
This paper presents a novel optimization model that explores the optimal investment strategies for DC pension plans with return of premium clauses. We have assumed that the financial market consists of a risk-free asset and a risky asset, where the price of the risky asset follows the CEV model. Under the expected utility criterion, the optimal investment strategies were derived by employing stochastic optimal control theory and the Legendre transformation method. Explicit expressions of the optimal investment strategy were provided when the utility function was specified as exponential, power, or logarithmic. Finally, numerical analysis was conducted to examine the impact of factors such as interest rate, return rate, and volatility of the risky asset on the optimal strategies.
Citation: Xiaoyi Tang, Wei Liu, Wanyin Wu, Yijun Hu. An optimal investment strategy for DC pension plans with costs and the return of premium clauses under the CEV model[J]. AIMS Mathematics, 2024, 9(8): 21731-21754. doi: 10.3934/math.20241057
[1] | Ling Chen, Xiang Hu, Mi Chen . Optimal investment and reinsurance for the insurer and reinsurer with the joint exponential utility under the CEV model. AIMS Mathematics, 2023, 8(7): 15383-15410. doi: 10.3934/math.2023786 |
[2] | Yinghui Dong, Chengjin Tang, Chunrong Hua . Optimal investment of DC pension plan under a joint VaR-ES constraint. AIMS Mathematics, 2024, 9(1): 2084-2104. doi: 10.3934/math.2024104 |
[3] | Zongqi Sun, Peng Yang, Ying Wang, Jing Lu . Research on the wealth management fees of defined contribution pensions during the pre-retirement stage. AIMS Mathematics, 2024, 9(12): 36102-36115. doi: 10.3934/math.20241713 |
[4] | Zongqi Sun, Peng Yang, Jing Wu, Yunpeng Fan . The defined contribution pension plan after retirement under the criterion of a revised loss considering the economic situation. AIMS Mathematics, 2024, 9(2): 4749-4761. doi: 10.3934/math.2024229 |
[5] | Liyuan Zhang, Limian Ci, Yonghong Wu, Benchawan Wiwatanapataphee . Blockchain asset portfolio optimization with proportional and fixed transaction fees. AIMS Mathematics, 2025, 10(3): 6694-6718. doi: 10.3934/math.2025306 |
[6] | Peng Yang . Robust optimal reinsurance-investment problem for n competitive and cooperative insurers under ambiguity aversion. AIMS Mathematics, 2023, 8(10): 25131-25163. doi: 10.3934/math.20231283 |
[7] | Jiuchao Ban, Yiran Wang, Bingjie Liu, Hongjun Li . Optimization of venture portfolio based on LSTM and dynamic programming. AIMS Mathematics, 2023, 8(3): 5462-5483. doi: 10.3934/math.2023275 |
[8] | Wuyuan Jiang, Zechao Miao, Jun Liu . Optimal investment and reinsurance for the insurer and reinsurer with the joint exponential utility. AIMS Mathematics, 2024, 9(12): 35181-35217. doi: 10.3934/math.20241672 |
[9] | Fuyun Sun, Yuelei Li . On the improved thinning risk model under a periodic dividend barrier strategy. AIMS Mathematics, 2021, 6(12): 13448-13463. doi: 10.3934/math.2021779 |
[10] | Ying Fang, Guo Cheng, Zhongfeng Qu . Optimal reinsurance for both an insurer and a reinsurer under general premium principles. AIMS Mathematics, 2020, 5(4): 3231-3255. doi: 10.3934/math.2020208 |
This paper presents a novel optimization model that explores the optimal investment strategies for DC pension plans with return of premium clauses. We have assumed that the financial market consists of a risk-free asset and a risky asset, where the price of the risky asset follows the CEV model. Under the expected utility criterion, the optimal investment strategies were derived by employing stochastic optimal control theory and the Legendre transformation method. Explicit expressions of the optimal investment strategy were provided when the utility function was specified as exponential, power, or logarithmic. Finally, numerical analysis was conducted to examine the impact of factors such as interest rate, return rate, and volatility of the risky asset on the optimal strategies.
Pension funds are one of the most important institutions in the financial markets because they can help to secure one's life after retirement. Therefore, studying the optimal management strategy for a pension fund is a research topic of theoretical and practical importance. Typically, there are mainly two kinds of pension plans, which are defined benefit (DB) and defined contribution (DC) pension plans. In recent years, DC pension plans have become more popular in the market because DC pension plans can meet the demographic development and the financial market. Obviously, DC pension plans are worth studying.
Over the last two decades, DC pension plans have been extensively studied in the literature. From the perspective of financial markets, for the case where the price of the stock is modeled by the geometric Brownian motion (GBM), see Boulier et al. [1], Deelstra et al. [2], Han and Hung [3], Guan and Liang [4], Guan and Liang [5], Chen et al. [6], Fombellida and Casado [7], Dong and Zheng [8], and the references therein. For the case where the stock price is assumed to be the constant elasticity variance (CEV) model, see Xiao et al. [9], Gao [10], Gao [11], Zhang et al. [12], Li et al. [13], Li et al. [14], and Chen et al. [15]. Since the GBM can not suitably capture the volatility smile of the stock, as the CEV model can do well, thus the CEV model is more suitable to model the prices of stock. Meanwhile, the GMB can be considered as a special case of the CEV model.
Recently, from a practical point of view, many researchers studied the optimal investment strategies for DC pension plans with the return of premium clauses. In a DC pension plan without return of premium clauses, if a pension fund member dies during the accumulation phase, then the pension plan is automatically terminated. Meanwhile, the premiums paid by the pension member will not be returned to the member. In contrast to DC pension plans without return of premium clauses, in a DC pension plan with return of premium clauses, the pension members withdraw their premiums when they die during the accumulation phase and the difference between the premium and the accumulation (negative or positive) is distributed among the surviving members. Obviously, DC pension plans with return of premium clauses can protect the rights and interests of pensioners well. Therefore, it is of great importance to study the optimal management strategies for DC pension plans with the return of premium clauses. In this direction, He and Liang [16] first introduced the return of premium clauses into a DC pension plan model. Under the mean-variance criterion, they studied the optimal investment strategies where the stock price was modeled by a GBM. Under the mean-variance criterion, Sheng and Rong [17] studied the optimal investment strategies for a DC pension plan with an annuity contract, where the price of the stock was modeled by Heston's stochastic volatility model. Under the mean-variance criterion, Sun and Li [18] investigated the optimal investment strategy for a DC pension plan, where the price of the stock was modeled by a jump-diffusion process. Under the multi-period mean-variance criterion, Bian et al. [19] studied the investment strategy for a DC pension plan in a discrete time setting. By maximising the mean-variance of terminal wealth, Li and Rong [14] studied the optimal investment strategy for DC pension plans, where the price of the stock was modeled by the CEV model and the invested bond was defaultable.
