Skip to main content
Log in

Optimal consumption/investment and retirement with necessities and luxuries

  • Original Article
  • Published:
Mathematical Methods of Operations Research Aims and scope Submit manuscript

Abstract

In this paper, we study an optimal consumption of necessary and luxury goods, investment, and voluntary retirement choice model. The felicity function is given by the weighted sum of quadratic and HARA utility functions. We use the duality/martingale method to derive a closed form solution for optimal consumption of necessity and luxury, and investment. We also explain properties of optimal policies by using numerical results.

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

Similar content being viewed by others

Notes

  1. Here the first good serves as a numeraire and the consumption of the luxury good is the total expenditure on the good.

  2. In Sect. 3, we see that the marginal utility of wealth at \({\bar{W}}\) is equal to \(\alpha _2 L^{-\gamma }\) and it tends to decrease with an increase in L, and from the inverse relationship between wealth and its marginal utility, we know \({\bar{W}}\) tends to increase with an increase in L.

  3. Notation \(X^w_t\) means that the wealth process is controlled by \(w=(c, g, \pi ,\tau )\).

  4. Since the agent retires from labor, labor income \(Y_t\) and disutility l are equal to 0 in this case. The time \(t=0\) and the initial wealth x correspond to the retirement date \(\tau \) and the wealth \(X_\tau \), respectively.

  5. In this figure, we only consider the case of positive wealth to observe a positive ratio.

References

  • Aït-Sahalia Y, Parker J, Yogo M (2004) Luxury goods and the equity premium. J Financ 59:2959–3004

  • Becker G (1974) A theory of social interactions. J Polit Econ 82(6):1063–1093

    Article  Google Scholar 

  • Blinder AS (1975) Distribution effects and the aggregate consumption function. J Polit Econ 83:447–475

    Article  Google Scholar 

  • Carroll C (2002) Why do the rich save so much? In: Slemrod J (ed) Does Atlas Shrug? Harvard University Press, The Economic Consequences of Taxing the Rich, pp 465–484

  • Choi KJ, Shim G (2006) Disutility, optimal retirement, and portfolio selection. Math Financ 16(2):443–467

    Article  MathSciNet  Google Scholar 

  • Choi KJ, Shim G, Shin YH (2008) Optimal portfolio, consumption-leisure and retirement choice problem with CES utility. Math Financ 18(3):445–472

    Article  MathSciNet  Google Scholar 

  • Ding J, Kingston G, Purcal S (2014) Dynamic asset allocation when bequests are luxury goods. J Econ Dyn Control 38:65–71

    Article  MathSciNet  Google Scholar 

  • Dybvig PH, Liu H (2010) Lifetime consumption and investment: retirement and constrained borrowing. J Econ Theor 145:885–907

    Article  MathSciNet  Google Scholar 

  • Farhi E, Panageas S (2007) Saving and investing for early retirement: a theoretical analysis. J Financ Econ 83:87–121

    Article  Google Scholar 

  • He H, Pages H (1993) Labor income, borrowing constraints, and equilibrium asset prices. Econ Theor 3(4):663–696

    Article  MathSciNet  Google Scholar 

  • Karatzas I, Wang H (2000) Utility maximazation with discretionary stopping. SIAM J. Control Optim 39(1):306–329

    Article  MathSciNet  Google Scholar 

  • Karatzas I, Lehoczky J, Shreve S (1987) Optimal portfolio and consumption decisions for a Small Investor on a finite horizon. SIAM J Control Optim 25(6):1557–1586

    Article  MathSciNet  Google Scholar 

  • Knudsen TS, Meister B, Zervos M (1998) Valuation of investments in real assets with implications for the stock prices. SIAM J. Control Optim 36(6):2082–2102

    Article  MathSciNet  Google Scholar 

  • Koo JL, Ahn SR, Koo BL, Koo HK, Shin YH (2016) Optimal consumption and portfolio selection with quadratic utility and a subsistence consumption constraint. Stoch Anal Appl 34(1):165–177

    Article  MathSciNet  Google Scholar 

  • Peskir G, Shiryaev A (2006) Optimal stopping and free-boundary problems. Birkha user Verlag, Basel

    MATH  Google Scholar 

  • Shin YH, Koo JL, Roh KH (2018) An optimal consumption and investment problem with quadratic utility and subsistence consumption constraints: a dynamic programming approach. Math Model Anal 23(4):627–638

    Article  MathSciNet  Google Scholar 

  • Wachter JA, Yogo M (2010) Why do household portfolio shares rise in wealth? Rev Financ Stud 23:3929–3965

    Article  Google Scholar 

  • Xu ZQ, Zheng H (2019) Optimal investment, heterogeneous consumption and the best time for retirement. working paper

  • Yang Z, Koo HK (2018) Optimal consumption and portfolio selection with early retirement option. Math Op Res 43(4):1378–1404

Download references

Acknowledgements

We thank the anonymous referees for their comments which greatly helped to improve our paper. A preliminary version of this paper was presented at Winter Workshop on Operations Research, Finance and Mathematics, Jozankei, Sapporo, Japan, February 20–24, 2017, The 5th Asian Quantitative Finance Conference (AQFC 2017), Seoul, Korea, April 24–26, 2017, The 3rd International Conference on Engineering and Computational Mathematics (ECM 2017), Stream 6: Financial Mathematics, The Hong Kong Polytechnic University, Hong Kong, May 31–June 02, 2017, KSIAM 2017 Spring Conference, Seoul National University, Seoul, Korea, June 24, 2017, The 49th ISCIE International Symposium on Stochastic Systems Theory and Its Applications (SSS17), Hiroshima Institute of Technology, Hiroshima, Japan, November 03–04, 2017, The 12th AIMS Conference on Dynamical Systems, Differential Equations and Applications (SS99: Problems and Challenges in Financial Engineering and Risk Management), National Taiwan University (NTU), Taipei, Taiwan, July 05–09, 2018, and Workshop on Financial Risks and Their Management, Ryukoku University, Kyoto, Japan, February 19–20, 2019.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kum-Hwan Roh.

Additional information

Publisher's Note

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

Hyeng Keun Koo was supported by the National Research Foundation of Korea (NRF) grant [Grant No. NRF-2020R1A2C1A01006134]. Kum-Hwan Roh was supported by the NRF grant [Grant No. NRF-2020R1F1A1A01076001]. Yong Hyun Shin was supported by the NRF grant [Grant No.s NRF-2019R1H1A2079177, NRF-2020R1A2C1A01006134]

Appendices

Budget constraint

Let \(w=((c_t)_{t=0}^{\infty }, (g_t)_{t=0}^{\infty },(\pi _t)_{t=0}^{\infty },\tau )\) be an admissible policy, and let \(\xi _T\) be a process defined by

$$\begin{aligned} \xi _T:= \int _0^T H_t(c_t+g_t)\, dt + H_T X^w_T. \end{aligned}$$

Suppose first that the agent has already retired, i.e., \(\tau \le 0\). By Itô’s formula, \(\xi _T\) is a local martingale. By (2.5),

$$\begin{aligned} H_T X^w_T\ge 0, \end{aligned}$$
(1.28)

and hence, \(\xi _T\) is a super-martingale. Therefore,

$$\begin{aligned} x= \xi _0\ge {\mathbb {E}}\, [\xi _T] = {\mathbb {E}}\, \left[ \int _0^T H_t (c_t+g_t)\, dt + H_T X^w_T\right] . \end{aligned}$$
(1.29)

By (1.28) and Fatou’s lemma, we have

$$\begin{aligned} \lim _{T\rightarrow \infty } {\mathbb {E}} [H_TX_T^w]\ge 0. \end{aligned}$$
(1.30)

By taking a limit in (1.29), and using (1.30) and the monotone convergence theorem, we obtain

$$\begin{aligned} x\ge {\mathbb {E}}\left[ \int _0^\infty H_t (c_t+g_t)\, dt\right] . \end{aligned}$$

Suppose that \(\xi _T\) is a martingale and \(\displaystyle \lim _{T\rightarrow \infty } {\mathbb {E}} [H_TX_T^w]=0\). Then, the budget constraint is satisfied with an equality:

$$\begin{aligned} x= {\mathbb {E}}\left[ \int _0^\infty H_t (c_t+g_t)\, dt\right] . \end{aligned}$$

We now consider the pre-retirement case, i.e., \(\tau >0\). Let \({\widetilde{\xi }}_T\) be a process defined by

$$\begin{aligned} {\widetilde{\xi }}_T:=\int _0^{T\wedge \tau } H_t (c_t+g_t)\, dt + H_{T\wedge \tau } \left( X^w_{T\wedge \tau }+ \frac{Y_{T\wedge \tau }}{r-\delta }\right) . \end{aligned}$$

