Skip to main content
Log in

High-dimensional inference using the extremal skew-t process

  • Published:
Extremes Aims and scope Submit manuscript

Abstract

Max-stable processes are a popular tool for the study of environmental extremes, and the extremal skew-t process is a general model that allows for a flexible extremal dependence structure. For inference on max-stable processes with high-dimensional data, exact likelihood-based estimation is computationally intractable. Composite likelihoods, using lower dimensional components, and Stephenson-Tawn likelihoods, using occurrence times of maxima, are both attractive methods to circumvent this issue for moderate dimensions. In this article we establish the theoretical formulae for simulations of and inference for the extremal skew-t process. We also incorporate the Stephenson-Tawn concept into the composite likelihood framework, giving greater statistical and computational efficiency for higher-order composite likelihoods. We compare 2-way (pairwise), 3-way (triplewise), 4-way, 5-way and 10-way composite likelihoods for models of up to 100 dimensions. Furthermore, we propose cdf approximations for the Stephenson-Tawn likelihood function, leading to large computational gains, and enabling accurate fitting of models in large dimensions in only a few minutes. We illustrate our methodology with an application to a 90-dimensional temperature dataset from Melbourne, Australia.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  • Arellano-Valle, R.B., Genton, M.G.: Multivariate extended skew-t distributions and related families. Metron 68(3), 201–234 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Azzalini, A., Arellano-Valle, R.B.: Maximum penalized likelihood estimation for skew-normal and skew-t distributions. J. Stat. Plan. Infer. 143(2), 419–433 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Azzalini, A., Capitanio, A.: Statistical applications of the multivariate skew normal distribution. J. R. Stat. Soc. Ser. B Stat. Methodol. 61(3), 579–602 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  • Azzalini, A., Genton, M.G.: Robust likelihood methods based on the skew-t and related distributions. Int. Stat. Rev. 76(1), 106–129 (2008)

    Article  MATH  Google Scholar 

  • Beranger, B., Padoan, S.: Extreme dependence models. In: Extreme Value Modeling and Risk Analysis. Chapman and Hall/CRC, pp 325–352 (2015)

  • Beranger, B., Padoan, S.A., Sisson, S.A.: Models for extremal dependence derived from skew-symmetric families. Scand. J. Stat. 44(1), 21–45 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Bienvenüe, A., Robert, C.Y.: Likelihood inference for multivariate extreme value distributions whose spectral vectors have known conditional distributions. Scand. J. Stat. 44(1), 130–149 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Blanchet, J., Davison, A.C.: Spatial modeling of extreme snow depth. Ann. Appl. Stat. 5(3), 1699–1725 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Brown, B.M., Resnick, S.I.: Extreme values of independent stochastic processes. J. Appl. Probability 14(4), 732–739 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  • Buishand, T.A., de Haan, L., Zhou, C.: On spatial extremes: with application to a rainfall problem. Ann. Appl. Stat. 2(2), 624–642 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Castruccio, S., Huser, R., Genton, M.G.: High-order composite likelihood inference for max-stable distributions and processes. J. Comput. Graph. Statist. 25(4), 1212–1229 (2016)

    Article  MathSciNet  Google Scholar 

  • Cooley, D., Cisewski, J., Erhardt, R.J., Jeon, S., Mannshardt, E., Omolo, B.O., Sun, Y.: A survey of spatial extremes: measuring spatial dependence and modeling spatial effects. REVSTAT 10(1), 135–165 (2012)

    MathSciNet  MATH  Google Scholar 

  • Davison, A.C., Gholamrezaee, M.M.: Geostatistics of extremes. In: Proceedings of the Royal Society of London Series A: Mathematical and Physical Sciences, vol. 468, pp 581–608 (2012)

  • Davison, A.C., Padoan, S.A., Ribatet, M.: Statistical modeling of spatial extremes. Statistical Science 27, 161–186 (2012)

    MathSciNet  MATH  Google Scholar 

  • de Fondeville, R., Davison, AC: High-dimensional peaks-over-threshold inference. Biometrika 105(3), 575–592 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  • de Haan, L.: A spectral representation for max-stable processes. Ann. Probab. 12(4), 1194–1204 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  • de Haan, L., Ferreira, A.: Extreme value theory springer series in operations research and financial engineering. Springer, New York (2006). An introduction

    Google Scholar 

  • Dieker, A., Mikosch, T.: Exact simulation of Brown-Resnick random fields at a finite number of locations. Extremes 18(2), 301–314 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Dombry, C., Engelke, S., Oesting, M.: Exact simulation of max-stable processes. Biometrika 103(2), 303–317 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Dombry, C., Eyi-Minko, F.: Regular conditional distributions of continuous max-infinitely divisible random fields. Electron. J. Probab. 18(7), 21 (2013)

    MathSciNet  MATH  Google Scholar 

  • Dombry, C., Éyi-Minko, F., Ribatet, M.: Conditional simulation of max-stable processes. Biometrika 100(1), 111–124 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Fernandez, C., Steel, M.: Multivariate Student-t regression models: Pitfalls and inference. Biometrika 86(1), 153–167 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  • Genton, M.G., Keyes, D.E., Turkiyyah, G.: Hierarchical decompositions for the computation of high-dimensional multivariate normal probabilities. J. Comput. Graph. Stat. 27(2), 268–277 (2018)

    Article  MathSciNet  Google Scholar 

  • Genton, M.G., Ma, Y., Sang, H.: On the likelihood function of Gaussian max-stable processes. Biometrika 98(2), 481–488 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Genz, A.: Numerical computation of multivariate normal probabilities. J. Comput. Graph. Stat. 1(2), 141–149 (1992)

    Google Scholar 

  • Genz, A.: A comparison of methods for numerical computation of multivariate Normal probabilities. Comput. Scie. Stat. 25, 400–405 (1993)

    Google Scholar 

  • Genz, A., Bretz, F.: Comparison of methods for the computation of multivariate t probabilities. J. Comput. Graph. Stat. 11(4), 950–971 (2002)

    Article  MathSciNet  Google Scholar 

  • Huser, R., Davison, A. C.: Composite likelihood estimation for the Brown-Resnick process. Biometrika 100(2), 511–518 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Huser, R., Davison, A.C., Genton, M.G.: Likelihood estimators for multivariate extremes. Extremes 19(1), 79–103 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Huser, R., Dombry, C., Ribatet, M., Genton, M. G.: Full likelihood inference for max-stable data. Stat 8(1) e218. e218 sta4.218 (2019)

  • Huser, R., Genton, M.G.: Non-stationary dependence structures for spatial extremes. J. Agric. Biol. Environ. Stat. 21(3), 470–491 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Jeffrey, S.J., Carter, J.O., Moodie, K.B., Beswick, A.R.: Using spatial interpolation to construct a comprehensive archive of Australian climate data. Environ. Modell. Softw. 16(4), 309–330 (2001)

    Article  Google Scholar 

  • Kabluchko, Z., Schlather, M., de Haan, L.: Stationary max-stable fields associated to negative definite functions. Ann. Probab. 37(5), 2042–2065 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Liu, Z., Blanchet, J.H., Dieker, A.B., Mikosch, T.: On optimal exact simulation of max-stable and related random fields, arXiv:1609.06001 (2016)

