Skip to main content
Log in

Robust Portfolio Optimization with Multi-Factor Stochastic Volatility

  • Published:
Journal of Optimization Theory and Applications Aims and scope Submit manuscript

Abstract

This paper studies a robust portfolio optimization problem under a multi-factor volatility model. We derive optimal strategies analytically under the worst-case scenario with or without derivative trading in complete and incomplete markets and for assets with jump risk. We extend our study to the case with correlated volatility factors and propose an analytical approximation for the robust optimal strategy. To illustrate the effects of ambiguity, we compare our optimal robust strategy with the strategies that ignore the information of uncertainty, and provide the welfare analysis. We also discuss how derivative trading affects the optimal strategies. Finally, numerical experiments are provided to demonstrate the behavior of the optimal strategy and the utility loss.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Cui, Z., Kirkby, J.L., Duy, N.: A general framework for discretely sampled realized variance derivatives in stochastic volatility models with jumps. Eur. J. Oper. Res. 262, 381–400 (2017)

    MathSciNet  MATH  Google Scholar 

  2. Cui, Z., Kirkby, J.L., Duy, N.: A general framework for time-changed Markov processes and applications. Eur. J. Oper. Res. 2, 785–800 (2019)

    MathSciNet  MATH  Google Scholar 

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

    Google Scholar 

  4. Ma, G., Zhu, S.P.: Optimal investment and consumption under a continuous-time cointegration model with exponential utility. Quant. Financ. 19(7), 1135–1149 (2019)

    MathSciNet  MATH  Google Scholar 

  5. Nandi, S.: How important is the correlation between returns and volatility in a stochastic volatility model? empirical evidence from pricing and hedging in the S&P 500 Index Options Market. J. Bank. Financ. 22, 589–610 (1998)

    Google Scholar 

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

    Google Scholar 

  7. Yang, B.Z., Yue, J., Huang, N.J.: Equilibrium price of variance swaps under stochastic volatility with Lévy jumps and stochastic interest rate. Int. J. Theor. Appl. Finance (IJTAF) 22(04), 1–33 (2019)

    MATH  Google Scholar 

  8. Yang, B.Z., He, X.J., Huang, N.J.: Equilibrium price and optimal insider trading strategy under stochastic liquidity with long memory. Appl. Math. Optim. (2020). https://doi.org/10.1007/s00245-020-09675-2

    Article  Google Scholar 

  9. Yue, J., Huang, N.J.: Neutral and indifference pricing with stochastic correlation and volatility. J Ind. Manag. Optim. 14(1), 199–229 (2018)

    MathSciNet  MATH  Google Scholar 

  10. Ma, G., Siu, C.C., Zhu, S.P.: Dynamic portfolio choice with return predictability and transaction costs. Eur. J. Oper. Res. 278(3), 976–988 (2019)

    MathSciNet  MATH  Google Scholar 

  11. Stein, J.C., Stein, E.M.: Stock price distributions with stochastic volatility: An analytic approach. Rev. Financ. Stud. 4(4), 727–752 (1991)

    MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  13. Schöbel, R., Zhu, J.: Stochastic volatility with an Ornstein–Uhlenbeck process: an extension. Rev. Financ. 3(1), 23–46 (1998)

    MATH  Google Scholar 

  14. 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)

    MATH  Google Scholar 

  15. Da Fonseca, J., Grasselli, M., Tebaldi, C.: A multifactor volatility Heston model. Quant. Financ. 8(6), 591–604 (2008)

    MathSciNet  MATH  Google Scholar 

  16. Li, G., Zhang, C.: On the number of state variables in option pricing. Manag. Sci. 56(11), 2058–2075 (2010)

    MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  18. Faidi, W., Mezghanni, H., Mnif, M.: Expected utility maximization problem under state constraints and model uncertainty. J. Optim. Theory Appl. 183(3), 1123–1152 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  19. Sun, X., Teo, K.L., Tang, L.: Dual approaches to characterize robust optimal solution sets for a class of uncertain optimization problems. J. Optim. Theory Appl. 182(3), 984–10000 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  20. Yuen, F.L., Yang, H.: Optimal asset allocation: a worst scenario expectation approach. J. Optim. Theory Appl. 153(3), 794–811 (2012)

    MathSciNet  MATH  Google Scholar 

  21. Aït-Sahalia, Y., Matthys, F.: Robust consumption and portfolio policies when asset prices can jump. J. Econ. Theory 179, 1–56 (2019)

    MathSciNet  MATH  Google Scholar 

  22. Ismail, A., Pham, H.: Robust Markowitz mean-variance portfolio selection under ambiguous covariance matrix. Math. Financ. 29(1), 174–207 (2019)

    MathSciNet  MATH  Google Scholar 

  23. Bergen, V., Escobar, M., Rubtsov, A., Zagst, R.: Robust multivariate portfolio choice with stochastic covariance in the presence of ambiguity. Quant. Financ. 18(8), 1265–1294 (2018)

    MathSciNet  MATH  Google Scholar 

  24. 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)

    MathSciNet  MATH  Google Scholar 

  25. Burns, P., Engle, R., Mezrich, J.: Correlations and volatilities of asynchronous data. J. Deriv. 5(4), 7–18 (1998)

    Google Scholar 

  26. Grzelak, L.A., Oosterlee, C.W.: On the Heston model with stochastic interest rates. SIAM J. Financ. Math. 2(1), 255–286 (2011)

    MathSciNet  MATH  Google Scholar 

  27. Grzelak, L.A., Oosterlee, C.W., Van Weeren, S.: Extension of stochastic volatility equity models with the Hull–White interest rate process. Quant. Financ. 12(1), 89–105 (2012)

    MathSciNet  MATH  Google Scholar 

  28. Uppal, R., Wang, T.: Model misspecification and under diversification. J. Financ. 58(6), 2465–2486 (2003)

    Google Scholar 

  29. Flor, C., Larsen, L.: Robust portfolio choice with stochastic interest rates. Annal. Financ. 10(2), 243–265 (2014)

    MathSciNet  MATH  Google Scholar 

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

    Google Scholar 

  31. Aït-Sahalia, Y., Kimmel, R.: Maximum-likelihood estimation for stochastic volatility models. J. Financ. Econ. 83(2), 413–452 (2007)

