
This paper is devoted to study a robust optimal excess-of-loss reinsurance and investment problem with p-thinning dependent risks for an ambiguity-averse insurer (AAI). Assume that the AAI's wealth process consists of two p-thinning dependent classes of insurance business. The AAI is allowed to purchase excess-of-loss reinsurance and invest in a financial market consisting of one risk-free asset and one risky asset, where risky asset's price follows CEV model. Under the criterion of maximizing the expected exponential utility of AAI's terminal wealth, the explicit expressions of the optimal excess-of-loss reinsurance and investment strategy are derived by employing techniques of stochastic control theory. Moreover, we provide the verification theorem and present some numerical examples to analyze the impacts of parameters on our optimal control strategies.
Citation: Lei Mao, Yan Zhang. Robust optimal excess-of-loss reinsurance and investment problem with p-thinning dependent risks under CEV model[J]. Quantitative Finance and Economics, 2021, 5(1): 134-162. doi: 10.3934/QFE.2021007
[1] | Daheng Peng, Fang Zhang . Mean-variance Optimal Reinsurance-investment Strategy in Continuous Time. Quantitative Finance and Economics, 2017, 1(3): 320-333. doi: 10.3934/QFE.2017.3.320 |
[2] | Jie Xiong, Geng Deng, Xindong Wang . Extension of SABR Libor Market Model to handle negative interest rates. Quantitative Finance and Economics, 2020, 4(1): 148-171. doi: 10.3934/QFE.2020007 |
[3] | Sang Phu Nguyen, Toan Luu Duc Huynh . Portfolio optimization from a Copulas-GJR-GARCH-EVT-CVAR model: Empirical evidence from ASEAN stock indexes. Quantitative Finance and Economics, 2019, 3(3): 562-585. doi: 10.3934/QFE.2019.3.562 |
[4] | Nitesha Dwarika . The risk-return relationship and volatility feedback in South Africa: a comparative analysis of the parametric and nonparametric Bayesian approach. Quantitative Finance and Economics, 2023, 7(1): 119-146. doi: 10.3934/QFE.2023007 |
[5] | Michael Jacobs Jr., Frank J. Sensenbrenner . A comparison of methodologies in the stress testing of credit risk – alternative scenario and dependency constructs. Quantitative Finance and Economics, 2018, 2(2): 294-324. doi: 10.3934/QFE.2018.2.294 |
[6] | Cemile Özgür, Vedat Sarıkovanlık . An application of Regular Vine copula in portfolio risk forecasting: evidence from Istanbul stock exchange. Quantitative Finance and Economics, 2021, 5(3): 452-470. doi: 10.3934/QFE.2021020 |
[7] | Oleg S. Sukharev . The restructuring of the investment portfolio: the risk and effect of the emergence of new combinations. Quantitative Finance and Economics, 2019, 3(2): 390-411. doi: 10.3934/QFE.2019.2.390 |
[8] | Michael Stutzer . Style investing and the ICAPM. Quantitative Finance and Economics, 2018, 2(3): 702-716. doi: 10.3934/QFE.2018.3.702 |
[9] | Elvira Caloiero, Massimo Guidolin . Volatility as an Alternative Asset Class: Does It Improve Portfolio Performance?. Quantitative Finance and Economics, 2017, 1(4): 334-362. doi: 10.3934/QFE.2017.4.334 |
[10] | Mara Madaleno, Elisabete Vieira . Volatility analysis of returns and risk: Family versus nonfamily firms. Quantitative Finance and Economics, 2018, 2(2): 348-372. doi: 10.3934/QFE.2018.2.348 |
This paper is devoted to study a robust optimal excess-of-loss reinsurance and investment problem with p-thinning dependent risks for an ambiguity-averse insurer (AAI). Assume that the AAI's wealth process consists of two p-thinning dependent classes of insurance business. The AAI is allowed to purchase excess-of-loss reinsurance and invest in a financial market consisting of one risk-free asset and one risky asset, where risky asset's price follows CEV model. Under the criterion of maximizing the expected exponential utility of AAI's terminal wealth, the explicit expressions of the optimal excess-of-loss reinsurance and investment strategy are derived by employing techniques of stochastic control theory. Moreover, we provide the verification theorem and present some numerical examples to analyze the impacts of parameters on our optimal control strategies.
Reinsurance and investment becomes more and more important in the insurance risk management. Recently, there has been a great deal of interest to investigate the optimal reinsurance and investment problems. For instance, Hipp and Plum (2000), Schmidli (2001) and Promislow and Young (2005) considered the optimal investment and/or reinsurance strategies for insurers to minimize the ruin probability; Irgens and Paulsen (2004), Yang and Zhang (2005), Liang et al. (2011), Guan and Liang (2014), Zhu (2015), Pan et al. (2019) and Zhang and Zhao (2020) studied the optimal problems under the criterion of maximize the survival probability or the expected utility of the insurer's terminal wealth; Bai and Zhang (2008), Zeng et al. (2013), Li et al. (2015), Zeng et al. (2016), Liang et al. (2016) and Tian et al. (2020) investigated the optimal reinsurance and investment problems under mean-variance criterion.
However, most of the above-mentioned literatures usually assume that there is only one class of insurance business in the insurer's wealth process. In reality, there usually are more than one classes of insurance businesses in the operation and management of an insurer, i.e., car claims and medical claims. Moreover, different insurance businesses are usually correlated through some way. A typical example is that a traffic accident (or fire accidents or earthquakes or aviation accidents and so on) may cause medical claims or property loss or death claims. Therefore, it is necessary to investigate the dependent risks in the actuarial literature. Liang and Yuen (2016) and Yuen et al. (2015) began to investigate optimal proportional reinsurance problems with dependent risks, they assumed that there are two or more dependent classes of insurance business in the insurer's wealth process and claim number processes are correlated through a common shock component, and their aims were to maximize the expected utility of the insurer's terminal wealth and derived the optimal proportional strategies. Then Bi et al. (2016) studied the optimal proportional investment and reinsurance problems under mean-variance criterion for the risk model with common shock dependence. Later, Zhang and Zhao (2020) considered the optimal proportional reinsurance and investment problem with p-thinning dependent risks. For other research about dependent risks, one can refer to Yuen et al. (2002), Gong et al. (2012), Liang and Wang (2012), Liang et al. (2016) and the references therein.
Moreover, ambiguity is still being worthy of further exploration. In practice, model uncertainties do exist widely in finance, especially in insurance and portfolio selection. In the recent years, optimal investment-reinsurance problems with ambiguity have been paid more attention. For example, Yi et al. (2013) and Zheng et al. (2016) optimized proportional reinsurance and investment problems with model uncertainty for an insurer. Huang et al. (2017) considered a robust optimal proportional reinsurance and investment problems for both an insurer and a reinsurer. Yang et al. (2017) took multiple dependent classes of insurance business into account and studied the optimal mean-variance proportional reinsurance-investment problem with delay. Zeng et al. (2016) analyzed a robust optimal proportional reinsurance-investment problem under the mean-variance criterion for an ambiguity-averse insurer (AAI) who worries about model uncertainty. Besides, excess-of-loss reinsurance is also more important, as shown in Asmussen et al. (2000), it is even better than the proportional reinsurance in most situations. Later, A and Li (2015), Li et al. (2017), A et al. (2018) and Zhang and Zhao (2019) analyzed robust optimal excess-of-loss reinsurance and investment problems for an AAI.
To the best of our knowledge, few researches focus on the robust optimal excess-of-loss reinsurance and investment problem with p-thinning dependent risks. Inspired by Zhang and Zhao (2019), this paper not only takes excess-of-loss reinsurance into account, but also considering the model uncertainly. Suppose that there are two p-dependent classes of insurance business in the insurer's wealth process. The insurer can purchase excess-of-loss reinsurance and invest in a risk-free asset and a risky asset, where the price process of risky asset follows CEV model. Firstly, a robust optimal control problem with p-thinning dependent risks is formulated. Secondly, the robust optimal excess-of-loss reinsurance and investment strategies are derived under the criterion of maximizing the expected exponential utility of AAI's terminal wealth. This paper has three following high contributions: (ⅰ) both p-thinning dependent risks and excess-of-loss reinsurance are considered in an optimal problem; (ⅱ) ambiguity is taken into account in this problem; (ⅲ) some special cases such as the case of investment-only and ambiguity-neutral insurer are given, which means that our results generalized some existing results, e.g., Gu et al. (2012), A et al. (2015), A et al. (2018) and Zhang and Zhao (2020).
The rest of this paper is organized as follows. Formulation of our model is presented in section 2. Section 3 derives robust optimal strategies by maximizing the utility of AAI's terminal wealth. We give the verification theorem in Section 4. Section 5 provides some special cases of our model. In Section 6, some numerical simulations are presented to illustrate our results. Section 7 concludes the paper.
Let (Ω,F,P) be a complete probability space with filtration{Ft,t∈[0,T]},where Ft stands for the information of the market available up to time t andTis a positive finite constant which represents the terminal time. All processes introduced below are assumed be well-defined and adapted processes in this space. Assume that trading takes place continuously, without taxes or transaction costs, and that all securities are infinitely divisible.
The insurer's wealth process {R(t),t∈[0,T]} with two dependent classes of insurance business is described by
R(t)=x0+ct−N(t)∑i=1Xi−Np(t)∑i=1Yi | (1) |
where x0⩾0 is the initial surplus, c>0 represents the premium rate; Xi is the ith claim size from the first class of business; {Xi,i⩾1} are assumed to be i.i.d. positive random variables with common distribution function denotes FX(⋅). Denote finite first moment E[Xi]=μX>0 and second moment E[X2i]=σ2X; Yi is the ith claim size from the second class of business and {Yi,i⩾1} are assumed to be i.i.d. positive random variables with common distribution function FY(⋅). Denote finite first moment E[Yi]=μY>0 and second moment denote E[Y2i]=σ2Y. The claim number process {N(t),t⩾0} is assumed to be a Poisson process with intensity λ>0 representing the number of claims occurring in time interval [0,t], while {Np(t),t⩾0} is a p-thinning process of {N(t),t⩾0} which is another Poisson process with intensity pλ. As we know, each claim Xi in reality may or may not cause another claim Yi, if yes, we assume that the claim Yi is caused with probability p. This paper assumes that the events whether each claim Xi causes another claim Yi or not are mutually independent. Thus, the claim number process Np(t) for the claims Yi is a p-thinning process of N(t) for claims Xi. The compound Poisson process ∑N(t)i=1Xi and ∑Np(t)i=1Yi are the cumulative amount of claims Xi and Yi in time interval [0,t], respectively. Moreover, we assume that {Xi,i⩾1},{Yi,i⩾1} and {N(t),t⩾0} are mutually independent.
As we all know, a serious epidemic like Covid-19 pandemic has been caused medical claims or death claims. The thinning-dependent structure in (1) can be interpreted as that medical insurance claims (i.e., Xi) causes death insurance claims (i.e., Yi) with probability p due to Covid-19 pandemic, then the death claim number process (i.e.Np(t)) is a p-thinning process of the claim number process of medical insurance (i.e.N(t)).
