This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Sensitivity of Optimal Retirement Problem to Liquidity Constraints

Guodong Ding1, Daniele Marazzina1,2
1 Department of Mathematics, Politecnico di Milano
Piazza Leonardo da Vinci 32, I-20133, Milano, Italy
2 Corresponding Author, daniele.marazzina@polimi.it

Abstract: In this work we analytically solve an optimal retirement problem, in which the agent optimally allocates the risky investment, consumption and leisure rate to maximise a gain function characterised by a power utility function of consumption and leisure, through the duality method. We impose different liquidity constraints over different time spans and conduct a sensitivity analysis to discover the effect of this kind of constraint.

Keywords: Liquidity Constraints, Retirement Stopping Time, Consumption-Portfolio-Leisure Controls, Duality Method, Variational Inequalities

1 Introduction

We study a stochastic control problem involving the consumption-portfolio-leisure policy and the optimal stopping time of retirement. By determining the continuous and stopping regions of the corresponding optimal stopping time problem, we prove that the optimal retirement time is the first hitting time of the wealth process X(t)X(t) upward to a critical wealth boundary. We implement different liquidity constraints over different time spans, which are X(t)RpreX(t)\geq R_{pre} and X(t)RpostX(t)\geq R_{post} separately for pre- and post-retirement periods. The numerical analysis shows that the wealth boundary triggering the retirement is decreasing to RpreR_{pre} but increasing to RpostR_{post}. The additional retirement option impels the agent to consume less and invest more as the wealth approaches the retirement boundary, and this incentive becomes weaker as RpreR_{pre} decreases.

The considered retirement mechanism is directly referred to [1, 2]. More precisely, [1] studied the optimal retirement model regarding the consumption-portfolio-leisure strategy, in which the leisure rate is limited to the binomial choice. [2] investigated a more complex optimization problem that endows the agent the flexibility in labour supply in the context of retirement planning. We extend their research and adopt a different utility function, a power utility function, as in [3], instead of the Constant Elasticity of Substitution (CES) function. Additionally, compared to [2], other extensions are i) the introduction of a continuous debt repayment the agent should face, ii) the different liquidity constraints before and after retirement, which is the main contribution of this work.

2 Problem Formulation

We deal with a financial market in which two kinds of investment are provided: the money market, concerning a fixed risk-free rate r>0r>0, and a risky asset, which dynamics is described by the stochastic differential equation dS(t)=μS(t)dt+σS(t)dB(t),S(0)=S0dS(t)=\mu S(t)dt+\sigma S(t)dB(t),\,S(0)=S_{0}, with μ\mu and σ\sigma representing the constant drift and diffusion coefficients. B(t)B(t) represents a standard Brownian motion on the filtered probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}), and {t,0t<}\{\mathcal{F}_{t},0\leq t<\infty\} is the augmented natural filtration on B(t)B(t). Moreover, by introducing the market price of risk as θμrσ\theta\triangleq\frac{\mu-r}{\sigma}, we can define the state-price density process as H(t)ξ(t)Z~(t)H(t)\triangleq\xi(t)\tilde{Z}(t) following [4], where ξ(t)ert\xi(t)\triangleq e^{-rt} and Z~(t)eθ22tθB(t)\tilde{Z}(t)\triangleq e^{-\frac{\theta^{2}}{2}t-\theta B(t)} indicate the discount process and an exponential martingale, respectively. Then we define the equivalent martingale measure ~\tilde{\mathbb{P}} by ~(A)𝔼[Z~(t)𝕀A]\tilde{\mathbb{P}}(A)\triangleq\mathbb{E}\left[\tilde{Z}(t)\mathbb{I}_{A}\right], At\forall A\in\mathcal{F}_{t}. Based on the Girsanov Theorem, a standard Brownian motion under ~\tilde{\mathbb{P}} measure can be defined as B~(t)B(t)+θt\tilde{B}(t)\triangleq B(t)+\theta t, t0\forall t\geq 0.

We now describe the optimization problem. The agent needs to optimally allocate the consumption c(t)c(t), the amount of money for the risky investment π(t)\pi(t) and the leisure rate l(t)l(t). The sum of labour and leisure rates equals the constant L¯\bar{L}. Furthermore, denoting the retirement time as τ\tau, the retirement mechanism can be elaborated as: 0l(t)L<L¯0\leq l(t)\leq L<\bar{L} on 0tτ0\leq t\leq\tau, i.e., the leisure rate, as the complement of labour rate, is upper bounded for keeping the employment state; and l(t)L¯l(t)\equiv\bar{L} on t>τt>\tau, since the agent enjoys the entire leisure L¯\bar{L} after declaring retirement. Then the dynamics of the wealth process X(t)X(t), i.e., the state variable of the optimization, is

dX(t)=[rX(t)+π(t)(μr)c(t)d+w(L¯l(t))]dt+σπ(t)dB(t),t0,dX(t)=\left[rX(t)+\pi(t)(\mu-r)-c(t)-d+w(\bar{L}-l(t))\right]dt+\sigma\pi(t)dB(t),\ \forall t\geq 0,

dd and ww are the constant debt repayment and the wage rate, respectively. The initial wealth is X(0)=xX(0)=x.
The considered optimal retirement problem (P)(P) is

V(x)sup(τ,{c(t),π(t),l(t)})𝒜(x)J(x;c,π,l,τ)=sup(τ,{c(t),π(t),l(t)})𝒜(x)𝔼[0eγtu(c(t),l(t))𝑑t],V(x)\triangleq\sup_{(\tau,\{c(t),\pi(t),l(t)\})\in\mathcal{A}(x)}J(x;c,\pi,l,\tau)=\sup_{(\tau,\{c(t),\pi(t),l(t)\})\in\mathcal{A}(x)}\mathbb{E}\left[\int_{0}^{\infty}e^{-\gamma t}u(c(t),l(t))dt\right], (PP)

in which γ\gamma is the subjective discount rate, and the utility is characterized by a power function

u(c,l)=(cδl1δ)1kδ(1k), 0<δ<1,k>1.u(c,l)\!=\!\frac{\left(c^{\delta}l^{1-\delta}\right)^{1-k}}{\delta(1-k)},\ 0\!<\!\delta\!<\!1,\ k\!>\!1.

The admissible control set 𝒜(x)\mathcal{A}(x) follows the standard definition, e.g., [5, Definition 2.1], imposing liquidity constraints: X(t)RpreX(t)\geq R_{pre} for 0t<τ0\leq t<\tau, X(τ)RpreRpostX(\tau)\geq R_{pre}\vee R_{post}, and X(t)RpostX(t)\geq R_{post} for t>τt>\tau a.s.. Notice that we must impose RpredwL¯rR_{pre}\geq\frac{d-w\bar{L}}{r} and RpostdrR_{post}\geq\frac{d}{r} to have the existence of an admissible solution, where dwL¯r\frac{d-w\bar{L}}{r} represents the discounted value of the full debt repayment minus the maximum amount to borrow against the future labour income (in the pre-retirement period).

3 Solution of Optimization Problem

Defining JPR(X(τ);c,π)𝔼[τeγ(sτ)u(c(s),L¯)𝑑s|τ]J_{\scriptscriptstyle PR}(X(\tau);c,\pi)\triangleq\mathbb{E}\left[\left.\int_{\tau}^{\infty}e^{-\gamma(s-\tau)}u(c(s),\bar{L})ds\right|\mathcal{F}_{\tau}\right], the gain function of Problem (P)(P) can be rewritten as the expectation of two separated terms representing the pre- and post-retirement part

J(x;c,π,l,τ)=𝔼[0τeγtu(c(t),l(t))𝑑t+eγτJPR(X(τ);c,π)],J(x;c,\pi,l,\tau)\!=\!\mathbb{E}\left[\!\int_{0}^{\tau}\!e^{\!-\!\gamma t}u(c(t),l(t))dt+e^{\!-\!\gamma\tau}\!J_{\scriptscriptstyle PR}(X(\tau);c,\pi)\right],

where the subscript PR\scriptstyle PR indicates that the corresponding variables and functions are related to the post-retirement problem.

The solutions of the pre- and post-retirement part are based on similar techniques, therefore in this letter we only report the solution of the post-retirement part, referring to the Online Appendix, Section A, for details. Depending on the value of RpostR_{post}, the solution of the post-retirement problem is divided into two different cases: one is Rpost=drR_{post}=\frac{d}{r}, in which the liquidity constraint has no restriction on the optimization, and the other is Rpost>drR_{post}>\frac{d}{r}, with the optimal solution being binded by the liquidity constraint.

Lemma 3.1.

The post-retirement value function

U(x)sup{c(t),π(t)}JPR(x;c,π),U(x)\triangleq\sup_{\{c(t),\pi(t)\}}J_{\scriptscriptstyle PR}(x;c,\pi),

for xRpostx\geq R_{post}, is given by:

U(x)={(xdr)δ(1k)K11δ(1k)L¯(1k)(1δ)1δ(1k),ifRpost=dr,B2,PR(λPR)n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)(λPR)δ(1k)δ(1k)1drλPR+λPRx,ifRpost>dr.U\left(x\right)=\begin{cases}\left(x-\frac{d}{r}\right)^{\delta(1-k)}K_{1}^{1-\delta(1-k)}\bar{L}^{(1-k)(1-\delta)}\frac{1}{\delta(1-k)},&\mbox{if}\quad R_{post}=\frac{d}{r},\\ B_{2,\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*})^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}(\lambda_{\scriptscriptstyle PR}^{*})^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}\lambda_{\scriptscriptstyle PR}^{*}+\lambda_{\scriptscriptstyle PR}^{*}x,&\mbox{if}\quad R_{post}>\frac{d}{r}.\end{cases}

The Legendre-Fenchel transform of U(x)U(x), U~(z)supxRpost[U(x)zx]\tilde{U}(z)\triangleq\sup\limits_{x\geq R_{post}}[U(x)-zx], is:

  • U~(z)=1δ(1k)δ(1k)zδ(1k)δ(1k)1K1L¯(1k)(1δ)1δ(1k)drz\tilde{U}(z)=\frac{1-\delta(1-k)}{\delta(1-k)}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}-\frac{d}{r}z, z>0z>0, if Rpost=drR_{post}\!=\!\frac{d}{r};

  • U~(z)={B2,PRz^PRn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z^PRδ(1k)δ(1k)1drz^PRRpost(zz^PR),zz^PR,B2,PRzn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz,0<z<z^PR,\tilde{U}(z)\!=\!\begin{cases}\!B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}}\!+\!\frac{1\!-\!\delta(1\!-\!k)}{\delta(1\!-\!k)}K_{1}\!\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!\!\!\!-\!\frac{d}{r}\hat{z}_{\scriptscriptstyle PR}\!\!-R_{post}(z\!-\!\hat{z}_{\scriptscriptstyle PR}),&z\!\geq\!\hat{z}_{\scriptscriptstyle PR},\\ \!B_{2,\scriptscriptstyle PR}z^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}z,&0\!\!<\!z\!\!<\!\hat{z}_{\scriptscriptstyle PR},\end{cases} if Rpost>drR_{post}\!>\!\frac{d}{r}.

Proof.

See the Online Appendix A for the proof and the definition of the constants λPR,K1,n2,z^PR\lambda_{\scriptscriptstyle PR}^{*},\,K_{1},\,n_{2},\,\hat{z}_{\scriptscriptstyle PR} and B2,PRB_{2,PR}. ∎

3.1 Pre-retirement Part

Based on the dynamic programming principle, we can only consider a subset of the admissible control set of Problem (P)(P), that is 𝒜1(x)𝒜(x)\mathcal{A}_{1}(x)\subset\mathcal{A}(x), in which any policy achieves the maximum of the post-retirement problem’s gain function. Hence we have

V(x)=sup(τ,{c(t),π(t),l(t)})𝒜1(x)𝔼[0τeγtu(c(t),l(t))𝑑t+eγτU(Xx,c,π,l(τ))],V(x)=\sup_{\left(\tau,\{c(t),\pi(t),l(t)\}\right)\in\mathcal{A}_{1}(x)}\mathbb{E}\left[\int_{0}^{\tau}e^{-\gamma t}u(c(t),l(t))dt+e^{-\gamma\tau}U\left(X^{x,c,\pi,l}(\tau)\right)\right],

where U(Xx,c,π,l(τ))sup{c(t),π(t),l(t)}𝒜1(x)𝔼[τeγ(sτ)u(c(s),L¯)𝑑s|τ]U\left(X^{x,c,\pi,l}(\tau)\right)\!\triangleq\!\!\sup\limits_{\{c(t),\pi(t),l(t)\}\in\mathcal{A}_{1}(x)}\!\!\mathbb{E}\!\left[\!\left.\int_{\tau}^{\infty}\!\!e^{\!-\!\gamma(s\!-\!\tau)}u(c(s),\bar{L})ds\right|\mathcal{F}_{\tau}\right] is given in the previous lemma.

We first define an admissible control set corresponding to a fixed stopping time τ𝒯\tau\in\mathcal{T}, with 𝒯\mathcal{T} representing the set of t\mathcal{F}_{t}-stopping times, as 𝒜τ(x){{c(t),π(t),l(t)}:(τ,{c(t),π(t),l(t)})𝒜(x)}\mathcal{A}_{\tau}(x)\triangleq\{\{c(t),\pi(t),l(t)\}:\left(\tau,\{c(t),\pi(t),l(t)\}\right)\in\mathcal{A}(x)\}, and the utility maximization problem

Vτ(x)sup{c(t),π(t),l(t)}𝒜τ(x)J(x;c,π,l,τ).V_{\tau}(x)\triangleq\sup\limits_{\{c(t),\pi(t),l(t)\}\in\mathcal{A}_{\tau}(x)}J(x;c,\pi,l,\tau). (PτP_{\tau})

Then, Problem (P)(P) is converted into an optimal stopping time problem, that is

V(x)=supτ𝒯Vτ(x).V(x)=\sup\limits_{\tau\in\mathcal{T}}V_{\tau}(x).

Similar to the post-retirement problem, the primal optimization problem’s solution depends on the value of RpreR_{pre}, which prompts us to solve it in two different cases. Before the discussion, we follow [5, Proposition 2.1] to provide the pre-retirement budget constraint, that is:

𝔼[0τH(t)(c(t)+d+wl(t)wL¯)𝑑t+H(τ)X(τ)]x.\mathbb{E}\left[\int_{0}^{\tau}H(t)\left(c(t)+d+wl(t)-w\bar{L}\right)dt+H(\tau)X(\tau)\right]\leq x. (3.1)

Additionally, we define the Legendre-Fenchel transform of u(c,l)u(c,l) by

u~(y)supc0, 0lL[u(c,l)(c+wl)y].\tilde{u}(y)\!\triangleq\!\!\!\sup\limits_{c\geq 0,\,0\leq l\leq L}\!\!\left[u(c,l)\!-\!(c\!+\!wl)y\right].

3.1.1 Duality Approach with Rpre=dwL¯rR_{pre}=\frac{d-w\bar{L}}{r}

Following the method from [6, Section 6], we first deduce an inequality of J(x;c,π,l,τ)J(x;c,\pi,l,\tau) by introducing a Lagrange multiplier λ>0\lambda>0 and using the budget constraint (3.1),

J(x;c,π,l,τ)𝔼[0τeγt(u~(λeγtH(t))(dwL¯)λeγtH(t))𝑑t+eγτU~(λeγτH(τ))]+λx.J(x;c,\pi,l,\tau)\leq\mathbb{E}\left[\int_{0}^{\tau}e^{-\gamma t}\left(\tilde{u}(\lambda e^{\gamma t}H(t))-(d-w\bar{L})\lambda e^{\gamma t}H(t)\right)dt+e^{-\gamma\tau}\tilde{U}(\lambda e^{\gamma\tau}H(\tau))\right]+\lambda x.

The inequality turns to equality if and only if the conditions

c(t)+wl(t)=u~(λeγtH(t)),t[0,τ],X(τ)=U~(λeγτH(τ)),a.s.,c(t)+wl(t)=-\tilde{u}^{\prime}(\lambda e^{\gamma t}H(t)),\quad\forall t\in[0,\tau],\quad X(\tau)=-\tilde{U}^{\prime}(\lambda e^{\gamma\tau}H(\tau)),\quad\mbox{a.s.,}

and 𝔼[0τ(c(t)+wl(t)+dwL¯)H(t)𝑑t+X(τ)H(τ)]=x\mathbb{E}\!\left[\int_{0}^{\tau}\!\left(c(t)\!+\!wl(t)\!+\!d\!-\!w\bar{L}\right)H(t)dt\!+\!X(\tau)H(\tau)\right]\!=\!x hold.

Additionally, Lemma 3.1 implies X(τ)=U~(λeγτH(τ))RpostdwL¯rX(\tau)\!=\!-\tilde{U}^{\prime}(\lambda e^{\gamma\tau}H(\tau))\!\geq\!R_{post}\!\geq\!\frac{d-w\bar{L}}{r}. Then the following lemma shows that under the above conditions, there always exists a portfolio policy to ensure Xx,c,π,l(t)dwL¯r=RpreX^{x,c,\pi,l}(t)\!\geq\!\frac{d-w\bar{L}}{r}\!=\!R_{pre}, t[0,τ]\forall t\!\in\![0,\tau], which implies the liquidity constraint is satisfied automatically.

Lemma 3.2.

For any given initial wealth xRprex\geq R_{pre}, any fixed stopping time τ𝒯\tau\in\mathcal{T}, any τ\mathcal{F}_{\tau}-measurable random variable KK with (KdwL¯r)=1\mathbb{P}(K\geq\frac{d-w\bar{L}}{r})=1 under the \mathbb{P} measure, and any given progressively measurable consumption and leisure processes c(t)0c(t)\geq 0, l(t)0l(t)\geq 0, t0\forall t\geq 0, satisfying 𝔼[0τH(t)(c(t)+wl(t)+dwL¯)𝑑t+H(τ)K]=x\mathbb{E}\left[\int_{0}^{\tau}H(t)(c(t)+wl(t)+d-w\bar{L})dt+H(\tau)K\right]=x, there exists a portfolio process π(t)\pi(t) making Xx,c,π,l(t)dwL¯rX^{x,c,\pi,l}(t)\geq\frac{d-w\bar{L}}{r}, t[0,τ]\forall t\in[0,\tau], and Xx,c,π,l(τ)=KX^{x,c,\pi,l}(\tau)=K hold almost surely.

Proof.

See Online Appendix B. ∎

Moreover, the Lagrange method indicates that Vτ(x)=infλ>0[J~τ(λ)+λx]V_{\tau}(x)=\inf\limits_{\lambda>0}[\tilde{J}_{\tau}(\lambda)+\lambda x] with introducing

J~τ(λ)𝔼[0τeγt(u~(λeγtH(t))(dwL¯)λeγtH(t))𝑑t+eγτU~(λeγτH(τ))],\tilde{J}_{\tau}(\lambda)\triangleq\mathbb{E}\left[\int_{0}^{\tau}e^{-\gamma t}\left(\tilde{u}(\lambda e^{\gamma t}H(t))-(d-w\bar{L})\lambda e^{\gamma t}H(t)\right)dt+e^{-\gamma\tau}\tilde{U}(\lambda e^{\gamma\tau}H(\tau))\right],

and the value function of Problem (P)(P) can be transformed as

V(x)=supτ𝒯Vτ(x)=supτ𝒯infλ>0[J~τ(λ)+λx]infλ>0supτ𝒯[J~τ(λ)+λx]=infλ>0[supτ𝒯J~τ(λ)+λx].V(x)=\sup_{\tau\in\mathcal{T}}V_{\tau}(x)=\sup_{\tau\in\mathcal{T}}\inf_{\lambda>0}[\tilde{J}_{\tau}(\lambda)+\lambda x]\leq\inf_{\lambda>0}\sup_{\tau\in\mathcal{T}}[\tilde{J}_{\tau}(\lambda)+\lambda x]=\inf_{\lambda>0}[\sup_{\tau\in\mathcal{T}}\tilde{J}_{\tau}(\lambda)+\lambda x].

Defining V~(λ)supτ𝒯J~τ(λ)\tilde{V}(\lambda)\triangleq\sup\limits_{\tau\in\mathcal{T}}\tilde{J}_{\tau}(\lambda), [6, Section 8, Theorem 8.5] shows that V(x)=infλ>0[V~(λ)+λx]V(x)=\inf\limits_{\lambda>0}[\tilde{V}(\lambda)+\lambda x] holds under the condition that the function V~(λ)\tilde{V}(\lambda) exists and it is differentiable for any λ>0\lambda>0. Then, the process to solve Problem (P)(P) is divided into two steps: the first is involved in the pure optimal stopping time problem of V~(λ)\tilde{V}(\lambda), and the second step mainly concerns finding the optimal Lagrange multiplier. We begin with the first step and introduce a new process, Z(t)λeγtH(t)Z(t)\triangleq\lambda e^{\gamma t}H(t). Then V~(λ)\tilde{V}(\lambda) can be rewritten as

V~(λ)=supτS𝔼[0τeγt(u~(Z(t))(dwL¯)Z(t))𝑑t+eγτU~(Z(τ))].\tilde{V}(\lambda)=\sup_{\tau\in S}\mathbb{E}\left[\int_{0}^{\tau}e^{-\gamma t}\left(\tilde{u}(Z(t))-(d-w\bar{L})Z(t)\right)dt+e^{-\gamma\tau}\tilde{U}(Z(\tau))\right].

We proceed with a generalized optimal stopping time problem

ϕ(t,z)=supτt𝔼[tτeγs(u~(Z(s))(dwL¯)Z(s))𝑑s+eγτU~(Z(τ))|Z(t)=z],\phi(t,z)=\sup_{\tau\geq t}\mathbb{E}\left[\left.\int_{t}^{\tau}e^{-\gamma s}\left(\tilde{u}(Z(s))-(d-w\bar{L})Z(s)\right)ds+e^{-\gamma\tau}\tilde{U}(Z(\tau))\right|Z(t)=z\right], (3.2)

which shows that V~(λ)=ϕ(0,λ)\tilde{V}(\lambda)=\phi(0,\lambda). The following lemma gives the continuous region and stopping region of the above optimal stopping time problem.

Lemma 3.3.

Considering the optimal stopping time problem (3.2) with the state variable Z(t)Z(t), the continuous region is Ω1={Z(t)>z¯}\Omega_{1}\!=\!\{Z(t)\!>\!\bar{z}\}, the stopping region is Ω2={0<Z(t)z¯}\Omega_{2}\!=\!\{0\!<\!Z(t)\!\leq\!\bar{z}\}, where z¯\bar{z} denotes the boundary that separates Ω1\Omega_{1} and Ω2\Omega_{2}.

Proof.

