Bivariate generalized Pareto distribution in practice: models and estimation

Similar documents
Bivariate generalized Pareto distribution

Models and estimation.

The extremal elliptical model: Theoretical properties and statistical inference

PREPRINT 2005:38. Multivariate Generalized Pareto Distributions HOLGER ROOTZÉN NADER TAJVIDI

Some conditional extremes of a Markov chain

Multivariate generalized Pareto distributions

Estimation of spatial max-stable models using threshold exceedances

Bayesian Inference for Clustered Extremes

Overview of Extreme Value Theory. Dr. Sawsan Hilal space

Sharp statistical tools Statistics for extremes

Multivariate generalized Pareto distributions

Modelling Multivariate Peaks-over-Thresholds using Generalized Pareto Distributions

A Conditional Approach to Modeling Multivariate Extremes

Statistics of Extremes

Extreme Value Analysis and Spatial Extremes

APPLICATION OF EXTREMAL THEORY TO THE PRECIPITATION SERIES IN NORTHERN MORAVIA

CONTAGION VERSUS FLIGHT TO QUALITY IN FINANCIAL MARKETS

Bayesian Point Process Modeling for Extreme Value Analysis, with an Application to Systemic Risk Assessment in Correlated Financial Markets

Modelação de valores extremos e sua importância na

On the Estimation and Application of Max-Stable Processes

Bayesian Modelling of Extreme Rainfall Data

Nonparametric Estimation of the Dependence Function for a Multivariate Extreme Value Distribution

Multivariate Pareto distributions: properties and examples

Extreme value statistics: from one dimension to many. Lecture 1: one dimension Lecture 2: many dimensions

Overview of Extreme Value Analysis (EVA)

ESTIMATING BIVARIATE TAIL

MULTIVARIATE EXTREMES AND RISK

Extreme Precipitation: An Application Modeling N-Year Return Levels at the Station Level

arxiv: v4 [stat.co] 18 Mar 2018

Estimating Bivariate Tail: a copula based approach

Financial Econometrics and Volatility Models Extreme Value Theory

On the occurrence times of componentwise maxima and bias in likelihood inference for multivariate max-stable distributions

Bayesian Model Averaging for Multivariate Extreme Values

Tail dependence in bivariate skew-normal and skew-t distributions

Semi-parametric estimation of non-stationary Pickands functions

COPULA FITTING TO AUTOCORRELATED DATA, WITH APPLICATIONS TO WIND SPEED MODELLING

Extreme Value Theory and Applications

Bayesian nonparametrics for multivariate extremes including censored data. EVT 2013, Vimeiro. Anne Sabourin. September 10, 2013

Journal of Environmental Statistics

A MODIFICATION OF HILL S TAIL INDEX ESTIMATOR

APPROXIMATING THE GENERALIZED BURR-GAMMA WITH A GENERALIZED PARETO-TYPE OF DISTRIBUTION A. VERSTER AND D.J. DE WAAL ABSTRACT

Technische Universität München Fakultät für Mathematik. Properties of extreme-value copulas

Estimation of the Angular Density in Multivariate Generalized Pareto Models

MFM Practitioner Module: Quantitiative Risk Management. John Dodson. October 14, 2015

Emma Simpson. 6 September 2013

Bayesian inference for multivariate extreme value distributions

arxiv: v3 [stat.co] 22 Jun 2017

Generalized additive modelling of hydrological sample extremes

Extreme Value Analysis of Multivariate High Frequency Wind Speed Data

Richard L. Smith Department of Statistics and Operations Research University of North Carolina Chapel Hill, NC

Chain Plot: A Tool for Exploiting Bivariate Temporal Structures

Max-stable Processes for Threshold Exceedances in Spatial Extremes

Abstract: In this short note, I comment on the research of Pisarenko et al. (2014) regarding the

Classical Extreme Value Theory - An Introduction

Financial Econometrics and Volatility Models Copulas

Regional Estimation from Spatially Dependent Data

Investigation of an Automated Approach to Threshold Selection for Generalized Pareto

HIERARCHICAL MODELS IN EXTREME VALUE THEORY

On the Estimation and Application of Max-Stable Processes

arxiv: v2 [stat.me] 25 Sep 2012

Zwiers FW and Kharin VV Changes in the extremes of the climate simulated by CCC GCM2 under CO 2 doubling. J. Climate 11:

Prediction for Max-Stable Processes via an Approximated Conditional Density

A simple graphical method to explore tail-dependence in stock-return pairs

A conditional approach for multivariate extreme values

Approximating the Conditional Density Given Large Observed Values via a Multivariate Extremes Framework, with Application to Environmental Data

A PRACTICAL WAY FOR ESTIMATING TAIL DEPENDENCE FUNCTIONS

SMOOTHED BLOCK EMPIRICAL LIKELIHOOD FOR QUANTILES OF WEAKLY DEPENDENT PROCESSES

EXTREMAL QUANTILES OF MAXIMUMS FOR STATIONARY SEQUENCES WITH PSEUDO-STATIONARY TREND WITH APPLICATIONS IN ELECTRICITY CONSUMPTION ALEXANDR V.