In addition, the insurer's premium rate is calculated according to the expected value principle, i.e.
c=λμX(1+η1)+λpμY(1+η2) |
where ηi>0 is the insurer's safety loading of the ith class of insurance business.
Denote by D1=sup{x:FX(x)⩽1}<+∞, then FX(0)=0,0<FX(x)<1 for 0<x<D1 and FX(x)=1 for x⩾D1; denote by D2=sup{y:FY(y)⩽1}<+∞, then FY(0)=0,0<FY(y)<1 for 0<y<D2, and FY(y)=1 for y⩾D2.
Assume that the insurer is allowed to purchase excess-of-loss reinsurance in order to reduce the underlying claims risk. Let mi(i=1,2) be (fixed) excess-of-loss retention levels, and let
ˉXm1i=min{Xi,m1}ˉYm2i=min{Yi,m2} |
be the parts of the first claims and the second claims held by the insurer, respectively. Then by (1), the wealth process {Xm(t),t∈[0,T]} after considering reinsurance with retention levels m=(m1,m2)
becomes
dXm(t)=cmdt−dN(t)∑i=1ˉXm1i−dNp(t)∑i=1ˉYm2i,Xm(0)=x0 | (2) |
with the premium rate
cm=c−λ(1+ξ1)(μX−E[ˉXm1i])−λp(1+ξ2)(μY−E[ˉYm2i])=λ((η1−ξ1)μX+(1+ξ1)E[ˉXm1i])+λp((η2−ξ2)μY+(1+ξ2)E[ˉYm2i]) |
where ξi is the reinsurer's safety loading of the ith class of insurance business. This paper assumes ξi>ηi which represents that the reinsurance is not cheap.
According to Grandell (1991) and Promislow and Young (2005), the claim process dN(t)∑i=1ˉXm1i can be approximated by a diffusion risk model as follows
dN(t)∑i=1ˉXm1i=λE[ˉXm1i]dt−γ1dWX(t) |
similarly,
dNp(t)∑i=1ˉYm2i=λpE[ˉYm2i]dt−γ2dWY(t) |
Then the wealth process (2) can be approximated by the following diffusion model
dXm(t)=(c1(m)+c2(m))dt+γ1(m)dWX(t)+γ2(m)dWY(t) |
d=(c1(m)+c2(m))dt+√γ21(m)+γ22(m)+2˜ργ1(m)γ2(m)dW0(t) | (3) |
where
c1(m)=λ((η1−ξ1)μX+ξ1E[ˉXm1i]),γ1(m)=√λE[(ˉXm1i)2] | (4) |
c2(m)=λp((η2−ξ2)μY+ξ2E[ˉYm2i]),γ2(m)=√λpE[(ˉYm2i)2] | (5) |
WX(t) and WY(t) are two standard Brownian motions, and their correlation coefficient is given by
˜ρ=λpγ1(m)γ2(m)E[ˉXm1i]E[ˉYm2i] |
and W0(t) is another standard Brownian motion which is dependent of WX(t) and WY(t).
Remark 1. If p=0, the model (1) will reduce to the classical C-L model in Zhang et al. (2020) and (3) will reduce to that in A and Li (2015), A et al. (2018) and Li et al. (2016), which implies that our model can generalize the optimal formulation of existing results to the case with p-thinning dependent risks.
For convenience, let
{gX(m1)=E[ˉXm1i]=∫m10ˉFX(x)dx,gY(m2)=E[ˉYm2i]=∫m20ˉFY(y)dyGX(m1)=E[(ˉXm1i)2]=∫m102xˉFX(x)dx,GY(m2)=E[(ˉYm2i)2]=∫m202yˉFY(y)dy | (6) |
where ˉFX(x)=1−FX(x),ˉFY(y)=1−FY(y).
Moreover, the insurer is allowed to invest in one risk-free asset (bound) and one risky asset (stock). The price process B(t) of the risk-free asset follows
{dB(t)=r0B(t)dt,t∈[0,T]B(0)=1 |
and the price process S(t) of risky asset is given by the following CEV model
{dS(t)=μS(t)dt+σSβ+1(t)dW1(t),t∈[0,T]S(0)=S0 | (7) |
where r0(>0) is the interest rate of the bond, μ(>r0) represents the expected instantaneous rate of the risky asset, σ(>0) is the volatility of the risky asset price and β(⩾0) is the elasticity parameter, respectively.{W1(t),t⩾0} is a standard Ft-adapted Brownian motion, independent of {Xi,i⩾1},{Yi,i⩾1} and {N(t),t⩾0}.
Remark 2. If β=0 in (7), the CEV model will reduce to the GBM model.
Let u:={u(t):=m1(t),m2(t),π(t),t∈[0,T]} be the reinsurance-investment strategy, where mi(t) is the excess-of-loss retention level for ith claim at time t, note that mi(t)=Di represents "no reinsurance", mi(t)=0 represents "full reinsurance", and π(t) represents the money amount invested in the risky asset at time t, so the amount of money invested in the risk-free asset at time t is Xu(t)−π(t), here Xu(t) represents the wealth of the insurer after adopting strategy u. Therefore, the evolution of Xu(t) is governed by
dXu(t)=[rXu(t)+(μ−r)π(t)+c1(m1(t))+c2(m1(t))]dt+π(t)σSβ(t)dW1(t) |
+√λGX(m1(t))+λpGY(m2(t))+2λpgX(m1(t))gY(m2(t))dW0(t) | (8) |
with Xu(0)=x0.
For notational convenience, we write mi(t) and π(t) as mi and π in below.
In traditional, the insurer is assumed to be ambiguity-neutral with the objective function as follows
supu∈˜∏Et,x,s[U(Xu(T))] | (9) |
where Et,x,s(⋅) denotes the expectation under P considering the dynamics of the process (Xu,S) with initial condition (t,x,s)∈[0,T]×R×R+, and ˜Π is the set of admissible strategies given by Definition 2.1.
However, most of insurers are ambiguity-averse and they always try to be against worse-case scenarios. Thus, it is reasonable to consider an ambiguity-averse insurer (AAI) in the field of insurance. In what follows, a robust portfolio choice with uncertainty will be presented for an AAI. Assume that there is a relatively good estimated model (also called reference model) for the AAI to describe the risky assets prices and claim process, but the AAI is always skeptical about the chosen reference model and always hopes to take alternative models into consideration. According to Anderson (1999), the alternative models can be defined by the following set of probability measures Q which are equivalent to the P:
Q:={Q|Q∼P} |
Definition 2.1. The strategy is said to be admissible if it is Ft-progressively measurable and satisfies
(ⅰ) mi(t)∈[0,Di],∀t∈[0,T],i=1,2;
(ⅱ) EQ∗(∫T0‖u(w)‖2dw)<∞, and EQ∗(∫T0π2(w)S2β(w)dw)<∞, where ‖u(t)‖2=m21(t)+m22(t)+π2(t);
(ⅲ) ∀(t,x,s)∈[0,T]×R×R+, the SDDE (8) has a pathwise unique solution {Xu(t),t∈[0,T]} with EQ∗t,x,s[U(Xu(T))]<∞, where Q∗ is the chosen model to describe the worst case, Et,x,s[⋅] is the condition expectation given Xu(t)=x∈R,S(t)=s∈R+. Let Π be the set of all admissible strategies.
Next, define a process {θ(t)=(θ0(t),θ1(t))|t∈[0,T]} satisfying that
(ⅰ) θ(t) is Ft-measurable, ∀t∈[0,T];
(ⅱ) E[exp{12∫T0‖θ(t)‖2dt}]<∞, where ‖θ(t)‖2=θ20(t)+θ21(t). We denote Θ for the space of all such processes θ.
For ∀θ∈Θ, we define a real-valued process {Λθ(t)|t∈[0,T]} on (Ω, F, P) by
Λθ(t)={−∫t0θ(w)dW(w)−12∫t0‖θ(w)‖2dw} |
where W(t)=(W0(t),W1(t))′. By Ito's differentiation rule,
dΛθ(t)=Λθ(t)[−θ(t)dW(t)] |
So, we know that Λθ(t) is a P-martingale and E[Λθ(t)]=1. For each θ∈Θ, define a new real-world probability measure Q absolutely continuous to P on FT as follows
dQdP|FT:=Λθ(t) |
Now, a family of real-world probability measures Q parameterized by θ∈Θ have been constructed. Applying Girsanov's theorem, we know that WQ(t)=(WQ0(t),WQ1(t))′ with dWQ(t)=dW(t)+θ(t)′dt under Q is a standard two-dimension Brownian motion.
Noting that the alternative models in class Q are different due to the drift terms. Thus, the risky asset's price (7) under Q is
{dS(t)=S(t)[μ−θ1(t)σSβ(t)]dt+σSβ+1(t)dWQ1(t),t∈[0,T]S(0)=s0 | (10) |
and the wealth process (8) under Q can be rewritten as
dXu(t)=[rXu(t)+(μ−r)π+c1+c2−θ0(t)√γ21+γ22+2λpgX(m1)gY(m2)−θ1(t)σπSβ(t)]dt+πσSβ(t)dWQ1(t)+√γ21+γ22+2λpgX(m1)gY(m2)dWQ0(t) | (11) |
by recalling that π=π(t),mi=mi(t),ci=ci(mi(t)),i=1,2,γ21=λGX(m1(t)) and γ22=λpGY(m2(t)).
Suppose that the insurer tries to seek a robust optimal control which is the best choice in some worst-case models. Inspired by Maenhout (2004) and Branger and Larsen (2013), the robust control problem to modify problem (9) can be formulated as follows
supu∈∏infQ∈QEQt,x,s{U(Xu(T))+∫TtΨ(w,Xu(w),S(w),θ(w))dw} | (12) |
where
Ψ(t,x,s,θ)=‖θ(t)‖22ϕ(t,x,s) | (13) |
and EQt,x,s is calculated under Q; ϕ is a strictly positive deterministic function and stands for the strength of the preference for robustness; the deviation from the reference model is penalized by the second term in the expectation, which depends on the relative entropy arising from the diffusion risks.
To solve (12), we define the optimal value function J(t,x,s) as
J(t,x,s)=supu∈∏infQ∈QEQt,x,s{U(Xu(T))+∫TtΨ(w,Xu(w),S(w),θ(w))dw} | (14) |
For convenience, some notations are first provided. Let O0⊂R×R+ be an open set and O=[0,T]×O0. Denoted by C1,2,2 the space of J(t,x,s) such that J and its partial derivatives Jt,Jx,Js,Jxx,Jxs,Jss are continuous on O. To solve the problem (14), applying dynamic programming principle, we can derive the following robust Hamilton-Jacobi-Bellman (HJB) equation (see Kraft(2005) and Maenhout(2006)):
supu∈[0,D1]×[0,D2]×Rinfθ∈R×R{Aθ,uJ(t,x,s)+Ψ(t,x,s,θ)}=0 | (15) |
with boundary condition
J(T,x,s)=U(x) | (16) |
where Aθ,u is the generator of (14) under Q given by
Aθ,uJ=Jt+[rx+(μ−r)π+c1+c2−θ0√γ21+γ22+2λpgX(m1)gY(m2)−θ1σπsβ]Jx |
+12[π2σ2s2β+γ21+γ22+2λpgX(m1)gY(m2)]Jxx+(μs−θ1σsβ+1)Js+12σ2s2β+2Jss+πσ2s2β+1Jxs | (17) |
Note that J is a short notation for J(t,x,s).