See Online Appendix C. ∎

Straight follows, with the operator =t+(γr)zz+θ22z22z2\mathcal{L}=\frac{\partial}{\partial t}+(\gamma-r)z\frac{\partial}{\partial z}+\frac{\theta^{2}}{2}z^{2}\frac{\partial^{2}}{\partial z^{2}}, the optimal stopping time problem (3.2) is equivalent to solving the free boundary problem below.
Variational Inequalities: Find a free boundary z¯>0\bar{z}\!>\!0 (Retirement level), and a function ϕ(t,z)C1((0,)×+)C2((0,)×(+{z¯}))\phi(t,z)\!\in\!C^{1}\!\left((0,\infty)\!\times\!\mathbb{R}^{+}\right)\cap C^{2}\left((0,\infty)\times\left(\mathbb{R}^{+}\setminus\{\bar{z}\}\right)\right) satisfying

{(V1)ϕ(t,z)+eγt(u~(z)(dwL¯)z)=0,z>z¯,(V2)ϕ(t,z)+eγt(u~(z)(dwL¯)z)0,0<zz¯,(V3)ϕ(t,z)eγtU~(z),z>z¯,(V4)ϕ(t,z)=eγtU~(z),0<zz¯,\begin{cases}(V1)\quad\mathcal{L}\phi(t,z)+e^{-\gamma t}\left(\tilde{u}(z)-(d-w\bar{L})z\right)=0,&z>\bar{z},\\ (V2)\quad\mathcal{L}\phi(t,z)+e^{-\gamma t}\left(\tilde{u}(z)-(d-w\bar{L})z\right)\leq 0,&0<z\leq\bar{z},\\ (V3)\quad\phi(t,z)\geq e^{-\gamma t}\tilde{U}(z),&z>\bar{z},\\ (V4)\quad\phi(t,z)=e^{-\gamma t}\tilde{U}(z),&0<z\leq\bar{z},\end{cases} (3.3)

for any t0t\geq 0, with the smooth fit conditions ϕ(t,z¯)=eγtU~(z¯)\phi(t,\bar{z})=e^{-\gamma t}\tilde{U}(\bar{z}) and ϕz(t,z¯)=eγtU~(z¯)\frac{\partial\phi}{\partial z}(t,\bar{z})=e^{-\gamma t}\tilde{U}^{\prime}(\bar{z}). The analytical solution of the above inequalities is presented in Online Appendix D.

Once ϕ\phi is computed, we recover V~(λ)=ϕ(0,λ)\tilde{V}(\lambda)=\phi(0,\lambda), and the value function is given by

V(x)=infλ>0[V~(λ)+λx]=V~(λ)+λx,V(x)=\inf_{\lambda>0}[\tilde{V}(\lambda)+\lambda x]=\tilde{V}(\lambda^{*})+\lambda^{*}x,

xx being the initial wealth. The retirement time is the first time the process Z(t)λeγtH(t)Z^{*}(t)\triangleq\lambda^{*}e^{\gamma t}H(t) touches the barrier z¯\bar{z} from above. The optimal strategies are reported at the end of the Online Appendix D.

Remark 3.1.

The optimal retirement time is the first time the process Z(t)Z^{*}(t) touches the lower barrier z¯\bar{z}. The same can be obtained with respect to the wealth level X(t)X(t). In fact, the optimal process ZZ^{*} is connected to the optimal wealth XX by the relation X(t)=v(Z(t))X(t)=-v^{\prime}(Z^{*}(t)), being ϕ(t,z)=eγtv(z),{\phi}(t,z)=e^{-\gamma t}v(z), see the online appendix. The convex property of v()v(\cdot), see [4, Section 3.4, Lemma 4.3], indicates that X(t)X(t) is a decreasing function of Z(t)Z^{*}(t), therefore, in this case the optimal retirement time is the first time the process X(t)X(t) touches an upper barrier x¯=v(z¯)\bar{x}=-v^{\prime}(\bar{z}).

3.1.2 Duality Approach with Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r}

Before proceeding to solve the problem, we present the following proposition to construct expectation form of the liquidity constraint related to X(t)RpreX(t)\!\geq\!R_{pre}, t[0,τ]\forall t\in[0,\tau].

Proposition 3.1.

The liquidity constraint of the considered problem is

𝔼[tτH(s)H(t)(c(s)+d+wl(s)wL¯)𝑑s+H(τ)H(t)X(τ)|t]Rpre,t[0,τ].\mathbb{E}\left[\left.\int_{t}^{\tau}\frac{H(s)}{H(t)}\left(c(s)+d+wl(s)-w\bar{L}\right)ds+\frac{H(\tau)}{H(t)}X(\tau)\right|\mathcal{F}_{t}\right]\geq R_{pre},\quad\forall t\in[0,\tau]. (3.4)
Proof.

See [5, Proposition 4.1]. ∎

Considering the budget and liquidity constraints, (3.1) and (3.4), and introducing a Lagrange multiplier λ>0\lambda>0 and a non-increasing process D(t)0D(t)\geq 0 [6, 7], the following inequality is obtained:

J(x;c,π,l,τ)𝔼[0τeγt(u~(λD(t)eγtH(t))(dwL¯)λeγtD(t)H(t))𝑑t+eγτU~(λD(τ)eγτH(τ))]+λ𝔼[0τRpreH(t)𝑑D(t)]+λx,\begin{split}J(x;c,\pi,l,\tau)&\!\leq\!\mathbb{E}\!\left[\!\int_{0}^{\tau}\!e^{\!-\!\gamma t}\left(\tilde{u}(\lambda D(t)e^{\gamma t}H(t))\!-\!(d\!-\!w\bar{L})\lambda e^{\gamma t}D(t)H(t)\right)\!dt\!+\!e^{\!-\!\gamma\tau}\tilde{U}(\lambda D(\tau)e^{\gamma\tau}\!H(\tau))\!\right]\\ &\qquad+\lambda\mathbb{E}\left[\int_{0}^{\tau}R_{pre}H(t)dD(t)\right]+\lambda x,\end{split}

which inspires us to define the dual individual’s shadow price problem

V~τ(λ)infD(t)𝒟𝔼[0τeγt(u~(λD(t)eγtH(t))(dwL¯)λeγtD(t)H(t))dt+eγτU~(λD(τ)eγτH(τ))]+λ𝔼[0τRpreH(t)dD(t)],\begin{split}\tilde{V}_{\tau}(\lambda)&\triangleq\inf_{D(t)\in\mathcal{D}}\mathbb{E}\bigg{[}\int_{0}^{\tau}e^{-\gamma t}\left(\tilde{u}(\lambda D(t)e^{\gamma t}H(t))-(d-w\bar{L})\lambda e^{\gamma t}D(t)H(t)\right)dt\\ &\qquad+e^{-\gamma\tau}\tilde{U}(\lambda D(\tau)e^{\gamma\tau}H(\tau))\bigg{]}+\lambda\mathbb{E}\left[\int_{0}^{\tau}R_{pre}H(t)dD(t)\right],\end{split} (SτS_{\tau})

where 𝒟\mathcal{D} is the set of non-negative, non-increasing and progressively measurable processes. Then we establish the duality between Problem (Sτ)(S_{\tau}) and (Pτ)(P_{\tau}).

Theorem 3.1.

(Duality Theorem) Suppose D(t)D^{*}(t) is the optimal solution to Problem (Sτ)(S_{\tau}), then c(t)+wl(t)=u~(Z(t))c^{*}(t)\!+\!wl^{*}(t)\!=\!-\!\tilde{u}^{\prime}(Z^{*}(t)) and Xx,c,π,l(τ)=U~(Z(τ))X^{x,c^{*},\pi^{*},l^{*}}(\tau)\!=\!-\tilde{U}^{\prime}(Z^{*}(\tau)) coincide with the optimal solution of Problem (Pτ)(P_{\tau}), and there exists Vτ(x)=infλ>0[V~τ(λ)+λx]V_{\tau}(x)\!=\!\inf\limits_{\lambda>0}\!\left[\!\tilde{V}_{\tau}(\lambda)\!+\!\lambda x\right], xRpre\forall x\!\geq\!R_{pre}. Here Z(t)=λeγtD(t)H(t)Z^{*}(t)=\lambda^{*}e^{\gamma t}D^{*}(t)H(t), where λ\lambda^{*} and D(t)D^{*}(t) are the parameters λ\lambda and D(t)D(t) giving the infimum.

Proof.

See [5, Theorem 4.1]. ∎

This duality theorem allows us to link Problem (P)(P) with the shadow price problem through

V(x)=supτ𝒯Vτ(x)=supτ𝒯infλ>0[V~τ(λ)+λx]infλ>0supτ𝒯[V~τ(λ)+λx]=infλ>0[supτ𝒯V~τ(λ)+λx].V(x)=\sup_{\tau\in\mathcal{T}}V_{\tau}(x)=\sup_{\tau\in\mathcal{T}}\inf_{\lambda>0}[\tilde{V}_{\tau}(\lambda)+\lambda x]\leq\inf_{\lambda>0}\sup_{\tau\in\mathcal{T}}[\tilde{V}_{\tau}(\lambda)+\lambda x]=\inf_{\lambda>0}[\sup_{\tau\in\mathcal{T}}\tilde{V}_{\tau}(\lambda)+\lambda x].

Defining V~(λ)supτ𝒯V~τ(λ)\tilde{V}(\lambda)\triangleq\sup\limits_{\tau\in\mathcal{T}}\tilde{V}_{\tau}(\lambda), [6, Section 8, Theorem 8.5] indicates that the last inequality takes the equal sign with the condition that V~(λ)\tilde{V}(\lambda) exists and is differentiable for any λ>0\lambda>0. Thereafter, the objective optimization problem can be divided into two parts:

{V~(λ)=supτSV~τ(λ),V(x)=infλ>0[V~(λ)+λx]V~(λ)+λx.\begin{cases}\tilde{V}(\lambda)=\sup\limits_{\tau\in S}\tilde{V}_{\tau}(\lambda),\\ V(x)=\inf\limits_{\lambda>0}[\tilde{V}(\lambda)+\lambda x]\triangleq\tilde{V}(\lambda^{*})+\lambda^{*}x.\end{cases}

We now consider the technique of [8] and insert an assumption on the process D(t)D(t) for acquiring a closed-form solution.

Assumption 3.1.

The non-increasing process D(t)D(t) is absolutely continuous with respect to tt. Hence, there is a non-negative process ψ(t)\psi(t) such that dD(t)=ψ(t)D(t)dtdD(t)=-\psi(t)D(t)dt.

Then, by means of a new defined process Z(t)λD(t)eγtH(t)Z(t)\triangleq\lambda D(t)e^{\gamma t}H(t), the value function of the individual’s shadow price problem can be written as

V~τ(λ)=infψ(t)0𝔼[0τeγt(u~(Z(t))(dwL¯)Z(t)Rpreψ(t)Z(t))𝑑t+eγτU~(Z(τ))],\tilde{V}_{\tau}(\lambda)=\inf_{\psi(t)\geq 0}\mathbb{E}\left[\int_{0}^{\tau}e^{-\gamma t}\left(\tilde{u}(Z(t))-(d-w\bar{L})Z(t)-R_{pre}\psi(t)Z(t)\right)dt+e^{-\gamma\tau}\tilde{U}(Z(\tau))\right],

where ψ(t)\psi(t) is the control variable, and Z(t)Z(t) is the state variable. Introducing a generalized problem

ϕ(t,z)supτtinfψ(t)0𝔼[tτeγs(u~(Z(s))(dwL¯)Z(s)Rpreψ(s)Z(s))𝑑s+eγτU~(Z(τ))|Z(t)=z],\phi(t,z)\!\triangleq\!\sup_{\tau\geq t}\!\inf_{\psi(t)\geq 0}\!\mathbb{E}\!\left[\!\left.\int_{t}^{\tau}\!\!\!\!\!e^{\!-\!\gamma s}\!\!\left(\tilde{u}(Z(s))\!-\!(d\!-\!w\bar{L})Z(s)\!-\!R_{pre}\psi(s)Z(s)\right)ds\!+\!e^{\!-\!\gamma\tau}\!\tilde{U}(Z(\tau))\right|\!Z(t)\!=\!z\!\right],

the solution of V~(λ)\tilde{V}(\lambda) is turned to ϕ(t,z)\phi(t,z) with V~(λ)=ϕ(0,λ)\tilde{V}(\lambda)\!=\!\phi(0,\lambda). We first handle the infimum part by defining

ϕinf(t,z)infψ(t)>0𝔼[tτeγs(u~(Z(s))(dwL¯)Z(s)Rpreψ(s)Z(s))𝑑s+eγτU~(Z(τ))|Z(t)=z].\phi_{\scriptscriptstyle inf}(t,z)\!\triangleq\!\inf_{\psi(t)>0}\mathbb{E}\!\left[\!\left.\int_{t}^{\tau}\!\!\!\!\!e^{\!-\!\gamma s}\!\left(\tilde{u}(Z(s))\!-\!(d\!-\!w\bar{L})Z(s)\!-\!R_{pre}\psi(s)Z(s)\right)\!ds\!+\!e^{\!-\!\gamma\tau}\!\tilde{U}(Z(\tau))\right|\!Z(t)\!=\!z\right].

The corresponding Bellman equation is

minψ0{ϕinf(t,z)+eγt(u~(z)(dwL¯)z)ψz[ϕinfz(t,z)+Rpreeγt]}=0.\min_{\psi\geq 0}\left\{\mathcal{L}\phi_{\scriptscriptstyle inf}(t,z)+e^{-\gamma t}\left(\tilde{u}(z)-(d-w\bar{L})z\right)-\psi z\left[\frac{\partial\phi_{\scriptscriptstyle inf}}{\partial z}(t,z)+R_{pre}e^{-\gamma t}\right]\right\}=0.

The optimum ψ\psi^{*} has the following characterization,

  • ϕinfz(t,z)+Rpreeγt=0ψ0\frac{\partial\phi_{\scriptscriptstyle inf}}{\partial z}(t,z)\!+\!R_{pre}e^{\!-\!\gamma t}\!=\!0\Rightarrow\psi^{*}\!\!\geq\!0 and ϕz(t,z)=ϕinfz(t,z)=Rpreeγt,zz^\frac{\partial\phi}{\partial z}(t,z)=\frac{\partial\phi_{\scriptscriptstyle inf}}{\partial z}(t,z)=-R_{pre}e^{-\gamma t},\quad z\geq\hat{z}.

  • ϕinfz(t,z)+Rpreeγt0ψ=0\frac{\partial\phi_{\scriptscriptstyle inf}}{\partial z}(t,z)\!+\!R_{pre}e^{\!-\!\gamma t}\!\leq\!0\Rightarrow\psi^{*}\!\!=\!0, then ϕ(t,z)\phi(t,z) switches to a pure optimal stopping time problem,

    ϕ(t,z)=supτt𝔼[tτeγs(u~(Z(s))(dwL¯)Z(s))𝑑s+eγτU~(Z(τ))|Z(t)=z],\phi(t,z)\!=\!\sup_{\tau\geq t}\mathbb{E}\!\left[\!\left.\int_{t}^{\tau}\!\!e^{\!-\!\gamma s}\left(\tilde{u}(Z(s))\!-\!(d\!-\!w\bar{L})Z(s)\right)ds\!+\!e^{\!-\!\gamma\tau}\tilde{U}(Z(\tau))\right|Z(t)\!=\!z\right],

    which has the same form as (3.2) but applies to the interval 0<z<z^0<z<\hat{z}.

Lemma 3.3 can be easily extended also in this case, therefore the optimal retirement time is the first time the process Z(t)Z^{*}(t) touches the lower barrier z¯\bar{z}. Therefore, we need to compare the value of z¯\bar{z} and z^\hat{z}, and split the discussion into two cases: the first one is z¯<z^\bar{z}<\hat{z}, which corresponds to the case where the liquidity constraint boundary, RpreR_{pre}, is lower than the retirement threshold.

Variational Inequalities assuming z¯<z^\bar{z}<\hat{z}: Find the free boundaries z¯>0\bar{z}>0 (retirement), z^>0\hat{z}>0 (RpreR_{pre}-wealth level), and a function ϕ(,)C1((0,)×+)C2((0,)×+{z¯}){\phi}(\cdot,\cdot)\in C^{1}((0,\infty)\times\mathbb{R}^{+})\cap C^{2}((0,\infty)\times\mathbb{R}^{+}\setminus\{\bar{z}\}) satisfying

{(V1)ϕz(t,z)+Rpreeγt=0,zz^,(V2)ϕz(t,z)+Rpreeγt0,0<z<z^,(V3)ϕ(t,z)+eγt(u~(z)(dwL¯)z)=0,z¯<z<z^,(V4)ϕ(t,z)+eγt(u~(z)(dwL¯)z)0,0<zz¯,(V5)ϕ(t,z)eγtU~(z),z¯<z<z^,(V6)ϕ(t,z)=eγtU~(z),0<zz¯,\begin{cases}(V1)\quad\frac{\partial{\phi}}{\partial z}(t,z)+R_{pre}e^{-\gamma t}=0,&z\geq\hat{z},\\ (V2)\quad\frac{\partial{\phi}}{\partial z}(t,z)+R_{pre}e^{-\gamma t}\leq 0,&0<z<\hat{z},\\ (V3)\quad\mathcal{L}{\phi}(t,z)+e^{-\gamma t}\left(\tilde{u}(z)-(d-w\bar{L})z\right)=0,&\bar{z}<z<\hat{z},\\ (V4)\quad\mathcal{L}{\phi}(t,z)+e^{-\gamma t}\left(\tilde{u}(z)-(d-w\bar{L})z\right)\leq 0,&0<z\leq\bar{z},\\ (V5)\quad{\phi}(t,z)\geq e^{-\gamma t}\tilde{U}(z),&\bar{z}<z<\hat{z},\\ (V6)\quad{\phi}(t,z)=e^{-\gamma t}\tilde{U}(z),&0<z\leq\bar{z},\end{cases} (3.5)

for any t0t\geq 0, with the smooth fit conditions

ϕz(t,z^)=Rpreeγt,2ϕz2(t,z^)=0,ϕ(t,z¯)=eγtU~(z¯),andϕz(t,z¯)=eγtU~(z¯).\frac{\partial{\phi}}{\partial z}(t,\hat{z})=-R_{pre}e^{-\gamma t},\quad\frac{\partial^{2}{\phi}}{\partial z^{2}}(t,\hat{z})=0,\quad{\phi}(t,\bar{z})=e^{-\gamma t}\tilde{U}(\bar{z}),\quad\mbox{and}\quad\frac{\partial{\phi}}{\partial z}(t,\bar{z})=e^{-\gamma t}\tilde{U}^{\prime}(\bar{z}).

The analytical solution of the variational equation (3.5) is reported in Online Appendix E. Once ϕ\phi and z¯\bar{z} are computed, the value function and the optimal retirement decision can be recovered as in Section 3.1.1. The optimal strategies are reported at the end of the Online Appendix E.

If the first case does not admit a solution, that is, the liquidity constraint boundary RpreR_{pre} is high enough (and larger than RpostR_{post}) to make the agent declare retirement at time 0 for any admissible initial wealth, we deal with an immediate retirement, and therefore V(x)=U(x)V(x)=U(x), and all the optimal strategies are the ones of the post-retirement problem.

4 Numerical Analysis

We now perform the sensitivity analysis to the liquidity constraint boundaries. All the input parameters are reported in Table 4.1. We change the values of RpreR_{pre}, RpostR_{post} and keep all other input parameters consistent with Table 4.1 to discover the different convergence phenomena of retirement wealth threshold concerning the pre- and post-retirement liquidity constraints.

Table 4.1: Input Parameters
δ\delta kk rr μ\mu σ\sigma γ\gamma dd ww RpreR_{pre} RpostR_{post} L¯\bar{L} LL
0.6 3 0.02 0.07 0.15 0.1 0.3 1.5 0 15 1 0.8

Figure 4.1 shows that the retirement wealth threshold x¯\bar{x} is a decreasing function of RpreR_{pre} due to the fact that the agent with higher RpreR_{pre} values prefers to set a lower wealth threshold to make sure entering in retirement ahead of schedule such that getting rid of the restriction caused by RpreR_{pre}. Whereas, the critical wealth level of retirement is increasing with respect to RpostR_{post}. Since the pre-retirement restriction keeps constant, a higher value of RpostR_{post}, which implies a more rigorous circumstance for the post-retirement period, impels the agent to step into retirement with a higher wealth level.

Figure 4.1: Convergence w.r.t. Liquidity Constraint Boundary of Pre- and Post-Retirement Part
Refer to caption

Moreover, we provide figures to illustrate the sensitivity of optimal consumption, portfolio and leisure fractions in terms of xdrx-\frac{d}{r} with respect to different values of RpreR_{pre} and RpostR_{post}. We begin this kind of analysis fixing the value of RpostR_{post} and arranging three values to RpreR_{pre}. The optimal control strategies for different cases are presented in Proposition A.3, Proposition D.1 and Proposition E.1 in the online appendix.

Figure 4.2: Optimal Control Fractions w.r.t. Liquidity Constraint Boundary of Pre-Retirement Part
Refer to caption

In Figure 4.2, RpostR_{post} is set equal to dr=15\frac{d}{r}=15, which implies that the post-retirement part is not restricted by the liquidity constraint. We can observe that the optimal consumption and portfolio fractions suffer a downward jump for various RpreR_{pre} values. This is due to the discontinuity of the leisure rate at the retirement time, which leads to a shrinkage of labour income and reduces the resources allocated to the consumption and investment. In fact, if x>256.6913x>256.6913 (Rpre=60)(R_{pre}=-60), x>164.5320x>164.5320 (Rpre=0)(R_{pre}=0), x>137.4776x>137.4776 (Rpre=10)(R_{pre}=10), i.e., the initial wealth is larger than the retirement threshold x¯\bar{x}, the agent is facing the post-retirement region, with l(t)=L¯=1l^{*}(t)=\bar{L}=1 (full leisure). In addition, it should be noted that for different RpreR_{pre} values, the jump happens at different wealth levels. As also shown in the left plot of Figure 4.1, the agent with a higher RpreR_{pre} value experiences the jump at a lower wealth threshold of retirement x¯\bar{x}. Moreover, since the value of RpostR_{post} keeps identical, the optimal consumption and portfolio fractions of different curves are coincident for the post-retirement part and equal to a constant, in line with the Merton classical problem.

Then we conduct a similar sensitivity analysis with respect to RpostR_{post}. Figure 4.3 shows that the retirement threshold is increasing with the value of RpostR_{post} (x¯=164.5320\bar{x}=164.5320 for Rpost=15R_{post}=15, x¯=171.1993\bar{x}=171.1993 for Rpost=20R_{post}=20, x¯=180.7943\bar{x}=180.7943 for Rpost=25R_{post}=25), in line with the right plot of Figure 4.1, and describes that the optimal control fractions for the post-retirement part of blue dashed and red dotted curves, whose RpostR_{post} values are greater than the boundary dr\frac{d}{r}, i.e., the liquidity constraints impose restrictions on optimal solutions, converge to the ones of the green curve ( Rpost=drR_{post}=\frac{d}{r}) as xx increases. It can be explained by the fact that the liquidity constraint plays a slighter role as the wealth becomes comparably larger and imposes a weaker restriction on the admissible control set. Moreover, we also notice that a high liquidity constraint for the post-retirement part induces the agent to take a large risk (high value of π\pi^{*}) when the retirement threshold is close.

