Skip to main content
Log in

Optimal convergence trading with unobservable pricing errors

  • S.I.: Recent Developments in Financial Modeling and Risk Management
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

We study a dynamic portfolio optimization problem related to convergence trading, which is an investment strategy that exploits temporary mispricing by simultaneously buying relatively underpriced assets and selling short relatively overpriced ones with the expectation that their prices converge in the future. We build on the model of Liu and Timmermann (Rev Financ Stud 26(4):1048–1086, 2013) and extend it by incorporating unobservable Markov-modulated pricing errors into the price dynamics of two co-integrated assets. We characterize the optimal portfolio strategies in full and partial information settings under the assumption of unrestricted and beta-neutral strategies. By using the innovations approach, we provide the filtering equation which is essential for solving the optimization problem under partial information. Finally, in order to illustrate the model capabilities, we provide an example with a two-state Markov chain.

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

Similar content being viewed by others

References

  • Allinger, D . F., & Mitter, S . K. (1981). New results on the innovations problem for non-linear filtering. Stochastics: An International Journal of Probability and Stochastic Processes, 4(4), 339–348.

    Article  Google Scholar 

  • Altay, S., Colaneri, K., & Eksi, Z. (2018). Pairs trading under drift uncertainty and risk penalization. International Journal of Theoretical and Applied Finance, 21(07), 1850046.

    Article  Google Scholar 

  • Angoshtari, B. (2016). On the market-neutrality of optimal pairs-trading strategies. Available at SSRN: https://ssrn.com/abstract=2831836.

  • Bain, A., & Crisan, D. (2009). Fundamentals of stochastic filtering (Vol. 3). Berlin: Springer.

    Book  Google Scholar 

  • Baran, N. A., Yin, G., & Zhu, C. (2013). Feynman–Kac formula for switching diffusions: Connections of systems of partial differential equations and stochastic differential equations. Advances in Difference Equations, 2013(1), 315.

    Article  Google Scholar 

  • Bäuerle, N., & Rieder, U. (2004). Portfolio optimization with Markov-modulated stock prices and interest rates. IEEE Transactions on Automatic Control, 49(3), 442–447.

    Article  Google Scholar 

  • Bäuerle, N., & Rieder, U. (2005). Portfolio optimization with unobservable Markov-modulated drift process. Journal of Applied Probability, 42(2), 362–378.

    Article  Google Scholar 

  • Björk, T., Davis, M. H. A., & Landén, C. (2010). Optimal investment under partial information. Mathematical Methods of Operations Research, 71(2), 371–399.

    Article  Google Scholar 

  • Cartea, Á., & Jaimungal, S. (2016). Algorithmic trading of co-integrated assets. International Journal of Theoretical and Applied Finance, 19(06), 1650038.

    Article  Google Scholar 

  • Chiu, M. C., & Wong, H. Y. (2011). Mean-variance portfolio selection of co-integrated assets. Journal of Economic Dynamics and Control, 35(8), 1369–1385.

    Article  Google Scholar 

  • Damian, C., Eksi, Z., & Frey, R. (2018). EM algorithm for Markov chains observed via Gaussian noise and point process information: Theory and case studies. Statistics & Risk Modeling, 35(1–2), 51–72.

    Article  Google Scholar 

  • De Jong, A., Rosenthal, L., & Van Dijk, M. A. (2009). The risk and return of arbitrage in dual-listed companies. Review of Finance, 13(3), 495–520.

    Article  Google Scholar 

  • Elliott, R. J. (1993). New finite-dimensional filters and smoothers for noisily observed Markov chains. IEEE Transactions on Information Theory, 39(1), 265–271.

    Article  Google Scholar 

  • Elliott, R. J., & Bradrania, R. (2017). Estimating a regime switching pairs trading model. Quantitative Finance, 18(5), 877–883.

    Article  Google Scholar 

  • Escobar, M., Neykova, D., & Zagst, R. (2015). Portfolio optimization in affine models with Markov switching. International Journal of Theoretical and Applied Finance, 18(05), 1550030.

    Article  Google Scholar 

  • Fleming, W., & Pardoux, E. (1982). Optimal control for partially observed diffusions. SIAM Journal on Control and Optimization, 20(2), 261–285.

    Article  Google Scholar 

  • Frazzini, A., & Pedersen, L. H. (2014). Betting against beta. Journal of Financial Economics, 111(1), 1–25.

