Skip to main content

Empirical Performances of Discrete Time Series Models

  • Chapter
  • First Online:
A Time Series Approach to Option Pricing

Abstract

Having in mind the previously developed models (Chap. 2) and pricing approaches (Chap. 3), this chapter intends to provide the reader with an empirical comparison of them. Even though part of the arguments in this book can seem rather theoretical, the underlying intention behind all of this is to provide a pricing technology that is based on realistic models of financial markets. In this chapter, we will be using a dataset of European options on the S&P500 to test and compare those models. We will not only compare the time series approaches that have been discussed earlier, but compare them to various “standards” of the industry, namely the calibrated and estimated versions of the Heston (1993) model (or at the very least their Heston and Nandi (2000) discrete time counterparts). By doing so, we aim at putting a greater emphasize on a key point most of quantitative analysts around the world are familiar with: rough scales of magnitude for option pricing errors that anyone could expect from a good option pricing model.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 54.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    See e.g. Heston and Nandi (2000), Christoffersen et al. (2006) and Barone-Adesi et al. (2008).

  2. 2.

    The value of an option is made of two different pieces. First, the intrinsic value that is the value of an immediate exercise of the option, either 0 or the difference between the current price of the underlying asset and the strike price in the case of a call option. The second piece is the time value, that is the value coming from the fact that the option is not exercised today. Usually, this time value is low.

  3. 3.

    In this case, the historical and the risk neutral volatilities coincide (see Proposition 3.2.2).

  4. 4.

    Computing the derivative of Eq. (4.3) with respect to σ, we obtain the so-called Véga (sensitivity of the price with respect to the volatility) that is strictly positive: the call option price is a strictly increasing (and thus invertible) function of the volatility. Given the integral involved in the formula, there is no closed-form expression to such an inverse function. It has to be numerically approximated. One possibility (that is used in the following) is to minimize the quadratic difference between the theoretical price and the actual market price. We also mention that very accurate approximations for N −1 exist (see Cody 1969) and are implemented in R via the qnorm function of the stats package.

  5. 5.

    By “series” here, we refer to a dataset of implied volatilities for a given moneyness and a fixed time to maturity, across the dates of our sample.

  6. 6.

    We refer the reader to Chamberlain and Rothschild (1983) and Bai and Ng (2002) for an introduction to factor models that have become increasingly important in both finance and macroeconomics.

  7. 7.

    N being the number of different (moneyness,time to maturity) pairs in our dataset and T the number of dates.

  8. 8.

    See Eq. (7), page 49 of Cont and da Fonseca (2002).

  9. 9.

    h is actually working as a smoothing parameter: the bigger this parameter is and the smoother is the function. An usual way to set the smoothing parameter h associated to a dataset x (here the moneyness or the maturity values) is to set it to be equal to \(\frac{1.8\sigma _{x}} {n^{2/5}}\), where n is the total number of observations in x and σ x the corresponding empirical standard deviation.

  10. 10.

    Here C(Tt, K) denotes the price of a call option with strike K and time to maturity Tt.

  11. 11.

    We have already seen in Sect. 2.7 that the Heston and Nandi model weakly converges under the historical probability toward the Heston diffusion when parametric constraints are well chosen.

  12. 12.

    For the GARCH(1,1) model, the long term variance is given by \(\frac{a_{0}} {1-(a_{1}+b_{1})}\) and the mean-reverting structure of the volatility a consequence of (2.16).

  13. 13.

    In this case, the computation of option prices is not as fast as in the case of closed-form formulas, but it remains much faster than when using Monte Carlo methods.

  14. 14.

    The purpose of this chapter is to provide the reader with a guide on how to use an option pricing model in the standard way: by standard, we mean that before computing option prices, we need to find a set of values for the parameters that fits the distribution of the returns. This step should be an essential one and is often overlooked, as rather liquid option markets made this step useless. When vanilla option markets are liquid, the parameters under the pricing measure can be recovered from option prices—a method usually referred to as “calibration”. However, liquidity traps and the need to price options on new underlying assets make this calibration step inadequate from time to time. What is more, we should mention that the understanding of the economic intuitions hidden behind an option pricing model and the study of historical patterns should make the quantitative analysts around the world more cautious about the ability of pricing models to be reliable under any circumstances.

  15. 15.

    An interested reader should browse the documentation associated with the following optimizers: optim, nlm and nlminb. Each of them provides optimization solutions depending on the nature of the optimization problem the user is faced with.

  16. 16.

    The order of the arguments of the optim function has been changing over the years depending on the version of R. We advise the interested reader to look at the documentation associated with the function by typing ?optim in the software console.

  17. 17.

    Remind that in the conditionally Gaussian case, the dynamics implied by the conditional Esscher transform and the extended Girsanov principle (see Sect. 3.3) coincide with the one coming from the LRNVR in Proposition 3.2.2.

  18. 18.

    On this point, see Bates (1997).

  19. 19.

