br />

� Zk,N 2 k(N−k)<br />

ing standard deviation: ˆσ (ˆχ) = N3 . The tail represents the<br />

most extreme 4% of the return values during a time interval from July<br />

1985 to April 2008. ’tail’ shows the calculation interval of L with c = 29<br />

and k = 229, and ∗ denotes ˆχ �= 1 and therefore ˆχ = 0. . . . . . . . . . 18<br />

3.3 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients ˆχ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for χc) of the return values during a time interval<br />

from January 1991 to December 2000. . . . . . . . . . . . . . . . . . . . 19<br />

3.4 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients ˆχ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for χc) of the return values during a time interval<br />

3.5<br />

from July 1985 to April 2008. . . . . . . . . . . . . . . . . . . . . . . .<br />

Comparison of tail index estimators ˆν (Hill’s estimator) and<br />

20<br />

ˆb γ n (estimator<br />

by Gabaix) for positive and negative tails of the S&P 500 index, 9<br />

included assets, and the residues ε obtained by regressing each asset on<br />

the S&P 500 index, during the time intervals ranging from January 1991<br />

to December 2000 with a 4% quantile of k = 100 values and from July<br />

1985 to April 2008 with a 4% quantile of k = 229 values. Values with ∗<br />

represent tail indexes, which can’t be considered equal to the respective<br />

S&P 500 index at the 95% condfidence level. . . . . . . . . . . . . . . . 28<br />

XI


3.6 Establishing the uncertainty of estimated upper and lower tail indexes<br />

ˆν by creating bs = 1000 bootstrap samples of historical return data for<br />

S&P 500 and included assets and calculation of bootstrap mean value<br />

νmean with rel. error from the original value, deviation quantiles and<br />

standard deviation from νmean, and most extreme deviation value. The<br />

tail represents the most extreme 4 % of the return values during a time<br />

interval ranging from January 1991 to December 2000 (k = 100 values)<br />

and from July 1985 to April 2008 (k = 229 values). . . . . . . . . . . . 29<br />

3.7 Establishing the uncertainty of estimated upper and lower tail indexes ˆ b γ n<br />

by creating bootstrap samples (bs) of historical return data for S&P 500<br />

and included assets and calculation of bootstrap mean value νmean with<br />

rel. error from the original value, deviation quantiles and standard deviation<br />

from bγ mean , and most extreme deviation value. The tail represents<br />

the most extreme 4 % of the return values during a time interval ranging<br />

from January 1991 to December 2000 (k = 100 values, bs = 1000) and<br />

from July 1985 to April 2008 (k = 229 values, bs = 650). . . . . . . . . 30<br />

3.8 Estimated values of upper and lower tail dependence for S&P 500 index<br />

with a set of nine major assets traded on the New York stock Exchange<br />

calculated by a non-parametric approach: ˆ λ +,− �<br />

= 1/ max 1, l<br />

�ν on β<br />

the left and by a parametric approach: ˆ λ +,− �<br />

= 1/ 1 + β−ν · Cε<br />

�<br />

on CY<br />

the right. The tail represents the most extreme 4% of the return values<br />

during a time interval from January 1991 to December 2000. ’tail’ shows<br />

the calculation interval of l with c = 13 and k = 100 and ∗ denotes<br />

negative ˆ β. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.9 Estimated values of upper and lower tail dependence for S&P 500 index<br />

with a set of nine major assets traded on the New York stock Exchange<br />

calculated by a non-parametric approach: ˆ λ +,− �<br />

= 1/ max 1, l<br />

� γ<br />

bn on β<br />

the left and by a parametric approach: ˆ λ +,− �<br />

= 1/ 1 + β−bγ �<br />

n Cε · on CY<br />

the right. The tail represents the most extreme 4% of the return values<br />

during a time interval from January 1991 to December 2000. ’tail’ shows<br />

the calculation interval of l with c = 13 and k = 100 and ∗ denotes<br />

negative ˆ β. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

3.10 Estimated values of upper and lower tail dependence for S&P 500 index<br />

with a set of nine major assets traded on the New York stock Exchange<br />

calculated by a non-parametric approach:<br />

35<br />

ˆ λ +,− �<br />

= 1/ max 1, l<br />

�ν on β<br />

the left and by a parametric approach: ˆ λ +,− �<br />

= 1/ 1 + β−ν · Cε<br />

�<br />

on CY<br />

the right. The tail represents the most extreme 4% of the return values<br />

during a time interval from July 1985 to April 2008. ’tail’ shows the<br />

calculation interval of l with c = 29 and k = 229. . . . . . . . . . . . . 35


3.11 Estimated values of upper and lower tail dependence for S&P 500 index<br />

with a set of nine major assets traded on the New York stock Exchange<br />

calculated by a non-parametric approach: ˆ λ +,− �<br />

= 1/ max 1, l<br />

� γ<br />

bn on β<br />

the left and by a parametric approach: ˆ λ +,− �<br />

= 1/ 1 + ˆ β−bγ �<br />

n Cε · on CY<br />

the right. The tail represents the most extreme 4% of the return values<br />

during a time interval from July 1985 to April 2008. ’tail’ shows the<br />

calculation interval of l with c = 29 and k = 229. . . . . . . . . . . . . 36<br />

3.12 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients ˆ λ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval<br />

from January 1991 to December 2000. ˆ βj have been calculated on the<br />

whole samples. ∗ denotes negative ˆ β and therefore ˆ λ = 0 and ’Inf’ denotes<br />

·/0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

3.13 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients<br />

40<br />

ˆ λ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval<br />

from July 1985 to April 2008. ˆ βj have been calculated on the whole<br />

samples. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

3.14 ARCH statistics applied to daily S&P 500 index return data for an interval<br />

ranging from January 1950 to April 2008 and for lags of q = 10,<br />

41<br />

15, and 20 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

3.15 Establishing the uncertainty of parametrically estimated upper and lower<br />

tail dependence coefficients<br />

56<br />

ˆ λ by creating 1000 bootstrap samples of historical<br />

return data tables for S&P 500 index and corresponding asset<br />

returns and calculation of quantiles, extreme values, and standard deviations<br />

of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval from<br />

January 1991 to December 2000. ˆ βj have been calculated on the whole<br />

samples. ∗ denotes negative ˆ β and therefore ˆ λ = 0 and ’Inf’ denotes ·/0.<br />

3.16 Establishing the uncertainty of of parametrically estimated upper and<br />

lower tail dependence coefficients<br />

69<br />

ˆ λ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values, and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval<br />

from July 1985 to April 2008. ˆ βj have been calculated on the whole<br />

samples. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70


3.17 Relative deviations between coefficients of upper and lower tail dependence<br />

estimated by the non-parametric approach of Sornette & Malevergne<br />

using ˆ β calculated for the whole sample and ˆ βSI calculated for the<br />

tail according to three different conditions for index S&P 500 and the<br />

nine assets. The tail represents the most extreme 4 % of data during<br />

a smaller time interval ranging from January 1991 to December 2000<br />

(k = 100) and during a bigger time interval ranging from July 1985 to<br />

Mars 2008 (k = 229). * denotes zero values of tail dependence calculated<br />

by ˆ βSI, ’NaN’ denotes 0/0, and ’inf’ denotes ·/0. . . . . . . . . . . . . . 76<br />

3.18 Results for ˆ βSI using the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k) and<br />

the second SI condition: Y ≥ Y (k) for upper and lower tails calculated<br />

by least squares method applied on linear additive single factor model:<br />

X = β · Y + ε and for resulting upper and lower tail dependence coefficients<br />

ˆ λ + and ˆ λ− with corrected ˆl (c . . .k) and uncorrected ˆl (1 . . .k).<br />

The tail represents the most extreme 4% of data during a time interval<br />

ranging from January 1991 to December 2000 (k = 100, c = 13). *<br />

denotes ˆ β < 0 → λ = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

3.19 Results for ˆ βSI using the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k) and<br />

the second SI condition: Y ≥ Y (k) for upper and lower tails calculated<br />

by least squares method applied on linear additive single factor model:<br />

X = β · Y + ε and for resulting upper and lower tail dependence coefficients<br />

ˆ λ + and ˆ λ− with corrected ˆl (c . . .k) and uncorrected ˆl (1 . . .k).<br />

The tail represents the most extreme 4% of data during a time interval<br />

ranging from July 1985 to Mars 2008 (k = 229, c = 29). * denotes<br />

ˆβ < 0 → λ = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

3.20 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients ˆ λ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values, and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval<br />

from January 1991 to December 2000. ˆ βj have been calculated on the<br />

first SI condition: Y ≥ Y (k) ∩ X ≥ X(k). ’inf’ denotes ·/0. . . . . . . . 95<br />

3.21 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients ˆ λ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values, and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval<br />

from July 1985 to Mars 2008. ˆ βj have been calculated on the first SI<br />

condition: Y ≥ Y (k) ∩ X ≥ X(k). ’inf’ denotes ·/0. . . . . . . . . . . . 96<br />

3.22 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients ˆ λ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values, and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval<br />

from January 1991 to December 2000. ˆ βj have been calculated on the<br />

second SI condition: Y ≥ Y (k). ’inf’ denotes ·/0. . . . . . . . . . . . . 97


3.23 Establishing the uncertainty of non-parametrically estimated upper and<br />

lower tail dependence coefficients ˆ λ by creating 1000 bootstrap samples<br />

of historical return data tables for S&P 500 index and corresponding<br />

asset returns and calculation of quantiles, extreme values, and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding<br />

highest 0.5% for λc) of the return values during a time interval<br />

from July 1985 to Mars 2008. ˆ βj have been calculated on the second SI<br />

condition: Y ≥ Y (k). ’inf’ denotes ·/0. . . . . . . . . . . . . . . . . . . 98<br />

3.24 Estimated values of upper and lower tail dependence for index S&P<br />

500 and a set of nine major assets traded on the New York stock Exchange<br />

calculated by non-parametric approaches according to Schmidt<br />

& Stadtmüller. The tail represents the most extreme 4% of the return<br />

values during a time interval from January 1991 to December 2000 with<br />

a threshold value of k = 0.04·m = 100.3. Columns denoted by ’w. noise’<br />

correspond to columns on their right and are performed by the same approaches<br />

but with a small random noise of amplitude (10−6 ) added to<br />

the returns that no two returns are equal anymore. . . . . . . . . . . . 107<br />

3.25 Estimated values of upper and lower tail dependence for index S&P<br />

500 and a set of nine major assets traded on the New York stock Exchange<br />

calculated by non-parametric approaches according to Schmidt &<br />

Stadtmüller. The tail represents the most extreme 4% of the return values<br />

during a time interval from July 1985 to April 2008 with a threshold<br />

value of k = 0.04·m = 229.4. Columns denoted by ’w. noise’ correspond<br />

to columns on their right and are performed by the same approaches but<br />

with a small random noise of amplitude (10−6 ) added to the returns that<br />

no two returns are equal anymore. . . . . . . . . . . . . . . . . . . . . . 108<br />

3.26 Establishing the uncertainty of non-parametrically estimated upper tail<br />

dependence coefficients ˆ λU,m and ˆEV T λ and lower tail dependence coef-<br />

ficients ˆ λL,m and ˆEV T λL,m U,m<br />

by creating 1000 bootstrap samples of historical<br />

return data tables for S&P 500 index and corresponding asset returns<br />

and calculation of quantiles, extreme values, and standard deviations of<br />

the results. The tails represent the most extreme 4% of the return values<br />

during a time interval from January 1991 to December 2000. . . . . . . 116<br />

3.27 Establishing the uncertainty of non-parametrically estimated upper tail<br />

dependence coefficients ˆ λU,m and ˆEV T λU,m and lower tail dependence coefficients<br />

ˆ λL,m and ˆEV T λL,m by creating 800 bootstrap samples of historical<br />

return data tables for S&P 500 index and corresponding asset returns<br />

and calculation of quantiles, extreme values, and standard deviations of<br />

the results. The tails represent the most extreme 4% of the return values<br />

during a time interval from July 1985 to April 2008. . . . . . . . . . . . 117<br />

4.1 Indexes that were used for the implementation of the different concepts<br />

given by their name, further used abbreviation, and country of origin. . 130<br />

4.2 Assets included in indexes ’AORD’ (ASX), ’CAC 40’, and ’DAX’ that<br />

were used for the implementation of the different concepts given by their<br />

name, further used abbreviation, and country of origin. . . . . . . . . . 131<br />

4.3 Assets included in indexes ’DOW’ (Dow Jones), ’FTSE 100’, and ’JKSE’<br />

that were used for the implementation of the different concepts given by<br />

their name, further used abbreviation, and country of origin. . . . . . . 132


4.4 Assets included in indexes ’MERVAL’, ’MIBTEL’, ’NASDAQ’, and ’SMI’<br />

that were used for the implementation of the different concepts given by<br />

their name, further used abbreviation, and country of origin. . . . . . . 133<br />

4.5 Assets included in indexes ’S&P 500’, ’SSEC’, and ’TA 100’ that were<br />

used for the implementation of the different concepts given by their name,<br />

further used abbreviation, and country of origin. Additional assets included<br />

in index S&P 500 that were not part of the reference samples<br />

presented in chapter (3) are listed below ’new assets’. . . . . . . . . . . 134<br />

4.6 Estimated upper and lower tail dependence ˆ λ +,− applying the non-parametric<br />

and parametric approaches according to Sornette & Malevergne to index<br />

S&P 500 and 16 assets included in dependence of their component<br />

weights in the index denoted by ’C.W.’ and β. The data samples contain<br />

N = 2000 daily price observations on a range from 12.04.2000 to<br />

31.03.2008. <strong>Tail</strong> index ˆν was calculated using Hill’s estimator, k =<br />

0.04 · N = 80, c = 0.005 · N = 10, and ∗ denotes negative ˆ β. . . . . . . 135<br />

4.7 Estimated upper and lower tail dependence ˆ λ +,− applying the non-parametric<br />

approach according to Sornette & Malevergne to index S&P 500 and 16<br />

assets included using βSI only calculated for the extreme tails by first<br />

and second β-smile conditions. Component weights of assets within the<br />

index are denoted by ’C.W.’, βSI for the first and second conditions and<br />

β calculated for all data are listed to observe their impact on the estimates.<br />

The data samples contain N = 2000 daily price observations on a<br />

time interval from 12.04.2000 to 31.03.2008. <strong>Tail</strong> index ˆν was calculated<br />

using Hill’s estimator. ’Cond 1’ denotes: Y ≥ Y (k) ∩ X ≥ X(k), ’Cond.<br />

2’ denotes Y ≥ Y (k), k = 0.04 · N = 80, c = 0.005 · N = 10, and ∗<br />

denotes negative ˆ β. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136<br />

4.8 Estimated upper and lower tail dependence ˆ λ +,− applying the parametric<br />

and the non-parametric approach according to Sornette & Malevergne<br />

to index Dow Jones Industrial Average and 12 assets included using β<br />

calculated by all data. Component weights of assets within the index<br />

are denoted by ’C.W.’, and β calculated for all data are listed to observe<br />

their impact on the estimates. The data samples contain N = 2000 daily<br />

price observations on a time interval from 16.08.2000 to 31.07.2008. <strong>Tail</strong><br />

index ˆν was calculated using Hill’s estimator, k = 0.04 · N = 80, and<br />

c = 0.005 · N = 10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

4.9 Estimated upper and lower tail dependence ˆ λ +,− applying the non-parametric<br />

approach according to Sornette & Malevergne to index Dow Jones Industrial<br />

Average and 12 assets included using βSI only calculated for<br />

the extreme tails by first and second β-smile conditions. Component<br />

weights of assets within the index denoted by ’C.W.’, βSI for the first<br />

and second conditions and β calculated for all data are listed to observe<br />

their impact on the estimates. The data samples contain N = 2000 daily<br />

price observations on a time interval from 16.08.2000 to 31.07.2008. <strong>Tail</strong><br />

index ˆν was calculated using Hill’s estimator, k = 0.04 · N = 80, and<br />

c = 0.005 · N = 10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138


4.10 Estimated upper and lower tail dependence ˆ λ +,− applying the parametric<br />

and the non-parametric approach according to Sornette & Malevergne to<br />

index NASDAQ Composite and 9 assets included using β calculated for<br />

all data and βSI only calculated for the extreme tails by first and second<br />

β-smile conditions. Component weights of assets within the index denoted<br />

by ’C.W.’, βSI for the first and second conditions and β calculated<br />

for all data are listed to observe their impact on the estimates. The data<br />

samples contain N = 2000 daily price observations on a time interval<br />

from 15.08.2000 to 31.07.2008. <strong>Tail</strong> index ˆν was calculated using Hill’s<br />

estimator, k = 0.04 ·N = 80, c = 0.005 ·N = 10, and ∗ denotes negative ˆ β.139<br />

4.11 Estimated upper and lower tail dependence according to Poon, Rockinger,<br />

and Tawn (ˆχ + · , ˆχ − · ) in the upper part and according to Schmidt &<br />

Stadtmüller in the lower part ( ˆ λ·,m, ˆEV T λ·,m ) for index AORD (ASX) and<br />

14 assets included including error bars estimated by bootstrap sampling<br />

with replacement for bs = 1000 bootstrap samples shown next to the<br />

estimates. The data samples contain N = 1398 daily price observations<br />

on a time interval from begin of February 2003 to end of August 2008. 140<br />

4.12 Estimated ˆ λ +,− applying the non-parametric approach according to Sornette<br />

& Malevergne to index AORD (ASX) and 14 assets included using<br />

βSI only calculated for the extreme tails by first and second β-smile conditions.<br />

Error bars estimated by bootstrap sampling with replacement<br />

for bs = 1000 bootstrap samples were large and therefore mostly denoted<br />

by *. The data samples contain N = 1398 data points from February<br />

2003 to end of August 2008. <strong>Tail</strong> index ˆν was calculated using Hill’s<br />

estimator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

4.13 Further used abbreviations of currency exchange rates that were used<br />

for the estimation of tail dependence. . . . . . . . . . . . . . . . . . . . 142<br />

4.14 Estimated ˆ λ +,− applying the non-parametric approach according to Sor-<br />

nette & Malevergne given by: ˆ λ +,− = 1/ max<br />

�<br />

1, l<br />

β<br />

� ν<br />

to exchange rates<br />

of various foreign currencies with the US$. The tail represents the most<br />

extreme 4% of the return values during a time interval ranging from<br />

20.10.2002 to 10.04.2008 consisting of N = 2000 daily observations. ’c’<br />

means that tails were corrected for 0.5% of most extreme data for the<br />

estimation of l. <strong>Tail</strong> index ˆν was calculated using Hill’s estimator. . . . 144<br />

4.15 Estimated ˆχ +,− applying the non-parametric approach according to Poon,<br />

Rockinger, and Tawn given by: ˆχ +,− = Zk,N ·k<br />

N to exchange rates of various<br />

foreign currencies with the US$. The tail represents the most extreme<br />

4% of the return values during a time interval ranging from 20.10.2002<br />

to 10.04.2008 consisting of N = 2000 daily observations. ’c’ means that<br />

tails were corrected for 0.5% of most extreme data for the estimation of<br />

L. <strong>Tail</strong> indexes ˆν were calculated using Hill’s estimator. . . . . . . . . . 145<br />

4.16 Estimated ˆ λL,m, ˆ λU,m, ˆEV T λL,m , and ˆEV T λU,m applying the non-parametric<br />

approaches according to Schmidt & Stadtmüller explained in section<br />

(3.4) to exchange rates of various foreign currencies with the US$. The<br />

tail represents the most extreme 4% of the return values during a time<br />

interval ranging from 20.10.2002 to 10.04.2008 consisting of N = 2000<br />

daily observations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146


4.17 Establishing bias and error bars of ˆ β calculated by all data and effect on<br />

ˆλ +,− estimated by applying the non-parametric approach according to<br />

Sornette & Malevergne to 1000 synthetic samples consisting of N = 2507<br />

data points. The bias is calculated by comparing mean values ˆ β all data<br />

and ˆ λ +,−<br />

( ˆ βalldata) to original values denoted by βorig and corresponding<br />

tail dependence estimator ˆ λ +,− (βorig) and for the estimation of error bars<br />

standard deviation ’std’ and 95% quantiles of the estimates are provided.<br />

Relative bias (’rel.bias’) is calculated as percentage of ˆ λ +,−<br />

(βorig). <strong>Tail</strong><br />

indexes α were set equal to three to avoid a second source of error and<br />

’inf’ means ·/0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150<br />

4.18 Establishing bias and error bars of ˆ βSI 2 calculated by second β-smile<br />

condition: � Y ≥ Y (k) � and effect on ˆ λ +,− estimated by applying the<br />

non-parametric approach according to Sornette & Malevergne to 1000<br />

synthetic samples consisting of N = 2507 data points. The bias is cal-<br />

culated by comparing mean values ˆ β alldata and ˆ λ +,−<br />

( ˆ βalldata) to original<br />

values denoted by βorig and corresponding tail dependence estimator<br />

ˆλ +,− (βorig) and for the estimation of error bars standard deviation ’std’<br />

and 95% quantiles of the estimates are provided. <strong>Tail</strong> indexes α were set<br />

equal to three to avoid a second source of error and ∗ means that relative<br />

bias (’rel.bias’) calculated as percentage of ˆ λ +,−<br />

(βorig) ≥ 100%. . . . . . 151<br />

4.19 Establishing bias and error bars of ˆν and ˆb γ n , estimated for 1000 synthetic<br />

samples consisting of N = 2507 data points. The bias is calculated by<br />

comparing mean values ˆν and ˆb γ<br />

n to original values denoted by α and for<br />

the estimation of error bars standard deviation ’std’ and 95% quantiles<br />

of the estimates are provided. Relative bias (’rel.bias’) is calculated as<br />

percentage of α. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

4.20 Establishing bias and error bars of ˆ λ +,− (ˆν, ˆ βalldata) with ˆ β calculated by<br />

all data and ˆ λ +,− (ˆν, ˆ βSI2) with ˆ β calculated by second β-smile condition:<br />

� Y ≥ Y (k) � by applying the non-parametric approach according<br />

to Sornette & Malevergne to 1000 synthetic samples consisting of<br />

N = 2507 data points. The bias is calculated by comparing mean<br />

values ˆ λ +,−<br />

(ˆν, ˆ βalldata) and ˆ λ +,−<br />

(ˆν, ˆ βSI 2) to original values denoted by<br />

ˆλ +,−<br />

(α, βorig) and for the estimation of error bars standard deviation<br />

’std’ and 95% quantiles of the estimates are provided. ∗ means that<br />

relative bias (’rel.bias’) calculated as percentage of ˆ λ +,−<br />

4.21 Comparison of estimates by non-parametric estimators ˆ λ·,m, ˆ λ<br />

(βorig) ≥ 100%. 153<br />

·,m<br />

cording to Schmidt & Stadtmüller, and ˆχ +,− according to Poon, Rockinger,<br />

and Tawn with ˆ λ +,− (βalldata) according to Sornette & Malevergne by<br />

applying the different concepts to 1000 synthetic samples consisting of<br />

N = 2507 data points created by different βorig. For the estimation of<br />

error bars standard deviation ’std’ and 95% quantiles of the estimates<br />

are calculated and for ˆχ +,− the proposed standard deviation ˆσ is given<br />

for comparison. <strong>Tail</strong> indexes α were estimated by the Hill estimator. . . 155<br />

EV T<br />

ac


5.1 Descriptive statistics of historical return series of S&P 500 and nine<br />

assets included during a time interval from January 1991 to December<br />

2000 (smaller reference sample of chapter (3)) . . . . . . . . . . . . . . 171<br />

5.2 Descriptive statistics of historical return series of S&P 500 and nine assets<br />

included during a time interval from July 1985 to April 2008 (bigger<br />

reference sample of chapter (3)) . . . . . . . . . . . . . . . . . . . . . . 171<br />

5.3 Descriptive statistics of historical return series of S&P 500 and nine assets<br />

included during a time interval from 12.04.2000 to 31.03.2008 (sample<br />

consisting of N = 2000 observations used in chapter (4)) . . . . . . 172<br />

5.4 Descriptive statistics of historical return series of Dow Jones Industrial<br />

Average and 12 assets included during a time interval from 16.08.2000 to<br />

31.07.2008 (sample consisting of N = 2000 observations used in chapter<br />

(4)) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172<br />

5.5 Descriptive statistics of historical return series of NASDAQ Composite<br />

and 9 assets included during a time interval from 15.08.2000 to 31.07.2008<br />

(sample consisting of N = 2000 observations used in chapter (4)) . . . 173<br />

5.6 Descriptive statistics of historical return series of AORD (ASX) and<br />

14 assets included during a time interval from February 2003 to end<br />

of August 2008 (sample consisting of N = 1398 observations used in<br />

chapter (4)) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173<br />

5.7 Descriptive statistics of historical currency exchange rate series during a<br />

time interval from 19.10.2002 to 10.04.2008 (sample consisting of N =<br />

2000 observations used in chapter (4)) . . . . . . . . . . . . . . . . . . 174<br />

5.8 Descriptive statistics of 6 exemplary synthetic samples (sample consisting<br />

of N = 2507 observations used in chapter (4)) . . . . . . . . . . . . 174


Chapter 1<br />

Introduction<br />

1.1 Motivation<br />

The importance of extreme events has become a major issue in financial risk management.<br />

Established risk measures like Pearson’s product moment correlation coefficient<br />

and Spearman’s rank order correlation coefficient are controlled by small movements<br />

around the mean and therefor fail to describe dependence between extreme events. Facing<br />

asset selection and allocation with respect to portfolio management extreme events<br />

are primarily represented by jump risks and default risks.<br />

Extreme events appear in the tails of return distributions of assets and because of the<br />

fact that asset return distributions are heavy tailed, extraordinary downside losses are<br />

more likely to happen than expected assuming Gaussian distributions. During market<br />

crashes or recessions, standard portfolio management and optimization strategies may<br />

break down. Recent events show that impact and frequency of stock market crashes<br />

have increased during the last two decades. Our target is to diversify away extreme<br />

risks by minimizing extreme dependence between assets within a portfolio. Therefore we<br />

focus on the concept of tail dependence, which was first introduced by Sibuya [1] (1960).<br />

The coefficient of tail dependence between two assets is defined as the probability that<br />

one of the two assets undergoes a large loss (or gain) assuming that the other asset also<br />

undergoes a large loss (or gain).<br />

Several approaches were developed for the estimation of the tail dependence coefficient.<br />

It is the purpose of this work to review and implement the most recent concepts<br />

for the estimation of tail dependence and to study their relevance for financial risk<br />

management as well as assets selection and allocation by applying them to a broadbase<br />

index such as the S&P 500 representing the market and an arbitrary asset that is<br />

included in the index, to measure extreme dependence between the two. Using historical<br />

time series and synthetic time series, we will provide an in-depth analysis of the<br />

different concepts. The results of this thesis should provide the investor with a tool<br />

that can directly be applied to real financial data and help ancillary to other criteria to<br />

decide about addition or removal of assets in their portfolios for the purpose of minimizing<br />

extreme dependence. It is important to add that a special effort has been made<br />

in writing down and presenting the results in a way that allows a non-expert reader to<br />

use this methods. Programs in the form of Matlab m-files are enclosed to the appendix<br />

as well as the attached data CD and throughout the whole work cross-references to the<br />

respective codes are provided to facilitate traceability and encourage application of the<br />

concepts.<br />

2


1.2 Thesis Outline<br />

The study of extreme co-movements requires the conjunction of the multivariate extreme<br />

value theory and the notion of copulas. In chapter (2) an introduction to fundamental<br />

concepts used within this work is provided.<br />

Chapter (3) is dedicated to the different concepts for the estimation of the coefficient<br />

of tail dependence.<br />

Starting in section (3.1) with concepts developed by Poon, Rockinger, and Tawn and<br />

published in [13] (2004) we will first provide an insight in their most important theoretical<br />

findings, then give a detailed description how their approaches were implemented,<br />

and finally analyze the results applying the methods to a bigger and a smaller reference<br />

data set of the index S&P 500 and nine major assets included in the index and additionally<br />

provide an estimation of error bars by bootstrap sampling with replacement.<br />

We will use these two reference data sets throughout the whole chapter.<br />

In section (3.2) we will come to the main topic of this thesis consisting of concepts<br />

according to Sornette & Malevergne presented in [21] (2006), [20] (2004), [19] (2002).<br />

We will also start with a subsection about the theoretical background and then continue<br />

with implementation and analysis of the different approaches including error bar<br />

estimations by bootstrap sampling with replacement. In subsection (3.3), we will expand<br />

the concepts by a so-called β-smile improvement condition. Here, we allude to<br />

the literature on the ’β-smile’ using a terminology borrowed from the volatility-smile 1<br />

in financial option theory pointing at the fact that parameter β may exhibit a change<br />

as a function of the quantiles on which it is estimated. For the non-parametric approach<br />

and the non-parametric approach including β-smile improvement we performed<br />

a rolling time window observation to analyze time dynamics and consistency over time<br />

of the estimators.<br />

Our last concept according to Schmidt & Stadtmüller was introduced in [24] (2005)<br />

and is presented in section (3.4). After presenting the theoretical background and<br />

details concerning the implementation of the concepts, we also provide an analysis of<br />

coefficients performed on the reference data sets and error bar estimations by bootstrap<br />

sampling with replacement. Finally, there is also a rolling time window observation performed.<br />

In section (3.5) we will compare the results yielded by the different methods<br />

and summarize the findings of the whole chapter.<br />

Chapter (4) shows the application of the different concepts to various data and<br />

purposes. Starting with the application of the concepts to major financial centers in the<br />

world in section (4.1) shows us interesting new insights into the methods and broadens<br />

our horizon on a global scale. On the other hand, it also reveals some weaknesses of<br />

our approaches.<br />

Section (4.2) provides us with applications of the concepts to a very different purpose:<br />

we use the tail dependence coefficients to estimate extreme dependences of international<br />

currency exchange rates.<br />

In section (4.3) the composition of a synthetic time series is presented in order to check<br />

our methods for bias and lower error bounds. Finally, some concluding remarks are<br />

provided in chapter (5).<br />

1 Pattern showing that at-the-money options tend to have lower implied volatilities than other options<br />

3


Chapter 2<br />

Important Concepts used in<br />

Multivariate Extreme Value Theory<br />

Extreme value theory (EVT) is a branch of statistics, which addresses extreme events.<br />

These are high impact events that happen with very low probability i.e. earthquakes or<br />

100 year flood etc. Extreme measures provide estimates for extreme but unlikely events,<br />

whereas classic general measures tend to focus on the whole empirical distribution and<br />

therefore often neglect low-probability events. Nevertheless, it is known that extreme<br />

events account for a large fraction in terms of effect.<br />

In the financial sector market crashes are extreme realizations of the respective return<br />

distribution. The traditional measure to assess the risk of an asset is the variance of<br />

its returns, introduced by Markowitz [2] (1952). But this measure assumes that the<br />

probability density function (pdf) is of Gaussian nature. Empirical data show that<br />

there is a ’fatness’ in the tails of pdf functions of returns. That means that extreme<br />

events occur more often than assumed by Gaussian approximations.<br />

Anyway, the determination of the precise shape of the tail of the distribution of<br />

returns has proven to be a challenging task. In order to assess risk of high probability<br />

levels of above 95 %, parametric and non-parametric estimations are established:<br />

Parametrization in our context means fitting the parameters of a general model to our<br />

return data to achieve a good approximation. That becomes necessary when risk at<br />

very high probability levels is estimated and non-parametric estimates fail because of<br />

lack of data, which is often the case facing daily return data as we will see later on.<br />

As mentioned in the title we are concerned with ’multivariate’ EVT. In contrast to<br />

univariate distributions, multivariate distributions of returns denote joint probabilities<br />

of realizations of multiple asset return series. This comprises both, risk explained by an<br />

asset’s univariate marginal distribution and risk due to dependence on other assets. To<br />

offer to the interested reader a very complete insight into univariate EVT and also on<br />

the very active and dynamical field of multivariate EVT we refer to [3] (2007) and [4]<br />

(2006) and references therein.<br />

Now we want to introduce some fundamental concepts of multivariate EVT to provide<br />

a wide overview of the field. The explanations will be based on [3] (2007), which<br />

provides a concise and recent overview of Extreme Value analysis.<br />

4


2.1 Multivariate Extreme Value Distributions<br />

In multivariate Extreme Value Analysis it is a standard practice to investigate vectors<br />

of component wise maxima or minima. Letting {(Xi,1, . . .,Xi,d)} for i = 1, . . ., n be<br />

independent and identically distributed (iid) random variables with distribution F the<br />

corresponding row-vector of component wise maxima Mn is given by:<br />

�<br />

Mn = (Mn,1, . . .,Mn,d) = max<br />

1≤i≤n {Xi,1},..., max<br />

1≤i≤n {Xi,d}<br />

�<br />

(2.1)<br />

and the corresponding row-vector of component wise minima � Mn is given by:<br />

�Mn = ( � Mn,1, . . ., � �<br />

Mn,d) = min<br />

1≤i≤n {Xi,1},..., min<br />

1≤i≤n {Xi,d}<br />

�<br />

(2.2)<br />

It is important to notice that the maximum or minimum of each of the different random<br />

variables may occur for different indexes i ∗ 1 , . . .,i∗ d , where ∗ denotes the maximum or<br />

minimum. Therefore Mn and � Mn do not necessarily correspond to observed values<br />

in the original series and the analysis of component wise maxima or minima may<br />

sometimes be of little use.<br />

A standard way to operate is to look for the existence of sequences of constants an,i<br />

and bn,i > 1 for 1 ≤ i ≤ d such that for all (x1, . . ., xd) ∈ Rd the function:<br />

�<br />

Mn,1 − an,1<br />

G (x1, . . .,xd) = lim Pr<br />

≤ x1, . . .,<br />

n→∞ Mn,d − an,d<br />

bn,1<br />

bn,d<br />

≤ xd<br />

= F n (an,1 + bn,1 · x1, . . .,an,d + bn,d · xd) (2.3)<br />

is a proper distribution function with non-degenerate marginals. For minima the following<br />

function holds:<br />

�<br />

�Mn,1 − ãn,1<br />

G(x1, . . .,xd) = lim Pr<br />

≥ x1, . . .,<br />

n→∞ ˜bn,1 � �<br />

Mn,d − ãn,d<br />

≥ xd<br />

˜bn,d �<br />

n<br />

= F ãn,1 + ˜bn,1 · x1, . . .,ãn,d + ˜ �<br />

bn,d · xd , (2.4)<br />

where G and F denote the survival functions associated with G and F defined below.<br />

For maxima the distribution F is said to belong to the Maximum Domain of Attraction<br />

(MDA) of the Multivariate Extreme Value (MEV) distribution G if there exist sequences<br />

of constants an,i and bn,i > 1 for 1 ≤ i ≤ d such that equation (2.3) is satisfied.<br />

A similar definition can be given for minima. The following equations hold for all<br />

(x1, . . .,xd) ∈ R d such that G(x) > 0 and G(x) > 0:<br />

lim<br />

n→∞ n[1 − F(an,1 + bn,1 · x1, . . .,an,d + bn,d · xd)] = − log G (x1, . . .,xd) (2.5)<br />

lim<br />

n→∞ n[1 − F(ãn,1 + ˜bn,1 · x1, . . .,ãn,d + ˜bn,d · xd)] = − log G(x1, . . .,xd) (2.6)<br />

From now on we only concentrate on the study of maxima because as we are using<br />

return series later on with mean around zero we can handle both in the same way by<br />

multiplying minima with factor (−1).<br />

5<br />


Setting all the xi’s but one to +∞ in equation (2.3) yields:<br />

n<br />

lim Fi (an,i + bn,i · xi) = Gi (xi) , fori = 1, . . ., d, (2.7)<br />

n→∞<br />

where Fi and Gi are the i-th marginals of F and G. In turn Fi ∈ MDA (Gi), where Gi<br />

is a Gumbel, Fréchet, or Weibull distribution 1 .<br />

In order to isolate the dependence features from the marginal distribution aspects [6]<br />

(2000), traditionally the components of both the distribution F and the corresponding<br />

MEV distribution G are transformed to standard marginals. It is customary to choose<br />

the standard Fréchet distribution as marginals i.e. the function φ : R → [0, 1] given<br />

by φ(x) = exp � �<br />

−1 for x > 0 and zero elsewhere and by its inverse function: y =<br />

x<br />

φ−1 (x) = −1/ log(x), where ’log’ denotes the natural logarithm. Defining X as a<br />

d-variate random row-vector with distribution F and continuous marginals<br />

Y = φ −1 1<br />

(F(X)) = − , (2.8)<br />

log F(X)<br />

i.e. Yi = φ−1 (Fi(Xi)) = −1/ log F(Xi) for all i ≤ 1 ≤ d. By the Probability Integral<br />

Transform it follows that Y has standard Fréchet marginals. Letting G be a<br />

multivariate distribution with continuous marginals Gi’s and defining:<br />

G ∗ �� � (−1) � � �<br />

(−1)<br />

−1<br />

−1<br />

(y1, . . .,yd) = G<br />

· yi, . . .,<br />

· yd , (2.9)<br />

log G1<br />

log Gd<br />

with y1 > 0, . . .,yd > 0, then G ∗ has standard Fréchet marginals, and G is a MEV distribution<br />

only if G ∗ is also MEV distributed. Thus the marginals of a MEV distribution<br />

can be standardized, yet preserving the extreme value properties.<br />

2.2 The Survival Function<br />

The survival function F associated with multivariate extreme value theory is given by:<br />

F = Pr {X1 > x1, . . .,Xd > xd} (2.10)<br />

In the univariate case d = 1 F = 1 − F(x). Unfortunately this does not hold generally<br />

in multivariate EVT [5] (1988).<br />

2.3 Multivariate <strong>Dependence</strong><br />

In the multivariate case the notions of dependence becomes numerous and complex.<br />

Below we introduce some basic concepts. For further information about how to measure<br />

dependence between random variables we refer to [7] (2006), [8] (2001), and [9] (1997).<br />

2.3.1 Positive Orthant <strong>Dependence</strong><br />

Assuming that X = (X1, . . .,Xd) is a d-variate random row-vector, X is positively<br />

lower orthant dependent (PLOD), if for all x = (x1, . . .,xd) ∈ R d ,<br />

Pr {X ≤ x} ≥<br />

d�<br />

Pr {Xi ≤ xi} (2.11)<br />

i=1<br />

1 This is shown in Theorem 1.7 of chapter 1.2 of [3] (2007)<br />

6


and X is positively upper orthant dependent (PUOD) if for all x = (x1, . . .,xd) ∈ R d ,<br />

Pr {X > x} ≥<br />

d�<br />

Pr {Xi > xi} (2.12)<br />

i=1<br />

Generally X is positively orthant dependent if for all x = (x1, . . .,xd) ∈ R d both<br />

equations (2.11) and (2.12) hold. The definition of negative lower orthant dependence<br />

(NLOD), negative upper orthant dependence (NUOD), and generally negative orthant<br />

dependence (NOD) can be introduced by reversing the sense of the inequalities.<br />

If X has joint distribution F with marginals Fi for i = 1, . . ., d then equation (2.11)<br />

is equivalent to:<br />

for all x ∈ R d and equation (2.12) is equivalent to:<br />

for all x ∈ R d .<br />

2.3.2 Quadrant <strong>Dependence</strong><br />

F (x1, . . .,xd) ≥ F1(x1) · · ·Fd(xd) (2.13)<br />

F (x1, . . .,xd) ≥ F 1(x1) · · ·Fd(xd) (2.14)<br />

Quadrant dependence given by positive quadrant dependence (PQD) and negative<br />

quadrant dependence (NQD) is a bivariate concept of dependence and for d = 2 the<br />

definitions of PUOD and PLOD are equivalent to PQD. Let X and Y be a pair of<br />

continuous random variables with joint distribution function FX,Y and marginals FX<br />

and FY . X and Y are positively quadrant dependent (PQD) if:<br />

and negatively quadrant dependent (NQD) if:<br />

∀(x, y) ∈ R 2 FX,Y ≥ FX(x) · FY (y) (2.15)<br />

∀(x, y) ∈ R 2 FX,Y ≤ FX(x) · FY (y), (2.16)<br />

whereas the quadrant dependence properties are invariant under strictly increasing<br />

transformations. Two random variables X and Y are PQD if Cov (f(X), g(Y )) for<br />

all increasing functions f and g for which the expectations E (f(x)), E (g(y)), and<br />

E (f(x) · g(y)) exist 2 .<br />

2.3.3 Associated Variables<br />

The random variables X1, . . .,Xd are said to be (positively) associated if, for every pair<br />

of a, b of non-decreasing real-valued functions defined on R d ,<br />

Cov (a (X1, . . .,Xd) , b (X1, . . .,Xd)) ≥ 0 (2.17)<br />

2 Cov denotes the covariance, which provides a measure of the strength of the correlation between two or<br />

more random variables and is given by: Cov(X, Y ) = E (X · Y ) − E (X) · E (Y ) assuming that X and Y are<br />

random variables on R<br />

7


whenever the relevant expectations exist [10] (1983). If the random variables X1, . . ., Xd<br />

have a MEV distributions, then they are associated.<br />

A MEV distribution G satisfies the condition:<br />

d�<br />

G (x1, . . .,xd) ≥ Gi (xi) (2.18)<br />

for all x ∈ R d . The forms of limiting multivariate distributions correspond to the cases<br />

of (asymptotic) total independence:<br />

G (x1, . . .,xd) =<br />

and (asymptotic) total dependence:<br />

i=1<br />

d�<br />

Gi (xi) (2.19)<br />

i=1<br />

G (x1, . . .,xd) = min {G1 (x1) , . . .,Gd (xd)} (2.20)<br />

between the component wise maxima for all x ∈ R d . For further details concerning<br />

asymptotic total dependence and asymptotic total independence we refer to [6] (2000)<br />

and [10] (1983).<br />

Pairwise independent random variables having a joint MEV distribution are mutually<br />

independent. Thus the study of asymptotic independence can be confined to the<br />

bivariate case.<br />

Asymptotic total independence arises only if equation (2.7) holds, and there exists<br />

x ∈ R d such that 0 < Gi (xi) < 1 for i = 1, . . ., d and<br />

F n (an,1 + bn,1 · x1, . . .,an,d + bn,d · xd) n→∞<br />

−→ G1 (x1) · · ·Gd (xd) (2.21)<br />

Moreover equation (2.19) only holds for any x ∈ R d if<br />

G(0, . . .,0) = G1(0), · · ·Gd(0) = exp(−d) (2.22)<br />

provided that Gi are standard Gumbel distributions or<br />

provided that Gi are Fréchet distributions or<br />

G(1, . . .,1) = G1(1), · · ·Gd(1) = exp(−d) (2.23)<br />

G(−1, . . .,−1) = G1(−1), · · ·Gd(−1) = exp(−d) (2.24)<br />

(2.25)<br />

provided that Gi are Weibull distributions 3 (2007).<br />

Similar conditions hold for the case of asymptotic total dependence. Asymptotic<br />

dependence arises only if equation (2.7) holds, and there exists x ∈ R d such that<br />

0 < G1 (x1) = · · · = Gd (xd) < 1 and<br />

F n (an,1 + bn,1 · x1, . . .,an,d + bn,d · xd) n→∞<br />

−→ G1 (x1) (2.26)<br />

3 The distinction between these three types of distributions belongs to the field of univariate extreme value<br />

Theory and is explained in i.e. Theorem 1.7 of chapter 1.2 of [3]<br />

8


We refer to Appendix B of [3] (2007) for a detailed handling of dependence between<br />

random variables. Now we come to notions of multivariate extreme value theory that<br />

are fundamental for the comprehension of the concepts that are discussed in chapter<br />

(3).<br />

2.4 Copulas<br />

Facing financial risk management, the diversification of risks between two or more assets<br />

is the major issue. To achieve diversification, it is fundamental to have notice of an<br />

accurate description of the dependence between different assets. Copulas introduced by<br />

Sklar [11] (1959) describe the general dependence structure of several random variables.<br />

For all further definitions I will focus on the bivariate case only.<br />

2.4.1 Copula and Survival Copula<br />

A bivariate distribution function FX,Y of two random variables X and Y with marginal<br />

distributions Fx(·) and Fy(·) can be written in the form:<br />

FX,Y (x, y) = Pr(X ≤ x, Y ≤ y)<br />

= C(Fx(x), Fy(y)), (2.27)<br />

where C(·, ·) with range in [0, 1] × [0, 1] is the copula of the two random variables X<br />

and Y , and is unique if the random variables have continuous marginal distributions.<br />

Moreover, the copula is invariant under strictly increasing transformation of the variables<br />

and is therefore an intrinsic or scale invariant measure of dependence. Denoting<br />

the joint survival function Pr(X > x, Y > y) by F X,Y (x, y) we may write:<br />

F X,Y (x, y) = 1 − FX(x) − FY (y) + FX,Y (x, y)<br />

= C{Fx(x), Fy(y)}, (2.28)<br />

where C(·, ·) with range in [0, 1]×[0, 1] is the survival copula of the two random variables<br />

X and Y .<br />

2.4.2 Empirical Copula<br />

If the bivariate distribution function FX,Y of two random variables X and Y and their<br />

respective marginal distributions Fx(·) and Fy(·) are not known we can calculate the<br />

empirical copula by counting of the number of pairs that satisfy given constraints.<br />

Let {(Rk, Sk)} be the ranks of the random sample {(Xk, Yk)}, then the corresponding<br />

empirical copula is defined as:<br />

Cn (Fx(x) = u, Fy(y) = v) = 1<br />

n<br />

n�<br />

I<br />

k=1<br />

� �<br />

Rk Sk<br />

≤ u and ≤ v , (2.29)<br />

n + 1 n + 1<br />

where u, v ∈ [0, 1] and I is an indicator function counting the cases, where the ’and’<br />

condition is fulfilled.<br />

For a summary of the most important copula families we refer to Appendix C of [3]<br />

(2007).<br />

9


2.5 <strong>Tail</strong> <strong>Dependence</strong><br />

In the previous section we introduced the concept of copula, which describes the general<br />

dependence structure between variables. There are many more specific measures<br />

to estimate dependence between two random variables. But standard measures - as<br />

mentioned above at the example of univariate distributions - are mostly controlled by<br />

relatively small moves of the asset prices around their mean. To study large and extreme<br />

co-movements, we introduce the coefficient of tail dependence between assets<br />

Xi and Xj, introduced by Sibuya [1] (1960) and advanced by Ledford and Tawn [12]<br />

(1996). It is defined as the probability for the asset Xi to incur a large loss (or gain)<br />

assuming that the asset Xj also undergoes a large loss (or gain) at the same probability<br />

level, in the limit where this probability explores the extreme tails of the distribution<br />

of returns of the two assets.<br />

By definition the coefficient of lower tail dependence λ −<br />

ij<br />

tail dependence λ +<br />

ij are defined by:<br />

and the coefficient of upper<br />

λ −<br />

ij = lim<br />

u→0 + Pr{Xi < F −1<br />

i (u)|Xj < F −1<br />

j (u)} (2.30)<br />

λ +<br />

ij<br />

= lim<br />

u→1 − Pr{Xi > F −1<br />

i (u)|Xj > F −1<br />

j (u)} (2.31)<br />

where F −1<br />

i (u) and F −1<br />

j (u) represent the quantiles of assets Xi and Xj at the level u.<br />

To give an example for the illustration of lower tail dependence: Assuming σ1 and σ2<br />

represent volatilities of two different stocks, λ − gives the probability that both stocks<br />

exhibit together very high losses.<br />

Since the measure of tail dependence can be expressed in terms of the copula of Xi<br />

and Xj as shown in equations (2.32) for the lower tail dependence and in equation<br />

(2.33) for the upper tail dependence, it is independent of the marginals and symmetric<br />

in Xi and Xj.<br />

λ −<br />

ij = limu→0 +<br />

C(u, u)<br />

u<br />

λ +<br />

ij = lim u→1 −<br />

= lim u→1 −<br />

C(u, u)<br />

1 − u<br />

1 − 2u + C(u, u)<br />

1 − u<br />

(2.32)<br />

(2.33)<br />

The values for the coefficients of tail dependence are known explicitly for a large number<br />

of different copulas. This can be found i.e. in [14] (2000) published by J.E. Heffernan.<br />

10


Chapter 3<br />

Concepts for the Estimation of <strong>Tail</strong><br />

<strong>Dependence</strong><br />

Many different concepts and approaches have been developed to calculate coefficients of<br />

tail dependence as described by equation (2.30) and (2.31). This chapter presents and<br />

describes in detail the most important concepts for the estimation of tail dependence<br />

as well as their implementation, and a detailed sensitivity analysis of the calculated<br />

coefficients.<br />

In section (3.1) an overview of several concepts introduced by Heffernan, Tawn,<br />

Coles, Pown and Rockinger is provided. Then section (3.2) covers approaches according<br />

to Sornette and Malevergne and in the third section a non-parametric estimator<br />

according to Schmidt and Stadtmüller (3.4) is presented.<br />

All the approaches were analyzed on real data. I downloaded historical price sheets<br />

of the index S&P 500 and nine major stocks included in the index to estimate and<br />

analyze the different coefficients of tail dependence. The data can be accessed on<br />

Yahoo finance 1 and is provided in Microsoft Office Excel-CSV format ranging from<br />

the first mentioned to the actual price. It is updated daily and therefore allows the<br />

ongoing adaptation of coefficients. The evaluation of the data sets was performed by<br />

SPSS 16.0, a commercial statistics software that allows all-in-one-step data analysis.<br />

All further calculations were performed by Matlab 7.0.<br />

In the present chapter the methods were applied on two data sets: A smaller data<br />

set ranging from January 1991 up to December 2000 in accordance with [21] (2002),<br />

and a bigger data set ranging from July 1985 up to April 2008. The period of the bigger<br />

data set contained all available data in order to provide as many large fluctuations of<br />

the returns as possible to achieve meaningful information about extreme dependencies.<br />

All the assets studied are traded on the New York stock Exchange and had to<br />

fulfill the following criteria: First they should be among the stocks with the largest<br />

capitalization and second, each of them should have a weight that is smaller than 1%<br />

in the S&P 500 index, so that the dependence does not stem from the overlap with the<br />

index. The nine stocks were composed of: Bristol-Myers Squibb Co. (BMY), Chevron<br />

Corp. (CVX), Hewlett-Packard Co. (HPQ), Coca-Cola Co. (KO), 3M Co. (MMM),<br />

Procter & Gamble Co. (PG), Schering-Plough Corp (SGP), Texas Instruments Inc.<br />

(TXN), and Walgreen Co. (WAG). Starting at this point of the thesis I will refer to the<br />

stocks by their abbreviations given within brackets. A detailed data analyis is provided<br />

in the appendix.<br />

Since Matlab uses vectors, matrices, and tensors for its calculations, the compiler<br />

