Skip to main content
Log in

Optimal Investment in an Illiquid Market with Search Frictions and Transaction Costs

  • Published:
Applied Mathematics & Optimization Submit manuscript

Abstract

We consider an optimal investment problem to maximize expected utility of the terminal wealth, in an illiquid market with search frictions and transaction costs. In the market model, an investor’s attempt of transaction is successful only at arrival times of a Poisson process, and the investor pays proportional transaction costs when the transaction is successful. We characterize the no-trade region describing the optimal trading strategy. Our asymptotic analysis implies that the effects of the transaction costs are more pronounced (more widening effect of the no-trade region and more diminishing effect of the value function) in the market with less search frictions.

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

Similar content being viewed by others

Notes

  1. For instance, [4, 12, 20, 37] study how asymmetric information effects price impact and optimal trading strategy in equilibrium. [2, 23, 46, 47] consider optimal order execution problems with exogenously given price impacts.

  2. Here, perfect liquidity assumption means that assets can be traded in any quantity and at any moment in time, without any transaction costs.

  3. In the similar model set up without the search frictions, it is well known that the width of the no-trade region is the order of \(\epsilon ^{\frac{1}{3}}\) and the decrease of the value function is the order of \(\epsilon ^{\frac{2}{3}}\) (e.g., see [7, 8, 10, 11, 25, 29, 45, 51]).

  4. To be more specific, let \(0<\lambda _1<\lambda _2\) and consider two markets \({\mathcal {M}}_1\) and \({\mathcal {M}}_2\) with search friction parameters \(\lambda _1\) and \(\lambda _2\), respectively. In words, \({\mathcal {M}}_2\) has less search frictions than \({\mathcal {M}}_1\). Our result implies that if we increase the transaction costs, the widening speed of the no-trade region in \({\mathcal {M}}_2\) is faster than that in \({\mathcal {M}}_1\). Similarly, if we increase the transaction costs, the diminishing speed of the optimal value in \({\mathcal {M}}_2\) is faster than that in \({\mathcal {M}}_1\).

  5. Therefore, bigger \(\lambda \) implies more frequent trading opportunities (less search frictions), on average.

  6. Indeed, for \(s>t\) and \(A=\{ W^{(1)}_{t}<0, W^{(0)}_{t} \ge 0\}\) (or \(A=\{W^{(1)}_{t} \ge 0, W^{(0)}_{t}<0\}\)), we observe that \({{\mathbb {P}}}(W_s<0 \, | \,A) \ge {{\mathbb {P}}}(P_t=P_s \text { and } W_s<0 \, | \, A)>0\).

  7. To check this, let \(\tau _n:=\inf \{ t\ge 0: P_t=n\}\) and \(\tau _0=0\). For \(\tau _n\le t<\tau _{n+1}\), the dynamics (2.3) produce \(W_t=W_{\tau _n}^{(0)} e^{r(t-\tau _n)} +W_{\tau _n}^{(1)} e^{(\mu -\frac{\sigma ^2}{2})(t-\tau _n)+\sigma (B_t-B_{\tau _n})}\). If \(W_{\tau _n-}>0, W_{\tau _n-}^{(0)}\ge 0, W_{\tau _n-}^{(1)}\ge 0\) and (2.4) are satisfied, then \(W_{\tau _n}>0, W_{\tau _n}^{(0)}\ge 0\) and \(W_{\tau _n}^{(1)}\ge 0\) hold. By this way, we can inductively show that \(W_{\tau _n}>0, W_{\tau _n}^{(0)}\ge 0\) and \(W_{\tau _n}^{(1)}\ge 0\) for all n, almost surely. Now the expression of \(W_t\) above implies that \(W_t>0\) for all \(t\ge 0\), almost surely.

  8. Indeed, the SDE in (2.5) without \(dP_t\) term has a unique (explicit) solution. The unique solution of (2.5) can be obtained by patching the unique solutions on time intervals between the jump times of the Poisson process, with the jump size described by the coefficient of \(dP_t\) term.

  9. The inequality (3.15) becomes strict on the event \(\left\{ \omega \in \Omega : \exists t\in [0,T] \text { such that }\Delta P_t(\omega )=1 \text { and }{\hat{M}}_t^0(\omega ) \hat{M}_t^1(\omega )<0\right\} \).

References

  1. Akian, M., Menaldi, J., Sulem, A.: On an investment-consumption model with transaction costs. SIAM J. Control. Optim. 34, 329–364 (1996)

    MathSciNet  MATH  Google Scholar 

  2. Almgren, R., Chriss, N.: Optimal execution of portfolio transactions. J. Risk 3, 5–40 (2001)

    Google Scholar 

  3. Ang, A., Papanikolaou, D., Westerfield, M.M.: Portfolio choice with illiquid assets. Manag. Sci. 60, 2737–2761 (2014)

    Google Scholar 

  4. Back, K.: Insider trading in continuous time. Rev. Financ. Stud. 5, 387–409 (1992)

    Google Scholar 

  5. Bayraktar, E., Dolinskyi, L., Dolinsky, Y.: Extended weak convergence and utility maximisation with proportional transaction costs. Financ. Stochast. 24, 1013–1034 (2020)

    MathSciNet  MATH  Google Scholar 

  6. Bertsekas, D.P., Shreve, S.: Stochastic optimal control: the discrete-time case (2004)

  7. Bichuch, M., Guasoni, P.: Investing with liquid and illiquid assets. Math. Financ. 28, 119–152 (2018)

    MathSciNet  MATH  Google Scholar 

  8. Bichuch, M., Shreve, S.E.: Utility maximization trading two futures with transaction costs. SIAM J. Financ. Math. 4, 26–85 (2013)

    MathSciNet  MATH  Google Scholar 

  9. Chen, X., Dai, M.: Characterization of optimal strategy for multiasset investment and consumption with transaction costs. SIAM J. Financ. Math. 4, 857–883 (2013)

    MathSciNet  MATH  Google Scholar 

  10. Choi, J.: Asymptotic analysis for Merton’s problem with transaction costs in power utility case. Stochastics 86, 803–816 (2014)

    MathSciNet  MATH  Google Scholar 

  11. Choi, J.H.: Optimal consumption and investment with liquid and illiquid assets. Math. Financ. 30, 621–663 (2020)

    MathSciNet  MATH  Google Scholar 

  12. Choi, J.H., Larsen, K., Seppi, D.J.: Information and trading targets in a dynamic market equilibrium. J. Financ. Econ. 132, 22–49 (2019)

    Google Scholar 

  13. Cretarola, A., Gozzi, F., Pham, H., Tankov, P.: Optimal consumption policies in illiquid markets. Financ. Stochast. 15, 85–115 (2011)

    MathSciNet  MATH  Google Scholar 

  14. Czichowsky, C., Muhle-Karbe, J., Schachermayer, W.: Transaction costs, shadow prices, and duality in discrete time. SIAM J. Financ. Math. 5, 258–277 (2014)

    MathSciNet  MATH  Google Scholar 

  15. Czichowsky, C., Schachermayer, W., et al.: Duality theory for portfolio optimisation under transaction costs. Ann. Appl. Probab. 26, 1888–1941 (2016)

    MathSciNet  MATH  Google Scholar 

  16. Dai, M., Jin, H., Liu, H.: Illiquidity, position limits, and optimal investment for mutual funds. J. Econ. Theory 146, 1598–1630 (2011)

    MathSciNet  MATH  Google Scholar 

  17. Dai, M., Li, P., Liu, H., Wang, Y.: Portfolio choice with market closure and implications for liquidity premia. Manag. Sci. 62, 368–386 (2016)

    Google Scholar 

  18. Dai, M., Yi, F.: Finite-horizon optimal investment with transaction costs: a parabolic double obstacle problem. J. Differ. Equ. 246, 1445–1469 (2009)

    MathSciNet  MATH  Google Scholar 

  19. Davis, M.H.A., Norman, A.R.: Portfolio selection with transaction costs. Math. Oper. Res. 15, 676–713 (1990)

    MathSciNet  MATH  Google Scholar 

  20. Foster, D., Viswanathan, S.: Strategic trading with asymmetrically informed traders and long-lived information. J. Financ. Quant. Anal. 29, 499–518 (1994)

    Google Scholar 

  21. Foucault, T., Kadan, O., Kandel, E.: Limit order book as a market for liquidity. Rev. Financ. Stud. 18, 1171–1217 (2005)