This section is devoted to derive the robust optimal strategy u∗ under the worst-case scenario. Suppose that the AAI has the exponential utility function given by as follows
U(x)=−1ve−vx | (18) |
where the constant v (>0) represents the absolute risk aversion coeffcient. As we all know, the exponential utility function plays an important role in insurance mathematics and actuarial practice. It is independent of the level of insurers' wealth and it is the only utility function under the principle of "zero utility" giving a fair premium (see Gerber, 1979).
In what follows, we set out to solve the HJB equation (15). At first, we show the form of ϕ. For analytical tractability, following Maenhout(2004, 2006), we set
ϕ(t,x,s)=−mvJ(t,x,s)⩾0 |
which is state-dependent, where m(⩾0) represents the ambiguity-aversion coefficient, which describes the AAI's attitude to the diffusion risk.
Then, to solve (15) preference parameter ϕ, we conjecture the form of the value function as follows
J(t,x,s)=−1vexp{−vxer(T−t)+G(t,s)} | (19) |
with G(T,s)=0. Let Gt,Gs,Gss be the partial derivatives of G(t,s). According to (19), we have
{Jt=[vxrer(T−t)+Gt]J,Jx=−ver(T−t)J,Js=GsJ,Jxx=v2e2r(T−t)J,Jxs=−ver(T−t)GsJ,Jss=(G2s+Gss)J | (20) |
Step1: Substituting (20) into (15) and rearranging terms, since we get
supu∈[0,D1]×[0,D2]×Rinfθ∈R×R{Gt+(μs−θ1σsβ+1)Gs+12σ2s2β+2(G2s+Gss)−[(μ−r)π+c1+c2−θ0√γ21+γ22+2λpgX(m1)gY(m2)−θ1σπsβ]ver(T−t)+12v2e2r(T−t)(π2σ2s2β+γ21+γ22+2λpgX(m1)gY(m2))−πσ2s2β+1ver(T−t)Gs−v2m(θ20+θ21)}=0. | (21) |
Step2: According to (21), fixing and maximizing over, we derive the first-order condition for the following minimum point :
{θ∗0=mer(T−t)√γ21+γ22+2λpgX(m1)gY(m2)θ∗1=−mvσsβ+1Gs+πmer(T−t)σsβ | (22) |
Observe that θ∗0 is a function on time and the retention levels (m1,m2), while θ∗1 is a function on the investment quote π and (t, s). Replacing (22) back into (21) leads to
Gt+μsGs+m+v2vσ2s2β+2G2s+12σ2s2β+2Gss−Mver(T−t)+infπ{v(m+v)2e2r(T−t)σ2s2βπ2−π(m+v)σ2s2β+1er(T−t)Gs−π(μ−r)ver(T−t)}infm1,m2{m+v2(γ21+γ22+2λpgX(m1)gY(m2))er(T−t)−λξ1gX(m1)−λpξ2gY(m2)}ver(T−t)=0 | (23) |
where
M=λ(η1−ξ1)μX+λp(η2−ξ2)μY | (24) |
Step3: According to the first-order condition for yields
π∗=μ−r(m+v)σ2s2βe−r(T−t)+svGse−r(T−t) | (25) |
Observe that π∗ is time and state-dependent since it is a function on (t, s). Plugging (25) into (23) derives,
Gt+rsGs+12σ2s2β+2Gss−v(μ−r)22(m+v)σ2s2β−Mver(T−t)+infm1,m2{m+v2(γ21+γ22+2λpgX(m1)gY(m2))er(T−t)−λξ1gX(m1)−λpξ2gY(m2)}ver(T−t)=0 | (26) |
For fixed t∈[0,T], let
f(m1,m2,t)=m+v2(γ21+γ22+2λpgX(m1)gY(m2))er(T−t)−λξ1gX(m1)−λpξ2gY(m2) | (27) |
To find the minimizer m∗1(t) and m∗2(t) of f, we assume that FX(x) and FY(y) are continuous and differentiable, and F′X(x)=fX(x),F′Y(y)=fY(y). According to (27) and recalling that γ21=λGX(m1(t)) and γ22=λpGY(m2(t)), for any t∈[0,T], differentiating f with respect to m1,m2 yields
{∂f∂m1=λˉFX(m1)[(m1+pgY(m2))(m+v)er(T−t)−ξ1]∂f∂m2=λpˉFY(m2)[(m2+gX(m1))(m+v)er(T−t)−ξ2] |
We first consider that m∗i(t)∈[0,Di), then 0<ˉFX(m∗1(t))⩽1 and 0<ˉFY(m∗2(t))⩽1. Suppose that there exists at least one point (m∗1,m∗2) satisfying the following equation
{m1+pgY(m2)−ξ1m+ve−r(T−t)=0m2+gX(m1)−ξ2m+ve−r(T−t)=0 | (28) |
Taking (m∗1,m∗2) into the second derivatives of f arriving at the following Hessian matrix
|∂2f∂m21∂2f∂m1∂m2∂2f∂m2∂m1∂2f∂m22|(m∗1,m∗2)=λ2p(m+v)2e2r(T−t)¯FX(m∗1)¯FY(m∗2)[1−p¯FX(m∗1)¯FY(m∗2)]>0 |
then we know that this Hessian matrix is positive definite at the point (m∗1,m∗2). Therefore, if the point (m∗1,m∗2) is found such that (28) holds, then the point (m∗1,m∗2) is indeed the minimizer of f.
In order to determine the point (m∗1,m∗2) clearly, we transform (28) into
m1+pgY(m2)m2+gX(m1)=ξ1ξ2 |
or equivalently,
ξ2m1−ξ1gX(m1)=ξ1m2−pξ2gY(m2) | (29) |
Define three following auxiliary functions
lX(x)=ξ2x−ξ1gX(x),lY(x)=ξ1x−pξ2gY(x),k(x)=ξ2(xξ1−e−r(T−t)m+v) | (30) |
For convenience, we assume that ξ1⩾ξ2. It's easy to verify that both lY(x) and k(x) are strictly increasing functions for x⩾0, so their inverse functions l−1Y(x) and k−1(x) exist. From (29), we get lX(m1)=lY(m2), then
m2=l−1Y(lX(m1)),m1∈(0,D1) | (31) |
Similarly, if ξ1⩽ξ2, we can easily verify that both lX(x) and k(x) are strictly increasing functions for x⩾0, so their inverse functions l−1X(x) and k−1(x) exist. Then
m1=l−1X(lY(m2)),m2∈(0,D2) |
Taking (31) for example, and inserting (31) into the second equation of (28) yields
l−1Y(lX(m1))+gX(m1)=ξ2m+ve−r(T−t) |
Let
h(x)=l−1Y(lX(x))+gX(x)−ξ2m+ve−r(T−t) | (32) |
If h(x)=0 has a solution on [0,D1], the solution will be indeed m∗1 we try to derive, as a result, we will easily determine the value of m∗2. Thus, the minimizer of f(m1,m2,t) in (27) are derived, which is the candidate robust optimal excess-of-loss reinsurance strategy for the optimal control problem (14). These results are summarized in the following theorem.
Theorem 3.1. Assume that and let
al=sup{x⩾0,lX(x)=0},ak=k−1(0)=ξ1m+ve−r(T−t) | (33) |
t0=T−1rlnξ1(m+v)al | (34) |
The candidate robust optimal excess-of-loss reinsurance strategy (m∗1(t),m∗2(t)) of the optimal control problem (14) is given as follows.
(ⅰ) If ξ1(m+v)al⩾1 or al=0, we have t0⩽T. For t0⩽t⩽T, we have (m∗1(t),m∗2(t))=(ˆm∗1(t),ˆm∗2(t)), where
{ˆm∗1(t)=x0(t)I{h(D1)⩾0}+D1I{h(D1)<0},ˆm∗2(t)=l−1Y(lX(x0(t)))I{h(D1)⩾0,lX(x0(t))⩽lY(D2)}+D2I{h(D1)⩾0,lX(x0(t))>lY(D2)}+(ξ2m+ve−r(T−t)−gX(D1))⋅I{h(D1)<0,ξ2m+ve−r(T−t)−gX(D1)<D2}+D2I{h(D1)<0,ξ2m+ve−r(T−t)−gX(D1)⩾D2} | (35) |
and x0(t)∈[al,D1] is the unique solution to h(x)=0 if h(D1)⩾0 holds.
For 0⩽t<t0, if the solution to h(x)=0 exists, we have (m∗1(t),m∗2(t))=(˜m∗1(t),˜m∗2(t)), where
{˜m∗1(t)=ξ1m+ve−r(T−t)I{ξ1m+ve−r(T−t)<D1}+D1I{ξ1m+ve−r(T−t)⩾D1}˜m∗2(t)=0 | (36) |
(ⅱ) If ξ1(m+v)al<1, we have t0>T. For 0⩽t⩽T<t0, we have (m∗1(t),m∗2(t))=(˜m∗1(t),˜m∗2(t)).
Proof. (ⅰ) If al=0 or ξ1(m+v)al≥1, then we have t0≤T.
For t0⩽t⩽T, we have 0<al⩽ak. Since lY(x) and k(x) are increasing functions on [al,+∞), it's easy to verify that h(x) increases on [al,+∞). What's more, we can get h(al)=k(al)⩽k(ak)=0 by the fact that lX(al)=0.
If h(D1)⩾0, the equation h(x)=0 admits a unique solution x0(t)∈[al,D1], thus the robust optimal excess-of-loss strategy is
m∗1(t)=x0(t) | (37) |
Since m2(t)⩽D1, we get
m∗2(t)=l−1Y(lX(x0(t)))I{h(D1)⩾0,lX(x0(t))⩽lY(D2)}+D2I{h(D1)⩾0,lX(x0(t))>lY(D2)} | (38) |
If h(D1)<0, we know that the equation h(x)=0 has a unique solution on (D1,+∞). Due to m1(t)⩽D1, we choose m∗1(t)=D1. At this time, substituting m∗1(t)=D1 back into f(m1,m2,t) in (27), and taking the first derivative of f(D1,m2,t) with respect to m2, the minimizer of f(D1,m2,t) is derived as follows
m∗2(t)=ξ2m+ve−r(T−t)−gX(D1) | (39) |
Consequently, from (36), (37) and (38), m∗1(t) and m∗2(t) can be expressed as ˆm∗1(t) and ˆm∗2(t) in (35) and (36), respectively.
For 0⩽t<t0, we know al>ak holds, then h(al)=k(al)>k(ak)=0. Meanwhile, h(x) is a strictly increasing function on [al,+∞) and h(x)>0,x∈[al,+∞). As a result, the equation h(x)=0 has no solution on [al,+∞). In other words, there does not exist the solution (m∗1,m∗2) satisfying (28) when m1∈(al,+∞) and m2∈[0,+∞). However, it doesn't mean that (28) has no solution on m1∈[0,al) and m2∈[0,+∞). It is not difficult to prove that lX(x) is a convex function on [0,al). So for 0⩽t<t0, if the solution of the equation h(x)=0 exists, it will only be obtained on x∈[0,al) which is indeed m∗1(t) we try to derive.
