Chapter 9
Statistics of Extremes: Challenges and Opportunities

M. de Carvalho*

Faculty of Mathematics, Pontificia Universidad Católica de Chile, Santiago, Chile

9.1 Introduction

My experience on discussing concepts of risk and statistics of extremes with practitioners started in 2009 while I was a visiting researcher at the Portuguese Central Bank (Banco de Portugal). At the beginning, colleague practitioners were intrigued about the methods I was applying; the questions were recurrent: “What is the difference between statistics of extremes and survival analysis (or duration analysis)?1 And why don't you apply empirical estimators?” The short answer is that when modeling rare catastrophic events, we need to extrapolate beyond observed data—into the tails of a distribution—and standard inference methods often fail to deal with this properly. To see this, suppose that we observe a random sample of losses c09-math-0001 and that we estimate the survivor function c09-math-0002, using the empirical survivor function, c09-math-0003, for c09-math-0004. Now, suppose that we want to assess what is the probability of observing a loss just c09-math-0005 larger than the maximum observed loss, c09-math-0006. Obviously, the probability of that event turns out to be zero [c09-math-0007, for all c09-math-0008], thus illustrating that the empirical survivor function fails to be able to extrapolate into the right tail of the loss distribution. As put simply by Taleb (2012, p. 46), “the fool believes that the tallest mountain in the world will be equal to the tallest one he has observed.”

In this chapter, I resume some viewpoints that I shared with the discussion group “Future of Statistics of Extremes” at the ESSEC Conference on Extreme Events in Finance, which took place in Royaumont Abbey, France, on 15–17 December 2014

extreme-events-in-finance.essec.edu

and which originated the invitation by the editor for writing this chapter. My goal is on providing a personal view on some recent concepts and methods of statistics of extremes and to discuss challenges and opportunities that could lead to potential future developments. The scope is far from encyclopedic, and many other interesting perspectives are found all over this monograph.

In Section 9.2, I note that a bivariate extreme value distribution is an example of what I call here a measure-dependent measure, I briefly review kernel density estimators for the spectral density, and I discuss families of spectral measures. In Section 9.3, I argue that the spectral density ratio model (de Carvalho and Davison, 2014), the proportional tails model (Einmahl et al., 2015), and the exponential families for heavy-tailed data (Fithian and Wager, 2015) share similar construction principles; in addition, I discuss en passant a new nonparametric estimator for the so-called scedasis function, which is one of the main estimation targets on the proportional tails model. Comments on potential future developments are scattered across the chapter and a miscellanea of topics are included in Section 9.4.

Throughout this chapter I use the acronym EVD to denote extreme value distribution.

9.2 Statistics of Bivariate Extremes

9.2.1 The Bivariate EVD is a Measure-dependent Measure

Let c09-math-0009 be a probability measure on c09-math-0010, and let c09-math-0011 be a parameter space. The family c09-math-0012 is a statistical model. Obviously not every statistical model is appropriate for modeling risk. As mentioned in Section 9.1, candidate statistical models should possess the ability to extrapolate into the tails of a distribution, beyond existing data.

See Coles (2001, Theorem 3.1.1). Here, c09-math-0025 and c09-math-0026 are location and scale parameters, while c09-math-0027 is a shape parameter that determines the rate decay of the tail: c09-math-0028, light tail (Gumbel); c09-math-0029, heavy tail (Fréchet); c09-math-0030, short tail (Weibull). The generalized EVD (c09-math-0031 in (9.1)) is a three-parameter family that plays an important role in statistics of univariate extremes.

In some cases we want to assess the risk of observing simultaneously large values of two random variables (say, two simultaneous large losses in a portfolio), and the mathematical basis for such modeling is that of statistics of bivariate extremes. In this context, “extremal dependence” is often interpreted as a synonym of risk. Moving from one dimension to two dimensions increases sharply the complexity of models for the extremes. The first challenge one faces when modeling bivariate extremes is that the estimation object of interest is infinite dimensional, whereas in the univariate case only three parameters c09-math-0032 are needed. The intuition is the following. When modeling bivariate extremes, apart from the marginal distributions, we are also interested in the extremal dependence structure of the data, and—as we shall see in Theorem 9.4—only an infinite-dimensional object is flexible enough to capture the “spectrum” of all possible types of dependence.