    Google Scholar 

  22. Gassiat, P., Gozzi, F., Pham, H.: Investment/consumption problem in illiquid markets with regime-switching. SIAM J. Control. Optim. 52, 1761–1786 (2014)

    MathSciNet  MATH  Google Scholar 

  23. Gatheral, J., Schied, A.: Optimal trade execution under geometric brownian motion in the almgren and chriss framework. Int. J. Theor. Appl. Financ. 14, 353–368 (2011)

    MathSciNet  MATH  Google Scholar 

  24. Gerhold, S., Guasoni, P., Muhle-Karbe, J., Schachermayer, W.: Transaction costs, trading volume, and the liquidity premium. Financ. Stoch. 18, 1–37 (2014)

    MathSciNet  MATH  Google Scholar 

  25. Gerhold, S., Muhle-Karbe, J., Schachermayer, W.: Asymptotics and duality for the Davis and Norman problem. Stochastics 84, 625–641 (2012)

    MathSciNet  MATH  Google Scholar 

  26. Guilbaud, F., Pham, H.: Optimal high-frequency trading with limit and market orders. Quant. Financ. 13, 79–94 (2013)

    MathSciNet  MATH  Google Scholar 

  27. Hobson, D., Tse, A.S.L., Zhu, Y.: A multi-asset investment and consumption problem with transaction costs. Financ. Stochast. 23, 641–676 (2019)

    MathSciNet  MATH  Google Scholar 

  28. Hugonnier, J., Kramkov, D.: Optimal investment with random endowments in incomplete markets. Ann. Appl. Probab. 14, 845–864 (2004)

    MathSciNet  MATH  Google Scholar 

  29. Janeček, K., Shreve, S.E.: Asymptotic analysis for optimal investment and consumption with transaction costs. Financ. Stochast. 8, 181–206 (2004)

    MathSciNet  MATH  Google Scholar 

  30. Karatzas, I., Lehoczky, J.P., Shreve, S.E.: Optimal portfolio and consumption decisions for a “small investor’’ on a finite horizon. SIAM J. Control. Optim. 25, 1557–1586 (1987)

    MathSciNet  MATH  Google Scholar 

  31. Karatzas, I., Lehoczky, J.P., Shreve, S.E., Xu, G.L.: Martingale and duality methods for utility maximization in an incomplete market. SIAM J. Control. Optim. 29, 702–730 (1991)

    MathSciNet  MATH  Google Scholar 

  32. Karatzas, I., Shreve, S.: Brownian Motion and Stochastic Calculus. Graduate Texts in Mathematics, Springer, New York (2014)

    MATH  Google Scholar 

  33. Karatzas, I., Žitković, G.: Optimal consumption from investment and random endowment in incomplete semimartingale markets. Ann. Probab. 31, 1821–1858 (2003)

    MathSciNet  MATH  Google Scholar 

  34. Kramkov, D., Schachermayer, W.: The asymptotic elasticity of utility functions and optimal investment in incomplete markets. Ann. Appl. Probab. 9, 904–950 (1999)

    MathSciNet  MATH  Google Scholar 

  35. Krylov, N., Society, A.M.: Lectures on Elliptic and Parabolic Equations in Holder Spaces. Graduate studies in mathematics. American Mathematical Society (1996)

  36. Kühn, C., Stroh, M.: Optimal portfolios of a small investor in a limit order market: a shadow price approach. Math. Financ. Econ. 3, 45–72 (2010)

    MathSciNet  MATH  Google Scholar 

  37. Kyle, A.S.: Continuous auctions and insider trading. Econometrica 53, 1315–1335 (1985)

    MATH  Google Scholar 

  38. Liu, H.: Optimal consumption and investment with transaction costs and multiple risky assets. J. Financ. 59, 289–338 (2004)

    Google Scholar 

  39. Magill, M.J.P., Constantinides, G.M.: Portfolio selection with transactions costs. J. Econ. Theory 13, 245–263 (1976)

    MathSciNet  MATH  Google Scholar 

  40. Matsumoto, K.: Optimal portfolio of low liquid assets with a log-utility function. Financ. Stochast. 10, 121–145 (2006)

    MathSciNet  MATH  Google Scholar 

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

    Google Scholar 

  42. Merton, R.C.: Optimum consumption and portfolio rules in a continuous-time model. J. Econ. Theory 3, 373–413 (1971)

    MathSciNet  MATH  Google Scholar 

  43. Muthuraman, K., Kumar, S.: Multidimensional portfolio optimization with proportional transaction costs. Math. Financ. 16, 301–335 (2006)

    MathSciNet  MATH  Google Scholar 

  44. Pham, H., Tankov, P.: A model of optimal consumption under liquidity risk with random trading times. Math. Financ. 18, 613–627 (2008)

    MathSciNet  MATH  Google Scholar 

  45. Possamaï, D., Soner, H.M., Touzi, N.: Homogenization and asymptotics for small transaction costs: the multidimensional case. Commun. Partial Differ. Equ. 40, 2005–2046 (2015)

    MathSciNet  MATH  Google Scholar 

  46. Predoiu, S., Shaikhet, G., Shreve, S.: Optimal execution in a general one-sided limit-order book. SIAM J. Financ. Math. 2, 183–212 (2011)

    MathSciNet  MATH  Google Scholar 

  47. Robert, E., Robert, F., Jeffrey, R.: Measuring and modeling execution cost and risk. J. Portf. Manag. 38, 14–28 (2012)

    Google Scholar 

  48. Rogers, L.C.G.: The relaxed investor and parameter uncertainty. Finance Stochast. 5, 131–154 (2001)

    MathSciNet  MATH  Google Scholar 

  49. Rogers, L.C.G., Zane, O.: A Simple Model of Liquidity Effects, pp. 161–176. Springer, Berlin (2002)

  50. Roşu, I.: A dynamic model of the limit order book. Rev. Financ. Stud. 22, 4601–4641 (2009)

    Google Scholar 

  51. Shreve, S.E., Soner, H.M.: Optimal investment and consumption with transaction costs. Ann. Appl. Probab. 4, 609–692 (1994)

    MathSciNet  MATH  Google Scholar 

  52. Yakov, A., Mendelson, H., Pederson, L.H.: Liquidity and asset prices. Found. Trends Financ. 1, 1–96 (2005)

    MATH  Google Scholar 

Download references

Funding

This work was supported by the National Research Foundation of Korea (NRF) Grant funded by the Korea government (MSIT) (No. 2019R1A5A1028324, No. 2020R1C1C1A01014142, and No. 2021R1I1A1A01050679).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jin Hyuk Choi.

Ethics declarations

Conflict of interest

The authors have no competing interests to declare that are relevant to the content of this article.

Additional information

Publisher's Note

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

Appendices

Appendix A. Proof of Lemma 3.1

Since the parabolic type PDE (3.2) is not uniformly elliptic, we change variable as \(x=h(z):=\frac{e^z}{1+e^{z}}\) and consider the PDE for v(th(z)).

To handle the nonlinear term in the PDE, we first consider the following map \(\phi \) from \(C_b([0,T]\times {{\mathbb {R}}})\) (equipped with the uniform norm) to itself:

$$\begin{aligned} \phi (f)(t,z):=\int _t^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ K_f(s,Z_s^{(t,z)}) \right] ds, \end{aligned}$$
(A.1)

where for \((s,z)\in [t,T]\times {{\mathbb {R}}}\),

$$\begin{aligned} \begin{aligned} Z_s^{(t,z)}&:=z+ \left( \mu -r-\tfrac{\sigma ^2}{2} \right) (s-t) + \sigma (B_s-B_t),\\ K_f(s,z)&:= (\mu -r)h(z)+r-\tfrac{1}{2}\sigma ^2 h(z)^2 \\&\qquad + \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left( f(s,\zeta )- \ln \left( \tfrac{1+\bar{\epsilon }h(\zeta )}{1+\bar{\epsilon }h(z)} \right) 1_{\{z<\zeta \}} -\ln \left( \tfrac{1-\underline{\epsilon }h(\zeta )}{1-\underline{\epsilon }h(z)} \right) 1_{\{z>\zeta \}} \right) . \end{aligned} \end{aligned}$$
(A.2)

We observe that for \(f,g \in C_b([0,T]\times {{\mathbb {R}}})\),

$$\begin{aligned} \begin{aligned} \Vert \phi (f)-\phi (g) \Vert _\infty&\le \lambda \int _t^T e^{-\lambda (s-t)} \left( \sup _{\zeta \in {{\mathbb {R}}}} \left| f(s,\zeta )-g(s,\zeta ) \right| \right) ds\\&\le \left( 1- e^{-\lambda T} \right) \Vert f-g \Vert _\infty , \end{aligned} \end{aligned}$$
(A.3)

where the first inequality is due to the triangular inequality for supremum. Therefore, the map \(\phi \) in (A.1) is a contraction map and there exists a unique \({\hat{u}}\in C_b([0,T]\times {{\mathbb {R}}})\) such that \(\phi ({\hat{u}})={\hat{u}}\), by the Banach fixed point theorem.

