Skip to main content
Log in

Portfolio Selection with Contrarian Strategy

  • Research
  • Published:
Methodology and Computing in Applied Probability Aims and scope Submit manuscript

Abstract

Compared with the extensive empirical literature on contrarian strategy, we develop a dynamic mean-variance model with geometric value-reversion asset prices, which implies a contrarian strategy. The model is solved (semi) explicitly under three asset price evaluations: constant valuation, exponential-varying valuation, and geometric average valuation. From a mathematical perspective, it is nontrivial to solve the extended HJB equations under stochastic opportunities. We demonstrate that our strategy exhibits the same monotonicity as that of the traditional constant relative risk-averse utility, and the welfare loss of using the dynamic mean-variance criterion is rather small, supporting that our model is a good approximation to the constant relative risk-averse utility. Empirical tests show that our strategies can help an investor achieve a less volatile wealth trajectory.

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
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Availability of Data and Material

All original datasets used come from public websites.

Notes

  1. For a function F of \(\left( t,P,R\right) \), the infinitesimal operator is \(\mathcal {D}F\left( t,P,R\right) =\frac{\partial F}{\partial t}+( \mu _{t}+\lambda ( \alpha _{t}-\ln P)) P\frac{\partial F}{\partial P}+\left[ r_{t}+\left( \mu _{t}-r_{t}+\lambda \left( \alpha _{t}-\ln P\right) \right) \pi _{t}-\frac{1}{2}\sigma _{t}^{2}\pi _{t}^{2}\right] \frac{\partial F}{\partial R}+\frac{1}{2}\left[ \sigma _{t}^{2}P^{2}\frac{\partial ^{2} F}{\partial P^{2}}+2\sigma _{t}^{2}\pi _{t}P\frac{\partial ^{2}F}{\partial P\partial R}\right. +\left. \sigma _{t}^{2}\pi _{t}^{2}\frac{\partial ^{2}F}{\partial R^{2} }\right] \). \(\mathcal {D}^{*}\) is the infinitesimal operator when replacing \(\pi \) by the equilibrium \(\hat{\pi }\) in \(\mathcal {D}\).

  2. By the structure of \(R_{T}\), it is easy to see that the expectation and variance, hence g and V, are separable with the starting return \(R_{t}\).

  3. For a function F of \(\left( t,P,\bar{P},R\right) \), the infinitesimal operator is \(\mathcal {D}F\left( t,P,\bar{P},R\right) =\frac{\partial F}{\partial t}+\left( \mu +\lambda \right. \left. (\ln \bar{P}-\ln P) \right) P\frac{\partial F}{\partial P}+\bar{P}\left[ \frac{\ln P-\ln \bar{P}}{t}\right] \frac{\partial F}{\partial \bar{P}}+\left[ r+\left( \mu -r+\lambda \left( \ln \bar{P}-\ln P\right) \right) \pi -\frac{1}{2}\sigma ^{2}\pi ^{2}\right] \frac{\partial F}{\partial R}+\frac{1}{2}\left[ \sigma ^{2}P^{2}\right. \left. \frac{\partial ^{2}F}{\partial P^{2}}+2\sigma ^{2}\pi P\frac{\partial ^{2}F}{\partial P\partial R}+\sigma ^{2}\pi ^{2}\frac{\partial ^{2}F}{\partial R^{2}}\right] \). \(\mathcal {D}^{*}\) is the infinitesimal operator when replacing \(\pi \) by the optimal \(\hat{\pi }\) in \(\mathcal {D}\).

References

  • Basak S, Chabakauri G (2010) Dynamic mean-variance asset allocation. Rev Financ Stud 23:2970–3016

    Article  Google Scholar 

  • Basu P, Vinod HD (1994) Mean reversion in stock prices: implications from a production based asset pricing model. Scand J Econ 96(1):51–65

    Article  Google Scholar 

  • Benth FE, Karlsen KH (2005) A note on Merton’s portfolio selection problem for the Schwartz mean-reversion model. Stoch Anal Appl 23(4):687–704

    Article  MathSciNet  Google Scholar 

  • Björk T, Murgoci A, Zhou XY (2014) Mean-variance portfolio optimization with state-dependent risk aversion. Math Financ 24(1):1–24

    Article  MathSciNet  Google Scholar 

  • Björk T, Khapko M, Murgoci A (2017) On time-inconsistent stochastic control in continuous time. Finance Stochast 21(2):331–360

    Article  MathSciNet  Google Scholar 

  • Chan KC (1988) On the contrarian investment strategy. J Bus 61(2):147–163

    Article  Google Scholar 

  • Chang RP, McLeavey DW, Rhee SG (1995) Short-term abnormal returns of the contrarian strategy in the Japanese stock market. J Bus Financ Acc 22(7):1035–1048

    Article  Google Scholar 

  • Chen Q, Hua X, Jiang Y (2018) Contrarian strategy and herding behaviour in the Chinese stock market. Eur J Financ 24(16):1552–1568

    Article  Google Scholar 

  • Chen J, Davison M, Escobar M, Zafari G (2021) Robust portfolios with commodities and stochastic interest rates. Quant Finance 21(6):991–1010

    Article  MathSciNet  Google Scholar 

  • Cheng S, Hameed A, Subrahmanyam A, Titman S (2017) Short-term reversals: the effects of past returns and institutional exits. J Financ Quant Anal 52(1):143–173

    Article  Google Scholar 

  • Chiu M, Wong H (2014) Mean-variance asset-liability management with asset correlation risk and insurance liabilities. Insur Math Econ 59:300–310

    Article  MathSciNet  Google Scholar 

  • Dai M, Jin H, Kou S, Xu YH (2021) A dynamic mean-variance analysis for log returns. Manage Sci 67(2):1093–1108

    Article  Google Scholar 

  • De Bondt WF, Thaler R (1985) Does the stock market overreact? J Financ 40(3):793–805

    Article  Google Scholar 

  • De Bondt WF, Thaler R (1987) Further evidence on investor overreaction and stock market seasonality. J Financ 42(3):557–581

    Article  Google Scholar 

  • Jegadeesh N (1990) Evidence of predictable behavior of security returns. J Financ 45(3):881–898

    Article  Google Scholar 

  • Kim H (2009) On the usefulness of the contrarian strategy across national stock markets: a grid bootstrap analysis. J Empir Financ 16(5):734–744

    Article  Google Scholar 

  • Lehmann BN (1990) Fads, martingales, and market efficiency. Q J Econ 105(1):1–28

    Article  MathSciNet  Google Scholar 

  • Li D, Ng WL (2000) Optimal dynamic portfolio selection: multiperiod mean-variance formulation. Math Financ 10:387–406

    Article  MathSciNet  Google Scholar 

  • Markowitz H (1952) Portfolio selection. J Financ 7(1):77–91

    Google Scholar 

  • Otchere I, Chan J (2003) Short-term overreaction in the Hong Kong stock market: can a contrarian trading strategy beat the market? J Behav Financ 4(3):157–171

    Article  Google Scholar 

  • Poterba JM, Summers LH (1988) Mean reversion in stock prices: evidence and implications. J Financ Econ 22(1):27–59

    Article  Google Scholar 

  • Schwartz ES (1997) The stochastic behavior of commodity prices: implications for valuation and hedging. J Financ 52(3):923–973

    Article  Google Scholar 

  • Zaitsev VF, Polyanin AD (2002) Handbook of exact solutions for ordinary differential equations chapter 2, table 14 and table 15. CRC Press

    Book  Google Scholar 

  • Zhou XY, Li D (2000) Continuous-time mean-variance portfolio selection: a stochastic LQ framework. Appl Math Optim 42:19–53

    Article  MathSciNet  Google Scholar 

Download references

Funding

Xu is supported by the Natural Science Foundation of China (No.12271391; No.11871050) and Tang Scholar Fund.