1 http://finance.yahoo.com<br />

11


can directly access the Excel sheets. In a first step the price data provided in the Excel<br />

sheets was converted into ”historical day-to-day return data” given by equation (3.1)<br />

to achieve a measure of relative volatility<br />

return(t) = log(price(t)) − log(price(t − 1)), (3.1)<br />

where log(·) denotes the natural logarithm, the inverse of the exponential function.<br />

The different methods are presented structured in subsections as follows:<br />

The first subsection respectively gives an insight into the underlying theory of the concepts<br />

or in other words, a summary of the most important findings in the papers. The<br />

second subsection respectively provides detailed description concerning the implementation,<br />

analysis and adaptation of variables for better robustness. Within the third<br />

subsection respectively, tail dependence estimates according to the different methods is<br />

analyzed i.e. by bootstrap sampling with replacement for an estimation of error bars.<br />

3.1 Approaches according to Poon, Rockinger, and Tawn<br />

In the first subsection I will start with some concepts for the estimation of dependence<br />

measures for multivariate extreme introduced by Poon, Rockinger, and Tawn published<br />

2004 in [13] (2004) with some further explanations provided by Heffernan 2000 in [14]<br />

(2000), by Coles, Heffernan and Tawn in [15] (1999), and by Ledford and Tawn in<br />

[12] (1996). Then I will go into some details concerning non-parametric estimation of<br />

theses measures and the fitting of parametric models. In the second subsection I will<br />

provide an overwiev on the implementation of non-parametric estimators for asymptotic<br />

dependence χ and asymptotic independence χ.<br />

3.1.1 Theoretical Background<br />

Within this subsection I will provide an overview on the different concepts and estimators<br />

that describe asymptotic dependence and asymptotic independence for the<br />

bivariate case.<br />

<strong>Dependence</strong> Measures for Multivariate Extreme<br />

In order to reduce the information contained in the copula C to a single parameter, two<br />

measures of extremal dependence χ and χ were introduced in [15] (1999). Both of them<br />

are needed in order to examine whether two variables are asymptotically dependent or<br />

asymptotically independent. First the bivariate returns X and Y are transformed to<br />

unit Fréchet marginals S and T because of fat-tail distributions for risk asset returns:<br />

S = −1/ log (FX(X)) T = −1/ log (FY (Y )), (3.2)<br />

where FX and FY are marginal distribution functions for X and Y . Variables S and<br />

T have the same dependence structure as X and Y and are now on a common scale,<br />

meaning that events of the form S > s and T > s for large s, correspond to equally<br />

extreme events for each variable. Asymptotic dependence χ was given by:<br />

χ = lim<br />

s→∞ Pr(T > s | S > s) (3.3)<br />

= lim<br />

s→∞<br />

Pr(T > s, S > s)<br />

Pr(S > s)<br />

12


or expressed through copula C:<br />

1 − 2u + C(u, u)<br />

= lim<br />

u→1 1 − u<br />

= lim<br />

u→1 2 −<br />

∼ lim<br />

→1 2 −<br />

1 − C(u, u)<br />

1 − u<br />

log C(u, u)<br />

log u<br />

(3.4)<br />

with 0 ≤ χ ≤ 1. χ is the same as the upper tail dependence coefficient given in equation<br />

(2.31). As mentioned in the previous chapter, S and T are asymptotically dependent if<br />

χ > 0, and they are perfectly dependent if χ = 1. If χ = 0, S and T are asymptotically<br />

independent.<br />

A complementary measure χ developed by Ledford and Tawn (1996) was given to<br />

measure extreme dependence of variables that are asymptotic independent, that is,<br />

where χ = 0:<br />

2 log (Pr(S > s))<br />

χ = lim<br />

− 1<br />

s→∞ log (Pr(S > s, T > s))<br />

(3.5)<br />

2 log(1 − u)<br />

= lim<br />

− 1,<br />

s→∞<br />

log Ĉ(u, u)<br />

(3.6)<br />

where −1 < χ ≤ 1, and χ is a measure of the rate at which Pr(T > t | S > s)<br />

approaches zero. For perfect dependence χ = 1 and for independence χ = 0. Hence<br />

values of χ > 0 indicate that S and T are positively associated, and χ < 0 indicates<br />

that S and T are negatively associated in the extremes. For the bivariate normal<br />

dependence structure, χ = ρ, denoting the coefficient of correlation. Other examples<br />

are listed in Heffernan [14] (2000).<br />

Before drawing conclusions about asymptotic dependence based on χ, it is important<br />

to test if χ = 1. For asymptotically dependent variables, χ = 1 with degree of<br />

dependence given by χ > 0 and for asymptotically independent variables, χ = 0 with<br />

degree of dependence given by χ.<br />

Non-Parametric Estimation of χ and χ<br />

The tail of an univariate heavy-tailed variable Y above a high threshold k is given to:<br />

Pr(Y > y) = L(y) · y −α<br />

where L(y) is a slowly varying function of y, that is<br />

for y > k, (3.7)<br />

L(k · y)<br />

lim = 1 forall y > 0. (3.8)<br />

k→∞ L(k)<br />

Then tail index α is estimated by Hill’s estimator:<br />

�<br />

1 �<br />

kj=1<br />

ˆν = log<br />

k<br />

Yj,N<br />

�−1 Yk,N<br />

13<br />

(3.9)


and L(y), assumed to be constant for y > k, is estimated by:<br />

ˆL(y) = k<br />

N<br />

· (Yk,N) ˆν<br />

(3.10)<br />

with Y1,N ≥ Y2,N ≥ · · · ≥ YN,N denoting ordered statistics of the sample containing N<br />

independent and identically distributed realizations of the market return vector Y and<br />

k denoting the number of the threshold value representing the least extreme value still<br />

counted to the tail of the distribution.<br />

Ledford and Tawn [12] (1996) characterized the joint tail behavior by constant η<br />

denoting the coefficient of tail dependence and slowly varying function L(s) and established<br />

that under weak conditions:<br />

Pr(S > s, T > s) ∼ L(s) · s −1/η<br />

with 0 < η ≤ 1. It follows from the representations in [15] (1999) that<br />

as s → ∞ (3.11)<br />

χ = 2η − 1<br />

⎧<br />

⎨ c if χ = 1 and L(s) → c > 0 as s → ∞<br />

(3.12)<br />

χ = 0 if χ = 1 and L(s) → 0 as s → ∞<br />

⎩<br />

0 if χ < 1<br />

(3.13)<br />

χ = 1 corresponds to η = 1 and yields χ = lims→∞ L(s). Thus the estimation of η<br />

and lims→∞ L(s) provides the basis for the estimation of χ and χ. According to the<br />

modeling assumption for the bivariate case, for Z = min(S, T):<br />

Pr(Z > z) = Pr (min(S, T) > z)<br />

= Pr(S > z, T > z) (3.14)<br />

= L(z) · z −1/η<br />

for z > Zk,N<br />

for some high threshold number k. As η is the tail index of the univariate variable Z, it<br />

can be estimated by equation (3.9) using Hill’s estimator restricted to the interval (0, 1]<br />

and lims→∞ L(s) can be estimated using equation (3.10). The following estimators are<br />

based on the assumption of independent observations on Z. χ was estimated to:<br />

ˆχ = 2<br />

�<br />

k� � �<br />

Zj,N<br />

log<br />

k Zk,N<br />

j=1<br />

�<br />

− 1 (3.15)<br />

var � ˆχ � �<br />

ˆχ + 1<br />

=<br />

�2 (3.16)<br />

k<br />

The estimator of χ, only calculated if there is no significant evidence to reject χ = 1,<br />

was proposed to be:<br />

ˆχ = Zk,N · k<br />

N<br />

var(ˆχ) = Zk,N 2 k(N − k)<br />

N3 (3.17)<br />

(3.18)<br />

ˆχ is a figure that could be compared to other estimators of tail dependence coefficients,<br />

which are explained in the following sections of chapter (3) because it has the same<br />

underlying definition as provided in equation (2.33) showing the formula of upper tail<br />

dependence.<br />

14


Parametric Joint-<strong>Tail</strong> Models<br />

<strong>Dependence</strong> models were estimated over the whole sample space. Conditional on being<br />

above the threshold kX and kY a generalized Pareto distribution was used and below<br />

the empirical distribution function ˜ F(x) was used. So the model for FX(x) (and also<br />

FY (y)) was given as:<br />

FX =<br />

� ˜F(x) if x < kX<br />

1 − 1 − ˜ F(kX)1 + αX(x − kX)/σX −1/αX if x ≥ kX,<br />

(3.19)<br />

where σX and αX are the generalized Pareto distribution scale and shape parameters.<br />

After estimation of the marginal parameter variables, X and Y are transformed to unit<br />

Fréchet form S and T by equation (3.2) to model the dependence structure by parametric<br />

dependence models. Model parameters are fitted by matching their associated<br />

χ and χ with the non-parametric estimates defined in equations (3.17) and (3.15).<br />

3.1.2 Implementation of Non-Parametric χ and χ<br />

Now I will provide a detailed description about the implementation of a non-parametric<br />

estimator for χ, which is a measure of asymptotic dependence and a non-parametric<br />

estimator for χ, which is a measure of asymptotic independence 2 . Both estimators<br />

were described in chapter (3.1.1).<br />

As already explained, we only calculate ˆχ if there is no significant evidence to reject<br />

ˆχ = 1. Or in other words, we only estimate χ under the assumption that χ = η = 1.<br />

Therefore we first calculate ˆχ and its estimated variance given by var � ˆχ � to check<br />

whether value 1 lies in the respective confidence interval of ˆχ. If that is the case, we<br />

go on and calculate ˆχ and its variance estimate var(ˆχ) to obtain a measure for tail<br />

dependence that can be compared to estimates by other methods. If ˆχ �= 1, we set<br />

ˆχ = 0 assuming asymptotic independence between return vectors X and Y .<br />

First we transform return vectors X and Y into vectors of respective Fréchet<br />

marginals given by S and T. Therefore we estimate the survival functions F X(x)<br />

and F Y (y) of the univariate heavy-tailed variables X an Y above a high threshold<br />

defined by threshold number k with k/N = 4% of most extreme return data by:<br />

F X(x) = L(x) · x −α<br />

(3.20)<br />

F Y (y) = L(y) · y −α , (3.21)<br />

where L is a slowly varying function estimated as mean value L in the extreme tail:<br />

L(x) = 1<br />

k<br />

L(y) = 1<br />

k<br />

k�<br />

j=1<br />

k�<br />

j=1<br />

j<br />

N<br />

· (Xj,N) α<br />

(3.22)<br />

j<br />

N · (Yj,N) α , (3.23)<br />

2 The matlab m-file for the implementation of ˆχ is denoted by: lambda porota.m and enclosed to the<br />

appendix and the data CD<br />

15


or estimated as mean value Lc, corrected by the most extreme outliers, with c =<br />

[0.005 ∗ N] and [·] denoting integer numbers:<br />

Lc(x) =<br />

Lc(y) =<br />

1<br />

k − c + 1<br />

1<br />

k − c + 1<br />

and α estimated by Hill’s estimator given by:<br />

k�<br />

j=c<br />

k�<br />

j=c<br />

j<br />

N<br />

· (Xj,N) α<br />

�<br />

1 �<br />

kj=1<br />

ˆνX = log<br />

k<br />

Xj,N<br />

�−1 Xk,N<br />

�<br />

1 �<br />

kj=1<br />

ˆνY = log<br />

k<br />

Yj,N<br />

�−1 (3.24)<br />

j<br />

N · (Yj,N) α , (3.25)<br />

Yk,N<br />

(3.26)<br />

(3.27)<br />

with X1,N ≥ X2,N ≥ · · · ≥ XN,N and Y1,N ≥ Y2,N ≥ · · · ≥ YN,N denoting ordered<br />

statistics of the sample containing N realizations of the market return vector Y and<br />

X. Now transformation to S and T is performed as follows:<br />

ˆS<br />

1<br />

= − �<br />

log 1 − ˆ � (3.28)<br />

F X(x)<br />

ˆT<br />

1<br />

= − �<br />

log 1 − ˆ � (3.29)<br />

F Y (y)<br />

Now we come to the estimation of χ and var � ˆχ � by first calculation of Z applying<br />

relation Z = min(S, T) element-wise to vectors X and Y , and then using:<br />

ˆχ = 2<br />

�<br />

k� � �<br />

Zj,N<br />

log<br />

k Zk,N<br />

j=1<br />

�<br />

− 1 (3.30)<br />

var � ˆχ � �<br />

ˆχ + 1<br />

=<br />

�2 (3.31)<br />

k<br />

The estimator of χ and var(χ), only calculated if {1} ∈ � ˆχ ± var � ˆχ �� by:<br />

ˆχ = Zk,N · k<br />

N<br />

var(ˆχ) = Zk,N 2 k(N − k)<br />

N3 (3.32)<br />

(3.33)<br />

To achieve estimates for the lower tail, we can just sort our data in ascending order<br />

and multiply the ordered sample by (−1). Tables (3.1) and (3.2) show results of the<br />

� Zk,N 2 k(N−k)<br />

estimated χ and ˆσ = N3 for the smaller and the bigger reference sample of<br />

the index S&P 500 and the nine assets.<br />

16


First of all we notice that only for one of the nine assets and the index S&P 500<br />

χ = 1 can be rejected. This can be seen in table (3.1) showing estimates for the smaller<br />

reference sample, where asset Chevron Corp. (CVX) in the negative uncorrected tail is<br />

marked with an asterisk. In both tables (3.1) and (3.2) only estimates of χ are provided<br />

because ˆχ underlies the same definition as the concept of upper tail dependence given in<br />

equation (2.33) and can therefore later on be compared to estimates by other methods.<br />

We also find that there is a big difference between tail dependence estimated for<br />

uncorrected tails on intervals 1, 2 . . .k and tail dependence estimates for corrected tails<br />

on intervals c, c + 1 . . .k, with c = [0.005 ∗ N] and [·] denoting integer numbers, in the<br />

lower tail of the smaller data set given in table (3.1). For upper tails of both data<br />

sets and smaller tails of the bigger data set the estimates for corrected and uncorrected<br />

tails are more consistent with each other and also more consistent over time because the<br />

bigger of the two data sets contains the smaller one and the two respective estimates<br />

are similar for most assets. A reason for these big discrepancies in the lower tail of the<br />

smaller sample might be distortion caused by extreme outliers. Indeed, looking at the<br />

return data sets it can be observed that most extreme values or outliers lie primarily<br />

in the negative tails. What furthermore can be observed is that in case of tail dependence<br />

coefficient estimates for corrected tails, left-tail dependence or tail dependence<br />

of negative tails tends to be stronger than right-tail dependence or tail dependence of<br />

positive tails. This is consistent with previous literature i.e. [13] (2004). Anyway, looking<br />

at estimates for uncorrected tails, we observe that right-tail dependence tends to be<br />

stronger than left-tail dependence. Estimated standard deviations ˆσ were calculated to<br />

be around 6%-7% of the estimates. Thich seems accurate enough.<br />

For comparison I calculated error bars by bootstrap sample estimates. Results for<br />

error bars estimated by bootstrap sampling with replacement are shown in table (3.3)<br />

for the smaller reference data set and in table (3.4) for the bigger reference data set.<br />

Looking at tails that are not corrected for outliers, deviations calculated by bootstrap<br />

sampling for both the smaller and the bigger data sample, primarily for negative<br />

tails, are significantly higher in most cases than calculated by the given relation:<br />

� Zk,N 2 k(N−k)<br />

ˆσ = N3 . Comparing standard deviations ’std bs’ with 90% quantiles and<br />

95%, we observe that error distributions are not Gaussian. Looking at 90% quantiles<br />

of estimates performed for corrected tails χc, the error bars are in most cases similar to<br />

ˆσ shown in tables (3.1) and (3.2), but for some assets also higher. I will come back to<br />

the estimation of error bars by bootstrap sampling with replacement later on and also<br />

provide an explanation about how it is performed 3 .<br />

3 The matlab m-file for the bootstrap sampling of ˆχ is denoted by: bootstr porota.m and enclosed to the<br />

data CD<br />

17


upper tail ˆσ upper tail ˆσ lower tail ˆσ lower tail ˆσ<br />

tail 1... k c... k 1... k c... k<br />

ˆχ<br />

BMY 0.94 0.09 0.92 0.09 0.50 0.05 0.94 0.09<br />

CVX 0.94 0.09 0.92 0.09 0.00* 0.00* 0.94 0.09<br />

HPQ 0.94 0.09 0.91 0.09 0.40 0.04 0.94 0.09<br />

KO 0.94 0.09 0.92 0.09 0.26 0.03 0.94 0.09<br />

MMM 0.94 0.09 0.92 0.09 0.42 0.04 0.94 0.09<br />

PG 0.94 0.09 0.92 0.09 0.46 0.05 0.94 0.09<br />

SGP 0.94 0.09 0.92 0.09 0.34 0.03 0.94 0.09<br />

TXN 0.94 0.09 0.92 0.09 0.52 0.05 0.94 0.09<br />

WAG 0.94 0.09 0.92 0.09 0.36 0.04 0.94 0.09<br />

Table 3.1: Estimated values of upper and lower tail dependence for S&P 500 index with a set<br />

of nine major assets traded on the New York stock Exchange calculated by a non-parametric<br />

approach: ˆχ +,− = Zk,N ·k<br />

� Zk,N 2 k(N−k)<br />

N and corresponding standard deviation: ˆσ (ˆχ) = N3 . The<br />

tail represents the most extreme 4% of the return values during a time interval from January<br />

1991 to December 2000. ’tail’ shows the calculation interval of L with c = 13 and k = 100,<br />

and ∗ denotes ˆχ �= 1 and therefore ˆχ = 0.<br />

upper tail ˆσ upper tail ˆσ lower tail ˆσ lower tail ˆσ<br />

tail 1... k c...k 1... k c... k<br />

ˆχ<br />

BMY 0.96 0.06 0.93 0.06 0.79 0.05 0.97 0.06<br />

CVX 0.96 0.06 0.93 0.06 0.35 0.02 0.97 0.06<br />

HPQ 0.96 0.06 0.92 0.06 0.74 0.05 0.97 0.06<br />

KO 0.93 0.06 0.93 0.06 0.54 0.04 0.97 0.06<br />

MMM 0.96 0.06 0.93 0.06 0.69 0.04 0.97 0.06<br />

PG 0.96 0.06 0.93 0.06 0.58 0.04 0.97 0.06<br />

SGP 0.96 0.06 0.93 0.06 0.74 0.05 0.97 0.06<br />

TXN 0.96 0.06 0.93 0.06 0.70 0.05 0.97 0.06<br />

WAG 0.96 0.06 0.93 0.06 0.73 0.05 0.97 0.06<br />

Table 3.2: Estimated values of upper and lower tail dependence for S&P 500 index with a set<br />

of nine major assets traded on the New York stock Exchange calculated by a non-parametric<br />

N and corresponding standard deviation: ˆσ (ˆχ) = N3 . The<br />

tail represents the most extreme 4% of the return values during a time interval from July<br />

1985 to April 2008. ’tail’ shows the calculation interval of L with c = 29 and k = 229, and ∗<br />

denotes ˆχ �= 1 and therefore ˆχ = 0.<br />

approach: ˆχ +,− = Zk,N ·k<br />

18<br />

� Zk,N 2 k(N−k)


19<br />

m=2507 bs=1000<br />

upper tail lower tail<br />

χbs,mean<br />

χbs,mean<br />

χ k=100 rel err max dev 95% q 90% q std bs k=100 rel err max dev 95% q 90% q std bs<br />

BMY 0.92 0.02 0.91 0.02 0.01 0.08 0.50 0.00 0.52 0.52 0.42 0.32<br />

CVX 0.93 0.01 0.94 0.01 0.01 0.04 0.35 Inf 0.64 0.61 0.60 0.41<br />

HPQ 0.93 0.01 0.93 0.01 0.01 0.07 0.35 0.11 0.63 0.53 0.41 0.22<br />

KO 0.91 0.04 0.90 0.04 0.03 0.20 0.36 0.39 0.64 0.63 0.63 0.30<br />

MMM 0.91 0.03 0.88 0.03 0.03 0.20 0.56 0.34 0.62 0.52 0.38 0.31<br />

PG 0.92 0.02 0.93 0.02 0.02 0.10 0.34 0.25 0.60 0.51 0.42 0.28<br />

SGP 0.92 0.02 0.89 0.02 0.02 0.10 0.34 0.01 0.62 0.44 0.28 0.19<br />

TXN 0.93 0.01 0.92 0.01 0.01 0.04 0.42 0.20 0.53 0.41 0.41 0.27<br />

WAG 0.91 0.03 0.94 0.03 0.03 0.10 0.29 0.20 0.60 0.33 0.32 0.21<br />

χc<br />

BMY 0.90 0.02 0.92 0.03 0.03 0.03 0.90 0.04 0.92 0.11 0.11 0.20<br />

CVX 0.90 0.02 0.12 0.03 0.03 0.02 0.83 0.11 0.81 0.82 0.22 0.31<br />

HPQ 0.90 0.01 0.93 0.03 0.03 0.03 0.91 0.03 0.91 0.14 0.12 0.22<br />

KO 0.91 0.01 0.14 0.04 0.03 0.02 0.90 0.04 0.88 0.09 0.09 0.21<br />

MMM 0.91 0.01 0.13 0.04 0.03 0.02 0.90 0.04 0.89 0.13 0.10 0.24<br />

PG 0.91 0.01 0.12 0.04 0.03 0.02 0.93 0.00 0.91 0.12 0.12 0.13<br />

SGP 0.90 0.01 0.91 0.04 0.03 0.03 0.87 0.07 0.91 0.90 0.12 0.20<br />

TXN 0.90 0.02 0.09 0.03 0.03 0.02 0.84 0.10 0.83 0.82 0.21 0.29<br />

WAG 0.91 0.01 0.10 0.04 0.03 0.02 0.93 0.01 0.87 0.14 0.12 0.10<br />

Table 3.3: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆχ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for χc) of the return values during a time interval from<br />

January 1991 to December 2000.


20<br />

m=5736 bs=1000<br />

upper tail lower tail<br />

χbs,mean<br />

χbs,mean<br />

χ k=229 rel err max dev 95% q 90% q std bs k=229 rel err max dev 95% q 90% q std bs<br />

BMY 0.92 0.03 0.92 0.04 0.03 0.20 0.74 0.06 0.74 0.71 0.21 0.21<br />

CVX 0.95 0.01 0.89 0.01 0.01 0.05 0.43 0.20 0.52 0.47 0.42 0.22<br />

HPQ 0.95 0.01 0.86 0.01 0.01 0.04 0.71 0.04 0.72 0.71 0.44 0.23<br />

KO 0.87 0.07 0.90 0.88 0.09 0.21 0.51 0.06 0.54 0.52 0.40 0.22<br />

MMM 0.79 0.20 0.82 0.79 0.80 0.40 0.68 0.02 0.68 0.31 0.20 0.23<br />

PG 0.94 0.01 0.91 0.01 0.01 0.08 0.44 0.20 0.52 0.40 0.41 0.34<br />

SGP 0.95 0.01 0.93 0.01 0.01 0.05 0.58 0.20 0.64 0.62 0.59 0.29<br />

TXN 0.94 0.02 0.90 0.02 0.02 0.10 0.59 0.20 0.63 0.62 0.61 0.28<br />

WAG 0.95 0.01 0.94 0.01 0.01 0.04 0.70 0.04 0.70 0.69 0.21 0.27<br />

χc<br />

BMY 0.92 0.01 0.93 0.03 0.02 0.07 0.94 0.02 0.91 0.05 0.04 0.13<br />

CVX 0.92 0.00 0.04 0.03 0.02 0.01 0.92 0.05 0.93 0.13 0.07 0.22<br />

HPQ 0.92 0.01 0.05 0.03 0.02 0.01 0.96 0.01 1.01 0.05 0.04 0.10<br />

KO 0.92 0.00 0.92 0.03 0.03 0.05 0.90 0.07 0.93 0.90 0.10 0.24<br />

MMM 0.90 0.04 0.90 0.06 0.05 0.20 0.95 0.02 0.93 0.12 0.04 0.09<br />

PG 0.92 0.01 0.91 0.03 0.02 0.03 0.95 0.02 1.03 0.06 0.05 0.10<br />

SGP 0.92 0.01 0.04 0.02 0.02 0.01 0.95 0.02 1.00 0.09 0.04 0.09<br />

TXN 0.92 0.01 0.87 0.03 0.02 0.06 0.92 0.04 0.93 0.13 0.07 0.22<br />

WAG 0.92 0.01 0.05 0.03 0.02 0.01 0.95 0.02 0.92 0.10 0.05 0.10<br />

Table 3.4: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆχ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for χc) of the return values during a time interval from<br />

July 1985 to April 2008.


3.2 Approaches according to Sornette and Malevergne<br />

A set of parametric and non-parametric estimators for lower and upper tail dependence<br />

is proposed in [19] (2006), [20] (2004), and [21] (2002) assuming a linear factor model<br />

between stock returns and market returns. Within this section the implementation and<br />

analysis of firstly the non-parametric approach, secondly the parametric approach, and<br />

thirdly a so-called β-smile improvement is discussed.<br />

3.2.1 Theoretical Background<br />

In this subsection a summary of the papers [20] (2004) and [21] (2002) is provided. First<br />

we present a non-parametric approach and secondly a parametric approach to calculate<br />

tail dependence between an asset and its market index. Finally, the calculation of tail<br />

dependence between two assets related by a factor model is explained.<br />

The Non-Parametric Approach by Sornette & Malevergne<br />

A linear additive single factor model was introduced to relate asset fluctuations to<br />

market fluctuations:<br />

X = β · Y + ε, (3.34)<br />

where X is a vector of asset returns and Y is the vector of corresponding index returns.<br />

β is the regression coefficient of the components of X to its corresponding components<br />

Y determined by the least squares method with ε denoting the vector of idiosyncratic<br />

noises assumed independent of Y .<br />

A general result concerning the tail dependence generated by factor models was<br />

derived from the definition of upper tail dependence given by equation (2.31). It is<br />

given by:<br />

� ∞<br />

with<br />

λ + =<br />

max{1, l<br />

β }<br />

f(x)dx, (3.35)<br />

l = lim<br />

u→1− F −1<br />

X (u)<br />

F −1<br />

Y (u),<br />

(3.36)<br />

t · PY (t · x)<br />

f(x) = lim ,<br />

t→∞ F Y (t)<br />

(3.37)<br />

where FX and FY are the marginal distribution functions of X and Y , and PY is the<br />

probability density function of Y . F Y = 1 − FY is the complementary cumulative<br />

distribution function of Y . A similar expression holds for the coefficient of lower tail<br />

dependence. A non-vanishing coefficient of tail dependence must have a limit function<br />

f(x) that is non-zero and the constant l must remain finite.<br />

In [20] (2004) it is explained that (3.35) even holds if factor Y and the idiosyncratic<br />

noise ε are dependent, provided that this dependence is not too strong. Another<br />

important statement in this paper is the absence of tail dependence for rapidly varying<br />

factors, which describes the Gaussian, exponential, and any other distribution decaying<br />

faster than any power law, for any arbitrary distribution of the idiosyncratic noise.<br />

21


A general result valid for any regularly varying distribution was provided. Let F Y<br />

follow a regularly varying distribution with tail index α:<br />

F Y (y) = L(y) ∗ y −α<br />

where L(y) is a slowly varying function, i.e.<br />

then:<br />

(3.38)<br />

L(t · y)<br />

lim = 1 for all y > 0, (3.39)<br />

t→∞ L(t)<br />

λ + =<br />

1<br />

�<br />

max 1, l<br />

β<br />

� α<br />

(3.40)<br />

with l given by equation (3.36). To obtain λ − the limit u → 1 − has to be replaced by<br />

u → 0 + .<br />

The Parametric Approach by Sornette & Malevergne<br />

Following the parametric approach according to Sornette & Malevergne we estimate<br />

a parametric form of tail distribution, which for extreme market risks is assumed to<br />

follow a power-law distribution. For ν = νY = νε we have F Y (y) = Ĉy ∗ y −ν and<br />

F ε(ε) = Ĉε ∗ ε −ν for large y and ε, then the coefficient of tail dependence is a simple<br />

function of the ratio Cε/CY of the scale factors:<br />

λ =<br />

1<br />

1 + β −α · Cε<br />

CY<br />

(3.41)<br />

If the tail indexes αY and αε of the distribution of the factors and the residue are<br />

different, then λ = 1 for αY < αε and λ = 0 for αY > αε.<br />

So far we have only considered a single asset with vector of returns X. Let us now<br />

consider a portfolio of assets with vectors of returns Xi, where each asset follows the<br />

linear factor model:<br />

Xi = βi · Y + εi<br />

(3.42)<br />

with independent noises εi, whose scale factors are Cεi . The portfolio X = � wiXi,<br />

with weights wi, also follows the factor model with a parameter β = � wiβi and noise<br />

ε, whose scale factor is Cε = � |wi| ν · Cεi . The tail dependence between the portfolio<br />

and the market index can now be obtained by equation (3.41):<br />

� �<br />

|wi|<br />

λ = 1 +<br />

α · Cεi<br />

( � wiβi) α �−1 (3.43)<br />

· CY<br />

<strong>Tail</strong> <strong>Dependence</strong> between two Assets related by a Factor Model<br />

The tail dependence between two assets related by a factor model given by:<br />

X1 = β1 · Y + ε1<br />

X2 = β2 · Y + ε2<br />

22<br />

(3.44)<br />

(3.45)


is equal to the weakest tail dependence between the assets X1 and X2, and their<br />

common factor Y :<br />

λij = min{λi, λj} (3.46)<br />

Since the dependence between two assets is due to their common factor, this dependence<br />

cannot be stronger than the weakest dependence between each of the assets and the<br />

factor. This result shows that it is sufficient to study the tail dependence between the<br />

assets and their common factor to obtain tail dependence between any pair of assets.<br />

Estimation of <strong>Tail</strong> Indexes<br />

For all our approaches according to Sornette & Malevergne we need to calculate a tail<br />

index α. This can be conducted independently of the method. As we assume stock<br />

and market returns to be asymptotically power-law distributed we can calculate the<br />

tail indexes for both positive and negative tails using Hill’s estimator given by:<br />

ˆν =<br />

�<br />

1 �<br />

kj=1<br />

log<br />

k<br />

Yj,N<br />

Yk,N<br />

� −1<br />

, (3.47)<br />

where Y1,N ≥ Y2,N ≥ · · · ≥ YN,N denote the ordered statistics of the sample containing<br />

N independent and identically distributed realizations of market return vector Y and<br />

k denotes the number of the threshold value representing the least extreme value still<br />

counted to the tail of the distribution. Assuming that the tail index of each asset αXi<br />

is the same as the tail index of the market index αY containing the asset, k denotes the<br />

number of the thteshold value of the N sorted return observation or the smallest value<br />

for each, market index and asset, still considered to belong to the tail. Hill’s estimator<br />

is asymptotically normal distributed with mean α and variance α 2 /k. But for finite k<br />

it is known that the estimator is biased.<br />

To find out whether tail indexes calculated by Hill’s estimator are accurate, I performed<br />

a test under conditions similar to our situation: Assuming heavy tail distributed<br />

market returns with a tail index of α = 3 and given a general formulation of heavy tail<br />

probability density distribution functions:<br />

p(y) ∼ α<br />

y 1+α,<br />

(3.48)<br />

I generated n = 100 random points on an uniformly distributed interval ranging from<br />

zero to one (p(r) = 1, for r U[0,1]).<br />

Then I performed a variable transformation applying p(y)dy = p(r)dr to convert the<br />

uniformly distributed random values into heavy tail distributed values given by relation<br />

(3.48). Putting in p(r) and p(y) yields:<br />

α<br />

y1+αdy = dr (3.49)<br />

and by integral of both sides: C − y −α = r, where constant C can be calculated using<br />

the probability condition:<br />

zeros(C − y −α 1<br />

−<br />

) = {C α }, for ∀C > 0<br />

� �<br />

Y2<br />

limY2→∞ p(y)dy = − 1<br />

yα �∞ = C = 1, for ∀α > 0<br />

C − 1 α<br />

23<br />

C − 1 α


and finally I obtained:<br />

y(r) =<br />

� � 1<br />

α 1<br />

, (3.50)<br />

1 − r<br />

which allowed me to transform our uniformly distributed data into data distributed<br />

according to relation (3.48) with α = 3.<br />

Putting now the transformed values into equation (3.47) to calculate Hill’s estimator,<br />

I could test whether the output was a tail index ˆν close enough to three representing<br />

the true value α of the tail index for our transformed data.<br />

Conducting this procedure several times allowed me to check whether the standard<br />

deviation fulfills the above described relation of:<br />

std(ˆν)<br />

α ∼ = 1<br />

√ k , (3.51)<br />

which, given α = 3 and n = 100, says that std(ˆν) ∼ = 0.3 and therefore all ˆν should be<br />

within an interval of 3 ±2∗0.3 if the error follows a Gaussian distribution as requested.<br />

I performed the above described sampling 10’000 times and received a mean value of<br />

mean(ˆν) = 3.06, which shows a small bias of 2%, and a 95% deviation-quantile of 0.63,<br />

which is close to the above calculated 0.6. Therefore I conclude that our smaller sample<br />

size with about 100 return data points perceived as tail data already agrees well with<br />

the general assumption of Hill’s estimator to be asymptotically normally distributed.<br />

However, performing this little affirmation we stressed the assumption of using independent<br />

and identically distributed (i.i.d.) data samples, which unfortunately is not<br />

the case for real financial data. Applying Engle’s test for the presence of ARCH effects<br />

[17] (1982) to our real financial time series the result shows significant evidence in<br />

support of GARCH effects (heteroscedasticity) 4 . As shown by Kearns and Pagan [22]<br />

(1997) for heteroskedastic time series the variance of the estimated tail index can be<br />

seven times larger than the variance given by the asymptotic normality assumption.<br />

For comparison I also implemented the OLS (ordinary least squares) log-log rank-size<br />

tail index estimate proposed by Gabaix [23] (2006) defined by:<br />

log(Rank − 1/2) = a − ˆb γ n · log(Size), (3.52)<br />

with the tail index given by ˆb γ n . A standard error of the OLS estimate ˆb γ � n of the slope<br />

2<br />

was estimated to n ˆb γ n in the paper. Assuming a tail index equal to three as above,<br />

this yields a standard error of 0.42, which is higher compared to Hill’s estimator.<br />

We can now perform the same little check as above to see how Gabaix’s ˆb γ n estimator<br />

performs for i.i.d. data samples. Sampling 10’000 times, I obtained a mean value of<br />

mean( ˆb γ n) = 3.015, which only constitutes 0.5% bias and a 95% deviation-quantile of<br />

0.83. This perfectly agrees with the estimated standard error given in the paper [23]<br />

(2006) but still is a little higher than the standard deviation of Hill’s estimator.<br />

Now we can apply both, Hill’s estimator ˆν and Gabaix’s ˆb γ n estimator, on real data<br />

for comparison and subsequently perform bootstrap sampling with replacement (Monte<br />

Carlo simulation) to find out, whether the fact, that real financial time series exhibit<br />

internal dependence, in terms of time varying volatility clustering (large price changes<br />

come in clusters), impact deviation quantiles or in other words: whether Hill’s estimator<br />

4 Engle’s ARCH test is described in subsection (3.2.3)<br />

24


ν(k)<br />

4<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3<br />

2.8<br />

2.6<br />

2.4<br />

S&P 500<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

2.2<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.1: <strong>Tail</strong> index estimators ˆν (Hill’s estimator) for the upper tails of index S&P 500<br />

and 9 major assets included plotted in dependence of threshold k on a time interval ranging<br />

from July 1985 to April 2008.<br />

still performs better in a deviation quantile sense compared to Gabaix’s ˆ b γ n estimator<br />

when it is applied to real financial time series.<br />

In figures (3.1) and (3.2) I plotted both estimators in dependence of threshold k for<br />

the upper tails and in figures (3.3) and (3.4) for the lower tails. In the positive tails<br />

the developing of the values with increasing k looks quite similar for the tail indexes<br />

calculated by Hill’s estimator and the tail indexes calculated by Gabaix’s estimator,<br />

although the ˆb γ n (k) estimator is less sensitive to k. In the negative tails ˆb γ n looks quite<br />

different from ˆν(k). Assets are increasing in k, whereas the index S&P 500 is slightly<br />

decreasing and at a quite high threshold of 15% they kind of converge. What all the<br />

is that they<br />

assets and the index S&P 500 have in common in the negative tails of ˆb γ n<br />

first tend to rise and then build something like a plateau before they drop. This plateau<br />

is also apparent in the positive tails, but there it is much less distinctive and at much<br />

smaller thresholds k.<br />

Now we have to define a relevant range for the estimation of the tail index in terms<br />

of threshold k. As k increases, the variance of the estimator decreases while its bias<br />

increases. It is reported in [20] (2004) that an accurate determination of the optimal k<br />

by optimization algorithms is rather difficult. Although a relevant range for k between<br />

1% and 5% of total data for the estimation of the tail index was given, I decided to<br />

search visually for a k that looks consistent for the 9 assets and the index S&P 500. The<br />

ˆγ bn estimators shown look quite smooth in the area of 4%. Some assets show something<br />

like a plateau around there or at least the tail index estimators seem to be stable there.<br />

Table (3.5) shows the results of tail indexes of all assets for both sample sizes,<br />

the index S&P 500, and the residues ε obtained by regressing each asset on the S&P<br />

500 index, assuming a threshold of k = 4%. It is interesting that Hill’s estimator is<br />

smaller than Gabaix ˆb γ n estimator for the upper tail and is systematically larger for the<br />

25


γ<br />

b<br />

n<br />

ν(k)<br />

4.2<br />

4<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3<br />

S&P 500<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

2.8<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.2: <strong>Tail</strong> index estimators ˆ b γ n (Gabaix’s estimator) for the upper tails of index S&P 500<br />

and 9 major assets included plotted in dependence of threshold k on a time interval ranging<br />

from July 1985 to April 2008.<br />

3.5<br />

3<br />

2.5<br />

S&P 500<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

2<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.3: <strong>Tail</strong> index estimators ˆν (Hill’s estimator) for the lower tails of index S&P 500 and<br />

9 major assets included plotted in dependence of threshold k on a time interval ranging from<br />

July 1985 to April 2008.<br />

26


γ<br />

b<br />

n<br />

3.2<br />

3<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

1.6<br />

S&P 500<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

1.4<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.4: <strong>Tail</strong> index estimators ˆ b γ n (Gabaix’s estimator) for the lower tails of index S&P 500<br />

and 9 major assets included plotted in dependence of threshold k on a time interval ranging<br />

from July 1985 to April 2008.<br />

lower tail. Furthermore, it is shown in [23] (2006) that although the standard error of<br />

the Hill’s estimator is less than that of the tail index estimate ˆb γ n, its (small sample)<br />

bias in the case of deviation from the power law is (typically) greater than that of ˆb γ n .<br />

Perhaps, we indeed under-estimate the lower tail such that the exponent is probably<br />

smaller than given by Hill’s estimator. To check whether our assumption of equal tail<br />

indexes of asset returns, residues, and index returns is accurate with our estimation<br />

methods and data sets, I performed a simple test of means at the 95% confidence level<br />

for any given quantile. Values that reject the equality hypothesis are asterisked. As<br />

we can see, the results were inconclusive. Anyway, the performed tests were based on<br />

the assumption of asymptotic normality in the distribution of the estimates with the<br />

standard deviations of � 2/n · ˆb γ n and ˆν/√ k, which might be too restrictive. The true<br />

distribution is probably ’wilder’ than Gaussian as explained above. This would expand<br />

the confidence interval. It is nevertheless interesting to notice that in the negative<br />

tails of ˆ b γ n<br />

, although the equality hypothesis was rejected for all assets and residues<br />

because estimates for assets and residues were unexeptionally lower than for the S&P<br />

500 index, tail index estimates of asset returns are nearly identical to tail indexes of<br />

their corresponding residues obtained by regression on the index S&P 500.<br />

As we can obtain in the upper parts of the tables (3.6) and (3.7) where the smaller<br />

samples with 100 data points considered as tail data are shown, 95% quantiles of the<br />

upper tails show good accordance to our i.i.d. estimates, whereas 95% quantiles of<br />

lower tails tend to be somewhat higher. In terms of deviation quantiles Hill’s estimator<br />

still outperforms Gabaix’s ˆ b γ n estimator. What we can observe in the lower parts of the<br />

tables is that deviation quantiles become obviously smaller when we augment sample<br />

sizes. This counts for both estimators.<br />

27


28<br />

+tail −tail<br />

ν b γ n ν b γ n ν b γ n ν b γ n<br />

k 100 100 100 100 229 229 229 229 100 100 100 100 229 229 229 229<br />

Asset ε Asset ε Asset ε Asset ε Asset ε Asset ε Asset ε Asset ε<br />

BMY 3.36 3.14 3.66 3.63 3.31 3.05 3.59 3.38 2.55 2.57 1.80* 1.76* 2.46* 2.46* 2.04* 1.97*<br />

CVX 3.73 3.67 4.26 3.89 3.49* 3.54* 3.83 3.60 3.61 3.61 2.30* 2.27* 3.14 3.20 2.33* 2.29*<br />

HPQ 3.42 3.02 3.94 3.42 2.88 2.65* 3.29 3.01 2.82 3.02 1.80* 1.70* 2.91 2.53* 2.35* 2.17*<br />

KO 3.50 3.08 3.70 3.36 3.48 3.40 3.36 3.24 2.79 3.08 1.80* 1.75* 2.48* 2.59* 1.80* 1.79*<br />

MMM 3.80 3.34 4.10 3.86 3.56* 3.22 3.69 3.54 2.89 3.34 2.23* 2.21* 2.48* 2.39* 2.04* 1.96*<br />

PG 3.40 3.14 3.57 3.76 3.19 3.03 3.60 3.39 2.45* 3.14 1.57* 1.55* 2.40* 2.41* 1.69* 1.66*<br />

SGP 3.37 3.11 3.78 3.72 3.36 2.97 3.88 3.45 2.69 3.11 1.70* 1.65* 2.40* 2.27* 1.84* 1.78*<br />

TXN 3.51 3.52 3.99 3.99 3.03 3.10 3.34 3.34 2.67 3.52 1.77* 1.77* 2.73 2.68* 2.05* 2.03*<br />

WAG 3.61 3.99* 3.92 4.22 3.33 3.40 3.81 3.58 2.42* 3.99* 1.46* 1.43* 2.44* 2.65* 1.93* 1.92*<br />

S&P 500 3.23 − 3.71 − 3.09 − 3.56 − 3.16 − 3.37 − 3.13 − 3.04 −<br />

Table 3.5: Comparison of tail index estimators ˆν (Hill’s estimator) and ˆ b γ n (estimator by Gabaix) for positive and negative tails of the S&P 500<br />

index, 9 included assets, and the residues ε obtained by regressing each asset on the S&P 500 index, during the time intervals ranging from January<br />

1991 to December 2000 with a 4% quantile of k = 100 values and from July 1985 to April 2008 with a 4% quantile of k = 229 values. Values with<br />

∗ represent tail indexes, which can’t be considered equal to the respective S&P 500 index at the 95% condfidence level.


29<br />

m=2507 bs=1000<br />

upper tail lower tail<br />

νmean<br />

ν k=100 rel err max dev 95% q 90% q std bs k=100 rel err max dev 95% q 90% q std bs<br />

Index 3.31 0.02 1.01 0.60 0.52 0.32 3.19 0.01 0.82 0.62 0.53 0.34<br />

BMY 3.42 0.02 1.22 0.64 0.52 0.31 2.62 0.03 1.10 0.72 0.60 0.28<br />

CVX 3.78 0.01 1.31 0.61 0.53 0.30 3.69 0.02 1.61 1.00 0.82 0.50<br />

HPQ 3.50 0.02 1.10 0.63 0.51 0.29 2.89 0.02 1.31 0.83 0.71 0.40<br />

KO 3.55 0.02 1.28 0.62 0.50 0.28 2.86 0.03 1.32 0.84 0.74 0.43<br />

MMM 3.85 0.01 1.36 0.67 0.58 0.40 2.94 0.02 1.20 0.71 0.62 0.41<br />

PG 3.45 0.02 1.32 0.59 0.53 0.31 2.51 0.03 1.24 0.70 0.61 0.44<br />

SGP 3.44 0.02 1.20 0.61 0.50 0.31 2.75 0.02 1.33 0.82 0.61 0.43<br />

TXN 3.56 0.01 1.21 0.60 0.52 0.32 2.74 0.03 1.24 0.70 0.62 0.42<br />

WAG 3.68 0.02 1.30 0.63 0.50 0.30 2.48 0.02 1.20 0.71 0.63 0.40<br />

m=5736 bs=1000<br />

ν k=229 k=229<br />

Index 3.10 0.01 0.62 0.33 0.32 0.21 3.16 0.01 0.58 0.42 0.33 0.24<br />

BMY 3.32 0.00 0.61 0.40 0.31 0.21 2.48 0.01 0.50 0.40 0.34 0.21<br />

CVX 3.50 0.00 0.71 0.42 0.34 0.21 3.17 0.01 0.82 0.53 0.42 0.22<br />

HPQ 2.89 0.00 0.50 0.33 0.32 0.22 2.93 0.01 0.71 0.44 0.42 0.20<br />

KO 3.50 0.01 0.74 0.43 0.41 0.22 2.51 0.01 0.64 0.39 0.31 0.21<br />

MMM 3.57 0.00 0.73 0.41 0.40 0.21 2.51 0.01 0.63 0.41 0.34 0.22<br />

PG 3.20 0.00 0.61 0.33 0.32 0.23 2.51 0.01 0.70 0.42 0.32 0.22<br />

SGP 3.37 0.00 0.61 0.40 0.28 0.23 2.42 0.01 0.63 0.40 0.34 0.20<br />

TXN 3.04 0.00 0.62 0.33 0.29 0.24 2.75 0.01 0.71 0.41 0.33 0.24<br />

WAG 3.34 0.00 0.62 0.40 0.31 0.23 2.46 0.01 0.60 0.40 0.32 0.20<br />

Table 3.6: Establishing the uncertainty of estimated upper and lower tail indexes ˆν by creating bs = 1000 bootstrap samples of historical return<br />

data for S&P 500 and included assets and calculation of bootstrap mean value νmean with rel. error from the original value, deviation quantiles<br />

and standard deviation from νmean, and most extreme deviation value. The tail represents the most extreme 4 % of the return values during a<br />

time interval ranging from January 1991 to December 2000 (k = 100 values) and from July 1985 to April 2008 (k = 229 values).<br />

νmean


30<br />

m=2507 bs=1000<br />

upper tail lower tail<br />

b γ mean<br />

b γ n k=100 rel err max dev 95% q 90% q std bs k=100 rel err max dev 95% q 90% q std bs<br />

Index 3.71 0.00 1.42 0.72 0.64 0.32 3.43 0.02 2.22 0.92 0.82 0.53<br />

BMY 3.66 0.00 1.50 0.72 0.52 0.31 1.93 0.07 2.91 1.21 0.81 0.60<br />

CVX 4.26 0.00 1.52 0.80 0.60 0.40 2.84 0.20 3.02 2.02 1.72 1.22<br />

HPQ 3.95 0.00 1.24 0.63 0.51 0.32 1.92 0.06 3.50 0.84 0.74 0.51<br />

KO 3.72 0.01 1.51 0.81 0.68 0.41 1.96 0.09 2.23 1.54 1.00 0.62<br />

MMM 4.12 0.01 1.52 0.80 0.70 0.43 2.45 0.11 2.00 1.23 1.11 0.74<br />

PG 3.59 0.00 1.44 0.81 0.61 0.42 1.67 0.06 2.61 0.71 0.64 0.42<br />

SGP 3.79 0.01 1.58 0.73 0.62 0.42 1.87 0.10 3.03 1.22 0.66 0.61<br />

TXN 4.00 0.00 1.54 0.79 0.66 0.43 1.86 0.05 2.12 0.93 0.63 0.40<br />

WAG 3.94 0.01 1.59 0.80 0.70 0.41 1.55 0.06 2.30 0.80 0.60 0.42<br />

m=5736 bs=650<br />

b γ n k=229 k=229<br />

Index 3.56 0.00 0.93 0.50 0.38 0.19 3.07 0.01 1.42 0.81 0.63 0.41<br />

BMY 3.60 0.00 1.10 0.42 0.42 0.18 2.07 0.02 1.40 0.60 0.52 0.31<br />

CVX 3.83 0.00 0.92 0.57 0.50 0.30 2.45 0.05 2.11 1.22 0.84 0.58<br />

HPQ 3.28 0.00 0.62 0.50 0.41 0.22 2.45 0.04 1.44 0.82 0.60 0.39<br />

KO 3.38 0.01 1.21 0.75 0.71 0.41 1.85 0.03 1.68 0.71 0.52 0.38<br />

MMM 3.69 0.00 0.84 0.42 0.40 0.21 2.10 0.03 1.39 0.73 0.64 0.39<br />

PG 3.58 0.01 0.83 0.50 0.39 0.22 1.74 0.03 1.40 0.58 0.40 0.37<br />

SGP 3.86 0.01 1.00 0.51 0.40 0.30 1.88 0.02 1.13 0.47 0.42 0.30<br />

TXN 3.34 0.00 0.62 0.42 0.40 0.21 2.13 0.04 1.90 0.70 0.51 0.32<br />

WAG 3.82 0.00 1.31 0.53 0.42 0.33 1.98 0.02 1.53 0.59 0.51 0.32<br />

Table 3.7: Establishing the uncertainty of estimated upper and lower tail indexes ˆ b γ n by creating bootstrap samples (bs) of historical return data<br />

for S&P 500 and included assets and calculation of bootstrap mean value νmean with rel. error from the original value, deviation quantiles and<br />

standard deviation from b γ mean, and most extreme deviation value. The tail represents the most extreme 4 % of the return values during a time<br />

interval ranging from January 1991 to December 2000 (k = 100 values, bs = 1000) and from July 1985 to April 2008 (k = 229 values, bs = 650).<br />

b γ mean


3.2.2 Implementation of the Non-Parametric Approach<br />

Within this subsection I provide an overwiev of the implementation of the non-parametric<br />

approach by Sornette & Malevergne. It is denoted non-parametric because of the estimation<br />

of the constant l by equation (3.36). We don’t give estimates of the scale<br />

factors of the tail distribution Cε and CY . Instead, we simply consider N sorted realizations<br />

of return vectors X and Y denoted by X1,N ≥ X2,N ≥ · · · ≥ XN,N and<br />

Y1,N ≥ Y2,N ≥ · · · ≥N,N and observe that the ratio of the empirical quantiles remains<br />

remarkably stable for small or large k.<br />

The coefficients of upper and lower tail dependence were estimated by the following<br />

equation:<br />

ˆλ +,− 1<br />

= �<br />

max 1, ˆl ˆβ<br />

where constant l was non-parametrically estimated by:<br />

ˆl = lim<br />

u→1− ,0 +<br />

F −1<br />

X (u)<br />

F −1<br />

Y (u)<br />

� α<br />

(3.53)<br />

Xk,N ∼ = , as k → N, 0 (3.54)<br />

As we can see already on equation (3.54) and equation (3.47) of Hill’s estimator,<br />

our non-parametric method is fully compatible with negative tails. Sorting return<br />

data in increasing order instead of decreasing order, as it is necessary to calculate tail<br />

dependence in negative tails, yields negative return data in both the index tail and the<br />

asset’s tail. Therefore the fraction in the definition of Hill’s estimator becomes positive<br />

and the natural logarithm gives out a real number. Constant ˆ l is accordingly always<br />

positive and if β < 0 for any asset and the index relation (3.53) cannot be applied<br />

since it assumes β > 0 and therefore λ = 0. To check whether constant ˆ li are sensitive<br />

to threshold k, I first plotted:<br />

ˆ l(k) = Xk,N/Yk,N for k = 1, 2, . . ., 150 for the smaller data set of all lower tales<br />

of the assets with the index. The 150 most extreme return values of the upper tails<br />

correspond to k/N = 6%.<br />

Figure (3.5) shows that ˆ l(k) becomes stable after about k/N = 0.5% and remains<br />

within narrow limits up to 6%.<br />

Using the average value of ˆ l(k), ˆ l(k) = 1<br />

k<br />

Yk,N<br />

�k Xj,N<br />

j=1 Yj,N<br />

on the one hands constitutes in<br />

a shift of emphasis to the extreme tail and on the other hand increases the robustness<br />

of constant ˆ l and ˆ λ to changes of threshold k, which also makes the necessity of finding<br />

k ∗ , the optimal threshold value, obsolete. Instead it becomes sufficient to determine<br />

an optimal interval, which in our case is: k = 3% . . .5%. Figure (3.6) shows on the<br />

example of negative tails, because there the effect is particularly strong, that in the<br />

area of the most extreme return values of an asset when k/N → 0, ˆ l(k) determined as<br />

average value over k becomes unstable. This could lead to distortion of our estimats.<br />

