Skip to main content
Log in

Skew exponential power stochastic volatility model for analysis of skewness, non-normal tails, quantiles and expectiles

  • Original Paper
  • Published:
Computational Statistics Aims and scope Submit manuscript

Abstract

This paper proposes a unified framework to analyse the skewness, tail heaviness, quantiles and expectiles of the return distribution based on a stochastic volatility model using a new parametrisation of the skew exponential power (SEP) distribution. The SEP distribution can express a wide range of distribution shapes through two shape parameters and one skewness parameter. Since the asymmetric Laplace and skew normal distributions are included as special cases, the proposed model is related to quantile regression and expectile regression. The efficient and simple Markov chain Monte Carlo estimation methods are also described. The proposed model is demonstrated using the simulated data and real data on daily return of foreign exchange rate.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Aas K, Haff IH (2006) The generalized hyperbolic skew student’s \(t\)-distribution. J Financ Econom 4:275–309

    Article  Google Scholar 

  • Aas K, Czado C, Frigessi A, Bakken H (2009) Pair-copula constructions of multiple dependence. Insur Math Econ 44:182–198

    Article  MathSciNet  MATH  Google Scholar 

  • Azzalini A (1985) A class of distributions which includes the normal ones. Scand J Stat 12:171–178

    MathSciNet  MATH  Google Scholar 

  • Azzalini A (1986) Further results on a class of distributions which includes the normal ones. Statistica 46:199–208

    MathSciNet  MATH  Google Scholar 

  • Balakrishnan N, Lai CD (2009) Continuous bivariate distributions. Springer, New York

    MATH  Google Scholar 

  • Bauwens L, Hafner CM, Laurent S (2012) Handbook of volatility models and their applications. Wiley, Hoboken

    Book  MATH  Google Scholar 

  • Bedford T, Cooke RM (2002) Vines: a new graphical model for dependent random variables. Ann Stat 30:1031–1068

    Article  MathSciNet  MATH  Google Scholar 

  • Bottazzi G, Secchi A (2011) A new class of asymmetric exponential power densities with applications to economics and finance. Ind Corp Change 20:991–1030

    Article  Google Scholar 

  • Cappuccio N, Lubian D, Raggi D (2004) MCMC Bayesian estimation of a skew-GED stochastic volatility model. Stud Nonlinear Dyn Econ 8:1–29

  • Chen CWS, Gerlach RH, Wei DCM (2009) Bayesian causal effects in quantiles: accounting for heteroscedasticity. Comput Stat Data Anal 53:1993–2007

    Article  MathSciNet  MATH  Google Scholar 

  • Chen CWS, Liu FC, So MKP (2013) Threshold variable selection of asymmetric stochastic volatility models. Comput Stat 28:2415–2447

    Article  MathSciNet  MATH  Google Scholar 

  • Chen Q, Gerlach RH, Lu Z (2012) Bayesian Value-at-Risk and expected shortfall forecasting via the asymmetric Laplace distribution. Comput Stat Data Anal 56:3498–3516

    Article  MathSciNet  MATH  Google Scholar 

  • Chib S (1995) Marginal likelihood from the Gibbs output. J Am Stat Assoc 90:1313–1321

    Article  MathSciNet  MATH  Google Scholar 

  • Chib S (2001) Markov chain Monte Carlo methods: computation and inference. In: Heckman JJ, Leamer E (eds) Handbook of econometrics. North Holland, Amsterdam, pp 3569–3649

    Google Scholar 

  • Chib S, Jeliazkov I (2001) Marginal likelihood from the Metropolis–Hastings output. J Am Stat Assoc 97:270–291

    Article  MathSciNet  MATH  Google Scholar 

  • Chib S, Nardari F, Shephard N (2002) Markov chain Monte Carlo methods for stochastic volatility models. J Econ 108:281–316

    Article  MathSciNet  MATH  Google Scholar 

  • Chollete L, Heinen A, Valdesogo A (2009) Modeling international financial returns with a multivariate regime-switching copula. J Financ Econ 7:237–480

    Google Scholar 

  • Choy STB, Wan WY, Chan CM (2008) Bayesian student-\(t\) stochastic volatility models via scale mixtures. Adv Econ 23:595–618

    MATH  Google Scholar 

  • Christoffersen P (1998) Evaluating interval forecasts. Int Econ Rev 39:841–862

    Article  MathSciNet  Google Scholar 

  • Christoffersen PF, Pelletier D (2004) Backtesting Value-at-Risk: a duration based approach. J Financ Econ 2:84–108

    Google Scholar 

  • de Jong P, Shephard N (1995) The simulation smoother for time series models. Biometrika 82:339–350

    Article  MathSciNet  MATH  Google Scholar 

  • De Rossi G, Harvey A (2009) Quantiles, expectiles, and splines. J Econ 152:179–185

    Article  MathSciNet  Google Scholar 

  • Doornik J (2007) Ox: object oriented matrix programming. Timberlake Consultants Press, London

    Google Scholar 

  • Dufour J-M (2006) Monte Carlo tests with nuisance parameters: a general approach to finite sample inference and nonstandard asymptotics. J Econ 133:443–477

    Article  MathSciNet  Google Scholar 

  • Durbin J, Koopman SJ (2002) A simple and efficient simulation smoother for state space time series analysis. Biometrika 89:603–615