Let c09-math-0033 c09-math-0034, where I assume that c09-math-0035 and c09-math-0036 are unit Fréchet c09-math-0037 marginally distributed, that is, c09-math-0038, for c09-math-0039. Similarly to the univariate case, the classical theory for characterizing the extremal behavior of bivariate extremes is based on block maxima, here given by the componentwise maxima c09-math-0040; note that the componentwise maxima c09-math-0041 needs not to be a sample point. Similarly to the univariate case, we focus on the standardized maxima, which for Fréchet marginals is given by the standardized componentwise maxima, that is, c09-math-0042. Next, I define a special type of statistical model that plays a key role on bivariate extreme value modeling.

What are the relevant statistical models for statistics of bivariate extremes? Is there an extension of the generalized EVD for the bivariate setting? The following is a bivariate analogue to Theorem 9.1.

See Coles (2001, Theorem 8.1). Throughout I refer to c09-math-0062 as a bivariate EVD. Note the similarities between (9.1) and (9.3): both start with an “exp,” but for bivariate EVD c09-math-0063, whereas for univariate EVD c09-math-0064. To understand why c09-math-0065 needs to be an element of c09-math-0066, let c09-math-0067 or c09-math-0068 in (9.3). Some further comments are in order. First, since (9.2) is the only constraint on c09-math-0069, neither c09-math-0070 nor c09-math-0071 can have a finite parameterization. Second, a bivariate extreme value distribution c09-math-0072 is an example of a measure-dependent measure, as introduced in Definition 9.2.

A pseudo-polar transformation is useful for understanding the role of c09-math-0073, which is the so-called spectral measure. Define c09-math-0074, and denote c09-math-0075 and c09-math-0076 as the radius and pseudo-angle, respectively. If c09-math-0077 is relatively large, then c09-math-0078; if c09-math-0079 is relatively large, then c09-math-0080. de Haan and Resnick (1977) have shown that c09-math-0081, as c09-math-0082. Thus, when the radius c09-math-0083 is large, the pseudo-angles c09-math-0084 are approximately distributed according to c09-math-0085. Perfect (extremal) dependence corresponds to c09-math-0086 being degenerate at c09-math-0087, whereas independence corresponds to c09-math-0088 being a binomial distribution function, with half of the mass in 0 and the other half in 1. The spectral probability measure c09-math-0089 determines the interactions between joint extremes and is thus an estimating target of interest; other functionals of the spectral measure are also often used, such as the spectral density c09-math-0090 or Pickands (1981) dependence function c09-math-0091, for c09-math-0092. The cases of extremal independence and extremal dependence, respectively, correspond to the bivariate EVDs, c09-math-0093 and c09-math-0094, for c09-math-0095.

9.2.2 Nonparametric Spectral Density Estimation

In practice, we have to deal with a statistical problem—lack of knowledge on c09-math-0096—and an inference challenge—that is, obtaining estimates that obey the marginal moment constraints and define a density on the unit interval. Indeed, as posed by Coles (2001, p. 146) “it is not straightforward to constrain nonparametric estimators to satisfy functional constraints of the type” of Eq. (9.2). Inference should be conducted by using c09-math-0097 pseudo-angles c09-math-0098, which are constructed from a sample of size c09-math-0099, thresholding the pseudo-radius at a sufficiently high threshold c09-math-0100. Kernel smoothing estimators for c09-math-0101 have been recently proposed by 2013 and are based on

9.4 equation

Here c09-math-0103 denotes the beta density with shape parameters c09-math-0104, and c09-math-0105 is a parameter responsible for the level of smoothing, which can be obtained through cross validation. Each beta density is centered around a pseudo-angle in the sense that c09-math-0106, for c09-math-0107. And how can we obtain the probability masses, c09-math-0108? There are at least two options. A simple one is to consider Euclidean likelihood methods (Owen, 2001, pp. 63–66), in which case the vector of probability masses c09-math-0109 solves

9.5 equation

By the method of Lagrange multipliers, we obtain c09-math-0111, where c09-math-0112, and c09-math-0113. This yields the following estimator, known as the smooth Euclidean likelihood spectral density:

Another option proposed by de Carvalho et al. (2013) is to consider a similar approach to that of Einmahl and Segers (2009), in which case the vector of probability masses c09-math-0115 solves the following empirical likelihood (Owen, 2001) problem:

Again by the method of Lagrange multipliers, the solution is c09-math-0117, for c09-math-0118, where c09-math-0119 is the Lagrange multiplier associated with the second equality constraint in (9.7), defined implicitly as the solution to the equation

equation

This yields the following estimator, known as the smooth empirical likelihood spectral density:

One can readily construct smooth estimators for the corresponding spectral measures; the smooth Euclidean spectral measure and smooth empirical likelihood spectral measure are, respectively, given by

equation

where c09-math-0123 is the regularized incomplete beta function, with c09-math-0124. By construction both estimators, (9.6) and (9.8), obey the moment constraint, so that, for example,

9.9 equation

Put differently, realizations of the random probability measures c09-math-0126 and c09-math-0127 are elements of c09-math-0128. Examples of applications of these estimators in finance can be found in Kiriliouk et al. (2015, Figure 4). At the moment, the large sample properties of these estimators remain unknown.

Other estimators for the spectral measure (obeying (9.2)) can be found in Boldi and Davison (2007), Guillotte et al. (2011), and Sabourin and Naveau (2014).

9.2.3 Predictor-Dependent Spectral Measures

Formally, c09-math-0129 is a set of predictor-dependent (henceforth pd) probability measures if the c09-math-0130 are probability measures on c09-math-0131, indexed by a covariate c09-math-0132; here c09-math-0133 is the Borel sigma-algebra on c09-math-0134. Analogously, I define the following:

And why do we care about pd spectral measures? Pd spectral measures allow us to assess how extremal dependence evolves over a certain covariate c09-math-0138, that is, they allow us to model nonstationary extremal dependence structures. Pd spectral measures are a natural probabilistic concept for modeling extremal dependence structures that may change according to a covariate. Indeed, in many settings of applied interest, it seems natural to regard risk from a covariate-adjusted viewpoint, and this leads us to ideas of “conditional risk.” However, if we want to develop ideas of “conditional risk” for bivariate extremes, that is, if we want to assess systematic variation of risk according to a covariate, we need to allow for nonstationary extremal dependence structures.

To describe how extremal dependence may change over a predictor, I now introduce the concept of spectral surface.

A simple spectral surface can be constructed with the pd spectral density c09-math-0143, where c09-math-0144. In Figure 9.1, I represent a spectral surface based on this model, with c09-math-0145, for c09-math-0146. (Larger values of the predictor c09-math-0147 lead to larger levels of extremal dependence.) Other spectral surfaces can be readily constructed from parametric models for the spectral density; see, for instance, Coles (2001, Section 8.2.1).

c09f001

Figure 9.1 (a) Example of a spectral density. (b) Spectral surface from a predictor-dependent beta family, with c09-math-0148, for c09-math-0149.

Let's now regard the subject of pd bivariate extremes from another viewpoint. Modeling nonstationarity in marginal distributions has been the focus of much recent literature in applied extreme value modeling; see for instance Coles (2001, Chapter 6). The simplest approach in this setting was popularized long ago by Davison and Smith (1990), and it is based on indexing the location and scale parameters of the generalized EVD by a predictor, say, by taking

And how to model “nonstationary bivariate extremes” if one must? Surprisingly, by comparison to the marginal case, approaches to modeling nonstationarity in the extremal dependence structure have received relatively little attention. These should be important to assess the dynamics governing extremal dependence of variables of interest. For example, has extremal dependence between returns of CAC 40 and DAX 30 been constant over time, or has this level been changing over the years?

By using pd spectral measures, we are essentially indexing the parameter of the bivariate extreme value distribution (c09-math-0151) with a covariate, and thus the approach can be regarded as an analogue of the Davison–Smith paradigm in (9.10), but for the bivariate setting. In the same way that (9.10) is a covariate-adjusted version of the generalized EVD (9.1), the following concept can be regarded as a pd version of the bivariate EVD in (9.3).