We can avoid this by calculation of ˆlc(k) = 1 �k Xj,N<br />

k−c+1 j=Y , where we chose an interval<br />

Yj,N<br />

ranging from c = [0.05·N] to k with [·] denoting integer numbers. This can be observed<br />

in figure (3.7). However ˆ l(k) seems to be stable on the interval we’re interested in.<br />

To summarize for upper tails: <strong>Tail</strong> indexes α were calculated by Hill’s estimator ˆν<br />

31


l(k)<br />

l(k) mean<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/N<br />

Figure 3.5: ˆ l(k) = Xk,N/Yk,N with k = 1... 150 (k/N = 0% ... 6%) for the lower tail of<br />

all assets with the index S&P 500 on the smaller data set ranging from January 1991 up to<br />

December 2000.<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/N<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

Figure 3.6: ˆl(k) = 1 �k Xj,N<br />

k j=1 with k = 1...150 (or k/N = 0% ...6%) for the lower tail of<br />

Yj,N<br />

all assets with the index S&P 500 on the smaller data set ranging from January 1991 up to<br />

December 2000.<br />

32


l(k) mean,c<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/N<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

Figure 3.7: ˆl(k)c = 1 �k Xj,N<br />

k−c+1 j=Y with k = 13... 150 (or k/N = 0.5% ... 6%), and c =<br />

Yj,N<br />

[0.05 · N] ([·] denotes integer numbers) applied to the lower tail of all assets with the index<br />

S&P 500 on the smaller data set ranging from January 1991 up to December 2000.<br />

and Gabaix’s ordinary least squares log-log rank-size tail index estimate ˆ b γ n<br />

ˆν =<br />

�<br />

1 �<br />

kj=1<br />

log<br />

k<br />

Yj,N<br />

Yk,N<br />

� −1<br />

given by:<br />

(3.55)<br />

log(Rank(Yj,N) − 1/2) = a − ˆ b γ n ∗ log(Yj,N), for j = 1 . . .k, (3.56)<br />

where Y1,N ≥ Y2,N ≥ · · · ≥ YN,N denote the ordered statistics of the sample containing<br />

N realizations of the market return vector Y and threshold number k assumed to<br />

represent the 4% quantile of Y .<br />

Constant ˆl was non-parametrically estimated by:<br />

k�<br />

ˆ<br />

1 Xj,N<br />

l(k) = (3.57)<br />

k<br />

ˆ lc(k) =<br />

j=1<br />

Yj,N<br />

1<br />

k − c + 1<br />

k�<br />

j=c<br />

Xj,N<br />

Yj,N<br />

(3.58)<br />

and ˆ β was calculated by ordinary least squares regression to the linear factor model:<br />

X = β · Y + ε with X denoting asset returns in descending chronologic order, with Y<br />

denoting corresponding index returns, and with ε denoting idiosyncratic noise.<br />

Putting all these parameters in equation (3.59) for upper tail dependence:<br />

ˆλ + 1<br />

= �<br />

max<br />

ˆλ + 1<br />

= �<br />

max<br />

33<br />

1, ˆ �ˆν l·<br />

ˆβ<br />

1, ˆ �ˆγ bn l·<br />

ˆβ<br />

(3.59)<br />

(3.60)


Non-Par Par<br />

up lo up lo up lo up lo<br />

ˆν 3.23 3.16 3.23 3.16 3.23 3.16 3.23 3.16<br />

tail 1... k 1...k Y ...k Y ... k 1... k 1... k Y ...k Y ...k<br />

ˆλ<br />

BMY 0.06 0.05 0.06 0.07 0.07 0.01 0.07 0.08<br />

CVX 0.02 0.02 0.02 0.02 0.02 0.00 0.02 0.02<br />

HPQ 0.10 0.07 0.10 0.09 0.13 0.02 0.14 0.12<br />

KO 0.06 0.05 0.06 0.07 0.07 0.01 0.08 0.08<br />

MMM 0.04 0.03 0.04 0.04 0.04 0.01 0.04 0.04<br />

PG 0.03 0.02 0.03 0.04 0.03 0.00 0.03 0.04<br />

SGP 0.04 0.04 0.04 0.05 0.05 0.01 0.05 0.06<br />

TXN 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00*<br />

WAG 0.09 0.06 0.09 0.11 0.11 0.01 0.10 0.12<br />

Table 3.8: Estimated values of upper and lower tail dependence for S&P 500 index with<br />

a set of nine major assets traded on the New York stock Exchange calculated by a nonparametric<br />

approach: ˆ λ +,− �<br />

= 1/max 1, l<br />

�ν β on the left and by a parametric approach:<br />

ˆλ +,− �<br />

= 1/ 1 + β−ν · Cε<br />

�<br />

on the right. The tail represents the most extreme 4% of the<br />

CY<br />

return values during a time interval from January 1991 to December 2000. ’tail’ shows the<br />

calculation interval of l with c = 13 and k = 100 and ∗ denotes negative ˆ β.<br />

and applying the approach to our reference data sets yields the estimates reported in<br />

tables (3.8) and (3.10) on the left side using Hill’s estimator ν and reported in tables<br />

(3.9) and (3.11) on the left side using Gabaix’s estimator b γ n 5 .<br />

To achieve estimates for the lower tail, we can just sort our data in ascending order<br />

and, for the calculation of Gabaix’s tail index estimator, multiply the ordered sample<br />

by (-1). ˆ β remains the same for negative tails, as it is calculated for the whole data<br />

sample within this approach.<br />

3.2.3 Analysis of TDC estimated by the Non-Parametric Approach<br />

Within this subsection I provide a sensitivity analysis of TDC estimated by the nonparametric<br />

approach of Sornette & Malevergne performed on the two reference data<br />

sets: the robustness of the results to the choice of threshold number k is first analyzed<br />

visually by plotting ˆ λ in dependence of k. In a second step estimation of error bars by<br />

bootstrap sampling with replacement of ˆ λ is provided. The aim is to find confidence<br />

intervals for the estimates and to get an insight into the underlying error distributions.<br />

Then as a third step variations of ˆ λ in time are studied by rolling time windows. Different<br />

sample sizes are shifted along the time axis in order to measure time dependence<br />

of ˆ λ.<br />

Figure (3.8) shows ˆ λ(k) for k<br />

k<br />

= 0% . . .6%, figure (3.9) λ(k) for = 0% . . .6%,<br />

N N<br />

and (3.10) shows λc(k) for k<br />

= 0.5% . . .6% for the lower tail of the smaller data<br />

N<br />

sample. Results for the positive tail look more stable. But also for negative tails the<br />

charts look consistent and stable. Looking at the interval of k = 3% . . .5% that is of<br />

N<br />

primary interest to us, the curve of λ is not stiff to moderate changes in k. Using the<br />

5 The matlab m-file for the implementation of λ +,− estimated by the non-parametric approach according to<br />

Sornette & Malevergne is denoted by: lambda np sor ma.m and enclosed to the appendix and the data CD<br />

34


Non-Par Par<br />

up lo up lo up lo up lo<br />

ˆ b γ n 3.71 3.37 3.71 3.37 3.71 3.37 3.71 3.37<br />

tail 1... k 1...k Y ...k Y ... k 1... k 1... k Y ...k Y ...k<br />

ˆλ<br />

BMY 0.04 0.04 0.04 0.05 0.05 0.00 0.05 0.07<br />

CVX 0.01 0.01 0.01 0.02 0.01 0.00 0.01 0.02<br />

HPQ 0.07 0.06 0.07 0.08 0.10 0.01 0.11 0.11<br />

KO 0.04 0.04 0.04 0.06 0.05 0.00 0.05 0.07<br />

MMM 0.02 0.02 0.02 0.03 0.03 0.00 0.03 0.03<br />

PG 0.02 0.02 0.02 0.03 0.02 0.00 0.02 0.03<br />

SGP 0.03 0.03 0.03 0.04 0.03 0.00 0.03 0.05<br />

TXN 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00*<br />

WAG 0.06 0.05 0.06 0.10 0.08 0.00 0.08 0.11<br />

Table 3.9: Estimated values of upper and lower tail dependence for S&P 500 index with<br />

a set of nine major assets traded on the New York stock Exchange calculated by a nonparametric<br />

approach: ˆ λ +,− �<br />

= 1/max 1, l<br />

� γ<br />

bn β on the left and by a parametric approach:<br />

ˆλ +,− �<br />

= 1/ 1 + β−bγn · Cε<br />

�<br />

on the right. The tail represents the most extreme 4% of the<br />

CY<br />

return values during a time interval from January 1991 to December 2000. ’tail’ shows the<br />

calculation interval of l with c = 13 and k = 100 and ∗ denotes negative ˆ β.<br />

Non-Par Par<br />

up lo up lo up lo up lo<br />

ˆν 3.09 3.13 3.09 3.13 3.09 3.13 3.09 3.13<br />

tail 1... k 1...k Y ...k Y ... k 1... k 1... k Y ...k Y ...k<br />

ˆλ<br />

BMY 0.05 0.04 0.05 0.05 0.06 0.02 0.06 0.06<br />

CVX 0.04 0.04 0.04 0.04 0.05 0.02 0.05 0.05<br />

HPQ 0.09 0.08 0.09 0.09 0.12 0.07 0.13 0.13<br />

KO 0.10 0.08 0.10 0.11 0.13 0.02 0.13 0.13<br />

MMM 0.06 0.05 0.06 0.06 0.07 0.02 0.07 0.08<br />

PG 0.02 0.02 0.02 0.03 0.03 0.00 0.03 0.03<br />

SGP 0.03 0.02 0.03 0.03 0.03 0.01 0.03 0.03<br />

TXN 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

WAG 0.07 0.06 0.07 0.08 0.09 0.02 0.09 0.09<br />

Table 3.10: Estimated values of upper and lower tail dependence for S&P 500 index with<br />

a set of nine major assets traded on the New York stock Exchange calculated by a nonparametric<br />

approach: ˆ λ +,− �<br />

= 1/max 1, l<br />

�ν β on the left and by a parametric approach:<br />

ˆλ +,− �<br />

= 1/ 1 + β−ν · Cε<br />

�<br />

on the right. The tail represents the most extreme 4% of the<br />

CY<br />

return values during a time interval from July 1985 to April 2008. ’tail’ shows the calculation<br />

interval of l with c = 29 and k = 229.<br />

35


Non-Par Par<br />

up lo up lo up lo up lo<br />

ˆ b γ n 3.56 3.04 3.56 3.04 3.56 3.04 3.56 3.04<br />

tail 1... k 1...k Y ...k Y ... k 1... k 1... k Y ...k Y ...k<br />

ˆλ<br />

BMY 0.03 0.05 0.03 0.05 0.04 0.03 0.04 0.07<br />

CVX 0.03 0.04 0.03 0.04 0.03 0.02 0.03 0.05<br />

HPQ 0.06 0.09 0.06 0.10 0.09 0.08 0.10 0.14<br />

KO 0.07 0.09 0.07 0.12 0.10 0.02 0.10 0.14<br />

MMM 0.04 0.05 0.04 0.06 0.05 0.02 0.05 0.08<br />

PG 0.01 0.02 0.01 0.03 0.02 0.01 0.02 0.04<br />

SGP 0.02 0.02 0.02 0.03 0.02 0.01 0.02 0.04<br />

TXN 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

WAG 0.05 0.07 0.05 0.08 0.06 0.03 0.06 0.10<br />

Table 3.11: Estimated values of upper and lower tail dependence for S&P 500 index with<br />

a set of nine major assets traded on the New York stock Exchange calculated by a nonparametric<br />

approach: ˆ λ +,− �<br />

= 1/max 1, l<br />

� γ<br />

bn β on the left and by a parametric approach:<br />

ˆλ +,− �<br />

= 1/ 1 + ˆ β−bγn · Cε<br />

�<br />

on the right. The tail represents the most extreme 4% of the<br />

CY<br />

return values during a time interval from July 1985 to April 2008. ’tail’ shows the calculation<br />

interval of l with c = 29 and k = 229.<br />

uncorrected mean value λ(k), estimates tend to be slightly lower than by choosing the<br />

corrected estimate λc(k). Nevertheless, I think that neglecting most extreme values as<br />

we did by corrected values does not have to yield better estimates because they might<br />

have high impanct on distributions as it is most notable for negative tails. This can be<br />

observed at the example of figure (3.10). Therefore I will give both estimators λ(k) and<br />

λc(k) in the result charts. Significant differences between these two estimators generally<br />

signalise strong impact of outliers, the values that are to a high degree uncommon.<br />

Estimation of Error Bars<br />

For the estimation of error bars or in other words: the level of uncertainty specified<br />

by a particular estimation method, I performed bootstrap sampling with replacement.<br />

The bootstrap method is a procedure that involves choosing random samples with<br />

replacement from a data set and analyzing each sample in the same way. Sampling<br />

with replacement means that every sample is returned to the data set after sampling or<br />

after being chosen. So a particular data point from the original data set could appear<br />

multiple times in a particular bootstrap sample. The number of elements in each<br />

bootstrap sample equals the number of elements in the original data set. The range<br />

of sample estimates obtained enables you to establish the uncertainty of the quantity<br />

you’re estimating.<br />

Considering the linear factor model: X = β·Y +ε, I created bootstrap samples (with<br />

replacement) of the index data table (i.e. S&P 500) and then picked out the respective<br />

data points of the assets (according in time to S&P 500 index data). Like that I only<br />

performed bootstrap sampling of one vector (S&P 500 index data) and searched for<br />

the corresponding asset data by algorithm. ’bootstrp’ is a predefined function handle<br />

in Matlab, which allows one to perform sampling by replacement. ˆ β, ˆν, and ˆ l as the<br />

36


λ(k)<br />

λ(k) mean<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/N<br />

Figure 3.8: ˆ λ(k) =<br />

1<br />

�<br />

max 1, l<br />

ˆβ<br />

� ˆν with k = 1...343 (or k/N = 0% ... 6%), applied to the lower<br />

tail of all assets with the index S&P 500 on the bigger data set ranging from July 1985 to<br />

April 2008.<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/N<br />

Figure 3.9: λ(k) =<br />

1<br />

�<br />

max 1, ˆl ˆβ<br />

� ˆν with k = 1... 343 (or k/N = 0% ... 6%), applied to the lower<br />

tail of all assets with the index S&P 500 on the bigger data set ranging from July 1985 to<br />

April 2008.<br />

37


λ(k) mean,c<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/N<br />

Figure 3.10: λ(k) =<br />

1<br />

�<br />

max 1, ˆlĉ β<br />

� ˆν with k = 29... 343 (or k/N = 0.5% ... 6%) and c = [0.005·N]<br />

([·] denotes integer numbers), applied to the lower tail of all assets with the index S&P 500<br />

on the bigger data set ranging from July 1985 to April 2008.<br />

mean value of lj = Xj,N<br />

Yj,N<br />

for j = 1 . . .k and j = c . . .k with c = 0.005 · N then were<br />

calculated independently for each sample. Table (3.12) shows results of bootstrap<br />

sampling with replacement applied to the smaller data set and (3.13) shows results of<br />

bootstrap sampling with replacement applied to the bigger data sample 6 . The mean<br />

values of the tail dependence coefficients calculated for the 1000 bootstrap samples were<br />

close to the ˆ λ of the original sample.<br />

I calculated relative deviations between average of estimated bootstrap sample tail<br />

dependence coefficients and original values (λ of original data sample) of maximally<br />

7%, which can be observed in columns denoted by ’rel err’. The quantiles describe<br />

the width of the error bar or in other words the extent of uncertainty: a 95% quantile<br />

for example gives the absolute deviation of 95% of the bootstrap tail dependence<br />

estimates from their mean value. Taking a look at estimated error bars we determine<br />

that standard deviations of the difference between tail dependence coefficient estimates<br />

calculated for the bs = 1000 bootstrap samples most of times are approximately half<br />

of the respective 95% quantiles. This is an indicator for the error distributions to be<br />

asymptitically Gaussian. Furthermore with respect to relative sizes of the quantiles<br />

compared to absolute values of estimators we can detect a certain pattern. Looking at<br />

the smaller sample bootstrap estimates, provided in table (3.12), we observe that most<br />

of the 95% quantiles calculated in the positive tails tend towards 70% to 80% of the<br />

corresponding total values of tail dependence coefficient estimates and for negative tails<br />

95% quantiles are about the size of the total value of corresponding tail dependence<br />

6 The matlab m-file for the bootstrap sampling with replacement of λ +,− estimated by the non-parametric<br />

approach according to Sornette & Malevergne is denoted by: bootstr nonpar alld.m and enclosed to the data<br />

CD<br />

38


coefficient estimates. For the bigger data sample quantiles are significantly smaller.<br />

95% quantiles here are mostly about 50% to 60% of corresponding total value of tail<br />

dependence coefficient estimates in the positive tails and slightly higher in the negative<br />

tails. This shows us that either the only way to improve our λ estimates by the given<br />

approach is to aggregate more data, or that error bars simply depend on the chosen<br />

time window and evolve according an ARCH process (autoregressive conditional heteroscedasticity)<br />

considering the variance of the current error term to be a function of<br />

the variances of the previous time period’s error terms.<br />

39


40<br />

m=2507 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=100 rel err max dev 95% q 90% q std bs k=100 rel err max dev 95% q 90% q std bs<br />

BMY 0.06 0.01 0.10 0.04 0.03 0.02 0.04 0.07 0.09 0.04 0.03 0.02<br />

CVX 0.02 0.00 0.04 0.02 0.01 0.01 0.02 0.01 0.05 0.02 0.02 0.01<br />

HPQ 0.10 0.00 0.20 0.07 0.06 0.04 0.07 0.05 0.10 0.06 0.05 0.03<br />

KO 0.06 0.01 0.10 0.04 0.04 0.02 0.05 0.02 0.20 0.06 0.05 0.03<br />

MMM 0.04 0.02 0.10 0.03 0.03 0.02 0.03 0.03 0.07 0.03 0.02 0.02<br />

PG 0.03 0.03 0.10 0.03 0.02 0.01 0.02 0.03 0.08 0.03 0.02 0.02<br />

SGP 0.04 0.01 0.10 0.03 0.02 0.02 0.03 0.05 0.08 0.03 0.03 0.02<br />

TXN 0.00* Inf 0.00 0.00 0.00 0.00 0.00* Inf 0.00 0.00 0.00 0.00<br />

WAG 0.09 0.02 0.20 0.08 0.06 0.04 0.06 0.03 0.20 0.06 0.05 0.03<br />

λc<br />

BMY 0.06 0.01 0.08 0.04 0.03 0.02 0.06 0.04 0.10 0.04 0.04 0.02<br />

CVX 0.02 0.00 0.04 0.02 0.01 0.01 0.02 0.03 0.04 0.02 0.02 0.01<br />

HPQ 0.10 0.00 0.20 0.07 0.06 0.04 0.09 0.03 0.20 0.07 0.05 0.04<br />

KO 0.06 0.01 0.08 0.04 0.04 0.02 0.07 0.03 0.10 0.06 0.05 0.03<br />

MMM 0.04 0.02 0.09 0.03 0.03 0.02 0.04 0.02 0.08 0.03 0.03 0.02<br />

PG 0.03 0.04 0.07 0.03 0.02 0.01 0.04 0.00 0.10 0.04 0.03 0.02<br />

SGP 0.04 0.01 0.08 0.03 0.02 0.02 0.05 0.03 0.10 0.04 0.03 0.02<br />

TXN 0.00* Inf 0.00 0.00 0.00 0.00 0.00* Inf 0.00 0.00 0.00 0.00<br />

WAG 0.09 0.02 0.20 0.08 0.06 0.04 0.11 0.03 0.20 0.08 0.07 0.04<br />

Table 3.12: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time interval from<br />

January 1991 to December 2000. ˆ βj have been calculated on the whole samples. ∗ denotes negative ˆ β and therefore ˆ λ = 0 and ’Inf’ denotes ·/0.


41<br />

m=5736 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=229 rel err max dev 95% q 90% q std bs k=229 rel err max dev 95% q 90% q std bs<br />

BMY 0.05 0.01 0.06 0.03 0.02 0.01 0.04 0.02 0.08 0.03 0.02 0.01<br />

CVX 0.04 0.01 0.05 0.02 0.02 0.01 0.04 0.05 0.08 0.03 0.03 0.02<br />

HPQ 0.09 0.00 0.07 0.04 0.03 0.02 0.09 0.02 0.12 0.04 0.04 0.02<br />

KO 0.10 0.02 0.10 0.05 0.04 0.03 0.09 0.05 0.15 0.07 0.06 0.04<br />

MMM 0.07 0.03 0.07 0.04 0.03 0.02 0.05 0.06 0.08 0.03 0.03 0.02<br />

PG 0.03 0.20 0.08 0.03 0.02 0.02 0.02 0.10 0.06 0.02 0.02 0.01<br />

SGP 0.03 0.10 0.05 0.03 0.02 0.02 0.02 0.10 0.06 0.02 0.02 0.01<br />

TXN 0.00 0.20 0.00 0.00 0.00 0.00 0.00 0.40 0.01 0.00 0.00 0.00<br />

WAG 0.07 0.00 0.07 0.04 0.03 0.02 0.06 0.03 0.08 0.03 0.03 0.02<br />

λc<br />

BMY 0.05 0.01 0.06 0.03 0.02 0.01 0.05 0.03 0.07 0.03 0.02 0.01<br />

CVX 0.04 0.01 0.05 0.02 0.02 0.01 0.04 0.05 0.07 0.03 0.02 0.02<br />

HPQ 0.09 0.00 0.07 0.04 0.03 0.02 0.10 0.03 0.11 0.04 0.03 0.02<br />

KO 0.10 0.01 0.12 0.05 0.04 0.03 0.11 0.02 0.13 0.07 0.06 0.04<br />

MMM 0.07 0.03 0.07 0.04 0.03 0.02 0.06 0.06 0.07 0.03 0.03 0.02<br />

PG 0.03 0.20 0.08 0.03 0.02 0.01 0.03 0.20 0.07 0.03 0.03 0.02<br />

SGP 0.03 0.10 0.05 0.03 0.02 0.01 0.03 0.10 0.08 0.03 0.02 0.01<br />

TXN 0.00 0.20 0.00 0.00 0.00 0.00 0.00 0.30 0.01 0.00 0.00 0.00<br />

WAG 0.07 0.00 0.07 0.04 0.03 0.02 0.08 0.03 0.08 0.04 0.03 0.02<br />

Table 3.13: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values and standard<br />

deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time interval from<br />

July 1985 to April 2008. ˆ βj have been calculated on the whole samples.


Observation of λ by Rolling Time Window<br />

As mentioned, it is the aim of the determination of tail dependence in terms of portfolio<br />

analysis to find assets with low tail dependence coefficients in order to diversify<br />

away extreme risks. Looking at tables (3.8), (3.10), (3.9), and (3.11) asset Texas<br />

Instruments Inc. is maybe the most interesting case. Depending on the sample size, ˆ β<br />

becomes either zero, which yields zero tail dependence by default criteria or diminutive,<br />

which yields vanishing tail dependence. The sample size reflects the time window of the<br />

estimate. It would be interesting to determine, how tail dependence estimates evolve<br />

with time. The hypothesis is that λ is not fixed but perhaps is evolving according to an<br />

ARCH process. Therefore I used a rolling time window to observe the time dependence<br />

of the estimate λ. The size of the time window should on the one hand be sufficiently<br />

large to allow for good precision according to the bootstrap procedure and should on<br />

the other hand be sufficiently small to reflect the possible change of values as a function<br />

of time. This had to be tested on different window sizes while I had to keep an<br />

eye on the individual parameters tail index α, β, and l to control whether the window<br />

size or generally the size of the sample allows for appropriate determination. I first<br />

plotted ˆ β, ˆ l, ˆν in dependence of N for three window sizes: S = 2500, S = 1600, and<br />

S = 800 shown in figures (3.11),. . .,(3.19). X-axis N has the dimension of observation<br />

days (daily observations) starting from N = 1 representing July 1 st 1985 and ending<br />

by N = 5736 representing Mars 31 st 2008<br />

First of all it is important to take notice of the different plot intervals in N given<br />

the different window sizes. When I chose S = 2500, I have to wait for the first 2500<br />

daily observations to take place, which literally corresponds to a time interval of 2500<br />

days from the first observation. Choosing S = 800 and a step size of one between<br />

successive measurements, indicating that the time window moves one data point to<br />

the right hand side per step (or chronologically one day), means that I already have<br />

2500 − 800 = 1700 measurements at observation day N = 2500, which is the starting<br />

point of measurements by choosing the biggest window size. That’s why at first sight<br />

the plots look kind of different.<br />

Looking at figures (3.11), (3.12), and (3.13) for the three window sizes S, we observe<br />

that the behavior of ˆ β looks consistent over time. The plots for different window sizes<br />

only differ in detailedness of the mapping of changes. The smaller we go, the bigger the<br />

peaks become. The same counts for constant ˆ l, which is shown in figures (3.14), (3.15),<br />

and (3.16) for the different window sizes. Sample peaks in ˆ l sometimes are smoothed<br />

out extensively when we use bigger window sizes S. This makes ˆ l look very stable over<br />

time when we plot it by choosing window size S = 2500. As furthermore observable for<br />

most of the assets, ˆ β, after having reached a peak, seems to have a declining tendency<br />

over time. The only asset that is completely incompatible with the declining trend<br />

of ˆ β is Texas Instruments Inc.(TXN). It reflects on the others by showing a countertrend.<br />

Asset TXN indicates a overall counter-movement with the index S&P 500 in the<br />

early phases indicated by negative ˆ β and recently a co-movement with the index, as ˆ β<br />

becomes positive or even close to one. The trend shown by the other assets would be<br />

the opposite of this: decreasing overall co movement of assets and index with increasing<br />

time.<br />

Looking at figures (3.17), (3.18), and (3.19) of ˆν for the three window sizes, estimates<br />

oscillate strongly but withing stable intervals depending on the window size. We<br />

know that Hill’s estimator performs better with bigger data sets. Therefore it is suggestive<br />

to choose a window size, such that upper and lower bounds of these oscillations<br />

remain small enough. This is the case for samples with S >1000.<br />

42


β<br />

β<br />

β<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

BMY<br />

0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

KO<br />

0.4<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

SGP<br />

0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

β<br />

β<br />

β<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

CVX<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

MMM<br />

0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

TXN<br />

−0.2<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

β<br />

β<br />

β<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

HPQ<br />

1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

PG<br />

0.2<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

WAG<br />

0.4<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.11: β(N) estimated for 9 assets X and index S&P 500 Y using the linear single<br />

factor model: X = β · Y + ε for rolling time horizon windows of S = 2500 considered data<br />

points from N = (1... S),(2... S +1),... ,(5736−S +1... 5736) or a total time interval from<br />

July 1985 to Mars 2008.<br />

β<br />

β<br />

β<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

BMY<br />

0.4<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

KO<br />

SGP<br />

0.2<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

β<br />

β<br />

β<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

CVX<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

MMM<br />

0.4<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

1.5<br />

1<br />

0.5<br />

0<br />

TXN<br />

−0.5<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

β<br />

β<br />

β<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

HPQ<br />

0.8<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

PG<br />

0.2<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

WAG<br />

0.4<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

Figure 3.12: β(N) estimated for 9 assets X and index S&P 500 Y using the linear single<br />

factor model: X = β · Y + ε for rolling time horizon windows of S = 1600 considered data<br />

points from N = (1... S),(2... S +1),... ,(5736−S +1... 5736) or a total time interval from<br />

July 1985 to Mars 2008.<br />

43


β<br />

β<br />

β<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

BMY<br />

0.4<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

KO<br />

0.2<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

1.5<br />

1<br />

0.5<br />

SGP<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

β<br />

β<br />

β<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

CVX<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

MMM<br />

0.4<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

1.5<br />

1<br />

0.5<br />

0<br />

TXN<br />

−0.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

β<br />

β<br />

β<br />

2<br />

1.5<br />

1<br />

HPQ<br />

0.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

PG<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

1.5<br />

1<br />

0.5<br />

WAG<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

Figure 3.13: β(N) estimated for 9 assets X and index S&P 500 Y using the linear single<br />

factor model: X = β · Y + ε for rolling time horizon windows of S = 800 considered data<br />

points from N = (1... S),(2... S +1),... ,(5736−S +1... 5736) or a total time interval from<br />

July 1985 to Mars 2008.<br />

l<br />

l<br />

l<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

BMY<br />

1.6<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.9<br />

1.8<br />

1.7<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

KO<br />

SGP<br />

1.8<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

l<br />

l<br />

l<br />

1.9<br />

1.8<br />

1.7<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

CVX<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.8<br />

1.7<br />

1.6<br />

1.5<br />

1.4<br />

MMM<br />

1.3<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

4<br />

3.5<br />

3<br />

TXN<br />

2.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

l<br />

l<br />

l<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

HPQ<br />

2<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2.2<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

PG<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

WAG<br />

1.6<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.14: Constant lU(N) for upper tails plotted in green and constant lL(N) for<br />

lower tails plotted in black estimated for 9 assets and index S&P 500 using relation:<br />

ˆ 1 �k l(k) = k j=1 Xj,S/Yj,S with k = 0.04 · S, Xj,S and Yj,S rank ordered return observations,<br />

and rolling time horizon window of S = 2500 considered data points from N =<br />

(1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to<br />

Mars 2008.<br />

44


l<br />

l<br />

l<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

BMY<br />

1.6<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

3<br />

2.5<br />

2<br />

1.5<br />

KO<br />

1<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

3<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

SGP<br />

1.8<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

l<br />

l<br />

l<br />

2.5<br />

2<br />

1.5<br />

CVX<br />

1<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

2.5<br />

2<br />

1.5<br />

MMM<br />

1<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

TXN<br />

2<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

l<br />

l<br />

l<br />

3.5<br />

3<br />

2.5<br />

2<br />

HPQ<br />

1.5<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

2.5<br />

2<br />

1.5<br />

1<br />

PG<br />

0.5<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

3<br />

2.5<br />

2<br />

1.5<br />

WAG<br />

1<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

Figure 3.15: Constant lU(N) for upper tails plotted in green and constant lL(N) for lower tails<br />

plotted in black estimated for 9 assets and index S&P 500 using: ˆl(k) = 1 �k k j=1 Xj,S/Yj,S with<br />

k = 0.04 ·S, Xj,S and Yj,S rank ordered return observations, and rolling time horizon window<br />

of S = 1600 considered data points from N = (1... S),(2... S +1),... ,(5736−S +1... 5736)<br />

or a total time interval from July 1985 to Mars 2008.<br />

l<br />

l<br />

l<br />

3<br />

2.5<br />

2<br />

1.5<br />

BMY<br />

1<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

KO<br />

SGP<br />

1.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

l<br />

l<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

CVX<br />

0.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

MMM<br />

0.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

l<br />

6<br />

5<br />

4<br />

3<br />

2<br />

TXN<br />

1<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

l<br />

l<br />

l<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

HPQ<br />

1.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

PG<br />

0.5<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

WAG<br />

1<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

Figure 3.16: Constant lU(N) for upper tails plotted in green and constant lL(N) for lower tails<br />

plotted in black estimated for 9 assets and index S&P 500 using: ˆl(k) = 1 �k k j=1 Xj,S/Yj,S with<br />

k = 0.04 ·S, Xj,S and Yj,S rank ordered return observations, and rolling time horizon window<br />

of S = 800 considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1...5736)<br />

or a total time interval from July 1985 to Mars 2008.<br />

45


ν S&P 500<br />

4<br />

3.5<br />

3<br />

2.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.17: νk,U(N) for upper tail of index S&P 500 plotted in red and νk,L(N) for lower tail<br />

�<br />

plotted in black using Hill’s estimator given by equation ˆν = kj=1 log Yj,S<br />

�−1 with k =<br />

0.04 · S and Xj,S and Yj,S rank ordered return observations for rolling time horizon windows<br />

of S = 2500 considered data points from N = (1... S),(2... S +1),... ,(5736−S +1... 5736)<br />

or a total time interval from July 1985 to Mars 2008.<br />

ν S&P 500<br />

5<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1500 2000 2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.18: νk,U(N) for upper tail of index S&P 500 plotted in red and νk,L(N) for lower tail<br />

�<br />

plotted in black using Hill’s estimator given by equation ˆν = kj=1 log Yj,S<br />

�−1 with k =<br />

0.04 · S and Xj,S and Yj,S rank ordered return observations for rolling time horizon windows<br />

of S = 1600 considered data points from N = (1... S),(2... S +1),... ,(5736−S +1... 5736)<br />

or a total time interval from July 1985 to Mars 2008.<br />

46<br />

�<br />

1<br />

k<br />

�<br />

1<br />

k<br />

Yk,N<br />

Yk,N<br />

ν U<br />

ν L<br />

ν U<br />

ν L


ν S&P 500<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

Figure 3.19: νk,U(N) for upper tail of index S&P 500 plotted in red and νk,L(N) for lower tail<br />

�<br />

plotted in black using Hill’s estimator given by equation ˆν = kj=1 log Yj,S<br />

�−1 with k =<br />

0.04 · S and Xj,S and Yj,S rank ordered return observations for rolling time horizon windows<br />

of S = 800 considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1...5736)<br />

or a total time interval from July 1985 to Mars 2008.<br />

Figures (3.20), (3.21), and (3.22) show plots of ˆ λU and ˆ λL for the different assets<br />

and for the three different window sizes. What most of these plots have in common is<br />

that upper tail dependence estimates ˆ λU show an increasingly constant behavior over<br />

time and lower tail dependence estimates ˆ λL overall show a decreasing tendency, which<br />

is more distinctive for window sizes with S >1000. Figures (3.23) and (3.24) show tail<br />

dependence estimated for rolling windows of the three different sizes altogether in the<br />

same figure and for the same range. It can be observed that tail dependence estimates<br />

for the different window sizes sometimes seem inconsistent with each other. It is not<br />

only because the plots with wider window sizes look smoothed, but also that some<br />

peaks, which are distinctive in red (S = 800) don’t even appear in blue (S = 1600) or<br />

black (S = 2500). This counts first of all for more recent estimates. The three assets<br />

Chevron Corp.(CVX), 3M Co.(MMM), and Texas Instruments Inc.(TXN) show a peak<br />

in the upper tail dependence at the end of the measurement period N >5000. The only<br />

asset that is very much contradictory with overall trends is again TXN, what makes it<br />

particularly interesting.<br />

Although it is difficult to draw conclusions because of the inconsistency of the estimates<br />

obtained by the different window sizes, something is clearly apparent in all of<br />

the plots: Whereas lower tail dependence ˆ λL in the beginning used to dominate over<br />

upper tail dependence ˆ λU, by some point typically around N = 4000 ˆ λU becomes bigger<br />

than ˆ λL. This is very interesting because it would mean that the tendency of extreme<br />

positive co-movement between the assets and index S&P 500 has become stronger than<br />

the tendency of extreme negative co-movements or in other words: extraordinary high<br />

gains of assets and index S&P 500 are more likely to happen simultaneously than it is<br />

the case that high losses of assets and index S&P 500 occur at the same time.<br />

47<br />

�<br />

1<br />

k<br />

Yk,N<br />

ν U<br />

ν L


48<br />

λ<br />

λ<br />

λ<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

KO<br />

SGP<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

CVX<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

MMM<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

x 10−3<br />

8<br />

6<br />

4<br />

2<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

HPQ<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

PG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.20: λU(N) for the upper tails of index S&P 500 and the nine assets plotted in blue and λL(N) for the lower tails plotted in red using<br />

non-parametric approach given by equation ˆ λ +,− �<br />

= 1/max 1, l<br />

�ˆν β with coefficients ˆν, l and β for rolling time horizon windows of S = 2500<br />

considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to Mars 2008.


49<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

BMY<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

KO<br />

SGP<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

CVX<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

MMM<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

TXN<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

HPQ<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

PG<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

WAG<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

N<br />

Figure 3.21: λU(N) for the upper tails of index S&P 500 and the nine assets plotted in blue and λL(N) for the lower tails plotted in red using<br />

non-parametric approach given by equation ˆ λ +,− �<br />

= 1/max 1, l<br />

�ˆν β with coefficients ˆν, l and β for rolling time horizon windows of S = 1600<br />

considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to Mars 2008.


50<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

BMY<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

KO<br />

SGP<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

λ<br />

λ<br />

λ<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

CVX<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

MMM<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

TXN<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

λ<br />

λ<br />

λ<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

HPQ<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

PG<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

WAG<br />

0<br />

0 1000 2000 3000<br />

N<br />

4000 5000 6000<br />

Figure 3.22: λU(N) for the upper tails of index S&P 500 and the nine assets plotted in blue and λL(N) for the lower tails plotted in red using<br />

non-parametric approach given by equation ˆ λ +,− �<br />

= 1/max 1, l<br />

�ˆν β with coefficients ˆν, l and β for rolling time horizon windows of S = 800<br />

considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to Mars 2008.


51<br />

λ<br />

λ<br />

λ<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

KO<br />

SGP<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

CVX<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

MMM<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

HPQ<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

PG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.23: λU(N) for the upper tails of index S&P 500 and the nine assets for rolling time horizon windows of size S = 2500 plotted in black,<br />

S = 1600 plotted in blue, and S = 800 plotted in red using non-parametric approach given by equation ˆ λ +,− = 1/max<br />

ˆν, l and β from N = (2501 − S ... S),(2502... S + 1),... ,(5736 − S + 1... 5736).<br />

�<br />

1, l<br />

β<br />

� ˆν<br />

with coefficients


52<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

KO<br />

SGP<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

CVX<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

MMM<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

HPQ<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

PG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.24: λL(N) for the lower tails of index S&P 500 and the nine assets for rolling time horizon windows of size S = 2500 plotted in black,<br />

S = 1600 plotted in blue, and S = 800 plotted in red using non-parametric approach given by equation ˆ λ +,− = 1/max<br />

ˆν, l and β from N = (2501 − S ... S),(2502... S + 1),... ,(5736 − S + 1... 5736).<br />

�<br />

1, l<br />

β<br />

� ˆν<br />

with coefficients


Error Bars in <strong>Dependence</strong> of Time<br />

To establish uncertainty of estimated upper and lower tail dependence I performed<br />

bootstrap sampling over the whole area: N = (1 . . .S), (2 . . .(S + 1)), . . .,(5736 − S +<br />

1 . . .5736) for the nine assets with index S&P 500 and all of the 3 window sizes 7 .<br />

Time order of the results is given top down. That means that the latest estimate is<br />

on top representing Mars 31 st , 2008. For the biggest window size of S = 2500 daily<br />

observations I calculated bs = 800 bootstrap sample per window and for each of the<br />

two smaller ones S = 1600 and S = 800 I calculated bs = 1000 bootstrap samples.<br />

As step size between the successive windows I choose five days. That means that the<br />

window moves five daily observations to the right per step considering the time axis<br />

(or five days into the future). This provides the window with five new observations per<br />

step, meanwhile it ignores the five oldest ones. As it can be observed in the Excel files,<br />

this makes [(5736 −2500)/5] = 647 windows for the biggest window size S = 2500 data<br />

points, 827 windows for samples of window size S = 1600, and 987 windows for the<br />

smallest window size S = 800.<br />

In terms of quantile estimates on the bootstrap samples in order to estimate error<br />

bars, bootstrap estimates of the biggest window size S = 2500 should theoretically yield<br />

similar results as obtained by bootstrap sampling of our smaller reference data set that<br />

ranges from January 1991 to December 2000 and consists of 2506 daily observations.<br />

In fact quantiles and standard deviations of the bootstrap estimates are in very good<br />

accordance with these further results given in table (3.12), whereas for upper tails on<br />

average 95% quantiles between 70% and 80% of the corresponding total values of tail<br />

dependence coefficient estimates were calculated and for lower tails 95% quantiles were<br />

about the size of the total values (100% of corresponding tail dependence coefficient<br />

estimates). Choosing a window size of S = 2500 yields average 95% quantiles of 70%<br />

up to 85% of the tail dependence estimates bs L np k up m (notation explained below)<br />

for positive tails and around 80% to 150% of corresponding bs L np k lo m for negative<br />

tails. Choosing corrected ˆ l, quantiles calculated for negative tails become lower.<br />

Here again the assumption of error distributions to be approximately Gaussian seems to<br />

hold well, as calculated standard deviations within and between the different bootstrap<br />

samples are very close of being equal to half of 95% quantiles. This counts for all<br />

window sizes. Looking at the smallest window size of S = 800 daily observations,<br />

95% quantiles on average tend to be 150% to 200% of corresponding tail dependence<br />

coefficient estimates for positive tails and up to 250% for negative tails. This means<br />

that standard deviations are around the size of the estimates, which is after my opinion<br />

too high. In between we have window size S = 1600 with average 95% quantiles of<br />

100% to 140% of corresponding tail dependence estimates for positive tails and 120%<br />

to 180% of corresponding tail dependence estimates for negative tails. This reveals<br />

that sample sizes shouldn’t be smaller than S = 1600 to achieve at least more or less<br />

accurate estimates.<br />

In the following list explanatory notes on the results of bootstrap sampling for the rolling<br />

window tail dependence estimations attached in Excel files provided at the example of<br />

non-parametric tail dependence estimates for upper tails and uncorrected ˆ l:<br />

bs L np k up m: mean value of the bootstrap tail dependence estimators ˆ λU by nonparametric<br />

approach calculated for each individual window consisting of bs samples<br />

and all of the nine assets given from left to right in the following order:<br />

BMY, CVX, HPQ, KO, MMM, PG, SGP, TXN, WAG. ’k’ means that ˆ l were<br />

7 Results are on the data CD submitted with the thesis in the folder: window results<br />

53


estimated for the whole threshold k given by X1,N ≥ X2,N ≥ · · · ≥ Xk,N and<br />

Y1,N ≥ Y2,N ≥ · · · ≥ Yk,N with k = 0.04 · N and . . .y . . . means that ˆ l were<br />

estimated for the corrected threshold from Xc,N ≥ Xc+1,N ≥ · · · ≥ Xk,N and<br />

Yc,N ≥ Yc+1,N ≥ · · · ≥ Yk,N with k = 0.04 · N and c = [0.05 · k] where [·] denotes<br />

integer numbers.<br />

rel 99k up: value of 99% quantile divided by bs L np k up m shows the magnitude<br />

of the quantile relative to the absolute average estimate. To obtain the absolute<br />

value of the quantile one simply multiplies the quantile by bs L np k up m.<br />

rel 95k up: value of 95% quantile divided by bs L np k up m shows the magnitude<br />

of the quantile relative to the absolute average estimate. To obtain the absolute<br />

value of the quantile one simply multiplies the quantile by bs L np k up m.<br />

rel 90k up: value of 90% quantile divided by bs L np k up m shows the magnitude<br />

of the quantile relative to the absolute average estimate. To obtain the absolute<br />

value of the quantile one simply multiplies the quantile by bs L np k up m.<br />

For 95% quantiles I additionally calculated mean values over time and bootstrap deviations<br />

over time (shown below data sets on Excel sheets).<br />

Overall the deviation estimates of the rolling time windows look consistent along the<br />

time axis. But first of all with respect to negative tails there are certain fluctuations<br />

in the quantile estimates in time. Therefore the time series were checked for serial<br />

correlation. This is shown at the example of a return series of index S&P 500 for an<br />

interval from January 1950 to April 2008 with a total of N = 14654 data points plotted<br />

over time in figure (3.25). It seems that large and small changes are clustered together.<br />

To confirm this, autocorrelation proposed by Box & Jenkins [16] (1994) for different<br />

lags is plotted in figure (3.26) for returns and squared returns 8 . There seems to be no<br />

autocorrelation in the return series itself but the squarred returns exhibit significant<br />

autocorrelation at least up to lag 5. Since the squared returns measure the second<br />

order moment of the original time series, this result indicates that the variance of S&P<br />

500 conditional on its past history may change over time, or that the time series may<br />

exhibit time varying conditional heteroskedasticity or volatility clustering. The serial<br />

correlation in squared returns, or conditional heteroskedasticity, can be expressed using<br />

an autoregressive process for squared residuals i.e. by the autoregressive conditional<br />

heteroscedasticity ARCH model of Engle [17] (1982) usually referred to as the ARCH(p)<br />

model given by:<br />

Yt = c + εt<br />

σ 2 t = α0 +<br />

p�<br />

αj · ε<br />

(3.61)<br />

2 t−j where α0 > 0, αj ≥ 0 (3.62)<br />

j=1<br />

Yt denote a stationary time series such as financial returns and is expressed by its mean<br />

c and iid Gaussian white noise εt with mean zero. p denotes the length of ARCH lags<br />

(number of time steps considered). Since εt has zero mean, Vart−1(ε) = Et−1 (ε 2 ) = σ 2 t ,<br />

the above equation can be rewritten as:<br />

rk =<br />

ε 2 t = α0 +<br />

p�<br />

j=1<br />

αj · ε 2 t−j<br />

+ ut<br />

(3.63)<br />

8 Autocorrelation denotes the correlation of a variable with itself over successive time intervals defined by:<br />

� N −k<br />

i=1 (Yi−Y )·(Yi+k−Y )<br />

�<br />

Ni=1(Y<br />

i−Y ) 2 with successive measurements Y1, . . . , YN and lag k<br />

54


Return<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

−0.1<br />

−0.15<br />

−0.2<br />

S&P 500 Daily Returns<br />

−0.25<br />

Jan 1950 Jan 1969 Jan 1988 Apr 2008<br />

Figure 3.25: Return data for index S&P 500 ranging from January 1950 to April 2008<br />

where ut = ε2 t − Et−q (ε2 t ) is a zero mean white noise process.<br />

To check the hypothesis that λ evolves according an ARCH process I applied Engle’s<br />

test [17] (1982) for the presence of ARCH effects to my historical day-to-day return<br />

time series. ’archtest’ is a predefined function handle in Matlab, which allows one to<br />

test for autocorrelation of univariate time series. Since an ARCH model can be written<br />

as an AR (autoregressive) model in terms of squared residuals as in equation (3.63),<br />

a Lagrange Multiplier (LM) test for ARCH effects can be constructed based on the<br />

auxiliary regression (3.63). Under the null hypothesis that there are no ARCH effects:<br />

α1 = α2 = . . . = αp = 0, the test statistic follows a χ2 distribution with p degrees of<br />

freedom:<br />

LM = T · R 2 ∼ χ 2 (p), (3.64)<br />

where T is the sample size and R 2 is computed from the regression (3.63) using estimated<br />

residuals 9 . The alternative hypothesis is that, in the presence of ARCH<br />

components, at least one of the estimated α coefficients must be significantly different<br />

from zero. Generally spoken if (T · R 2 ) is greater than the χ 2 table value, we reject the<br />

null hypothesis and conclude there are ARCH effects. If (T · R 2 ) is smaller than the<br />

χ 2 value, we accept the null hypothesis assuming there are no ARCH effects. These<br />

explainations are drawn on Ziwot & Wang [18] (2007). They based their explainations<br />

on S-Plus, which is a statistics software similar to Matlab.<br />

Test statistics at the example of index S&P 500 for an interval from January 1950<br />

to April 2008 with a total of N = 14654 is given in table (3.14). The results show<br />

significant evidence in support of GARCH effects for all the three lags.<br />

p: size of lags<br />

9 We refer to Engle [17] (1982) for details<br />

55


Sample Autocorrelation<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

ACF with Bounds for Raw Return Series<br />

−0.2<br />

0 5 10<br />

Lag<br />

15 20<br />

Sample Autocorrelation<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

ACF of the Squared Returns<br />

−0.2<br />

0 5 10<br />

Lag<br />

15 20<br />

Figure 3.26: Autocorrelation for a return series on the left and for a squared return series on<br />

the right of index S&P 500 ranging from January 1950 to April 2008<br />

H: Boolean decision variable: 0 indicates acceptance of null hypothesis that no ARCH<br />

effects exist; i.e. there is no heteroskedasticity at the corresponding length of lags.<br />

1 indicates rejection of the null hypothesis.<br />

P-Value: P-values (significance levels) at which the ARCH test rejects the null hypothesis<br />

of no existing ARCH effects<br />

ARCH Test Stat: ARCH test statistics for each number of lags equal to (T · R 2 ) with<br />

T denoting the sample size and R 2 computed from the regression (3.63) using<br />

estimated residuals<br />

Critical Value: critical values of the χ 2 distribution for comparison with the corresponding<br />

ARCH test statistics<br />

p H P-Value ARCH Test Stat. Critical Value<br />

10 1 0 805 18<br />

15 1 0 815 25<br />

20 1 0 824 31<br />

Table 3.14: ARCH statistics applied to daily S&P 500 index return data for an interval<br />

ranging from January 1950 to April 2008 and for lags of q = 10, 15, and 20<br />

As we can see on the example of index S&P 500 for the chosen time interval, the p-value<br />

is zero for all of the three lags, which is smaller than the conventional 5% rejection level,<br />

so we reject the null hypothesis that there are no ARCH effects.<br />

56


3.2.4 Implementation of the Parametric Approach<br />

Now I come to the Implementation of the parametric approach by Sornette & Malevergne.<br />

This approach assumes that in the extreme tails the survival distributions<br />

functions of asset returns X and the corresponding residues ε, assumed as idiosyncratic<br />

noise independent of index returns Y and given by the linear factor model:<br />

X = β · Y + ε, can be approximated by power-laws given by: Pr{X > x} ∼ C · x −α ,<br />

which constitutes the parametrization.<br />

Considering N sorted realizations of return vector Y and residual vector ε with<br />

Y1,N ≥ Y2,N ≥ · · · ≥ YN,N and ε1,N ≥ ε2,N ≥ · · · ≥ εN,N, the coefficient of tail<br />

dependence is given as function of the ratio of the scale factors:<br />

as:<br />

ˆλ =<br />

CY<br />

ˆ = k<br />

N · (Yk,N) α , as k → N, 0 (3.65)<br />

Ĉε = k<br />

N · (εk,N) α , as k → N, 0 (3.66)<br />

1<br />

1 + ˆ β −α · Ĉε<br />

Ĉy<br />

=<br />

1 +<br />

1<br />

� ˆεk,N<br />

ˆβ·Yk,N<br />

� α, for k → N, 0 (3.67)<br />

where ε and β are estimated in a least squares sense. If β < 0 for any asset relation,<br />

relation (3.67) cannot be applied, since it assumes β > 0 and therefore ˆ λ = 0.<br />

We can check whether the scale factors Ĉ can be consistently estimated given a<br />

certain threshold k by plotting ĈY and Ĉε for the nine assets in dependence of k using<br />

Hill’s estimator ˆν and Gabaix’s estimator ˆ b γ n. Figure (3.27) shows ĈY (k) and Ĉε(k) for<br />

k=1, 2,..., 458 (this constitutes the interval k/N = 0% . . .8% of the bigger data set)<br />

calculated by Hill’s estimator, and figure (3.28) shows ĈY (k) and Ĉε(k) for k=1, 2,...,<br />

458 calculated by Gabaix’s estimator, both for the upper tails of the bigger data set.<br />

Both plots show robust behaviour in the range of 3 to 5%.<br />

To provide a coefficient that is independent of the choice of the tail coefficient, I also<br />

plotted:<br />

� Ĉε(k)<br />

ĈY (k)<br />

� 1/α<br />

= εk,N<br />

. Figure (3.29) shows<br />

Yk,N<br />

� Ĉε(k)<br />

ĈY (k)<br />

� 1/α<br />

for k=1, 2,..., 458 also<br />

for the upper tails of the bigger data set. All these plots show that asset TXN (Texas<br />

Instruments Inc.) seems to show somewhat a different behavior than the other assets<br />

and the index.<br />

Also for the parametric approach, in order to put emphasis on the extreme tails and<br />

in order to increase robustness of ĈY and Ĉε and hence of tail dependence coefficients<br />

ˆλ to changes of threshold k, I used the average values of ĈY , ĈY = 1<br />

dependence of k and the average value of Ĉε, Ĉε = 1<br />

k<br />

k<br />

� k<br />

j=1 Ĉy,j in<br />

� k<br />

j=1 Ĉε,j in dependence of k.<br />

Figure (3.30) shows on the example of the positive tails of the bigger sample that<br />

� �1/α Ĉε<br />

ĈY<br />

for the nine assets is stable in k for k<br />

N<br />

= 3% . . .5%.<br />

Figure (3.28) shows that for small k/N values of ĈY and Ĉε could be comparatively<br />

� �1/α Ĉε(k)<br />

large, which also has a certain, even though weaker effect on . To avoid<br />

ĈY (k)<br />

distortion of our results it might be better to exclude the most extreme values (k →<br />

0, N) in order to estimate the mean of the scale factors of the power law distributions.<br />

Therefore we calculate ĈY,c = 1 �k k−c+1 j=c ĈY,j and Ĉε,c = 1 �k k−c+1 j=c Ĉε,j with c =<br />

[0.05 · k] where [·] denotes integer numbers.<br />

57


C Y , C ε<br />

C Y , C ε<br />

x 10−6<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

S&P 500<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08<br />

k/N<br />

Figure 3.27: Scale factors ĈY (k) = k<br />

N · (Yk,N) ˆν of the index S&P 500 and Ĉε(k) = k<br />

N · (εk,N)<br />

ˆν<br />

of the nine assets’ residues plotted for the upper tails of the bigger data set ranging from July<br />

1985 to April 2008 with k=1, 2,. . ., 458 (k/N = 0% ... 8%).<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

x 10−6<br />

2<br />

S&P 500<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08<br />

k/N<br />

Figure 3.28: Scale factors ĈY (k) = k<br />

N<br />

·(Yk,N) ˆ<br />

b γ n of the index S&P 500 and Ĉε(k) = k<br />

N<br />

·(εk,N) ˆ<br />

b γ n<br />

of the nine assets’ residues plotted for the upper tails of the bigger data set ranging from July<br />

1985 to April 2008 with k=1, 2,. . ., 458 (k/N = 0% ... 8%).<br />

58


(C ε /C Y ) 1/α<br />

(C ε,mean /C Y,mean ) 1/α<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08<br />

Figure 3.29: Fraction of scale factors<br />

� Ĉε(k)<br />

ĈY (k)<br />

k/N<br />

�1/α = εk,N<br />

Yk,N<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

of the nine assets’ residues and the<br />

index S&P 500, plotted for the upper tails of the bigger data set ranging from July 1985 to<br />

April 2008 with k=1, 2,. .., 458 (k/N = 0% ... 8%).<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

1<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08<br />

k/N<br />

�1/α Figure 3.30: Fraction of scale factors<br />

� Ĉε<br />

ĈY<br />

of the nine assets’ residues and the index S&P<br />

500, plotted for the upper tails of the bigger data set ranging from July 1985 to April 2008<br />

with k=1, 2,. .., 458 (k/N = 0% ...8%).<br />

59


In general it is an advantage to follow a parametric approach, if the assumed parametric<br />

form of the distributions is not too far from the true one. We can check this<br />

visually by comparing the empirical complementary cumulative distribution functions<br />

F i,empirical = k<br />

for k/N = 0% . . .4% with the parametric complementary cumulative<br />

N<br />

distribution functions F i,parametric = C· · x−α of the tails. Figure (3.31) shows the two<br />

functions in dependence of return values for the upper tails of the bigger sample of εi<br />

for all assets using Hill’s estimator and Ĉε for (k/N = 4%) and figure (3.32) using<br />

Ĉε,c the mean value of Ĉε(k) for k/N = 0.5% . . .4%. Figure (3.33) and figure (3.34)<br />

show the same functions using Gabaix’s estimator. Results are consistent and show<br />

that by building the mean values of constants C the parametrization becomes more<br />

precise independent of the tail index estimator we chose. Anyway, it is difficult to say<br />

in general, which tail index estimator works better. Therefore it makes sense to still<br />

calculate both for comparison.<br />

60


F par ,F emp<br />

F par ,F emp<br />

61<br />

F par ,F emp<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06<br />

returns<br />

KO<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

0<br />

SGP<br />

0.04 0.05<br />

returns<br />

0.06 0.07<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

CVX<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

MMM<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

0<br />

TXN<br />

0.06 0.07 0.08 0.09<br />

returns<br />

0.1 0.11<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

HPQ<br />

0.04 0.05 0.06<br />

returns<br />

0.07 0.08<br />

PG<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

WAG<br />

0<br />

0.03 0.035 0.04 0.045 0.05 0.055<br />

returns<br />

Figure 3.31: Empirical complementary cumulative distribution functions F ε,empirical = k<br />

N plotted in red and parametric complementary cumulative<br />

distribution functions F ε,parametric = Ĉε · ε−ˆν with Ĉε for (k/N = 4%) plotted in black for k/N = 4% ...0% (from left to right) in dependence of<br />

returns for the upper tails of the nine assets.