Figure 4.3: Optimal Control Fractions w.r.t. Liquidity Constraint Boundary of Post-Retirement Part
Refer to caption

Finally, we conduct the sensitivity analysis of optimal control strategies to both the liquidity constraint boundary and the retirement option. In Figure 4.4, we fix the value of RpostR_{post} to dr\frac{d}{r} and plot the curves of optimal consumption and portfolio fractions in terms of xdrx-\frac{d}{r} under different situations. The dashed lines represent the optimal control fractions of different RpreR_{pre} values with retirement option, while the solid lines represent the corresponding optimal control fractions without retirement option (and therefore with fixed liquidity constraint R=RpreR=R_{pre}).111The theoretical solutions of optimal consumption-portfolio problem without retirement comes from [5, Section 5] by replacing the liquidity constraint boundary F+ηF+\eta with RpreR_{pre}. From all the dashed lines, we can see that the optimal consumption and portfolio fractions suffer a downward jump for various RpreR_{pre} values. This is due to the discontinuity of the leisure rate at the retirement time, which leads to a shrinkage of labour income and reduces the resources allocated to the consumption and investment. Comparing the solid and dashed lines with the same colour, the agent with the additional retirement option tends to consume less and invest more in the risky asset for the motivation of arriving at the retirement wealth threshold and enjoying the full leisure rate faster. This kind of difference becomes more significant as the wealth approaches the critical level. Furthermore, the degree of this motivation is related to the liquidity constraint boundary. Observing the convexity of the pre-retirement part of different dashed lines, the optimal control fraction with a higher RpreR_{pre} value takes a larger convexity, which is because stricter liquidity constraints give the agent a stronger motivation to achieve the critical wealth level to get rid of this restriction.

Figure 4.4: Optimal Control Fractions w.r.t. Liquidity Constraint Boundary and Retirement Option.
Refer to caption

References

  • [1] E. Farhi and S. Panageas, “Saving and investing for early retirement: A theoretical analysis,” Journal of Financial Economics, vol. 83, no. 1, pp. 87–121, 2007.
  • [2] K. J. Choi, G. Shim, and Y. H. Shin, “Optimal portfolio, consumption-leisure and retirement choice problem with CES utility,” Mathematical Finance: An International Journal of Mathematics, Statistics and Financial Economics, vol. 18, no. 3, pp. 445–472, 2008.
  • [3] G. Shim and Y. H. Shin, “An optimal job, consumption/leisure, and investment policy,” Operations Research Letters, vol. 42, no. 2, pp. 145–149, 2014.
  • [4] I. Karatzas and S. E. Shreve, Methods of mathematical finance, vol. 39. Springer, 1998.
  • [5] G. Ding and D. Marazzina, “Effect of Labour Income on the Optimal Bankruptcy Problem,” arXiv e-prints, p. arXiv:2106.15426, June 2021.
  • [6] I. Karatzas and H. Wang, “Utility maximization with discretionary stopping,” SIAM Journal on Control and Optimization, vol. 39, no. 1, pp. 306–329, 2000.
  • [7] H. He and H. F. Pages, “Labor income, borrowing constraints, and equilibrium asset prices,” Economic Theory, vol. 3, no. 4, pp. 663–696, 1993.
  • [8] M. H. Davis and A. R. Norman, “Portfolio selection with transaction costs,” Mathematics of operations research, vol. 15, no. 4, pp. 676–713, 1990.

Sensitivity of Optimal Retirement Problem to Liquidity Constraints - Online Appendix

Guodong Ding1, Daniele Marazzina1,2
1 Department of Mathematics, Politecnico di Milano
Piazza Leonardo da Vinci 32, I-20133, Milano, Italy
2 Corresponding Author, daniele.marazzina@polimi.it

Appendix A Post Retirement Part

Assuming τ=0\tau=0^{-}, we deal with the post-retirement problem, which is an infinite-time optimization problem with two control variables, the consumption and portfolio processes. Introducing uPR(c)u(c,L¯)=cδ(1k)L¯(1δ)(1k)δ(1k)u_{\scriptscriptstyle PR}(c)\triangleq u(c,\bar{L})=\frac{c^{\delta(1-k)}\bar{L}^{(1-\delta)(1-k)}}{\delta(1-k)}, the corresponding value function, denoted as (PPR)(P_{\scriptscriptstyle PR}), is

VPR(x)sup{c(t),π(t)}𝒜PR(x)JPR(x;c,π).V_{\scriptscriptstyle PR}(x)\triangleq\sup_{\{c(t),\pi(t)\}\in\mathcal{A}_{\scriptscriptstyle PR}(x)}J_{\scriptscriptstyle PR}(x;c,\pi). (PPRP_{\scriptscriptstyle PR})

The admissible control set 𝒜PR(x)\mathcal{A}_{\scriptscriptstyle PR}(x) takes the compatible definition with 𝒜(x)\mathcal{A}(x), except that the condition for stopping time is abolished, and the condition for liquidity constraint is given by X(t)RpostX(t)\!\geq\!R_{post}, a.s., t0\forall t\!\geq\!0. Then we derive the derivative function uPR(c)=cδ(1k)1L¯(1k)(1δ)u_{\scriptscriptstyle PR}^{\prime}(c)=c^{\delta(1-k)-1}\bar{L}^{(1-k)(1-\delta)}, which is positive and strictly decreasing and has the inverse function IPR(z)z1δ(1k)1L¯(1k)(1δ)1δ(1k)I_{\scriptscriptstyle PR}(z)\triangleq z^{\frac{1}{\delta(1-k)-1}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}. Furthermore, referring to [4, Section 3, Definition 4.2], we define the Legendre-Fenchel transform of uPR(z)u_{\scriptscriptstyle PR}(z) as u~PR(z)supc0[uPR(c)cz]\tilde{u}_{\scriptscriptstyle PR}(z)\triangleq\sup\limits_{c\geq 0}\left[u_{\scriptscriptstyle PR}(c)-cz\right], which has the explicit expression

u~PR(z)=uPR(IPR(z))zIPR(z)=1δ(1k)δ(1k)zδ(1k)δ(1k)1L¯(1k)(1δ)1δ(1k).\tilde{u}_{\scriptscriptstyle PR}(z)=u_{\scriptscriptstyle PR}(I_{\scriptscriptstyle PR}(z))-zI_{\scriptscriptstyle PR}(z)=\frac{1-\delta(1-k)}{\delta(1-k)}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}.
Proposition A.1.

The infinite horizon budget constraint of the post-retirement problem is

𝔼[0H(t)(c(t)+d)𝑑t]x.\mathbb{E}\left[\int_{0}^{\infty}H(t)(c(t)+d)dt\right]\leq x. (A.1)
Proof.

The proof can be accomplished directly by replacing l(t)l(t) in [5, Proposition 3.1] with the constant L¯\bar{L}, meanwhile inserting a constant term dd in the integral. ∎

Depending on the value of RpostR_{post}, the solution of Problem (PPR)(P_{\scriptscriptstyle PR}) is divided into two different cases. One is Rpost=drR_{post}=\frac{d}{r}, in which the liquidity constraint has no restriction on the optimization, and the other is Rpost>drR_{post}>\frac{d}{r}, with the optimal solution being binded by the liquidity constraint.

As in [4, Chapter 3, Example 9.22], the optimal wealth process under the condition Rpost=drR_{post}\!=\!\frac{d}{r} is X(t)=(xdr)e11δ(1k)(rγ+θ22)t+θ1δ(1k)B(t)+drX^{*}(t)\!=\!\left(x\!-\!\frac{d}{r}\right)e^{\frac{1}{1\!-\!\delta(1\!-\!k)}\left(r\!-\!\gamma\!+\!\frac{\theta^{2}}{2}\right)t+\frac{\theta}{1\!-\!\delta(1\!-\!k)}B(t)}\!+\!\frac{d}{r}. The optimal consumption-portfolio polices are c(t)=1K1[X(t)dr]c^{*}(t)\!=\!\frac{1}{K_{1}}\!\!\left[\!X^{*}(t)\!-\!\frac{d}{r}\!\right] and π(t)=θσ(1δ(1k))[X(t)dr]\pi^{*}(t)\!=\!\frac{\theta}{\sigma(1\!-\!\delta(1\!-\!k))}\!\!\left[\!X^{*}(t)\!-\!\frac{d}{r}\!\right], with K11δ(1k)γrδ(1k)θ22δ(1k)1δ(1k)>0K_{1}\triangleq\frac{1-\delta(1-k)}{\gamma-r\delta(1-k)-\frac{\theta^{2}}{2}\frac{\delta(1-k)}{1-\delta(1-k)}}>0. And the value function of Problem (PPR)(P_{\scriptscriptstyle PR}) can be obtained as

VPR(x)=(xdr)δ(1k)K11δ(1k)L¯(1k)(1δ)δ(1k).V_{\scriptscriptstyle PR}(x)=\left(x-\frac{d}{r}\right)^{\delta(1-k)}K_{1}^{1-\delta(1-k)}\frac{\bar{L}^{(1-k)(1-\delta)}}{\delta(1-k)}. (A.2)
Remark A.1.

Analogous to the solution of the Merton problem, under the infinite time horizon, the optimal fraction invested in the risky asset in terms of the wealth minus the debt, i.e., π(t)X(t)dr\frac{\pi^{*}(t)}{X^{*}(t)-\frac{d}{r}} keeps constant as θσ(δ(1k)1)=μrσ2(1δ(1k))-\frac{\theta}{\sigma(\delta(1-k)-1)}=\frac{\mu-r}{\sigma^{2}(1-\delta(1-k))}, and the optimal fractional consumption c(t)X(t)dr\frac{c^{*}(t)}{X^{*}(t)-\frac{d}{r}} takes a constant ratio as 1K1\frac{1}{K_{1}}.

Hereafter, we impose a stricter liquidity constraint on the wealth process, X(t)Rpost>drX(t)\geq R_{post}>\frac{d}{r}. The following proposition provides the expectation form of the liquidity constraint, which will be accessible to deduce the duality problem subsequently.

Proposition A.2.

The infinite horizon liquidity constraint of the post-retirement problem is

𝔼[tH(s)H(t)(c(s)+d)𝑑s|t]Rpost.\mathbb{E}\left[\left.\int_{t}^{\infty}\frac{H(s)}{H(t)}(c(s)+d)ds\right|\mathcal{F}_{t}\right]\geq R_{post}. (A.3)
Proof.

See [5, Proposition 3.2]. ∎

Referring to [6, 7], we introduce a real number λPR>0\lambda_{\scriptscriptstyle PR}>0, the Lagrange multiplier, and a non-increasing process DPR(t)0D_{\scriptscriptstyle PR}(t)\geq 0, then rewrite the post-retirement gain function as

JPR(x;c,π)𝔼[0eγt(u~PR(λPReγtH(t)DPR(t))dλPReγtDPR(t)H(t))𝑑t]+λPRx+λPR𝔼[0RpostH(t)𝑑DPR(t)].\begin{split}J_{\scriptscriptstyle PR}(x;c,\pi)\leq\mathbb{E}&\left[\int_{0}^{\infty}e^{-\gamma t}\left(\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}e^{\gamma t}H(t)D_{\scriptscriptstyle PR}(t))-d\lambda_{\scriptscriptstyle PR}e^{\gamma t}D_{\scriptscriptstyle PR}(t)H(t)\right)dt\right]\\ &+\lambda_{\scriptscriptstyle PR}x+\lambda_{\scriptscriptstyle PR}\mathbb{E}\left[\int_{0}^{\infty}R_{post}H(t)dD_{\scriptscriptstyle PR}(t)\right].\end{split}

The derivation of this inequality involves the budget constraint (A.1) and the liquidity constraint (A.3). In line with [7, Section 4], the post-retirement individual’s dual shadow price problem, labelled (SPR)(S_{\scriptscriptstyle PR}), can be defined as

V~PR(λPR)infDPR(t)𝒟𝔼[0eγt(u~PR(λPReγtH(t)DPR(t))dλPReγtDPR(t)H(t))𝑑t]+λPR𝔼[0RpostH(t)𝑑DPR(t)],\begin{split}\tilde{V}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR})\triangleq\inf_{D_{\scriptscriptstyle PR}(t)\in\mathcal{D}}&\mathbb{E}\left[\int_{0}^{\infty}e^{-\gamma t}\left(\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}e^{\gamma t}H(t)D_{\scriptscriptstyle PR}(t))-d\lambda_{\scriptscriptstyle PR}e^{\gamma t}D_{\scriptscriptstyle PR}(t)H(t)\right)dt\right]\\ &+\lambda_{\scriptscriptstyle PR}\mathbb{E}\left[\int_{0}^{\infty}R_{post}H(t)dD_{\scriptscriptstyle PR}(t)\right],\end{split} (SPRS_{\scriptscriptstyle PR})

where 𝒟\mathcal{D} is the set of non-negative, non-increasing and progressively measurable processes. Then the duality between Problem (PPR)(P_{\scriptscriptstyle PR}) and Problem (SPR)(S_{\scriptscriptstyle PR}) is put forward.

Theorem A.1.

(Duality Theorem) Suppose DPR(t)D_{\scriptscriptstyle PR}^{*}(t) is the optimal solution to the dual shadow price problem (SPR)(S_{\scriptscriptstyle PR}), then c(t)=IPR(λPReγtDPR(t)H(t))c^{*}(t)=I_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D_{\scriptscriptstyle PR}^{*}(t)H(t)) is the optimal consumption solution to the problem (PPR)(P_{\scriptscriptstyle PR}). And we have the relation VPR(x)=infλPR>0[V~PR(λPR)+λPRx]V_{\scriptscriptstyle PR}(x)\!=\!\inf\limits_{\lambda_{\scriptscriptstyle PR}>0}[\tilde{V}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR})\!+\!\lambda_{\scriptscriptstyle PR}x], with λPR\lambda_{\scriptscriptstyle PR}^{*} attaining the infimum.

Proof.

See Appendix A.1. ∎

The Duality Theorem enables us to transform the solution of Problem (PPR)(P_{\scriptscriptstyle PR}) to its duality, (SPR)(S_{\scriptscriptstyle PR}). Besides, adopting the technique from [8], the subsequent assumption should be imposed for solving the problem explicitly.

Assumption A.1.

The non-increasing process DPR(t)D_{\scriptscriptstyle PR}(t) is absolutely continuous with respect to t. Hence, there exists a process ψPR(t)\psi_{\scriptscriptstyle PR}(t) such that dDPR(t)=ψPR(t)DPR(t)dtdD_{\scriptscriptstyle PR}(t)=-\psi_{\scriptscriptstyle PR}(t)D_{\scriptscriptstyle PR}(t)dt.

Introducing ZPR(t)λPReγtDPR(t)H(t)Z_{\scriptscriptstyle PR}(t)\!\triangleq\!\lambda_{\scriptscriptstyle PR}e^{\gamma t}\!D_{\scriptscriptstyle PR}(t)H(t), the value function of Problem (SPR)(S_{\scriptscriptstyle PR}) is converted into

V~PR(λPR)=infψPR(t)0𝔼[0eγt(u~PR(ZPR(t))dZPR(t)ψPR(t)ZPR(t)Rpost)𝑑t].\tilde{V}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR})=\inf_{\psi_{\scriptscriptstyle PR}(t)\geq 0}\mathbb{E}\left[\int_{0}^{\infty}e^{-\gamma t}(\tilde{u}_{\scriptscriptstyle PR}(Z_{\scriptscriptstyle PR}(t))-dZ_{\scriptscriptstyle PR}(t)-\psi_{\scriptscriptstyle PR}(t)Z_{\scriptscriptstyle PR}(t)R_{post})dt\right].

Then we define

ϕPR(t,z)infψPR(t)0𝔼[teγs(u~PR(ZPR(s))dZPR(s)ψPR(s)ZPR(s)Rpost)𝑑s|ZPR(t)=z],\phi_{\scriptscriptstyle PR}(t,z)\!\triangleq\!\inf\limits_{\psi_{\scriptscriptstyle PR}(t)\geq 0}\!\mathbb{E}\left[\left.\!\int_{t}^{\infty}\!e^{\!-\!\gamma s}\left(\tilde{u}_{\scriptscriptstyle PR}(Z_{\scriptscriptstyle PR}(s))\!-\!dZ_{\scriptscriptstyle PR}(s)\!-\!\psi_{\scriptscriptstyle PR}(s)Z_{\scriptscriptstyle PR}(s)R_{post}\right)ds\right|Z_{\scriptscriptstyle PR}(t)\!=\!z\right],

and observe that V~PR(λPR)=ϕPR(0,λPR)\tilde{V}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR})=\phi_{\scriptscriptstyle PR}(0,\lambda_{\scriptscriptstyle PR}). The associated Bellman equation to ϕPR(t,z)\phi_{\scriptscriptstyle PR}(t,z) follows

minψPR0{~ϕPR(t,z)+eγt(u~PR(z)dz)ψPRz[ϕPRz(t,z)+eγtRpost]}=γϕPR(t,z),\min_{\psi_{\scriptscriptstyle PR}\geq 0}\left\{\tilde{\mathcal{L}}\phi_{\scriptscriptstyle PR}(t,z)+e^{-\gamma t}(\tilde{u}_{\scriptscriptstyle PR}(z)-dz)-\psi_{\scriptscriptstyle PR}z\left[\frac{\partial\phi_{\scriptscriptstyle PR}}{\partial z}(t,z)+e^{-\gamma t}R_{post}\right]\right\}=\gamma\phi_{\scriptscriptstyle PR}(t,z),

with the operator ~=(γr)zz+12θ2z22z2\tilde{\mathcal{L}}=(\gamma-r)z\frac{\partial}{\partial z}+\frac{1}{2}\theta^{2}z^{2}\frac{\partial^{2}}{\partial z^{2}}. From the characterization of optimum ψPR\psi_{\scriptscriptstyle PR}^{*}:

ϕPRz(t,z)+eγtRpost=0ψPR0;ϕPRz(t,z)+eγtRpost0ψPR=0,\frac{\partial\phi_{\scriptscriptstyle PR}}{\partial z}(t,z)+e^{-\gamma t}R_{post}=0\Rightarrow\psi_{\scriptscriptstyle PR}^{*}\geq 0;\qquad\frac{\partial\phi_{\scriptscriptstyle PR}}{\partial z}(t,z)+e^{-\gamma t}R_{post}\leq 0\Rightarrow\psi_{\scriptscriptstyle PR}^{*}=0,

the Bellman equation is equivalent to

min{~ϕPR(t,z)γϕPR(t,z)+eγt(u~PR(z)dz),[ϕPRz(t,z)+eγtRpost]}=0,\min\left\{\tilde{\mathcal{L}}\phi_{\scriptscriptstyle PR}(t,z)-\gamma\phi_{\scriptscriptstyle PR}(t,z)+e^{-\gamma t}(\tilde{u}_{\scriptscriptstyle PR}(z)-dz),-\left[\frac{\partial\phi_{\scriptscriptstyle PR}}{\partial z}(t,z)+e^{-\gamma t}R_{post}\right]\right\}=0,

which results in the consequent modified variational inequalities: Find a free boundary z^PR>0\hat{z}_{\scriptscriptstyle PR}>0, which makes RpostR_{post}-wealth level, and a function ϕPR(,)C2((0,)×+)\phi_{\scriptscriptstyle PR}(\cdot,\cdot)\in C^{2}((0,\infty)\times\mathbb{R}^{+}) satisfying

{(V1)ϕPRz(t,z)+eγtRpost=0,zz^PR,(V2)ϕPRz(t,z)+eγtRpost0,0<z<z^PR,(V3)~ϕPR(t,z)γϕPR(t,z)+eγt(u~PR(z)dz)=0,0<z<z^PR,(V4)~ϕPR(t,z)γϕPR(t,z)+eγt(u~PR(z)dz)0,zz^PR,\begin{cases}(V1)\quad\frac{\partial\phi_{\scriptscriptstyle PR}}{\partial z}(t,z)+e^{-\gamma t}R_{post}=0,&z\geq\hat{z}_{\scriptscriptstyle PR},\\ (V2)\quad\frac{\partial\phi_{\scriptscriptstyle PR}}{\partial z}(t,z)+e^{-\gamma t}R_{post}\leq 0,&0<z<\hat{z}_{\scriptscriptstyle PR},\\ (V3)\quad\tilde{\mathcal{L}}\phi_{\scriptscriptstyle PR}(t,z)-\gamma\phi_{\scriptscriptstyle PR}(t,z)+e^{-\gamma t}(\tilde{u}_{\scriptscriptstyle PR}(z)-dz)=0,&0<z<\hat{z}_{\scriptscriptstyle PR},\\ (V4)\quad\tilde{\mathcal{L}}\phi_{\scriptscriptstyle PR}(t,z)-\gamma\phi_{\scriptscriptstyle PR}(t,z)+e^{-\gamma t}(\tilde{u}_{\scriptscriptstyle PR}(z)-dz)\geq 0,&z\geq\hat{z}_{\scriptscriptstyle PR},\end{cases} (A.4)

for any t0t\geq 0, with the smooth fit conditions ϕPRz(t,z^PR)=Rposteγt\frac{\partial\phi_{\scriptscriptstyle PR}}{\partial z}(t,\hat{z}_{\scriptscriptstyle PR})=-R_{post}e^{-\gamma t} and 2ϕPRz2(t,z^PR)=0\frac{\partial^{2}\phi_{\scriptscriptstyle PR}}{\partial z^{2}}(t,\hat{z}_{\scriptscriptstyle PR})=0.

Proposition A.3.