All of the above references to optimal investment strategies for DC pension plans with return of premium clauses used mean-variance criteria. However, it is well-known that besides the popular mean-variance criterion, another popular optimality criterion is maximizing the expected utility of the terminal wealth. Therefore, this paper will focus on the expected utility criterion. On the other hand, in reality, financial markets usually involve costs such as trading fees and taxes, while these economic factors have not been discussed in the literature for a DC pension plan with return of premium clauses. Consequently, this paper is mainly motivated by the following two concerns:
(1) How can we best incorporate the trading fees and taxes into the model of DC pension plans with the return of premiums clauses?
(2) Furthermore, under the above first concern, what is the optimal investment strategy by that maximizes the expected utility of the terminal wealth?
To the best of our knowledge, the above two concerns have not been discussed in the literature. In the present paper, we aim to address the above two concerns. We assume that the financial market consists of a risk-free asset and a risky asset (called a stock), where the price of the stock is modeled by the CEV model. We also assume that there are trading fees and taxes in the financial market, which may affect the asset process. Instead of mean-variance criterion, by maximizing the expected utility of the terminal wealth, we obtain the optimal investment strategy for a DC pension plan with the return of premium clauses via stochastic optimal control theory and the Legendre transformation method. When the utility function is specified in terms of exponential, power, and logarithmic utility functions, the explicit expressions of the optimal investment strategies are given. Finally, the effect of the involved parameters on the optimal investment strategies is analyzed, and these parameters are the rate of interest and the return rate and volatility of the stock.
The remainder of this paper is organized as follows. In Section 2, we introduce the financial market and the optimization model for a DC pension plan with return of premium clauses. In Section 3, we provide a general solution to the optimization model by solving the Hamilton-Jacobi-Bellman (HJB) equation. In Section 4, we give the explicit expressions of the optimal investment strategies when the utility function is specified in terms of the exponential, power, and logarithmic utility functions. In Section 5, we provide a numerical analysis of the effect of parameters on the optimal investment strategies. These parameters include the interest rate and the return and volatility of the stock. Finally, Section 6 summarizes the conclusions.
In this section, we introduce the financial market and the optimization program.
Let (Ω,F,{Ft}t∈[0,T],P) be a complete filtered probability space satisfying the usual conditions and Ft represents the information up to time t in the market. The accumulation phase of the DC pension plan occurs within [0,T]. The DC pension fund starts at time 0 and T>0 is the retirement time. We assume that all of the processes introduced below are well-defined and also adapted to {Ft}t∈[0,T].
Suppose that the market structure consists of two financial assets, a risk-free asset and a risky asset. We denote the price of the risk-free asset (i.e., cash) at time t by B(t), and the price process by
dB(t)B(t)=rdt, | (2.1) |
where r is a constant rate of interest.
The price process of the risky asset (i.e., stock) at time t by S(t), is described by the CEV model
dS(t)S(t)=μdt+kSβ(t)dW(t), | (2.2) |
where μ is an instantaneous expected return on the stock and satisfies the general condition μ>r, kSβ(t) is the instantaneous volatility, and β is the elasticity parameter and satisfies the general condition β≥0. {Wt:t≥0} is a standard Brownian motion defined on a complete probability space (Ω,F,{Ft}t∈[0,T],P).
Remark 2.1. Note that if the elasticity parameter β=0 in(2), then it reduces to a GBM. If β<0, the instantaneous volatility kSβt increases as the stock price decreases, and can generate a distribution with a fatter left tail. If β>0, the situation is reversed.
For a DC pension, the holder of the pension fund in the accumulation stage needs to pay a certain amount of premium in advance. In the model, the premium paid per unit of time is P (P is constant) and the starting point of the accumulation phase is ω0. The final moment is ω0+T. According to the rules of endowment insurance, if a member dies before retirement, the balance of the individual contribution in the personal account and the interest earned are returned to the heir in a lump sum. In order to protect the rights and interests of pensioners who die during the accumulation phase, this paper introduces the premium return clause into the pension plan. Under actuarial terms, a deceased annuitant can withdraw premiums. qΔtqω0+t is an actuarial notation that denotes the probability of dying at age ω0+t and in time [ω0+t,ω0+t+Δt]. tP denotes the accumulation of premiums at time t, so that the return of the death annuity at age ω0+t during the period [0,Δt] is tPqΔtqω0+t and the difference between the income after the premiums are returned and the the difference between the accumulation phase will be distributed equally to the survivors.
To increase the wealth value of the pension fund, the manager of the pension fund invests the insurance premium in a risky asset (stock) and a risk-free asset (cash), where the proportion allocated of π and (1−π) respectively. ρ is the trading fees of the stock trading, and θ is the tax when the holder of the pension fund withdraws.
Consider the discrete form of wealth in [t,t+1n]:
Xπ(t+1n)=[Xπ(t)(πS(t+1n)S(t)+(1−π)B(t+1n)B(t))−Xπ(t)(πρ+θ)1n+P1n−aPtq1nqω0+t]×(11−q1nqω0+t). | (2.3) |
To compare the optimal strategies of pension funds with and without the return of premium clauses, we introduce parameters, the parameters a=1, which means there is a return of premium clause, and a=0, which means there is no return of premium clause. The last item of the above formula means that the difference between the income and the accumulate phase amount will be distributed equally to the survivors.
