Log in

Mean–Variance Portfolio Selection Under Volterra Heston Model

  • Published:
Applied Mathematics & Optimization Aims and scope Submit manuscript

Abstract

Motivated by empirical evidence for rough volatility models, this paper investigates continuous-time mean–variance (MV) portfolio selection under the Volterra Heston model. Due to the non-Markovian and non-semimartingale nature of the model, classic stochastic optimal control frameworks are not directly applicable to the associated optimization problem. By constructing an auxiliary stochastic process, we obtain the optimal investment strategy, which depends on the solution to a Riccati–Volterra equation. The MV efficient frontier is shown to maintain a quadratic curve. Numerical studies show that both roughness and volatility of volatility materially affect the optimal strategy.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

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. See, for example, Oxford-Man Institute’s realized library at https://realized.oxford-man.ox.ac.uk/data.

  2. There are several equivalent formulations.

  3. Available at https://github.com/differint/differint.

References

  1. Gatheral, J., Jaisson, T., Rosenbaum, M.: Volatility is rough. Quant. Financ. 18, 933–949 (2018)

    Article  MathSciNet  Google Scholar 

  2. El Euch, O., Rosenbaum, M.: The characteristic function of rough Heston models. Math. Financ. 29, 3–38 (2019)

    Article  MathSciNet  Google Scholar 

  3. Guennoun, H., Jacquier, A., Roome, P., Shi, F.: Asymptotic behavior of the fractional Heston model. SIAM J. Financ. Math. 9, 1017–1045 (2018)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  5. Fukasawa, M.: Asymptotic analysis for stochastic volatility: martingale expansion. Financ. Stoch. 15, 635–654 (2011)

    Article  MathSciNet  Google Scholar 

  6. El Euch, O., Fukasawa, M., Rosenbaum, M.: The microstructural foundations of leverage effect and rough volatility. Financ. Stoch. 22, 241–280 (2018)

    Article  MathSciNet  Google Scholar 

  7. El Euch, O., Rosenbaum, M.: Perfect hedging in rough Heston models. Ann. Appl. Probab. 28, 3813–3856 (2018)

    MathSciNet  MATH  Google Scholar 

  8. Abi Jaber, E., Larsson, M., Pulido, S.: Affine Volterra processes. Ann. Appl. Probab. 29, 3155–3200 (2019)

    Article  MathSciNet  Google Scholar 

  9. Keller-Ressel, M., Larsson, M., Pulido, S.: Affine rough models. ar**v:1812.08486 (2018)

  10. Fouque, J.P., Hu, R.: Optimal portfolio under fast mean-reverting fractional stochastic environment. SIAM J. Financ. Math. 9, 564–601 (2018)

    Article  MathSciNet  Google Scholar 

  11. Fouque, J.P., Hu, R.: Optimal portfolio under fractional stochastic environment. Math. Financ. 29, 697–734 (2018)

    Article  MathSciNet  Google Scholar 

  12. Bäuerle, N., Desmettre, S.: Portfolio optimization in fractional and rough Heston models. ar**v:1809.10716 (2018)

  13. Zhou, X.Y., Li, D.: Continuous-time mean–variance portfolio selection: a stochastic LQ framework. Appl. Math. Optim. 42, 19–33 (2000)

    Article  MathSciNet  Google Scholar 

  14. Lim, A.E., Zhou, X.Y.: Mean–variance portfolio selection with random parameters in a complete market. Math. Oper. Res. 27, 101–120 (2002)

    Article  MathSciNet  Google Scholar 

  15. Lim, A.E.: Quadratic hedging and mean–variance portfolio selection with random parameters in an incomplete market. Math. Oper. Res. 29, 132–161 (2004)

    Article  MathSciNet  Google Scholar 

  16. Černý, A., Kallsen, J.: Mean-variance hedging and optimal investment in Heston’s model with correlation. Math. Financ. 18, 473–492 (2008)

    Article  MathSciNet  Google Scholar 

  17. Jeanblanc, M., Mania, M., Santacroce, M., Schweizer, M.: Mean-variance hedging via stochastic control and BSDEs for general semimartingales. Ann. Appl. Probab. 22, 2388–2428 (2012)

    Article  MathSciNet  Google Scholar 

  18. Shen, Y.: Mean–variance portfolio selection in a complete market with unbounded random coefficients. Automatica 55, 165–175 (2015)