  • Nikoloulopoulos, A.K., Joe, H., Li, H.: Extreme value properties of multivariate t copulas. Extremes 12(2), 129–148 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Oesting, M., Schlather, M., Zhou, C.: Exact and fast simulation of max-stable processes on a compact set using the normalized spectral representation. Bernoulli 24(2), 1497–1530 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  • Opitz, T.: Extremal t processes: Elliptical domain of attraction and a spectral representation. Journal of Multivariate Analysis 122, 409–413 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Padoan, S.A.: Extreme dependence models based on event magnitude. J. Multivariate Anal. 122(0), 1–19 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Padoan, S.A., Ribatet, M., Sisson, S.A.: Likelihood-Based inference for max-stable processes. J. Am. Stat. Assoc. 105(489), 263–277 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Reich, B.J., Shaby, B.A.: A hierarchical max-stable spatial model for extreme precipitation. Ann. Appl. Stat. 6(4), 1430–1451 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Ribatet, M.: Spatial extremes: Max-stable processes at work. J. SFdS 154(2), 156–177 (2013)

    MathSciNet  MATH  Google Scholar 

  • Sang, H., Genton, M.G.: Tapered composite likelihood for spatial max-stable models. Spat. Stat. 8, 86–103 (2014)

    Article  MathSciNet  Google Scholar 

  • Schlather, M.: Models for stationary max-stable random fields. Extremes 5(1), 33–44 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  • Smith, R.L.: Maximum likelihood estimation in a class of nonregular cases. Biometrika 72(1), 67–90 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  • Smith, R.L.: Max-stable processes and spatial extremes. University of Surrey 1990 technical report (1990)

  • Stephenson, A., Tawn, J.: Exploiting occurrence times in likelihood inference for componentwise maxima. Biometrika 92(1), 213–227 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  • Thibaud, E., Aalto, J., Cooley, D.S., Davison, A.C., Heikkinen, J.: Bayesian inference for the brown–resnick process, with an application to extreme low temperatures. Ann. Appl. Stat. 10(4), 2303–2324 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Thibaud, E., Opitz, T.: Efficient inference and simulation for elliptical Pareto processes. Biometrika 102(4), 855–870 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Wadsworth, J.L.: On the occurrence times of componentwise maxima and bias in likelihood inference for multivariate max-stable distributions. Biometrika 102(3), 705–711 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Wadsworth, J.L., Tawn, J.A.: Efficient inference for spatial extreme value processes associated to log-Gaussian random functions. Biometrika 101 (1), 1–15 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  • Wang, Y., Stoev, S.A.: Conditional sampling for spectrally discrete max-stable random fields. Adv. Appl. Probab. 43(2), 461–483 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Warne, D.J., Sisson, S.A., Drovandi, C.C.: Acceleration of expensive computations in Bayesian statistics using vector operations. arXiv:1902.09046 (2019)

  • Whitaker, T., Beranger, B., Sisson, S.A.: Composite likelihood methods for histogram-valued random variables. Stat Comput (2020). https://doi.org/10.1007/s11222-020-09955-5 (2019)

Download references

Acknowledgments

This research was undertaken with the assistance of resources and services from the National Computational Infrastructure (NCI), which is supported by the Australian Government. The authors acknowledge Research Technology Services at UNSW Sydney for supporting this project with compute resources. SAS and BB are supported by the Australian Centre of Excellence for Mathematical and Statistical Frontiers (ACEMS; CE140100049) and the Australian Research Council Discovery Project scheme (FT170100079).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to B. Beranger.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Technical details

1.1 A.1 The non-central extended skew-t distribution (Beranger et al. 2017)

Definition 1

Y is a d-dimensional, non-central extended skew-t distributed random vector, denoted by \(Y\sim \mathcal {S}\mathcal {T}_{d}(\mu , {\Omega }, \alpha , \tau ,\kappa ,\nu )\), if for \(y\in \mathbb {R}^{d}\) it has pdf

$$ \psi_{d}(y;\mu,{\Omega},\alpha,\tau,\kappa,\nu)=\frac{\psi_{d}(y;\mu,{\Omega},\nu)} {{\Psi}\left( \frac{\tau}{\sqrt{1+Q_{\bar{\Omega}}(\alpha)}};\frac{\kappa}{\sqrt{1+Q_{\bar{\Omega}}(\alpha)}},\nu\right)} {\Psi} \left\{(\alpha^{\top} z+\tau) \sqrt{\frac{\nu + d}{\nu + Q_{\bar{\Omega}^{-1}}(z)}};\kappa,\nu+d\right\}, $$

where ψd(y; μ,Ω, ν) is the pdf of a d-dimensional t-distribution with location \(\mu \in \mathbb {R}^{d}\), d × d scale matrix Ω and \(\nu \in \mathbb {R}^{+}\) degrees of freedom, Ψ(⋅; a, ν) denotes a univariate non-central t cdf with non-centrality parameter \(a\in \mathbb {R}\) and ν degrees of freedom, and \(Q_{\bar {\Omega }^{-1}}(z)=z^{\top } \bar {\Omega }^{-1} z\), z = (yμ)/ω, ω = diag(Ω)1/2, \(\bar {\Omega }=\omega ^{-1} {\Omega } \omega ^{-1}\) and \(Q_{\bar {\Omega }}(\alpha )=\alpha ^{\top } \bar {\Omega }\alpha \). The associated cdf is

$$ {\Psi}_{d}(y;\mu,{\Omega},\alpha,\tau,\kappa,\nu) = \frac{{\Psi}_{d+1} \left\{ \bar{z}; {\Omega}^{\ast},\kappa^{\ast}, \nu \right\}} {{\Psi}\left( \bar{\tau}; \bar{\kappa},\nu\right)}, $$

where \(\bar {z}=(z^{\top },\bar {\tau })^{\top }\), Ψd+ 1 is a (d + 1)-dimensional (non-central) t cdf with covariance matrix and non-centrality parameters

$$ {\Omega}^{\ast}=\left( \begin{array}{cc} \bar{\Omega} & - \delta\\ - \delta^{\top} & 1 \end{array} \right), \quad \kappa^{\ast}=\left( \begin{array}{c} 0 \\ \bar{\kappa} \end{array} \right), $$

and ν degrees of freedom, and where

$$ \delta = \left\{ 1+Q_{\bar{\Omega}}(\alpha) \right\}^{-1/2} \bar{\Omega}\alpha,\quad \bar{\kappa} = \left\{ 1+Q_{\bar{\Omega}}(\alpha) \right\}^{-1/2} \kappa, \quad \bar{\tau} = \left\{ 1+Q_{\bar{\Omega}}(\alpha) \right\}^{-1/2} \tau. $$

1.2 A.2 Parameters of the exponent function of the extremal skew-t

First of all, we define mj+m+(sj) as follows

$$ m_{j+} = {\int}_{0}^{\infty} {y}_{j}^{\nu} \phi(y_{j};{\alpha}_{j}^{\ast},{\tau}_{j}^{\ast})\mathrm{d} y_{j} = \frac{2^{(\nu-2)/2} {\Gamma}\{(\nu+1)/2\} {\Psi}({\alpha}_{j}^{\ast} \sqrt{\nu+1};-{\tau}_{j}^{\ast},\nu+1)} {\sqrt{\pi}{\Phi}[\tau\{1+Q_{\bar{\Omega}}(\alpha)\}^{-1/2}]}, $$