    Article  Google Scholar 

  • Frey, R., Gabih, A., & Wunderlich, R. (2012). Portfolio optimization under partial information with expert opinions. International Journal of Theoretical and Applied Finance, 15(1), 1250009.

    Article  Google Scholar 

  • Fujisaki, M., Kallianpur, G., & Kunita, H. (1972). Stochastic differential equations for the non linear filtering problem. Osaka Journal of Mathematics, 9(1), 19–40.

    Google Scholar 

  • Hansen, B. E., & Seo, B. (2002). Testing for two-regime threshold cointegration in vector error-correction models. Journal of Econometrics, 110(2), 293–318.

    Article  Google Scholar 

  • Jacod, J., & Shiryaev, A. N. (1987). Limit theorems for stochastic processes (Vol. 2003). Berlin: Springer.

    Book  Google Scholar 

  • Jurek, J. W., & Yang, H. (2007). Dynamic portfolio selection in arbitrage. EFA 2006 Meetings Paper. Available at SSRN: https://ssrn.com/abstract=882536.

  • Krolzig, H. M. (1997). Statistical analysis of cointegrated VAR processes with Markovian regime shifts. Oxford: Nuffield College, Unpublished preprint.

  • Kuwana, Y. (1995). Certainty equivalence and logarithmic utilities in consumption/investment problems. Mathematical Finance, 5(4), 297–309.

    Article  Google Scholar 

  • Lee, S., & Papanicolaou, A. (2016). Pairs trading of two assets with uncertainty in co-integration’s level of mean reversion. International Journal of Theoretical and Applied Finance, 19(08), 1650054.

    Article  Google Scholar 

  • Lei, Y., & Xu, J. (2015). Costly arbitrage through pairs trading. Journal of Economic Dynamics and Control, 56, 1–19.

    Article  Google Scholar 

  • Liu, J., & Longstaff, F. A. (2003). Losing money on arbitrage: Optimal dynamic portfolio choice in markets with arbitrage opportunities. The Review of Financial Studies, 17(3), 611–641.

    Article  Google Scholar 

  • Liu, J., & Timmermann, A. (2013). Optimal convergence trade strategies. The Review of Financial Studies, 26(4), 1048–1086.

    Article  Google Scholar 

  • Mudchanatongsuk, S., Primbs, J. A., & Wong, W. (2008). Optimal pairs trading: A stochastic control approach. In 2008 American control conference (pp. 1035–1039). IEEE.

  • Sotomayor, L. R., & Cadenillas, A. (2009). Explicit solutions of consumption-investment problems in financial markets with regime switching. Mathematical Finance, 19(2), 251–279.

    Article  Google Scholar 

  • Teschl, G. (2012). Ordinary differential equations and dynamical systems (Vol. 140). Philadelphia: American Mathematical Society.

    Book  Google Scholar 

  • Tourin, A., & Yan, R. (2013). Dynamic pairs trading using the stochastic control approach. Journal of Economic Dynamics and Control, 37(10), 1972–1981.

    Article  Google Scholar 

  • Vijh, A. M. (1994). S&P 500 trading strategies and stock betas. The Review of Financial Studies, 7(1), 215–251.

    Article  Google Scholar 

  • Zhou, X. Y., & Yin, G. (2003). Markowitz’s mean-variance portfolio selection with regime switching: A continuous-time model. SIAM Journal on Control and Optimization, 42(4), 1466–1482.

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to acknowledge two anonymous referees for their valuable comments and suggestions. We also would like to thank participants of 9th General AMaMeF Conference, Vienna Congress on Mathematical Finance 2019, and Brown Bag seminar at Institute for Statistics and Mathematics at WU Wien for constructive discussions. Part of this work has been done while Sühan Altay was affiliated with the Department of Financial and Actuarial Mathematics of TU Wien, and Katia Colaneri was affiliated with the School of Mathematics of the University of Leeds. Sühan Altay gratefully acknowledges financial support from the funds of the OeNB Jubiläumsfonds 17804. The work of Katia Colaneri has been partially supported by INdAM GNAMPA through the Project U-UFMBAZ-2019-000436.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Katia Colaneri.

