Skip to main content
Log in

A realized volatility approach to option pricing with continuous and jump variance components

  • Published:
Decisions in Economics and Finance Aims and scope Submit manuscript

Abstract

Stochastic and time-varying volatility models typically fail to correctly price out-of-the-money put options at short maturity. We extend realized volatility option pricing models by adding a jump component which provides a rapidly moving volatility factor and improves on the fitting properties under the physical measure. The change of measure is performed by means of an exponentially affine pricing kernel which depends on an equity and two variance risk premia, associated with the continuous and jump components of realized volatility. Our choice preserves analytical tractability and offers a new way of estimating variance risk premia by combining high-frequency returns and option data in a multicomponent pricing model.

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

Similar content being viewed by others

Notes

  1. As shown in “Appendix B,” the more general specification

    $$\begin{aligned} y_{t}=r+\lambda _c \mathrm {RV}^{c}_{t} + \lambda _j \mathrm {RV}^{j}_{t} +\sqrt{ \mathrm {RV}^{c}_{t} + \mathrm {RV}^{j}_{t} }\epsilon _{t} \end{aligned}$$
    (3)

    admits consistency with the no-arbitrage principle if and only if \(\lambda _c=\lambda _j=\lambda \).

  2. “Perhaps surprisingly, the results indicate that neither of the jump-adjusted standardized series are systematically closer to Gaussian than the nonadjusted realized volatility standardized returns. [...] One reason is that jumps largely self-standardize: a large jump tends to inflate the (absolute) value of both the return (numerator) and the realized volatility (denominator) of standardized returns, so the impact is muted” Andersen et al. (2010).

  3. For example, in Gagliardini et al. (2011), Corsi et al. (2013), Christoffersen et al. (2013), Majewski et al. (2015) and Bandi and Renò (2016).

  4. For practical implementation, we refer to the updated version available on SSRN including some corrections to the published version. Link: http://ssrn.com/abstract=2494379.

References

  • Andersen, T.G., Bollerslev, T., Diebold, F., Ebens, H.: The distribution of stock returns volatilities. J. Financ. Econ. 61, 43–76 (2001)

    Article  Google Scholar 

  • Andersen, T.G., Bollerslev, T., Frederiksen, P., Ørregaard Nielsen, M.: Continuous-time models, realized volatilities, and testable distributional implications for daily stock returns. J. Appl. Econom. 25(2), 233–261 (2010)

    Article  Google Scholar 

  • Bandi, F.M., Renò, R.: Price and volatility co-jumps. J. Financ. Econ. 119, 107–146 (2016)

    Article  Google Scholar 

  • Barndorff-Nielsen, O.E., Shephard, N.: Power and bipower variation with stochastic volatility and jumps. J. Financ. Econom. 2(1), 1–37 (2004)

    Article  Google Scholar 

  • Barndorff-Nielsen, O.E., Shephard, N.: Econometrics of testing for jumps in financial economics using bipower variation. J. Financ. Econom. 4(1), 1–30 (2006)

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Bates, D.: Post-’87 crash fears in the S&P 500 futures option market. J. Econom. 94(1–2), 181–238 (2000)

    Article  Google Scholar 

  • Bates, D.: Maximum likelihood estimation of latent affine processes. Rev. Financ. Stud. 19, 909–965 (2006)

    Article  Google Scholar 

  • Bormetti, G., Calcagnile, L.M., Treccani, M., Corsi, F., Marmi, S., Lillo, F.: Modelling systemic price cojumps with Hawkes factor models. Quant. Finance 15(7), 1137–1156 (2015)

    Article  Google Scholar 

  • Broadie, M., Detemple, J.B.: Anniversary article—option pricing: valuation models and applications. Manag. Sci. 50(9), 1145–1177 (2004)

    Article  Google Scholar 

  • Broadie, M., Chernov, M., Johannes, M.: Model specification and risk premia: evidence from futures options. J. Finance 62, 1453–1490 (2007)

    Article  Google Scholar 

  • Bühlmann, H., Delbaen, F., Embrechts, P., Shiryaev, A.N.: No-arbitrage, change of measure and conditional esscher transforms. CWI Q. 9(4), 291–317 (1996)

    Google Scholar 

  • Cai, N., Kou, S.G.: Option pricing under a mixed-exponential jump diffusion model. Manag. Sci. 57(11), 2067–2081 (2011)

    Article  Google Scholar 

  • Calcagnile, L.M., Bormetti, G., Treccani, M., Marmi, S., Lillo, F.: Collective synchronization and high frequency systemic instabilities in financial markets. Quant. Finance 18(2), 237–247 (2018)

    Article  Google Scholar 

  • Christoffersen, P., Jacobs, K., Ornthanalai, C., Wang, Y.: Option valuation with long-run and short-run volatility components. J. Financ. Econ. 90(3), 272–297 (2008)