Author information

Authors and Affiliations

Authors

Contributions

The authors are listed in alphabetical order. L.Z. performed the numerical solutions for Sections 4.2 and 4.3, P.P. derived all the mathematical results, X.Y. contributed to writing, reviewing, and providing guidance throughout the paper, and Z.W. prepared the figures for Sections 4.1 and 4.4. All authors reviewed the manuscript.

Corresponding author

Correspondence to Yuhong Xu.

Ethics declarations

Competing Interests

The authors declare no competing interests.

Additional information

Publisher's Note

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

Appendices

Appendix 1 Proof of Theorem 2.1

(i) We assume that V and g are seperableFootnote 2 in R, i.e., they have the following structures:

$$\begin{aligned} V(t,P,R)&=\tilde{V}(t,P)+R+r\left( T-t\right) ;\end{aligned}$$
(A.1)
$$\begin{aligned} g\left( t,P,R\right)&=f\left( t,P\right) +R+r\left( T-t\right) . \end{aligned}$$
(A.2)

Substituting (A.1) and (A.2) into PDE (2.6) and PDE (2.7), we obtain:

$$\begin{aligned} \underset{\pi }{\max }\left\{ \mathcal {D}\tilde{V}\left( t,P\right) +\left( \mu -r+\lambda \left( \alpha -\ln P\right) \right) \pi -\frac{1}{2}\sigma ^{2}\pi ^{2}\right.&\\ -\left. \frac{\gamma }{2}\left[ \sigma ^{2}P^{2}(\frac{\partial f}{\partial P})^{2}+2\sigma ^{2}\pi P\frac{\partial f}{\partial P}+\sigma ^{2}\pi ^{2}\right] \right\}&=0,\\ \mathcal {D}^{*}f\left( t,P\right) +\left( \mu -r+\lambda \left( \alpha -\ln P\right) \right) \pi -\frac{1}{2}\sigma ^{2}\pi ^{2}&=0, \end{aligned}$$

subject to \(\tilde{V}(T,P)=0\), \(f(T,P)=0\). By the first-order condition, we derive a representation of the equilibrium \(\hat{\pi }\) as (2.8). It follows putting \(\hat{\pi }\) into f that, the function f satisfies PDE (2.9), subject to terminal condition \(f\left( T,P\right) =0\). Then, following the line of Björk et al. (2017), it is easy to verify that the strategy (2.10) is really an equilibrium of the original problem.

(ii) First we make a transform on the variable \(x=\ln P\). Then PDE (2.9) becomes

$$\begin{aligned} \left\{ \begin{array}{r} \frac{\partial f}{\partial t} +\left( \mu +\lambda \left( \alpha -x\right) -\frac{\gamma ^{2}(\mu -r+\lambda \left( \alpha -x\right) )}{\left( 1+\gamma \right) ^{2}}\right) \frac{\partial f}{\partial x}\\ +\frac{1}{2} \sigma ^{2}\left( -\frac{\partial f}{\partial x}+\frac{\partial ^{2}f}{\partial x^{2}}-\frac{\gamma ^{2}}{\left( 1+\gamma \right) ^{2}}(\frac{\partial f}{\partial x})^{2}\right) \\ +\frac{\left( 2\gamma +1\right) (\mu -r+\lambda \left( \alpha -x\right) )^{2} }{2\left( \gamma +1\right) ^{2}\sigma ^{2}}=0,\\ f(T,x)=0. \end{array} \right. \end{aligned}$$
(A.3)

Let \(f\left( t,x\right) =-\frac{\left( 1+\gamma \right) ^{2}}{\gamma ^{2} }\ln u\left( t,x\right) \). Then PDE (A.3) is rewritten as

$$\begin{aligned} \left\{ \begin{array}{r} \frac{\partial u}{\partial t}+\left( \mu +\lambda \left( \alpha -x\right) -\frac{\gamma ^{2}(\mu -r+\lambda \left( \alpha -x\right) )}{\left( 1+\gamma \right) ^{2}}-\frac{1}{2}\sigma ^{2}\right) \frac{\partial u}{\partial x}\\ +\frac{1}{2}\sigma ^{2}\frac{\partial ^{2}u}{\partial x^{2}} -\frac{\left( 2\gamma +1\right) \gamma ^{2}(\mu -r+\lambda \left( \alpha -x\right) )^{2}}{2\left( \gamma +1\right) ^{4}\sigma ^{2}}u=0,\\ u\left( T,x\right) =1. \end{array} \right. \end{aligned}$$
(A.4)

Let \(\omega =\alpha +\frac{\mu -r}{\lambda }-x\), \(\tilde{u}(t,\omega )=u(t,x).\) Putting \(\frac{\partial u}{\partial x}=-\frac{\partial \tilde{u}}{\partial \omega }\), \(\frac{\partial ^{2}u}{\partial x^{2}}=\frac{\partial ^{2}\tilde{u}}{\partial \omega ^{2}}\) into PDE (A.4), we get

$$\begin{aligned} \left\{ \begin{array}{c} \frac{\partial \tilde{u}}{\partial t}-\left( r-\frac{1}{2}\sigma ^{2} +\frac{(1+2\gamma )\lambda }{\left( 1+\gamma \right) ^{2}}\omega \right) \frac{\partial \tilde{u}}{\partial \omega }+\frac{1}{2}\sigma ^{2}\frac{\partial ^{2}\tilde{u}}{\partial \omega ^{2}}-\frac{\left( 2\gamma +1\right) \lambda ^{2}\gamma ^{2}}{2\left( \gamma +1\right) ^{4}\sigma ^{2}}\omega ^{2}\tilde{u}=0,\\ \tilde{u}\left( T,\omega \right) =1. \end{array} \right. \end{aligned}$$