Additional information

Publisher's Note

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

Proofs

Proofs

Proof of Theorem 1

Existence We denote by \({\mathcal {L}}^h_{\mathbb {G}}\) the generator of the process (tWXY), that is

$$\begin{aligned} {\mathcal {L}}^h_{\mathbb {G}}F(t,w,x,i)&=F_t(t,w,x,i) + \left( \varGamma _1 +\lambda _1^i\alpha _1^i+\lambda _2^i\alpha _2^i- (\lambda _1^i + \lambda _2^i)x \right) F_x(t,w,x,i)\\&+w \left( r+ \mu _m \left( h^{(m)}+ h^{(1)}\beta _1 + h^{(2)}\beta _2 \right) \right. \\&+ \left. h^{(2)}\lambda _2^i(x-\alpha _2^i) - h^{(1)}\lambda _1^i(x-\alpha _1^i) \right) F_{w}(t,w,x,i) \\&+\frac{1}{2} w^2 \left( \sigma _m^2 \left( h^{(m)}+ h^{(1)}\beta _1 + h^{(2)}\beta _2 \right) ^2 + \sigma ^2 \left( h^{(1)}+ h^{(2)}\right) ^2 \right. \\&+\left. (h^{(1)}b_1)^2 + (h^{(2)}b_2)^2 \right) F_{ww}(t,w,x,i)\\&+ w \left( \sigma _m^2 ( \beta _1 - \beta _2 ) \left( h^{(m)}+ h^{(1)}\beta _1 + h^{(2)}\beta _2 \right) + h^{(1)}b_1 - h^{(2)}b_2 \right) F_{wx}(t,w,x,i) \\&+ \frac{1}{2}\varGamma _2F_{xx}(t,w,x,i)+\sum _{j=1}^K F (t,w,x,j)q^{ij}, \end{aligned}$$

for every function \(F(\cdot , i)\in C^{1,2,2}([0,T]\times {\mathbf {R}}_{+}\times {\mathbf {R}})\), i.e. bounded, differentiable with respect to t and twice differentiable with respect to w and x, for every \(i\in \{1, \dots , K\}\).

Suppose that the value function \(V(\cdot , i)\in C^{1,2,2}([0,T]\times {\mathbf {R}}_{+}\times {\mathbf {R}})\) for every \(i\in \{1,\dots ,K\}\). Then it solves the HJB equation given by

$$\begin{aligned} 0=\underset{h\in {\mathcal {A}}}{\sup }{\mathcal {L}}^h V(t,w,x,i) \end{aligned}$$
(45)

for every \(i\in \{1,\dots ,K \}\), subject to the terminal condition \(V(T,w,x,i)=\log (w)\), for all \((w,x)\in {\mathbf {R}}_{+}\times {\mathbf {R}}\) and \(i\in \{1,\dots ,K \}\). It follows from the form of the utility function that for all \(i\in \{1,\dots ,K \}\) the value function can be rewritten as \(V(t,w,x,i)=\log (w)+\nu (t,x,i)\), for some function \(\nu (t,x,i)\) such that \(\nu (T,x,i)=0\). Inserting the ansatz for the value function in Eq. (45) and taking first order conditions leads to

$$\begin{aligned} 0&=\frac{\mu _m}{\sigma _m^2}-h^{(1)}\beta _1-h^{(2)}\beta _2-h^{(m)},\\ 0&= \beta _1\mu _m- \lambda _1^i(x -\alpha _1^i) -\beta _1\sigma _m^2\left( h^{(m)}+ h^{(1)}\beta _1 + h^{(2)}\beta _2 \right) -\sigma ^2(h^{(1)}+h^{(2)})-h^{(1)}b_1^2 ,\\ 0&=\beta _2\mu _m+\lambda _2^i(x-\alpha _2^i)-\beta _2\sigma _m^2\left( h^{(m)}+h^{(1)}\beta _1+h^{(2)}\beta _2\right) -\sigma ^2(h^{(1)}+h^{(2)})-h^{(2)}b_2^2. \qquad {} \end{aligned}$$

Second order conditions imply that portfolio weights given in (10)–(12) are candidates to be optimal strategies. Next, we insert the optimal portfolio weights in the HJB equation. This yields the following PDE:

$$\begin{aligned} 0&=\nu _t(t,x,i)+\varTheta _1^ix^2-\varTheta _2^ix+\varTheta _3^i+\sum _{j=1}^K \nu (t,x,j)q^{ij}+ \frac{1}{2}\varGamma _2 \nu _{xx}(t,x,i)\nonumber \\&+\left( \varGamma _1+\lambda _1^i\alpha _1^i+\lambda _2^i\alpha _2^i-(\lambda _1^i+\lambda _2^i)x \right) \nu _x(t,x,i). \end{aligned}$$
(46)

We conjecture that \(\nu (t,x,i)=m(t,i)x^2+n(t,i)x+u(t,i)\). Substituting this ansatz in (46) results in a quadratic equation for x. Setting the coefficients of the terms \(x^2\), x and the independent term to zero yields that the functions m, n and u solve the system of ODEs given in (14)–(16) see, e.g., (2012, Theorem 3.9).

