Skip to main content
Log in

Dynamic portfolio strategies under a fully correlated jump-diffusion process

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

Abstract

This paper presents the first continuous-time model to feature a flexible dependence structure among jump intensity, stock variance, and stock returns. In particular, it addresses a gap in the financial portfolio optimization literature concerning the non-trivial correlation between stock return variance and the intensity of price jumps. The model permits closed-form representations for the optimal strategy and value functions in an expected utility theory setting. It also produces analytical expressions for the value function associated with relevant suboptimal strategies. Such an analytical setting allows for the first wealth-equivalent utility loss (WEL) analysis of the pitfalls of ignoring the aforementioned dependence. The model and results can be easily extended to the pair intensity-covariance in multi-assets. The WEL analysis is carried out for three different suboptimal classes: tailor-made incomplete markets, misspecifications in the parameters of the model, and time-independent (myopic) strategies. For the numerical section, we focus on the correlation between jump intensity and stock variance, which is assumed to be either zero or one in the existing literature. We demonstrate that simplistic assumptions like perfect dependence or independence could lead to wealth-equivalent losses of up to 61%. Similarly, a failure to hedge these variances and intensity drivers could cause losses of up to 95% (in particular, up to 60% due to the factors driving the dependence).

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

Similar content being viewed by others

Notes

  1. The absence of transaction costs and the assumption of liquidity make this feasible. Otherwise investor would be very concern on which derivatives to choose.

  2. All relevant suboptimal strategies that we could thought of are fortunately independent of the state variables: \(\mathbb {v}_t\) and \(v^{(d+2)}_t\).

  3. This measure of utility loss is different from the one used in Liu and Pan (2003), \(R^W=-\frac{\gamma }{T}\ln {\left( 1-L^{\varPi }\right) }\). However, as is shown in Larsen and Munk (2012), there is a one-to-one correspondence between the two measures.

References

  • Andersen, T., Benzoni, L., Lund, J.: An empirical investigation of continuous-time equity return models. J Finance 57(3), 1239–1284 (2002)

    Article  Google Scholar 

  • Anderson, T.W.: An Introduction to Multivariate Statistical Analysis: Hoboken: Wiley (2003)

  • Bakshi, G., Cao, C., Chen, Z.: Empirical performance of alternative option pricing models. J Finance 52(5), 2003–2049 (1997)

    Article  Google Scholar 

  • Branger, N., Larsen, L.: Robust portfolio choice with uncertainty about jump and diffusion risk. J Bank Finance 37(12), 5036–5047 (2013)

    Article  Google Scholar 

  • Branger, N., Muck, M., Seifried, F., Weisheit, S.: Optimal portfolios when variances and covariances can jump. J Econ Dyn Control 85, 59–89 (2017)

    Article  Google Scholar 

  • Branger, N., Schlag, C., Schneider, E.: Optimal portfolios when volatility can jump. J Bank Finance 32(6), 1087–1097 (2008)

    Article  Google Scholar 

  • Christoffersen, P., Heston, S., Jacobs, K.: The shape and term structure of the index option smirk: why multifactor stochastic volatility models work so well. Manag Sci 55(12), 1914–1932 (2009)

    Article  Google Scholar 

  • Christoffersen, P., Kris, J., Chayawat, O.: Dynamic jump intensities and risk premiums: evidence from S&P500 returns and options. J Financ Econ 106(3), 447–472 (2012)

    Article  Google Scholar 

  • Das, S., Uppal, R.: Systemic risk and international portfolio choice. J Finance 59(6), 2809–2834 (2004)

    Article  Google Scholar 

  • De Col, A., Gnoatto, A., Grasselli, M.: Smiles all around: FX joint calibration in a multi-Heston model. J Bank Finance 37(10), 3799–3818 (2013)

    Article  Google Scholar 

  • Engle, R.: Autoregressive conditional heteroscedasticity with estimates of the variance of United Kingdom inflation. Econometrica 50(4), 987–1007 (1982)