    Article  Google Scholar 

  • Christoffersen, P., Elkamhi, R., Feunou, B., Jacobs, K.: Options valuation with conditional heteroskedasticity and non-normality. Rev. Financ. Stud. 23, 2139–2183 (2010)

    Article  Google Scholar 

  • Christoffersen, P., Heston, S., Jacobs, K.: Capturing option anomalies with a variance-dependent pricing kernel. Rev. Financ. Stud. 26(8), 1962–2006 (2013)

    Article  Google Scholar 

  • Christoffersen, P., Feunou, B., Jacobs, K., Meddahi, N.: The economic value of realized volatility: using high-frequency returns for option valuation. J. Financ. Quant. Anal. 49(03), 663–697 (2014)

    Article  Google Scholar 

  • Christoffersen, P., Feunou, B., Jeon, Y.: Option valuation with observable volatility and jump dynamics. J. Bank. Finance 61(Supplement 2), S101–S120 (2015)

    Article  Google Scholar 

  • Corsi, F.: A simple approximate long-memory model of realized-volatility. J. Financ. Econom. 7, 174–196 (2009)

    Article  Google Scholar 

  • Corsi, F., Renò, R.: Discrete-time volatility forecasting with persistent leverage effect and the link with continuous-time volatility modeling. J. Bus. Econ. Stat. 30(3), 368–380 (2012)

    Article  Google Scholar 

  • Corsi, F., Pirino, D., Renò, R.: Threshold bipower variation and the impact of jumps on volatility forecasting. J. Econom. 159(2), 276–288 (2010)

    Article  Google Scholar 

  • Corsi, F., Fusari, N., La Vecchia, D.: Realizing smiles: options pricing with realized volatility. J. Financ. Econ. 107(2), 284–304 (2013)

    Article  Google Scholar 

  • Darolles, S., Gourieroux, C., Jasiak, J.: Structural laplace transform and compound autoregressive models. J. Time Ser. Anal. 27, 477–503 (2006)

    Article  Google Scholar 

  • Duan, J.C.: The GARCH option pricing model. Math. Finance 5, 13–32 (1995)

    Article  Google Scholar 

  • Duan, J.C., Ritchken, P., Sun, Z.: Approximating garch-jump models, jump-diffusion processes, and option pricing. Math. Finance 16, 21–52 (2006)

    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 

  • Engle, R., Lee, G.: A permanent and transitory component model of stock return volatility. In: Engle, R., White, H. (eds.) Cointegration, Causality, and Forecasting: A Festschrift in Honor of Clive WJ Granger. Oxford University Press, Oxford (1999)

    Google Scholar 

  • Eraker, B.: Do stock prices and volatility jump? Reconciling evidence from spot and option prices. J. Finance 59, 1367–1403 (2004)

    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 

  • Fang, F., Oosterlee, C.W.: A novel pricing method for european options based on Fourier-cosine series expansions. SIAM J. Sci. Comput. 31, 826–848 (2008)

    Article  Google Scholar 

  • Gagliardini, P., Gouriéroux, C., Renault, E.: Efficient derivative pricing by the extended method of moments. Econometrica 79(4), 1181–1232 (2011)

    Article  Google Scholar 

  • Gerber, H.U., Shiu, E.S.: Option pricing by esscher transforms. Trans. Soc. Actuar. 46(99), 140 (1994)

    Google Scholar 

  • Gourieroux, C., Jasiak, J.: Autoregressive gamma process. J. Forecast. 25, 129–152 (2006)

    Article  Google Scholar 

  • Hansen, P.R., Huang, Z., Shek, H.H.: Realized GARCH: a joint model for returns and realized measures of volatility. J. Appl. Econom. 27(6), 877–906 (2012)

    Article  Google Scholar 

  • Heston, S.: Options with stochastic volatility with applications to bond and currency options. Rev. Fianc. Stud. 6, 327–343 (1993)

    Article  Google Scholar 

  • Heston, S., Nandi, S.: A closed-form GARCH option valuation model. Rev. Financ. Stud. 13(3), 585–625 (2000)

    Article  Google Scholar 

  • Huang, X., Tauchen, G.: The relative contribution of jumps to total price variance. J. Financ. Econom. 3(4), 456–499 (2005)

    Article  Google Scholar 

  • Huang, X., Wu, L.: Specification analysis of option pricing model s base on time-changed Lévy processes. J. Finance 59, 1405–1439 (2004)

    Article  Google Scholar 

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

  • Majewski, A.A., Bormetti, G., Corsi, F.: Smile from the past: a general option pricing framework with multiple volatility and leverage components. J. Econom. 187(2), 521–531 (2015)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Pan, J.: The jump-risk premia implicit in options: evidence from an integrated time-series study. J. Financ. Econ. 63, 3–50 (2002)

    Article  Google Scholar 

  • Renault, E.: Econometric models of option pricing errors. Econom. Soc. Monogr. 28, 223–278 (1997)

    Google Scholar 

  • Zhang, L., Aït-Sahalia, Y., Mykland, P.A.: A tale of two time scales: determining integrated volatility with noisy high frequency data. J. Am. Stat. Assoc. 100, 1394–1411 (2005)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giacomo Bormetti.