Claim: For all \(\delta \in (0,1)\), \(K_{\hat{u}}\in C^{\frac{\delta }{2},\delta }([0,T]\times {{\mathbb {R}}})\).

(Proof of Claim): We first check that \(K_{{\hat{u}}}\) is bounded. Since \(0<h(z)<1\), we observe that

$$\begin{aligned} |K_{{\hat{u}}}(t,z) | \le |\mu |+2|r| +\tfrac{\sigma ^2}{2}+ \lambda \left( \Vert {\hat{u}} \Vert _\infty + \ln \left( \tfrac{1+\bar{\epsilon }}{1-\underline{\epsilon }} \right) \right) , \end{aligned}$$
(A.4)

i.e., \(\Vert K_{{\hat{u}}}\Vert _\infty <\infty \). Therefore, to prove the claim, it is enough to check that \(K_{{\hat{u}}}\) is uniformly Lipschitz with respect to z variable and \(\tfrac{1}{2}\)-Hölder continuous with respect to t variable. For \(\Delta >0\), we obtain the following inequalities:

$$\begin{aligned} \begin{aligned}&|K_{{\hat{u}}}(t,z+\Delta )- K_{{\hat{u}}}(t,z) | \\&\quad \le (|\mu -r|+\sigma ^2)\Delta + \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left| 1_{\{z+\Delta<\zeta \}}\cdot \ln \left( \tfrac{1+\bar{\epsilon }h(z+\Delta )}{1+\bar{\epsilon }h(z)}\right) \right. \\ {}&\qquad \left. + 1_{\{z<\zeta \le z+\Delta \}} \cdot \ln \left( \tfrac{1+\bar{\epsilon }h(\zeta )}{1+\bar{\epsilon }h(z)} \right) \right| \\&\quad \quad + \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left| 1_{\{z+\Delta >\zeta \}} \cdot \ln \left( \tfrac{1-\underline{\epsilon }h(z+\Delta )}{1-\underline{\epsilon }h(z)}\right) - 1_{\{z\le \zeta < z+\Delta \}} \cdot \ln \left( \tfrac{1-\underline{\epsilon }h(\zeta )}{1-\underline{\epsilon }h(z)} \right) \right| \\&\quad \le \left( |\mu -r| +\sigma ^2 + \lambda \left( \bar{\epsilon }+ \tfrac{\underline{\epsilon }}{1-\underline{\epsilon }} \right) \right) \Delta , \end{aligned} \end{aligned}$$
(A.5)

where we used the mean value theorem and the bounds \(0<h<1\) and \(0<h'<1\). We can treat \(\Delta <0\) by the same way, and conclude that \(K_{\hat{u}}\) is uniformly Lipschitz with respect to z.

For \(\Delta >0\), using (A.5) and the mean value theorem, we observe that

$$\begin{aligned} \begin{aligned}&{{\mathbb {E}}}\left| e^{\lambda \Delta } K_{{\hat{u}}}(s,Z_s^{(t+\Delta ,z)})-K_{{\hat{u}}}(s,Z_s^{(t,z)}) \right| \\&\quad \le \lambda e^{\lambda T} \Vert K_{{\hat{u}}}\Vert _\infty \Delta + {{\mathbb {E}}}\left| K_{{\hat{u}}}(s,Z_s^{(t+\Delta ,z)})-K_{{\hat{u}}}(s,Z_s^{(t,z)}) \right| \\&\quad \le \lambda e^{\lambda T}\Vert K_{{\hat{u}}}\Vert _\infty \Delta + \left( |\mu -r| +\sigma ^2 + \lambda \left( \bar{\epsilon }+ \tfrac{\underline{\epsilon }}{1-\underline{\epsilon }} \right) \right) {{\mathbb {E}}}\left| Z_s^{(t+\Delta ,z)}-Z_s^{(t,z)} \right| \\&\quad \le \lambda e^{\lambda T}\Vert K_{{\hat{u}}}\Vert _\infty \Delta + \left( |\mu -r| +\sigma ^2 + \lambda \left( \bar{\epsilon }+ \tfrac{\underline{\epsilon }}{1-\underline{\epsilon }} \right) \right) \\ {}&\qquad \left( \left| \mu -r-\tfrac{\sigma ^2}{2}\right| \Delta + \sigma \sqrt{\tfrac{2}{\pi }} \, \Delta ^{\frac{1}{2}} \right) , \end{aligned} \end{aligned}$$
(A.6)

where the last inequality is due to

$$\begin{aligned} {{\mathbb {E}}}\left| Z_s^{(t+\Delta ,z)}-Z_s^{(t,z)} \right|\le & {} \left| \mu -r-\tfrac{\sigma ^2}{2}\right| \Delta + \sigma \, {{\mathbb {E}}}\left| B_{t+\Delta }-B_t \right| \\ {}= & {} \left| \mu -r-\tfrac{\sigma ^2}{2}\right| \Delta + \sigma \sqrt{\tfrac{2}{\pi }} \, \Delta ^{\frac{1}{2}}. \end{aligned}$$

Using \({\hat{u}}= \phi ({\hat{u}})\) and triangular inequality, we obtain

$$\begin{aligned}&|K_{{\hat{u}}}(t+\Delta ,z)- K_{{\hat{u}}}(t,z) |\le \lambda \sup _{\zeta \in R} \left| {\hat{u}}(t+\Delta ,\zeta )-{\hat{u}}(t,\zeta ) \right| \\&\quad =\lambda \sup _{\zeta \in {{\mathbb {R}}}} \left| \int _{t+\Delta }^T e^{-\lambda (s-t-\Delta )} {{\mathbb {E}}}\left[ K_{{\hat{u}}}(s,Z_s^{(t+ \Delta ,\zeta )}) \right] ds \right. \\&\qquad \left. -\int _{t}^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ K_{{\hat{u}}}(s,Z_s^{(t,\zeta )}) \right] ds \right| \\&\quad \le \lambda \sup _{\zeta \in {{\mathbb {R}}}} \bigg ( \int _{t+\Delta }^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left| e^{\lambda \Delta } K_{{\hat{u}}}(s,Z_s^{(t+ \Delta ,\zeta )})-K_{{\hat{u}}}(s,Z_s^{(t,\zeta )}) \right| ds \\&\qquad +\int _t^{t+\Delta } e^{-\lambda (s-t)} {{\mathbb {E}}}\left| K_{{\hat{u}}}(s,Z_s^{(t,\zeta )}) \right| ds \bigg ) \le C \Delta ^{\frac{1}{2}}, \end{aligned}$$

where the generic constant C only depends on the market parameters and \(\Vert {\hat{u}}\Vert _{\infty }\) and the last inequality is due to (A.4) and (A.6). We can treat \(\Delta <0\) by the same way, and conclude that \(K_{{\hat{u}}}\) is \(\tfrac{1}{2}\)-Hölder continuous with respect to t variable. (End of the proof of Claim).

Let \(\delta \in (0,1)\) be fixed. Then, the above claim and Theorem 9.2.3 in [35] ensure that there exists a unique function \(u\in C^{1+\frac{\delta }{2},2+\delta }((0,T)\times {{\mathbb {R}}})\) such that it satisfies the following PDE on \((t,z)\in (0,T)\times {{\mathbb {R}}}\).

$$\begin{aligned} \begin{aligned} {\left\{ \begin{array}{ll} 0=u(T,z),\\ 0= u_t + (\mu -r-\frac{\sigma ^2}{2} )u_z+\frac{1}{2} \sigma ^2 u_{zz} - \lambda u + K_{{\hat{u}}} \end{array}\right. } \end{aligned} \end{aligned}$$
(A.7)

Since u admits a unique continuous extension on \([0,T]\times {{\mathbb {R}}}\) (i.e., see chapter 8.5 in [35]), we let \(u\in C^{1+\frac{\delta }{2},2+\delta }([0,T]\times {{\mathbb {R}}})\). By the Feynman-Kac formula (i.e., see Theorem 5.7.6 in [32]), the solution u of the parabolic PDE (A.7) has the stochastic representation \(u=\phi ({\hat{u}})\), where \(\phi \) is defined in (A.1). Since \({\hat{u}}\) is chosen as the unique fixed point of the map \(\phi \), we conclude that \(u={\hat{u}}\).