    Article  Google Scholar 

  • Escobar, M., Christoph, G.: Parameters recovery via calibration in the Heston model: a comprehensive review. Wilmott 86, 60–81 (2016)

    Article  Google Scholar 

  • Escobar, M., Ferrando, S., Rubtsov, A.: Robust portfolio choice with derivative trading under stochastic volatility. J Bank Finance 61, 142–157 (2015)

    Article  Google Scholar 

  • Escobar, M., Ferrando, S., Rubtsov, A.: Optimal investment under multi-factor stochastic volatility. Quant Finance 17(2), 241–260 (2017)

    Article  Google Scholar 

  • Fama, E., French, K.: A five-factor asset pricing model. J Financ Econ 116, 1–2 (2015)

    Article  Google Scholar 

  • Flor, C., Larsen, L.: Robust portfolio choice with stochastic interest rates. Ann Finance 10(2), 243–265 (2014)

    Article  Google Scholar 

  • Heston, S.: A closed-form solution for options with stochastic volatility with applications to bonds and currency options. Rev Financ Stud 6(2), 327–343 (1993)

    Article  Google Scholar 

  • Hull, J., Predescu, M., White, A.: The valuation of correlation-dependent credit derivatives using a structural model. J Credit Risk 6(3), 99 (2010)

    Article  Google Scholar 

  • Jeanblanc, M., Yor, M., Chesney, M.: Mathematical Methods for Financial Markets: Berlin: Springer (2009)

  • Jorion, P.: On jump processes in the foreign exchange and stock markets. Rev Financ Stud 1(4), 427–445 (1988)

    Article  Google Scholar 

  • Keller-Ressel, M., Mayerhofer, E.: Exponential moments of affine processes. Ann Appl Probab 25(2), 714–752 (2015)

    Article  Google Scholar 

  • Kraft, H.: Optimal portfolios and Heston’s stochastic volatility model: an explicit solution for power utility. Quant Finance 5(3), 303–313 (2005)

    Article  Google Scholar 

  • Lando, D.: On cox processes and credit risky securities. Rev Deriv Res 2(2), 99–120 (1998)

    Google Scholar 

  • Larsen, L., Munk, C.: The cost of suboptimal dynamic asset allocation: general results and applications to interest rate risk, stock volatility risk, and growth/value tilts. J Econ Dyn Control 36(2), 266–293 (2012)

    Article  Google Scholar 

  • Li, G., Zhang, C.: On the relationship between conditional jump intensity and diffusive volatility. J Empir Finance 37, 196–213 (2016)

    Article  Google Scholar 

  • Liu, J.: Portfolio selection in stochastic environments. Rev Financ Stud 20(1), 1–39 (2007)

    Article  Google Scholar 

  • Liu, J., Pan, J.: Dynamic derivative strategies. J Financ Econ 69(3), 401–430 (2003)

    Article  Google Scholar 

  • Merton, R.: Lifetime portfolio selection under uncertainty: the continuous-time case. Rev Econ Stat 51(3), 247–257 (1969)

    Article  Google Scholar 

  • Merton, R.: Option pricing when underlying stock returns are discontinuous. J Financ Econ 3(1), 125–144 (1976)

    Article  Google Scholar 

  • Roll, R.: The international crash of October 1987. Financ Anal J 44(5), 19–35 (1988)

    Article  Google Scholar 

  • Santa-Clara, P., Yan, S.: Jump and volatility risk and risk premia: a new model and lessons from S&P 500 options. No. w10912. National Bureau of Economic Research (2004). https://scholar.google.ca/scholar?cluster=16731161349935270384&hl=en&as_sdt=0,5

Download references

Acknowledgements

Harold A. Moreno-Franco was supported by the Russian Academic Excellence Project ‘5-100’.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcos Escobar-Anel.