Additional information

Publisher's Note

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

Appendices

A MGF under \(\varvec{\mathbb {P}}\) measure

The relations which follow are derived for the log-return dynamics specified in Eq. (2). For the ease of computation, expression (4) is rewritten as

$$\begin{aligned} \varTheta ({\mathbf{RV}}^{{\varvec{c}}}_{t},{\mathbf{L}}_{t})=d+\sum \limits _{i=1}^{22}\beta _i\mathrm {RV}_{t+1-i}^c+\sum \limits _{i=1}^{22}\alpha _i\left( \epsilon _{t+1-i}-\gamma \sqrt{\mathrm {RV}^c_{t+1-i}+\mathrm {RV}^j_{t+1-i}}\right) ^2\,, \end{aligned}$$

with

$$\begin{aligned} \beta _i= {\left\{ \begin{array}{ll} \beta _d &{}\text {for}\,i=1\\ \beta _w/4 &{}\text {for}\,2\le i\le 5\\ \beta _w/17 &{}\text {for}\,6\le i\le 22\\ \end{array}\right. } \alpha _i= {\left\{ \begin{array}{ll} \alpha _d &{}\text {for}\,i=1\\ \alpha _w/4 &{}\text {for}\,2\le i\le 5\\ \alpha _w/17 &{}\text {for}\,6\le i\le 22\\ \end{array}\right. }. \end{aligned}$$
(8)

We start showing that JLHARG processes satisfy the affine relation

$$\begin{aligned} {\mathbb {E}}\left[ e^{zy_{s+1}+{\mathbf {b}}\cdot {\mathbf {RV}}_{s+1}+c\ell _{s+1}}|{\mathcal {F}}_s\right] =e^{{\mathcal {A}}\left( z,{\mathbf {b}},c\right) +\sum \limits _{i=1}^p{\varvec{{\mathcal {B}}}}_i\left( z,{\mathbf {b}},c\right) \cdot {\mathbf {RV}}_{s+1-i}+\sum \limits _{j=1}^q{\mathcal {C}}_j\left( z,{\mathbf {b}},c\right) \ell _{s+1-j}}\,, \end{aligned}$$
(9)

for some functions \({\mathcal {A}}\,:\,{\mathbb {R}}\times {\mathbb {R}}^2\times {\mathbb {R}}\rightarrow {\mathbb {R}}\), \({\varvec{{\mathcal {B}}}}_i\,:\,{\mathbb {R}}\times {\mathbb {R}}^2\times {\mathbb {R}}\rightarrow {\mathbb {R}}^2\), \({\varvec{{\mathcal {C}}}}_j\,:\,{\mathbb {R}}\times {\mathbb {R}}^2\times {\mathbb {R}}\rightarrow {\mathbb {R}}\), where \({\mathbf {RV}}_t=(\mathrm {RV}_t^c,\mathrm {RV}_t^j)\), \({\mathbf {b}}\in {\mathbb {R}}^2\), \(c\in {\mathbb {R}}\), and \(\cdot \) is the scalar product in \({\mathbb {R}}^2\). To derive the explicit form of the functions \({\mathcal {A}},\,{\mathcal {B}}_i,\,{\mathcal {C}}_j\) which allows to characterize the MGF we show that

$$\begin{aligned}&{\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{zy_{t}+{\mathbf {b}}\cdot {\mathbf {RV}}_t+c\ell _{t}}|{\mathcal {F}}_{t-1}\right] \nonumber \\&\quad ={\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{z(r+\lambda _c\mathrm {RV}^c_{t}+\lambda _j\mathrm {RV}^j_{t}+\sqrt{\mathrm {RV}^c_{t}+\mathrm {RV}^j_{t}}\epsilon _{t})+{\mathbf {b}}\cdot {\mathbf {RV}}_t+c\ell _{t}}|{\mathcal {F}}_{t-1}\right] \nonumber \\&\quad ={\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{z(r+\lambda _c\mathrm {RV}^c_{t}+\lambda _j\mathrm {RV}^j_{t})+{\mathbf {b}}\cdot {\mathbf {RV}}_t} {\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{z\sqrt{\mathrm {RV}^c_{t}+\mathrm {RV}^j_{t}}\epsilon _{t}+c(\epsilon _{t}-\gamma \sqrt{\mathrm {RV}^c_{t}+\mathrm {RV}^j_{t}})^2}|{\mathbf {RV}}_t\right] |{\mathcal {F}}_{t-1}\right] \nonumber \\&\quad ={\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{z(r+\lambda _c\mathrm {RV}^c_{t}+\lambda _j\mathrm {RV}^j_{t})+b_1\mathrm {RV}_t^c+b_2\mathrm {RV}_t^j-\frac{1}{2}\ln (1-2c)+\left( \frac{\frac{z^2}{2}+\gamma ^2c-2c\gamma z}{1-2c}\right) (\mathrm {RV}^c_{t}+\mathrm {RV}^j_{t})}|{\mathcal {F}}_{t-1}\right] \nonumber \\&\quad ={\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{zr-\frac{1}{2}\ln (1-2c)+\left( z\lambda _c+b_1+\frac{\frac{z^2}{2}+\gamma ^2c-2c\gamma z}{1-2c}\right) \mathrm {RV}^{c}_{t}+\left( z\lambda _j+b_2+\frac{\frac{z^2}{2}+\gamma ^2c-2c\gamma z}{1-2c}\right) \mathrm {RV}^j_{t}}|{\mathcal {F}}_{t-1}\right] \nonumber \\&\quad =\mathrm {e}^{zr-\frac{1}{2}\ln (1-2c)}{\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{\left( z\lambda _c+b_1+\frac{\frac{z^2}{2}+\gamma ^2c-2c\gamma z}{1-2c}\right) \mathrm {RV}^{c}_{t}}|{\mathcal {F}}_{t-1}\right] \nonumber \\&\qquad {\mathbb {E}}^{{\mathbb {P}}}\left[ \mathrm {e}^{\left( z\lambda _j+b_2+\frac{\frac{z^2}{2}+\gamma ^2c-2c\gamma z}{1-2c}\right) \mathrm {RV}^j_{t}}|{\mathcal {F}}_{t-1}\right] . \end{aligned}$$
(10)

In the third line we have used the result that if \(Z\sim {\mathcal {N}}(0,1)\) then

$$\begin{aligned} {\mathbb {E}}\left[ \exp \left( x\left( Z+y\right) ^2\right) \right] =\exp \left( -\frac{1}{2}\ln \left( 1-2x\right) +\frac{xy^2}{1-2x}\right) . \end{aligned}$$

For a noncentered gamma random variable, from Gourieroux and Jasiak (2006) we know that

$$\begin{aligned} {\mathbb {E}}^{{\mathbb {P}}}\left[ e^{x_1\mathrm {RV}^{c}_{t}}|{\mathcal {F}}_{t-1}\right] =\exp \left( -\delta {\mathcal {W}}\left( x_1,\theta \right) +{\mathcal {V}}\left( x_1,\theta \right) \left( d+\sum \limits _{i=1}^{p}\beta _i\mathrm {RV}_{s-i}^c+\sum \limits _{j=1}^{q}\alpha _j\ell _{s-j}\right) \right) \,, \end{aligned}$$

where

$$\begin{aligned} {\mathcal {V}}(x_1,\theta )=\frac{\theta x_1}{1-\theta x_1}, \;\;\; {\mathcal {W}}(x_1,\theta )=\ln (1-x_1\theta ), \end{aligned}$$

and

$$\begin{aligned} x_1(z,b_1,c)=z\lambda _c+b_1+\frac{\frac{1}{2}z^2+\gamma ^2c-2c\gamma z}{1-2c}\,. \end{aligned}$$
(11)

For the computation of the last expectation in the final line of (10), we use the property that if \(Z_t\) is a compound Poisson process with rate \(\omega \) and i.i.d. jump sizes \(D_i\), then

$$\begin{aligned} {\mathbb {E}}\left[ e^{xZ_t}|{\mathcal {F}}_{t-1}\right] =\exp \left( \omega \left( M_D(x)-1\right) \right) , \end{aligned}$$
(12)

where \(M_D(x)\) is the MGF of the jump size random variable D. Since sizes are distributed according to a gamma distribution, we have

$$\begin{aligned} M_D(x)=\frac{1}{\left( 1-x{\tilde{\theta }}\right) ^{{\tilde{\delta }}} }\,. \end{aligned}$$
(13)

From expressions (12) and (13) we obtain

$$\begin{aligned} {\mathbb {E}}^{{\mathbb {P}}}\left[ e^{x_2 \mathrm {RV}^{j}_{t}}|{\mathcal {F}}_{t-1}\right] =\exp \left( {\tilde{\varTheta }}{\mathcal {J}}\left( x_2,{\tilde{\theta }},{\tilde{\delta }}\right) \right) , \end{aligned}$$

where

$$\begin{aligned} {\mathcal {J}}(x_2,{\tilde{\theta }},{\tilde{\delta }})=\frac{1-(1-{\tilde{\theta }}x_2)^{{\tilde{\delta }}}}{(1-{\tilde{\theta }}x)^{{\tilde{\delta }}}} \ \ \text{ and } \ \ x_2(z,b_2,c)=z\lambda _j+b_2+\frac{\frac{1}{2}z^2+\gamma ^2c-2c\gamma z}{1-2c}\,. \end{aligned}$$

Gathering all the previous results, we finally conclude

$$\begin{aligned} \begin{aligned}&{\mathbb {E}}^{{\mathbb {P}}}\left[ e^{zy_{t}+{\mathbf {b}}\cdot {\mathbf {RV}}_t+c\ell _{t}}|{\mathcal {F}}_{t-1}\right] \\&\quad =\exp \left[ zr-\frac{1}{2}\ln (1-2c)+{\mathcal {V}}(x_1,\theta )\left( d+\sum \limits _{i=1}^{p}\beta _i\mathrm {RV}_{t-i}^c+\sum \limits _{j=1}^{q}\alpha _j\ell _{t-j}\right) \right. \\&\qquad \left. -\delta {\mathcal {W}}(x_1,\theta )+{\tilde{\varTheta }}{\mathcal {J}}(x_2,{\tilde{\theta }},{\tilde{\delta }}) \phantom {{\mathcal {V}}(x_1,\theta )\left( d+\sum \limits _{i=1}^{p}\beta _i\mathrm {RV}_{t-i}^c+\sum \limits _{j=1}^{q}\alpha _j\ell _{t-j}\right) }\right] \,. \end{aligned} \end{aligned}$$

The direct comparison of the last expression with (9) allows to derive the following explicit expressions

$$\begin{aligned}&{\mathcal {A}}(z,{\mathbf {b}},c)=zr-\frac{1}{2}\ln (1-2c)-\delta {\mathcal {W}}(x_1,\theta )+d{\mathcal {V}}(x_1,\theta )+{\tilde{\varTheta }}{\mathcal {J}}(x_2,{\tilde{\theta }},{\tilde{\delta }}) \,, \end{aligned}$$
(14)
$$\begin{aligned}&{\mathcal {B}}_i(z,b_1,c)={\mathcal {V}}(x_1,\theta )\beta _i \,, \end{aligned}$$
(15)
$$\begin{aligned}&{\mathcal {C}}_j(z,b_1,c)={\mathcal {V}}(x_1,\theta )\alpha _j \,. \end{aligned}$$
(16)

As shown in Majewski et al. (2015), once we have above expressions we obtain

$$\begin{aligned} \phi ^{{\mathbb {P}}}\left( t,T,z\right) ={\mathbb {E}}^{{\mathbb {P}}}\left[ e^{zy_{t,T}}|{\mathcal {F}}_t\right] =\exp \left( \mathrm {a}_t+\sum \limits _{i=1}^{p}\mathrm {b}_{t,i}\mathrm {RV}_{t+1-i}^c+\sum \limits _{i=1}^{q}\mathrm {c}_{t,i}\ell _{t+1-i}\right) \end{aligned}$$

where

$$\begin{aligned}&\mathrm {a}_s=\mathrm {a}_{s+1}+zr-\frac{1}{2}\log (1-2\mathrm {c}_{s+1,1})+d{\mathcal {V}}(\mathrm {x}_{s+1}^{c},\theta )-\delta {\mathcal {W}}(\mathrm {x}_{s+1}^{c},\theta )+{\tilde{\varTheta }}{\mathcal {J}}(\mathrm {x}_{s+1}^{j},{\tilde{\theta }}) \nonumber \\&\mathrm {b}_{s,i}= {\left\{ \begin{array}{ll} \mathrm {b}_{s+1,i}+{\mathcal {V}}(\mathrm {x}_{s+1}^{c},\theta )\beta _i &{} \text {for} \, 1\le i \le p-1\\ {\mathcal {V}}(\mathrm {x}_{s+1}^{c},\theta )\beta _i &{}\text {for} \, i=p \end{array}\right. } \nonumber \\&\mathrm {c}_{s,i}= {\left\{ \begin{array}{ll} \mathrm {c}_{s+1,i}+{\mathcal {V}}(\mathrm {x}_{s+1}^{c},\theta )\alpha _i &{}\text {for} \, 1\le i \le q-1\\ {\mathcal {V}}(\mathrm {x}_{s+1}^{c},\theta )\alpha _i &{}\text {for} \, i=q \end{array}\right. } \end{aligned}$$
(17)

with

$$\begin{aligned}&\mathrm {x}_{s+1}^{c}=z\lambda _c+\mathrm {b}_{s+1,1}+\frac{\frac{1}{2}z^2+\gamma ^2\mathrm {c}_{s+1,1}-2\mathrm {c}_{s+1,1}\gamma z}{1-2\mathrm {c}_{s+1,1}}\,, \end{aligned}$$
(18)
$$\begin{aligned}&\mathrm {x}_{s+1}^{j}=z\lambda _j+\frac{\frac{1}{2}z^2+\gamma ^2\mathrm {c}_{s+1,1}-2\mathrm {c}_{s+1,1}\gamma z}{1-2\mathrm {c}_{s+1,1}}\,. \end{aligned}$$
(19)

The functions \({\mathcal {V}}\), \({\mathcal {W}}\) and \({\mathcal {J}}\) are defined as above. The terminal conditions read \(\mathrm {a}_T=\mathrm {b}_{T,i}=\mathrm {c}_{T,j}=0\) for \(i=1,2,...,p\) and \(j=1,2,...,q\).