Analysis of Bivariate Extreme Values

Accounting for extreme-value dependence in multivariate data

Extreme Value Theory as a Theoretical Background for Power Law Behavior

Estimation de mesures de risques à partir des L p -quantiles

THE MODELLING OF HYDROLOGICAL JOINT EVENTS ON THE MORAVA RIVER USING AGGREGATION OPERATORS

Modulation of symmetric densities

Modelling extreme-value dependence in high dimensions using threshold exceedances

Extreme Value Theory An Introduction

Lecture 2 APPLICATION OF EXREME VALUE THEORY TO CLIMATE CHANGE. Rick Katz

The Behavior of Multivariate Maxima of Moving Maxima Processes

New Classes of Multivariate Survival Functions

Assessing Dependence in Extreme Values

Behaviour of multivariate tail dependence coefficients

Modelling Dependence in Space and Time with Vine Copulas

Spatial and temporal extremes of wildfire sizes in Portugal ( )

EVANESCE Implementation in S-PLUS FinMetrics Module. July 2, Insightful Corp

Bivariate Analysis of Extreme Wave and Storm Surge Events. Determining the Failure Area of Structures

Does k-th Moment Exist?

Simulation of Max Stable Processes

A Closer Look at the Hill Estimator: Edgeworth Expansions and Confidence Intervals

Multivariate Distribution Models

A Brief Introduction to Copulas

MULTIDIMENSIONAL COVARIATE EFFECTS IN SPATIAL AND JOINT EXTREMES

Bayesian Multivariate Extreme Value Thresholding for Environmental Hazards

Construction and estimation of high dimensional copulas

Probability Distributions and Estimation of Ali-Mikhail-Haq Copula

Copulas, a novel approach to model spatial and spatio-temporal dependence

Change Point Analysis of Extreme Values

EXTREMAL MODELS AND ENVIRONMENTAL APPLICATIONS. Rick Katz

EVA Tutorial #2 PEAKS OVER THRESHOLD APPROACH. Rick Katz

Statistics of Extremes

Latent process modelling of threshold exceedances in hourly rainfall series

Transcription:

Bivariate generalized Pareto distribution in practice: models and estimation November 9, 2011 Pál Rakonczai Department of Probability Theory and Statistics, Eötvös University, Hungary e-mail: paulo@cs.elte.hu András Zempléni Department of Probability Theory and Statistics, Eötvös University, Hungary e-mail: zempleni@ludens.elte.hu Abstract There is often interest in understanding how the extremely high/low values of different processes are related to each other. One possible way to tackle this problem is via an asymptotic approach, which involves fitting the multivariate generalized Pareto distribution (MGPD) to data that exceed a suitably high threshold. There are two possible definitions of the MGPD. The first, which has classically been used, consider values that jointly exceed the thresholds for all components. The second definition considers those values which exceed a threshold for at least one component. The first type of definition is widely investigated, so here we focus on the second type of definition (MGPD type II). One aim of this paper is 1

to investigate the applicability of classical parametric dependence models within MGPD type II. Since in this framework the set of applicable asymmetric dependence models is more restricted, a general transformation is proposed for creating asymmetric models from the well-known symmetric ones. We apply the proposed approach to the exceedances of bivariate wind speed data and outline methods for calculating prediction regions as well as evaluating goodness-of-fit. Key words: Asymmetric dependence models, goodness-of-fit, multivariate threshold exceedances, prediction regions, wind speed data. 1 Introduction Univariate extreme value theory provides limit results for the distribution of extremes of a single stationary process. Thus block maxima (monthly, yearly etc.) of stationary observations usually can be modeled using the generalized extreme value distribution (GEV), which has a distribution function (df.) of the form { ( G ξ,µ,σ (x) = exp 1+ξ x µ σ } ) 1 ξ, (1) where 1 + ξ x µ σ > 0. µ R is called the location parameter, σ > 0 the scale parameter and ξ R the shape parameter. Similarly, threshold exceedances (e.g. over a high quantile of the observations) can be approximated by the gen- 2