Our next task is to define \(u(t,\pm \infty )\) for \(t\in [0,T]\). Using \(\lim _{z\rightarrow \infty }h(Z_s^{(t,z)})=1\) and \(\lim _{z\rightarrow -\infty }h(Z_s^{(t,z)})=0\) almost surely, we obtain

$$\begin{aligned} \begin{aligned} {\left\{ \begin{array}{ll} \lim _{z\rightarrow \infty } K_u(s,Z_s^{(t,z)}) = \mu -\tfrac{1}{2}\sigma ^2 + \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left( u(s,\zeta ) -\ln \left( \frac{1-\underline{\epsilon }h(\zeta )}{1-\underline{\epsilon }} \right) \right) \\ \lim _{z\rightarrow -\infty } K_u(s,Z_s^{(t,z)}) = r+ \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left( u(s,\zeta )- \ln \left( 1+\bar{\epsilon }h(\zeta ) \right) \right) \end{array}\right. } a.s. \end{aligned}\nonumber \\ \end{aligned}$$
(A.8)

The above convergence and \(\Vert K_u\Vert _\infty <\infty \) enable us to apply the dominated convergence theorem:

$$\begin{aligned} \begin{aligned} \lim _{z\rightarrow \pm \infty } u(t,z)&=\lim _{z\rightarrow \pm \infty } \phi (u)(t,z)=\lim _{z\rightarrow \pm \infty } \int _t^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ K_u(s,Z_s^{(t,z)}) \right] ds\\&= {\left\{ \begin{array}{ll} \int _t^T e^{-\lambda (s-t)} \left( \mu -\tfrac{1}{2}\sigma ^2 \right. \\ \quad \left. + \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left( u(s,\zeta ) -\ln \left( \frac{1-\underline{\epsilon }h(\zeta )}{1-\underline{\epsilon }} \right) \right) \right) ds, &{}\hbox { for}\ z\rightarrow \infty \\ \int _t^T e^{-\lambda (s-t)} \left( r+ \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left( u(s,\zeta ) \right. \right. \\ \quad \left. \left. - \ln \left( 1+\bar{\epsilon }h(\zeta ) \right) \right) \right) ds, &{}\hbox { for}\ z\rightarrow -\infty \ \end{array}\right. }. \end{aligned} \end{aligned}$$

Therefore, we can continuously extend u to \(z=\pm \infty \) and \(u(t,\infty )\) and \(u(t,-\infty )\) are defined by the above limit. We observe that for \(z\in \{ \infty , -\infty \}\), u(tz) satisfies

$$\begin{aligned} \begin{aligned} 0&=u_t(t,z) + (\mu -r)h(z)+r-\tfrac{1}{2}\sigma ^2 h(z)^2 - \lambda u(t,z) \\&\qquad + \lambda \sup _{\zeta \in {{\mathbb {R}}}} \left( u(t,\zeta )- \ln \left( \tfrac{1+\bar{\epsilon }h(\zeta )}{1+\bar{\epsilon }h(z)} \right) 1_{\{z<\zeta \}} -\ln \left( \tfrac{1-\underline{\epsilon }h(\zeta )}{1-\underline{\epsilon }h(z)} \right) 1_{\{z>\zeta \}} \right) , \end{aligned} \end{aligned}$$
(A.9)

where the function h is continuously extended as \(h(\infty ):=1\) and \(h(-\infty ):=0\).

Now we define v as \(v(t,x):=u(t,h^{-1}(x))\) for \((t,x)\in [0,T]\times [0,1]\). Such v is well-defined because \(h:{{\mathbb {R}}}\cup \{ \infty , -\infty \} \rightarrow [0,1]\) is bijective. We observe that for \((t,x)\in (0,T)\times (0,1)\) and \(z=h^{-1}(x)\),

$$\begin{aligned} \begin{aligned} v_t(t,x)&=u_t(t,z),\\ x(1-x)v_x(t,x)&=u_z(t,z),\\ x^2(1-x)^2 v_{xx}(t,x)&=u_{zz}(t,z)-(1-2x)u_z(t,z). \end{aligned} \end{aligned}$$
(A.10)

The PDE for u in (A.7) with \({\hat{u}}\) replaced by u and the equalities in (A.10) produce the PDE for v, which is (3.2). Therefore, statement (i) is valid.

To check statement (ii), we observe that \(v(t,0)=u(t,-\infty )\) and \(v(t,1)=u(t,\infty )\). Then, the continuous differentiability of v(t, 0) and v(t, 1) with respect to t is followed by that of \(u(t,-\infty )\) and \(u(t,\infty )\), and (A.9) produces (3.3).

Finally, statement (iii) is a direct consequence of (A.10) and \(u\in C^{1+\frac{\delta }{2},2+\delta }([0,T]\times {{\mathbb {R}}})\).

Appendix B. Proof of Lemma 5.3

(i) When \(\epsilon =0\), Lemma 4.1 and (5.1) produce (5.2).

To prove (5.3), we first check that \(Y_s^{(t,x)}\) in (4.9) safisfies

$$\begin{aligned} dY_s^{(t,x)}=Y_s^{(t,x)} (1-Y_s^{(t,x)}) \left( (\mu -r-\sigma ^2 Y_s^{(t,x)})ds + \sigma dB_s \right) . \end{aligned}$$
(B.1)

Then application of Ito’s formula produces that for \((s,x)\in [t,T) \times (0,1)\),

$$\begin{aligned} \left( Y_s^{(t,x)}-x\right) ^2&= \int _t^s Y_u^{(t,x)} (1-Y_u^{(t,x)})\left( 2(Y_u^{(t,x)}-x )( \mu -r-\sigma ^2 Y_u^{(t,x)}) \right. \\ {}&\qquad \left. + \sigma ^2 Y_u^{(t,x)} (1-Y_u^{(t,x)}) \right) du \\&\qquad + \int _t^s 2\sigma (Y_u^{(t,x)}-x ) Y_u^{(t,x)} (1-Y_u^{(t,x)}) dB_u. \end{aligned}$$

Since \(0<Y_s^{(t,x)}<1\), the local martingale part above is a true martingale and we obtain

$$\begin{aligned}&\tfrac{\partial }{\partial s} \left( {{\mathbb {E}}}\left[ \left( Y_s^{(t,x)}-x\right) ^2 \right] \right) \nonumber \\ {}&\quad ={{\mathbb {E}}}\left[ Y_s^{(t,x)} (1-Y_s^{(t,x)})\left( 2(Y_s^{(t,x)}-x )( \mu -r-\sigma ^2 Y_s^{(t,x)})\right. \right. \nonumber \\ {}&\quad \left. \left. + \sigma ^2 Y_s^{(t,x)} (1-Y_s^{(t,x)}) \right) \right] =-x^2(1-x)^2 \nonumber \\ {}&\quad \cdot {{\mathbb {E}}}\left[ \tfrac{\partial }{\partial x}\left( \tfrac{Y_s^{(t,x)} \left( 1-Y_s^{(t,x)}\right) \left( \mu -r-\sigma ^2 Y_s^{(t,x)}\right) }{x(1-x)} \right) \right] , \end{aligned}$$
(B.2)

where the second equality is from elementary computations using the definition of \(Y_s^{(t,x)}\) in (4.9).