    Google Scholar 

  32. Bakshi, G., Kapadia, N.: Delta-hedged gains and the negative market volatility risk premium. Rev. Financ. Stud. 16(2), 527–566 (2003)

    Google Scholar 

  33. Black, F.: Studies in stock price volatility changes. In: Proceedings of the 1976 Business Meeting of the Business and Economic Statistics Section, American Statistical Association, pp. 177–181 (1976)

  34. Anderson, E., Hansen, L.P., Sargent, T.J.: A quartet of semi-groups for model specification, robustness, prices of risk, and model detection. J Eur. Econ. Assoc. 1(1), 68–123 (2003)

    Google Scholar 

  35. Kraft, H.: Optimal portfolios with stochastic interest rates and defaultable assets. Quant. Financ. 5(3), 303–313 (2004)

    MATH  Google Scholar 

  36. Grzelak, L.A., Oosterlee, C.W. Van Weeren, S.: Efficient option pricing with multi-factor equity-interest rate hybrid models. Delft University of Technology (2009)

  37. Johnson, N.L., Kotz, S., Balakrishnan, N.: Continuous Univariate Distributions, vol. 2, 2nd edn. Wiley, New York (1994)

    MATH  Google Scholar 

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

    Google Scholar 

  39. Faria, G., Correia-da-Silva, J.: A closed-form solution for options with ambiguity about stochastic volatility. Rev. Deriv.Res. 17(2), 125–159 (2014)

    MATH  Google Scholar 

  40. Fouque, J.P., Papanicolau, G., Sircar, K.R.: Derivatives in Financial Markets with Stochastic Volatility. Cambridge University Press, Cambridge (2000)

    MATH  Google Scholar 

  41. Lobo, B.J.: Jump risk in the US stock market: evidence using political information. Rev. Financ. Econ. 8(2), 149–163 (1999)

    Google Scholar 

  42. Naik, V., Lee, M.: General equilibrium pricing of otions on the market portfolio with discontinuous returns. Rev. Financ. Stud. 3(4), 493–521 (1990)

    Google Scholar 

  43. Yan, S.: Jump risk, stock returns, and slope of implied volatility smile. J. Financ. Econ. 99(1), 216–233 (2011)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xiaoping Lu.

Additional information

Communicated by Kok Lay Teo.

Publisher's Note

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

Appendices

Appendix A Proof of Proposition 3.1

Solving optimal problem (15) with respect to \(e^S_j\) and \(e_j^V\), we have

$$\begin{aligned} (e^S_j)^*=\varPsi _j^S(x\beta ^S_jJ_x+\rho _j\sigma _jJ_{v_j})\sqrt{v_j},\quad (e^V_j)^*=\varPsi _j^V(x\beta ^V_jJ_x+\sqrt{1-\rho _j^2}\sigma _jJ_{v_j})\sqrt{v_j}.\nonumber \\ \end{aligned}$$
(54)

Substituting (54) into (15), we obtain the following equation:

$$\begin{aligned} \begin{aligned}&\sup _{\beta ^S_i,\beta ^V_i} \bigg \{J_t+x\big (r+\sum _{j=1}^{2}(\beta ^{S}_j\lambda _jv_j +\beta ^{V}_j\mu _jv_j\big )J_x \\&\quad +\frac{1}{2}x^2\sum _{j=1}^{2}[(\beta ^{S}_j)^2+(\beta ^{V}_j)^2]v_jJ_{xx} +\sum _{j=1}^{2}\big (\kappa _j(\theta _j-v_j)\big )J_{v_j}\\&\quad +\frac{1}{2}\sum _{j=1}^{2}\sigma _j^2v_jJ_{v_jv_j}+\sum _{j=1}^{2}\sigma _jv_jx\big (\beta ^S_j\rho _j+\beta ^V_j\sqrt{1-\rho _j^2}\big )J_{v_jx}\\&\quad -\sum _{j=1}^{2}\frac{1}{2}\varPsi _j^Sv_j\left[ x^2(\beta ^S_j)^2J_x^2\right. \left. +2x\beta ^S_j\rho _j\sigma _jJ_xJ_{V_j}+\rho _j^2\sigma _j^2J^2_{v_j}\right] \\&\quad -\sum _{j=1}^{2}\frac{1}{2}\varPsi _j^Sv_j\left[ x^2(\beta ^V_j)^2J_x^2+2x\beta ^V_j\sqrt{1-\rho _j^2}\sigma _jJ_xJ_{v_j}+(1-\rho _j^2)\sigma _j^2J^2_{v_j}\right] \bigg \}=0. \end{aligned} \end{aligned}$$
(55)

Assuming that the solution J is of the form \(J(t,x,v_1,v_2)=\frac{x^{1-\gamma }}{1-\gamma }\exp (H_1(T-t)v_1+H_2(T-t)v_2+h(T-t))\) and choosing

$$\begin{aligned} \varPsi ^S_j=\frac{\phi _j^S}{(1-\gamma )J(t,x,v_1,v_2)}, \quad \varPsi ^V_j=\frac{\phi _j^V}{(1-\gamma )J(t,x,v_1,v_2)}, \quad \phi _j^S, \ \phi _j^V>0, \end{aligned}$$

Equation (55) yields

$$\begin{aligned} \begin{aligned}&\sup _{\beta ^S_i,\beta ^V_i} \bigg \{ -h'-\sum _{j=1}^{2}(H_j)'v_j+(1-\gamma )\big (r+\sum _{j=1}^{2}(\beta ^{S}_j\lambda _jv_j +\beta ^{v}_j\mu _jv_j\big )\big )\\&\quad -\frac{\gamma (1-\gamma )}{2}\sum _{j=1}^{2}[(\beta ^{S}_j)^2+(\beta ^{v}_j)^2]v_j\\&\quad +\sum _{j=1}^{2}\big (\kappa _j(\theta _j-v_j))H_j +\frac{1}{2}\sum _{j=1}^{2}\sigma _j^2v_j(H_j)^2+\sum _{j=1}^{2}\sigma _jv_jx\big (\beta ^S_j\rho _j+\beta ^V_j\sqrt{1-\rho _j^2}\big )H_j\\&\quad -\sum _{j=1}^{2}\frac{1}{2}(1-\gamma )\phi _j^Sv_j\left[ (\beta ^S_j)^2+2\beta ^S_j\rho _j\sigma _j\frac{H_j}{1-\gamma }+\rho _j^2\sigma _j^2\frac{(H_j)^2}{(1-\gamma )^2}\right] \\&\quad -\sum _{j=1}^{2}\frac{1}{2}(1-\gamma )\phi _j^Vv_j\left[ (\beta ^V_j)^2+2\beta ^V_j\sqrt{1-\rho _j^2}\sigma _j\frac{H_j}{1-\gamma }+(1-\rho _j^2)\sigma _j^2\frac{(H_j)^2}{(1-\gamma )^2}\right] \bigg \}=0, \end{aligned} \end{aligned}$$
(56)