B No-arbitrage condition

The no-arbitrage conditions are

$$\begin{aligned} {\mathbb {E}}^{{\mathbb {P}}}\left[ M_{s,s+1}|{\mathcal {F}}_s\right]&=1\; \text {for}\; s\in {\mathbb {N}}\,,\nonumber \\ {\mathbb {E}}^{{\mathbb {P}}}\left[ M_{s,s+1}e^{y_{s+1}}|{\mathcal {F}}_s\right]&=\mathrm {e}^r\; \text {for}\; s\in {\mathbb {N}} . \end{aligned}$$
(20)

The first relation is satisfied by definition of \(M_{s,s+1}\). From a general result in Majewski et al. (2015), condition (20) is satisfied if, and only if

$$\begin{aligned} \begin{aligned}&{\mathcal {A}} (1-\nu _y,-{\varvec{\nu }},0)=r+{\mathcal {A}} (-\nu _y,-{\varvec{\nu }},0)\,,\\&{\mathcal {B}}_i(1-\nu _y,-{\varvec{\nu }},0)= {\mathcal {B}}_i(-\nu _y,-{\varvec{\nu }},0)\,,\\&{\mathcal {C}}_j(1-\nu _y,-{\varvec{\nu }},0)= {\mathcal {C}}_j(-\nu _y,-{\varvec{\nu }},0)\,, \end{aligned} \end{aligned}$$

with \({\varvec{\nu }}=(\nu _c,\nu _j)\). To conclude, it is sufficient to show under which conditions the following two relations hold true

$$\begin{aligned} x_1(1-\nu _y,-\nu _c,0)=x_1(-\nu _y,-\nu _c,0)\,,\\ x_2(1-\nu _y,-\nu _j,0)=x_2(-\nu _y,-\nu _j,0)\,. \end{aligned}$$

Simple computations show that the latter equations are satisfied if and only if

$$\begin{aligned} \nu _y=\lambda _c +\frac{1}{2} = \lambda _j +\frac{1}{2}. \end{aligned}$$

Remarkably, the only specification for the log-return dynamics in Eq. (2) which ensures consistency with no-arbitrage is the dynamics where the equity premia \(\lambda _c\) and \(\lambda _j\) are equal and coincide to \(\lambda \). Then, we obtain

$$\begin{aligned} \nu _y = \lambda + \frac{1}{2}\,. \end{aligned}$$

It is important to notice that the no-arbitrage condition for the equity premium does not constrain the value of the variance risk premia \(\nu _c\) and \(\nu _j\).

C Risk-neutral dynamics

JLHARG models imply a risk-neutral MGF for log-returns whose exponential affine terms can be re-parameterized in order to obtain an expression formally equivalent to the physical MGF. Firstly we observe that the risk-neutral MGF can be expressed with a recursive set of expressions, involving a combination of the functions \({\mathcal {A}},\,{\mathcal {B}}_i,\,{\mathcal {C}}_j\). Then, recalling the results given in Majewski et al. (2015), the MGF for JLHARG model under measure \({\mathbb {Q}}\) has the following form

$$\begin{aligned} \phi ^{{\mathbb {Q}}}_{\nu _c\,\nu _j\,\nu _y}\left( t,T,z\right) ={\mathbb {E}}^{{\mathbb {Q}}}\left[ e^{zy_{t,T}}|{\mathcal {F}}_t\right] =\exp \left( \mathrm {a}^*_t+\sum \limits _{i=1}^{p}\mathrm {b}_{t,i}^{*}\mathrm {RV}_{t+1-i}^c+\sum \limits _{i=1}^{q}\mathrm {c}^*_{t,i}\ell _{t+1-i}\right) \,, \end{aligned}$$

where

$$\begin{aligned}&\mathrm {a}^*_s=\mathrm {a}^*_{s+1}+zr-\frac{1}{2}\log (1-2\mathrm {c}^*_{s+1,1})+d{\mathcal {V}}(\mathrm {x}_{s+1}^{c\,*},\theta )-d{\mathcal {V}}(\mathrm {y}_{s+1}^{c\,*},\theta ) \nonumber \\&\qquad -\delta {\mathcal {W}}(\mathrm {x}_{s+1}^{c\,*},\theta )+\delta {\mathcal {W}}(\mathrm {y}_{s+1}^{c\,*},\theta )+{\tilde{\varTheta }}{\mathcal {J}}(\mathrm {x}_{s+1}^{j\,*},{\tilde{\theta }})-{\tilde{\varTheta }}{\mathcal {J}}(\mathrm {y}_{s+1}^{j\,*},{\tilde{\theta }}) \nonumber \\&\mathrm {b}_{s,i}^{*}= {\left\{ \begin{array}{ll} \mathrm {b}_{s+1,i}^{*}+\left( {\mathcal {V}}(\mathrm {x}_{s+1}^{c\,*},\theta )-{\mathcal {V}}(\mathrm {y}_{s+1}^{c\,*},\theta )\right) \beta _i &{}\text {for} \, 1\le i \le p-1\\ \left( {\mathcal {V}}(\mathrm {x}_{s+1}^{c\,*},\theta )-{\mathcal {V}}(\mathrm {y}_{s+1}^{c\,*},\theta )\right) \beta _i &{}\text {for} \, i=p \end{array}\right. } \nonumber \\&\mathrm {c}^*_{s,j}= {\left\{ \begin{array}{ll} \mathrm {c}_{s+1,j}^{*}+\left( {\mathcal {V}}(\mathrm {x}_{s+1}^{c\,*},\theta )-{\mathcal {V}}(\mathrm {y}_{s+1}^{c\,*},\theta )\right) \alpha _j &{}\text {for} \, 1\le j \le q-1\\ \left( {\mathcal {V}}(\mathrm {x}_{s+1}^{c\,*},\theta )-{\mathcal {V}}(\mathrm {y}_{s+1}^{c\,*},\theta )\right) \alpha _j &{}\text {for} \, j=q \end{array}\right. } \end{aligned}$$
(21)