For \(x\in (0,1)\), we observe that

$$\begin{aligned} \tfrac{Y_s^{(t,x)}-x}{x(1-x)} = \tfrac{A^{(t,s)}-1}{A^{(t,s)} x+ (1-x)} \quad \text {with} \quad A^{(t,s)}:=e^{(\mu -r-\frac{\sigma ^2}{2})(s-t)+\sigma (B_s-B_t)}, \end{aligned}$$
(B.3)

and the above expression is decreasing in x, therefore,

$$\begin{aligned} 1- \tfrac{1}{A^{(t,s)}}<\tfrac{Y_s^{(t,x)}-x}{x(1-x)}<A^{(t,s)}-1 \quad \text {for} \quad 0<x<1. \end{aligned}$$
(B.4)

When \(\epsilon =0\), the representation of \(v_x\) in (4.11) becomes

$$\begin{aligned} v_x^0(t,x)=\int _t^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ \tfrac{Y_s^{(t,x)}\left( 1-Y_s^{(t,x)}\right) \left( \mu -r-\sigma ^2 Y_s^{(t,x)} \right) }{x(1-x)} \right] ds, \end{aligned}$$
(B.5)

We take derivative with respect to x in the above expression. Then, the mean value theorem and the dominated convergence theorem, together with the inequalities (4.14) and (B.4), allow us to take derivative inside of the expectation and obtain that for \((t,x)\in [0,T)\times (0,1)\),

$$\begin{aligned} v_{xx}^0(t,x)&=\int _t^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ \tfrac{\partial }{\partial x}\left( \tfrac{Y_s^{(t,x)} \left( 1-Y_s^{(t,x)}\right) \left( \mu -r-\sigma ^2 Y_s^{(t,x)}\right) }{x(1-x)} \right) \right] ds \end{aligned}$$
(B.6)
$$\begin{aligned}&=- \int _t^T e^{-\lambda (s-t)} \tfrac{\partial }{\partial s} \left( {{\mathbb {E}}}\left[ \left( \tfrac{Y_s^{(t,x)}-x}{x(1-x)}\right) ^2 \right] \right) \, ds \nonumber \\&=- e^{-\lambda (T-t)} \, {{\mathbb {E}}}\left[ \left( \tfrac{Y_T^{(t,x)}-x}{x(1-x)}\right) ^2 \right] - \lambda \int _t^T e^{-\lambda (s-t)} \, {{\mathbb {E}}}\left[ \left( \tfrac{Y_s^{(t,x)}-x}{x(1-x)}\right) ^2 \right] \, ds, \end{aligned}$$
(B.7)

where the second equality is due to (B.2), and the third equality is from integration by parts. Obviously (B.7) implies that \(v_{xx}^0(t,x)<0\) for \((t,x)\in [0,T)\times (0,1)\).

To conclude (5.3), it only remains to check that \(\lim _{x\uparrow 1} v_{xx}^0(t,x)<0\) and \(\lim _{x\downarrow 0} v_{xx}^0(t,x)<0\). Indeed, (B.3) and (B.4) enable us to apply the dominated convergence theorem to (B.7) and obtain

$$\begin{aligned} \lim _{x\uparrow 1} v_{xx}^0(t,x)&= - e^{-\lambda (T-t)} \, {{\mathbb {E}}}\Big [ \left( 1- \tfrac{1}{A^{(t,T)}}\right) ^2 \Big ] - \lambda \int _t^T e^{-\lambda (s-t)} \, {{\mathbb {E}}}\Big [ \left( 1- \tfrac{1}{A^{(t,s)}}\right) ^2 \Big ] ds ,\\ \lim _{x\downarrow 0} v_{xx}^0(t,x)&= - e^{-\lambda (T-t)} \, {{\mathbb {E}}}\Big [ \left( A^{(t,T)}-1\right) ^2 \Big ] \\ {}&\quad - \lambda \int _t^T e^{-\lambda (s-t)} \, {{\mathbb {E}}}\Big [ \left( A^{(t,s)}-1\right) ^2 \Big ] ds, \end{aligned}$$

and we conclude that \(\lim _{x\uparrow 1} v_{xx}^0(t,x)<0\) and \(\lim _{x\downarrow 0} v_{xx}^0(t,x)<0\).

(ii) The SDE for \(Y_s^{(t,x)}\) in (B.1) and \(0<Y_s^{(t,x)}<1\) imply that for \((s,x)\in [t,T) \times (0,1)\),

$$\begin{aligned} \tfrac{\partial }{\partial s} \left( {{\mathbb {E}}}\left[ Y_s^{(t,x)} \right] \right) = {{\mathbb {E}}}\left[ Y_s^{(t,x)} \left( 1- Y_s^{(t,x)}\right) \left( \mu - r - \sigma ^2 Y_s^{(t,x)} \right) \right] \end{aligned}$$
(B.8)

We divide both sides of (B.8) by \(x(1-x)\) and take derivative with respect to x. Then, we can put the derivative inside of the expectation as in the proof of part (i), and obtain

$$\begin{aligned} {{\mathbb {E}}}\left[ \tfrac{\partial }{\partial x} \left( \tfrac{Y_s^{(t,x)} \left( 1- Y_s^{(t,x)}\right) \left( \mu - r - \sigma ^2 Y_s^{(t,x)} \right) }{x(1-x)} \right) \right]&=\tfrac{\partial }{\partial s} \left( {{\mathbb {E}}}\left[ \tfrac{\partial }{\partial x} \left( \tfrac{Y_s^{(t,x)} }{x(1-x)}\right) \right] \right) \\&=\tfrac{\partial }{\partial s} \left( \tfrac{ {{\mathbb {E}}}\left[ \frac{\partial }{\partial x}Y_s^{(t,x)} \right] }{x(1-x)} - \tfrac{(1-2x){{\mathbb {E}}}\left[ Y_s^{(t,x)} \right] }{x^2(1-x)^2} \right) . \end{aligned}$$

We rearrange the above equation and use (B.8), (B.5), and (B.7) to obtain

$$\begin{aligned} \int _t^T e^{-\lambda (s-t)} \tfrac{\partial }{\partial s} \left( {{\mathbb {E}}}\left[ \tfrac{\partial }{\partial x} Y_s^{(t,x)} \right] \right) ds = x(1-x)v_{xx}^0(t,x) + (1-2x) v_x^0(t,x). \end{aligned}$$
(B.9)

Now we conclude that \(F(t, {\hat{y}}^0(t))<1\) by the following way:

$$\begin{aligned} F(t, {\hat{y}}^0(t))&\le \lambda \int _t^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ \tfrac{\partial }{\partial x} Y_s^{(t,x)} \right] ds \, \Big |_{x={\hat{y}}^0(t)} \\&=\left( 1- e^{-\lambda (T-t)} {{\mathbb {E}}}\left[ \tfrac{\partial }{\partial x} Y_T^{(t,x)} \right] + x(1-x)v_{xx}^0(t,x) \right. \\ {}&\quad \left. + (1-2x) v_x^0(t,x) \right) \Big |_{x={\hat{y}}^0(t)} <1, \end{aligned}$$

where the first inequality is from the definition of F in (5.4) and the positivity of \(\frac{\partial }{\partial x} Y_s^{(t,x)}\) (see (4.14)), and the equality is due to integration by parts and (B.9), and the last inequality is due to the positivity of \(\frac{\partial }{\partial x} Y_T^{(t,x)}\), \(v_x^0(t,{\hat{y}}^0(t))=0\), and \(v_{xx}^0(t,\hat{y}^0(t))<0\).

We can check that \(F(t, {\hat{y}}^0(t))>-1\) by the same way as above. (iii) The expression in (4.11) produces

$$\begin{aligned} \begin{aligned} v_{x}^\epsilon (t,x_\epsilon )-v_x^0(t,x_\epsilon )&= \lambda \int _t^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ \left( \tfrac{\partial }{\partial x} Y_s^{(t,x_\epsilon )} \right) L_y^\epsilon (s, Y_s^{(t,x_\epsilon )} ) \right] ds. \end{aligned} \end{aligned}$$
(B.10)

In the above expression, when we take limit as \(\epsilon \downarrow 0\), the inequalities (4.13) and (4.14) enable us to use the dominated convergence theorem to conclude that

$$\begin{aligned} \lim _{\epsilon \downarrow 0} \left( v_{x}^\epsilon (t,x_\epsilon )-v_x^0(t,x_\epsilon ) \right) =0. \end{aligned}$$

The above limit and the continuity of \(v_{x}^0\) imply (5.6).