Under the assumption ϕPR(t,z)=eγtvPR(z)\phi_{\scriptscriptstyle PR}(t,z)=e^{-\gamma t}v_{\scriptscriptstyle PR}(z), the variational inequalities (A.4) takes the solution

vPR(z)={B2,PRz^PRn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z^PRδ(1k)δ(1k)1drz^PRRpost(zz^PR),zz^PR,B2,PRzn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz,0<z<z^PR,v_{\scriptscriptstyle PR}(z)=\begin{cases}B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}\hat{z}_{\scriptscriptstyle PR}-R_{post}(z-\hat{z}_{\scriptscriptstyle PR}),&z\geq\hat{z}_{\scriptscriptstyle PR},\\ B_{2,\scriptscriptstyle PR}z^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}z,&0<z<\hat{z}_{\scriptscriptstyle PR},\end{cases}

with

n2=γrθ22θ2+(γrθ22θ2)2+2γθ2,n_{2}=-\frac{\gamma-r-\frac{\theta^{2}}{2}}{\theta^{2}}+\sqrt{\left(\frac{\gamma-r-\frac{\theta^{2}}{2}}{\theta^{2}}\right)^{2}+\frac{2\gamma}{\theta^{2}}},
z^PR=L¯(1k)(1δ)[(1n2)(1δ(1k))n2(δ(1k)1)δ(1k)(Rpostdr)K1]δ(1k)1>0,\hat{z}_{\scriptscriptstyle PR}=\bar{L}^{(1-k)(1-\delta)}\left[\frac{(1-n_{2})(1-\delta(1-k))}{n_{2}(\delta(1-k)-1)-\delta(1-k)}\frac{\left(R_{post}-\frac{d}{r}\right)}{K_{1}}\right]^{\delta(1-k)-1}>0,

and

B2,PR==K1(δ(1k)1)(n21)L¯(1k)(1δ)(1n2)n2(n21)(δ(1k)1)[(1n2)(1δ(1k))n2(δ(1k)1)δ(1k)(Rpostdr)]δ(1k)n2(δ(1k)1)<0.B_{2,\scriptscriptstyle PR}=\!=\!\frac{K_{1}^{(\delta(1\!-\!k)\!-\!1)(n_{2}\!-\!1)}\!\bar{L}^{(1\!-\!k)(1\!-\!\delta)(1\!-\!n_{2})}}{n_{2}(n_{2}-1)(\delta(1\!-\!k)\!-\!1)}\!\!\left[\!\frac{(1\!-\!n_{2})(1\!-\!\delta(1\!-\!k))}{n_{2}(\delta(1\!-\!k)\!-\!1)\!-\!\delta(1\!-\!k)}\!\!\left(\!R_{\scriptscriptstyle post}\!-\!\frac{d}{r}\!\right)\!\right]^{\delta(1\!-\!k)\!-\!n_{2}(\delta(1\!-\!k)\!-\!1)}\!\!\!<\!0.

Furthermore, for a given initial wealth xRpostx\geq R_{\scriptscriptstyle post}, the value function of the post-retirement problem is

VPR(x)=B2,PR(λPR)n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)(λPR)δ(1k)δ(1k)1drλPR+λPRx,V_{\scriptscriptstyle PR}(x)\!=\!B_{2,\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*})^{n_{2}}\!+\!\frac{1\!-\!\delta(1\!-\!k)}{\delta(1\!-\!k)}K_{1}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}(\lambda_{\scriptscriptstyle PR}^{*})^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!-\!\frac{d}{r}\lambda_{\scriptscriptstyle PR}^{*}\!+\!\lambda_{\scriptscriptstyle PR}^{*}x, (A.5)

with n2B2,PR(λPR)n21+K1L¯(1k)(1δ)1δ(1k)(λPR)1δ(1k)1+dr=x-n_{2}B_{2,\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*})^{n_{2}-1}+K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}(\lambda_{\scriptscriptstyle PR}^{*})^{\frac{1}{\delta(1-k)-1}}+\frac{d}{r}=x. Taking ZPR(t)λPReγtH(t)Z_{\scriptscriptstyle PR}^{*}(t)\triangleq\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}H(t), the optimal wealth process follows

X(t)=n2B2,PR(ZPR(t))n21+K1L¯(1k)(1δ)1δ(1k)(ZPR(t))1δ(1k)1+dr,0<ZPR(t)z^PR,X^{*}(t)=-n_{2}B_{2,\scriptscriptstyle PR}(Z_{\scriptscriptstyle PR}^{*}(t))^{n_{2}-1}+K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}(Z_{\scriptscriptstyle PR}^{*}(t))^{\frac{1}{\delta(1-k)-1}}+\frac{d}{r},\quad 0<Z_{\scriptscriptstyle PR}^{*}(t)\leq\hat{z}_{\scriptscriptstyle PR},

and the corresponding optimal consumption and portfolio strategies are

c(t)=IPR(λPReγtH(t))=(ZPR(t))1δ(1k)1L¯(1k)(1δ)1δ(1k),c^{*}(t)=I_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}H(t))=(Z_{\scriptscriptstyle PR}^{*}(t))^{\frac{1}{\delta(1-k)-1}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}},
π(t)=θσ[n2(n21)B2,PR(ZPR(t))n21K11δ(1k)1L¯(1k)(1δ)1δ(1k)(ZPR(t))1δ(1k)1].\pi^{*}(t)=\frac{\theta}{\sigma}\left[n_{2}(n_{2}-1)B_{2,\scriptscriptstyle PR}(Z_{\scriptscriptstyle PR}^{*}(t))^{n_{2}-1}-K_{1}\frac{1}{\delta(1-k)-1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}(Z_{\scriptscriptstyle PR}^{*}(t))^{\frac{1}{\delta(1-k)-1}}\right].
Proof.

See Appendix A.2. ∎

Then based on the dynamic programming principle, we can only consider a subset of the admissible control set of Problem (P)(P), that is 𝒜1(x)𝒜(x)\mathcal{A}_{1}(x)\subset\mathcal{A}(x), in which any policy achieves the maximum of the post-retirement problem’s gain function. Hence, for any (τ,{c(t),π(t),l(t)})𝒜1(x)(\tau,\{c(t),\pi(t),l(t)\})\in\mathcal{A}_{1}(x), we have 𝔼[τeγtu(c(t),L¯)𝑑t]=𝔼[eγτVPR(Xx,c,π,l(τ))𝕀{τ<}]\mathbb{E}\left[\int_{\tau}^{\infty}e^{-\gamma t}u(c(t),\bar{L})dt\right]=\mathbb{E}\left[e^{-\gamma\tau}V_{\scriptscriptstyle PR}(X^{x,c,\pi,l}(\tau))\mathbb{I}_{\{\tau<\infty\}}\right]. Afterwards, the whole optimization problem can be rewritten as

V(x)=sup(τ,{c(t),π(t),l(t)})𝒜1(x)𝔼[0τeγtu(c(t),l(t))𝑑t+eγτU(Xx,c,π,l(τ))],V(x)=\sup_{\left(\tau,\{c(t),\pi(t),l(t)\}\right)\in\mathcal{A}_{1}(x)}\mathbb{E}\left[\int_{0}^{\tau}e^{-\gamma t}u(c(t),l(t))dt+e^{-\gamma\tau}U\left(X^{x,c,\pi,l}(\tau)\right)\right],

denoting U(Xx,c,π,l(τ))sup{c(t),π(t),l(t)}𝒜1(x)𝔼[τeγ(sτ)u(c(s),L¯)𝑑s|τ]=VPR(Xx,c,π,l(τ))U\left(X^{x,c,\pi,l}(\tau)\right)\!\triangleq\!\!\sup\limits_{\{c(t),\pi(t),l(t)\}\in\mathcal{A}_{1}(x)}\!\!\mathbb{E}\!\left[\!\left.\int_{\tau}^{\infty}\!\!e^{\!-\!\gamma(s\!-\!\tau)}u(c(s),\bar{L})ds\right|\mathcal{F}_{\tau}\right]\!=\!V_{\scriptscriptstyle PR}\left(X^{x,c,\pi,l}(\tau)\right). Finally, we summarize the two different forms of U(x)U(x) and introduce its Legendre-Fenchel transform under the definition U~(z)supxRpost[U(x)xz]\tilde{U}(z)\!\triangleq\!\!\sup\limits_{x\geq R_{post}}\!\!\left[U(x)\!-\!xz\right], 0<z<0\!<\!z\!<\!\infty following [4, Chapter 3, Definition 4.2].

Lemma A.1.

The post-retirement value function U(x)U(x), for xRpostx\geq R_{post}, is given in two separate cases, the partition being based on the value of threshold in the liquidity constraint, i.e., RpostR_{post}.

U(x)={(xdr)δ(1k)K11δ(1k)L¯(1k)(1δ)1δ(1k),ifRpost=dr,B2,PR(λPR)n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)(λPR)δ(1k)δ(1k)1drλPR+λPRx,ifRpost>dr.U\left(x\right)=\begin{cases}\left(x-\frac{d}{r}\right)^{\delta(1-k)}K_{1}^{1-\delta(1-k)}\bar{L}^{(1-k)(1-\delta)}\frac{1}{\delta(1-k)},&\mbox{if}\quad R_{post}=\frac{d}{r},\\ B_{2,\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*})^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}(\lambda_{\scriptscriptstyle PR}^{*})^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}\lambda_{\scriptscriptstyle PR}^{*}+\lambda_{\scriptscriptstyle PR}^{*}x,&\mbox{if}\quad R_{post}>\frac{d}{r}.\end{cases}

Furthermore, the Legendre-Fenchel transform of U(x)U(x) is:

  • U~(z)=1δ(1k)δ(1k)zδ(1k)δ(1k)1K1L¯(1k)(1δ)1δ(1k)drz\tilde{U}(z)=\frac{1-\delta(1-k)}{\delta(1-k)}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}-\frac{d}{r}z, z>0z>0, if Rpost=drR_{post}\!=\!\frac{d}{r};

  • U~(z)={B2,PRz^PRn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z^PRδ(1k)δ(1k)1drz^PRRpost(zz^PR),zz^PR,B2,PRzn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz,0<z<z^PR,\tilde{U}(z)\!=\!\begin{cases}\!B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}}\!+\!\frac{1\!-\!\delta(1\!-\!k)}{\delta(1\!-\!k)}K_{1}\!\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!\!\!\!-\!\frac{d}{r}\hat{z}_{\scriptscriptstyle PR}\!\!\!-\!R_{post}(z\!-\!\hat{z}_{\scriptscriptstyle PR}),&z\!\geq\!\hat{z}_{\scriptscriptstyle PR},\\ \!B_{2,\scriptscriptstyle PR}z^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}z,&0\!\!<\!z\!\!<\!\hat{z}_{\scriptscriptstyle PR},\end{cases} if Rpost>drR_{post}\!>\!\frac{d}{r}.

Proof.

See Appendix A.3. ∎

A.1 Proof of Theorem A.1

We first provide a lemma for proving Theorem A.1.

Lemma A.2.

For any given initial wealth x>Rpostx>R_{\scriptscriptstyle post}, and any given progressively measurable consumption process c(t)0c(t)\geq 0 satisfying supτ𝒯𝔼[0τH(t)(c(t)+d)𝑑t]xRpost\sup\limits_{\tau\in\mathcal{T}}\mathbb{E}\left[\int_{0}^{\tau}H(t)(c(t)+d)dt\right]\leq x-R_{\scriptscriptstyle post}, with 𝒯\mathcal{T} standing for the set of \mathcal{F}-stopping times, there exists a portfolio process π(t)\pi(t) making Xx,c,π(t)RpostX^{x,c,\pi}(t)\geq R_{\scriptscriptstyle post}, t0\forall t\geq 0, holds almost surely.

Proof.

Adopting the technique of [7, Appendix, Lemma 1], we introduce K(t)0t(c(s)+d)H(s)𝑑sK(t)\triangleq\int_{0}^{t}(c(s)+d)H(s)ds and show that {K(τ)}τ𝒯\{K(\tau)\}_{\tau\in\mathcal{T}} is uniformly integrable from the fact 𝔼[K(t)]<\mathbb{E}[K(t)]<\infty. Then, Dellacherie & Meyer (1982), Appendix I,111C. Dellacherie and P. Meyer, Probabilities and potential b, theory of martingales, North–Holland Mathematics Studies, 1982. indicates that there exists a Snell envelope of K(t)K(t) denoted as K¯(t)\bar{K}(t), which is a super-martingale under the \mathbb{P} measure and satisfies K¯(0)=supτ𝒯𝔼[K(τ)]\bar{K}(0)=\sup\limits_{\tau\in\mathcal{T}}\mathbb{E}[K(\tau)], K¯()=K()\bar{K}(\infty)=K(\infty). The Doob-Meyer Decomposition Theorem of Karatzas & Shreve (1998), Section 1.4, Theorem 4.10,222I. Karatzas and S. E. Shreve, Brownian Motion and Stochastic Calculus. Second edition. Springer-Verlag, 1998. enables us to represent the super-martingale K¯(t)\bar{K}(t) as K¯(t)=K¯(0)+M¯(t)A¯(t)\bar{K}(t)=\bar{K}(0)+\bar{M}(t)-\bar{A}(t), with a uniformly integrable martingale under the \mathbb{P} measure M¯(t)\bar{M}(t) satisfying M¯(0)=0\bar{M}(0)=0 and a strictly increasing process A¯(t)\bar{A}(t) satisfying A¯(0)=0\bar{A}(0)=0. Moreover, the Martingale Representation Theorem from Bjork (2009), Chapter 11, Theorem 11.2,333T. Bjork, Arbitrage theory in continuous time. Oxford university press, 2009. makes M¯(t)\bar{M}(t) take expression M¯(t)=0tρ¯(s)𝑑B(s)\bar{M}(t)=\int_{0}^{t}\bar{\rho}(s)dB(s), where ρ¯(t)\bar{\rho}(t) is an 𝔽\mathbb{F}-adapted process satisfying 0ρ¯2(s)𝑑s<\int_{0}^{\infty}\bar{\rho}^{2}(s)ds<\infty a.s..

Defining a new process X¯(t)1H(t)[xK¯(0)+K¯(t)K(t)+A¯(t)]Rpost\bar{X}(t)\triangleq\frac{1}{H(t)}\left[x-\bar{K}(0)+\bar{K}(t)-K(t)+\bar{A}(t)\right]-R_{\scriptscriptstyle post}, it can be observed that X¯(t)\bar{X}(t) is a non-negative process with the initial wealth X¯(0)=xRpost\bar{X}(0)=x-R_{\scriptscriptstyle post}, because of

K¯(0)=supτ𝒯𝔼[K(τ)]=supτ𝒯𝔼[0τH(t)(c(t)+d)𝑑t]xRpost.\bar{K}(0)=\sup\limits_{\tau\in\mathcal{T}}\mathbb{E}[K(\tau)]=\sup\limits_{\tau\in\mathcal{T}}\mathbb{E}\left[\int_{0}^{\tau}H(t)(c(t)+d)dt\right]\leq x-R_{\scriptscriptstyle post}.

Then X¯(t)\bar{X}(t) can be expressed with M¯(t)\bar{M}(t) as

X¯(t)=1H(t)[x+M¯(t)K(t)]Rpost=1H(t)[x+0tρ¯(s)𝑑B(s)0t(c(s)+d)H(s)𝑑s]Rpost.\bar{X}(t)=\frac{1}{H(t)}\left[x+\bar{M}(t)-K(t)\right]-R_{\scriptscriptstyle post}=\frac{1}{H(t)}\left[x+\int_{0}^{t}\bar{\rho}(s)dB(s)-\int_{0}^{t}(c(s)+d)H(s)ds\right]-R_{\scriptscriptstyle post}.

Applying the Itô’s formula to H(t)Xx,c,π(t)H(t)X^{x,c,\pi}(t), we can get

d(H(t)Xx,c,π(t))=H(t)Xx,c,π(t)θdB(t)(c(t)+d)H(t)dt+σπ(t)H(t)dB(t).d(H(t)X^{x,c,\pi}(t))=-H(t)X^{x,c,\pi}(t)\theta dB(t)-(c(t)+d)H(t)dt+\sigma\pi(t)H(t)dB(t).

Considering the portfolio strategy π(t)=ρ¯(t)σH(t)+θXx,c,π(t)σ\pi(t)=\frac{\bar{\rho}(t)}{\sigma H(t)}+\frac{\theta X^{x,c,\pi}(t)}{\sigma}, the wealth process takes

Xx,c,π(t)=1H(t)[x+0tρ¯(s)𝑑B(s)0t(c(s)+d)H(s)𝑑s],X^{x,c,\pi}(t)=\frac{1}{H(t)}\left[x+\int_{0}^{t}\bar{\rho}(s)dB(s)-\int_{0}^{t}(c(s)+d)H(s)ds\right],

which indicates that X¯(t)=Xx,c,π(t)Rpost\bar{X}(t)=X^{x,c,\pi}(t)-R_{\scriptscriptstyle post}, a.s.. The non-negativity of X¯(t)\bar{X}(t) makes clear that Xx,c,π(t)RpostX^{x,c,\pi}(t)\geq R_{\scriptscriptstyle post}, a.s., t0.\forall t\geq 0.

Now we turn back to the proof of Theorem A.1. Following [7, Section 4, Theorem 1], the proof mainly contains two aspects: the first part is to show the admissibility of c(t)c^{*}(t), and the second part is to claim that c(t)c^{*}(t) is the optimal consumption strategy to Problem (PPR)(P_{\scriptscriptstyle PR}).

(1) We first prove that c(t)=IPR(λPReγtDPR(t)H(t))c^{*}(t)=I_{\scriptscriptstyle PR}^{*}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D_{\scriptscriptstyle PR}^{*}(t)H(t)) is an admissible consumption policy. Taking any stopping time τ\tau from 𝒯\mathcal{T} and a positive constant ϵ\epsilon, we can introduce DPRϵ(t)DPR(t)+ϵ𝕀[0,τ)(t)D_{\scriptscriptstyle PR}^{\epsilon}(t)\triangleq D_{\scriptscriptstyle PR}^{*}(t)+\epsilon\mathbb{I}_{[0,\tau)}(t), which evidently satisfies DPRϵ(t)𝒟D_{\scriptscriptstyle PR}^{\epsilon}(t)\in\mathcal{D}. Then defining a function

𝔏(D(t))𝔼[0eγt(u~PR(λPReγtD(t)H(t))dλPReγtD(t)H(t))𝑑t]+λPR𝔼[0RpostH(t)𝑑D(t)]+λPR(xRpost)D(0),\begin{split}\mathfrak{L}(D(t))\triangleq\mathbb{E}&\left[\int_{0}^{\infty}e^{-\gamma t}\big{(}\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D(t)H(t))-d\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D(t)H(t)\big{)}dt\right]\\ &+\lambda_{\scriptscriptstyle PR}^{*}\mathbb{E}\left[\int_{0}^{\infty}R_{\scriptscriptstyle post}H(t)dD(t)\right]+\lambda_{\scriptscriptstyle PR}^{*}(x-R_{\scriptscriptstyle post})D(0),\end{split}

an inequality, 𝔏(DPR(t))𝔏(DPRϵ(t))\mathfrak{L}(D_{\scriptscriptstyle PR}^{*}(t))\leq\mathfrak{L}(D_{\scriptscriptstyle PR}^{\epsilon}(t)), is obtained from the facts that DPR(t)D_{\scriptscriptstyle PR}^{*}(t) is the optimal solution of Problem (SPR)(S_{\scriptscriptstyle PR}) and xRpostx\geq R_{\scriptscriptstyle post}. This inequality gives us

lim supϵ0𝔼[0τ(eγtu~PR(λPReγtDPRϵ(t)H(t))u~PR(λPReγtDPR(t)H(t))ϵdλPRH(t))𝑑t]+λPR(xRpost)0,\limsup_{\epsilon\downarrow 0}\mathbb{E}\!\left[\!\int_{0}^{\tau}\!\!\!\!\!\!\big{(}e^{\!-\!\gamma t}\!\frac{\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}\!D_{\scriptscriptstyle PR}^{\epsilon}(t)H(t))\!\!-\!\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}\!D_{\scriptscriptstyle PR}^{*}(t)H(t))}{\epsilon}\!-\!d\lambda_{\scriptscriptstyle PR}^{*}H(t)\big{)}dt\!\right]\!+\!\lambda_{\scriptscriptstyle PR}^{*}(x\!-\!R_{\scriptscriptstyle post})\!\!\geq\!0,

considering dDϵ(t)=dD(t)dD^{\epsilon}(t)=dD^{*}(t), t(0,τ)\forall t\in(0,\tau). The decreasing property of u~PR()\tilde{u}_{\scriptscriptstyle PR}(\cdot) and the Fatou’s lemma endows us with

𝔼[0τeγtu~PR(λPReγtDPR(t)H(t))λPReγtH(t)dt]lim supϵ0𝔼[0τeγtu~PR(λPReγtDPRϵ(t)H(t))u~PR(λPReγtDPR(t)H(t))ϵ𝑑t].\begin{split}\mathbb{E}\bigg{[}\!\int_{0}^{\tau}&e^{-\gamma t}\tilde{u}_{\scriptscriptstyle PR}^{\prime}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D_{\scriptscriptstyle PR}^{*}(t)H(t))\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}H(t)dt\bigg{]}\geq\\ &\limsup_{\epsilon\downarrow 0}\mathbb{E}\left[\int_{0}^{\tau}e^{-\gamma t}\frac{\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}\!D_{\scriptscriptstyle PR}^{\epsilon}(t)H(t))-\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D_{\scriptscriptstyle PR}^{*}(t)H(t))}{\epsilon}dt\right].\end{split}

Then u~PR()=IPR()\tilde{u}_{\scriptscriptstyle PR}^{\prime}(\cdot)=-I_{\scriptscriptstyle PR}(\cdot) indicates that 𝔼[0τH(t)(c(t)+d)𝑑t]xRpost\mathbb{E}\left[\int_{0}^{\tau}H(t)(c^{*}(t)+d)dt\right]\leq x-R_{\scriptscriptstyle post}. Since τ\tau can be any stopping time in the set 𝒯\mathcal{T}, Lemma A.2 claims that there exists a portfolio strategy π(t)\pi^{*}(t) making the related wealth process satisfying Xx,c,π(t)RpostX^{x,c^{*},\pi^{*}}(t)\geq R_{\scriptscriptstyle post}, t0\forall t\geq 0.