By Itô’s formula, \({\widetilde{\xi }}_T\) is a local martingale. By (2.5),

$$\begin{aligned} H_T \left( X^w_T+\frac{Y_T}{r-\delta }\right) \ge \frac{\varGamma }{r} H_T \ge 0, \end{aligned}$$

and hence, \({\widetilde{\xi }}_T\) is a super-martingale. Therefore,

$$\begin{aligned} x+\frac{Y}{r-\delta } = {\widetilde{\xi }}_0\ge {\mathbb {E}} \left[ {\widetilde{\xi }}_T\right] = {\mathbb {E}} \left[ \int _0^{T\wedge \tau } H_t (c_t+g_t)\, dt + H_{T\wedge \tau } \left( X^w_{T\wedge \tau }+\frac{Y_{T\wedge \tau }}{r-\delta }\right) \right] .\nonumber \\ \end{aligned}$$
(1.31)

By Fatou’s lemma, we have

$$\begin{aligned} \liminf _{T\rightarrow \infty } {\mathbb {E}} \left[ H_{T\wedge \tau } \left( X^w_{T\wedge \tau }+\frac{Y_{T\wedge \tau }}{r-\delta }\right) \right] \ge {\mathbb {E}}\left[ H_\tau \left( X^w_\tau +\frac{Y_\tau }{r-\delta }\right) \right] . \end{aligned}$$
(1.32)

By taking a limit in (1.31), and using (1.32) and the monotone convergence theorem, we obtain budget constraint (3.18):

$$\begin{aligned} x+\frac{Y}{r-\delta }\ge {\mathbb {E}} \left[ \int _0^{\tau } H_t (c_t+g_t)\, dt + H_\tau \left( X^w_\tau +\frac{Y_\tau }{r-\delta }\right) \right] . \end{aligned}$$

Proof of lemma 5

The dual utility \({\widetilde{u}}_1(y)\) is given as follows:

$$\begin{aligned} {\widetilde{u}}_1(y)=\max _{c\ge \varGamma }\left\{ u(c)-\frac{1}{\alpha _1}yc\right\} =c^*-R(c^*)^2-\frac{1}{\alpha _1}yc^*, \end{aligned}$$

where \(c^*\) satisfies the equation \(1-2Rc^*=\frac{1}{\alpha _1}y.\) Since \(c^*\ge \varGamma \), the optimal consumption rate of the necessary good is

$$\begin{aligned} c_t^*=\left\{ \begin{array}{ll} \frac{1-\frac{1}{\alpha _1}y_t}{2R}, &{} \text{ if } 0<y_t\le \alpha _1(1-2R{\varGamma }),\\ {\varGamma }, &{} \text{ if } y_t> \alpha _1(1-2R{\varGamma }). \end{array}\right. \end{aligned}$$

Similarly, we obtain the optimal consumption rate of the luxury good.

Proof of proposition 9

From ODE (3.13), we can obtain a general solution \(\phi \) as follows:

$$\begin{aligned} \phi (y_t)=\left\{ \begin{array}{ll} a_1y_t^{m_+}+a_2 y_t^{m_-}+\frac{\gamma }{1-\gamma }\left( \frac{1}{\alpha _2}\right) ^{-\frac{1}{\gamma }}\frac{1}{K}y_t^{-\frac{1-\gamma }{\gamma }}&{}\\ \qquad \qquad -\frac{1}{4\alpha _1R(\rho -2r+\theta ^2)}y_t^2+\frac{2LR-1}{2rR}y_t+\frac{\alpha _1}{4\rho R}, &{} \text{ if } 0<y_t\le \alpha _2L^{-\gamma },\\ b_1y_t^{m_+}+b_2y_t^{m_-}-\frac{1}{4\alpha _1R(\rho -2r+\theta ^2)}y_t^2&{}\\ \qquad \qquad -\frac{1}{2rR}y_t+\frac{\alpha _1}{4\rho R}+\frac{\alpha _2L^{1-\gamma }}{\rho (1-\gamma )}, &{} \text{ if } \alpha _2L^{-\gamma }<y_t\le \alpha _1(1-2R\varGamma ),\\ c_1 y_t^{m_+}+c_2 y_t^{m_-}-\frac{\varGamma }{r}y_t+\frac{1}{\rho }&{}\\ \qquad \qquad \left\{ \alpha _1(\varGamma -R\varGamma ^2)+\frac{\alpha _2L^{1-\gamma }}{1-\gamma }\right\} , &{} \text{ if } y_t> \alpha _1(1-2R\varGamma ), \end{array}\right. \end{aligned}$$
(3.33)

where \(m_+\) and \(m_-\) are two real roots of the quadratic equation (3.15). By the transversality condition in (3.11) constants \(a_2\) and \(c_1\) should be zero. By using the smooth-pasting condition at \(y=\alpha _2 L^{-\gamma }\) and \(y=\alpha _1(1-2R\varGamma )\), we determine the coefficients as follows:

$$\begin{aligned} b_1&=\frac{\alpha _1^{1-m_+}(1-2R\varGamma )^{2-m_+}}{2R(m_+-m_-)}\left\{ \frac{2-m_-}{2(\rho -2r+\theta ^2)}+\frac{1-m_-}{r}+\frac{m_-}{2\rho }\right\} ,\\ b_2&=\frac{\alpha _2^{1-m_-} L^{1-\gamma +\gamma m_-}}{m_+ -m_-}\left\{ \left( \frac{\gamma m_+}{1-\gamma }+1\right) \frac{1}{K}+\frac{m_+-1}{r}-\frac{m_+}{\rho (1-\gamma )}\right\} ,\\ a_1&=b_1+\frac{\alpha _2^{1-m_+}L^{1-\gamma +\gamma m_+}}{m_+-m_-}\left\{ \left( \frac{\gamma m_-}{1-\gamma }+1\right) \frac{1}{K}+\frac{m_--1}{r}-\frac{m_-}{\rho (1-\gamma )}\right\} ,\\ c_2&=b_2+\frac{\alpha _1^{1-m_-}(1-2R\varGamma )^{2-m_-}}{2R(m_+-m_-)}\left\{ \frac{2-m_+}{2(\rho -2r+\theta ^2)}+\frac{1-m_+}{r}+\frac{m_+}{2\rho }\right\} . \end{aligned}$$

By Assumption 3, \(\phi (y)\) satisfies the tranversality condition (3.11). The functional form of \(\phi (y)\) in (3.33) implies that \(\phi \) satisfies the growth condition (3.10). Since \(M_t\) defined in (3.12) is a martingale, Itô’s formula together with the transversality condition (3.11) implies

$$\begin{aligned} \phi (y)={\mathbb {E}}^{y_0=y}\left[ \int _0^\infty e^{-\rho s}\left\{ \alpha _1 {\widetilde{u}}_1(y_s)+\alpha _2 {\widetilde{u}}_2(y_s)\right\} ds\right] . \end{aligned}$$

Proof of lemma 10

Remark 22

From the quadratic function (3.14) and the quadratic equation (3.15), we see that

$$\begin{aligned} f_1\left( -\frac{1-\gamma }{\gamma }\right) =-K<0 \end{aligned}$$

by Assumption 3. This implies

$$\begin{aligned} m_-<-\frac{1-\gamma }{\gamma }<m_+. \end{aligned}$$
(4.34)

Remark 23

From the quadratic equation (3.15), we see that

$$\begin{aligned}&m_++m_-=-\frac{2\left( \rho -r-\frac{1}{2}\theta ^2\right) }{\theta ^2},\quad m_+m_-=-\frac{2\rho }{\theta ^2},\\&(m_+-1)+(m_--1)=-\frac{2\left( \rho -r+\frac{1}{2}\theta ^2\right) }{\theta ^2},\quad (m_+-1)(m_--1)=-\frac{2r}{\theta ^2}. \end{aligned}$$

Remark 24

From Remark 23, we obtain that

$$\begin{aligned} K&=-\frac{\theta ^2(\gamma m_++1-\gamma )(\gamma m_-+1-\gamma )}{2\gamma ^2}>0,\\ \rho&=-\frac{\theta ^2 m_+m_-}{2},\qquad r=-\frac{\theta ^2 (m_+-1)(m_--1)}{2}. \end{aligned}$$

Lemma 25

We see that

$$\begin{aligned}&\left( \frac{\gamma m_+}{1-\gamma }+1\right) \frac{1}{K}+\frac{m_+-1}{r}-\frac{m_+}{\rho (1-\gamma )}=-\frac{2}{\theta ^2m_-(m_--1)(\gamma m_-+1-\gamma )}>0,\\&\left( \frac{\gamma m_-}{1-\gamma }+1\right) \frac{1}{K}+\frac{m_--1}{r}-\frac{m_-}{\rho (1-\gamma )}=-\frac{2}{\theta ^2m_+(m_+-1)(\gamma m_++1-\gamma )}<0. \end{aligned}$$

Proof

First we use Remark 24 to obtain

$$\begin{aligned}&\left( \frac{\gamma m_+}{1-\gamma }+1\right) \frac{1}{K}+\frac{m_+-1}{r}-\frac{m_+}{\rho (1-\gamma )}\\&\quad =-\frac{2\gamma ^2}{\theta ^2(1-\gamma )(\gamma m_-+1-\gamma )}-\frac{2}{\theta ^2(m_--1)}+\frac{2}{\theta ^2 m_-(1-\gamma )}\\&\quad =-\frac{2}{\theta ^2m_-(m_--1)(\gamma m_-+1-\gamma )}>0, \end{aligned}$$

since \(\gamma m_-+1-\gamma <0\) (see the first inequality in (4.34)). Similarly, we can obtain the second identity. \(\square \)