Verification. In the sequel we verify martingale conditions that ensure that V in (13) is indeed the value function. To this, let v(twxi) be a solution of the HJB equation (45) and \(h\in {\mathcal {A}}\) and admissible control. By Itô’s formula we get

$$\begin{aligned}&v(T, W_T^h, X_T, Y_T)=v(t,w,x,i)+\int _t^T {\mathcal {L}}{\mathcal {v}}(r, W^h_r, X_r, Y_r)\, \mathrm dr\\&\qquad +\int _t^T \sigma _m v_w(r, W^h_r, X_r, Y_r) W^h_r \left( h^{(m)}_r+h^{(1)}_r\beta _1+h^{(2)}_r\beta _2\right) \mathrm dB^{(m)}_r\\&\qquad +\int _t^T \sigma _m+ v_x(r, W^h_r, X_r, Y_r) \left( \beta _1-\beta _2\right) \mathrm dB^{(m)}_r\\&\qquad +\int _t^T v_w(r, W^h_r, X_r, Y_r) W^h_r \sigma \left( h^{(1)}_r+h^{(2)}_r \right) \mathrm dB^{(0)}_r\\&\qquad +\int _t^T \left( v_w(r, W^h_r, X_r, Y_r) W^h_r b_1h^{(1)}_r + v_x(r, W^h_r, X_r, Y_r) b_1\right) \mathrm dB^{(1)}_r\\&\qquad +\int _t^T \left( v_w(r, W^h_r, X_r, Y_r) W^h_r b_2h^{(2)}_r - v_x(r, W^h_r, X_r, Y_r) b_2\right) \mathrm dB^{(2)}_r\\&\qquad +\int _t^T\sum _{j=1}^Kv(r, W^h_r, X_r, j)-v(r, W^h_r, X_r, Y_{r^-}) (m-\nu )(\mathrm dr \times \{j\}). \end{aligned}$$

The last term in the expression above corresponds to the compensated integral with respect to the jump measure of Y, that is

$$\begin{aligned}&\int _t^T\sum _{j=1}^Kv(r, W^h_r, X_r, j) -v(r, W^h_r, X_r, Y_{r^-}) (m-\nu )(\mathrm dr \times \{j\})\\&\quad =\sum _{t\le r\le T}\varDelta v(r, W^h_r, X_r, Y_r)- \int _t^T\sum _{j=1}^Kv(r, W^h_r, X_r, j)-v(r, W^h_r, X_r, Y_{r^-}) q^{Y_{r^-} j}\, \mathrm dr. \end{aligned}$$

where \(\varDelta v(t, W^h_t, X_t, Y_t)=v(t, W^h_t, X_t, Y_t)-v(t, W^h_t, X_t, Y_{t^-})\) for every \(t \in [0,T]\),

$$\begin{aligned} m([0,t]\times \{j\}):=\sum _{n \ge 1} {{\mathbf {1}}}_{\{Y_{T_n}=j\}}{{\mathbf {1}}}_{\{T_n\le t\}},\quad j\in \{1,\dots ,K\},\,\,t\in [0,T], \end{aligned}$$

is the jump measure of Markov chain Y with the compensator

$$\begin{aligned} \nu ([0,t]\times \{j\})=\int _0^t\sum _{i \ne j} q^{ij}{{\mathbf {1}}}_{\{Y_{r^-}=i\}}\,\mathrm dr,\quad j\in \{1,\dots ,K\},\,\,t\in [0,T]. \end{aligned}$$

and \(\{T_n\}_{n \in {\mathbb {N}}}\) is the sequence of jump times of Y. Since v satisfies equation (45) we get

