Chapter 7
Extreme Values Statistics for Markov Chains with Applications to Finance and Insurance

Patrice Bertail1, Stéphan Clémençon2 and Charles Tillier1

1MODAL'X, Université Paris-Ouest, Nanterre, France

2TSI, TelecomParisTech, Paris, France

AMS 2000 Mathematics Subject Classification: 60G70, 60J10, 60K20.

7.1 Introduction

Extremal events for (strongly or weakly) dependent data have received an increasing attention in the statistical literature in the last past years (see (Newell, 1964); (Loynes, 1965); (O'Brien, 1974), (O'Brien, 1987); (Hsing, 1988), (Hsing, 1991), (Hsing, 1993); (Resnick and Stărică, 1995); (Rootzén, 2009), for instance). A major issue for evaluating risks and understanding extremes and their possible replications is to take into account some dependencies. Indeed, whereas extreme values naturally occur in an isolated fashion in the identically independent distributed (i.i.d.) setup, since extreme values may be highly correlated, they generally tend to take place in small clusters for weakly dependent sequences. Most methods for statistical analysis of extremal events in weakly dependent setting rely on (fixed length) blocking techniques, which consist, roughly speaking, in dividing an observed data series into (overlapping or nonoverlapping) blocks of fixed length. Examining how extreme values occur over these data segments allows to capture the tail and the dependency structure of extreme values.

As originally pointed out in Rootzén (1988), the extremal behavior of instantaneous functionals c07-math-0002 of a Harris recurrent Markov chain c07-math-0003 may be described through the regenerative properties of the underlying chain. This chapter emphasizes the importance of renewal theory and regeneration from the perspective of statistical inference for extremal events. Indeed, as observed by Rootzén (1988) (see also (Asmussen, 1998a); (Asmussen, 1998b); (Haiman et al., 1995); (Hansen and Jensen, 2005)), certain parameters of extremal behavior features of Harris Markov chains may be also expressed in terms of regeneration cycles, namely, data segments between consecutive regeneration times c07-math-0004, that is, random times at which the chain completely forgets its past. Following in the footsteps of the seminal contribution of Rootzén (1988) (see also (Asmussen, 1998a)), Bertail et al. (2009) and Bertail et al. (2013) have recently investigated the performance of regeneration-based statistical procedures for estimating key parameters, related to the extremal behavior analysis in a Markovian setup. In the spirit of the works of Bertail and Clémençon (2006b) (refer also to (Bertail and Clémençon, 2004a); (Bertail and Clémençon, 2004b); (Bertail and Clémençon, 2006a)), they developed a statistical methodology, called the “pseudoregenerative method,” based on approximating the pseudoregeneration properties of general Harris Markov chains, for tackling various estimation problems in a Markovian setup. Most of their works deal with regular differentiable functionals like the mean (see (Bertail and Clémençon, 2004a), (Bertail and Clémençon, 2007)), the variance, quantiles, c07-math-0005-statistics and their robustified versions (Bertail et al., 2015), as well as c07-math-0006-statistics (Bertail et al., 2011). Bootstrap versions of these estimates have also been proposed. For regular functionals, they possess the same nice second-order properties as the bootstrap in the i.i.d. case, that is, the rate of the convergence of the bootstrap distribution which is close to c07-math-0007, for regular Markov chains, instead of c07-math-0008 for the asymptotic (Gaussian) benchmark (see (Bertail and Clémençon, 2006b)).

The purpose of this chapter is to review and give some extensions of this approach in the framework of extreme values for general Markov chains. The proposed methodology consists in splitting up the observed sample path into regeneration data blocks (or into data blocks drawn from a distribution approximating the regeneration cycle's distribution, in the general case when regeneration times cannot be observed). We mention that the estimation principle exposed in this chapter is by no means restricted to the sole Markovian setup, but indeed applies to any process for which a regenerative extension can be constructed and simulated from available data (see Chapter 10 in Thorisson (2000)). Then, statistical tools are built over the sequence of maxima over the resulting data segments, as if these maxima were i.i.d. In order to illustrate the interest of this technique, we focus on the question of estimating the sample maximum's tail, the extremal dependence index, and the tail index by means of the (pseudo)regenerative method. To motivate this approach in financial and insurance applications (as well as queuing or inventory models), we illustrate how these tools may be used in order to estimate ruin probabilities or extremal index, in ruin models with a dividend barrier, exhibiting some regenerative properties. Such applications have also straightforward extensions (for continuous Markov chains) in the field of finance, for instance for put option pricing (for which the “strike” plays here the role of the ruin level).

7.2 On the (pseudo) Regenerative Approach for Markovian Data

Here and throughout, c07-math-0009 denotes a c07-math-0010-irreducible aperiodic time-homogeneous Markov chain, valued in a (countable generated) measurable space c07-math-0011 with transition probability c07-math-0012 and initial distribution c07-math-0013. We recall that the Markov property means that, for any set c07-math-0014, such that c07-math-0015, for any sequence c07-math-0016 in c07-math-0017,

equation

For homogeneous Markov chains, the transition probability does not depend on c07-math-0019. Refer to Revuz (1984) and Meyn and Tweedie (1996) for basic concepts of the Markov chain theory. For sake of completeness, we specify the two following notions:

  • The chain is irreducible if there exists a c07-math-0020-finite measure c07-math-0021 such that for all set c07-math-0022, when c07-math-0023, the chain visits c07-math-0024 with a strictly positive probability, no matter what the starting point.
  • Assuming c07-math-0025-irreducibility, there are c07-math-0026 and disjointed sets c07-math-0027, c07-math-0028 (c07-math-0029) weighted by c07-math-0030 such that c07-math-0031 and c07-math-0032, c07-math-0033. The period of the chain is the greatest common divisor c07-math-0034 of such integers. It is aperiodic if c07-math-0035.
  • The chain is said to be recurrent if any set c07-math-0036 with positive measure c07-math-0037, if and only if (i.f.f.) the set c07-math-0038 is visited an infinite number of times.

The first notion formalizes the idea of a communicating structure between subsets, and the second notion considers the set of time points at which such communication may occur. Aperiodicity eliminates deterministic cycles. If the chain satisfies these three properties, it is said to be Harris recurrent.

In what follows, c07-math-0039 (respectively, c07-math-0040 for c07-math-0041 in c07-math-0042) denotes the probability measure on the underlying space such that c07-math-0043 (resp., conditioned upon c07-math-0044), c07-math-0045 the c07-math-0046-expectation (resp. c07-math-0047 the c07-math-0048-expectation), and c07-math-0049 the indicator function of any event c07-math-0050. We assume further that c07-math-0051 is positive recurrent and denote by c07-math-0052 its (unique) invariant probability distribution.

7.2.1 Markov Chains with Regeneration Times: Definitions and Examples

A Markov chain c07-math-0053 is said regenerative when it possesses an accessible atom, that is, a measurable set c07-math-0054 such that c07-math-0055 and c07-math-0056 for all c07-math-0057, c07-math-0058 in c07-math-0059. A recurrent Markov chain taking its value in a finite set is always atomic since each visited point is itself an atom. Queuing systems or ruin models visiting an infinite number of time the value 0 (the empty queue) or a given level (for instance, a barrier in the famous Cramér–Lundberg model; see Embrechts et al. (1997) and the following examples) are also naturally atomic. Refer also to Asmussen (2003) for regenerative models involved in queuing theory, and see also the examples and the following applications.