Similarly to Section 2.2, n practice we need to obtain estimates that obey the marginal moment constraint and define a density on the unit interval, for all c09-math-0155. It is not straightforward to construct nonparametric estimators able to yield valid pd spectral measures. Indeed, any such estimator, c09-math-0156, needs to obey the moment constraint, that is, c09-math-0157, for all c09-math-0158. Castro and de Carvalho (2016) and Castro et al. (2015) are currently developing models for these contexts, but there are still plenty of opportunities here.3

Needless to say that other pd objects of interest can be readily constructed. For example, a pd version of Pickands (1981) dependence function can be defined as c09-math-0160, and a pd c09-math-0161 can also be constructed. Using the fact that c09-math-0162 (de Carvalho and Ramos, 2012, p. 91), the pd c09-math-0163 can be defined as c09-math-0164, for c09-math-0165.

9.2.4 Other Families of Spectral Measures

Beyond pd spectral measures other families of spectral measures are of interest. In a recent paper, de Carvalho and Davison (2014) proposed a model for a family of spectral measures c09-math-0166. The applied motivation for the concept was to track the effect of explanatory variables on joint extremes. Put differently, their main concern was on the joint modeling of extremal events when data are gathered from several populations, to each of which corresponds a vector of covariates. Thus, conceptually, there are already in de Carvalho and Davison (2014) some of the ingredients of pd spectral measures and related modeling objectives. Each element in the family should be regarded as a “distorted version,” of a baseline spectral measure c09-math-0167, in a sense that I will precise in the succeeding text. Formally, spectral density ratio families are defined as follows.

From (9.11), we can write all the normalization and moment constraints for this family as a function of the baseline spectral measure and the tilting parameters, that is,

Inference is based on the combined sample c09-math-0185 from the spectral distributions c09-math-0186. Details on estimation and inference through empirical likelihood methods can be found in de Carvalho and Davison (2011), de Carvalho and Davison (2014). An extremely appealing feature of their model is that it allows for borrowing strength across samples, in the sense that the estimate of c09-math-0187 is based on c09-math-0188 pseudo-angles, instead of simply c09-math-0189. Although flexible, their approach requires however a substantial computational investment; in particular, inference entails intensive constrained optimization problems—even for a moderate c09-math-0190—so that estimates of c09-math-0191 obey empirical versions of the normalization and moment constraints in (9.12). Their approach allows for modeling extremal dependence in settings such as Figure 9.2a, but it excludes data configurations such as Figure 9.2b. The pd-based approach of Castro et al. (2015) allows for inference to be conducted in both settings in Figure 9.2.

c09f002

Figure 9.2 Scatter plots presenting two configurations of data (predictor, pseudo-angles): one (a) where there are sample pseudo-angles per each observed covariate and another (b) where to each observed covariate may correspond a single pseudo-angle.

9.3 Models Based on Families of Tilted Measures

The main goal of this section is on describing the link between the specifications underlying the spectral density ratio model, discussed in Section 2.4, the proportional tails model (Einmahl et al., 2015), and the exponential families for heavy-tailed data (Fithian and Wager, 2015).

9.3.1 Proportional Tails Model

The proportional tails model is essentially an approach for modeling nonstationary extremes. Suppose that at time points c09-math-0192 we gather independent observations c09-math-0193, respectively, sampled from the continuous distribution functions c09-math-0194, all with a common right end point c09-math-0195. Suppose further that there exists a (time-invariant) baseline distribution function c09-math-0196, also with right end point c09-math-0197, and a continuous function c09-math-0198, such that

Here c09-math-0200 is the so-called scedasis density, and following Einmahl et al. (2015) I assume the following normalization constraint c09-math-0201. Equation (9.13) is the key specification of the proportional tails model. Roughly speaking, the scedasis density tells us how much more/less mass there is on the tail c09-math-0202, relatively to the baseline tail, c09-math-0203, for a large c09-math-0204; uniform scedasis corresponds to a constant frequency of extremes over time.