$$\begin{aligned}&v(T, W_T^h, X_T, Y_T) \\&\quad \le v(t,w,x,i)+ \int _t^T \sigma _m v_w(r, W^h_r, X_r, Y_r) W^h_r\left( h^{(m)}_r+h^{(1)}_r\beta _1+h^{(2)}_r\beta _2\right) \mathrm dB^{(m)}_r\\&\qquad +\int _t^T \sigma _m+ v_x(r, W^h_r, X_r, Y_r) \left( \beta _1-\beta _2\right) \mathrm dB^{(m)}_r\\&\qquad +\int _t^T v_w(r, W^h_r, X_r, Y_r) W^h_r \sigma \left( h^{(1)}_r+h^{(2)}_r \right) \mathrm dB^{(0)}_r\\&\qquad +\int _t^T \left( v_w(r, W^h_r, X_r, Y_r) W^h_r b_1h^{(1)}_r + v_x(r, W^h_r, X_r, Y_r) b_1\right) \mathrm dB^{(1)}_r\\&\qquad +\int _t^T \left( v_w(r, W^h_r, X_r, Y_r) W^h_r b_2h^{(2)}_r- v_x(r, W^h_r, X_r, Y_r) b_2\right) \mathrm dB^{(2)}_r\\&\qquad +\int _t^T\sum _{j=1}^Kv(r, W^h_r, X_r, j)-v(r, W^h_r, X_r, Y_{r^-}) (m-\nu )(\mathrm dr \times \{j\}). \end{aligned}$$

The form of v and integrability condition (7) ensure that integrals with respects to Brownian motions \(B^{(m)}, B^{(0)},B^{(1)},B^{(2)} \) and the compensated jump measure \(m-\nu \) are true \(({\mathbb {G}}, {\mathbf {P}})\)-martingales. Then, taking expectations we get that

$$\begin{aligned} V(t,w,x,i)\le v(t,w,x,i), \end{aligned}$$

and the equality holds if h is a maximizer of Eq. (45). \(\square \)

Proof of Proposition 1

In the following we use the notation \(\widehat{g(Y_t)}={\mathbf {E}}\left[ g(Y_t)|{\mathcal {F}}_t\right] \), \(t\in [0,T]\). Consider the semimartingale decomposition of f(Y) given by

$$\begin{aligned} f(Y_t)= f(Y_0)+\int _0^t \langle Q{\mathbf {f}}, Y_{u^-}\rangle \, \mathrm du + M^{(1)}_t,\quad t \in [0,T], \end{aligned}$$

where \(M^{(1)}\) is a \(({\mathbb {G}}, {\mathbf {P}})\)-martingale. Now, projecting over \({\mathbb {F}}\) leads to

$$\begin{aligned} \widehat{f(Y_t)}- \widehat{f(Y_0)}-\int _0^t \langle Q{\mathbf {f}}, {\widehat{Y}}_{u^-}\rangle \, \mathrm du = M^{(2)}_t,\quad t \in [0,T], \end{aligned}$$

where \(M^{(2)}\) is an \(({\mathbb {F}}, {\mathbf {P}})\)-martingale. Using the martingale representation in (29) we get

$$\begin{aligned} \widehat{f(Y_t)}- \widehat{f(Y_0)}-\int _0^t \langle Q{\mathbf {f}}, {\widehat{Y}}_{u^-}\rangle \, \mathrm du = \int _0^t\gamma _u\, \mathrm dI_u,\quad t \in [0,T]. \end{aligned}$$

Let \(m_t=I_t+\int _0^tX_u \varSigma ^{-1} \widehat{A(X_u,Y_u)}\,\mathrm du\), for every \(t \in [0,T]\). Computing the product \(f(Y)\cdot m\) and projecting on \({\mathbb {F}}\), we obtain

$$\begin{aligned} \widehat{f(Y_t)\cdot m_t}= \int _0^t m_u \langle Q{\mathbf {f}}, {\widehat{Y}}_u\rangle \, \mathrm du + \int _0^t X_u \varSigma ^{-1}\left( \widehat{f(Y_u) A(X_u,Y_u)}\right) \,\mathrm du + M^{(3)}_t,\,\, \quad {}\end{aligned}$$
(47)

for every \( t \in [0,T]\) and for some \(({\mathbb {F}}, {\mathbf {P}})\)-martingale \(M^{(3)}\). The hat in the second integrand of Eq. (47) stands for \(\widehat{f(Y_u) A(X_u,Y_u)}={\mathbf {E}}\left[ f(Y_u) A(X_u,Y_u)|{\mathcal {F}}_u\right] \).