Denote
Δδ1nt:=πS(t+1n)−S(t)S(t)+(1−π)B(t+1n)−B(t)B(t), | (2.4) |
then
Xπ(t+1n)=[X(t)π(1+Δδ1nt)−Xπ(t)(πρ+θ)1n+P1n−aPtq1nqω0+t]×(1+q1nqω0+t1−q1nqω0+t). | (2.5) |
In (2.5) u(t) is the death force function, and qtqx is an actuarial symbol standing for the probability that the person who is alive at the age of x will be dead in the following t time period, and satisfies qtqx=1−ptpx=1−e−∫t0u(x+s)ds. So
q1nqω0+t=1−e−∫1n0u(ω0+t+s)ds≈u(ω0+t)1n=O(1n). | (2.6) |
Let t→0, then u(ω0+t) is small in the accumulation stage and
q1nqω0+t1−q1nqω0+t=1−e−∫1n0u(ω0+t+s)dse−∫0u(ω0+t+s)ds=e∫1n0u(ω0+t+s)ds−1≈u(ω0+t)1n=O(1n). | (2.7) |
Hence, it is easy to get
Δδt⋅O(1n)=o(1n),q1nqω0+t⋅O(1n)=o(1n). |
Then (2.5) can be written as follows
Xπ(t+1n)=Xπ(t)(1+Δδ1nt)+Xπ(t)u(ω0+t)1n−Xπ(t)(πρ+θ)1n+P1n−aPtu(ω0+t)1n+o(1n). | (2.8) |
When t→0, and we have
dXπ(t)=Xπ(t)[πdS(t)S(t)+(1−π)dB(t)B(t)]+X(t)u(ω0+t)dt−Xπ(t)(πρ+θ)dt+Pdt−aPtu(ω0+t)dt. | (2.9) |
Substitute (2.1) and (2.2) into (2.9), we have
dXπ(t)=Xπ(t)[π(μ−ρ−r)+(r−θ)+u(ω0+t)]dt+Pdt−aPtu(ω0+t)dt+Xπ(t)πkSβ(t)dW(t). | (2.10) |
To simplify the model, we have chosen the Abraham De Moivre model (refer to Kohler and Kohler [20]) to characterize the force of mortality function, u(t), in the following form:
u(t)=1ω−t,0≤t<ω, | (2.11) |
where ω is the limit age of the person, (2.10) of the wealth process of the pension fund can be written as follows
{dXπ(t)=Xπ(t)[π(μ−ρ−r)+(r−θ)+1ω−ω0−t]dt+Pω−ω0−(1+a)tω−ω0−tdt+Xπ(t)πkSβ(t)dW(t),X(0)=x0. | (2.12) |
In this paper, a dynamic asset allocation strategy π(t) is an asset allocation proportion invested in the stock at time t. A strategy Π={π(t)∈[0,∞],(0≤t≤T)} is called admissible. We denote the set of all admissible strategies as the following.
Definition 2.1. For any fixed t∈[0,T], a strategy π is said to be admissible if
(1) π is F-predictable;
(2) For any v∈[t,T],E[∫Tt(x(v)π(v)ks2β)2]<+∞;
(3) For any v∈R×R, Eq (2.12) has a unique solution Xπ(v)v∈[t,T] with Xπ(t)=x,S(t)=s.
Let U(x):R+→R be a utility function. Assume that U(x) has the finite first- and second-order derivatives, and assume that U′(x)>0 and U″(x)<0 for any x. Denote by RAC(x) the risk aversion coefficient of the utility function U(x), that is RAC(x):=−U″(x)U′(x). The RAC(x) reflects the risk preference of an investor. The larger the RAC(x) is, the more risk-averse the investor. In other words, the smaller the RAC(x), the more risk-appetite the investor is.
To study a DC pension plan is to choose an optimal investment strategy π(t) that maximizes the expected utility of the terminal wealth E[U(X(T))]. In other words, the optimization problem studied in this paper is described as following:
maxπt∈ΠE[U(X(T))]. | (2.13) |
This section begins by deriving the general framework of the optimization problem (2.13) using the maximum principle. It then proceeds to find the explicit solution of the optimal investment strategy that maximizes the terminal expected utility function using the Legendre transformation.
By using the classical tools of stochastic optimal control, we define the value function
H(t,s,x)=supπt∈ΠE{U(X(T))|S(t)=s,X(t)=x}. | (3.1) |
Assume that V(t,s,x)∈C1,2,2([0,T]×R×R), then we can define the following variational operator on any function V(t,s,x):
Aπ(t)V(t,s,x)=Ht+{x[π(μ−ρ−r)+(r−θ)+1ω−ω0−t]+Pω−ω0−(1+a)tω−ω0−t}Hx |
+μsHs+12k2s2β+2Hss+12x2π2k2s2βHxx+xπk2s2β+1Hxs, |
where Vt,Vx,Vxx,Vs,Vss, and Vsx represent first-order and second-order partial derivatives with respect to the variables t,s, and x respectively.
According to the principle of stochastic dynamic programming, we can easily derive the following Hamilton-Jacobi-Bellman (HJB) equation
Ht+μsHs+12k2s2β+2Hss+{x[(r−θ)+1ω−ω0−t]+Pω−ω0−(1+a)tω−ω0−t}Hxsupπt∈Π{xπ(μ−ρ−r)Hx+12x2π2k2s2βHxx+xπk2s2β+1Hxs}=0. | (3.2) |
The optimal investment strategy π∗ satisfies
π∗=−(μ−ρ−r)Hx+k2s2β+1Hxsxk2s2βHxx. | (3.3) |
Substituting (3.3) into (3.2), we obtain another partial differential equation for the value function H
Ht+μsHs+{x[(r−θ)+1ω−ω0−t]+Pω−ω0−(1+a)tω−ω0−t}Hx+12k2s2β+2Hss−12[(μ−ρ−r)Hx+k2s2β+1Hxs]2k2s2βHxx=0. | (3.4) |
(3.4) represents a non-linear second-order partial differential equation, which can be challenging to solve. However, the use of the Legendre transformation can transform it into a linear partial differential equation, making it easier to solve.
The following theorem verifies that the investment strategy given by (3.3) is indeed optimal for optimization problem (2.13) and its proof can be found in the Appendix.
Theorem 3.1. If V(t,s,x)∈C1,2,2([0,T]×R×R) is a solution of the HJB equation (3.2), i.e., V(t,s,x) satisfies the following HJB equation
supπt∈Π{Aπ(t)V(t,s,x)}=0,V(T,s,x)=U(x(T)), |
then for all admissible strategies π(t)∈Π, one has H(t,s,x)≤V(t,s,x); if ˜π(t) satisfies
˜π(t)∈argsupπt∈ΠE{U(X(T))|S(t)=s,X(t)=x}, |
then one has H(t,s,x)=V(t,s,x), and it implies that ˜π(t) is the optimal solution of optimization problem (2.13). Consequently, the investment strategy given by (3.3) is the optimal solution to the optimization problem (2.13).
Definition 3.1. Let f:Rn⟶R be a convex function. For z>0, define the Legendre transform
L(z)=maxxf(x)−zx, |
and the function L(z) is called the Legendre dual of the function f(x).