Additional information

Publisher's Note

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

Appendix: Proofs

Appendix: Proofs

Proof of Lemma 2

Let \(g^{(j)}\) be in \({{\,\mathrm{C}\,}}^{1,2}(\mathbb {R}_{+}\times \mathbb {R}_{+}\times \mathbb {R}^{d+1}_{+}\times \mathbb {R}_{+})\), with \(j=1, \ldots ,2d+3\). Using (8), it can be verified that the price of the jth stock option satisfies the following identity

$$\begin{aligned} \mathcal {L}g^{(j)}&=rg^{(j)}+\eta _{1}\mathbb {b}_{1}^{2} \mathbb {v}^{{{\,\mathrm{T}\,}}}sg^{(j)}_{s}+[\eta _{1}\mathbb {a}_{1} +\mathbb {a}_{2}]{{\,\mathrm{diag}\,}}[\mathbb {v}]\left[ {{\,\mathrm{D}\,}}^{1}_{\mathbb {v}}g^{(j)} \right] ^{{{\,\mathrm{T}\,}}}\nonumber \\&\quad +\eta _{3}\sigma _{2}v^{(d+2)}g^{(j)}_{v^{(d+2)}} +\left[ \lambda -\lambda ^{Q}\right] \left[ \mathbb {b}_{2}\mathbb {v}^{{{\,\mathrm{T}\,}}} +av^{(d+1)}\right] \varDelta g^{(j)}. \end{aligned}$$
(34)

The operator \(\mathcal {L}\) is the infinitesimal generator of the process \(\left( S_{t},\mathbb {V}^{(1)}_{t},V^{(2)}_{t}\right) \) which is given by

$$\begin{aligned} \mathcal {L} f= & {} \partial _{t}f+s\left[ r+\eta _{1}\mathbb {b}_{1}^{2} \mathbb {v}^{{{\,\mathrm{T}\,}}}-\mu \lambda ^{Q}\left[ \mathbb {b}_{2}\mathbb {v}^{{{\,\mathrm{T}\,}}} +av^{(d+2)}\right] \right] f_{s}+\mathbb {k}_{1}{{\,\mathrm{diag}\,}}\left[ \bar{\vartheta }_{1} -\mathbb {v}\right] {{\,\mathrm{D}\,}}^{1}_{\mathbb {v}}f^{{{\,\mathrm{T}\,}}}\nonumber \\&+\kappa _{2}\left[ \vartheta _{2}-\mathbb {v}\right] f_{v^{(d+2)}} +{{\,\mathrm{tr}\,}}\left[ \tilde{a}{{\,\mathrm{D}\,}}^{2}f\right] +\lambda \left[ \mathbb {b}_{2} \mathbb {v}^{{{\,\mathrm{T}\,}}}+av^{(d+2)}\right] \varDelta f, \end{aligned}$$
(35)

where \(f\in {{\,\mathrm{C}\,}}^{1,2}(\mathbb {R}_{+}\times \mathbb {R}_{+}\times \mathbb {R}^{d+1}_{+}\times \mathbb {R}_{+})\), \({{\,\mathrm{D}\,}}^{2}\) is the matrix operator of the second derivatives with respect \(\left( s,\mathbb {v},v^{(d+2)}\right) \), \(\mathbb {k}_{1}:=(\kappa _{1,1},\ldots ,\kappa _{1,d+1})\), \(\bar{\vartheta }_{1}:=(\vartheta _{1,1},\ldots ,\vartheta _{1,d+1})\) and \(\tilde{a}:=\dfrac{1}{2}\tilde{\sigma }\tilde{\sigma }^{{{\,\mathrm{T}\,}}}\), with \(\tilde{\sigma }=(\tilde{\sigma }_{i,j})\) a matrix function of size \((d+3)\times (2d+3)\) whose components are given by