We now compute the product \(\widehat{f(Y)}\cdot m\) as

$$\begin{aligned} \widehat{f(Y_t)} \cdot m_t= \int _0^t m_u \langle Q\mathbf{f}, {\widehat{Y}}_u\rangle \,\mathrm du + \int _0^t X_u \varSigma ^{-1}\widehat{f(Y_u)} \widehat{A(X_u,Y_u)}\, \mathrm du + \int _0^t \gamma _u \,\mathrm du + M^{(4)}_t. \end{aligned}$$
(48)

for every \(t \in [0,T]\), where \(M^{(4)}\) is an \(({\mathbb {F}}, {\mathbf {P}})\)-martingale. Comparing the finite variation terms in (47) and (48), we get

$$\begin{aligned} \gamma ^{(1)}_t&=\frac{\widehat{f(Y_t) \mu _1(X_t,Y_t)}-\widehat{f(Y_t)}\widehat{\mu _1(X_t,Y_t)}}{\sigma _1},\\ \gamma ^{(2)}_t&=\frac{\sigma _1 (\widehat{f(Y_t)\mu _2(X_t,Y_t)}-\widehat{f(Y_t)}\widehat{\mu _2(X_t,Y_t)})-\sigma _2\rho (\widehat{f(Y_t)\mu _1(X_t,Y_t)}-\widehat{f(Y_t)}\widehat{\mu _1(X_tY_t)})}{\sigma _1\sigma _2\sqrt{1-\rho ^2}}, \end{aligned}$$

for every \(t \in [0,T]\). By taking \(f(Y_t)={{\mathbf {1}}}_{\{Y_t=e_i\}}\), we obtain the result. Finally, since the drift and diffusion coefficients in (30) are continuous, bounded and locally Lipschitz, we get that \(\varvec{\pi }=(\pi ^1, \dots , \pi ^K)\) is the unique strong solution of the system (30) . \(\square \)

Proof of Theorem 3

Existence For notational ease we set \(\sigma _1=\sqrt{\sigma ^2+b_1^2}\) and \(\sigma _2=\sqrt{\sigma ^2+b_2^2}\). Assume first that function \(V(t,w,x,{\mathbf {p}})\) is regular. Then it satisfies the following HJB equation

$$\begin{aligned} 0= \underset{h\in {\mathcal {A}}^{{\mathbb {F}}}}{\sup }{\mathcal {L}}_{\mathbb {F}}^h V(t,w,x,{\mathbf {p}}) \end{aligned}$$
(49)

subject to the terminal condition \(V(T,w,x,{\mathbf {p}})=\log (w)\), for all \(w>0\), \(x\in {\mathbb {R}}\) and for every \({\mathbf {p}}\in \varDelta _K\), where \({\mathcal {L}}_{\mathbb {F}}^h\) is given by

$$\begin{aligned}&{\mathcal {L}}_{\mathbb {F}}^h f(t,w,x,{\mathbf {p}})=\left\{ f_t + f_x\left( \varGamma _1+\varvec{\mu }_1(x)^\top {\mathbf {p}}-\varvec{\mu }_2(x)^\top {\mathbf {p}}\right) + \sum _{i,j=1}^K f_{p^i} q^{ji}p^j\right. \nonumber \\&\qquad \left. + \left( r+\left( h^{(m)}+ h^{(1)}\beta _1+h^{(2)}\beta _2 \right) \mu _m +h^{(1)}\varvec{\mu }_1(x)^\top {\mathbf {p}}+ h^{(2)}\varvec{\mu }_2(x)^\top {\mathbf {p}}\right) w f_w(t,w,x,{\mathbf {p}})\right. \nonumber \\&\qquad \left. + \frac{1}{2} f_{xx}\varGamma _2 + \frac{1}{2} \sum _{i,j=1}^Kf_{p^ip^j} (H^{(i),1}({\mathbf {p}}) H^{(j),1}({\mathbf {p}})+H^{(i),2}({\mathbf {p}}) H^{(j),2}({\mathbf {p}}))\right. \nonumber \\&\qquad \left. +\frac{1}{2} f_{ww}w^2 \left( (h^{(m)}+ h^{(1)}\beta _1+h^{(2)}\beta _2 )^2\sigma _m^2 +(\sigma _1 h^{(1)}+\rho \sigma _2h^{(2)})^2 + \sigma _2^2 (1-\rho ^2){h^{(2)}}^2\right) \right. \nonumber \\&\qquad \left. + f_{wx}w \left( \sigma _m^2(\beta _1-\beta _2)(h^{(m)}+ h^{(1)}\beta _1+h^{(2)}\beta _2 )+\sigma _1^2 h^{(1)}{-} \sigma _2^2 h^{(2)}{-}\rho \sigma _1\sigma _2 (h^{(1)}-h^{(2)})\right) \right. \nonumber \\&\qquad \left. + \sum _{i=1}^K f_{wp^i}w \left( H^{(i),1}({\mathbf {p}}) (\sigma _1 h^{(1)}+\rho \sigma _2h^{(2)})+ H^{(i),2}({\mathbf {p}})\sqrt{1-\rho ^2}h^{(2)}\sigma _2\right) \right. \nonumber \\&\qquad \left. + \sum _{i=1}^K f_{xp^i} \left( (\sigma _1-\rho \sigma _2) H^{(i),1}({\mathbf {p}}) - H^{(i),2}({\mathbf {p}})\sigma _2\sqrt{1-\rho ^2}\right) \right\} \end{aligned}$$
(50)