eralized Pareto distribution (GPD), which has a df. of the form H ξ,σ (x) = 1 ( 1+ξ x ) 1 ξ, (2) σ where 1 + ξ x σ > 0 and σ > 0. The above distributions are strongly linked, since the (limit) distribution of the exceedances is GPD if and only if the (limit) distribution of the maxima is GEV. Moreover the shape parameter is the same for both distributions. Multivariate extremes are often modeled by applying block maxima methods (BMM) to the componentwise maxima of the series being studied. An important drawback of this approach is that it ignores whether the extremes of the different processes occurred simultaneously or not. The annual maxima may of the processes may, for example, all have arisen on the same day, but it is also possible that they all occurred in completely different methods. The BMM approach does not distinguish between these situations. This problem can be avoided by modeling the sizes of all observations that exceed a given high threshold, rather than modeling the highest value within a particular block of time. There are several methods for analyzing multivariate threshold exceedances depending on which observations are considered as exceedances. In a classical framework only those observations which exceed the threshold in all components are involved in modeling. The methods using this definition (which we call type I) are widely investigated in the literature, see e.g. Smith (1994) for one of its first applications or the PhD thesis of Michel (2006) for a review. However the type I definition can be 3

very restrictive in some cases, as it still ignores many potentially important observations: those having at least one component below the threshold. For illustration see Fig 2 (simulated data) or the upper panels in Fig 3 (wind data). Here we consider an alternative, which was originally developed in the PhD thesis of Tajvidi (1996), and has been further developed for more general cases in Rootzén and Tajvidi (2006). The main advantage of this approach is that it includes all observations that are extreme in at least one component (later: type II). The above two works have demonstrated that the multivariate exceedances of a random vector can be approximated by a multivariate generalized Pareto distribution (MGPD). We subsequently use the abbreviation MGPD to refer to the type II MGPD distribution. Although the mathematical theory for the MGPD model has been developed, the statistical properties of this model are not yet fully understood. An initial investigation about the performance of a basic bivariate model (BGPD) in this setting can be found in Rakonczai and Tajvidi (2010). Here we consider a wider range of BGPD models, which assume various dependence structure. We investigate their applicability and consider further extensions where possible. Standard models for multivariate extremes are presented in Section 2, along with their role and properties in MGPD models. A new general method for constructing asymmetric versions of known symmetric models is also presented. In Section 3 we discuss a practical application of the proposed models to German wind speed data. Here we compare the performance of different BGPD models fitted to the real data and demonstrate the improvements due to using the model 4

construction algorithm of the previous section. Some useful conclusions and further directions are summarized in Section 4. 2 Models In this section we present the currently used models for maxima of random vectors and discuss their applicability for threshold exceedances. Analogously to the univariate case it can be shown that if the componentwise maxima has a limit distribution, it is necessarily a multivariate extreme value distribution (MEVD). Its univariate margins are GEV distributions and there are several different (but equivalent) ways of characterizing the underlying dependence structure. 2.1 Dependence measures As it is shown by Resnick (1987), the MEVD assuming unit Fréchet margins can be written as ( ) G Fréchet (t 1,...,t d ) = exp V(t 1,...,t d ), (3) with V(t 1,...,t d ) = ν(([0,t 1 ] [0,t d ]) c ) = d S d i=1 ( wi t i ) S(dw). (4) 5

Here S is a finite measure on the d-dimensional simplex S d = {w R d : w = 1} satisfying the equations w i S(dw) = 1 for i = 1,...,d, S d where V and S are called exponent measure and spectral measure respectively. The symbol stands for the maximum (analogously, a b = min(a,b)). In particular, the total mass of S is always S(S d ) = d. In this context S may have a density not only on the interior of S d but also on each of the lower-dimensional subspaces of S d. Remark 1. Note that even if G is absolutely continuous in some cases the spectral measure S can put positive mass on the vertices; for instance, when the margins ofgare independent thens({e i }) = 1 for alle i vertices of the simplex. Moreover for some absolutely continuous subclasses of the MEVD S may put positive mass to the vertices at arbitrary level of association (not only for independence), e.g. asymmetric logistic and asymmetric negative logistic models. In such cases there are further difficulties arising in threshold exceedance models, as we will see later in Section 2.2. Another characterization of the MEVD, due to Pickands (1981), uses the socalled dependence function. In the bivariate setting the dependence functiona(t) must satisfy the following three properties: A(t) is convex,(1 t) t A(t) 1 and A(0) = A(1) = 1. Finally, the exponent measure can be formulated by A(t) 6