If f(x) is strictly convex, the maximum in the above equation will be attained at just one point, which we denote by x0. It is attained at the unique solution to the first-order condition
df(x)dx−z=0. |
So we can write
L(z)=f(x0)−zx0. |
We also use the assumed convexity of the value function to define the Legendre transform
ˆH(t,s,z)=supx>0{H(t,s,x)−zx}, | (3.5) |
where z>0 denotes the variable dual to x. The value of x where this optimum is denoted by g(t,s,z), so that
g(t,s,z)=infx>0{x|H(t,s,x)≥zx+ˆH(t,s,z)}. | (3.6) |
The function ˆH is related to g by
g(t,s,z)=−ˆHz(t,s,z). | (3.7) |
Hence we can take either one of the two functions g and ˆH as the dual of H.
As the Legendre transform describes, we have that
Hx(t,s,g(t,s,z))=z, | (3.8) |
and this leads to
ˆH(t,s,z)=H(t,s,g)−zg. | (3.9) |
By differentiating (3.5) and (3.6) with respect to t,s, and z, the transformation rules for the derivatives of the value function H and the dual function ˆH can be given as (e.g., Jonsson and Sircar[21])
Hx=z,Ht=ˆHt,Hss=ˆHss−ˆH2szˆHzz, |
Hs=ˆHs,Hxs=−ˆHszˆHzz,Hxx=−1ˆHzz. | (3.10) |
At the terminal time, we define
ˆU(z)=supx>0{U(x)−zx},G(z)=infx>0{X|U(x)≥zx+ˆU(z)}. | (3.11) |
In this way, the problem can be turned into a dual problem.
Substituting expression (3.10), we rewrite (3.4) as
ˆHt+μsˆHs+x(r−θ+1ω−ω0−t)z+P(ω−ω0−(1+a)tω−ω0−t)z |
+12k2s2β+2(ˆHss−ˆH2szˆHzz)−12[(μ−ρ−r)z+k2s2β+1(−ˆHszˆHzz)]2k2s2β(−1ˆHzz)=0, | (3.12) |
where φ1:=r−θ+1ω−ω0−t, φ2:=P(ω−ω0−(1+a)tω−ω0−t), φ3:=μ−ρ−r.
Now we have the following partial differential equation
ˆHt+μsˆHs+xφ1z+φ2z+12k2s2β+2ˆHss+12φ23z2k2s2βˆHzz−φ3szˆHsz=0. | (3.13) |
The derivative with respect to z is
ˆHtz+(μ−φ3)sˆHsz+xφ1+φ1zxz+φ2+12k2s2β+2ˆHssz+12φ23z2k2s2βˆHzzz+φ23zk2s2βˆHzz−φ3szˆHszz=0. | (3.14) |
For any t,s, let x=g(t,x,z) in (3.14), and by using (3.7), we have
gt+(μ−φ3)sgs+(φ23k2s2β−φ1)zgz−gφ1−φ2+12k2s2β+2gss+12φ23z2k2s2βgzz−φ3szgsz=0. | (3.15) |
The focus is typically on the optimal investment strategy rather than the value function. Using Eq (3.4), we can calculate the optimal stock holding through a feedback formula based on the derivatives of the value function. The optimal stock holding can be expressed in terms of the dual function g.
π∗=−φ3Hx+k2s2β+1Hxsxk2s2βHxx=−φ3zgz+k2s2β+1gsgk2s2β. | (3.16) |
Thus, for a given optimal problem, we solve the linear PDE for g and obtain the investment amount π∗.
This section presents the explicit expressions for the solutions to the optimization problem (2.13) when the utility function U(x) is specified as the exponential (constant absolute risk aversion (CARA)), power (constant relative risk aversion (CRRA)), and logarithmic utility functions.
The exponential utility function is given by
U(x)=−1qe−qx, |
where q>0 is the risk aversion coefficient.
The Eq (3.15) is then solved.
g(T,s,z)=(U′)−1(z)=−1qlnz. |
The solution to Eq (3.15) is sought as follows.
g(t,s,z)=−1q[b(t)(lnz+m(t,s))]+α(t) | (4.1) |
with the boundary conditions given by m(T,s)=0,α(T)=0, and b(T)=1.
Then
gt=−1q{b′(t)[lnz+m(t,s)]+b(t)mt}+α′(t),gs=−1qb(t)ms,gz=−b(t)qz,gss=−1qb(t)mss,gsz=0,gzz=b(t)qz2. | (4.2) |
Plugging (4.1) in (3.15), we have
−1qlnz{b′(t)−b(t)φ1}−1qb(t){mb′(t)b(t)+mt+(u−φ3)sms−mφ1−φ1+12k2s2β+2mss+12φ32k2s2β}+α′(t)−α(t)φ1−φ2=0. | (4.3) |
The equation can be split into three separate equations.
α′(t)−α(t)φ1−φ2=0, | (4.4) |
b′(t)−b(t)φ1=0, | (4.5) |
mt+(ρ+r)sms+12k2s2β+2mss+12φ23k2s2β−φ1=0. | (4.6) |
Taking into account the boundary condition α(T)=0, we have
α(t)=P(1+a)[1−e(r−θ)(t−T)](ω−ω0−t)(r−θ)2−P(ω−ω0−(1+a)t)(ω−ω0−t)(r−θ)+P(ω−ω0−(1+a)T)(ω−ω0−t)(r−θ)e(r−θ)(t−T). | (4.7) |
Taking into account the boundary condition b(T)=1, we have
b(t)=ω−ω0−tω−ω0−Te(r−θ)(t−T). | (4.8) |
Note that since (4.5) includes coefficient variables s,s2β+2, and s2β, it is difficult to solve this equation. Therefore, we can use a power transformation and a variable change technique to transform the non-linear equation into a linear one.
m(t,s)=h(t,y),y=s−2β, |
so that
mt=ht,ms=−2βs−2β−1hy,mss=4β2s−4β−2hyy+2β(2β+1)s−2β−2hy. | (4.9) |
Substituting these derivatives into (4.5), we obtain the following linear PDE
ht−2β(ρ+r)yhy+2β2k2yhyy+k2β(2β+1)hy+12φ23yk2−φ1=0. | (4.10) |
We try to find a solution of (4.9) with the following structure
h(t,y)=˜A(t)+˜B(t)y, | (4.11) |
with ˜A(T)=0,˜B(T)=0.