    Article  MathSciNet  Google Scholar 

  19. Glasserman, P., He, P.: Buy rough, sell smooth. Quant. Financ. 1–16 (2019)

  20. Revuz, D., Yor, M.: Continuous martingales and Brownian motion, vol. 293. Springer, New York (1999)

    Book  Google Scholar 

  21. Gripenberg, G., Londen, S.O., Staffans, O.: Volterra Integral and Functional Equations, vol. 34. Cambridge University Press, Cambridge (1990)

    Book  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  23. Zeng, X., Taksar, M.: A stochastic volatility model and optimal portfolio selection. Quant. Financ. 13, 1547–1558 (2013)

    Article  MathSciNet  Google Scholar 

  24. Shen, Y., Zeng, Y.: Optimal investment-reinsurance strategy for mean–variance insurers with square-root factor process. Insur. Math. Econom. 62, 118–137 (2015)

    Article  MathSciNet  Google Scholar 

  25. Abi Jaber, E., El Euch, O.: Multifactor approximation of rough volatility models. SIAM J. Financ. Math. 10, 309–349 (2019)

    Article  MathSciNet  Google Scholar 

  26. Mytnik, L., Salisbury, T.S.: Uniqueness for Volterra-type stochastic integral equations. ar**v:1502.05513 (2015)

  27. Gerhold, S., Gerstenecker, C., Pinter, A.: Moment explosions in the rough Heston model. Decis. Econ. Financ. 42, 575–608 (2019)

    Article  MathSciNet  Google Scholar 

  28. Yong, J., Zhou, X.Y.: Stochastic Controls: Hamiltonian Systems and HJB Equations, vol. 43. Springer, New York (1999)

    Book  Google Scholar 

  29. Luenberger, D.G.: Optimization by Vector Space Methods. Wiley, Hoboken (1968)

    Google Scholar 

  30. Veraar, M.: The stochastic Fubini theorem revisited. Stochastics 84, 543–551 (2012)

    Article  MathSciNet  Google Scholar 

  31. Diethelm, K., Ford, N.J., Freed, A.D.: A predictor-corrector approach for the numerical solution of fractional differential equations. Nonlinear Dynam. 29, 3–22 (2002)

    Article  MathSciNet  Google Scholar 

  32. Diethelm, K., Ford, N.J., Freed, A.D.: Detailed error analysis for a fractional Adams method. Numer. Algorithms 36, 31–52 (2004)

    Article  MathSciNet  Google Scholar 

  33. Li, C., Tao, C.: On the fractional Adams method. Comput. Math. Appl. 58, 1573–1588 (2009)

    Article  MathSciNet  Google Scholar 

  34. Abi Jaber, E.: Lifting the Heston model. Quant. Financ. 19, 1995–2013 (2019)

    Article  MathSciNet  Google Scholar 

  35. Brunner, H.: Volterra Integral Equations: An Introduction to Theory and Applications, vol. 30. Cambridge University Press, Cambridge (2017)

    Book  Google Scholar 

  36. Gatheral, J., Keller-Ressel, M.: Affine forward variance models. Financ. Stoch. 23, 501–533 (2019)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors would like to thank two anonymous referees and the Editor for their careful reading and valuable comments, which have greatly improved the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bingyan Han.

Ethics declarations

Conflict of interest

The authors declare that they have no conflicts of interest.

Additional information

Publisher's Note

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

Appendices

Appendix A: Solutions of Riccati–Volterra Equations

To demonstrate the existence and uniqueness of the solution to a Riccati–Volterra equation, we first rephrase the following result from a recent monograph [35] with more general assumptions. The underlying idea of the proof is the Picard iteration.

Theorem A.1

Suppose kernel \(K(\cdot )\) is bounded or is the fractional kernel with \(\alpha \in (0, 1)\). Let \(c_0, c_1, c_2\) be constant. Then there exists \(\delta >0\) such that

$$\begin{aligned} f(t) = \int ^t_0 K(t-s) \Big [ c_0 + c_1 f(s) + c_2 f^2(s) \Big ] ds \end{aligned}$$
(A.1)

has a unique continuous solution f on \([0, \delta ]\).

Proof

Note that quadratic function is locally Lipschitz; then according to Theorem 3.1.2 and Theorem 3.1.4 in [35], the claim holds. \(\square \)

However, \(\delta \) in Theorem A.1 is not explicit. Tighter results exist if more assumptions are imposed.