where \(h'\) represents the time derivative of function h(t). From (56), we derive the optimal exposures as

$$\begin{aligned} (\beta ^S_j)^*= & {} \frac{\lambda _j}{\gamma +\phi ^S_j}+\frac{(1-\gamma -\phi ^S_j)\sigma _j\rho _j}{(1-\gamma )(\gamma +\phi ^S_j)}H_j(T-t),\nonumber \\ (\beta ^V_j)^*= & {} \frac{\mu _j}{\gamma +\phi ^V_j}+\frac{(1-\gamma -\phi ^V_j)\sigma _j\sqrt{1-\rho _j^2}}{(1-\gamma )(\gamma +\phi ^V_j)}H_j(T-t). \end{aligned}$$
(57)

To determine \(H_1\) and \(H_2\) and h, we substitute (57) into (56) to obtain

$$\begin{aligned} \begin{aligned}&-h'-\sum _{j=1}^{2}(H_j)'v_j+(1-\gamma )( r+\sum _{j=1}^{2}\big (\kappa _j(\theta _j-v_j))H_j) \\&+\frac{1}{2}\sum _{j=1}^{2}\sigma _j^2v_j(H_j)^2-\sum _{j=1}^{2}\frac{1}{2}\phi _j^Sv_j\rho _j^2\sigma _j^2\frac{(H_j)^2}{1-\gamma }\\&-\sum _{j=1}^{2}\frac{1}{2}\phi _j^Vv_j(1-\rho _j^2)\sigma _j^2\frac{(H_j)^2}{1-\gamma }+\sum _{j=1}^{2}\left[ \frac{(1-\gamma )\lambda _1^2}{2(\gamma +\phi _j^S)}+\frac{\lambda _1(1-\gamma -\phi _j^S)\sigma _j\rho _iH_j}{\gamma +\phi ^S_j} \right. \\&\left. +\frac{(1-\gamma -\phi _i^S)^2\sigma ^2_i\rho ^2_i(H_j)^2}{2(1-\gamma )(\gamma +\phi ^S_i)} \right] v_j\\&+\sum _{j=1}^{2}\left[ \frac{(1-\gamma )\lambda _2^2}{2(\gamma +\phi _1^V)}+\frac{\lambda _2(1-\gamma -\phi _j^V)\sigma _j\sqrt{1-\rho ^2_i}H_j}{\gamma +\phi ^S_j} \right. \\&\left. +\frac{(1-\gamma -\phi _j^V)^2\sigma ^2_j(1-\rho ^2_j)(H_j)^2}{2(1-\gamma )(\gamma +\phi ^V_j)} \right] v_j=0. \end{aligned} \end{aligned}$$
(58)

Comparing the terms concerned with \(v_j\), we get the following Riccati equations:

$$\begin{aligned} \left\{ \begin{aligned}&H_j'=a_jH_j+b_j(H_j)^2+c_j,\quad H_j(0)=0,\ j=1, 2, \\&h'=\kappa _1\theta _1H_1+\kappa _2\theta _2H_2+(1-\gamma )r,\quad h(0)=0. \end{aligned} \right. \end{aligned}$$
(59)

Let \(d_j=\sqrt{a_j^2-4b_jc_j}\)  , where

$$\begin{aligned} a_j= & {} -\kappa _j+\frac{\lambda _1(1-\gamma -\phi _j^S)\sigma _j\rho _j}{\gamma +\phi _j^S}+\frac{\lambda _2(1-\gamma -\phi _j^V)\sigma _j}{\gamma +\phi _j^V}\sqrt{1-\rho _j^2},\\ b_j= & {} \frac{\sigma _j^2}{2}-\frac{\phi ^S_j\rho _j^2\sigma _j^2}{2(1-\gamma )}-\frac{\phi ^V_j(1-\rho _j^2)\sigma _j^2}{2(1-\gamma )}+\frac{(1-\gamma -\phi _j^S)\sigma ^2_j\rho ^2_j}{2(1-\gamma )(\gamma +\phi _j^S)}\\&+\frac{(1-\gamma -\phi _j^V)^2\sigma ^2_j(1-\rho _j^2)}{2(1-\gamma )(\gamma +\phi _j^V)},\\ c_j= & {} \frac{(1-\gamma )\lambda _1^2}{2(\gamma +\phi ^S_j)}+\frac{(1-\gamma )\lambda _2^2}{2(\gamma +\phi ^V_j)}. \end{aligned}$$

Functions \(H_j\) and h can be obtained from (59), as shown in (17). \(\square \)

Appendix B Proof of Proposition 3.2

Proof

It is sufficient to show that the Novikov’s condition is satisfied for the worst-case probability measure, that is,

$$\begin{aligned} \mathrm {E}^{{\mathbb {P}}}\left[ \exp \left\{ \frac{1}{2}\int _{0}^{T}((e^S_1(t))^*)^2+((e^S_2(t))^*)^2+((e^V_1(t))^*)^2+((e^V_2(t))^*)^2\mathrm {d}t\right\} \right] <\infty . \end{aligned}$$

Applying (19) in Proposition 3.1, one obtains \(\mathrm {E}^{{\mathbb {P}}}\big [\exp \big \{\frac{1}{2}\int _{0}^{T}\big (K_1(T-t)V_1(t)+K_2(T-t)V_2(t)\big )\mathrm {d}t\big \}\big ]<\infty ,\) where

$$\begin{aligned} K_j(T-t)= & {} \left( \frac{\lambda _j}{\gamma +\phi ^S_j}+\frac{\sigma _j\rho _j H_j(T-t)}{(1-\gamma )(\gamma +\phi ^S_j)}\right) ^2(\phi ^S_j)^2\\&+\left( \frac{\mu _j}{\gamma +\phi ^V_j}+\frac{\sigma _j\sqrt{1-\rho _j^2} H_j(T-t)}{(1-\gamma )(\gamma +\phi ^V_j)}\right) ^2(\phi ^V_j)^2,\ j= 1, 2. \end{aligned}$$