(2) We move to show the optimality of c(t)c^{*}(t) to Problem (PPR)(P_{\scriptscriptstyle PR}). The proof of Lemma A.2 indicates that for an arbitrary consumption strategy c(t)𝒜PR(x)c(t)\in\mathcal{A}_{\scriptscriptstyle PR}(x), there exists a process ζ(t)\zeta(t) satisfying

0t(c(s)+d)H(s)𝑑s+H(t)Xx,c,π(t)=x+0tζ(s)𝑑B(s).\int_{0}^{t}(c(s)+d)H(s)ds+H(t)X^{x,c,\pi}(t)=x+\int_{0}^{t}\zeta(s)dB(s). (A.6)

The property Xx,c,π(t)RpostX^{x,c,\pi}(t)\geq R_{\scriptscriptstyle post} a.s. gives us the subsequent inequality with any process D(t)𝒟D(t)\in\mathcal{D},

0T0t(c(s)+d)H(s)𝑑s𝑑D(t)+0TRpostH(t)𝑑D(t)0T[x+0tζ(s)𝑑B(s)]𝑑D(t),\int_{0}^{T}\int_{0}^{t}(c(s)+d)H(s)dsdD(t)+\int_{0}^{T}R_{\scriptscriptstyle post}H(t)dD(t)\geq\int_{0}^{T}\left[x+\int_{0}^{t}\zeta(s)dB(s)\right]dD(t),

where TT is any time meeting TtT\!\geq\!t. Since D(t)D(t) is bounded variational, integrating by parts gives us

0TD(s)(c(s)+d)H(s)ds0TD(s)ζ(s)𝑑B(s)D(0)x+D(T)[0T(c(s)+d)H(s)𝑑sx0Tζ(s)𝑑B(s)]+0TRpostH(s)𝑑D(s).\begin{split}\int_{0}^{T}D(s)&(c(s)+d)H(s)ds-\int_{0}^{T}D(s)\zeta(s)dB(s)\leq\\ &D(0)x+D(T)\left[\int_{0}^{T}(c(s)+d)H(s)ds\!-\!x\!-\!\int_{0}^{T}\zeta(s)dB(s)\right]+\int_{0}^{T}R_{\scriptscriptstyle post}H(s)dD(s).\end{split}

Then we can take the expectation under the \mathbb{P} measure on both sides and replace Equation (A.6) to get 𝔼[0TD(s)(c(s)+d)H(s)𝑑s]D(0)x+𝔼[0TRpostH(s)𝑑D(s)]\mathbb{E}\left[\int_{0}^{T}D(s)(c(s)+d)H(s)ds\right]\leq D(0)x+\mathbb{E}\left[\int_{0}^{T}R_{\scriptscriptstyle post}H(s)dD(s)\right]. Then the Lebesgue’s Monotone Convergence Theorem indicates that

𝔼[0D(s)(c(s)+d)H(s)𝑑s]D(0)x+𝔼[0RpostH(s)𝑑D(s)],\mathbb{E}\left[\int_{0}^{\infty}D(s)(c(s)+d)H(s)ds\right]\leq D(0)x+\mathbb{E}\left[\int_{0}^{\infty}R_{\scriptscriptstyle post}H(s)dD(s)\right],

which holds for any admissible consumption policy c(t)c(t) and any non-negative, non-increasing process D(t)D(t). Furthermore, it will be proved that the above inequality becomes equalized with the given c(t)c^{*}(t) and DPR(t)D_{\scriptscriptstyle PR}^{*}(t). Introducing D¯PRϵ(t)DPR(t)(1+ϵ)𝒟\bar{D}_{\scriptscriptstyle PR}^{\epsilon}(t)\triangleq D_{\scriptscriptstyle PR}^{*}(t)(1+\epsilon)\in\mathcal{D} with a small enough constant ϵ\epsilon and defining a new function as

𝔏~(D(t))𝔼[0eγt(u~PR(λPReγtD(t)H(t))dλPReγtD(t)H(t))𝑑t]+λPR𝔼[0RpostH(t)𝑑D(t)]+λPRxD(0),\begin{split}\tilde{\mathfrak{L}}(D(t))\triangleq\mathbb{E}&\left[\int_{0}^{\infty}e^{-\gamma t}\big{(}\tilde{u}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D(t)H(t))-d\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D(t)H(t)\big{)}dt\right]\\ &+\lambda_{\scriptscriptstyle PR}^{*}\mathbb{E}\left[\int_{0}^{\infty}R_{\scriptscriptstyle post}H(t)dD(t)\right]+\lambda_{\scriptscriptstyle PR}^{*}xD(0),\end{split}

we get 𝔏~(D¯PRϵ(t))𝔏~(DPR(t))\tilde{\mathfrak{L}}(\bar{D}_{\scriptscriptstyle PR}^{\epsilon}(t))\geq\tilde{\mathfrak{L}}(D_{\scriptscriptstyle PR}^{*}(t)). Following the same argument with the first part, we apply the Fatou’s lemma to obtain separately

𝔼[0DPR(t)H(t)(c(t)+d)𝑑t]xDPR(0)+𝔼[0RpostH(t)𝑑DPR(t)],𝔼[0DPR(t)H(t)(c(t)+d)𝑑t]xDPR(0)+𝔼[0RpostH(t)𝑑DPR(t)],\begin{split}&\mathbb{E}\left[\int_{0}^{\infty}D_{\scriptscriptstyle PR}^{*}(t)H(t)(c^{*}(t)+d)dt\right]\leq xD_{\scriptscriptstyle PR}^{*}(0)+\mathbb{E}\left[\int_{0}^{\infty}R_{\scriptscriptstyle post}H(t)dD_{\scriptscriptstyle PR}^{*}(t)\right],\\ &\mathbb{E}\left[\int_{0}^{\infty}D_{\scriptscriptstyle PR}^{*}(t)H(t)(c^{*}(t)+d)dt\right]\geq xD_{\scriptscriptstyle PR}^{*}(0)+\mathbb{E}\left[\int_{0}^{\infty}R_{\scriptscriptstyle post}H(t)dD_{\scriptscriptstyle PR}^{*}(t)\right],\end{split}

which claims 𝔼[0DPR(t)H(t)(c(t)+d)𝑑t]=xDPR(0)+𝔼[0RpostH(t)𝑑DPR(t)]\mathbb{E}\left[\int_{0}^{\infty}\!D_{\scriptscriptstyle PR}^{*}(t)H(t)(c^{*}(t)\!+\!d)dt\right]\!=\!xD_{\scriptscriptstyle PR}^{*}(0)\!+\!\mathbb{E}\left[\int_{0}^{\infty}\!R_{\scriptscriptstyle post}H(t)dD_{\scriptscriptstyle PR}^{*}(t)\right]. Afterwards, we define a new optimization problem named (PPR)(P_{\scriptscriptstyle PR}^{\prime}) as

maxc(t)0𝔼[0eγtuPR(c(t))𝑑t]\max_{c(t)\geq 0}\mathbb{E}\left[\int_{0}^{\infty}e^{-\gamma t}u_{\scriptscriptstyle PR}(c(t))dt\right] (PPRP_{\scriptscriptstyle PR}^{\prime})
s.t.𝔼[0DPR(t)H(t)(c(t)+d)dt]xDPR(0)+𝔼[0RpostH(t)dDPR(t)].s.t.\quad\mathbb{E}\left[\int_{0}^{\infty}D_{\scriptscriptstyle PR}^{*}(t)H(t)(c(t)+d)dt\right]\leq xD_{\scriptscriptstyle PR}^{*}(0)+\mathbb{E}\left[\int_{0}^{\infty}R_{\scriptscriptstyle post}H(t)dD_{\scriptscriptstyle PR}^{*}(t)\right].

The Lagrange method implies that the optimal consumption solution of the above problem, denoted as c~(t)\tilde{c}^{*}(t), satisfies eγtuPR(c~(t))=λ~PRDPR(t)H(t)e^{-\gamma t}u_{\scriptscriptstyle PR}^{\prime}(\tilde{c}^{*}(t))=\tilde{\lambda}_{\scriptscriptstyle PR}D_{\scriptscriptstyle PR}^{*}(t)H(t), with λ~PR>0\tilde{\lambda}_{\scriptscriptstyle PR}>0 as the Lagrange multiplier. The condition λ~PR=λPR\tilde{\lambda}_{\scriptscriptstyle PR}=\lambda_{\scriptscriptstyle PR}^{*} makes the constraint of Problem (PPR)(P_{\scriptscriptstyle PR}^{\prime}) takes equality. And the condition uPR(c~(t))=λPReγtDPR(t)H(t)u_{\scriptscriptstyle PR}^{\prime}(\tilde{c}^{*}(t))=\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D_{\scriptscriptstyle PR}^{*}(t)H(t) implies that c~(t)=I(λPReγtDPR(t)H(t))=c(t)\tilde{c}^{*}(t)=I(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}D_{\scriptscriptstyle PR}^{*}(t)H(t))=c^{*}(t), which shows that c(t)c^{*}(t) is the optimal consumption policy of Problem (PPR)(P_{\scriptscriptstyle PR}^{\prime}). Finally, since the maximum utility of Problem (PPR)(P_{\scriptscriptstyle PR}) is upper bounded by the maximum utility of (PPR)(P_{\scriptscriptstyle PR}^{\prime}), we can conclude that c(t)c^{*}(t) is also the optimal consumption solution of the primal problem (PPR)(P_{\scriptscriptstyle PR}).

A.2 Proof of Proposition A.3

Referring to [2, Appendix A], the function ϕPR(t,z)\phi_{\scriptscriptstyle PR}(t,z) is assumed to be time-independent, that is, ϕPR(t,z)=eγtvPR(z)\phi_{\scriptscriptstyle PR}(t,z)=e^{-\gamma t}v_{\scriptscriptstyle PR}(z). Then the condition (V3)(V3) of (A.4) leads to a differential equation

γvPR(z)+(γr)zvPR(z)+12θ2z2vPR′′(z)+u~PR(z)dz=0,0<z<z^PR,-\gamma v_{\scriptscriptstyle PR}(z)+(\gamma-r)zv_{\scriptscriptstyle PR}^{\prime}(z)+\frac{1}{2}\theta^{2}z^{2}v_{\scriptscriptstyle PR}^{\prime\prime}(z)+\tilde{u}_{\scriptscriptstyle PR}(z)-dz=0,\quad 0<z<\hat{z}_{\scriptscriptstyle PR}, (A.7)

which has the solution

vPR(z)=B1,PRzn1+B2,PRzn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz,0<z<z^PR.v_{\scriptscriptstyle PR}(z)=B_{1,\scriptscriptstyle PR}z^{n_{1}}+B_{2,\scriptscriptstyle PR}z^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}z,\quad 0<z<\hat{z}_{\scriptscriptstyle PR}.

n1n_{1} and n2n_{2} are the roots of the second-order equation θ22n2+(γrθ22)nγ=0\frac{\theta^{2}}{2}n^{2}+\left(\gamma-r-\frac{\theta^{2}}{2}\right)n-\gamma=0, and satisfy

n1,2=γrθ22θ2(γrθ22θ2)2+2γθ2,n1<0,n2>1.n_{1,2}=-\frac{\gamma-r-\frac{\theta^{2}}{2}}{\theta^{2}}\mp\sqrt{\left(\frac{\gamma-r-\frac{\theta^{2}}{2}}{\theta^{2}}\right)^{2}+\frac{2\gamma}{\theta^{2}}},\qquad n_{1}<0,\quad n_{2}>1.

Since n1<0n_{1}<0, the term zn1z^{n_{1}} will suffer the explosion as zz goes to 0. Therefore, we set the coefficient B1,PR=0B_{1,\scriptscriptstyle PR}=0 by the boundedness assumption. Considering the smooth conditions at z^PR\hat{z}_{\scriptscriptstyle PR}, we can construct a two-equations system to determine the parameters B2,PRB_{2,\scriptscriptstyle PR} and z^PR\hat{z}_{\scriptscriptstyle PR}.

  • 𝒞1\mathcal{C}^{1} condition at z=z^PRz=\hat{z}_{\scriptscriptstyle PR}: n2B2,PRz^PRn21K1L¯(1k)(1δ)1δ(1k)z^PR1δ(1k)1dr+Rpost=0.n_{2}B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}-1}-K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{1}{\delta(1-k)-1}}-\frac{d}{r}+R_{post}=0.

  • 𝒞2\mathcal{C}^{2} condition at z=z^PRz=\hat{z}_{\scriptscriptstyle PR}: n2(n21)B2,PRz^PRn22K11δ(1k)1L¯(1k)(1δ)1δ(1k)z^PR2δ(1k)δ(1k)1=0.n_{2}(n_{2}-1)B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}-2}-K_{1}\frac{1}{\delta(1-k)-1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{2-\delta(1-k)}{\delta(1-k)-1}}=0.

By multiplying the 𝒞2\mathcal{C}^{2} condition with z^PR\hat{z}_{\scriptscriptstyle PR} and then adding with the 𝒞1\mathcal{C}^{1} condition, we have

B2,PRz^PRn21=K1δ(1k)δ(1k)11n22L¯(1k)(1δ)1δ(1k)z^PR1δ(1k)1+1n22(drRpost).B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}-1}=K_{1}\frac{\delta(1-k)}{\delta(1-k)-1}\frac{1}{n_{2}^{2}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{1}{\delta(1-k)-1}}+\frac{1}{n_{2}^{2}}\left(\frac{d}{r}-R_{post}\right).

Then, substituting the above expression into the 𝒞1\mathcal{C}^{1} condition, we get the exact value of z^PR\hat{z}_{\scriptscriptstyle PR} as

z^PR=L¯(1k)(1δ)[(1n2)(1δ(1k))n2(δ(1k)1)δ(1k)(Rpostdr)K1]δ(1k)1>0,\hat{z}_{\scriptscriptstyle PR}=\bar{L}^{(1-k)(1-\delta)}\left[\frac{(1-n_{2})(1-\delta(1-k))}{n_{2}(\delta(1-k)-1)-\delta(1-k)}\frac{\left(R_{post}-\frac{d}{r}\right)}{K_{1}}\right]^{\delta(1-k)-1}>0,

and B2,PRB_{2,\scriptscriptstyle PR} can also be solved by bringing z^PR\hat{z}_{\scriptscriptstyle PR} into the expression B2,PRz^PRn21B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}-1},

B2,PR=K1(δ(1k)1)(n21)L¯(1k)(1δ)(1n2)n2(n21)(δ(1k)1)[(1n2)(1δ(1k))n2(δ(1k)1)δ(1k)(Rpostdr)]δ(1k)n2(δ(1k)1)<0.B_{2,\scriptscriptstyle PR}\!=\!\frac{K_{1}^{(\delta(1\!-\!k)\!-\!1)(n_{2}\!-\!1)}\!\bar{L}^{(1\!-\!k)(1\!-\!\delta)(1\!-\!n_{2})}}{n_{2}(n_{2}-1)(\delta(1\!-\!k)\!-\!1)}\!\!\left[\!\frac{(1\!-\!n_{2})(1\!-\!\delta(1\!-\!k))}{n_{2}(\delta(1\!-\!k)\!-\!1)\!-\!\delta(1\!-\!k)}\!\!\left(\!R_{\scriptscriptstyle post}\!-\!\frac{d}{r}\!\right)\!\right]^{\delta(1\!-\!k)\!-\!n_{2}(\delta(1\!-\!k)\!-\!1)}\!\!\!<\!0.

Moreover, the piecewise function of vPR(z)v_{\scriptscriptstyle PR}(z) is completely determined as

vPR(z)={B2,PRz^PRn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z^PRδ(1k)δ(1k)1drz^PRRpost(zz^PR),zz^PR,B2,PRzn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz,0<z<z^PR.v_{\scriptscriptstyle PR}(z)=\begin{cases}B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}\hat{z}_{\scriptscriptstyle PR}-R_{post}(z-\hat{z}_{\scriptscriptstyle PR}),&z\geq\hat{z}_{\scriptscriptstyle PR},\\ B_{2,\scriptscriptstyle PR}z^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}z,&0<z<\hat{z}_{\scriptscriptstyle PR}.\end{cases}

Since vPR(z)v_{\scriptscriptstyle PR}(z) is a piecewise polynomial function with smoothing merging conditions and differentiable everywhere, [6, Section 8, Theorem 8.5] indicates that VPR(x)=infλPR>0[V~PR(λPR)+λPRx]V_{\scriptscriptstyle PR}(x)=\inf\limits_{\lambda_{\scriptscriptstyle PR}>0}[\tilde{V}_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR})+\lambda_{\scriptscriptstyle PR}x] keeps true for any given initial wealth xRpostx\geq R_{post}. Thereafter, the closed-form of VPR(x)V_{\scriptscriptstyle PR}(x) is

VPR(x)=B2,PR(λPR)n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)(λPR)δ(1k)δ(1k)1drλPR+λPRx,xx^PR,V_{\scriptscriptstyle PR}(x)\!=\!B_{2,\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*})^{n_{2}}\!+\!\frac{1\!-\!\delta(1\!-\!k)}{\delta(1\!-\!k)}K_{1}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}(\lambda_{\scriptscriptstyle PR}^{*})^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!-\!\frac{d}{r}\lambda_{\scriptscriptstyle PR}^{*}\!+\!\lambda_{\scriptscriptstyle PR}^{*}x,\quad x\!\geq\!\hat{x}_{\scriptscriptstyle PR},

with n2B2,PR(λPR)n21+K1L¯(1k)(1δ)1δ(1k)(λPR)1δ(1k)1+dr=x-n_{2}B_{2,\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*})^{n_{2}-1}+K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}(\lambda_{\scriptscriptstyle PR}^{*})^{\frac{1}{\delta(1-k)-1}}+\frac{d}{r}=x, xx^PRx\geq\hat{x}_{\scriptscriptstyle PR}. x^PR\hat{x}_{\scriptscriptstyle PR} is the critical wealth level corresponding to z^PR\hat{z}_{\scriptscriptstyle PR} and follows

x^PR=vPRz|z=z^PR=n2B2,PRz^PRn21+K1L¯(1k)(1δ)1δ(1k)z^PR1δ(1k)1+dr.\hat{x}_{\scriptscriptstyle PR}=\left.-\frac{\partial v_{\scriptscriptstyle PR}}{\partial z}\right|_{z=\hat{z}_{\scriptscriptstyle PR}}=-n_{2}B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}-1}+K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{1}{\delta(1-k)-1}}+\frac{d}{r}.

Moreover, the optimal wealth process takes the form

X(t)=vPR(ZPR(t))=n2B2,PR(ZPR(t))n21+K1L¯(1k)(1δ)1δ(1k)(ZPR(t))1δ(1k)1+dr,0<ZPR(t)z^PR,X^{*}(t)\!=\!-\!v_{\scriptscriptstyle PR}^{\prime}(Z_{\scriptscriptstyle PR}^{*}(t))\!=\!-\!n_{2}B_{2,\scriptscriptstyle PR}(Z_{\scriptscriptstyle PR}^{*}(t))^{n_{2}\!-\!1}\!+\!K_{1}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}(Z_{\scriptscriptstyle PR}^{*}(t))^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\!+\!\frac{d}{r},\quad 0\!<\!Z_{\scriptscriptstyle PR}^{*}(t)\!\leq\!\hat{z}_{\scriptscriptstyle PR},

and the related optimal consumption-portfolio strategies are

c(t)=IPR(λPReγtH(t))=(ZPR(t))1δ(1k)1L¯(1k)(1δ)1δ(1k),c^{*}(t)=I_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR}^{*}e^{\gamma t}H(t))=(Z_{\scriptscriptstyle PR}^{*}(t))^{\frac{1}{\delta(1-k)-1}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}},
π(t)=θσ[n2(n21)B2,PR(ZPR(t))n21K11δ(1k)1L¯(1k)(1δ)1δ(1k)(ZPR(t))1δ(1k)1],\pi^{*}(t)=\frac{\theta}{\sigma}\left[n_{2}(n_{2}-1)B_{2,\scriptscriptstyle PR}(Z_{\scriptscriptstyle PR}^{*}(t))^{n_{2}-1}-K_{1}\frac{1}{\delta(1-k)-1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}(Z_{\scriptscriptstyle PR}^{*}(t))^{\frac{1}{\delta(1-k)-1}}\right],

the optimal portfolio strategy is obtained from [7, Section 5, Theorem 3].

A.3 Proof of Lemma A.1

The form of function U(x)U(x) is directly summarized from Equation (A.2) and (A.5), hence the proof here only focuses on the derivation of the Legendre-Fenchel transform of U(x)U(x), which is also divided into two cases. We first extend the supremum in the definition of Legendre-Fenchel transform U~(z)\tilde{U}(z) by enlarging the range of xx to \mathbb{R}, that is, U~(z)=supx[U(x)xz]\tilde{U}(z)=\sup\limits_{x\in\mathbb{R}}[U(x)-xz], for 0<z<0<z<\infty. Moreover, it can be proved the optimal solution xx^{*} attaining the supremum automatically satisfies xRpostx^{*}\geq R_{post}.
(1) Rpost=drR_{post}=\frac{d}{r}: From the first-order condition, we have

z=U(x)=K11δ(1k)L¯(1k)(1δ)(xdr)δ(1k)1,z=U^{\prime}(x^{*})=K_{1}^{1-\delta(1-k)}\bar{L}^{(1-k)(1-\delta)}\left(x^{*}-\frac{d}{r}\right)^{\delta(1-k)-1},

which entails that x=(zK1δ(1k)1L¯(1k)(1δ))1δ(1k)1+drx^{*}=\left(zK_{1}^{\delta(1-k)-1}\bar{L}^{-(1-k)(1-\delta)}\right)^{\frac{1}{\delta(1-k)-1}}+\frac{d}{r}. Then x>Rpost=drx^{*}>R_{post}=\frac{d}{r} is obviously satisfied for z>0z>0. Taking the above relationship back to the dual transform definition, U~(z)\tilde{U}(z) is directly acquired after elementary calculation,

U~(z)=1δ(1k)δ(1k)zδ(1k)δ(1k)1K1L¯(1k)(1δ)1δ(1k)drz.\tilde{U}(z)=\frac{1-\delta(1-k)}{\delta(1-k)}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}-\frac{d}{r}z.

(2) Rpost>drR_{post}>\frac{d}{r}: Considering the fact VPR(x)=infλPR>0[vPR(λPR)+λPRx]V_{\scriptscriptstyle PR}(x)\!=\!\inf\limits_{\lambda_{\scriptscriptstyle PR}>0}[v_{\scriptscriptstyle PR}(\lambda_{\scriptscriptstyle PR})\!+\!\lambda_{\scriptscriptstyle PR}x], it can be obtained that vPR(z)v_{\scriptscriptstyle PR}(z) is the Legendre-Fenchel transform of VPR(x)V_{\scriptscriptstyle PR}(x) from [4, Chapter 3, Lemma 4.3]. Then the identical forms of functions VPR(x)V_{\scriptscriptstyle PR}(x) and U(x)U(x) enable us to deduce the solution as U~(z)=vPR(z)\tilde{U}(z)=v_{\scriptscriptstyle PR}(z). The last step is to claim xRpostx^{*}\geq R_{post}, which can be resorted to the condition x=vPR(z)vPR(z^PR)=Rpostx^{*}=-v_{\scriptscriptstyle PR}^{\prime}(z)\geq-v_{\scriptscriptstyle PR}^{\prime}(\hat{z}_{\scriptscriptstyle PR})=R_{post}.