$$\begin{aligned} \tilde{\sigma }_{ij}(s,\mathbb {v},v^{(d+2)})= {\left\{ \begin{array}{ll} b_{1,j}s\sqrt{v^{(j)}}, &{}\text {if}\ i=1,\ j=1,\ldots ,d+1,\\ \sigma _{1,i-1}\rho _{1,i-1}\sqrt{v^{(i-1)}}, &{}\text {if}\ i=2,\ldots ,d+2,\ j=i-1,\\ \sigma _{1,i-1}[1-\rho _{1,i-1}^{2}]^{\frac{1}{2}}\sqrt{v^{(i-1)}}, &{}\text {if}\ i=2,\ldots ,d+2,\ j=d+i,\\ \sigma _{2}\sqrt{v^{(d+2)}}, &{}\text {if}\ i=d+3,\ j=2d+3,\\ 0,&{}\text {otherwise}. \end{array}\right. } \end{aligned}$$
(36)

Applying Itô formula in \(g^{(j)}(t,S_{t},\mathbb {V}^{(1)}_{t},V^{(2)}_{t})\) and using (34)–(36), we see that

$$\begin{aligned} \mathrm {d} g^{(j)}&=\mathcal {L} g^{(j)}\mathrm {d} t+{{\,\mathrm{D}\,}}^{1}g^{(j)}\tilde{\sigma }\mathrm {d} B_{t}^{{{\,\mathrm{T}\,}}}+\varDelta g^{(j)}\left[ \mathrm {d} N_{t}-\lambda Z_{t}\mathrm {d} t\right] \\&=rg^{(j)}\mathrm {d} t+\left[ S_{t}g_{s}\mathbb {b}_{1}+\dfrac{\mathbb {a}_{1}}{\mathbb {b}_{1}} {{\,\mathrm{diag}\,}}\left[ {{\,\mathrm{D}\,}}_{\mathbb {v}}g^{(j)}\right] \right] \nonumber \\&\quad \times \left[ \eta _{1}\left[ \mathbb {b}_{1}{{\,\mathrm{diag}\,}}\left[ \mathbb {V}^{(1)}_{t}\right] \right] ^{{{\,\mathrm{T}\,}}}\mathrm {d} t+\sqrt{{{\,\mathrm{diag}\,}}{\mathbb {V}^{(1)}_{t}}}\mathrm {d} \mathbb {B}^{(1){{\,\mathrm{T}\,}}}_{t}\right] \\&\quad +\dfrac{\mathbb {a}_{2}}{\bar{\eta }_{2}}{{\,\mathrm{diag}\,}}\left[ {{\,\mathrm{D}\,}}_{\mathbb {v}}g^{(j)}\right] \left[ \left[ \bar{\eta }_{2} {{\,\mathrm{diag}\,}}\left[ \mathbb {V}^{(1)}_{t}\right] \right] ^{{{\,\mathrm{T}\,}}}\mathrm {d} t+ \sqrt{{{\,\mathrm{diag}\,}}{\mathbb {V}^{(1)}_{t}}}\mathrm {d} \mathbb {B}^{(2){{\,\mathrm{T}\,}}}_{t}\right] \\&\quad +\sigma _{2}g^{(j)}_{v^{(d+2)}}\left[ \eta _{3}V^{(2)}_{t}\mathrm {d} t+ \sqrt{V^{(2)}_{t}}\mathrm {d} B^{(3)}_{t}\right] \nonumber \\&\quad +\left[ \left[ \lambda -\lambda ^{Q}\right] Z_{t}\mathrm {d} t+\mathrm {d} N_{t}-\lambda Z_{t}\mathrm {d} t\right] \varDelta g^{(j)}, \end{aligned}$$

where \(B:=\left( \mathbb {B}^{(1)},\mathbb {B}^{(2)},B^{(3)}\right) \) and \({{\,\mathrm{D}\,}}^{1}\) is the gradient with respect to \(\left( s,\mathbb {v},v^{(d+2)}\right) \). \(\square \)