F par ,F emp<br />

F par ,F emp<br />

62<br />

F par ,F emp<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06<br />

returns<br />

KO<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

0<br />

SGP<br />

0.04 0.05<br />

returns<br />

0.06 0.07<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

CVX<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

MMM<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

0<br />

TXN<br />

0.06 0.07 0.08 0.09<br />

returns<br />

0.1 0.11<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

HPQ<br />

0.04 0.05 0.06<br />

returns<br />

0.07 0.08<br />

PG<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

WAG<br />

0<br />

0.03 0.035 0.04 0.045 0.05 0.055<br />

returns<br />

Figure 3.32: Empirical complementary cumulative distribution functions F ε,empirical = k<br />

N plotted in red and parametric complementary cumulative<br />

distribution functions F ε,parametric = Ĉε,c · ε−ˆν with Ĉε,c = 1 �k k−c j=c Ĉε for c = 0.005 · N plotted in black for k/N = 4% ...0% (from left to right)<br />

in dependence of returns for the upper tails of the nine assets.


F par ,F emp<br />

F par ,F emp<br />

63<br />

F par ,F emp<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06<br />

returns<br />

KO<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

0<br />

SGP<br />

0.04 0.05<br />

returns<br />

0.06 0.07<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

CVX<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

MMM<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

0<br />

TXN<br />

0.06 0.07 0.08 0.09<br />

returns<br />

0.1 0.11<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0.03<br />

0.02<br />

0.01<br />

0.03<br />

0.02<br />

0.01<br />

HPQ<br />

0.04 0.05 0.06<br />

returns<br />

0.07 0.08<br />

PG<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

WAG<br />

0<br />

0.03 0.035 0.04 0.045 0.05 0.055<br />

returns<br />

Figure 3.33: Empirical complementary cumulative distribution functions F εi,empirical = k<br />

N plotted in green and parametric complementary cumulative<br />

distribution functions F ε,parametric = Ĉε ·ε−ˆb γ n with Ĉε for (k/N = 4%) plotted in black for the upper tails of k/N = 4% ... 0% in dependence<br />

of returns for the nine assets.


F par ,F emp<br />

F par ,F emp<br />

64<br />

F par ,F emp<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.02<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06<br />

returns<br />

KO<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

0<br />

SGP<br />

0.04 0.05<br />

returns<br />

0.06 0.07<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

CVX<br />

0<br />

0.025 0.03 0.035<br />

returns<br />

0.04 0.045<br />

MMM<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

0<br />

TXN<br />

0.06 0.07 0.08 0.09<br />

returns<br />

0.1 0.11<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.04<br />

0.02<br />

0<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

HPQ<br />

0.04 0.05 0.06<br />

returns<br />

0.07 0.08<br />

PG<br />

0<br />

0.025 0.03 0.035 0.04 0.045<br />

returns<br />

WAG<br />

0<br />

0.03 0.035 0.04 0.045 0.05 0.055<br />

returns<br />

Figure 3.34: Empirical complementary cumulative distribution functions F εi,empirical = k<br />

N plotted in green and parametric complementary cumulative<br />

distribution functions F ε,parametric = Ĉε,c ·ε−ˆb γ n with Ĉε,c = 1 �k k−c j=c Ĉε for c = 0.005 ·N plotted in black for k/N = 4% ... 0% in dependence<br />

of returns for the upper tails of the nine assets.


To summarize for upper tails: tail indexes α were calculated by Hill’s estimator ˆν<br />

and Gabaix’s OLS log-log rank-size tail index estimate ˆb γ n given by:<br />

ˆν =<br />

�<br />

1 �<br />

kj=1<br />

log<br />

k<br />

Yj,N<br />

�−1 Yk,N<br />

(3.68)<br />

log(Rank(Yj,N) − 1/2) = a − ˆ b γ n · log(Yj,N), for j = 1 . . .k (3.69)<br />

where Y1,N ≥ Y2,N ≥ · · · ≥ YN,N denote ordered statistics of the sample consisting of<br />

N realizations of the market return vector Y and threshold k is assumed to represent<br />

the 4% quantile of Y .<br />

β and idiosyncratic noise ε were calculated by ordinary least squares regression to<br />

the linear factor model: X = β · Y + ε with X denoting asset returns in descending<br />

chronologic order, and with Y denoting corresponding index returns.<br />

Constants ĈY and Ĉε were parametrically estimated by:<br />

and<br />

ĈY = 1<br />

k<br />

ĈY,c =<br />

Ĉε = 1<br />

k<br />

Ĉε,c =<br />

k�<br />

j=1<br />

j<br />

N<br />

1<br />

k − c + 1<br />

k�<br />

j=1<br />

j<br />

N<br />

1<br />

k − c + 1<br />

· (Yj,N) α<br />

k�<br />

j=c<br />

j<br />

N<br />

· (εj,N) α<br />

k�<br />

j=c<br />

j<br />

N<br />

· (Yj,N) α<br />

· (εj,N) α<br />

(3.70)<br />

(3.71)<br />

(3.72)<br />

(3.73)<br />

considering N sorted realizations of index return vector Y and residual vector ε denoted<br />

by Y1,N ≥ Y2,N ≥ · · · ≥ YN,N and ε1,N ≥ ε2,N ≥ · · · ≥ εN,N.<br />

Putting all these parameters in equation (3.74) for upper tail dependence:<br />

ˆλ + =<br />

ˆλ + =<br />

1<br />

1 + ˆ β−ˆν · Ĉε<br />

ĈY<br />

1<br />

1 + ˆ β −ˆ b γ n · Ĉε<br />

ĈY<br />

(3.74)<br />

(3.75)<br />

and applying the approach to our reference data sets yields the estimates reported in<br />

tables (3.8) and (3.10) on the left side using Hill’s estimator ν and reported in tables<br />

(3.9) and (3.11) on the left side using Gabaix’s estimator b γ n 10 .<br />

To achieve estimates for the lower tail, we can just sort our data in ascending order<br />

and multiply all of the ordered samples by (-1). β and ε remains the same for negative<br />

tails, as it is calculated for the whole data sample within this approach.<br />

10 The matlab m-file for the implementation of λ +,− estimated by the parametric approach according to<br />

Sornette & Malevergne is denoted by: lambda par sor ma.m and enclosed to the appendix and the data CD<br />

65


λ(k)<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08<br />

k/N<br />

Figure 3.35: ˆ λ =<br />

1<br />

1+ ˆ β −ˆν · Cε<br />

Cy<br />

with k = 1... 458 (or k/N = 0% ...8%), applied to the lower tails<br />

of all assets with the index S&P 500 on the bigger data set ranging from July 1985 to April<br />

2008.<br />

3.2.5 Analysis of TDC estimated by the Parametric Approach<br />

Also for the parametric approach by Sornette & Malevergne I provide a sensitivity<br />

analysis of TDC estimates performed on the two reference data sets: the robustness<br />

of the results to the choice of threshold k is first analyzed visually by plotting ˆ λ in<br />

dependence of k. In a second step estimation of error bars of ˆ λ is provided by bootstrap<br />

sampling with replacement. The aim is to find confidence intervals for the estimates<br />

and to get an insight into the underlying error distributions.<br />

Figure (3.35) shows ˆ λ(k) for k<br />

k<br />

= 0% . . .8%, figure (3.36) λ(k) for = 0% . . .8%,<br />

N N<br />

and figure (3.37) shows λc(k) for k = 0.5% . . .8% for the lower tail of the smaller data<br />

N<br />

sample. Results of estimators for positive tails look less fluctuating because in positive<br />

tails outliers are less extreme. For the parametric approach also counts that on the<br />

interval: k = 3% . . .5%, that is of primary interest to us, the curve of λ is not stiff to<br />

N<br />

moderate changes in k. Using the uncorrected mean value λ(k), estimates tend to be<br />

slightly lower than by choosing the corrected estimate λc(k), which means neglecting<br />

most extreme outliers.<br />

The results of the bootstrap sampling are shown in table (3.15) for the smaller sample<br />

and in table (3.16) for the bigger sample 11 . bs denotes the number of samples, which<br />

has been set to thousand for both sample sizes and N shows the number of daily observations<br />

of the samples. The procedure is exactly the same as for the non-parametric<br />

approach. Especially conspicuous is the column of relative errors in the uncorrected<br />

negative tail calculated for the smaller sample and located in the upper right parts of<br />

table (3.15). ’rel err’ are calculated as follows: |λoriginal − 1/bs · �bs j=1 λj|/λoriginal. An<br />

explanation for this could be that extreme outliers, as already discussed at the example<br />

of the non-parametric approach, have an even higher stake in terms of distorting tail<br />

11 The matlab m-file for the bootstrap sampling with replacement of λ +,− estimated by the parametric<br />

approach according to Sornette & Malevergne is denoted by: bootstr par alld.m and enclosed to the data CD<br />

66


λ(k) mean<br />

λ(k) mean,c<br />

0.1<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08<br />

k/N<br />

Figure 3.36: λ =<br />

1<br />

1+ ˆ β −ˆν · Ĉε<br />

Ĉy<br />

with k = 1... 458 (or k/N = 0% ... 8%), applied to the lower tails<br />

of all assets with the index S&P 500 on the bigger data set ranging from July 1985 to April<br />

2008.<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

CVX<br />

HPQ<br />

KO<br />

MMM<br />

PG<br />

SGP<br />

TXN<br />

WAG<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08<br />

k/N<br />

Figure 3.37: λ =<br />

1<br />

1+ ˆ β −ˆν · Ĉε,c<br />

Ĉy,c<br />

with k = 29... 458 (or k/N = 0.5% ... 8%), and c = [0.005 · N]<br />

([·] denotes integer numbers) applied to the lower tails of all assets with the index S&P 500<br />

on the bigger data set ranging from July 1985 to April 2008.<br />

67


dependence estimates in the parametric approach. If we correct for these most extreme<br />

outliers results become accurate. To have a clear advantage by using a parametric<br />

approximation, a certain amount of data is needed to determine the coefficients with<br />

satisfying accuracy. This amount might be higher than for a non-parametric approach.<br />

Looking at the relative errors of the bigger sample we see that relative errors are<br />

much lower compared to the smaller sample. We can check how the parametrization<br />

works for the lower tail of the smaller sample by comparing the empirical complementary<br />

cumulative distribution functions F i,empirical = k<br />

for k/N = 0% . . .4% with the<br />

N<br />

parametric complementary cumulative distribution functions F i,parametric = C· ∗ x−α of the tails. Figure (3.38) shows the two functions in dependence of absolute return<br />

values for the lower tails of the smaller sample of εi for all assets using Hill’s estimator<br />

and Ĉε the mean value of Ĉε(k) for k/N = 0% . . .4% and figure (3.39) using Hill’s<br />

estimator and Ĉε,c the mean value of Ĉε(k) for k/N = 0.5% . . .4%.<br />

Comparing the two plots shows that the extraordinary high outliers in the negative<br />

tails of our data samples in connection with a sample size that seems to be less<br />

than the critical or necessary size indeed seems to distort the approximation of the<br />

complementary cumulative distribution function in the region of k/N ≈ 4%. Looking<br />

at figure (3.39) we can detect that in the domain of k/N → 0 in most cases<br />

F εi,parametric = Ĉε,c · ε −ˆν seems to overestimate F εi,empirical = k/N. To see this more<br />

clearly we can plot the complementary cumulative distribution functions on a logarithmic<br />

scale to the base ten. Figures (3.40) and (3.41) show the semi-log plots<br />

corresponding to figures (3.38) and (3.39). As we can observe on the semi-log plots<br />

the uncorrected parametric complementary cumulative distribution functions perform<br />

better in the area k/N → 0 and the corrected parametric complementary cumulative<br />

distribution functions perform better in the area k/N → 4%. This is intuitively clear<br />

because if we do not correct for the most extreme outliers we shift the emphasis of<br />

the parametrization towards these most extreme outliers. This shows the limitations<br />

of our parametrization approach but anyway only poses a problem in the negative tails<br />

because of the extraordinary outliers we find there. Because of the fact that both approaches<br />

have a domain where they dominate, it still makes sense to calculate both of<br />

them first of all for negative tails.<br />

Anyway, for the non-parametric approach sample sizes should be chosen around the<br />

size of the bigger reference data sample that consists of 5736 data points to achieve<br />

satisfying accuracy. In spite of relative errors the magnitude of quantile and standard<br />

deviation estimates relative to the total tail dependence estimates is comparable to<br />

estimates by the non-parametric approach.<br />

Because we need very high sample sizes here to obtain trustworthy results and because<br />

it has been shown that for bigger sample sizes there is a loss of significant peaks, I<br />

think that the non-parametric approach is better to analyze tail dependence estimates<br />

in dependence of time. Therefore I don’t go into observations of λ by rolling time windows<br />

within this approach. Error bars are of similar extent as for the non-parametric<br />

approach but the procedure seems to request more data to provide accurate results.<br />

68


69<br />

m=2507 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=100 rel err max dev 95% q 90% q std bs k=100 rel err max dev 95% q 90% q std bs<br />

BMY 0.07 0.01 0.12 0.04 0.04 0.02 0.02 0.90 0.10 0.06 0.04 0.03<br />

CVX 0.02 0.01 0.04 0.02 0.02 0.01 0.01 1.60 0.10 0.03 0.02 0.01<br />

HPQ 0.13 0.02 0.20 0.08 0.07 0.04 0.03 0.50 0.20 0.08 0.05 0.04<br />

KO 0.07 0.01 0.10 0.06 0.05 0.03 0.02 2.00 0.20 0.09 0.07 0.04<br />

MMM 0.04 0.01 0.09 0.03 0.03 0.02 0.02 0.90 0.09 0.04 0.03 0.02<br />

PG 0.03 0.02 0.05 0.03 0.02 0.02 0.01 1.40 0.09 0.02 0.01 0.01<br />

SGP 0.05 0.01 0.07 0.04 0.03 0.02 0.01 0.90 0.10 0.04 0.02 0.02<br />

TXN 0.00 Inf 0.00 0.00 0.00 0.00 0.00* Inf 0.00 0.00 0.00 0.00<br />

WAG 0.11 0.00 0.15 0.08 0.06 0.04 0.01 0.90 0.20 0.05 0.02 0.02<br />

λc<br />

BMY 0.07 0.01 0.20 0.04 0.04 0.02 0.08 0.03 0.13 0.05 0.05 0.03<br />

CVX 0.02 0.01 0.06 0.02 0.02 0.01 0.02 0.01 0.08 0.03 0.02 0.01<br />

HPQ 0.13 0.02 0.20 0.08 0.07 0.04 0.12 0.04 0.19 0.07 0.06 0.04<br />

KO 0.08 0.02 0.10 0.05 0.05 0.03 0.08 0.02 0.13 0.06 0.05 0.03<br />

MMM 0.04 0.01 0.08 0.03 0.03 0.02 0.04 0.00 0.10 0.04 0.03 0.02<br />

PG 0.03 0.02 0.09 0.03 0.02 0.02 0.04 0.02 0.10 0.04 0.03 0.02<br />

SGP 0.05 0.01 0.09 0.04 0.03 0.02 0.06 0.02 0.09 0.05 0.04 0.02<br />

TXN 0.00 Inf 0.01 0.00 0.00 0.00 0.00* inf 0.02 0.00 0.00 0.00<br />

WAG 0.10 0.00 0.10 0.07 0.06 0.04 0.12 0.02 0.20 0.08 0.07 0.04<br />

Table 3.15: Establishing the uncertainty of parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values, and<br />

standard deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time<br />

interval from January 1991 to December 2000. ˆ βj have been calculated on the whole samples. ∗ denotes negative ˆ β and therefore ˆ λ = 0 and ’Inf’<br />

denotes ·/0.


70<br />

m=5736 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=229 rel err max dev 95% q 90% q std bs k=229 rel err max dev 95% q 90% q std bs<br />

BMY 0.06 0.01 0.06 0.03 0.03 0.02 0.03 0.40 0.08 0.04 0.03 0.02<br />

CVX 0.05 0.00 0.05 0.03 0.02 0.01 0.02 0.50 0.09 0.04 0.03 0.02<br />

HPQ 0.12 0.00 0.08 0.05 0.04 0.03 0.08 0.10 0.10 0.07 0.06 0.04<br />

KO 0.13 0.02 0.10 0.06 0.05 0.03 0.03 1.10 0.30 0.09 0.05 0.04<br />

MMM 0.08 0.02 0.10 0.04 0.04 0.02 0.03 0.80 0.20 0.06 0.04 0.03<br />

PG 0.03 0.10 0.06 0.03 0.03 0.02 0.01 1.00 0.07 0.02 0.01 0.01<br />

SGP 0.03 0.10 0.06 0.03 0.03 0.02 0.01 0.60 0.08 0.02 0.01 0.01<br />

TXN 0.00 0.30 0.01 0.00 0.00 0.00 0.00 0.80 0.01 0.00 0.00 0.00<br />

WAG 0.09 0.09 0.07 0.04 0.04 0.02 0.03 0.50 0.08 0.03 0.03 0.02<br />

λc<br />

BMY 0.06 0.01 0.05 0.03 0.03 0.02 0.07 0.04 0.06 0.03 0.03 0.02<br />

CVX 0.05 0.00 0.05 0.03 0.02 0.01 0.05 0.05 0.07 0.04 0.03 0.02<br />

HPQ 0.13 0.00 0.08 0.05 0.04 0.03 0.13 0.02 0.09 0.05 0.04 0.03<br />

KO 0.13 0.01 0.10 0.06 0.05 0.03 0.14 0.02 0.10 0.07 0.06 0.04<br />

MMM 0.08 0.02 0.09 0.04 0.04 0.02 0.08 0.05 0.10 0.04 0.04 0.02<br />

PG 0.03 0.10 0.06 0.03 0.03 0.02 0.04 0.10 0.09 0.04 0.03 0.02<br />

SGP 0.03 0.10 0.06 0.03 0.03 0.02 0.04 0.10 0.06 0.03 0.03 0.02<br />

TXN 0.00 0.20 0.01 0.00 0.00 0.00 0.00 0.40 0.01 0.00 0.00 0.00<br />

WAG 0.09 0.01 0.07 0.04 0.03 0.02 0.10 0.02 0.08 0.04 0.04 0.02<br />

Table 3.16: Establishing the uncertainty of of parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values, and<br />

standard deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time<br />

interval from July 1985 to April 2008. ˆ βj have been calculated on the whole samples.


F par ,F emp<br />

F par ,F emp<br />

71<br />

F par ,F emp<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

BMY<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

KO<br />

SGP<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

CVX<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

MMM<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

TXN<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

HPQ<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

PG<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

1<br />

0.5<br />

WAG<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

Figure 3.38: Empirical complementary cumulative distribution functions F εi,empirical = k<br />

N plotted in red and parametric complementary cumulative<br />

distribution functions F ε,parametric = Ĉε · ε−ˆν with Ĉε = 1 �k k j=1 Ĉε plotted in black for k/N = 4% ... 0% (from left to right) in dependence of<br />

returns for the lower tails of the nine assets for a time interval ranging from January 1991 to December 2000.


F par ,F emp<br />

F par ,F emp<br />

72<br />

F par ,F emp<br />

0.06<br />

0.04<br />

0.02<br />

0.06<br />

0.04<br />

0.02<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

KO<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

SGP<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.06<br />

0.04<br />

0.02<br />

0.06<br />

0.04<br />

0.02<br />

CVX<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

MMM<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

TXN<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

0.06<br />

0.04<br />

0.02<br />

0.06<br />

0.04<br />

0.02<br />

0.06<br />

0.04<br />

0.02<br />

HPQ<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

PG<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

WAG<br />

0<br />

0 0.2 0.4<br />

returns<br />

0.6 0.8<br />

Figure 3.39: Empirical complementary cumulative distribution functions F εi,empirical = k<br />

N plotted in red and parametric complementary cumulative<br />

distribution functions F εi,parametric = Ĉε, c ·ε−ˆν with Ĉε,c = 1 �k k−c j=c Ĉε for c = 0.005 ·N plotted in black for k/N = 4% ... 0% (from left to right)<br />

in dependence of returns for the lower tails of the nine assets for a time interval ranging from January 1991 to December 2000.


F par ,F emp<br />

F par ,F emp<br />

73<br />

F par ,F emp<br />

10 0<br />

BMY<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

10 0<br />

KO<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

10 0<br />

SGP<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

10 0<br />

10 −2<br />

10 −4<br />

CVX<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−6<br />

returns<br />

10 0<br />

10 −5<br />

MMM<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−10<br />

returns<br />

10 0<br />

TXN<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

10 0<br />

HPQ<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

10 0<br />

PG<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

10 0<br />

WAG<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

Figure 3.40: Empirical complementary cumulative distribution functions F εi,empirical = k<br />

N plotted in red and parametric complementary cumulative<br />

distribution functions F ε,parametric = Ĉε · ε−ˆν with Ĉε = 1 �k k j=1 Ĉε plotted in black for k/N = 4% ... 0% in dependence of returns for the lower<br />

tails of the nine assets for a time interval ranging from January 1991 to December 2000.


F par ,F emp<br />

F par ,F emp<br />

74<br />

F par ,F emp<br />

10 0<br />

10 −2<br />

10 −4<br />

BMY<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−6<br />

returns<br />

10 0<br />

10 −5<br />

KO<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−10<br />

returns<br />

10 0<br />

10 −5<br />

SGP<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−10<br />

returns<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

10 0<br />

10 −5<br />

CVX<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−10<br />

returns<br />

10 0<br />

10 −5<br />

MMM<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−10<br />

returns<br />

10 0<br />

TXN<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−5<br />

returns<br />

F par ,F emp<br />

F par ,F emp<br />

F par ,F emp<br />

10 0<br />

10 −2<br />

10 −4<br />

HPQ<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−6<br />

returns<br />

10 0<br />

10 −5<br />

PG<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−10<br />

returns<br />

10 0<br />

10 −5<br />

WAG<br />

10<br />

0 0.2 0.4 0.6 0.8<br />

−10<br />

returns<br />

Figure 3.41: Empirical complementary cumulative distribution functions F εi,empirical = k<br />

N plotted in red and parametric complementary cumulative<br />

distribution functions F εi,parametric = Ĉε, c · ε−ˆν with Ĉε,c = 1 �k k−c j=c Ĉε for c = 0.005 · N plotted in black for k/N = 4% ...0% in dependence of<br />

returns for the lower tails of the nine assets for a time interval ranging from January 1991 to December 2000.


3.3 β-Smile Improvement<br />

As an addition to my Sornette & Malevergne approaches, where the regression coefficient<br />

β of the components of vector Y on its corresponding components of vector X<br />

was determined by the least squares method applying a linear additive single factor<br />

model to the whole return data samples of a given time interval given by:<br />

X = β · Y + ε (3.76)<br />

with ε denoting the vector of error terms, a simple improvement was proposed. Considering<br />

that the linear factor model given by equation (3.76) only holds for components<br />

X and Y large enough on a certain condition, typically when both X and corresponding<br />

Y belong to the extreme tails of their probability density distributions, which by<br />

assumption denotes the most extreme 4 % of the return data samples. ˆ βSI is then<br />

calculated only for extreme tail data in dependence of threshold value k. There are<br />

basically four possible conditions for the calculation of ˆ βSI of the upper tails 12 :<br />

1 st condition : Y ≥ Y (k) ∩ X ≥ X(k)<br />

2 nd condition : Y ≥ Y (k)<br />

3 rd condition : X ≥ X(k)<br />

4 th condition : Y ≥ Y (k) ∪ X ≥ X(k)<br />

The first condition is to calculate β by only considering return values, where X and Y<br />

belong to the most extreme 4% of the sample, the second and the third conditions are<br />

to calculate β by either considering Y large enough or X large enough. As a fourth<br />

condition I wanted to use the Boolean ’or’ (∪) as an extension to the first condition given<br />

by Boolean ’and’ (∩). Table (3.17) shows relative deviations between tail dependence<br />

coefficient estimates calculated using β of the whole data sample and tail dependence<br />

coefficient estimates that were calculated using ˆ βSI applying the first, second, or third<br />

condition. The fourth condition yields β < 0 and therefor zero tail dependence for all<br />

the nine assets.<br />

In terms of implementation, we just have to build vectors of data that fulfill the<br />

chosen condition and then calculate ˆ βSI (smile improvement) with this data. The<br />

other coefficients are calculated in the same way as described in section (3.4.2), where<br />

the implementation of the non-parametric approach was discussed.<br />

What we can observe in table (3.17) is that the first condition: Y ≥ Y (k)∩X ≥ X(k)<br />

seems to fit best or at least yields values that are closer to my classic β approach than<br />

the two other conditions. Anyway for some assets the second condition: Y ≥ Y (k)<br />

yields estimates, which are more accurate than estimated by the first condition in<br />

terms of being consistent with the classic β approach. Therefor I decided to further<br />

analyze the first and the second condition. Table (3.18) shows results of ˆ βSI calculated<br />

by first and second SI condition in comparison to results of ˆ β calculated by all data and<br />

results of ˆ λ +,− calculated by the first and second SI conditions in comparison to results<br />

of ˆ λ +,− with ˆ β calculated by all data for the smaller data set and table (3.19) shows<br />

the same for the bigger data set 13 . <strong>Tail</strong> indexes were calculated by Hill’s estimator ˆν.<br />

12 I will describe the procedure only on the example of upper tails. For ˆ βSI of the lower tails the same<br />

relation hold using absolute values of return data.<br />

13 The matlab m-file for the implementation of λ +,− estimated by the non-parametric approach according<br />

to Sornette & Malevergne with ˆ β calculated by SI conditions is denoted by: Lambda np SI.m and enclosed to<br />

the data CD<br />

75


condition {X ≥ X(k)} ∩ X ≥ X(k) Y ≥ Y (k) {X ≥ X(k)} ∩ X ≥ X(k) Y ≥ Y (k)<br />

{Y ≥ Y (k)} {Y ≥ Y (k)}<br />

upper tail lower tail<br />

m=2507<br />

ν 3.23 3.23 3.23 3.16 3.16 3.16<br />

λ k=100 k=100 k=100 k=100 k=100 k=100<br />

BMY 0.27 0.96 0.19 0.92 1.00* 2.78<br />

CVX 1.00* 1.00* 0.93 0.73 1.00* 2.06<br />

HPQ 1.00 1.00* 0.33 1.00* 1.00* 0.10<br />

KO 0.86 0.99 0.68 0.21 1.00* 1.36<br />

MMM 1.00* 1.00* 0.91 0.20 0.97 7.76<br />

PG 9.11 0.95 0.96 0.35 1.00* 2.38<br />

SGP 0.26 0.93 0.32 0.97 1.00* 1.91<br />

TXN Inf Inf NaN* NaN* NaN* NaN*<br />

WAG 0.32 0.96 0.79 1.00* 0.85 0.33<br />

m=5736<br />

ν 3.09 3.09 3.09 3.13 3.13 3.13<br />

λ k=229 k=229 k=229 k=229 k=229 k=229<br />

BMY 1.26 0.99 1.00* 0.88 1.00 0.54<br />

CVX 1.17 1.00* 1.00* 0.24 1.00* 0.82<br />

HPQ 0.68 1.00 0.81 0.90 1.00 0.60<br />

KO 8.78 0.79 0.09 1.93 1.00* 3.35<br />

MMM 0.06 0.98 1.00 1.00 1.00* 0.77<br />

PG 1.74 1.00* 1.00* 19.58 1.00* 1.00*<br />

SGP 0.93 1.00* 1.00* 0.27 1.00* 1.00*<br />

TXN 93.63 0.70 1.00* 0.96 1.00* 1.00*<br />

WAG 0.39 1.00* 0.81 0.94 0.91 0.69<br />

Table 3.17: Relative deviations between coefficients of upper and lower tail dependence estimated<br />

by the non-parametric approach of Sornette & Malevergne using ˆ β calculated for the<br />

whole sample and ˆ βSI calculated for the tail according to three different conditions for index<br />

S&P 500 and the nine assets. The tail represents the most extreme 4 % of data during a<br />

smaller time interval ranging from January 1991 to December 2000 (k = 100) and during a<br />

bigger time interval ranging from July 1985 to Mars 2008 (k = 229). * denotes zero values of<br />

tail dependence calculated by ˆ βSI, ’NaN’ denotes 0/0, and ’inf’ denotes ·/0.<br />

76


77<br />

β β +<br />

SI β −<br />

SI λ + λ +<br />

SI λ +<br />

SI λ− λ −<br />

SI λ −<br />

ν<br />

condition all data<br />

3.23 3.23 3.23 3.16 3.16<br />

SI<br />

3.16<br />

� X ≥ X(k) � ∩ � X ≥ X(k) � ∩ 1... k Y ≥ Y (k) � X ≥ X(k) � ∩ 1... k Y ≥ Y (k) � X ≥ X(k) � m=2507<br />

� �<br />

Y ≥ Y (k)<br />

� �<br />

Y ≥ Y (k)<br />

� �<br />

Y ≥ Y (k)<br />

�<br />

∩<br />

�<br />

Y ≥ Y (k)<br />

BMY 0.82 0.74 0.37 0.06 0.07 0.04 0.05 0.18 0.00<br />

CVX 0.45 0.02 0.30 0.02 0.00 0.00 0.02 0.06 0.01<br />

HPQ 1.10 0.22 -0.16 0.10 0.13 0.00 0.07 0.06 0*<br />

KO 0.69 0.38 0.64 0.06 0.02 0.01 0.05 0.12 0.04<br />

MMM 0.55 -0.18 0.51 0.04 0.00 0.00* 0.03 0.26 0.02<br />

PG 0.58 1.20 0.50 0.03 0.00 0.30 0.02 0.07 0.01<br />

SGP 0.85 0.78 0.29 0.04 0.03 0.03 0.04 0.10 0.00<br />

TXN -0.09 5.10 -0.65 0.00* 0.00* 1.00 0.00* 0.00 0.00*<br />

WAG<br />

condition<br />

0.97<br />

all data<br />

1.10<br />

Y ≥ Y (k)<br />

-0.35<br />

Y ≥ Y (k)<br />

0.09<br />

c... k<br />

0.16<br />

Y ≥ Y (k)<br />

0.12 0.06 0.04 0*<br />

� X ≥ X(k) � ∩ c... k Y ≥ Y (k) � X ≥ X(k) � m=2507<br />

� �<br />

Y ≥ Y (k)<br />

�<br />

∩<br />

�<br />

Y ≥ Y (k)<br />

BMY 0.82 0.87 1.20 0.06 0.07 0.04 0.07 0.25 0.01<br />

CVX 0.45 0.20 0.64 0.02 0.00 0.00 0.02 0.07 0.01<br />

HPQ 1.10 1.30 1.10 0.10 0.13 0.00 0.09 0.09 0.00*<br />

KO 0.69 0.48 0.91 0.06 0.02 0.01 0.07 0.17 0.06<br />

MMM 0.55 0.26 1.10 0.04 0.00 0.00* 0.04 0.31 0.03<br />

PG 0.58 0.22 0.85 0.03 0.00 0.30 0.04 0.13 0.03<br />

SGP 0.85 0.76 1.20 0.04 0.03 0.03 0.05 0.15 0.00<br />

TXN -0.09 -1.40 -0.10 0.00* 0.00* 1.00 0.00* 0.00 0.00*<br />

WAG 0.97 1.20 0.86 0.09 0.16 0.12 0.11 0.07 0.00*<br />

Table 3.18: Results for ˆ βSI using the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k) and the second SI condition: Y ≥ Y (k) for upper and lower tails<br />

calculated by least squares method applied on linear additive single factor model: X = β ·Y +ε and for resulting upper and lower tail dependence<br />

coefficients ˆ λ + and ˆ λ− with corrected ˆl (c... k) and uncorrected ˆl (1... k). The tail represents the most extreme 4% of data during a time interval<br />

ranging from January 1991 to December 2000 (k = 100, c = 13). * denotes ˆ β < 0 → λ = 0.


78<br />

β β +<br />

SI β −<br />

SI λ + λ +<br />

SI λ +<br />

SI λ− λ −<br />

SI λ −<br />

ν<br />

condition all data<br />

3.09 3.09 3.09 3.13 3.13<br />

SI<br />

3.13<br />

� X ≥ X(k) � ∩ � X ≥ X(k) � ∩ 1... k Y ≥ Y (k) � X ≥ X(k) � ∩ 1... k Y ≥ Y (k) � X ≥ X(k) � m=5736<br />

� �<br />

Y ≥ Y (k)<br />

� �<br />

Y ≥ Y (k)<br />

� �<br />

Y ≥ Y (k)<br />

�<br />

∩<br />

�<br />

Y ≥ Y (k)<br />

BMY 0.69 0.90 0.35 0.05 0.00* 0.12 0.04 0.02 0.00<br />

CVX 0.51 0.66 0.47 0.04 0.00* 0.09 0.04 0.07 0.03<br />

HPQ 1.10 0.74 0.51 0.09 0.02 0.03 0.08 0.03 0.01<br />

KO 0.70 1.60 0.98 0.10 0.09 1.00 0.08 0.36 0.24<br />

MMM 0.58 0.56 0.09 0.06 0.00 0.06 0.05 0.01 0.00<br />

PG 0.44 0.61 1.20 0.02 0.00* 0.07 0.02 0.00* 0.39<br />

SGP 0.64 0.27 0.58 0.03 0.00 0.00 0.02 0.00* 0.01<br />

TXN 0.30 1.30 0.10 0.00 0.00* 0.08 0.00 0.00 0.00<br />

WAG<br />

condition<br />

0.77<br />

all data<br />

0.85<br />

Y ≥ Y (k)<br />

0.31<br />

Y ≥ Y (k)<br />

0.07<br />

c... k<br />

0.01<br />

Y ≥ Y (k)<br />

0.10 0.06 0.02 0.00<br />

� X ≥ X(k) � ∩ c... k Y ≥ Y (k) � X ≥ X(k) � m=5736<br />

� �<br />

Y ≥ Y (k)<br />

�<br />

∩<br />

�<br />

Y ≥ Y (k)<br />

BMY 0.69 -0.10 0.54 0.05 0.00* 0.12 0.05 0.02 0.01<br />

CVX 0.51 -0.61 0.62 0.04 0.00* 0.09 0.04 0.07 0.03<br />

HPQ 1.10 0.62 0.79 0.09 0.02 0.03 0.09 0.04 0.01<br />

KO 0.70 0.68 1.10 0.10 0.09 1.00 0.11 0.48 0.32<br />

MMM 0.58 0.10 0.36 0.06 0.00 0.06 0.06 0.01 0.00<br />

PG 0.44 -0.17 -0.06 0.02 0.00* 0.06 0.03 0.00* 0.57<br />

SGP 0.64 0.01 -0.11 0.03 0.00 0.00 0.03 0.00* 0.02<br />

TXN 0.30 -0.41 0.09 0.00 0.00* 0.08 0.00 0.00 0.00<br />

WAG 0.77 0.45 0.53 0.07 0.01 0.10 0.08 0.02 0.00<br />

Table 3.19: Results for ˆ βSI using the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k) and the second SI condition: Y ≥ Y (k) for upper and lower tails<br />

calculated by least squares method applied on linear additive single factor model: X = β ·Y +ε and for resulting upper and lower tail dependence<br />

coefficients ˆ λ + and ˆ λ− with corrected ˆl (c... k) and uncorrected ˆl (1... k). The tail represents the most extreme 4% of data during a time interval<br />

ranging from July 1985 to Mars 2008 (k = 229, c = 29). * denotes ˆ β < 0 → λ = 0.


I also implemented the β-smile improvement to the estimation of tail dependence<br />

by the parametric approach according to Sornette & Malevergne discussed in section<br />

(3.2.4). But to estimate ˆ λ by the parametric approach we first need to calculate:<br />

ε = X − β · Y with ε, X, Y as vectors of N elements in chronologic order and scale<br />

factor Cε estimated by: Ĉε = 1<br />

k<br />

� k<br />

j=1<br />

j<br />

N · εν j,N with ε1,N ≥ ε2,N ≥ . . . ≥ εk,N and k<br />

denoting the threshold. The problem now is that when ˆ βSI by one of the SI conditions<br />

is calculated, the yielded ˆ βSI is only valid for tail data that fulfills the respective SI<br />

condition. Therefore values for ε can only be calculated for return values where ˆ βSI is<br />

valid, which is data that fulfills the respective SI condition. Looking at the equation<br />

for the estimation of scale factor Cε we need rank ordered ε-values of the whole data<br />

sample. But return data, which fulfills the SI conditions does not necessarily have<br />

high ε-values and therefore the scale factor Ĉε adapted to the SI conditions is not the<br />

correct one because with ˆ βSI ε cannot be calculated for the whole sample and the<br />

ε-elements that we can calculate do not have to be the most extreme ones. Because<br />

of that application of β-smile improvement to the parametric approach did not yield<br />

meaningful results. That’s why I don’t present it.<br />

Analysis of Coefficients<br />

It is interesting to see the convergence of ˆ βSI(k) to β (calculated for the whole data set)<br />

if we plot it according to the different conditions in dependence of increasing threshold k.<br />

Figures (3.42), (3.43), (3.44), and (3.45) show ˆ βSI(k) plotted for the smaller reference<br />

sample on first, on second, on third, and on fourth condition.<br />

79


β, β SI<br />

β, β SI<br />

80<br />

β, β SI<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

BMY<br />

−1.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

KO<br />

SGP<br />

−0.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

CVX<br />

−0.4<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0<br />

−1<br />

−2<br />

MMM<br />

−3<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

60<br />

40<br />

20<br />

0<br />

TXN<br />

−20<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

2<br />

0<br />

−2<br />

−4<br />

−6<br />

HPQ<br />

−8<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

PG<br />

−1<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

2<br />

1<br />

0<br />

−1<br />

WAG<br />

−2<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

Figure 3.42: ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in black calculated by least square method applied on linear<br />

additive single factor model: X = β ·Y + ε and first SI condition: Y ≥ Y (k) ∩ X ≥ X(k) in dependence of threshold k. Reference ˆ β calculated for<br />

all data is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset return vector of the nine assets for a time interval<br />

ranging from January 1991 to December 2000.


81<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

BMY<br />

0.4<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1.5<br />

1<br />

0.5<br />

0<br />

KO<br />

SGP<br />

−0.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

CVX<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1.5<br />

1<br />

0.5<br />

MMM<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

2<br />

1<br />

0<br />

−1<br />

TXN<br />

−2<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

1.5<br />

1<br />

0.5<br />

0<br />

HPQ<br />

−0.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

2<br />

1.5<br />

1<br />

0.5<br />

PG<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

WAG<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

Figure 3.43: ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in black calculated by least square method applied on linear<br />

additive single factor model: X = β · Y + ε and second SI condition: Y ≥ Y (k) in dependence of threshold k. Reference ˆ β calculated for all data<br />

is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset return vector of the nine assets for a time interval ranging<br />

from January 1991 to December 2000.


β, β SI<br />

82<br />

β, β SI<br />

β, β SI<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

BMY<br />

−1.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

KO<br />

SGP<br />

−1<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

0.5<br />

0<br />

−0.5<br />

CVX<br />

−1<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

MMM<br />

−0.4<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

TXN<br />

−1.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

HPQ<br />

−1.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

PG<br />

−1<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

WAG<br />

−1<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

Figure 3.44: ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in black calculated by least square method applied on linear<br />

additive single factor model: X = β · Y + ε and third SI condition: X ≥ X(k) in dependence of threshold k. Reference ˆ β calculated for all data<br />

is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset return vector of the nine assets for a time interval ranging<br />

from January 1991 to December 2000.


β, β SI<br />

83<br />

β, β SI<br />

β, β SI<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

BMY<br />

−2<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

KO<br />

SGP<br />

−2<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

CVX<br />

−1.5<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

MMM<br />

−1<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

0<br />

−1<br />

−2<br />

−3<br />

TXN<br />

−4<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

β, β SI<br />

β, β SI<br />

β, β SI<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

HPQ<br />

−3<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

PG<br />

−2<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

1<br />

0<br />

−1<br />

−2<br />

WAG<br />

−3<br />

0 0.2 0.4 0.6 0.8 1<br />

k/N<br />

Figure 3.45: ˆ βSI(k) of the upper tails plotted in blue and of the lower tails plotted in black calculated by least square method applied on linear<br />

additive single factor model: X = β · Y + ε and fourth SI condition: Y ≥ Y (k) ∪ X ≥ X(k) in dependence of threshold k. Reference ˆ β calculated<br />

for all data is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset return vector of the nine assets for a time interval<br />

ranging from January 1991 to December 2000.


ˆβSI(k) calculated by the first and the second condition seem to converge best. For<br />

some assets we have strong fluctuations in the extreme tails. Anyway, relative deviations<br />

calculated for k/N = 4% given in table (3.17) agree well with the convergence plots.<br />

To see it in detail I also plotted ˆ βSI for the first and for the third condition just in<br />

the threshold area of extreme tails, which we determined to be around 3% to 5% of<br />

total return data. Figure (3.46) shows ˆ βSI plotted in dependence of threshold k for<br />

k/N = 3% . . .6% calculated by first and third condition with ˆ β calculated for all data<br />

for comparison for upper tails and figure (3.47) shows the same for lower tails both for<br />

the smaller data sample. Figures (3.48) and (3.49) show ˆ βSI calculated by the first<br />

and by the second condition for the upper and for the lower tails of the bigger reference<br />

sample also for k/N = 3% . . .6%. The green line in all of these plots denotes the<br />

reference ˆ β calculated for the whole data sample and is therefore constant on the whole<br />

interval. As we can see, there is no general pattern between ˆ β calculated for all data<br />

and ˆ βSI calculated by first and second SI conditions for the different assets. For some<br />

cases the three are similar and for others not. But in many cases there is something<br />

like a convergence of the ˆ βSI and ˆ β around k/N = 4%. What we observe furthermore is<br />

that in most cases the differences between ˆ βSI calculated by first and second conditions<br />

and ˆ β calculated for all data is within certain limits. It is interesting to see to what<br />

extent ˆ λ +,− is affected by these discrepancies. Therefore I plotted ˆ λ +,− (k) calculated<br />

by the first and by the second conditions for k/N = 3% . . .10%. Figure (3.50) shows<br />

plots of ˆ λ(k) calculated by ˆ βSI according to the first condition: Y ≥ Y (k) ∩ X ≥ X(k)<br />

for upper tails and figure (3.51) for lower tails. Figures (3.52) and (3.53) show plots<br />

of ˆ λ(k) calculated by ˆ βSI according to the second condition: Y ≥ Y (k) for upper and<br />

lower tails. Also here differences between ˆ λ calculated by ˆ βSI plotted in black and<br />

ˆλ calculated by ˆ β plotted in green in most cases are not ’huge’. In the positive tails<br />

Texas Instruments Inc. (TXN), Hewlett Packard Co. (HPQ), and Procter & Gamble<br />

Co. (PG) are exceptions, where we have big differences of ˆ λ calculated using ˆ βSI by<br />

first condition and ˆ λ using ˆ β of all data. In the negative tails we don’t find these big<br />

discrepancies around k/N = 4%. Looking at ˆ λ calculated by the second condition we<br />

find some large discrepancies for asset 3M Co. (MMM) and Bristol-Myers Squibb Co.<br />

(BMY) in the negative tails and in the positive tails we don’t find ’huge’ discrepancies<br />

around k/N = 4%.<br />

84


85<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

BMY<br />

0.4<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

KO<br />

SGP<br />

0.2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

CVX<br />

−0.2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

MMM<br />

−0.4<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

6<br />

4<br />

2<br />

0<br />

TXN<br />

−2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

1.5<br />

1<br />

0.5<br />

HPQ<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1.5<br />

1<br />

0.5<br />

PG<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

WAG<br />

0.8<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

Figure 3.46: ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted in blue and by the second condition: Y ≥ Y (k) plotted in<br />

black for the upper tail calculated by least squares method applied on linear additive single factor model: X = β ·Y +ε in dependence of threshold<br />

k for k/N = 3% ... 6%. Reference ˆ β calculated for all data is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from January 1991 to December 2000.