Then
ht=˜At+˜Bty,hy=˜B,hyy=0, | (4.12) |
and introducing this in (4.9), we derive
˜At+k2β(2β+1)˜B−φ1+y{˜Bt−2β(ρ+r)˜B+φ232k2}=0. | (4.13) |
We can decompose (4.12) into two conditions in order to eliminate the dependence in y and t:
˜At+k2β(2β+1)˜B−φ1=0, | (4.14) |
˜Bt−2β(ρ+r)˜B+φ232k2=0. | (4.15) |
Taking into account the boundary condition ˜B(T)=0, the solution to (4.15) is
˜B(t)=φ234k2β(ρ+r)(1−e2β(ρ+r)(t−T)). | (4.16) |
Taking into account the boundary condition ˜A(T)=0 and ˜B(t), the solution to (4.13) is
˜A(t)=(−(2β+1)φ234(ρ+r)+r−θ)(t−T)+(2β+1)φ238β(ρ+r)e2β(ρ+r)(t−T)+lnω−ω0−Tω−ω0−t. | (4.17) |
Finally, we can obtain the optimal solution of problem (2.13) under the CARA utility. The optimal strategy invested in the stock is given by
π∗(t)=φ3b(t)+2βk2b(t)˜B(t)qxk2s2β, | (4.18) |
which is summarized in the following theorem,
Theorem 4.1. For the optimization problem (2.13) with wealth process (2.12) and the exponential utility function, the optimal investment strategy is given by
π∗(t)=φ3b(t)+2βk2b(t)˜B(t)qxk2s2β, |
where
φ3=μ−ρ−r |
b(t)=ω−ω0−tω−ω0−Te(r−θ)(t−T), |
˜B(t)=φ234k2β(ρ+r)(1−e2β(ρ+r)(t−T)). |
Remark 4.1. For the CARA utility function, the return of premium clauses have no effect on the optimal investment strategy. If elasticity parameter β=0, the CEV model reduces to the GBM model, and the optimal investment strategy is given by π∗(t)=φ3b(t)qxk2, which also implies that the optimal investment strategy has nothing to do with the price of the stock.
The power utility function is given by
U(x)=1γxγ, |
where γ<1,γ≠0.
Next, we solve Eq (3.15).
g(T,s,z)=(U′)−1(z)=z1γ−1. |
We try to find a solution to (3.15) in the following way
g(t,s,z)=z1γ−1h(t,s)+α(t), | (4.19) |
with the boundary conditions given by h(T,s)=1,α(T)=0.
Then
gt=htz1γ−1+α′(t),gs=hsz1γ−1,gz=hγ−1z1γ−1−1,gss=hssz1γ−1,gsz=hsγ−1z1γ−1−1,gzz=(2−γ)h(γ−1)2z1γ−1−2. | (4.20) |
Plugging (4.19) in (3.15), we have
z1γ−1[ht+(u−γγ−1φ3)shs−γγ−1hφ1+12k2s2β+2hss+12φ23k2s2β(2−γ)h(γ−1)2+φ23k2s2βhγ−1]+α′(t)−α(t)φ1−φ2=0. | (4.21) |
We can split (4.20) into equations in order to eliminate the dependence on z1γ−1:
α′(t)−α(t)φ1−φ2=0, | (4.22) |
ht+(u−γγ−1φ3)shs−hγγ−1φ1+12k2s2β+2hss+φ23k2s2βγh2(γ−1)2. | (4.23) |
Taking into account the boundary condition α(T)=0, we have
α(t)=P(1+a)[1−e(r−θ)(t−T)](ω−ω0−t)(r−θ)2−P(ω−ω0−(1+a)t)(ω−ω0−t)(r−θ)+P(ω−ω0−(1+a)T)(ω−ω0−t)(r−θ)e(r−θ)(t−T). | (4.24) |
To solve (4.23), similar to the exponential case, we define let h(t,s)=f(t,y) and y=s−2β so that
ht=ft,hs=−2βs−2β−1fy,hss=4β2s−4β−2fyy+2β(2β+1)s−2β−2fy. | (4.25) |
Substituting (4.25) into (4.23), we obtain the following:
ft−γfγ−1(φ1−φ232k2(γ−1))+2k2β2yfyy+βfy{−2uy+k2(2β+1)+γγ−12φ3y}=0. | (4.26) |
We try to find a solution of (4.26) with the following:
f(t,y)=A(t)eB(t)y, | (4.27) |
with A(T)=1,B(T)=0.
Then
ft=A′(t)eB(t)y+A(t)B′(t)yeB(t)y,fy=A(t)B(t)eB(t)y,fyy=A(t)B2(t)eB(t)y. | (4.28) |
Introducing this in (4.26), we derive
AtA+Bty−γγ−1(φ1−φ232k2(γ−1))+2k2β2B2y+βB{−2μy+k2(2β+1)+γγ−12φ3y}=0. | (4.29) |
We can decompose (4.29) into two conditions in order to eliminate the dependence in y and t:
AtA−γγ−1φ1+Bk2β(2β+1)=0,A(T)=1, | (4.30) |
Bt+φ23γ2k2(γ−1)2+2k2β2B2+2βBμ−γ(ρ+r)γ−1=0,B(T)=0. | (4.31) |
The solution to (4.30) and (4.31) is
A(t)=(ω−ω0−Tω−ω0−t)γγ−1(λ2−λ1)β(2β+1)2β2[λ2−λ1e2β2(λ1−λ2)(T−t)]−β(2β+1)2β2e[γγ−1(r−θ)−β(2β+1)λ1](t−T), | (4.32) |
B(t)=k−2I(t), | (4.33) |
where I(t)=λ1−λ1e2β2(λ1−λ2)(T−t)1−λ1λ2e2β2(λ1−λ2)(T−t), λ1,2=μ−γ(ρ+r)±√(1−γ)(μ2−(ρ+r)2γ)2β(1−γ).
Finally, we can obtain the optimal solution of problem (2.13) under the CRRA utility. The optimal strategy invested in the stock is given by
π∗(t)=−φ3(x−α(t))(γ−1)xk2s2β−2β(x−α(t))I(t)xk2s2β, | (4.34) |
which is summarized in the following theorem.