To prove (5.7), we first observe that for \((t,x)\in [0,T)\times (0,1)\) and \((s,z)\in (t,T)\times (0,1)\), the density function for \(Y_s^{(t,x)}\) is given by

$$\begin{aligned} \begin{aligned} \varphi (s,z;t,x)&:= \tfrac{\partial }{\partial z} {{\mathbb {P}}}\left( Y_s^{(t,x)} \le z \right) =\tfrac{\exp \left( -\frac{1}{2\sigma ^2 (s-t)} \left( (r-\mu +\frac{\sigma ^2}{2})(s-t) + \ln \left( \frac{z(1-x)}{(1-z)x} \right) \right) ^2 \right) }{\sigma z(1-z)\sqrt{2\pi (s-t)}}. \end{aligned}\nonumber \\ \end{aligned}$$
(B.11)

Then, the expression in (4.11) and \(\frac{\partial }{\partial x} Y_s^{(t,x)}=\frac{Y_s^{(t,x)}(1-Y_s^{(t,x)})}{x(1-x)}\) imply that

$$\begin{aligned} \begin{aligned} v_x^\epsilon (t,x)-v_x^0(t,x)&= \lambda \int _t^T e^{-\lambda (s-t)} {{\mathbb {E}}}\left[ \tfrac{Y_s^{(t,x)}\left( 1-Y_s^{(t,x)}\right) }{x(1-x)} L_y^\epsilon (s, Y_s^{(t,x)} ) \right] ds\\&=\lambda \int _t^T e^{-\lambda (s-t)} \left( \int _0^1 \tfrac{z(1-z)}{x(1-x)} L_y^\epsilon (s, z) \varphi (s,z;t,x) dz \right) ds. \end{aligned} \end{aligned}$$
(B.12)

For \(x\in (0,1)\), direct computations produce

$$\begin{aligned} \begin{aligned} \tfrac{\partial }{\partial x} \left( \tfrac{z(1-z)}{x(1-x)} \varphi (s,z;t,x) \right) = \tfrac{(1-z)z \left( r-\mu + (2x-\frac{1}{2})\sigma ^2 + \frac{1}{s-t} \ln \left( \frac{z(1-x)}{(1-z)x} \right) \right) \varphi (s,z;t,x) }{(1-x)^2 x^2 \sigma ^2}. \end{aligned} \end{aligned}$$
(B.13)

Assumption  implies that

$$\begin{aligned} \left| r-\mu + \tfrac{\sigma ^2}{2} \right| \le \sigma ^2, \quad \left| r-\mu + (2x-\tfrac{1}{2})\sigma ^2 \right| \le 2\sigma ^2 \quad \text {for} \quad x\in (0,1). \end{aligned}$$
(B.14)

Then, (B.11) and (B.13) produce the following:

$$\begin{aligned} \begin{aligned}&\int _0^1\left| \frac{\partial }{\partial x} \left( \tfrac{z(1-z)}{x(1-x)} \varphi (s,z;t,x) \right) \right| \, dz\\&\quad \le \int _0^1\tfrac{\left( 2\sigma ^2 + \frac{1}{s-t} \left| \ln \left( \tfrac{z(1-x)}{(1-z)x} \right) \right| \right) }{\sigma ^3(1-x)^2 x^2 \sqrt{2\pi (s-t)}}\exp \left( -\tfrac{\left( \ln \left( \frac{z(1-x)}{(1-z)x} \right) \right) ^2}{2\sigma ^2 (s-t)} + \left| \ln \left( \tfrac{z(1-x)}{(1-z)x} \right) \right| \right) \, dz \\&\quad = \int _{-\infty }^\infty \tfrac{\left( 2\sigma ^2 + \frac{1}{s-t} \left| \zeta \right| \right) e^{ -\frac{\zeta ^2}{2\sigma ^2 (s-t)} + \left| \zeta \right| }}{\sigma ^3(1-x)^2 x^2 \sqrt{2\pi (s-t)}} \cdot \tfrac{\frac{x}{1-x}e^\zeta }{(1+\frac{x}{1-x} e^\zeta )^2} \, d\zeta \\&\quad \le \int _{-\infty }^\infty \tfrac{\left( 2\sigma ^2 + \frac{1}{s-t} \left| \zeta \right| \right) e^{ -\frac{\zeta ^2}{4\sigma ^2 (s-t)} + 4\sigma ^2(s-t)}}{\sigma ^3(1-x)^3 x \sqrt{2\pi (s-t)}} \, d\zeta \\&\quad =\tfrac{2\sqrt{2} }{(1-x)^3 x}\left( 1 + \tfrac{1}{\sigma \sqrt{\pi (s-t)}} \right) e^{4\sigma ^2(s-t)}, \end{aligned} \end{aligned}$$
(B.15)

where the first inequality is due to (B.14), and the first equality is obtained by change of variables as \(\zeta =\ln \left( \frac{z(1-x)}{(1-z)x} \right) \). The second inequality is due to the inequality of arithmetic and geometric means, and the second equality is obtained by direct computations.

By (B.15), we conclude that

$$\begin{aligned} \int _t^T e^{-\lambda (s-t) }\int _0^1\left| \tfrac{\partial }{\partial x} \left( \tfrac{z(1-z)}{x(1-x)} \varphi (s,z;t,x) \right) \right| \, dz \, ds <\infty , \end{aligned}$$
(B.16)

and the function \(H:[0,T)\times (0,1) \rightarrow {{\mathbb {R}}}\) given by

$$\begin{aligned} H(t,x):=\lambda \int _t^T e^{-\lambda (s-t) }\int _0^1 \tfrac{\partial }{\partial x} \left( \tfrac{z(1-z)}{x(1-x)} \varphi (s,z;t,x) \right) L_y^\epsilon (s,z) \, dz \, ds \end{aligned}$$
(B.17)

is well-defined due to (B.16) and the boundedness \(| L_y^\epsilon |\le \frac{\epsilon }{1-\epsilon }\) in (4.13). Then, (B.15) implies that

$$\begin{aligned} \left| H(t,x)\right| \le \tfrac{2\sqrt{2}\lambda e^{4\sigma ^2 T}}{(1-x)^3 x} \left( T + \tfrac{2\sqrt{T}}{\sigma \sqrt{\pi }} \right) \cdot \tfrac{\epsilon }{1-\epsilon } \quad \text {for} \quad (t,x)\in [0,T)\times (0,1). \end{aligned}$$
(B.18)

Now, let’s check that

$$\begin{aligned} H(t,x)=v_{xx}^\epsilon (t,x)-v_{xx}^0(t,x). \end{aligned}$$
(B.19)

Indeed, for \((t,x)\in [0,T)\times (0,1)\),

$$\begin{aligned} H(t,x)&=\lim _{\delta \rightarrow 0} \frac{1}{\delta }\int _x^{x+\delta }H(t,\eta )d\eta \\&=\lim _{\delta \rightarrow 0} \frac{1}{\delta } \,\, \lambda \int _t^T e^{-\lambda (s-t)} \int _0^1 \left( \tfrac{z(1-z)\varphi (s,z;t,x+\delta ) }{(x+\delta )(1-x-\delta )} - \tfrac{z(1-z)\varphi (s,z;t,x) }{x(1-x)} \right) \\ {}&\quad L_y^\epsilon (s,z) \, dz \, ds \\&=\lim _{\delta \rightarrow 0} \frac{1}{\delta } \left( \left( v_x^\epsilon (t,x+\delta )-v_x^0(t,x+\delta )\right) -\left( v_x^\epsilon (t,x)-v_x^0(t,x)\right) \right) \\&=v_{xx}^\epsilon (t,x)-v_{xx}^0(t,x), \end{aligned}$$

where the second equality is due to Fubini’s theorem and the fundamental theorem of calculus, and the third equality is from (B.12).

Finally, we conclude (5.7) by the following observation:

$$\begin{aligned}&\limsup _{\epsilon \downarrow 0} \left| v_{xx}^\epsilon (t,x_\epsilon )-v_{xx}^0(t,x_0)\right| \\ {}&\quad \le \limsup _{\epsilon \downarrow 0} \left| v_{xx}^\epsilon (t,x_\epsilon )-v_{xx}^0(t,x_\epsilon )\right| + \limsup _{\epsilon \downarrow 0} \left| v_{xx}^0(t,x_\epsilon )-v_{xx}^0(t,x_0)\right| \\&\quad \le \limsup _{\epsilon \downarrow 0}\tfrac{2\sqrt{2}\lambda e^{4\sigma ^2 T}}{(1-x_\epsilon )^3 x_\epsilon } \left( T + \tfrac{2\sqrt{T}}{\sigma \sqrt{\pi }} \right) \cdot \tfrac{\epsilon }{1-\epsilon } =0, \end{aligned}$$

where the second inequality is due to (B.18), (B.19), and the continuity of \(v_{xx}^0\).