Denote then by c07-math-0060 the hitting time on c07-math-0061 or first return time to c07-math-0062. Put also c07-math-0063 for the so-called successive return times to c07-math-0064, corresponding to the time of successive visits to the set c07-math-0065.

In the following c07-math-0066 denotes the expectation conditioned on the event c07-math-0067. When the chain is Harris recurrent, for any starting distribution, the probability of returning infinitely often to the atom c07-math-0068 is equal to one. Then, for any initial distribution c07-math-0069, by the strong Markov property, the sample paths of the chain may be divided into i.i.d. blocks of random length corresponding to consecutive visits to c07-math-0070, generally called regeneration cycles:

equation

taking their values in the torus c07-math-0072. The renewal sequence c07-math-0073 defines successive times at which the chain forgets its past, termed regeneration times.

Example 1: Queuing system or storage process with an empty queue

We consider here a storage model (or a queuing system), evolving through a sequence of input times c07-math-0074 (with c07-math-0075 by convention), at which the storage is refilled. Such models appear naturally in not only many domains like hydrology and operation research but also for modeling computer CPU occupancy.

Let c07-math-0076 be the size of the input into the storage system at time c07-math-0077. Between each input time, it is assumed that withdrawals are done from the storage system at a constant rate c07-math-0078. Then, in a time period c07-math-0079, c07-math-0080, the amount of stored contents which disappears is equal to c07-math-0081. If c07-math-0082 denotes the amount of contents immediately before the input time c07-math-0083, we have, for all c07-math-0084,

equation

with c07-math-0086, c07-math-0087 by convention and c07-math-0088 for all c07-math-0089 and c07-math-0090 is sometimes called the waiting time period.

This model can be seen as a reflected random walk on c07-math-0091. Assume that, conditionally to c07-math-0092, c07-math-0093, the amounts of input c07-math-0094, c07-math-0095 are independent from each other and independent from the interarrival times c07-math-0096, c07-math-0097 and that the distribution of c07-math-0098 is given by c07-math-0099 for c07-math-0100. Under the further assumption that c07-math-0101 is an i.i.d. sequence, independent from c07-math-0102, the storage process c07-math-0103 is a Markov chain. The case with exponential input–output has been extensively studied in Asmussen (1998a).

It is known that the chain c07-math-0104 is irreducible as soon as c07-math-0105 has an infinite tail for all c07-math-0106 and if in addition c07-math-0107, c07-math-0108 is an accessible atom of the chain c07-math-0109. Moreover, if c07-math-0110 has exponential tails, then the chain is exponentially geometrically ergodic. The case with heavy tails has been studied in detail by Asmussen (1998b) and Asmussen et al. (2000). Under some technical assumptions, the chain is recurrent positive, and the times at which the storage process c07-math-0111 reaches the value 0 are regeneration times. This property allows to define regeneration blocks dividing the sample path into independent blocks, as shown in the following. Figure 7.1 represents the storage process with c07-math-0112 and c07-math-0113 with c07-math-0114 distribution and c07-math-0115 The horizontal line corresponds to the atom c07-math-0116 and the vertical lines are the corresponding renewal times (visit to the atom).

c07f001

Figure 7.1 Splitting a reflected random walk, with an atom at {0}; vertical lines corresponds to regeneration times, at which the chain forgets its past. A block is a set of observations between two lines (it may be reduced to {0} in some case).

Notice that the blocks are of random size. Some are rather long (corresponding to large excursion of the chain); others reduce to the point c07-math-0117 if the chain stays at 0 for several periods. In this example, for the given values of the parameters, the mean length of a block is close to 50.5. It is thus clear that we need a lot of observations to get enough blocks. The behavior of the maximum of this process for subexponential arrivals has been studied at length in Asmussen (1998b).

Example 2: Cramér–Lundberg with a dividend barrier

Ruin models, used in insurance, are dynamic models in continuous time which describe the behavior of the reserve of a company as a function of:

  1. Its initial reserve c07-math-0120 (which may be chosen by the insurer)
  2. The claims which happen at some random times (described by a arrival claims process)
  3. The premium rate which is the price paid by customers per unit of time

In the classical Cramér–Lundberg model (Figure 7.2), the claims arrival processc07-math-0121 is supposed to be an homogeneous Poisson process with rate c07-math-0122, modeling the number of claims in an interval c07-math-0123. The claims sizes c07-math-0124,…,c07-math-0125, which an insurance company has to face, are assumed to be strictly positive and independent, with cumulative distribution function (c.d.f.) c07-math-0126 The premium rate is supposed to be constant equal to c07-math-0127 Then, the total claims process, given by c07-math-0128 c07-math-0129, is a compound Poisson process. Starting with an initial reserve c07-math-0130, the reserve of the company evolves as

equation

One of the major problems in ruin models for insurance company is how to choose the initial amount to avoid the ruin or at least ensure that the probability of ruin over a finite horizon (or an infinite one) is small, equal to some given error of first kind, for instance, c07-math-0132. The probability of ruin for an initial reserve c07-math-0133 over an horizon c07-math-0134 is given by

equation

Notice that this model is very close to the queuing process considered in Example 1. The input times c07-math-0136 correspond here to the times of the claims. It is easy to see that under the given hypotheses, the interarrival times c07-math-0137 are i.i.d. with exponential distribution c07-math-0138 (with c07-math-0139. However, most of the time, for a given company, we only observe (at most) one ruin (since it is an absorbing state), and the reserve is not allowed to grow over a given barrier. Actually, if the process c07-math-0140 crosses a given threshold c07-math-0141, the money is redistributed in some way to the shareholders of the company. This threshold is called a dividend barrier. In this case the process of interest is rather 1

equation

where c07-math-0143 designs the infimum between c07-math-0144 and c07-math-0145. Of course, the existence of a barrier reinforces the risk of ruin especially if the claims size may be large in particular if their distributions have a fat tail. The embedded chain is defined as the value of c07-math-0146 at the claim times, say, c07-math-0147 then it is easy to see that we have

equation
c07f002

Figure 7.2 Cramér–Lundberg model with a dividend barrier at c07-math-0117 (where the chain is reflected); ruin occurs at c07-math-0118 when the chain goes below 0.

Otherwise, the probability of no ruin is clearly linked to the behavior of c07-math-0149

In comparison to Example 1, this model is simply a mirror process, with this time an atom at c07-math-0150 instead of c07-math-0151 as shown in the following two graphics. In this example, the c07-math-0152 are exponential and the claims with exponential tails; the initial reserve is 5 and the barrier at 9. In this simulation the “ruin” is attained at time c07-math-0153.

The embedded chain shows that the barrier is attained several times and allows to build regeneration times (vertical lines) and independent blocks just as in the first example. Because of the choice of the parameters (fat tail for the claims), the number of blocks is small on this short period, but in practical insurance applications, we may hope to have more regenerations (Figure 7.3).

c07f003

Figure 7.3 Splitting the embedded chain of a Cramér–Lundberg model with a dividend barrier. Vertical lines corresponds to regeneration times (when the chain attains the barrier c07-math-0154). The blocks of observations between two vertical lines are independent.