where

$$\begin{aligned}&x^{c\,*}_{s+1}=(z-\nu _y)\lambda +\mathrm {b}_{s+1,1}^{*}-\nu _c+\frac{\frac{1}{2}(z-\nu _y)^2+\gamma ^2 \mathrm {c}^*_{s+1,1}-2\mathrm {c}^*_{s+1,1}\gamma (z-\nu _y)}{1-2c^*_{s+1,1}}\\&x^{j\,*}_{s+1}=(z-\nu _y)\lambda -\nu _j+\frac{\frac{1}{2}(z-\nu _y)^2+\gamma ^2 \mathrm {c}^*_{s+1,1}-2\mathrm {c}^*_{s+1,1}\gamma (z-\nu _y)}{1-2c^*_{s+1,1}}\\&y^{l\,*}_{s+1}=-\nu _y\lambda -\nu _l+\frac{1}{2}\nu _y^2, \end{aligned}$$

with \(l=c,j\) and the terminal conditions are \(\mathrm {a}^*_T=\mathrm {b}_{T,i}^{*}=\mathrm {c}^*_{T,j}=0\) for \(i=1,2,...,p\) and \(j=1,2,...,q\).

The first passage consists in comparing expression (21) with (17). We have to find a set of new parameters for which the recursive expressions for \(\mathrm {a}^*_t,\mathrm {b}^{*}_t,\mathrm {c}^*_t \) under \({\mathbb {Q}}\) correspond to the expressions under \({\mathbb {P}}\). We start defining

$$\begin{aligned}&x^{c\,**}_{s+1,i}=z\lambda ^*+b^{*}_{s+1,1}+\frac{\frac{1}{2}z^2+(\gamma ^*)^2 c^*_{s+1,1}-2c^*_{s+1,1}\gamma ^* z}{1-2c^*_{s+1,1}}\,,\\&x^{j\,**}_{s+1,i}=z\lambda ^*+\frac{\frac{1}{2}z^2+(\gamma ^*)^2 c^*_{s+1,1}-2c^*_{s+1,1}\gamma ^* z}{1-2c^*_{s+1,1}}\,. \end{aligned}$$

Then, the following relations have to hold

$$\begin{aligned} \delta \left( {\mathcal {W}}\left( \mathrm {x}_{s+1}^{c\,*},\theta \right) -{\mathcal {W}}\left( \mathrm {y}^{c\,*},\theta \right) \right)&=\delta ^*{\mathcal {W}}\left( \mathrm {x}_{s+1}^{c\,**},\theta ^*\right) \end{aligned}$$
(22)
$$\begin{aligned} \beta _i\left( {\mathcal {V}}\left( \mathrm {x}_{s+1}^{c\,*},\theta \right) -{\mathcal {V}}\left( \mathrm {y}^{c\,*},\theta \right) \right)&=\beta ^*_i{\mathcal {V}}\left( \mathrm {x}_{s+1}^{c\,**},\theta ^*\right) \end{aligned}$$
(23)
$$\begin{aligned} \alpha _j\left( {\mathcal {V}}\left( \mathrm {x}_{s+1}^{c\,*},\theta \right) -{\mathcal {V}}\left( \mathrm {y}^{c\,*},\theta \right) \right)&=\alpha ^*_j{\mathcal {V}}\left( \mathrm {x}_{s+1}^{c\,**},\theta ^*\right) \end{aligned}$$
(24)
$$\begin{aligned} {\tilde{\varTheta }}\left( {\mathcal {J}}\left( \mathrm {x}_{s+1}^{j\,*},{\tilde{\theta }}\right) -{\mathcal {J}}\left( \mathrm {y}^{j\,*},{\tilde{\theta }}\right) \right)&={\tilde{\varTheta }}^*{\mathcal {J}}\left( \mathrm {x}_{s+1}^{j\,**},{\tilde{\theta }}^*\right) \end{aligned}$$
(25)

with \(\mathrm {y}^{c\,*}=-\lambda ^2/2-\nu _c+\frac{1}{8}\) and \(\mathrm {y}^{j\,*}=-\lambda ^2/2-\nu _j+\frac{1}{8}\).