where \(\alpha ^{\ast }_{j}\) and \(\tau _{j}^{\ast }\) are respectively the marginal shape and extension parameters and \(Q_{\bar {\Omega }}(\alpha ) = \alpha ^{\top } \bar {\Omega } \alpha \). This then allows us to obtain the exponent function of the extremal skew-t given in Eq. 4 as the sum of (d − 1)-dimensional non-central extended skew-t distribution with correlation matrix \(\bar {\Omega }_{i}^{\circ }=\omega _{I_{i}I_{i}\cdot i}^{-1} \bar {\Omega }_{I_{i}I_{i}\cdot i} \omega _{I_{i}I_{i}\cdot i}^{-1}\), where \(\omega _{I_{i}I_{i}\cdot i}=\text {diag}(\bar {\Omega }_{I_{i}I_{i}\cdot i})^{1/2}\), \(\bar {\Omega }_{I_{i}I_{i}\cdot i}=\bar {\Omega }_{I_{i}I_{i}}-\bar {\Omega }_{I_{i}i}\bar {\Omega }_{iI_{i}}\), I = {1,…, d}, Ii = Ii, \(\bar {\Omega }=\omega ^{-1} {\Omega } \omega ^{-1}\), ω = diag(Ω)1/2. The shape parameter is \(\alpha _{i}^{\circ }=\omega _{I_{i}I_{i}\cdot i} \alpha _{I_{i}} \in \mathbb R^{d-1}\), the extension parameter \(\tau _{j}^{\circ }=(\bar {\Omega }_{jI_{j}}\alpha _{I_{j}}+\alpha _{j})(\nu +1)^{1/2} \in \mathbb R\), the non-centrality \(\kappa ^{\circ }_{j}=-\{1+Q_{\bar {\Omega }_{I_{j}I_{j}\cdot j}}(\alpha _{I_{j}})\}^{-1/2}\tau \in \mathbb R\)\(Q_{\bar {\Omega }_{I_{j}I_{j}\cdot j}}(\alpha _{I_{j}}) = \alpha _{I_{j}}^{\top } \bar {\Omega }_{I_{j}I_{j}\cdot j} \alpha _{I_{j}}\), and the degrees of freedom ν + 1.

1.3 A.3 Proof of proposition 1

Lemma 1

The finite d-dimensional distribution of the random process \((W(s) / W(s_{0}))_{s \in \mathcal {S}}\) under the transformed probability measure \(\widehat {\Pr } = \{ W(s_{0})\}_{+}^{\nu } / m_{+}(s_{0}) \mathrm {d}\Pr \) is equal to the distribution of a non-central extended skew-t process with mean μd, d × d scale matrix \(\hat {\Sigma }_{d}\), slant vector \(\hat {\alpha }_{d}\), extension τd, non-centrality κd = −τ and νd = ν + 1 degrees of freedom, given by

$$ \begin{array}{@{}rcl@{}} \mu_{d} &=& {\Sigma}_{d;0}, \quad \hat{\Sigma}_{d} = \frac{{\Sigma}_{d} - {\Sigma}_{d;0} {\Sigma}_{0;d} }{\nu +1 }, \quad \hat{\alpha}_{d} = \sqrt{\nu +1} \hat{\omega} \omega_{d}^{-1} \alpha,\\ \tau_{d} &=& (\alpha_{0} + {\Sigma}_{0;d} \omega_{d}^{-1} \alpha ) \sqrt{\nu + 1}, \end{array} $$

and where Σd = (K(xi, xj))1≤i, j, ≤d, \({\Sigma }_{d;0} = {\Sigma }_{0;d}^{\top } = (K(x_{0}, x_{i})) {1 \leq i \leq d}\), α = (α1, … , αd), ωd = diag(Σd)1/2 and \(\hat {\omega } = \text {diag}(\hat {\Sigma }_{d})^{1/2}\).

Proof of Lemma 1

The proof runs along the same lines as the proof of Lemma 2 in the supplementary material of Dombry et al. (2016). We consider finite dimensional distributions only. Let d ≥ 1 and \(s_{1}, \ldots , s_{d} \in \mathcal {S}\). We assume that the covariance matrix \(\tilde {\Sigma } = (K(s_{i}, s_{j}))_{0 \leq i,j \leq d}\) is non singular so that (W(si))0≤id has density

$$ \begin{array}{@{}rcl@{}} \tilde{g}(y) &=& (2 \pi)^{-(d+1)/2} \det (\tilde{\Sigma})^{-1/2} \exp\left\{ -\frac{1}{2} y^{\top} \tilde{\Sigma}^{-1} y \right\} \frac{\Phi(\tilde{\alpha}^{\top} \tilde{\omega}^{-1} y + \tau)}{\Phi(\tau/ \sqrt{1 + Q_{\tilde{\bar{\Sigma}}^{-1}}(\tilde{\alpha})}) },\\ &&y \in \mathbb {R}^{d+1}, \end{array} $$

where \(\tilde {\alpha } = (\alpha _{0}, \alpha _{1}, \ldots , \alpha _{d} )\), \(Q_{\tilde {\bar {\Sigma }}^{-1}}(\tilde {\alpha }) = \tilde {\alpha }^{\top } \tilde {\bar {\Sigma }} \tilde {\alpha }\), \(\tilde {\bar {\Sigma }} = \tilde {\omega }^{-1} \tilde {\Sigma } \tilde {\omega }^{-1}\) and \(\tilde {\omega } = \text {diag}(\tilde {\Sigma })^{1/2}\). Setting z = (yi/y0)1≤id, for all Borel sets \(A_{1}, \ldots , A_{d} \subset \mathbb {R}\),

$$ \begin{array}{@{}rcl@{}} \widehat{\Pr} \left\{ \frac{W(s_{i})}{W(s_{0})} \in A_{i}, i=1, \ldots, d \right\}&=& {\int}_{\mathbb {R}^{d+1}} \mathbb{I} \{ y_{i}/y_{0} \in A_{i}, i=1, \ldots, d \} \frac{(y_{0})_{+}^{\nu} \tilde{g}(y)}{m_{0+}} \mathrm{d} y \\ &=&{\int}_{\mathbb {R}^{d}} \mathbb{I} \{ z_{i} \in A_{i}, i=1, \ldots, d \} \left\{ {\int}_{0}^{\infty} \frac{(y_{0})_{+}^{\nu} \tilde{g}((y_{0}, y_{0} z)^{\top})}{m_{0+}} {y_{0}^{d}} \mathrm{d} y_{0} \right\} \mathrm{d} z. \end{array} $$

We deduce that under \(\widehat {\Pr }\) the random vector (W(si)/W(s0))1≤id has density

$$ \begin{array}{@{}rcl@{}} g(z)&=& {\int}_{0}^{\infty} \frac{y_{0}^{d+\nu}}{m_{0+}} \tilde{g}((y_{0}, y_{0} z)^{\top}) \mathrm{d} y_{0} \\ &=& \frac{(2\pi)^{-(d+1)/2} \det(\tilde{\Sigma})^{-1/2}} {m_{0+} {\Phi}(\tau/ \sqrt{1 + Q_{\tilde{\bar{\Sigma}}^{-1}}(\tilde{\alpha})})} {\int}_{0}^{\infty} y_{0}^{d+\nu} \exp\left\{ -\frac{1}{2} \tilde{z}^{\top} \tilde{\Sigma}^{-1} \tilde{z} {y_{0}^{2}} \right\} {\Phi}(\tilde{\alpha}^{\top} \tilde{\omega}^{-1} (y_{0}, y_{0} z)^{\top} + \tau) \mathrm{d} y_{0}, \end{array} $$

where \(\tilde {z} = (1, z)^{\top }\). Through the change of variable \(u = \left (\tilde {z}^{\top } \tilde {\Sigma }^{-1} \tilde {z} \right )^{1/2} y_{0}\), we obtain