Proof of Proposition 5

Suppose that a solution f to the HJB equation (14) has the form of Eq. (16), i.e., \(f=J\). Take first and second derivatives to \(f(t,w,\mathbb {v},v^{(d+2)})\). Now, applying these in (14) and dividing by \(\gamma f\), the following identity can be verified

$$\begin{aligned}&\max _{\tilde{\theta }}\biggr \{-\left[ h^{\prime }+H^{(1)\prime }\mathbb {v}^{{{\,\mathrm{T}\,}}}+H^{(2)\prime }v_{(d+2)}\right] +\mathbb {k}_{1}{{\,\mathrm{diag}\,}}\left[ \bar{\vartheta }_{1}-\mathbb {v}\right] \left[ H^{(1)}\right] ^{{{\,\mathrm{T}\,}}}+\kappa _{2}\left[ \vartheta _{2}-v^{(d+2)} \right] H^{(2)}\nonumber \\&\quad +\dfrac{\lambda \left[ \mathbb {b}_{2}\mathbb {v}^{{{\,\mathrm{T}\,}}} +av^{(d+2)}\right] }{\gamma }\left[ \left[ 1+\mu \theta ^{N}\right] ^{1-\gamma } \exp \left[ \gamma \left[ H^{(1)}\bar{\mu }_{1}^{{{\,\mathrm{T}\,}}}+\mu _{2}H^{(2)}\right] \right] -1\right] \nonumber \\&\quad +\dfrac{1-\gamma }{\gamma }\Big [r-\mu \lambda ^{Q}\left[ \mathbb {b}_{2} \mathbb {v}^{{{\,\mathrm{T}\,}}}+av^{(d+2)}\right] \theta ^{N}+\eta _{1} \bar{\theta }^{(1)}\left[ \mathbb {b}_{1}{{\,\mathrm{diag}\,}}\left[ \mathbb {v}\right] \right] ^{{{\,\mathrm{T}\,}}}+\bar{\theta }^{(2)}[\bar{\eta }_{2} {{\,\mathrm{diag}\,}}[\mathbb {v}]]^{{{\,\mathrm{T}\,}}}\nonumber \\&\quad +\eta _{3}v^{(d+2)}\theta ^{(3)}\Big ]+\dfrac{\gamma }{2} \left[ \bar{\sigma }_{1}^{2}{{\,\mathrm{diag}\,}}[\mathbb {v}]\left[ [H^{(1)}]^{2} \right] ^{{{\,\mathrm{T}\,}}}+\sigma _{2}^{2}v^{(d+2)}\left[ H^{(2)}\right] ^{2}\right] \nonumber \\&\quad -\dfrac{[1-\gamma ]}{2}[\bar{\theta }^{(1)}{{\,\mathrm{diag}\,}}[\mathbb {v}] \bar{\theta }^{(1){{\,\mathrm{T}\,}}}+\bar{\theta }^{(2)}{{\,\mathrm{diag}\,}}[\mathbb {v}] \bar{\theta }^{(2){{\,\mathrm{T}\,}}}+[\theta ^{(3)}]^{2}v^{(d+2)}] \nonumber \\&\quad +[1-\gamma ]\biggr [\biggr [\dfrac{\mathbb {a}_{1}}{\mathbb {b}_{1}}{{\,\mathrm{diag}\,}}[\bar{\theta }^{(1)}]+\dfrac{\mathbb {a}_{2}}{\bar{\eta }_{2}}{{\,\mathrm{diag}\,}}[\bar{\theta }^{(2)}]\biggl ] {{\,\mathrm{diag}\,}}[\mathbb {v}]\left[ H^{(1)}\right] ^{{{\,\mathrm{T}\,}}}+\sigma _{2}v^{(d+2)} \theta ^{(3)}H^{(2)}\biggl ]\biggl \}=0. \end{aligned}$$
(37)