Theorem 4.2. For the optimization problem (2.13) with wealth process (2.12) and the power utility function, the optimal investment strategy is given by
π∗(t)=−φ3(x−α(t))(γ−1)xk2s2β−2β(x−α(t))I(t)xk2s2β, |
where
φ3=μ−ρ−r |
α(t)=P(1+a)[1−e(r−θ)(t−T)](ω−ω0−t)(r−θ)2−P(ω−ω0−(1+a)t)(ω−ω0−t)(r−θ)+P(ω−ω0−(1+a)T)(ω−ω0−t)(r−θ)e(r−θ)(t−T), |
I(t)=λ1−λ1e2β2(λ1−λ2)(T−t)1−λ1λ2e2β2(λ1−λ2)(T−t),λ1,2=u−γ(ρ+r)±√(1−γ)(u2−(ρ+r)2γ)2β(1−γ). |
Remark 4.2. If elasticity parameter β=0, the CEV model reduces to the GBM model, and the optimal investment strategy is given by π∗(t)=φ3(x−α(t))(1−γ)xk2, which also implies that the optimal investment strategy has nothing to do with the price of the stock.
The logarithm utility function is given by
U(x)=lnx. |
Next, we solve the Eq (3.15).
g(T,s,z)=(U′)−1(z)=1z. |
We try to find a solution of (3.15) in the following way:
g(t,s,z)=1zf(s)+α(t), | (4.35) |
where we denote the s value of the terminal time T as sT, with f(sT)=1,α(T)=0, and we can obtain
gt=α′(t),gs=1zf′(s),gz=−1z2f(s);gss=1zf″(s),gzz=2z3f(s),gsz=−1z2f′(s). | (4.36) |
Plugging (4.35) in (3.15), we have
α′(t)+μs1zf′(s)−φ1α(t)−φ2+12k2s2β+21zf″(s)=0. | (4.37) |
We can decompose this equation into two conditions in order to eliminate the dependence on s:
{α′(t)−φ1α(t)−φ2=0,μs1zf′(s)+12k2s2β+21zf″(s)=0. | (4.38) |
The solution to (4.37), which takes into account the initial conditions
φ(T)=0,f(sT)=1, |
is then
α(t)=P(1+a)[1−e(r−θ)(t−T)](ω−ω0−t)(r−θ)2−P(ω−ω0−(1+a)t)(ω−ω0−t)(r−θ)+P(ω−ω0−(1+a)T)(ω−ω0−t)(r−θ)e(r−θ)(t−T), | (4.39) |
f(s)=1. | (4.40) |
Finally, this leads to
g=1z+P(1+a)[1−e(r−θ)(t−T)](ω−ω0−t)(r−θ)2−P(ω−ω0−(1+a)t)(ω−ω0−t)(r−θ)+P(ω−ω0−(1+a)T)(ω−ω0−t)(r−θ)e(r−θ)(t−T). | (4.41) |
Introducing (4.41) in (3.16) and using (4.36), we obtain the optimal solution in the case of the logarithm utility
π∗(t)=φ3[x−α(t)]xk2s2β, | (4.42) |
which is summarized in the following theorem.
Corollary 1. For the optimization problem (2.13) with wealth process (2.12) and the logarithm utility function, the optimal investment strategy is given by
π∗(t)=φ3(x−α(t))xk2s2β, |
where
φ3=μ−ρ−r, |
α(t)=P(1+a)[1−e(r−θ)(t−T)](ω−ω0−t)(r−θ)2−P(ω−ω0−(1+a)t)(ω−ω0−t)(r−θ)+P(ω−ω0−(1+a)T)(ω−ω0−t)(r−θ)e(r−θ)(t−T). |
Remark 4.3. If elasticity parameter β=0, the CEV model reduces to the GBM model, and the optimal investment strategy is given by π∗(t)=φ3[x−α(t)]xk2. Therefore, the optimal investment strategy has nothing to do with the price of the stock.
In this section, we will numerically analyze the effect of parameters on the optimal investment strategies. Throughout this section, unless otherwise stated, the following parameters are given as P=1, s=5, x=1, ω=100, ω0=20, T=40, r=0.02, u=0.1, k=0.5, β=1, θ=0.005, ρ=0.01, q=5, and γ=0.5.
Figures 1 and 2 show that, compared with the case without return of premium clauses, the proportion of money invested into the stock becomes less under the case with return of premium clauses. This phenomenon can be economically interpreted as follows. The possible return of premium during the accumulation phase could reduce the fund size level. Hence, under the expected utility investment criterion and to hedge the loss of the premium income, the fund manager would prefer to invest less money in the stock. In other words, under the expected utility investment criterion, the risk preference of the fund manager is risk-averse.
Figure 3 shows that the proportion of money invested into the stock is the same for both of the cases of with and without return of premiums. This phenomenon can be economically interpreted as follows. In contrast to the power and logarithmic utility functions, the risk aversion coefficient RAC(x) of the exponential utility function is a constant, which implies that the risk preference of an investor does not depend on its wealth. Therefore, in the case of the exponential utility function, the proportions of the money invested into the stock stay the same regardless of being with or without return of premiums.
Figure 4 shows that the proportion of money invested into the stock under the logarithmic utility function is less than that under the power utility function. This phenomenon coincides with the following intuition. Since the RAC(x) is 1x for the logarithmic utility function and 1−γx for the power utility function, hence the intuition is that the fund manager is more risk-averse under the logarithmic utility function than that under the power utility function. Therefore, the fund manager would prefer to invest less money in the stock in the case of the logarithmic utility function. Note also that the risk aversion coefficients for the logarithmic and power utility functions become almost the same when the wealth is large enough. Meanwhile, the fund size should become larger and larger when the time moves into the future. Consequently, when the time moves into the future, the proportions of money invested into the stock under both the logarithmic and power utility functions become the same.
In this subsection, we examine the sensitivity of the parameters involved in the model. These parameters include taxes, trading fees, the rate of interest, and the return rate and volatility of the stock.
Figure 5 shows that under any one of the exponential, power, and logarithmic utility functions, the larger the tax, the more money invested into the stock. This phenomenon can be economically interpreted as follows. Tax is a kind of income tax. Therefore, in order to hedge this loss of income tax, the fund manager would like to invest more money into the stock, because the stock is of higher return although it is of higher risk.
Figure 6 shows that under any one of the exponential, power, and logarithmic utility functions, the larger the trading fee, the less the money invested into the stock. This phenomenon can be interpreted from the behavioral finance point of view as follows. The trading fee is a kind of market management cost. Therefore, in order to reduce this loss of market management cost, the fund manager would like to invest less money into the stock.
Figure 7 shows that in the case of the exponential utility function, the larger the risk aversion coefficient q, the less proportion of money invested into the stock. This phenomenon also coincides with the fact that the RAC(x) is now just the risk aversion coefficient q. On the other hand, since the RAC(x) does not depend on the wealth, the impacts of the different values of q on the proportion of money invested into the stock appear more sensitive when the wealth is small, but less sensitive when the wealth is large enough.