as V(t 1,t 2 ) = ( 1 + 1 ) ( t1 ) A. t 1 t 2 t 1 +t 2 2.2 MGPD of type II Recalling the definition proposed by Rootzén and Tajvidi (2006) lety = (Y 1,...,Y d ) denote a random vector, u = (u 1,...,u d ) be a suitably high threshold vector and X = Y u = (Y 1 u 1,...,Y d u d ) be the vector of exceedances. Then the multivariate generalized Pareto distribution (MGPD) for the X exceedances can be written by a MEVDGwith non-degenerate margins as H(x 1,...,x d ) = 1 logg ( 0,...,0 ) log ( ) G x1,...,x d G ( ), (5) x 1 0,...,x d 0 where 0 < G(0,..., 0) < 1. By this definition, (5) provides a model for observations that are extreme in at least one component. Similarly to (3) we may switch to unit Fréchet margins by the transformation t i = t i (x i ) = 1 logg ξi,µ i,σ i (x i ) = (1+ξ i(x i µ i )/σ i ) 1/ξ i, (6) 7

with 1 + ξ i (x i µ i )/σ i > 0 and σ i > 0 as in (1) for i = 1,...,d. The MGPD density, if exists, is the form of h(x) = = H (x) = x 1... x d d i=1 t i ( (x i) V t 1 (0),...,t d (0) ( x 1 x d ) 1 logg(x) ) log G(0) V ( t 1 (x 1 ),...,t d (x d ) t 1 t d ). (7) Remark 2. Note that H 1 (x) = H(x, ) is not a one dimension GPD as in (2), only the conditional distribution ofx 1 X 1 > 0 is GPD. Remark 3. The interpretation of the parameters is non-standard since the parameters µ i, σ i and ξ i in (5) after performing (6) are not only related to the characteristics of the marginal distribution of thei-th component ofx, but influence the j-th (j i) margins as well, due to the constant logg(0,...,0) appearing in the distribution function. 2.3 Construction of absolutely continuous MGPD It is reasonable to require that in the MGPD model there is no positive probability mass on the boundary of the distribution, because otherwise the model would not remain absolutely continuous. A model, being not absolutely continuous, is hardly realistic and causes further complications for maximum likelihood estimation. The following lemma is a useful construction of an absolutely continuous MGPD model (as in (7)) from a known absolutely continuous MEVD. Lemma 1. Let H be a MGPD represented by an absolutely continuous MEVDG 8

with spectral measures. H is absolutely continuous if and only ifs(int(s d )) = d holds, i.e. all mass is put on the interior of the simplex. See Appendix for the proof. By Remark 1. in Section 2.1 the class of absolutely continuous MGPD type II models does not include the case of independent margins. Since MGPD models are defined by the underlying MEVD model, and since MEVD models are defined by their dependence structures, the most popular parametric families for describing the dependence structures of MEVD models may also be considered for the MGPD models. The characteristics of these models are summarized in Table 1. The list is not exhaustive, but covers a rather wide range of families. The first and second column together show which specific parametric cases include an absolutely continuous asymmetric model, based on Lemma 1. In the next section we propose a general approach to constructing asymmetric models from symmetric ones. 2.4 Construction of new asymmetric models From the third column of Table 1 we can see that there are a lack of easily computable asymmetric models, especially if all probability mass is required to be put on the interior ofs 2, ensuring the absolute continuity of the model. The bilogistic and negative bilogistic models do not have an explicit formula for their exponent measure. The C-T model only have an explicit formula for the spectral density. In order to solve this problem we propose a methodology, which allows for the 9

construction of new dependence models with extra asymmetry parameter(s) from any valid models (denoted by in Table 1). As the result of this method we may obtain more flexible asymmetric models defining a new class of absolutely continuous MGPD. Because of its mathematical simplicity we illustrate the method in the bivariate setting using the dependence function, but the same methodology can be eliminated to the higher dimensional cases as well. The algorithm is fairly simple: i. take an absolutely continuous baseline dependence model from Table 1; ii. take a strictly monotonic transformation Ψ(x) : [0,1] [0,1], such that Ψ(0) = 0, Ψ(1) = 1; iii. construct a new dependence model from the baseline model A Ψ (t) = A(Ψ(t)); iv. check the constraints: A Ψ (0) = 1, A Ψ (1) = 1 and A Ψ 0 (convexity). For the construction of a feasible transformation it is natural to assume that it has a form of Ψ(t) = t+f(t) (8) hence ( A(Ψ(t))) = A (Ψ(t))Ψ (t) = A (Ψ(t)) [1+f (t)] (9) ( A(Ψ(t))) = A (t+f(t))(1+2f (t))+a (t+f(t))f (t) 10