Appendix C. Proof of Lemma 5.10

(i) Using \(\Gamma \) in (5.25), the equations (B.5) and (B.6) can be written as

$$\begin{aligned} v_x^0(t,x)=\int _0^{T-t} e^{-\lambda s} \Gamma (s,x)ds, \quad v_{xx}^0(t,x)=\int _0^{T-t} e^{-\lambda s} \Gamma _x(s,x)ds. \end{aligned}$$
(C.1)

We can do the similar argument to obtain representations for \(v_{xxx}^{0}\) and partial derivatives of \(v_x^{0}\) and \(v_{xx}^{0}\) with respect to \(\lambda \), with the observation that \(\Gamma , \Gamma _x, \Gamma _{xx}\) are continuous & bounded maps on \([0,T]\times (0,1)\). The result is summarized as follows:

$$\begin{aligned} \begin{aligned}&\tfrac{\partial }{\partial \lambda } v_x^{0}(t,x)=- \int _0^{T-t} e^{-\lambda s} s \Gamma (s,x)ds, \quad \tfrac{\partial }{\partial \lambda } v_{xx}^{0}(t,x)=-\int _0^{T-t} e^{-\lambda s}s \Gamma _x(s,x)ds \\&v_{xx}^{0}(t,x)=\int _0^{T-t} e^{-\lambda s} \Gamma _x(s,x)ds,\qquad \quad v_{xxx}^{0}(t,x)=\int _0^{T-t} e^{-\lambda s} \Gamma _{xx}(s,x)ds. \end{aligned} \end{aligned}$$
(C.2)

Using \(Y_t^{(t,x)}=x\), direct computations produce

$$\begin{aligned} \begin{aligned}&\Gamma (0,x)=\mu -r-\sigma ^2 x, \quad \Gamma _x(0,x) = - \sigma ^2, \quad \Gamma _{xx}(0,x)=0. \end{aligned} \end{aligned}$$
(C.3)

With (C.3) and (5.24), we apply Lemma D.2 to (C.2) and conclude (5.61).

(ii) The mean value theorem and (5.2) produce

$$\begin{aligned} \begin{aligned}&{\hat{y}}^{0,\lambda +\delta }(t) - {\hat{y}}^{0,\lambda }(t) \\ {}&\quad = - \tfrac{ v_x^{0,\lambda +\delta }(t,{\hat{y}}^{0,\lambda }(t)) - v_x^{0,\lambda }(t,\hat{y}^{0,\lambda }(t))}{v_{xx}^{0,\lambda +\delta }(t, z(\lambda ,\delta ))} \quad \text {for} z(\lambda ,\delta ) \text {between} \hat{y}^{0,\lambda }(t) \text {and} {\hat{y}}^{0,\lambda +\delta }(t), \end{aligned} \end{aligned}$$

where we specify the dependence on \(\lambda \) for clarity. Since \(\frac{\partial }{\partial \lambda } v_x^{0}\) exists (see (C.2)), the above equality and (5.3) ensure that \({\hat{y}}^{0,\lambda }(t)\) is differentiable with respect to \(\lambda \) and

$$\begin{aligned} \begin{aligned} \tfrac{\partial }{\partial \lambda } {\hat{y}}^{0}(t) = - \frac{ \frac{\partial }{\partial \lambda } v_x^{0}(t,x)\big |_{x=\hat{y}^{0}(t)} }{ v_{xx}^{0}(t,{\hat{y}}^{0}(t))}. \end{aligned} \end{aligned}$$
(C.4)

Observe that the bounds for \(\frac{\partial }{\partial x}Y_t^{(0,x)}\) and \(\frac{Y_t^{(0,x)}-x}{x(1-x)}\) in (4.14) and (B.4) do not depend on the variable x. Therefore, the following convergence is uniform on \(x \in (0,1)\):

$$\begin{aligned} \Gamma _x(t,x)&=-{{\mathbb {E}}}\left[ \sigma ^2 \left( \tfrac{\partial }{\partial x}Y_t^{(0,x)}\right) ^2 \right. \\ {}&\quad \left. + 2\left( \mu -r-\sigma ^2 Y_t^{(0,x)} \right) \left( \tfrac{Y_t^{(0,x)}-x}{x(1-x)}\right) \tfrac{\partial }{\partial x}Y_t^{(0,x)} \right] \xrightarrow [ t \downarrow 0]{} -\sigma ^2. \end{aligned}$$

Hence, there is a constant \({\tilde{T}}\in (0,T)\) such that \(\Gamma _x(t,x)\le -\frac{\sigma ^2}{2}\) for \((t,x)\in [0,{\tilde{T}}] \times (0,1)\). This observation and the expression of \(v_{xx}\) in (C.2) imply

$$\begin{aligned} \lambda v_{xx}(t,\hat{y}^0(t)) \le -\tfrac{\sigma ^2}{2} \left( 1-e^{-\lambda (T-t)} \right) \quad \text {for} \quad (t,x)\in [T-{\tilde{T}}, T) \times (0,1). \end{aligned}$$
(C.5)

We obtain \(\Vert \Gamma _{xt} \Vert _\infty <\infty \) by using Ito’s formula with (B.1) and the bounds (4.14) and (B.4). Then, (C.2) and (C.3) imply

$$\begin{aligned} \left| \lambda v_{xx}(t,\hat{y}^0 (t)) + \sigma ^2 \right|&= \left| \lambda \int _0^{T-t} e^{-\lambda s} s \, \left( \tfrac{\Gamma _x(s,\hat{y}^0(t)) -\Gamma _x(0,\hat{y}^0(t))}{s} \right) ds + \sigma ^2 e^{-\lambda (T-t)} \right| \\&\le \Vert \Gamma _{xt} \Vert _\infty \left( \tfrac{1-e^{-\lambda (T-t)}}{\lambda } -(T-t) e^{-\lambda (T-t)} \right) + \sigma ^2 e^{-\lambda (T-t)}. \end{aligned}$$

This implies that there exists a constant \({\tilde{\Lambda }}\) (may depend on \({\tilde{T}}\)) such that

$$\begin{aligned} \lambda v_{xx}(t,\hat{y}^0 (t)) \le -\tfrac{\sigma ^2}{2} \quad \text {for} \quad (t,\lambda )\in [0,T-{\tilde{T}}] \times [\tilde{\Lambda },\infty ). \end{aligned}$$
(C.6)

Using (C.2) and (C.3), we obtain

$$\begin{aligned}&\left| \lambda ^3 \tfrac{\partial }{\partial \lambda } v_x^{0}(t,x)\big |_{x={\hat{y}}^{0}(t)} \right| \nonumber \\&\quad \le \left| \lambda ^3 \int _0^{T-t} e^{-\lambda s} s^2 \left( \tfrac{\Gamma (s, \hat{y}^0(t))-\Gamma (0, \hat{y}^0(t))}{s}\right) ds \right| \nonumber \\ {}&\qquad + \left| \sigma ^2(y_\infty - \hat{y}^0(t))\lambda ^3 \int _0^{T-t} e^{-\lambda s} s \,ds \right| \nonumber \\&\quad \le C\left( 1-e^{-\lambda (T-t)} + \lambda (T-t)\big (1+\lambda (T-t) \big ) e^{-\lambda (T-t)} \right) \nonumber \\ {}&\quad \text {for} \quad (t,\lambda )\in [0,T]\times [1,\infty ), \end{aligned}$$
(C.7)

where the second inequality is due to \(\Vert \Gamma _t \Vert _\infty <\infty \) and (5.24).

From (C.4), we obtain the boundedness of \(\left| \lambda ^2 \tfrac{\partial }{\partial \lambda } {\hat{y}}^{0}(t) \right| \):

$$\begin{aligned}&\sup _{(t,\lambda )\in [0,T-{\tilde{T}}]\times [{\tilde{\Lambda }}, \infty )} \left| \lambda ^2 \tfrac{\partial }{\partial \lambda } {\hat{y}}^{0}(t) \right|<\infty \quad \text {due to} (C.6), (C.7), \sup _{x>0} x(1+x) e^{-x}<\infty , \\&\sup _{(t,\lambda )\in [0,T-{\tilde{T}}]\times [1, {\tilde{\Lambda }}]} \left| \lambda ^2 \tfrac{\partial }{\partial \lambda } {\hat{y}}^{0}(t) \right|<\infty \quad \text {due to the continuity on compact set},\\&\sup _{(t,\lambda )\in [T-{\tilde{T}},T)\times [1,\infty )} \left| \lambda ^2 \tfrac{\partial }{\partial \lambda } {\hat{y}}^{0}(t) \right|<\infty \quad \text {due to} (C.5), (C.7), \sup _{x>0}\tfrac{ x(1+x) e^{-x}}{1-e^{-x}}<\infty . \end{aligned}$$

Therefore, we conclude (5.62).