β SI , β<br />

β SI , β<br />

86<br />

β SI , β<br />

1.5<br />

1<br />

0.5<br />

BMY<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.95<br />

0.9<br />

0.85<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1.5<br />

1<br />

0.5<br />

KO<br />

SGP<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

CVX<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

MMM<br />

0.2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

TXN<br />

−1.5<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

HPQ<br />

−1.5<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

PG<br />

0.4<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

2<br />

1<br />

0<br />

−1<br />

WAG<br />

−2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

Figure 3.47: ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted in blue and by the second condition: Y ≥ Y (k) plotted in<br />

black for the lower tail calculated by least squares method applied on linear additive single factor model: X = β ·Y +ε in dependence of threshold<br />

k for k/N = 3% ... 6%. Reference ˆ β calculated for all data is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from January 1991 to December 2000.


β SI , β<br />

87<br />

β SI , β<br />

β SI , β<br />

1<br />

0.5<br />

0<br />

BMY<br />

−0.5<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

KO<br />

SGP<br />

−0.2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

1<br />

0.5<br />

0<br />

−0.5<br />

CVX<br />

−1<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

MMM<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

TXN<br />

−1<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

HPQ<br />

0.4<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

PG<br />

−0.2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

WAG<br />

0.2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

Figure 3.48: ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted in blue and by the second condition: Y ≥ Y (k) plotted in<br />

black for the upper tail calculated by least squares method applied on linear additive single factor model: X = β ·Y +ε in dependence of threshold<br />

k for k/N = 3% ... 6%. Reference ˆ β calculated for all data is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from July 1985 to Mars 2008.


β SI , β<br />

88<br />

β SI , β<br />

β SI , β<br />

0.65<br />

0.6<br />

0.55<br />

0.5<br />

0.45<br />

0.4<br />

0.35<br />

BMY<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1.3<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1<br />

0.5<br />

0<br />

KO<br />

SGP<br />

−0.5<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

0.65<br />

0.6<br />

0.55<br />

0.5<br />

0.45<br />

CVX<br />

0.4<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

MMM<br />

0<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

TXN<br />

−0.4<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

β SI , β<br />

β SI , β<br />

β SI , β<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

HPQ<br />

0.4<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1.5<br />

1<br />

0.5<br />

0<br />

PG<br />

−0.5<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

WAG<br />

0.2<br />

0.03 0.035 0.04 0.045<br />

k/N<br />

0.05 0.055 0.06<br />

Figure 3.49: ˆ βSI(k) calculated by the first condition: Y ≥ Y (k) ∩ X ≥ X(k) plotted in blue and by the second condition: Y ≥ Y (k) plotted in<br />

black for the lower tail calculated by least squares method applied on linear additive single factor model: X = β ·Y +ε in dependence of threshold<br />

k for k/N = 3% ... 6%. Reference ˆ β calculated for all data is given in green. Y denotes the index return vector of S&P 500 and Y denotes asset<br />

return vector of the nine assets for a time interval ranging from July 1985 to Mars 2008.


As an additional information, ˆν was always taken constant for the whole intervals<br />

and estimated at k/N = 4%. Otherwise when ˆν was adapted to the threshold k, fluctuations<br />

of ˆ λ in k were much stronger and dominated by ˆν. In order to understand<br />

this, we reconsider figures (3.1) showing ˆν for the index S&P 500 and the nine assets<br />

for the positive tails and (3.3) showing ˆν for the index S&P 500 and the nine assets<br />

for the negative tails. In the positive tail of index S&P 500 we have a fluctuation<br />

spectrum ranging from around 3.2 for k/N = 3% to 2.2 for k/N = 10% for the bigger<br />

data sample, which is very high. In the negative tails the situation is similar. Fluctuations<br />

furthermore become stronger as k approaches zero and Hill’s estimator cannot be<br />

trusted anymore because of lack of data in that area. The situation for Gabaix’s b γ n is<br />

slightly better but as I wanted to observe the impact of β in k tail index α was assumed<br />

to be constant for the considered interval of threshold k calculated at k/N = 4%.<br />

We already know that adaptation of ˆ l to threshold k does not change ˆ λ significantly.<br />

This can also be observed in figures (3.50), (3.51), (3.52), and (3.53), where plots in<br />

green show ˆ l(k) and plots in red show ˆ l(k/N = 4%). In our area of interest k/N =<br />

3% . . .5% the two plots in red and green respectively stay close together showing us<br />

that ˆ l can be assumed more or less constant in the tails up to 10% of total data. For<br />

ˆ l(k) we used means that were not corrected for most extreme 0.5%: ˆ l(k) = 1<br />

k<br />

�k Xj,N<br />

j=1 Yj,N .<br />

Anyway the difference between ˆ λ calculated by uncorrected means and ˆ λ calculated by<br />

corrected means is small compared to differences between ˆ λ calculated by the different<br />

β-smile conditions and ˆ λ with ˆ β calculated by all data. Another approach would be<br />

to calculate ˆ l only by data that fulfills the ˆ βSI conditions. I also implemented this<br />

adapted ˆ lSI ’improvement’ for the first ˆ βSI condition. Plots are shown in figures (3.50)<br />

and (3.51) given in blue. Implementing adapted ˆ lSI for the second condition sometimes<br />

yielded negative ˆ l indicating that extreme positive return movements of assets happen<br />

simultaneously with extreme negative return movements of index S&P 500 or vice versa.<br />

This certainly yields meaningless results for ˆ λ. The differences between ˆ λ calculated by<br />

adapted ˆ lSI and ˆ λ calculated by ˆ l were also small. Another indication that ˆ λ is robust<br />

to the choice of ˆ l or that the assumption of constant ˆ l in the tails is appropriate.<br />

We can conclude that β seems to be the crucial parameter within these approaches.<br />

Sometimes the two or three estimates (for different β approaches) are consistent and<br />

for other cases they are not. These latter cases might diagnose the limits of the linear<br />

β model to describe the dependence structure between asset and index.<br />

89


90<br />

λ<br />

λ<br />

λ<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

KO<br />

SGP<br />

0.02<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

CVX<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

MMM<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

TXN<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

HPQ<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

PG<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

WAG<br />

0.05<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.50: ˆ λ + (k) for the upper tail of the nine assets and the index S&P 500 is plotted in black for ˆ β(k) calculated by least squares method<br />

applied to linear additive single factor model: X = β · Y + ε and adapted to the first condition: Y ≥ Y (k) ∩ X ≥ X(k), ˆν(k = 4%) and ˆ l(k) in<br />

dependence of threshold k for k/N = 3% ... 10%, in blue for ˆ β(k) and ˆ l(k) adapted to the first SI condition for comparison, in green for Reference<br />

ˆβ and ˆ l(k), and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index return vector of S&P 500 and X denotes asset return vector of the<br />

nine assets for a time interval ranging from January 1991 to December 2000.


91<br />

λ<br />

λ<br />

λ<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

KO<br />

SGP<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

CVX<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.015<br />

MMM<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.01<br />

0.005<br />

TXN<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

HPQ<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

PG<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

WAG<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.51: ˆ λ + (k) for the lower tail of the nine assets and the index S&P 500 is plotted in black for ˆ β(k) calculated by least squares method<br />

applied to linear additive single factor model: X = β · Y + ε and adapted to the first condition: Y ≥ Y (k) ∩ X ≥ X(k), ˆν(k = 4%) and ˆ l(k) in<br />

dependence of threshold k for k/N = 3% ... 10%, in blue for ˆ β(k) and ˆ l(k) adapted to the first SI condition for comparison, in green for Reference<br />

ˆβ and ˆ l(k), and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index return vector of S&P 500 and X denotes asset return vector of the<br />

nine assets for a time interval ranging from January 1991 to December 2000.


92<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

KO<br />

SGP<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

CVX<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

MMM<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

1<br />

0.5<br />

0<br />

−0.5<br />

TXN<br />

−1<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0.025<br />

0.015<br />

0.005<br />

HPQ<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.03<br />

0.02<br />

0.01<br />

PG<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

WAG<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.52: ˆ λ + (k) for the upper tail of the nine assets and the index S&P 500 is plotted in black for ˆ β(k) calculated by least squares method<br />

applied to linear additive single factor model: X = β · Y + ε and adapted to the second condition: Y ≥ Y (k), ˆν(k = 4%) and ˆ l(k) in dependence<br />

of threshold k for k/N = 3% ... 10%, in blue for ˆ β(k) and ˆ l(k) adapted to the first SI condition for comparison, in green for Reference ˆ β and ˆ l(k),<br />

and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index return vector of S&P 500 and X denotes asset return vector of the nine assets<br />

for a time interval ranging from January 1991 to December 2000.


93<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

BMY<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

KO<br />

SGP<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

CVX<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

MMM<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

2.5<br />

1.5<br />

0.5<br />

x 10−7<br />

3<br />

2<br />

1<br />

TXN<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

λ<br />

λ<br />

λ<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

HPQ<br />

0.04<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

PG<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

WAG<br />

0<br />

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1<br />

k/N<br />

Figure 3.53: ˆ λ + (k) for the lower tail of the nine assets and the index S&P 500 is plotted in black for ˆ β(k) calculated by least squares method<br />

applied to linear additive single factor model: X = β · Y + ε and adapted to the second condition: Y ≥ Y (k), ˆν(k = 4%) and ˆ l(k) in dependence<br />

of threshold k for k/N = 3% ... 10%, in blue for ˆ β(k) and ˆ l(k) adapted to the first SI condition for comparison, in green for Reference ˆ β and ˆ l(k),<br />

and for Reference ˆ λ + (k = 4%) is given in red. Y denotes the index return vector of S&P 500 and X denotes asset return vector of the nine assets<br />

for a time interval ranging from January 1991 to December 2000.


Estimation of Error Bars<br />

To establish uncertainty of upper and lower tail dependence coefficients I included<br />

bootstrap approximations 14 . The sampling has been performed by creating bootstrap<br />

samples of historical day-by-day return data of index S&P 500 and searching for corresponding<br />

asset return data. I always sampled the whole data sets because I think that<br />

this is more representative than just sampling of tails. Tables (3.20) and (3.21) show<br />

quantiles, extreme values, and standard deviations (’std bs’) of the results of bootstrap<br />

sampling of λ +,− for the smaller and bigger reference samples using ˆ βSI calculated by<br />

the first SI condition: {Y ≥ Y (k)} ∩ {X ≥ X(k)} and tables (3.22) and (3.23) show<br />

the results of bootstrap sampling of λ +,− for the smaller and bigger reference samples<br />

using ˆ βSI calculated by the second SI condition: Y ≥ Y (k). ˆ λ +,− were calculated using<br />

the mean values of ˆl = Xj,N/Yj,N for j = 1 . . .k with k = 100 for the smaller sample<br />

and with k = 229 for the bigger sample and ˆ λ +,−<br />

c were calculated using the mean values<br />

of ˆl from c . . .k with c = 13 and k = 100 for the smaller sample and for k = c . . .k with<br />

c = 29 and k = 229 for the bigger sample.<br />

Relative errors of average bootstrap tail dependence coefficients ˆ λ +,−<br />

bs compared to<br />

tail dependence coefficients ˆ λ +,− of the original sample were unusually big in most<br />

cases. If tail dependence coefficient estimates were very small (≈ 10 −3 ) relative errors<br />

could have become multiples of the estimates. Quantiles of the bigger data samples in<br />

some cases look very consistent and are approximately Gaussian distributed. In some<br />

other cases first of all if ˆ βSI are close to zero accuracy of the results become poor as<br />

it can be observed at the example of assets 3M Co. (MMM) and Texas-Instruments<br />

inc. (TXN) in the lower tails of the bigger data set applying the first SI condition<br />

and at the example of Schering & Plough Corp. In the positive tail also of the bigger<br />

data sample applying the first SI condition. 95% quantiles for example are around the<br />

size of the estimated total value of the respective tail dependence coefficient estimates<br />

provided ˆ λ is not too small. For the smaller data set it can be observed that for<br />

small estimates of the tail dependence coefficient by the first as well as the second<br />

condition, quantiles get relatively high. Accuracy here seems to be poor. Overall the<br />

situation concerning bootstrap quantiles of β-smile improvement estimates calculated<br />

by the non-parametric approach according to Sornette & Malevergne, applying the first<br />

or the second condition looks worse compared to the classic non-parametric approach<br />

with ˆ β calculated by all data. What makes the difference is that estimates calculated<br />

by SI conditions at times become very small and therefore relative error bars become<br />

very high. ˆ βSI therefore does not seem to have a negative stake on the error bars<br />

of ˆ λ. Another reason might be that because ˆ βSI shows strong fluctuations in k, it is<br />

a further source of uncertainty that distorts tail dependence estimates. Maybe if we<br />

use much bigger data sets with more data in the tails the estimates would become<br />

more stable. The problem is that this has not much practical application because our<br />

bigger data set already contains more than 5000 daily observations and for many assets<br />

there is not more data available. Looking at the second SI condition the situation is<br />

even worse. Relative errors are mostly within limits but tail dependence coefficient<br />

estimates are very inaccurate. Standard deviations are at least of the size of the total<br />

estimate. Anyways adaptation of β to the tails seems to add too much uncertainty to<br />

the estimates.<br />

14 The matlab m-file for the bootstrap sampling with replacement of λ +,− estimated by the non-parametric<br />

approach according to Sornette & Malevergne with ˆ β calculated by first and second SI conditions is denoted<br />

by: bootstr simple impr alldata.m and enclosed to the data CD<br />

94


95<br />

constr 1 m=2507 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=100 rel err max dev 95% q 90% q std bs k=100 rel err max dev 95% q 90% q std bs<br />

BMY 0.06 0.42 0.93 0.12 0.08 0.08 0.01 1.79 0.52 0.03 0.02 0.02<br />

CVX 0.00 Inf 0.45 0.01 0.00 0.02 0.02 2.61 0.38 0.12 0.03 0.03<br />

HPQ 0.00 9.28 0.77 0.01 0.00 0.03 0.01 Inf 1.02 0.01 0.01 0.06<br />

KO 0.04 3.07 1.04 0.10 0.10 0.09 0.09 1.18 0.61 0.33 0.20 0.13<br />

MMM 0.04 Inf 1.00 0.23 0.04 0.17 0.04 0.55 1.00 0.12 0.06 0.07<br />

PG 0.37 0.21 0.63 0.58 0.55 0.40 0.05 2.58 0.56 0.17 0.10 0.08<br />

SGP 0.11 2.39 0.91 0.42 0.19 0.21 0.01 2.89 0.08 0.02 0.01 0.01<br />

TXN 0.79 0.24 0.84 0.82 0.78 0.40 0.04 Inf 1.01 0.10 0.04 0.21<br />

WAG 0.24 1.04 0.78 0.78 0.45 0.34 0.03 Inf 0.60 0.14 0.04 0.08<br />

λc<br />

BMY 0.06 0.43 0.92 0.12 0.10 0.09 0.02 1.86 0.51 0.05 0.03 0.03<br />

CVX 0.00 Inf 0.60 0.01 0.00 0.02 0.02 2.71 0.34 0.06 0.03 0.04<br />

HPQ 0.00 9.25 0.88 0.01 0.00 0.03 0.01 Inf 1.01 0.01 0.01 0.09<br />

KO 0.04 3.09 1.04 0.11 0.10 0.10 0.13 1.21 0.70 0.40 0.30 0.20<br />

MMM 0.04 Inf 0.96 0.20 0.04 0.19 0.05 0.60 1.00 0.11 0.07 0.08<br />

PG 0.37 0.24 0.61 0.59 0.58 0.40 0.09 2.48 0.66 0.30 0.20 0.10<br />

SGP 0.11 2.41 0.90 0.26 0.20 0.20 0.01 2.67 0.10 0.02 0.01 0.01<br />

TXN 0.79 0.23 0.83 0.81 0.74 0.42 0.04 Inf 1.04 0.10 0.04 0.20<br />

WAG 0.25 1.02 0.81 0.80 0.47 0.31 0.03 Inf 0.78 0.18 0.06 0.09<br />

Table 3.20: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values, and<br />

standard deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time<br />

interval from January 1991 to December 2000. ˆ βj have been calculated on the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k). ’inf’ denotes ·/0.


96<br />

constr 1 m=5736 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=229 rel err max dev 95% q 90% q std bs k=229 rel err max dev 95% q 90% q std bs<br />

BMY 0.15 0.29 0.92 0.31 0.19 0.09 0.01 0.52 0.09 0.01 0.01 0.01<br />

CVX 0.18 1.01 0.78 0.60 0.30 0.21 0.03 0.10 0.29 0.03 0.03 0.03<br />

HPQ 0.04 0.32 0.49 0.12 0.11 0.12 0.01 0.01 0.00 0.01 0.01 0.01<br />

KO 0.65 0.41 0.59 0.58 0.58 0.38 0.21 0.08 0.31 0.20 0.19 0.12<br />

MMM 0.09 0.48 0.91 0.19 0.10 0.09 0.02 201.00 0.78 0.09 0.04 0.06<br />

PG 0.13 1.00 0.92 0.41 0.18 0.19 0.45 0.11 0.61 0.61 0.55 0.37<br />

SGP 0.02 9.03 0.49 0.09 0.05 0.05 0.05 3.02 0.92 0.20 0.08 0.12<br />

TXN 0.16 1.01 0.81 0.57 0.31 0.22 0.00 Inf 0.10 0.01 0.01 0.01<br />

WAG 0.12 0.32 0.90 0.23 0.16 0.10 0.02 6.96 0.57 0.09 0.02 0.07<br />

λc<br />

BMY 0.15 0.28 0.92 0.28 0.21 0.09 0.01 0.56 0.10 0.02 0.01 0.01<br />

CVX 0.18 1.00 0.79 0.57 0.30 0.21 0.03 0.10 0.31 0.04 0.03 0.03<br />

HPQ 0.04 0.29 0.47 0.13 0.11 0.10 0.01 0.03 0.03 0.01 0.01 0.01<br />

KO 0.65 0.40 0.59 0.57 0.56 0.39 0.28 0.09 0.33 0.32 0.20 0.10<br />

MMM 0.09 0.49 0.89 0.21 0.09 0.08 0.03 137.00 1.00 0.10 0.06 0.08<br />

PG 0.13 1.00 0.91 0.43 0.21 0.19 0.54 0.05 0.53 0.51 0.50 0.40<br />

SGP 0.02 9.02 0.52 0.10 0.04 0.05 0.07 3.02 0.91 0.30 0.11 0.19<br />

TXN 0.16 1.01 0.80 0.59 0.29 0.21 0.00 Inf 0.17 0.02 0.01 0.01<br />

WAG 0.12 0.31 0.91 0.26 0.21 0.09 0.03 6.01 0.79 0.11 0.03 0.08<br />

Table 3.21: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values, and<br />

standard deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time<br />

interval from July 1985 to Mars 2008. ˆ βj have been calculated on the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k). ’inf’ denotes ·/0.


97<br />

constr 2 m=2507 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=100 rel err max dev 95% q 90% q std bs k=100 rel err max dev 95% q 90% q std bs<br />

BMY 0.14 0.95 0.88 0.29 0.21 0.22 0.19 0.09 0.82 0.21 0.19 0.10<br />

CVX 0.02 11.03 0.59 0.10 0.03 0.04 0.07 0.20 0.41 0.13 0.12 0.09<br />

HPQ 0.19 0.39 0.80 0.32 0.23 0.20 0.09 0.37 0.94 0.22 0.10 0.11<br />

KO 0.05 2.02 0.73 0.21 0.08 0.09 0.16 0.32 0.77 0.38 0.31 0.20<br />

MMM 0.01 2.96 0.21 0.05 0.03 0.03 0.27 0.03 0.71 0.39 0.18 0.18<br />

PG 0.05 37.95 0.87 0.18 0.09 0.11 0.11 0.48 0.89 0.22 0.19 0.13<br />

SGP 0.10 3.03 0.88 0.39 0.21 0.20 0.12 0.12 0.62 0.21 0.08 0.10<br />

TXN 0.00 Inf 0.81 0.00 0.00 0.03 0.00 Inf 0.04 0.00 0.00 0.00<br />

WAG 0.28 0.68 0.73 0.57 0.38 0.28 0.11 1.81 0.91 0.40 0.19 0.19<br />

λc<br />

BMY 0.15 0.96 0.90 0.40 0.20 0.21 0.27 0.10 0.69 0.32 0.20 0.16<br />

CVX 0.02 11.02 0.58 0.11 0.03 0.04 0.08 0.22 0.68 0.14 0.11 0.10<br />

HPQ 0.19 0.48 0.80 0.30 0.21 0.20 0.14 0.56 0.89 0.28 0.09 0.19<br />

KO 0.05 1.62 0.72 0.22 0.09 0.08 0.23 0.28 0.81 0.61 0.37 0.20<br />

MMM 0.01 2.73 0.19 0.05 0.03 0.03 0.33 0.04 0.70 0.48 0.26 0.23<br />

PG 0.05 40.04 0.89 0.26 0.08 0.12 0.18 0.41 0.79 0.40 0.20 0.21<br />

SGP 0.10 2.47 0.89 0.40 0.22 0.19 0.17 0.10 0.81 0.21 0.09 0.12<br />

TXN 0.00 Inf 0.86 0.00 0.00 0.03 0.00 Inf 0.13 0.00 0.00 0.00<br />

WAG 0.28 0.73 0.72 0.68 0.50 0.31 0.16 1.23 0.80 0.53 0.28 0.18<br />

Table 3.22: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values, and<br />

standard deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time<br />

interval from January 1991 to December 2000. ˆ βj have been calculated on the second SI condition: Y ≥ Y (k). ’inf’ denotes ·/0.


98<br />

constr 2 m=5736 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ k=229 rel err max dev 95% q 90% q std bs k=229 rel err max dev 95% q 90% q std bs<br />

BMY 0.01 Inf 0.18 0.04 0.01 0.02 0.04 0.88 0.39 0.09 0.08 0.05<br />

CVX 0.01 Inf 0.30 0.04 0.01 0.03 0.08 0.21 0.47 0.08 0.07 0.10<br />

HPQ 0.06 2.03 0.68 0.20 0.10 0.09 0.04 0.20 0.21 0.04 0.03 0.02<br />

KO 0.19 1.02 0.79 0.79 0.38 0.30 0.34 0.07 0.70 0.31 0.21 0.10<br />

MMM 0.04 123.95 0.51 0.10 0.09 0.07 0.08 6.03 0.72 0.33 0.23 0.11<br />

PG 0.01 Inf 0.47 0.03 0.01 0.02 0.10 Inf 0.88 0.48 0.22 0.20<br />

SGP 0.02 Inf 1.02 0.07 0.02 0.08 0.04 Inf 0.82 0.21 0.10 0.09<br />

TXN 0.00 Inf 0.21 0.00 0.00 0.01 0.00 Inf 0.12 0.00 0.00 0.00<br />

WAG 0.04 2.01 0.47 0.13 0.07 0.07 0.05 2.01 0.87 0.13 0.05 0.09<br />

λc<br />

BMY 0.01 Inf 0.26 0.04 0.01 0.02 0.05 1.04 0.46 0.10 0.08 0.06<br />

CVX 0.01 Inf 0.31 0.04 0.01 0.03 0.09 0.19 0.47 0.09 0.07 0.06<br />

HPQ 0.06 2.02 0.68 0.20 0.11 0.09 0.05 1.00 0.30 0.07 0.04 0.03<br />

KO 0.19 1.02 0.81 0.81 0.40 0.29 0.45 0.06 0.61 0.32 0.28 0.18<br />

MMM 0.04 123.97 0.52 0.10 0.09 0.07 0.10 6.03 0.88 0.34 0.21 0.20<br />

PG 0.01 Inf 0.49 0.03 0.01 0.03 0.14 Inf 0.89 0.77 0.44 0.31<br />

SGP 0.02 Inf 1.01 0.07 0.02 0.08 0.06 Inf 0.90 0.21 0.20 0.10<br />

TXN 0.00 Inf 0.20 0.00 0.00 0.01 0.00 Inf 0.11 0.01 0.00 0.01<br />

WAG 0.04 1.98 0.49 0.13 0.07 0.07 0.06 2.02 0.92 0.20 0.06 0.08<br />

Table 3.23: Establishing the uncertainty of non-parametrically estimated upper and lower tail dependence coefficients ˆ λ by creating 1000 bootstrap<br />

samples of historical return data tables for S&P 500 index and corresponding asset returns and calculation of quantiles, extreme values, and<br />

standard deviations of the results. The tails represent the most extreme 4% (excluding highest 0.5% for λc) of the return values during a time<br />

interval from July 1985 to Mars 2008. ˆ βj have been calculated on the second SI condition: Y ≥ Y (k). ’inf’ denotes ·/0.


Observation of λ by Rolling Time Window<br />

I wanted to include a rolling time window observation of the β-smile improvement<br />

estimates just to see whether ˆ λ calculated by SI conditions performs similarly as ˆ λ<br />

calculated by ˆ β of the whole sample. Figures (3.54) and (3.55) show ˆ βSI for the window<br />

size of S = 2500 in dependence of time calculated by the first and by the second SI<br />

conditions and figures (3.56) and (3.57) show ˆ λ +,− calculated by ˆ βSI for the window<br />

size of S = 2500 in dependence of time by the first and by the second SI conditions.<br />

As we can see fluctuations of ˆ λ are quite strong and a window size of S = 2500 seems<br />

to be the minimum size to obtain more or less reasonable estimates. Choosing smaller<br />

window sizes yields instable results for both ˆ βSI and ˆ λ. It is interesting to observe that<br />

whereas ˆ βSI calculated by first condition seems to be high in the early period up to<br />

N = 3000 for most assets, ˆ βSI calculated by the second condition is mostly negative<br />

in that area. The spectron of fluctuations on the interval (−1, 1.5) seems to be similar<br />

for both ˆ βSI. Hewlett-Packard Co. (HPQ) is a special case because it shows for both<br />

conditions a significant difference between ˆ βSI calculated for the upper tail and ˆ βSI<br />

calculated for the lower tail. The case, where the linear dependences of asset Xj and<br />

index Yj are different in the positive and the negative tail, is an aspect that ˆ β calculated<br />

by all data is not able to display.<br />

What furthermore can be observed in all of the four figures is that the overall trend<br />

for my shown ˆ βSI and corresponding ˆ λ is the change in time of the positive tails<br />

dominating over negative tail to a tendency of negative tails to dominate the positive<br />

tails. This counts primarily for the second condition with assets Hewlett-Packard Co.<br />

(HPQ) and Texas Instruments Inc. (TXN) being exceptions that show a counter-trend.<br />

Looking at figure (3.57) showing ˆ λ calculated by the second SI condition, it seems that<br />

both upper and lower tail dependence have increased over time.<br />

I also applied bootstrap sampling over time to ˆ λ +,− calculated by first and second<br />

β-smile conditions in order to see whether my impression of unusual high quantiles is<br />

consistent over time 15 . This seems to be the case. Assets with more or less limited<br />

quantile deviations show this behavior over time and others remain inaccurate all the<br />

time. Furthermore it seems that there are windows were the quantiles become much<br />

bigger and then after some time get back to their normal extent.<br />

It is a pity that estimates of ˆ λ calculated by β-smile conditions are not more robust<br />

because the approach is very promising. Nevertheless the β-smile improvement conditions<br />

can be helpful as a complement to other approaches because they allow for new<br />

insights into the structure of extreme dependence of an asset and the index, representing<br />

the market factor, particularly in terms of showing differences between negative and<br />

positive tails. The classic non-parametric approach describes the dependence of asset<br />

and index by a linear relation that is assumed to be constant over the whole range of<br />

the distribution. This assumption might be to restrictive in some cases. In order to find<br />

and to describe these discrepancies by an advanced concept with the aspect of being<br />

more general in describing dependence structures, the β-smile improvement conditions<br />

can be applied.<br />

15 Results of bootstrap sampling of the rolling time window estimates for first and second SI conditions with<br />

window size S = 2500 data points can be observed on the data CD submitted with the thesis in the folder:<br />

window results<br />

99


100<br />

β<br />

β<br />

β<br />

1.5<br />

1<br />

0.5<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

KO<br />

SGP<br />

−1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

β<br />

β<br />

β<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

CVX<br />

−0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.5<br />

1<br />

0.5<br />

0<br />

MMM<br />

−0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

6<br />

4<br />

2<br />

0<br />

−2<br />

TXN<br />

−4<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

β<br />

β<br />

β<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

HPQ<br />

−1.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

PG<br />

−0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2<br />

1<br />

0<br />

−1<br />

WAG<br />

−2<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.54: ˆ βSI(N) plotted in black for the upper tails and plotted in green for the lower tails for 9 assets given by X and index S&P 500 given<br />

by Y using the linear single factor model: X = β · Y + ε and the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k) for rolling time horizon windows of<br />

S = 2500 considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to Mars 2008.


101<br />

β<br />

β<br />

β<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

BMY<br />

−1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

KO<br />

SGP<br />

−1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

β<br />

β<br />

β<br />

2<br />

1<br />

0<br />

−1<br />

CVX<br />

−2<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

MMM<br />

−1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2<br />

1<br />

0<br />

−1<br />

TXN<br />

−2<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

β<br />

β<br />

β<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

HPQ<br />

−0.5<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

PG<br />

−1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

2<br />

1.5<br />

1<br />

0.5<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.55: ˆ βSI(N) plotted in black for the upper tails and plotted in green for the lower tails for 9 assets given by X and index S&P 500 given<br />

by Y using the linear single factor model: X = β · Y + ε and the second SI condition: Y ≥ Y (k) for rolling time horizon windows of S = 2500<br />

considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to Mars 2008.


102<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

KO<br />

SGP<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

CVX<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

MMM<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

HPQ<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

PG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.56: λ + (N) for upper tails of index S&P 500 and the nine assets plotted in blue and λ − (N) for lower tails plotted in red using non-<br />

parametric approach given by equation ˆ λ +,− = 1/max<br />

�<br />

1, l<br />

ˆβ +,−<br />

SI<br />

� ˆν<br />

with coefficients ˆν, l, and ˆ β +,−<br />

SI for rolling time horizon windows of S = 2500<br />

considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to Mars 2008. ˆ β +,−<br />

SI were<br />

calculated using the linear single factor model: X = β · Y + ε and the first SI condition: Y ≥ Y (k) ∩ X ≥ X(k).


103<br />

λ<br />

λ<br />

λ<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

KO<br />

SGP<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

CVX<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

MMM<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

HPQ<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

PG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

N<br />

Figure 3.57: λ + (N) for upper tail of index S&P 500 and the nine assets plotted in blue and λ − (N) for lower tail plotted in red using non-parametric<br />

approach given by equation ˆ λ +,− = 1/max<br />

�<br />

l 1, ˆβ +,−<br />

�ˆν SI<br />

with coefficients ˆν, l, and ˆ β +,−<br />

SI for rolling time horizon windows of S = 2500 considered<br />

data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1...5736) or a total time interval from July 1985 to Mars 2008. ˆ β +,−<br />

SI were calculated<br />

using the linear single factor model: X = β · Y + ε and the second SI condition: Y ≥ Y (k).


3.4 Non-Parametric Approaches according to<br />

Schmidt & Stadtmüller<br />

Within this section several concepts of non-parametric approaches for the estimation<br />

of tail dependence according to Schmidt & Stadtmüller published in [24] (2005) are<br />

discussed. In the first sub-section a presentation of the most important concepts and<br />

assumptions of the paper is provided. Then the second sub-section again is devoted to<br />

the implementation of the concepts and in the third sub-section results are analyzed.<br />

3.4.1 Theoretical Background<br />

In [24] (2005) a set of non-parametric estimators is proposed for the upper and lower<br />

tail copula ΛU(x, y) and ΛL(x, y) (x, y) ′ ∈ R 2 +. Non-parametric estimation is used if<br />

no general finite-dimensional parametrization of tail copulas exists.<br />

Let Cm denote the empirical copula defined by:<br />

Cm(u, v) = Fm(G −1<br />

m (u), H −1<br />

m (v)), (u, v) ′ ∈ [0, 1] 2<br />

(3.77)<br />

with Fm, Gm, Hm being the empirical distribution functions corresponding to F, G, H.<br />

Analogously we define the ’empirical survival copula’ by: Cm(u, v) = F m(G −1<br />

m (u), H−1 m (v)),<br />

(u, v) ′ ∈ [0, 1] 2 with<br />

F m(x, y) = 1<br />

m<br />

m�<br />

j=1<br />

1 {X (j) >x,Y (j) >y}<br />

(3.78)<br />

and Gm = 1 − Gm, Hm = 1 − Hm. Let R (j)<br />

m1 and R (j)<br />

m2 denote the rank of X (j) and<br />

Y (j) , j = 1, . . .,m respectively. A set of estimators is given by:<br />

ˆΛL,m(x, y) := m<br />

k<br />

ˆΛU,m(x, y) := m<br />

k<br />

Cm( kx<br />

m<br />

Cm( kx<br />

m<br />

ky 1<br />

, ) ≈<br />

m k<br />

ky 1<br />

, ) ≈<br />

m k<br />

m�<br />

j=1<br />

m�<br />

j=1<br />

1 (j)<br />

{R m1≤kx and R(j) ms≤ky}<br />

1 (j)<br />

{R m1 >m−kx and R(j) ms>m−ky}<br />

(3.79)<br />

(3.80)<br />

with some parameter k ∈ {1, . . .,m} chosen by the statistician. The estimators<br />

ˆΛU,m(x, y) and ˆ ΛL,m(x, y) are referred to as ’empirical tail copulas’. For the asymptotic<br />

result we assume that k = k(m) → ∞ and k(m) → 0 as m → ∞.<br />

A related estimator was introduced by Huang in [25] (1992), [26] (1998), and [27]<br />

(1998). The relation between the upper tail copula and the stable tail dependence l is<br />

given by<br />

ΛU(x, y) = x + y − l(x, y). The Corresponding estimator for ΛL(x, y) on R2 + is:<br />

ˆΛ<br />

EV T<br />

L,m (x, y) = x + y − m<br />

�<br />

k<br />

≈ x + y − 1<br />

k<br />

Cm<br />

m�<br />

j=1<br />

� ��<br />

kx ky<br />

,<br />

m m<br />

1 (j)<br />

{R m1≤kx or R(j)<br />

m2≤ky}, (x, y) ∈ R2 + (3.81)<br />

104


and the corresponding estimator for ΛU(x, y) on R 2 + is:<br />

ˆΛ<br />

EV T<br />

U,m<br />

� � ��<br />

m kx ky<br />

(x, y) = x + y − Cm ,<br />

k m m<br />

≈ x + y − 1<br />

k<br />

m�<br />

j=1<br />

1 (j)<br />

{R m1 >m−kx or R(j)<br />

m2 >m−ky}, (x, y) ∈ R2 + (3.82)<br />

with k = k(m) → ∞ and k(m) → 0 as m → ∞. An important practical problem<br />

again arises on the optimal choice of the threshold parameter k, which leads to the<br />

usual variance-bias problem. Based on the above estimates of the lower and upper tail<br />

copula, for the upper tail dependence coefficient<br />

ˆλU,m := ˆ ΛU,m(1, 1), and ˆEV T<br />

λ<br />

EV T<br />

U,m := ˆ ΛU,m (1, 1) (3.83)<br />

are proposed as non-parametric estimators and analogous estimates for the lower tail<br />

dependence coefficient.<br />

3.4.2 Implementation of Non-Parametric Approaches<br />

It is the purpose of this sub-section to provide an insight into details concerning the<br />

implementation of the above introduced concepts into Matlab m-files and apply the<br />

function handles to our reference data sets of historical daily return data of the index<br />

S&P 500 and the nine assets for the calculation of upper and lower tail dependence coefficient<br />

estimates. Results can then be compared to further calculations, i.e. estimations<br />

of tail dependence λ +,− according to Sornette & Malevergne.<br />

Calculation of general Rank-Order Statistics<br />

For the calculation of tail copulas ˆ ΛU,m, ˆ ΛL,m, ˆEV T ΛU,m , and ˆEV T ΛL,m between the index S&P<br />

500 and an asset as described in (3.79), (3.80), (3.81), and (3.82) we need rank-ordered<br />

statistics of index returns X (j) and asset returns Y (j) denoted by R (j)<br />

m1 and R (j)<br />

m2 with<br />

j = 1, . . .,m. Then we define a count variable, which passes through the rank vectors<br />

checking for the required conditions given by Boolean ’and’ and ’or’.<br />

Given daily return data for any asset as from January 1991 to December 2000<br />

(m=2506) sorted in descending order raises the question, how to rank equal returns.<br />

The simplest way to calculate R (j)<br />

m,1 is to write an algorithm that first creates a returndata<br />

table sorted in descending order from which it top down picks out values and<br />

searches for equivalents in the historical return data table and then displaces the found<br />

equivalents by increasing rank numbers. Equal return values receive equal rank, described<br />

by natural numbers, whereas the step size always equals one. Looking at the<br />

resulting rank tables for different assets, we observe that maximal rank numbers vary.<br />

Therefore we would have to adapt threshold k to the different assets.<br />

Another way to calculate R (j)<br />

m1 and R (j)<br />

m2, which after my opinion is more reasonable,<br />

is to adapt the rank numbers to the weight of values, described by the number of equal<br />

return values. This means that equal return values still receive equal ranks but larger<br />

step sizes to preceding and successive ranks indicating the number (weight) of equal<br />

return values. To achieve this, I included the following relation to the above described<br />

code:<br />

105


R eq<br />

m,· =<br />

� n<br />

j=1 Rpre<br />

m,· + j<br />

, (3.84)<br />

n<br />

where R pre<br />

m,· denotes the rank previous to the set of n equal return-values ranked by R eq<br />

m,·<br />

This has the advantage that for each return table with equal amount of data measured<br />

simultaneously (i.e. m = 2506), maximal ranks are equal to m, which leaves threshold<br />

k consistent with all samples. To have equal k for different assets and index S&P 500<br />

is a big advantage in terms of the implementation because in all of the four equations<br />

(3.79), (3.80), (3.81), and (3.82) tail copulas are estimated by the summing up of<br />

indicator functions by certain conditions and then by division of the resulting sum by<br />

threshold k. In case of different k for asset and index S&P 500, we would have to divide<br />

by the mean of km,1 and km,2. This would lead to distortion of the estimate.<br />

To check for consistency of my results I added a small noise of amplitude (10 −6 )<br />

to all my returns that none of them were equal anymore. In this way I removed the<br />

degeneracy evolved from the little modification described by equation (3.84). The<br />

results remained consistent up to the decimal places that were of interest. Therefore I<br />

can be pretty sure that there is no significant distortion caused by this modification.<br />

Table (3.24) shows the results of tail dependence coefficients ˆ λU,m, ˆ λL,m, ˆEV T λU,m , and<br />

ˆλ EV T<br />

L,m of the index S&P 500 and the nine assets calculated for the shorter reference<br />

period and table (3.25) for the longer reference period, both at a threshold k equal to<br />

4% of total data m. Results achieved by adding some random noise are also shown for<br />

comparison 16 . As mentioned I analysed exactly the same data as for the calculations<br />

according to Sornette & Malevergne but decided not to present a version that neglects<br />

most extreme values because with the actual approaches quantitative return data does<br />

not affect the result, as by definition of the estimations based on the findings of Schmidt<br />

& Stadtmüller we only consider rank vectors representing relative magnitudes within<br />

historical return tables. Therefore extreme outliers don’t distort the results at all.<br />

To summarize for both upper and lower tails: upper and lower <strong>Tail</strong> dependence coefficients<br />

according to Schmidt & Stadtmüller were calculated by the following relations:<br />

and<br />

ˆλL,m = 1<br />

k<br />

ˆλU,m = 1<br />

k<br />

ˆλ<br />

EV T<br />

L,m<br />

ˆλ<br />

EV T<br />

U,m<br />

m�<br />

j=1<br />

m�<br />

j=1<br />

= 2 − 1<br />

k<br />

= 2 − 1<br />

k<br />

1 (j)<br />

R m1≤k and R(j)<br />

m2≤k 1 (j)<br />

R m1 >m−k and R(j)<br />

m2 >m−k<br />

m�<br />

j=1<br />

m�<br />

j=1<br />

1 (j)<br />

R m1≤k or R(j)<br />

m2≤k 1 (j)<br />

R m1 >m−k or R(j)<br />

m2 >m−k<br />

16 The matlab m-file for the implementation of λU,m, λL,m, λ EV T<br />

(3.85)<br />

(3.86)<br />

(3.87)<br />

(3.88)<br />

U,m , and λL,m estimated by approaches<br />

according to Schmidt & Stadtmüller is denoted by: Lambda Schmidt.m and enclosed to the appendix and to<br />

the data CD<br />

106<br />

EV T


ˆλU,m<br />

ˆλL,m<br />

ˆλ EV T<br />

U,m<br />

ˆλ EV T<br />

L,m<br />

m=2507<br />

k 100.28 100.28 100.28 100.28<br />

w. noise w. noise w. noise w. noise<br />

BMY 0.19 0.19 0.21 0.21 0.18 0.19 0.22 0.21<br />

CVX 0.18 0.17 0.20 0.20 0.17 0.17 0.21 0.20<br />

HPQ 0.25 0.25 0.27 0.27 0.23 0.25 0.27 0.27<br />

KO 0.27 0.26 0.23 0.23 0.25 0.26 0.23 0.23<br />

MMM 0.16 0.15 0.24 0.24 0.15 0.15 0.24 0.24<br />

PG 0.22 0.21 0.19 0.19 0.21 0.21 0.20 0.19<br />

SGP 0.14 0.13 0.23 0.23 0.13 0.13 0.23 0.23<br />

TXN 0.11 0.11 0.09 0.09 0.10 0.11 0.10 0.09<br />

WAG 0.19 0.19 0.27 0.27 0.18 0.19 0.27 0.27<br />

Table 3.24: Estimated values of upper and lower tail dependence for index S&P 500 and a set<br />

of nine major assets traded on the New York stock Exchange calculated by non-parametric<br />

approaches according to Schmidt & Stadtmüller. The tail represents the most extreme 4% of<br />

the return values during a time interval from January 1991 to December 2000 with a threshold<br />

value of k = 0.04 · m = 100.3. Columns denoted by ’w. noise’ correspond to columns on their<br />

right and are performed by the same approaches but with a small random noise of amplitude<br />

(10 −6 ) added to the returns that no two returns are equal anymore.<br />

whereas m denotes the size of the return sample i.e. number of daily observations, k<br />

again denotes the threshold typically assumed as 4% of m, and R (j)<br />

m1 and R (j)<br />

m2 with<br />

j = 1, . . .,m represent rank ordered statistics of index returns Y (j) and asset returns<br />

X (j) calculated by: R eq<br />

m,· =<br />

�n j=1 Rpre m,·+j<br />

denoting the ranks of equal returns and R pre<br />

m,·<br />

to the set of considered equal returns.<br />

3.4.3 Analysis of Coefficients<br />

n<br />

if we have n equal return values with R eq<br />

m,·<br />

denoting the rank of the return(s) previous<br />

Looking at equations (3.85), (3.86), (3.87), and (3.88) we can observe that ˆ λU,m, ˆ λL,m,<br />

ˆλ EV T<br />

U,m , and ˆEV T λL,m all depend on threshold k, which is an indicator of what is assumed<br />

to represent the extreme tail of a certain data set. In our result tables, table (3.24)<br />

and table (3.25) we assumed the most extreme 4 % of data ordered by increasing or<br />

decreasing values to constitute the extreme tail. It might now be interesting to check for<br />

sensitivity of my tail dependence coefficient estimates to different choices of threshold<br />

k. For this purpose I plotted ˆ λL,m, ˆEV T λL,m , ˆ λU,m, and ˆEV T λU,m in dependence of k on an<br />

interval ranging from 0 % up to 6 % of total data m. Figure (3.58) shows the two plots<br />

at the example of the nine assets for the upper tails and figure (3.59) for the lower tails.<br />

It is interesting to observe that ˆ λL,m seems to be a lower bound of ˆEV T λ<br />

L,m and ˆ λU,m seems<br />

to be a upper bound of ˆEV T λU,m . Furthermore ˆ λL,m and ˆ λU,m seem to be less sensitive to<br />

small changes of threshold k than ˆEV T λL,m and ˆEV T λU,m . In fact when I choose step sizes<br />

equal to my available data (low data density), that means without scaling of my x-axis<br />

in terms of refining the k-density, then the curves of ˆEV T λU,m and ˆ λU,m become nearly<br />

identical. This holds also for negative tails and might be an indication for the Boolean<br />

107


ˆλU,m<br />

ˆλL,m<br />

ˆλ EV T<br />

U,m<br />

ˆλ EV T<br />

L,m<br />

m=5736<br />

k 229.40 229.40 229.40 229.40<br />

w. noise w. noise w. noise w. noise<br />

BMY 0.22 0.22 0.28 0.28 0.22 0.22 0.28 0.28<br />

CVX 0.17 0.17 0.27 0.27 0.17 0.17 0.27 0.27<br />

HPQ 0.28 0.28 0.29 0.29 0.28 0.28 0.29 0.29<br />

KO 0.27 0.27 0.28 0.28 0.26 0.27 0.29 0.28<br />

MMM 0.25 0.25 0.27 0.27 0.25 0.25 0.27 0.27<br />

PG 0.19 0.19 0.20 0.20 0.19 0.19 0.20 0.20<br />

SGP 0.17 0.17 0.26 0.26 0.17 0.17 0.27 0.26<br />

TXN 0.17 0.17 0.16 0.16 0.17 0.17 0.17 0.16<br />

WAG 0.21 0.21 0.29 0.29 0.20 0.21 0.29 0.29<br />

Table 3.25: Estimated values of upper and lower tail dependence for index S&P 500 and a set<br />

of nine major assets traded on the New York stock Exchange calculated by non-parametric<br />

approaches according to Schmidt & Stadtmüller. The tail represents the most extreme 4%<br />

of the return values during a time interval from July 1985 to April 2008 with a threshold<br />

value of k = 0.04 · m = 229.4. Columns denoted by ’w. noise’ correspond to columns on their<br />

right and are performed by the same approaches but with a small random noise of amplitude<br />

(10 −6 ) added to the returns that no two returns are equal anymore.<br />

’or’ criterion applied to real financial data to exhibit higher fluctuations compared to<br />

the Boolean ’and’ criterion.<br />

108


EVT<br />

λ , λ<br />

U,m U,m<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

109<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

BMY<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

KO<br />

SGP<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

CVX<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

MMM<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/m<br />

EV T<br />

Figure 3.58: ˆ λU,m for upper tails of index S&P 500 and the nine assets plotted in black and ˆ λ<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

TXN<br />

U,m<br />

EVT<br />

λ , λ<br />

U,m U,m<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

HPQ<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

PG<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

WAG<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

for upper tails plotted in red, both in dependence<br />

of threshold k on an interval from k/m = 0% ... 6% for m = 2507 daily return values during a time interval ranging from January 1991 to December<br />

2000.


110<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

BMY<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

KO<br />

SGP<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

CVX<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

MMM<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

TXN<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

k/m<br />

EV T<br />

L,m<br />

Figure 3.59: ˆ λL,m for lower tails of index S&P 500 and the nine assets plotted in black and ˆ λ<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

EVT<br />

λ , λ<br />

L,m L,m<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

HPQ<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

PG<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

WAG<br />

0<br />

0 0.01 0.02 0.03<br />

k/m<br />

0.04 0.05 0.06<br />

for lower tails plotted in red, both in dependence<br />

of threshold k on an interval from k/m = 0% ... 6% for m = 2507 daily return values during a time interval ranging from January 1991 to December<br />

2000.