The question arises naturally: “If the scedasis density provides an indication of the ‘relative frequency’ of extremes over time, would it seem natural that such function could be somehow connected to the intensity measure of the point process characterization of univariate extremes (Coles 2001, Section 7.3)?” To have an idea on how the concepts relate, I sketch here a heuristic argument. I insist, the argument is heuristic, and my aim here does not go beyond shedding some light on how these ideas connect. Consider the following artificial setting. Suppose that we could gather a large sample from c09-math-0205, say, c09-math-0206, and that at each time point we could also collect a large sample from c09-math-0207, say, c09-math-0208, for c09-math-0209. For concreteness let's focus on c09-math-0210. Then, the definition of scedasis in (9.13) and similar arguments as in Coles (2001, Section 4.2.2) suggest that for a sufficiently large c09-math-0211,

where c09-math-0213, for c09-math-0214, is the intensity measure of the limiting Poisson process for univariate extremes (cf Coles 2001, Theorem 7.1.1). Thus, it can be seen from (9.14) that in this artificial setting the scedasis density can be literally interpreted as a measure of the relative intensity of the extremes at period c09-math-0215, with respect to a (time-invariant) baseline.

Another important question is: “How can we estimate the scedasis density?” Einmahl et al. (2015) propose a kernel-based estimator

where c09-math-0217, with c09-math-0218 being a bandwidth and c09-math-0219 being a kernel; in addition, c09-math-0220 are the order statistics of c09-math-0221. Specifically, Einmahl et al. (2015) recommend c09-math-0222 to be a symmetric kernel on c09-math-0223. A conceptual problem with using a kernel on c09-math-0224 is that it allows for the scedasis density to put mass outside c09-math-0225.4 Using similar ideas to the ones involved in the construction of the smooth spectral density estimators in Section 2.2, I propose here the following estimator:

Indeed, each beta density is centered close to c09-math-0227 in the sense that c09-math-0228, for c09-math-0229, where c09-math-0230 is the parameter controlling the level of smoothing. My goal here will not be on trying to recommend an estimator over the other, but rather on providing a brief description of strengths and limitations with both approaches. In Figure 9.3, I illustrate how the two estimators, (9.15) and (9.16), perform on the same data used by Einmahl et al. (2015) and on simulated data (single-run experiment).5 The data consist of daily negative returns of the Standard and Poor's index from 1988 to 2007 (c09-math-0231), and I use the same value for c09-math-0232 c09-math-0233 and the same bandwidth (c09-math-0234) and (biweight) kernel [c09-math-0235, for c09-math-0236] as authors; I also follow the author's settings for the simulated data. Finally, I consider c09-math-0237 for illustration.

c09f003

Figure 9.3 Scedasis density estimates. The solid line represents the beta kernel estimate from (9.16), whereas the dashed line represents the estimate from (9.15). (a) Daily Standard and Poor's index from 1988 to 2007; the gray rectangles correspond to contraction periods in the US economy. (b) Simulated data illustration from c09-math-0238, for c09-math-0239, with c09-math-0240 and c09-math-0241; the grey line represents the true scedasis c09-math-0242, for c09-math-0243, and c09-math-0244, for c09-math-0245.

In the Standard and Poor's example in Figure 9.3a, it can be seen that both estimators capture similar dynamics; the gray rectangles represent contraction periods of the US economy as dated by the National Bureau of Economic Research (NBER). It is interesting to observe that the local maxima of the scedasis density are relatively close to economic contraction periods. Indeed, “turning points” (local maxima and minima) of the scedasis density seem like an interesting estimation target for many settings of applied interest.

The estimator in (9.16) has the appealing feature of putting all mass of the scedasis density inside the c09-math-0246 interval, and some further numerical experiments suggest that it tends to have a similar behavior to that in (9.15) except at the boundary. However, a shortcoming with the method in (9.16) is that it may not be defined at the vertex 0 or 1, and hence it could be inappropriate for forecasting purposes.

The proportional tails model is extremely appealing and simple to fit. A possible shortcoming is that it does not allow for c09-math-0247 to change over time. For applications in which we suspect that c09-math-0248 may change over time, the generalized additive approach by Chavez-Demoulin and Davison (2005) is a sensible alternative, and although the model is more challenging to implement, it can be readily fitted with the R package QRM by typing in the command game.

A problem that seems relevant for practice is that of cluster analysis for the proportional tails model. To see this, suppose that one estimates the scedasis density and tail index for several stocks. It seems natural to wonder: “How can we cluster stocks whose scedasis looks more alike, or—perhaps more interestingly—how can we cluster stocks with a similar scedasis and tail index?”