7.2.2 Basic Regeneration Properties

When an accessible atom exists, the stochastic stability properties of the chain are reduced to properties concerning the speed of return time to the atom only. Theorem 10.2.2 in Meyn and Tweedie (1996) shows, for instance, that the chain c07-math-0155 is positive recurrent i.f.f. c07-math-0156. The (unique) invariant probability distribution c07-math-0157 is then the Pitman occupation measure given by

In the case c07-math-0159, if there exists c07-math-0160 such that c07-math-0161 and c07-math-0162, for any c07-math-0163, the chain is said c07-math-0164, and there exists an invariant measure (not a probability) for the chain. The splitting into independent blocks still holds (see, for instance, Tjöstheim, 1990; Karlsen and Tjöstheim, 2001). This includes the case of the random walk (with c07-math-0165, and such procedure may be useful for studying the properties of the maximum for Markovian processes which have somehow the same kind of behavior as long-range memory processes. We will not consider this more technical case here. For atomic chains, limit theorems can be derived from the application of the corresponding results to the i.i.d. blocks c07-math-0166 (see (Smith, 1992) and the references therein). For instance, using this kind of techniques, Meyn and Tweedie (1996) have proved the law of large number (LLN), the central limit theorem (CLT), and laws of iterated logarithm (LIL) for Markov chains. Bolthausen (1980) obtained a Berry–Esseen-type theorem, and Malinovskibreive (1985), Malinovskibreive (1987, 1989); Bertail and Clémençon (2006b) have proved other refinements of the CLT, in particular Edgeworth expansions. The same technique can also be applied to establish moment and probability inequalities, which are not asymptotic results (see (Clémençon, 2001); (Bertail and Clémençon, 2010)).

Recall that a set c07-math-0167 is said to be small for c07-math-0168 if there exist c07-math-0169, c07-math-0170 and a probability measure c07-math-0171 supported by c07-math-0172 such that, for all c07-math-0173, c07-math-0174,

denoting by c07-math-0176 the c07-math-0177th iterate of the transition kernel c07-math-0178. In the sequel, (7.2) is referred to as the minorization condition c07-math-0179. Recall that accessible small sets always exist for c07-math-0180-irreducible chains: any set c07-math-0181 such that c07-math-0182 contains such a set (cf (Jain and Jamison, 1967)). In many models of interest c07-math-0183 but even if it is not the case it is possible to vectorize the Markov chains to reduce the study of this condition to c07-math-0184. Even if it entails replacing the initial chain c07-math-0185 by the chain c07-math-0186, we now suppose c07-math-0187. From a practical point of view, the minorizing probability measure may be chosen by the user. For instance, c07-math-0188 may be the uniform distribution over a given small set, typically a compact set which is often visited by the chain; then in this case c07-math-0189 may simply be seen as the minimum of the c07-math-0190 over c07-math-0191. Of course in practice c07-math-0192 is unknown but easily estimable so that plug-in estimators of these quantities may be easily constructed (see following text).

7.2.3 The Nummelin Splitting Trick and A Constructive Approximation

We now precise how to construct the atomic chain onto which the initial chain c07-math-0193 is embedded. Suppose that c07-math-0194 satisfies c07-math-0195 for c07-math-0196 such that c07-math-0197. The sample space is expanded so as to define a sequence c07-math-0198 of independent Bernoulli random variables (r.v.'s) with parameter c07-math-0199 by defining the joint distribution c07-math-0200 whose construction relies on the following randomization of the transition probability c07-math-0201 each time the chain hits c07-math-0202. If c07-math-0203 and

  • if c07-math-0204 (with probability c07-math-0205), then c07-math-0206.
  • if c07-math-0207, then c07-math-0208.

The key point of the construction relies on the fact that c07-math-0209 is an atom for the bivariate Markov chain c07-math-0210, which inherits all its communication and stochastic stability properties from c07-math-0211 (refer to Chapter 14 in Meyn and Tweedie (1996)).

Here we assume further that the conditional distributions c07-math-0212 and the initial distribution c07-math-0213 are dominated by a c07-math-0214-finite measure c07-math-0215 of reference, so that c07-math-0216 and c07-math-0217 for all c07-math-0218. For simplicity, we suppose that condition c07-math-0219 is fulfilled with c07-math-0220. Hence, c07-math-0221 is absolutely continuous with respect to c07-math-0222 too, and, setting c07-math-0223,

If we were able to generate binary random variables c07-math-0225, c07-math-0226, so that c07-math-0227 be a realization of the split chain described previously, then we could divide the sample path c07-math-0228, c07-math-0229 into regeneration blocks. Given the sample path c07-math-0230, it may be shown that the c07-math-0231's are independent r.v.'s and the conditional distribution of c07-math-0232 is the Bernoulli distribution with parameter

Therefore, knowledge of c07-math-0234 over c07-math-0235 is required to draw c07-math-0236,c07-math-0237 by this way.

A natural way of mimicking the Nummelin splitting construction consists in computing first an estimate c07-math-0238 of the transition density over c07-math-0239, based on the available sample path and such that c07-math-0240 a.s. for all c07-math-0241, and then generating independent Bernoulli random variables c07-math-0242 given c07-math-0243, the parameter of c07-math-0244 being obtained by plugging c07-math-0245 into (7.4) in place of c07-math-0246. We point out that, from a practical point of view, it actually suffices to draw the c07-math-0247's only at times c07-math-0248 when the chain hits the small set c07-math-0249. c07-math-0250 indicates whether the trajectory should be cut at time point c07-math-0251 or not. Proceeding this way, one gets the sequence of approximate regeneration times, namely, the successive time points at which c07-math-0252 visits the set c07-math-0253. Setting c07-math-0254 for the number of splits (i.e., the number of visits of the approximated split chain to the artificial atom), one gets a sequence of approximate renewal times,

7.5 equation

with c07-math-0256 by convention and forms the approximate regeneration blocks c07-math-0257.

The knowledge of the parameters c07-math-0258, c07-math-0259, c07-math-0260 of condition (7.3) is required for implementing this approximation method. A practical method for selecting those parameters in a fully data-driven manner is described at length in Bertail and Clémençon (2007). The idea is essentially to select a compact set around the mean of the time series and to increase its size. Indeed, if the small set is too small, then there will be no data in it and the Markov chain could not be split. On the contrary, if the small set is too large, the minimum c07-math-0261 over the small set will be very small, and there is little change that the we observe c07-math-0262 As the size increases, the number of regenerations increases up to an optimal value and then decreases; the choice of the small set and of the corresponding splitting is then entirely driven by the observations. To illustrate these ideas, we apply the method to a financial time series, assuming that it is Markovian (even if there are some structural changes, the Markovian nature still remains).

Example 3: Splitting a nonregenerative financial time series

Many financial time series exhibit some nonlinearities and structural changes both in level and variance. To illustrate how it is possible to divide such kind of data into “almost” independent blocks, we will study a particular model exhibiting such behavior.

Consider the following Smooth Exponential Threshold AutoRegressive Model with AutoRegressive Conditional Heteroskedasticity (SETAR(1)-ARCH(1)) model defined by

equation

where the noise c07-math-0264 are i.i.d with variance c07-math-0265. See Fan and Yao (2003) for a detailed description of these kinds of nonlinear models. It may be used to model log returns or log prices. Notice that this Markov chain (of order 1) may be seen as a continuous approximation of a threshold model. Assume that c07-math-0266, then for large values of c07-math-0267, it is easy to see that in mean c07-math-0268 behaves like a simple c07-math-0269 model with coefficient c07-math-0270 (ensuring that the process will come back to its mean, equal to 0). Conversely, for small values of c07-math-0271 (close to 0), the process behaves like an c07-math-0272 model with coefficient c07-math-0273 (eventually explosive if c07-math-0274). This process is thus able to engender bursting bubbles. The heteroscedastic part implies that the conditional variance c07-math-0275 c07-math-0276 may be strongly volatile when large values (the bubble) of the series occur. To ensure stationarity, we require c07-math-0277.

