Skip to main content
Log in

A switching self-exciting jump diffusion process for stock prices

  • Research Article
  • Published:
Annals of Finance Aims and scope Submit manuscript

Abstract

This study proposes a new Markov switching process with clustering effects. In this approach, a hidden Markov chain with a finite number of states modulates the parameters of a self-excited jump process combined to a geometric Brownian motion. Each regime corresponds to a particular economic cycle determining the expected return, the diffusion coefficient and the long-run frequency of clustered jumps. We study first the theoretical properties of this process and we propose a sequential Monte-Carlo method to filter the hidden state variables. We next develop a Markov Chain Monte-Carlo procedure to fit the model to the S&P 500. We find that self-exciting jumps occur mainly during economic recession and nearly disappear in periods of economic growth. Finally, we analyse the impact of such a jump clustering on implied volatilities of European options.

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
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. Notice that stochastic volatility models also explain assymmetry and high kurtosis. However Cont and Tankov (2004) mention on page 6 that Brownian models with stochastic volatility cannot explain the presence of jumps in price due to the continuity of sample paths.

  2. Notice that if \(C\in [u^{*},\omega _{2})\), the function \(F_{\omega _{1}}(C)\) is equal to

    $$\begin{aligned} F_{\omega _{1}}(C):&=-\int _{C}^{\omega _{2}}\frac{du}{-\beta (\omega _{1})-\alpha u+\psi \left( \omega _{1}\,,\,\eta \,u\right) }. \end{aligned}$$
  3. We can think to relate thresholds to regimes. For example, if we denote by \(\tilde{\sigma }_{i}^{ML}\) the volatility of the SGBM in the most likely state at time \(t_{i}\), we can assume that \(g(\alpha _{k},i)=\sqrt{\Delta }\tilde{\sigma }_{i}^{ML}\Phi ^{-1}(\alpha _{k})\quad k=1,2\). However, when this method is applied to the S&P 500 data set, fewer jumps are detected during recessions than in periods of growth. This counter intuitive result is explained by the fact that thresholds are proportional to the volatility in each regime. As the volatility is much important during recessions than in other periods, thresholds are also much higher. Consequence: less log-returns exceed thresholds during bad economic times. This observation motivates us to not relate thresholds to regimes and to use instead the standard deviation of the whole sample. This reduces the accuracy of the POT method. However, given that we use it to find an acceptable starting point for the MCMC algorithm, this loss of accuracy has a limited impact on final conclusions.

  4. In theory the acceptance rate can be improved to any desired level. Increasing the number of particles raises considerably the computational time of the estimation procedure.

References

  • Alexander, S.S.: Price movements in speculative markets: trends or random walks. In: Cootner, 338–372 (1964)

  • Andersen, T., Benzoni, L., Lund, J.: An empirical investigation of continuous-time equity return models. J Finance 57, 1239–1284 (2002)

    Article  Google Scholar 

  • Ait-Sahalia, Y., Cacho-Diaz, J., Laeven, R.J.A.: Modeling financial contagion using mutually exciting jump processes. J Financ Econom 117, 585–606 (2015)

    Article  Google Scholar 

  • Ait-Sahalia, Y., Laeven, R.J.A., Pelizzon, L.: Mutual excitation in eurozone sovereign CDS. J Econom 183(2), 151–167 (2014)

    Article  Google Scholar 

  • Al-Anaswah, N., Wilfing, B.: Identification of speculative bubbles using state-space models with Markov-switching. J Bank Finance 35(5), 1073–1086 (2011)

    Article  Google Scholar 

  • Bates, D.: Post-87 crash fears in S&P500 futures options. J Econom 94, 181–238 (2000)

    Article  Google Scholar 

  • Bollerslev, T.: Generalized autoregressive conditional heteroskedasticity. J Econom 31, 307–27 (1986)

    Article  Google Scholar 

  • Boyarchenko, M., Boyarchenko, S.: Double barrier options in regime-switching hyper-exponential jump-diffusion models. Int J Theor Appl Finance 14(7), 1005–1044 (2011)

    Article  Google Scholar 

  • Boyarchenko, M., Levendorskii, S.: American options in regime switching models. SIAM J Control Optim 48(3), 1353–1376 (2009)

    Article  Google Scholar 

  • Calvet, L., Fisher, A.: Forecasting multifractal volatility. J Econom 105, 17–58 (2001)

    Article  Google Scholar 

  • Calvet, L., Fisher, A.: How to forecast long-Run volatility: regime switching and the estimation of multifractal processes. J Financ Econom 2, 49–83 (2004)

    Article  Google Scholar 

  • Carr, P., Madan, B.: Option valuation using the fast Fourier transform. J Comput Finance 2, 61–73 (1999)

    Article  Google Scholar 

  • Carr P., Wu L.: Leverage effect, volatility feedback, and self-exciting market disruptions. J Financ Quant Anal (forthcoming) (2016)

  • Chen, K., Poon, S.-H.: Variance swap premium under stochastic volatility and self-exciting jumps. Working paper SSRN-id2200172 (2013)

  • Chevallier, J., Goutte, S.: Detecting jumps and regime switches in international stock markets returns. Appl Econ Lett 22(13), 1011–1019 (2015)

    Article  Google Scholar 

  • Cholette, L., Heinen, A., Al, V.: Modelling international financial returns with a multivariate regime switching copula. J Financ Econom 7(4), 437–480 (2009)

    Article  Google Scholar 

  • Chopin, N., Jacob, P.E., Papaspiliopoulos, O.: SMC2: an efficient algorithm for sequential analysis of state space models. J R Stat Soc Ser B Stat Methodol 75(3), 397–426 (2013)

    Article  Google Scholar 

  • Cont, R., Tankov, P.: Financial modelling with jump processes. Chapman & Hall/CRC Financial Mathematics Series (2004)

  • Doucet, A., De Freitas, J.F.G., Gordon, N.: Sequential Monte Carlo Methods in Practice. Cambridge University Press, Cambridge (2000)

    Google Scholar 

  • Duffie, D., Singleton, K.: Simulated moments estimation of Markov models of asset prices. Econometrica 61, 929–952 (1993)

    Article  Google Scholar 

  • Duffie, D., Pan, J., Singleton, K.: Transform analysis and asset pricing for affine jump diffusions. Econometrica 68, 1343–1376 (2000)

    Article  Google Scholar 

  • Elliott, R.J., Aggoun, L., Moore, J.B.: Hidden Markov Models: Estimation and Control. Springer, New York (1995)

    Google Scholar 

  • Elliott, R.J., Chan, L., Siu, T.K.: Option pricing and Esscher transform under regime switching. Ann Finance 1, 423–432 (2005)

    Article  Google Scholar 

  • Elliott, R.J., Siu, T.K., Chan, L., Lau, J.W.: Pricing options under a generalized Markov-modulated jump-diffusion Model. Stoch Anal Appl 25(4), 821–843 (2007)

    Article  Google Scholar 

  • Embrechts, P., Liniger, T., Lu, L.: Multivariate Hawkes processes: an application to financial data. J Appl Probab 48(A), 367–378 (2011)

  • Engle, R.: Autoregressive conditional heteroskedasticity with estimates of the variance of United Kingdom inflation. Econometrica 50, 987–1007 (1982)

    Article  Google Scholar 

  • Eraker, B., Johannes, M., Polson, N.: The impact of jumps in volatility and returns. J Finance 58, 1269–1300 (2003)

    Article  Google Scholar 

  • Evans, K.: Intraday jumps and US macroeconomic news announcements. J Bank Finance 35(10), 2511–2527 (2011)

    Article  Google Scholar 

  • Fulop, A., Li, J.: Efficient learning via simulation: a marginalized resample-move approach. J Econom 176(2), 146–161 (2013)

    Article  Google Scholar 

  • Fulop, A., Duan, J.-C.: Density-tempered marginalized sequential Monte Carlo samplers. J Bus Econom Stat 33, 192–202 (2015)

    Article  Google Scholar 

  • Gallant, A.R., Tauchen, G.: Which moments to match? Econom Theory 12, 657–681 (1996)

    Article  Google Scholar 

  • Gourieroux, C., Monfort, A., Renault, E.: Indirect inference. J Appl. Econom 8, S85–S118 (1993)

    Google Scholar 

  • Fulop, A., Li, J., Yu, J.: Self-exciting jumps, learning, and asset pricing implications. Rev Financ Stud 28, 876–912 (2015)

    Article  Google Scholar 

  • Gerber, H.U., Shiu, E.S.W.: Options pricing by Esscher transform. Trans Soc Actuaries 26, 99–191 (1994)

    Google Scholar 

  • Guidolin, M., Timmermann, A.: Economic implications of bull and bear regimes in UK stock and bond returns. Econom J 115, 11–143 (2005)

    Google Scholar 

  • Guidolin, M., Timmermann, A.: Asset allocation under multivariate regime switching. J Econom Dyn Control 31(11), 3503–3544 (2007)

    Article  Google Scholar 

  • Guidolin, M., Timmermann, A.: International asset allocation under regime switching, skew, and kurtosis preferences. Rev Financ Stud 21(2), 889–935 (2008)

    Article  Google Scholar 

  • Hainaut, D.: A model for interest rates with clustering effects. Quant Finance 16(8), 1203–1218 (2016)

    Article  Google Scholar 

  • Hainaut, D., MacGilchrist, R.: Strategic asset allocation with switching dependence. Ann Finance 8(1), 75–96 (2012)

    Article  Google Scholar 

  • Hainaut, D., Moraux, F.: Hedging of options in the presence of jump clustering. J Comput Finance 22(3), 1–35 (2018)

    Google Scholar 

  • Hamilton, J.D.: A New approach to the economic analysis of nonstationary time series and the business cycle. Econometrica 57(2), 357–384 (1989)

    Article  Google Scholar 

  • Hardy, M.: A Regime-switching model of long-term stock returns. N Am Actuar J 5(2), 41–53 (2001)

    Article  Google Scholar 

  • Hawkes, A.: Point sprectra of some mutually exciting point processes. J R Stat Soc Ser B 33, 438–443 (1971)

    Google Scholar 

  • Hawkes, A.: Spectra of some self-exciting and mutually exciting point processes. Biometrika 58, 83–90 (1971)

    Article  Google Scholar 

  • Hawkes, A., Oakes, D.: A cluster representation of a self-exciting process. J Appl Probab 11, 493–503 (1974)

    Article  Google Scholar 

  • Kalman, R.E.: A new approach to linear filtering and prediction problems. J Basic Eng 82(1), 35–45 (1960)

    Article  Google Scholar 

  • Kelly, F., Yudovina, E.: A Markov model of a limit order book: thresholds, recurrence, and trading strategies. Math Oper Res (2017). https://doi.org/10.1287/moor.2017.0857

    Google Scholar 

  • Kou, S.G.: A jump-diffusion model for option pricing. Manage Sci 48(8), 1086–1101 (2002)

    Article  Google Scholar 

  • Kyle, A.: Continuous auctions and insider trading. Econometrica 53, 1315–1335 (1985)

    Article  Google Scholar 

  • Lee, S., Mykland, P.: Jumps in financial markets: a new nonparametric test and jump dynamics. Rev Financ Stud 21(6), 2535–2563 (2008)

    Article  Google Scholar 

  • Maheu, J., Chan, W.: Conditional jump dynamics in stock market returns. J Bus Econom Stat 20(3), 377–389 (2002)

    Article  Google Scholar 

  • Maheu, J., McCurdy, T.: News arrival, jump dynamics and volatility components for individual stock returns. J Finance 59, 755–793 (2004)

    Article  Google Scholar 

  • Mandelbrot, B.: The Variation of certain speculative prices. J Bus 36, 394–419 (1963)

    Article  Google Scholar 

  • Merton, R.C.: Option pricing when underlying stock returns are discontinuous. J Financ Econom 3, 125–144 (1976)

    Article  Google Scholar 

  • Nakajima, J.: Stochastic volatility model with regime-switching skewness in heavy-tailed errors for exchange rate returns. Stud Nonlinear Dyn Econom 17(5), 499–520 (2013)

    Google Scholar 

  • Osborne, M.F.M.: Brownian motion in the stock market. Oper Res 7, 145–173 (1959)

    Article  Google Scholar 

  • Protter, P.: Stochastic Integration and Differential Equations, 2nd edn. Springer, Heidelberg (2005)

    Book  Google Scholar 

  • Schoutens, W.: Lévy Processes in Finance. Pricing of Financial Derivatives. Wiley, New York (2003)

    Book  Google Scholar 

  • Shen, Y., Siu, T.K., Fan, K.: Option valuation under a double regime-dwitching model. J Fut Markets 34(5), 451–478 (2014)

    Article  Google Scholar 

  • Siu, T.K.: A self-exciting threshold jump-diffusion model for option valuation. Insur Math Econom 69, 168–193 (2016)

    Article  Google Scholar 

  • Wang, T., Bebbington, M., Harte, D.: Markov-modulated Hawkes process with stepwise decay. Ann Inst Stat Math 64, 521–544 (2012)

    Article  Google Scholar 

Download references

Acknowledgements

Donatien Hainaut thanks for its support the Chair “Data Analytics and Models for insurance” of BNP Paribas Cardif, hosted by ISFA (Université Claude Bernard, Lyon) and managed by the “Fondation Du Risque”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Donatien Hainaut.

Appendices

Appendix A: Proofs

Proof of proposition 2.1

As \(\mathcal {F}_{0}\subset \mathcal {F}_{0}\vee \mathcal {G}_{t}\), the expectation of \(\lambda _{t}\) is nested as follows:

$$\begin{aligned} \mathbb {E}(\lambda _{t}|\mathcal {F}_{0})= & {} \mathbb {E}\left( \mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0} \vee \mathcal {G}_{t}\right) |\mathcal {F}_{0}\right) \end{aligned}$$

and if we remember the expression (9) for the intensity, we infer that

$$\begin{aligned} \mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right)= & {} \lambda _{0}-\alpha \int _{0}^{t}e^{\alpha (s-t)}\left( \lambda _{0} -\theta _{s}\right) ds+\mathbb {E}\left( \int _{0}^{t}\eta e^{\alpha (s-t)}dL_{s}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) . \end{aligned}$$

Let us denote by \(\chi (t,l)\), the random measure of \(L_{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\). \(\chi (t,l)\) is such that \(L_{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}=\int _{0}^{t}\int _{0}^{\infty }\chi (ds,dl)\) and \(d\left( L_{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) =\int _{0}^{\infty }\chi (ds,dl)\). Given that the integrand \(\eta e^{\alpha (s-t)}\) is time integrable on \(\mathbb {R}^{+}\), independent from the random measure \(\chi \), we infer that

$$\begin{aligned} \mathbb {E}\left( \int _{0}^{t}\eta e^{\alpha (s-t)}dL_{s}|\mathcal {F}_{0} \vee \mathcal {G}_{t}\right)= & {} \int _{0}^{t}\int _{0}^{\infty }\eta e^{\alpha (s-t)} \chi (ds,dl)\\= & {} \int _{0}^{t}\eta e^{\alpha (s-t)}\int _{0}^{\infty }\chi (ds,dl)\\= & {} \int _{0}^{t}\eta e^{\alpha (s-t)}\mathbb {E}\left( dL_{s}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) . \end{aligned}$$

According to the Itô’s lemma for semi-martingale and by construction of \(L_{s}\), \(L_{s}\) is solution of the following SDE:

$$\begin{aligned} dL_{s}&=\int _{0}^{\infty }(L_{s}+|z|-L_{s})\chi (ds,dl)=|J|dN_{s}. \end{aligned}$$

Given that the jump J is independent from \(N_{t}\), the expectation of \(dL_{s}\) with respect to \(\mathcal {F}_{0}\vee \mathcal {G}_{t}\) is then

$$\begin{aligned} \mathbb {E}\left( dL_{s}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right)&=\mathbb {E}\left( \left| J\right| \right) \times \mathbb {E}\left( dN_{s}\,|\, \mathcal {F}_{0}\vee \mathcal {G}_{s}\right) \end{aligned}$$
(33)

for \(s\le t\). Furthermore, conditionally to the filtration of \(\lambda _{t}\) denoted \(\left( \mathcal {L}_{t}\right) _{t\ge 0}\), \(\left( N_{t}\right) _{t\ge 0}\) is a Poisson random variable of intensity and expectation equal to \(\int _{0}^{t-}\lambda _{u}du\). Using nested expectations allows us to infer that

$$\begin{aligned} \mathbb {E}\left( dN_{s}\,|\,\mathcal {F}_{0}\vee \mathcal {G}_{s}\right) \,= & {} \mathbb {E}\left( \mathbb {E}\left( dN_{s}\,|\,\mathcal {F}_{0} \vee \mathcal {G}_{s}\vee \mathcal {L}_{s}\right) \,|\,\mathcal {F}_{0} \vee \mathcal {G}_{s}\right) \nonumber \\= & {} \mathbb {E}\left( \lambda _{s-}\,|\,\mathcal {F}_{0}\vee \mathcal {G}_{s}\right) . \end{aligned}$$
(34)

Combining (33) and (34) leads then to

$$\begin{aligned} \mathbb {E}\left( dL_{s}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right)&=\mathbb {E}\left( \left| J\right| \right) \times \mathbb {E}\left( \lambda _{s-}\,|\,\mathcal {F}_{0}\vee \mathcal {G}_{s}\right) ds \end{aligned}$$

and

$$\begin{aligned} \mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right)= & {} \lambda _{0}-\alpha \int _{0}^{t}e^{\alpha (s-t)}\\&\times \,\left( \lambda _{0}-\theta _{s}\right) ds+\eta \mathbb {E}\left( |J|\right) \int _{0}^{t}e^{\alpha (s-t)}\mathbb {E}\left( \lambda _{s-}\,|\,\mathcal {F}_{0}\vee \mathcal {G}_{s}\right) ds \end{aligned}$$

If we derive this last expression with respect to time, we find that \(\mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) \) is solution of an ordinary differential equation (ODE):

$$\begin{aligned} \frac{\partial }{\partial t}\mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right)= & {} -\,\alpha \left( \lambda _{0}-\theta _{t}\right) +\alpha ^{2}\int _{0}^{t}e^{\alpha (s-t)}\left( \lambda _{0}-\theta _{s}\right) ds+\eta \mathbb {E}\left( |J|\right) \mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) \nonumber \\&-\,\alpha \eta \mathbb {E}\left( |J|\right) \int _{0}^{t}e^{-\alpha (t-s)}\mathbb {E}\left( \lambda _{s-}\,|\,\mathcal {F}_{0}\vee \mathcal {G}_{s}\right) ds\nonumber \\= & {} \left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) \mathbb {E}\left( \lambda _{t}| \mathcal {F}_{0}\vee \mathcal {G}_{t}\right) +\alpha \theta _{t}. \end{aligned}$$
(35)

The solution of this ODE is the following function:

$$\begin{aligned} \mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right)= & {} \int _{0}^{t}\alpha \theta _{s}e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) (t-s)}ds+\lambda _{0}\,e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t}, \end{aligned}$$

and the \(\mathcal {F}_{0}\) conditional expectation is given by:

$$\begin{aligned} \mathbb {E}\left( \mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) |\mathcal {F}_{0}\right)= & {} \mathbb {E}\left( \int _{0}^{t}\alpha \theta _{s}e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) (t-s)}ds+\lambda _{0}\,e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t}|\mathcal {F}_{0}\right) \nonumber \\= & {} \alpha \int _{0}^{t}e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) (t-s)}\mathbb {E}\left( \theta _{s}|\mathcal {F}_{0}\right) ds+\lambda _{0}\,e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t}.\nonumber \\ \end{aligned}$$
(36)

The expected level of mean reversion at time s is the sum of \(\bar{\theta }_{j}\) for \(j=1\) to N weighted by the probabilities of transition:

$$\begin{aligned} \mathbb {E}\left( \theta _{s}|\mathcal {F}_{0}\right)= & {} \delta (0)'\,\exp \left( Q_{0}s\right) \,\bar{\theta } \end{aligned}$$

and if I is the \(N\times N\) identity matrix, the integral in (36) may be rewritten as follows:

$$\begin{aligned}&\int _{0}^{t}e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) (t-s)}\mathbb {E}\left( \theta _{s}|\mathcal {F}_{0}\right) ds=e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t}\int _{0}^{t}e^{-\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) s}\delta (0)'\,\exp \left( Q_{0}s\right) \,\bar{\theta }ds\\&\qquad =e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t}\int _{0}^{t}\delta (0)'\,\exp \left( \left( Q_{0}-I\,(\eta \mathbb {E}\left( |J|\right) -\alpha )\right) s\right) \,\bar{\theta }ds\\&\qquad =e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t}\left[ \delta (0)'\,\left( Q_{0}-I(\eta \mathbb {E}\left( |J|\right) -\alpha )\right) ^{-1}\exp \left( \left( Q_{0}-I(\eta \mathbb {E}\left( |J|\right) -\alpha )\right) s\right) \,\bar{\theta }\right] _{s=0}^{s=t}\\&\qquad =\delta (0)'\,\left( Q_{0}-I(\eta \mathbb {E}\left( |J|\right) -\alpha )\right) ^{-1}\left[ \exp \left( Q_{0}t\right) -\exp \left( I\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t\right) \right] \,\bar{\theta } \end{aligned}$$

\(\square \)

Proof of Corollary 2.2

Conditionally to \(\mathcal {G}_{t}\), this expectation is equal to

$$\begin{aligned}&\mathbb {E}\left( \theta _{t}\lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) =\theta _{t}\mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) \\&\quad =\theta _{t}\int _{0}^{t}\alpha \theta _{s}e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) (t-s)}ds+\lambda _{0}\theta _{t}\,e^{\left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) t}. \end{aligned}$$

We infer from this relation that the derivative with respect to time is given by

$$\begin{aligned} \frac{\partial }{\partial t}\mathbb {E}\left( \theta _{t}\lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right)= & {} \theta _{t}\frac{\partial }{\partial t}\mathbb {E}\left( \lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) \\= & {} \left( \eta \mathbb {E}\left( |J|\right) -\alpha \right) \mathbb {E}\left( \theta _{t}\lambda _{t}|\mathcal {F}_{0}\vee \mathcal {G}_{t}\right) +\alpha \left( \theta _{t}\right) ^{2}. \end{aligned}$$

and we can conclude that \(\mathbb {E}\left( \theta _{t}\lambda _{t}|\mathcal {F}_{0}\right) \) is well given by Eq. (12). \(\square \)

Proof of proposition 2.3

Let us introduce the following notations: \(f(t,N_{t},\lambda _{t},\delta _{t})=\lambda _{t}^{2}\), according to the Itô’s lemma for semi-martingale, the infinitesimal generator of this function is given by:

$$\begin{aligned} \mathcal {A}f(t,N_{t},\lambda _{t},\delta _{t})= & {} \alpha (\theta _{t}-\lambda _{t})\,2\lambda _{t}+\lambda _{t} \int _{-\infty }^{+\infty }\left( \,\lambda _{t}+\eta \,|z|\right) ^{2} -\lambda _{t}^{2}\,d\nu (z)\\= & {} \alpha (\theta _{t}-\lambda _{t})\,2\lambda _{t}+\lambda _{t} \int _{-\infty }^{+\infty }\left( \,\lambda _{t}^{2}+2\eta \,|z| \lambda _{t}+\eta ^{2}|z|^{2}\right) -\lambda _{t}^{2}\,d\nu (z)\\= & {} 2\alpha \theta _{t}\lambda _{t}+2\left( \eta \,\mathbb {E}\left( |J|\right) -\alpha \right) \lambda _{t}^{2}+\eta ^{2}\lambda _{t}\mathbb {E}\left( |J|^{2}\right) \end{aligned}$$

On the other hand, we can prove that

$$\begin{aligned} \mathbb {E}\left( f(t,N_{t},\lambda _{t},\delta _{t})|\mathcal {F}_{0}\right)= & {} f(0,N_{0},\lambda _{0},\delta _{0})+\mathbb {E}\left( \int _{0}^{t} \mathcal {A}f(s,N_{s},\lambda _{s},\delta _{s})ds|\mathcal {F}_{0}\right) \nonumber \\= & {} f(0,N_{0},\lambda _{0},\delta _{0})+\int _{0}^{t}\mathbb {E} \left( \mathcal {A}f(s,N_{s},\lambda _{s},\delta _{s})|\mathcal {F}_{0}\right) ds. \end{aligned}$$
(37)

The derivative of this expectation with respect to time is equal to its expected infinitesimal generator:

$$\begin{aligned} \frac{\partial }{\partial t}\mathbb {E}\left( f(t,N_{t},\lambda _{t},\delta _{t})|\mathcal {F}_{0}\right)= & {} \mathbb {E}\left( \mathcal {A} f(t,N_{t},\lambda _{t},\theta _{t})|\mathcal {F}_{0}\right) , \end{aligned}$$
(38)

that allows us to infer Eq. (13). It is easy to check that the right term of Eq. (38) is linear and Lipschitz. Then, according to the theorem of Cauchy-Lipschitz the solution exists and is unique. \(\square \)

Proof of proposition 2.4

If we note \(f(t,N_{t},\lambda _{t},\delta _{t})=\mathbb {E}\left( \omega ^{N_{s}}\,|\,\mathcal {F}_{t}\right) \), f is solution of an Itô’s equation for semi martingale. If \(\delta _{t}=e_{i}\) then:

$$\begin{aligned} 0= & {} f_{t}+\alpha (\bar{\theta }_{i}-\lambda _{t})\,f_{\lambda }+\lambda _{t}\int _{-\infty }^{+\infty }f\left( t,N_{t}+1,\,\lambda _{t}+\eta \,|z|,\,e_{i}\right) -f(.)\,d\nu (z)\nonumber \\&+\sum _{j\ne i}^{N}q_{i,j}\left( f(t,N_{t},\lambda _{t},e_{j})-f(t,N_{t},\lambda _{t},e_{i})\right) \end{aligned}$$
(39)

If we assume that f is an exponential affine function of \(\lambda _{t}\) and \(N_{t}\):

$$\begin{aligned} f= & {} exp\left( A(t,s,e_{i})+B(t,s)\lambda _{t}+C(t,s)N_{t}\right) , \end{aligned}$$

where \(A(t,s,e_{i})\) for \(i=1\) to N, B(ts), C(ts) are time dependent functions, the partial derivatives of f are given by:

$$\begin{aligned}&f_{t} = \left( \frac{\partial }{\partial t}A(t,s,e_{j})+\frac{\partial }{\partial t}B(t,s)\lambda _{t}+\frac{\partial }{\partial t}C(t,s)N_{t}\right) f,\\&f_{\lambda }=B(t,s)f \end{aligned}$$

The integral in Eq. (39) is rewritten as follows:

$$\begin{aligned}&\int _{-\infty }^{+\infty }f(t,N_{t}+1,\lambda _{t}+\eta |z|,e_{j})-f(t,N_{t},\lambda _{t},e_{i})\,d\nu (z)\\&\quad =\int _{0}^{+\infty }exp\left( A(t,s,e_{j})+B(t,s)\lambda _{t}+C(t,s)N_{t}\right) \left( e^{B(t,s)\eta |z|+C(t,s)}-1\right) \,d\nu (z)\\&\quad =f\left[ e^{C}\psi \left( 0\,,\,B(t,s)\eta \right) -1\right] . \end{aligned}$$

As \(q_{ii}=-\sum _{i\ne j}^{N}q_{i,j}\), the last term of Eq. (39) becomes

$$\begin{aligned} \sum _{j\ne i}^{N}q_{i,j}\left( f(t,N_{t},\lambda _{t},e_{j})-f(t,N_{t},\lambda _{t},e_{i})\right)= & {} \sum _{j=1}^{N}q_{i,j}f(t,N_{t},\lambda _{t},e_{j}). \end{aligned}$$

Then

$$\begin{aligned}&0=\left( \frac{\partial }{\partial t}A+\frac{\partial }{\partial t}B\,\lambda _{t}+\frac{\partial }{\partial t}C\,N_{t}\right) e^{A(t,s,e_{i})}+\alpha (\bar{\theta }_{i}-\lambda _{t})\,B\,e^{A(t,s,e_{i})}\\&\qquad +\lambda _{t}\left[ e^{C}\psi \left( 0\,,\,B(t,s)\eta \right) -1\right] e^{A(t,s,e_{i})}+\sum _{j=1}^{N}q_{i,j}e^{A(t,s,e_{j})}\,, \end{aligned}$$

from which we deduce that \(C(t,s)=\ln \omega \). Regrouping terms allows to infer that

$$\begin{aligned}&0=\left( \frac{\partial }{\partial t}A\right) e^{A(t,s,e_{i})}+\alpha \,\bar{\theta }_{i}\,B\,e^{A(t,s,e_{i})}+\sum _{j=1}^{N}q_{i,j}e^{A(t,s,e_{j})}\\&\qquad +\lambda _{t}\left( \frac{\partial }{\partial t}B-\alpha B+\left[ \omega \psi \left( 0\,,\,B(t,s)\eta \right) -1\right] \right) e^{A(t,s,e_{i})} \end{aligned}$$

or that

$$\begin{aligned} \frac{\partial }{\partial t}B= & {} \alpha B-\left[ \omega \psi \left( 0\,,\,B(t,s)\eta \right) -1\right] \\ \left( \frac{\partial }{\partial t}A\right) e^{A(t,s,e_{i})}= & {} -\alpha \,\bar{\theta }_{i}\,B\,e^{A(t,s,e_{i})}-\sum _{j=1}^{N}q_{i,j}e^{A(t,s,e_{j})} \end{aligned}$$

If we define \(\tilde{A}(t,s,e_{i})=e^{A(t,s,e_{i})}\), the first equations can finally be put in matrix form as:

$$\begin{aligned} \frac{\partial \tilde{A}(t)}{\partial t}+\left( \text {diag}\left( \alpha \bar{\theta }B\right) +Q_{0}\right) \tilde{A}(t)=0, \end{aligned}$$
(40)

\(\square \)

Proof of proposition 2.5

If we note \(f(t,X_{t},\lambda _{t},\delta _{t})=\mathbb {E}\left( e^{\omega _{1}X_{s}+\omega _{2}\lambda _{s}}\,|\,\mathcal {F}_{t}\right) \), f is solution of an Itô’s equation for semi-martingale. If \(\delta _{t}=e_{i}\) then:

$$\begin{aligned}&0=f_{t}+f_{X}\left( \bar{\mu }_{i}-\frac{\bar{\sigma }_{i}^{2}}{2}-\lambda _{t}\,\mathbb {E}\left( e^{J}-1\right) \right) +f_{XX}\frac{\bar{\sigma }_{i}^{2}}{2}+\alpha (\bar{\theta }_{i}-\lambda _{t})\,f_{\lambda }\nonumber \\&\qquad + \lambda _{t}\int _{-\infty }^{+\infty }f\left( t,X_{t} +z,\,\lambda _{t}+\eta \,|z|,\,e_{i}\right) -f(.)\,\nu (dz)\nonumber \\&\qquad +\sum _{j\ne i}^{N}q_{i,j}\left( f(t,X_{t},\lambda _{t},e_{j})-f(t,X_{t},\lambda _{t},e_{i})\right) \end{aligned}$$
(41)

If we assume that f is an exponential affine function of \(\lambda _{t}\) and \(X_{t}\):

$$\begin{aligned} f= & {} exp\left( A(t,s,e_{i})+B(t,s)\lambda _{t}+C(t,s)X_{t}\right) , \end{aligned}$$

where \(A(t,s,e_{i})\) for \(i=1\) to N, B(ts), C(ts) are time dependent functions, the partial derivatives of f are given by:

$$\begin{aligned} f_{t}= & {} \left( \frac{\partial }{\partial t}A(t,s,e_{j})+\frac{\partial }{\partial t}B(t,s)\lambda _{t}+\frac{\partial }{\partial t}C(t,s)X_{t}\right) f,\\ f_{X}= & {} C(t,s)f\,f_{XX}=C(t,s)^{2}f \; f_{\lambda }=B(t,s)f \end{aligned}$$

and the integrand in Eq. (41) is rewritten as follows:

$$\begin{aligned} \int _{-\infty }^{+\infty }f\left( t,X_{t}+z,\,\lambda _{t}+\eta \,|z|,e_{i}\right) -f(.)\,\nu (dz)= & {} f\left[ \psi \left( C(t,s)\,,\,B(t,s)\eta \right) -1\right] . \end{aligned}$$

As \(q_{ii}=-\sum _{i\ne j}^{N}q_{i,j}\), the last term of the Itô equation is also equal to

$$\begin{aligned} \sum _{j\ne i}^{N}q_{i,j}\left( f(t,X_{t},\lambda _{t},e_{j})-f(t,X_{t},\lambda _{t},e_{i})\right)= & {} \sum _{j=1}^{N}q_{i,j}f(t,X_{t},\lambda _{t},e_{j}) \end{aligned}$$

injecting these expressions into Eq. (41), leads to the following relation:

$$\begin{aligned} 0= & {} \left( \frac{\partial }{\partial t}A+\frac{\partial }{\partial t}B\,\lambda _{t}+\frac{\partial }{\partial t}C\,X_{t}\right) e^{A(t,s,e_{i})}+C\,\left( \bar{\mu }_{i}-\frac{\bar{\sigma }_{i}^{2}}{2}-\lambda _{t}\mathbb {E}\left( e^{J}-1\right) \right) e^{A(t,s,e_{i})}\\&+C^{2}\,\frac{\bar{\sigma }_{i}^{2}}{2}e^{A(t,s,e_{i})}+\alpha (\bar{\theta }_{i}-\lambda _{t})Be^{A(t,s,e_{i})}+\lambda _{t} \left[ \psi \left( C\,,\,B\eta \right) -1\right] e^{A(t,s,e_{i})}+\sum _{j=1}^{N}q_{i,j}e^{A(t,s,e_{j})}\,, \end{aligned}$$

from which we deduce that \(C(t,s)=\omega _{1}\). Regrouping terms allows to infer that A and B are solutions of a system of ODE’s:

$$\begin{aligned} 0= & {} \left( \frac{\partial }{\partial t}A+\frac{\partial }{\partial t} B\,\lambda _{t}\right) e^{A(t,s,e_{i})}+\omega _{1}\,\left( \bar{\mu }_{i} -\frac{\bar{\sigma }_{i}^{2}}{2}-\lambda _{t}\mathbb {E}\left( e^{J}-1\right) \right) e^{A(t,s,e_{i})}\\&+\omega _{1}^{2}\,\frac{\bar{\sigma }_{i}^{2}}{2}e^{A(t,s,e_{i})} +\alpha (\bar{\theta }_{i}-\lambda _{t})Be^{A(t,s,e_{i})} +\lambda _{t}\left[ \psi \left( \omega _{1}\,,\,B\eta \right) -1\right] e^{A(t,s,e_{i})}\\&+\sum _{j=1}^{N}q_{i,j}e^{A(t,s,e_{j})}\,, \end{aligned}$$

From this last relation, we deduce that

$$\begin{aligned} {\left\{ \begin{array}{ll} 0= &{} \frac{\partial }{\partial t}A\,e^{A(t,s,e_{i})}+\omega _{1}\,\left( \bar{\mu }_{i}-\frac{\bar{\sigma }_{i}^{2}}{2}\right) e^{A(t,s,e_{i})}+\omega _{1}^{2}\,\frac{\bar{\sigma }_{i}^{2}}{2}e^{A(t,s,e_{i})}\\ &{} +\alpha \bar{\theta }_{i}B\,e^{A(t,s,e_{i})}+\sum _{j=1}^{N}q_{i,j}e^{A(t,s,e_{j})}\;for\,i=1\ldots N,\\ 0= &{} \frac{\partial }{\partial t}B-\alpha B-\omega _{1}\,\mathbb {E}\left( e^{J}-1\right) +\left[ \psi \left( \omega _{1}\,,\,B\eta \right) -1\right] \end{array}\right. } \end{aligned}$$

If we define \(\tilde{A}(t,s)=(e^{A(t,s,e_{i})})_{i=1,\ldots ,N}\), the first equations can finally be put in matrix form as:

$$\begin{aligned} \frac{\partial \tilde{A}(t,s)}{\partial t}+\left( \text {diag}\left( \omega _{1}\,\left( \bar{\mu }-\frac{\bar{\sigma }^{2}}{2}\right) +\omega _{1}^{2}\,\frac{\bar{\sigma }^{2}}{2}+\alpha \bar{\theta }B\right) +Q_{0}\right) \tilde{A}(t,s)=0, \end{aligned}$$
(42)

\(\square \)

Proof of proposition

From the previous results, we know that B(ts) is solution of the ODE:

$$\begin{aligned} \frac{\partial }{\partial t}B=\alpha B+\omega _{1}\left( \psi \left( 1,0\right) -1\right) -\left[ \psi \left( \omega _{1}\,,\,B\,\eta \right) -1\right] \end{aligned}$$

with the terminal condition: \(B(s,s)=\omega _{2}\). If we set \(B(t,s)=C(s-t)\) and \(\tau =s-t\). Then \(\frac{\partial }{\partial t}B=\frac{\partial }{\partial \tau }C\frac{\partial \tau }{\partial t}=-\frac{\partial }{\partial \tau }C\) and

$$\begin{aligned} \frac{\partial }{\partial \tau }C(\tau )= & {} -\alpha C(\tau )+\psi \left( \omega _{1}\,,\,C(\tau )\,\eta \right) -\underbrace{\left[ \omega _{1}\left( \psi \left( 1,0\right) -1\right) +1\right] }_{\beta (\omega _{1})}\nonumber \\= & {} -\alpha C(\tau )+\psi \left( \omega _{1}\,,\,C(\tau )\,\eta \right) -\beta (\omega _{1}) \end{aligned}$$
(43)

The left hand side is denoted h(C). Due to the convexity of \(\psi \left( .\right) \) there is only one point \(u^{*}\) such that \(h(u)=0\). This equation is indeed equivalent to

$$\begin{aligned} \psi \left( \omega _{1}\,,\,u\,\eta \right)= & {} \beta \left( \omega _{1}\right) +\alpha \,u \end{aligned}$$

We rewrite the Eq. (43) as follows,

$$\begin{aligned} \frac{dC}{-\beta -\alpha C+\psi \left( \omega _{1}\,,\,C\,\eta \right) }= & {} d\tau . \end{aligned}$$

As \(C(0)=\omega _{2}\), by direct integration, we have that

$$\begin{aligned} \int _{\omega _{2}}^{C}\frac{du}{-\beta -\alpha u+\psi \left( \omega _{1}\,,\,\eta \,u\right) }= & {} \tau \end{aligned}$$

with \(C\in [\omega _{2},u^{*})\) or \(C\in [u^{*},\omega _{2})\). Remark that if \(C=u^{*}\) then \(\tau =+\infty \) as the numerator converges to zero. If we define the function on the left hand side as

$$\begin{aligned} F_{\omega _{1}}(C):= & {} \int _{\omega _{2}}^{C}\frac{du}{-\beta -\alpha u+\psi \left( \omega _{1}\,,\,\eta \,u\right) } \end{aligned}$$

then \(F_{\omega _{1}}(C)=\tau \) and \(C=F_{\omega _{1}}^{-1}(\tau )\) or \(B(t)=F_{\omega _{1}}^{-1}(s-t)\). \(\square \)

Appendix B: Particle filter

The structure of the particle filter algorithm is the following:

  1. 1.

    Initial step: draw M values of \(v_{0}^{(i)}\) for \(i=1,\ldots ,M\), from an initial distribution \(p(v_{0})\)

  2. 2.

    For \(j=1\,:\,T\)

    Prediction step: draw a sample of \(\Delta L_{j}^{(i)}\) and \(\delta _{j}^{(i)}\) and update \(\lambda _{j}^{(i)}\), \(\mu _{j}^{b\,(i)}\)\(\sigma _{j}^{(i)}\) , \(\theta _{j}^{(i)}\) using the relations

    $$\begin{aligned}&\lambda _{j}^{(i)}=\lambda _{j-1}^{(i)}+\alpha (\theta _{j-1}^{(i)}-\lambda _{j-1}^{(i)})\Delta +\eta {\varvec{\Delta }}L_{j}^{(i)}\\&\mu _{j}^{(i)}=\delta _{j}^{(i)}\bar{\mu }\quad \sigma _{j}^{(i)}=\delta _{j}^{(i)}\bar{\sigma }\quad \theta _{j}^{(i)}=\delta _{j}^{(i)}\bar{\theta } \end{aligned}$$

    Correction step: the particle \(v_{j}^{(i)}\) has a probability of occurrence equal to \(w_{j}^{(i)}=\frac{p(x_{j}\,|\,v_{j})}{\sum _{i=1:M}p(x_{j}\,|\,v_{j})}\) where

    $$\begin{aligned} p(x_{j}\,|\,v_{j}^{(i)})&\sim \mathcal {N}\left( \left( \mu _{j}^{(i)}-\frac{\sigma _{j}^{(i)2}}{2}-\lambda _{j}^{(i)}\mathbb {E}\left( e^{J}-1\right) \right) \Delta -{\varvec{\Delta }}L_{j}^{'(i)}\,,\,\sigma _{j}^{(i)}\sqrt{\Delta }\right) \end{aligned}$$

    Resampling step: resample with replacement M particles according to the importance weights \(w_{j}^{(i)}\). The new importance weights are set to \(w_{j}^{(i)}=\frac{1}{M}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hainaut, D., Moraux, F. A switching self-exciting jump diffusion process for stock prices. Ann Finance 15, 267–306 (2019). https://doi.org/10.1007/s10436-018-0340-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10436-018-0340-5

Keywords

JEL Classification

Navigation