Appendix B Proof of Lemma 3.2

Following [6, Section 6, Lemma 6.3], we first define a new continuous process as

G(t)1ξ(t)𝔼~[tτξ(s)(c(s)+wl(s)+dwL¯)𝑑s+ξ(τ)K|t],t[0,τ],G(t)\triangleq\frac{1}{\xi(t)}\tilde{\mathbb{E}}\left[\left.\int_{t}^{\tau}\xi(s)(c(s)+wl(s)+d-w\bar{L})ds+\xi(\tau)K\right|\mathcal{F}_{t}\right],\quad\forall t\in[0,\tau],

where 𝔼~[]\tilde{\mathbb{E}}[\cdot] representing the expectation under ~\tilde{\mathbb{P}} measure. From the property of the random variable KK, this process satisfies G(t)=1ξ(t)𝔼~[tτξ(s)(c(s)+wl(s))𝑑s+ξ(τ)(KdwL¯r)|t]+dwL¯rdwL¯rG(t)\!=\!\frac{1}{\xi(t)}\tilde{\mathbb{E}}\!\left[\!\left.\int_{t}^{\tau}\!\xi(s)(c(s)\!+\!wl(s))ds\!+\!\xi(\tau)\left(K\!-\!\frac{d-w\bar{L}}{r}\right)\right|\mathcal{F}_{t}\!\right]\!+\!\frac{d\!-\!w\bar{L}}{r}\!\geq\!\frac{d\!-\!w\bar{L}}{r}, a.s.. Then, making use of the condition 𝔼[0τH(t)(c(t)+wl(t)+dwL¯)𝑑t+H(τ)K]=x\mathbb{E}\left[\int_{0}^{\tau}H(t)(c(t)+wl(t)+d-w\bar{L})dt+H(\tau)K\right]=x, we get G(τ)=KG(\tau)=K and

G(0)=𝔼~[0τξ(s)(c(s)+wl(s)+dwL¯)𝑑s+ξ(τ)K]=𝔼[0τH(s)(c(s)+wl(s)+dwL¯)𝑑s+H(τ)K]=x,\begin{split}G(0)&=\tilde{\mathbb{E}}\left[\int_{0}^{\tau}\xi(s)(c(s)+wl(s)+d-w\bar{L})ds+\xi(\tau)K\right]\\ &=\mathbb{E}\left[\int_{0}^{\tau}H(s)(c(s)+wl(s)+d-w\bar{L})ds+H(\tau)K\right]=x,\end{split}

the above derivation involves changing the measure from the ~\tilde{\mathbb{P}} measure with the pricing kernel as ξ(t)\xi(t) to the \mathbb{P} measure with the pricing kernel as H(t)H(t). Meanwhile, we define a new process

M(t)=ξ(t)G(t)+0tξ(s)(c(s)+wl(s)+dwL¯)𝑑s,t[0,τ].M(t)=\xi(t)G(t)+\int_{0}^{t}\xi(s)(c(s)+wl(s)+d-w\bar{L})ds,\quad\forall t\in[0,\tau].

Based on the fact

𝔼~[M(t)]=𝔼~[𝔼~[tτξ(s)(c(s)+wl(s)+dwL¯)𝑑s+ξ(τ)K|t]+0tξ(s)(c(s)+wl(s)+dwL¯)𝑑s]=𝔼~[0τξ(s)(c(s)+wl(s)+dwL¯)𝑑s+ξ(τ)K]=x=M(0),\begin{split}\tilde{\mathbb{E}}[M(t)]&=\!\tilde{\mathbb{E}}\left[\tilde{\mathbb{E}}\left[\left.\int_{t}^{\tau}\xi(s)(c(s)\!+\!wl(s)\!+\!d\!-\!w\bar{L})ds\!+\!\xi(\tau)K\right|\!\mathcal{F}_{t}\right]\!+\!\int_{0}^{t}\xi(s)(c(s)\!+\!wl(s)\!+\!d\!-\!w\bar{L})ds\right]\\ &=\tilde{\mathbb{E}}\left[\int_{0}^{\tau}\xi(s)(c(s)+wl(s)+d-w\bar{L})ds+\xi(\tau)K\right]=x=M(0),\end{split}

M(t)M(t) is a ~\tilde{\mathbb{P}}-martingale. According to the Martingale Representation Theorem from Bjork (2009), Chapter 11, Theorem 11.2, it can be expressed as M(t)=x+0tρ(s)𝑑B~(s)M(t)=x+\int_{0}^{t}\rho(s)d\tilde{B}(s), t[0,τ]\forall t\in[0,\tau], with an 𝔽\mathbb{F}-adapted process ρ(t)\rho(t) satisfying 0ρ2(s)𝑑s<\int_{0}^{\infty}\rho^{2}(s)ds<\infty a.s.. Furthermore, adopting the portfolio strategy π(t)ρ(t)σξ(t)\pi(t)\triangleq\frac{\rho(t)}{\sigma\xi(t)}, the wealth process becomes

dXx,c,π,l(t)=rXx,c,π,l(t)dt+π(t)(μr)dt(c(t)+wl(t)+dwL¯)dt+σπ(t)dB(t)=rXx,c,π,l(t)dt(c(t)+wl(t)+dwL¯)dt+σπ(t)dB~(t)=rXx,c,π,l(t)dt(c(t)+wl(t)+dwL¯)dt+ρ(t)ξ(t)dB~(t),\begin{split}dX^{x,c,\pi,l}(t)&=rX^{x,c,\pi,l}(t)dt+\pi(t)(\mu-r)dt-(c(t)+wl(t)+d-w\bar{L})dt+\sigma\pi(t)dB(t)\\ &=rX^{x,c,\pi,l}(t)dt-(c(t)+wl(t)+d-w\bar{L})dt+\sigma\pi(t)d\tilde{B}(t)\\ &=rX^{x,c,\pi,l}(t)dt-(c(t)+wl(t)+d-w\bar{L})dt+\frac{\rho(t)}{\xi(t)}d\tilde{B}(t),\end{split}

the second equality also comes from changing the measure by B~(t)B(t)+θt\tilde{B}(t)\triangleq B(t)+\theta t. We can observe that G(t)=Xx,c,π,l(t)G(t)=X^{x,c,\pi,l}(t) a.s. on [0,τ][0,\tau], which concludes the proof of this lemma.

Appendix C Proof of Lemma 3.3

Remark C.1.

In this section we prove Lemma 3.3. Moreover, we also show that the conditions z¯<y~,\bar{z}\!<\!\tilde{y}, defined below, and z¯<z^PR\bar{z}\!<\!\hat{z}_{\scriptscriptstyle PR} hold true.

The proof here refers to Oksendal (2013) Section 10, Example 10.3.1.444B. Oksendal, Stochastic differential equations: an introduction with applications. Springer Science & Business Media, 2013. First of all, [5, Lemma 2.1] shows that

u~(z)=[A1zδ(1k)δ(1k)1wLz]𝕀{0<z<y~}+[A2z1kk]𝕀{zy~},\tilde{u}(z)\!=\!\left[A_{1}z^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!-\!wLz\right]\mathbb{I}_{\{0<z<\tilde{y}\}}\!+\!\left[A_{2}z^{\!-\!\frac{1\!-\!k}{k}}\right]\mathbb{I}_{\{z\geq\tilde{y}\}},

with A11δ+δkδ(1k)L(1k)(1δ)1δ(1k)A_{1}\!\triangleq\!\frac{1\!-\!\delta\!+\!\delta k}{\delta(1\!-\!k)}L^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}, A2kδ(1k)(1δδw)(1k)(1δ)kA_{2}\triangleq\frac{k}{\delta(1\!-\!k)}\left(\frac{1\!-\!\delta}{\delta w}\right)^{\frac{(1\!-\!k)(1\!-\!\delta)}{k}}, and y~Lk(1δδw)1δ(1k)\tilde{y}\triangleq L^{\!-k}\left(\frac{1\!-\!\delta}{\delta w}\right)^{1\!-\!\delta(1\!-\!k)}.

Introducing two functions

g(t,z)eγtU~(z),G(t,z,w¯)g(t,z)+w¯=eγtU~(z)+w¯,g(t,z)\triangleq e^{-\gamma t}\tilde{U}(z),\quad G(t,z,\bar{w})\triangleq g(t,z)+\bar{w}=e^{-\gamma t}\tilde{U}(z)+\bar{w},

and an operator 𝒜PG(t,z,w¯)Gt+(γr)zGz+θ22z22Gz2+eγtu~(z)eγt(dwL¯)z\mathcal{A}_{P}G(t,z,\bar{w})\triangleq\frac{\partial G}{\partial t}+(\gamma-r)z\frac{\partial G}{\partial z}+\frac{\theta^{2}}{2}z^{2}\frac{\partial^{2}G}{\partial z^{2}}+e^{-\gamma t}\tilde{u}(z)-e^{-\gamma t}(d-w\bar{L})z, we can determine the continuous region as Ω1={(t,z,w¯):𝒜PG(t,z,w¯)>0}\Omega_{1}=\{(t,z,\bar{w}):\mathcal{A}_{P}G(t,z,\bar{w})>0\}. Moreover, since

𝒜PG(t,z,w¯)=γeγtU~(z)+(γr)zeγtU~(z)+θ22z2eγtU~′′(z)+eγt(u~(z)(dwL¯)z),\mathcal{A}_{P}G(t,z,\bar{w})=-\gamma e^{-\gamma t}\tilde{U}(z)+(\gamma-r)ze^{-\gamma t}\tilde{U}^{\prime}(z)+\frac{\theta^{2}}{2}z^{2}e^{-\gamma t}\tilde{U}^{\prime\prime}(z)+e^{-\gamma t}\left(\tilde{u}(z)-(d-w\bar{L})z\right),

defining a new function h(z)=γU~(z)+(γr)zU~(z)+θ22z2U~′′(z)+u~(z)(dwL¯)zh(z)=-\gamma\tilde{U}(z)+(\gamma-r)z\tilde{U}^{\prime}(z)+\frac{\theta^{2}}{2}z^{2}\tilde{U}^{\prime\prime}(z)+\tilde{u}(z)-(d-w\bar{L})z, the continuous region can be rewritten as Ω1={z>0:h(z)>0}\Omega_{1}=\{z>0:h(z)>0\}. Since the function U~(z)\tilde{U}(z) takes two different forms based on the value of RpostR_{post}, we split the remaining discussion also into two cases: Rpost=drR_{post}=\frac{d}{r} and Rpost>drR_{post}>\frac{d}{r}.

(1) For Rpost=drR_{post}\!=\!\frac{d}{r}, we have U~(z)=1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz\tilde{U}(z)\!=\!\frac{1\!-\!\delta(1\!-\!k)}{\delta(1\!-\!k)}K_{1}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\!z^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!-\!\frac{d}{r}z. After the basic calculation, we get

h(z)=δ(1k)1δ(1k)L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1+wL¯z+u~(z)=δ(1k)1δ(1k)L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1+wL¯z+[A1zδ(1k)δ(1k)1wLz]𝕀{0<z<y~}+[A2z1kk]𝕀{zy~}.\begin{split}h(z)&=\frac{\delta(1-k)-1}{\delta(1-k)}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+w\bar{L}z+\tilde{u}(z)\\ &=\frac{\delta(1\!-\!k)\!-\!1}{\delta(1\!-\!k)}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}z^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!+\!w\bar{L}z\!+\!\left[A_{1}z^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!-\!wLz\right]\!\mathbb{I}_{\{0<z<\tilde{y}\}}\!+\!\left[A_{2}z^{-\frac{1\!-\!k}{k}}\right]\!\mathbb{I}_{\{z\geq\tilde{y}\}}.\end{split} (C.1)

h(z)h(z) inherits the piecewise form from the function u~(z)\tilde{u}(z). Afterwards, determining the continuous region corresponds to characterize the features of the zero of h(z)h(z). We begin claiming its convexity by the second derivative function. On the interval 0<z<y~0<z<\tilde{y}, we can directly determine the sign of h′′(z)h^{\prime\prime}(z) with h′′(z)=1δ(1k)1z2δ(1k)δ(1k)1[L¯(1k)(1δ)1δ(1k)L(1k)(1δ)1δ(1k)]>0h^{\prime\prime}(z)=\frac{1}{\delta(1-k)-1}z^{\frac{2-\delta(1-k)}{\delta(1-k)-1}}\left[\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}-L^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\right]>0. As for the interval z>y~z>\tilde{y}, the corresponding second derivative function h′′(z)h^{\prime\prime}(z) is

h′′(z)=1δ(1k)1L¯(1k)(1δ)1δ(1k)z2δ(1k)δ(1k)1+1δk(1δδw)(1k)(1δ)kz1+kk.h^{\prime\prime}(z)=\frac{1}{\delta(1-k)-1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{2-\delta(1-k)}{\delta(1-k)-1}}+\frac{1}{\delta k}\left(\frac{1-\delta}{\delta w}\right)^{\frac{(1-k)(1-\delta)}{k}}z^{-\frac{1+k}{k}}.

By solving the inequality, 1δ(1k)1L¯(1k)(1δ)1δ(1k)z2δ(1k)δ(1k)1+1δk(1δδw)(1k)(1δ)kz1+kk>0\frac{1}{\delta(1-k)-1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{2-\delta(1-k)}{\delta(1-k)-1}}+\frac{1}{\delta k}\left(\frac{1-\delta}{\delta w}\right)^{\frac{(1-k)(1-\delta)}{k}}z^{-\frac{1+k}{k}}>0, we get

z>L¯k(1δ(1k)δk)k(δ(1k)1)(1k)(δ1)(1δδw)1δ(1k).z>\bar{L}^{-k}\left(\frac{1-\delta(1-k)}{\delta k}\right)^{\frac{k(\delta(1-k)-1)}{(1-k)(\delta-1)}}\left(\frac{1-\delta}{\delta w}\right)^{1-\delta(1-k)}.

Since

y~L¯k(1δ(1k)δk)k(δ(1k)1)(1k)(δ1)(1δδw)1δ(1k)=(1δδw)1δ(1k)[Lk(1δ(1k)δk)k(δ(1k)1)(1k)(δ1)L¯k]>0,\tilde{y}-\bar{L}^{\!-\!k}\left(\!\frac{1\!-\!\delta(1\!-\!k)}{\delta k}\!\right)^{\frac{k(\delta(1\!-\!k)\!-\!1)}{(1\!-\!k)(\delta\!-\!1)}}\!\!\left(\!\frac{1\!-\!\delta}{\delta w}\!\right)^{1\!-\!\delta(1\!-\!k)}\!\!\!=\!\left(\!\frac{1\!-\!\delta}{\delta w}\!\right)^{1\!-\!\delta(1\!-\!k)}\!\!\left[\!L^{\!-\!k}\!-\!\left(\!\frac{1\!-\!\delta(1\!-\!k)}{\delta k}\!\right)^{\frac{k(\delta(1\!-\!k)\!-\!1)}{(1\!-\!k)(\delta\!-\!1)}}\!\!\bar{L}^{\!-\!k}\!\right]\!\!>\!0,

h′′(z)>0h^{\prime\prime}(z)>0 keeps true for z>y~z>\tilde{y}. Besides, considering the condition limzy~h′′(z)=limzy~h′′(z)\lim\limits_{z\uparrow\tilde{y}}h^{\prime\prime}(z)=\lim\limits_{z\downarrow\tilde{y}}h^{\prime\prime}(z), we can conclude that the function h(z)h(z) is strictly convex on the interval z>0z>0. Then we move to claim h(y~)>0h(\tilde{y})>0: before this, a new function is introduced as

f(z)=δ(1k)1δ(1k)L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1+kδ(1k)(1δδw)(1k)(1δ)kz1kk+wL¯z,z>0,f(z)=\frac{\delta(1-k)-1}{\delta(1-k)}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{k}{\delta(1-k)}\left(\frac{1-\delta}{\delta w}\right)^{\frac{(1-k)(1-\delta)}{k}}\!\!\!z^{-\frac{1-k}{k}}+w\bar{L}z,\quad z>0,

and its derivative functions are

f(z)=z1δ(1k)1L¯(1k)(1δ)1δ(1k)1δ(1δδw)(1k)(1δ)kz1k+wL¯,f^{\prime}(z)=z^{\frac{1}{\delta(1-k)-1}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}-\frac{1}{\delta}\left(\frac{1-\delta}{\delta w}\right)^{\frac{(1-k)(1-\delta)}{k}}z^{-\frac{1}{k}}+w\bar{L},
f′′(z)=1δ(1k)1z2δ(1k)δ(1k)1L¯(1k)(1δ)1δ(1k)+1δk(1δδw)(1k)(1δ)kz1+kk.f^{\prime\prime}(z)=\frac{1}{\delta(1-k)-1}z^{\frac{2-\delta(1-k)}{\delta(1-k)-1}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}+\frac{1}{\delta k}\left(\frac{1-\delta}{\delta w}\right)^{\frac{(1-k)(1-\delta)}{k}}z^{-\frac{1+k}{k}}.

Defining y~1=L¯k(1δδw)1δ(1k)\tilde{y}_{1}=\bar{L}^{-k}\left(\frac{1-\delta}{\delta w}\right)^{1-\delta(1-k)}, it can be obtained that

f(y~1)=δ(1k)1δ(1k)L¯1k(1δδw)δ(1k)+kδ(1k)L¯1k(1δδw)δ(1k)+wL¯1k(1δδw)1δ(1k)=0,f(\tilde{y}_{1})\!=\!\frac{\delta(1\!-\!k)\!-\!1}{\delta(1\!-\!k)}\bar{L}^{1\!-\!k}\!\left(\frac{1\!-\!\delta}{\delta w}\right)^{\!-\!\delta(1\!-\!k)}\!\!+\!\frac{k}{\delta(1\!-\!k)}\bar{L}^{1\!-\!k}\!\left(\frac{1\!-\!\delta}{\delta w}\right)^{\!-\!\delta(1\!-\!k)}\!\!+\!w\bar{L}^{1\!-\!k}\left(\frac{1\!-\!\delta}{\delta w}\right)^{1\!-\!\delta(1\!-\!k)}\!\!\!=\!0,
f(y~1)=(L¯k(1δδw)1δ(1k))1δ(1k)1L¯(1k)(1δ)1δ(1k)1δ(1δδw)(1k)(1δ)k(L¯k(1δδw)1δ(1k))1k+wL¯=0.f^{\prime}(\tilde{y}_{1})\!=\!\!\left(\!\bar{L}^{\!-\!k}\!\left(\!\frac{1\!-\!\delta}{\delta w}\!\right)^{1\!-\!\delta(1\!-\!k)}\!\right)^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\!\!\!\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\!\!-\!\frac{1}{\delta}\!\left(\!\frac{1\!-\!\delta}{\delta w}\!\right)^{\frac{(1\!-\!k)(1\!-\!\delta)}{k}}\!\!\!\left(\!\bar{L}^{\!-\!k}\!\left(\!\frac{1\!-\!\delta}{\delta w}\!\right)^{1\!-\!\delta(1\!-\!k)}\!\right)^{\!-\!\frac{1}{k}}\!+\!w\bar{L}\!\!=\!0.

For the second derivative, f′′(z)>0f^{\prime\prime}(z)\!>\!0 is equivalent to z>(1δ(1k)δk)k(δ(1k)1)(1k)(δ1)(1δδw)1δ(1k)L¯kz\!>\!\left(\frac{1\!-\!\delta(1\!-\!k)}{\delta k}\right)^{\frac{k(\delta(1-k)-1)}{(1-k)(\delta-1)}}\left(\frac{1\!-\!\delta}{\delta w}\right)^{1\!-\!\delta(1\!-\!k)}\bar{L}^{\!-\!k}. Since 0<(1δ(1k)δk)k(δ(1k)1)(1k)(δ1)<10\!<\!\left(\frac{1-\delta(1-k)}{\delta k}\right)^{\frac{k(\delta(1-k)-1)}{(1-k)(\delta-1)}}\!\!<\!1, we have y~1>(1δ(1k)δk)k(δ(1k)1)(1k)(δ1)(1δδw)1δ(1k)L¯k\tilde{y}_{1}\!>\!\left(\frac{1\!-\!\delta(1\!-\!k)}{\delta k}\right)^{\frac{k(\delta(1-k)-1)}{(1-k)(\delta-1)}}\left(\frac{1\!-\!\delta}{\delta w}\right)^{1\!-\!\delta(1\!-\!k)}\bar{L}^{-k}, which results in f′′(z)>0f^{\prime\prime}(z)>0 for z>y~1z>\tilde{y}_{1}. Then the fact f(y~1)=0f^{\prime}(\tilde{y}_{1})=0 indicates that f(z)>0f^{\prime}(z)>0 for z>y~1z>\tilde{y}_{1}, which means f(z)f(z) is strictly increasing on the corresponding interval. Considering the relationship y~=Lk(1δδw)1δ(1k)>y~1\tilde{y}=L^{-k}\left(\frac{1-\delta}{\delta w}\right)^{1-\delta(1-k)}>\tilde{y}_{1}, we can observe the positive value of h(y~)h(\tilde{y}) through

0=f(y~1)<f(y~)=h(y~).0=f(\tilde{y}_{1})<f(\tilde{y})=h(\tilde{y}). (C.2)

Finally, in view of the limitations

limz0h(z)=limz0[δ(1k)1δ(1k)zδ(1k)δ(1k)1(L¯(1k)(1δ)1δ(1k)L(1k)(1δ)1δ(1k))+wz(L¯L)]=0,\lim\limits_{z\downarrow 0}h(z)=\lim\limits_{z\downarrow 0}\left[\frac{\delta(1-k)-1}{\delta(1-k)}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}\left(\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}-L^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\right)+wz(\bar{L}-L)\right]=0,
limz0h(z)=limz0[z1δ(1k)1(L¯(1k)(1δ)1δ(1k)L(1k)(1δ)1δ(1k))+w(L¯L)]=,\lim\limits_{z\downarrow 0}h^{\prime}(z)=\lim\limits_{z\downarrow 0}\left[z^{\frac{1}{\delta(1-k)-1}}\left(\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}-L^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\right)+w(\bar{L}-L)\right]=-\infty,

and the properties h′′(z)>0h^{\prime\prime}(z)>0 for z>0z>0, h(y~)>0h(\tilde{y})>0, we can conclude that there is a unique zero of h(z)h(z), which is denoted as z¯\bar{z}, satisfying z¯<y~\bar{z}<\tilde{y} and h(z¯)=L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1+wL¯+u~(z¯)h^{\prime}(\bar{z})=\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{1}{\delta(1-k)-1}}+w\bar{L}+\tilde{u}^{\prime}(\bar{z}). Hence, the continuous region is Ω1={(t,z,w¯):𝒜PG(s,z,w¯)>0}={z:h(z)>0}={z>z¯}\Omega_{1}=\{(t,z,\bar{w}):\mathcal{A}_{P}G(s,z,\bar{w})>0\}=\{z:h(z)>0\}=\{z>\bar{z}\}.

(2) For Rpost>drR_{post}>\frac{d}{r}, we have

U~(z)={B2,PRz^PRn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z^PRδ(1k)δ(1k)1drz^PRRpost(zz^PR),zz^PR,B2,PRzn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz,0<z<z^PR,\tilde{U}(z)=\begin{cases}B_{2,\scriptscriptstyle PR}\hat{z}_{\scriptscriptstyle PR}^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}\hat{z}_{\scriptscriptstyle PR}-R_{post}(z-\hat{z}_{\scriptscriptstyle PR}),&z\!\geq\!\hat{z}_{\scriptscriptstyle PR},\\ B_{2,\scriptscriptstyle PR}z^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}z,&0\!<\!z\!<\!\hat{z}_{\scriptscriptstyle PR},\end{cases}

then the function h(z)h(z) on the interval 0<z<z^PR0<z<\hat{z}_{\scriptscriptstyle PR} is obtained as

h(z)=B2,PRzn2[γ+(γr)n2+θ22n2(n21)]+δ(1k)1δ(1k)L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1+wL¯z+u~(z).h(z)\!=\!B_{2,\scriptscriptstyle PR}z^{n_{2}}\left[-\!\gamma\!+\!(\gamma\!-\!r)n_{2}\!+\!\frac{\theta^{2}}{2}n_{2}(n_{2}\!-\!1)\right]+\frac{\delta(1-k)-1}{\delta(1-k)}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+w\bar{L}z+\tilde{u}(z).