Define \(k_1=\sup _{t\in [0,T]}K_1(T-t)\) and \(k_2=\sup _{t\in [0,T]}K_2(T-t)\), and then,

$$\begin{aligned} \begin{aligned}&\mathrm {E}^{{\mathbb {P}}}\left[ \exp \left\{ \frac{1}{2}\int _{0}^{T}\left( K_1(T-t)V_1(t)+K_2(T-t)V_2(t)\right) \mathrm {d}t\right\} \right] \\&\quad< \sum \limits _{i=1}^2\mathrm {E}^{{\mathbb {P}}}\left[ \exp \left\{ \frac{k_i}{2}\int _{0}^{T}V_i(t)\mathrm {d}t\right\} \right] <\infty . \end{aligned} \end{aligned}$$
(60)

The first inequity in (60) holds because \(V_1\) and \(V_2\) are independent, while the second holds if \(k_1\le \frac{\kappa _1^2}{\sigma _1^2}\) and \(k_2\le \frac{\kappa _2^2}{\sigma _2^2}\) (see [35]). Recall that \(\gamma >1\), \(\lambda _j > 0\), \(\rho _j<0\) and \(\mu _j<0\). Thus, \(K_j\) reaches its maximum value at \(t=T\) because \(H_j\) is maximum at \(t= T\). Therefore,

$$\begin{aligned} k_1= & {} (\phi _1^S)^2\frac{\lambda _1^2}{(\gamma +\phi ^S_1)^2}+(\phi _1^V)^2\frac{\mu ^2_1}{(\gamma +\phi ^V_1)^2}, \\ k_2= & {} (\phi _2^S)^2\frac{\lambda ^2_2}{(\gamma +\phi ^S_2)^2}+(\phi _2^V)^2\frac{\mu ^2_2}{(\gamma +\phi ^V_2)^2}, \end{aligned}$$

which yield the desired results immediately. \(\square \)

Appendix C Proof of Proposition 3.3

Proof

In the incomplete market, solving the optimal problem (20)–(21), we obtain the general optimal wealth invested in stock

$$\begin{aligned} \pi ^S=\frac{\sum _{j=1}^{2}\left[ (1-\gamma )v_j(\lambda _j+\rho _j\sigma _j{\bar{H}}_j)-\phi _j^S\rho _j\sigma _jv_j{\bar{H}}_j\right] }{\sum _{j=1}^{2}(1-\gamma )v_j(\gamma +\phi _j^S)}. \end{aligned}$$
(61)

Substituting (61) into (21), we obtain the following general HJB equation:

$$\begin{aligned}&{\bar{h}}'+\sum _{j=1}^{2}({\bar{H}}_j)'v_j-(1-\gamma )\big (r+\sum _{j=1}^{2}\pi ^S\lambda _jv_j \big )\nonumber \\&\quad +\frac{\gamma (1-\gamma )}{2}\sum _{j=1}^{2}(\pi ^S)^2v_j+\sum _{j=1}^{2}\big (\kappa _j(\theta _j-v_j)){\bar{H}}_j \nonumber \\&\quad -\frac{1}{2}\sum _{j=1}^{2}\sigma _j^2v_j({\bar{H}}_j)^2-\sum _{j=1}^{2}\sigma _jv_jx\rho _j{\bar{H}}_j\pi ^S\nonumber \\&\quad +\sum _{j=1}^{2}\frac{1}{2}(1-\gamma )\phi _j^Sv_j\left[ (\pi ^S)^2+2\pi ^S({\bar{H}}_j)\rho _j\sigma _j\frac{{\bar{H}}_j}{1-\gamma }\right. \nonumber \\&\quad \left. +\rho _j^2\sigma _j^2\frac{({\bar{H}}_j)^2}{(1-\gamma )^2}\right] \nonumber \\&\quad +\sum _{j=1}^{2}\frac{1}{2}(1-\gamma )\phi _j^Vv_j(1-\rho _j^2)\sigma _j^2\frac{({\bar{H}}_j)^2}{(1-\gamma )^2}=0. \end{aligned}$$
(62)

If there only exists one risk factor, say, \(W_1\), from (61), we have

$$\begin{aligned} \pi ^S=\frac{\lambda _1}{\gamma +\phi ^S_1}+\frac{(1-\gamma -\phi ^S_1)\sigma _1\rho _1}{(1-\gamma )(\gamma +\phi ^S_1)}{\bar{H}}_1(\tau ). \end{aligned}$$

Alternatively, if the only risk factor is \(W_2\), \(\pi ^S=\frac{\lambda _2}{\gamma +\phi ^S_2}+\frac{(1-\gamma -\phi ^S_2)\sigma _2\rho _2}{(1-\gamma )(\gamma +\phi ^S_2)}{\bar{H}}_2(\tau ).\)

Further, if the risk factors are the same, \(W_1=W_2\), then

$$\begin{aligned} \pi ^S=\frac{\lambda _1}{\gamma +\phi ^S_1}+\frac{(1-\gamma -\phi ^S_1)\sigma _1\rho _1}{(1-\gamma )(\gamma +\phi ^S_1)}{\bar{H}}_1(\tau )=\frac{\lambda _2}{\gamma +\phi ^S_2}+\frac{(1-\gamma -\phi ^S_2)\sigma _2\rho _2}{(1-\gamma )(\gamma +\phi ^S_2)}{\bar{H}}_2(\tau ). \end{aligned}$$

Therefore, under the circumstances that there are only one single risk or two identical risk factors we can write, without loss of generality,

$$\begin{aligned} \begin{aligned} \pi ^S=\pi _j({\bar{H}}_j) = \frac{\lambda _j}{\gamma +\phi ^S_j}+\frac{(1-\gamma -\phi ^S_j)\sigma _j\rho _j}{(1-\gamma )(\gamma +\phi ^S_j)}{\bar{H}}_j(\tau ),\ j = 1, 2. \end{aligned} \end{aligned}$$