Obviously by choosing f such that f (0) = f (1) = 0, the A Ψ (0) = 1 and A Ψ (1) = 1 constraints are automatically fulfilled. Only the convexity needs to be checked. In general we found that the following functional form leads to a wide class of valid models f ψ1,ψ 2 (t) = ψ 1 (t(1 t)) ψ 2, fort [0,1], (10) where ψ 1 R and ψ 2 1 are asymmetry parameters. If ψ 1 = 0 we get back the baseline model. Solving the convexity inequation provides the valid range for the asymmetry parameters, see Fig 1 for illustration. Remark 4. Of course either f can be chosen differently from (10) or the baseline dependence functions having explicit dependence functions 1 can differ from logistic or negative logistic as long as the above constraints remained fulfilled. More complicated baselines as C-T model (where numerical integral is needed to obtain dependence function) or Tajvidi s generalized logistic model (having two dependence parameters instead of one) haven t been applied yet. Remark 5. Note that this transformation has particular impact beyond the BGPD models as it can be universally used also for BEVD, BGPD of type I or generally in the copula of any bivariate distributions. Remark 6. Although the transformations Ψ was defined now for the bivariate case only, it can be extended for the higher dimensional cases as well as e.g. in the 1 Therefore the bilogistic and negative bilogistic models are excluded from the list of applicable models. 11

trivariate case it can have a form of Ψ(t 1,t 2 ) = (t 1 +ψ 1,1 (t 1 ([1 t 2 ] t 1 )) ψ 2,1,t 2 +ψ 2,1 (t 2 ([1 t 1 ] t 2 )) ψ 2,2 ). The investigation of the properties and application of this model to data from 3 sites will be the subject of planned future work. We plan to come back to this issue in another paper soon. When using these transformations we put the prefix "Ψ"- before the name of the baseline dependence model, as e.g.ψ-logistic model. As a further illustration of the above method there has been a small simulation study performed for theψnegative logistic BGPD model. 200 samples (n = 2000) have been generated by different knownαandψ 1 parameters (with fixed marginal parameters andψ 2 = 2) and then maximum likelihood (re-)estimation of the parameters was performed. The results are summarized in the left side of Table 2 where the small deviations from the known parameter values and the relatively small standard errors verify the capability of the model construction. We also found that maximum likelihood methods perform well in general. However there is strong pairwise correlation between the parameters µ i,σ i, i = 1,2, which can be extinguished in practice by keeping e.g. one location parameter fixed during the optimization process. Estimation within this subclass gives stable convergence results. 12

3 Statistical inference on wind speed data We illustrate the practical application of the proposed methods using daily windspeed maxima at two locations in northern Germany (Bremerhaven and Schleswig) over the past five decades (1957-2007). Originally, average hourly windspeeds were recorded, but in order to reduce the serial correlation within the series, we calculated the maxima of the original observations for each day. The data for the above cities are rather strongly correlated, due to the relatively short distance between them (Kendall s correlation is 0.57). From Fig 3 we can see that the limit of χ(q) is significantly above zero, and it appears that χ(q) tends towards one as q approaches one, so that the wind speeds of the two sites may be considered to be asymptotically dependent. (The coefficients χ(q) and χ(q) can be interpreted as quantile dependent measures 2 of dependence, and asymptotic dependence can be measured by their limits at one.) Previous studies suggest that this level of association is strong enough to ensure reasonable performance for BGPD models (Rakonczai and Tajvidi, 2010). After choosing a suitably high quantile (95%) as a threshold level for fitting the univariate GPD (13.8m/s for Bremerhaven, 10.3m/s for Schleswig) the main aim is modeling threshold exceedances occurring at any measuring station. The bivariate data and the threshold levels are displayed in Fig 3. Though observations are considered as being stationary, non-stationarity can also be handled e.g. by choosing time-dependent model parameters as in Rakonczai et al. (2010). 2 For formulas and exact interpretation see Section 8.4 of Coles (2001). 13