Now we will prove Lemma 10 as follows: From Proposition 9, we see that

$$\begin{aligned} \phi '(y)=\left\{ \begin{array}{ll} m_+a_1 y^{m_+-1}-\alpha _2^{\frac{1}{\gamma }}\frac{1}{K}y^{-\frac{1}{\gamma }} &{}\\ \quad -\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}y+\frac{2LR-1}{2rR}, &{} \text{ if } 0<y< \alpha _2L^{-\gamma },\\ m_+b_1y^{m_+-1}+m_-b_2 y^{m_--1} &{}\\ \quad -\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}y-\frac{1}{2rR}, &{} \text{ if } \alpha _2L^{-\gamma }<y< \alpha _1(1-2R{\varGamma }),\\ m_-c_2 y^{m_--1}-\frac{{\varGamma }}{r}, &{} \text{ if } y> \alpha _1(1-2R{\varGamma }). \end{array}\right. \end{aligned}$$

Let us consider

$$\begin{aligned}&\phi '(\alpha _2L^{-\gamma }+)-\phi '(\alpha _2L^{-\gamma }-)\\&\quad =m_+(b_1-a_1)\alpha _2^{m_+-1}L^{-\gamma m_++\gamma }+m_-b_2 \alpha _2^{m_--1}L^{-\gamma m_-+\gamma }+\frac{1}{K}L-\frac{L}{r}\\&\quad =\frac{L}{m_+ -m_-}\left[ -\left( m_+-m_-\right) \frac{1}{K}+\frac{m_+-m_-}{r}\right] +\frac{1}{K}L-\frac{L}{r}\\&\quad =0 \end{aligned}$$

and

$$\begin{aligned}&\phi '(\alpha _1(1-2R{\varGamma })+)-\phi '(\alpha _1(1-2R{\varGamma })-)\\&\quad =-m_+b_1\alpha _1^{m_+-1}(1-2R{\varGamma })^{m_+-1}+m_-(c_2-b_2) \alpha _1^{m_--1}(1-2R{\varGamma })^{m_--1}\\&\qquad +\frac{1-2R{\varGamma }}{2R(\rho -2r+\theta ^2)}+\frac{1-2R{\varGamma }}{2rR}\\&\quad =\frac{1-2R\varGamma }{2R(m_+-m_-)}\left[ -\frac{m_+-m_-}{\rho -2r+\theta ^2}- \frac{m_+-m_-}{r} \right] \\&\qquad +\frac{1-2R\varGamma }{2R}\left( \frac{1}{\rho -2r+\theta ^2}+\frac{1}{r}\right) \\&\quad =0. \end{aligned}$$

Thus \(\phi (y)\) is differentiable for \(y>0\) and \(\phi '(y)\) is continuous for \(y>0\).

Also we see that

$$\begin{aligned} \phi ''(y)=\left\{ \begin{array}{ll} m_+(m_+-1)a_1 y^{m_+-2}+\frac{1}{\gamma }\alpha _2^{\frac{1}{\gamma }}\frac{1}{K}y^{-\frac{1}{\gamma }-1}&{}\\ \quad -\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}, &{} \text{ if } 0<y< \alpha _2L^{-\gamma },\\ m_+(m_+-1)b_1y^{m_+-2}+m_-(m_--1)b_2 y^{m_--2}&{}\\ \quad -\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}, &{} \text{ if } \alpha _2L^{-\gamma }<y< \alpha _1(1-2R{\varGamma }),\\ m_-(m_--1)c_2 y^{m_--2}, &{} \text{ if } y> \alpha _1(1-2R{\varGamma }). \end{array}\right. \end{aligned}$$

Let us consider

$$\begin{aligned}&\phi ''(\alpha _2L^{-\gamma }+)-\phi ''(\alpha _2L^{-\gamma }-)\\&\quad =m_+(m_+-1)(b_1-a_1)\alpha _2^{m_+-2}L^{-\gamma (m_+-2)} +m_-(m_--1)b_2\alpha _2^{m_--2}L^{-\gamma (m_--2)}\\&\qquad -\frac{1}{\gamma }\frac{1}{K}\alpha _2^{-1}L^{1+\gamma }\\&\quad =-\frac{m_+(m_+-1)}{m_+-m_-}\left\{ \left( \frac{\gamma m_-}{1-\gamma }+1\right) \frac{1}{K}+\frac{m_--1}{r}-\frac{m_-}{\rho (1-\gamma )}\right\} \alpha _2^{-1}L^{1+\gamma }\\&\qquad +\frac{ m_-(m_--1)}{m_+ -m_-}\left\{ \left( \frac{\gamma m_+}{1-\gamma }+1\right) \frac{1}{K}+\frac{m_+-1}{r}-\frac{m_+}{\rho (1-\gamma )}\right\} \alpha _2^{-1}L^{1+\gamma }\\&\qquad -\frac{1}{\gamma }\frac{1}{K}\alpha _2^{-1}L^{1+\gamma }\\&\quad =\alpha _2^{-1}L^{1+\gamma }\left[ -\frac{1}{K}\left\{ \frac{2 \rho \gamma }{(\gamma -1)\theta ^2}-\frac{2\left( \rho -r-\frac{1}{2}\theta ^2\right) }{\theta ^2}-1+\frac{1}{\gamma }\right\} +\frac{2\gamma }{(\gamma -1)\theta ^2}\right] =0 \end{aligned}$$

and

$$\begin{aligned}&\phi ''(\alpha _1(1-2R{\varGamma })+)-\phi ''(\alpha _1(1-2R{\varGamma })-)\\&\quad =m_-(m_--1)(c_2-b_2) \alpha _1^{m_--2}(1-2R{\varGamma })^{m_--2}\\&\qquad -m_+(m_+-1)b_1\alpha _1^{m_+-2}(1-2R{\varGamma })^{m_+-2}\\&\qquad +\frac{1}{2\alpha _1R(\rho -2r+\theta ^2)}\\&\quad =\frac{m_-(m_--1)}{2\alpha _1R(m_+-m_-)}\left\{ \frac{2-m_+}{2(\rho -2r+\theta ^2)}+\frac{1-m_+}{r}+\frac{m_+}{2\rho }\right\} \\&\qquad -\frac{m_+(m_+-1)}{2\alpha _1R(m_+-m_-)}\left\{ \frac{2-m_-}{2(\rho -2r+\theta ^2)}+\frac{1-m_-}{r}+\frac{m_-}{2\rho }\right\} +\frac{1}{2\alpha _1R(\rho -2r+\theta ^2)}\\&\quad =\frac{1}{2\alpha _1R}\left[ \frac{(m_+-1)(m_--1)-(m_++m_-)+1}{2(\rho -2r+\theta ^2)}+\frac{(m_+-1)(m_--1)}{r}\right. \\&\qquad \left. -\frac{m_+m_-}{2\rho }+\frac{1}{\rho -2r+\theta ^2}\right] =0. \end{aligned}$$

Thus \(\phi '(y)\) is differentiable for \(y>0\) and \(\phi ''(y)\) is continuous for \(y>0\).

Proof of theorem 11

Let us define

$$\begin{aligned} \widehat{{\mathcal {X}}}(y):=-\phi '(y),~\text{ for }~y>0. \end{aligned}$$

Then we see that

$$\begin{aligned}&\lim _{y\rightarrow 0+}\widehat{{\mathcal {X}}}(y)=-\lim _{y\rightarrow 0+}\left[ m_+a_1 y^{m_+-1}-\alpha _2^{\frac{1}{\gamma }}\frac{1}{K}y^{-\frac{1}{\gamma }}-\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}y\right. \\&\quad \left. +\frac{2LR-1}{2rR}\right] =+\infty ,\\&\lim _{y\rightarrow +\infty }\widehat{{\mathcal {X}}}(y)=-\lim _{y\rightarrow +\infty }\left[ m_-c_2 y^{m_--1}-\frac{{\varGamma }}{r}\right] =\frac{{\varGamma }}{r},\\&\lim _{y\rightarrow 0+}\widehat{{\mathcal {X}}}'(y)=-\lim _{y\rightarrow 0+}\left[ m_+(m_+-1)a_1 y^{m_+-2}+\frac{1}{\gamma }\alpha _2^{\frac{1}{\gamma }}\frac{1}{K}y^{-\frac{1}{\gamma }-1}\right. \\&\qquad \left. -\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}\right] =-\infty ,\\&\lim _{y\rightarrow +\infty }\widehat{{\mathcal {X}}}'(y)=-\lim _{y\rightarrow +\infty }m_-(m_--1)c_2 y^{m_--2}=0. \end{aligned}$$