Estimation of optimal Threshold k<br />

To estimate the optimal threshold k for my tail dependence estimates according to<br />

Schmidt & Stadtmüller it is proposed in chapter seven of paper [24] (2005) to use a<br />

simple plateau finding algorithm after smoothing the plots of ˆ λL,m(k), ˆEV T λL,m (k), ˆ λU,m(k),<br />

and ˆEV T λU,m (k) by some box kernel. To make the results comparable to my further<br />

estimates, I wanted to choose a threshold value k around 4% of total data. Therefore I<br />

first wanted to smooth out my function of estimated tail dependence in dependence of<br />

threshold k by using a kernel with a bell-type shape and unit integral (i.e. the function<br />

is normalized). Basically, I wanted to replace a data point by the function centered on<br />

its abscissa times an amplitude equal to the ordinate of that data point by choosing a<br />

Gaussian or just a window function (zero outside of an interval and constant inside the<br />

window). The general formula of the approach is as follows:<br />

λsmoothed(k) = ˆ f(k) · λ(k) (3.89)<br />

n�<br />

� �<br />

with f(k) ˆ<br />

1 k − kj<br />

= K<br />

(3.90)<br />

nh h<br />

The kernel estimator ˆ f with kernel K is a sum of boxes or ’bumps’ placed at the<br />

observations. If K satisfies the condition: � ∞<br />

−∞ K(k)dk = 1, ˆ f represents a probability<br />

density function of my estimates that weights the observations with window width h<br />

and n denoting the number of observations.<br />

After some tries I realized that whatever width of the filter I used, the supposedly<br />

smooth function remained sharp toothed. The problem is that the local density of data<br />

in my case is constant over the whole plot interval. ˆ f by definition only considers values<br />

of the x-axis for the weighting. Because these values in my case are equally distributed,<br />

the only solution of ˆ f, if the parameter of window width is adapted properly, could be<br />

equal to 1 and therefore λsmoothed becomes equal to λ indicating that this smoothing<br />

approach has no effect.<br />

Therefore I think that we should choose another smoothing method: A locally<br />

weighted scatter plot smooth using least squares quadratic polynomial fitting yielded<br />

the best results. It has a strong smoothing effect and still leaves the plots accurate<br />

enough. The regression weights for each data point in the span d(k) are given by the<br />

� � �<br />

� �<br />

tricube function: wi = 1− � 3�3 , whereas a second degree least squares regression<br />

was performed:<br />

S =<br />

n�<br />

j=1<br />

r 2 j =<br />

� k−ki<br />

d(k)<br />

j=1<br />

n�<br />

wj(yj − ˆyj) 2 with ˆyj = p1 · k 2 j + p2 · kj + p3 (3.91)<br />

j=1<br />

Figure (3.60) shows smoothing of ˆ λL,m(k) and ˆ λU,m(k) for all assets and the smaller<br />

interval and figure (3.61) for the bigger interval by setting d(k) = 0.125 of the considered<br />

interval k/m = 0% . . .8% or in my case of constant local data density a span of<br />

25 values (out of 2507) for the smaller sample and a span of 57 values (out of 5736)<br />

for the bigger sample. k = 4% seems to be a good choice for the thresholds also for<br />

estimates according to Schmidt & Stadtmüller. Most of the plots show stiff behaviour<br />

betwenn k/m = 3% . . .5%. Therefore and to be able to compare the results estimated<br />

by methods according to Sornette & Malevergne I decided to stick with k = 4%.<br />

111


λ BMY<br />

λ KO<br />

112<br />

λ SGP<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.02 0.04 0.06 0.08<br />

k/m<br />

λ CVX<br />

λ MMM<br />

λ TXN<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

0 0.02 0.04 0.06 0.08<br />

k/m<br />

λ HPQ<br />

λ PG<br />

λ WAG<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.02 0.04 0.06 0.08<br />

Figure 3.60: Smoothed ˆ λL,m plotted in black and smoothed ˆ λU,m plotted in green for the nine assets with the index S&P 500 given for the most<br />

extreme 8% during a time interval from January 1991 to December 2000. Smoothing is performed by locally weighted scatter plot smooth using<br />

least squares quadratic polynomial fitting filters with 25 considered data points by step or [0.125 · 0.08 · m] values with a total of m = 2507 data<br />

points and [·] denoting integer numbers.<br />

k/m


λ BMY<br />

λ KO<br />

113<br />

λ SGP<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

0 0.02 0.04 0.06 0.08<br />

k/m<br />

λ CVX<br />

λ MMM<br />

λ TXN<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

−0.05<br />

0 0.02 0.04 0.06 0.08<br />

k/m<br />

λ HPQ<br />

λ PG<br />

λ WAG<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

0 0.02 0.04<br />

k/m<br />

0.06 0.08<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.02 0.04 0.06 0.08<br />

Figure 3.61: Smoothed ˆ λL,m plotted in black and smoothed ˆ λU,m plotted in green for the nine assets with the index S&P 500 given for the most<br />

extreme 8% during a time interval from July 1985 to April 2008. Smoothing is performed by locally weighted scatter plot smooth using least<br />

squares quadratic polynomial fitting filters with 57 considered data points by step or [0.125 · 0.08 · m] values with a total of m = 5736 data points<br />

and [·] denoting integer numbers.<br />

k/m


1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 1 2 3 4 5 6 7 8<br />

θ<br />

λ L<br />

0.2<br />

0.18<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

0 1 2 3 4 5 6 7 8<br />

Figure 3.62: Lower tail dependence λL(θ) = 2−1/θ and corresponding asymptotic variance<br />

σ2 L (θ) = 2−1/θ − 3<br />

24−1/θ + 1<br />

28−1/θ for the Pareto copula.<br />

Estimation of Error Bars<br />

Schmidt & Stadtmüller proposed in chapter (5) of paper [24] (2005) a method to approximate<br />

the unknown asymptotic variance of tail dependence estimates obtained by<br />

their methods. They choose the Pareto copula as a simple but flexible parametric copula,<br />

calculated its tail copula, and utilized the corresponding variance functional σ 2 L (θ)<br />

as an approximation of the unknown asymptotic variance. The tail copula of the Pareto<br />

copula is given by:<br />

C(u, v) = max � [u −θ + v −θ − 1] −1/θ , 0 � , θ ∈ [−1, ∞) \ 0.<br />

Further the lower tail copula exists for θ > 0 and can be expressed by:<br />

The asymptotic variance was given by:<br />

ΛL(x, y) = (x −θ + y −θ ) −1/θ .<br />

σ 2 L(θ) = 2 −1/θ − 3<br />

2 4−1/θ + 1<br />

2 8−1/θ , (3.92)<br />

where θ was replaced by the maximum likelihood estimator ˆ θ.<br />

It can be shown that the Pareto copula is lower tail dependent with lower tail<br />

dependence coefficient λL = ΛL(1, 1) = 2−1/θ for θ > 0, (i.e. page 77 of paper [28]<br />

(2008)). Given our estimated results of ˆ λU,m and ˆEV T λU,m , σ2 L (ˆ θ) can be calculated. Figure<br />

(3.62) shows plots of λL(θ) and σ2 L (θ).<br />

Looking at the result table for my tail dependence estimates according to Schmidt &<br />

Stadtmüller in tables (3.24) and (3.25) we observe that all of the estimates are between<br />

0.1 and 0.3, which yields ˆ log(2)<br />

θ = −log{λ∈(0.1,0.3)}<br />

∈ (0.3, 0.6) and finally σ2 L (ˆ θ) ∈ (0.08, 0.18).<br />

114<br />

θ<br />

2<br />

σ<br />

L


The right hand side of figure (3.62) shows that this is exactly the area where σ 2 L (θ)<br />

peaks. To achieve results of higher accuracy, meaning that the variance functional<br />

σ 2 L (ˆ θ) becomes lower, we would need very high estimates of lower tail dependence<br />

coefficients ˆ λL,m. Even if our estimates were around 0.7 we would still calculate standard<br />

deviations of around 50% of the estimates by the variance functional. This, as we will<br />

see shortly, would still be about twice the amount of bootstrap sampling standard<br />

deviation estimates. Therefore I don’t think that this approximation approach has<br />

meaningful appliance to my data.<br />

Now I come to the estimation of error bars by the bootstrap method. I built bootstrap<br />

samples of S&P 500 index return data Y1,m, Y2,m, . . .,Ym,m, where m denotes the<br />

total number of values, and searched for the asset return data corresponding in time.<br />

For the smaller sample consisting of m = 2507 data points I built 1000 bootstrap samples<br />

and for the bigger sample consisting of m = 5736 data points I built 800 bootstrap<br />

samples because this already required considerable computing power. Table (3.26)<br />

shows results of bootstrap sampling for the smaller data set and table (3.27) shows the<br />

results for the bigger data set 17 .<br />

Relative errors of the mean values of bootstrap sample estimates compared to original<br />

sample estimates denoted by ’rel err’ were small for each asset by both approaches ˆ λ<br />

and λ ˆ EV T : the biggest relative error was calculated for λbs,mean of the upper tail of<br />

asset Minnesota Mining & MFG (MMM) shown in table (3.26) for smaller sample<br />

bootstrap estimates. It showed 12 % relative deviation from the real value. Performing<br />

this calculation again yielded similar values. Here again the assumption of errors to<br />

be approximately Gaussian distributed seems to hold as calculated standard deviations<br />

within and between the different bootstrap samples are very close of being equal to<br />

half of 95% quantiles. This holds for both sample sizes. The standard deviations<br />

between the bootstrap sample estimates of the smaller sample were found to be around<br />

18% up to maximally 30% of total tail dependence estimates. The lower the tail<br />

dependence estimates the less accurate were the quantiles and standard deviations. For<br />

the bigger sample relative deviations were significantly lower. Here standard deviations<br />

were around 10% to 15% of total tail dependence estimates.<br />

The results obtained by approaches according to Schmidt & Stadtmüller seem to be<br />

very consistent. Tables (3.24) and (3.25) also show that, compared to other methods,<br />

estimates are close to each other, whereas estimates for lower tails tend to be higher<br />

than estimates for upper tails like it used to be the case for the non-parametric and<br />

the parametric approach according to Sornette & Malevergne applied to the smaller<br />

data set and using corrected ˆl and Ĉ given by ˆlc and Ĉc. This also holds for the nonparametric<br />

approach according to Poon, Rockinger, and Tawn using corrected ˆ L given<br />

by ˆ Lc.<br />

17 The matlab m-file for the bootstrap sampling with replacement of λU,m, λL,m, λ EV T<br />

U,m , and λL,m estimated<br />

by approaches according to Schmidt & Stadtmüller is denoted by: bootstr Schmidt.m and enclosed to the data<br />

CD<br />

115<br />

EV T


116<br />

m=2507 bs=1000<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ·,m k=100.28 rel err max dev 95% q 90% q std bs k=100.28 rel err max dev 95% q 90% q std bs<br />

BMY 0.20 0.03 0.11 0.07 0.06 0.04 0.21 0.00 0.19 0.08 0.07 0.04<br />

CVX 0.18 0.01 0.09 0.08 0.06 0.04 0.19 0.02 0.20 0.07 0.06 0.04<br />

HPQ 0.25 0.01 0.10 0.08 0.07 0.04 0.27 0.00 0.09 0.08 0.07 0.04<br />

KO 0.26 0.05 0.19 0.08 0.07 0.04 0.22 0.02 0.10 0.08 0.07 0.04<br />

MMM 0.17 0.08 0.08 0.08 0.07 0.04 0.24 0.00 0.11 0.08 0.06 0.04<br />

PG 0.21 0.03 0.11 0.08 0.06 0.04 0.19 0.00 0.12 0.07 0.06 0.04<br />

SGP 0.14 0.02 0.10 0.07 0.06 0.04 0.23 0.02 0.09 0.08 0.06 0.04<br />

TXN 0.11 0.01 0.07 0.06 0.05 0.03 0.09 0.04 0.08 0.05 0.05 0.03<br />

WAG 0.18 0.04 0.10 0.08 0.06 0.04 0.27 0.01 0.12 0.08 0.07 0.04<br />

λEV T<br />

·,m<br />

BMY 0.19 0.06 0.12 0.07 0.06 0.04 0.22 0.02 0.11 0.08 0.07 0.04<br />

CVX 0.17 0.02 0.10 0.08 0.07 0.04 0.20 0.00 0.09 0.07 0.06 0.04<br />

HPQ 0.24 0.02 0.09 0.08 0.07 0.04 0.28 0.01 0.10 0.08 0.07 0.04<br />

KO 0.25 0.04 0.20 0.08 0.07 0.04 0.23 0.00 0.11 0.08 0.07 0.04<br />

MMM 0.16 0.12 0.11 0.08 0.07 0.04 0.25 0.02 0.12 0.07 0.06 0.04<br />

PG 0.20 0.01 0.12 0.08 0.06 0.04 0.20 0.02 0.08 0.07 0.06 0.04<br />

SGP 0.13 0.06 0.10 0.07 0.06 0.04 0.24 0.04 0.17 0.08 0.06 0.04<br />

TXN 0.10 0.03 0.09 0.06 0.05 0.03 0.10 0.08 0.10 0.06 0.05 0.03<br />

WAG 0.17 0.01 0.10 0.08 0.07 0.04 0.28 0.02 0.13 0.09 0.07 0.04<br />

Table 3.26: Establishing the uncertainty of non-parametrically estimated upper tail dependence coefficients ˆ λU,m and ˆ λEV T<br />

U,m and lower tail dependence<br />

coefficients ˆ λL,m and ˆ λEV T<br />

L,m by creating 1000 bootstrap samples of historical return data tables for S&P 500 index and corresponding asset<br />

returns and calculation of quantiles, extreme values, and standard deviations of the results. The tails represent the most extreme 4% of the return<br />

values during a time interval from January 1991 to December 2000.


117<br />

m=5736 bs=800<br />

upper tail lower tail<br />

λbs,mean<br />

λbs,mean<br />

λ·,m k=229.40 rel err max dev 95% q 90% q std bs k=229.40 rel err max dev 95% q 90% q std bs<br />

BMY 0.22 0.01 0.07 0.05 0.04 0.02 0.28 0.00 0.10 0.06 0.05 0.03<br />

CVX 0.18 0.03 0.09 0.05 0.04 0.02 0.27 0.01 0.09 0.06 0.05 0.03<br />

HPQ 0.28 0.01 0.10 0.05 0.04 0.03 0.28 0.02 0.10 0.05 0.04 0.03<br />

KO 0.27 0.03 0.11 0.06 0.05 0.03 0.29 0.02 0.08 0.05 0.04 0.03<br />

MMM 0.25 0.02 0.09 0.05 0.04 0.03 0.26 0.01 0.09 0.05 0.04 0.03<br />

PG 0.20 0.02 0.09 0.05 0.04 0.03 0.20 0.01 0.09 0.05 0.04 0.02<br />

SGP 0.17 0.02 0.08 0.05 0.04 0.02 0.26 0.00 0.09 0.05 0.05 0.03<br />

TXN 0.17 0.00 0.07 0.05 0.04 0.02 0.16 0.02 0.07 0.05 0.04 0.02<br />

WAG 0.21 0.01 0.08 0.05 0.04 0.03 0.28 0.03 0.10 0.06 0.05 0.03<br />

λEV T<br />

·,m<br />

BMY 0.22 0.00 0.08 0.05 0.04 0.03 0.28 0.01 0.10 0.06 0.05 0.03<br />

CVX 0.17 0.04 0.08 0.05 0.04 0.02 0.27 0.00 0.09 0.06 0.05 0.03<br />

HPQ 0.28 0.01 0.09 0.05 0.05 0.03 0.29 0.01 0.11 0.05 0.04 0.03<br />

KO 0.27 0.04 0.10 0.06 0.05 0.03 0.29 0.02 0.08 0.05 0.05 0.03<br />

MMM 0.24 0.01 0.09 0.05 0.04 0.03 0.27 0.00 0.09 0.05 0.04 0.03<br />

PG 0.19 0.03 0.09 0.05 0.04 0.03 0.20 0.00 0.09 0.05 0.04 0.03<br />

SGP 0.17 0.01 0.08 0.05 0.04 0.02 0.27 0.00 0.09 0.05 0.05 0.03<br />

TXN 0.17 0.01 0.07 0.05 0.04 0.02 0.16 0.00 0.07 0.05 0.04 0.02<br />

WAG 0.21 0.02 0.08 0.05 0.04 0.03 0.29 0.02 0.09 0.06 0.05 0.03<br />

Table 3.27: Establishing the uncertainty of non-parametrically estimated upper tail dependence coefficients ˆ λU,m and ˆ λEV T<br />

U,m and lower tail dependence<br />

coefficients ˆ λL,m and ˆ λEV T<br />

L,m by creating 800 bootstrap samples of historical return data tables for S&P 500 index and corresponding asset<br />

returns and calculation of quantiles, extreme values, and standard deviations of the results. The tails represent the most extreme 4% of the return<br />

values during a time interval from July 1985 to April 2008.


Observation of λ by Rolling Time Window<br />

I also implemented the rolling horizon time window to the Schmidt & Stadtmüller<br />

estimator for the three window sizes S = 2500, S = 1600, and S = 800. Figures (3.63),<br />

(3.64), and (3.65) show the results. We can see that for most of the time left-tail<br />

dependence ˆ λL,m plotted in black dominates right-tail dependence ˆ λU,m given in green.<br />

The bigger the time window, the more obvious this becomes. In terms of trends it is<br />

difficult to draw conclusion because plots for different window sizes look a little different<br />

from each other. Assuming that plots with window size S = 2500 given in figure (3.63)<br />

represent something like a smoothed version of the other plots with smaller window<br />

sizes we get the impression that tail dependence for both positive and negative tails<br />

have become remarkably constant recently. Comparing the right part of the plots,<br />

representing estimates from 1985 to 1995, to the left part of the plots, representing<br />

estimates from 2000 to 2008, the left side is more fluctuating than the right side. For<br />

most assets it is also this area, where ˆ λL,m grows clearly beyond ˆ λU,m. It is not clear<br />

whether there is in general a decreasing or an increasing tendency of ˆ λL,m and ˆ λU,m.<br />

Also here asset Texas Instruments inc. (TXN) looks different from the others as it<br />

shows a clear increasing trend over time. Overall it is remarkable how consistent most<br />

of the assets remain over time. Except asset TXN, all assets are fluctuating around a<br />

constant value. Even if we look at figure (3.65), which shows estimates for the smallest<br />

windows of size S = 800 fluctuations remain limited. This was not the case for i.e. first<br />

and second β-smile conditions.<br />

Error Bars in <strong>Dependence</strong> of Time<br />

To establish uncertainty of estimated upper and lower tail dependence estimates in<br />

dependence of time I performed bootstrap sampling for the rolling time windows 18 .<br />

Therefore I used a window size of S = 1600. The Schmidt and Stadtmüller approaches<br />

request the highest computing power of all implemented concepts.<br />

Because bootstrap estimates showed good accuracy I think that S = 1600 daily observations<br />

per estimate is enough for this approach. Results provided on the data CD<br />

show that bootstrap quantiles remain stable and low enough for the chosen window<br />

size. 95% quantiles are around 40% to 50% of total estimates for upper and lower tails<br />

of most assets. This yields standard deviations around 20% to 25% of the estimates,<br />

which is comparably low. ˆEV T λL,m and ˆEV T λU,m perform similarly to ˆ λL,m and ˆ λU,m in terms<br />

of bootstrap quantile estimations.<br />

18 Results of bootstrap sampling of the rolling time window estimates for λU,m, λL,m, λ EV T<br />

U,m , and λL,m and<br />

window size of S = 1600 data points can be observed on the data CD submitted with the thesis in the folder:<br />

window results<br />

118<br />

EV T


119<br />

λ<br />

λ<br />

λ<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

BMY<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

m<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

KO<br />

m<br />

SGP<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

λ<br />

λ<br />

λ<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

CVX<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

m<br />

MMM<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

m<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

λ<br />

λ<br />

λ<br />

0.32<br />

0.3<br />

0.28<br />

0.26<br />

0.24<br />

0.22<br />

HPQ<br />

0.2<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

m<br />

PG<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

m<br />

WAG<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

m<br />

m<br />

Figure 3.63: λU,m(m) for upper tails of index S&P 500 and the nine assets plotted in green and λL,m(m) for lower tails plotted in black using a<br />

non-parametric approach given by equation ˆ λL,m = 1 �m k j=1 1 R (j)<br />

m1≤k and R(j)<br />

m2≤k and ˆ λU,m = 1 �m k j=1 1 R (j)<br />

for rolling time horizon<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

windows of S = 2500 considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to<br />

Mars 2008.


120<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

BMY<br />

0.1<br />

1000 2000 3000 4000 5000 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

m<br />

0.1<br />

1000 2000 3000 4000 5000 6000<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

KO<br />

m<br />

SGP<br />

0.1<br />

1000 2000 3000 4000 5000 6000<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

CVX<br />

0.1<br />

1000 2000 3000 4000 5000 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

MMM<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

TXN<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

λ<br />

λ<br />

λ<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

HPQ<br />

0.1<br />

1000 2000 3000 4000 5000 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

PG<br />

0<br />

1000 2000 3000 4000 5000 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

m<br />

WAG<br />

0.1<br />

1000 2000 3000 4000 5000 6000<br />

m<br />

m<br />

m<br />

Figure 3.64: λU,m(m) for upper tails of index S&P 500 and the nine assets plotted in green and λL,m(m) for lower tails plotted in black using a<br />

non-parametric approach given by equation ˆ λL,m = 1 �m k j=1 1 R (j)<br />

m1≤k and R(j)<br />

m2≤k and ˆ λU,m = 1 �m k j=1 1 R (j)<br />

for rolling time horizon<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

windows of S = 1600 considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to<br />

Mars 2008.


121<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

BMY<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

m<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

KO<br />

m<br />

SGP<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

CVX<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

MMM<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

TXN<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

HPQ<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

PG<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

WAG<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

m<br />

m<br />

m<br />

Figure 3.65: λU,m(m) for upper tails of index S&P 500 and the nine assets plotted in green and λL,m(m) for lower tails plotted in black using a<br />

non-parametric approach given by equation ˆ λL,m = 1 �m k j=1 1 R (j)<br />

m1≤k and R(j)<br />

m2≤k and ˆ λU,m = 1 �m k j=1 1 R (j)<br />

for rolling time horizon<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

windows of S = 800 considered data points from N = (1... S),(2... S + 1),... ,(5736 − S + 1... 5736) or a total time interval from July 1985 to<br />

Mars 2008.


122<br />

λ<br />

λ<br />

λ<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

KO<br />

m<br />

SGP<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

λ<br />

λ<br />

λ<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

CVX<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

MMM<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

λ<br />

λ<br />

λ<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

HPQ<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

PG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

m<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

m<br />

m<br />

Figure 3.66: λL,m(N) for lower tails of index S&P 500 and the nine assets for rolling time horizon windows of size S = 2500 plotted in blue,<br />

S = 1600 plotted in green, and S = 800 plotted in red using a non-parametric approach given by equation ˆ λU,m = 1 �m k j=1 1 R (j)<br />

m1 >m−k and R(j)<br />

m2 >m−k<br />

from N = (2501 − S ...S),(2502... S + 1),... ,(5736 − S + 1... 5736).


123<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

BMY<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

KO<br />

SPG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

λ<br />

λ<br />

λ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

CVX<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

MMM<br />

0.1<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

TXN<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

λ<br />

λ<br />

λ<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

HPQ<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

PG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

WAG<br />

0<br />

2500 3000 3500 4000 4500 5000 5500 6000<br />

m<br />

Figure 3.67: λU,m(N) for lower tails of index S&P 500 and the nine assets for rolling time horizon windows of size S = 2500 plotted in blue,<br />

S = 1600 plotted in green, and S = 800 plotted in red a using non-parametric approach given by equation ˆ λL,m = 1<br />

k<br />

N = (2501 − S ...S),(2502... S + 1),... ,(5736 − S + 1...5736).<br />

� m<br />

j=1 1 R (j) from<br />

m1≤k and R(j)<br />

m2≤k


3.5 Conclusions<br />

Within this section I want to compare the results of the different concepts and conclude<br />

about which approaches have more practical application and which approaches are more<br />

of theoretical or supplemental interest.<br />

It is interesting to observe that by the three main concepts: approaches according to<br />

Poon, Rockinger, and Tawn with respective non-parametric upper and lower tail dependence<br />

estimator denoted by ˆχ, approaches according to Sornette & Malevergne with<br />

respective parametric and non-parametric tail dependence estimators λ +,− including<br />

β-smile improvement conditions, and approaches according to Schmidt & Stadtmüller<br />

with respective non-parametric tail dependence estimators denoted by λL,m and λU,m<br />

presented in this chapter all differ significantly in their results.<br />

The estimator ˆχ that was covered in the first section (3.1) yielded results reported<br />

in tables (3.1) and (3.2), that were all close to one. This would mean that all of our<br />

considered assets show a nearly perfect extremal dependence with the index S&P 500<br />

for corrected upper and lower tails. Furthermore, most of the estimates were close<br />

together or even identical. Therefore results are not of much use. The only column<br />

that is different from the others reports estimates for ˆχ yielded by approximations<br />

of univariate survival functions that were estimated without correction for the most<br />

extreme outliers for tail values with rank (1 . . .k) and k denoting the rank number of<br />

least extreme tail value, which is somewhat questionable.<br />

Results according to Sornette & Malevergne calculated by ˆ β of the whole data set<br />

reported in tables (3.8) and (3.10) using Hill’s estimator and reported in tables (3.9)<br />

and (3.11) using Gabaix’s estimator are consistent, whereas parametric estimators tend<br />

to be somewhat higher than non-parametric estimators. The only estimates where the<br />

performance of the parametric and the non-parametric approach was significantly different<br />

from each other was yielded applying the parametric approach to negative tails by<br />

approximation of corresponding tail distributions by univariate survival functions without<br />

correction for the most extreme outliers, namely by tail values with rank (1 . . .k),<br />

where k denotes the rank number of the least extreme tail values. Anyway plots of<br />

the parametrized survival function compared to the empirical complementary cumulative<br />

distribution function for the smaller data set plotted in figure (3.38) show big<br />

differences between the two distribution functions and again call the parametrization<br />

of survival copulas with uncorrected average values for the estimation of scale factors<br />

ĈY and Ĉε into question. Using the bigger data set we can overcome these difficulties<br />

by some extent, which can be observed in figures (3.31) and (3.33) and also in<br />

tables (3.10) and (3.11) showing estimates for the bigger data sets. In case of both<br />

non-parametric and parametric estimators we observe that estimates with corrected l<br />

and corrected C, calculated for tail values with rank c . . .k excluding the most extreme<br />

0.5%, are in most cases higher than estimates for the positive tails. Using the Gabaix<br />

estimator shown in tables (3.11) and (3.9) differences between upper and lower tail<br />

dependence estimates are more distinctive. This finding agrees well with previous literature.<br />

Comparing differences between tail dependence estimates for index S&P 500<br />

and the nine assets of the smaller and the bigger data sets the estimates are consistent<br />

in most cases. Overall, estimates between the different assets and the index are not as<br />

close together as for ˆχ, although the biggest values were estimated for assets Hewlett-<br />

Packard Co. (HPQ) and Coca-Cola Co (KO) around 12% to 14%, which intuitively<br />

seems very small.<br />

β-smile estimates for the first and second SI conditions given in table (3.18) and<br />

(3.19) sometimes were similar to the approaches using ˆ β calculated by all data and<br />

124


sometimes were significantly different. The problem here is that ˆ βSI are fluctuating<br />

and show large error bars. Mostly the β-smile estimators tend to be more radical than<br />

estimates calculated by classic ˆ β in terms of yielding more extreme results (close to<br />

one or close to zero). It might be good to have the β-smile improvement approaches<br />

as additional indication for very low or high tail dependence coefficient estimates and<br />

complex dependence structures but I would not trust in these estimates because of the<br />

high error bars obtained by bootstrap sampling.<br />

Coming to concepts according to Schmidt & Stadtmüller with results reported in<br />

tables (3.25) and (3.24) we observe that all the estimates lie in-between the two other<br />

concepts. They are bigger than estimates according to Sornette & Malevergne and<br />

smaller than estimates according to Poon, Rockinger, and Tawn. Also here, lower tail<br />

dependence estimates tend to be generally higher than tail dependence estimates of<br />

the upper tails. Estimates are consistent for the different time windows and they are<br />

rather close together for all assets. Studying the definition of the estimators proposed<br />

by Schmidt & Stadtmüller we find that the dependence structure in the tails is fully<br />

explained by the rank order of the input data. Absolute return values have no impact<br />

on the results. That means that highly inconvenient outliers are just weighted by their<br />

ranks and not in terms of their absolute value relative to the rest of data. This is<br />

different to the other estimators where we always have a certain impact of the extent<br />

of extreme outliers in terms of a proportion i.e. by ˆl in the non-parametric estimator or<br />

in terms of a proportion of scale factors Ĉε/ ĈY in the parametric approach proposed<br />

by Sornette & Malevergne. The χ estimator uses Z, which is the minimum value of tail<br />

distribution approximations by unit Fréchet marginals S and T of assets and index.<br />

Now we come to the error bar estimations of the different concepts: bootstrap estimates<br />

for error bars of ˆχ are given in tables (3.3) and (3.4). Looking at the smaller<br />

sample, quantiles are much bigger in the negative tails than in the positive tails for both<br />

uncorrected tails and tails corrected for most extreme outliers. For the bigger sample<br />

this only counts for uncorrected tails. Anyway quantiles for corrected tails seem to be<br />

low indicating that estimates are accurate. For uncorrected tails quantiles were very<br />

�<br />

Zk,N<br />

high for some assets. Using the proposed relation ˆσ =<br />

2k(N−k) N3 for the calculation<br />

of error bars is only recommended for tails that are corrected for outliers because these<br />

outliers seem to have a strong impact on the accuracy of the estimate for concepts<br />

according to Poon, Rockinger, and Tawn.<br />

Looking at the non-parametric and parametric approaches proposed by Sornette &<br />

Malevergne with bootstrap estimates of error bars for smaller and bigger data sets<br />

provided in tables (3.12) and (3.13) for the non-parametric approach and provided in<br />

tables (3.15) and (3.16) for the parametric approach the situation looks very similarly.<br />

Errors seem to be approximately Gaussian distributed for both parametric and nonparametric<br />

estimators. The only thing we have to keep an eye on is that the parametric<br />

approach seems to perform poorly with uncorrected negative tails primarily for the<br />

smaller data set. Correcting the tails for outliers helps to achieve stable estimates. The<br />

non-parametric approach seems to be less sensitive to outliers.<br />

<strong>Tail</strong> dependence estimates calculated by β-smile improvement conditions perform<br />

very poorly in terms of error bars that were estimated by bootstrap sampling with<br />

replacement shown in tables (3.20) and (3.21), for the non-parametric approach with<br />

ˆβSI calculated by the first SI condition applied to the smaller and the bigger data sets,<br />

and in tables (3.22) and (3.23), for the non-parametric approach with ˆ βSI calculated<br />

by the second SI condition applied to the smaller and the bigger data sets. Standard<br />

deviations ’std bs’ are mostly around the size of the estimates for both conditions.<br />

125


Now we come to bootstrap estimates performed for the Schmidt & Stadtmüller<br />

approaches shown in tables (3.24) and (3.25) for the smaller and the bigger data<br />

sets. Errors seem to be approximately Gaussian distributed and consistent without<br />

exception. Furthermore ’std bs’ seem to be comparably low for both smaller and bigger<br />

data sets.<br />

As a last point in this chapter I wanted to compare the rolling time window estimates<br />

of the different methods: it is remarkable how similar the estimators ˆ λU,m and<br />

ˆλL,m provided by Schmidt & Stadtmüller compared to the estimators ˆ λ +,− provided<br />

by Sornette & Malevergne look for some assets if we just consider the shapes of their<br />

respective plots in dependence of time. Also the extent of movements is very similar.<br />

The main difference seems to be that ˆ λU,m and ˆ λL,m generally perform on a higher<br />

estimation level. Dynamics in dependence of time seem to be similar to ˆ λ +,− . ˆ λ +,−<br />

estimated by SI conditions shows much stronger fluctuations than the other methods.<br />

For the implementation part of the thesis presented in the next chapter I will provide<br />

results estimated by all the different methods to observe the impact on their<br />

performance applying the concepts to different data sets. Anyway from this point it<br />

seems that the non-parametric estimator by Sornette & Malevergne and the also nonparametric<br />

estimator by Schmidt & Stadtmüller have more practical application than<br />

the rest of the estimators because of their accuracy in terms of error bars and their<br />

consistent performance for different time windows.<br />

126


Chapter 4<br />

Application of Concepts to various<br />

Data and Purposes<br />

After implementation of the different methods I wanted to extend the scope of this work<br />

by applying the concepts to various data and purposes and also show some interesting<br />

applications.<br />

In a first step described in section (4.1), I applied the concepts to indexes from<br />

major finance centers in the world and respective assets in a comparable way as on the<br />

example of S&P 500 and the nine assets to gain further experience by comparing the<br />

results. In section (4.2) I wanted to calculate tail dependence coefficient estimates for<br />

market indexes from different countries, to find out to what extent extreme measures<br />

depend on geographical distances, and then compare the results with tail dependence<br />

coefficient estimates of exchange rates to check whether they accord well with each<br />

other. Finally in section (4.3) I will implement the different approaches to synthetic<br />

time series to find out what the best result can be. Furthermore it will show me biases<br />

and give me the lower bounds for the errors of the different concepts.<br />

4.1 Application to major Financial Centers in the World<br />

I downloaded a lot of asset and index data for the implementation of the different<br />

concepts to the major financial centers in the world. Table (4.1) shows the indexes<br />

including abbreviations adopted from 1 and tables (4.2), (4.3), (4.4), and (4.5) give<br />

respective assets included in the indexes plus abbreviations also adopted from Yahoo<br />

finance. Most of the assets used for the calculations are among the stocks with the<br />

largest capitalization within their respective indexes, whereas their weight in the indexes<br />

should not be too high in order that dependence does not stem from their overlap with<br />

the market factor. Anyway I tried to choose a certain spectrum of different weights<br />

to observe the impact on the estimates 2 . A problem was that for most indexes I did<br />

not find exact index component weights. Only exceptions were Dow Jones Industrial<br />

Average and S&P 500 3 . One has to remember that even if I knew the exact component<br />

weights of the indexes, they could be altering in time as I sometimes consider time<br />

intervals covering several decades. Anyway this is primarily of high interest for my<br />

approaches according to Sornette & Malevergne because these approaches depend on<br />

1 http://finance.yahoo.com<br />

2 Asset and index data is provided on the data CD<br />

3 Information about index component weights for assets of S&P 500 and most important assets of Dow Jones<br />

Industrial Average index is provided on: http://www.indexarb.com/indexComponentWtSwitch.html updated<br />

on September-16-2008.<br />

127


the regression coefficient β of the linear additive single factor model X = β · Y + ε,<br />

which above has been found to be the crucial parameter for this type of models. If<br />

β is small, tail dependence estimators ˆ λ +,− (β) are low as well and vice versa. This<br />

has been recognized as a weakness of the model, which lead to the introduction of the<br />

β-smile improvement conditions, where βSI(k) was calculated only by tail return data<br />

up to threshold number k but unfortunately yielded highly volatile estimates. Using<br />

the classic β that is calculated by all data, though yields stable solutions, focuses on<br />

linear dependences over the whole return data range, whereas moderate return data is<br />

weighted equally like extreme data.<br />

First of all I defined a time interval for which I had complete data for all assets<br />

and indexes. The biggest ’common’ time interval ranges from begin of February 2003<br />

to end of August 2008 and contains, depending on the index of the financial center,<br />

from 1372 data points for Mercado de Valores (MERVAL) to 1433 data points for SSE<br />

Composite index (SSEC). Data for index Tel Aviv 100 (TA 100) and included assets<br />

was available not till end of August 2003 but the impact on tail dependence coefficients<br />

of the absent of these early data should be marginal. Anyway for some indexes there is<br />

much more data available i.e. Dow Jones Industrial Average (DOW) price data starting<br />

from January 1962 up to now consisting of more than 11’700 observations for assets<br />

Alcoa Inc. (AA), Boeing Co. (BA), and Caterpillar Inc. (CAT) or S%P 500 price data<br />

for assets Hewlett-Packard Co. (HPQ) and Coca-Cola Co. (KO) also starting from<br />

January 1962 up to now consisting of more than 11’600 observations. These huge data<br />

sets can be used to check for the consistency of the estimators over time. Anyway the<br />

target is to be able to draw conclusions with far less data as for example by our above<br />

described five and a half year interval from early 2003 up to now.<br />

First I implemented the parametric and the non-parametric approaches according to<br />

Sornette & Malevergne and already made a very unexpected finding: tail dependence<br />

estimates for all assets and indexes outside of the U.S. were significantly smaller than<br />

0.001. That means that their extreme dependence of the market factor would be close<br />

to zero, what intuitively seems to be highly questionable. All the U.S. indexes instead<br />

exhibited more or less reasonable estimates. My first idea was that maybe the market<br />

capitalization of the chosen assets might be too small but then, performing the estimates<br />

on assets with different known weights (assets of S&P 500 and some assets of Dow<br />

Jones Industrial Average), I realized that the only cause of vanishing tail dependence<br />

estimates for all but the U.S. indexes and corresponding assets were differences in<br />

estimated β because the estimated tail dependence coefficients are very sensitive to<br />

β calculated by the linear single factor model. If β < 0.2 ˆ λ +,− → 0. That means<br />

that when we observe a low dependence of moderate changes between asset and index,<br />

extreme dependence vanishes as well. This can be observed on table (4.6), which shows<br />

results for parametric and non-parametric ˆ λ +,− for a data set consisting of N = 2000<br />

daily returns of index S%P 500 and assets with different component weights denoted<br />

by ’C.W.’.<br />

I also observed on index S&P 500 that assets, which have a low weight in the index<br />

(< 0.5%) tend to have smaller β and therefore low ˆ λ +,− . What is interesting now<br />

is that this is no longer the case if we calculate βSI(k) by SI conditions: estimates<br />

of λ +,− that were close to zero when we use β of the whole sample could become<br />

significantly higher than zero when calculated using βSI calculated by first or second<br />

β-smile conditions. This can be observed in table (4.7) on the same data set as above<br />

consisting of N = 2000 data points for index S&P 500 and corresponding assets and<br />

would imply that assets that overall have a very low tendency to move simultaneously<br />

128


with the index can have much stronger dependence in their extreme tails.<br />

Additionally, I performed bootstrap sampling with replacement of my biggest data<br />

sets up to 11’640 data points to find out whether the estimates become more accurate if<br />

we boost the data sets. The relative errors between the results of the original sample and<br />

the means of all bootstrap samples became small but the standard deviations between<br />

the samples for all my estimates were still precisely of the size of the estimates for both<br />

the first and the second SI conditions. For comparison I added the results of the tail<br />

dependence estimates of the other U.S indexes Dow Jones Industrial Average (DOW)<br />

and respective assets included in table (4.8) for the parametric and non-parametric<br />

approaches with β calculated by all data and in table (4.9) for the non-parametric<br />

approach with βSI only calculated for tail data. NASDAQ Composite (NASDAQ) and<br />

respective assets are included in table (4.10) for both, β calculated by all data and βSI.<br />

The data sets also consists of N = 2000 data points, which has been identified to be<br />

the smallest size to provide stable estimates for the approaches according to Sornette &<br />

Malevergne. Assets with biggest weights in the NASDAQ Composite Index are i.e. Microsoft<br />

Corp. (MSFT) with around 9%, Intel Corp. (INTC) with around 6%, and Cisco<br />

Systems Inc. (CSCO) with around 5.4% but as mentioned, unfortunately I couldn’t<br />

access more precise data. Error bar results of bootstrap sampling with replacement for<br />

the indexes Dow Jones Industrial Average (DOW) and NASDAQ Composite (NAS-<br />

DAQ) were fully consistent with my S%P 500 results presented in subsections (3.2.3)<br />

and (3.2.5).<br />

Now I come to the other approaches according to Poon, Rockinger, and Tawn and<br />

according to Schmidt & Stadtmüller. In table (4.11) I show results at the example of<br />

Australian Securities Exchange index (AORD) and corresponding assets for the common<br />

data samples ranging from begin of February 2003 to end of August 2008 and<br />

consisting of N = 1398 daily return observations. That was the maximum common<br />

interval with data available for assets and index. Table (4.12) shows estimates by the<br />

same data but by the non-parametric approach according to Sornette & Malevergne<br />

with first and second β-smile conditions. 95% and 90% bootstrap error quantiles show<br />

on table (4.11) that for approaches according to Poon, Rockinger, and Tawn and according<br />

to Schmidt & Stadtmüller errors are mostly lower than the size of the estimates<br />

but still for many assets uncertainty in terms of error bars seems to be very high. This<br />

is similar for all samples with less than N = 2000 data points showing us that we need<br />

an interval of at least eight years to obtain accurate results by any of the methods<br />

presented. The problem is that the data that is available for assets traded outside<br />

the U.S. mostly does not go back that far in time. Anyway in most cases estimates<br />

yielded by smaller data sets are close to estimates yielded by bigger ones but there are<br />

some exceptions. Considering the limit of N = 2000 there are only a few assets of my<br />

gathered sample data remaining that are traded outside the U.S.:<br />

�Australian Securities Exchange index (AORD): Billabong International Limited<br />

(BBG) and Commonwealth Bank of Australia (CBA) with N = 2059 daily observations<br />

up to now<br />

�Financial Times Stock Exchange index (FTSE 100): United Utilities Gr. (UU)<br />

with N = 2117 daily observations up to now<br />

�MIBTEL: Tiscali (TIS) with N = 2123 daily observations up to now<br />

Applying Schmidt & Stadtmüller approaches to these bigger data sets yields lower<br />

bootstrap quantiles than to the smaller data sets consisting of N = 1398 data points.<br />

129


Index Abbrev. Country<br />

Australian Securities Exchange (ASX) AORD Australia<br />

CAC 40 CAC 40 France<br />

Deutscher Aktien Index DAX Germany<br />

Dow Jones Industrial Average DOW U.S.<br />

Financial Times Stock Exchange Index FTSE 100 U.K<br />

Jakarta Composite Index JKSE Indonesia<br />

Mercado de Valores MERVAL Argentina<br />

MIBTEL MIBTEL Italy<br />

NASDAQ Composite NASDAQ U.S.<br />

Swiss Market Index SMI Swiss<br />

SSE Composite Index SSEC China<br />

Tel Aviv 100 TA 100 Israel<br />

Table 4.1: Indexes that were used for the implementation of the different concepts given by<br />

their name, further used abbreviation, and country of origin.<br />

This is fully consistent with my bootstrap sampling estimates for the two reference data<br />

sets of S&P 500 presented in subsection (3.4.3).<br />

Applying approaches according to Poon, Rockinger and Tawn with tail dependence<br />

coefficient given by χ +,− , estimates sometimes yield lower bootstrap quantiles but for<br />

several assets i.e. Commonwealth Bank of Australia (CBA) errors remain high. This is<br />

also consistent with my further results presented in subsection (3.1.2), where bootstrap<br />

error bars for the different assets included in the index S&P 500 were also inconsistent.<br />

130


Index Asset Abbrev.<br />

AORD Australian Agricultural Company Limited AAC<br />

Abacus Property Group ABP<br />

AGL Energy Limited AGK<br />

Asciano Group AIO<br />

Axa Asia Pacific Holdings Limited AXA<br />

Billabong International Limited BBG<br />

Bioral Limited BLD<br />

Commonwealth Bank of Australia CBA<br />

Centennial Coal Company Limited CEY<br />

CSR Limited CSR<br />

Futuris Corporation Limited FCL<br />

Ing Industrial Fund IIF<br />

Macquarie Office Trust MOF<br />

Newcrest Mining Limited NCM<br />

Nido Petroleum Limited NDO<br />

CAC Alstom ALO<br />

Michelin ML<br />

EDF EDF<br />

Peugeot UG<br />

L’Oreal OR<br />

France Telecom FTE<br />

Sanofi-Aventis SAN<br />

Alcatel-Lucent ALU<br />

PPR PP<br />

Renault RNO<br />

DAX Adidas ADS<br />

Bayer BAY<br />

Deutsche Bank N DBK<br />

Hypo Real Estate HRX<br />

Linde LIN<br />

Metro MEO<br />

Merck MRK<br />

Munich Re Group MUV2<br />

SAP SAP<br />

Siemens SIE<br />

Tui TUI1<br />

Table 4.2: Assets included in indexes ’AORD’ (ASX), ’CAC 40’, and ’DAX’ that were used for<br />

the implementation of the different concepts given by their name, further used abbreviation,<br />

and country of origin.<br />

131


Index Asset Abbrev. Comp. Weight<br />

DOW Alcoa Inc. AA 1.95%<br />

American Int’l. Group AIG 0.28%<br />

Boeing Co. BA 4.54%<br />

Caterpillar Inc. CAT 4.77%<br />

General Motors GM 0.80%<br />

The Home Depot Inc. HD 2.03%<br />

Honeywell International Inc. HON -<br />

Intel Corporation INTC 1.43%<br />

Johnson and Johnson JNJ 5.14%<br />

Mc Donald’s MCD 4.73%<br />

Wal Mart Stores Inc. WMT 4.58%<br />

Exxon Mobil Corp. XOM 5.63%<br />

FTSE 100 Anglo American AAL<br />

Aviva AV<br />

Bae Systems BA<br />

Barclays BARC<br />

British American Tobacco BATS<br />

BG Group BG<br />

Cable and Wireless CW<br />

Glaxo Smith Klein GSK<br />

International Power IPR<br />

Lonmin LMI<br />

Tui Travel TT<br />

United Utilities Gr. UU<br />

JKSE Astra Agro Lestari Tbk AALI<br />

Tiga Pilar Sejahtera Food Tbk AISA<br />

AKR Corporindo Tbk AKRA<br />

Aneka Tambam Tbk ANTM<br />

Bank Rakyat Indonesia BBRI<br />

Bank Mandiri Tbk BMRI<br />

Central Proteinaprima Tbk CPRO<br />

Gozco Plantations Tbk GZCO<br />

Hexindo Adiperkasa Tbk HEXA<br />

Laguna Cipta Griya Tbk LCGP<br />

Sampoerna Agro Tbk SGRO<br />

Sorini Agro Asia Corporindo SOBI<br />

Table 4.3: Assets included in indexes ’DOW’ (Dow Jones), ’FTSE 100’, and ’JKSE’ that<br />

were used for the implementation of the different concepts given by their name, further used<br />

abbreviation, and country of origin.<br />

132


Index Asset Abbrev.<br />

MERVAL Acindar-Escriturales ACIN<br />

Petrobras Ordinarias APBR<br />

Cresud CRES<br />

Siderar ’A’ Voto Escri ERAR<br />

Irsa Investments and Representations Inc. IRSA<br />

Mirgor-Ord.Esc MIRG<br />

MIBTEL Ascopiave ASC<br />

Alitalia AZA<br />

Benetton Group BEN<br />

Buzzi Unicem BZU<br />

Banca Carige CRG<br />

Ducati Motor Hold. DMH<br />

Eni ENI<br />

Mediobanca MB<br />

Milano Ass. MI<br />

Risanamento RN<br />

ST Microelectronics STM<br />

Tiscali TIS<br />

Terna TRN<br />

NASDAQ Composite Apple Inc. AAPL<br />

Abraxis BioScience Inc. ABII<br />

Alliance Bankshare Corporation ABVA<br />

Cephalon Inc. CEPH<br />

Cico Systems Inc. CSCO<br />

Henry Schein Inc. HSIC<br />

Intel Corporation INTC<br />

Microsoft Corporation MSFT<br />

On Semiconductor Corp. ONNN<br />

Pacific Ethanol Inc. PEIX<br />

Perry Ellis International Inc. PERY<br />

Power Shares QQQ QQQQ<br />

SMI Julius Baer Holding N BAER<br />

Adecco N ADEN<br />

Ciba Holding N CIBN<br />

The Swatch Group UHR<br />

Zurich finl. Services ZURN<br />

Syngenta N SYNN<br />

Swiss Re N RUKN<br />

Holcim N HOLN<br />

Richemont Units A CFR<br />

Table 4.4: Assets included in indexes ’MERVAL’, ’MIBTEL’, ’NASDAQ’, and ’SMI’ that<br />

were used for the implementation of the different concepts given by their name, further used<br />

abbreviation, and country of origin.<br />

133


Index Asset Abbrev. Comp. Weight<br />

S&P 500 Bristol-Myers Squibb Co. BMY 0.40%<br />

Chevron CVX 1.59%<br />

Hewlett-Packard Co. HPQ 1.12%<br />

Coca-Cola Co. KO 1.03%<br />

3M MMM 0.46%<br />

Procter and Gamble Co. PG 2.07%<br />

Schering-Plough Corp. SGP 1.12%<br />

Texas Instruments Inc. TXN 0.28%<br />

Walgreen Co. WAG 0.31%<br />

new assets:<br />

Citygroup C 0.81%<br />

McDonalds MCD 0.69%<br />

Occidental Petroleum OXY 0.54%<br />

Dell DE 0.30%<br />

PNC Financial Services PNC 0.25%<br />

Noble Energy Inc. NBL 0.10%<br />

Parker-Hannifin PH 0.09%<br />

SSEC Northeast Express 600003<br />

Anhui Expressway 600012<br />

Henan Zhongyuan Ex 600020<br />

Shanghai Electricity Power 600021<br />

Jinan Iron and Steel 600022<br />

Huadian Power International 600027<br />

Hubai Chutian Expr. 600035<br />

CNTIC Trading 600056<br />

Minmetals Development 600058<br />

Zhengzhou Yutong 600066<br />

TA 100 AFR-ISR INV ILS AFIL<br />

Alony Hetz PTY and I ALHE<br />

Audiocodes AUDC<br />

Cellcom Israel CEL<br />

Global Industries Ltd. GLOB<br />

HSG and Constr. Holding HUCN<br />

Nice Systems NICE<br />

Scailex Corp. SCIX<br />

Table 4.5: Assets included in indexes ’S&P 500’, ’SSEC’, and ’TA 100’ that were used for<br />

the implementation of the different concepts given by their name, further used abbreviation,<br />

and country of origin. Additional assets included in index S&P 500 that were not part of the<br />

reference samples presented in chapter (3) are listed below ’new assets’.<br />

134


C. W. β Non-Par Par<br />

up lo up lo up lo up lo<br />

ˆν 3.16 3.69 3.16 3.69 3.16 3.69 3.16 3.69<br />

tail 1... k 1... k c... k c... k 1...k 1... k c...k c... k<br />

ˆλ<br />

BMY 0.40% 0.52 0.03 0.01 0.03 0.01 0.03 0.01 0.03 0.01<br />

CVX 1.59% 0.46 0.06 0.01 0.06 0.02 0.07 0.00 0.07 0.03<br />

HPQ 1.12% 1.07 0.07 0.04 0.07 0.05 0.10 0.02 0.10 0.07<br />

KO 1.03% 0.33 0.02 0.01 0.02 0.01 0.02 0.01 0.02 0.01<br />

MMM 0.46% 0.58 0.08 0.03 0.08 0.05 0.10 0.00 0.11 0.08<br />

PG 2.07% 0.25 0.01 0.00 0.01 0.00 0.01 0.00 0.01 0.00<br />

SGP 1.12% 0.51 0.02 0.00 0.02 0.00 0.02 0.00 0.02 0.01<br />

TXN 0.28% 0.87 0.01 0.01 0.02 0.01 0.02 0.00 0.02 0.01<br />

WAG 0.31% 0.48 0.03 0.01 0.03 0.01 0.03 0.01 0.03 0.02<br />

MCD 0.69% 0.01 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

C 0.81% 0.04 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

OXY 0.54% 0.06 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

DE 0.30% -0.11 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00*<br />

PNC 0.25% 0.16 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

NBL 0.10% 0.09 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

PH 0.09% 0.22 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

Table 4.6: Estimated upper and lower tail dependence ˆ λ +,− applying the non-parametric and<br />

parametric approaches according to Sornette & Malevergne to index S&P 500 and 16 assets<br />

included in dependence of their component weights in the index denoted by ’C.W.’ and β.<br />

The data samples contain N = 2000 daily price observations on a range from 12.04.2000<br />

to 31.03.2008. <strong>Tail</strong> index ˆν was calculated using Hill’s estimator, k = 0.04 · N = 80, c =<br />

0.005 · N = 10, and ∗ denotes negative ˆ β.<br />

135


136<br />

C. W. β β +<br />

SI1 Cond. 1<br />

up up<br />

β+<br />

SI2 Cond. 2<br />

up up<br />

β−<br />

SI1 Cond. 1<br />

lo lo<br />

β+<br />

SI2 Cond. 2<br />

lo lo<br />

ˆν 3.16 3.16 3.16 3.16 3.69 3.69 3.69 3.69<br />

tail<br />

ˆλ<br />

1... k c... k 1... k c... k 1... k c...k 1... k c...k<br />

BMY 0.40% 0.52 0.19 0.00 0.00 -0.34 0.00* 0.00* -0.26 0.00* 0.00* 0.92 0.06 0.08<br />

CVX 1.59% 0.46 0.35 0.03 0.02 -0.20 0.00* 0.00* -0.23 0.00* 0.00* 0.79 0.11 0.16<br />

HPQ 1.12% 1.07 1.25 0.11 0.11 2.04 0.51 0.54 -0.16 0.00* 0.00* 0.13 0.00 0.00<br />

KO 1.03% 0.33 0.51 0.07 0.07 0.42 0.04 0.04 0.11 0.00 0.00 0.60 0.07 0.07<br />

MMM 0.46% 0.58 0.81 0.23 0.24 0.78 0.21 0.22 0.43 0.01 0.02 1.19 0.46 0.70<br />

PG 2.07% 0.25 0.12 0.00 0.00 0.04 0.00 0.00 1.59 1.00 1.00 1.06 0.44 0.77<br />

SGP 1.12% 0.51 -0.20 0.00* 0.00* -0.01 0.00* 0.00* 0.09 0.00 0.00 0.84 0.02 0.03<br />

TXN 0.28% 0.87 1.75 0.14 0.14 -0.28 0.00* 0.00* 0.51 0.00 0.00 0.24 0.00 0.00<br />

WAG 0.31% 0.48 0.66 0.07 0.08 -0.03 0.00* 0.00* 0.30 0.00 0.00 1.06 0.21 0.24<br />

MCD 0.69% 0.01 1.84 1.00 1.00 -0.62 0.00* 0.00* -0.69 0.00* 0.00* 0.14 0.00 0.00<br />

C 0.81% 0.04 -0.06 0.00* 0.00* -0.53 0.00* 0.00* 5.45 1.00 1.00 0.13 0.00 0.00<br />

OXY 0.54% 0.06 -0.72 0.00* 0.00* -0.60 0.00* 0.00* -0.51 0.00* 0.00* 0.65 0.03 0.03<br />

DE 0.30% -0.11 -0.22 0.00* 0.00* -1.09 0.00* 0.00* -0.42 0.00* 0.00* -0.01 0.00* 0.00*<br />

PNC 0.25% 0.16 -0.02 0.00* 0.00* 0.11 0.00 0.00 0.26 0.00 0.00 -0.01 0.00* 0.00*<br />

NBL 0.10% 0.09 -0.80 0.00* 0.00* -0.82 0.00* 0.00* 0.77 0.04 0.04 1.15 0.16 0.16<br />

PH 0.09% 0.22 1.20 0.30 0.31 -0.43 0.00* 0.00* 3.02 1.00 1.00 0.42 0.01 0.01<br />

Table 4.7: Estimated upper and lower tail dependence ˆ λ +,− applying the non-parametric approach according to Sornette & Malevergne to index<br />

S&P 500 and 16 assets included using βSI only calculated for the extreme tails by first and second β-smile conditions. Component weights of assets<br />

within the index are denoted by ’C.W.’, βSI for the first and second conditions and β calculated for all data are listed to observe their impact on<br />

the estimates. The data samples contain N = 2000 daily price observations on a time interval from 12.04.2000 to 31.03.2008. <strong>Tail</strong> index ˆν was<br />

calculated using Hill’s estimator. ’Cond 1’ denotes: Y ≥ Y (k) ∩ X ≥ X(k), ’Cond. 2’ denotes Y ≥ Y (k), k = 0.04 · N = 80, c = 0.005 · N = 10,<br />

and ∗ denotes negative ˆ β.