It is easy to realize that the solution \(\tilde{\theta }^{*}\) to the maximizing problem in (37) is determined by (17). Substituting (17) in (37) and grouping terms multiplied by \(\mathbb {v}\) and \(v^{(d+2)}\), respectively, we have that

$$\begin{aligned}&-h^{\prime }+\dfrac{1-\gamma }{\gamma }r+\mathbb {k}_{1}{{\,\mathrm{diag}\,}}\left[ \bar{\vartheta }_{1}\right] \left[ H^{(1)}\right] ^{{{\,\mathrm{T}\,}}}+\kappa _{2}\vartheta _{2}H^{(2)}+\frac{\gamma }{2}\bar{\sigma }_{1}^{2}{{\,\mathrm{diag}\,}}\left[ \left[ H^{(1)}\right] ^{2}\right] \nonumber \\&\quad +\biggr [-H^{(1)\prime }+\biggr [[1-\gamma ]\biggr [\dfrac{\mathbb {a}_{1}}{\mathbb {b}_{1}} {{\,\mathrm{diag}\,}}\left[ \bar{\theta }^{(1)*}\right] +\dfrac{\mathbb {a}_{2}}{\bar{\eta }_{2}}{{\,\mathrm{diag}\,}}\left[ \bar{\theta }^{(2)*}\right] \biggl ] -\mathbb {k}_{1}\biggl ]{{\,\mathrm{diag}\,}}\left[ H^{(1)}\right] \nonumber \\&\quad +\dfrac{1}{\gamma }\left[ \lambda \left[ \left[ 1+\mu \theta ^{N*} \right] ^{1-\gamma }\exp \left[ \gamma \left[ H^{(1)}\bar{\mu }_{1}^{{{\,\mathrm{T}\,}}} +\mu _{2}H^{(2)}\right] \right] -1\right] -[1-\gamma ]\mu \lambda ^{Q} \theta ^{N*}\right] \mathbb {b}_{2}\nonumber \\&\quad +\frac{1-\gamma }{\gamma }\left[ \eta _{1}\mathbb {b}_{1}{{\,\mathrm{diag}\,}}\left[ \bar{\theta }^{(1)*}\right] +\bar{\eta }_{2}{{\,\mathrm{diag}\,}}\left[ \bar{\theta }^{(2)*}\right] \right] -\dfrac{[1-\gamma ]}{2} \left[ \left[ \bar{\theta }^{(1)*}\right] ^{2}+\left[ \bar{\theta }^{(2)*} \right] ^{2}\right] \biggl ]\mathbb {v}^{{{\,\mathrm{T}\,}}} \nonumber \\&\quad +v^{(d+2)}\biggr [-H^{(2)\prime }+\left[ [1-\gamma ]\sigma _{2} \theta ^{(3)*}-\kappa _{2}\right] H^{(2)}+\dfrac{\gamma \sigma _{2}^{2}}{2} \left[ H^{(2)}\right] ^{2}\nonumber \\&\quad +\dfrac{a}{\gamma }\left[ \lambda \left[ \left[ 1+\mu \theta ^{N*} \right] ^{1-\gamma }\exp \left[ \gamma \left[ H^{(1)}\bar{\mu }_{1}^{{{\,\mathrm{T}\,}}} +\mu _{2}H^{(2)}\right] \right] -1\right] -[1-\gamma ]\mu \lambda ^{Q} \theta ^{N*}\right] \nonumber \\&\quad +\dfrac{1-\gamma }{\gamma }\eta _{3}\theta ^{(3)*}-\dfrac{[1-\gamma ]}{2}\left[ \theta ^{(3)*}\right] ^{2} \biggl ]=0. \end{aligned}$$
(38)

Notice that Eq. (38) is equivalent to solving the generalized Riccati system in (18). By Theorem 2.14 of Keller-Ressel and Mayerhofer (2015), it follows that (18) has unique solutions which are given by h, \(H^{(1)}=(H^{(1,1)},\ldots ,H^{(1,d+1)})\) and \(H^{(2)}\). \(\square \)

Proof of Proposition 6

Let \(f^{\varPi }\) be as the right side of (20). We shall prove that \(f^{\varPi }\) satisfies (19). To see that \(J^{\varPi }=f^{\varPi }\), where \(J^{\varPi }\) is given by (13), we remit the reader to Escobar et al. (2015). Notice that (19) is equivalent to

$$\begin{aligned}&-h_{\varPi }^{\prime }+\dfrac{1-\gamma }{\gamma }r+\mathbb {k}_{1} {{\,\mathrm{diag}\,}}[\bar{\vartheta }_{1}]H_{\varPi }^{(1){{\,\mathrm{T}\,}}}+\kappa _{2}\vartheta _{2} H_{\varPi }^{(2)} \\&\quad +\biggr [-H_{\varPi }^{(1)\prime }+\biggr [[1-\gamma ] \biggr [\dfrac{\mathbb {a}_{1}}{\mathbb {b}_{1}}{{\,\mathrm{diag}\,}}[\bar{\theta }^{(1)}] +\dfrac{\mathbb {a}_{2}}{\bar{\eta }_{2}}{{\,\mathrm{diag}\,}}[\bar{\theta }^{(2)}]\biggl ] -\mathbb {k}_{1}\biggl ]{{\,\mathrm{diag}\,}}[H_{\varPi }^{(1)}]\\&\quad +\frac{\gamma }{2}\bar{\sigma }_{1}^{2}{{\,\mathrm{diag}\,}}[[H_{\varPi }^{(1)}]^{2}] +\dfrac{1}{\gamma }[\lambda [[1+\mu \theta ^{N}]^{1-\gamma } \exp [\gamma [H_{\varPi }^{(1)}\bar{\mu }_{1}^{{{\,\mathrm{T}\,}}}+\mu _{2}H_{\varPi }^{(2)}]]-1]\\&\quad -[1-\gamma ]\mu \lambda ^{Q}\theta ^{N}]\mathbb {b}_{2}\\&\quad +\frac{1-\gamma }{\gamma }[\eta _{1}\mathbb {b}_{1} {{\,\mathrm{diag}\,}}[\bar{\theta }^{(1)}]+\bar{\eta }_{2}{{\,\mathrm{diag}\,}}[\bar{\theta }^{(2)}]] -\dfrac{[1-\gamma ]}{2}[[\bar{\theta }^{(1)}]^{2}+[\bar{\theta }^{(2)}]^{2}] \biggl ]\mathbb {v}^{{{\,\mathrm{T}\,}}}\\&\quad +v^{(d+2)}\biggr [-H_{\varPi }^{(2)\prime }+[[1-\gamma ]\sigma _{2} \theta ^{(3)}-\kappa _{2}]H_{\varPi }^{(2)}+\dfrac{\gamma \sigma _{2}^{2}}{2} [H_{\varPi }^{(2)}]^{2}\\&\quad +\dfrac{a}{\gamma }[\lambda [[1+\mu \theta ^{N}]^{1-\gamma } \exp [\gamma [H_{\varPi }^{(1)}\bar{\mu }_{1}^{{{\,\mathrm{T}\,}}}+\mu _{2}H_{\varPi }^{(2)}]] -1]-[1-\gamma ]\mu \lambda ^{Q}\theta ^{N}]\\&\quad +\dfrac{1-\gamma }{\gamma }\eta _{3}\theta ^{(3)}-\dfrac{[1 -\gamma ]}{2}[\theta ^{(3)}]^{2} \biggl ]=0. \end{aligned}$$

From here we see that \(f^{\varPi }\) is a solution of (19) if \(h_{\varPi },H^{(1)}_{\varPi }, H^{(2)}_{\varPi }\) are solutions to the Riccati system given in (21). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Escobar-Anel, M., Moreno-Franco, H.A. Dynamic portfolio strategies under a fully correlated jump-diffusion process. Ann Finance 15, 421–453 (2019). https://doi.org/10.1007/s10436-019-00350-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10436-019-00350-3

Keywords

JEL Classification

Navigation