3.1 Results for bivariate observations The large number of model parameters (7-9), which are difficult to interpret (location and scale parameters of the underlying BEVD of the BGPD are not location and scale parameters for the BGPD margins any more) requires some transparent illustration. Hence we used bivariate prediction regions as an alternative for visualizing the estimates. A prediction region (Hall and Tajvidi, 2000) is the smallest region containing an observation with a given (usually high) probability γ. The uncertainty in the estimation can also be visualized using these prediction regions. Fig 2 shows the contours of similar regions for the BGPD type I and type II models, illustrating that indeed in the case of BGPD type II the estimators are more stable. The prediction regions of some fitted models are displayed in Fig 4. The maximum likelihood optimization has been performed by fixing the location parameter µ x = 0, as it has been suggested in the last paragraph of Section 2.4. (It turned out that for other fixed values of µ x the algorithm has found the same maximum and resulted in indistinguishable prediction regions). In addition, in the new (Ψ) models we have chosen fixedψ 2 = 2 values for simplicity. In these models the asymmetry parameters (ψ 1 s) are significantly greater than zero (ψ 1 = 0 gives back the baseline model), providing some evidence of asymmetry in the data, whereas the dependence parameters are very close to those estimated for the baseline models. The effect of asymmetry parameters can clearly be seen on the top right panel of Fig 4, where the Ψ-negative logistic model has a relevant torsion in its regions to the left. Similar torsion occurs for the regions calculated from the Coles-Tawn dependence model. It is still difficult to make comparisons 14

based on visually very similar regions, hence a formal test would be very useful. The test we propose in this case can be easily performed as a simple byproduct of the prediction region method. The regions can be viewed as a partition of the plane. As the expected number of observations in each of these partitions is known, we can compare the theoretical frequencies with the realization e.g. byχ 2 statistics. The results for Bremerhaven and Schleswig can be found in Table 3. It seems that the C-T (χ 2 CT = 9.1) and Ψ-Neglog models (χ2 Ψ-NegLog = 10.7) perform the best. In general the new asymmetric (Ψ) models are significantly better than their symmetric baseline versions, χ 2 Log = 21.5 reduces to χ 2 Ψ-Log = 16.9 and χ 2 NegLog = 14 reduces to χ2 Ψ-NegLog = 10.7. Tajvidi s generalized symmetric logistic model(χ 2 Tajvidi s = 11.5) is the best among the symmetric ones. In order to investigate the uncertainty of the estimated models a bootstrap study has been performed. Since the original distribution of the data is unknown, 100 samples have been generated and then the BGPD distribution has been re-estimated in every case. The bootstrapped dependence parameters for theψ-neglog model are presented in the right side of Table 2. The standard errors are about the same as the ones we got by a parametric simulation study (on the left side of the same table) where known parameters have been re-estimated. The average difference between the log-likelihood values of the Neglog and Ψ-NegLog models (fitted to the bootstrap samples) is around 10, which shows a significant improvement due to the2new parameters in the likelihood ratio as well. 15

4 Discussion and further extensions We have presented several models for bivariate exceedances and illustrated the practical application of these models using a real dataset on wind speed. An R library for fitting the proposed models, mgpd, has been developed as is available on CRAN. For a visual (and empirical) model evaluation we suggested comparing the cover rates of different prediction regions. While this is a valuable tool, the development of more formal tests, like the ones known for copula modeling (Rakonczai and Zempléni, 2008) would definitely be a further improvement. Alternative methods for comparing models, assessing the statistical significance of individual terms in the model, and assessing predictive performance should also be explored. The suggested asymmetric models are promising, as with their help one becomes able to fit asymmetric families without heavy computational burden. Our results have shown that they are real competitors for the previously used models in terms of coverage probabilities. Especially, both Ψ models perform significantly better than their baseline counterparts. Probably the approach can be successfully extended into higher dimensions as well. In this paper we have focused on the new MGPD type II distributions, including the definition of new asymmetric models. Thus it was beyond the scope of this approach to evaluate the differences between the two types of MGPD. However, it turned out that the model we proposed can be estimated and effectively used for real data. 16

Acknowledgements The European Union and the European Social Fund have provided financial support to the project under the grant agreement no. TÁMOP 4.2.1/B-09/KMR-2010-0003. We wish to express particular thanks to Adam Butler (Biomathematics and Statistics Scotland) as well as to the anonymous referees of Environmetrics for their very helpful suggestions and comments. References [1] Beirlant, J., Goegebeur, Y., Segers, J. and Teugels, J. 2004. Statistics of Extremes: Theory and Applications. Wiley: West Sussex: England. [2] Coles, S. and Tawn, J., 1991. Modelling extreme multivariate events. J. R. Statist. Soc. Ser. B 53: 377-392. [3] Coles, S. and Tawn, J., 1994. Statistical Methods for Multivariate Extremes: an Application to Structural Design Appl. Statistics 43 1-48. [4] Coles, S. 2001. An Introduction to Statistical Modelling of Extreme Values. Springer-Verlag. [5] Joe, H. 1990. Families of min-stable multivariate exponential and multivariate extreme value distributions. Statist. Probab. Letters, 9: 75-81. [6] Hall, P. and Tajvidi, N. 2000. Prediction regions for bivariate extreme events. Aust. N. Z. J. Stat. 46(1): 99-112. [7] Michel, R. 2006. Simulation and Estimation in Multivariate Generalized Pareto Models PhD thesis, University of Wuerzburg, Germany [8] Pickands, J. 1981. Multivariate extreme value distributions. In Bulletin of the International Statistical Institute, Proceedings of the 43rd Session: 859-878. [9] Rakonczai, P. 2011. mgpd package (bgpd version) available at http://cran.rproject.org/web/packages/mgpd/index.html 17