Figure 8 shows that under the power utility function, the proportion of money invested into the stock increases when γ increases. This phenomenon can be interpreted as follows. Since the RAC(x) is 1−γx, the fund manager is more risk-appetite when γ becomes large. Hence the fund manager would prefer to invest more money into the stock when γ increases. Note also that the RAC(x) becomes almost the same when the wealth x is larger enough. Meanwhile, the fund size should become larger when the time moves into the future. Consequently, when the time moves into the future, the proportion of money invested into the stock under the power utility function becomes the same.
Figure 9 shows that under any one of the exponential, power, and logarithmic utility functions, the larger the interest rate, the smaller the proportion of money invested into the stock. This phenomenon coincides with people's common behavior. Indeed, when the interest rate increases, the fund manager would have no incentive to invest more money into the stock in order to maintain the profit target. In other words, the fund manager can simply invest more money into the risk-free asset to reach his/her profit target.
Figure 10 shows that under any one of the exponential, power, and logarithmic utility functions, the larger the return rate of the stock, the greater the proportion of money invested into the stock. This phenomenon coincides with people's common behavior. Indeed, when the return rate of the stock increases, the fund manager would have the incentive to invest more money into the stock in order to achieve higher profit.
Figure 11 shows that under any one of the exponential, power, and logarithmic utility functions, the larger the volatility of the stock, the smaller the proportion of money invested into the stock. This phenomenon can be interpreted as follows. Larger volatility of the stock represents higher risk. Hence, fund managers who are not risk-appetite would like to invest less money into the stock, and at the same time invest more money into the risk-free asset.
In this paper, we have proposed a new optimal DC pension model with the return of premium by incorporating two practical factors, tax and trading fee, into the model. In the model, the pension is allowed to invest in a risk-free asset and a risky asset, where the price of a risky asset is assumed to follow the CEV model. The optimal investment strategy is obtained by maximizing the expected utility of the terminal wealth. Explicit expressions of the optimal investment strategies are also given when the utility function is specified to be the exponential, the power, and the logarithmic utility functions, respectively. Finally, numerical studies are provided to show the impacts of the parameters involved on the optimal investment strategies with the interpretations in the context of economics. In the future, the return of premium clauses can be considered in more complex financial models to obtain the optimal investment strategy for DC pensions.
Let Q=R×R+, and we choose a sequence of bounded open sets Qi satisfying Qi⊂Qi+1⊂Q, i=1,2,⋯, and Q=⋃∞i=1Qi. For (s,x)∈Qi, assume that the exit time of (S(t),X(t)) from Qi is denoted by τi. When i→∞, we get τi∧T→T.
(ⅰ) Consider an arbitrary admissible strategy π(t). By applying Itˆo's formula for V(t,s,x) on [t,T], we obtain
V(T,S(T),X(T))=V(t,s,x)+∫TtAπ(t)V(ν,s(ν),x(ν))dν+∫TtxπksβVx(ν,s(ν),x(ν))dW(ν)+∫TtksβVs(ν,s(ν),x(ν))dW(ν). |
Taking supπ(t)∈Π{Aπ(t)V(t,s,x)}=0 into consideration, which displays that the variational inequality Aπ(t)V(t,s,x)≤0, we have
V(T,S(T),X(T))≤V(t,s,x)+∫TtxπksβVx(ν,s(ν),x(ν)))dW(ν)+∫TtksβVs(ν,s(ν),x(ν)))dW(ν). |
The last three terms on the right-hand side of the above inequality are square-integrable martingales and their expectations are equal to zero. Hence, we get
E(V(T,S(T),X(T))∣S(t)=s,X(t)=x)≤V(t,s,x). |
Further, taking the supremum, we obtain
supπt∈ΠE(V(T,S(T),X(T))∣S(t)=s,X(t)=x)≤V(t,s,x), |
and it implies that
H(t,r,x)≤V(t,r,x). |
(ⅱ) E(V(τi∧T,S(τi∧T),X(τi∧T)))<∞ for a specific strategy π∗(t). Applying Itˆo's formula to V(t,s,x) on [0,τi∧T] once again, we have
V(τi∧T,S(τi∧T),X(τi∧T))=V(0,s0,x0)+∫τi∧T0Aπ∗(t)V(ν,s(ν),x(ν))dν+∫τi∧T0xπksβVx(ν,s(ν),x(ν))dW(ν)+∫τi∧T0ksβVs(ν,s(ν),x(ν))dW(ν). |
For a specific strategy π∗(t) satisfies (15), i.e., Aπ∗(t)V(t,S(ν),X(ν))=0, and the last three terms are also square-integrable martingales. Hence, taking the expectation on both sides on the above equation, we obtain
E(V(τi∧T,S(τi∧T),X(τi∧T)))=V(0,s0,x0)<∞. |
(ⅲ) H(t,s,x)=V(t,s,x) for a specific strategy π∗(t). Using Itˆo's formula for V(t,s,x) on [t,τi∧T] once more, similarly, we derive
V(τi∧T,S(τi∧T),X(τi∧T))=V(t,s,x)+∫τi∧T0Aπ∗(t)V(ν,s(ν),x(ν))dν+∫τi∧T0xπksβVx(ν,s(ν),x(ν))dW(ν)+∫τi∧T0ksβVs(ν,s(ν),x(ν))dW(ν). |
Taking the conditional expectation, we yield
V(t,s,x)=E(V(τi∧T,S(τi∧T),X(τi∧T))∣S(t)=s,X(t)=x). |
Taking the limitation once more, we get
V(t,s,x)=limi→∞E(V(τi∧T,S(τi∧T),X(τi∧T))∣S(t)=s,X(t)=x). |
In addition, we derive
H(t,s,x)=supπt∈ΠE(V(τi∧T,S(τi∧T),X(τi∧T))∣S(t)=s,X(t)=x)=limi→∞E(V(τi∧T,S(τi∧T),X(τi∧T))∣S(t)=s,X(t)=x)=V(t,s,x). |
Therefore, it implies that π∗(t) is indeed the optimal investment strategy for the problem (2.13).
Xiaoyi Tang: Conceptualization; Wei Liu: Supervision; Wanyin Wu: Validation; Yijun Hu: Supervision. All authors have read and approved the final version of the manuscript for publication.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work is supported by the National Natural Science Foundation of China (Nos: 11961064, 12271415).
All authors declare no conflicts of interest in this paper.