for every function \(f:[0,T]\times {\mathbb {R}}^+ \times {\mathbb {R}}\times \varDelta _K \rightarrow {\mathbb {R}}\), which is bounded, differentiable with respect to time and twice differentiable with respect to \((w,x,{\mathbf {p}})\) with bounded derivatives. By the form of the utility function we have that the value function has the form \(V(t,w,x,\pi )=\log (w)+v(t,x, \pi )\), for some function \(v(t,x,\pi )\), such that \(v(T,x,{\mathbf {p}})=0\) for all \((x,{\mathbf {p}})\in ({\mathbb {R}}\times \varDelta _K)\). By inserting the first ansatz in Eq. (50) and considering the first order condition we get that the candidate for an optimal strategy is given by (37), (38),(39). Since \(V(t,w,x,{\mathbf {p}})\) is concave and increasing in w, the second order condition implies that (37),(38) and (39) is the maximizer and the optimal portfolio strategy. Here, we choose v of the form \(v(t,x,{\mathbf {p}})={{\bar{m}}}({\mathbf {p}})x^2+{{\bar{n}}} (t,{\mathbf {p}})x+{{\bar{u}}}(t,{\mathbf {p}})\). Inserting this ansatz in Eq. (50) leads to the system of linear partial differential equations in (41), (42), (43).

Verification. To conclude that V is the value function, we show a verification result. Let \({\widetilde{V}}(t,w,x,{\mathbf {p}})\) be a solution of (49) with the boundary condition \({\widetilde{V}}(T,w,x,{\mathbf {p}})=\log (w)\). Let \(h\in {\mathcal {A}}^{{\mathbb {F}}}\) be an \({\mathbb {F}}\)-admissible control, let \(W^h\) the solution to Eq. (34). By applying Itô’s formula we get

$$\begin{aligned}&{\widetilde{V}}(T, W_T^h, X_T,\varvec{\pi }_T)= {\widetilde{V}}(t,w,x,{\mathbf {p}})+\int _t^T {\mathcal {L}}_{\mathbb {F}}^h {\widetilde{V}}(u, W^h_u, X_u, \varvec{\pi }_u)\, \mathrm du \\&\qquad + \int _t^T \left( {\widetilde{V}}_w(u, W^h_u, X_u, \varvec{\pi }_u) W^h_u (h^{(m)}_u+ h^{(1)}_u\beta _1+ h^{(2)}_u \beta _2 )\right. \\&\qquad \left. + {\widetilde{V}}_x(u, W^h_u, X_u, \varvec{\pi }_u) (\beta _i-\beta _2) \right) \sigma _m \mathrm dB^{(m)}_u\\&\qquad +\int _t^T \left( {\widetilde{V}}_w(u, W^h_u, X_u, \varvec{\pi }_u) W^h_u (\sigma _1h^{(1)}{+}\rho \sigma _2 h^{(2)}) {+} (\sigma _1-\rho \sigma _2) {\widetilde{V}}_x(u, W^h_u, X_u, \varvec{\pi }_u)\right) \mathrm dI^{(1)}_u \\&\qquad +\int _t^T\sum _{i=1}^K{\widetilde{V}}_{p^i}(u, W^h_u, X_u, \varvec{\pi }_u) {{\bar{H}}}^{(i),1}(\varvec{\pi }_t) \mathrm dI^{(1)}_u\\&\qquad + \int _t^T \left( \sigma _2 \sqrt{1-\rho ^2}( {\widetilde{V}}_w(u, W^h_u, X_u, \varvec{\pi }_u) W^h_u h^{(2)}_u -{\widetilde{V}}_x(u, W^h_u, X_u, \varvec{\pi }_u)) \right) \mathrm dI^{(2)}_u\\&\qquad + \int _t^T \sum _{i=1}^K{\widetilde{V}}_{p^i}(u, W^h_u, X_u, \varvec{\pi }_u) H^{(i), 2} (\pi _t) \mathrm dI^{(2)}_u. \end{aligned}$$