Set \(\tilde{u}=\exp \left\{ k(T-t)+\eta \omega +\theta \omega ^{2}\right\} V(t,\omega )\). Substituting

$$\begin{aligned} \tilde{u}_{t} =&\ (-kV+V_{t})\exp \left\{ k(T-t)+\eta \omega +\theta \omega ^{2}\right\} ,\\ \tilde{u}_{\omega } =&\ \left[ (\eta +2\theta \omega )V+V_{\omega }\right] \exp \left\{ k(T-t)+\eta \omega +\theta \omega ^{2}\right\} ,\\ \tilde{u}_{\omega \omega } =&\ [ (\eta ^{2}+2\theta +4\theta ^{2}\omega ^{2}+4\theta \eta \omega )V\\ {}&+2(\eta +2\theta \omega )V_{\omega }+V_{\omega \omega }] \exp \left\{ k(T-t)+\eta \omega +\theta \omega ^{2}\right\} , \end{aligned}$$

into the above PDE, we obtain

$$\begin{aligned} \begin{aligned}&\left( -k+(\frac{\sigma ^{2}}{2}-r)\eta +\frac{1}{2}\sigma ^{2}\eta ^{2}+\sigma ^{2}\theta \right) V+\left( 2(\frac{\sigma ^{2}}{2}-r)\theta -\frac{(1+2\gamma )\lambda }{(1+\gamma )^{2}}\eta +2\sigma ^{2}\eta \theta \right) \omega V\\&+\left( 2\sigma ^{2}\theta ^{2}-2\frac{\lambda (1+2\gamma )}{(1+\gamma )^{2} }\theta -\frac{\lambda ^{2}\gamma ^{2}(1+2\gamma )}{2(1+\gamma )^{4}\sigma ^{2} }\right) \omega ^{2}V+V_{t}\\&+\left( (-r+\frac{1}{2}\sigma ^{2}+\sigma ^{2}\eta )+(2\sigma ^{2}\theta -\frac{(1+2\gamma )\lambda }{(1+\gamma )^{2}})\omega \right) V_{\omega }+\frac{1}{2}\sigma ^{2}V_{\omega \omega }=0. \end{aligned} \end{aligned}$$
(A.5)

Let the coefficient of V be 0 by separating the variable \(\omega \), we get the following algebraic equations:

$$\begin{aligned} \left\{ \begin{array}{l} -k+a\eta +\frac{1}{2}\sigma ^{2}\eta ^{2}+\sigma ^{2}\theta =0,\\ 2a\theta -b\eta +2\sigma ^{2}\eta \theta =0,\\ 2\sigma ^{2}\theta ^{2}-2b\theta -c=0, \end{array} \right. \end{aligned}$$
(A.6)

where

$$\begin{aligned} a=\frac{\sigma ^{2}}{2}-r\text {, }b=\frac{\lambda (1+2\gamma )}{(1+\gamma )^{2} }\text {, }c=\frac{\lambda ^{2}\gamma ^{2}(1+2\gamma )}{2(1+\gamma )^{4}\sigma ^{2} }\text {.} \end{aligned}$$
(A.7)

From the last and the second equations, we deduce that

$$\begin{aligned} \left\{ \begin{array}{l} \theta =\frac{b\pm q}{2\sigma ^{2}},\\ \eta =\mp \frac{a(b\pm q)}{\sigma ^{2}q}, \end{array} \right. \end{aligned}$$
(A.8)

where \(q=\sqrt{b^{2}+2\sigma ^{2}c}=\frac{\lambda \sqrt{1+2\gamma }}{(1+\gamma )}.\) Thus (A.5) satisfies

$$\begin{aligned} \left\{ \begin{array}{c} \frac{\partial V}{\partial t}+\left( (-r+\frac{1}{2}\sigma ^{2}+\sigma ^{2} \eta )+(2\sigma ^{2}\theta -\frac{(1+2\gamma )\lambda }{(1+\gamma )^{2}} )\omega \right) \frac{\partial V}{\partial \omega }+\frac{1}{2}\sigma ^{2} \frac{\partial ^{2}V}{\partial \omega ^{2}}=0,\\ V\left( T,\omega \right) =e^{-\eta \omega -\theta \omega ^{2}}. \end{array} \right. \end{aligned}$$
(A.9)

Then we denote \(\hat{a}:=a+\sigma ^{2}\eta =\mp \frac{ab}{q}\), \(\hat{b}:=2\sigma ^{2}\theta -b=\pm q\). So (A.9) is rewritten as

$$\begin{aligned} \left\{ \begin{array}{l} \frac{\partial V}{\partial t}+\left( \hat{a}+\hat{b}\omega \right) \frac{\partial V}{\partial \omega }+\frac{1}{2}\sigma ^{2}\frac{\partial ^{2} V}{\partial \omega ^{2}}=0,\\ V\left( T,\omega \right) =e^{-\eta \omega -\theta \omega ^{2}}. \end{array} \right. \end{aligned}$$
(A.10)

From the terminal structure of \(V\left( T,\omega \right) ,\) it is natural to suppose \(V\left( t,\omega \right) =\exp \{A\left( t\right) +B\left( t\right) \omega +C\left( t\right) \omega ^{2}\}\). Substituting \(V_{t}=\left( A_{t}+B_{t}\omega +C_{t}\omega ^{2}\right) \cdot V\), \(V_{\omega }=\left( B+2C\omega \right) \cdot V\), \(V_{\omega \omega }=\left( B^{2}+2C+4BC\omega +4C^{2}\omega ^{2}\right) \cdot V\) into (A.10) and by separating variables, we obtain the following system of ordinary differential equations (ODEs):

$$\begin{aligned} \left\{ \begin{array}{ll} A_{t}+\hat{a}B+\frac{1}{2}\sigma ^{2}\left( B^{2}+2C\right) =0\text {,} &{} A\left( T\right) =0,\\ B_{t}+\hat{b}B+2\hat{a}C+2\sigma ^{2}BC=0\text {,} &{} B\left( T\right) =-\eta ,\\ C_{t}+2\hat{b}C+2\sigma ^{2}C^{2}=0\text {,} &{} C\left( T\right) =-\theta . \end{array} \right. \end{aligned}$$
(A.11)

Solving the last equation, we have

$$\begin{aligned} C\left( t\right) =-\frac{\hat{b}\theta e^{2\hat{b}\left( T-t\right) } }{\theta \sigma ^{2}\left( e^{2\hat{b}\left( T-t\right) }-1\right) +\hat{b} }. \end{aligned}$$

Then

$$\begin{aligned} \bar{\Theta }&:=\theta +C\\&=\frac{\left( e^{2\hat{b}\left( T-t\right) }-1\right) \theta \left( \theta \sigma ^{2}-\hat{b}\right) }{\frac{1}{2}\left( b\pm q\right) \left( e^{2\hat{b}\left( T-t\right) }-1\right) \pm q}=\frac{-c\left( e^{\pm 2q\left( T-t\right) }-1\right) }{\left( b\pm q\right) \left( e^{\pm 2q\left( T-t\right) }-1\right) \pm 2q}\\&=\frac{-c\left( e^{2q\left( T-t\right) }-1\right) }{\left( b+q\right) \left( e^{2q\left( T-t\right) }-1\right) +2q}\le 0. \end{aligned}$$

Denote \(\bar{B}=\eta +B\). Then it follows (A.11) that,

$$\begin{aligned} \begin{array}{ll} \bar{B}_{t}+\left( 2\sigma ^{2}\bar{\Theta }-b\right) \bar{B}+2a\bar{\Theta }=0\text {,}&\bar{B}\left( T\right) =0. \end{array} \end{aligned}$$
(A.12)

Note that the coefficient of \(\bar{B}\) and the nonhomogeneous term are functions of time t, it is not easy to get the solution of ODE (A.12). Since (A.12) is a linear equation, we can get the solution first in integral form, then by a quite daunting and tedious calculation, the explicit solution is obtained as follows:

$$\begin{aligned} \bar{B}\left( t\right) =\frac{-2ac}{q}\left( e^{q\left( T-t\right) }-1\right) ^{2}\left[ \left( b+q\right) \left( e^{2q\left( T-t\right) }-1\right) +2q\right] ^{-1}. \end{aligned}$$

From the deformations of the above series of formulas, it yields

$$\begin{aligned} \frac{\partial f}{\partial P}=\frac{(1+\gamma )^{2}}{\gamma ^{2}P}[\bar{B}(t)+2\bar{\Theta }(t)(\alpha +\frac{\mu -r}{\lambda }-\ln P)]. \end{aligned}$$
(A.13)

Then the equilibrium strategy (2.8) can be expressed explicitly by

$$\begin{aligned} \hat{\pi }(t,P)=\frac{\mu -r+\lambda \left( \alpha -\ln P\right) }{(1+\gamma )\sigma ^{2}}-\frac{2(1+\gamma )(\alpha +\frac{\mu -r}{\lambda }-\ln P)}{\gamma }\bar{\Theta }(t)-\frac{1+\gamma }{\gamma }\bar{B}\left( t\right) . \end{aligned}$$

We then rewrite the above \(\hat{\pi }(t,P)\) as (2.10).

The proof is complete. \(\square \)

Appendix 2 Proof of Corollary 2.1

(i) From the expression of \(\hat{\pi }\), it’s not difficult to find that \(\hat{\pi }\) decreases as \(\ln P\) increases.

(ii) When \(\ln P<\alpha +\frac{\mu -r}{\lambda }\) and \(r<\frac{\sigma ^{2}}{2}\), we find that \(\frac{\lambda \gamma (1+2\gamma )\left( \mu -r+\lambda (\alpha -\ln P)\right) }{(1+\gamma )^{3}\sigma ^{2}}>0\) and \(\frac{\lambda ^{2} \gamma (1+2\gamma )(\sigma ^{2}-2r)}{2(1+\gamma )^{3}\sigma ^{2}}>0\). On the contrary, when \(\ln P>\alpha +\frac{\mu -r}{\lambda }\) and \(r>\frac{\sigma ^{2} }{2}\), it is easy to get that \(\frac{\lambda \gamma (1+2\gamma )\left( \mu -r+\lambda (\alpha -\ln P)\right) }{(1+\gamma )^{3}\sigma ^{2}}<0\) and \(\frac{\lambda ^{2}\gamma (1+2\gamma )(\sigma ^{2}-2r)}{2(1+\gamma )^{3}\sigma ^{2} }<0\). At the same time, \(\hat{\Theta }(\tau )\) is increasing in the time horizon \(\tau .\) Moreover,

$$\begin{aligned} \hat{B}\left( \tau \right)&=\frac{\left( e^{q\tau }-1\right) ^{2} }{q\left[ (b+q)(e^{2q\tau }-1)+2q\right] }\\&=\frac{1}{q\left[ (b+q)(1+\frac{2}{e^{q\tau }-1})+\frac{2q}{(e^{q\tau }-1)^{2}}\right] } \end{aligned}$$

is also increasing in the time horizon \(\tau \). Consequently, we obtain the desired results. \(\square \)

Appendix 3 Proof of Corollary 2.2

(i) In order to calculate the limit of \(\hat{\pi }(t,P)\) when \(\tau \) tends to infinity, we first calculate the limits of \(\hat{\Theta } (\tau )\) and \(\hat{B}\left( \tau \right) \).

Note that

$$\begin{aligned} \underset{\tau \rightarrow \infty }{\lim }\hat{\Theta }(\tau )&=\underset{\tau \rightarrow \infty }{\lim }\frac{e^{2q\tau }-1}{(b+q)(e^{2q\tau }-1)+2q} =\frac{1}{b+q},\\ \underset{\tau \rightarrow \infty }{\lim }\hat{B}\left( \tau \right)&=\underset{\tau \rightarrow \infty }{\lim }\frac{\left( e^{q\tau }-1\right) ^{2} }{q\left[ (b+q)(e^{2q\tau }-1)+2q\right] }\\ {}&=\underset{\tau \rightarrow \infty }{\lim }\frac{1}{q[(b+q)(1+\frac{2}{e^{q\tau }-1})+\frac{2q}{(e^{q\tau }-1)^{2} }]}=\frac{1}{q(b+q)}. \end{aligned}$$

Therefore,

$$\begin{aligned} \underset{\tau \rightarrow \infty }{\lim }\hat{\pi }(t,P) =&\ \frac{\mu -r+\lambda \left( \alpha -\ln P\right) }{(1+\gamma )\sigma ^{2}}+\frac{\lambda \gamma (1+2\gamma )\left( \mu -r+\lambda (\alpha -\ln P)\right) }{(1+\gamma )^{3}\sigma ^{2}}\frac{1}{b+q}\\&+\frac{\lambda ^{2}\gamma (1+2\gamma )(\sigma ^{2}-2r)}{2(1+\gamma )^{3}\sigma ^{2}}\frac{1}{q(b+q)}\\ =&\ \frac{\mu -r+\lambda \left( \alpha -\ln P\right) }{(1+\gamma )\sigma ^{2} }\\ {}&+\frac{\lambda ^{2}\gamma (1+2\gamma )}{[(1+\gamma )^{3}\sigma ^{2}](b+q)}\left( \left( \alpha +\frac{\mu -r}{\lambda }-\ln P\right) +\frac{\sigma ^{2}-2r}{2q}\right) . \end{aligned}$$

(ii) Denote

$$\begin{aligned} I=&\ \frac{\lambda \gamma (1+2\gamma )}{(1+\gamma )^{2}}\frac{e^{2q\tau } -1}{(b+q)(e^{2q\tau }-1)+2q}\\ {}&+\frac{\lambda \gamma (1+2\gamma )(\sigma ^{2} -2r)}{2(1+\gamma )^{2}\left( \alpha +\frac{\mu -r}{\lambda }-\ln P\right) } \frac{\left( e^{q\tau }-1\right) ^{2}}{q\left[ (b+q)(e^{2q\tau }-1)+2q\right] }. \end{aligned}$$

Hence,

$$\begin{aligned} \hat{\pi }(\tau ,P)=\frac{\mu -r+\lambda \left( \alpha -\ln P\right) }{(1+\gamma )\sigma ^{2}}\left[ 1+I\right] , \end{aligned}$$

and

$$\begin{aligned} \hat{R}^{\infty }(\tau ):=\underset{\gamma \rightarrow \infty }{\lim }\frac{\hat{H}(\tau )}{\hat{\pi }(\tau )}=\underset{\gamma \rightarrow \infty }{\lim }\frac{I}{I+1}. \end{aligned}$$

To compute the hedging ratio \(\frac{I}{I+1}\), we first calculate the limit of I

$$\begin{aligned} \underset{\gamma \rightarrow \infty }{\lim }I&=\underset{\gamma \rightarrow \infty }{\lim }(\frac{\lambda \gamma (1+2\gamma )}{(1+\gamma )^{2}}\frac{e^{2q\tau }-1}{(b+q)(e^{2q\tau }-1)+2q}\\&\ \ \ +\frac{\lambda \gamma (1+2\gamma )(\sigma ^{2}-2r)}{2(1+\gamma )^{2}\left( \alpha +\frac{\mu -r}{\lambda }-\ln P\right) }\frac{\left( e^{q\tau }-1\right) ^{2}}{q\left[ (b+q)(e^{2q\tau }-1)+2q\right] })\\&=\underset{\gamma \rightarrow \infty }{\lim }\left( \frac{2\lambda \tau }{(b+q)\tau +1}+\frac{\lambda (\sigma ^{2}-2r)\tau ^{2}}{2\left[ (b+q)\tau +1\right] \left( \alpha +\frac{\mu -r}{\lambda }-\ln P\right) }\right) \\&=2\lambda \tau +\frac{\lambda ^{2}(\sigma ^{2}-2r)\tau ^{2}}{2(\mu -r+\lambda \left( \alpha -\ln P\right) )}\\&=\frac{\lambda \tau \left[ \lambda \tau (\sigma ^{2}-2r)+4(\mu -r+\lambda \left( \alpha -\ln P\right) )\right] }{2(\mu -r+\lambda \left( \alpha -\ln P\right) )}. \end{aligned}$$

Thus,

$$\begin{aligned} \underset{\gamma \rightarrow \infty }{\lim }\frac{I}{I+1}=\frac{\lambda \tau \left[ \lambda \tau (\sigma ^{2}-2r)+4(\mu -r+\lambda \left( \alpha -\ln P\right) )\right] }{\lambda \tau \left[ \lambda \tau (\sigma ^{2}-2r)+4(\mu -r+\lambda \left( \alpha -\ln P\right) )\right] +2(\mu -r+\lambda \left( \alpha -\ln P\right) )}. \end{aligned}$$

The proof is complete. \(\square \)

Appendix 4 Proof of Proposition 4.1

Benth and Karlsen (2005) provide a semi-explicit solution for the CRRA optimization with geometric mean reversion and a risk aversion coefficient \(\tilde{\gamma }\in \left( 0,1\right) \). In our paper, \(\tilde{\gamma }=\gamma +1\) should be greater than 1 because \(\gamma \ge 0\) in the MV model. Here, we outline the proof and provide an explicit solution. Utilizing the classical dynamic programming principle, we derive the HJB equation for the value function:

$$\begin{aligned} \begin{aligned} \frac{\partial v}{\partial t}&+\underset{\pi }{\max }\left\{ \left( \left( \mu -r+\lambda \left( \alpha -\ln P\right) \right) \pi +r\right) W\frac{\partial v}{\partial W}\right. \\ {}&+\left. \frac{1}{2}\sigma ^{2}\pi ^{2}W^{2}\frac{\partial ^{2}v}{\partial W^{2}}+\sigma ^{2}\pi WP\frac{\partial ^{2}v}{\partial WP}\right\} \\&+\left( \mu +\lambda \left( \alpha -\ln P\right) \right) P\frac{\partial v}{\partial P}+\frac{1}{2}\sigma ^{2}P^{2}\frac{\partial ^{2}v}{\partial P^{2} }=0, \end{aligned} \end{aligned}$$
(D.1)

subject to \(v(T,W,P)=\frac{W^{1-\tilde{\gamma }}-1}{1-\tilde{\gamma }}.\)

Assuming there exists a sufficiently smooth solution v to the HJB equation, the first-order condition for an optimal strategy gives

$$\begin{aligned} \pi ^{*}=-\frac{\left( \mu -r+\lambda \left( \alpha -\ln P\right) \right) \frac{\partial v}{\partial W}+\sigma ^{2}P\frac{\partial ^{2}v}{\partial W\partial P}}{\sigma ^{2}W\frac{\partial ^{2}v}{\partial W^{2}}}. \end{aligned}$$

Then substituting \(\pi ^{*}\) into (D.1), we get the following PDE:

$$\begin{aligned} \begin{aligned} \frac{\partial v}{\partial t}&+rw\frac{\partial v}{\partial W}+\left( \lambda \left( \alpha -\ln P\right) +\mu \right) P\frac{\partial v}{\partial P}\\ {}&+\frac{1}{2}\sigma ^{2}P^{2}\frac{\partial ^{2}v}{\partial P^{2}}-\frac{1}{2}\sigma ^{2}P^{2}\left( \frac{\partial ^{2}v}{\partial W\partial P}\right) ^{2}/\frac{\partial ^{2}v}{\partial W^{2}}\\&-\frac{\left( \lambda \left( \alpha -\ln P\right) +\mu -r\right) ^{2} }{2\sigma ^{2}}\left( \frac{\partial v}{\partial W}\right) ^{2} /\frac{\partial ^{2}v}{\partial W^{2}}\\ {}&-\left( \lambda \left( \alpha -\ln P\right) +\mu -r\right) P\frac{\partial v}{\partial W}\frac{\partial ^{2} v}{\partial W\partial P}/\frac{\partial ^{2}v}{\partial W^{2}}=0. \end{aligned} \end{aligned}$$
(D.2)

As in Benth and Karlsen (2005), we make the assumption that the solution of (D.2) takes the form

$$\begin{aligned} v(t,W,P)=\frac{W^{1-\tilde{\gamma }}}{1-\tilde{\gamma }}g(t,P)^{\tilde{\gamma } }. \end{aligned}$$

Then from (D.2) we get

$$\begin{aligned} \left\{ \begin{array}{r} \frac{\partial g}{\partial t}+\frac{1}{\tilde{\gamma }}\left( \lambda \left( \alpha -\ln P\right) +\mu -\left( 1-\tilde{\gamma }\right) r\right) P\frac{\partial g}{\partial P}+\frac{1}{2}\sigma ^{2}P^{2}\frac{\partial ^{2} g}{\partial P^{2}}\\ +\left( \frac{\left( ^{1-\tilde{\gamma }}\right) r}{\tilde{\gamma }}+\frac{\left( 1-\tilde{\gamma }\right) \left( \lambda \left( \alpha -\ln P\right) +\mu -r\right) ^{2}}{2\sigma ^{2}\tilde{\gamma }^{2} }\right) g=0,\\ g(T,P)=\left( \frac{W^{1-\tilde{\gamma }}-1}{W^{1-\tilde{\gamma }}}\right) ^{\frac{1}{\tilde{\gamma }}}, \end{array} \right. \end{aligned}$$
(D.3)

which indeed is a one-dimensional linear parabolic equation. Similar to the baseline model in Section 2, we assume \(g(t,P)=\exp \left( f_{0}(t)+f_{1}(t)\ln P+f_{2}(t)(\ln P)^{2}\right) .\) Then the optimal control can be rewritten as:

$$\begin{aligned} \pi ^{*}(t,P)=\frac{\mu -r+\lambda (\alpha -\ln P)}{\sigma ^{2}\tilde{\gamma } }+f_{1}(t)+2f_{2}(t)\ln P, \end{aligned}$$

where \(f_{1},\) \(f_{2}\) satisfy the following system of ODEs:

$$\begin{aligned} \left\{ \begin{array}{ll} f_{1}^{\prime }-\frac{\lambda }{\tilde{\gamma }}f_{1}-\left( \sigma ^{2} -2r-\frac{2\left( \alpha \lambda +\mu -r\right) }{\tilde{\gamma }}\right) f_{2}+2\sigma ^{2}f_{1}f_{2}-\frac{\lambda \left( 1-\tilde{\gamma }\right) \left( \alpha \lambda +\mu -r\right) }{\sigma ^{2}\tilde{\gamma }^{2}}=0, &{} f_{1}\left( T\right) =0,\\ f_{2}^{\prime }+2\sigma ^{2}f_{2}^{2}-\frac{2\lambda }{\tilde{\gamma }}f_{2} +\frac{\lambda ^{2}\left( 1-\tilde{\gamma }\right) }{2\sigma ^{2}\tilde{\gamma }^{2}}=0, &{} f_{2}\left( T\right) =0. \end{array} \right. \end{aligned}$$

Through a series of calculations, we get

$$\begin{aligned} f_{1}(t)&=\frac{\lambda (\tilde{\gamma }-1)\left( \alpha \lambda +\mu -r\right) }{\sigma ^{2}\tilde{\gamma }^{2}}\hat{\Theta }\left( t\right) +\frac{\lambda ^{2}\left( \sigma ^{2}-2r\right) (\tilde{\gamma }-1)}{2\sigma ^{2}\tilde{\gamma }^{2}}\hat{B}\left( t\right) ,\\ f_{2}(t)&=-\frac{\lambda ^{2}(\tilde{\gamma }-1)}{2\sigma ^{2}\tilde{\gamma }^{2}}\hat{\Theta }\left( t\right) , \end{aligned}$$

where \(\hat{\Theta }\left( t\right) ,\) \(\hat{B}\left( t\right) \) are defined by (2.12), (2.13) with \(b=\frac{\lambda }{\tilde{\gamma }} \), \(q=\frac{\lambda }{\sqrt{\tilde{\gamma }}}\). \(\square \)

Appendix 5 Proof of Theorem 3.1

Let \(\alpha _{t}:=\ln \bar{P}_{t}=\ln \bar{P}_{0}+\beta t\). Similar to the baseline model in Section 2, we can get an equilibrium strategy by

$$\begin{aligned} \hat{\pi }(t,P)=\frac{\mu -r+\lambda \left( \alpha _{t}-\ln P\right) }{(1+\gamma )\sigma ^{2}}-\frac{\gamma P}{1+\gamma }\frac{\partial f}{\partial P}, \end{aligned}$$
(E.1)

where the function f satisfies the following PDE:

$$\begin{aligned} \left\{ \begin{array}{r} \frac{\partial f}{\partial t}+\left( \mu +\lambda \left( \alpha _{t}-\ln P\right) -\frac{\gamma ^{2}\left( \mu -r+\lambda \left( \alpha _{t}-\ln P\right) \right) }{\left( 1+\gamma \right) ^{2}}\right) P\frac{\partial f}{\partial P}\\ +\frac{1}{2}\sigma ^{2}P^{2}\left( \frac{\partial ^{2}f}{\partial P^{2}}-\frac{\gamma ^{2}}{\left( 1+\gamma \right) ^{2}}(\frac{\partial f}{\partial P})^{2}\right) \\ +\frac{\left( 2\gamma +1\right) \left( \mu -r+\lambda \left( \alpha _{t}-\ln P\right) \right) ^{2}}{2\left( \gamma +1\right) ^{2}\sigma ^{2}}=0,\\ f\left( T,P\right) =0. \end{array} \right. \end{aligned}$$
(E.2)

First we make a transform on the variable \(x=\ln P-\ln \bar{P}_{0}\). Then PDE (E.2) adapts to

$$\begin{aligned} \left\{ \begin{array}{r} \frac{\partial f}{\partial t}+\left( \mu +\lambda \left( \beta t-x\right) -\frac{\gamma ^{2}(\mu -r+\lambda \left( \beta t-x\right) )}{\left( 1+\gamma \right) ^{2}}\right) \frac{\partial f}{\partial x}\\ +\frac{1}{2} \sigma ^{2}\left( -\frac{\partial f}{\partial x}+\frac{\partial ^{2}f}{\partial x^{2}}-\frac{\gamma ^{2}}{\left( 1+\gamma \right) ^{2}}(\frac{\partial f}{\partial x})^{2}\right) \\ +\frac{\left( 2\gamma +1\right) (\mu -r+\lambda \left( \beta t-x\right) )^{2}}{2\left( \gamma +1\right) ^{2}\sigma ^{2}}=0,\\ f(T,x)=0. \end{array} \right. \end{aligned}$$
(E.3)

Let f(x) \(=-\frac{\left( 1+\gamma \right) ^{2}}{\gamma ^{2}}\ln u\left( t,x\right) \). Then PDE (E.3) satisfies

$$\begin{aligned} \left\{ \begin{array}{r} \frac{\partial u}{\partial t}+\left( \mu +\lambda \left( \beta t-x\right) -\frac{\gamma ^{2}(\mu -r+\lambda \left( \beta t-x\right) )}{\left( 1+\gamma \right) ^{2}}-\frac{1}{2}\sigma ^{2}\right) \frac{\partial u}{\partial x}\\ +\frac{1}{2}\sigma ^{2}\frac{\partial ^{2}u}{\partial x^{2}} -\frac{\left( 2\gamma +1\right) \gamma ^{2}(\mu -r+\lambda \left( \beta t-x\right) )^{2}}{2\left( \gamma +1\right) ^{4}\sigma ^{2}}u=0,\\ u\left( T,x\right) =1. \end{array} \right. \end{aligned}$$
(E.4)

Let \(\omega =\frac{\mu -r}{\lambda }-x\), \(\tilde{u}(t,\omega )=u(t,x).\) Then it yields

$$\begin{aligned} \left\{ \begin{array}{r} \frac{\partial \tilde{u}}{\partial t}-\left( r-\frac{1}{2}\sigma ^{2} +\frac{(1+2\gamma )\lambda }{\left( 1+\gamma \right) ^{2}}(\omega +\beta t)\right) \frac{\partial \tilde{u}}{\partial \omega }\\ +\frac{1}{2}\sigma ^{2} \frac{\partial ^{2}\tilde{u}}{\partial \omega ^{2}}-\frac{\left( 2\gamma +1\right) \lambda ^{2}\gamma ^{2}}{2\left( \gamma +1\right) ^{4}\sigma ^{2} }(\omega +\beta t)^{2}\tilde{u}=0,\\ \tilde{u}\left( T,\omega \right) =1. \end{array} \right. \end{aligned}$$

Set \(\tilde{u}=\exp \left\{ k(T-t)+\eta (\omega +\beta t)+\theta (\omega +\beta t)^{2}\right\} V(t,\omega )\). We obtain

$$\begin{aligned} \begin{aligned}&(-k+(\frac{\sigma ^{2}}{2}-r+\beta )\eta +\frac{1}{2}\sigma ^{2}\eta ^{2} +\sigma ^{2}\theta )V+(2(\frac{\sigma ^{2}}{2}-r+\beta )\theta \\&\quad -\frac{(1+2\gamma )\lambda }{(1+\gamma )^{2}}\eta +2\sigma ^{2}\eta \theta )(\omega +\beta t)V +(2\sigma ^{2}\theta ^{2}-2\frac{\lambda (1+2\gamma )}{(1+\gamma )^{2}} \theta \\ {}&\quad -\frac{\lambda ^{2}\gamma ^{2}(1+2\gamma )}{2(1+\gamma )^{4}\sigma ^{2} })(\omega +\beta t)^{2}V+V_{t}+\frac{1}{2}\sigma ^{2}V_{\omega \omega }\\&\left( (-r+\frac{1}{2}\sigma ^{2}+\sigma ^{2}\eta )+(2\sigma ^{2}\theta -\frac{(1+2\gamma )\lambda }{(1+\gamma )^{2}})(\omega +\beta t)\right) V_{\omega }=0. \end{aligned} \end{aligned}$$
(E.5)

Let the coefficient of V be 0 by separating the variable \((\omega +\beta t)\). Then we get the following ODE systems:

$$\begin{aligned} \left\{ \begin{array}{c} -k+(\frac{\sigma ^{2}}{2}-r+\beta )\eta +\frac{1}{2}\sigma ^{2}\eta ^{2}+\sigma ^{2}\theta =0,\\ 2(\frac{\sigma ^{2}}{2}-r+\beta )\theta -\frac{(1+2\gamma )\lambda }{(1+\gamma )^{2}}\eta +2\sigma ^{2}\eta \theta =0,\\ 2\sigma ^{2}\theta ^{2}-2\frac{\lambda (1+2\gamma )}{(1+\gamma )^{2}}\theta -\frac{\lambda ^{2}\gamma ^{2}(1+2\gamma )}{2(1+\gamma )^{4}\sigma ^{2}}=0. \end{array} \right. \end{aligned}$$

Set

$$\begin{aligned} a=\frac{\sigma ^{2}}{2}-r+\beta \text {, }b\text { and }c\text { are the same as in (A.7)}. \end{aligned}$$

Then we get the ODE system the same as (A.6). Hence \(\theta \), \(\eta \) have the same forms with (A.8). Consequently, (E.5) is reduced to

$$\begin{aligned} \left\{ \begin{array}{c} \frac{\partial V}{\partial t}+\left( (-r+\frac{1}{2}\sigma ^{2}+\sigma ^{2} \eta )+(2\sigma ^{2}\theta -\frac{(1+2\gamma )\lambda }{(1+\gamma )^{2}} )(\omega +\beta t)\right) \frac{\partial V}{\partial \omega }+\frac{1}{2} \sigma ^{2}\frac{\partial ^{2}V}{\partial \omega ^{2}}=0,\\ V\left( T,\omega \right) =e^{-\eta (\omega +\beta T)-\theta (\omega +\beta T)^{2}}. \end{array} \right. \end{aligned}$$

Then we denote \(\hat{a}:=a+\sigma ^{2}\eta =\mp \frac{ab}{q}\), \(\hat{b}:=2\sigma ^{2}\theta -b=\pm q\). Hence the above PDE becomes

$$\begin{aligned} \left\{ \begin{array}{l} \frac{\partial V}{\partial t}+\left( \hat{a}-\beta +\hat{b}(\omega +\beta t)\right) \frac{\partial V}{\partial \omega }+\frac{1}{2}\sigma ^{2} \frac{\partial ^{2}V}{\partial \omega ^{2}}=0,\\ V\left( T,\omega \right) =e^{-\eta (\omega +\beta T)-\theta (\omega +\beta T)^{2}}. \end{array} \right. \end{aligned}$$
(E.6)

We assume V has the following structure:

$$\begin{aligned} V\left( t,\omega \right) =\exp \left\{ A\left( t\right) +B\left( t\right) (\omega +\beta t)+C\left( t\right) (\omega +\beta t)^{2}\right\} . \end{aligned}$$

Substituting it into (E.6), and by separating variables, we get the same ODE system as (A.11). By solving the ODE system, we get an equilibrium strategy. The proof is complete. \(\square \)

Appendix 6 Proof of Theorem 3.2

By the general extended HJB in Björk et al. (2017), we have,

$$\begin{aligned} \underset{\pi }{\max }\left\{ \mathcal {D}V\left( t,P,\bar{P},R\right) -\frac{\gamma }{2}\left[ \sigma ^{2}P^{2}(\frac{\partial g}{\partial P} )^{2}+\sigma ^{2}\pi ^{2}(\frac{\partial g}{\partial R})^{2}+2\sigma ^{2}\pi P\frac{\partial g}{\partial P}\frac{\partial g}{\partial R}\right] \right\}&=0,\end{aligned}$$
(F.1)
$$\begin{aligned} \mathcal {D}^{*}g\left( t,P,\bar{P},R\right)&=0, \end{aligned}$$
(F.2)

subject to \(V(T,P,\bar{P},R)=R\), \(g(T,P,\bar{P},R)=R\), where \(\mathcal {D},\mathcal {D}^{*}\) are the infinitesimal operatorsFootnote 3. Now we assume that V and g are separable in R, i.e., they have the following structures:

$$\begin{aligned} V(t,P,\bar{P},R))&=\tilde{V}(t,P,\bar{P})+R+r\left( T-t\right) ;\end{aligned}$$
(F.3)
$$\begin{aligned} g\left( t,P,\bar{P},R)\right)&=f\left( t,P,\bar{P}\right) +R+r\left( T-t\right) . \end{aligned}$$
(F.4)

Substituting (F.3) and (F.4) into PDE (F.1) and PDE (F.2), it yields that

$$\begin{aligned}&\underset{\pi }{\max }\left\{ \mathcal {D}\tilde{V}\left( t,P,\bar{P}\right) +\left( \mu -r+\lambda \left( \ln \bar{P}-\ln P\right) \right) \pi -\frac{1}{2}\sigma ^{2}\pi ^{2}\right. \\ {}&\quad \qquad -\left. \frac{\gamma }{2}\left[ \sigma ^{2}P^{2}(\frac{\partial f}{\partial P})^{2}+2\sigma ^{2}\pi P\frac{\partial f}{\partial P}+\sigma ^{2}\pi ^{2}\right] \right\} =0,\\&\mathcal {D}^{*}f\left( t,P,\bar{P}\right) +\left( \mu -r+\lambda \left( \ln \bar{P}-\ln P\right) \right) \hat{\pi }-\frac{1}{2}\sigma ^{2}\hat{\pi }^{2} =0, \end{aligned}$$

subject to \(\tilde{V}(T,P,\bar{P})=0\), \(f(T,P,\bar{P})=0\).

Similar to the baseline model in Section 2, by the first-order condition, we get an equilibrium strategy by

$$\begin{aligned} \hat{\pi }(t,P,\bar{P})=\frac{\mu -r+\lambda \left( \ln \bar{P}-\ln P\right) }{(1+\gamma )\sigma ^{2}}-\frac{\gamma P}{1+\gamma }\frac{\partial f}{\partial P}, \end{aligned}$$

where the function f satisfies the following PDE

$$\begin{aligned} \begin{aligned} \frac{\partial f}{\partial t}&+\left( \mu +\lambda \left( \ln \bar{P}-\ln P\right) -\frac{\gamma ^{2}\left( \mu -r+\lambda \left( \ln \bar{P}-\ln P\right) \right) }{\left( 1+\gamma \right) ^{2}}\right) P\frac{\partial f}{\partial P}\\ {}&+\frac{1}{2}\sigma ^{2}P^{2}\left( \frac{\partial ^{2}f}{\partial P^{2}}-\frac{\gamma ^{2}}{\left( 1+\gamma \right) ^{2}}(\frac{\partial f}{\partial P})^{2}\right) \\&+\bar{P}\left[ \frac{\ln P-\ln \bar{P}}{t}\right] \frac{\partial f}{\partial \bar{P}}+\frac{\left( 2\gamma +1\right) \left( \mu -r+\lambda \left( \ln \bar{P}-\ln P\right) \right) ^{2}}{2\left( \gamma +1\right) ^{2}\sigma ^{2}}=0, \end{aligned} \end{aligned}$$
(F.5)

subject to terminal condition \(f\left( T,\bar{P},P\right) =0\).

Let \(f\left( t,\bar{P},P\right) =-\frac{\left( 1+\gamma \right) ^{2}}{\gamma ^{2}}\ln u\left( t,\bar{P},P\right) \). Then PDE (F.5) becomes

$$\begin{aligned} \frac{\partial u}{\partial t}&+\left( \mu +\lambda \left( \ln \bar{P}-\ln P\right) -\frac{\gamma ^{2}\left( \mu -r+\lambda \left( \ln \bar{P}-\ln P\right) \right) }{\left( 1+\gamma \right) ^{2}}\right) P\frac{\partial f}{\partial P}+\frac{1}{2}\sigma ^{2}P^{2}\frac{\partial ^{2}u}{\partial P^{2} }\\&+\bar{P}\left[ \frac{\ln P-\ln \bar{P}}{t}\right] \frac{\partial f}{\partial \bar{P}}-\frac{\left( 2\gamma +1\right) \gamma ^{2}\left( \mu -r+\lambda \left( \ln \bar{P}-\ln P\right) \right) ^{2} }{2\left( \gamma +1\right) ^{4}\sigma ^{2}}u=0, \end{aligned}$$

with terminal condition \(u\left( T,\bar{P},P\right) =1\). Let \(x=\ln \bar{P}-\ln P\), \(\tilde{u}(t,x):=u(t,\bar{P},P)\). Then we deduce

$$\begin{aligned} \left\{ \begin{aligned} \frac{\partial \tilde{u}}{\partial t}&-\left( \mu +\lambda x-\frac{\gamma ^{2}(\mu -r+\lambda x)}{\left( 1+\gamma \right) ^{2}}-\frac{1}{2}\sigma ^{2}+\frac{x}{t}\right) \frac{\partial \tilde{u}}{\partial x}\\&+\frac{1}{2}\sigma ^{2}\frac{\partial ^{2}\tilde{u}}{\partial x^{2}}-\frac{\left( 2\gamma +1\right) \gamma ^{2}(\mu -r+\lambda x)^{2}}{2\left( \gamma +1\right) ^{4}\sigma ^{2}}\tilde{u}=0, \quad \tilde{u}(T,x)=1. \end{aligned} \right. \end{aligned}$$
(F.6)

As in the baseline model, we set \(\tilde{u}(t,x)=\exp (A(t)+B(t)x+C(t)x^{2}).\) Substituting it into PDE (F.6) and by separating variables, we obtain

$$\begin{aligned} \left\{ \begin{array}{ll} A_{t}+\hat{a}B+\frac{1}{2}\sigma ^{2}\left( B^{2}+2C\right) -\frac{\gamma ^{2}(2\gamma +1)(\mu -r)^{2}}{2(\gamma +1)^{4}\sigma ^{2}}=0\text {,} &{} A\left( T\right) =0\\ B_{t}+(\hat{b}-\frac{1}{t})B+2\hat{a}C+2\sigma ^{2}BC-\frac{\lambda \gamma ^{2}(2\gamma +1)(\mu -r)}{(\gamma +1)^{4}\sigma ^{2}}=0\text {,} &{} B\left( T\right) =0\\ C_{t}+2(\hat{b}-\frac{1}{t})C+2\sigma ^{2}C^{2}-\frac{\lambda ^{2}\gamma ^{2}(2\gamma +1)}{2(\gamma +1)^{4}\sigma ^{2}}=0\text {,} &{} C\left( T\right) =0 \end{array} \right. \end{aligned}$$
(F.7)

where we denote \(\hat{a}=-(\mu -\frac{\gamma ^{2}(\mu -r)}{(1+\gamma )^{2}} -\frac{1}{2}\sigma ^{2})\), \(\hat{b}=-\frac{\lambda (1+2\gamma )}{(1+\gamma )^{2}}\).

From the above transformations, it’s easy to get that

$$\begin{aligned} f(t,\bar{P},P)=-\frac{\left( 1+\gamma \right) ^{2}}{\gamma ^{2}}\left[ A(t)+B(t)(\ln \bar{P}-\ln P)+C(t)(\ln \bar{P}-\ln P)^{2}\right] , \end{aligned}$$
(F.8)

where B(t) and C(t) satisfy (F.7). Therefore,

$$\begin{aligned} \hat{\pi }(t,P,\bar{P})=\frac{\mu -r+\lambda \left( \ln \bar{P}-\ln P\right) }{(1+\gamma )\sigma ^{2}}-\frac{1+\gamma }{\gamma }[B(t)+2C(t)(\ln \bar{P}-\ln P)]. \end{aligned}$$

The proof is complete. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lu, Z., Pang, P., Xu, Y. et al. Portfolio Selection with Contrarian Strategy. Methodol Comput Appl Probab 26, 16 (2024). https://doi.org/10.1007/s11009-024-10085-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11009-024-10085-y

Keywords

JEL Classification

JEL Classification

Navigation