Furthermore we see that \(\phi (y)\) is convex since \({\widetilde{u}}_1\) and \({\widetilde{u}}_2\) are convex (see Remark 4), that is, for any \(\eta \in [0,1]\),

$$\begin{aligned}&\phi (y)=\phi (\eta y+(1-\eta ) y)\\&\quad ={\mathbb {E}}\left[ \int _0^\infty e^{-\rho t}\left\{ \alpha _1\cdot {\widetilde{u}}_1(\eta y_t+(1-\eta ) y_t)+\alpha _2\cdot {\widetilde{u}}_2(\eta y_t+(1-\eta ) y_t)\right\} dt\right] \\&\quad \le {\mathbb {E}}\left[ \int _0^\infty e^{-\rho t}\left\{ \alpha _1\cdot \eta {\widetilde{u}}_1(y_t)+\alpha _1\cdot (1-\eta ) {\widetilde{u}}_1(y_t)+\alpha _2\cdot \eta {\widetilde{u}}_2(y_t)+ \alpha _2\cdot (1-\eta ){\widetilde{u}}_2(y_t)\right\} dt\right] \\&\quad =\eta \phi (y)+(1-\eta ) \phi (y). \end{aligned}$$

Thus \(\widehat{{\mathcal {X}}}(y)\) is decreasing and one-to-one mapping from \((0,+\infty )\) to \((\varGamma /r,+\infty )\) and we see that there exists a unique \(y^*\in (0,+\infty )\) such that

$$\begin{aligned} y^*=(\phi ')^{-1}(-x). \end{aligned}$$

Hence

$$\begin{aligned} x= \widehat{{\mathcal {X}}}(y^*). \end{aligned}$$
(5.35)

Let \(y^*_t= y^* e^{\rho t}H_t\) and define \({\widetilde{M}}_{t,s}\), for \(t\le s,\) by

$$\begin{aligned} {\widetilde{M}}_{t,s} := \int _t^s H_v (c_v^*+g_v^*) dv + H_s \widehat{{\mathcal {X}}}(y^*_s), \end{aligned}$$

where \(c_s^*, g^*_s\) are as in Lemma 5 with \(y_t=y^*_t.\) Since \(\widehat{{\mathcal {X}}}' (y)=-\phi ''(y)\) satisfies a growth condition of the form (3.10), Itô’s formula implies that the process \(({\widetilde{M}}_{t,s})_{s\ge t}\) is a martingale. Thus,

$$\begin{aligned} H_t\widehat{{\mathcal {X}}}(y^*_t)= {\mathbb {E}} \left[ \left. \int _t^s H_v (c_v^*+g_v^*) dv + H_s \widehat{{\mathcal {X}}}(y^*_s) \,\right| \, {{\mathcal {F}}}_t \right] . \end{aligned}$$
(5.36)

Assumption 3 implies that

$$\begin{aligned} \lim _{s\rightarrow \infty } {\mathbb {E}} \left[ \left. H_s \widehat{{\mathcal {X}}}(y^*_s)\, \right| \, {{\mathcal {F}}}_t \right] =0. \end{aligned}$$

Thus, by (5.36) and the monotone convergence theorem,

$$\begin{aligned} H_t\widehat{{\mathcal {X}}}(y^*_t)= {\mathbb {E}} \left[ \left. \int _t^{\infty } H_s (c_s^*+g_s^*) ds \,\right| \, {{\mathcal {F}}}_t \right] . \end{aligned}$$

In particular, by (5.35)

$$\begin{aligned} x= {\mathbb {E}} \left[ \int _0^{\infty } H_s (c_s^*+g_s^*) ds \right] . \end{aligned}$$

We can now derive the algebraic equation for each region in Theorem 11 by the relationship \(x=-\phi '(y^*)\). Furthermore, the proof of (3.16) is standard and can be given in a way similar to that in Karatzas et al. (1987).

Proof of theorem 12

The post-retirement value function U in (3.16) attains a minimum value when the equation

$$\begin{aligned} x=-\phi '(y) \end{aligned}$$

is satisfied. Thus the optimal wealth \(x^*\) are given as follows:

$$\begin{aligned} x^*=\left\{ \begin{array}{ll} -m_+a_1 y^{m_+-1}+\left( \frac{1}{\alpha _2}\right) ^{-\frac{1}{\gamma }}\frac{1}{K}y^{-\frac{1}{\gamma }}&{}\\ \quad +\frac{1}{2\alpha _1R(\rho -2r+\theta ^2)}y-\frac{2LR-1}{2rR}, &{} \text{ if } 0<y\le \alpha _2L^{-\gamma },\\ -m_+b_1y^{m_+-1}-m_-b_2 y^{m_--1}&{}\\ \quad +\frac{1}{2\alpha _1R(\rho -2r+\theta ^2)}y+\frac{1}{2rR}, &{} \text{ if } \alpha _2L^{-\gamma }<y\le \alpha _1(1-2R\varGamma ),\\ -m_-c_2 y^{m_--1}+\frac{\varGamma }{r}, &{} \text{ if } y> \alpha _1(1-2R\varGamma ). \end{array}\right. \end{aligned}$$

Applying Itô’s formula to the equation

$$\begin{aligned} X_t=-\phi '(y_t) \end{aligned}$$

which expresses the relationships between wealth process \(X_t\) and its dual variable process \(y_t\), then we obtain the following optimal portfolio \(\pi _t^*\) by comparing with the wealth equation (2.4) and equating coefficients:

$$\begin{aligned} \pi ^*_t=\frac{\theta }{\sigma }y_t \phi ''(y_t). \end{aligned}$$

Let

$$\begin{aligned} \widehat{{\mathcal {X}}}(y_t):=-\phi '(y_t)=X_t,~\text{ for }~y_t>0, \end{aligned}$$

then we see that \(\widehat{{\mathcal {X}}}(y_t)\) is decreasing and one-to-one mapping from \((0,+\infty )\) to \((\varGamma /r,+\infty )\). Also note that

$$\begin{aligned} \widehat{{\mathcal {X}}}(\alpha _1(1-2R\varGamma ))=&-m_-c_2 (\alpha _1(1-2R\varGamma ))^{m_--1}+\frac{\varGamma }{r}\\ =&-m_+b_1(\alpha _1(1-2R\varGamma ))^{m_+-1}-m_-b_2 (\alpha _1(1-2R\varGamma ))^{m_--1}\\&+\frac{1-2R\varGamma }{2 R(\rho -2r+\theta ^2)}+\frac{1}{2rR}\\ =:&{\widehat{x}}_{\alpha _1(1-2R{\varGamma })} \end{aligned}$$

and

$$\begin{aligned} \widehat{{\mathcal {X}}}(\alpha _2L^{-\gamma })=&-m_+b_1(\alpha _2L^{-\gamma })^{m_+-1}-m_-b_2 (\alpha _2L^{-\gamma })^{m_--1}\\&+\frac{\alpha _2L^{-\gamma }}{2\alpha _1R(\rho -2r+\theta ^2)}+\frac{1}{2rR}\\ =&-m_+a_1 (\alpha _2L^{-\gamma })^{m_+-1}+\frac{L}{K}+\frac{\alpha _2L^{-\gamma }}{2\alpha _1R(\rho -2r+\theta ^2)}-\frac{2LR-1}{2rR}\\ =:&{\widehat{x}}_{\alpha _2L^{-\gamma }} \end{aligned}$$

since \(\phi '(y)\) is continuous for \(y>0\). Thus \(\widehat{{\mathcal {X}}}(y_t)\) is one-to-one mapping from \((0,\alpha _2L^{-\gamma })\) to \(({\widehat{x}}_{\alpha _2L^{-\gamma }},+\infty )\) and from \((\alpha _2L^{-\gamma },\alpha _1(1-2R\varGamma ))\) to \(({\widehat{x}}_{\alpha _1(1-2R{\varGamma })},{\widehat{x}}_{\alpha _2L^{-\gamma }})\). Therefore if \({\widehat{x}}_{\alpha _1(1-2R{\varGamma })}<X_t<{\widehat{x}}_{\alpha _2L^{-\gamma }}\), then \({\widehat{\xi }}_t\) is the unique solution to the following equation

$$\begin{aligned}&X_t=-m_+b_1{\widehat{\xi }}_t^{m_+-1}-m_-b_2 {\widehat{\xi }}_t^{m_--1}+\frac{1}{2\alpha _1R(\rho -2r+\theta ^2)}{\widehat{\xi }}_t\\&\quad +\frac{1}{2rR}, ~\text{ for } \alpha _2L^{-\gamma }<{\widehat{\xi }}_t< \alpha _1(1-2R{\varGamma }), \end{aligned}$$

and, if \(X_t>{\widehat{x}}_{\alpha _2L^{-\gamma }}\), then \({\widehat{\zeta }}_t\) is the unique solution to the following equation

$$\begin{aligned}&X_t=-m_+a_1 {\widehat{\zeta }}_t^{m_+-1}+\left( \frac{1}{\alpha _2}\right) ^{-\frac{1}{\gamma }}\frac{1}{K}{\widehat{\zeta }}_t^{-\frac{1}{\gamma }}+\frac{1}{2\alpha _1R(\rho -2r+\theta ^2)}{\widehat{\zeta }}_t\\&\quad -\frac{2LR-1}{2rR}, ~ \text{ for } 0<{\widehat{\zeta }}_t<\alpha _2L^{-\gamma }. \end{aligned}$$

Also the optimal consumption rates of necessity and luxury goods are derived during calculating the dual utility functions \({\widetilde{u}}_1\) and \({\widetilde{u}}_2\) (see Lemma 5).