[10] Rakonczai, P., Butler, A. and Zempléni, A. 2010. Modeling temporal trend within bivariate generalized Pareto models of logistic type (working paper available at http://www.math.elte.hu// paulo//pdf//) [11] Rakonczai, P. and Tajvidi, N. 2010. On prediction of bivariate extremes. International Journal of Intelligent Technologies and Applied Statistics, 3(2): 115-139. [12] Rootzén, H. and Tajvidi, N. 2006. The multivariate generalized Pareto distribution. Bernoulli 12: 917-930. [13] Smith, R. L. 1990. Extreme value theory. In Handbook of Applicable Mathematics (ed. W. Ledermann), vol. 7: 437-471. [14] Stephenson, A. G. 2002. evd: Extreme Value Distributions. R News, 2(2): 31-32. [15] Tajvidi, N. 1996. Multivariate generalized Pareto distributions. Article in PhD thesis, Department of Mathematics, Chalmers, Göteborg. [16] Tawn, J. A. 1988. Bivariate extreme value theory: Models and estimation. Biometrika, 75(3): 397-415. 18

Model Asym. Density Complications Sym. Logistic Asym. Logistic (Tawn, 1988) Ψ-logistic convexity constraints Sym. Neg. Logistic Asym. Neg. Logistic (Joe, 1990) Ψ-negative logistic convexity constraints Bilogistic (Smith, 1990) not explicit Neg. Bilogistic (C. and T., 1994) not explicit Tajvidi s (Tajvidi, 1996) C-T (Coles and Tawn, 1991) only s(w) is given Mixed (Tawn, 1988) if ψ = 1 Asym. Mixed Table 1: Summary of bivariate dependence models. The asymmetric logistic and negative logistic are not absolutely continuous. Asymmetric models with valid density are the bilogistic, negative bilogistic and Dirichlet models, but all of them have further complications in calculation. New models denoted by are proposed in Section 2.4. Simulation study for Ψ-negative logistic BGPD Bootstrap α = 1.3 α = 1.4 unknown α,ψ 1 ˆα ˆψ1 ˆα ˆψ1 ˆα ˆψ1 ψ 1 = 0.0 1.30 (0.04) 0.01 (0.05) 1.40 (0.04) 0.00 (0.05) 1.32(0.05) 0.25(0.05) ψ 1 = 0.1 1.30 (0.04) 0.11 (0.05) 1.41 (0.04) 0.12 (0.04) ψ 1 = 0.2 1.32 (0.04) 0.23 (0.05) 1.41 (0.05) 0.20 (0.04) ψ 1 = 0.3 1.31 (0.05) 0.30 (0.07) n.a n.a Table 2: Parameter estimation and standard errors for simulated samples (n=2000), number of repetitions is 200 (columns on the left side); bootstrap estimates from the wind speed data (rightmost columns). 19

GoF with χ 2 -statistics: Bremerhaven and Schleswig γ-range 0.99< 0.95-0.99 0.9-0.95 0.75-0.9 0.5-0.75 <0.5 χ 2 Expected 12.6 50.3 62.8 188.6 314.2 628.5 - Log 7 30 66 216 350 588 21.5 Ψ-Log 9 33 60 215 349 591 16.9 Neglog 11 31 66 215 334 600 14.0 Ψ-NegLog 11 34 63 210 337 602 10.7 Mix 12 46 71 247 331 550 30.3 C-T 12 34 65 209 330 607 9.1 Tajvidi s 11 35 62 212 340 597 11.5 Bilog 6 29 62 220 354 586 25.6 NegBilog 13 32 61 220 337 594 15.5 Table 3: Number of observations between the γ = 0.99, 0.95, 0.9, 0.75, 0.5- prediction regions (graphical illustration in Fig 4.) at Bremerhaven and Schleswig. According to theχ 2 test statistics (last column) the C-T andψ-neglog models are the closest to the observations. 20