    Article  MathSciNet  MATH  Google Scholar 

  • Embrechts P, Kaufmann R, Patie P (2005) Strategic long-term financial risks: single risk factors. Comput Optim Appl 32:61–90

    Article  MathSciNet  MATH  Google Scholar 

  • Engle RF, Russell JR (1998) Autoregressive conditional duration: a new model for irregularly spaced transaction data. Econometrica 66:1127–1162

    Article  MathSciNet  MATH  Google Scholar 

  • Engle RF, Manganelli S (2004) CAViaR: conditional autoregressive value at risk by regression quantiles. J Bus Econ Stat 22:367–381

    Article  MathSciNet  Google Scholar 

  • Fernández C, Steel MFJ (1995) On Bayesian modeling of fat tails and skewness. J Am Stat Assoc 93:359–371

    MathSciNet  MATH  Google Scholar 

  • Gerlach RH, Chen CWS, Chan NYC (2011) Bayesian time-varying quantile forecasting for Value-at-Risk in financial markets. J Bus Econ Stat 29:481–492

    Article  MathSciNet  Google Scholar 

  • Gerlach RH, Chen CWS, Lin L (2012) Bayesian semi-parametric expected shortfall forecasting in financial markets. The University of Sydney Business School, BA Working Paper 01/2012

  • Hafner CM, Manner H (2012) Dynamic stochastic copula models: estimation, inference and applications. J Appl Econ 27:269–295

    Article  MathSciNet  Google Scholar 

  • Ishihara T, Omori Y (2012) Efficient Bayesian estimation of a multivariate stochastic volatility model with cross leverage and heavy-tailed errors. Comput Stat Data Anal 56:3674–3689

    Article  MathSciNet  MATH  Google Scholar 

  • Jones MC (1994) Expectiles and M-quantiles are quantiles. Stat Probab Lett 20:149–153

    Article  MathSciNet  MATH  Google Scholar 

  • Koenker R, Bassett G (1978) Regression quantiles. Econometrica 46:33–50

    Article  MathSciNet  MATH  Google Scholar 

  • Koenker R (2005) Quantile regression. Cambridge University Press, New York

    Book  MATH  Google Scholar 

  • Kozumi H, Kobayashi G (2011) Gibbs sampling methods for Bayesian quantile regression. J Stat Comput Simul 81:1565–1578

    Article  MathSciNet  MATH  Google Scholar 

  • Kupiec P (1995) Techniques for verifying the accuracy of risk measurement models. J Deriv 3:73–84

    Article  Google Scholar 

  • Li S (2011) Three essays on econometrics: asymmetric exponential power distribution, econometric computation, and multifactor model. Ph.D. Thesis, Rutgers The State University of New Jersey, New Brunswick

  • Nakajima J, Omori Y (2009) Leverage, heavy-tails and correlated jumps in stochastic volatility models. Comput Stat Data Anal 53:2535–2553

    MathSciNet  MATH  Google Scholar 

  • Nakajima J, Omori Y (2012) Stochastic volatility model with leverage and asymmetrically heavy-tailed error using GH skew student’s-\(t\) distribution. Comput Stat Data Anal 56:3690–3704

    Article  MathSciNet  MATH  Google Scholar 

  • Nakajima J (2013) Stochastic volatility model with regime-switching skewness in heavy-tailed errors for exchange rate returns. Stud Nonlinear Dyn Econ 17:499–520

    MathSciNet  Google Scholar 

  • Nakajima J (2014) Bayesian analysis of multivariate stochastic volatility with skew distribution. Econ Rev. doi:10.1080/07474938.2014.977093

  • Naranjo L, Pérez CJ, Martín J (2012) Bayesian analysis of a skewed exponential power distribution. In: Proceedings of COMPSTAT 2012, 20th international conference on computational statistics, pp 641–652

  • Naranjo L, Pérez CJ, Martín J (2015) Bayesian analysis of some models that use the asymmetric exponential power distribution. Stat Comput 25:497–514

    Article  MathSciNet  MATH  Google Scholar 

  • Nelsen RB (2006) An introduction to copulas. Springer, Yew York

    MATH  Google Scholar 

  • Newey WK, Powell JL (1987) Asymmetric least squares estimation and testing. Econometrica 55:819–847

    Article  MathSciNet  MATH  Google Scholar 

  • Omori Y, Chib S, Shephard N, Nakajima J (2007) Stochastic volatility with leverage: fast likelihood inference. J Econ 140:425–449

    Article  MathSciNet  MATH  Google Scholar 

  • Omori Y, Watanabe T (2008) Block sampler and posterior mode estimation for asymmetric stochastic volatility models. Comput Stat Data Anal 52:2892–2910

    Article  MathSciNet  MATH  Google Scholar 

  • Park T, van Dyk D (2008) Partially collapsed Gibbs samplers: theory and methods. J Am Stat Assoc 103:790–796

    Article  MathSciNet  MATH  Google Scholar 

  • Pitt MK, Shephard N (1999) Filtering via simulation: auxiliary particle filter. J Am Stat Assoc 94:590–599

    Article  MathSciNet  MATH  Google Scholar 

  • Rubio FJ, Steel MFJ (2013) Bayesian modelling of skewness and kurtosis with two-piece scale and shape transformations. CRiSM working paper 13–10, University of Warwick