In the following simulation, we choose c07-math-0278, c07-math-0279, c07-math-0280, c07-math-0281, and c07-math-0282. The following graph panel shows the Nadaraya estimator of the transition density and the number of blocks obtained as the size of the small set increases. For a small set of the form c07-math-0283, we obtain c07-math-0284 pseudoblocks and the mean length of a block is close to c07-math-0285 The estimated lower bound for the density over the small set c07-math-0286 is c07-math-0287 The third graphic shows the level sets of the density and the corresponding optimal small set (containing the possible points at which the times series may be split). The last graph shows the original time series and the corresponding pseudoblocks obtained for an optimal data-driven small set.

Beyond the consistency property of the estimators that we will later study, this method has an important advantage that makes it attractive from a practical perspective: blocks are here entirely determined by the data (up to the approximation step), in contrast to standard blocking techniques based on fixed length blocks. Indeed, it is well known that the choice of the block length is crucial to obtain satisfactory results and is a difficult technical task.

7.2.4 Some Hypotheses

The validity of this approximation has been tackled in Bertail and Clémençon (2006a) using a coupling approach. Precisely, the authors established a sharp bound for the deviation between the distribution of c07-math-0296 and the one of the c07-math-0297 in the sense of Wasserstein distance. The coupling “error” essentially depends on the rate of the mean squared error (MSE) of the estimator of the transition density

with the sup norm over c07-math-0299 as a loss function under the next conditions:

  1. A1. The parameters c07-math-0300 and c07-math-0301 in (7.3) are chosen so that c07-math-0302.
  2. A2. c07-math-0303 and c07-math-0304-almost surely c07-math-0305.

Throughout the next sections, c07-math-0306 denotes a fixed real-valued measurable function defined on the state space c07-math-0307. To study the properties of the block, we will also need the following usual moment conditions on the time return (Figure 7.4).

  1. A3 (Regenerative case)
    equation

    and their analog versions in the nonregenerative case.

  2. A4 (General Harris recurrent case)
    equation
c07f004

Figure 7.4 Splitting a Smooth Exponential Threshold Arch time series, with c07-math-0288, c07-math-0289, c07-math-0290, c07-math-0291 and c07-math-0292. (a) Estimator of the transition density. (b) Visit of the chain to the small set c07-math-0293 and the level sets of the transition density estimator: the optimal small set should contain a lot of points in a region with high density. (c) Number of regenerations according to the size c07-math-0294 of the small set, optimal for c07-math-0295. (d) Splitting (vertical bars) of the original time series, with horizontal bars corresponding to the optimal small set.

7.3 Preliminary Results

Here we begin by briefly recalling the connection between the (pseudo) regeneration properties of a Harris chain c07-math-0310 and the extremal behavior of sequences of type c07-math-0311, firstly pointed out in the seminal contribution of Rootzén (1988) (see also (Asmussen, 1998b); (Hansen and Jensen, 2005)).

7.3.1 Cycle Submaxima for Regenerative Markov Chains

We first consider the case when c07-math-0312 possesses a known accessible atom c07-math-0313. In the following we denote c07-math-0314 For c07-math-0315, define the submaximum over the c07-math-0316th cycle of the sample path:

7.7 equation

In the following c07-math-0318 denotes the number of visits of c07-math-0319 to the regeneration set c07-math-0320 until time c07-math-0321 . c07-math-0322 denotes the maximum over the first cycle (starting from an initial distribution c07-math-0323 Because of the “initialization” phase, its distribution is different from the others and essentially depends on c07-math-0324 denotes the maximum over the last nonregenerative data block (meaning by that it may be an incomplete block, since we may not observe the return to the atom A) with the usual convention that maximum over an empty set equals to c07-math-0325.

With these definitions, it is easy to understand that the maximum value c07-math-0326, taken by the sequence c07-math-0327 over a trajectory of length c07-math-0328, may be naturally expressed in terms of submaxima over cycles

7.8 equation

By the strong Markov property and independence of the blocks, the c07-math-0330's are i.i.d. r.v.'s with common distribution function (d.f.) c07-math-0331. Moreover, by Harris recurrence, the number of blocks is of order c07-math-0332 c07-math-0333-almost surely as c07-math-0334. Thus, c07-math-0335 behaves like the maximum of c07-math-0336 i.i.d. r.v.c07-math-0337. The following result established in Rootzén (1988) shows that the limiting distribution of the sample maximum of c07-math-0338 is entirely determined by the tail behavior of the d.f. c07-math-0339 and relies on this crucial asymptotic independence of the blocks.

In the terminology of O'Brien (see (O'Brien, 1974, 1987), c07-math-0344 may be seen as a so-called phantom distribution, that is, an artificial distribution which gives the same distribution for the maximum as in the i.i.d. case. Indeed the preceding theorem shows that the distribution of the maximum behaves exactly as if the observations were independent with distribution c07-math-0345. As a consequence, the limiting behavior of the maximum in this dependent setting may be simply retrieved by using the famous Fischer–Tippett–Gnedenko theorem (obtained in the i.i.d. case), with the marginal distribution replaced by the phantom distribution c07-math-0346. Then, the asymptotic behavior of the sample maximum is entirely determined by the tail properties of the d.f. c07-math-0347. In particular, the limiting distribution of c07-math-0348 (for a suitable normalization) is the generalized extreme value distribution function c07-math-0349 with parameter c07-math-0350, given by

equation

In the following c07-math-0352 will be referred as extreme value index. When c07-math-0353, we will also call it the tail index, corresponding to a Pareto-like distribution. The smaller c07-math-0354, the heavier the tail is.

In the following, we assume that c07-math-0364 belongs to the maximum domain of attraction c07-math-0365say, c07-math-0366 (refer to (Resnick, 1987) for basics in extreme value theory). Then, there exist some sequences c07-math-0367 and c07-math-0368 such that c07-math-0369 as c07-math-0370 and we have c07-math-0371 as c07-math-0372, with c07-math-0373.

7.3.1.1 Estimation of the cycle submaximum cumulative distribution function

In the atomic case, the c.d.f. c07-math-0374 of the cycle submaxima, c07-math-0375 with c07-math-0376, may be naturally estimated by the empirical counterpart d.f. from the observation of a random number c07-math-0377 of complete regenerative cycles, namely,

7.11 equation

with c07-math-0379 by convention when c07-math-0380. Notice that the first and the last (nonregenerative blocks) are dropped in this estimator. As a straightforward consequence of Glivenko-Cantelli's theorem for i.i.d. data, we have that

7.12 equation

Furthermore, by the LIL, we also have c07-math-0382 a.s.

7.3.1.2 Estimation of submaxima in the pseudoregenerative case

Cycle submaxima of the split chain are generally not observable in the general Harris case, since Nummelin extension depends on the true underlying transition probability. However, our regeneration-based statistical procedures may be directly applied to the submaxima over the approximate regeneration cycles. Define the pseudoregenerative block maxima by

7.13 equation

for c07-math-0384. The empirical d.f. counterpart is now given by

7.14 equation

with, by convention, c07-math-0386 if c07-math-0387. As shown by the next theorem, using the approximate cycle submaxima instead of the “true” ones does not affect the convergence under assumption A1. Treading in the steps of Bertail and Clémençon (2004a), the proof essentially relies on a coupling argument.

For smooth Markov chains with smooth c07-math-0397 transition kernel density, the rate of convergence of c07-math-0398 will be close to c07-math-0399. Under standard Hölder constraints of order c07-math-0400, the typical rate for the MSE (7.6) is of order c07-math-0401 so that c07-math-0402 =c07-math-0403.

7.4 Regeneration-based Statistical Methods for Extremal Events

The core of this paragraph is to show that, in the regenerative setup, consistent statistical procedures for extremal events may be derived from the application of standard inference methods introduced in the i.i.d. setting.

In the case when assumption (7.9) holds, one may straightforwardly derive from (7.10) estimates of c07-math-0404 as c07-math-0405 and c07-math-0406 based on the observation of (a random number of) submaxima c07-math-0407 over a sample path of length c07-math-0408, as proposed in Glynn and Zeevi (2000). Because of the estimation step, we will require that c07-math-0409 Indeed, if we want to obtain convergent estimators of the distribution of the maximum, we need to subsample the size of the maximum to ensure that the empirical estimation procedure does not alter the limiting distribution. For this, put

with c07-math-0411. The next limit result establishes the asymptotic validity of estimator (7.16) for an adequate choice of c07-math-0412 depending both on the number of regenerations and of the size c07-math-0413, extending this way Proposition 3.6 of Glynn and Zeevi (2000). If computations are carried out with the pseudoregeneration cycles, under some additional technical assumptions taking into account the order of the approximation of the transition kernel, the procedure remains consistent. In this case, one would simply consider estimates of the form c07-math-0414. The following theorem is a simple adaption of a theorem given in Bertail et al. (2009).

This result indicates that, in the most favorable case, we can recover the behavior of the maximum only over c07-math-0439 observations with c07-math-0440 much smaller than c07-math-0441. However, it is still possible to estimate the tail behavior of c07-math-0442 by extrapolation techniques (as it is done, for instance, in Bertail et al. (2004)). If, in addition, one assumes that c07-math-0443 belongs to some specific domain of attraction c07-math-0444, for instance, to the Fréchet domain with c07-math-0445 , it is possible to use classical inference procedures (refer to Section 6.4 in (Embrechts et al., 1997), for instance) based on the submaxima c07-math-0446 or the estimated submaxima over pseudocycles to estimate the shape parameter c07-math-0447, as well as the normalizing constants c07-math-0448 and c07-math-0449.

7.5 The Extremal Index

The problem of estimating the extremal index of some functionals of this quantity has been the subject of many researches in the strong mixing framework (see, for instance, (Hsing, 1993); (Ferro and Segers, 2003); and more recently (Robert, 2009); (Robert et al., 2009)). However, we will show that in a Markov chain setting, the estimators are much more simpler to study. Recall that c07-math-0450 is the mean return to the atom c07-math-0451In the following, when the regenerative chain c07-math-0452 is positive recurrent, we denote c07-math-0453, the empirical distribution function of the limiting stationary measurec07-math-0454 given by (7.1). It has been shown (see (Leadbetter and Rootzén, 1988), for instance) that there exists some index c07-math-0455, called the extremal index of the sequence c07-math-0456, such that

for any sequence c07-math-0458 such that c07-math-0459. Once again, c07-math-0460 may be seen as an another phantom distribution. The inverse of the extremal index measures the clustering tendency of high threshold exceedances and how the extreme values cluster together. It is a very important parameter to estimate in risk theory, since it indicates somehow how many times (in mean) an extremal event will reproduce, due to the dependency structure of the data.

As notice in Rootzén (1988), because of the nonunicity of the phantom distribution, it is easy to see from Proposition 7.1 and (7.20) that

7.22 equation
7.23 equation

The last equality is followed by a simple Taylor expansion. In the i.i.d. setup, by taking the whole state space as an atom (c07-math-0464, so that c07-math-0465), one immediately finds that c07-math-0466. In the dependent case, the index c07-math-0467 may be interpreted as the proportionality constant between the probability of exceeding a sufficiently high threshold within a regenerative cycle and the mean time spent above the latter between consecutive regeneration times.

It is also important to notice that Proposition 7.1 combined with (7.20) also entail that, for all c07-math-0468 in c07-math-0469, c07-math-0470 and c07-math-0471 belong to the same domain of attraction (when one of them is in a domain attraction of the maximum). Their tail behavior only differs from the slowly varying functions appearing in the tail behavior. We recall that a slowly varying function is a function c07-math-0472 such that c07-math-0473 as c07-math-0474 for any c07-math-0475. For instance, c07-math-0476, iterated logarithmc07-math-0477 are slowly varying functions.

Suppose that c07-math-0478 and c07-math-0479 belong to the Fréchet domain of attraction; then it is known (cf Theorem 8.13.2 in Bingham et al. (1987)) that there exist c07-math-0480 and two slowly varying functions c07-math-0481 and c07-math-0482 such that c07-math-0483 and c07-math-0484. In this setup, the extremal index is thus simply given by the limiting behavior of

equation

However, estimating slowly varying functions is a difficult task, which requires a lot of data (see (Bertail et al., 2004)). Some more intuitive empirical estimators of c07-math-0486 will be proposed in the following.

In the regenerative case, a simple estimator of c07-math-0487 is given by the empirical counterpart of expression (7.21). c07-math-0488 is a natural a.s. convergent empirical estimate of c07-math-0489. Recalling that c07-math-0490 a.s.c07-math-0491 define for a given threshold c07-math-0492,

7.24 equation

with the convention that c07-math-0494 if c07-math-0495. For general Harris chains, the empirical counterpart of Eq. (7.21) computed from the approximate regeneration blocks is now given by

7.25 equation

with c07-math-0497 by convention when c07-math-0498. The following result has been recently proved in Bertail et al. (2013). Other estimators based on fixed length blocks in the framework of strong mixing processes are given in Robert (2009) and Robert et al. (2009).

7.6 The Regeneration-Based Hill Estimator

As pointed out in Section 7.5, provided that the extremal index of c07-math-0542 exists and is strictly positive, the equivalence c07-math-0543 holds true, in particular in the Fréchet case, for c07-math-0544. Classically, the d.f. c07-math-0545 belongs to c07-math-0546 i.f.f. it fulfills the tail regularity condition

7.32 equation

where c07-math-0548 is a slowly varying function. Statistical estimation of the tail risk index c07-math-0549 of a regularly varying d.f. based on i.i.d. data has been the subject of a good deal of attention since the seminal contribution of Hill (1975). Most methods that boil down to computing a certain functional of an increasing sequence of upper order statistics have been proposed for dealing with this estimation problem, just like the celebrated Hill estimator, which can be viewed as a conditional maximum likelihood approach. Given i.i.d. observations c07-math-0550 with common distribution c07-math-0551, the Hill estimator is

7.33 equation

where c07-math-0553 denotes the c07-math-0554th largest order statistic of the data sample c07-math-0555. The asymptotic behavior of this estimator has been extensively investigated when stipulating that c07-math-0556 goes to c07-math-0557 at a suitable rate. Strong consistency is proved when c07-math-0558 and c07-math-0559 as c07-math-0560 in Deheuvels et al. (1988). Its asymptotic normality is established in Goldie (1991): under further conditions on c07-math-0561 (referred to as second-order regular variation) and c07-math-0562, we have the convergence in distribution c07-math-0563, c07-math-0564.

The regeneration-based Hill estimator based on the observation of the c07-math-0565 submaxima c07-math-0566, denoting by c07-math-0567 the c07-math-0568th largest submaximum, is naturally defined as

with c07-math-0570 when c07-math-0571. Observing that, as c07-math-0572, c07-math-0573 with c07-math-0574 probability one, limit results holding true for i.i.d. data can be immediately extended to the present setting (cf assertion (i) of Proposition 7.9). In the general Harris situation, an estimator of exactly the same form can be used, except that approximate submaxima are involved in the computation:

with c07-math-0576 when c07-math-0577. As shown by the next result, the approximation stage does not affect the consistency of the estimator, on the condition that the estimator c07-math-0578 involved in the procedure is sufficiently accurate. For the purpose of building Gaussian asymptotic confidence intervals (CI) in the nonregenerative case, the estimator c07-math-0579 is also considered, still given by Eq. (7.35).

Before showing how the extreme value regeneration-based statistics reviewed in the present article practically perform on several examples, a few comments are in order.

The tail index estimator (7.34) is proved strongly consistent under mild conditions in the regenerative setting, whereas only (weak) consistency has been established for the alternative method proposed in Resnick and Stărică (1995) under general strong mixing assumptions. The condition stipulated in assertion (ii) may not be satisfied for some c07-math-0598. When the slowly varying function c07-math-0599 equals, for instance, c07-math-0600, it cannot be fulfilled. Indeed in this case, c07-math-0601 should be chosen of order c07-math-0602 according to the von Mises conditions. In contrast, choosing a subsampling size c07-math-0603 such that the conditions stipulated in assertion (iii) hold is always possible. The issue of picking c07-math-0604 in an optimal fashion in this case remains open.

Given the number c07-math-0605 (c07-math-0606 or c07-math-0607) of (approximate) regeneration times observed within the available data series, the tuning parameter c07-math-0608 can be selected by means of standard methods in the i.i.d. context. A possible solution is to choose c07-math-0609 so as to minimize the estimated MSE

equation

where c07-math-0611 is a bias-corrected version of the Hill estimator. Either the jackknife method or else an analytical method (see (Feuerverger and Hall, 1999) or (Beirlant et al., 1999)) can be used for this purpose. The randomness of the number of submaxima is the sole difference here.

7.7 Applications to Ruin Theory and Financial Time Series

As an illustration, we now apply the inference methods described in the previous section to two models from the insurance and the financial fields.

7.7.1 Cramér–Lundberg model with a barrier: Example 2

Considering Example 2, we apply the preceding results to obtain an approximation of the distribution of the subminimum, the global minimum (i.e., the probability of ruin over a given period), and the extremal index. We will not consider here the subexponential case (heavy-tailed claims) for which it is known that the extremal index is equal to c07-math-0612, corresponding to infinite clusters of extreme values (see (Asmussen, 1998b)). Recall that the continuous process of interest is given by

equation

and that the embedded chain satisfies

equation

Notice that if the barrier c07-math-0615 is too high in comparison to the initial reservec07-math-0616, then the chain will regenerate very rarely (unless the price c07-math-0617 is very high) and the method will not be useful. But if the barrier is attained at least one time, then the probability of ruin will only depend on c07-math-0618 not on c07-math-0619. Assume that c07-math-0620 is c07-math-0621 and the claims are distributed as c07-math-0622 with c07-math-0623 . The safety loading is then given by c07-math-0624 and is assumed to be nonnegative to ensure that the probability of ruin is not equal to 1 a.s.

Using well-known results in the case of i.i.d. exponential inputs and outputs, the extremal index is given by c07-math-0625 c07-math-0626. In our simulation we choose c07-math-0627 and c07-math-0628 with c07-math-0629 so that the extremal index is given here by c07-math-0630. We emphasize the fact that we need to observe the times series over a very long period (5000 days) so as to observe enough cycles. The barrier is here at c07-math-0631 with a initial reserve c07-math-0632

For c07-math-0633 and if we choose c07-math-0634 of order c07-math-0635, with proposition 3 by calculating the quantile of c07-math-0636 of order c07-math-0637 for c07-math-0638,we obtain that c07-math-0639. This is an indicator that in the next 70 days there is a rather high probability of being ruined. Inversely, some straightforward inversions (here c07-math-0640 show that the probability of ruin

equation

and that

equation

This strongly suggests that the dividend barrier and the initial reserve are too low.

As far as the extremal index is concerned, we obtain a rather good estimator of c07-math-0643 as shown in Figure 7.5 (see also the simulation results in Bertail et al. (2013) in a slightly different setting (M/M/1 queues)). It represents the value of c07-math-0644 for a sequence of high value of the threshold. The stable part of c07-math-0645 for a large range of value of levels corresponding to c07-math-0646 is very close to the true value. It should be noticed that when c07-math-0647 is too high, the quantiles ofc07-math-0648 are badly estimated, resulting in a very bad estimation of c07-math-0649. Although we did not present in this chapter the validity of the regenerative bootstrap (i.e., bootstrapping regenerative blocks) as shown in Bertail et al. (2013), we represent the corresponding bootstrap CI on the graphics. It is also interesting to notice that the change in width of the CI is a good indicator in order to choose the adequate level c07-math-0650.

c07f005

Figure 7.5 Estimator (continuous line) and bootstrap confidence interval (dotted lines) of the extremal index c07-math-0651, for a sequence of high values of the threshold c07-math-0652 (seen as a quantile of the c07-math-0653-coordinate). True value of c07-math-0654.

7.7.2 Pseudoregenerative financial time series: extremal index and tail estimation

We will consider the model exhibited in Example 3 for a much more longer stretch of observations. Recall that the process is given by the nonlinear autoregressive form

equation

Indeed the methods used here will only be of interest when c07-math-0656 and the number of pseudoregeneration is not too small. The rate of convergence of the Hill estimator is also strongly influenced by the presence of the slowly varying function (here in the distribution of the submaxima). Recall that if the slowly varying function belongs to the Hall's family, that is, is of the form, for some c07-math-0657 and c07-math-0658,

equation

then the optimal rate of convergence of the Hill estimator is of order at most c07-math-0660 (see (Goldie, 1991)). Thus, if c07-math-0661 is small, the rate of convergence of the Hill estimator may be very slow. In practice, we rarely estimate the slowly varying function, but the index is determined graphically by looking at range c07-math-0662 of extreme values, where the index is quite stable. We also use the bias correction methods (Feuerverger and Hall, (1999) or Beirlant et al., (1999)) mentioned before, which greatly improve the stability of the estimators.

We now present in Figure 7.6 a path of an SETAR-ARCH process, with a large value of c07-math-0663. We choose c07-math-0664, c07-math-0665, and c07-math-0666, which ensure stationarity of the process. This process clearly exhibits the features of many log returns encountered in finance. The optimal small set (among those of the form c07-math-0667) is given by c07-math-0668, which is quite large, because of the variability of the time series, with a corresponding value of c07-math-0669.

c07f006

Figure 7.6 Simulation of the SETAR-ARCH process for c07-math-0670, c07-math-0671, c07-math-0672, and c07-math-0673, exhibiting strong volatility and large excursions.

The true value of c07-math-0678 (obtained by simulating several very long time series c07-math-0679) is close to 0.50. This means that maxima clusterize by pair. Figure 7.7 presents the dependence index estimator for a range of values of the threshold (the level of the quantile is given on the axe). The estimator is rather unstable for large quantiles, but we clearly identify a zone of stability near the true value of c07-math-0680. Bootstrap CI lead to an estimator of c07-math-0681, between c07-math-0682 and 0.587 at the level 95% (which is in the range of the limit theorem given before). The problem of choosing the optimal value of c07-math-0683 in this case is still an open problem.

c07f007

Figure 7.7 Estimator (continuous line) and confidence intervals (dotted lines) of the extremal index as a function of the quantile level c07-math-0674, for a sequence of high values of the threshold c07-math-0675 (seen as a quantile of the c07-math-0676-coordinate). True value of c07-math-0677 close to 0.5.

Figure 7.8 presents the regenerative Bootstrap distribution (Bertail and Clémençon (2006b)) of the Hill estimator, with a choice of the optimal fraction c07-math-0684 obtained by minimizing the MSE. This leads to a CI for the tail (with a error rate of 5%) of the distribution given by c07-math-0685. This suggests that for this process, the tail may be quite heavy, since even the moment of order 3 may not exist.

c07f008

Figure 7.8 Bootstrap distribution of the pseudoregenerative Hill estimator (smoothed with a Gaussian kernel), based on c07-math-0686 bootstrap replications. Mode around 2.8.

7.8 An Application to the CAC40

We will apply our method to the daily log return of the CAC40, from 10/07/1987 to 06/16/2014, assuming that this time series follows a Markov chain. Such hypothesis has been tested by several authors (using discretization methods) on other periods, suggesting that the usual stochastic volatility models (Black and Scholes) may not be appropriate in this case (see, for instance, McQueen and Thorley (1991), Jondeaua and Rockinger (2003), Bhat and Kuma (2010), and Cont (2001)). The log returns are plotted in Figure 7.9. Notice that the time series exhibits the same features as the SETAR-ARCH model studied before. However, we do not assume here any specific model for the underlying Markov chain. We observe a lot of regeneration blocks (1567 over 6814 observations) in a small set of optimal size close to c07-math-0687 (a minorizing constant close to c07-math-0688, yielding blocks of mean size 4.35 (Figure 7.9).

c07f009

Figure 7.9 Log returns of the CAC40, from 10/07/1987 to 06/16/2014.

We have used two different specifications for the Markov chains, a Markov chain of order 1 and 2. The results are very similar, and we thus present only the results for a specification of a Markov chain of order 1. We distinguish between the behavior of the Markov chain for the minimum and the maximum, for which both the tail index and the extremal index may be different, leading to an asymmetric behavior between gains and losses. The following table summarizes the main results: we give the value of the estimators of the tail and extremal index as well as Bootstrap CI, respectively, for the minimum and the maximum of the time series.

Estimators/left and right tail Min (left tail) Max (right tail)
Hill tail index estimator 0.307 0.328
Lower CI tail index 2.5% 0.242 0.273
Upper CI tail index 97.5% 0.361 0.389
Extremal index estimator 0.440 0.562
Lower CI extremal index 2.5% 0.359 0.494
Upper CI extremal index 97.5% 0.520 0.614

Estimators and confidence intervals for tails and extremal indexes.

The extremal index estimators are very stable when the threshold c07-math-0689 is changed, yielding very robust estimatorsc07-math-0690 We emphasize that the tail of the process is very heavy since we are close to the nonexistence of the moment of order 4. A simple test based on the Bootstrap CI allows us to accept the hypothesis that c07-math-0691 against c07-math-0692 but reject the existence of the moment of order four, c07-math-0693 against c07-math-0694 for a type I error of c07-math-0695.

A striking feature of these results is seemingly some asymmetry in the times series between the minimum and the maximum log returns. In both case, the process has heavy tails with a much more heavy tail for positive log returns, but with a dynamic which creates smaller clusters of extremum values for maximum (of mean size 1.78) than for minimum (of mean size 2.27). This means that losses may be less strong than gains but may be more persistent. However, a simple test (consisting in comparing the confidence regions) yields that we do not reject the hypothesis of similar tail. This goes in the same direction as Jondeaua and Rockinger (2003) on a different period with different method, rather based on the notion of weak dependence.

7.9 Conclusion

Given the ubiquity of the Markov assumption in time series modeling and applied probability models, we review in this chapter several statistical methods, specifically tailored for the Markovian framework with a view toward the extremal behavior of such processes. Precisely, this chapter looks at statistical inference for extremal events from the renewal theory angle. We recalled that certain extremal behavior features of Harris Markov chains may be also expressed in terms of regeneration cycles, namely, data segments between consecutive regeneration times c07-math-0696 (i.e., random times at which the chain forgets its past). Working on this approach, the methodology proposed in this chapter boils down to split up the observed sample path into regeneration data blocks (or into data blocks drawn from a distribution approximating the regeneration cycle's distribution, in the general case when regeneration times cannot be observed). Then the analysis boils down to examining the sequence of maxima over the resulting data segments, as if they were i.i.d., via standard statistical methods. In order to illustrate the interest of this technique, we have concentrated on several important inference problems concerning the question of estimating the sample maximum's tail, the extremal dependence index, and the tail index. However many other parameters of interest may be investigated in the same manner. The two examples given here—ruin models in insurance and times series exhibiting threshold and/or strong conditional heteroscedasticity—clearly show the potentiality of such methods in these fields. An illustration of the estimation methods to the CAC40 shows the potential of the method for real data applications.

References

  1. Asmussen, S. Extreme value theory for queues via cycle maxima. Extremes 1998a;1(2):137–168.
  2. Asmussen, S. Subexponential asymptotics for stochastic processes: extremal behavior, stationary distributions and first passage probabilities. Adv Appl Probab 1998b;8(2):354–374.
  3. Asmussen, S. Applied Probability and Queues. New York: Springer-Verlag; 2003.
  4. Asmussen, S., Binswanger, K., Höjgaard, B. Rare events simulation for heavy-tailed distributions. Bernoulli 2000;6:303–322.
  5. Beirlant, J., Dierckx, G., Goegebeur, Y., Matthys, G. Tail index estimation and an exponential regression model. Extremes 1999;2(2):177–200.
  6. Bertail, P., Clémençon, S. Edgeworth expansions for suitably normalized sample mean statistics of atomic Markov chains. Probab Theory Relat Fields 2004a;130(3):388–414.
  7. Bertail, P., Clémençon, S. Note on the regeneration-base bootstrap for atomic Markov chains. Test 2004b;16:109–122.
  8. Bertail, P., Clémençon, S. Regeneration-based statistics for Harris recurrent Markov chains. In: Bertail, P., Doukhan, P., Soulier, P., editors. Probability and Statistics for Dependent Data. Volume 187 of Lecture Notes in Statistics. Springer-Verlag; 2006a. pp 3–54.
  9. Bertail, P., Clémençon, S. Regenerative-block bootstrap for Markov chains. Bernoulli 2006b;12(4):689–712.
  10. Bertail, P., Clémençon, S. Approximate regenerative block-bootstrap for Markov chains. Computational Statistics and Data Analysis 2007;52(5):2739–2756.
  11. Bertail, P., Clémençon, S. Sharp bounds for the tails of functionals of Markov chains; 2010. 54(3):505–515.
  12. Bertail, P., Clémençon, S., Tressou, J. Extreme value statistics for Markov chains via the (pseudo-)regenerative method. Extremes 2009;12:327–360.
  13. Bertail, P., Clémençon, S., Tressou, J. A renewal approach to Markovian U-statistics. Mathematical Methods of Statistics 2011;20:79–105.
  14. Bertail, P., Clémençon, S., Tressou, J. Regenerative block-bootstrap confidence intervals for the tail and extremal indexes. Electronic Journal of Statistics 2013;7:1224–1248.
  15. Bertail, P., Clémençon, S., Tressou, J. Bootstrapping robust statistics for Markovian data applications to regenerative R- and L-statistics. J Times Ser Anal 2015;36(3):462–480.
  16. Bertail, P., Haeffke, C., Politis, D., White, H. A subsampling approach to estimating the distribution of diverging statistics with applications to assessing financial market risks. Journal of Econometrics 2004;120:295–326.
  17. Bhat, H., Kuma, N. Markov tree options pricing. Proceedings of the 4th SIAM Conference on Mathematics for Industry; 2010.
  18. Bingham, N.H., Goldie, C.M., Teugels, J.L. Regular variation. Encyclopedia of Mathematics and its Applications. Cambridge: Cambridge University Press; 1987.
  19. Bolthausen, E. The Berry-Esseen theorem for functionals of discrete Markov chains. Z Wahrsch Verw Geb 1980;54(1):59–73.
  20. Clémençon, S. Moment and probability inequalities for sums of bounded additive functionals of a regular Markov chains via the Nummelin splitting technique. Statistics and Probability Letters 2001;55:227–238.
  21. Cont, R. Empirical properties of asset returns: stylized facts and statistical issues. Quantitative Finance 2001;1:223–236.
  22. Deheuvels, P., Häusler, E., Mason, D. Almost sure convergence of the Hill estimator. Math Proc Camb Philos Soc 1988;104:371–381.
  23. Embrechts, P., Klüppelberg, C., Mikosch, T. Modelling Extremal Events for Insurance and Finance. Applications of Mathematics. Berlin: Springer-Verlag; 1997.
  24. Fan, J., Yao, Q. Nonlinear Time Series - Nonparametric and Parametric Methods. New York: Springer-Verlag; 2003. 576pp.
  25. Ferro, C., Segers, J. Inference for clusters of extreme values. J R Stat Soc Ser B Stat Methodol 2003;65(2):545–556.
  26. Feuerverger, A., Hall, P. Estimating a tail exponent by modelling departure from a Pareto Distribution. Ann Stat 1999;27:760–781.
  27. Glynn, P., Zeevi, A. Estimating tail probabilities in queues via extremal statistics. In: McDonald, D., Turner, S., editors. Analysis of Communication Networks: Call Centres, Traffic, and Performance. Providence (RI): AMS Press; 2000. p 135–158.
  28. Goldie, C. Implicit renewal theory and tails of solutions of random equations. Ann Appl Probab 1991;1:126–166.
  29. Haiman, G., Kiki, M., Puri, M. Extremes of Markov sequences. Journal of Statistical Planning Inference 1995;45:185–201.
  30. Hansen, N., Jensen, A. The extremal behaviour over regenerative cycles for Markov additive processes with heavy tails. Stoch Proc Appl 2005;115:579–591.
  31. Hill, B. A simple approach to inference about the tail of a distribution. Ann Stat 1975;3:1163–1174.
  32. Hsing, T. On the extreme order statistics for a stationary sequence. Stoch Proc Appl 1988;29(1):155–169.
  33. Hsing, T. On tail estimation using dependent data. Ann Stat 1991;19:1547–1569.
  34. Hsing, T. Extremal index estimation for a weakly dependent stationary sequence. Ann Stat 1993;21(4):2043–2071.
  35. Jain, J., Jamison, B. Contributions to Doeblin's theory of Markov processes. Z Wahrsch Verw Geb 1967;8:19–40.
  36. Jondeaua, E., Rockinger, M. Testing for differences in the tails of stock-market returns. Journal of Empirical Finance 2003;10(5):559–581.
  37. Karlsen, H.A., Tjöstheim, D. Nonparametric estimation in null recurrent time series. Ann Stat 2001;29:372–416.
  38. Leadbetter, M., Rootzén, H. Extremal theory for stochastic processes. Ann Probab 1988;16:431–478.
  39. Loynes, R. Extreme values in uniformly mixing stochastic processes. Ann Math Stat 1965;36:993–999.
  40. McQueen, G., Thorley, S. Are stock returns predictable? A test using Markov chains. The Journal of Finance 1991;46:239–63.
  41. Malinovskii, V. On some asymptotic relations and identities for Harris recurrent Markov chains. In: Statistics and Control of Stochastic Processes (Steklov seminar, Moscow, 1984). New York: Optimization Software; 1985. p 317–336.
  42. Malinovskii, V. Limit theorems for Harris Markov chains I. Theory Probab Appl 1987;31:269–285.
  43. Malinovskii, V. Limit theorems for Harris Markov chains II. Theory Probab Appl 1989;34:252–265.
  44. Meyn, S., Tweedie, R. Markov Chains and Stochastic Stability. London: Springer-Verlag; 1996.
  45. Newell, G. Asymptotic extremes for c07-math-0697-dependent random variables. Ann Math Stat 1964;35:1322–1325.
  46. O'Brien, G. The maximum term of uniformly mixing stationary processes. Z Wahrsch Verw Geb 1974;30:57–63.
  47. O'Brien, G. Extreme values for stationary and Markov sequences. Ann Probab 1987;15:281–291.
  48. Resnick, S. Extreme Values, Point Processes and Regular Variation. New York: Springer-Verlag; 1987.
  49. Resnick, S., Stărică, C. Consistency of Hill's estimator for dependent data. J Appl Probab 1995;32:139–167.
  50. Revuz, D. Markov Chains. 2nd ed. Amsterdam: North-Holland; 1984.
  51. Robert, C,Y. Inference for the limiting cluster size distribution of extreme values. Ann Stat 2009;37(1):271–310.
  52. Robert, C., Segers, J., Ferro, C. A sliding blocks estimator for the extremal index. Electron J Stat 2009;3:993–1020.
  53. Rootzén, H. Maxima and exceedances of stationary Markov chains. Adv Appl Probab 1988;20:371–390.
  54. Rootzén, H. Weak convergence of the tail empirical process for dependent sequences. Stoch Proc Appl 2009;119(2):468–490.
  55. Smith, R. The extremal index for a Markov chain. J Appl Probab 1992;29(4):37–45.
  56. Thorisson, H. Coupling Stationarity and Regeneration, Probability and its Applications. New York: Springer-Verlag; 2000.
  57. Tjöstheim, D. Non-linear time series and Markov chains. Adv Appl Probab 1990;22:587–611.
..................Content has been hidden....................

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