[1] |
J. F. Boulier, S. Huang, G. Tailand, Optimal management under stochastic interest rates: The case of a protected defined contribution pension funds, Insur. Math. Econ., 28 (2001), 173–189. https://doi.org/10.1016/S0167-6687(00)00073-1 doi: 10.1016/S0167-6687(00)00073-1
![]() |
[2] |
G. Deelstra, M. Grasselli, P. F. Koehl, Optimal investment strategies in the presence of a minimum guarantee, Insur. Math. Econ., 33 (2003), 189–207. https://doi.org/10.1016/S0167-6687(03)00153-7 doi: 10.1016/S0167-6687(03)00153-7
![]() |
[3] |
N. Han, M. Hung, Optimal asset allocation for the DC pension plans under inflation, Insur. Math. Econ., 51 (2012), 172–181. https://doi.org/10.1016/j.insmatheco.2012.03.003 doi: 10.1016/j.insmatheco.2012.03.003
![]() |
[4] |
G. H. Guan, Z. X. Liang, Optimal management of DC pension plan in a stochastic interest rate and stochastic volatility framework, Insur. Math. Econ., 57 (2014), 58–66. https://doi.org/10.1016/j.insmatheco.2014.05.004 doi: 10.1016/j.insmatheco.2014.05.004
![]() |
[5] |
G. H. Guan, Z. X. Liang, Optimal management of DC pension plan under loss aversion and value-at-risk constraints, Insur. Math. Econ., 69 (2016), 224–237. https://doi.org/10.1016/j.insmatheco.2016.05.014 doi: 10.1016/j.insmatheco.2016.05.014
![]() |
[6] |
Z. Chen, Z. F. Li, Y. Zeng, J. Y. Sun, Asset allocation under loss aversion and minimum performance constraint in a dc pension plan with inflation risk, Insur. Math. Econ., 75 (2017), 137–150. https://doi.org/10.1016/j.insmatheco.2017.05.009 doi: 10.1016/j.insmatheco.2017.05.009
![]() |
[7] |
R. Josa-Fombellida, P. López-Casado, J. P. Rincón-Zapatero, Portfolio optimization in a defined benefit pension plan where the risky assets are processes with constant elasticity of variance, Insur. Math. Econ., 82 (2018), 73–86. https://doi.org/10.1016/j.insmatheco.2018.06.011 doi: 10.1016/j.insmatheco.2018.06.011
![]() |
[8] |
Y. Dong, H. Zheng, Optimal investment of DC pension plan under short-selling constraints and portfolio insurance, Insur. Math. Econ., 85 (2019), 47–59. https://doi.org/10.1016/j.insmatheco.2018.12.005 doi: 10.1016/j.insmatheco.2018.12.005
![]() |
[9] |
J. Xiao, Z. Hong, C. Qin, The constant elasticity of variance (CEV) model and the Legendre transform-dual solution for annuity contracts, Insur. Math. Econ., 40 (2007), 302–310. https://doi.org/10.1016/j.insmatheco.2006.04.007 doi: 10.1016/j.insmatheco.2006.04.007
![]() |
[10] | J. Gao, Optimal investment for the defined-contribution pension with stochastic salary under a CEV model, Insur. Math. Econ., 44 (2009), 479–490. |
[11] |
J. Gao, Optimal investment strategy for annuity contracts under the constant elasticity of variance (CEV) model, Insur. Math. Econ., 45 (2009), 9–18. https://doi.org/10.1016/j.insmatheco.2009.02.006 doi: 10.1016/j.insmatheco.2009.02.006
![]() |
[12] | C. B. Zhang, X. M. Rong, H. Zhao, R. J. Hou, Optimal investment for the defined-contribution pension with stochastic salary under a CEV model, Appl. Math. J. Chinese Univ., 28 (2013), 187–203. |
[13] |
D. P. Li, X. M. Rong, H. Zhao, Time-consistent investment strategy for DC pension plan with stochastic salary under CEV model, J. Syst. Sci. Complex., 29 (2016), 428–454. https://doi.org/10.1007/s11424-016-3171-3 doi: 10.1007/s11424-016-3171-3
![]() |
[14] |
D. P. Li, X. M. Rong, H. Zhao, B. Yi, Equilibrium investment strategy for DC pension plan with default risk and return of premiums clauses under CEV model, Insur. Math. Econ., 72 (2017), 6–20. https://doi.org/10.1016/j.insmatheco.2016.10.007 doi: 10.1016/j.insmatheco.2016.10.007
![]() |
[15] | H. Chen, Z. Yin, T. Xie, Determining equivalent administrative charges for defined contribution pension plans under CEV model, Math. Probl. Eng., 2018, 1–10. https://doi.org/10.1155/2018/6278353 |
[16] |
L. He, Z. Liang, Optimal investment strategy for the DC plan with the return of premiums clauses in a mean-variance framework, Insur. Math. Econ., 53 (2013), 643–649. https://doi.org/10.1016/j.insmatheco.2013.09.002 doi: 10.1016/j.insmatheco.2013.09.002
![]() |
[17] | D. L. Sheng, X. M. Rong, Optimal time-consistent investment strategies for a DC pension plan with the return of premiums clauses and annuity contracts, Discrete Dyn. Nat. Soc., 2014, 1–13. https://doi.org/10.1155/2014/862694 |
[18] |
J. Y. Sun, Z. F. Li, Y. Zeng, Precommitment and equilibrium investment strategies for defined contribution pension plans under a jump-diffusion model, Insur. Math. Econ., 67 (2016), 158–172. https://doi.org/10.1016/j.insmatheco.2016.01.005 doi: 10.1016/j.insmatheco.2016.01.005
![]() |
[19] |
L. H. Bian, Z. F. Li, Pre-commitment and equilibrium investment strategies for the DC pension plan with regime switching and a return of premiums clause, Insur. Math. Econ., 81 (2018), 78–94. https://doi.org/10.1016/j.insmatheco.2018.05.005 doi: 10.1016/j.insmatheco.2018.05.005
![]() |
[20] | P. H. Kohler, I. Kohler, Frailty modeling for adult and old age mortality: The application of a modified De Moivre Hazard function to sex differentials in mortality, Demogr. Res., 3 (2000). Available from: http://www.demographic-research.org/Volumes/Vol3/8/. |
[21] | M. Jonsson, R. Sircar, Optimal investment problems and volatility homogenization approximations, In: Modern Methods in Scientific Computing and Applications NATO Science Series II, 75 (2002), 255–281. |