    Roughly speaking, the FFT is an efficient algorithm due to Cooley and Tukey (1965) which allows to compute the sums F 0, , F N−1 where

    $$\displaystyle{ F_{k} =\sum \limits _{ j=0}^{N-1}e^{\frac{-2i\pi \mathit{jk}} {N} }x(j) }$$
    (4.21)

    using only \(\mathrm{O}(\mathit{Nlog}(N))\) operations instead of O(N 2).

  20. 20.

    In practice we will take α = 2, value that correctly works in our empirical study. See Carr and Madan (1999) for a discussion on the choice of this parameter.

  21. 21.

    In R, the complex number i is denoted by 1i.

  22. 22.

    When under the risk neutral distribution, we refer to γ (see Eq. (4.16)) and to γ under the historical distribution (see Eq. (4.12)).

  23. 23.

    The price of an European call option can be decomposed into an intrinsic value and a time value. The intrinsic value at time t is simply equal to (S t K)+, that is what the bearer of such an option could obtain if he could exercise it immediately. The time value is defined as the difference between the current price of the option and the intrinsic value of the option. It represents how much the market values the investor’s patience until the time to maturity, hence its name.

  24. 24.

    Remind that the risk neutral dynamics ((4.15), (4.16)) is implied by the conditional Esscher transform and that, in the conditionally Gaussian case, it coincides with the one given by the extended Girsanov principle (see Sect. 3.3) and the one coming from the LRNVR in Proposition 3.2.2.

  25. 25.

    In our empirical results, we use the squared value of the VIX as a measure of the risk neutral variance to determine the value of π.

  26. 26.

    We do not study the GJR specification in this section because we have seen in Table 2.4 that it is a special case of the APARCH model.

  27. 27.

    As introduced in Chaps. 2 and 3, we will refer to these mix of variance dynamics and distributions using EGARCH and APARCH for the two volatility structures and GH for the Generalized Hyperbolic distribution, MN for the Mixture of Gaussian distributions and “Jumps” for the jump component. The “EGARCH-Jump” model will therefore be the mixture of an EGARCH variance dynamics with the jump component as a conditional distribution.

  28. 28.

    An interested reader will read with interest Chorro et al. (2014).

  29. 29.

    Say we deal with a time series model for the log-returns whose estimated conditional density at time t is \(f(Y _{t}\vert \underline{Y _{t-1}},\hat{\theta }_{1})\), where \(\underline{Y _{t-1}} = (Y _{1},\ldots,Y _{t-1})\) and \(\hat{\theta }_{1}\) is the vector of parameters describing the shape of this conditional distribution and the volatility structure estimated from methodology 1. We compare this estimation method to another one defined by the conditional density \(f(Y _{t}\vert \underline{Y _{t-1}},\hat{\theta }_{2})\), with \(\hat{\theta }_{2}\) being the estimated parameters obtained from this second method. The null hypothesis of the test is “methods 1 and 2 provide a similar fit of the log-return’s conditional distribution using the same underlying model”. The corresponding test statistic is then:

    $$\displaystyle{t_{1,2} = \frac{1} {n}\sum _{t=1}^{n}\left (\log f(Y _{t}\vert \underline{Y _{ t-1}},\hat{\theta }_{1}) -\log f(Y _{t}\vert \underline{Y _{t-1}},\hat{\theta }_{2})\right )}$$

    where n is the total number of observations available. Under the null hypothesis

    $$\displaystyle{ \frac{t_{1,2}} {\hat{\sigma }_{n}} \sqrt{n}\mathop{ \rightarrow }\limits_{ n \rightarrow +\infty }\mathcal{N}(0, 1), }$$
    (4.53)

    where \(\hat{\sigma }_{n}\) is a properly selected estimator for the statistic volatility. Here, as proposed in Amisano and Giacomini (2007), we use a Newey-West estimator, with a lag empirically retained to be large (around 25).

  30. 30.

    Thus, the Carr and Madan (1999) methodology mentioned in Sect. 3.7 cannot be used in practice to price European call options.

  31. 31.

    The poor pricing performances of the extended Girsanov principle approach (see Sect. 3.3) are now well-documented in the literature especially for long maturity options as remarked in Badescu and Kulperger (2008) and Badescu et al. (2011). This leads us to focus on the Esscher dynamics.

  32. 32.

    Explicit expressions of the conditional moment generating functions for the distributions considered in this empirical part are given in Sect. 3.4.2.

  33. 33.

    See e.g. Andersen et al. (2001).

  34. 34.

    Those results are still based on the use of the REC estimates.