Because lX(0)=lX(al)=0, and lX(x) is a convex function on [0,al), we can derive that lX(m1)⩽0 on m1∈[0,al). What's more, l−1Y(x) is a strictly increasing function and l−1Y(0)=0, so we can obtain from (31) that m2=l−1Y(lX(m1))⩽0. Due to m2⩾0, we get m∗2(t)=0. Plugging m∗2(t)=0 into f(m1,m2,t) in (27) and taking the first derivative of f(m1,0,t) with respect to m1, we can derive the following minimizer of f(m1,0,t)
m∗1(t)=ξ1m+ve−r(T−t)=ak⩽al |
To sum up, for the case of 0⩽t<t0, we derive (36).
(ⅱ) If ξ1(m+v)al<1, we have t0>T. Thus for 0⩽t⩽T<t0, the inequality al⩾ak holds. The optimal problem is similar to that of 0⩽t<t0 in case (i). At this time, we can obtain the candidate robust optimal excess-of-loss reinsurance strategy for 0⩽t⩽T is (36). This ends the proof of Theorem 3.1.
In order to give the expressions of π∗ and the value function J(t,x,s), we first have to derive the expression of G(t,s) in (25). According to (27), we can rewrite (26) as
Gt+rsGs+12σ2s2β+2Gss−Mver(T−t)−v(μ−r)22(m+v)σ2s2β+ver(T−t)f(m∗1,m∗2,t)=0 | (40) |
Now this problem should be discussed in two cases as follows:
Case Ⅰ: If for the candidate optimal reinsurance strategy of the problem (14) is (35). Denote by the function in (40), we have
G1t+rsG1s+12σ2s2β+2G1ss−Mver(T−t)−v(μ−r)22(m+v)σ2s2β+ver(T−t)f(ˆm∗1,ˆm∗2,t)=0 | (41) |
In what follows, we employ power transformation technique along with variable change method to solve the problem. Let
G1(t,s)=K1(t)+L1(t)y,y=s−2β | (42) |
with K1(T)=0,L1(T)=0. Then substituting (42) into (41) leads to
K′1(t)+σ2β(1+2β)L1(t)+y(L′1(t)−2rβL1(t)−v(μ−r)22(m+v)σ2)−Mver(T−t)+ver(T−t)f(ˆm∗1,ˆm∗2,t)=0 |
By matching coefficients, we derive
{K′1(t)+σ2β(1+2β)L1(t)−Mver(T−t)+ver(T−t)f(ˆm∗1,ˆm∗2,t)=0L′1(t)−2rβL1(t)−v(μ−r)22(m+v)σ2=0 | (43) |
and we have the solution of (43)
L1(t)=−v(μ−r)24rβσ2(m+v)(1−e−2rβ(T−t)) | (44) |
K1(t)=σ2β(1+2β)∫TtL1(w)dw+Mvr(1−er(T−t))+∫Ttver(T−w)f(ˆm∗1(w),ˆm∗2(w),w)dw | (45) |
Similarly, for 0⩽t<t0, the candidate optimal excess-of-loss strategy is (36). Denote the function G by G2 in (40), and let
G2(t,s)=K2(t)+L2(t)y,y=s−2β | (46) |
Similar to (43)–(45), we obtain
L2(t)=−v(μ−r)24rβσ2(m+v)+C1e2rβt | (47) |
and
K2(t)=v(1+2β)(μ−r)2t4r(m+v)−σ2(1+2β)2rC1e2rβt−Mvrer(T−t)−∫t0ver(T−w)f(˜m∗1(w),0,w)dw+C2 | (48) |
Hence for ξ1(m+v)al⩾1, we have
J(t,x,s)={−1vexp{−vxer(T−t)+K1(t)+L1(t)s−2β},t0⩽t⩽T−1vexp{−vxer(T−t)+K2(t)+L2(t)s−2β},0⩽t<t0 |
Since J(t,x,s) is continuous at t=t0, we have
K1(t0)=K2(t0),L1(t0)=L2(t0) | (49) |
and derive
C1=v(μ−r)24rβσ2(m+v)e−2rβT | (50) |
Based on (46)–(49), it is not difficult to find that L1(t)=L2(t). Since K1(t0)=K2(t0), we get
C2=σ2β(1+2β)∫T0L1(w)dw+Mvr+∫t00ver(T−w)f(˜m∗1(w),0,w)dw+∫Tt0ver(T−w)f(ˆm∗1(w),ˆm∗2(w),w)dw | (51) |
As a result, we can rewrite (48) as
K2(t)=σ2β(1+2β)∫TtL1(w)dw−Mr(er(T−t)−1)−∫t0ver(T−w)f(˜m∗1(w),0,w)dw+∫t00ver(T−w)f(˜m∗1(w),0,w)dw+∫Tt0ver(T−w)f(ˆm∗1(w),ˆm∗2(w),w)dw | (52) |
By (25), (42) and (44), it can easily be seen that the candidate optimal investment strategy π∗ for 0⩽t<t0 is the same as that for t0⩽t⩽T, which is
π∗=2r(μ−r)+(μ−r)2(1−e−2rβ(T−t))2rσ2s2βe−r(T−t)m+v | (53) |
Case Ⅱ: If we have For the candidate optimal excess-of-loss strategy is (36). Denote by the function in (40), we have
G3t+rsG3s+12σ2s2β+2G3ss−Mver(T−t)−v(μ−r)22(m+v)σ2s2β+ver(T−t)f(˜m∗1,˜m∗2,t)=0 | (54) |
with boundary condition G3(T,s)=0. Similar to the analysis for t0⩽t⩽T in case I, we conjecture a solution to (54) of the following form
G3(t,s)=K3(t)+s−2βL3(t) | (55) |
with {K_3}(T) = 0, {L_3}(T) = 0. Then by the same method as that in Case I and a direct calculation, we have {L_3}(t) = {L_1}(t), and
{K_3}(t) = {\sigma ^2}\beta (1 + 2\beta)\int_t^T {{L_1}(w)} {\rm{d}}w + \frac{{Mv}}{r}(1 - {e^{r(T - t)}}) + \int_t^T {v{e^{r(T - w)}}f(\tilde m_1^*(w), 0, w)} {\rm{d}}w | (56) |
Therefore, the expression of the candidate optimal investment strategy for 0 \leqslant t \leqslant T can be derived, which is the same as (53) and the corresponding candidate value function can also be obtained.
We summarize the above analysis in following theorem.
Theorem3.2. Recall functions defined in (42), (46) and (55), respectively. For the problem (14), the candidate robust optimal investment strategy is given by
{\pi ^*}(t) = \frac{{2r(\mu - r) + {{(\mu - r)}^2}(1 - {e^{ - 2r\beta (T - t)}})}}{{2r{\sigma ^2}{S^{2\beta }}(t)}}\frac{{{e^{ - r(T - t)}}}}{{m + v}}, {\rm{0}} \leqslant t \leqslant T | (57) |
and (ⅰ) if \frac{{{\xi _1}}}{{(m + v){a_l}}} \geqslant 1 holds, the candidate optimal value function is
J(t, x, s) = \left\{ {\begin{array}{*{20}{c}} { - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {K_{\rm{1}}}(t) + {s^{ - {\rm{2}}\beta }}{L_{\rm{1}}}(t)\}, {t_{\rm{0}}} \leqslant t \leqslant T} \\ { - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {K_2}(t) + {s^{ - {\rm{2}}\beta }}{L_1}(t)\}, {\rm{0}} \leqslant t < {t_{\rm{0}}}} \end{array}} \right. |
(ⅱ) if \frac{{{\xi _1}}}{{(m + v){a_l}}} < 1 holds, the candidate optimal value function is
J(t, x, s) = - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {K_3}(t) + {s^{ - {\rm{2}}\beta }}{L_{\rm{1}}}(t)\}, {\rm{0}} \leqslant t \leqslant T |
Remark 3. According to Theorem 3.1 and Theorem 3.2, we can see that
(1) the candidate robust optimal reinsurance strategy in Theorem 3.1 is similar to Theorem 13 in Zhang and Zhao (2019). Moreover, if p = 0, we have {\eta _2} = {\xi _2} = 0, then Theorem 3.1 will coincide with Corollary 3.6 in A and Li (2015) and A et al. (2018).
(2) if m = {\rm{0}}, the candidate optimal investment strategy in Theorem 3.2 will be the same as Theorem 3.1 in Gu et al. (2012) and Theorem 3.10 in Zhang and Zhao (2020). This implies that our results generalize the existing results to the case of dependent risks and ambiguity.
(3) the insurer's wealth has no influence on the optimal strategies due to the exponential utility function.
This section will apply the result of Kraft (2004, Corollary 1.2) to verify the candidate optimal strategies {m_i} = m_i^*(t), {\theta _i} = \theta _i^*, i = 1, 2 and \pi = {\pi ^*}(t) are indeed optimal, and the value function given by (20) is just the value function J(t, x, s) defined by (14). The main theorem is summarized as follows.
Theorem 4.1. For the optimal control problem (14), if there exists a function and a measurable function which satisfy the HJB Equation (15) and the parameters satisfy
\frac{{4v}}{{m + v}} \cdot \frac{{\sqrt {(\mu - r)(4\mu + 3r)} }}{r} \leqslant \frac{\mu }{{\mu - r}} | (58) |
then ({u^*}(t) = \hat u(t, {S_t}), {\theta ^*}(t) = \hat \theta (t, {S_t})), is an optimal strategy and V(t, x, s) is the corresponding value function.
Proof. From Kraft (2004, Corollary 1.2), the above theorem will hold if ({u^*}, {\theta ^*}) and the corresponding candidate value function V(t, x, s) have the following three properties:
(1) {u^*} is an admissible strategy and {Q^*} is well-defined by {\mathit \Lambda ^{\theta *}}(t) with \theta _0^* and \theta _1^*;
(2) {E^{{Q^*}}}\left[ {\mathop {\sup }\limits_{t \in [0, T]} {{\left| {V(t, {X^{{u^*}}}(t), S(t))} \right|}^4}} \right] < \infty ;
(3) {E^{{Q^*}}}\left[ {\mathop {\sup }\limits_{t \in [0, T]} {{\left| {{\rm{\Psi }}(t, {X^{{u^*}}}(t), S(t), {\theta ^*}(t))} \right|}^2}} \right] < \infty.
Next, we shall verify the properties (1)–(3), respectively.
Proof of (1). m_1^*(t) and m_2^*(t) are deterministic and bounded on [0, T], thus condition (i) in Definition 2.1 is met. By Itô's lemma and (7), we have
{\rm{d}}{S^{ - {\rm{2}}\beta }}(t) = [\beta ({\rm{2}}\beta + {\rm{1}}){\sigma ^{\rm{2}}} - {\rm{2}}\beta \mu {S^{ - {\rm{2}}\beta }}(t)]{\rm{d}}t - {\rm{2}}\beta \sigma {S^{ - \beta }}(t){\rm{d}}{W_{\rm{1}}}(t) | (59) |
According to Zhao et al. (2017) and Jeanblanc et al. (2009), we know that (59) has a unique strong solution. Therefore, for \beta \geqslant {\rm{0, }} (7) has a unique solution such that S(t) \in ({\rm{0}}, + \infty). Moreover, by (57), condition (ⅱ) in Definition 2.1 is met. Condition (ⅲ) in Definition 2.1 can be obtained by Property (2).