From n1+n2=γrθ22θ22n_{1}+n_{2}=-\frac{\gamma-r-\frac{\theta^{2}}{2}}{\frac{\theta^{2}}{2}} and n1n2=γθ22n_{1}n_{2}=-\frac{\gamma}{\frac{\theta^{2}}{2}}, we can deduce γ+(γr)n2+θ22n2(n21)=0-\gamma+(\gamma-r)n_{2}+\frac{\theta^{2}}{2}n_{2}(n_{2}-1)=0. Hence, the function h(z)h(z) is reduced as

h(z)=δ(1k)1δ(1k)L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1+wL¯z+u~(z),0<z<z^PR.h(z)=\frac{\delta(1-k)-1}{\delta(1-k)}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+w\bar{L}z+\tilde{u}(z),\quad 0<z<\hat{z}_{\scriptscriptstyle PR}.

Compared to Equation (C.1), we can observe that h(z)h(z) adopts the same form but applies to the different intervals. As for the interval zz^PRz\geq\hat{z}_{\scriptscriptstyle PR}, considering that U~(z)=vPR(z^PR)Rpost(zz^PR)\tilde{U}(z)=v_{\scriptscriptstyle PR}(\hat{z}_{\scriptscriptstyle PR})-R_{post}(z-\hat{z}_{\scriptscriptstyle PR}), and the condition (A.7) is applicable at the point z=z^PRz=\hat{z}_{\scriptscriptstyle PR}, we get

h(z)=γ(vPR(z^PR)Rpost(zz^PR))(γr)zRpost+u~(z)(dwL¯)z=(rRpostd)(zz^PR)u~PR(z^PR)+u~(z)+wL¯z.\begin{split}h(z)&=-\gamma\left(v_{\scriptscriptstyle PR}(\hat{z}_{\scriptscriptstyle PR})-R_{post}(z-\hat{z}_{\scriptscriptstyle PR})\right)-(\gamma-r)zR_{post}+\tilde{u}(z)-(d-w\bar{L})z\\ &=(rR_{post}-d)(z-\hat{z}_{\scriptscriptstyle PR})-\tilde{u}_{\scriptscriptstyle PR}(\hat{z}_{\scriptscriptstyle PR})+\tilde{u}(z)+w\bar{L}z.\end{split}

h′′(z)=u~′′(z)>0h^{\prime\prime}(z)\!=\!\tilde{u}^{\prime\prime}(z)\!>\!0 shows that h(z)h(z) is strictly convex on (z^PR,)(\hat{z}_{\scriptscriptstyle PR},\infty). Then a contradiction is constructed to prove z¯<z^PR\bar{z}\!<\!\hat{z}_{\scriptscriptstyle PR}. We first use r=θ22(1n2)(n11)r\!=\!\frac{\theta^{2}}{2}(1\!-\!n_{2})(n_{1}\!-\!1) and n2>1>δ(1k)δ(1k)1>0n_{2}\!>\!1\!>\!\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}\!>\!0 to derive the condition

(1n2)(1δ(1k))n2(δ(1k)1)δ(1k)1K1=rγrδ(1k)θ22δ(1k)1δ(1k)γrδ(1k)+θ22n2<r,\frac{(1-n_{2})(1-\delta(1-k))}{n_{2}(\delta(1-k)-1)-\delta(1-k)}\frac{1}{K_{1}}=r\frac{\gamma-r\delta(1-k)-\frac{\theta^{2}}{2}\frac{\delta(1-k)}{1-\delta(1-k)}}{\gamma-r\delta(1-k)+\frac{\theta^{2}}{2}n_{2}}<r,

which gives us

z^PR1δ(1k)1=L¯(1k)(1δ)δ(1k)1(1n2)(1δ(1k))n2(δ(1k)1)δ(1k)(Rpostdr)K1<L¯(1k)(1δ)δ(1k)1(rRpostd).\hat{z}_{\scriptscriptstyle PR}^{\frac{1}{\delta(1-k)-1}}=\bar{L}^{\frac{(1-k)(1-\delta)}{\delta(1-k)-1}}\frac{(1-n_{2})(1-\delta(1-k))}{n_{2}(\delta(1-k)-1)-\delta(1-k)}\frac{\left(R_{post}-\frac{d}{r}\right)}{K_{1}}<\bar{L}^{\frac{(1-k)(1-\delta)}{\delta(1-k)-1}}\left(rR_{post}-d\right).

Assuming z¯z^PR\bar{z}\geq\hat{z}_{\scriptscriptstyle PR}, we can observe the contradiction through

h(z¯)=L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1+wL¯+u~(z¯)L¯(1k)(1δ)1δ(1k)z^PR1δ(1k)1+wL¯+u~(z¯)<L¯(1k)(1δ)1δ(1K)L¯(1k)(1δ)δ(1K)1(rRpostd)+wL¯+u~(z¯)=(rRpostd)+wL¯+u~(z¯)=h(z¯).\begin{split}h^{\prime}(\bar{z})&=\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{1}{\delta(1-k)-1}}+w\bar{L}+\tilde{u}^{\prime}(\bar{z})\\ &\leq\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\hat{z}_{\scriptscriptstyle PR}^{\frac{1}{\delta(1-k)-1}}+w\bar{L}+\tilde{u}^{\prime}(\bar{z})\\ &<\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-K)}}\bar{L}^{\frac{(1-k)(1-\delta)}{\delta(1-K)-1}}(rR_{post}-d)+w\bar{L}+\tilde{u}^{\prime}(\bar{z})\\ &=(rR_{post}-d)+w\bar{L}+\tilde{u}^{\prime}(\bar{z})=h^{\prime}(\bar{z}).\end{split}

Then the condition z¯<z^PR\bar{z}<\hat{z}_{\scriptscriptstyle PR} implies h(z^PR)>0h(\hat{z}_{\scriptscriptstyle PR})>0 and

limzz^PRh(z)=limzz^PR[L¯(1k)(1δ)1δ(1k)z1δ(1k)1+u~(z)+wL¯]=u~PR(z^PR)+u~(z^PR)+wL¯0.\lim\limits_{z\uparrow\hat{z}_{\scriptscriptstyle PR}}h^{\prime}(z)=\lim\limits_{z\uparrow\hat{z}_{\scriptscriptstyle PR}}\left[\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{1}{\delta(1-k)-1}}+\tilde{u}^{\prime}(z)+w\bar{L}\right]=-\tilde{u}_{\scriptscriptstyle PR}^{\prime}(\hat{z}_{\scriptscriptstyle PR})+\tilde{u}^{\prime}(\hat{z}_{\scriptscriptstyle PR})+w\bar{L}\geq 0.

Afterwards, we have

limzz^PRh(z)=rRpostd+u~(z^PR)+wL¯rRpostd+u~PR(z^PR)=rRpostdz^PR1δ(1k)1L¯(1k)(1δ)1δ(1k)=rRpostd(1n2)(1δ(1k))n2(δ(1k)1)δ(1k)1K1(Rpostdr)>0,\begin{split}\lim\limits_{z\downarrow\hat{z}_{\scriptscriptstyle PR}}h^{\prime}(z)&=rR_{post}-d+\tilde{u}^{\prime}(\hat{z}_{\scriptscriptstyle PR})+w\bar{L}\\ &\geq rR_{post}-d+\tilde{u}_{\scriptscriptstyle PR}^{\prime}(\hat{z}_{\scriptscriptstyle PR})\\ &=rR_{post}-d-\hat{z}_{\scriptscriptstyle PR}^{\frac{1}{\delta(1-k)-1}}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\\ &=rR_{post}-d-\frac{(1-n_{2})(1-\delta(1-k))}{n_{2}(\delta(1-k)-1)-\delta(1-k)}\frac{1}{K_{1}}\left(R_{post}-\frac{d}{r}\right)>0,\end{split}

which indicates that h(z)h(z) is strictly increasing for z>z^PRz>\hat{z}_{\scriptscriptstyle PR} regarding the convex property already shown. Therefore, z¯\bar{z} is the unique zero of function h(z)h(z), and satisfies z¯<z^PR\bar{z}<\hat{z}_{\scriptscriptstyle PR}. The last step is to claim z¯<y~\bar{z}<\tilde{y} under this case, which is equivalent to h(y~)>0h(\tilde{y})>0 and discussed in two different situations. If y~z^PR\tilde{y}\leq\hat{z}_{\scriptscriptstyle PR}, using the result (C.2), we have

h(y~)=δ(1k)1δ(1k)L¯(1k)(1δ)1δ(1k)y~δ(1k)δ(1k)1+wL¯y~+u~(y~)=f(y~)>0,h(\tilde{y})=\frac{\delta(1-k)-1}{\delta(1-k)}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\tilde{y}^{\frac{\delta(1-k)}{\delta(1-k)-1}}+w\bar{L}\tilde{y}+\tilde{u}(\tilde{y})=f(\tilde{y})>0,

otherwise, if y~>z^PR\tilde{y}\!>\!\hat{z}_{\scriptscriptstyle PR}, we use the increasing property of h(z)h(z) on (z^PR,)(\hat{z}_{\scriptscriptstyle PR},\infty) to directly obtain h(y~)>0h(\tilde{y})\!>\!0.

Appendix D Calculation of Variational Inequalities (3.3)

The solution of (3.3) is split into two different cases based on the value of RpostR_{post}, namely Rpost=drR_{post}\!=\!\frac{d}{r} and Rpost>drR_{post}\!>\!\frac{d}{r}. Following [2, Appendix A], we take the time-separated form of function ϕ(t,z)=eγtv(z){\phi}(t,z)\!=\!e^{\!-\!\gamma t}v(z) for solving the above variational inequalities explicitly.

We recall that [5, Lemma 2.1] shows that u~(z)=[A1zδ(1k)δ(1k)1wLz]𝕀{0<z<y~}+[A2z1kk]𝕀{zy~}\tilde{u}(z)\!=\!\left[A_{1}z^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!-\!wLz\right]\mathbb{I}_{\{0<z<\tilde{y}\}}\!+\!\left[A_{2}z^{\!-\!\frac{1\!-\!k}{k}}\right]\mathbb{I}_{\{z\geq\tilde{y}\}}, with A11δ+δkδ(1k)L(1k)(1δ)1δ(1k)A_{1}\!\triangleq\!\frac{1\!-\!\delta\!+\!\delta k}{\delta(1\!-\!k)}L^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}, A2kδ(1k)(1δδw)(1k)(1δ)kA_{2}\triangleq\frac{k}{\delta(1\!-\!k)}\left(\frac{1\!-\!\delta}{\delta w}\right)^{\frac{(1\!-\!k)(1\!-\!\delta)}{k}}, and y~Lk(1δδw)1δ(1k)\tilde{y}\triangleq L^{\!-k}\left(\frac{1\!-\!\delta}{\delta w}\right)^{1\!-\!\delta(1\!-\!k)}. Moreover, n1n_{1} and n2n_{2} are the roots of the second-order equation θ22n2+(γrθ22)nγ=0\frac{\theta^{2}}{2}n^{2}+\left(\gamma-r-\frac{\theta^{2}}{2}\right)n-\gamma=0, and satisfy

n1,2=γrθ22θ2(γrθ22θ2)2+2γθ2,n1<0,n2>1.n_{1,2}=-\frac{\gamma-r-\frac{\theta^{2}}{2}}{\theta^{2}}\mp\sqrt{\left(\frac{\gamma-r-\frac{\theta^{2}}{2}}{\theta^{2}}\right)^{2}+\frac{2\gamma}{\theta^{2}}},\qquad n_{1}<0,\quad n_{2}>1.
Case 1. Rpre=dwL¯rR_{pre}=\frac{d-w\bar{L}}{r} &\& Rpost=drR_{post}=\frac{d}{r}

From the condition (V1)(V1) of (3.3), the following differential equation holds in the region z>z¯z>\bar{z},

γv(z)+(γr)zv(z)+12θ2z2v′′(z)+u~(z)(dwL¯)z=0,-\gamma v(z)+(\gamma-r)zv^{\prime}(z)+\frac{1}{2}\theta^{2}z^{2}v^{\prime\prime}(z)+\tilde{u}(z)-(d-w\bar{L})z=0, (D.1)

whose solution takes a form as

v(z)={B11zn1+B21zn2+A1Γ1zδ(1k)δ(1k)1+w(L¯L)drz,z¯<z<y~,B12zn1+B22zn2+A2Γ2z1kk+wL¯drz,zy~.v(z)=\begin{cases}B_{11}z^{n_{1}}+B_{21}z^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)-d}{r}z,&\bar{z}<z<\tilde{y},\\ B_{12}z^{n_{1}}+B_{22}z^{n_{2}}+\frac{A_{2}}{\Gamma_{2}}z^{-\frac{1-k}{k}}+\frac{w\bar{L}-d}{r}z,&z\geq\tilde{y}.\end{cases}

Since n2>0n_{2}>0, for the sake of avoiding the explosion of the term zn2z^{n_{2}} as zz goes to \infty, we set B22=0B_{22}=0. Then, the condition (V4)(V4) of (3.3) enables us to obtain

v(z¯)=U~(z¯)=1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1drz¯,v(\bar{z})=\tilde{U}(\bar{z})=\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}\bar{z},

the second equality results from the condition Rpost=drR_{post}=\frac{d}{r}. Furthermore, combining with the smooth condition at the point z=y~z=\tilde{y}, we can construct a four-equations system to determine the parameters B11B_{11}, B21B_{21}, B12B_{12} and z¯\bar{z}.

  • 𝒞0\mathcal{C}^{0} condition at z=z¯z=\bar{z}

    B11z¯n1+B21z¯n2+A1Γ1z¯δ(1k)δ(1k)1+w(L¯L)rz¯=1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1;B_{11}\bar{z}^{n_{1}}+B_{21}\bar{z}^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)}{r}\bar{z}=\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}};
  • 𝒞1\mathcal{C}^{1} condition at z=z¯z=\bar{z}

    n1B11z¯n11+n2B21z¯n21+δ(1k)δ(1k)1A1Γ1z¯1δ(1k)1+w(L¯L)r=K1L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1;n_{1}B_{11}\bar{z}^{n_{1}\!-\!1}\!+\!n_{2}B_{21}\bar{z}^{n_{2}\!-\!1}\!+\!\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\!+\!\frac{w(\bar{L}\!-\!L)}{r}=\!-\!K_{1}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}};
  • 𝒞0\mathcal{C}^{0} condition at z=y~z=\tilde{y}

    B11y~n1+B21y~n2+A1Γ1y~δ(1k)δ(1k)1wLry~=B12y~n1+A2Γ2y~1kk;B_{11}\tilde{y}^{n_{1}}+B_{21}\tilde{y}^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}\tilde{y}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{wL}{r}\tilde{y}=B_{12}\tilde{y}^{n_{1}}+\frac{A_{2}}{\Gamma_{2}}\tilde{y}^{-\frac{1-k}{k}};
  • 𝒞1\mathcal{C}^{1} condition at z=y~z=\tilde{y}

    n1B11y~n11+n2B21y~n21+δ(1k)δ(1k)1A1Γ1y~1δ(1k)1wLr=n1B12y~n111kkA2Γ2y~1k.n_{1}B_{11}\tilde{y}^{n_{1}-1}+n_{2}B_{21}\tilde{y}^{n_{2}-1}+\frac{\delta(1-k)}{\delta(1-k)-1}\frac{A_{1}}{\Gamma_{1}}\tilde{y}^{\frac{1}{\delta(1-k)-1}}-\frac{wL}{r}=n_{1}B_{12}\tilde{y}^{n_{1}-1}-\frac{1-k}{k}\frac{A_{2}}{\Gamma_{2}}\tilde{y}^{-\frac{1}{k}}.
Case 2. Rpre=dwL¯rR_{pre}=\frac{d-w\bar{L}}{r} &\& Rpost>drR_{post}>\frac{d}{r}

Then we move to the second case with a different condition Rpost>drR_{post}>\frac{d}{r} compared to Case 1, which mainly affects the post-retirement part and leads to a different form of U~(z)\tilde{U}(z). Lemma 3.1 shows that the corresponding Legendre-Fenchel transform of post-retirement value function U~(z)\tilde{U}(z) is

U~(z)=B2,PRzn2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)zδ(1k)δ(1k)1drz.\tilde{U}(z)=B_{2,\scriptscriptstyle PR}z^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}z.

Meanwhile, the dual transform involving the pre-retirement part u~(z)\tilde{u}(z) stays the same; hence Equation (V1)(V1) from (3.3) takes the identical solution. Afterwards, using the smooth fit conditions at z=z¯z\!=\!\bar{z} and z=y~z\!=\!\tilde{y}, we construct a four-equations system to achieve the unknowns, B11B_{11}, B21B_{21}, B12B_{12} and z¯\bar{z}.

  • 𝒞0\mathcal{C}^{0} condition at z=z¯z=\bar{z}

    B11z¯n1+B21z¯n2+A1Γ1z¯δ(1k)δ(1k)1+w(L¯L)rz¯=B2,PRz¯n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1;B_{11}\bar{z}^{n_{1}}\!+\!B_{21}\bar{z}^{n_{2}}\!+\!\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!+\!\frac{w(\bar{L}\!-\!L)}{r}\bar{z}\!=\!B_{2,\scriptscriptstyle PR}\bar{z}^{n_{2}}\!+\!\frac{1\!-\!\delta(1\!-\!k)}{\delta(1\!-\!k)}K_{1}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\bar{z}^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}};
  • 𝒞1\mathcal{C}^{1} condition at z=z¯z=\bar{z}

    n1B11z¯n11+n2B21z¯n21+δ(1k)δ(1k)1A1Γ1z¯1δ(1k)1+w(L¯L)r=n2B2,PRz¯n21K1L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1;n_{1}B_{11}\bar{z}^{n_{1}\!-\!1}\!+\!n_{2}B_{21}\bar{z}^{n_{2}\!-\!1}\!+\!\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}\!\frac{A_{1}}{\Gamma_{1}}\!\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\!+\!\frac{w(\bar{L}\!-\!L)}{r}\!=\!n_{2}B_{2,\scriptscriptstyle PR}\bar{z}^{n_{2}\!-\!1}\!-\!K_{1}\!\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\!\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}};
  • 𝒞0\mathcal{C}^{0} condition at z=y~z=\tilde{y}

    B11y~n1+B21y~n2+A1Γ1y~δ(1k)δ(1k)1wLry~=B12y~n1+A2Γ2y~1kk;B_{11}\tilde{y}^{n_{1}}+B_{21}\tilde{y}^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}\tilde{y}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{wL}{r}\tilde{y}=B_{12}\tilde{y}^{n_{1}}+\frac{A_{2}}{\Gamma_{2}}\tilde{y}^{-\frac{1-k}{k}};
  • 𝒞1\mathcal{C}^{1} condition at z=y~z=\tilde{y}

    n1B11y~n11+n2B21y~n21+δ(1k)δ(1k)1A1Γ1y~1δ(1k)1wLr=n1B12y~n111kkA2Γ2y~1k.n_{1}B_{11}\tilde{y}^{n_{1}-1}+n_{2}B_{21}\tilde{y}^{n_{2}-1}+\frac{\delta(1-k)}{\delta(1-k)-1}\frac{A_{1}}{\Gamma_{1}}\tilde{y}^{\frac{1}{\delta(1-k)-1}}-\frac{wL}{r}=n_{1}B_{12}\tilde{y}^{n_{1}-1}-\frac{1-k}{k}\frac{A_{2}}{\Gamma_{2}}\tilde{y}^{-\frac{1}{k}}.

After obtaining the closed forms of v(z)v(z) separately in Case 1 and Case 2, and given the initial wealth xRprex\geq R_{pre}, the optimal Lagrange multiplier λ\lambda^{*} can be acquired through solving the equation x=v(λ)x=-v^{\prime}(\lambda^{*}), due to the fact that V(x)=infλ>0[V~(λ)+λx]=infλ>0[v(λ)+λx]=v(λ)+λxV(x)=\inf\limits_{\lambda>0}\big{[}\tilde{V}(\lambda)+\lambda x\big{]}=\inf\limits_{\lambda>0}\big{[}v(\lambda)+\lambda x\big{]}=v(\lambda^{*})+\lambda^{*}x holds under the differentiable property of v()v(\cdot). Then the optimal dual process of wealth follows Z(t)=λeγtH(t)Z^{*}(t)=\lambda^{*}e^{\gamma t}H(t).

Proposition D.1.

For Case 1 and Case 2, the optimal retirement time is τ=inft0{Z(t)z¯}\tau^{*}=\inf\limits_{t\geq 0}\{Z^{*}(t)\leq\bar{z}\}, the optimal consumption-portfolio-leisure plan {c(t),π(t),l(t)}\{c^{*}(t),\pi^{*}(t),l^{*}(t)\} before retirement is given by

c(t)={L(1k)(1δ)δ(1k)1(Z(t))1δ(1k)1,z¯<Z(t)<y~,(1δδw)(1δ)(1k)k(Z(t))1k,Z(t)y~,c^{*}(t)=\begin{cases}L^{-\frac{(1-k)(1-\delta)}{\delta(1-k)-1}}(Z^{*}(t))^{\frac{1}{\delta(1-k)-1}},&\bar{z}<Z^{*}(t)<\tilde{y},\\ \left(\frac{1-\delta}{\delta w}\right)^{\frac{(1-\delta)(1-k)}{k}}(Z^{*}(t))^{-\frac{1}{k}},&Z^{*}(t)\geq\tilde{y},\\ \end{cases}
l(t)={L,z¯<Z(t)<y~,(1δδw)δ(1k)1k(Z(t))1k,Z(t)y~,l^{*}(t)=\begin{cases}L,&\bar{z}<Z^{*}(t)<\tilde{y},\\ \left(\frac{1-\delta}{\delta w}\right)^{-\frac{\delta(1-k)-1}{k}}(Z^{*}(t))^{-\frac{1}{k}},&Z^{*}(t)\geq\tilde{y},\\ \end{cases}
π(t)={θσ[n1(n11)B11(Z(t))n11+n2(n21)B21(Z(t))n21+δ(1k)(δ(1k)1)2A1Γ1(Z(t))1δ(1k)1],z¯<Z(t)<y~,θσ[n1(n11)B12(Z(t))n11+1kk2A2Γ2(Z(t))1k],Z(t)y~.\pi^{*}(t)=\begin{cases}\frac{\theta}{\sigma}\left[n_{1}(n_{1}-1)B_{11}(Z^{*}(t))^{n_{1}-1}+n_{2}(n_{2}-1)B_{21}(Z^{*}(t))^{n_{2}-1}\right.&\\ \qquad\qquad\qquad\qquad\left.+\frac{\delta(1-k)}{(\delta(1-k)-1)^{2}}\frac{A_{1}}{\Gamma_{1}}(Z^{*}(t))^{\frac{1}{\delta(1-k)-1}}\right],&\bar{z}<Z^{*}(t)<\tilde{y},\\ \frac{\theta}{\sigma}\left[n_{1}(n_{1}\!-\!1)B_{12}(Z^{*}(t))^{n_{1}-1}+\frac{1-k}{k^{2}}\frac{A_{2}}{\Gamma_{2}}\right.(Z^{*}\left.(t))^{-\frac{1}{k}}\right],&Z^{*}(t)\geq\tilde{y}.\end{cases}
Proof.