δ 2 ta αlog=2(ψ(t) δ 2 ta αneglog=1.2(ψ(t) 0.0 1.0 2.0 3.0 Psi Logistic ψ 1 = 0.3 ψ 1 = 0.15 ψ 1 = 0 ψ 1 = 0.15 ψ 1 = 0.3 0.0 1.0 2.0 3.0 Psi NegLog ψ 1 = 0.3 ψ 1 = 0.15 ψ 1 = 0 ψ 1 = 0.15 ψ 1 = 0.3 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 t t Constraints of convexity for psi logistic models Constraints of convexity for psi negative logistic models ψ 1 0.4 0.2 0.0 0.2 0.4 ψ 2 = 1.8 ψ 2 = 2 ψ 2 = 2.2 ψ 1 4 2 0 2 4 ψ 2 = 1.8 ψ 2 = 2 ψ 2 = 2.2 1.4 1.6 1.8 2.0 2.2 0.9 1.0 1.1 1.2 1.3 1.4 1.5 α α Figure 1: Upper panels: second derivatives of Ψ-logistic and Ψ-negative logistic dependence functions with ψ 2 = 2 and different parameters ψ 1. Lower panels: valid parameters ψ 1 and α are between the identically colored curves for a given ψ 2. BGPD I BGPD II 5 0 5 10 15 5 0 5 10 15 5 0 5 10 15 5 0 5 10 15 Figure 2: Prediction regions for simulated data from a logistic model. Left:BGPD type I, right: BGPD type II 21

fbvpot bgpd by evd 10 15 20 density curves 20 15 Schleswig (m/s) 5 10 15 10 5 Schleswig (m/s) 20 25 Daily wind speed 25???? 10 Bremerhaven (m/s) 20 25 30 Bremerhaven (m/s) Chi Bar 0.0 0.4 0.0 0.4 0.8 Chi Bar Plot 0.8 Chi Plot Chi 15 0.0 0.2 0.4 0.6 0.8 1.0 0.0 Quantile 0.2 0.4 0.6 0.8 1.0 Quantile Figure 3: Upper left: Wind speed data, with thresholds. Upper right: Density contours for BGPD type I, fitted by the evd package, no (parametric) inference on the sides. (See Stephenson (2002) and reference manual at CRAN.) Lower panels: χ and χ -plots, respectively. 22

NegLog Psi NegLog Schleswig(m/s) 5 10 15 20 25 Regions γ= 0.99 γ= 0.95 γ= 0.9 Schleswig(m/s) 5 10 15 20 25 Regions γ= 0.99 γ= 0.95 γ= 0.9 5 10 15 20 25 30 5 10 15 20 25 30 Bremerhaven(m/s) Bremerhaven(m/s) Coles Tawn Tajvidi s Schleswig(m/s) 5 10 15 20 25 Regions γ= 0.99 γ= 0.95 γ= 0.9 Schleswig(m/s) 5 10 15 20 25 Regions γ= 0.99 γ= 0.95 γ= 0.9 5 10 15 20 25 30 5 10 15 20 25 30 Bremerhaven(m/s) Bremerhaven(m/s) Figure 4: Prediction regions for the 4 best models (having the smallest χ 2 - statistics in Table 3). 23

Appendix Proof of Lemma 1. Suppose that (X,Y) H is a BGPD with lower endpoints l x,l y for its margins. The marginal transformations mappingx andy into the unit Fréchet scale are t 1,t 2, as in (6). Let the value of the first component of (X,Y) converge tol x assuming fixed y > 0 value for the second one, then logg(x,y) logg(x 0,y 0) lim H(x,y) = lim = (11) x l x x lx G(0, 0) logg (t 1 (x),t 2 (y)) logg (t 1 (x),t 2 (0)) = lim = x lx G (t 1 (0),t 2 (0)) = G 1 (t 1 (0),t 2 (0)) lim[v(t 1 (x),t 2 (0)) V(t 1 (x),t 2 (y))] = x lx { ( ) = G 1 (t w 1(0),t 2 (0)) lim max x lx S 2 t 1 (x), 1 w S(dw) t 2 (0) ( ) } w max S 2 t 1 (x), 1 w S(dw) = t 2 (y) { = G 1 1 w (t 1 (0),t 2 (0)) S({0,1})+ lim t 2 (0) x l x S 2 \{0,1} t 1 (x) S(dw) } 1 w S({0,1}) lim t 2 (y) x l x S 2 \{0,1} t 1 (x) S(dw) = = G 1 (t 1 (0),t 2 (0)) = 0 S({0,1}) = 0 { 1 t 2 (0) 1 t 2 (y) } S({0,1}) = Consequently, independence among the margins leads to a degenerate MGPD model without having density. This case is discussed in detail in the Section 3 of Rootzén and Tajvidi (2006). 24