Derivation of equation (3.19)

The first equality of (3.19) is the definition of V(x) and we need to prove the second equality. The “\(\le \)," i.e., the RHS is greater than or equal to the LHS follows easily from the property of maximum. We will show “\(\ge \)," i.e., the RHS is less than or equal to the LHS.

We assume here the existence of strong Markov optimal policies and will establish the existence of such optimal policies and these policies are strong Markov processes in Appendix J. Suppose that \(w^b=((c^b_t)_{t=0}^{\tau ^*}, (g^b_t)_{t=0}^{\tau ^*}, (\pi ^b_t)_{t=0}^{\tau ^*}, \tau ^*)\) is optimal for the maximization problem in the RHS. Suppose also that \(((c^{a}_t)_{t=0}^{\infty }, (g^a_t)_{t=0}^{\infty }, (\pi ^a_t)_{t=0}^{\infty })\) is optimal for the post-retirement problem with \(x=X^{w^b}_{\tau ^*}\). Define

$$\begin{aligned} c^*_t= {\left\{ \begin{array}{ll} c^b_t &{} \hbox {if}\ t\le \tau ^* \\ c^a_{t-\tau ^*} &{} \hbox {if}\ t>\tau ^* \end{array}\right. }, \quad g^*_t= {\left\{ \begin{array}{ll} g^b_t &{} \hbox {if}\ t\le \tau ^* \\ g^a_{t-\tau ^*} &{} \hbox {if}\ t>\tau ^* \end{array}\right. }, \quad \pi ^*_t= {\left\{ \begin{array}{ll} \pi ^b_t &{} \hbox {if}\ t\le \tau ^* \\ \pi ^a_{t-\tau ^*} &{} \hbox {if}\ t>\tau ^* \end{array}\right. }. \end{aligned}$$

Clearly, \(w= ((c^{*}_t)_{t=0}^{\infty }, (g^*_t)_{t=0}^{\infty }, (\pi ^*_t)_{t=0}^{\infty }, \tau ^*)\) is an admissible policy. We know

$$\begin{aligned} V&=\max _{(c,g,\pi ,\tau )\in {\mathcal {A}}(x)}{\mathbb {E}}\left[ \int _0^\tau e^{-\rho t} \left\{ \alpha _1 {u(c_t)}+\alpha _2\frac{(g_t+L)^{1-\gamma }}{1-\gamma }-l\right\} dt\right. \\&\quad \left. +\int _\tau ^\infty e^{-\rho t} \left\{ \alpha _1 {u(c_t)}+\alpha _2\frac{(g_t+L)^{1-\gamma }}{1-\gamma }\right\} dt\right] \\&\ge {\mathbb {E}}\left[ \int _0^{\tau ^*} e^{-\rho t} \left\{ \alpha _1 {u(c^*_t)}+\alpha _2\frac{(g^*_t+L)^{1-\gamma }}{1-\gamma }-l\right\} dt \right. \\&\quad \left. +e^{-\rho \tau ^*}\int _{\tau ^*}^\infty e^{-\rho (t-\tau ^*)} \left\{ \alpha _1 {u(c^*_t)}+\alpha _2\frac{(g^*_t+L)^{1-\gamma }}{1-\gamma }\right\} dt\right] \\&= {\mathbb {E}}\left[ \int _0^{\tau ^*} e^{-\rho t} \left\{ \alpha _1 {u(c^*_t)}+\alpha _2\frac{(g^*_t+L)^{1-\gamma }}{1-\gamma }-l\right\} dt \right. \\&\quad \left. +e^{-\rho \tau ^*}{\mathbb {E}}\left[ \left. \int _{\tau ^*}^\infty e^{-\rho (t-\tau ^*)} \left\{ \alpha _1 {u(c^*_t)}+\alpha _2\frac{(g^*_t+L)^{1-\gamma }}{1-\gamma }\right\} dt\right| {{\mathcal {F}}}_{\tau ^*}\right] \right] \\&= {\mathbb {E}}\left[ \int _0^{\tau ^*} e^{-\rho t} \left\{ \alpha _1 {u(c^*_t)}+\alpha _2\frac{(g^*_t+L)^{1-\gamma }}{1-\gamma }-l\right\} dt \right. \\&\quad \left. +e^{-\rho \tau ^*}{\mathbb {E}}\left[ \left. \int _{\tau ^*}^\infty e^{-\rho (t-\tau ^*)} \left\{ \alpha _1 {u(c^*_t)}+\alpha _2\frac{(g^*_t+L)^{1-\gamma }}{1-\gamma }\right\} dt\right| X_{\tau ^*}\right] \right] \\&=\max _{(c,g,\pi ,\tau )\in {\mathcal {A}}(x)}{\mathbb {E}}\left[ \int _0^\tau e^{-\rho t} \left\{ \alpha _1 {u(c_t)}+\alpha _2\frac{(g_t+L)^{1-\gamma }}{1-\gamma }-l\right\} dt+e^{-\rho \tau } U(X_\tau )\right] ,\!\!\! \end{aligned}$$

where we have used the strong Markov property of optimal policies and the last equality follows from the definition of \(w= ((c^{*}_t)_{t=0}^{\infty }, (g^*_t)_{t=0}^{\infty }, (\pi ^*_t)_{t=0}^{\infty }, \tau ^*)\). This proves the assertion.

Proof of proposition 19

It can be easily obtained that

$$\begin{aligned} \left( \frac{\partial }{\partial t}+ {\mathcal {L}}\right) {\widehat{U}}(y,Y_t)+\alpha _1{\widetilde{u}}_1(y)+\alpha _2 {\widetilde{u}}_2(y)-l= & {} \left( \frac{\partial }{\partial t}+ {\mathcal {L}}\right) \left[ \phi (y)-\frac{Y_t}{r-\delta }y\right] \\&+\,\alpha _1{\widetilde{u}}_1(y)+\alpha _2 {\widetilde{u}}_2(y)-l\\= & {} Y_ty-l. \end{aligned}$$

Let \(P(t,y):={\widetilde{\varphi }}(t,y)-{\widehat{U}}(y,Y_t)\), then we see that

$$\begin{aligned}&\left( \frac{\partial }{\partial t}+ {\mathcal {L}}\right) {\widetilde{\varphi }}(t,y)+\alpha _1{\widetilde{u}}_1(y)+\alpha _2 {\widetilde{u}}_2(y)-l\\&\quad =\left( \frac{\partial }{\partial t}+ {\mathcal {L}}\right) P(t,y)+\left( \frac{\partial }{\partial t}+ {\mathcal {L}}\right) {\widehat{U}}(y,Y_t)+\alpha _1{\widetilde{u}}_1(y)+\alpha _2 {\widetilde{u}}_2(y)-l\\&\quad =\left( \frac{\partial }{\partial t}+ {\mathcal {L}}\right) P(t,y)+Y_ty-l. \end{aligned}$$

Thus, Variational Inequality (VI) (3.22) can be considered as

$$\begin{aligned} \max \left\{ \left( \frac{\partial }{\partial t}+{\mathcal {L}}\right) P(t,y)+Y_ty-l,-P(t,y)\right\} =0. \end{aligned}$$
(8.37)

We will now proceed to solve VI (8.37). Suppose

$$\begin{aligned} P(t,y_t)=Q(z_t), \end{aligned}$$

where \(z_t= Y_ty_t= e^{\delta t}Yy_t.\) Then, VI (8.37) becomes

$$\begin{aligned} \max \left\{ \bar{{\mathcal {L}}}Q(z)+z-l,-Q(z)\right\} =0, \end{aligned}$$
(8.38)

where

$$\begin{aligned} \bar{{\mathcal {L}}}:=\frac{1}{2}\theta ^2z^2\frac{d^2}{d z^2}+(\rho +\delta -r)z\frac{d}{d z}-\rho . \end{aligned}$$

Thus the optimal retirement threshold \({\bar{z}}\) is given by

$$\begin{aligned} {\bar{z}}=\inf \left\{ z>0\,|\,Q(z)>0\right\} . \end{aligned}$$

Now we can take the following free boundary value problem (FBVP) into consideration:

$$\begin{aligned} \left\{ \begin{array}{ll} \bar{{\mathcal {L}}}Q(z)+z-l=0,&{}~\text{ for } z>{\bar{z}},\\ Q(z)=0,&{}~\text{ for } 0<z\le {\bar{z}},\\ Q({\bar{z}})=Q'({\bar{z}})=0. \end{array}\right. \end{aligned}$$
(8.39)

From the FBVP (8.39), we obtain

$$\begin{aligned} Q(z)=\left\{ \begin{array}{ll} A_2z^{n_-}+\frac{1}{r-\delta }z-\frac{l}{\rho },&{}~\text{ for } z>{\bar{z}},\\ 0,&{}~\text{ for } 0<z\le {\bar{z}}. \end{array}\right. \end{aligned}$$
(8.40)

Also, from the free boundary conditions in (8.39), we derive

$$\begin{aligned} {\bar{z}}=\frac{n_-}{n_--1}\frac{(r-\delta )l}{\rho }>0\quad \text{ and }\quad A_2=\frac{1}{1-n_-}\frac{l}{\rho }{\bar{z}}^{-n_-}>0. \end{aligned}$$

Thus the inequality (3.25) in Assumption 17 implies

$$\begin{aligned} {\bar{z}}<\alpha _2 L^{-\gamma }Y\le \alpha _2 L^{-\gamma }Y_t,~\text{ for } t\ge 0. \end{aligned}$$

Lemma 26

Q(z) in (8.40) satisfies VI (8.38).