Substituting \(\pi ^S=\pi _j({\bar{H}}_j)\) into (21), we obtain the following HJB equation:

$$\begin{aligned} \begin{aligned} 0=&{\bar{h}}'+\sum _{j=1}^{2}({\bar{H}}_j)'v_j-(1-\gamma )\big (r+\sum _{j=1}^{2}\pi _j({\bar{H}}_j)\lambda _jv_j \big )\\&+\frac{\gamma (1-\gamma )}{2}\sum _{j=1}^{2}\pi ^2_j({\bar{H}}_j)v_j+\sum _{j=1}^{2}\big (\kappa _j(\theta _j-v_j)){\bar{H}}_j\\&-\frac{1}{2}\sum _{j=1}^{2}\sigma _j^2v_j({\bar{H}}_j)^2-\sum _{j=1}^{2}\sigma _jv_jx\rho _j{\bar{H}}_j\pi _j({\bar{H}}_j)\\&+\sum _{j=1}^{2}\frac{1}{2}(1-\gamma )\phi _j^Vv_j(1-\rho _j^2)\sigma _j^2\frac{({\bar{H}}_j)^2}{(1-\gamma )^2}\\&+\sum _{j=1}^{2}\frac{1}{2}(1-\gamma )\phi _j^Sv_j\left[ (\pi _j({\bar{H}}_j))^2+2\pi _j({\bar{H}}_j)\rho _j\sigma _j\frac{{\bar{H}}_j}{1-\gamma }+\rho _j^2\sigma _j^2\frac{({\bar{H}}_j)^2}{(1-\gamma )^2}\right] . \end{aligned} \end{aligned}$$
(63)

By matching coefficients, we obtain the ordinary differential equations for the solution of \({\bar{H}}_j\) and \({\bar{h}}\) as follows:

$$\begin{aligned} \left\{ \begin{aligned} {\bar{H}}'_j&=(1-\gamma )\pi _j({\bar{H}}_j)\lambda _j -\frac{\gamma (1-\gamma )}{2}\pi ^2_j({\bar{H}}_j)-\kappa _j{\bar{H}}_j +\frac{1}{2}\sigma _j^2({\bar{H}}_j)^2+(1-\gamma )\sigma _j\rho _j{\bar{H}}_j\pi _j({\bar{H}}_j)\\&-\frac{1}{2}(1-\gamma )\phi _j^S\left[ (\pi _j({\bar{H}}_j))^2+2\pi _j({\bar{H}}_j)\rho _j\sigma _j\frac{{\bar{H}}_j}{1-\gamma }+\rho _j^2\sigma _j^2\frac{({\bar{H}}_j)^2}{(1-\gamma )^2}\right] \\&-\frac{1}{2}(1-\gamma )\phi _j^V(1-\rho _j^2)\sigma _j^2\frac{({\bar{H}}_j)^2}{(1-\gamma )^2},\quad&{\bar{H}}_j(0)=0,\\ {\bar{h}}'&=\kappa _1\theta _1{\bar{H}}_1+\kappa _2\theta _2{\bar{H}}_2+(1-\gamma )r,\quad&{\bar{h}}(0)=0. \end{aligned} \right. \end{aligned}$$
(64)

The worst-case measures can be computed directly by (24), and thus, the proof is complete.\(\square \)

Appendix D Proof of Proposition 3.4

Proof

Without loss of generality, we assume the indirect utility function of all suboptimal strategy \(\varPi \) as

$$\begin{aligned} J^{\varPi }=\frac{x^{1-\gamma }}{1-\gamma }\exp \left( H^{\varPi }_1(\tau )v_1+H^{\varPi }_2(\tau )v_2+h^{\varPi }(\tau )\right) . \end{aligned}$$
(65)

For mathematical tractability, we choose

$$\begin{aligned} \displaystyle \varPsi ^S_j=\frac{{\widetilde{\phi }}_j^S}{(1-\gamma )J(t,x,v_1,v_2)} \quad \text{ and } \quad \varPsi ^V_j=\frac{{\widetilde{\phi }}_j^V}{(1-\gamma )J(t,x,v_1,v_2)}, \end{aligned}$$

where the ambiguity aversion parameters \({\widetilde{\phi }}_j^S\), \({\widetilde{\phi }}_j^V>0\).

The functions \(H^{\varPi }_1\), \(H^{\varPi }_2\) and \(h^{\varPi }\) in (65) should satisfy the HJB PDE (26), which leads to

$$\begin{aligned} \begin{aligned} 0=&-(h^{\varPi })'-\sum _{j=1}^{2}(H^{\varPi }_j)'v_j+(1-\gamma )\big (r+\sum _{j=1}^{2}(\beta ^{S}_j\lambda _jv_j \\&+\beta ^{v}_j\mu _jv_j\big )\big ) -\frac{\gamma (1-\gamma )}{2}\sum _{j=1}^{2}[(\beta ^{S}_j)^2+(\beta ^{v}_j)^2]v_j+\sum _{j=1}^{2}\big (\kappa _j(\theta _j-v_j))H^{\varPi }_j\\&+\frac{1}{2}\sum _{j=1}^{2}\sigma _j^2v_j(H^{\varPi }_j)^2+\sum _{j=1}^{2}\sigma _jv_jx\big (\beta ^S_j\rho _j+\beta ^V_j\sqrt{1-\rho _j^2}\big )H^{\varPi }_j\\&-\sum _{j=1}^{2}\frac{1}{2}(1-\gamma ){\widetilde{\phi }}_j^Sv_j\left[ (\beta ^S_j)^2+2\beta ^S_j\rho _j\sigma _j\frac{H^{\varPi }_j}{1-\gamma }+\rho _j^2\sigma _j^2\frac{(H^{\varPi }_j)^2}{(1-\gamma )^2}\right] \\&-\sum _{j=1}^{2}\frac{1}{2}(1-\gamma ){\widetilde{\phi }}_j^Vv_j\left[ (\beta ^V_j)^2+2\beta ^V_j\sqrt{1-\rho _j^2}\sigma _j\frac{H^{\varPi }_j}{1-\gamma }+(1-\rho _j^2)\sigma _j^2\frac{(H^{\varPi }_j)^2}{(1-\gamma )^2}\right] . \end{aligned} \end{aligned}$$
(66)