Proof of (2). Substituting ({u^*}, {\theta ^*}) into (11), we have the wealth process under ({u^*}, {\theta ^*})
\begin{array}{l} {X^{{u^*}}}(t) = {x_{\rm{0}}}{e^{rt}} + \int_{\rm{0}}^t {{e^{r(t - w)}}} ({c_{\rm{1}}} + {c_{\rm{2}}} + \frac{{v(\mu - r)}}{{m + v}}{\pi ^*}(w))dw \\ \quad \quad \quad - \int_{\rm{0}}^t {{e^{r(t - w)}}} \theta _{\rm{0}}^*(w)\sqrt {\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + {\rm{2}}\lambda p{g_X}(m_{\rm{1}}^*{\rm{(}}w{\rm{)}}){g_Y}(m_{\rm{2}}^{\rm{*}}{\rm{(}}w{\rm{)}})} dw \\ \quad \quad \quad + \int_{\rm{0}}^t {{e^{r(t - w)}}} {\pi ^*}(w)\sigma {S^\beta }(w){\rm{d}}W_1^{{Q^*}}(w) \\ \quad \quad \quad + \int_{\rm{0}}^t {{e^{r(t - w)}}} \sqrt {\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + {\rm{2}}\lambda p{g_X}(m_{\rm{1}}^*{\rm{(}}w{\rm{)}}){g_Y}(m_{\rm{2}}^{\rm{*}}{\rm{(}}w{\rm{)}})} {\rm{d}}W_{\rm{0}}^{{Q^*}}(w) \\ \end{array} |
\begin{array}{l} = {x_{\rm{0}}}{e^{rt}} + \int_{\rm{0}}^t {{e^{r(t - w)}}} ({c_{\rm{1}}} + {c_{\rm{2}}})dw \\ + {e^{ - r(T - t)}}\int_{\rm{0}}^t {\frac{{v(\mu - r)}}{{m + v}}} {B_1}(w){S^{ - 2\beta }}(w)dw \\ - \int_{\rm{0}}^t {{e^{r(t - w)}}} \theta _{\rm{0}}^*(w)\sqrt {\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + {\rm{2}}\lambda p{g_X}(m_{\rm{1}}^*{\rm{(}}w{\rm{)}}){g_Y}(m_{\rm{2}}^{\rm{*}}{\rm{(}}w{\rm{)}})} dw \\ + {e^{ - r(T - t)}}\int_{\rm{0}}^t {\sigma {B_1}(w){S^{ - \beta }}(w){\rm{d}}W_1^{{Q^*}}(w)} \\ \end{array} |
+ \int_{\rm{0}}^t {{e^{r(T - w)}}} \sqrt {\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + {\rm{2}}\lambda p{g_X}(m_{\rm{1}}^*{\rm{(}}w{\rm{)}}){g_Y}(m_{\rm{2}}^{\rm{*}}{\rm{(}}w{\rm{)}})} {\rm{d}}W_{\rm{0}}^{{Q^*}}(w) | (60) |
where B(t) = \frac{{{{(\mu - r)}^2}(1 - {e^{ - 2r\beta (T - t)}})}}{{4r\beta {\sigma ^2}(m + v)}}, {B_1}(t) = \frac{{\mu - r}}{{{\sigma ^2}(m + v)}} + 2\beta B(t).
Inserting (60) into (19), we obtain the following estimate with appropriate constants {M_1} < {M_2},
\begin{array}{l} {\left| {V(t, {X^{{u^*}}}(t), S(t))} \right|^4} \\ \quad = \frac{1}{{{v^4}}}\exp \{ - 4v{e^{r(T - t)}}{X^{{u^*}}}(t) + 4{L_1}(t){S^{ - {\rm{2}}\beta }}(t) + 4K(t)]\} \\ \quad \leqslant {M_1}\exp \{ - 4v{e^{r(T - t)}}{X^{{u^*}}}(t)\} \\ \quad \leqslant {M_2}\exp \{ - 4v\int_{\rm{0}}^t {\frac{{v(\mu - r)}}{{m + v}}} {B_1}(w){S^{ - 2\beta }}(w) - 4v\int_{\rm{0}}^t {\sigma {B_1}(w){S^{ - \beta }}(w){\rm{d}}W_1^{{Q^*}}(w)} \} \\ \quad \cdot \exp \{ - 4v\int_{\rm{0}}^t {{e^{r(T - w)}}} \sqrt {\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + {\rm{2}}\lambda p{g_X}(m_{\rm{1}}^*{\rm{(}}w{\rm{)}}){g_Y}(m_{\rm{2}}^{\rm{*}}{\rm{(}}w{\rm{)}})} {\rm{d}}W_{\rm{0}}^{{Q^*}}(w)\} \end{array} |
= {M_2}\prod\limits_{i = 1}^4 {\exp \{ {E_i}(t)\} }, | (61) |
where
\begin{array}{l} K(t) = {K_1}(t)I\{ \frac{{{\xi _1}}}{{(m + v){a_l}}} \geqslant 1, {t_{\rm{0}}} \leqslant t \leqslant T\} + {K_2}(t)I\{ \frac{{{\xi _1}}}{{(m + v){a_l}}} \geqslant 1, 0 \leqslant t \leqslant {t_{\rm{0}}}\} \\ \quad \quad + {K_3}(t)I\{ \frac{{{\xi _1}}}{{(m + v){a_l}}} < 1\} \\ \end{array} |
and
\left\{ \begin{array}{l} {E_1}(t) = - 4v\int_{\rm{0}}^t {\sigma {B_1}(w){S^{ - \beta }}(w){\rm{d}}W_1^{{Q^*}}(w)} - \int_{\rm{0}}^t {32{v^2}{\sigma ^2}B_1^2(w){S^{ - 2\beta }}(w){\rm{d}}w, } \\ {E_2}(t) = \int_{\rm{0}}^t {4{v^2}B_1^2(w)\left({8{\sigma ^2}{B_1}(w) - \frac{{\mu - r}}{{m + v}}} \right)} {S^{ - 2\beta }}(w){\rm{d}}w, \\ {E_3}(t) = - 4v\int_{\rm{0}}^t {{e^{r(T - w)}}} \sqrt {\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + {\rm{2}}\lambda p{g_X}(m_{\rm{1}}^*{\rm{(}}w{\rm{)}}){g_Y}(m_{\rm{2}}^{\rm{*}}{\rm{(}}w{\rm{)}})} {\rm{d}}W_{\rm{0}}^{{Q^*}}(w) - {E_4}(t) \\ {E_4}(t) = 32{v^2}\int_{\rm{0}}^t {{e^{2r(T - w)}}} \left({\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + {\rm{2}}\lambda p{g_X}(m_{\rm{1}}^*{\rm{(}}w{\rm{)}}){g_Y}(m_{\rm{2}}^{\rm{*}}{\rm{(}}w{\rm{)}})} \right){\rm{d}}w \\ \end{array} \right. | (62) |
Because {L_1}(t) and K(t) are deterministic and bounded on [0, T] and S(t) \in ({\rm{0}}, + \infty), we obtain that the first estimate in (61) is valid. Since m_{\rm{1}}^*{\rm{(}}t{\rm{), }}\, m_{\rm{2}}^*{\rm{(}}t{\rm{), }}\theta _{\rm{0}}^*{\rm{(}}t{\rm{)}} and {e^{r(T - t)}} are deterministic and bounded on [0, T], the second inequality holds.
In what follows, we consider the four integrals about \exp \{ {E_i}(t)\}, i = 1, 2, 3, 4.
Firstly, note that m_{\rm{1}}^*{\rm{(}}t{\rm{)}} and m_{\rm{2}}^*{\rm{(}}t{\rm{)}} are bounded on [0, T], it easy to check that
{E^{{Q^*}}}[\exp \{ 4{E_4}(t)\} ] < \infty | (63) |
Secondly, because {B_1}(t) is deterministic and bounded on [0, T], then by the Lemma 4.3 in Zeng and Taksar (2013), it is easily be seen that \exp \{ 4{E_1}(t)\} and \exp \{ 4{E_3}(t)\} are martingales under {Q^*}, consequently,
{E^{{Q^*}}}[\exp \{ 4{E_1}(t)\} ] < \infty \;{\rm{and}}\;{E^{{Q^*}}}[\exp \{ 4{E_3}(t)\} ] < \infty | (64) |
Thirdly, according to Theorem5.1 in Zeng and Taksar (2013), we obtain a sufficient condition for
{E^{{Q^*}}}[\exp \{ 2{E_2}(t)\} ] < \infty | (65) |
is
128{v^2}{\sigma ^2}B_1^2(w) - 16{v^2}\frac{{\mu - r}}{{m + v}}{B_1}(w) \leqslant \frac{{{\mu ^2}}}{{2{\sigma ^2}}} | (66) |
for \forall w \in [0, T].
Note that \frac{{\mu - r}}{{(m + v){\sigma ^2}}} \leqslant {B_1}(t) \leqslant \frac{{{\mu ^2} - {r^2}}}{{2r(m + v){\sigma ^2}}}, \forall t \in [0, T] and T \in [0, \infty), and by (58), then (66) holds for \forall t \in [0, T] and T \in [0, \infty), because of the property of quadratic function. According to (63)–(65) and Cauchy-Schwarz inequality, we have
\begin{array}{l} {E^{{Q^*}}}{\left| {V(t, {X^{{u^*}}}(t), S(t))} \right|^4} \\ \; \leqslant {M_2}{E^{{Q^*}}}[\exp \{ {E_1}(t) + {E_2}(t) + {E_3}(t) + {E_4}(t)\} ] \\ \; \leqslant {M_2}{\left({{E^{{Q^*}}}[\exp \{ 2{E_1}(t) + 2{E_2}(t)\} ]} \right)^{\frac{1}{2}}}{\left({{E^{{Q^*}}}[\exp \{ 2{E_3}(t) + 2{E_4}(t)\} ]} \right)^{\frac{1}{2}}} \\ \; \leqslant {M_2}{\left({{E^{{Q^*}}}[\exp \{ 4{E_1}(t)\} ] \cdot {E^{{Q^*}}}[\exp \{ 4{E_2}(t)\} ]} \right)^{\frac{1}{4}}} \cdot {\left({{E^{{Q^*}}}[\exp \{ 4{E_1}(t)\} ] \cdot {E^{{Q^*}}}[\exp \{ 4{E_2}(t)\} ]} \right)^{\frac{1}{4}}} \\ \end{array} |
< \infty | (67) |
Hence, property (2) holds.
Proof of (3). By (22), with an appropriate constant {M_3}, we get
E\left[ {\exp\{ \frac{1}{2}\int_0^T {{{\left\| {{\theta ^*}(t)} \right\|}^2}{\rm{d}}t} \} } \right]\; \leqslant {M_3}E\left[ {\exp\{ \frac{1}{2}\int_0^T {\frac{{{m^2}{{(\mu - r)}^2}}}{{{{(m + v)}^2}{\sigma ^2}}}} {S^{ - 2\beta }}(w){\rm{d}}w\} } \right] |
\leqslant {M_3}E\left[ {\exp\{ \frac{1}{2}\int_0^T {\frac{{{{(\mu - r)}^2}}}{{{\sigma ^2}}}{S^{ - 2\beta }}(w){\rm{d}}w\} } } \right] < \infty | (68) |
where the first estimate in (68) follows from the deterministic and bounded \theta _0^*(t) on [0, T], and by (59) and Theorem 5.1 in Zeng and Taksar (2013), the last estimate is easily derived.
Moreover, according to, we can see that {Q^*} is well-defined.
Inserting {\theta ^*}(t) and {u^*}(t) into (13) arrives at
\begin{array}{l} {E^{{Q^*}}}{\left| {{\rm{\Psi }}(t, {X^{{u^*}}}(t), S(t))} \right|^2} \\ \;\; = {E^{{Q^*}}}\left[ {\frac{{{v^2}}}{{{m^2}}}{{\left| {V(t, {X^{{u^*}}}(t), S(t))} \right|}^2} \cdot {{\left({\frac{1}{2}{{\left\| {{\theta ^*}(t)} \right\|}^2}} \right)}^2}} \right] \\ \end{array} |
\leqslant \frac{{{v^2}}}{{{m^2}}}{\left[ {{E^{{Q^*}}}{{\left| {V(t, {X^{{u^*}}}(t), S(t))} \right|}^4} \cdot {E^{{Q^*}}}{{\left({\frac{1}{2}{{\left\| {{\theta ^*}(t)} \right\|}^2}} \right)}^4}} \right]^{1/2}} < \infty | (69) |
where the first estimate follows from the Cauchy-Schwarz inequality and the last estimate from (67) and (68). Thus, property (3) holds.
With all the properties are satisfied, the result of Kraft (2004, Corollary 1.2) guarantees ({u^*}, {\theta ^*}) is an optimal strategy and V(t, x, s) is the corresponding value function.
This section considers some special cases of the problem (14), such as the investment-only case and the case with ambiguity-neutral insurer (ANI), respectively. These results here are only provided without giving the proofs.
If there is no reinsurance, i.e., {m_{\rm{1}}}(t) = {D_{\rm{1}}}, {m_{\rm{2}}}(t) = {D_{\rm{2}}}, then X_i^m = {X_i}, Y_i^m = {Y_i}, {\xi _1} = {\xi _2} = 0, and (3) can be rewritten as
\operatorname{d} \tilde X(t) = \lambda ({\eta _1}{\mu _X} + p{\eta _2}{\mu _Y})\operatorname{d} t + \sqrt {\lambda \sigma _X^2 + \lambda p\sigma _Y^2 + 2\lambda p{\mu _X}{\mu _Y}} {\rm{d}}{W_0}(t) |
According to (11), the above wealth process for an AAI under the probability Q is
\begin{array}{l} {\rm{d}}\tilde X(t)\; = [r\tilde X(t) + (\mu - r)\pi + \lambda ({\eta _1}{\mu _X} + p{\eta _2}{\mu _Y}) - {\theta _{\rm{0}}}(t)\sqrt {\lambda \sigma _X^{\rm{2}} + \lambda p\sigma _Y^{\rm{2}} + {\rm{2}}\lambda p{\mu _X}{\mu _Y}} \\ \quad \quad \;\;\; - {\theta _1}(t)\sigma \pi {S^\beta }(t)]dt\; + \pi \sigma {S^\beta }(t){\rm{d}}W_1^Q(t) \\ \;\quad \quad \;\; + \sqrt {\lambda \sigma _X^{\rm{2}} + \lambda p\sigma _Y^{\rm{2}} + {\rm{2}}\lambda p{\mu _X}{\mu _Y}} {\rm{d}}W_{\rm{0}}^Q(t) \\ \end{array} | (70) |
and the HJB equation is given by
\mathop {\sup }\limits_{\pi \in R} \mathop {\inf }\limits_{\theta \in R \times R} \left\{ {{A^{\theta, \pi }}\tilde J(t, \tilde X(t), s{\rm{) + \Psi }}(t, \tilde X(t), \theta (t))} \right\} = 0 | (71) |
where \tilde J is a short notation for \tilde J(t, x, s{\rm{), }} representing the optimal value function of the investment-only problem with the boundary condition \tilde J(T, x, s{\rm{)}} = U(x), and
{\mathcal{A}^{\theta, \pi }}\tilde J = {\tilde J_t} + [rx + (\mu - r)\pi + \lambda ({\eta _1}{\mu _X} + p{\eta _2}{\mu _Y}) - {\theta _{\rm{0}}}(t)\sqrt {\lambda \sigma _X^{\rm{2}} + \lambda p\sigma _Y^{\rm{2}} + {\rm{2}}\lambda p{\mu _X}{\mu _Y}} - {\theta _1}\sigma \pi {s^\beta }]{\tilde J_x} |
\begin{array}{l} + \tfrac{1}{2}[{\pi ^2}{\sigma ^2}{s^{2\beta }} + \lambda \sigma _X^{\rm{2}} + \lambda p\sigma _Y^{\rm{2}} + {\rm{2}}\lambda p{\mu _X}{\mu _Y}]{{\tilde J}_{xx}} + (\mu s - {\theta _1}\sigma {s^{\beta + 1}}){{\tilde J}_s} \\ + \tfrac{1}{2}{\sigma ^2}{s^{2\beta + 2}}{{\tilde J}_{ss}} + \pi {\sigma ^2}{s^{2\beta + 1}}{{\tilde J}_{xs}} \\ \end{array} | (72) |
Theorem 5.1. For the investment-only problem (71) under exponential utility (18), i.e. in (15), then the robust optimal investment strategy is
\pi _0^{\rm{*}}(t) = \frac{{2r(\mu - r) + {{(\mu - r)}^2}(1 - {e^{ - 2r\beta (T - t)}})}}{{2r{\sigma ^2}{S^{2\beta }}(t)}}\frac{{{e^{ - r(T - t)}}}}{{m + v}}, {\rm{0}} \leqslant t \leqslant T, | (73) |
and the optimal value function is given by
\tilde J(t, x, s) = - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {\tilde K_{\rm{1}}}(t) + {s^{ - {\rm{2}}\beta }}{L_{\rm{1}}}(t)\}, {\rm{0}} \leqslant t \leqslant T, |
where {L_{\rm{1}}}(t) is given by (44),
{\tilde K_1}(t) = {\sigma ^2}\beta (1 + 2\beta)\int_t^T {{L_1}(w)} {\rm{d}}w + \frac{{\tilde Mv}}{r}(1 - {e^{r(T - t)}}) + \int_t^T {v{e^{r(T - w)}}{{\tilde f}_1}(w)\, } {\rm{d}}w |
and \tilde M = \lambda {\eta _1}{\mu _X} + \lambda p{\eta _{\rm{2}}}{\mu _Y} and {\tilde f_1}(t) is given by
{\tilde f_1}(t) = \frac{{m + v}}{2}(\sigma _X^{\rm{2}} + \lambda p\sigma _Y^{\rm{2}} + {\rm{2}}\lambda p{\mu _X}{\mu _Y}){e^{r(T - t)}}\; - \lambda {\xi _1}{\mu _X} - \lambda p{\xi _2}{\mu _Y} |
In addition, if p = 0, {m_{\rm{1}}}(t) = {D_{\rm{1}}}, it means that only one class of insurance business is considered, in other words, there are no dependent risks.
Corollary 5.2. If for the investment-only problem, the robust optimal investment strategy is the same as (73), and the optimal value function is
\tilde J(t, x, s) = - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {\tilde K_{\rm{2}}}(t) + {s^{ - {\rm{2}}\beta }}{L_{\rm{1}}}(t)\}, {\rm{0}} \leqslant t \leqslant T, |
where
{\tilde K_2}(t) = {\sigma ^2}\beta (1 + 2\beta)\int_t^T {{L_1}(w)} {\rm{d}}w + \frac{{\lambda {\eta _1}{\mu _X}v}}{r}(1 - {e^{r(T - t)}}) + \int_t^T {v{e^{r(T - w)}}{{\tilde f}_2}(w)\, } {\rm{d}}w |
and
{\tilde f_2}(t) = \frac{{m + v}}{2}\sigma _X^{\rm{2}}{e^{r(T - t)}}\; - \lambda {\xi _1}{\mu _X} - \lambda p{\xi _2}{\mu _Y} |
If ambiguity-aversion coefficient m = \rm{0}, our model will reduce to an optimal control problem for an ANI. Then, the wealth process under probability measurers P is given by (8). Let the optimal value function be
\bar J(t, x, s) = \mathop {\sup }\limits_{\bar u \in \prod } {E_{t, x, s}}\left[ {U(\bar X(T))} \right] | (74) |
where \bar u = \left\{ {{\rm{(}}{{\bar m}_1}{\rm{(}}t{\rm{), }}{{\bar m}_2}{\rm{(}}t{\rm{)), }}\bar \pi {\rm{(}}t{\rm{), }}t \in [0, T]} \right\}, and HJB equation is
\begin{array}{l} \mathop {\sup }\limits_{\bar u \in [0, {D_1}] \times [0, {D_2}] \times R} {\mathcal{A}^{\bar u}}\left\{ {{{\bar J}_t} + [rx + (\mu - r)\pi + {c_{\rm{1}}} + {c_{\rm{2}}}]{{\bar J}_x} + \tfrac{1}{2}[{\pi ^2}{\sigma ^2}{s^{2\beta }} + \gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + 2\lambda p{g_X}({m_{\rm{1}}}){g_Y}({m_{\rm{2}}})]{{\bar J}_{xx}}} \right. \\ \quad \;\;\left. { + \mu s{{\bar J}_s} + \tfrac{1}{2}{\sigma ^2}{s^{2\beta + 2}}{{\bar J}_{ss}} + \pi {\sigma ^2}{s^{2\beta + 1}}{{\bar J}_{xs}}} \right\} = 0 \\ \end{array} |
Note that \bar J is the short notation for \bar J(t, x, s) with \bar J(T, x, s) = U(x). Let
\bar k(x) = {\xi _2}(\frac{x}{{{\xi _1}}} - \frac{{{e^{ - r(T - t)}}}}{v}), {\bar a_k} = {\bar k^{ - 1}}(0) = \frac{{{\xi _1}}}{v}{e^{ - r(T - t)}}, {\bar t_0} = T - \frac{1}{r}\ln \frac{{{\xi _1}}}{{v{{\bar a}_l}}} |
\bar h(x) = l_Y^{ - 1}({l_X}(x)) + {g_X}(x) - \frac{{{\xi _2}}}{v}{e^{ - r(T - t)}} |
Note that, if {\xi _1}/v{\bar a_l} \geqslant 1 and \bar h({D_1}) \geqslant 0, according to the proof of Theorem 3.1, by the similar analysis of that the equation \bar h(x) = 0 in (32) has a unique solution {\bar x_0}(t) \in [{\bar a_l}, {D_1}] when \bar h({D_1}) \geqslant 0, it's easy to prove that equation \bar h(x) = 0 has a unique solution {\bar x_0}(t) \in [{\bar a_l}, {D_1}].
Theorem 5.3. When for the optimal control problem (74) of an ANI who ignores ambiguity with utility (18), the optimal investment strategy is
\pi _1^{\rm{*}}(t) = \frac{{2r(\mu - r) + {{(\mu - r)}^2}(1 - {e^{ - 2r\beta (T - t)}})}}{{2r{\sigma ^2}{S^{2\beta }}(t)}}\frac{{{e^{ - r(T - t)}}}}{v}, \;{\rm{0}} \leqslant t \leqslant T, | (75) |
and (ⅰ) if {\bar a_l} = 0 or \frac{{{\xi _1}}}{{v{{\hat a}_l}}} \geqslant 1, {\bar t_0} \leqslant Tholds. For {\bar t_0} \leqslant t \leqslant T, the optimal reinsurance strategy (\bar m_1^*(t), \bar m_2^*(t)) is given by
\left\{ \begin{array}{l} \bar m_1^*(t) = {{\bar x}_0}(t)I\{ \bar h({D_1}) \geqslant 0\} + {D_1}I\{ \bar h({D_1}) < 0\}, \\ \bar m_2^*(t) = l_Y^{ - 1}({l_X}({{\bar x}_0}(t)))I\{ \bar h({D_1}) \geqslant 0, {l_X}({{\bar x}_0}(t)) \leqslant {l_Y}({D_2})\} \\ \quad \quad \;\; + {D_2}I\{ \bar h({D_1}) \geqslant 0, {l_X}({{\bar x}_0}(t)) > {l_Y}({D_2})\} \\ \quad \quad \;\; + \left({\frac{{{\xi _2}}}{v}{e^{ - r(T - t)}} - {g_X}({D_1})} \right) \\ \quad \quad \;\; \cdot I\{ \bar h({D_1}) < 0, \frac{{{\xi _2}}}{v}{e^{ - r(T - t)}} - {g_X}({D_1}) < {D_2}\} \\ \quad \quad \;\; + {D_2}I\{ \bar h({D_1}) < 0, \frac{{{\xi _2}}}{v}{e^{ - r(T - t)}} - {g_X}({D_1}) \geqslant {D_2}\} \\ \end{array} \right. |
For 0 \leqslant t < {\bar t_0}, if the solution to \bar h(x) = 0 exists, the optimal reinsurance strategy ({\tilde {\bar m}}_1^*(t), {\tilde {\bar m}}_2^*(t)) is expressed by
\left\{ \begin{array}{l} {\tilde {\bar m}}_1^*(t) = \frac{{{\xi _1}}}{v}{e^{ - r(T - t)}}I\{ \frac{{{\xi _1}}}{v}{e^{ - r(T - t)}} < {D_1}\} + {D_1}I\{ \frac{{{\xi _1}}}{v}{e^{ - r(T - t)}} \geqslant {D_1}\} \\ {\tilde {\bar m}}_2^*(t) = 0 \\ \end{array} \right. | (76) |
and the optimal value function is
\bar J(t, x, s) = \left\{ {\begin{array}{*{20}{c}} { - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {{\bar K}_{\rm{1}}}(t) + {s^{ - {\rm{2}}\beta }}{{\bar L}_{\rm{1}}}(t)]\}, {{\bar t}_{\rm{0}}} \leqslant t \leqslant T} \\ { - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {{\bar K}_{\rm{2}}}(t) + {s^{ - {\rm{2}}\beta }}{{\bar L}_{\rm{1}}}(t)\}, {\rm{0}} \leqslant t < {{\bar t}_{\rm{0}}}} \end{array}} \right. |
(ⅱ) if \frac{{{\xi _1}}}{{v{{\bar a}_l}}} < 1, in this case {\bar t_0} > T. For 0 \leqslant t \leqslant T < {\bar t_0}, the optimal reinsurance strategy is expressed by (76), and the optimal value function is given by
\bar J(t, x, s) = - \frac{{\rm{1}}}{v}\exp \{ - vx{e^{r(T - t)}} + {\bar K_{\rm{3}}}(t) + {s^{ - {\rm{2}}\beta }}{\bar L_{\rm{1}}}(t)\}, {\rm{0}} \leqslant t \leqslant T,
where
{\bar L_1}(t) = - \frac{{v{{(\mu - r)}^2}}}{{4rv\beta {\sigma ^2}}}(1 - {e^{ - 2r\beta (T - t)}}) |
{\bar K_1}(t) = {\sigma ^2}\beta (1 + 2\beta)\int_t^T {{{\bar L}_1}(w)} {\rm{d}}w + \frac{{Mv}}{r}(1 - {e^{r(T - t)}}) + \int_t^T {v{e^{r(T - w)}}\bar f(\bar m_1^*(w), \bar m_2^*(w), w)} {\rm{d}}w |
\bar f({m_1}, {m_2}, t) = \frac{v}{2}(\gamma _{\rm{1}}^{\rm{2}} + \gamma _{\rm{2}}^{\rm{2}} + 2\lambda p{g_X}({m_{\rm{1}}}){g_Y}({m_{\rm{2}}})){e^{r(T - t)}}\; - \lambda {\xi _1}{g_X}({m_{\rm{1}}}) - \lambda p{\xi _2}{g_Y}({m_2}) |
\begin{array}{l} {{\bar K}_2}(t) = {\sigma ^2}\beta (1 + 2\beta)\int_t^T {{{\bar L}_1}(w)} {\rm{d}}w - \frac{M}{r}({e^{r(T - t)}} - 1) - \int_0^t {v{e^{r(T - w)}}\bar f({\tilde {\bar m}}_1^*(w), 0, w)} {\rm{d}}w \\ {\kern 1pt} \quad \quad + \int_0^{{{\bar t}_0}} {v{e^{r(T - w)}}\bar f({\tilde {\bar m}}_1^*(w), 0, w)} {\rm{d}}w + \int_{{{\bar t}_0}}^T {v{e^{r(T - w)}}\bar f(\bar m_1^*(w), \bar m_2^*(w), w)} {\rm{d}}w \\ \end{array} |
{\bar K_3}(t) = {\sigma ^2}\beta (1 + 2\beta)\int_t^T {{{\bar L}_1}(w)} {\rm{d}}w + \frac{{Mv}}{r}(1 - {e^{r(T - t)}}) + \int_t^T {v{e^{r(T - w)}}\bar f({\tilde {\bar m}}_1^*(w), 0, w)} {\rm{d}}w |
and M is given by (24).
Furthermore, if p = 0 in Theorem 5.3, the optimal reinsurance-investment will coincide with Corollary 4.5 in Gu et al. (2012) and A et al. (2018). This means that our model extends the results in Gu et al. (2012) and A et al. (2018) to the case of robust optimal formulation under p-thinning dependent risks.
This section investigates the impacts of parameters on the optimal strategies. Here only the analysis for the case of {\xi _1}/((m + v){a_l}) \geqslant 1 is taken for consideration. For drawing convenience, the claim sizes {X_i} and {Y_i} are assumed follow uniform distribution U(0, 1). Throughout this section, unless otherwise stated, the basic parameters are given by {\xi _1} = 2, {\xi _2} = 1, r = 0.05, \mu = 0.09, m = 0.5, v = 0.5, t = 10, T = 20, p = 0.2, s = 5, \sigma = 0.16, \beta = 1/3. The conclusions are drawn from Figures 1–10.
(1) Effect of parameters m and v on the optimal reinsurance-investment strategy.
Form Figures 1–2, we can see that both the optimal excess-of-loss reinsurance strategy m_i^{\rm{*}} and the optimal investment strategies {\pi ^{\rm{*}}} decrease with m. As we know, m is the ambiguity-aversion coeffcient representing the insurer's attitude toward the uncertainty of the model, which is reflected in the claim process and the risky assets' price process. The larger m is, the more uncertain risk aversion the insurer is, so the insurer tends to adopt a lower reinsurance-investment strategy. Then, the risk from model uncertainty is spread to the reinsurer.
Figures 3–4 shows that m_i^{\rm{*}}, {\pi ^*} and \pi _{\rm{1}}^{\rm{*}} decrease with respect to v. Note that ν represents the insurer's risk aversion coefficient, the larger ν is, the more risk averse the insurer is, thus, as v increases, the insurer would like to reduce the investment amount to avoid risk, meanwhile, he/she prefers to reduce the retention level of reinsurance so as to transfer more underlying risks to a reinsurer.
(2) Effect of parameters t, r, p on the optimal reinsurance strategy.
From Figure 5, we can see that m_i^* increases with respect to the time t. As time passes, the insurer will obtain much wealth from insurance business and investment, so he/she is able to take more insurance risk and prefers to raise the retention level of reinsurance.
Figure 6 demonstrates that m_i^* decreases with the risk-free interest rate r. The larger r is, the more attractive the risk-free asset is, so as r increases, the insurer prefer to invest more wealth in the risk-free asset rather than purchase more reinsurance.
Figure 7 indicates that m_i^* decreases with respect to p. Since a larger p implies greater values of two expected claim numbers, so the insurer prefers to retain a less share of each claim.
(3) Effect of parameters \mu, \sigma, s on the optimal investment strategy.
Figure 8 reports that \mu exerts positive effects on the optimal investment strategies {\pi ^{\rm{*}}} and \pi _1^{\rm{*}}. Note that \mu is the rates of the risky assets' return, as \mu increases, the insurer will gain more from investment. Therefore, the insurer would like to increase the investment.
As shown in Figure 9, {\pi ^{\rm{*}}} and \pi _1^{\rm{*}} decrease with respect to the standard volatility \sigma. As \sigma increases, the volatility of risky asset will fluctuate a little drastically, then the insurer prefer to decrease the investment so as to avoid risks.
We see from Figure 10 that {\pi ^{\rm{*}}} and \pi _1^{\rm{*}} increase with respect to s. A larger s means higher the risky asset's price, hence the insurer will wish to increase the investment in the risky asset.
In addition, in Figure 4 and Figures 8–10, we compare that {\pi ^{\rm{*}}} with model uncertainty and \pi _1^{\rm{*}} without model uncertainty, and we find that \pi _1^{\rm{*}} is larger than {\pi ^{\rm{*}}}. It means that it is necessary to study the robust optimal control problems which can help us make more reasonable decisions.
This paper investigates a robust optimal excess-of-loss reinsurance and investment problem with p-thinning dependent risks for an AAI under CEV model. We aim to maximize the expected exponential utility of AAI's terminal wealth. Applying the stochastic control theory, we obtain the explicit expressions of the optimal excess-of-loss reinsurance and investment strategies and provide some special cases. Finally, we present some numerical examples to illustrate our results.
There are still some problems needed to be investigated in this direction. Firstly, we only deal with the excess-of-loss reinsurance, in fact, one can consider other reinsurance such as stop-loss reinsurance in our model. Secondly, other kinds of dependent risks can be taken into consideration in our problem, in addition, this paper assume the risky asset's price process and wealth process are independent, it is more realistic to consider the correlation between the risky asset's price process and wealth. Thirdly, it is meaningful to discuss other objectives such as the general utility function and mean-variance criterion in our model. Although these problems are challenging, they are meaningful and interesting to be investigated, thus we will focus on these optimal problems so as to enrich our research in the future.