  • Shephard N, Pitt MK (1997) Likelihood analysis of non-Gaussian measurement time series. Biometrika 84:653–667

    Article  MathSciNet  MATH  Google Scholar 

  • Steel MFJ (1998) Bayesian analysis of stochastic volatility models with flexible tails. Econ Rev 17:109–143

    Article  MathSciNet  MATH  Google Scholar 

  • Takahashi M, Omori Y, Watanabe T (2009) Estimating stochastic volatility models using daily returns and realized volatility simultaneously. Comput Stat Data Anal 53:2404–2426

    Article  MathSciNet  MATH  Google Scholar 

  • Takahashi M, Watanabe T, Omori Y (2014) Volatility and quantile forecasts by realized stochastic volatility models with generalized hyperbolic distribution. CIRJE discussion papers 949

  • Tsiotas G (2012) On generalised asymmetric stochastic volatility models. Comput Stat Data Anal 56:151–172

    Article  MathSciNet  MATH  Google Scholar 

  • Watanabe T, Omori Y (2004) A multi-move sampler for estimating non-Gaussian time series models: comments on Shephard and Pitt (1997). Biometrika 91:246–248

    Article  MathSciNet  MATH  Google Scholar 

  • Wichitaksorn N,Wang JJJ, ChoySTB, Gerlach R (2014) Analyzing return asymmetry and quantiles through stochastic volatility models using asymmetric Laplace error via uniform scale mixtures. Appl Stoch Models Bus Ind. doi: 10.1002/asmb/2062

  • Yu K, Moyeed RA (2001) Bayesian quantile regression. Stat Probab Lett 54:437–447

    Article  MathSciNet  MATH  Google Scholar 

  • Zhu D, Zinde-Walsh V (2009) Properties and estimation of asymmetric exponential power distribution. J Econ 148:86–99

    Article  MathSciNet  Google Scholar 

  • Zhu D, Galbraith JW (2011) Modeling and forecasting expected shortfall with the generalized asymmetric student-\(t\) and asymmetric exponential power distributions. J Empir Financ 18:765–778

    Article  Google Scholar 

Download references

Acknowledgments

The author sincerely thanks the reviewers for their valuable suggestions regarding improvements in this paper. This work is partly supported by JSPS KAKENHI Grant Numbers 15K17036 and 25245035, and Grant-in-Aid for JSPS Fellows PD (13J06738). The computational results are obtained using Ox version 6.21 (Doornik 2007).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Genya Kobayashi.

Appendices

Appendix 1: SEP distribution and double two-piece family

Rubio and Steel (2013) proposed the double two-piece (DTP) family. Let \(f(y|\delta ,\sigma ,\theta )=\frac{1}{\sigma }f(\frac{y-\delta }{\sigma }|\theta )\) be the probability density function for the family of of continuous, unimodal, symmetric distribution where \(\delta \in \mathbb {R}\) is the mode and the location parameter, \(\sigma >0\) is the scale parameter, and \(\theta \) is the shape parameter. Then the probability density function of the DTP family is given by

$$\begin{aligned} f_{DTP}(y|\delta ,\sigma _1,\sigma _2,\theta _1,\theta _2)= & {} \frac{2\epsilon }{\sigma _1}f\left( \frac{y-\delta }{\sigma _1}\Big |\theta _1\right) I(y<\delta )\nonumber \\&+\frac{2(1-\epsilon )}{\sigma _2}f\left( \frac{y-\delta }{\sigma _2}\Big |\theta _2\right) I(y\ge \delta ), \end{aligned}$$
(9)

where \(I(\cdot )\) is the indicator function and

$$\begin{aligned} \epsilon =\frac{\sigma _1f(0|\theta _2)}{\sigma _1f(0|\theta _2)+\sigma _2f(0|\theta _1)}. \end{aligned}$$

The density (9) is continuous, unimodal with mode at \(\delta \), and the amount of mass to the left of its mode is given by \(\epsilon \).

By letting \(\sigma _1=\frac{\sigma }{(1-p)^{1/\theta _1}}\) and \(\sigma _2=\frac{\sigma }{p^{1/\theta _2}}\), the probability density function of the SEP distribution (1) becomes the parameterisation of Naranjo et al. (2012):

$$\begin{aligned} f_{SEP}(y|\delta , \sigma _1,\sigma _2, \theta _1,\theta _2)= & {} c^{-1} \left\{ \begin{array}{ll} \exp \left\{ -\left| \frac{y-\delta }{\sigma _1}\right| ^{\theta _1} \right\} , \ &{}\quad \text {if} \quad y<\delta , \\ \exp \left\{ -\left| \frac{y-\delta }{\sigma _2}\right| ^{\theta _2} \right\} , \ &{}\quad \text {otherwise}, \\ \end{array} \right. \nonumber \\ c= & {} \sigma _1\Gamma (1+1/\theta _1) + \sigma _2\Gamma (1+1/\theta _2). \end{aligned}$$
(10)

It can be immediately seen that (10) belongs to the DTP family based on the exponential power distribution given by (6).