$$ \begin{array}{@{}rcl@{}} {\int}_{0}^{\infty} y_{0}^{d+\nu} &&\exp \left\{ -\frac{1}{2} \tilde{z}^{\top} \tilde{\Sigma}^{-1} \tilde{z} {y_{0}^{2}} \right\} {\Phi}(\tilde{\alpha}^{\top} \tilde{\omega}^{-1} (y_{0}, y_{0} z) + \tau) \mathrm{d} y_{0} \\ &=& \sqrt{2\pi} \left( \tilde{z} \tilde{\Sigma}^{-1} \tilde{z} \right)^{-\frac{d + \nu +1}{2}} {\int}_{0}^{\infty} u^{d + \nu} \phi(u) {\Phi} \left( \frac{\tilde{\alpha}^{\top} \tilde{\omega}^{-1} \tilde{z} } { \sqrt{ \tilde{z} \tilde{\Sigma}^{-1} \tilde{z} } } u + \tau \right) \mathrm{d} u \\ &=& \left( \tilde{z} \tilde{\Sigma}^{-1} \tilde{z} \right)^{-\frac{d_{\nu}}{2}} 2^{\frac{\nu - 3}{2}} \pi^{-\frac{d+2}{2}} {\Gamma}\left( \frac{d_{\nu}}{2}\right) {\Psi} \left( \frac{ \tilde{\alpha}^{\top} \tilde{\omega}^{-1} \tilde{z} } {\sqrt{ \tilde{z} \tilde{\Sigma}^{-1} \tilde{z} } } \sqrt{d_{\nu}}; -\tau, d_{\nu} \right), \end{array} $$

where α = (α1,…, αd), dν = d + ν + 1 and Ψ(⋅; κ, ν) is the cdf of the non-central t distribution with non-centrality parameter κ and ν degrees of freedom. Thus applying the definition of m0+, \({\alpha }^{\ast }_{0}\) and \({\tau }^{\ast }_{0}\) from Beranger et al. (2017) we get

$$ g(z) = \frac{\pi^{-d/2} \det(\tilde{\Sigma})^{-1/2} } { {\Psi}(\alpha^{\ast}_{0} \sqrt{\nu +1}; -\tau^{\ast}_{0}, \nu +1) } \left( \tilde{z} \tilde{\Sigma}^{-1} \tilde{z} \right)^{-\frac{d_{\nu}}{2}} \frac{{\Gamma}\left( \frac{d_{\nu}}{2}\right)}{{\Gamma}\left( \frac{\nu + 1}{2}\right)} {\Psi} \left( \frac{\tilde{\alpha}^{\top} \tilde{\omega}^{-1} \tilde{z}} {\sqrt{\tilde{z} \tilde{\Sigma}^{-1} \tilde{z}} } \sqrt{d_{\nu}}; -\tau, d_{\nu} \right), $$

and the block decomposition \(\tilde {\Sigma } = \left (\begin {array}{cc} 1 & {\Sigma }_{0;d} \\ {\Sigma }_{d;0} & {\Sigma }_{d} \end {array} \right )\), allows us to write

$$ g(z) = \frac{\psi_{d}(z; \mu_{d}, \hat{\Sigma}_{d}, \nu + 1 ) {\Psi} \left( \frac{\alpha_{0} + \alpha^{\top} \omega_{d}^{-1} z} {\sqrt{ \tilde{z} \tilde{\Sigma}^{-1} \tilde{z} } } \sqrt{d_{\nu}} ; -\tau, d_{\nu} \right)} { {\Psi}(\alpha_{0}^{\ast} \sqrt{\nu +1}; -\tau_{0}^{\ast}, \nu +1)}, $$

where μd = Σd;0, \(\hat {\Sigma }_{d} = ({\Sigma }_{d} - {\Sigma }_{d;0} {\Sigma }_{0;d}) / (\nu + 1)\) and ωd = diag(Σd)1/2. Noting that

$$ \begin{array}{@{}rcl@{}} {\Psi}(\alpha^{\ast}_{0} \sqrt{\nu +1}; -\tau^{\ast}_{0}, \nu +1) &=& {\Psi} \left( \frac{\tau_{d}} {\sqrt{1 + \hat{\alpha}_{d}^{\top} \hat{\bar{\Sigma}}_{d} \hat{\alpha}_{d} }}; \frac{-\tau}{\sqrt{1 + \hat{\alpha}_{d}^{\top} \hat{\bar{\Sigma}}_{d} \hat{\alpha}_{d} }}; \nu + 1 \right) \\ {\Psi} \left( \frac{\alpha_{0} + \alpha^{\top} \omega_{d}^{-1} z } {\sqrt{ \tilde{z} \tilde{\Sigma}^{-1} \tilde{z} } } \sqrt{d_{\nu}} ; -\tau, d_{\nu} \right) &=& {\Psi} \left( \frac{ \hat{\alpha}_{d} z^{\prime} + \tau_{d} } {\sqrt{\nu + 1 + z^{{\prime}{\top}} \hat{\bar{\Sigma}}_{d}^{-1} z^{\prime} }} \sqrt{d_{\nu} }; - \tau, d_{\nu} \right), \end{array} $$

where Ψ(⋅; κ, ν) denotes the cdf of the univariate non-central t distribution with non-centrality κ and ν degrees of freedom, \(z^{\prime } = \hat {\omega }^{-1} (z - {\Sigma }_{d;0})\), \(\hat {\alpha }_{d} = \sqrt {\nu + 1} \hat {\omega } \omega _{d}^{-1} \alpha \), \(\tau _{d} = (\alpha _{0} + \alpha ^{\top } {\Sigma }_{d;0}) \sqrt {\nu + 1 }\), \(\hat {\bar {\Sigma }}_{d} = \hat {\omega }^{-1} \hat {\Sigma } \hat {\omega }^{-1}\), \(\hat {\omega } = \text {diag}(\hat {\Sigma })^{1/2}\) which leads us to the conclusion that g(z) is the density of a non-central extended skew-t distribution with parameters μd = Σd;0, \(\hat {\Sigma }_{d}\), \(\hat {\alpha }_{d}\), τd and κd = −τ. □

In order to prove Proposition 1, let \(\mathcal {C}_{+} = \mathcal {C} \{\mathcal {S}, [0, \infty ]\}\) denote the space of continuous non-negative functions on \(\mathcal {S}\), and σ represent the distribution of the \(\{W(s_{i})\}_{+}^{\nu } / m_{+}\), and consider the set \( A = \{ f \in \mathcal {C}_{0} : f(s_{1}) \in A_{1}, \ldots , f(s_{d}) \in A_{d} \}. \) Then by (Dombry and Eyi-Minko 2013, Proposition 4.2) we have

$$ \begin{array}{@{}rcl@{}} P_{s_{0}}(A)&=&{\int}_{\mathcal{C}} \mathbb{I} \{ f/f(s) \in A \} f(s) \sigma(\mathrm{d} f) \\ &=& \mathbb{E} \left[\{W(s_{0})\}_{+}^{\nu} / m_{0+} \mathbb{I} \left\{ \frac{m_{0+} \{W(s_{i})\}_{+}^{\nu}}{m_{i+} \{ W(s_{0})\}_{+}^{\nu}} \in A_{i}; i = 1, \ldots, d\right\} \right] \\ &=& \widehat{\Pr} \left\{ \frac{m_{0+} \{W(s_{i})\}_{+}^{\nu}}{m_{i+} \{ W(s_{0})\}_{+}^{\nu}} \in A_{i}; i = 1, \ldots, d \right\} \\ &=& \Pr \left\{ \frac{m_{0+}}{m_{i+}} (T_{i})_{+}^{\nu} \in A_{i}; i = 1, \ldots, d \right\}, \end{array} $$

where T = (T1,…, Td), Ti = W(si)/W(s0) which, from Lemma 1, is distributed as

$$ T \sim {\Psi}_{d} \left( {\Sigma}_{d;0}, \frac{{\Sigma}_{d} - {\Sigma}_{d;0} {\Sigma}_{0;d} }{\nu +1 }, \hat{\alpha}_{d}, \tau_{d}, - \tau, \nu + 1 \right). $$

1.4 A.4 Proof of Proposition 2

The following Lemma is required in order to complete the proof.