Equation (22) can be explicitly written as

$$\begin{aligned} \delta \log \left[ 1-\frac{\theta }{1-\theta \mathrm {y}^{c\,*}}\left( \mathrm {x}_{s+1}^{c\,*}-\mathrm {y}^{c\,*}\right) \right] =\delta ^*\log \left( 1-\theta ^*\mathrm {x}_{s+1}^{c\,**}\right) , \end{aligned}$$

which implies the following three sufficient conditions

$$\begin{aligned} \delta ^*&=\delta \nonumber \\ \theta ^*&=\frac{\theta }{1-\theta \mathrm {y}^{c\,*}} \nonumber \\ \mathrm {x}_{s+1}^{c\,**}&=\mathrm {x}_{s+1}^{c\,*}-\mathrm {y}^{c\,*} . \end{aligned}$$
(26)

It can be easily verified that last condition (26) is satisfied by substituting

$$\begin{aligned} \lambda _c^*&=-\frac{1}{2} \,,\\ \gamma ^*&=\gamma +\lambda +\frac{1}{2} . \end{aligned}$$

Equation (23) can be equivalently expressed in the form

$$\begin{aligned} \frac{\beta _i}{1-\theta \mathrm {y}^{c\,*}}\frac{\theta }{1-\theta \mathrm {y}^{c\,*}}\frac{\mathrm {x}_{s+1}^{c\,*}-\mathrm {y}^{c\,*}}{1-\theta /(1-\theta \mathrm {y}^{c\,*})\left( \mathrm {x}_{s+1}^{c\,*}-\mathrm {y}^{c\,*}\right) }=\beta _i^*\frac{\theta ^*\mathrm {x}_{s+1}^{c\,**}}{1-\theta ^*\mathrm {x}_{s+1}^{c\,**}} \end{aligned}$$

which gives another sufficient condition for the mapping

$$\begin{aligned} \beta _i^*=\frac{\beta _i}{1-\theta \mathrm {y}^{c\,*}}\,. \end{aligned}$$

An analogous consideration about third condition (24) allows to obtain the condition on \(\alpha _i^*\),

$$\begin{aligned} \alpha _i^*=\frac{\alpha _i}{1-\theta \mathrm {y}^{c\,*}}\,. \end{aligned}$$

Relation (8) gives us the expressions for \(\beta _d^*\), \(\beta _w^*\), \(\beta _m^*\), \(\alpha _d^*\), \(\alpha _w^*\) and \(\alpha _m^*\). Finally, equation (25) provides the last sufficient condition

$$\begin{aligned} \frac{{\tilde{\varTheta }}}{\left( 1-{\tilde{\theta }}\mathrm {y}^{j\,*}\right) ^{{\tilde{\delta }}}}\frac{1-\left( \left( 1-{\tilde{\theta }}\mathrm {x}^{j\,*}_{s+1}\right) /\left( 1-{\tilde{\theta }}\mathrm {y}^{j\,*}\right) \right) ^{{\tilde{\delta }}}}{\left( \left( 1-{\tilde{\theta }}\mathrm {x}^{j\,*}_{s+1}\right) /\left( 1-{\tilde{\theta }}\mathrm {y}^{j\,*}\right) \right) ^{{\tilde{\delta }}}}={\tilde{\varTheta }}^*\frac{1-(1-{\tilde{\theta }}^*\mathrm {x}_{s+1}^{j\,**})^{{\tilde{\delta }}^*}}{(1-{\tilde{\theta }}^*\mathrm {x}_{s+1}^{j\,**})^{{\tilde{\delta }}^*}}, \end{aligned}$$

which is satisfied if

$$\begin{aligned} {\tilde{\delta }}^*&={\tilde{\delta }}\,,\nonumber \\ {\tilde{\varTheta }}^*&=\frac{{\tilde{\varTheta }}}{\left( 1-{\tilde{\theta }}\mathrm {y}^{j\,*}\right) ^{{\tilde{\delta }}}}\,,\nonumber \\ {\tilde{\theta }}^*&=\frac{{\tilde{\theta }}}{1-{\tilde{\theta }}\mathrm {y}^{j\,*}}\,,\nonumber \\ \mathrm {x}_{s+1}^{j\,**}&=\mathrm {x}_{s+1}^{j\,*}-\mathrm {y}^{j\,*} \,. \end{aligned}$$
(27)

As it can be seen last condition (27) is redundant when compared to condition (26).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Alitab, D., Bormetti, G., Corsi, F. et al. A realized volatility approach to option pricing with continuous and jump variance components. Decisions Econ Finan 42, 639–664 (2019). https://doi.org/10.1007/s10203-019-00241-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10203-019-00241-2

Keywords

JEL Classification

Navigation