Proof

For \(z>{\bar{z}}\), we see that

$$\begin{aligned} Q'(z)=n_-A_2z^{n_--1}+\frac{1}{r-\delta }\quad \text{ and }\quad Q''(z)=n_-(n_--1)A_2z^{n_--2}>0. \end{aligned}$$
(8.41)

So \(Q'(z)\) is an increasing function for \(z>{\bar{z}}\) and Q(z) is also an increasing function for \(z>{\bar{z}}\) since \(Q'({\bar{z}})=0\). Thus \(Q(z)>0\) for \(z>{\bar{z}}\) since \(Q({\bar{z}})=0\). Therefore we have, for \(z>{\bar{z}}\),

$$\begin{aligned} \bar{{\mathcal {L}}}Q(z)+z-l=0\quad \text{ and }\quad -Q(z)<0. \end{aligned}$$

For \(0<z\le {\bar{z}}\), we see that \(Q(z)=0\) and consequently \(\bar{{\mathcal {L}}}Q(z)=0\). The inequality (3.24) implies

$$\begin{aligned} \frac{n_-}{n_--1}<\frac{\rho }{r-\delta }~\Rightarrow ~{\bar{z}}=\frac{n_-}{n_--1}\frac{(r-\delta )l}{\rho }<l, \end{aligned}$$

that is, \(z-l<0\) for \(0<z\le {\bar{z}}\). Therefore we have, for \(0<z\le {\bar{z}}\),

$$\begin{aligned} \bar{{\mathcal {L}}}Q(z)+z-l<0\quad \text{ and }\quad -Q(z)=0. \end{aligned}$$

\(\square \)

Let

$$\begin{aligned} \varPhi (t,y_t):=P(t,y_t)+{\widehat{U}}(y_t,Y_t)=Q(z_t)+\phi (y_t)-\frac{z_t}{r-\delta }. \end{aligned}$$
(8.42)

Then, for \({\bar{z}}/Y_t<y_t<\alpha _2L^{-\gamma }\), we have

$$\begin{aligned} \varPhi (t,y_t)= & {} a_1 y_t^{m_+}+A_2 y_t^{n_-}Y_t^{n_-}+\frac{\gamma }{1-\gamma }\left( \frac{1}{\alpha _2}\right) ^{-\frac{1}{\gamma }}\frac{1}{K}y_t^{-\frac{1-\gamma }{\gamma }}\\&-\,\frac{1}{4\alpha _1R(\rho -2r+\theta ^2)}y_t^2+\frac{2LR-1}{2rR}y_t+\frac{\alpha _1}{4\rho R}-\frac{l}{\rho }. \end{aligned}$$

For \(\alpha _2L^{-\gamma }<y_t\le \alpha _1(1-2R\varGamma )\), we have

$$\begin{aligned} \varPhi (t,y_t)= & {} b_1y_t^{m_+}+b_2y_t^{m_-}+A_2 y_t^{n_-}Y_t^{n_-}-\frac{1}{4\alpha _1R(\rho -2r+\theta ^2)}y_t^2-\frac{1}{2rR}y_t\\&+\,\frac{\alpha _1}{4\rho R}+\frac{\alpha _2L^{1-\gamma }}{\rho (1-\gamma )}-\frac{l}{\rho }. \end{aligned}$$

For \(y_t> \alpha _1(1-2R\varGamma )\), we have

$$\begin{aligned} \varPhi (t,y_t)=c_2 y_t^{m_-}+A_2 y_t^{n_-}Y_t^{n_-}-\frac{\varGamma }{r}y_t+\frac{1}{\rho }\left\{ \alpha _1(\varGamma -R\varGamma ^2)+\frac{\alpha _2L^{1-\gamma }}{1-\gamma }-l\right\} . \end{aligned}$$

Also note that

$$\begin{aligned} \varPhi (t,y_t)=\phi (y_t)-\frac{1}{r-\delta }Y_ty_t,~\text{ for }~0<y_t\le {\bar{z}}/Y_t. \end{aligned}$$

Since \(\phi (y)\) is a \(C^2\)-function for \(y>0\), it can be easily checked that \(\varPhi (t,y)\) is also a \(C^2\)-function for \(y>0\) except at \(y_t={\bar{z}}/Y_t\) (we see that \(\varPhi (t,y)\) is a \(C^1\)-function at \(y_t={\bar{z}}/Y_t\)).

Notice that \(Q(z)-\frac{z}{r-\delta }\) satisfies growth condition of the form (3.10) and \(z_t\) follows the dynamics

$$\begin{aligned} \frac{dz_t}{z_t}=(\rho +\delta -r) dt -\theta dB(t). \end{aligned}$$

Define the process \(N_t\) by

$$\begin{aligned} N_t:=\int _0^{t}e^{-\rho s} \theta y_s \frac{\partial \varPhi }{\partial y} (t,y_s) ds. \end{aligned}$$

Then, by the growth condition, \(N_t\) is a martingale.

Equation (8.42) implies that the following transversality condition is satisfied:

$$\begin{aligned} \lim _{t\rightarrow \infty } e^{-\rho t}{\mathbb {E}} \left[ |\varPhi (t,y_t)|\right] =0. \end{aligned}$$

Hence, by Theorem 3.2 of Knudsen et al. (1998) we know

$$\begin{aligned}&\varPhi (t,y) = {\mathbb {E}}^{y_0=y}\left[ \int _0^{ \tau ^*}e^{-\rho t} \left\{ \alpha _1{\widetilde{u}}_1(y_t)+\alpha _2 {\widetilde{u}}_2(y_t)-l\right\} dt \right. \\&\left. + e^{-\rho \tau ^*} \left( \phi (y_{\tau ^*}) -\frac{Y_{\tau ^*}}{r-\delta }y_{\tau ^*}\right) \right] ={\widetilde{V}}. \end{aligned}$$

Proof of theorem 20

Let us define

$$\begin{aligned} {\mathcal {X}}(t,y):=-\frac{\partial }{\partial y}\varPhi (t,y)-\frac{Y_t}{r-\delta }, ~\text{ for }~y> 0. \end{aligned}$$

Then we see that

$$\begin{aligned}&\lim _{y\rightarrow 0+}{\mathcal {X}}(t,y)\\&\quad =-\lim _{y\rightarrow 0+}\left[ m_+a_1 y^{m_+-1}-\alpha _2^{\frac{1}{\gamma }}\frac{1}{K}y^{-\frac{1}{\gamma }}-\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}y+\frac{2LR-1}{2rR}\right] \\&\quad =+\infty ,\\&\lim _{y\rightarrow +\infty }{\mathcal {X}}(t,y)\\&\quad =-\lim _{y\rightarrow +\infty }\left[ m_-c_2y^{m_--1}+n_-A_2 y^{n_--1}Y_t^{n_-}-\frac{{\varGamma }}{r}+\frac{Y}{r-\delta }\right] =\frac{{\varGamma }}{r}-\frac{Y}{r-\delta },\\&\lim _{y\rightarrow 0+}{\mathcal {X}}'(t,y)\\&\quad =-\lim _{y\rightarrow 0+}\left[ m_+(m_+-1)a_1 y^{m_+-2}+\frac{1}{\gamma }\alpha _2^{\frac{1}{\gamma }}\frac{1}{K}y^{-\frac{1}{\gamma }-1}-\frac{2}{4\alpha _1R(\rho -2r+\theta ^2)}\right] \\&\quad =-\infty ,\\&\lim _{y\rightarrow +\infty }{\mathcal {X}}'(t,y)\\&\quad =-\lim _{y\rightarrow +\infty }\left[ m_-(m_--1)c_2 y^{m_--2}+n_-(n_--1)A_2 y^{n_--2}Y_t^{n_-} \right] =0. \end{aligned}$$

We see that \(Q''(z)>0\), for \(z>{\bar{z}}\) from (8.41) in Lemma 26. Thus

$$\begin{aligned} \frac{\partial ^2}{\partial y^2}P(t,y_t)=Y_t^2Q''(z_t)>0,~\text{ for } y_t>\frac{{\bar{z}}}{Y_t}. \end{aligned}$$

This implies that \(P(\cdot ,y)\) is a strictly convex function for \(y>{\bar{z}}/Y\). Also we have shown that \(\phi (y)\) is a convex function for \(y>0\) in Appendix E. Since

$$\begin{aligned} \varPhi (t,y_t)=P(t,y_t)+\phi (y_t)-\frac{Y_t}{r-\delta }y_t,~\text{ for }~~y_t>\frac{{\bar{z}}}{Y_t}, \end{aligned}$$

it is also clear that \(\varPhi (\cdot ,y)\) is a convex function for \(y>{\bar{z}}/Y\). Furthermore since

$$\begin{aligned} \varPhi (t,y_t)=\phi (y_t)-\frac{Y_t}{r-\delta }y_t,~\text{ for }~0<y_t\le \frac{{\bar{z}}}{Y_t}, \end{aligned}$$

it is clear that \(\varPhi (\cdot ,y)\) is a convex function for \(0<y\le {\bar{z}}/Y \).