C. W. β N-Par Par<br />

up lo up lo up lo up lo<br />

ˆν 3.15 3.84 3.15 3.84 3.15 3.84 3.15 3.84<br />

tail 1... k 1...k c... k c...k 1... k 1... k c... k c... k<br />

ˆλ<br />

AA 1.95% 1.35 0.27 0.19 0.26 0.19 0.37 0.31 0.37 0.31<br />

BA 4.54% 1.08 0.23 0.13 0.22 0.14 0.29 0.21 0.29 0.23<br />

CAT 4.77% 1.18 0.29 0.26 0.28 0.27 0.47 0.40 0.47 0.47<br />

GM 0.80% 1.38 0.17 0.12 0.17 0.13 0.24 0.20 0.25 0.21<br />

HON - 1.37 0.28 0.23 0.29 0.27 0.43 0.36 0.46 0.44<br />

JNJ 5.14% 0.51 0.08 0.05 0.07 0.05 0.08 0.05 0.08 0.07<br />

MCD 4.73% 0.66 0.07 0.04 0.07 0.05 0.08 0.05 0.08 0.05<br />

WMT 4.58% 0.88 0.18 0.17 0.17 0.17 0.25 0.24 0.25 0.25<br />

XOM 5.63% 0.79 0.24 0.11 0.24 0.11 0.29 0.19 0.28 0.20<br />

HD 2.03% 1.31 0.25 0.18 0.26 0.20 0.40 0.19 0.41 0.33<br />

INTC 1.43% 1.59 0.21 0.12 0.21 0.14 0.33 0.17 0.34 0.23<br />

AIG 0.28% 0.35 0.01 0.00 0.01 0.00 0.01 0.00 0.01 0.00<br />

Table 4.8: Estimated upper and lower tail dependence ˆ λ +,− applying the parametric and the<br />

non-parametric approach according to Sornette & Malevergne to index Dow Jones Industrial<br />

Average and 12 assets included using β calculated by all data. Component weights of assets<br />

within the index are denoted by ’C.W.’, and β calculated for all data are listed to observe<br />

their impact on the estimates. The data samples contain N = 2000 daily price observations<br />

on a time interval from 16.08.2000 to 31.07.2008. <strong>Tail</strong> index ˆν was calculated using Hill’s<br />

estimator, k = 0.04 · N = 80, and c = 0.005 · N = 10.<br />

137


138<br />

C. W. β β +<br />

SI1 Cond. 1<br />

up up<br />

β+<br />

SI2 Cond. 2<br />

up up<br />

β−<br />

SI1 Cond. 1<br />

lo lo<br />

β+<br />

SI2 Cond. 2<br />

lo lo<br />

ˆν 3.15 3.15 3.15 3.15 3.84 3.84 3.84 3.84<br />

tail<br />

ˆλ<br />

1... k c... k 1... k c... k 1... k c...k 1... k c...k<br />

AA 1.95% 1.35 0.23 0.00 0.00 1.20 0.19 0.19 0.83 0.03 0.03 1.50 0.27 0.27<br />

BA 4.54% 1.08 0.38 0.01 0.01 0.56 0.03 0.03 1.88 1.00 1.00 2.30 1.00 1.00<br />

CAT 4.77% 1.18 0.54 0.03 0.02 1.20 0.27 0.26 0.48 0.01 0.01 0.81 0.06 0.06<br />

GM 0.80% 1.38 0.24 0.00 0.00 1.20 0.10 0.10 1.04 0.04 0.04 1.70 0.27 0.29<br />

HON - 1.37 0.91 0.08 0.08 1.50 0.35 0.36 2.21 1.00 1.00 2.40 1.00 1.00<br />

JNJ 5.14% 0.51 0.69 0.20 0.19 0.93 0.51 0.49 2.12 1.00 1.00 0.80 0.28 0.29<br />

MCD 4.73% 0.66 0.08 0.00 0.00 -0.09 0.00 0.00 -0.45 0.00 0.00 0.42 0.01 0.01<br />

WMT 4.58% 0.88 0.48 0.03 0.03 0.89 0.18 0.18 0.11 0.00 0.00 0.60 0.04 0.04<br />

XOM 5.63% 0.79 0.88 0.34 0.33 0.83 0.27 0.27 0.01 0.00 0.00 0.66 0.06 0.05<br />

HD 2.03% 1.31 0.76 0.05 0.05 1.30 0.26 0.26 1.86 0.68 0.79 2.10 1.00 1.00<br />

INTC 1.43% 1.59 0.67 0.01 0.01 1.80 0.28 0.28 -0.08 0.00 0.00 0.49 0.00 0.00<br />

AIG 0.28% 0.35 0.44 0.01 0.01 -0.51 0.00 0.00 -0.24 0.00 0.00 -0.17 0.00 0.00<br />

Table 4.9: Estimated upper and lower tail dependence ˆ λ +,− applying the non-parametric approach according to Sornette & Malevergne to index<br />

Dow Jones Industrial Average and 12 assets included using βSI only calculated for the extreme tails by first and second β-smile conditions.<br />

Component weights of assets within the index denoted by ’C.W.’, βSI for the first and second conditions and β calculated for all data are listed<br />

to observe their impact on the estimates. The data samples contain N = 2000 daily price observations on a time interval from 16.08.2000 to<br />

31.07.2008. <strong>Tail</strong> index ˆν was calculated using Hill’s estimator, k = 0.04 · N = 80, and c = 0.005 · N = 10.


C. W. β N-Par Par<br />

up lo up lo up lo up lo<br />

ˆν 2.59 3.83 2.59 3.83 2.59 3.83 2.59 3.83<br />

tail 1... k 1... k c... k c... k 1... k 1... k c...k c... k<br />

ˆλ<br />

AAPL - 1.01 0.26 0.12 0.25 0.16 0.28 0.04 0.28 0.28<br />

INTC - 1.25 0.52 0.26 0.49 0.31 0.65 0.43 0.64 0.62<br />

MSFT - 0.85 0.41 0.26 0.40 0.29 0.50 0.49 0.50 0.57<br />

CSCO - -0.08 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00* 0.00*<br />

CEPH - 0.34 0.02 0.00 0.02 0.00 0.02 0.00 0.02 0.00<br />

PERY - 0.21 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00<br />

HSIC - 0.26 0.02 0.00 0.02 0.00 0.02 0.00 0.02 0.00<br />

QQQQ - 1.12 0.87 0.63 0.86 0.62 0.96 0.99 0.96 0.99<br />

ONNN - 0.57 0.01 0.00 0.01 0.00 0.01 0.00 0.01 0.00<br />

ˆλ β +<br />

SI1 β−<br />

SI1<br />

AAPL 0.28 -0.60 0.01 0.00* 0.01 0.00* - - - -<br />

INTC 0.68 0.69 0.11 0.03 0.10 0.03 - - - -<br />

MSFT 0.69 0.62 0.23 0.07 0.23 0.08 - - - -<br />

CSCO -0.10 0.09 0.00* 0.00 0.00* 0.00 - - - -<br />

CEPH 0.20 0.21 0.00 0.00 0.00 0.00 - - - -<br />

PERY 1.55 0.69 0.50 0.02 0.50 0.03 - - - -<br />

HSIC 0.58 1.46 0.15 1.00 0.14 1.00 - - - -<br />

QQQQ 0.99 0.89 0.64 0.26 0.63 0.26 - - - -<br />

ONNN -3.24 0.53 0.00* 0.00 0.00* 0.00 - - - -<br />

ˆλ β +<br />

SI2 β−<br />

SI2<br />

AAPL 0.38 -0.43 0.02 0.00* 0.02 0.00* - - - -<br />

INTC 0.82 0.19 0.18 0.00 0.17 0.00 - - - -<br />

MSFT 0.82 0.39 0.37 0.01 0.37 0.02 - - - -<br />

CSCO -0.11 -1.00 0.00* 0.00* 0.00* 0.00* - - - -<br />

CEPH -0.66 -0.57 0.00* 0.00* 0.00* 0.00* - - - -<br />

PERY -0.25 -1.67 0.00* 0.00* 0.00* 0.00* - - - -<br />

HSIC -0.15 0.34 0.00* 0.01 0.00* 0.01 - - - -<br />

QQQQ 1.04 1.04 0.73 0.47 0.72 0.47 - - - -<br />

ONNN -1.11 -1.84 0.00* 0.00* 0.00* 0.00* - - - -<br />

Table 4.10: Estimated upper and lower tail dependence ˆ λ +,− applying the parametric and the<br />

non-parametric approach according to Sornette & Malevergne to index NASDAQ Composite<br />

and 9 assets included using β calculated for all data and βSI only calculated for the extreme<br />

tails by first and second β-smile conditions. Component weights of assets within the index<br />

denoted by ’C.W.’, βSI for the first and second conditions and β calculated for all data<br />

are listed to observe their impact on the estimates. The data samples contain N = 2000<br />

daily price observations on a time interval from 15.08.2000 to 31.07.2008. <strong>Tail</strong> index ˆν was<br />

calculated using Hill’s estimator, k = 0.04·N = 80, c = 0.005·N = 10, and ∗ denotes negative<br />

ˆβ.<br />

139


140<br />

ˆχ + 95% Q 90% Q ˆχ + c 95% Q 90% Q ˆχ − 95% Q 90% Q ˆχ − c 95% Q 90% Q<br />

CBA 0.91 0.82 0.10 0.87 0.05 0.04 0.30 0.49 0.47 0.88 0.09 0.07<br />

BBG 0.91 0.83 0.09 0.92 0.09 0.07 0.91 0.06 0.05 0.88 0.06 0.05<br />

AAC 0.92 0.05 0.05 0.91 0.05 0.04 0.91 0.79 0.79 0.88 0.07 0.06<br />

IIF 0.92 0.03 0.03 0.92 0.05 0.04 0.91 0.84 0.08 0.88 0.05 0.04<br />

MOF 0.91 0.79 0.79 0.89 0.05 0.04 0.91 0.05 0.04 0.88 0.05 0.04<br />

NCM 0.91 0.79 0.79 0.89 0.05 0.04 0.91 0.05 0.04 0.88 0.05 0.04<br />

ABP 0.92 0.05 0.04 0.92 0.05 0.04 0.91 0.06 0.06 0.88 0.06 0.05<br />

AGK 0.91 0.84 0.07 0.88 0.05 0.04 0.66 0.53 0.53 0.88 0.07 0.06<br />

BLD 0.92 0.04 0.04 0.90 0.04 0.04 0.91 0.04 0.04 0.88 0.05 0.04<br />

NDO 0.84 0.46 0.46 0.78 0.65 0.65 0.84 0.50 0.50 0.79 0.59 0.59<br />

CEY 0.91 0.75 0.75 0.86 0.06 0.05 0.91 0.77 0.77 0.88 0.06 0.05<br />

CSR 0.85 0.29 0.13 0.92 0.06 0.05 0.31 0.51 0.50 0.88 0.10 0.08<br />

FCL 0.83 0.71 0.71 0.92 0.75 0.75 0.88 0.79 0.13 0.88 0.06 0.05<br />

AXA 0.91 0.83 0.08 0.92 0.07 0.06 0.91 0.83 0.08 0.88 0.07 0.05<br />

ˆλU,m 95% Q 90% Q ˆ λEV T<br />

U,m 95% Q 90% Q ˆ λL,m 95% Q 90% Q ˆ λEV T<br />

L,m 95% Q 90% Q<br />

CBA 0.23 0.10 0.08 0.23 0.11 0.09 0.16 0.11 0.07 0.19 0.10 0.08<br />

BBG 0.13 0.10 0.08 0.12 0.09 0.08 0.14 0.09 0.08 0.17 0.10 0.08<br />

AAC 0.07 0.08 0.06 0.06 0.09 0.07 0.07 0.07 0.06 0.10 0.08 0.06<br />

IIF 0.20 0.10 0.09 0.19 0.11 0.09 0.22 0.11 0.09 0.24 0.11 0.08<br />

MOF 0.23 0.12 0.10 0.23 0.11 0.09 0.25 0.11 0.10 0.28 0.11 0.10<br />

NCM 0.23 0.12 0.10 0.23 0.11 0.09 0.25 0.11 0.10 0.28 0.11 0.10<br />

ABP 0.13 0.09 0.07 0.12 0.09 0.09 0.18 0.09 0.07 0.21 0.10 0.08<br />

AGK 0.05 0.05 0.05 0.05 0.06 0.05 0.13 0.09 0.07 0.15 0.09 0.08<br />

BLD 0.09 0.07 0.06 0.08 0.08 0.06 0.04 0.05 0.04 0.06 0.05 0.05<br />

NDO 0.00 0.03 0.01 0.03 0.09 0.07 0.02 0.05 0.04 0.01 0.08 0.06<br />

CEY 0.05 0.05 0.05 0.05 0.06 0.06 0.07 0.07 0.06 0.10 0.08 0.07<br />

CSR 0.09 0.08 0.06 0.08 0.07 0.07 0.07 0.07 0.06 0.10 0.08 0.06<br />

FCL 0.02 0.03 0.03 0.01 0.05 0.04 0.07 0.07 0.07 0.10 0.08 0.06<br />

AXA 0.13 0.09 0.07 0.12 0.09 0.08 0.11 0.09 0.07 0.14 0.08 0.08<br />

Table 4.11: Estimated upper and lower tail dependence according to Poon, Rockinger, and Tawn (ˆχ + · , ˆχ− ·<br />

Schmidt & Stadtmüller in the lower part ( ˆ λ·,m, ˆ λEV T<br />

·,m<br />

) in the upper part and according to<br />

) for index AORD (ASX) and 14 assets included including error bars estimated by bootstrap<br />

sampling with replacement for bs = 1000 bootstrap samples shown next to the estimates. The data samples contain N = 1398 daily price<br />

observations on a time interval from begin of February 2003 to end of August 2008.


141<br />

ˆλ +<br />

SI1 95% Q 90% Q ˆ λ +<br />

SI1,c 95% Q 90% Q ˆ λ −<br />

SI1 95% Q 90% Q ˆ λ −<br />

SI1,c 95% Q 90% Q<br />

CBA 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

BBG 0.03 * * 0.03 * * 0.00 * * 0.00 * *<br />

AAC 0.00 * * 0.00 * * 0.02 * * 0.02 * *<br />

IIF 0.10 * * 0.10 * * 0.10 * * 0.10 * *<br />

MOF 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

NCM 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

ABP 0.06 0.06 0.02 0.06 * * 0.00 * * 0.00 * *<br />

AGK 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

BLD 0.44 * * 0.42 * * 0.00 * * 0.00 * *<br />

NDO 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

CEY 0.05 * * 0.05 * * 0.00 * * 0.00 * *<br />

CSR 0.91 0.58 0.58 0.90 * * 0.40 * * 0.70 * *<br />

FCL 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

AXA 0.03 * * 0.03 * * 0.04 * * 0.04 * *<br />

ˆλ +<br />

SI2 95% Q 90% Q ˆ λ +<br />

SI2,c 95% Q 90% Q ˆ λ −<br />

SI2 95% Q 90% Q ˆ λ −<br />

SI2,c 95% Q 90% Q<br />

CBA 0.00 * * 0.00 * * 0.02 * * 0.05 * *<br />

BBG 0.01 * * 0.01 * * 0.00 * * 0.00 * *<br />

AAC 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

IIF 0.28 * * 0.29 * * 0.11 * * 0.11 * *<br />

MOF 0.12 * * 0.11 * * 0.06 * * 0.07 * *<br />

NCM 0.12 * * 0.11 * * 0.06 * * 0.07 * *<br />

ABP 0.02 * * 0.02 * * 0.00 * * 0.00 * *<br />

AGK 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

BLD 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

NDO 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

CEY 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

CSR 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

FCL 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

AXA 0.00 * * 0.00 * * 0.00 * * 0.00 * *<br />

Table 4.12: Estimated ˆ λ +,− applying the non-parametric approach according to Sornette & Malevergne to index AORD (ASX) and 14 assets<br />

included using βSI only calculated for the extreme tails by first and second β-smile conditions. Error bars estimated by bootstrap sampling with<br />

replacement for bs = 1000 bootstrap samples were large and therefore mostly denoted by *. The data samples contain N = 1398 data points from<br />

February 2003 to end of August 2008. <strong>Tail</strong> index ˆν was calculated using Hill’s estimator.


Abbreviation Currency exchange rate<br />

A USD-Australian Dollar<br />

Eur USD-Euro<br />

GBP USD-British Pound<br />

BrR USD-Brazilian Real<br />

CHFR USD-Swiss Frank<br />

CHY USD-Chinese Yuan<br />

indoRU USD-Indonesian Rupiah<br />

indRu USD-Indian Rupee<br />

JPY USD-Japanese Yen<br />

Table 4.13: Further used abbreviations of currency exchange rates that were used for the<br />

estimation of tail dependence.<br />

4.2 Application to Exchange Rates<br />

I wanted to apply the different concepts to index data of different markets. Here the<br />

advantage is that in terms of historical index prices there is much more data available<br />

than it is the case for individual asset prices. But the problem is that the data provided<br />

does not accord in time. Choosing a sample size of N = 2000 data points for the different<br />

indexes, which has been shown to be the minimum sample size in order to obtain<br />

stable estimates, already shows shifts of two months time between different indexes 4 .<br />

Because the data is not synchronous it makes no sense to apply tail dependence.<br />

Now I come to the application of the concepts for the estimation of tail dependence to<br />

exchange rates of various foreign currencies with the US$. An explanation of abbreviations<br />

of different currency exchange rates is provided in table (4.13) 5 . The dependence<br />

between different currency exchange rates is an indicator for the dependence between<br />

different economies. The existence of a dependence between foreign exchange markets<br />

could affect firms’ investment decisions. Knowledge of short- and long-term extreme<br />

relations could furthermore have important implications for risk management.<br />

For the implementation I wanted to use relatively small, preferably up-to-date data<br />

sets. Fortunately the provided data is consistent and perfectly corresponding in time.<br />

I choose sample sizes of N = 2000 daily observations on a time interval ranging from<br />

20.10.2002 to 10.04.2008 for several currency exchange rates between the US$ and<br />

major foreign currencies. Table (4.14) shows tail dependence coefficients estimated by<br />

the non-parametric approach according to Sornette & Malevergne for the chosen time<br />

interval. The results look quite interesting because the spectrum is large. Sometimes<br />

dependences seem very strong and in other cases we estimate zero tail dependence.<br />

Regarding the results intuitively, they look reasonable to me. It is interesting that<br />

extreme changes of exchange rates of the US$ to any of the chosen currencies seem to<br />

have no significant influence on the exchange rate of the US$ to the Chinese Yuan so<br />

far. It looks as if the Chinese Yuan has been approximately independent of the other<br />

currencies in terms of extreme changes up to now. The Indian Rupee shows a similar<br />

situation. It looks as if these huge countries remained somewhat autonomic of any<br />

other economy. Another interesting finding is that by far the strongest tail dependence<br />

is estimated between the currency exchange rate of the US$ and the Euro and the<br />

4<br />

This is shown in Excel file: Index data.xls enclosed on the data CD, where several indexes are listed in<br />

parallel<br />

5<br />

Data tables of daily currency exchange rates can be downloaded at FXHistory:<br />

http://www.oanda.com/convert/fxhistory<br />

142


exchange rate of the US$ and the Swiss Franc. This seems to be reasonable.<br />

I also performed the estimations by the other methods for comparison. This can be<br />

seen in table (4.15) for the χ approach according to Poon, Rockinger, and Tawn [13]<br />

(2004) and in table (4.16) for the λ approaches according to Schmidt & Stadtmüller [24]<br />

(2005).<br />

Comparing the results achieved by the different methods, at first sight the results<br />

seem very different, but again, if we take into account relative differences, table (4.14)<br />

showing results according to Sornette & Malevergne and table (4.16) showing results<br />

according to Schmidt & Stadtmüller make similar statements i.e. tail dependence for<br />

currencies within Europe is larger than outside of Europe. Looking for example at<br />

tail dependence coefficient estimates of currency exchange rates in US$ between Euro<br />

and Swiss Frank, absolute results are nearly identical and in both tables represent<br />

the maximum estimates achieved. Another example is the Australian $: Comparing<br />

currency exchange rates of the Australian $ and the Euro with currency exchange rates<br />

of the Australian $ and the British Pound and currency exchange rates of the Australian<br />

$ and the Swiss Frank we notice that in both tables tail dependence estimates for all of<br />

the three exchange rates are relatively high whereas estimates for the Euro are slightly<br />

higher than for the Pound and estimates for the Pound are slightly higher than for<br />

the Swiss Frank. The rest of the currency exchange rate estimates with the Australian<br />

$ are significantly lower with estimates for Japanese Yen being higher than the rest.<br />

Many more of these equivalences can be found.<br />

Comparing the χ-approach shown in table (4.15) with results achieved by the other<br />

two methods it is much more difficult to draw inferences because also applied to currency<br />

exchange rates results of the χ-approach are very close together, which makes it hard<br />

to assess differences between the estimates because differences are mostly smaller than<br />

estimated error bars. What we notice is that for all cases ˆχ = 1 could not be rejected<br />

and therefore it can be concluded that asymptotic dependence exists between all of the<br />

chosen currency exchange rates. For further details we refer to section (3.1).<br />

Although the Sornette-Malevergne factor model is not so natural for the application<br />

to currency exchange rates because we do not expect a ’common factor’ to appear,<br />

the results look reasonable also compared to the results yielded by the Schmidt &<br />

Stadtmüller approaches.<br />

143


A Eur GBP BrR CHFR CHY indoRu indRu JPY<br />

A λ + * 0.29 0.23 0.01 0.20 0.00 0.01 0.00 0.06<br />

λ− 0.13 0.13 0.00 0.07 0.00 0.00 0.00 0.02<br />

λ + c<br />

λ<br />

0.28 0.22 0.01 0.19 0.00 0.01 0.00 0.06<br />

− Eur<br />

c<br />

λ<br />

0.13 0.14 0.00 0.07 0.00 0.00 0.00 0.02<br />

+ * 0.33 0.00 0.77 0.00 0.00 0.00 0.11<br />

λ− 0.46 0.00 0.71 0.00 0.00 0.00 0.10<br />

λ + c<br />

λ<br />

0.31 0.00 0.76 0.00 0.00 0.00 0.11<br />

− GBP<br />

c<br />

λ<br />

0.46 0.00 0.71 0.00 0.00 0.00 0.11<br />

+ * 0.00 0.33 0.00 0.00 0.00 0.06<br />

λ− 0.00 0.18 0.00 0.00 0.00 0.02<br />

λ + c<br />

λ<br />

0.00 0.34 0.00 0.00 0.00 0.06<br />

− BrR<br />

c<br />

λ<br />

0.00 0.18 0.00 0.00 0.00 0.02<br />

+ * 0.00 0.00 0.00 0.00 0.00<br />

λ− 0.00 0.00 0.01 0.01 0.00<br />

λ + c<br />

λ<br />

0.00 0.00 0.00 0.00 0.00<br />

− CHFR<br />

c<br />

λ<br />

0.00 0.00 0.01 0.01 0.00<br />

+ * 0.00 0.00 0.00 0.19<br />

λ− 0.00 0.00 0.00 0.12<br />

λ + c<br />

λ<br />

0.00 0.00 0.00 0.19<br />

− CHY<br />

c<br />

λ<br />

0.00 0.00 0.00 0.12<br />

+ * 0.00 0.00 0.00<br />

λ− 0.00 0.00 0.00<br />

λ + c<br />

λ<br />

0.00 0.00 0.00<br />

− indoRu<br />

c<br />

λ<br />

0.00 0.00 0.00<br />

+ * 0.00 0.01<br />

λ− 0.00 0.01<br />

λ + c<br />

λ<br />

0.00 0.01<br />

− indRu<br />

c<br />

λ<br />

0.00 0.01<br />

+ * 0.01<br />

λ− 0.00<br />

λ + c<br />

0.01<br />

0.00<br />

λ − c<br />

Table 4.14: Estimated ˆ λ +,− applying the non-parametric approach according to Sornette &<br />

Malevergne given by: ˆ λ +,− �<br />

= 1/max 1, l<br />

�ν β to exchange rates of various foreign currencies<br />

with the US$. The tail represents the most extreme 4% of the return values during a time<br />

interval ranging from 20.10.2002 to 10.04.2008 consisting of N = 2000 daily observations. ’c’<br />

means that tails were corrected for 0.5% of most extreme data for the estimation of l. <strong>Tail</strong><br />

index ˆν was calculated using Hill’s estimator.<br />

144


A Eur GBP BrR CHFR CHY indoRu indRu JPY<br />

A χ + * 0.92 0.91 0.92 0.91 0.92 0.92 0.92 0.92<br />

χ− 0.91 0.91 0.91 0.91 0.84 0.91 0.91 0.91<br />

χ + c<br />

χ<br />

0.88 0.86 0.87 0.86 0.89 0.89 0.87 0.89<br />

− Eur<br />

c<br />

χ<br />

0.86 0.86 0.86 0.86 0.86 0.86 0.86 0.86<br />

+ * 0.91 0.92 0.91 0.92 0.92 0.92 0.92<br />

χ− 0.91 0.91 0.91 0.84 0.91 0.91 0.91<br />

χ + c<br />

χ<br />

0.86 0.87 0.86 0.88 0.88 0.87 0.88<br />

− GBP<br />

c<br />

χ<br />

0.86 0.86 0.86 0.86 0.86 0.86 0.86<br />

+ * 0.91 0.91 0.91 0.91 0.91 0.91<br />

χ− 0.93 0.92 0.84 0.93 0.92 0.93<br />

χ + c<br />

χ<br />

0.86 0.86 0.86 0.86 0.86 0.86<br />

− BrR<br />

c<br />

χ<br />

0.90 0.89 0.90 0.90 0.88 0.90<br />

+ * 0.91 0.92 0.92 0.92 0.92<br />

χ− 0.92 0.84 0.93 0.92 0.93<br />

χ + c<br />

χ<br />

0.86 0.87 0.87 0.87 0.87<br />

− CHFR<br />

c<br />

χ<br />

0.89 0.91 0.90 0.88 0.90<br />

+ * 0.91 0.91 0.91 0.91<br />

χ− 0.84 0.92 0.92 0.92<br />

χ + c<br />

χ<br />

0.86 0.86 0.86 0.86<br />

− CHY<br />

c<br />

χ<br />

0.89 0.89 0.88 0.89<br />

+ * 0.93 0.92 0.93<br />

χ− 0.84 0.84 0.84<br />

χ + c<br />

χ<br />

0.91 0.87 0.89<br />

− indoRu<br />

c<br />

χ<br />

0.90 0.88 0.90<br />

+ * 0.92 0.93<br />

χ− 0.92 0.93<br />

χ + c<br />

χ<br />

0.87 0.89<br />

− indRu<br />

c<br />

χ<br />

0.88 0.90<br />

+ * 0.92<br />

χ− 0.92<br />

χ + c<br />

0.87<br />

0.88<br />

χ − c<br />

Table 4.15: Estimated ˆχ +,− applying the non-parametric approach according to Poon,<br />

Rockinger, and Tawn given by: ˆχ +,− = Zk,N ·k<br />

N to exchange rates of various foreign currencies<br />

with the US$. The tail represents the most extreme 4% of the return values during a time<br />

interval ranging from 20.10.2002 to 10.04.2008 consisting of N = 2000 daily observations. ’c’<br />

means that tails were corrected for 0.5% of most extreme data for the estimation of L. <strong>Tail</strong><br />

indexes ˆν were calculated using Hill’s estimator.<br />

145


A Eur GBP BrR CHFR CHY indoRu indRu JPY<br />

A λU,m * 0.40 0.36 0.14 0.34 0.08 0.13 0.05 0.29<br />

λL,m 0.41 0.33 0.10 0.43 0.04 0.13 0.15 0.26<br />

λEV T<br />

U,m<br />

λEV T<br />

L,m<br />

0.40<br />

0.44<br />

0.36<br />

0.35<br />

0.14<br />

0.12<br />

0.34<br />

0.45<br />

0.07<br />

0.06<br />

0.12<br />

0.15<br />

0.05<br />

0.17<br />

0.29<br />

0.29<br />

Eur λU,m * 0.53 0.13 0.73 0.09 0.08 0.04 0.29<br />

λL,m 0.43 0.09 0.69 0.08 0.14 0.09 0.24<br />

λEV T<br />

U,m<br />

λEV T<br />

L,m<br />

0.52<br />

0.45<br />

0.12<br />

0.11<br />

0.72<br />

0.71<br />

0.09<br />

0.10<br />

0.07<br />

0.16<br />

0.04<br />

0.11<br />

0.29<br />

0.26<br />

GBP λU,m * 0.09 0.44 0.06 0.06 0.08 0.25<br />

λL,m 0.08 0.43 0.06 0.09 0.09 0.26<br />

λEV T<br />

U,m<br />

λEV T<br />

L,m<br />

0.09<br />

0.10<br />

0.44<br />

0.45<br />

0.06<br />

0.09<br />

0.06<br />

0.11<br />

0.07<br />

0.11<br />

0.25<br />

0.29<br />

BrR λU,m * 0.08 0.06 0.09 0.10 0.10<br />

λL,m 0.06 0.04 0.09 0.06 0.06<br />

λEV T<br />

U,m<br />

λEV T<br />

L,m<br />

0.07<br />

0.09<br />

0.06<br />

0.06<br />

0.09<br />

0.11<br />

0.10<br />

0.09<br />

0.10<br />

0.09<br />

CHFR λU,m * 0.08 0.10 0.05 0.36<br />

λL,m 0.08 0.11 0.08 0.31<br />

λEV T<br />

U,m<br />

λEV T<br />

L,m<br />

0.07<br />

0.10<br />

0.10<br />

0.14<br />

0.05<br />

0.10<br />

0.36<br />

0.34<br />

CHY λU,m * 0.04 0.10 0.05<br />

λL,m 0.04 0.05 0.09<br />

λEV T<br />

U,m<br />

λEV T<br />

L,m<br />

0.04<br />

0.06<br />

0.10<br />

0.07<br />

0.05<br />

0.11<br />

indoRu λU,m * 0.10 0.10<br />

λL,m 0.08 0.06<br />

λEV T<br />

U,m<br />

λEV T<br />

L,m<br />

0.10<br />

0.10<br />

0.10<br />

0.09<br />

indRu λU,m * 0.10<br />

λL,m<br />

0.11<br />

λEV T<br />

U,m<br />

λEV T<br />

0.10<br />

0.14<br />

L,m<br />

Table 4.16: Estimated ˆ λL,m, ˆ λU,m, ˆ λEV T<br />

L,m , and ˆ λU,m applying the non-parametric approaches<br />

according to Schmidt & Stadtmüller explained in section (3.4) to exchange rates of various<br />

foreign currencies with the US$. The tail represents the most extreme 4% of the return values<br />

during a time interval ranging from 20.10.2002 to 10.04.2008 consisting of N = 2000 daily<br />

observations.<br />

EV T<br />

146


4.3 Application to Synthetic Time Series<br />

To check for the bias of the parameters ˆ β and tail index ˆα and to get a sense of what<br />

the best result could be by confining resulting errors on the estimators to statistical<br />

errors in order to find lower error bounds, I implemented the different concepts to<br />

synthetic time series. The difficulty creating a synthetic sample on the one hand is<br />

that it should be as simple as possible enabling me to control everything and ensuring<br />

exact knowledge about the structure of the data and on the other hand not to simplify<br />

the model too much because we want it to become realistic.<br />

I approximated the real return time series by a distribution consisting of two parts:<br />

the distribution body was modeled Gaussian and the tails up to a certain threshold were<br />

modeled by heavy tailed distribution functions. The probability density distribution<br />

function is given by:<br />

�<br />

N(µ, σ) if |y| < Yk,N<br />

p(y) = α<br />

y1+α if |y| ≥<br />

(4.1)<br />

Yk,N<br />

with Yk,N denoting the threshold value or the least extreme return value still counted<br />

to the tails.<br />

Now I will explain the implementation of the distribution 6 : First I created two<br />

samples of N uniformly distributed random values on the interval (0, 1) denoted by<br />

’u1’ and ’u2’, whereas N denotes the sample size. Then I sorted the two samples of<br />

uniformly distributed random values in descending order (’u1desc’ and ’u2desc’) and<br />

applied the inverse error function for transformation to standard normal distributed<br />

values N(0,1) with zero mean and variance equal to one given by equation:<br />

z1 = √ 2 · erfinv(2 · u1desc − 1) (4.2)<br />

z2 = √ 2 · erfinv(2 · u2desc − 1) (4.3)<br />

Now I built the tail samples by first creating four samples of Z uniformly distributed<br />

random values on the interval (0, 1) with Z = [0.04 · N] and then transforming the<br />

uniformly distributed tail values U(0,1) to heavy tail distributed values by applying<br />

relation: � �<br />

1 1/α<br />

to get positive tails and by multiplying the same relation by factor<br />

1−x<br />

(−1) for negative tails with α denoting the tail index of the heavy tails. The heavy tailed<br />

parts where then sorted by decreasing order and scaled in a way that the least extreme<br />

values still counted to the tails were equal to the most extreme values counted to the<br />

body of the respective distribution. Then the tails of the standard normal distributed<br />

samples ’u1desc’ and ’u2desc’ were replaced by the heavy tails, scaled to typical size<br />

in order to obtain synthetic return series that are comparable to original return series,<br />

and named Y and ε denoting vectors of index returns and idiosyncratic noise. Error<br />

terms ε calculated for real time series were typically of similar size as respective return<br />

values. Therefore the scaling of the two vectors Y and ε was performed equally. As<br />

a next step asset return vector X had to be built by Y , ε, and a predefined factor<br />

β applying the linear single factor model: X = β · Y + ε, whereas firstly the vectors<br />

were brought back into a random sequence by uniform sampling without replacement<br />

to avoid violation of the i.i.d (independent and identically distributed) criterion. Then<br />

after comparison of the synthetic samples with real historic return series, concepts to<br />

6 the m-file for the estimation of tail dependence coefficients on synthetic samples is enclosed to the appendix<br />

and to the data CD and denoted by: synthdata all.m<br />

147


estimate tail dependence were applied. Figure (4.1) shows non-parametric probability<br />

density distribution functions of the real index S&P 500 and asset CVX (Chevron Corp.)<br />

return series on a time interval ranging from July 1985 to Mars 2008 representing our<br />

smaller reference samples and a synthetic sample created by the m-file described within<br />

this section. Both were estimated by a Gaussian box kernel described in subsection<br />

(3.4.3) and the data sets consist of N = 2507 daily observations. We can see that the<br />

tails of the synthetic distribution were chosen similar to the tails of index S&P 500,<br />

whereas some assets show a slightly different tail behavior. As we are interested in the<br />

tails of the distributions I also plotted the distribution on a semi-log scale to the base<br />

ten shown in figure (4.2). First of all we notice the bumps caused by the Gaussian<br />

box kernel. The three bumps on the left that can be observed on the non-parametric<br />

probability density distribution function of asset CVX represent the most extreme<br />

outliers and therefore are reasonable. Looking at the non-parametric probability density<br />

distribution function of the sample of the synthetic time series plotted in red we observe<br />

that it lies in-between the two real data distributions. This is exactly what we intended.<br />

The transition from the Gaussian part to the heavy tailed part of the distribution seems<br />

to be as smooth as on the example of real time series and should not generate problems<br />

for the calculation of tail dependence coefficient estimates. Random returns Xi and<br />

random errors εi have the same distribution and tails modeled by the same tail index<br />

α as it is requested for the concepts according to Sornette & Malevergne.<br />

Implementing first the concept of non-parametric tail dependence, estimated by β<br />

calculated for all data, to synthetic time series we determine that there is no significant<br />

bias created by the estimation of β. This can be seen on table (4.17) showing that<br />

the maximum bias of the β-estimator, calculated as an average value performing 1000<br />

samples with identical inputs in dependence of the size of the original β that was predefined<br />

for the synthetic samples, was always smaller than 0.5% of the original value<br />

βorig. Additionally to show the effect on ˆ λ +,− of the bias created by ˆ β, tail dependence<br />

estimators calculated by the original β denoted by ˆ λ +,− (βorig) were compared<br />

�<br />

to tail<br />

ˆβalldata<br />

dependence estimators in dependence of estimated β denoted by ˆ λ +,−<br />

�<br />

. Only<br />

for small ˆ λ +,−<br />

� �<br />

ˆβalldata the bias is significantly different from zero and reaches a maximum<br />

of two percent of the estimate. It has to be added that the bias is independent<br />

of the size of the synthetic sample and that tail index α was always set equal to three<br />

to avoid a second source of errors. An estimation of error bars is also provided in table<br />

(4.17) by quoting standard deviations and 95% quantiles within the 1000 samples.<br />

Standard deviations of ˆ βalldata and ˆ λ +,−<br />

�<br />

resulting from ˆ β compaired with the<br />

� ˆβall data<br />

95% quantiles show that errors seem to be approximately Gaussian distributed.<br />

Using βSI calculated by β-smile conditions yields surprising results: as we can observe<br />

in table (4.18) there is no bias in the estimation of β when we apply the second<br />

β-smile condition given by: � Y ≥ Y (k) � to 1000 synthetic samples with identical inputs<br />

in dependence of the size of the original β. This is surprising insofar as it only<br />

counts for the second β-smile condition. The first and the third β-smile conditions both<br />

yield β-estimates that are strongly biased. As discussed in section (3.3) we still face<br />

the problem of wide error bars for tail dependence coefficient estimates performed by<br />

β-smile conditions. Even if βSI 2 is not biased itself, table (4.18) shows that primarily<br />

for small βorig we obtain a significant bias in ˆ λ +,− with ˆ β estimated by the second βsmile<br />

condition. Also here tail index α was always set equal to three to avoid a second<br />

source of errors.<br />

148


Density<br />

Density<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Fitted synthetic time series<br />

Fitted S&P 500<br />

Fitted asset CVX<br />

0<br />

−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2<br />

Data<br />

Figure 4.1: Non-parametric probability density distribution functions of index S&P 500, assets<br />

CVX (Chevron Corp.), and a synthetic sample estimated using a Gaussian box kernel for a<br />

time interval of N = 2507 data points ranging from January 1991 to December 2000.<br />

10 0<br />

10 −5<br />

10 −10<br />

10 −15<br />

10 −20<br />

10 −25<br />

10 −30<br />

Fitted synthetic time series<br />

Fitted S&P 500<br />

Fitted asset CVX<br />

−0.8 −0.6 −0.4 −0.2 0 0.2<br />

Data<br />

Figure 4.2: Non-parametric probability density distribution functions of index S&P 500, assets<br />

CVX (Chevron Corp.), and a synthetic sample plotted on a semi-log scale and estimated using<br />

a Gaussian box kernel for a time interval of N = 2507 data points ranging from January 1991<br />

to December 2000.<br />

149


150<br />

βorig<br />

ˆβall data bias rel. bias std 95% quantile<br />

+,−<br />

λ ˆ (βorig)<br />

+,−<br />

λ ˆ ( ˆ α = 3<br />

βall data) bias rel. bias std 95% quantile<br />

0.05 0.05 0.00 0.00 0.02 0.04 0.00 0.01 0.01 inf 0.10 0.01<br />

0.10 0.10 0.00 0.00 0.02 0.03 0.00 0.00 0.00 0.00 0.00 0.00<br />

0.20 0.20 0.00 0.00 0.02 0.03 0.01 0.01 0.00 0.01 0.00 0.00<br />

0.30 0.30 0.00 0.00 0.01 0.01 0.02 0.02 0.00 0.02 0.01 0.01<br />

0.40 0.40 0.00 0.00 0.02 0.03 0.05 0.05 0.00 0.00 0.01 0.02<br />

0.50 0.50 0.00 0.00 0.02 0.03 0.09 0.09 0.00 0.00 0.01 0.03<br />

0.60 0.60 0.00 0.00 0.02 0.03 0.13 0.13 0.00 0.00 0.02 0.04<br />

0.70 0.70 0.00 0.00 0.02 0.03 0.18 0.18 0.00 0.00 0.03 0.04<br />

0.80 0.80 0.00 0.00 0.02 0.04 0.23 0.23 0.00 0.00 0.03 0.05<br />

0.90 0.90 0.00 0.00 0.02 0.03 0.28 0.28 0.00 0.00 0.03 0.06<br />

1.00 1.00 0.00 0.01 0.02 0.03 0.33 0.33 0.00 0.00 0.04 0.06<br />

1.10 1.10 0.00 0.00 0.02 0.03 0.38 0.38 0.00 0.00 0.04 0.06<br />

1.20 1.20 0.00 0.00 0.02 0.03 0.42 0.42 0.00 0.00 0.04 0.07<br />

1.30 1.30 0.00 0.00 0.02 0.03 0.46 0.46 0.00 0.00 0.04 0.07<br />

Table 4.17: Establishing bias and error bars of ˆ β calculated by all data and effect on ˆ λ +,− estimated by applying the non-parametric approach<br />

according to Sornette & Malevergne to 1000 synthetic samples consisting of N = 2507 data points. The bias is calculated by comparing mean<br />

values ˆ βall data and ˆ λ +,−<br />

( ˆ βall data) to original values denoted by βorig and corresponding tail dependence estimator ˆ λ +,− (βorig) and for the estimation<br />

of error bars standard deviation ’std’ and 95% quantiles of the estimates are provided. Relative bias (’rel.bias’) is calculated as percentage of<br />

ˆλ +,−<br />

(βorig). <strong>Tail</strong> indexes α were set equal to three to avoid a second source of error and ’inf’ means ·/0.


151<br />

βorig<br />

ˆβSI 2 bias rel. bias std 95% quantile<br />

+,−<br />

λ ˆ (βorig)<br />

+,−<br />

λ ˆ ( ˆ α = 3<br />

βSI 2) bias rel. bias std 95% quantile<br />

0.05 0.05 0.00 0.00 0.11 0.17 0.00 0.28 0.28 * 0.45 0.71<br />

0.10 0.10 0.00 0.00 0.11 0.15 0.00 0.12 0.12 * 0.32 0.87<br />

0.20 0.20 0.00 0.00 0.10 0.16 0.01 0.04 0.03 * 0.17 0.02<br />

0.30 0.30 0.00 0.00 0.10 0.16 0.02 0.04 0.02 0.62 0.08 0.05<br />

0.40 0.40 0.00 0.00 0.10 0.16 0.05 0.06 0.01 0.20 0.06 0.07<br />

0.50 0.50 0.00 0.00 0.10 0.16 0.08 0.09 0.01 0.13 0.06 0.10<br />

0.60 0.06 0.00 0.00 0.11 0.17 0.13 0.14 0.01 0.08 0.08 0.12<br />

0.70 0.70 0.00 0.00 0.11 0.16 0.18 0.19 0.01 0.06 0.08 0.14<br />

0.80 0.80 0.00 0.00 0.11 0.17 0.23 0.23 0.00 0.00 0.10 0.17<br />

0.90 0.90 0.00 0.00 0.11 0.17 0.28 0.28 0.00 0.00 0.10 0.17<br />

1.00 1.00 0.00 0.00 0.11 0.16 0.33 0.33 0.00 0.00 0.11 0.19<br />

1.10 1.10 0.00 0.00 0.10 0.16 0.38 0.39 0.01 0.03 0.11 0.19<br />

1.20 1.20 0.00 0.00 0.10 0.16 0.42 0.43 0.01 0.02 0.12 0.19<br />

1.30 1.30 0.00 0.00 0.10 0.17 0.46 0.47 0.01 0.02 0.11 0.20<br />

Table 4.18: Establishing bias and error bars of ˆ βSI 2 calculated by second β-smile condition: � Y ≥ Y (k) � and effect on ˆ λ +,− estimated by applying<br />

the non-parametric approach according to Sornette & Malevergne to 1000 synthetic samples consisting of N = 2507 data points. The bias is<br />

calculated by comparing mean values ˆ βalldata and ˆ λ +,−<br />

( ˆ βall data) to original values denoted by βorig and corresponding tail dependence estimator<br />

ˆλ +,− (βorig) and for the estimation of error bars standard deviation ’std’ and 95% quantiles of the estimates are provided. <strong>Tail</strong> indexes α were set<br />

equal to three to avoid a second source of error and ∗ means that relative bias (’rel.bias’) calculated as percentage of ˆ λ +,−<br />

(βorig) ≥ 100%.


152<br />

α ˆν bias rel. bias std 95% quantile ˆ b γ n bias rel. bias std 95% quantile<br />

1.50 1.53 0.03 0.02 0.16 0.27 1.50 0.00 0.00 0.21 0.35<br />

2.00 2.05 0.05 0.03 0.21 0.34 2.02 0.02 0.01 0.29 0.51<br />

2.50 2.56 0.06 0.02 0.26 0.44 2.53 0.03 0.01 0.36 0.65<br />

3.00 3.07 0.07 0.02 0.31 0.52 2.99 0.01 0.03 0.41 0.76<br />

3.50 3.57 0.07 0.02 0.38 0.62 3.50 0.00 0.00 0.49 0.81<br />

4.00 4.07 0.07 0.02 0.46 0.79 3.98 0.02 0.01 0.59 1.05<br />

4.50 4.59 0.09 0.02 0.45 0.80 4.50 0.00 0.00 0.60 1.00<br />

5.00 5.12 0.12 0.02 0.53 0.95 5.04 0.04 0.01 0.73 1.29<br />

Table 4.19: Establishing bias and error bars of ˆν and ˆb γ n, estimated for 1000 synthetic samples consisting of N = 2507 data points. The bias is<br />

calculated by comparing mean values ˆν and ˆb γ<br />

n to original values denoted by α and for the estimation of error bars standard deviation ’std’ and<br />

95% quantiles of the estimates are provided. Relative bias (’rel.bias’) is calculated as percentage of α.


153<br />

βorig<br />

ˆλ +,−<br />

(α,βorig)<br />

+,−<br />

λ ˆ (ˆν, ˆ βall data) bias rel. bias std 95% quantile<br />

+,−<br />

λ ˆ (ˆν, ˆ α = 3<br />

βSI 2) bias rel. bias std 95% quantile<br />

0.05 0.00 0.01 0.01 * 0.08 0.01 0.27 0.27 * 0.44 0.73<br />

0.10 0.00 0.00 0.00 0.30 0.00 0.00 0.12 0.12 * 0.32 0.88<br />

0.20 0.01 0.01 0.00 0.05 0.01 0.01 0.05 0.04 * 0.18 0.03<br />

0.30 0.02 0.02 0.00 0.01 0.01 0.02 0.03 0.01 0.39 0.05 0.05<br />

0.40 0.05 0.05 0.00 0.01 0.02 0.04 0.06 0.01 0.21 0.05 0.08<br />

0.50 0.09 0.08 0.00 0.00 0.03 0.06 0.09 0.01 0.11 0.07 0.11<br />

0.60 0.13 0.12 0.00 0.02 0.04 0.07 0.14 0.01 0.08 0.07 0.14<br />

0.70 0.18 0.17 0.00 0.01 0.04 0.07 0.19 0.01 0.06 0.09 0.16<br />

0.80 0.23 0.23 0.00 0.00 0.05 0.08 0.24 0.01 0.06 0.10 0.16<br />

0.90 0.28 0.27 0.00 0.00 0.06 0.10 0.28 0.01 0.02 0.11 0.19<br />

1.00 0.33 0.32 0.01 0.02 0.06 0.10 0.33 0.01 0.02 0.12 0.20<br />

1.10 0.38 0.37 0.00 0.00 0.06 0.10 0.38 0.00 0.01 0.12 0.18<br />

1.20 0.42 0.42 0.01 0.01 0.06 0.11 0.42 0.00 0.00 0.12 0.20<br />

1.30 0.46 0.46 0.01 0.01 0.06 0.10 0.47 0.01 0.01 0.13 0.21<br />

Table 4.20: Establishing bias and error bars of ˆ λ +,− (ˆν, ˆ βall data) with ˆ β calculated by all data and ˆ λ +,− (ˆν, ˆ βSI 2) with ˆ β calculated by second<br />

β-smile condition: � Y ≥ Y (k) � by applying the non-parametric approach according to Sornette & Malevergne to 1000 synthetic samples consisting<br />

of N = 2507 data points. The bias is calculated by comparing mean values ˆ λ +,−<br />

(ˆν, ˆ βall data) and ˆ λ +,−<br />

(ˆν, ˆ βSI 2) to original values denoted by<br />

ˆλ +,−<br />

(α,βorig) and for the estimation of error bars standard deviation ’std’ and 95% quantiles of the estimates are provided. ∗ means that relative<br />

bias (’rel.bias’) calculated as percentage of ˆ λ +,−<br />

(βorig) ≥ 100%.