Comparing the terms with and without multiplier \(v_j\), one can certainly obtain the following Riccati equations:

$$\begin{aligned} \left\{ \begin{aligned}&(H^{\varPi }_j)'=a_jH^{\varPi }_j+b_j(H^{\varPi }_j)^2+c_j,\quad&H^{\varPi }_j(0)=0,\ j=1, 2, \\&(h^{\varPi })'=\kappa _1\theta _1H^{\varPi }_1+\kappa _2\theta _2H^{\varPi }_2+(1-\gamma )r,\quad&h^{\varPi }(0)=0, \end{aligned} \right. \end{aligned}$$
(67)

where

$$\begin{aligned} \begin{aligned} a_j=&-\kappa _j+\sigma _j(1-\gamma )(\beta ^S_j\rho +\beta ^V_j\sqrt{1-\rho ^2})-{\widetilde{\phi }}^S_j\beta ^S_j\rho \sigma _j-{\widetilde{\phi }}^V_j\beta ^V_j\sqrt{1-\rho ^2}\sigma _j,\\ b_j=&\frac{\sigma _j^2}{2}-\frac{{\widetilde{\phi }}^S_j\rho _j^2\sigma _j^2}{2(1-\gamma )}-\frac{{\widetilde{\phi }}^V_j(1-\rho _j^2)\sigma _j^2}{2(1-\gamma )},\\ c_j=&(1-\gamma )(\beta ^S_j\lambda _1+\beta ^V_j\lambda _2)-\frac{1}{2}\gamma (1-\gamma )((\beta ^S_j)^2+(\beta ^V_j)^2)\\&- \frac{(1-\gamma )(\beta ^S_j)^2{\widetilde{\phi }}^S_j}{2}-\frac{(1-\gamma )(\beta ^V_j)^2{\widetilde{\phi }}^V_j}{2}. \end{aligned} \end{aligned}$$

\(\square \)

Appendix E Proof of Proposition 3.5

Proof

Recall that the utility loss \(L^\varPi =1-\exp \big \{\frac{1}{1-\gamma }\big [(H_1^\varPi -H_1)v_1+(H_2^\varPi -H_2)v_2+(h^\varPi -h)\big ]\big \}\). If the investor chooses strategy \(\varPi _1\), then \(H_2^{\varPi _1}=H_2\) since the strategy does not affect on the uncertainty about the second component of the ambiguity. Assume that \(\beta ^S_1=p^S_1+p^S_2{\widehat{H}}_1\) and \(\beta ^V_1=p^V_1+p^V_2{\widehat{H}}_1\), where \(p^S_j\) and \(p^V_j\)\((j=1,2)\) are constants. Let \({\widehat{H}}_1\) be the function satisfying the equation \(({\widehat{H}}_1)'={\widehat{a}}{\widehat{H}}_1+{\widehat{b}}\left( {\widehat{H}}_1\right) ^2+{\widehat{c}}\), where \({\widehat{a}},{\widehat{b}}\) and \({\widehat{c}}\) are all constants, and we obtain the following system of equations in terms of \(H^{\varPi }_1\) and \({\widehat{H}}_1\):

$$\begin{aligned} \left\{ \begin{aligned}&(H^{\varPi }_1)'=P_1+P_2{\widehat{H}}_1+P_3(H_1^{\varPi })^2+P_4H_1^{\varPi }+P_5{\widehat{H}}_1H_1^{\varPi }+P_6(H_1^{\varPi })^2, \quad&H_1^{\varPi }(0)=0,\\&({\widehat{H}}_1)'={\widehat{a}}{\widehat{H}}_1+{\widehat{b}}{\widehat{H}}_1+{\widehat{c}}, \quad&{\widehat{H}}_1(0)=0, \end{aligned}\right. \end{aligned}$$
(68)

where

$$\begin{aligned} \left\{ \begin{aligned} P_1&=(1-\gamma )(\lambda _1p_1^S+\lambda _2p^V_2-\frac{1}{2}(\gamma +{\widehat{\phi }}_1^S))(p_1^S)^2-\frac{1}{2}(\gamma +{\widehat{\phi }}_1^V))(p_1^V)^2,\\ P_2&=(1-\gamma )(\lambda _1p_1^S+\lambda _2p^V_2-p_1^Sp_2^S(\gamma +{\widehat{\phi }}_1^S)-p_1^Vp_2^V(\gamma +{\widehat{\phi }}_1^V)),\\ P_3&=-\frac{1}{2}(1-\gamma )((\gamma +{\widehat{\phi }}_1^S)(p_2^S)^2+(\gamma +{\widehat{\phi }}_1^V)(p_2^V)^2),\\ P_4&=-\kappa _1+\sigma _1(\rho _1(1-\gamma -{\widetilde{\phi }}_1^S)p_1^S+\sqrt{1-\rho ^2_1}(1-\gamma -{\widetilde{\phi }}_1^V)p_1^V),\\ P_5&=\sigma _1(\rho _1(1-\gamma -{\widetilde{\phi }}_1^S)p_2^S+\sqrt{1-\rho ^2_1}(1-\gamma -{\widetilde{\phi }}_1^V)p_2^V),\\ P_6&=\frac{1}{2(1-\gamma )}\sigma _1^2(1-\gamma -\rho _1^2{\widetilde{\phi }}_1^S-(1-\rho _1^2){\widetilde{\phi }}_1^V). \end{aligned} \right. \end{aligned}$$
(69)

It follows from (59) and (68) that

$$\begin{aligned} (H^{\varPi _1}_1-H_1)'+E(t)(H^{\varPi _1}_1-H_1)'=F(t), \end{aligned}$$
(70)

where \(E=-P_4-P_6(H^{\varPi _1}_1+H_1)-P_5{\widehat{H}}_1\) and \(F=P_1-c+(P_4-a)H_1+(K_6-b)H_1^2+P_2{\widehat{H}}_1+P_3{\widehat{H}}_1^2+P_5{\widehat{H}}_1H_1.\)

Solving (70), we have

$$\begin{aligned} H^{\varPi _1}_1(t)-H_1(t)=e^{-\int _{t}^{T}E(s)ds}\int _{t}^{T}e^{\int _{s}^{T}E(\tau )d\tau }F(s)ds. \end{aligned}$$