The optimal consumption and leisure strategies come from [5, Lemma 2.1], and the optimal portfolio strategy is derived by π(t)=θσZ(t)v′′(Z(t))\pi^{*}(t)=\frac{\theta}{\sigma}Z^{*}(t)v^{\prime\prime}(Z^{*}(t)) from [7, Section 5, Theorem 3]. ∎

Appendix E Calculation of Variational Inequalities (3.5)

Recalling the condition z¯<y~\bar{z}<\tilde{y} in Lemma 3.3, the problem to be solved is split into four different cases depending on the relationship between y~\tilde{y} with z^\hat{z}, and RpostR_{post} with dr\frac{d}{r}. We provide a diagram for a clear classification.

0<z¯<y~z^0<\bar{z}<\tilde{y}\leq\hat{z} Case 4. 0<z¯<y~z^0<\bar{z}<\tilde{y}\leq\hat{z}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost>drR_{post}>\frac{d}{r} Case 3. 0<z¯<y~z^0<\bar{z}<\tilde{y}\leq\hat{z}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost=drR_{post}=\frac{d}{r}
0<z¯<z^<y~0<\bar{z}<\hat{z}<\tilde{y} Case 6. 0<z¯<z^<y~0<\bar{z}<\hat{z}<\tilde{y}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost>drR_{post}>\frac{d}{r} Case 5. 0<z¯<z^<y~0<\bar{z}<\hat{z}<\tilde{y}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost=drR_{post}=\frac{d}{r}

Additionally, we assume that ϕ(t,z){\phi}(t,z) takes the time-separated form, ϕ(t,z)=eγtv(z){\phi}(t,z)\!=\!e^{\!-\!\gamma t}v(z), as in [2, Appendix A].

Case 3. 0<z¯<y~z^0<\bar{z}<\tilde{y}\leq\hat{z}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost=drR_{post}=\frac{d}{r}

We begin with the condition (V3)(V3) in (3.5), the following differential equation is obtained,

γv(z)+(γr)zv(z)+12θ2z2v′′(z)+u~(z)(dwL¯)z=0,-\gamma v(z)+(\gamma-r)zv^{\prime}(z)+\frac{1}{2}\theta^{2}z^{2}v^{\prime\prime}(z)+\tilde{u}(z)-(d-w\bar{L})z=0,

which is identical with Equation (D.1), hence shares the same solution as

v(z)={B11zn1+B21zn2+A1Γ1zδ(1k)δ(1k)1+w(L¯L)drz,z¯<z<y~,B12zn1+B22zn2+A2Γ2z1kk+wL¯drz,y~z<z^.v(z)=\begin{cases}B_{11}z^{n_{1}}+B_{21}z^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)-d}{r}z,&\bar{z}<z<\tilde{y},\\ B_{12}z^{n_{1}}+B_{22}z^{n_{2}}+\frac{A_{2}}{\Gamma_{2}}z^{-\frac{1-k}{k}}+\frac{w\bar{L}-d}{r}z,&\tilde{y}\leq z<\hat{z}.\end{cases} (E.1)

As follows, a six-equations system is established to obtain the unknown parameters B11B_{11}, B21B_{21}, B12B_{12}, B22B_{22}, z¯\bar{z} and z^\hat{z}.

  • 𝒞0\mathcal{C}^{0} condition at z=z¯z=\bar{z}

    B11z¯n1+B21z¯n2+A1Γ1z¯δ(1k)δ(1k)1+w(L¯L)rz¯=1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1.B_{11}\bar{z}^{n_{1}}+B_{21}\bar{z}^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)}{r}\bar{z}=\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}.
  • 𝒞1\mathcal{C}^{1} condition at z=z¯z=\bar{z}

    n1B11z¯n11+n2B21z¯n21+δ(1k)δ(1k)1A1Γ1z¯1δ(1k)1+w(L¯L)r=K1L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1.n_{1}B_{11}\bar{z}^{n_{1}-1}\!+\!n_{2}B_{21}\bar{z}^{n_{2}-1}\!+\!\frac{\delta(1-k)}{\delta(1-k)-1}\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{1}{\delta(1-k)-1}}\!+\!\frac{w(\bar{L}-L)}{r}\!=\!-K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{1}{\delta(1-k)-1}}.
  • 𝒞0\mathcal{C}^{0} condition at z=y~z=\tilde{y}

    B11y~n1+B21y~n2+A1Γ1y~δ(1k)δ(1k)1wLry~=B12y~n1+B22y~n2+A2Γ2y~1kk.B_{11}\tilde{y}^{n_{1}}+B_{21}\tilde{y}^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}\tilde{y}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{wL}{r}\tilde{y}=B_{12}\tilde{y}^{n_{1}}+B_{22}\tilde{y}^{n_{2}}+\frac{A_{2}}{\Gamma_{2}}\tilde{y}^{-\frac{1-k}{k}}.
  • 𝒞1\mathcal{C}^{1} condition at z=y~z=\tilde{y}

    n1B11y~n11+n2B21y~n21+δ(1k)δ(1k)1A1Γ1y~1δ(1k)1wLr=n1B12y~n11+n2B22y~n211kkA2Γ2y~1k.n_{1}B_{11}\tilde{y}^{n_{1}\!-\!1}\!+\!n_{2}B_{21}\tilde{y}^{n_{2}\!-\!1}\!+\!\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}\!\frac{A_{1}}{\Gamma_{1}}\tilde{y}^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\!-\!\frac{wL}{r}\!=\!n_{1}B_{12}\tilde{y}^{n_{1}\!-\!1}\!+\!n_{2}B_{22}\tilde{y}^{n_{2}\!-\!1}\!-\!\frac{1\!-\!k}{k}\!\frac{A_{2}}{\Gamma_{2}}\tilde{y}^{\!-\!\frac{1}{k}}.
  • 𝒞1\mathcal{C}^{1} condition at z=z^z=\hat{z}

    n1B12z^n11+n2B22z^n211kkA2Γ2z^1k+wL¯dr+Rpre=0.n_{1}B_{12}\hat{z}^{n_{1}-1}+n_{2}B_{22}\hat{z}^{n_{2}-1}-\frac{1-k}{k}\frac{A_{2}}{\Gamma_{2}}\hat{z}^{-\frac{1}{k}}+\frac{w\bar{L}-d}{r}+R_{pre}=0.
  • 𝒞2\mathcal{C}^{2} condition at z=z^z=\hat{z}

    n1(n11)B12z^n12+n2(n21)B22z^n22+1kk2A2Γ2z^1+kk=0.n_{1}(n_{1}-1)B_{12}\hat{z}^{n_{1}-2}+n_{2}(n_{2}-1)B_{22}\hat{z}^{n_{2}-2}+\frac{1-k}{k^{2}}\frac{A_{2}}{\Gamma_{2}}\hat{z}^{-\frac{1+k}{k}}=0.
Case 4. 0<z¯<y~z^0<\bar{z}<\tilde{y}\leq\hat{z}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost>drR_{post}>\frac{d}{r}

The only difference between this case and the previous one occurs in z=z¯z=\bar{z}. Since Rpost>drR_{post}>\frac{d}{r}, Lemma 3.1 shows that the Legendre-Fenchel transform of post-retirement value function U~(z¯)\tilde{U}(\bar{z}) is

U~(z¯)=B2,PRz¯n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1drz¯.\tilde{U}(\bar{z})=B_{2,\scriptscriptstyle PR}\bar{z}^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}-\frac{d}{r}\bar{z}.

As the same before, we set up a six-equation system to achieve the unknowns, B11B_{11}, B21B_{21}, B12B_{12}, B22B_{22}, z¯\bar{z} and z^\hat{z}. Compared with the first case, only 𝒞0\mathcal{C}^{0} and 𝒞1\mathcal{C}^{1} conditions at z=z¯z=\bar{z} change, whereas all the others keep true.

  • 𝒞0\mathcal{C}^{0} condition at z=z¯z=\bar{z}

    B11z¯n1+B21z¯n2+A1Γ1z¯δ(1k)δ(1k)1+w(L¯L)rz¯=B2,PRz¯n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1.B_{11}\bar{z}^{n_{1}}\!+\!B_{21}\bar{z}^{n_{2}}\!+\!\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}\!+\!\frac{w(\bar{L}\!-\!L)}{r}\bar{z}\!=\!B_{2,\scriptscriptstyle PR}\bar{z}^{n_{2}}\!+\!\frac{1\!-\!\delta(1\!-\!k)}{\delta(1\!-\!k)}K_{1}\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\bar{z}^{\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}}.
  • 𝒞1\mathcal{C}^{1} condition at z=z¯z=\bar{z}

    n1B11z¯n11+n2B21z¯n21+δ(1k)δ(1k)1A1Γ1z¯1δ(1k)1+w(L¯L)r=n2B2,PRz¯n21K1L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1.n_{1}\!B_{11}\!\bar{z}^{n_{1}\!-\!1}\!+\!n_{2}\!B_{21}\!\bar{z}^{n_{2}\!-\!1}\!+\!\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}\!\frac{A_{1}}{\Gamma_{1}}\!\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\!+\!\frac{w(\bar{L}\!-\!L)}{r}\!=\!n_{2}\!B_{2,\scriptscriptstyle PR}\!\bar{z}^{n_{2}\!-\!1}\!-\!K_{1}\!\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\!\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}}.
Case 5. 0<z¯<z^<y~0<\bar{z}<\hat{z}<\tilde{y}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost=drR_{post}=\frac{d}{r}

Firstly, the interval 0<z¯<z<z^<y~0<\bar{z}<z<\hat{z}<\tilde{y}, where the condition (V3)(V3) of (3.5) holds, is considered. Also adopting the time-independent form of ϕ(t,z)=eγtv(z){\phi}(t,z)=e^{-\gamma t}v(z), the following differential equation is obtained γv(z)+(γr)zv(z)+12θ2z2v′′(z)+u~(z)(dwL¯)z=0-\gamma v(z)+(\gamma-r)zv^{\prime}(z)+\frac{1}{2}\theta^{2}z^{2}v^{\prime\prime}(z)+\tilde{u}(z)-(d-w\bar{L})z=0. The dual transform of u(c,l)u(c,l) is u~(z)=A1zδ(1k)δ(1k)1wLz\tilde{u}(z)=A_{1}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}-wLz in the considered interval; therefore, the above differential equation takes the identical form of the one in 0<z¯<z<y~<z^0<\bar{z}<z<\tilde{y}<\hat{z} of Case 3. The solution of v(z)v(z) is given directly from (E.1), only changing the parameters’ notations from B11B_{11} to B1B_{1} and B21B_{21} to B2B_{2} respectively,

v(z)=B1zn1+B2zn2+A1Γ1zδ(1k)δ(1k)1+w(L¯L)drz,z¯<z<z^.v(z)=B_{1}z^{n_{1}}+B_{2}z^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)-d}{r}z,\quad\bar{z}<z<\hat{z}.

Next, a four-equations system is set up to derive the desired parameters B1B_{1}, B2B_{2}, z¯\bar{z}, z^\hat{z}. The same arguments with Case 3, only 𝒞1\mathcal{C}^{1} and 𝒞2\mathcal{C}^{2} conditions in z=z^z=\hat{z} changes.

  • 𝒞0\mathcal{C}^{0} condition at z=z¯z=\bar{z}

    B1z¯n1+B2z¯n2+A1Γ1z¯δ(1k)δ(1k)1+w(L¯L)rz¯=1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1.B_{1}\bar{z}^{n_{1}}+B_{2}\bar{z}^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)}{r}\bar{z}=\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}.
  • 𝒞1\mathcal{C}^{1} condition at z=z¯z=\bar{z}

    n1B1z¯n11+n2B2z¯n21+δ(1k)δ(1k)1A1Γ1z¯1δ(1k)1+w(L¯L)r=K1L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1.n_{1}B_{1}\bar{z}^{n_{1}-1}+n_{2}B_{2}\bar{z}^{n_{2}-1}+\frac{\delta(1-k)}{\delta(1-k)-1}\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{1}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)}{r}=-K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{1}{\delta(1-k)-1}}.
  • 𝒞1\mathcal{C}^{1} condition at z=z^z=\hat{z}

    n1B1z^n11+n2B2z^n21+δ(1k)δ(1k)1A1Γ1z^1δ(1k)1+w(L¯L)dr+Rpre=0.n_{1}B_{1}\hat{z}^{n_{1}-1}+n_{2}B_{2}\hat{z}^{n_{2}-1}+\frac{\delta(1-k)}{\delta(1-k)-1}\frac{A_{1}}{\Gamma_{1}}\hat{z}^{\frac{1}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)-d}{r}+R_{pre}=0.
  • 𝒞2\mathcal{C}^{2} condition at z=z^z=\hat{z}

    n1(n11)B1z^n12+n2(n21)B2z^n22+δ(1k)(δ(1k)1)2A1Γ1z^2δ(1k)δ(1k)1=0.n_{1}(n_{1}-1)B_{1}\hat{z}^{n_{1}-2}+n_{2}(n_{2}-1)B_{2}\hat{z}^{n_{2}-2}+\frac{\delta(1-k)}{(\delta(1-k)-1)^{2}}\frac{A_{1}}{\Gamma_{1}}\hat{z}^{\frac{2-\delta(1-k)}{\delta(1-k)-1}}=0.
Case 6. 0<z¯<z^<y~0<\bar{z}<\hat{z}<\tilde{y}, Rpre>dwL¯rR_{pre}>\frac{d-w\bar{L}}{r} &\& Rpost>drR_{post}>\frac{d}{r}

We now move to Case 6. The only difference from the previous case happens on the condition Rpost>drR_{post}>\frac{d}{r}, which is mainly involved in the post-retirement part; hence, the solution of the partial differential equation corresponding to Condition (V3)(V3) in (3.5) remains unchanged, that is,

v(z)=B1zn1+B2zn2+A1Γ1zδ(1k)δ(1k)1+w(L¯L)drz,z¯<z<z^.v(z)=B_{1}z^{n_{1}}+B_{2}z^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}z^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)-d}{r}z,\quad\bar{z}<z<\hat{z}.

Considering the smooth fit conditions at z¯\bar{z} and z^\hat{z}, we construct a four-equations system to deduce the values of unknown parameters B1B_{1}, B2B_{2}, z¯\bar{z} and z^\hat{z}.

  • 𝒞0\mathcal{C}^{0} condition at z=z¯z=\bar{z}

    B1z¯n1+B2z¯n2+A1Γ1z¯δ(1k)δ(1k)1+w(L¯L)rz¯=B2,PRz¯n2+1δ(1k)δ(1k)K1L¯(1k)(1δ)1δ(1k)z¯δ(1k)δ(1k)1.B_{1}\bar{z}^{n_{1}}+B_{2}\bar{z}^{n_{2}}+\frac{A_{1}}{\Gamma_{1}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)}{r}\bar{z}=B_{2,\scriptscriptstyle PR}\bar{z}^{n_{2}}+\frac{1-\delta(1-k)}{\delta(1-k)}K_{1}\bar{L}^{\frac{(1-k)(1-\delta)}{1-\delta(1-k)}}\bar{z}^{\frac{\delta(1-k)}{\delta(1-k)-1}}.
  • 𝒞1\mathcal{C}^{1} condition at z=z¯z=\bar{z}

    n1B1z¯n11+n2B2z¯n21+δ(1k)δ(1k)1A1Γ1z¯1δ(1k)1+w(L¯L)r=n2B2,PRz¯n21K1L¯(1k)(1δ)1δ(1k)z¯1δ(1k)1.n_{1}B_{1}\bar{z}^{n_{1}\!-\!1}\!+\!n_{2}B_{2}\bar{z}^{n_{2}\!-\!1}\!+\!\frac{\delta(1\!-\!k)}{\delta(1\!-\!k)\!-\!1}\!\frac{A_{1}}{\Gamma_{1}}\!\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\!+\!\frac{w(\bar{L}\!-\!L)}{r}\!=\!n_{2}B_{2,\scriptscriptstyle PR}\bar{z}^{n_{2}\!-\!1}\!-\!K_{1}\!\bar{L}^{\frac{(1\!-\!k)(1\!-\!\delta)}{1\!-\!\delta(1\!-\!k)}}\!\bar{z}^{\frac{1}{\delta(1\!-\!k)\!-\!1}}.
  • 𝒞1\mathcal{C}^{1} condition at z=z^z=\hat{z}

    n1B1z^n11+n2B2z^n21+δ(1k)δ(1k)1A1Γ1z^1δ(1k)1+w(L¯L)dr+Rpre=0.n_{1}B_{1}\hat{z}^{n_{1}-1}+n_{2}B_{2}\hat{z}^{n_{2}-1}+\frac{\delta(1-k)}{\delta(1-k)-1}\frac{A_{1}}{\Gamma_{1}}\hat{z}^{\frac{1}{\delta(1-k)-1}}+\frac{w(\bar{L}-L)-d}{r}+R_{pre}=0.
  • 𝒞2\mathcal{C}^{2} condition at z=z^z=\hat{z}

    n1(n11)B1z^n12+n2(n21)B2z^n22+δ(1k)(δ(1k)1)2A1Γ1z^2δ(1k)δ(1k)1=0.n_{1}(n_{1}-1)B_{1}\hat{z}^{n_{1}-2}+n_{2}(n_{2}-1)B_{2}\hat{z}^{n_{2}-2}+\frac{\delta(1-k)}{(\delta(1-k)-1)^{2}}\frac{A_{1}}{\Gamma_{1}}\hat{z}^{\frac{2-\delta(1-k)}{\delta(1-k)-1}}=0.

Same argument with Case 1 and Case 2 in Appendix D, given the initial wealth xRprex\geq R_{pre} and solving x=v(λ)x=-v^{\prime}(\lambda^{*}), we can obtain the optimal Lagrange multiplier λ\lambda^{*} and then the optimal process Z(t)=λeγtH(t)Z^{*}(t)=\lambda^{*}e^{\gamma t}H(t).

Proposition E.1.

Under the condition y~z^\tilde{y}\leq\hat{z}, corresponding to Case 3 and Case 4, the optimal consumption-portfolio-leisure plan {c(t),π(t),l(t)}\{c^{*}(t),\pi^{*}(t),l^{*}(t)\} before retirement is given by

c(t)={(1δδw)(1δ)(1k)k(Z(t))1k,y~Z(t)z^,L(1k)(1δ)δ(1k)1(Z(t))1δ(1k)1,z¯<Z(t)<y~,c^{*}(t)=\begin{cases}\left(\frac{1-\delta}{\delta w}\right)^{\frac{(1-\delta)(1-k)}{k}}(Z^{*}(t))^{-\frac{1}{k}},&\tilde{y}\leq Z^{*}(t)\leq\hat{z},\\ L^{-\frac{(1-k)(1-\delta)}{\delta(1-k)-1}}(Z^{*}(t))^{\frac{1}{\delta(1-k)-1}},&\bar{z}<Z^{*}(t)<\tilde{y},\end{cases}
l(t)={(1δδw)δ(1k)1k(Z(t))1k,y~Z(t)z^,L,z¯<Z(t)<y~,l^{*}(t)=\begin{cases}\left(\frac{1-\delta}{\delta w}\right)^{-\frac{\delta(1-k)-1}{k}}(Z^{*}(t))^{-\frac{1}{k}},&\tilde{y}\leq Z^{*}(t)\leq\hat{z},\\ L,&\bar{z}<Z^{*}(t)<\tilde{y},\end{cases}
π(t)={θσ[n1(n11)B12(Z(t))n11+n2(n21)B22(Z(t))n21+1kk2A2Γ2(Z(t))1k],y~Z(t)z^,θσ[n1(n11)B11(Z(t))n11+n2(n21)B21(Z(t))n21+δ(1k)(δ(1k)1)2A1Γ1(Z(t))1δ(1k)1],z¯<Z(t)<y~.\pi^{*}(t)=\begin{cases}\frac{\theta}{\sigma}\bigg{[}n_{1}(n_{1}-1)B_{12}(Z^{*}(t))^{n_{1}-1}+n_{2}(n_{2}-1)B_{22}(Z^{*}(t))^{n_{2}-1}&\\ \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad+\frac{1-k}{k^{2}}\frac{A_{2}}{\Gamma_{2}}(Z^{*}(t))^{-\frac{1}{k}}\bigg{]},&\tilde{y}\leq Z^{*}(t)\leq\hat{z},\\ \frac{\theta}{\sigma}\bigg{[}n_{1}(n_{1}-1)B_{11}(Z^{*}(t))^{n_{1}-1}+n_{2}(n_{2}-1)B_{21}(Z^{*}(t))^{n_{2}-1}&\\ \qquad\qquad\qquad\qquad\qquad\quad+\frac{\delta(1-k)}{(\delta(1-k)-1)^{2}}\frac{A_{1}}{\Gamma_{1}}(Z^{*}(t))^{\frac{1}{\delta(1-k)-1}}\bigg{]},&\bar{z}<Z^{*}(t)<\tilde{y}.\end{cases}

Meanwhile, under the condition z^<y~\hat{z}<\tilde{y}, corresponding to Case 5 and Case 6, the optimal consumption-portfolio-leisure plan {c(t),π(t),l(t)}\{c^{*}(t),\pi^{*}(t),l^{*}(t)\} before retirement is given by

c(t)=L(1k)(1δ)δ(1k)1(Z(t))1δ(1k)1,l(t)=L,c^{*}(t)=L^{-\frac{(1-k)(1-\delta)}{\delta(1-k)-1}}\left(Z^{*}(t)\right)^{\frac{1}{\delta(1-k)-1}},\qquad l^{*}(t)=L,
π(t)=θσ[n1(n11)B1(Z(t))n11+n2(n21)B2(Z(t))n21+δ(1k)(δ(1k)1)2A1Γ1(Z(t))1δ(1k)1].\pi^{*}(t)\!=\!\frac{\theta}{\sigma}\!\left[\!n_{1}(n_{1}\!-\!1)B_{1}(Z^{*}(t))^{n_{1}\!-\!1}\!+\!n_{2}(n_{2}\!-\!1)B_{2}(Z^{*}(t))^{n_{2}\!-\!1}\!+\!\frac{\delta(1\!-\!k)}{(\delta(1\!-\!k)\!-\!1)^{2}}\frac{A_{1}}{\Gamma_{1}}(Z^{*}(t))^{\frac{1}{\delta(1\!-\!k)\!-\!1}}\right].
Proof.

Follow the lines of Proposition D.1. ∎