In this paper, we consider a market model where the risky asset is a jump diffusion whose drift, volatility and jump coefficients are influenced by market regimes and history of the asset itself. Since the trajectory of the risky asset is discontinuous, we modify the delay variable so that it remains defined in this discontinuous setting. Instead of the actual path history of the risky asset, we consider the continuous approximation of its trajectory. With this modification, the delay variable, which is a sliding average of past values of the risky asset, no longer breaks down. We then use the resulting stochastic process in formulating the state variable of a portfolio optimization problem. In this formulation, we obtain the dynamic programming principle and Hamilton Jacobi Bellman equation. We also provide a verification theorem to guarantee the optimal solution of the corresponding stochastic optimization problem. We solve the resulting finite time horizon control problem and show that close form solutions of the stochastic optimization problem exist for the cases of power and logarithmic utility functions. In particular, we show that the HJB equation for the power utility function is a first order linear partial differential equation while that of the logarithmic utility function is a linear ordinary differential equation.
Citation: Dennis Llemit, Jose Maria Escaner IV. Value functions in a regime switching jump diffusion with delay market model[J]. AIMS Mathematics, 2021, 6(10): 11595-11609. doi: 10.3934/math.2021673
[1] | Yuanfu Shao . Dynamics and optimal harvesting of a stochastic predator-prey system with regime switching, S-type distributed time delays and Lévy jumps. AIMS Mathematics, 2022, 7(3): 4068-4093. doi: 10.3934/math.2022225 |
[2] | Lin Xu, Linlin Wang, Hao Wang, Liming Zhang . Optimal investment game for two regulated players with regime switching. AIMS Mathematics, 2024, 9(12): 34674-34704. doi: 10.3934/math.20241651 |
[3] | Guocheng Li, Pan Zhao, Minghua Shi, Gensheng Li . A hybrid framework for mean-CVaR portfolio selection under jump-diffusion processes: Combining cross-entropy method with beluga whale optimization. AIMS Mathematics, 2024, 9(8): 19911-19942. doi: 10.3934/math.2024972 |
[4] | Yongxiang Zhu, Min Zhu . Well-posedness and order preservation for neutral type stochastic differential equations of infinite delay with jumps. AIMS Mathematics, 2024, 9(5): 11537-11559. doi: 10.3934/math.2024566 |
[5] | Wanlu Zhang, Hui Meng . Robust optimal investment-reinsurance strategies with the preferred reinsurance level of reinsurer. AIMS Mathematics, 2022, 7(6): 10024-10051. doi: 10.3934/math.2022559 |
[6] | Min Han, Bin Pei . An averaging principle for stochastic evolution equations with jumps and random time delays. AIMS Mathematics, 2021, 6(1): 39-51. doi: 10.3934/math.2021003 |
[7] | Gunduz Caginalp . Fat tails arise endogenously from supply/demand, with or without jump processes. AIMS Mathematics, 2021, 6(5): 4811-4846. doi: 10.3934/math.2021283 |
[8] | Xiuxian Chen, Zhongyang Sun, Dan Zhu . Mean-variance investment and risk control strategies for a dynamic contagion process with diffusion. AIMS Mathematics, 2024, 9(11): 33062-33086. doi: 10.3934/math.20241580 |
[9] | Xipan Zhang, Changchun Shen, Dingju Xu . Reachable set estimation for neutral semi-Markovian jump systems with time-varying delay. AIMS Mathematics, 2024, 9(4): 8043-8062. doi: 10.3934/math.2024391 |
[10] | Qiang Yu, Na Xue . Asynchronously switching control of discrete-time switched systems with a $ \Phi $-dependent integrated dwell time approach. AIMS Mathematics, 2023, 8(12): 29332-29351. doi: 10.3934/math.20231501 |
In this paper, we consider a market model where the risky asset is a jump diffusion whose drift, volatility and jump coefficients are influenced by market regimes and history of the asset itself. Since the trajectory of the risky asset is discontinuous, we modify the delay variable so that it remains defined in this discontinuous setting. Instead of the actual path history of the risky asset, we consider the continuous approximation of its trajectory. With this modification, the delay variable, which is a sliding average of past values of the risky asset, no longer breaks down. We then use the resulting stochastic process in formulating the state variable of a portfolio optimization problem. In this formulation, we obtain the dynamic programming principle and Hamilton Jacobi Bellman equation. We also provide a verification theorem to guarantee the optimal solution of the corresponding stochastic optimization problem. We solve the resulting finite time horizon control problem and show that close form solutions of the stochastic optimization problem exist for the cases of power and logarithmic utility functions. In particular, we show that the HJB equation for the power utility function is a first order linear partial differential equation while that of the logarithmic utility function is a linear ordinary differential equation.
The classical Merton portfolio optimization problem, in its basic form, involves an investor who is limited to investing to two types of assets-risky and non-risky. An investor's goal is to determine the optimal allocation strategy such that the wealth-dependent performance criterion is a maximum. Under certain conditions, Merton found out that the optimal strategy is to keep a constant fraction of the wealth in the risky asset [1].
Merton's portfolio optimization problem belongs to a wider set of stochastic optimal control problems which are known to be generally hard to solve and explicit solutions are rare. Despite this, several authors have improved on Merton's work by relaxing particular assumptions of the original paper. For instance, Davis and Norman [8] incorporated proportional transaction costs and successfully obtained optimal solutions from their formulation. In addition, they found out that the solution space can be divided into three regions-no transaction, sell, and buy. Framstad, Oksendal and Sulem [2] studied the case when the risky asset is a jump-diffusion in a portfolio problem with transaction costs. Pagliarani and Vargiolu [12] examined the case when risky assets are defaultable Levy processes.
Meanwhile, the series of works involving Pang ([6,7,14]) considered portfolio problems that incorporate the path history of the risky asset. These type of problems are called stochastic systems with delay or memory. The rationale for considering delay stems from the tendency of market participants to look at the past performance of assets and decide based on these information. While the concept of stochastic systems with delay or memory seems to run counter to the Markovian nature of dynamic programming, Larssen [4] provided the settings where dynamic programming still applies.
Bauerle and Rieder [11] considered a market model and portfolio problem where the drift and volatility of the price process are driven by a continuous time Markov chain. The work of Valdez and Vargiolu [13] provided a framework for dealing with a portfolio problem involving multidimensional risky assets which are diffusions with switching coefficients. Azevedo et al. [9] considered Markov switching jump-diffusions, thereby expanding the framework into discontinuous settings. The main motivation for incorporating regime switching in portfolio optimization problems is to take into account the effect of market regimes or states of the economy in the dynamics of asset prices.
Several tweaks and modifications have been made to the original Merton portfolio probem in order to capture market realities. However, to the best of our knowledge, there has been no study, in the context of the classical Merton portfolio optimization problem, that considered systems with delay and regime switching in discontinuous models. This is due to the fact that the memory or delay variable, as a time integral, becomes undefined in the face of jumps or discontinuities. Hence, this paper aims to supplement this gap and provides a way to overcome the problem of the delay variable breaking down in discontinuous settings.
Let T>0 be finite. Let (Ω,F,P) be a complete filtered probability space where the filtration F={Ft:t∈[0,T]} satisfies all the usual conditions.
To model market regimes, we set {αt:t∈[0,T]} to be a continuous time Markov chain defined on a finite state space M={a1,a2,⋯,an} with generator Q=(qij)i,j∈M. Let
Kijt=∑0<s≤t1{αs−=i}1{αs=j} |
be a counting process that counts the number of jumps of the Markov chain αt from state i to state j up to time t. Corresponding to this counting process is the intensity process
λijt=qij1{αt−=i} |
such that the purely discontinuous, square-integrable process
Mijt=Kijt−∫t0λijsds |
is a martingale.
As in Larssen and Risebro [5] or in the case of Pang and Hussain [7], we define the delay variable, which is a form of sliding average of past values, to be
Yt:=∫0−∞eρuχt+udu | (2.1) |
where ρ∈R and χt is the continuous approximation of the path of the risky asset until time t. With this definition (2.1) is defined even when the risky asset has discontinuities. The motivation for considering (2.1) in our market model is to account for the tendency of market participants to look at past performances of stocks before making investment decisions which in turn affect stock prices [7].
The investment opportunities in our model are the non-risky asset Pt that follows
dPt=r(t,Yt,αt−)Ptdt | (2.2) |
and the risky asset Xt which is assumed to be a Levy process given by
dXt=Xt−[μ(t,Yt,αt−)dt+σ(t,Yt,αt−)dBt+∫+∞−1Γ(t,Yt,αt−,z)N(dt,dz)], | (2.3) |
where r:[0,T]×R×M→R, μ:[0,T]×R×M→R, σ:[0,T]×R×M→R and Γ:[0,T]×R×M×R→R are uniformly continuous functions representing the risk-free rate, drift, volatility, and jump coefficient, respectively, where z=ΔXt=Xt−Xt− is the jump size of the risky asset. N(t,A) is the compensated Poisson random measure given by
N(t,A)=¯N(t,A)−tν(A) |
where ¯N(t,A) is the Poisson random measure that counts the number of jumps of the risky asset up to time t and ν(A) is the Levy measure for each A∈B0, where B0 is a Borel σ-field generated by the open subsets O of R0=R∖{0} whose closure does not contain the zero element.
We assume that the Brownian motion Bt, the Markov chain αt and the compensated Poisson random measure N(t,A) are all independent and adapted to the filtration F.
We also assume that
E[∫T0(|σ(s,y,a)|2+∫+∞−1|Γ(s,y,a,z)|2ν(dz))ds]<∞ |
for every (y,a)∈R×M.
Let πt be F-progressively measurable and that for a fixed T>0,
∫T0|πt|2dt<∞a.s. |
This πt represents the proportion of wealth Wt invested by an agent in the risky asset while the balance 1−πt is allocated to the non-risky asset. A self-financing portfolio resulting from these investments evolves according to
dWt=Wt−[(1−πt)r(t,Yt,αt−)dt+πtdRt] | (2.4) |
where
dRt=μ(t,Yt,αt−)dt+σ(t,Yt,αt−)dBt+∫+∞−1Γ(t,Yt,αt−,z)N(dt,dz). | (2.5) |
The corresponding state variable is the wealth process
{dWt=Wt−[r(t,Yt,αt−)(1−πt)dt+πtdRt]Ws=w>0,Ys=y,αs=a. | (2.6) |
We define the performance criterion to be
Gπ(s,w,y,a):=Es,w,y,a[U(Wπ;s,w,y,aT)] | (2.7) |
and the value function
V(s,w,y,a):=Gπ∗(s,w,y,a)=supπ∈A[s,T]Es,w,y,a[U(Wπ;s,w,y,aT)], | (2.8) |
where U(⋅) is a utility function, Es,w,y,a[⋅] is the conditional expectation conditioned on the initial data (s,w,y,a), and A[s,T] is the set of admissible controls such that (2.6) has a unique strong solution Wπ;s,w,y,at for every t∈[s,T].
Lemma 2.1. Let Yt be defined as in (2.1). Then
dYt=(χt−ρYt)dt. |
Proof: The proof follows [7]. We have
ddtYt=ddt[∫0−∞eρuχt+udu]=ddt[∫t−∞eρ(θ−t)χθdθ],where θ=t+u=ddt[limτ→−∞∫tτeρ(θ−t)χθdθ]=χt−limτ→−∞[ρ∫tτeρ(θ−t)χθdθ]=χt−ρ[∫0−∞eρuχt+udu]=χt−ρYt. |
Thus, the delay satisfies
dYt=(χt−ρYt)dt. | (2.9) |
The next result is the equivalent of Ito's Lemma for the particular portfolio optimization problem.
Lemma 2.2. Let V(s,w,y,a) such that V(⋅,⋅,⋅,a)∈C1,2,1([0,T]×R+×R) for every a∈M. Then, we have
V(T,WπT,YT,αT)=V(0,Wπ0,Y0,α0)+∫T0A(s,w,y,a,πs)ds+∫T0L(s,w,y,a)dBs+∫T0D(s,w,y,a)dMajs+∫T0∫+∞−1Z(s,w,y,a,πs)N(ds,dz) | (2.10) |
where
A(s,w,y,a,πs)=∂V∂s+w[πsμ(s,y,a)+r(s,y,a)(1−πs)]∂V∂w+12π2sw2σ2(s,y,a)∂2V∂w2+∫+∞−1(V(s,w+πswΓ(s,y,a,z),y,a)−V(s,w,y,a)−πswΓ(s,y,a,z)∂V∂w)ν(dz)+(χs−ρy)∂V∂y+∑j≠aqa,j(V(s,w,y,j)−V(s,w,y,a)), | (2.11) |
L(s,w,y,a)=σ(s,y,a)∂V∂w, | (2.12) |
D(s,w,y,a)=∑j≠a(V(s,w,y,j)−V(s,w,y,a)), | (2.13) |
and
Z(s,w,y,a,πs)=V(s,w+πswΓ(s,y,a,z),y,a)−V(s,w,y,α). | (2.14) |
Proof: Applying the change of variable rules (see [18,9]) and using Lemma (2.1) for the delay, we get
dV(s,w,y,a)=∂V∂sds+w[πsμ(s,y,a)+r(s,y,a)(1−πs)]∂V∂wds+πswσ(s,y,a)∂V∂wdBs+∫+∞−1[V(s,w+πswΓ(s,y,a,z),y,a)−V(s,w,y,a)]N(ds,dz)+∫+∞−1(V(s,w+πswΓ(s,y,a,z),y,a)−V(s,w,y,a)−πswΓ(s,y,a,z)∂V∂w)ν(dz)ds+∂V∂y(χs−ρy)ds+∑j≠a(V(s,w,y,j)−V(s,w,y,a))dMajs+∑j≠aqaj(V(s,w,y,j)−V(s,w,y,a))ds. |
Combining all terms with ds, we get
A(s,w,y,a,πs)=∂V∂s+w[πsμ(s,y,a)+r(s,y,a)(1−πs)]∂V∂w+12π2sw2σ2(s,y,a)∂2V∂w2+∫+∞−1(V(s,w+πswΓ(s,y,a,z),y,a)−V(s,w,y,a)−πswΓ(s,y,a,z)∂V∂w)ν(dz)+(χs−ρy)∂V∂y+∑j≠aqa,j(V(s,w,y,j)−V(s,w,y,a)) |
and setting the rest as
L(s,w,y,a,πs)=πswσ(s,y,a)∂V∂w, |
D(s,w,y,a)=∑j≠a(V(s,w,y,j)−V(s,w,y,a)), |
and
Z(s,w,y,a,πs)=V(s,w+πswΓ(s,y,a,z),y,a)−V(s,w,y,a). |
Thus,
dV(s,w,y,a)=A(s,w,y,a,πs)ds+L(s,w,y,a,πs)dBs+D(s,w,y,a)dMajs+∫+∞−1Z(s,w,y,a,πs)N(ds,dz). |
Integrating over [0,T], we obtain
V(T,WπT,YT,αT)=V(0,Wπ0,Y0,α0)+∫T0A(s,w,y,a,πs)ds+∫T0L(s,w,y,a,πs)dBs+∫T0D(s,w,y,a)dMajs+∫T0∫+∞−1Z(s,w,y,a,πs)N(ds,dz) |
We employ dynamic programming to solve the particular portfolio optimization problem.
Theorem 3.1 (Dynamic programming principle). Assuming that the value function as given by (2.8) is continuous over the space of controls A with the state variable (2.6), then for any (s,w,y,a)∈[0,T]×R+×R×M, we have that
V(s,w,y,a)=supπ∈A[s,T]Es,w,y,a[V(s+h,Wπs+h,Ys+h,αs+h)] | (3.1) |
for all s+h∈[s,T].
Proof: For any (s,w,y,a)∈[0,T)×R+×R×M and any arbitrary admissible control π, we have
Gπ(s,w,y,a)=Es,w,y,a[U(Wπ;s,w,y,aT)]=Es,w,y,a[Es+h,Wπs+h,Ys+h,αs+h[U(Wπ;s,w,y,aT)]]=Es,w,y,a[Gπ(s+h,Wπs+h,Ys+h,αs+h)]≤Es,w,y,a[V(s+h,Wπs+h,Ys+h,αs+h)] |
Taking supremum over admissible controls, we obtain
V(s,w,y,a)≤supπ∈A[s,T]Es,w,y,a[V(s+h,Wπs+h,Ys+h,αs+h)]. |
For the other direction, we follow Cartea et al. [23] by considering an ϵ-optimal control. Let ϵ>0 and take an admissible control πϵ∈A[s,T] such that
V(s,w,y,a)≥Gπϵ(s,w,y,a)≥V(s,w,y,a)−ϵ |
This is guaranteed because the value function is continuous over A[s,T]. Next, we consider the modification of the ϵ-optimal control
˜πϵ=π1t≤s+h+πϵ1t>s+h,t∈[s,T]. |
Then we have
G˜πϵ(s,w,y,a)=Es,w,y,a[G˜πϵ(s+h,W˜πϵs+h,Ys+h,αs+h)]=Es,w,y,a[Gπϵ(s+h,Wπs+h,Ys+h,αs+h)]≥Es,w,y,a[V(s+h,Wπs+h,Ys+h,α+h)]−ϵ. |
Taking limits as ϵ→0,
V(s,w,y,a)≥Es,w,y,a[V(s+h,Wπs+h,Ys+h,αs+h)]. |
Taking supremum over admissible controls,
V(s,w,y,a)≥supπ∈A[s,T]Es,w,y,a[V(s+h,Wπs+h,Ys+h,αs+h)]. |
Since both inequalities are true for any arbitrary admissible control π, we conclude that
V(s,w,y,a)=supπ∈A[s,T]Es,w,y,a[V(s+h,Wπs+h,Ys+h,αs+h)] |
for all s+h∈[s,T].
Theorem 3.2 (Hamilton-Jacobi-Bellman equation). Assume that V(⋅,⋅,⋅,α)∈C1,2,1([0,T]×R+×R) for every α∈M. Then for each α∈M, the value function V(⋅,⋅,⋅,α) defined on [0,T]×R+×R is the solution to the Hamilton-Jacobi-Bellman (HJB) equation
{∂V∂t+supπ∈A[s,T]H(t,W,Y,α,π,∂V∂W,∂2V∂W2,∂V∂Y)=0V(T,WπT,YT,αT)=U(Wπ;s,w,y,aT), | (3.2) |
where the Hamiltonian function is defined to be
H(t,W,Y,α,π,∂V∂W,∂2V∂W2,∂V∂Y)=Wt[πtμ(t,Yt,α)+r(t,Yt,α)(1−πt)]∂V∂W+12π2tW2tσ2(t,Yt,α)∂2V∂W2+∫+∞−1(V(t,Wt+πtWtΓ(t,Yt,α,z),Yt,α)−V(t,Wt,Yt,α)−πtWtΓ(t,Wt,Yt,α,z)∂V∂W)ν(dz)+(χt−ρYt)∂V∂Y+∑j≠αqα,j(V(t,Wt,Yt,j)−V(t,Wt,Yt,α)). | (3.3) |
Proof: From the dynamic progamming principle (Theorem 3.1), we have that for s+h∈[s,T],
V(s,w,y,a)≥Es,w,y,a[V(s+h,Wπs+h,Ys+h,αs+h)]. |
Using Lemma 2.2, we have for t∈[s,s+h]
V(s,w,y,a)≥Es,w,y,a[V(s,w,y,a)+∫s+hsA(t,Wπt,Yt,αt,πt)dt+∫s+hsL(t,Wπt,Yt,αt)dBt+∫s+hsD(t,Wπt,Yt,αt)dMαjt+∫s+hs∫+∞−1Z(t,Wπt,Yt,αt,πt)N(dt,dz)] |
From the assumption on Mαjt and from Davis [24] and Fleming and Soner [25] for Bt and N(t,z), it follows that
0=Es,w,y,a[∫s+hsL(t,Wπt,Yt,αt)dBt]=Es,w,y,a[∫s+hsD(t,Wπt,Yt,αt,)dMαjt]=Es,w,y,a[∫s+hs∫+∞−1Z(t,Wπt,Yt,αt,πt)N(dt,dz)]. |
We get,
0≥Es,w,y,a[∫s+hsA(t,Wπt,Yt,αt,πt)dt]. |
Dividing by h, taking limits as h→0, and by Mean Value Theorem,
0≥limh→0Es,w,y,a[1h∫s+hsA(t,Wπt,Yt,αt,πt)dt]. |
We conclude that
0≥A(t,Wπt,Yt,αt,πt)=∂V∂t+[Wt[πtμ(t,Yt,α)+r(t,Yt,α)(1−πt)]∂V∂W+12π2tW2tσ2(t,Yt,α)∂2V∂W2+∫R(V(t,Wt+πtWtΓ(t,Yt,α,z),Yt,α)−V(t,Wt,Yt,α)−πtWtΓ(t,Yt,α,z)∂V∂W)ν(dz)+(χt−ρYt)∂V∂Y+∑j≠αqαj(V(t,Wt,Yt,j)−V(t,Wt,Yt,α))]=∂V∂t+H(t,W,Y,α,π,∂V∂W,∂2V∂W2,∂V∂Y). |
Taking supremum over all admissible controls, we get
0=∂V∂t+supπ∈A[s,T]H(t,W,Y,α,π,∂V∂W,∂2V∂W2,∂V∂Y). |
Theorem 3.3 (Verification Theorem). Let V(⋅,⋅,⋅,α)∈C1,2,1([0,T]×R+×R) for every α∈M. If V(s,w,y,a) is the solution to the HJB Eq (3.2), then
V(s,w,y,a)≥Gπ(s,w,y,a) |
holds for every π∈A[s,T] and (s,w,y,a)∈[0,T]×R+×R×M. Moreover, π∗∈A[s,T] is optimal if and only if
∂V∂t+H(t,W,Y,α,π∗,∂V∂W,∂2V∂W2,∂V∂Y)=0 |
for a.e. t∈[s,T].
Proof: For any π∈A[s,T] and by Lemma 2.2, we have that
V(s,w,y,a)=Es,w,y,a[V(T,WπT,YT,αT)−∫TsA(t,Wπt,Yt,αt,πt)dt]=Es,w,y,a[U(Wπ;s,w,y,aT)]−Es,w,y,a[∫TsA(t,Wπt,Yt,αt,πt)dt]=Gπ(s,w,y,a)−Es,w,y,a[∫Ts∂V∂t+H(t,W,Y,α,πt,∂V∂W,∂2V∂W2,∂V∂Y)dt]. | (3.4) |
By Theorem 3.2, the integral of the last line is at most zero for any π∈A[s,T]. It follows that
V(s,w,y,a)≥Gπ(s,w,y,a). |
For the second part, we assume that π∗∈A[s,T] is optimal. Then from the last equality of (3.4),
V(s,w,y,a)≥Gπ∗(s,w,y,a)−Es,w,y,a[∫Ts∂V∂t+H(t,W,Y,α,π∗,∂V∂W,∂2V∂W2,∂V∂Y)dt]0≥−Es,w,y,a[∫Ts∂V∂t+H(t,W,Y,α,π∗,∂V∂W,∂2V∂W2,∂V∂Y)dt]. |
Again, using the fact that the integral is at most zero by Theorem 3.2, it follows that
∂V∂t+H(t,W,Y,α,π∗,∂V∂W,∂2V∂W2,∂V∂Y)=0. |
Before taking on a particular utility function, we consider at time s a portion of the Hamiltonian H that involves wealth w and control πs,
w[πsμ(s,y,a)+r(s,y,a)(1−πs)]∂V∂w+12π2sw2σ2(s,y,a)∂2V∂w2+∫+∞−1(V(s,w+πswΓ(s,y,a,z),y,a)−V(s,w,y,a)−πtwΓ(s,y,a,z)∂V∂w)ν(dz). | (4.1) |
Theorem 4.1. If the utility function is U(w)=wγaγa, 0<γa<1, then the value function is
V(s,w,y,a)=ζa(s,y)wγaγa, | (4.2) |
where ζa(s,y) is the solution to the boundary-value problem
{∂ζa(s,y)∂s+∂ζa(s,y)∂y(χs−ρy)+γaζa(s,y)F(π∗;s,y,a)+∑j≠aqa,j(ζj(s,y)−ζa(s,y))=0ζa(T,YT)=1 | (4.3) |
with
F(π∗;s,y,a)=[π∗μ(s,y,a)+r(s,y,a)(1−π∗)]−12(1−γa)π∗2σ2(s,y,a)+1γa∫+∞−1[(1+π∗Γ(s,y,a,z))γa−1−γaπ∗Γ(s,y,a,z)]ν(dz). |
Proof: With this value function (4.1) becomes
ζa(s,y)w[πsμ(s,y,a)+r(s,y,a)(1−πs)]wγa−1+ζa(s,y)12π2sw2σ2(s,y,a)(γa−1)wγa−2+ζa(s,y)wγaγa∫+∞−1[(1+πsΓ(s,y,a,z))γa−1−γaπtΓ(s,y,a,z)]ν(dz)=ζa(s,y)wγa{[πsμ(s,y,a)+r(s,y,a)(1−πs)]−12(1−γa)π2sσ2(s,y,a)+1γa∫+∞−1[(1+πsΓ(s,y,a,z))γa−1−γaπsΓ(s,y,a,z)]ν(dz)}=ζa(s,y)wγaF(πs;s,y,a). | (4.4) |
The HJB Eq (3.2) now becomes
0=∂ζa(s,y)∂swγaγa+∂ζa(s,y)∂y(χs−ρy)wγaγa+ζa(s,y)wγa{supπ∈A[s,T]F(πs;s,y,a)}+wγaγa∑j≠aqa,j(ζj(s,y)−ζa(s,y)). | (4.5) |
Since wγa>0, Eq (4.5) further becomes
0=∂ζa(s,y)∂s+∂ζa(s,y)∂y(χs−ρy)+γaζa(s,y){supπ∈A[s,T]F(πs;s,y,a)}+∑j≠aqa,j(ζj(s,y)−ζa(s,y)). | (4.6) |
By applying the first order condition on F(πs;s,y,a), we have
0=∂F∂π=μ(s,y,a)−r(s,y,a)−πs(1−γa)σ2(s,y,a)−∫+∞−1[(1+πsΓ(s,y,a,z))γa−1−1]Γ(s,y,a,z)ν(dz). |
Solving for πs, we get
πs=1(1−γa)σ2(s,y,a)(μ(s,y,a)−r(s,y,a)−∫+∞−1[(1+πsΓ(s,y,a,z))γa−1−1]Γ(s,y,a,z)ν(dz)). | (4.7) |
Since for every (s,y,a)∈[0,T]×R×M,
∂2F∂π2=−(1−γa)σ2(s,y,a)−∫+∞−1(1+πsΓ(s,y,a,z))γa−2Γ2(s,y,a,z)ν(dz)<0, |
it follows that (4.7) is a maximum and F(π∗;s,y,a) is optimal. Thus, (4.6) becomes
0=∂ζa(s,y)∂s+∂ζa(s,y)∂y(χs−ρy)+γaζa(s,y)F(π∗;s,y,a)+∑j≠aqa,j(ζj(s,y)−ζa(s,y)). | (4.8) |
Note that (4.8) is a first order linear partial differential equation. Hence, it is solvable and solutions exist. To obtain a unique solution for (4.8) we impose the boundary condition
ζa(T,YT)=1 | (4.9) |
and also to be consistent with (3.2).
Theorem 4.2. If the utility function is U(w)=ln(w), then the value function is
V(s,w,y,a)=ξa(s)lnw+ϱ(y)lnw+ζa(s), | (4.10) |
where
ϱ(y)=1−ξa(s)+∫Ts∑j≠aqa,j(ξj(u)−ξa(u))du | (4.11) |
and ζa(s) is the solution to the coupled ordinary differential equation terminal-value problem:
{ζ′a(s)+∑j≠aqa,j(ζj(s)−ζa(s))+[ξa(s)+ϱ(y)]F(π∗;s,y,a)=0ζa(T)=0,ξa(T)=12,ϱ(YT)=12 | (4.12) |
with
F(π∗;s,y,a)=[π∗μ(s,y,a)+r(s,y,a)(1−π∗)]−12π∗2σ2(s,y,a)+∫+∞−1(ln(1+π∗Γ(s,y,a,z))−π∗Γ(s,y,a,z))ν(dz). |
Proof: With this value function (4.1) becomes
[ξa(s)+ϱ(y)][πsμ(s,y,a)+r(s,y,a)(1−πs)]−[ξa(s)+ϱ(y)]12π2sσ2(s,y,a)+[ξa(s)+ϱ(y)]∫+∞−1(ln(1+πsΓ(s,y,a,z))−πsΓ(s,y,a,z))ν(dz)=[ξa(s)+ϱ(y)]{[πsμ(s,y,a)+r(s,y,a)(1−πs)]−12π2sσ2(s,y,a)+∫+∞−1(ln(1+πsΓ(s,y,a,z))−πsΓ(s,y,a,z))ν(dz)}=[ξa(s)+ϱ(y)]F(πs;s,y,a). | (4.13) |
And the HJB Eq (3.2) now reads
0=ξ′a(s)lnw+ζ′a(s)+[ξa(s)+ϱ(y)]{supπ∈A[s,T]F(πs;s,y,a)}+ϱ′(y)(χs−ρy)lnw+lnw∑j≠aqa,j(ξj(s)−ξa(s))+∑j≠aqa,j(ζj(s)−ζa(s)). | (4.14) |
We split (4.14) into
ξ′a(s)lnw+lnw∑j≠aqa,j(ξj(s)−ξa(s))+ϱ′(y)(χt−ρy)lnw=0 | (4.15) |
and
ζ′a(s)+∑j≠aqa,j(ζj(s)−ζa(s))+[ξa(s)+ϱ(y)]{supπ∈A[s,T]F(πs;s,y,a)}=0. | (4.16) |
For lnw≠0, (4.15) becomes
ξ′a(s)+∑j≠aqa,j(ξj(s)−ξa(s))=−ϱ′(y)(χs−ρy). | (4.17) |
Since the delay is a function of time, we integrate (4.17) with respect to t to obtain
ξa(u)|Ts+∫Ts∑j≠aqa,j(ξj(u)−ξa(u))du=−ϱ(Yu)|Ts. | (4.18) |
We impose the terminal condition ξa(T)=12=ϱ(YT) so that (4.10) remains consistent with (3.2). Hence (4.18) becomes
12−ξa(s)+∫Ts∑j≠aqa,j(ξj(u)−ξa(u))du=−12+ϱ(y) | (4.19) |
which can be expressed as
ϱ(y)=1−ξa(s)+∫Ts∑j≠aqa,j(ξj(u)−ξa(u))du | (4.20) |
and which can be substituted into (4.16).
Similarly, by applying the first order condition on F(πs;s,y,a), we have
0=∂F∂π=μ(s,y,a)−r(s,y,a)−πsσ2(s,y,a)−∫+∞−1πsΓ2(s,y,a,z)1+πsΓ(s,y,a,z)ν(dz). |
Solving for πs, we get
πs=1σ2(s,y,a)[μ(s,y,a)−r(s,y,a)−∫+∞−1πsΓ2(s,y,a,z)1+πsΓ(s,y,a,z)ν(dz)]. | (4.21) |
Now, for every (s,y,a)∈[0,T]×R×M,
∂2F∂π2=−σ2(s,y,a)−∫+∞−1Γ2(s,y,a,z)(1+πsΓ(s,y,a,z))2ν(dz)<0. |
It follows that (4.21) is a maximum and F(π∗;s,y,a) is optimal. Thus (4.16) becomes
ζ′a(s)+∑j≠aqa,j(ζj(s)−ζa(s))+[ξa(s)+ϱ(y)]F(π∗;s,y,a)=0. | (4.22) |
Since both the Markov chain αt and delay Yt are dependent on the time variable, (4.22) is a first order linear ordinary differential equation. Thus, solutions exist for (4.22). We impose the terminal condition
ζa(T)=0 | (4.23) |
in order for (4.22) to have a unique solution and to be consistent with (3.2).
We considered a portfolio optimization problem where the riskless asset and the coefficients of the risky asset, represented by a Levy price process, depends on time t, delay Yt and Markov chain αt. We formulated a finite time horizon Merton-type optimization problem and came up with a version of the stochastic chain rule for the system we are working. This chain rule serves as the main machinery in solving the optimization problem which generates a Hamilton-Jacobi-Bellman (HJB) equation bearing the aforementioned variables. In the world of dynamic programming involving portfolio optimization, the HJB equation we came up with is novel in the sense that it incorporated both the delay and regime switching, a result deemed unlikely at first given the discontinuous nature of Levy processes. This obstacle was overcame by fixing the delay variable in order for it to be defined even in stochastic systems with jumps.
The main results of this paper are the optimal portfolio π∗, which we obtained for logarithmic and power utility functions, and the value functions which represent the solution to the portfolio optimization problem. The optimal portfolio was found by invoking the first order condition on the HJB equation. We found that for a logarithmic utility function, the solution consists of four functions, three of which are interrelated via a coupled differential equation. The value function is simpler for the case of a power utility in the sense that the solution is represented by the product of wealth and a function which solves a first order linear partial differential equation.
The authors declare that they have no competing interests.
[1] |
R. C. Merton, Lifetime portfolio selection under uncertainty: The continuous-time case, Rev. Econ. Stat., 51 (1969), 247–257. doi: 10.2307/1926560
![]() |
[2] |
N. C. Framstad, B. Oksendal, A. Sulem, Optimal consumption and portfolio in a jump diffusion market with proportional transaction costs, J. Math. Econ., 35 (2001), 233–257. doi: 10.1016/S0304-4068(00)00067-7
![]() |
[3] |
I. Elsanosi, B. Oksendal, A. Sulem, Some solvable stochastic control problems with delay, Stochs. Stoch. Rpts., 71 (2000), 69–89. doi: 10.1080/17442500008834259
![]() |
[4] | B. Larssen, Dynamic programming in stochastic control of systems with delay, Stoch. Anal. Apps., 74 (2002), 651–673. |
[5] |
B. Larssen, H. Risebro, When are HJB equations in stochastic control of delay systems finite dimensional, Stoch. Anal. Apps., 21 (2003), 643–671. doi: 10.1081/SAP-120020430
![]() |
[6] |
M. Chang, T. Pang, Y. Yang, A stochastic portfolio optimization model with bounded memory, Maths. Ops. Res., 36 (2011), 604–619. doi: 10.1287/moor.1110.0508
![]() |
[7] |
T. Pang, A. Hussain, A stochastic portfolio optimization model with complete memory, Stoch. Anal. Apps., 35 (2017), 742–766. doi: 10.1080/07362994.2017.1299629
![]() |
[8] |
M. Davis, A. Norman, Portfolio selection with transaction costs, Maths. Ops. Res., 15 (1990), 676–713. doi: 10.1287/moor.15.4.676
![]() |
[9] |
N. Azevedo, D. Pinheiro, G. Weber, Dynamic programming for a Markov-switching jump–diffusion, J. Comput. App. Math., 267 (2014), 1–19. doi: 10.1016/j.cam.2014.01.021
![]() |
[10] |
L. Pasin, T. Vargiolu, Optimal portfolio for CRRA utility functions when risky assets are exponential additive processes, Econ. Notes, 39 (2010), 65–90. doi: 10.1111/j.1468-0300.2010.00222.x
![]() |
[11] |
N. Bäuerle, U. Rieder, Portfolio optimization with Markov-modulated stock prices and interest rates, IEEE T. Automat. Cnotr., 49 (2004), 442–447. doi: 10.1109/TAC.2004.824471
![]() |
[12] |
S. Pagliarani, T. Vargiolu, Portfolio optimization in a defaultable Lévy-driven market model, OR Spec., 37 (2015), 617-654. doi: 10.1007/s00291-014-0374-7
![]() |
[13] | A. Valdez, T. Vargiolu, Optimal portfolio in a regime switching model, Seminar on Stochastic Analysis, Random Fields and Applications VII. Prog. Prob., 67 (2013), 435–449. |
[14] | T. Pang, Y. Yong, A new stochastic model for stock price with delay effects, Proceedings of the Conference on Control and its Applications, 2019,110–117. |
[15] | N. C. Framstad, B. Oksendal, A. Sulem, Optimal consumption and portfolio in a jump diffusion market, A. Shyriaev et al (eds): Workshop on Mathematical Finance, 1998, 9–20. |
[16] | S. Mohammed, Stochastic differential systems with memory: Theory, examples and applications, Stochastic Analysis and Related Topics VI. The Geilo Workshop, Prog. Prob., 42 (1998), 1–77. |
[17] | G. Yin, C. Zhu, Hybrid Switching Diffusions: Properties and Applications, Springer, 2010. |
[18] | R. Cont, P. Tankov, Financial Modeling with Jump Processes, Chapman and Hall, 2004. |
[19] | B. Oksendal, A. Sulem, Applied Stochastic Control of Jump Diffusions, Springer, 2005. |
[20] | A. Shiryaev, J. Jacod, Limit Theorems for Stochastic Processes, 2Eds., Springer, 2002. |
[21] | D. Applebaum, Levy Processes and Stochastic Calculus (Cambridge Studies in Advanced Mathematics), 2Eds., Cambridge University Press, 2009. |
[22] | P. Protter, Stochastic Integration and Differential Equations, Springer, 2010. |
[23] | A. Cartea, S. Jaimungal, J. Penalva, Algorithmic and High-Frequency Trading, Cambridge University Press, 2015. |
[24] | M. Davis, Markov Models and Optimization, Chapman and Hall, 1993. |
[25] | W. Fleming, H. Soner, Controlled Markov Processes and Viscosity Solutions, 2nd ed., Springer-Verlag, 2006. |
[26] | H. Pham, Continuous-time Stochastic Control and Optimization with Financial Applications, Springer, 2009. |