Lastly, I would like to comment that it seems conceivable that Bernstein polynomials could be used for scedasis density estimation. In particular, a natural question is “Would it be possible to construct a prior over the space of all integrated scedasis functions?” Random Bernstein polynomials could seem like the way to go; see Petrone (1999) and references therein.

9.3.2 Exponential Families for Heavy-Tailed Data

In this section I sketch some basic ideas on exponential families for heavy-tailed data; I will be more brief here than in Section 3.1. My goal is mainly on introducing the model specification and to move on; further details can be found in Fithian and Wager (2015).

The starting point for the Fithian–Wager approach is on modeling the conditional right tail law from a population, c09-math-0249, as an exponential family with carrier measure c09-math-0250, for a sufficiently large threshold c09-math-0251. Two random samples are assumed to be available, c09-math-0252 and c09-math-0253, with c09-math-0254; hence, the applied setting of interest is one where the size of the sample from c09-math-0255 is much larger than the one from c09-math-0256. The model specification is

where the sufficient statistic c09-math-0258 is of the form c09-math-0259 for a certain c09-math-0260; the functional form of c09-math-0261 is motivated from the case where c09-math-0262 and c09-math-0263 are generalized Pareto distributions (cf. Fithian and Wager, 2015, p. 487).

In common with the spectral density ratio model, the Fithian–Wager model is motivated by the gains from borrowing strength across samples. Fithian and Wager are not however concerned about spectral measures, but rather on estimating a (small-sample) mean of a heavy-tailed distribution, by borrowing information from a much larger sample from a related population with the same c09-math-0264. More concretely, the authors propose a semiparametric method for estimating the mean of c09-math-0265, by using the decomposition c09-math-0266, where c09-math-0267, c09-math-0268, c09-math-0269, and c09-math-0270. The Fithian–Wager estimator for the (small-sample) mean can be written as

where c09-math-0272 and c09-math-0273, for a large threshold c09-math-0274. Here c09-math-0275 can be computed through a logistic regression with an intercept and predictor c09-math-0276, as a consequence of results on imbalanced logistic regression (Owen, 2007). As it can be observed from (9.18), the main trick on the estimation of c09-math-0277 is on the exponential tilt-based estimator for the mean residual lifetime c09-math-0278.

9.3.3 Families of Tilted Measures

From previous sections it may have became obvious that the common link underlying the specification of the spectral density ratio model, the proportional tails model, and the exponential families for heavy-tailed data was the assumption that all members in a family of interest were obtained through a suitable “distortion” of a certain baseline measure. In this section I make this link more precise.

Some examples are presented in the succeeding text.

9.4 Miscellanea

9.4.1 Asymptotic (in)dependence

Here, I comment on the need for further developing models compatible with both asymptotic dependence and asymptotic independence. In two influential papers, Poon et al. (2003, 2004) put forward that asymptotic independence was observed on many pairs of stock market returns. This had important consequences in finance, mostly because inferences in a seminal paper (Longin and Solnik, 2001) had been based on the assumption of asymptotic dependence, and hence perhaps risk had been overestimated earlier. However, an important questions is: “What if pairs of financial losses can move over time from asymptotic independence to asymptotic dependence and the other way around?” Some markets are believed to be more integrated these days than in the past, so for such markets it is relevant to ask whether they could have entered an “asymptotic dependence regime.” An accurate answer to this question would require however models able to allow for smooth transitions from asymptotic independence to asymptotic dependence, and vice versa, but as already mentioned in Section 2.3 at the moment, there is a shortage of models for nonstationary extremal dependence structures. Wadsworth et al. (2016) presents an interesting approach for modeling asymptotic (in)dependence.

9.4.2 Spatial Multivariate Extremes

An important reference here is Genton et al. (2015), but there is a wealth of problems to work in this direction, so I stop my comment here.