Lemma 2

Under the assumptions of Proposition 2, the intensity function of the extremal skew-t is

$$ \lambda_{\boldsymbol{s}} (\boldsymbol{v}) = \frac{ 2^{(\nu-2)/2} \nu^{-d+1} {\Gamma} \left( \frac{d+\nu}{2} \right) {\Psi} \left( \tilde{\alpha}_{\boldsymbol{s}} \sqrt{d+\nu}; -\tau_{\boldsymbol{s}}, d+\nu \right) {\prod}_{i=1}^{d} \left( m_{i+} v_{i}^{1-\nu} \right)^{1/\nu} } { \pi^{d/2} |\bar{\Omega}_{\boldsymbol{s}}|^{1/2} Q_{\bar{\Omega}_{\boldsymbol{s}}}(\boldsymbol{v}^{\circ})^{(d+\nu)/2} {\Phi}(\tau_{\boldsymbol{s}} \{ 1 + Q_{\bar{\Omega}^{-1}_{\boldsymbol{s}}}(\alpha_{\boldsymbol{s}}) \}^{-1/2} ) }, $$

where \(\boldsymbol {v}^{\circ } = (\boldsymbol {v} m_{+}(\boldsymbol {s}))^{1/\nu } \in \mathbb {R}^{d}\), \(\tilde {\alpha }_{\boldsymbol {s}} = \alpha _{\boldsymbol {s}}^{\top } \boldsymbol {v}^{\circ } Q_{\bar {\Omega }_{\boldsymbol {s}}}(\boldsymbol {v}^{\circ })^{-1/2} \in \mathbb {R}\) and \(\alpha _{\boldsymbol {s}} \in \mathbb {R}^{d}\).

Proof

By definition of the intensity measure (5), for all \(\boldsymbol {s} \in \mathcal {S}^{d}\) and Borel set \(A \subset \mathbb {R}^{d}\),

$$ {\Lambda}_{\boldsymbol{s}}(A) = {\int}_{0}^{\infty} \int\limits_{\mathbb {R}^{d}} \mathbb{I} \left\{ \zeta \boldsymbol{t}^{\nu} / m_{+}(\boldsymbol{s}) \in A \right\} g_{\boldsymbol{s}}(\boldsymbol{t}) \mathrm{d} \boldsymbol{t} \zeta^{-2} \mathrm{d} \zeta, $$

where gs is the density of the random vector W(s), i.e. a centred extended skew normal random vector with correlation matrix \(\bar {\Omega }_{\boldsymbol {s}}\), slant αs and extension τs. The change of variable \(\boldsymbol {v} = (m_{+}(\boldsymbol {s}))^{-1} \zeta \boldsymbol {t}^{\nu }\) leads to \(\mathrm {d} \boldsymbol {t} = \nu ^{-d} \zeta ^{-d/\nu } {\prod }_{i=1}^{d} m_{i+}^{1/\nu } v_{i}^{(1-\nu )/\nu } \mathrm {d} \boldsymbol {v}\) and

$$ {\Lambda}_{\boldsymbol{s}}(A) = \nu^{-d} {\int}_{0}^{\infty} {\int}_{A} \prod\limits_{i=1}^{d} \left( m_{i+} {v}_{i}^{(1-\nu)}\right)^{1/\nu} \frac{ \phi_{d} \left( \boldsymbol{v}^{\circ} \zeta^{-1/\nu}; \bar{\Omega}_{\boldsymbol{s}} \right) {\Phi} \left( \alpha_{\boldsymbol{s}}^{\top} \boldsymbol{v}^{\circ} \zeta^{-1/\nu} + \tau_{\boldsymbol{s}} \right)} { {\Phi}\left( \tau_{\boldsymbol{s}} \{ 1 + Q_{\bar{\Omega}^{-1}_{\boldsymbol{s}}}(\alpha_{\boldsymbol{s}}) \}^{-1/2} \right) } \zeta^{-\frac{d}{\nu}-2} \mathrm{d} \boldsymbol{v} \mathrm{d} \zeta, $$
(9)

where v = (vm+(s))1/ν. Now, through the consecutive change of variables t = ζ− 1/ν and \(u = t Q_{\bar {\Omega }_{\boldsymbol {s}}}(\boldsymbol {v}^{\circ })^{1/2}\) we obtain

$$ \begin{array}{@{}rcl@{}} {\int}_{0}^{\infty} \phi_{d} &&\left( \boldsymbol{v}^{\circ} \zeta^{-1/\nu}; \bar{\Omega}_{\boldsymbol{s}} \right) {\Phi} \left( \alpha_{\boldsymbol{s}}^{\top} \boldsymbol{v}^{\circ} \zeta^{-1/\nu} + \tau_{\boldsymbol{s}} \right) \zeta^{-\frac{d}{\nu}-2} \mathrm{d}\zeta \\ \quad &=& \nu {\int}_{0}^{\infty} \phi_{d} \left( \boldsymbol{v}^{\circ} t ; \bar{\Omega}_{\boldsymbol{s}} \right) {\Phi} \left( \alpha_{\boldsymbol{s}}^{\top} \boldsymbol{v}^{\circ} t + \tau_{\boldsymbol{s}} \right) t^{d+\nu-1} \mathrm{d} t \\ \quad &=& (2\pi)^{-d/2} |\bar{\Omega}_{\boldsymbol{s}}|^{-1/2} \nu {\int}_{0}^{\infty} t^{d+\nu-1} \exp \left\{ -t^{2} \frac{ Q_{\bar{\Omega}_{\boldsymbol{s}}}(\boldsymbol{v}^{\circ}) }{2} \right\} {\Phi} \left( \alpha_{\boldsymbol{s}}^{\top} \boldsymbol{v}^{\circ} t + \tau_{\boldsymbol{s}} \right) \mathrm{d} t \\ \quad &=& (2\pi)^{-d/2} |\bar{\Omega}_{\boldsymbol{s}} |^{-1/2} \nu Q_{\bar{\Omega}_{\boldsymbol{s}}}(\boldsymbol{v}^{\circ})^{-(d+\nu)/2} (2\pi)^{1/2} {\int}_{0}^{\infty} u^{d+\nu-1} \phi(u) {\Phi} \left( \tilde{\alpha}_{{s}} u + \tau_{{s}} \right) \mathrm{d} u , \end{array} $$
(10)

where \(\tilde {\alpha }_{\boldsymbol {s}} = \alpha _{\boldsymbol {s}}^{\top } \boldsymbol {v}^{\circ } Q_{\bar {\Omega }_{\boldsymbol {s}}}(\boldsymbol {v}^{\circ })^{-1/2} \in \mathbb R\).

The remaining integral is linked to the moments of the extended skew-normal distribution. Beranger et al. (2017, Appendix A.4) derives the result

$$ {\int}_{0}^{\infty} y^{\nu} \phi(y) {\Phi}(\alpha y + \tau) \mathrm{d} y = 2^{(\nu-2)/2} \pi^{-1/2} {\Gamma} \left( \frac{\nu+1}{2} \right) {\Psi} \left( \alpha \sqrt{\nu+1}; -\tau, \nu+1 \right), $$

and thus Eq. 10 is equal to

$$ 2^{(\nu-2)/2} \pi^{-d/2} |\bar{\Omega}|^{-1/2} \nu \left( \boldsymbol{v}^{\circ\top} \bar{\Omega}_{\boldsymbol{s}}^{-1} \boldsymbol{v}^{\circ}\right)^{-(d+1)/2} {\Gamma} \left( \frac{\nu + d}{2} \right) {\Psi} \left( \tilde{\alpha}_{\boldsymbol{s}} \sqrt{\nu+d}; -\tau_{\boldsymbol{s}}, \nu+d \right). $$
(11)

Substituting (11) into Eq. 9 completes the proof. □

Assume that \((W(\boldsymbol {t}), W(\boldsymbol {s})) \sim \mathcal {S}\mathcal {N}_{m+d}(\bar {\Omega }_{(\boldsymbol {t},\boldsymbol {s})}, \alpha _{(\boldsymbol {t},\boldsymbol {s})}, \tau _{(\boldsymbol {t},\boldsymbol {s})})\), then according to Beranger et al. (2017, Proposition 1) we have that \(W(\boldsymbol {s}) \sim \mathcal {S}\mathcal {N}_{d} (\bar {\Omega }_{\boldsymbol {s}}, {\alpha }^{\ast }_{\boldsymbol {s}}, {\tau }^{\ast }_{\boldsymbol {s}})\) with

$$ \begin{array}{ccc} {\alpha}^{\ast}_{\boldsymbol{s}} = \frac{\alpha_{\boldsymbol{s}} + \bar{\Omega}_{\boldsymbol{s}}^{-1} \bar{\Omega}_{\boldsymbol{s}\boldsymbol{t}} \alpha_{\boldsymbol{t}}} {\sqrt{1 + Q_{\tilde{\Omega}^{-1}}(\alpha_{\boldsymbol{t}})}}, & {\tau}^{\ast}_{\boldsymbol{s}} = \frac{\tau_{(\boldsymbol{t},\boldsymbol{s})}}{\sqrt{1 + Q_{\tilde{\Omega}^{-1}}(\alpha_{\boldsymbol{t}})}}, & \tilde{\Omega} = \bar{\Omega}_{\boldsymbol{t}} - \bar{\Omega}_{\boldsymbol{t} \boldsymbol{s}} \bar{\Omega}_{\boldsymbol{s}}^{-1} \bar{\Omega}_{\boldsymbol{s} \boldsymbol{t}}. \end{array} $$

Additionally let u = (um+(t))1/ν, v = (vm+(s))1/ν, m+(t) = (m+(t1),…, m+(tm)), m+(s) = (m+(s1),…, m+(sd)), \(\boldsymbol {u} \in \mathbb {R}^{m}\), \(\boldsymbol {v} \in \mathbb {R}^{d}\). Noting that \({\Phi } (\tau _{(\boldsymbol {t},\boldsymbol {s})} (1+Q_{\bar {\Omega }^{-1}_{(\boldsymbol {t},\boldsymbol {s})}}(\alpha _{(\boldsymbol {t},\boldsymbol {s})}))^{-1/2})\) is equal to \({\Phi } (\tau ^{\ast }_{\boldsymbol {s}} (1+Q_{\bar {\Omega }^{-1}_{\boldsymbol {s}}}(\alpha ^{\ast }_{\boldsymbol {s}}))^{-1/2})\) and applying Lemma 2 to Eq. 6 leads to

$$ \begin{array}{@{}rcl@{}} \lambda_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}(\boldsymbol{u}){} \!&=&\!{} \pi^{-m/2} \nu^{-m} \frac{|\bar{\Omega}_{(\boldsymbol{t},\boldsymbol{s})}|^{-1/2}}{|\bar{\Omega}_{\boldsymbol{s}}|^{-1/2}} \left\{{} \frac{Q_{\bar{\Omega}_{(\boldsymbol{t},\boldsymbol{s})}} (\boldsymbol{u}^{\circ}, \boldsymbol{v}^{\circ}) } {Q_{\bar{\Omega}_{\boldsymbol{s}}} (\boldsymbol{v}^{\circ})} \right\}^{-\frac{\nu+d+m}{2}} Q_{\bar{\Omega}_{\boldsymbol{s}}} (\boldsymbol{v}^{\circ})^{-\frac{m}{2}} \frac{{\Gamma}(\frac{\nu+d+m}{2})}{{\Gamma}(\frac{\nu+d}{2})} \\ \quad &&\times \frac{ {\Psi}\left( \tilde{\alpha}_{(\boldsymbol{t}, \boldsymbol{s})} \sqrt{\nu + d + m}; -\tau_{(\boldsymbol{t}, \boldsymbol{s})}, \nu + d + m \right) } { {\Psi} \left( \tilde{\alpha}_{\boldsymbol{s}} \sqrt{\nu + d}; -\tau^{\ast}_{\boldsymbol{s}}, \nu + d \right) } \prod\limits_{i=1}^{m} \left( m_{+}(t_{i}) u_{i}^{1-\nu} \right)^{1/\nu}, \end{array} $$

where \( \tilde {\alpha }_{(\boldsymbol {t}, \boldsymbol {s})} = \alpha ^{\top }_{(\boldsymbol {t}, \boldsymbol {s})} (\boldsymbol {u}^{\circ }, \boldsymbol {v}^{\circ }) Q_{\bar {\Omega }_{(\boldsymbol {t},\boldsymbol {s})}} (\boldsymbol {u}^{\circ }, \boldsymbol {v}^{\circ })^{-1/2} \) and \( \tilde {\alpha }_{\boldsymbol {s}} = {\alpha }^{\ast \top }_{\boldsymbol {s}} \boldsymbol {v}^{\circ } Q_{\bar {\Omega }_{\boldsymbol {s}}} (\boldsymbol {v}^{\circ })^{-1/2}. \)

Following Dombry et al. (2013) and Ribatet (2013) we can show that

$$ \frac{\lvert\bar{\Omega}_{(\boldsymbol{t},\boldsymbol{s})}\rvert}{\lvert\bar{\Omega}_{\boldsymbol{s}}\rvert} {}= {}\left\{ {}\frac{\nu + d}{ Q_{\bar{\Omega}} (\boldsymbol{v}^{\circ}) }{} \right\}^{m}{} \lvert {\Omega}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}} \rvert, {} \frac{Q_{\bar{\Omega}_{(\boldsymbol{t},\boldsymbol{s})}} (\boldsymbol{u}^{\circ}, \boldsymbol{v}^{\circ}) } {Q_{\bar{\Omega}_{\boldsymbol{s}}} (\boldsymbol{v}^{\circ})} {}={} 1 + \frac{Q_{{\Omega}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}} (\boldsymbol{u}^{\circ} - \mu_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}) } {\nu + d}, {\Omega}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}} = \frac{ Q_{\bar{\Omega}_{\boldsymbol{s}}} (\boldsymbol{v}^{\circ}) }{\nu + d} \tilde{\Omega}, $$

and \( \mu _{\boldsymbol {t} | \boldsymbol {s}, \boldsymbol {v}} = \bar {\Omega }_{\boldsymbol {t}\boldsymbol {s}} \bar {\Omega }_{\boldsymbol {s}}^{-1} \boldsymbol {v}^{\circ }. \) Thus we have

$$ \begin{array}{@{}rcl@{}} \lambda_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}(\boldsymbol{u}) &= \pi^{-m/2} (\nu + d)^{-m/2} \lvert{\Omega}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}\rvert^{-1/2} \left\{ 1 + \frac{Q_{{\Omega}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}} (\boldsymbol{u}^{\circ} - \mu_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}) } {\nu + d} \right\}^{-\frac{\nu + d + m}{2}} \\ & \quad \times \frac{{\Gamma}(\frac{\nu+d+m}{2})}{{\Gamma}(\frac{\nu+d}{2})} \frac{{\Psi} \left( \tilde{\alpha}_{(\boldsymbol{t}, {s})} \sqrt{\nu + d + m}; -\tau_{(\boldsymbol{t}, {s})}, \nu + d + m \right)} {{\Psi} \left( \tilde{\alpha}_{\boldsymbol{s}} \sqrt{\nu + d}; -\tau^{\ast}_{\boldsymbol{s}}, \nu + d \right)} \nu^{-m} \prod\limits_{i=1}^{m} \left( m_{+}(t_{i}) u_{i}^{1-\nu} \right)^{1/\nu}, \end{array} $$

and we recognise the m-dimensional Student-t density with mean μt|s, v, dispersion matrix Ωt|s, v and degree of freedom ν + d.

Finally, by considering \( \alpha _{\boldsymbol {t} | \boldsymbol {s}, \boldsymbol {v}} = \tilde {\omega }^{-1} \alpha _{\boldsymbol {t}} \), \( \tilde {\omega } = \text {diag}(\tilde {\Omega })^{1/2}, \)\( \tau _{\boldsymbol {t} | \boldsymbol {s}, \boldsymbol {v}} = \left (\alpha _{\boldsymbol {s}} + \bar {\Omega }_{\boldsymbol {s}}^{-1} \bar {\Omega }_{\boldsymbol {s} \boldsymbol {t}} \alpha _{\boldsymbol {t}} \right )^{\top } \boldsymbol {v}^{\circ } (d+\nu )^{1/2} Q_{\bar {\Omega }} (\boldsymbol {v}^{\circ })^{-1/2} \) and κt|s, v = −τ(t, s) then it is easy to show that

$$ \begin{array}{@{}rcl@{}} {\Psi} \left( \tilde{\alpha}_{\boldsymbol{s}} \sqrt{\nu + d}; -\tau^{\ast}_{\boldsymbol{s}}, \nu + d \right) {}&=& {}{\Psi} \left( \frac{\tau_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}}{\sqrt{1+Q_{\bar{\Omega}^{-1}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}}(\alpha_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}})}}; -\frac{\tau_{(\boldsymbol{t},\boldsymbol{s})}}{\sqrt{1+Q_{\bar{\Omega}^{-1}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}}(\alpha_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}})}}, \nu_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}} \right) \\ &&{\Psi} \left( \tilde{\alpha}_{(\boldsymbol{t}, \boldsymbol{s})} \sqrt{\nu + d + m}; -\tau_{(\boldsymbol{t}, \boldsymbol{s})}, \right. \left. \nu + d + m \right)\\ &=& {}{\Psi} \left( {} \left( {}\alpha_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}^{\top} z + \tau_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}} \right) \sqrt{\frac{\nu_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}} + m}{\nu_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}} + Q_{\bar{\Omega}_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}}(z)}}; -\tau_{(\boldsymbol{t},\boldsymbol{s})}, \nu_{\boldsymbol{t} | \boldsymbol{s}, \boldsymbol{v}}+m \right) , \end{array} $$

where z = uμt|s, v, completes the proof. Note that \(\bar {\Omega }_{\boldsymbol {t} | \boldsymbol {s}, \boldsymbol {v}}\) reduces to \(\tilde {\omega }^{-1} \tilde {\Omega } \tilde {\omega }^{-1}\).

1.5 A.5 Partial derivatives of the V function of the extremal skew-t (Lemma 3)

Consider the conditional intensity function of the extremal skew-t model given in Proposition 2 with s = (s1,…, sm) ≡s1:m, t = (sm+ 1,…, sd) ≡sm+ 1:d, v = z1:m and u = zm+ 1:d. In the following the matrix notation \({\Sigma }_{a:b}= {\Sigma }_{\boldsymbol {s}_{a:b}}, {\Sigma }_{a:b;c:d}= {\Sigma }_{\boldsymbol {s}_{a:b} \boldsymbol {s}_{c:d}}\) will be used. Integration w.r.t. zm+ 1:d gives

$$ {\Psi}_{d-m} \left( \boldsymbol{z}_{m+1:d}^{\circ}; \mu_{c}, {\Omega}_{c}, \alpha_{c}, \tau_{c}, \kappa_{c}, \nu_{c} \right) $$
(12)

where the index c represents sm+ 1:d|s1:m, z1:m such that the parameters are

$$ \begin{array}{@{}rcl@{}} \mu_{c} &=& \bar{\Omega}_{m+1:d ; 1:m} \bar{\Omega}_{1:m}^{-1} \boldsymbol{z}_{1:m}^{\circ}, \quad {\Omega}_{c} = \frac{Q_{\bar{\Omega}_{1:m}} (\boldsymbol{z}_{1:m}^{\circ})}{\nu_{c}} \tilde{\Omega}_{c}, \\ \tilde{\Omega}_{c} &=& \bar{\Omega}_{m+1:d} - \bar{\Omega}_{m+1:d; 1:m} \bar{\Omega}_{1:m}^{-1} \bar{\Omega}_{1:m ; m+1:d}, \\ \tau_{c} &=& (\alpha_{1:m} + \bar{\Omega}_{1:m}^{-1} \bar{\Omega}_{1:m ; m+1:d} \alpha_{m+1:d} )^{\top} \boldsymbol{z}_{1:m}^{\circ} (\nu+m)^{1/2} Q_{\bar{\Omega}_{1:m}} (\boldsymbol{z}_{1:m}^{\circ})^{-1/2}, \\ \alpha_{c} &=& \omega_{c} \alpha_{m+1:d}, \quad \tilde{\omega}_{c} = \text{diag}(\tilde{\Omega}_{c})^{1/2}, \quad \kappa_{c} = -\tau, \quad \nu_{c} = \nu + m, \end{array} $$

with \(\boldsymbol {z}_{1:m}^{\circ } = (\boldsymbol {z}_{1:m} m_{+}(\boldsymbol {s}_{1:m}))^{1/\nu }\) and \(\boldsymbol {z}_{m+1:d}^{\circ } = (\boldsymbol {z}_{m+1:d} m_{+}(\boldsymbol {s}_{m+1:d}))^{1/\nu }\). According to Lemma 2, the m-dimensional marginal density is

$$ \frac{ 2^{(\nu-2)/2} \nu^{-m+1} {\Gamma} \left( \frac{m+\nu}{2} \right) {\Psi} \left( \tilde{\alpha}_{1:m} \sqrt{m+\nu}; - \tau_{1:m}^{\ast}, m+\nu \right) {\prod}_{i=1}^{m} \left( m_{i+} z_{i}^{1-\nu} \right)^{1/\nu} } { \pi^{m/2} |\bar{\Omega}_{1:m}|^{1/2} Q_{\bar{\Omega}_{1:m}} (\boldsymbol{z}_{1:m}^{\circ} )^{(m+\nu)/2} {\Phi}(\tau_{1:m}^{\ast} \{ 1 + Q_{\bar{\Omega}^{-1}_{1:m}}(\alpha^{\ast}_{1:m}) \}^{-1/2} ) }, $$
(13)

where \(\tilde {\alpha }_{1:m} = \alpha _{1:m}^{\ast \top } \boldsymbol {z}_{1:m}^{\circ } Q_{\bar {\Omega }_{1:m}} (\boldsymbol {z}_{1:m}^{\circ } )^{-1/2} \in \mathbb {R}\), and m-dimensional marginal parameters

$$ \begin{array}{ll} \alpha_{1:m}^{\ast} = \frac{\alpha_{1:m} + \bar{\Omega}_{1:m}^{-1} \bar{\Omega}_{1:m; m+1:d} \alpha_{m+1:d} } {\sqrt{1+ Q_{\bar{\Omega}^{-1}_{c}} (\alpha_{m+1:d})}}, & \tau_{1:m}^{\ast} = \frac{\tau} {\sqrt{1+ Q_{\bar{\Omega}^{-1}_{c}}(\alpha_{m+1:d})}}. \end{array} $$

Combining (12) and Eq. 13 completes the proof.

Setting τs = 0 corresponds to an extremal skew-t model constructed from a skew-normal random field rather than an extended skew-normal field. Then

$$ m_{j+} = 2^{\nu/2} \pi^{-1/2} {\Gamma}\left( \frac{\nu + 1}{2}\right) {\Psi}(\alpha^{\ast}_{j} \sqrt{\nu +1}; \nu +1), $$

with

$$ \alpha^{\ast}_{j} = \frac{\alpha_{j} + \bar{\Omega}^{-1}_{jj} \bar{\Omega}_{j I_{j}} \alpha_{I_{j}} } {\sqrt{1 + \alpha_{I_{j}}^{\top} \left( \bar{\Omega}_{I_{j} I_{j}} - \bar{\Omega}_{I_{j} j} \bar{\Omega}^{-1}_{jj} \bar{\Omega}_{j I_{j}} \right) \alpha_{I_{j}}} }, $$

and the associated partial derivatives of the V function are equal to

$$ {\Psi}_{d-m} \left( \boldsymbol{z}_{m+1:d}^{\circ}; \mu_{c}, {\Omega}_{c}, \alpha_{c}, \tau_{c}, 0, \nu_{c} \right) \frac{ 2^{\nu/2} {\Gamma} \left( \frac{m+\nu}{2} \right) {\Psi} \left( \tilde{\alpha}_{1:m} \sqrt{m+\nu}; m+\nu \right) {\prod}_{i=1}^{m} \left( m_{i+} z_{i}^{1-\nu} \right)^{1/\nu} } { \pi^{m/2} \nu^{m-1} |\bar{\Omega}_{1:m}|^{1/2} Q_{\bar{\Omega}_{1:m}} (\boldsymbol{z}_{1:m}^{\circ} )^{(m+\nu)/2} }, $$

with parameters defined as in Lemma 3.

Setting αs = α1:d = 0 and τs = 0 leads to the extremal-t model for which

$$ m_{i+} = 2^{(\nu-2)/2} \pi^{-1/2} {\Gamma}\left( \frac{\nu + 1}{2}\right) \equiv m_{+}, $$

and the partial derivatives of the V function are

$$ {\Psi}_{d-m} \left( \boldsymbol{z}_{m+1:d}^{1/\nu}; \mu^{\prime}_{c}, {\Omega}^{\prime}_{c}, \nu_{c} \right) \frac{ 2^{(\nu-2)/2} \nu^{-m+1} {\Gamma} \left( \frac{m+\nu}{2} \right) {\prod}_{i=1}^{m} z_{i}^{(1-\nu)/\nu} } { \pi^{m/2} |\bar{\Omega}_{1:m}|^{1/2} Q_{\bar{\Omega}_{1:m}} (\boldsymbol{z}_{1:m}^{1/ \nu} )^{(m+\nu)/2} m_{+} }, $$

where \(\mu ^{\prime }_{c} = \bar {\Omega }_{m+1:d ; 1:m} \bar {\Omega }_{1:m}^{-1} \boldsymbol {z}_{1:m}^{1/ \nu }\) and \({\Omega }^{\prime }_{c} = \frac {Q_{\bar {\Omega }_{1:m}} (\boldsymbol {z}_{1:m}^{1/\nu })}{\nu _{c}} \bar {\Omega }_{c}\).

Appendix B: Simulation Tables

Table 4 Absolute biases \(|\bar {\hat {\theta }}_j - \theta _j|\) for \(\hat {\theta }_j = (\hat {\eta }_j, \hat {r}_j)\) and \(\hat {\theta }_j = (\hat {\eta }_j, \hat {r}_j, \hat {\beta }_{1j}, \hat {\beta }_{2j})\) the parameter vectors of the extremal-t and extremal skew-t models, using the full likelihood Type I and Type II approximations given in Table 1 when d = 20,50 and 100 sites are considered. Calculations are based on 500 replicate maximisations
Table 5 RMSEs for \(\hat {\theta }_j = (\hat {\eta }_j, \hat {r}_j)\) and \(\hat {\theta }_j = (\hat {\eta }_j, \hat {r}_j, \hat {\beta }_{1j}, \hat {\beta }_{2j})\) the parameter vectors of the extremal-t and extremal skew-t models using the j-wise composite likelihood when v = 1 and d = 20. The case j = d corresponds to full likelihood estimation using approximation Type II from Table 1 . Calculations are based on 500 replicate maximisations
Table 6 Absolute biases \(|\bar {\hat {\theta }}_j - \theta _j|\) for \(\hat {\theta }_j = (\hat {\eta }_j, \hat {r}_j)\) and \(\hat {\theta }_j = (\hat {\eta }_j, \hat {r}_j, \hat {\beta }_{1j}, \hat {\beta }_{2j})\) the parameter vectors of the extremal-t and extremal skew-t models using the j-wise composite likelihood when v = 1 and d = 20. The case j = d corresponds to full likelihood estimation using approximation Type II from Table 1. Calculations are based on 500 replicate maximisations.

Appendix C: Exact simulation of Extremal skew-t Max Stable Process with Hitting Scenarios

Below we provide pseudo-code for exact simulation of extremal skew-t max stable processes with unit Fréchet marginal distributions using Algorithm 2 of Dombry et al. (2016), extended to include the hitting scenario in the output. This requires the simulation of an extended skew-t distribution; here we use rejection sampling and the stochastic representation given in Arellano-Valle and Genton (2010). The simpler extremal-t max stable process only requires the simulation of a multivariate t-distribution and therefore does not use rejection sampling; this simpler algorithm is also given below.

When simulating N independent replicates for d sites with the Dombry et al. (2016) algorithm, it is much more efficient to have the sites in the outer loop and the replicates in the inner loop, because derivations of quantities from the distribution of \((W(s) / W(s_{0}))_{s \in \mathcal {S}}\) are then only performed once for each site (lines 3 to 7 in the skew-t code), irrespective of the number of replicates required. In practice these quantities should be calculated on the log scale to avoid numerical issues.

In the algorithm below, the input Σd is derived from the correlation function ρ(h). The normalization in line 6 is needed for the simulation of an extended skew-t distribution. Matrix multiplication is not needed here because ω is a diagonal matrix. The term Exp(1) refers to a standard exponential distribution, \(t_{\nu _{d}}\) is a univariate t-distribution with νd degrees of freedom, N(0, 1) is a standard univariate normal distribution, and \(\chi ^{2}_{\nu _{d}}\) is a chi-squared distribution with νd degrees of freedom. The function Ψ(⋅; νd) is the distribution function of a univariate t-distribution, as used in Eq. (A.5) The code in line 16 simulates from a multivariate t-distribution with shape matrix \(\hat {\bar {\Sigma }}_{d} \) and νd degrees of freedom. Lines 20 and 24 are identical by intent.

The index j in the code corresponds to the s0 site. We recommend the use of the eigendecomposition, which is more stable than the Cholesky decomposition. Moreover, \(\bar {\Sigma }^{}_{d}\) is positive semi-definite as the j th row and columns are zero by construction. If the Cholesky decomposition were used then the code would need to handle the singular component explicitly. The eigendecomposition is slower, but it can be evaluated outside the loop over the observations (in line 7) and therefore only d decompositions are required for any N.

The do-while loop in line 12 is the rejection sampling needed to simulate from a multivariate extended skew-t distribution. The Dombry et al. (2016) algorithm also has a rejection step, with B = 0 in the code indicating rejection via exceeding an observation on an already simulated site (i.e. on a site with index less than j). If the simulation is not rejected (line 22) then the outputs are set. A simulated process will always update the value on the j th site, because there is a singular component X[j] ≡ 1 and therefore the code would otherwise not enter the while loop at line 10. If the code enters the while loop (line 10), it breaks out of it when \(\tilde {E}\) is small enough that the j th site simulation can never exceed the existing value. The vector V counts the number of times the while loop executes for each replicate. This ultimately provides the hitting scenario H.

figure a
figure b

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Beranger, B., Stephenson, A.G. & Sisson, S.A. High-dimensional inference using the extremal skew-t process. Extremes 24, 653–685 (2021). https://doi.org/10.1007/s10687-020-00376-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10687-020-00376-1

Keywords

AMS 2000 Subject Classifications

Navigation