Appendix 2: MCMC methods for SEPSV

1.1 Single-move sampler with MH updates for \(\psi _0\) and \(\psi _1\)

Using the mixture representation for the SEP distribution (7), the return equation can be rewritten as

$$\begin{aligned} y_t|u_{1t}&\sim U\left( \psi _0+\psi _1y_{t-1}-\frac{e^{h_t/2}}{s_{\varvec{\theta }}}\left( \frac{u_{1t}}{1-p}\right) ^{1/\theta _1}, \psi _0+\psi _1y_{t-1}\right) \ \quad \text {with probability } \pi _1,\nonumber \\ y_t|u_{2t}&\sim U\left( \psi _0+\psi _1y_{t-1}, \psi _0+\psi _1y_{t-1}+\frac{e^{h_t/2}}{s_{\varvec{\theta }}}\left( \frac{u_{2t}}{p}\right) ^{1/\theta _2}\right) \ \quad \text {with probability } \pi _2, \end{aligned}$$
(11)

where \(u_1\sim Ga(1+1/\theta _1, 1)\), \(u_2\sim Ga(1+1/\theta _2, 1)\), and \(\pi _1\) and \(\pi _2\) are defined in (2).

The sampling scheme described here samples the values of \(\psi _0\), \(\psi _1\), \(\theta _1\), \(\theta _2\), p, \(\mu \), \(\phi \), \(\tau ^2\), \(\{u_{1t}\}_{t=1}^T\), \(\{u_{2t}\}_{t=1}^T\), and \(\{h_t\}_{t=1}^T\) alternately from their full conditional distributions (FCDs). The single-move sampler samples each \(h_t\) at a time from its FCD.

Sampling of \(\psi _0\), \(\psi _1\), \(\theta _1\), \(\theta _2\), and p We integrate out the latent variables \(u_{1t}\) and \(u_{2t}\) (see, e.g., Park and van Dyk 2008) and implement the Metropolis–Hastings (MH) algorithm which samples the following three blocks: \(\psi _0\), \(\psi _1\), and \({\varvec{\theta }}=(\theta _1,\theta _2,p)\). The acceptance probability of the MH algorithm for \(\psi _0\) is given by

$$\begin{aligned} \min \left\{ 1,\frac{\prod _{t=1}^Tf_{{\textit{SEP}}}(y_t|\psi _0^*+\psi _1y_{t-1},s_{{\varvec{\theta }}}^{-1} e^{h_t/2}, \theta _1,\theta _2,p)\pi (\psi _0^*)q(\psi _0|\psi ^*)}{\prod _{t=1}^Tf_{{\textit{SEP}}}(y_t|\psi _0+\psi _1y_{t-1},s_{{\varvec{\theta }}}^{-1} e^{h_t/2}, \theta _1,\theta _2,p)\pi (\psi _0)q(\psi _0^*|\psi _0)}\right\} , \end{aligned}$$

where \(\psi _0^*\) is the proposed value, \(q(\cdot |\cdot )\) is the proposal density, and \(\pi (\cdot )\) is the prior density. Similarly, the acceptance probabilities are given by

$$\begin{aligned} \min \left\{ 1,\frac{\prod _{t=1}^Tf_{{\textit{SEP}}}(y_t|\psi _0+\psi _1^*y_{t-1},s_{{\varvec{\theta }}}^{-1} e^{h_t/2}, \theta _1,\theta _2,p)\pi (\psi _1^*)q(\psi _1|\psi _1^*)}{\prod _{t=1}^Tf_{{\textit{SEP}}}(y_t|\psi _0+\psi _1y_{t-1},s_{{\varvec{\theta }}}^{-1} e^{h_t/2}, \theta _1,\theta _2,p)\pi (\psi _1)q(\psi _1^*|\psi _1)}\right\} , \end{aligned}$$

for \(\psi _1\), and

$$\begin{aligned} \min \left\{ 1,\frac{\prod _{t=1}^Tf_{{\textit{SEP}}}(y_t|\psi _0+\psi _1y_{t-1},s_{{\varvec{\theta }}^*}^{-1} e^{h_t/2}, \theta _1^*,\theta _2^*,p^*)\pi ({\varvec{\theta }}^*)q({\varvec{\theta }}|{\varvec{\theta }}^*)}{\prod _{t=1}^Tf_{{\textit{SEP}}}(y_t|\psi _0+\psi _1y_{t-1},s_{{\varvec{\theta }}}^{-1} e^{h_t/2}, \theta _1,\theta _2,p)\pi ({\varvec{\theta }})q({\varvec{\theta }}^*|{\varvec{\theta }})}\right\} , \end{aligned}$$

for \({\varvec{\theta }}\), where \(\psi _1^*\) and \({\varvec{\theta }}^*\) are the proposed values. We use the normal distribution centred at the current state as the proposal distribution. The variance of the proposal distribution is chosen such that the acceptance rate is between 0.2 and 0.4.