We investigate g(at) in (2.8) first. Based on [36, Theorem A.5], we have

Lemma A.1

Suppose Assumption 2.1 holds and \(\kappa ^2 - 2a\sigma ^2 > 0\). Then (2.8) has a unique global continuous solution. Moreover,

$$\begin{aligned} 0< g(a, t) \le r_2(t) < w_*, \quad \forall \; t > 0, \end{aligned}$$
(A.2)

where \(w_* \triangleq \frac{\kappa - \sqrt{\kappa ^2 - 2a \sigma ^2}}{\sigma ^2}\) and \(r_2(t) \triangleq Q^{-1}_2 \Big ( \int ^t_0 K(s) ds \Big )\); that is, the inverse function of \(Q_2\), given by

$$\begin{aligned} Q_2(w) = \int ^w_0 \frac{du}{a - \kappa u + \frac{\sigma ^2}{2} u^2}. \end{aligned}$$
(A.3)

Proof

To apply the result in [36, Theorem A.5], we define

$$\begin{aligned} H(w) = a - \kappa w + \frac{\sigma ^2}{2} w^2. \end{aligned}$$

Then H(w) satisfies Assumption A.1 in [36] with \(w_{max} \triangleq \frac{\kappa }{\sigma ^2}\) and \(w_*\) defined above. The claim follows from [36, Theorem A.5(c)] with \(a(t) \equiv 0\) in their theorem. \(\square \)

For the specific fractional kernel \(K(t) = \frac{t^{\alpha -1}}{\varGamma (\alpha )}\), [7, Theorem 3.2] obtains the following tighter results and the proof is based on the scaling limits of the Hawkes processes.

Lemma A.2

If \(K(t) = \frac{t^{\alpha -1}}{\varGamma (\alpha )}\), \(\alpha \in (1/2, 1)\), then g(at) in (2.8) satisfies

$$\begin{aligned} g(a, t) \le \frac{c}{\sigma ^2} \Big [ \kappa + \frac{t^{-\alpha }}{\varGamma (1-\alpha )} + \sigma \sqrt{ a_0(t) - a} \Big ], \end{aligned}$$
(A.4)

with \(a_0(t) = \frac{1}{2\sigma ^2} \Big [ \kappa + \frac{t^{-\alpha }}{\varGamma (1-\alpha )} \Big ]^2\) and a constant \(c>0\). In other words, if \( a < a_0(T)\), then Assumption 2.3 is satisfied.

Next, we study \(\psi (\cdot )\) in (4.7). (4.7) has a unique continuous solution on some interval \([0, \delta ]\) if the conditions in Theorem A.1 are satisfied. Without Theorem A.1, we also have the following result.

Lemma A.3

Suppose Assumption 2.1 holds.

  1. (1)

    If \(1 - 2\rho ^2 > 0\), then (4.7) has a unique global continuous solution \(\psi \in L^2_{loc}({\mathbb {R}}_+, {\mathbb {R}})\) and \(\psi < 0\) for \(t>0\).

  2. (2)

    If \(1 - 2\rho ^2 = 0\), then (4.7) is linear and has a unique continuous solution on [0, T].

  3. (3)

    If \(1 - 2\rho ^2 < 0\), further assume \(\lambda >0\) and \(\lambda ^2 + 2(1-2\rho ^2)\theta ^2\sigma ^2 >0\). Then (4.7) has a unique global continuous solution. Moreover,

    $$\begin{aligned} \frac{{\bar{w}}_*}{1-2\rho ^2}< \frac{{\bar{r}}_2(t)}{1-2\rho ^2} \le \psi (t) < 0, \quad \forall \; t > 0, \end{aligned}$$
    (A.5)

    with \({\bar{w}}_* = \frac{\lambda - \sqrt{\lambda ^2 + 2(1-2\rho ^2)\theta ^2\sigma ^2}}{ \sigma ^2}\) and \({\bar{r}}_2(t) \triangleq {\bar{Q}}^{-1}_2 \Big ( \int ^t_0 K(s) ds \Big )\), where

    $$\begin{aligned} {\bar{Q}}_2(w) = \int ^w_0 \frac{du}{ \frac{\sigma ^2}{2} u^2 - \lambda u - (1-2\rho ^2)\theta ^2}. \end{aligned}$$
    (A.6)