(iii) The bounds (4.14) and (5.62) enable us to use Leibniz integral rule to obtain

$$\begin{aligned} \tfrac{\partial }{\partial \lambda } {{\mathbb {E}}}\left[ \left| Y_{s}^{(t, \hat{y}^{0}(t))} - \hat{y}^{0}(s) \right| \right]&= {{\mathbb {E}}}\left[ \left( \tfrac{\partial }{\partial \lambda }\hat{y}^0(t) \cdot \tfrac{\partial }{\partial x} Y_s^{(t,x)} \Big |_{x=\hat{y}^0(t)}- \tfrac{\partial }{\partial \lambda }\hat{y}^0(s) \right) \cdot \right. \\ {}&\quad \left. {{\,\textrm{sgn}\,}}\left( Y_{s}^{(t, \hat{y}^{0}(t))} - \hat{y}^{0}(s) \right) \right] . \end{aligned}$$

The above expression, together with the bounds (4.14) and (5.62), implies (5.63).

Appendix D. Supplementary Lemmas

Lemma D.1

Let \(F:[0,T]\times [0,1]^2\rightarrow {{\mathbb {R}}}\) be a continuous function. We define \(f:[0,T]\times [0,1]\rightarrow [0,1]\) as

$$\begin{aligned} f(t,x):=\max \left\{ z: z\in \mathop {\textrm{argmax}}\limits _{y\in [0,1]} F(t,x,y) \right\} , \end{aligned}$$
(D.1)

then f is upper semicontinuous (which is obviously Borel-measurable).

Proof

This type of result is well-known (e.g., see p. 153 in [6]), but we give a short proof here for the sake of self-containedness.

Since F is continuous, \( \mathop {\textrm{argmax}}\limits _{y\in [0,1]} F(t,x,y) \) is a nonempty closed subset of [0, 1], so the maximum element of \( \mathop {\textrm{argmax}}\limits _{y\in [0,1]} F(t,x,y) \) exists and f is well-defined. Let \(\{(t_n,x_n)\}_{n\in {{\mathbb {N}}}}\subset [0,T]\times [0,1]\) be a sequence converging to \((t_\infty ,x_\infty )\) such that \(\lim _{n\rightarrow \infty } f(t_n,x_n)\) exists. Then, by definition of f,

$$\begin{aligned} F(t_n,x_n,f(t_\infty ,x_\infty ))\le F(t_n,x_n,f(t_n,x_n)). \end{aligned}$$
(D.2)

We let \(n\rightarrow \infty \) above and using the continuity of F to obtain

$$\begin{aligned} F(t_\infty ,x_\infty ,f(t_\infty ,x_\infty )) \le F(t_\infty ,x_\infty ,\lim _{n\rightarrow \infty } f(t_n,x_n)). \end{aligned}$$
(D.3)

This implies that \(\lim _{n\rightarrow \infty } f(t_n,x_n) \in \mathop {\textrm{argmax}}\limits _{y\in [0,1]} F(t_\infty ,x_\infty ,y)\), and the definition of f ensures

$$\begin{aligned} f(t_\infty ,x_\infty )\ge \lim _{n\rightarrow \infty } f(t_n,x_n). \end{aligned}$$

Therefore, f is upper semicontinuous. \(\square \)

Lemma D.2

Let \(f(s,x): [0,t] \times (0,1) \rightarrow {{\mathbb {R}}}\) be a continuous function, and \(g(\lambda ): [1,\infty ) \rightarrow (0,1)\) be a function satisfying \(\lim _{\lambda \rightarrow \infty } g(\lambda )=x_\infty \in (0,1)\). Then, for \(\alpha \in \{0,1,2,3,4\}\) and \(t > 0\),

$$\begin{aligned} \begin{aligned}&\lim _{\lambda \rightarrow \infty } \lambda ^{\frac{\alpha }{2}+1} \int _0^{t} e^{-\lambda s}s^{\frac{\alpha }{2}} f(s,g(\lambda )) ds = c_\alpha \cdot f(0,x_\infty ), \end{aligned} \end{aligned}$$
(D.4)

where \(c_0=1\), \(c_1= \frac{\sqrt{\pi }}{2}\), \(c_2= 1\), \(c_3= \frac{3\sqrt{\pi }}{4}\), \(c_4= 2\). Also, there exists a constant C such that

$$\begin{aligned} \lambda ^{\frac{\alpha }{2}+1} \int _0^t e^{-\lambda s} s^{\frac{\alpha }{2}} ds \le C, \quad \text {for} \quad (t,\lambda ,\alpha )\in [0,\infty )\times [1,\infty )\times \{0,1,2,3,4\}.\nonumber \\ \end{aligned}$$
(D.5)

Proof

Let \(\eta >0\) be a given constant. The uniform continuity of f on a compact set containing the point \((0,x_\infty )\), together with \(\lim _{\lambda \rightarrow \infty } g(\lambda )=x_\infty \in (0,1)\), implies that there exists \(\delta >0\) such that

$$\begin{aligned} \left| f(s,g(\lambda )) - f(0,x_\infty ) \right| \le \eta \quad \text {for any} \quad (s,\lambda ) \in [0, \delta ] \times \left[ \tfrac{1}{\delta }, \infty \right) . \end{aligned}$$
(D.6)

Simple computations and (D.6) produce

$$\begin{aligned} \begin{aligned}&\limsup _{\lambda \rightarrow \infty } \Big | \lambda \int _0^t e^{-\lambda s } f(s,g(\lambda )) ds - f(0,x_\infty ) \Big | \\ {}&\quad = \limsup _{\lambda \rightarrow \infty } \bigg | \int _0^{\delta } \lambda e^{-\lambda s} (f(s,g(\lambda )) - f(0,x_\infty ))ds \\&+\int _{ \delta }^t \lambda e^{-\lambda s} (f(s,g(\lambda )) - f(0,x_\infty )) ds - e^{-\lambda t} f(0,x_\infty ) \bigg | \le \eta . \end{aligned} \end{aligned}$$

Since \(\eta >0\) can be arbitrary small, we conclude (D.4) for the case of \(\alpha =0\). The other cases in (D.4) can be obtained by the same way as above, using the following expressions:

$$\begin{aligned} \begin{aligned}&\lambda ^{\frac{3}{2}} \int _0^t e^{-\lambda s} s^{\frac{1}{2}} ds = -\sqrt{\lambda t} e^{-\lambda t} + \int _0^{\sqrt{\lambda t}} e^{-s^2} ds \xrightarrow [\lambda \rightarrow \infty ]{} \frac{\sqrt{\pi }}{2}, \\&\lambda ^2 \int _0^t e^{-\lambda s}s \,ds =1 -(1+\lambda t) e^{-\lambda t}\xrightarrow [\lambda \rightarrow \infty ]{} 1, \\&\lambda ^{\frac{5}{2}} \int _0^t e^{-\lambda s} s^{\frac{3}{2}} ds = -\tfrac{(3+2\lambda t)\sqrt{\lambda t} e^{-\lambda t}}{2} + \frac{3}{2}\int _0^{\sqrt{\lambda t}} e^{-s^2} ds \xrightarrow [\lambda \rightarrow \infty ]{} \frac{3\sqrt{\pi }}{4},\\&\lambda ^3 \int _0^t e^{-\lambda s}s^2 ds =2 -(2+\lambda t(2+\lambda t)) e^{-\lambda t}\xrightarrow [\lambda \rightarrow \infty ]{} 2. \end{aligned} \end{aligned}$$
(D.7)

One can easily observe that \(\sup _{x>0} x^{\frac{\alpha }{2}} e^{-x} <\infty \) for \(\alpha \in \{0,1,2,3,4\}\). This observation and the explicit expressions in (D.7) produce the bound (D.5). \(\square \)

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

Gang, T.U., Choi, J.H. Optimal Investment in an Illiquid Market with Search Frictions and Transaction Costs. Appl Math Optim 88, 3 (2023). https://doi.org/10.1007/s00245-023-09971-7

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00245-023-09971-7

Keywords

Navigation