References

  • Ait-Sahalia Y, Lo A (1998) Nonparametric estimation of state-price densities implicit in financial asset prices. J Financ 53:499–547

    Article  Google Scholar 

  • Amisano G, Giacomini R (2007) Comparing density forecasts via weighted likelihood ratio tests. J Bus Econ Stat Am Stat Assoc 25:177–190

    Article  Google Scholar 

  • Andersen TG, Bollerslev T, Diebold FX Labys P (2001) The distribution of realized exchange rate volatility. J Am Stat Assoc 96:42–55

    Article  Google Scholar 

  • Badescu A, Kulperger R (2008) GARCH option pricing: a semiparametric approach. Insur Math Econ 43(1):69–84

    Article  Google Scholar 

  • Badescu A, Elliott RJ, Kulperger R, Miettinen J, Siu TK (2011) A comparison of pricing kernels for GARCH option pricing with generalized hyperbolic distributions. Int J Theor Appl Financ 14(5):669–708

    Article  Google Scholar 

  • Bai J, Ng S (2002) Determining the number of factors in approximate factor models. Econometrica 70:191–221

    Article  Google Scholar 

  • Barone-Adesi G, Engle R, Mancini L (2008) A GARCH option pricing model with filtered historical simulation. Rev Financ Stud 21:1223–1258

    Article  Google Scholar 

  • Bates D (1988) Pricing options under jump-diffusion processes. Working Paper

    Google Scholar 

  • Bates D (1996) Jumps and stochastic volatility: exchange rate processes implicit in deutsche mark options. Rev Financ Stud 9(1):69–107

    Article  Google Scholar 

  • Bates D (1997) The Skewness premium: option pricing under asymmetric processes. Adv Futures Options Res 9:51–82

    Google Scholar 

  • Bertholon H, Monfort A, Pegoraro F (2008) Econometric Asset pricing modelling. J Financ Econ 6(4):407–458

    Google Scholar 

  • Black F, Scholes M (1973) The pricing of options and corporate liabilities. J Polit Econ 81:637–659

    Article  Google Scholar 

  • Carr P, Madan D (1999) Option valuation using the fast fourier transform. J Comput Financ 2(4):61–73

    Google Scholar 

  • Chacko G, Viceira LM (2003) Spectral GMM estimation of continuous-time processes. J Econ 116(1–2):259–292

    Article  Google Scholar 

  • Chamberlain G, Rothschild M (1983) Arbitrage, factor structure, and mean variance analysis on large asset markets. Econometrica 51:1281–1304

    Article  Google Scholar 

  • Chorro C, Guégan D, Ielpo F (2010) Martingalized historical approach for option pricing. Financ Res Lett 7(1):24–28

    Article  Google Scholar 

  • Chorro C, Guégan D, Ielpo F, Lalaharison H (2014) Testing for leverage effect in financial returns. CES Working Papers, 2014.22

    Google Scholar 

  • Christoffersen P, Heston SL, Jacobs K (2006) Option valuation with conditional skewness. J Econ 131:253–284

    Article  Google Scholar 

  • Cody WJ (1969) Rational Chebyshev approximations for the error function. Math Comput 23:631–637

    Article  Google Scholar 

  • Cont R, da Fonseca J (2002) Dynamics of implied volatility surfaces. Quant Finan 2(1):45–60

    Article  Google Scholar 

  • Cooley JW, Tukey JW (1965) An algorithm for the machine calculation of complex Fourier series. Math Comput 19:297–301

    Article  Google Scholar 

  • Duan JC, Simonato JG (1998) Empirical martingale simulation for asset prices. Manag Sci 44:1218–1233

    Article  Google Scholar 

  • Duffie D, Kan R (1996) A yield-factor model of interest rates. Math Financ 6:379–406

    Article  Google Scholar 

  • Guégan D, Ielpo F, Lalaharison H (2013) Option pricing with discrete time jump processes. J Econ Dyn Control 37(12):2417–2445

    Article  Google Scholar 

  • Heston S (1993) A closed-form solution for options with stochastic volatility with applications to bond and currency options. Rev Financ Stud 6(2):27–343

    Article  Google Scholar 

  • Heston SL, Nandi S (2000) A closed-form garch option valuation. Rev Financ Stud 13:585–625

    Article  Google Scholar 

  • Jiang GJ, Knight JL (2002) Estimation of continuous-time processes via the empirical characteristic function. J Bus Econ Stat 20(2):198–212

    Article  Google Scholar 

  • Jiang GJ, Tian Y (2005) The model-free implied volatility and its information content. Rev Financ Stud 18(4):1305–1342

    Article  Google Scholar 

  • Rockinger M, Semenova M (2005) Estimation of jump-diffusion process via empirical characteristic function. FAME Research Paper Series rp150, International Center for Financial Asset Management and Engineering.

    Google Scholar 

  • Shimko D (1993) Bounds of probability. RISK 6:33–37

    Google Scholar 

  • Vuong QH (1989) Likelihood ratio tests for model selection and non-nested hypotheses. Econometrica 57:307–333

    Article  Google Scholar 

  • Walker JS (1996) Fast Fourier transforms, 2nd edn. CRC Press, Boca Raton

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Chorro, C., Guégan, D., Ielpo, F. (2015). Empirical Performances of Discrete Time Series Models. In: A Time Series Approach to Option Pricing. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-662-45037-6_4

Download citation

  • DOI: https://doi.org/10.1007/978-3-662-45037-6_4

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Print ISBN: 978-3-662-45036-9

  • Online ISBN: 978-3-662-45037-6

  • eBook Packages: Business and EconomicsEconomics and Finance (R0)

Publish with us

Policies and ethics