9.4.3 Dimension Reduction for Multivariate Extremes

Is there a way to reduce dimension in such a way that the interesting features of the data—in terms of tails of multivariate distributions—are preserved?6 I think it is fair to say that, apart from some remarkable exceptions, most models for multivariate extremes have been applied only to low-dimensional settings. I remember that at a seminal workshop on high-dimensional extremes, organized by Anthony Davison, at the Ecole Polytechnique Fédérale de Lausanne (September 14–18, 2009), for most talks high dimensional actually meant “two dimensional,” and all speakers were top scientists in the field.

Principal component analysis (PCA) itself would seem inappropriate, since principal axes are constructed in a way to find the directions that account for most variation, and for our axes of interest (whatever they are …), variation does not seem to be the most reasonable objective. A naive approach could be to use PCA for compositional data (Jolliffe, 2002, Section 13.3) and apply it to the pseudo-angles themselves. Such approach could perhaps provide a simple way to disentangle dependence into components that could be of practical interest.

9.4.4 Should the Journal Extremes Include an Applications and Case Studies Section?

Theory and methods are the backbone of our field, without regular variation we wouldn't have gone far anyway. But, beyond theory, should our community be investing even more than it already is, in modeling and applications? As put simply by Box (1979), “all models are wrong, but some are useful.” However, while most of us agree that models only provide an approximation to reality, we seem to be very demanding about the way that we develop theory about such—wrong yet useful—models. Some models entail ingenious approximations to reality and yet are very successful in practice. Should we venture more on this direction in the future? Applied work can also motivate new, and useful, theories. Should we venture more on collaborating with researchers from other fields or on creating more conferences such as the ESSEC Conference on Extreme Events in Finance, where one has the opportunity to regard risk and extremes from a broader perspective, so to think out of the box? Should the journal Extremes include an Applications and Case Studies section?

9.4.5 Communicating Risk and Extremes

What has our community been supplying in terms of communication of risk and extremes? Silence, for the most part. Definitely there have been some noteworthy initiatives, but perhaps mostly from people outside of our field such as those of David Spiegelhalter and David Hand. My own view is that it would be excellent if, in a recent future, leading scientists in our field could be more involved in communicating risk and extremes to the general public, either by writing newspaper and magazine articles or by promoting science vulgarization. Our community is becoming more and more aware of this need, I think. I was happy to see Paul Embrechts showing recently his concern about this matter at EVA 2015 in Ann Arbor.

9.4.6 Prior Elicitation in Contexts Where a Conflict of Interest Exists

How can we accurately elicit prior information when modeling extreme events in finance, in cases where a conflict of interest may exist? Suppose that a regulator requires a bank to report an estimate. If prior information is gathered from a bank expert—and if the bank is better off by misreporting—then how can we trust in the accuracy of the inferences? In such cases, I think the only Bayesian analysis a regulator should be willing to accept would be an objective Bayes-based analysis; see Berger (2006) for a review on objective Bayes.