Under the assumption that the suboptimal strategies are admissible, the integral \(\displaystyle e^{-\int _{t}^{T}E(s)ds}\) is bounded and positive. Furthermore, the choice of strategy \(\varPi _1\) means that \({\widetilde{\phi }}_1^S = 0 \) and \({\widetilde{\phi }}_1^V = 0\). Thus,

$$\begin{aligned} \begin{aligned} F=&\frac{\gamma -1}{2\gamma ^2(\phi ^S_1+\gamma )} \left( \lambda _1\phi ^S_1+(\gamma +\phi ^S_1)\rho _1\sigma _1({\widehat{H}}_1-\frac{\gamma (\gamma +\phi _1^S-1)}{(\gamma -1)(\gamma +\phi _1^S)}H_1)\right) ^2\\&+\frac{\gamma -1}{2\gamma ^2(\phi ^V_1+\gamma )} \left( \lambda _1\phi ^V_1+(\gamma +\phi ^V_1)\sqrt{1-\rho _1^2}\sigma _1({\widehat{H}}_1-\frac{\gamma (\gamma +\phi _1^V-1)}{(\gamma -1)(\gamma +\phi _1^V)}H_1)\right) ^2>0. \end{aligned} \end{aligned}$$

This leads to \(H_1^\varPi -H_1>0\). Following (59), we obtain \(\displaystyle h^{\varPi _1}-h= \kappa _1\theta _1\int _0^t \left( H^{\varPi _1}_1(s)-H_1(s)\right) ds>0.\) Therefore, we have

$$\begin{aligned} L^{\varPi _1}=1-\exp \left\{ \frac{1}{1-\gamma }\left[ (H_1^\varPi -H_1)v_1+(h^\varPi -h)\right] \right\} >0. \end{aligned}$$
(71)

Due to symmetry, \(L^{\varPi _2}>0\) can be proved in a similar way.

If \(L^{\varPi _3}\) is chosen, from Propositions 3.1 and 3.3, we have

$$\begin{aligned} H^{\varPi _3}_j(t)-H_j(t)=e^{-\int _{t}^{T}E_j(s)ds}\int _{t}^{T}e^{\int _{s}^{T}E_j(\tau )d\tau }F_j(s)ds, \ j=1, 2, \end{aligned}$$

where \(F_j=\frac{\gamma -1}{2(\gamma +\phi ^V_j)}\left( \lambda _j+\frac{\sigma _jH^{\varPi _3}_j\sqrt{1-\rho _j^2}(\gamma +\phi ^V_j-1)}{\gamma -1}\right) .\) It is straightforward to see that \(F_j>0\), and thus, \(L^{\varPi _3}>0\). \(\square \)

Appendix F HJB in Sect. 4

In complete markets, the value function (47) satisfies the robust HJB PDE:

$$\begin{aligned} \begin{aligned}&\sup _{\beta ^S_i,\beta ^V_i}\inf _{e^S_i,e^V_i} \bigg \{ J_t+x\big (r+\sum _{j=1}^{2}(\beta ^{S}_j\lambda _jv_j-\beta ^{S}_ju_je^S_j \\&\quad +\beta ^{V}_j\mu _jv_j-\beta ^{V}_ju_je^V_j)\big )J_x +\sum _{j=1}^{2}\big (\kappa _j(\theta _j-v_j)-\rho _j\sigma _ju_je^S_j\\&\quad -\sqrt{1-\rho _j^2}\sigma _ju_je^V_j\big )J_{v_j}-\sum _{j=1}^{2}\left( \mu _j^U-\rho _j\psi ^U_je^S_j-\sqrt{1-\rho _j^2}\psi ^U_je^V_j\right) J_{u_j} \\&\quad +\big (\mu ^Y-\rho _2\psi _2^Ue^S_2u_1-\rho _1\psi _1^Ue^S_1u_2\\&\quad -\sqrt{1-\rho _2^2}\psi _2^Ue^V_2u_1-\sqrt{1-\rho _1^2}\psi _1^Ue^V_1u_2\big )J_y+\frac{1}{2}x^2\sum _{j=1}^{2}[(\beta ^{S}_j)^2\\&\quad +(\beta ^{V}_j)^2]v_jJ_{xx} +\rho \rho _1\psi ^U_1\beta ^S_2u_2xJ_{xu_1} \\&\quad +\sum _{j=1}^{2}\sigma _jv_jx\big (\beta ^S_j\rho _j+\beta ^V_j\sqrt{1-\rho _j^2}\big )J_{x v_j}+\rho \rho _2\psi ^U_2\beta ^S_1u_1xJ_{xu_2}\\&\quad +x\bigg ( \beta ^S_1u_1(\rho \rho _2u_1\psi ^U_2+\rho _1u_2\psi ^U_1)\\&\quad +\beta ^S_2u_2(\rho _2\psi _2^Uu_1+\rho \rho _1\psi ^U_1u_2)+\sqrt{1-\rho _1^2}\beta ^V_1\psi ^U_1v_1\\&\quad +\sqrt{1-\rho _2^2}\beta ^V_2\psi ^U_2v_2 \bigg )J_{xy} +\frac{1}{2}\sum _{j=1}^{2}{\varvec{\varSigma }}_{jj}J_{v_jv_j}+ {\varvec{\varSigma }}_{12}J_{v_1v_2}\\&\quad +\sum _{j=1}^{2}{\varvec{\varSigma }}_{j,j+2}J_{v_ju_j}+\varvec{\varSigma _{14}}J_{v_1u_2}+{\varvec{\varSigma }}_{23}J_{v_2u_1}\\&\quad +{\varvec{\varSigma }}_{15}J_{v_1y} +{\varvec{\varSigma }}_{25}J_{v_2y}+\frac{1}{2}\sum _{j=3}^{4}{\varvec{\varSigma }}_{jj}J_{u_ju_j} +{\varvec{\varSigma }}_{34}J_{u_1u_2}+{\varvec{\varSigma }}_{35}J_{u_1y}\\&\quad +{\varvec{\varSigma }}_{45}J_{u_2y}+{\varvec{\varSigma }}_{55}J_{yy}+\sum _{j=1}^{2}\frac{(e^S_j)^2}{2\varPsi ^S_j} +\frac{(e^V_j)^2}{2\varPsi ^V_j}\bigg \}=0, \end{aligned} \end{aligned}$$
(72)

where the symmetrical matrix \({\varvec{\varSigma }}(t)\) is given as follows:

$$\begin{aligned} \begin{bmatrix} \frac{1}{2}\sigma _1^2v_1 &{} \rho \rho _1\rho _2\sigma _1\sigma _2&{} \sigma _1u_1\psi ^U_1&{} \rho \rho _1\sqrt{1-\rho _2^2}\sigma _1\psi ^U_2u_1&{} \sigma _1\rho _1u_1(\rho _1u_2\psi ^U_1+\rho \rho _2u_1\psi ^U_2)+(1-\rho _1^2)\sigma _1\psi _1^U\\ &{} \frac{1}{2}\sigma _2^2v_2 &{} \rho \rho _1\rho _2\sigma _2\psi ^U_1u_2 &{} \sigma _2u_2\psi _2^U &{} \sigma _2\rho _2u_2(\rho _2u_1\psi ^U_2+\rho \rho _1u_2\psi ^U_1)+(1-\rho _2^2)\sigma _2\psi _2^U\\ &{} &{} \frac{1}{2}(\psi ^U_1)^2 &{} \rho \rho _1\rho _2\psi ^U_1\psi ^U_2 &{} (\psi ^U_1)^2u_2+\rho \rho _1\rho _2\psi _1^U\psi _2^Uu_1\\ &{} &{} &{}\frac{1}{2}(\psi ^U_2)^2&{} (\psi ^U_2)^2u_1+\rho \rho _1\rho _2\psi _1^U\psi _2^Uu_2\\ &{} &{} &{} &{} \frac{1}{2}\left( u_1^2(\psi ^U_2)^2+u_2^2(\psi ^U_1)^2+\rho \rho _1\rho _2\psi ^U_1\psi ^U_2\right) \end{bmatrix}. \end{aligned}$$

Appendix G Detection-Error Probabilities

Define the conditional characteristic functions

$$\begin{aligned} f_1(\omega ,t,T)=\mathrm {E}^{{\mathbb {P}}}[\exp (i\omega \xi _1(T))|{\mathcal {F}}_t^{S,V_1,V_2}]=\mathrm {E}^{{\mathbb {P}}}[({\mathcal {Z}}_1(T))^{i\omega }|{\mathcal {F}}_t^{S,V_1,V_2}] \end{aligned}$$

and

$$\begin{aligned} f_2(\omega ,t,T)= & {} \mathrm {E}^{{\mathbb {P}}^e}[\exp (i\omega \xi _1(T))|{\mathcal {F}}_t^{S,V_1,V_2}]=\mathrm {E}^{{\mathbb {P}}^e}[({\mathcal {Z}}_1(T))^{i\omega }|{\mathcal {F}}_t^{S,V_1,V_2}]\\= & {} \mathrm {E}^{{\mathbb {P}}}[({\mathcal {Z}}_1(T))^{i\omega +1}|{\mathcal {F}}_t^{S,V_1,V_2}], \end{aligned}$$

where \(i^2=-1\) and \(\omega \) is the transform variable. Denote \(\mathbf {e_t}=[e^S_1,e^S_2,e_1^V,e^V_2]\) and

$$\begin{aligned} {\varvec{\sigma }}=\left[ \sigma _1\sqrt{V_1}\rho _1,\sigma _2\sqrt{V_2}\rho _2,\sigma _1\sqrt{V_1}\sqrt{1-\rho _1^2},\sigma _2\sqrt{V_2}\sqrt{1-\rho _2^2}\right] . \end{aligned}$$

Since the conditional characteristic functions are martingales, the Feynman–Kac theorem implies that \(f_1\) and \(f_2\) satisfy the same PDE:

$$\begin{aligned}&\frac{\partial f}{\partial t}+\sum _{j=1}^{2}\kappa _j(\theta _j-V_j)\frac{\partial f}{\partial V_j}+\frac{1}{2}{\mathcal {Z}}_1(t)^2||\mathbf {e_t}||^2 \frac{\partial ^2 f}{\partial {\mathcal {Z}}_1^2}+\frac{1}{2}\sum _{j=1}^{2}\sigma ^2_jV_j \frac{\partial ^2 f}{\partial V_j^2}\nonumber \\&-\sum _{j=1}^{2}{\mathcal {Z}}_1(t){\varvec{\sigma }}{\mathbf {e_t}}{}^{T}\frac{\partial ^2 f}{\partial {{\mathcal {Z}}_1}\partial V_j}=0 \end{aligned}$$
(73)

with different terminal conditions \(f_1(\omega ,T,T)={\mathcal {Z}}_1^{i\omega }(T)\) and \(f_2(\omega ,T,T)={\mathcal {Z}}_1^{i\omega +1}(T)\), respectively. Conjecturing the solution in the form \(f={\mathcal {Z}}_1^{i\omega }(t)\exp (C_1(t)V_1+C_2(t)V_2+D(t))\) and substituting it into (73), we obtain

$$\begin{aligned}&C^{'}_{j}(t)-\kappa _jC_j(t)+\frac{1}{2}i\omega (i\omega -1)(q^S_j+q^V_j)^2\\&\quad +\frac{1}{2}\sigma _j^2C_j^2(t)-i\omega C_j(t)\sigma _i(q^S_j\rho _j+q^V_j\sqrt{1-\rho _j^2})=0,\ C_j(T)=0,\ j=1, 2, \\&D^{'}(t)+\kappa _1\theta _1C_1(t)+\kappa _2\theta _2C_2(t)=0,\ D(T)=0. \end{aligned}$$

Similarly, we can obtain the solution for \(f_2\) by replicating \(i\omega \) with \(i\omega +1\) in the above equations. Thus, we have the detection-error probability

$$\begin{aligned} \varepsilon _T(\phi ^S_1,\phi ^S_2,\phi ^V_1,\phi ^V_2)=\frac{1}{2}-\frac{1}{2\pi }\int _{0}^{\infty }\left( \mathrm {Re}\left[ \frac{f_1(\omega ,0,T)}{i\omega }\right] -\mathrm {Re}\left[ \frac{f_2(\omega ,0,T)}{i\omega }\right] \right) d\omega . \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, BZ., Lu, X., Ma, G. et al. Robust Portfolio Optimization with Multi-Factor Stochastic Volatility. J Optim Theory Appl 186, 264–298 (2020). https://doi.org/10.1007/s10957-020-01687-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10957-020-01687-w

Keywords

Mathematics Subject Classification

Navigation