Sampling of \(u_{1t}\) and \(u_{2t}\) The FCD of \(u_{1t}\) is the exponential distribution with mean 1 truncated to the region in which the value of \(u_{1t}\) satisfies the mixture representation as

$$\begin{aligned} u_{1t}|-\sim Exp(1) I\left[ (1-p)\left( \frac{s_{\varvec{\theta }}(\psi _0+\psi _1y_{t-1}-y_t)}{e^{h_t/2}}\right) ^{\theta _1}<u_{1t}\right] ,\quad \text {if}\quad y_t<\psi _0+\psi _1y_{t-1}, \end{aligned}$$
(12)

for \(t=1,\dots ,T\) where \(u_{1t}|-\) denotes \(u_{1t}\) conditional on all other variables. The value of \(u_{2t}\) is sampled in a similar manner from

$$\begin{aligned} u_{2t}|-\sim Exp(1) I\left[ p\left( \frac{s_{\varvec{\theta }}(y_t-\psi _0-\psi _1y_{t-1})}{e^{h_t/2}}\right) ^{\theta _2}<u_{2t}\right] ,\quad \text {if}\quad y_t\ge \psi _0+\psi _1y_{t-1}. \end{aligned}$$
(13)

See also Naranjo et al. (2012, 2015).

Sampling of \(h_t\) Utilising the mixture representation for the SEP distribution, the FCD for \(h_t\) is the truncated normal distribution. Therefore, no MH type update is required.

The FCD of \(h_t\) is given by

$$\begin{aligned} h_t|-\sim TN_{(\underline{h_t},\infty )}(m_t,s_t), \quad t=1,\dots ,T, \end{aligned}$$

where \(TN_{(a,b)}(\mu ,\sigma )\) denotes the normal distribution with mean \(\mu \) and variance \(\sigma \) truncated on the interval (ab),

$$\begin{aligned} m_t= & {} \left\{ \begin{array}{ll} \mu +\phi (h_2-\mu )-\frac{\tau ^2}{2}, \ &{}t=1, \\ \mu +\frac{\phi ((h_{t-1}-\mu ) + ( h_{t+1} -\mu )) -\tau ^2/2}{1+\phi ^2}, \ &{}\quad t = 2,\dots ,T-1, \\ \mu + \phi (h_{T-1}-\mu )-\frac{\tau ^2}{2}, \ &{}\quad t = T \end{array} \right. \\ s_t= & {} \left\{ \begin{array}{ll} \tau ^2, \ &{}t=1, T \\ \frac{\tau ^2}{1+\phi ^2}, \ &{}\quad t = 2,\dots ,T-1, \\ \end{array} \right. \\ \underline{h_t}= & {} \left\{ \begin{array}{ll} 2\left[ \log (s_{\varvec{\theta }})+\log (\psi _0+\psi _1y_{t-1}-y_t) +\frac{1}{\theta _1}\left( \log (1-p) - \log u_{1t} \right) \right] , &{}\quad \text {if} \ y_t<\psi _0+\psi _1y_{t-1},\\ 2\left[ \log (s_{\varvec{\theta }})+\log ( y_t-\psi _0-\psi _1y_{t-1}) + \frac{1}{\theta _2}\left( \log p-\log u_{2t}\right) \right] , &{}\quad \text {otherwise.} \\ \end{array} \right. \end{aligned}$$

A similar method can be found, e.g., in Choy et al. (2008), where the SV model with Student’s t distribution was estimated using the scale mixture of uniforms representation.

Sampling of \(\mu \) The FCD of \(\mu \) is given by

$$\begin{aligned} \mu |-\sim N(m_1, M_1), \end{aligned}$$

where

$$\begin{aligned} m_1= & {} M_1\left[ \frac{1-\phi ^2}{\tau ^2}h_1 + \frac{1-\phi }{\tau ^2}\sum _{t=2}^T(h_t-\phi h_{t-1})+ \frac{m_0}{M_0}\right] ,\\ \quad \text {and}\quad M_1= & {} \left[ \frac{(T-1)(1-\phi )^2}{\tau ^2} + \frac{(1-\phi ^2)}{\tau ^2} + \frac{1}{M_0}\right] ^{-1}. \end{aligned}$$

Sampling of \(\tau ^2\) The FCD of \(\tau ^2\) is given by

$$\begin{aligned} \tau ^2|-\sim IG\left( \frac{n_1}{2}, \frac{S_1}{2}\right) , \end{aligned}$$

where

$$\begin{aligned} n_1\!=\!n_0 + T, \quad \text {and}\quad S_1\!=\!S_0 + (h_1-\mu )^2(1-\phi ^2) \!+\! \sum _{t=2}^T(h_t-\mu -\phi (h_{t-1}-\mu ))^2. \end{aligned}$$

Sampling of \(\phi \) The FCD of \(\phi \) is given by

$$\begin{aligned} \pi (\phi ^*|-)\propto \pi (\phi )\times \sqrt{1-\phi ^2} \times N( k_1, t_1), \end{aligned}$$

where

$$\begin{aligned} k_1= \frac{t_1}{\tau ^2} \sum _{t=2}^T ( h_t-\mu )( h_{t-1}-\mu ), \quad \text {and} \quad t_1 = \frac{\tau ^2}{\sum _{t=2}^{T-1}( h_t-\mu )^2}. \end{aligned}$$

We can sample from this FCD using the independence MH algorithm with the \(N(k_1,t_1)\) proposal distribution and the acceptance probability is given by

$$\begin{aligned} \min \left\{ 1,\frac{\pi (\phi ^*)\sqrt{1-\phi ^{*2}}}{\pi (\phi )\sqrt{1-\phi ^2}}\right\} . \end{aligned}$$

1.2 Single-move sampler with Gibbs updates for \(\psi _0\) and \(\psi _1\)

An alternative sampling method draws the values of \(\psi _0\) and \(\psi _1\) from the FCDs derived directly from the mixture representation (11) as Naranjo et al. (2012, 2015). The rest of the steps for \(\theta _1\), \(\theta _2\), p, \(\mu \), \(\phi \), \(\tau ^2\), \(\{u_{1t}\}_{t=1}^T\), \(\{u_{2t}\}_{t=1}^T\), and \(\{h_t\}_{t=1}^T\) is identical to the sampler described previously.

Sampling of \(\psi _0\) The FCD of \(\psi _0\) is the truncated normal distribution given by

$$\begin{aligned} \psi _0|-\sim TN_{(\underline{\psi _0}, \overline{\psi _0})}(v_0, V_0), \end{aligned}$$

where

$$\begin{aligned} \underline{\psi _0}= & {} \max \left\{ \max _{t:u_{2t}>0}\left\{ y_t-\psi _1y_{t-1}-\frac{e^{h_t/2}}{s_{\varvec{\theta }}}\left( \frac{u_{2t}}{p}\right) ^{1/\theta _2}\right\} ,\quad \max _{t:u_{1t}>0,\#\left\{ u_{2t}>0\right\} =0}\left\{ y_t-\psi _1y_{t-1}\right\} \right\} , \\ \overline{\psi _0}= & {} \min \left\{ \min _{t:u_{1t}>0}\left\{ y_t-\psi _1y_{t-1}+\frac{e^{h_t/2}}{s_{\varvec{\theta }}}\left( \frac{u_{1t}}{1-p}\right) ^{1/\theta _1}\right\} ,\quad \min _{t:u_{2t}>0,\#\left\{ u_{1t}>0\right\} =0}\left\{ y_t-\psi _1y_{t-1}\right\} \right\} . \end{aligned}$$

Sampling of \(\psi _1\) The FCD of \(\psi _1\) is given by

$$\begin{aligned} \psi _1|-\sim U(\underline{\psi _1}, \overline{\psi _1}) \end{aligned}$$

where

$$\begin{aligned} \underline{\psi _1}= & {} \max \left\{ \max _{t:u_{1t}>0,y_{t-1}<0 }\left\{ \frac{y_t-\psi _0+\frac{e^{h_t/2}}{s_{\varvec{\theta }}} \left( \frac{u_{1t}}{1-p}\right) ^{1/\theta _1}}{y_{t-1}}\right\} ,\right. \\&\left. \max _{t:u_{2t}>0,y_{t-1}>0}\left\{ \frac{y_t-\psi _0-\frac{e^{h_t/2}}{s_{\varvec{\theta }}}\left( \frac{u_{2t}}{p}\right) ^{1/\theta _2}}{y_{t-1}}\right\} , -1\right\} , \\ \overline{\psi _1}= & {} \min \left\{ \min _{t:u_{1t}>0, y_{t-1}>0}\left\{ \frac{y_t-\psi _0+\frac{e^{h_t/2}}{s_{\varvec{\theta }}}\left( \frac{u_{1t}}{1-p}\right) ^{1/\theta _1}}{y_{t-1}}\right\} , \right. \\&\left. \min _{t:u_{2t}>0,y_{t-1}<0}\left\{ \frac{y_t-\psi _0-\frac{e^{h_t/2}}{s_{\varvec{\theta }}} \left( \frac{u_{2t}}{p}\right) ^{1/\theta _2}}{y_{t-1}}\right\} , 1\right\} . \end{aligned}$$

1.3 Block sampler

A block sampler (Shephard and Pitt 1997; Watanabe and Omori 2004; Omori and Watanabe 2008) samples \((h_s,\dots ,s_{s+m-1})\) in blocks in stead of sampling each \(h_t\) at a time. Let \(\alpha _t=h_t-\mu \) and \(\psi _t=\psi _0+\psi _1y_{t-1}\). We divide \((\alpha _1,\dots ,\alpha _T)\) into \(K+1\) blocks, \((\alpha _{k_{i-1}+1},\dots ,\alpha _{k_i})\) for \(i=1,\dots ,K+1\), with \(k_0=0\), \(k_{K+1}=T\), and \(k_i-k_{i-1}\ge 2\). How to determine K and \(k_i\) are described below.

The posterior distribution of \({\varvec{\eta }}=( \eta _s,\dots ,\eta _{s+m-1})\) conditional on \(\Xi =(\alpha _s, \alpha _{s+m}, \psi _0,\psi _1,\theta _1,\theta _2,p,\mu ,\tau ^2,\phi , \mathbf{y})\) is given by

$$\begin{aligned} \pi ({\varvec{\eta }}|\Xi ) \propto \prod _{t=s}^{s+m-1}f_{{\textit{SEP}}}(y_t|\psi _t, e^{h_t/2}s^{-1}_{\varvec{\theta }},\theta _1,\theta _2,p) \prod _{t=s}^{s+m-1}f(\eta _t) \times f(\alpha _{s+m}), \end{aligned}$$
(14)

where

$$\begin{aligned} f(\eta _t)=\left\{ \begin{array}{ll} \exp \left\{ -\frac{-(1-\phi ^2)\eta _1^2}{2\tau ^2}\right\} , &{}\quad \text {if}\quad t=1, \\ \exp \left\{ -\frac{\eta _t^2}{2\tau ^2}\right\} , &{} \quad \text {otherwise},\\ \end{array} \right. \end{aligned}$$

and

$$\begin{aligned} f(\alpha _{s+m+1})=\left\{ \begin{array}{c@{\quad }l} \exp \left\{ -\frac{(\alpha _{s+m+1}-\phi \alpha _{s+m})^2}{2\tau ^2}\right\} , \quad &{}\text {if} \quad s+m<T,\\ 1\quad &{}\text {if} \quad s+m=T. \end{array} \right. \end{aligned}$$

We sample \({\varvec{\eta }}\) from (14) using the acceptance–rejection (AR) MH algorithm. To construct a proposal distribution, consider the second-order Taylor expansion of

$$\begin{aligned} \ell _t= & {} \log f_{{\textit{SEP}}}(y_t|\psi _t,e^{h_t/2}s^{-1}_{\varvec{\theta }},\theta _1,\theta _2,p) \\= & {} \text {const} - \frac{\alpha _t}{2}-e^{-\frac{\theta _1\alpha _t}{2}}g_1(y_t)I(y_t<\psi _t) -e^{-\theta _2\alpha _t/2}g_2(y_t)I(y_t\ge \psi _t), \end{aligned}$$

around the mode \(\hat{\alpha }_t\) (hence \(\hat{\eta }_t\)) such that

$$\begin{aligned} \ell _t \approx \widehat{\ell }_t + (\alpha _t-\hat{\alpha }_t)\widehat{\ell }_t' +\frac{1}{2}(\alpha _t-\hat{\alpha }_t)^2\widehat{\ell }''_t, \end{aligned}$$

where

$$\begin{aligned}&\ell '_t=\frac{\partial \ell _t}{\partial \alpha _t} =-\frac{1}{2}+\frac{\theta _1}{2}e^{-\frac{\theta _1\alpha _t}{2}}g_1(y_t)I(y_t<\psi _t)+\frac{\theta _2}{2}e^{-\frac{\theta _2\alpha _t}{2}}g_2(y_t)I(y_t\ge \psi _t),\\&\ell ''_t=\frac{\partial ^2\ell _t}{\partial \alpha _t^2} =-\frac{\theta _1^2}{4}e^{-\frac{\theta _1\alpha _t}{2}}g_1(y_t)I(y_t<\psi _t) - \frac{\theta _2^2}{4}e^{-\frac{\theta _2\alpha _t}{2}}g_2( y_t)I( y_t\ge \psi _t),\\&g_1(y_t)=e^{-\frac{\theta _1\mu }{2}}(1-p)\left| s_{\varvec{\theta }}( y_t-\psi _t)\right| ^{\theta _1}, \\&g_2(y_t)=e^{-\frac{\theta _2\mu }{2}}p\left| s_{\varvec{\theta }}(y_t-\psi _t)\right| ^{\theta _2}, \end{aligned}$$

\(\widehat{\ell }_t=\ell _t|_{\alpha _t=\hat{\alpha }_t}\), \(\widehat{\ell }'_t=\ell '_t|_{\alpha _t=\hat{\alpha }_t}\), and \(\widehat{\ell }''_t=\ell ''_t|_{\alpha _t=\hat{\alpha }_t}\). How to obtain the mode \(\hat{\alpha }_t\) is described below. Note that we do not use the mixture representation (11) to construct the block sampler for the SEPSV model. This is because if we use (11), the resulting \(\ell ''_t\) is not strictly negative as required by Shephard and Pitt (1997).

The proposal distribution is formed as

$$\begin{aligned} q({\varvec{\eta }}|\Xi )\propto \prod _{t=s}^{s+m}\exp \left\{ -\frac{(\alpha _t^*-\alpha _t)^2}{2\sigma _t^{2*}}\right\} \prod _{t=s}^{s+m-1} f(\eta _t), \end{aligned}$$
(15)

where

$$\begin{aligned} \alpha ^*_t = \hat{\alpha }_t+\sigma _t^*\widehat{\ell }'_t,\quad \sigma ^*_t=-1/\widehat{\ell }''_t, \quad t=s,\dots ,s+m-1\quad \text {and} \quad t=T, \end{aligned}$$

and

$$\begin{aligned} \alpha ^*_t= \sigma _t^*\left( \widehat{\ell }'_t-\widehat{\ell }''_t\hat{\alpha }_t + \frac{\phi }{\tau ^2}\alpha _{t+1}\right) , \quad \sigma ^*_t = \left( \frac{\phi ^2}{\tau ^2}-\widehat{\ell }''_t\right) ^{-1},\quad t = s+m < T. \end{aligned}$$

Then consider the linear Gaussian state space model given by

$$\begin{aligned} \alpha ^*_t&= \alpha _t + \epsilon _t,\quad \epsilon _t\sim N(0, \sigma ^*_t),\quad t = s,\dots , s+m-1,\nonumber \\ \alpha _{t}&=\phi \alpha _{t-1} + \eta _t,\quad \eta _t\sim N(0,\tau ^2), \quad t = s+1,\dots ,s+m-1, \end{aligned}$$
(16)

with \(\alpha _1\sim N(0, \tau ^2/(1-\phi ^2))\). To sample from (15), we first use the Kalman filter and disturbance smoother with respect to (16) to obtain the mode of the posterior distribution. Then the simulation smoother (de Jong and Shephard 1995; Durbin and Koopman 2002) with the accept–rejection (AR) algorithm is applied to sample values around the posterior mode. Finally, the candidate values are accepted based on the MH ratio.

Suppose the current state is \({\varvec{\eta }}\). The algorithm can be summarised as follows.

  1. 1.

    Initialise \((\hat{\eta }_s,\dots ,\hat{\eta }_{s+m-1})\) and \((\hat{\alpha }_s,\dots ,\hat{\alpha }_{s+m-1})\).

  2. 2.

    Find mode \((\hat{\alpha }_s,\dots ,\hat{\alpha }_{s+m-1})\) by repeating the following steps several times:

    1. (a)

      Compute \((\alpha ^*_s,\dots ,\alpha ^*_{s+m-1})\) and \((\sigma ^{2*}_s,\dots ,\sigma ^{2*}_{s+m-1})\).

    2. (b)

      Apply Kalman filter and disturbance smoother to the linear Gaussian state space model (16) and obtain the posterior mode \((\hat{\alpha }_s,\dots ,\hat{\alpha }_{s+m-1})\).

  3. 3.

    Given the posterior mode, compute \((\alpha ^*_s,\dots ,\alpha ^*_{s+m-1})\) and \((\sigma ^{2*}_s,\dots ,\sigma ^{2*}_{s+m-1})\).

  4. 4.

    (AR-step) Repeat the following until a candidate is accepted.

    1. (a)

      Apply simulation smoother to (16) to draw a candidate \(\tilde{{\varvec{\eta }}}=(\tilde{\eta }_s,\dots ,\tilde{\eta }_{s+m-1})\).

    2. (b)

      Accept \(\tilde{{\varvec{\eta }}}\) with probability

      $$\begin{aligned} \frac{\min \left\{ \pi (\tilde{{\varvec{\eta }}}|\Xi ),cq(\tilde{{\varvec{\eta }}}|\Xi )\right\} }{cq(\tilde{{\varvec{\eta }}}|\Xi )}, \end{aligned}$$

      where \(c>0\) is a constant.

  5. 5.

    (MH-step) Accept \(\tilde{{\varvec{\eta }}}\) with probability

    $$\begin{aligned} \min \left\{ 1,\frac{\pi (\tilde{{\varvec{\eta }}}|\Xi )\min \left\{ \pi ({\varvec{\eta }}|\Xi ),cq({\varvec{\eta }})\right\} }{\pi ({\varvec{\eta }}|\Xi )\min \left\{ \pi (\tilde{{\varvec{\eta }}}|\Xi ),cq(\tilde{{\varvec{\eta }}})\right\} } \right\} . \end{aligned}$$

To determine the blocks, we employ the stochastic knots (Shephard and Pitt 1997) based on \(k_i=\text {int}\left[ T(i+U_i)/(K+2)\right] \) where \(U_i\sim U(0,1)\). It is known that the stochastic knots approach can improve the efficiency of MCMC by randomly changing the conditioning sets over the iterations. The tuning parameter K is determined such that the size of block is neither too small nor too large (see, e.g., Omori and Watanabe 2008). When K is too small, the sampler would resemble the single move sampler using an MH algorithm and would be inefficient. Choosing K too large would lead to low acceptance rate and to inefficiency of the sampler as well. Therefore, we choose the value of K such that the number of points in a block is 30 on average. This choice of K works well in the numerical examples in this paper.

The sampling scheme for the rest, \(\psi _0\), \(\psi _1\), \(\theta _1\), \(\theta _2\), p, \(\mu \), \(\phi \), and \(\tau ^2\), is the same as in the case of the single-move sampler with MH updates.

1.4 Sampling from the posterior predictive distribution

To compute one-day-ahead VaR and ES forecasts, the following steps are added after a sweep of an MCMC scheme described above. We generate daily return forecast utilising the mixture representation (11).

  1. 1.

    Generate \(h_{t+1}\) from \(N(\mu +\phi (h_t-\mu ), \tau ^2)\).

  2. 2.

    Generate u from U(0, 1).

    • if \(\pi _1>u\)

      1. (a)

        Generate \(u_1\) from \(Ga(1+1/\theta _1,1)\).

      2. (b)

        Generate \(y_{t+1}\) from \(U\left( \psi _0+\psi _1y_t-\frac{e^{h_{t+1}/2}}{s_{\varvec{\theta }}}\left( \frac{u_1}{1-p}\right) ^{1/\theta _1}, \psi _0+\psi _1y_t\right) \).

    • else

      1. (a)

        Generate \(u_2\) from \(Ga(1+1/\theta _2,1)\).

      2. (b)

        Generate \(y_{t+1}\) from \(U\left( \psi _0+\psi _1y_t, \psi _0+\psi _1y_t+\frac{e^{h_{t+1}/2}}{s_{\varvec{\theta }}}\left( \frac{u_2}{p}\right) ^{1/\theta _2}\right) \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kobayashi, G. Skew exponential power stochastic volatility model for analysis of skewness, non-normal tails, quantiles and expectiles. Comput Stat 31, 49–88 (2016). https://doi.org/10.1007/s00180-015-0596-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00180-015-0596-4

Keywords

Navigation