Thus \({\mathcal {X}}(\cdot ,y)\) is decreasing and one-to-one mapping from \((0,+\infty )\) to \((\varGamma /r-Y/(r-\delta ),+\infty )\) and we see that there exists a unique \({\mathcal {Y}}^*\in (0,+\infty )\) such that

$$\begin{aligned} x= {\mathcal {X}}(0,{\mathcal {Y}}^*). \end{aligned}$$

From the inequality (3.20), we obtain

$$\begin{aligned} V\le \min _{y>0}\left[ \varPhi (0,y)+y\left( x+\frac{Y}{r-\delta }\right) \right] . \end{aligned}$$

Let \(y^*_t= {\mathcal {Y}}^* e^{\rho t}H_t\).

We will now show that the budget constraint holds with equality for the optimal policies by proceeding similarly to the proof of Theorem 3.2 part b), Knudsen et al. (1998). Define

$$\begin{aligned} {\widetilde{N}}_{t}:= \frac{1}{{{\mathcal {Y}}}^*}\int _0^{t} e^{-\rho s} \theta y_s^* \left( {\mathcal {X}}(s,y^*_s)+ \frac{Y_s}{r-\delta }+ y_s^*\frac{\partial }{\partial y}{\mathcal {X}}(s,y^*_s)\right) \,dB(s). \end{aligned}$$

We know \(H_t= e^{-\rho t}\frac{1}{{\mathcal {Y}}^*} y^*_t ,\) and \( y_t^* {\mathcal {X}}(t,y^*_t)\) and \((y_t^*)^2\frac{\partial }{\partial y}{\mathcal {X}}(t,y^*_t)\) are composed of parts satisfying growth conditions of the form (3.10). Hence, the process \({\widetilde{N}}_{t}\) is a martingale. Furthermore, we know

$$\begin{aligned} \lim _{t\rightarrow \infty } {\mathbb {E}}\left[ H_t \left( {\mathcal {X}}(t,y^*_t)+\frac{Y_t}{r-\delta }\right) \right] =0. \end{aligned}$$
(9.43)

Then, Itô’s formula implies

$$\begin{aligned} \int _0^{t\wedge \tau ^*} H_s (c^*_s+g^*_s)ds + H_{t\wedge \tau ^*}\left( {\mathcal {X}}_{t\wedge \tau ^*}+ \frac{Y_{t\wedge \tau ^*}}{r-\delta }\right) = {\mathcal {X}}(0,{{\mathcal {Y}}}^*) +\frac{Y}{r-\delta } - {\widetilde{N}}_{t\wedge \tau ^*}, \end{aligned}$$

where \(c^*_s\) and \(g^*_s\) are as in Lemma 5 with \(y_t=y_t^*.\) Hence,

$$\begin{aligned}&\int _0^{t\wedge \tau ^*} H_s (c^*_s+g^*_s)ds+\mathbf{1}_{\{\tau ^*<t\}} H_{\tau ^*}\left( {\mathcal {X}}_{\tau ^*}+ \frac{Y_{\tau ^*}}{r-\delta }\right) \\&\quad = {\mathcal {X}}(0,{{\mathcal {Y}}}^*)+\frac{Y}{r-\delta }- \mathbf{1}_{\{\tau ^*\ge t\}} H_t \left( {\mathcal {X}}(t,y^*_t)+\frac{Y_t}{r-\delta }\right) - {\widetilde{N}}_{t\wedge \tau ^*}. \end{aligned}$$

Taking expectations, we have

$$\begin{aligned} \begin{aligned} {\mathbb {E}}&\left[ \int _0^{t\wedge \tau ^*} H_s (c^*_s+g^*_s)ds + \mathbf{1}_{\{\tau ^*<t\}} H_{\tau ^*}\left( {\mathcal {X}}_{\tau ^*}+ \frac{Y_{\tau ^*}}{r-\delta }\right) \right] \\&\qquad =\, {\mathcal {X}}(0,{{\mathcal {Y}}}^*)+\frac{Y}{r-\delta } -{\mathbb {E}}\left[ \mathbf{1}_{\{\tau ^*\ge t\}} H_t \left( {\mathcal {X}}(t,y^*_t)+\frac{Y_t}{r-\delta }\right) \right] . \end{aligned} \end{aligned}$$
(9.44)

By (9.43)

$$\begin{aligned} 0\le \lim _{t\rightarrow \infty }{\mathbb {E}} \left[ \mathbf{1}_{\{\tau ^*\ge t\}} H_t \left( {\mathcal {X}}(t,y^*_t)+\frac{Y_t}{r-\delta }\right) \right] \le \lim _{t\rightarrow \infty } {\mathbb {E}}\left[ H_t \left( {\mathcal {X}}(t,y^*_t)+\frac{Y_t}{r-\delta }\right) \right] =0.\nonumber \\ \end{aligned}$$
(9.45)

Thus, by (9.44) and (9.45), the monotone convergence theorem implies

$$\begin{aligned} \begin{aligned}&x +\frac{Y}{r-\delta }= {\mathcal {X}}(0,{{\mathcal {Y}}}^*)+\frac{Y}{r-\delta }\\&\quad = \lim _{t\rightarrow \infty }{\mathbb {E}} \left[ \int _0^{t\wedge \tau ^*} H_s (c^*_s+g^*_s)ds + \mathbf{1}_{\{\tau ^*<t\}} H_{\tau ^*}\left( {\mathcal {X}}_{\tau ^*}+ \frac{Y_{\tau ^*}}{r-\delta }\right) \right] \\&\quad = {\mathbb {E}}\left[ \int _0^{\tau ^*} H_t(c_t^*+g_t^*)dt+H_{ \tau ^*}\left( {\mathcal {X}}\left( \tau ^*,y_{\tau ^*}\right) +\frac{Y_{\tau ^*}}{r-\delta }\right) \right] . \end{aligned} \end{aligned}$$
(9.46)

Then,

$$\begin{aligned} V&\ge {\mathbb {E}}\left[ \int _0^{\tau ^*} e^{-\rho t} \left\{ \alpha _1 u(c_t^*)+\alpha _2\frac{(g_t^*+L)^{1-\gamma }}{1-\gamma }-l\right\} dt+e^{-\rho \tau ^*} U(X_{\tau ^*}^{(c^*,g^*,\pi ^*,\tau ^*)})\right] \\&\quad ={\mathbb {E}}\left[ \int _0^{\tau ^*} e^{-\rho t} \left\{ \alpha _1 u(c_t^*)+\alpha _2\frac{(g_t^*+L)^{1-\gamma }}{1-\gamma }-l\right\} dt+e^{-\rho \tau ^*} U(X_{\tau ^*}^{(c^*,g^*,\pi ^*,\tau ^*)})\right] \\&\qquad -{\mathcal {Y}}^*\,{\mathbb {E}}\left[ \int _0^{\tau ^*} H_t(c_t^*+g_t^*)dt+H_{\tau ^*}\left( X_{\tau ^*}^{(c^*,g^*,\pi ^*,\tau ^*)}+\frac{Y_{\tau ^*}}{r-\delta }\right) \right] +{\mathcal {Y}}^*\left( x+\frac{Y}{r-\delta }\right) \\&\quad = \varPhi (0,{\mathcal {Y}}^*)+{\mathcal {Y}}^*\left( x+\frac{Y}{r-\delta }\right) \\&\quad \ge \min _{y>0}\left[ \varPhi (0,y)+y\left( x+\frac{Y}{r-\delta }\right) \right] ,\\ \end{aligned}$$

where the first equality follows from (9.46).