References

  1. Barrientos, A.F., Jara, A., Quintana, F.A. Fully nonparametric regression for bounded data using dependent Bernstein polynomials. J Am Stat Assoc 2016. In press.
  2. Berger, J. The case for objective bayesian analysis. Bayesian Anal 2006;1:385–402.
  3. Boldi, M.-O., Davison, A.C. A mixture model for multivariate extremes. Journal of the Royal Statistical Society, Series B 2007;69:217–229.
  4. Box, G.E.P. Some problems of statistics and everyday life. Journal of the American Statistical Association 1979;74:1–4.
  5. Castro, D., de Carvalho, M. Spectral density regression for bivariate extremes. Stochastic Environmental Research and Risk Assessment 2016. In press. DOI: 0.1007/s00477-016-1257-z.
  6. Castro, D., de Carvalho, M., Wadsworth, J. Time-varying extremal dependence with application to leading european stock markets. 2015. Submitted.
  7. Chautru, E. Dimension reduction in multivariate extreme value analysis. Electronic Journal of Statistics 2015;9:383–418.
  8. Chavez-Demoulin, V., Davison, A.C. Generalized additive modelling of sample extremes. Journal of the Royal Statistical Society, Ser. C, 2005;54:207–222.
  9. Chen, S.X. Beta Kernel estimators for density functions. Comput Stat Data Anal 1999;31:131–145.
  10. Coles, S. An Introduction to Statistical Modeling of Extreme Values. New York: Springer-Verlag; 2001.
  11. Davison, A.C., Smith, R.L. Models for exceedances over high thresholds (with Discussion). Journal of the Royal Statistical Society, Ser. B 1990;52:393–442.
  12. de Carvalho, M., Davison, A.C. Semiparametric estimation for c09-math-0310-sample multivariate extremes. Proceedings 58th World Statistical Congress; 2011. 2961–2969.
  13. de Carvalho, M., Davison, A.C. Spectral density ratio models for multivariate extremes. J Am Stat Assoc 2014;109:764–776.
  14. de Carvalho, M., Oumow, B., Segers, J., Warchoł, M. A Euclidean likelihood estimator for bivariate tail dependence. Communications in Statistics—Theory and Methods 2013;42:1176–1192.
  15. de Carvalho, M., Ramos, A. Bivariate extreme statistics, II. RevStat—Statistical Journal 2012;10:81–104.
  16. de Haan, L., Resnick, S.I. Limit theory for multivariate sample extremes. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete 1977;40:317–377.
  17. Einmahl, J.H.J., de Haan, L., Zhou, C. Statistics of heteroscedastic extremes. Journal of the Royal Statistical Society, Ser. B 2015;78(1):31–51. DOI: 10.1111/rssb.12099.
  18. Einmahl, J.H.J., Segers, J. Maximum empirical likelihood estimation of the spectral measure of an extreme-value distribution. The Annals of Statistics 2009;37:2953–2989.
  19. Fithian, W., Wager, S. Semiparametric exponential families for heavy-tailed data. Biometrika 2015;102:486–493.
  20. Geenens, G. Probit transformation for kernel density estimation on the unit interval. Journal of the American Statistical Association, 2014;109:346–359.
  21. Genton, M.G., Padoan, S.A., Sang, H. Multivariate max-stable spatial processes. Biometrika 2015;102:215–230.
  22. Guillotte, S., Perron, F., Segers, J. Non-parametric bayesian inference on bivariate extremes. Journal of the Royal Statistical Society, Ser. B, 2011;73:377–406.
  23. Jolliffe, I.T. Principal Component Analysis. New York: Springer-Verlag; 2002.
  24. Jones, M.C., Henderson, D.A. Kernel-type density estimation on the unit interval. Biometrika 2007;94:977–984.
  25. Kiriliouk, A., Segers, J., Warchoł, M. Nonparametric estimation of extremal dependence. In: Dey, D.K., Yan, J., editors. Extreme Value Modelling and Risk Analysis: Methods and Applications. Boca Raton (FL): Chapman and Hall/CRC; 2015.
  26. Longin, F., Solnik, B. Extreme correlation of international equity markets. The Journal of Finance 2001;56:649–676.
  27. Owen, A.B. Empirical Likelihood. Boca Raton (FL): Chapman and Hall/CRC; 2001.
  28. Owen, A.B. Infinitely imbalanced logistic regression. Journal of Machine Learning Research 2007;8:761–773.
  29. Petrone, S. Random Bernstein polynomials. Scand inavian Journal of Statistics 1999;26:373–393.
  30. Pickands, J. Multivariate extreme value distributions. Proceedings 43rd World Statistical Congress; 1981. p 859–878.
  31. Poon, S.-H., Rockinger, M., Tawn, J. Modelling extreme-value dependence in international stock markets. Statistica Sinica 2003;13:929–953.
  32. Poon, S.-H., Rockinger, M., Tawn, J. Extreme value dependence in financial markets: diagnostics, models, and financial implications. Review of Financial Studies 2004;17:581–610.
  33. Sabourin, A., Naveau, P. Bayesian dirichlet mixture model for multivariate extremes: a re-parametrization. Comput Stat Data Anal 2014;71:542–567.
  34. Taleb, N. Antifragile: Things that Gain from Disorder. New York: Random House; 2012.
  35. Wadsworth, J.L., Tawn, J.A., Davison, A.C., Elton, D.M. Modelling across extremal dependence classes. J R Stat Soc, Ser B 2016. In press.
  36. Wooldridge, J.M. Econometric Analysis of Cross Section and Panel Data. 2nd ed. Cambridge (MA): MIT Press; 2015.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset