Skip to main content
Log in

Option Pricing Model Biases: Bayesian and Markov Chain Monte Carlo Regression Analysis

  • Published:
Computational Economics Aims and scope Submit manuscript

Abstract

We investigate systematic and unsystematic option pricing biases in (a) pure jump Lévy, (b) jump-diffusion, (c) stochastic volatility, and (d) GARCH models applied to the Black–Scholes–Merton model. We use options data for trades on the S&P500 index from the CBOE. In addition to standard ordinary least square regression, we employ Bayesian regression and Markov Chain Monte Carlo regression to investigate the moneyness and maturity biases of the models. We demonstrate the usefulness of these advanced methodologies as compared to the benchmark techniques.

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

Similar content being viewed by others

Notes

  1. The Bayesian approach differs fundamentally from the classical, frequentist approach in that the former views parameters as random variables whereas the latter views them as fixed characteristics of the population.

  2. A comprehensive survey of Lévy processes in finance can be found in Schoutens (2003), Cont and Tankov (2004), and Wu (2006). Accessible theoretical treatments of Lévy processes can be found in Kyprianou (2006), Sato (1999), and Applebeum (2004).

  3. For more details see Cont and Tankov (2004) and Kyprianou (2006).

References

  • Applebeum, D. (2004). Lévy processes and stochastic calculus. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Bakshi, G., Cao, Z., & Chen, Z. (1997). Empirical performance of alternative option pricing models. Journal of Finance, 52, 2003–2049.

    Article  Google Scholar 

  • Black, F., & Scholes, M. (1973). The pricing of options and corporate liabilities. Journal of Political Economy, 81, 637–659.

    Article  Google Scholar 

  • Carr, P., Geman, H., Madan, D. B., & Yor, M. (2002). The fine structure of asset returns: An empirical investigation. Journal of Business, 75(2), 305–332.

    Article  Google Scholar 

  • Carr, P., & Madan, D. (1999). Option valuation using the fast Fourier transform. Journal of Computational Finance, 2, 61–73.

    Article  Google Scholar 

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

    Google Scholar 

  • Cox, J. C., Ingersoll, J. E., & Ross, S. A. (1985). A theory of the term structure of interest rates. Econometrica, 53, 385–408.

    Article  Google Scholar 

  • Duffy, D. J. (2006). Finite difference methods in financial engineering: A partial differential equation approach. Hoboken: Wiley.

    Book  Google Scholar 

  • Geman, H. (2002). Pure jump Lévy processes for asset price modelling. Journal of Banking & Finance, 26, 1297–1316.

    Article  Google Scholar 

  • Heston, S. L. (1993). A closed form solutions for options with stochastic volatility with applications to bond and currency options. Review of Financial Studies, 6, 327–343.

    Article  Google Scholar 

  • Heston, S. L., & Nandi, S. (2000). A closed form GARCH option valuation model. Review of Financial Studies, 13, 585–625.

    Article  Google Scholar 

  • Kou, S. (2002). A jump diffusion model for option pricing. Management Science, 48, 1086–1101.

    Article  Google Scholar 

  • Kyprianou, A. (2006). Introductory lectures on fluctuations of Lévy processes with applications. Berlin: Springer.

    Google Scholar 

  • Merton, R. C. (1973). Theory of rational option pricing. Bell Journal of Economics and Management Science, 4, 141–183.

    Article  Google Scholar 

  • Ramezani, C. A., & Zeng, Y. (1999). Maximum likelihood estimation of asymmetric jump-diffusion processes: Application to security prices. Working paper, Department of Statistics, University of Winsconsin, Madison, WI.

  • Sato, K. (1999). Lévy processes and infinitely divisible distributions. Cambridge: Cambridge University Press.

    Google Scholar 

  • Schouten, W. (2003). Lévy processes in finance: Pricing financial derivatives. Hoboken: Wiley.

    Book  Google Scholar 

  • Shiryaev, A. N. (1999). Essentials of stochastic finance: Facts, models, theory. Singapore: World Scientific.

    Book  Google Scholar 

  • Wu, L. (2006). Modelling financial security returns using Lévy processes. Working paper, Zicklin School of Business, City University of New York.

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sharif Mozumder.

Additional information

Publisher's Note

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

Appendix: Fundamental Models of Previous Studies and Their Dynamics

Appendix: Fundamental Models of Previous Studies and Their Dynamics

This Appendix provides a short description of the models corresponding to different approaches to incorporating skewness and kurtosis.

The Black–Scholes–Merton (BSM) framework uses geometric Brownian motion to model an underlying asset. Thereby, the BSM provides a closed form solution to European options. The price of a European option is determined (Shiryaev 1999):

$$\nu \left( {0,S_{0} } \right) = {\text{e}}^{ - rT} E\left[ {V\left( {S_{T} } \right)} \right]$$
(7)

where V (ST) represents the payoff of the option at expiration, T, r is the continuously-compounded rate of interest, and E denotes the expectation under the risk-neutral probability that is derived from the risk-neutral process:

$$\frac{{{\text{d}}S_{t} }}{{S_{t} }} = r{\text{d}}t + \sigma {\text{d}}B$$
(8)

The BSM considers a normal distribution for log-returns and therefore fails to incorporate any effect resulting from skewness and kurtosis of returns. A more realistic pricing model is obtained by replacing simple Brownian motion with a richer Lévy process that incorporates the empirical evidence of jumps in returns.Footnote 2 In practice, of these Lévy model prices must be computed through (time-consuming) numerical inversion of characteristic functions.

Here, we briefly revisit the other option pricing models where the skewness is characterized quite differently. For pricing options, we use risk-neutral characterization of each model.

1.1 Heston (1993) Stochastic Volatility Model

The Heston (1993) stochastic volatility model assumes a diffusion process for the stock price given, using the usual symbols, by

$$\frac{{{\text{d}}S_{t} }}{{S_{t} }} = \mu {\text{d}}t + \sqrt \vartheta_{t} {\text{d}}B_{t}^{1}$$
(9)

coupled with a CIR process for the volatility \(\sqrt {\vartheta_{t} }\) given by

$${\text{d}}\vartheta_{t} = \kappa \left[ {\theta - \vartheta_{t} } \right]{\text{d}}t + \sigma \sqrt {\vartheta_{t} } {\text{d}}B_{t}$$
(10)

with

$${\text{d}}B_{t} {\text{d}}B_{t}^{1} = \rho {\text{d}}t$$
(11)

The model entails a flexible distributional structure, where the correlation (ρ) between volatility and asset returns controls the level of asymmetry, and the volatility variation coefficient (σ) controls the level of kurtosis. A non-zero correlation incorporates skewness into the model. The solution, expressed in standard symbols, is given by:

$$c_{{{\text{HS}}}} = S_{t} \left( {\frac{1}{2} + \frac{1}{\pi }\int_{0}^{\infty } {{\text{Re}} \left[ {\frac{{K^{ - iz} f_{1} }}{iz}} \right]{\text{d}}z} } \right) - K{\text{e}}^{rt} \left( {\frac{1}{2} + \frac{1}{\pi }\int_{0}^{\infty } {{\text{Re}} \left[ {\frac{{K^{ - iz} f_{2} }}{iz}} \right]{\text{d}}z} } \right)$$
(12)

where \(f_{j} = \exp \left\{ {C_{j} + D_{j} \vartheta + izx} \right\}\) with

$$\begin{aligned} & x = x_{t} = \log \left( {S_{t} } \right) \\ & \vartheta = \vartheta_{t} \\ & C_{j} = irz\left( {T - t} \right) + \frac{\kappa \theta }{{\sigma^{2} }}\left\{ {\left( {b_{j} - i\rho \sigma z + d_{j} } \right)\left( {T - t} \right) - 2\log \left[ {\frac{{1 - g_{j} {\text{e}}^{{d_{j} \left( {T - t} \right)}} }}{{1 - g_{j} }}} \right]} \right\} \\ & D_{j} = \frac{{b_{j} - iz\rho \sigma + d_{j} }}{{\sigma^{2} }}\left[ {\frac{{1 - {\text{e}}^{{d_{j} \left( {T - t} \right)}} }}{{1 - g_{j} {\text{e}}^{{d_{j} \left( {T - t} \right)}} }}} \right] \\ & g_{j} = \frac{{b_{j} - iz\rho \sigma + d_{j} }}{{b_{j} - iz\rho \sigma - d_{j} }} \\ & d_{j} = \sqrt {\left( {iz\rho \sigma - b_{j} } \right)^{2} - \left( {2iu_{j} z - z^{2} } \right)\sigma^{2} } \\ & u_{1} = \frac{1}{2},u_{2} = - \frac{1}{2} \\ & b_{1} = \kappa + \lambda - \rho \sigma ,b_{2} = \kappa + \lambda \\ \end{aligned}$$
(13)

1.2 Heston-Nandi (2000) GARCH Model

For a particular type of GARCH process, Heston and Nandi (2000) provide a closed form pricing formula for European options, where returns are generated by the process:

$$\begin{aligned} & \log \left( {\frac{{S_{t + 1} }}{{S_{t} }}} \right) = r + \lambda \sigma_{t}^{2} + \sigma_{t + 1} z_{t + 1} ;\quad \, z_{t + 1} \sim N\left( {0,1} \right) \\ & \sigma_{t + 1}^{2} = \omega + \alpha \left( {z_{t} - \theta \sigma_{t} } \right)^{2} + \beta \sigma_{t}^{2} \\ \end{aligned}$$
(14)

The variance persistence of return process can be derived as \(\beta + \alpha \theta^{2}\) from the above GARCH characterization, and the process becomes mean-reverting if \(\beta + \alpha \theta^{2} < 1\). Heston and Nandi (2000) show that a risk-neutral characterization can be obtained if \(\lambda = - \tfrac{1}{2}\), and \(\theta^{*} = \theta + \lambda + \tfrac{1}{2}\). It can further be shown that α determines kurtosis and θ determines skewness in the model. Given a moment generating function of the form:

$$f\left( z \right) = S_{t}^{z} \exp \left\{ {A\left( {t;t + T,z} \right) + B\left( {t;t + T,z} \right)\sigma_{t + 1}^{2} } \right\}$$
(15)

the price can be obtained in a closed form GARCH (CFG) as

$$c_{{{\text{CFG}}}} = S_{t} \left( {\frac{1}{2} + \frac{1}{\pi }\int_{0}^{\infty } {{\text{Re}} \left[ {\frac{{K^{ - iz} f^{*} \left( {iz + 1} \right)}}{{izf^{*} \left( 1 \right)}}} \right]{\text{d}}z} } \right) - K{\text{e}}^{rt} \left( {\frac{1}{2} + \frac{1}{\pi }\int_{0}^{\infty } {{\text{Re}} \left[ {\frac{{K^{ - iz} f^{*} \left( i \right)}}{iz}} \right]{\text{d}}z} } \right)$$
(16)

where f* is the risk-neutral version of f, and A(t;t + T,z) and B(t;t + T,z) are given by the recursive relations:

$$A\left( {t;t + T,z} \right) = A\left( {t + 1;t + T,z} \right) + zr + B\left( {t + 1;t + T,z} \right)\omega - \frac{1}{2}\ln \left( {1 - 2\alpha B\left( {t + 1;t + T,z} \right)} \right)$$
(17)
$$B\left( {t;t + T,z} \right) = z\left( {\lambda + \theta } \right) - \frac{1}{2}\theta^{2} + \beta B\left( {t + 1;t + T,z} \right) + \frac{{\tfrac{1}{2}\left( {z - \theta } \right)^{2} }}{{1 - 2\alpha B\left( {t + 1;t + T,z} \right)}}$$
(18)

1.3 Pure Jump Lévy Models

The Lévy models we consider here assume that all possible movements in stock price are caused by jumps. Thus, the Lévy measure of these processes ensures that the frequent arrival of small jumps sufficiently captures the diffusion process. Hence, the process is effectively a pure jump processes (Geman 2002).

Consistent with the Lévy–Kintchine formula, the distribution of \(X_{{\left( {t_{2} - t_{1} } \right)}} = \log \left( {\tfrac{{S_{{t_{2} }} }}{{S_{{t_{1} }} }}} \right)\) is characterized by the characteristic function of an infinitely divisible random variable:

$$E\left[ {{\text{e}}^{{isX_{{\left( {t_{2} - t_{1} } \right)}} }} } \right] = \exp \left\{ {\left( {t_{2} - t_{1} } \right)\left[ {ias - \frac{1}{2}s^{2} b^{2} } \right] + \int_{{\Re \backslash \left\{ 0 \right\}}} {\left[ {{\text{e}}^{isx} - 1 - isx{\rm I}_{{\left\{ { - 1,1} \right\}}} \left( x \right)} \right]\nu \left( {{\text{d}}x} \right)} } \right\}$$
(19)

where t1 can naturally be zero. Scalars \(a,b \in \Re\) and the measure ν satisfies \(\nu \left( {\left\{ 0 \right\}} \right) = 0\) and \(\int_{{\Re \backslash \left\{ 0 \right\}}} {\left( {\left| x \right|^{2} \wedge 1} \right)} \nu \left( {{\text{d}}x} \right) < \infty\), which means that though numerous small jumps may not be integrable, squares of those jumps are always integrable—the condition which helps us extract a square integrable martingale process in the limit. In the case of pure jump processes, b is always zero.Footnote 3 For example, the variance gamma process characterizes the random variable X1 through the parameters (σ, θ, γ) and the Lévy measure:

$$\nu_{vg} \, = \frac{1}{\gamma \left| x \right|}\exp \left( {\frac{x\theta }{{\sigma^{2} }} - \frac{\left| x \right|}{\sigma }\sqrt {\frac{2}{\gamma } + \frac{{\theta^{2} }}{{\sigma^{2} }}} } \right){\text{d}}x$$
(20)

When integrated for jumps of all possible sizes, Eq. (20) implies that the total rate is infinite, i.e. \(\int_{0}^{\infty } {\nu_{vg} } \left( {{\text{d}}x} \right) = \infty\). However, for any \(\varepsilon > 0\), we have \(\int_{\varepsilon }^{\infty } {\nu_{vg} } \left( {{\text{d}}x} \right) < \infty\), implying that sufficiently small jumps are numerous while jumps exceeding any threshold \(\varepsilon > 0\) are finite, arriving in compound Poisson fashion. The Lévy measure, when used in Eq. (21) with a = b = 0, yields the following closed form characteristic function of the process Xt:

$$\Phi_{{X_{t} }} \left( s \right) = \left( {\frac{1}{{1 - is\theta \gamma + \tfrac{1}{2}s^{2} \sigma^{2} \gamma }}} \right)^{{\frac{t}{\gamma }}}$$
(21)

For this pure jump Lévy model the skewness of log returns over an interval of length one, is given by

$${\text{skew}}\left( {X_{1} } \right) = \frac{{\theta \gamma \left( {3\sigma^{2} + 2\gamma \sigma^{2} } \right)}}{{\left( {\sigma^{2} + \gamma \theta^{2} } \right)^{\frac{3}{2}} }}$$
(22)

The risk-neutral version of the characteristic function (21) required in the formula of Carr and Madan (1999) for pricing the options is given by

$$\Phi_{{X_{t} }}^{rn} \left( s \right) \, = \exp \left\{ {i\left[ {r + \frac{1}{\gamma }\ln \left( {1 - \theta \gamma - \frac{1}{2}\sigma^{2} \gamma } \right)} \right]st - \frac{t}{\gamma }\ln \left( {1 - is\theta \gamma + \frac{1}{2}s^{2} \sigma^{2} \gamma } \right)} \right\}$$
(23)

This risk-neutral form results from mean-correction of the drift part, thereby introducing a drift to a driftless process (Schouten 2003).

1.4 A Jump-Diffusion Model

We choose Kou’s double exponential model, motivated by the findings of Ramezani and Zeng (1999), who find that a double exponential jump-diffusion model fits stock market data better than the normal-diffusion model of Merton. Kou (2002) assumes that, in addition to drifted diffusion, the log-return has occasional jumps following a double exponential distribution DE (p, η1, η2). Here p is the probability of an upward jump and η1 and η2 govern the decay of the tails for the distribution of negative and positive jumps, respectively. The Lévy measure is given by

$$\nu \left( {{\text{d}}x} \right) = \left[ {p\lambda \eta_{1} {\text{e}}^{{ - \eta_{1} x}} I_{x < 0} + \left( {1 - p} \right)\lambda \eta_{2} {\text{e}}^{{ - \eta_{2} \left| x \right|}} I_{x > 0} } \right]{\text{d}}x$$
(24)

where \(\int_{ - \infty }^{\infty } \nu \left( {{\text{d}}x} \right) < \infty\), and unlike pure jump processes, Kou’s jump diffusion model is a finite activity model. The Lévy measure, through Eq. (13), this time with non-zero a and b, provides a closed form characteristic function:

$$\Phi_{{X_{t} }} \left( s \right) = \exp \left\{ {t\left( {ias - \frac{1}{2}b^{2} s^{2} + is\lambda \left[ {\frac{{p\eta_{1} }}{{\eta_{1} + is}} + \frac{{\left( {1 - p} \right)\eta_{2} }}{{\eta_{2} + is}} - 1} \right]} \right)} \right\}$$
(25)

The skewness in this model is not explicitly characterized. However Kou suggests that the feature of heavier tails becomes more pronounced with an increase of either the jump size expectation (1/ηj) or jump rate (λ). The mean-corrected characteristic function is obtained as

$$\Phi_{{X_{t} }}^{rn} \left( s \right) = \exp \left\{ \begin{gathered} i\left( {r - \frac{1}{2}b^{2} - \lambda \left[ {\frac{{p\eta_{1} }}{{\eta_{1} + 1}} + \frac{{\left( {1 - p} \right)\eta_{2} }}{{\eta_{2} + 1}} - 1} \right]} \right)st \hfill \\ - \frac{1}{2}b^{2} s^{2} t + is\lambda t\left[ {\frac{{p\eta_{1} }}{{\eta_{1} + is}} + \frac{{\left( {1 - p} \right)\eta_{2} }}{{\eta_{2} + is}} - 1} \right] \hfill \\ \end{gathered} \right\}$$
(26)

We consider the logarithm of the prices, \(s_{t} = \log \left( {S_{t} } \right)\) and k = log(K) where K is the strike price of the option. As in Carr and Madan (1999) the value of a European call with maturity T can be expressed as a function of k:

$$C_{T} \left( k \right) = \int_{k}^{\infty } {e^{ - rT} } \left( {e^{s} - e^{k} } \right)q_{T} \left( s \right){\text{d}}s$$
(27)

Here qT(s) is the risk-neutral density of the log prices. In order to overcome the non-square integrability of CT(k), Carr and Madan (1999) introduce modified call prices:

$$c_{T} \left( k \right) = {\text{e}}^{\alpha k} C_{T} \left( k \right),\alpha > 0$$
(28)

where α is the dampening factor. Following Carr and Madan (1999), an analytic expression for the pricing formula (21) can be obtained as

$$C_{T} \left( k \right) = \frac{{{\text{e}}^{ - \alpha k} }}{\pi }\int_{0}^{\infty } {{\text{e}}^{ - iuk} } \psi_{T} \left( u \right){\text{d}}u$$
(29)

where ψT has an analytic expression:

$$\psi_{T} \left( u \right) = \frac{{{\text{e}}^{ - rT} \Phi_{T} \left( {u - \left( {\alpha + 1} \right)i} \right)}}{{\alpha^{2} + \alpha - u^{2} - i\left( {2\alpha + 1} \right)u}}$$
(30)

here Ф is the characteristic function of the model for which prices are computed. In our empirical study, we shall consider Ф under risk-neutral dynamics for all our Lévy models.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mozumder, S., Choudhry, T. & Dempsey, M. Option Pricing Model Biases: Bayesian and Markov Chain Monte Carlo Regression Analysis. Comput Econ 57, 1287–1305 (2021). https://doi.org/10.1007/s10614-020-10029-x

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10614-020-10029-x

Keywords

Navigation