Thus we have

$$\begin{aligned} V=\min _{y>0}\left[ \varPhi (0,y)+y\left( x+\frac{Y}{r-\delta }\right) \right] . \end{aligned}$$

Proof of theorem 21

The main value function V in (3.27) attains a minimum value when the equation

$$\begin{aligned} x=-\frac{\partial }{\partial y}\varPhi (t,y)-\frac{Y}{r-\delta } \end{aligned}$$

is satisfied. Thus, from \(\varPhi (t,y)\) in (3.26), the optimal wealth \(x^*\) are derived as follows:

$$\begin{aligned} x^*=\left\{ \begin{array}{ll} - m_+a_1 y^{m_+-1}-n_-A_2 y^{n_- -1}Y^{n_-}+\left( \frac{1}{\alpha _2}y\right) ^{-\frac{1}{\gamma }} \frac{1}{K} &{}\\ \qquad \qquad +\frac{1}{2 \alpha _1R(\rho -2r+\theta ^2)}y-\frac{2LR-1}{2r R}-\frac{Y}{r-\delta }, &{} \text{ if } {\bar{z}}/Y<y\le \alpha _2 L^{-\gamma },\\ -m_+b_1 y^{m_+-1}- m_-b_2y^{m_- -1}-n_-A_2 y^{n_--1}Y^{n_-}&{}\\ \qquad \qquad +\frac{1}{2 \alpha _1R(\rho -2r+\theta ^2)}y+\frac{1}{2r R}-\frac{Y}{r-\delta }, &{} \text{ if } \alpha _2 L^{-\gamma }<y\le \alpha _1(1-2R\varGamma ) ,\\ -m_- c_2 y^{m_--1}-n_-A_2 y^{n_--1}Y^{n_-}-\left( \frac{Y}{r-\delta }-\frac{\varGamma }{r}\right) , &{} \text{ if } y>\alpha _1(1-2R\varGamma ). \end{array}\right. \end{aligned}$$

Applying Itô’s formula to the equation

$$\begin{aligned} X_t=-\frac{\partial }{\partial y}\varPhi (t,y_t)-\frac{Y_t}{r-\delta } \end{aligned}$$

which expresses the relation of wealth process \(X_t\) and its dual variable process \(y_t\), we obtain the following optimal portfolio \(\pi _t^*\) by comparing with the wealth equation (2.4) and equating coefficients:

$$\begin{aligned} \pi ^*_t=\frac{\theta }{\sigma }y_t \frac{\partial ^2}{\partial y^2}\varPhi (t,y_t). \end{aligned}$$

Let

$$\begin{aligned} {\mathcal {X}}(t,y_t):=-\frac{\partial }{\partial y}\varPhi (t,y_t)-\frac{Y_t}{r-\delta }=X_t, ~\text{ for }~y_t> 0, \end{aligned}$$

then we see that \({\mathcal {X}}(\cdot ,y_t)\) is decreasing and one-to-one mapping from \((0,+\infty )\) to \((-Y_t/(r-\delta )+\varGamma /r,+\infty )\). Also note that

$$\begin{aligned}&{\mathcal {X}}(t,\alpha _1(1-2R\varGamma ))=-m_- c_2 \{\alpha _1(1-2R\varGamma )\}^{m_--1}-n_-A_2 \{\alpha _1(1-2R\varGamma )\}^{n_--1}Y_t^{n_-}\\&\qquad -\left( \frac{Y_t}{r-\delta }-\frac{\varGamma }{r}\right) \\&\quad =-m_+b_1 \{\alpha _1(1-2R\varGamma )\}^{m_+-1}- m_-b_2\{\alpha _1(1-2R\varGamma )\}^{m_- -1}\\&\qquad -n_-A_2 \{\alpha _1(1-2R\varGamma )\}^{n_--1}Y_t^{n_-}\\&\qquad +\frac{(1-2R\varGamma )}{2 R(\rho -2r+\theta ^2)} +\frac{1}{2r R}-\frac{Y_t}{r-\delta }\\&\quad =:x_{\alpha _1(1-2R\varGamma ),t},\\&{\mathcal {X}}(t,\alpha _2L^{-\gamma })=-m_+b_1 (\alpha _2 L^{-\gamma })^{m_+-1}- m_-b_2(\alpha _2 L^{-\gamma })^{m_- -1}-n_-A_2 (\alpha _2 L^{-\gamma })^{n_--1}Y_t^{n_-}\\&\qquad +\frac{\alpha _2 L^{-\gamma }}{2 \alpha _1R(\rho -2r+\theta ^2)} +\frac{1}{2r R}-\frac{Y_t}{r-\delta }\\&\quad =-m_+a_1 ( \alpha _2 L^{-\gamma })^{m_+-1}-n_- A_2 ( \alpha _2 L^{-\gamma })^{n_- -1}Y_t^{n_-}\\&\qquad + \frac{L}{K}+\frac{\alpha _2 L^{-\gamma }}{2 \alpha _1R(\rho -2r+\theta ^2)}-\frac{2LR-1}{2r R}-\frac{Y_t}{r-\delta }\\&\quad =:x_{\alpha _2L^{-\gamma },t}, \end{aligned}$$

and

$$\begin{aligned}&{\mathcal {X}}(t,{\bar{z}}/Y_t)=-m_+ a_1 \left( \frac{{\bar{z}}}{Y_t}\right) ^{m_+-1}-n_- A_2{\bar{z}}^{n_- -1}Y_t+\left( \frac{{\bar{z}}}{\alpha _2Y_t}\right) ^{-\frac{1}{\gamma }} \frac{1}{K}\\&\quad +\frac{1}{2 \alpha _1 R(\rho -2r+\theta ^2)}\frac{{\bar{z}}}{Y_t}-\frac{2LR-1}{2r R}-\frac{Y_t}{r-\delta }=:{\bar{x}}_t \end{aligned}$$

since \(\frac{\partial }{\partial y}\varPhi (\cdot ,y)\) is continuous for \(y>0\). Thus \({\mathcal {X}}(\cdot ,y_t)\) is one-to-one mapping from \((\alpha _1(1-2R\varGamma ),+\infty )\) to \((-(Y_t/(r-\delta )-\varGamma /r),x_{\alpha _1(1-2R{\varGamma }),t})\), from \((\alpha _2L^{-\gamma },\alpha _1(1-2R\varGamma ))\) to \((x_{\alpha _1(1-2R{\varGamma }),t},x_{\alpha _2L^{-\gamma },t})\), and from \(({\bar{z}}/Y_t,\alpha _2L^{-\gamma })\) to \((x_{\alpha _2L^{-\gamma },t},{\bar{x}}_t)\). Therefore if \(-(Y_t/(r-\delta )-\varGamma /r)<X_t<x_{\alpha _1(1-2R{\varGamma }),t}\), then \(\varXi _t\) is a unique solution to the following equation

$$\begin{aligned} X_t=-m_-c_2 \varXi _t^{m_--1}-n_-A_2 \varXi _t^{n_- -1}Y_t^{n_-}+\frac{\varGamma }{r}-\frac{Y_t}{r-\delta }, ~\text{ for } \varXi _t> \alpha _1(1-2R\varGamma ), \end{aligned}$$

if \(x_{\alpha _1(1-2R{\varGamma }),t}<X_t<x_{\alpha _2L^{-\gamma },t}\), then \(\xi _t\) is a unique solution to the following equation

$$\begin{aligned}&X_t=-m_+b_1 \xi _t^{m_+-1}-m_-b_2\xi _t^{m_- -1}-n_-A_2 \xi _t^{n_- -1}Y_t^{n_-}+\frac{1}{2 \alpha _1R(\rho -2r+\theta ^2)}\xi _t\\&\quad +\frac{1}{2r R}-\frac{Y_t}{r-\delta }, ~\text{ for } \alpha _2L^{-\gamma }<\xi _t< \alpha _1(1-2R\varGamma ), \end{aligned}$$

and, if \(x_{\alpha _2L^{-\gamma },t}<X_t<{\bar{x}}_t\), then \(\zeta _t\) is a unique solution to the following equation

$$\begin{aligned}&X_t=-m_+a_1 \zeta _t^{m_+-1}-n_- A_2 \zeta _t^{n_- -1}Y_t^{n_-}+\left( \frac{1}{\alpha _2}\zeta _t\right) ^{-\frac{1}{\gamma }} \frac{1}{K}\\&\quad +\frac{1}{2 \alpha _1R(\rho -2r+\theta ^2)}\zeta _t-\frac{2LR-1}{2r R}-\frac{Y_t}{r-\delta }, ~ \text{ for } {\bar{z}}/Y_t<\zeta _t< \alpha _2L^{-\gamma }. \end{aligned}$$

The optimal consumption rates of necessity and luxury goods can be derived during calculating the dual utility functions \({\widetilde{u}}_1\) and \({\widetilde{u}}_2\) (see Lemma 5 and Remark 14).

We have shown that the wealth \(X_t\) is determined as a function of \(y_t\) in Appendices E and I. Since \(y_t\) is a geometric Brownian motion, and hence, strongly Markovian. Thus, \(X_t\) is a strong Markov process. The optimal policies are determined as functions of wealth in Theorem 21, and are strong Markov processes.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Koo, H.K., Roh, KH. & Shin, Y.H. Optimal consumption/investment and retirement with necessities and luxuries. Math Meth Oper Res 94, 281–317 (2021). https://doi.org/10.1007/s00186-021-00758-6

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00186-021-00758-6

Keywords

Navigation