By Eq. (50) we get

$$\begin{aligned}&{\widetilde{V}}(T, W_T^h, X_T,\varvec{\pi }_T)\le {\widetilde{V}}(t,w,x,{\mathbf {p}}) \nonumber \\&\qquad + \int _t^T \left( {\widetilde{V}}_w(u, W^h_u, X_u, \varvec{\pi }_u) W^h_u (h^{(m)}_u+ h^{(1)}_u\beta _1+ h^{(2)}_u \beta _2 )\right. \nonumber \\&\qquad \left. + {\widetilde{V}}_x(u, W^h_u, X_u, \varvec{\pi }_u) (\beta _i-\beta _2) \right) \sigma _m \mathrm dB^{(m)}_u\nonumber \\&\qquad +\int _t^T \left( {\widetilde{V}}_w(u, W^h_u, X_u, \varvec{\pi }_u) W^h_u (\sigma _1h^{(1)}{+}\rho \sigma _2 h^{(2)}) {+} (\sigma _1-\rho \sigma _2) {\widetilde{V}}_x(u, W^h_u, X_u, \varvec{\pi }_u)\right) \mathrm dI^{(1)}_u \nonumber \\&\qquad +\int _t^T\sum _{i=1}^K{\widetilde{V}}_{p^i}(u, W^h_u, X_u, \varvec{\pi }_u) {{\bar{H}}}^{(i),1}(\varvec{\pi }_t) \mathrm dI^{(1)}_u\nonumber \\&\qquad + \int _t^T \left( \sigma _2 \sqrt{1-\rho ^2}( {\widetilde{V}}_w(u, W^h_u, X_u, \varvec{\pi }_u) W^h_u h^{(2)}_u -{\widetilde{V}}_x(u, W^h_u, X_u, \varvec{\pi }_u)) \right) \mathrm dI^{(2)}_u\nonumber \\&\qquad + \int _t^T \sum _{i=1}^K{\widetilde{V}}_{p^i}(u, W^h_u, X_u, \varvec{\pi }_u) H^{(i), 2} (\varvec{\pi }_t) \mathrm dI^{(2)}_u. \end{aligned}$$
(51)

Note that stochastic integrals with respect to \(B^{(m)},I^{(1)}\) and \(I^{(2)}\) are true martingales. This is a consequence of the fact that function \({\widetilde{V}}(t,w,x,{\mathbf {p}})=\log (w)+{{\bar{m}}}(t)x^2+\bar{n}(t,{\mathbf {p}})x+{{\bar{u}}}(t, {\mathbf {p}})\) solves the HJB equation, that \((h^{(m)}, h^{(1)}, h^{(2)})\) is an \({\mathbb {F}}\)-admissible strategy and that functions \(\bar{m}(t),{{\bar{n}}}(t,{\mathbf {p}}), {{\bar{u}}}(t,{\mathbf {p}})\) and their derivatives are bounded over the compact interval \([0,T]\times \varDelta _K\). Then taking the expectation on both sides of inequality (51) implies that \(V(t,w,x,{\mathbf {p}})\le {\widetilde{V}}(t,w,x,{\mathbf {p}})\). Moreover if \(({h^{(m)}}^*,{h^{(1)}}^*,{h^{(2)}}^ *) \) is a maximizer of Eq. (49), then we obtain the equality \(V(t,w,x,{\mathbf {p}})= {\widetilde{V}}(t,w,x,{\mathbf {p}})\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Altay, S., Colaneri, K. & Eksi, Z. Optimal convergence trading with unobservable pricing errors. Ann Oper Res 299, 133–161 (2021). https://doi.org/10.1007/s10479-020-03647-z

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-020-03647-z

Keywords

Navigation