The authors declare that there are no conflicts of interest regarding the publication of this paper.
[1] |
A C, Li Z (2015) Optimal investment and excess-of-loss reinsurance problem with delay for an insurer under Heston's SV model. Insur Math Econ 61: 181-196. doi: 10.1016/j.insmatheco.2015.01.005
![]() |
[2] |
A C, Lai Y, Shao Y (2018) Optimal excess-of-loss reinsurance and investment problem with delay and jump-diffusion risk process under the CEV model. J Comput Appl Math 342: 317-336. doi: 10.1016/j.cam.2018.03.035
![]() |
[3] | Anderson E, Hansen L, Sargent T (1999) Robustness detection and the price of risk. Working Paper. Available from: https://files.nyu.edu/ts43/public/research/.svn/text-base/ahs3.pdf.svn-base. |
[4] |
Asmussen S, Højgaard B, Taksar M (2000) Optimal risk control and dividend distribution policies: example of excess-of-loss reinsurance for an insurance corporation. Financ Stoch 4: 299-324. doi: 10.1007/s007800050075
![]() |
[5] |
Bai L, Zhang H (2008) Dynamic mean-variance problem with constrained risk control for the insurers. Math Meth Oper Res 68: 181-205. doi: 10.1007/s00186-007-0195-4
![]() |
[6] |
Bi J, Liang Z, Xu F (2016) Optimal mean-variance investment and reinsurance problems for the risk model with common shock dependence. Insur Math Econ 70: 245-258. doi: 10.1016/j.insmatheco.2016.06.012
![]() |
[7] |
Branger N, Larsen L (2013) Robust portfolio choice with uncertainty about jump and diffusion risk. J Bank Financ 37: 5036-5047. doi: 10.1016/j.jbankfin.2013.08.023
![]() |
[8] | Gerber H (1979) An Introduction to Mathematical Risk Theory, In: S.S. Huebner Foundation Monograph, Series No. 8, Irwin, Homewood, Ⅲ, 1979. |
[9] |
Gong L, Badescu A, Cheung ECK (2012) Recursive methods for a multidimensional risk process with common shocks. Insur Math Econ 50: 109-120. doi: 10.1016/j.insmatheco.2011.10.007
![]() |
[10] | Grandell J (1991) Aspects of Risk Theory, New York: Springer-Verlag, 1-32. |
[11] |
Gu A, Guo X, Li Z, et al. (2012) Optimal control of excess-of-loss reinsurance and investment for insurers under a CEV model. Insur Math Econ 51: 674-684. doi: 10.1016/j.insmatheco.2012.09.003
![]() |
[12] |
Guan G, Liang Z (2014) Optimal reinsurance and investment strategies for insurer under interest rate and inflation risks. Insur Math Econ 55: 105-115. doi: 10.1016/j.insmatheco.2014.01.007
![]() |
[13] |
Hipp C, Plum M (2000) Optimal investment for insurers. Insur Math Econ 27: 215-228. doi: 10.1016/S0167-6687(00)00049-4
![]() |
[14] |
Huang Y, Yang X, Zhou J (2017) Robust optimal investment and reinsurance problem for a general insurance company under Heston model. Math Meth Oper Res 85: 1-22. doi: 10.1007/s00186-017-0578-0
![]() |
[15] |
Irgens C, Paulsen J (2004) Optimal control of risk exposure, reinsurance and investments for insurance portfolios. Insur Math Econ 35: 21-51. doi: 10.1016/j.insmatheco.2004.04.004
![]() |
[16] | Jeanblanc M, Yor M, Chesney M (2009) Mathematical Methods for Financial Markets, London: Springer-Verlag. |
[17] |
Kraft H (2005) Optimal portfolios and Heston stochastic volatility model: an explicit solution for power utility. Quant Financ 5: 303-313. doi: 10.1080/14697680500149503
![]() |
[18] | Kraft H (2004) Optimal portfolios with stochastic interest rates and defaultable assets, Springer. |
[19] |
Li D, Rong X, Zhao H (2015) Time-consistent reinsurance-investment strategy for an insurer and a reinsurer with mean-variance criterion under the CEV model. J Comput Appl Math 283: 142-162. doi: 10.1016/j.cam.2015.01.038
![]() |
[20] | Li D, Zeng Y, Yang H (2017) Robust optimal excess-of-loss reinsurance and investment strategy for an insurer in a model with jumps. Scand Actuar J, 1-27. |
[21] |
Liang Z, Yuen KC (2016) Optimal dynamic reinsurance with dependent risks: variance premium principle. Scand Actuar J 1: 18-36. doi: 10.1080/03461238.2014.892899
![]() |
[22] |
Liang Z, Yuen KC, Guo J (2011) Optimal proportional reinsurance and investment in a stock market with Ornstein-Uhlenbeck process. Insur Math Econ 49: 207-215. doi: 10.1016/j.insmatheco.2011.04.005
![]() |
[23] | Liang Z, Bi J, Yuen K, et al. (2016) Optimal mean-variance reinsurance and investment in a jump-diffusion financial market with common shock dependence. Math Methods Oper Res 1: 1-27. |
[24] |
Liang X, Wang G (2012) On a reduced form credit risk model with common shock and regime switching. Insur Math Econ 51: 567-575. doi: 10.1016/j.insmatheco.2012.07.010
![]() |
[25] |
Maenhout PJ (2006) Robust portfolio rules and dectection-error probabilities for a mean-reverting risk premium. J Econ Theory 128: 136-163. doi: 10.1016/j.jet.2005.12.012
![]() |
[26] |
Maenhout PJ (2004) Robust portfolio rules and asset pricing. Rev Financ Stud 17: 951-983. doi: 10.1093/rfs/hhh003
![]() |
[27] | Pan J, Hu S, Zhou X (2019) Optimal investment strategy for asset-liability management under the Heston model. Optimization. |
[28] |
Promislow SD, Young VR (2005) Minimizing the probability of ruin when claims follow Brownian motion with drift. N Am Actuar J 9: 110-128. doi: 10.1080/10920277.2005.10596214
![]() |
[29] |
Schmidli H (2002) On minimizing the ruin probability by investment and reinsurance. Ann Appl Probab 12: 890-907. doi: 10.1214/aoap/1031863173
![]() |
[30] | Tian Y, Guo J, Sun Z (2020) Optimal mean-variance reinsurance in a financial market with stochastic rate of return. J Ind Manga Optim. |
[31] |
Yang H, Zhang L (2005) Optimal investment for insurer with jump-diffusion risk process. Insur Math Econ 37: 615-634. doi: 10.1016/j.insmatheco.2005.06.009
![]() |
[32] |
Yang X, Liang Z, Zhang C (2017) Optimal mean-variance reinsurance with delay and multiple classes of dependent risks (in Chinese). Sci Sin Math 47: 723-756. doi: 10.1360/SCM-2016-0388
![]() |
[33] |
Yi B, Li Z, Viens FG, et al. (2013) Robust optimal control for an insurer with reinsurance and investment under Heston's stochastic volatility model. Insur Math Econ 53: 601-614. doi: 10.1016/j.insmatheco.2013.08.011
![]() |
[34] |
Yuen KC, Guo J, Wu X (2002) On a correlated aggregate claim model with Poisson and Erlang risk process. Insur Math Econ 31: 205-214. doi: 10.1016/S0167-6687(02)00150-6
![]() |
[35] |
Yuen KC, Guo J, Wu X (2006) On the first time of ruin in the bivariate compound Poisson model. Insur Math Econ 38: 298-308. doi: 10.1016/j.insmatheco.2005.08.011
![]() |
[36] |
Yuen KC, Liang Z, Zhou M (2015) Optimal proportional reinsurance with common shock dependence. Insur Math Econ 64: 1-13. doi: 10.1016/j.insmatheco.2015.04.009
![]() |
[37] | Zhang Y, Zhao P, Teng X, et al. (2020) Optimal reinsurance and investment strategies for an insurer and a reinsurer under Hestons SV model: HARA utility and Legendre transform. J Ind Manga Optim. |
[38] | Zhang Y, Zhao P (2020) Optimal Reinsurance and Investment problem with dependent risks Based on Legendre transform. J Ind Manga Optim 16: 1457-1479. |
[39] | Zhang Y, Zhao P (2019) Robust optimal excess-of-loss reinsurance and investment problem with delay and dependent risks. Discrete Dyn Nat Soc 2019: 1-21. |
[40] |
Zhao H, Weng C, Shen Y, et al. (2017) Time-consistent investment-reinsurance strategies towards joint interests of the insurer and the reinsurer under CEV models. Sci China Math 60: 317-344. doi: 10.1007/s11425-015-0542-7
![]() |
[41] |
Zeng X, Taksar M (2013) A stochastic volatility model and optimal portfolio selection. Quant Financ 13: 1547-1558. doi: 10.1080/14697688.2012.740568
![]() |
[42] |
Zeng Y, Li Z, Lai Y (2013) Time-consistent investment and reinsurance strategies for mean-variance insurers with jumps. Insur Math Econ 52: 498-507. doi: 10.1016/j.insmatheco.2013.02.007
![]() |
[43] |
Zeng Y, Li D, Gu A (2016) Robust equilibrium reinsurance-investment strategy for a mean-variance insurer in a model with jumps. Insur Math Econ 66: 138-152. doi: 10.1016/j.insmatheco.2015.10.012
![]() |
[44] |
Zheng X, Zhou J, Sun Z (2016) Robust optimal portfolio and proportional reinsurance for an insurer under a CEV model. Insur Math Econ 67: 77-87. doi: 10.1016/j.insmatheco.2015.12.008
![]() |
[45] |
Zhu H, Deng C, Yue S, et al. (2015) Optimal reinsurance and investment problem for an insurer with counterparty risk. Insur Math Econ 61: 242-254. doi: 10.1016/j.insmatheco.2015.01.013
![]() |
1. | Jingyun Sun, Haixiang Yao, Zhongfei Li, Robust optimal investment strategy for a DC pension plan in the market with mispricing and constant elasticity of variance, 2023, 0, 1547-5816, 0, 10.3934/jimo.2023008 | |
2. | Xia Zhao, Mengjie Li, Qinrui Si, Optimal investment-reinsurance strategy with derivatives trading under the joint interests of an insurer and a reinsurer, 2022, 30, 2688-1594, 4619, 10.3934/era.2022234 | |
3. | Xinyue Wei, Cuilian You, Yujie Zhang, European Option Pricing Under Fuzzy CEV Model, 2023, 196, 0022-3239, 415, 10.1007/s10957-022-02108-w | |
4. | Min Zhou, Xiaoqun Liu, Overnight-Intraday Mispricing of Chinese Energy Stocks: A View from Financial Anomalies, 2022, 9, 2296-598X, 10.3389/fenrg.2021.807881 | |
5. | Yalian Zhang, Qihang Zhang, Hexiao Hu, Cao Wang, Xin Guo, Accountability audit of natural resource, government environmental regulation and pollution abatement: An empirical study based on difference-in-differences model, 2023, 410, 09596526, 137205, 10.1016/j.jclepro.2023.137205 | |
6. | Yue Ma, Zhongfei Li, Robust portfolio choice with limited attention, 2023, 31, 2688-1594, 3666, 10.3934/era.2023186 |