We also find a bias of Hill’s estimator denoted by ˆν reported in table (4.19). Com-<br />

pared to Gabaix’s estimator denoted by ˆb γ n the bias of the Hill estimator is significantly<br />

bigger but ˆb γ n instead has a wider error bar.<br />

In table (4.20) results of ˆ λ +,−<br />

(ˆν, ˆ βalldata) and ˆ λ +,−<br />

(ˆν, ˆ βSI2) estimated as mean values<br />

of 1000 synthetic samples with identical inputs are shown. We notice that the bias of the<br />

tail dependence estimators does not significantly change when we calculate the tail index<br />

by the Hill estimator compared to results reported in tables (4.17) and (4.18), where we<br />

used correct α. Error bars instead become larger because as shown in (4.19) tail index<br />

estimators add another source of uncertainty to tail dependence estimators. Anyway<br />

the calculation of Hill’s estimator is faster and more simple compared to Gabaix’s<br />

estimator and as tail dependence estimators are not sensitive to moderate changes of<br />

tail index α but are very sensitive to changes in β, Hill’s estimator might be accurate<br />

enough here because on synthetic time series bias resulting from the Hill estimator has<br />

negligible effect on the tail dependence coefficient estimates.<br />

We conclude that, applied to our synthetic time series, which are generally built of<br />

Gaussian distributions with tails modeled by heavy tailed distribution parts, concepts<br />

according to Sornette & Malevergne with ˆ β calculated by all data allow to estimate<br />

λ +,− in a accurate way showing us the best solution that we can expect. Furthermore<br />

we found that the second β-smile condition performs well for βorig ≥ 0.5 with respect<br />

to relative bias as well as error bars.<br />

Now, for comparison, I implemented concepts according to Poon, Rockinger, and<br />

Tawn and concepts according to Schmidt & Stadtmüller to synthetic samples. For<br />

these concepts there is no need to calculate any linear factor, but for example we<br />

can compare the results to the tail dependence estimators according to Sornette &<br />

Malevergne applying their non-parametric approach and compare the resulting error<br />

bars to error bars estimated by bootstrap sampling with replacement. Table (4.21)<br />

shows tail dependence coefficients estimated by the different concepts including error<br />

bars calculated for 1000 synthetic samples compared to their average values. <strong>Tail</strong><br />

indexes were calculated by Hill’s estimator. For comparison I added βorig and ˆ λ +,− (βorig)<br />

with tail indexes also calculated by Hill’s estimator. Respective standard deviations<br />

and 95% quantiles, also given in table (4.21), show us that error bars resulting from<br />

both factors, ˆν and ˆ β are around double size of errors bars just resulting from ˆ β given<br />

in table (4.17) and of same extent as error bars estimated by bootstrap sampling with<br />

replacement in subsection (3.2.3). For the methods according to Schmidt & Stadtmüller<br />

a relation for the calculation of the estimates’ variance σ2 L (θ) was proposed in their<br />

paper [24] (2005) that was already discussed in subsection (3.4.3). Applied to real<br />

historical data of index S&P 500 and corresponding assets the error bar estimates given<br />

by relation (3.92) were high compared to bootstrap quantiles. Compared to the error<br />

bars calculated for the 1000 synthetic samples the proposed estimator works better.<br />

σL(θ) are close to standard deviations ’std’ calculated for the 1000 samples compared<br />

to their average values. Another standard deviation estimator ˆσ, which was introduced<br />

in the paper [13] (2004) to provide error bar estimates for ˆχ +,− seems significantly lower<br />

than bootstrap estimates of real historical data provided in subsection (3.1.2) as well as<br />

standard deviations calculated for the 1000 samples compared to their average values<br />

for synthetic samples provided in table (4.21). Looking at total values of the estimates,<br />

the different estimators still yield mostly different results. In terms of error bars the<br />

three presented methods all yield reasonable results that seem to be approximately<br />

Gaussian distributed and agree well with results presented in chapter (3).<br />

154


155<br />

βorig<br />

ˆλ +,− EV T<br />

(ˆν,βall data) std 95% quantile λ·,m<br />

ˆ std 95% quantile λˆ ·,m std 95% quantile ˆχ ˆσ std 95% quantile<br />

0.05 0.01 0.08 0.01 0.05 0.02 0.04 0.05 0.02 0.04 0.87 0.09 0.13 0.06<br />

0.1 0.00 0.00 0.00 0.06 0.02 0.04 0.06 0.02 0.04 0.87 0.09 0.14 0.07<br />

0.2 0.01 0.01 0.01 0.10 0.03 0.05 0.10 0.03 0.05 0.87 0.09 0.14 0.07<br />

0.3 0.02 0.01 0.02 0.15 0.03 0.05 0.15 0.03 0.05 0.87 0.08 0.14 0.06<br />

0.4 0.05 0.02 0.04 0.19 0.04 0.06 0.19 0.06 0.05 0.87 0.09 0.14 0.07<br />

0.5 0.08 0.03 0.06 0.24 0.04 0.06 0.25 0.04 0.06 0.88 0.09 0.15 0.07<br />

0.6 0.12 0.04 0.07 0.28 0.04 0.06 0.28 0.04 0.06 0.88 0.09 0.14 0.06<br />

0.7 0.17 0.04 0.07 0.32 0.04 0.07 0.32 0.04 0.07 0.88 0.09 0.13 0.06<br />

0.8 0.23 0.05 0.08 0.36 0.04 0.07 0.36 0.04 0.07 0.89 0.09 0.12 0.06<br />

0.9 0.27 0.06 0.10 0.40 0.04 0.07 0.40 0.04 0.07 0.89 0.09 0.13 0.06<br />

1 0.32 0.06 0.10 0.43 0.04 0.07 0.43 0.04 0.07 0.89 0.09 0.12 0.05<br />

1.1 0.37 0.06 0.10 0.46 0.04 0.07 0.46 0.04 0.07 0.89 0.09 0.12 0.05<br />

1.2 0.42 0.06 0.11 0.49 0.04 0.07 0.49 0.04 0.07 0.89 0.09 0.12 0.05<br />

1.3 0.46 0.06 0.10 0.51 0.04 0.07 0.52 0.04 0.07 0.89 0.09 0.12 0.05<br />

Table 4.21: Comparison of estimates by non-parametric estimators ˆ λ·,m, ˆ EV T<br />

λ·,m according to Schmidt & Stadtmüller, and ˆχ+,− according to Poon,<br />

Rockinger, and Tawn with ˆ λ +,− (βall data) according to Sornette & Malevergne by applying the different concepts to 1000 synthetic samples consisting<br />

of N = 2507 data points created by different βorig. For the estimation of error bars standard deviation ’std’ and 95% quantiles of the estimates<br />

are calculated and for ˆχ +,− the proposed standard deviation ˆσ is given for comparison. <strong>Tail</strong> indexes α were estimated by the Hill estimator.


Chapter 5<br />

Concluding Remarks<br />

Now we summarize the findings of chapter (4) and finally draw some conclusions for<br />

practical use of the different concepts.<br />

In section (4.1) we have seen that all tail dependence coefficients ˆ λ +,− estimated by<br />

the non-parametric approach and by the parametric approach according to Sornette<br />

& Malevergne for assets and indexes traded outside the U.S. were vanishing while all<br />

estimates of the crucial parameter ˆ β were comparably small. This seems implausible.<br />

Furthermore we found on U.S. assets and indexes that for ˆ β < 0.2 tail dependence<br />

coefficient estimates ˆ λ → 0. However, for index AORD (ASX) and included assets<br />

results estimated by concepts according to Schmidt & Stadtmüller reported in table<br />

(4.11) look reasonable. To achieve diversification of extreme financial risks within a<br />

portfolio we are looking for small or vanishing tail dependence coefficients, which we<br />

expect to be rare.<br />

In contrast, as shown in section (4.2), applied to exchange rates of various foreign<br />

currencies with the US$ results by the methods according to Sornette & Malevergne<br />

were more reasonable. Concepts according to Sornette & Malevergne yielded mostly<br />

vanishing results while results by concepts according to Schmidt & Stadtmüller were<br />

all close together and around 0.9. Approaches according to Poon, Rockinger, and Tawn<br />

again yielded nearly identical results for all cases.<br />

In section (4.3) we found that the non-parametric approach according to Sornette<br />

& Malevergne with ˆ β estimated for all data applied to synthetic time series succeeded<br />

in providing estimates that were unbiased and of satisfying accuracy. This at least<br />

provides us with lower error bounds (the optimum solution we can expect) and is a<br />

promising result for application of the concepts to real data. Estimation of tail index<br />

α by Hill’s estimator has shown a small bias on our synthetic sample surveys reported<br />

in table (4.19). But the impact on the final estimators, as shown in table (4.20), was<br />

marginal. Furthermore, for real data we can apply the Gabaix estimator to calculate<br />

tail index ˆα for comparison because in spite of wider error bars the Gabaix estimator<br />

should perform better than Hill’s estimator with respect to bias facing heteroscedastic<br />

time series.<br />

Anyway, at the example of synthetic time series we were assuming that assets and index<br />

have a common factor that is constant over the whole distribution interval, whereas<br />

by the β-smile improvement conditions we are generally assuming that dependence in<br />

the extreme parts of the distributions might be different compared to dependence in<br />

the body of the distributions. As we were able to calculate the common factors β only<br />

by tail data applying the second β-smile condition reported in table (4.18), we should<br />

always check on real data whether βalldata is close to βSI· to get a feeling about dependences<br />

between assets and index. Applying the β-smile improvement conditions to<br />

156


synthetic time series we detected a strong bias in the first and in the third SI conditions.<br />

The second SI condition has shown to perform well for (β > 0.5) with respect to bias.<br />

Tables (4.6), (4.8), and (4.10) show ˆ β-values calculated for all data on historical time<br />

series that are much bigger than all of our synthetic sample estimates ˆ β reported in<br />

table (4.17), which all fulfill ˆ β < 0.5. This might indicate that the bias of historic time<br />

series could be different from the bias of synthetic time series. Now we come to the final<br />

comparison of the different concepts on synthetic time series: in table (4.21) we see<br />

all the different concepts including error bar estimates applied to synthetic time series.<br />

Estimates by the different methods are mostly inconsistent as it was also the case for<br />

historic time series. Whereas ˆχ is always big and within a narrow range, the two other<br />

estimators ˆ λ +,− and ˆ λ·,m are on a wider range but still different from each other. What<br />

all methods have in common is that they are increasing in increasing β, whereas ˆ λ·,m<br />

tends to be slightly higher than ˆ λ +,− primarily for small ˆ β. This is consistent with<br />

real data results i.e. reported in section (4.1) where the different concepts were applied<br />

to major financial centers in the world. Comparing estimates of λ +,− to estimates of<br />

λ·,m on the historical reference data series of the index S&P 500 and the nine assets<br />

provided in tables (3.8) and (3.10) for the Sornette & Malevergne estimators and in<br />

tables (3.24) and (3.25) for the Schmidt & Stadtmüller estimators we detected that<br />

within different ranges the relative proportions agree in many cases. Asset Texas Instruments<br />

Inc. (TXN) is the smallest observation in positive and negative tails observed<br />

by both methods and sample sizes and Hewlett-Packard Co. (HPQ) and Coca-Cola Co.<br />

(KO) both belong to the biggest observations in the positive and negative tails also for<br />

both reference data sets. There are many of these conformities showing us that the<br />

two estimators qualitatively point into the same direction. As it is our purpose to find<br />

differences between the tail dependences of the different assets and the index for small<br />

ˆβ, we should always calculate λ·,m according to Schmidt & Stadtmüller for comparison<br />

because λ +,− according to Sornette & Malevergne mostly yield zero tail dependence in<br />

that case.<br />

It cannot be concluded that one of the presented concepts for the estimation of<br />

tail dependence outperforms the others by all criteria. Anyway, having the methods<br />

on our hands that were discussed within this report, I would recommend to apply<br />

all of the available concepts for the estimation of the tail dependence coefficient to<br />

various data representing alternative choices out of the field of interest and then draw<br />

conclusions considering the differences between the estimates of alternative portfolio<br />

choices. Having a pool of alternatives evaluated by the different concepts might allow<br />

the investor to pick out the method that fits best to his respective purpose and helps<br />

him to get an intuition about dependences between assets and indexes over the whole<br />

range of the distribution. Finally it might enable him to choose between the considered<br />

alternatives.<br />

The coefficient of tail dependence provides the investor with a tool to minimize<br />

extreme correlations within his portfolio, which first of all should help him to endure<br />

bearish markets by optimized diversification criteria. Globalization might have lead<br />

to increasing extreme dependences across and within economies making extreme lefttail<br />

events more and more likely to happen. To be responsive to this movements the<br />

coefficient of tail dependence represents a simple and concrete extension to financial<br />

risk management.<br />

157


Bibliography<br />

[1] Sibuya, M. (1960). Bivariate extreme statistics. Ann. Inst. Statist. Math. 11, 195-<br />

210.<br />

[2] Markovitz, H.M. (1952). Portfolio Selection. Journal of Finance 7, 77-91.<br />

[3] Salvadori, G., De Michele, C., Kottegoda, N.T. and Rosso R. (2007). Extremes in<br />

Nature, An Approach Using Copulas. Springer, Netherlands.<br />

[4] De Haan, L., Ferreira, A. (2006). Extreme Value Theory, An Introduction.<br />

Springer, New York.<br />

[5] Castillo, E (1988). Extreme Value Theory in Engineering. Academic Press, San<br />

Diego, California.<br />

[6] Kotz, S. and Nadarajah, S. (2000). Extreme Value Distributions, Theory and Applications.<br />

Imperial College Press, London.<br />

[7] Nelsen, R.B. (2006). An Introduction to Copulas. Springer, New York, second<br />

edition.<br />

[8] Drouet-Mari, D. and Kotz, S. (2001). Correlation and <strong>Dependence</strong>. Imperial College<br />

Press, London.<br />

[9] Joe, H. (1997). Multivariate Models and <strong>Dependence</strong> Concepts. Chapman & Hall,<br />

London.<br />

[10] Marshall, A. and Olkin, I. (1983). Domains of attraction of multivariate Extreme<br />

Value distributions. Ann. Prob., 11 (1), 168177.<br />

[11] Slar, A. (1959). Fonctions de répartition à n dimensions et leurs marges. Publ. Inst.<br />

Statist. Univ. Paris, 8, 229-231.<br />

[12] Ledford, A.W. and Tawn, J.A. (1996). Statistics for near independence in multivariate<br />

extreme dependence. Biometrika 83, 169-187.<br />

[13] Poon, S-H., Rockinger, M. and Tawn, J. (2004). Extreme Value <strong>Dependence</strong> in<br />

Financial Markets: Diagnostics, Models, and Financial Implications. The Review<br />

of Financial Studies 17 (2), 581-610.<br />

[14] Heffernan, J.E. (2000). A Directory of <strong>Tail</strong> <strong>Dependence</strong>. Extremes 3, 279-290.<br />

[15] Coles, S.G., Heffernan, J. and Tawn, J.A. (1999). <strong>Dependence</strong> Measures for Extreme<br />

Value Analyses. Extremes 3, 5-38.<br />

[16] Box, G.E.P., Jenkins, G.M. and Reinsel, G.C. (1994). Time Series Analysis,<br />

Forecasting and Control. 3rd ed. Prentice Hall, Englewood Clifs, New Jersey.<br />

158


[17] Engle, R.F. (1982). Autoregressive Conditional Heteroskedasticity with Estimates<br />

of Variance of United Kingdom Inflation. Econometrica 50, 987-1008.<br />

[18] Zivot, E. and Wang, J. (2007). Modeling Financial Time Series with S-Plus. 2nd<br />

ed. Springer, New York.<br />

[19] Malevergne, Y. and Sornette, D. (2006). Extreme Financial <strong>Risks</strong>, From <strong>Dependence</strong><br />

to Risk Management. Springer, New York.<br />

[20] Malevergne, Y. and Sornette, D. (2004). How to account for extreme co-movements<br />

between individual stocks and the market. J. Risk 6, 71-116.<br />

[21] Malevergne, Y. and Sornette, D. (2002). Minimizing extremes. Risk 15, 129-132.<br />

[22] Kearns, P. and Pagan, A. (1997). Estimating the density tail index for financial<br />

time series. Review of Economics & Statistics 79, 171-175.<br />

[23] Gabaix, X. and Ibragimow, R. (2006). Rank-1/2: A simple Way to improve the<br />

OLS Estimation of <strong>Tail</strong> Exponents. 1-31.<br />

[24] Schmidt, R. and Stadtmüller, U. (2005). Non-parametric Estimation of <strong>Tail</strong> <strong>Dependence</strong>.<br />

Board of the Foundation of the Scandinavian Journal of Statistics, USA<br />

33, 307-335.<br />

[25] Huang, X. (1992). Statistics of bivariate extreme values, PhD thesis. Tinbergen Institute<br />

Research Series 22, Thesis Publishers and Tinbergen Institute Rotterdam.<br />

[26] Peng, L. (1998). Second order condition and extreme value theory. PhD thesis,<br />

Tinbergen Institute Research Series 178, Thesis Publishers and Tinbergen Institute<br />

Rotterdam.<br />

[27] Drees, H. and Huang, X. (1998). Best attainable rates of convergence for estimates<br />

of the stable tail dependence function. J. Multivariate Anal. 64, 109-126.<br />

[28] Klüppelberg, C. and Resnick, S. (2008). The Pareot Copula, Aggregation of risks,<br />

and the emperor’s socks. J. Appl. Prob. 45, 67-84.<br />

159


Appendix<br />

M-Files<br />

In the first part of the appendix Matlab m-files are provided. Comments are parenthesized<br />

in: %...%, which is also read as a comment by the Matlab solver. Implementation<br />

works by copy and paste, whereas data sets should be provided in Excel format (xls or<br />

xlsx).<br />

Import of Data<br />

clear all;<br />

exl = actxserver(’excel.application’);<br />

exlWkbk = exl.Workbooks;<br />

exlFile = exlWkbk.Open([’%set path for opening%’]);<br />

nn=%number of rows%;<br />

n=%number of columns%;<br />

exlSheet=exlFile.Sheets.Item(’table1’); %choice of table%;<br />

dat_range = [’B3:e’ num2str(nn)]; %range in Excel file%;<br />

rngObj=exlSheet.Range(dat_range);<br />

exlData=rngObj.Value; %end of read-in%;<br />

Prices(:,:)=exlData;<br />

Prices=cell2mat(Prices); %define Prices as matrix%;<br />

for j=1:n;<br />

%for i=nn-2:-1:1;<br />

for i=nn-3:-1:1;<br />

Returns(i,j)=log(Prices(i,j))-log(Prices(i+1,j));<br />

end<br />

end %transform daily prices to returns%<br />

a=1;<br />

b=nn-3; %data range%<br />

Data can also be imported by:<br />

Prices=xlsread(’testdata.xls’, 1, ’A4:B5’) %1 denotes table1%<br />

But then the Excel-data sheet has to be copied into the folder where the m-file is safed.<br />

160


Approach according to Poon, Rockinger, and Tawn<br />

retdesc=sort(Returns(a:b,1:n),’descend’);<br />

retasc=sort(Returns(a:b,1:n)); %sort Returns%<br />

Z=roundn((b-a)*0.04,0); %define ’k’%<br />

ZZ=roundn(0.005*(b-a),0); %define ’c’%<br />

%calculate tail indexes by the Hill estimator%;<br />

for j=1:n;<br />

vk(j)=(1/(Z)*sum(log(retdesc(1:Z,j)))-log(retdesc(Z,j)))^-1;<br />

vka(j)=(1/(Z)*sum(log(-retasc(1:Z,j)))-log(-retasc(Z,j)))^-1;<br />

end<br />

%calculate l%<br />

t=linspace(1,Z,Z);<br />

p=size(t);<br />

p=p(2);<br />

for j=1:n;<br />

for i=1:p;<br />

f=t(i);<br />

d=@(f)(f/(b-a)*(retdesc(f,j))^vk(j));<br />

dd=@(f)(f/(b-a)*(-retasc(f,j))^vka(j));<br />

lpos(i,j)=d(f);<br />

lneg(i,j)=dd(f);<br />

end<br />

end<br />

lmk_pos=mean(lpos(1:Z,:));<br />

lmy_pos=mean(lpos(ZZ:Z,:));<br />

lmk_neg=mean(lneg(1:Z,:));<br />

lmy_neg=mean(lneg(ZZ:Z,:));<br />

%calculate S and T%<br />

for j=1:n-1;<br />

for i=1:Z;<br />

S(i,1)=-1/log(1-lmk_pos(1)*retdesc(i,1)^(-vk(1)));<br />

T(i,j)=-1/log(1-lmk_pos(j+1)*retdesc(i,j+1)^(-vk(j+1)));<br />

S_y(i,1)=-1/log(1-lmy_pos(1)*retdesc(i,1)^(-vk(1)));<br />

T_y(i,j)=-1/log(1-lmy_pos(j+1)*retdesc(i,j+1)^(-vk(j+1)));<br />

Sn(i,1)=-1/log(1-lmk_neg(1)*(-retasc(i,1))^(-vka(1)));<br />

Tn(i,j)=-1/log(1-lmk_neg(j+1)*(-retasc(i,j+1))^(-vka(j+1)));<br />

Sn_y(i,1)=-1/log(1-lmy_neg(1)*(-retasc(i,1))^(-vka(1)));<br />

Tn_y(i,j)=-1/log(1-lmy_neg(j+1)*(-retasc(i,j+1))^(-vka(j+1)));<br />

end<br />

end<br />

%calculate Z%<br />

for j=1:n-1;<br />

for i=1:Z;<br />

z(i,j)=min(S(i,1),T(i,j));<br />

z_y(i,j)=min(S_y(i,1),T_y(i,j));<br />

zn(i,j)=min(Sn(i,1),Tn(i,j));<br />

zn_y(i,j)=min(Sn_y(i,1),Tn_y(i,j));<br />

end<br />

end<br />

161


%estimate overline_chi and corresponding sigma%<br />

for j=1:n-1;<br />

ovchi(j)=2/Z*(sum(log(z(:,j)./z(Z,j))))-1;<br />

sigovchi(j)=(ovchi(j)+1)/sqrt(Z);<br />

ovchi_y(j)=2/Z*(sum(log(z_y(:,j)./z_y(Z,j))))-1;<br />

sigovchi_y(j)=(ovchi_y(j)+1)/sqrt(Z);<br />

ovchin(j)=2/Z*(sum(log(zn(:,j)./zn(Z,j))))-1;<br />

sigovchin(j)=(ovchin(j)+1)/sqrt(Z);<br />

ovchin_y(j)=2/Z*(sum(log(zn_y(:,j)./zn_y(Z,j))))-1;<br />

sigovchin_y(j)=(ovchin_y(j)+1)/sqrt(Z);<br />

end<br />

%output%<br />

ovchi<br />

sigovchi<br />

%estimate chi and corresponding sigma if overline_chi=1%<br />

for j=1:n-1;<br />

if abs(ovchi(j)-1)


Non-Parametric Approach according to Sornette & Malevergne<br />

retdesc=sort(Returns(a:b,1:n),’descend’);<br />

retasc=sort(Returns(a:b,1:n)); %sort Returns%<br />

Z=roundn((b-a)*0.04,0); %define ’k’%<br />

Y=roundn(0.005*(b-a),0); %define ’c’%<br />

kretdesc=retdesc(1:Z,:); %from 1 to ’k’;<br />

yretdesc=retdesc(Y:Z,:); %from ’c’ to ’k’;<br />

kretasc=retasc(1:Z,:);<br />

yretasc=retasc(Y:Z,:);<br />

%calculate beta for the whole data set%<br />

for q=1:n-1;<br />

coeff=polyfit(Returns(a:b,1),Returns(a:b,q+1),1);<br />

bbeta(q)=coeff(1);<br />

end<br />

%estimate tail indexes by the Hill estimator%<br />

vk=(1/(Z)*sum(log(kretdesc(:,1)))-log(kretdesc(Z,1)))^-1;<br />

vka=(1/(Z)*sum(log(-kretasc(:,1)))-log(-kretasc(Z,1)))^-1;<br />

%estimate ’l’ and lambda%<br />

for i=1:n-1;<br />

ly(i)=mean(yretdesc(:,i+1)./(yretdesc(:,1)));<br />

lk(i)=mean(kretdesc(:,i+1)./(kretdesc(:,1)));<br />

lya(i)=mean(yretasc(:,i+1)./(yretasc(:,1)));<br />

lka(i)=mean(kretasc(:,i+1)./(kretasc(:,1)));<br />

if(bbeta(i)>0)<br />

L_nonpar_y_up(i)=1/(max(1,ly(i)/bbeta(i)))^vk;<br />

L_nonpar_k_up(i)=1/(max(1,lk(i)/bbeta(i)))^vk;<br />

L_nonpar_y_lo(i)=1/(max(1,lya(i)/bbeta(i)))^vka;<br />

L_nonpar_k_lo(i)=1/(max(1,lka(i)/bbeta(i)))^vka;<br />

elseif(bbeta(i)


Parametric Approach according to Sornette & Malevergne<br />

retdesc=sort(Returns(a:b,1:n),’descend’);<br />

retasc=sort(Returns(a:b,1:n)); %sort Returns%<br />

Z=roundn((b-a)*0.04,0); %define ’k’%<br />

ZZ=roundn(0.005*(b-a),0); %define ’c’%<br />

%estimate tail indexes by the Hill estimator%<br />

vk=(1/(Z)*sum(log(retdesc(1:Z,1)))-log(retdesc(Z,1)))^-1<br />

vka=(1/(Z)*sum(log(-retasc(1:Z,1)))-log(-retasc(Z,1)))^-1<br />

%calculate beta for the whole data set%<br />

for q=1:n-1;<br />

coeff=polyfit(Returns(a:b,1),Returns(a:b,q+1),1);<br />

bbeta(q)=coeff(1);<br />

end<br />

%calculate error epsilon%<br />

for i=1:n-1;<br />

eps=Returns(:,i+1)-(bbeta(1,i)*Returns(:,1));<br />

Eps(:,i)=eps;<br />

end<br />

%estimate Cy and mean value of Cy for most extr. 4 per cent%<br />

t=linspace(1,Z,Z);<br />

p=size(t);<br />

p=p(2);<br />

for i=1:p;<br />

f=t(i);<br />

d=@(f)(f/(b-a)*(retdesc(f,1))^vk);<br />

dd=@(f)(f/(b-a)*(-(retasc(f,1)))^vka);<br />

cypos(i,1)=d(f);<br />

cyneg(i,1)=dd(f);<br />

end;<br />

cymk_pos=mean(cypos(1:Z));<br />

cymy_pos=mean(cypos(ZZ:Z));<br />

cymk_neg=mean(cyneg(1:Z));<br />

cymy_neg=mean(cyneg(ZZ:Z));<br />

%estimate Ceps and mean value of Ceps for most extr. 4 per cent%<br />

epsdesc=sort(Eps(a:b,:),’descend’);<br />

epsasc=sort(Eps(a:b,:));<br />

for i=1:n-1;<br />

for j=1:p;<br />

f=t(j);<br />

d=@(f)(f/(b-a)*(epsdesc(f,i))^vk);<br />

dd=@(f)(f/(b-a)*(-(epsasc(f,i)))^vka);<br />

cepos(j,i)=d(f);<br />

ceneg(j,i)=dd(f);<br />

end<br />

end<br />

cemk_pos=mean(cepos(1:Z,:));<br />

cemy_pos=mean(cepos(ZZ:Z,:));<br />

cemk_neg=mean(ceneg(1:Z,:));<br />

cemy_neg=mean(ceneg(ZZ:Z,:));<br />

164


%estimate lambda%<br />

for j=1:n-1;<br />

if(bbeta(j)>0)<br />

L_par_k_up(1,j)=(1+(bbeta(1,j)^-vk)*cemk_pos(1,j)/cymk_pos)^-1;<br />

L_par_y_up(1,j)=(1+(bbeta(1,j)^-vk)*cemy_pos(1,j)/cymy_pos)^-1;<br />

L_par_k_lo(1,j)=(1+(bbeta(1,j)^-vka)*cemk_neg(1,j)/cymk_neg)^-1;<br />

L_par_y_lo(1,j)=(1+(bbeta(1,j)^-vka)*cemy_neg(1,j)/cymy_neg)^-1;<br />

elseif(bbeta(j)


β-Smile Improvement Conditions<br />

A=Returns(a:b,1:n);<br />

retdesc=sort(A,’descend’);<br />

retasc=sort(A); %sort Returns%<br />

Z=roundn((b-a+1)*0.04,0); %define ’k’%<br />

Y=roundn(0.005*(b-a+1),0); %define ’c’%<br />

%Implementation of chosen condition and calculation of beta_SI%<br />

for u=1:n-1;<br />

h=1;<br />

ha=1;<br />

for i=1:b-a+1;<br />

if (A(i,1)>=retdesc(Z,1))&&(A(i,u+1)>=retdesc(Z,u+1));<br />

%if (A(i,1)>=retdesc(Z,1));<br />

%if (A(i,u+1)>=retdesc(Z,u+1));<br />

D(h,1)=A(i,1);<br />

D(h,2)=A(i,u+1);<br />

h=h+1;<br />

end<br />

if (A(i,1)


Approaches according to Schmidt & Stadtmüller<br />

A=Returns(a:b,1:n);<br />

p=size(A);<br />

B=sort(A);<br />

C=zeros(p(1,1),p(1,2));<br />

for k=1:p(1,2); along columns%<br />

pp=1;<br />

for i=1:p(1,1); %along rows%<br />

v=find(A(:,k)==B(i,k));<br />

t=size(v);<br />

t=t(1,1);<br />

if t>1;<br />

for z=1:t;<br />

C(v(z),k)=0.5-t/2+i;<br />

end<br />

else C(v,k)=i;<br />

v=[];<br />

end<br />

end<br />

end<br />

k=0.04*(b-a+1);<br />

sum_L(1,1:n-1)=0;<br />

sum_U(1,1:n-1)=0;<br />

sum_L_EVT(1,1:n-1)=0;<br />

sum_U_EVT(1,1:n-1)=0;%set sums to zero%<br />

%conditions for the different sums%<br />

for j=1:n-1;<br />

for i=1:b-a+1;<br />

if(C(i,1)(b-a+1)-k);<br />

sum_U(1,j)=sum_U(1,j)+1;<br />

end<br />

if(C(i,1)>(b-a+1)-k) || (C(i,j+1)>(b-a+1)-k);<br />

sum_U_EVT(1,j)=sum_U_EVT(1,j)+1;<br />

end<br />

if(C(i,1)


<strong>Tail</strong> Indexes by the Gabaix Estimator<br />

A=Returns(a:b,1:n);<br />

AA=sort(A,’descend’);<br />

B=sort(-A,’descend’); %sort Returns%<br />

p=size(A);<br />

C=NaN(p(1,1),p(1,2));<br />

c=NaN(p(1,1),p(1,2));<br />

for k=1:p(1,2); %along columns%<br />

for i=1:p(1,1); %along rows%<br />

v=find(-A(:,k)==B(i,k));%negative tail%<br />

vv=find(A(:,k)==AA(i,k));%positive tail%<br />

t=size(v);<br />

t=t(1,1);<br />

tt=size(vv);<br />

tt=tt(1,1);<br />

if t>1;<br />

for z=1:t;<br />

C(v(z),k)=0.5-t/2+i;%negative tail%<br />

end<br />

else C(v,k)=i;<br />

end<br />

if tt>1;<br />

for z=1:tt;<br />

c(vv(z),k)=0.5-tt/2+i;%positive tail%<br />

end<br />

else c(vv,k)=i;<br />

end<br />

end<br />

end<br />

Z=round(0.04*(b-a+1)); %define ’k’%<br />

Y=round(0.005*(b-a)); %define ’c’%<br />

AAA=log(AA(1:Z,:));<br />

BB=log(B(1:Z,:));<br />

CC=log(sort(C)-0.5);<br />

cc=log(sort(c)-0.5);<br />

for i=1:n;<br />

p=polyfit(AAA(:,i),cc(1:Z,i),1);%positive tail<br />

pp=polyfit(BB(:,i),CC(1:Z,i),1);%negative tail<br />

bnn(i)=-(pp(1,1));<br />

bn(i)=-(p(1,1));<br />

end<br />

%output%<br />

bn %upper tail indexes%<br />

bnn %lower tail indexes%<br />

168


Composition of Synthetic Data Set<br />

clear all<br />

for jj=1:1000;<br />

p=2507; %sample size%<br />

vvvv=3; %tail index%<br />

bbeta=0.9; %predefined beta%<br />

u1=rand(p,1);<br />

u2=rand(p,1);<br />

Z=round(0.04*p);<br />

u1desc=sort(u1,’descend’);<br />

u2desc=sort(u2,’descend’);<br />

%Box-Muller trafo;<br />

z1(:,1)=sqrt(2) *erfinv (2*u1desc - 1);<br />

z2(:,1)=sqrt(2) *erfinv (2*u2desc - 1);<br />

%here we have the standard normal distributed part for eps and x%<br />

u1desc_up(:,1)=rand(Z,1);<br />

u2desc_up(:,1)=rand(Z,1);<br />

u1desc_lo(:,1)=rand(Z,1);<br />

u2desc_lo(:,1)=rand(Z,1);<br />

for i=1:Z;<br />

z1_up(i,1)=(1/(1-u1desc_up(i,1)))^(1/vvvv);<br />

z2_up(i,1)=(1/(1-u2desc_up(i,1)))^(1/vvvv);<br />

z1_lo(i,1)=-((1/(1-u1desc_lo(i,1)))^(1/vvvv));<br />

z2_lo(i,1)=-((1/(1-u2desc_lo(i,1)))^(1/vvvv));<br />

end<br />

%heavy tail parts are created%<br />

%set the different distribution parts together into a single total%<br />

%distribution%<br />

z1_up=sort(z1_up,’descend’);<br />

z2_up=sort(z2_up,’descend’);<br />

z1_lo=sort(z1_lo,’descend’);<br />

z2_lo=sort(z2_lo,’descend’);<br />

z1_up=z1(101,1)/z1_up(Z,1)*z1_up;<br />

z2_up=z2(101,1)/z2_up(Z,1)*z2_up;<br />

z1_lo=z1(p-Z)/z1_lo(1,1)*z1_lo;<br />

z2_lo=z2(p-Z)/z2_lo(1,1)*z2_lo;<br />

z1(1:Z,1)=z1_up;<br />

z2(1:Z,1)=z2_up;<br />

z1(p-Z+1:p,1)=z1_lo;<br />

z2(p-Z+1:p,1)=z2_lo;<br />

y=z1;<br />

eps=z2;<br />

169


%define mu and sigma if requested%<br />

y_s=(0.05*10^(-3)+y(:).*0.02)*0.5;<br />

eps_s=(0.05*10^(-3)+eps(:).*0.02)*0.5;<br />

%now we have to build the y values by x, eps, and beta%<br />

%The problem now is that both x_s and eps_s are in descending order, which%<br />

%means that high return-values correspond to high eps-values. This is a%<br />

%violation of i.i.d. Therefore we bring back the random sequence to our%<br />

%samples by uniform sampling w.o. replacement%<br />

o1=randperm(p)’;<br />

o2=randperm(p)’;<br />

for i=1:p<br />

ys(i,1)=y_s(o1(i),1);<br />

epss(i,1)=eps_s(o2(i),1);<br />

x_s(i,1)=ys(i,1)*bbeta+epss(i,1);<br />

end<br />

xdesc=sort(x_s,’descend’);<br />

epsdesc=sort(epss,’descend’);<br />

ydesc=sort(ys,’descend’);<br />

xasc=sort(x_s);<br />

epsasc=sort(epss);<br />

yasc=sort(ys);<br />

170


Descriptive Statistics<br />

The kurtosis is a measure of the extent to which data is concentrated in the peak versus<br />

the tail. A positive value (leptokurtic) indicates that data is concentrated in the peak;<br />

a negative value (platykurtic) indicates that data is concentrated in the tail. For all<br />

assets and indexes we observe a kurtosis greater than zero, which is inconsistent with the<br />

assumption of Gaussian distributed returns. The negative skewness for all assets means<br />

that there is an elongated tail on the left of the probability density distribution function<br />

meaning that we find more data in the left tail compared to Gaussian distributed<br />

returns.<br />

N Mean Std Skew. Kurt.<br />

Statistics Statistics Statistics Statistics Std. error Statistik Std. error<br />

ret SP500 2506 2.43E-04 4.05E-03 -3.21E-01 4.90E-02 5.42E+00 9.80E-02<br />

ret BMY 2506 1.40E-05 1.20E-02 -1.41E+01 4.90E-02 3.61E+02 9.80E-02<br />

ret CVX 2506 2.30E-05 8.73E-03 -1.60E+01 4.90E-02 5.46E+02 9.80E-02<br />

ret HPQ 2506 1.00E-05 1.38E-02 -9.81E+00 4.90E-02 1.96E+02 9.80E-02<br />

ret KO 2506 5.10E-05 1.06E-02 -1.61E+01 4.90E-02 4.40E+02 9.80E-02<br />

ret MMM 2506 2.70E-05 8.65E-03 -1.54E+01 4.90E-02 5.18E+02 9.80E-02<br />

ret PG 2506 -2.40E-05 1.17E-02 -1.49E+01 4.90E-02 3.62E+02 9.80E-02<br />

ret SGP 2506 4.60E-05 1.38E-02 -1.28E+01 4.90E-02 2.83E+02 9.80E-02<br />

ret TXN 2506 3.00E-05 1.81E-02 -7.49E+00 4.90E-02 1.26E+02 9.80E-02<br />

ret WAG 2506 -2.70E-05 1.44E-02 -1.38E+01 4.90E-02 2.82E+02 9.80E-02<br />

Table 5.1: Descriptive statistics of historical return series of S&P 500 and nine assets included<br />

during a time interval from January 1991 to December 2000 (smaller reference sample of<br />

chapter (3))<br />

N Mean Std Skew. Kurt.<br />

Statistic Statistic Statistic Statistic Std. Error Statistic Std. Error<br />

ret SP500 5736 1.46E-04 4.61E-03 -1.94E+00 3.20E-02 4.15E+01 6.50E-02<br />

ret BMY 5736 -7.90E-05 1.08E-02 -1.27E+01 3.20E-02 3.54E+02 6.50E-02<br />

ret CVX 5736 6.30E-05 8.71E-03 -1.43E+01 3.20E-02 4.87E+02 6.50E-02<br />

ret HPQ 5736 2.10E-05 1.20E-02 -6.49E+00 3.20E-02 1.51E+02 6.50E-02<br />

ret KO 5736 -1.00E-05 1.10E-02 -2.02E+01 3.20E-02 6.87E+02 6.50E-02<br />

ret MMM 5736 0.00E+00 9.32E-03 -1.73E+01 3.20E-02 5.53E+02 6.50E-02<br />

ret PG 5736 1.60E-05 1.06E-02 -1.69E+01 3.20E-02 4.61E+02 6.50E-02<br />

ret SGP 5736 -8.80E-05 1.30E-02 -1.16E+01 3.20E-02 2.66E+02 6.50E-02<br />

ret TXN 5736 -8.00E-05 1.66E-02 -8.69E+00 3.20E-02 2.09E+02 6.50E-02<br />

ret WAG 5736 2.40E-05 1.12E-02 -1.29E+01 3.20E-02 3.38E+02 6.50E-02<br />

Table 5.2: Descriptive statistics of historical return series of S&P 500 and nine assets included<br />

during a time interval from July 1985 to April 2008 (bigger reference sample of chapter (3))<br />

171


N Mean Std Skew. Kurt.<br />

Statistic Statistic Statistic Statistic Std. Error Statistic Std. Error<br />

ret SP500 2000 -2.85E-05 4.82E-03 5.34E-02 5.47E-02 2.48E+00 1.09E-01<br />

ret BMY 2000 -2.38E-04 9.03E-03 -2.08E+00 5.47E-02 2.44E+01 1.09E-01<br />

ret CVX 2000 -1.09E-05 9.13E-03 -1.78E+01 5.47E-02 5.78E+02 1.09E-01<br />

ret HPQ 2000 -2.51E-04 1.32E-02 -4.37E+00 5.47E-02 9.46E+01 1.09E-01<br />

ret KO 2000 5.27E-05 5.74E-03 -1.83E-01 5.47E-02 2.78E+00 1.09E-01<br />

ret MMM 2000 -4.79E-05 9.33E-03 -1.86E+01 5.47E-02 6.20E+02 1.09E-01<br />

ret PG 2000 5.12E-06 8.73E-03 -2.07E+01 5.47E-02 7.15E+02 1.09E-01<br />

ret SGP 2000 -2.36E-04 1.03E-02 -2.10E+00 5.47E-02 2.40E+01 1.09E-01<br />

ret TXN 2000 -3.56E-04 1.62E-02 -3.17E+00 5.47E-02 6.61E+01 1.09E-01<br />

ret WAG 2000 6.89E-05 7.69E-03 -4.58E-01 5.47E-02 6.81E+00 1.09E-01<br />

ret MCD 2000 1.24E-04 7.43E-03 -1.92E-01 5.47E-02 5.15E+00 1.09E-01<br />

ret C 2000 -1.06E-04 8.58E-03 -2.01E-01 5.47E-02 6.97E+00 1.09E-01<br />

ret OXY 2000 4.61E-04 7.88E-03 -1.21E-01 5.47E-02 8.12E-01 1.09E-01<br />

ret DE 2000 -2.06E-04 1.15E-02 -1.18E-01 5.47E-02 6.47E+00 1.09E-01<br />

ret PNC 2000 1.42E-04 7.33E-03 -4.90E-01 5.47E-02 7.33E+00 1.09E-01<br />

ret NBL 2000 3.44E-04 9.01E-03 -2.36E-01 5.47E-02 1.77E+00 1.09E-01<br />

ret PH 2000 2.19E-04 8.36E-03 -7.08E-02 5.47E-02 3.94E+00 1.09E-01<br />

Table 5.3: Descriptive statistics of historical return series of S&P 500 and nine assets included<br />

during a time interval from 12.04.2000 to 31.03.2008 (sample consisting of N = 2000<br />

observations used in chapter (4))<br />

N Mean Std Skew. Kurt.<br />

Statistic Statistic Statistic Statistic Std. Error Statistic Std. Error<br />

ret DOW 2000 6.02E-06 4.63E-03 -3.83E-02 5.47E-02 3.55E+00 1.09E-01<br />

ret AA 2000 3.09E-05 9.80E-03 -2.96E-03 5.47E-02 2.09E+00 1.09E-01<br />

ret BA 2000 8.33E-05 8.36E-03 -7.80E-01 5.47E-02 7.34E+00 1.09E-01<br />

ret CAT 2000 3.18E-04 8.13E-03 -2.46E-01 5.47E-02 4.49E+00 1.09E-01<br />

ret GM 2000 -3.06E-04 1.11E-02 5.35E-02 5.47E-02 4.90E+00 1.09E-01<br />

ret HON 2000 1.17E-04 9.50E-03 2.18E-02 5.47E-02 1.59E+01 1.09E-01<br />

ret JNJ 2000 1.10E-04 5.31E-03 -1.20E+00 5.47E-02 2.18E+01 1.09E-01<br />

ret MCD 2000 1.65E-04 7.17E-03 -2.89E-01 5.47E-02 5.39E+00 1.09E-01<br />

ret WMT 2000 4.62E-05 6.78E-03 2.57E-01 5.47E-02 2.78E+00 1.09E-01<br />

ret XOM 2000 1.84E-04 6.38E-03 -2.69E-01 5.47E-02 2.53E+00 1.09E-01<br />

ret HD 2000 -1.54E-04 9.73E-03 -1.77E+00 5.47E-02 2.92E+01 1.09E-01<br />

ret INTC 2000 -2.25E-04 1.22E-02 -6.62E-01 5.47E-02 8.39E+00 1.09E-01<br />

ret AIG 2000 -2.52E-04 8.47E-03 -2.79E-01 5.47E-02 6.08E+00 1.09E-01<br />

Table 5.4: Descriptive statistics of historical return series of Dow Jones Industrial Average<br />

and 12 assets included during a time interval from 16.08.2000 to 31.07.2008 (sample consisting<br />

of N = 2000 observations used in chapter (4))<br />

172


N Mean Std Skew. Kurt.<br />

Statistic Statistic Statistic Statistic Std. Error Statistic Std. Error<br />

ret NASDAQ 2000 -1.10E-04 7.37E-03 3.81E-01 5.50E-02 4.73E+00 1.09E-01<br />

ret AAPL 2000 4.17E-04 1.44E-02 -5.43E+00 5.50E-02 1.19E+02 1.09E-01<br />

ret INTC 2000 -2.25E-04 1.22E-02 -6.62E-01 5.50E-02 8.39E+00 1.09E-01<br />

ret MSFT 2000 -3.50E-05 8.70E-03 1.74E-01 5.50E-02 7.79E+00 1.09E-01<br />

ret CSCO 2000 -2.29E-04 1.29E-02 2.82E-01 5.50E-02 6.67E+00 1.09E-01<br />

ret CEPH 2000 1.02E-04 1.15E-02 -3.63E-01 5.50E-02 4.50E+00 1.09E-01<br />

ret PERY 2000 2.76E-04 1.40E-02 -2.20E-01 5.50E-02 1.53E+01 1.09E-01<br />

ret HSIC 2000 3.87E-04 8.38E-03 2.07E-01 5.50E-02 4.76E+00 1.09E-01<br />

ret QQQQ 2000 -1.51E-04 8.66E-03 1.92E-01 5.50E-02 5.05E+00 1.09E-01<br />

ret ONNN 2000 -1.64E-04 2.16E-02 2.02E-01 5.50E-02 6.32E+00 1.09E-01<br />

Table 5.5: Descriptive statistics of historical return series of NASDAQ Composite and 9 assets<br />

included during a time interval from 15.08.2000 to 31.07.2008 (sample consisting of N = 2000<br />

observations used in chapter (4))<br />

N Mean Std Skew. Kurt.<br />

Statistic Statistic Statistic Statistic Std. Error Statistic Std. Error<br />

ret AORD 1398 1.69E-04 3.77E-03 -5.59E-01 6.54E-02 7.26E+00 1.31E-01<br />

ret CBA 1398 -4.00E-04 2.16E-02 -3.39E+01 6.54E-02 1.23E+03 1.31E-01<br />

ret BBG 1398 1.62E-04 7.50E-03 5.27E-02 6.54E-02 4.72E+00 1.31E-01<br />

ret AAC 1398 3.68E-04 9.67E-03 2.74E-01 6.54E-02 4.39E+00 1.31E-01<br />

ret IIF 1398 -4.79E-05 6.18E-03 -2.72E-01 6.54E-02 8.10E+00 1.31E-01<br />

ret MOF 1398 -8.33E-05 6.99E-03 -9.28E-01 6.54E-02 1.09E+01 1.31E-01<br />

ret NCM 1398 -8.33E-05 6.99E-03 -9.28E-01 6.54E-02 1.09E+01 1.31E-01<br />

ret ABP 1398 9.02E-05 6.13E-03 -3.07E-02 6.54E-02 5.47E+00 1.31E-01<br />

ret AGK 1398 8.17E-05 7.05E-03 -8.13E+00 6.54E-02 1.69E+02 1.31E-01<br />

ret BLD 1398 7.02E-05 6.97E-03 1.36E-01 6.54E-02 1.79E+00 1.31E-01<br />

ret NDO 1398 1.10E-03 5.53E-02 2.16E-01 6.54E-02 2.16E+01 1.31E-01<br />

ret CEY 1398 4.57E-04 1.07E-02 -1.14E+00 6.54E-02 1.34E+01 1.31E-01<br />

ret CSR 1398 -3.21E-04 1.84E-02 -2.58E+01 6.54E-02 8.49E+02 1.31E-01<br />

ret FCL 1398 -2.35E-05 9.85E-03 -1.69E+00 6.54E-02 3.58E+01 1.31E-01<br />

ret AXA 1398 2.43E-04 7.77E-03 7.45E-01 6.54E-02 8.15E+00 1.31E-01<br />

Table 5.6: Descriptive statistics of historical return series of AORD (ASX) and 14 assets<br />

included during a time interval from February 2003 to end of August 2008 (sample consisting<br />

of N = 1398 observations used in chapter (4))<br />

173


N Mean Std Skew. Kurt.<br />

Statistic Statistic Statistic Statistic Std. Error Statistic Std. Error<br />

A 2000 -1.14E-04 2.30E-03 5.24E-01 5.50E-02 3.43E+00 1.09E-01<br />

Eur 2000 -1.05E-04 1.93E-03 1.00E-03 5.50E-02 2.63E+00 1.09E-01<br />

GBP 2000 -5.30E-05 1.74E-03 5.30E-02 5.50E-02 2.14E+00 1.09E-01<br />

BrR 2000 -1.80E-04 3.88E-03 -1.15E+00 5.50E-02 2.36E+01 1.09E-01<br />

CHFR 2000 -8.80E-05 2.18E-03 -1.24E-01 5.50E-02 2.58E+00 1.09E-01<br />

CHY 2000 -3.60E-05 3.05E-04 -1.25E+01 5.50E-02 3.39E+02 1.09E-01<br />

indoRu 2000 0.00E+00 2.41E-03 1.03E+00 5.50E-02 3.20E+01 1.09E-01<br />

indRu 2000 -4.10E-05 9.75E-04 -8.20E-02 5.50E-02 7.63E+00 1.09E-01<br />

JPY 2000 -4.40E-05 1.93E-03 -1.97E-01 5.50E-02 2.64E+00 1.09E-01<br />

Table 5.7: Descriptive statistics of historical currency exchange rate series during a time<br />

interval from 19.10.2002 to 10.04.2008 (sample consisting of N = 2000 observations used in<br />

chapter (4))<br />

N Mean Std Skew. Kurt.<br />

Statistic Statistic Statistic Statistic Std. Error Statistic Std. Error<br />

ret synth1 2507 6.86E-05 1.45E-02 -1.59E+00 4.89E-02 3.45E+01 9.77E-02<br />

ret synth2 2507 2.93E-04 1.83E-02 -5.45E-01 4.89E-02 1.20E+01 9.77E-02<br />

ret synth3 2507 2.30E-04 1.30E-02 1.70E-01 4.89E-02 1.12E+01 9.77E-02<br />

ret synth4 2507 -7.74E-05 1.28E-02 1.17E-01 4.89E-02 5.03E+00 9.77E-02<br />

ret synth5 2507 6.61E-05 1.72E-02 -3.41E-02 4.89E-02 3.26E+00 9.77E-02<br />

ret synth6 2507 1.35E-04 1.30E-02 1.74E-01 4.89E-02 7.80E+00 9.77E-02<br />

Table 5.8: Descriptive statistics of 6 exemplary synthetic samples (sample consisting of N =<br />

2507 observations used in chapter (4))<br />

174

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!