Proof

The claim in (1) follows from [8, Theorem 7.1]. The continuity follows from the uniqueness of the global solution and [21, Theorem 12.1.1]. The claim in (2) is classic and can be found in [35, Theorem 1.2.3]. For (3), we consider \( {\tilde{\psi }} = (1 - 2 \rho ^2) \psi \). Then \({\tilde{\psi }} \) satisfies

$$\begin{aligned} {\tilde{\psi }} = K * \Big ( \frac{\sigma ^2}{2} {\tilde{\psi }}^2 - \lambda {\tilde{\psi }} - (1 - 2\rho ^2) \theta ^2 \Big ). \end{aligned}$$
(A.7)

Define

$$\begin{aligned} H(w) = \frac{\sigma ^2}{2} w^2 - \lambda w - (1 - 2\rho ^2) \theta ^2. \end{aligned}$$
(A.8)

Then \({\bar{w}}_*\) is the unique root of \(H(w) = 0\) on \((-\infty , {\bar{w}}_{max}]\) with \({\bar{w}}_{max} = \frac{\lambda }{\sigma ^2}\). H(w) satisfies Assumption A.1 in [36]. Therefore, [36, Theorem A.5 (c)] with \(a(t) \equiv 0\) implies (A.7) has a unique global continuous solution and

$$\begin{aligned} 0< {\tilde{\psi }}(t) \le {\bar{r}}_2(t) < {\bar{w}}_*, \quad \forall \; t > 0. \end{aligned}$$
(A.9)

Note \( {\tilde{\psi }} = (1 - 2 \rho ^2) \psi \). This gives the result desired. \(\square \)

Appendix B: Positivity of Integrals with Forward Variance

Lemma B.1

Suppose Assumption 2.1 holds. The forward variance \(\xi _t(s)\) in (4.3) satisfies \(\int ^T_t \xi _t(s) ds > 0\), \({\mathbb {P}}\)-\({\mathrm{a.s.}}\), for every \( t \in [0, T)\).

Proof

As \(\int ^T_t \xi _t(s) ds = \mathbb {{\tilde{E}}}[\int ^T_t V_s ds| {{\mathcal {F}}}_t]\) and \(V_s\) is non-negative by Theorem 2.2, it is sufficient to show that \(\int ^T_t V_s ds > 0\), \({\mathbb {P}}\)-\({\mathrm{a.s.}}\).

Given \(t \in [0, T)\), for \(\omega \in \varOmega \) such that \(V_s(\omega )\) is continuous in s, we suppose \(\int ^T_t V_s(\omega ) ds = 0\). By the continuity of \(V_s(\omega )\), \(V_s(\omega ) = 0\) for \(s \in [t, T]\). Using the argument given in [8, Theorem 3.5, Equation (3.8)], for \(0< h < T - t\), we have

$$\begin{aligned} V_{t + h}(\omega )&= V_0 + \int _{0}^{t} K(t+h -s)\left( \kappa \phi - \lambda V_s (\omega ) \right) d s \nonumber \\&\quad + \int _{0}^{t} K(t+h-s) \sigma \sqrt{V_{s} (\omega )} d {\tilde{B}}_{s} (\omega ) \nonumber \\&\quad + \int _{t}^{t+h} K(t+h -s)\left( \kappa \phi - \lambda V_{s} (\omega ) \right) d s \nonumber \\&\quad + \int _{t}^{t+h} K(t+h-s) \sigma \sqrt{V_{s} (\omega ) } d {\tilde{B}}_{s} (\omega ) \nonumber \\&\ge \int _{t}^{t+h} K(t+h -s)\left( \kappa \phi - \lambda V_{s} (\omega ) \right) d s \nonumber \\&\quad + \int _{t}^{t+h} K(t+h-s) \sigma \sqrt{V_{s} (\omega ) } d {\tilde{B}}_{s}(\omega ). \end{aligned}$$
(B.1)

As \(V_s (\omega ) = 0,\, s \in [t, t+h]\), then

$$\begin{aligned} V_{t + h} (\omega ) \ge \kappa \phi \int _{t}^{t+h} K(t+h - s) ds > 0. \end{aligned}$$
(B.2)

This contradiction implies that \(\int ^T_t V_s ds > 0\), \({\mathbb {P}}\)-\({\mathrm{a.s.}}\), and the claim follows. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Han, B., Wong, H.Y. Mean–Variance Portfolio Selection Under Volterra Heston Model. Appl Math Optim 84, 683–710 (2021). https://doi.org/10.1007/s00245-020-09658-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00245-020-09658-3

Keywords

Mathematics Subject Classification

Navigation