t | T | λ1 | λ2 | λ | μ11 | μ12 | μ21 | μ22 | m | β1 | β2 | r | δ | ν | η1 | η2 | τ |
0 | 10 | 0.8 | 0.5 | 0.04 | 2.2 | 2.7 | 1.6 | 2.3 | 0.2 | 0.2 | 0.4 | 0.02 | 0.25 | 0.4 | 0.15 | 0.5 | 0.5 |
This paper investigates the robust optimal reinsurance strategy, which simultaneously takes into account the ambiguity aversion, the correlated claims and the joint interests of an insurer and a reinsurer. The correlated claims mean that future claims are correlated with historical claims, which are measured by an extrapolative bias. The joint interests of the insurer and the reinsurer are reflected by the competition between them. To better reflect competition, we assume that the insurer and the reinsurer are engaged in related insurance business. The insurer is allowed to purchase proportional reinsurance or acquire a new business. Under ambiguity aversion and the criterion of maximizing the expected utility of terminal wealth, we obtain explicit solutions for the robust optimal reinsurance strategy and the corresponding value function by using the stochastic dynamic programming approach. Furthermore, we obtain the optimal reinsurance strategy under four typical cases. A series of numerical experiments were carried out to illustrate how the robust optimal reinsurance strategy varies with model parameters, and the result analyses reveal some interesting phenomena and provide useful guidance for reinsurance in reality.
Citation: Peng Yang. Robust optimal reinsurance strategy with correlated claims and competition[J]. AIMS Mathematics, 2023, 8(7): 15689-15711. doi: 10.3934/math.2023801
[1] | Peng Yang . Robust optimal reinsurance-investment problem for $ n $ competitive and cooperative insurers under ambiguity aversion. AIMS Mathematics, 2023, 8(10): 25131-25163. doi: 10.3934/math.20231283 |
[2] | 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 |
[3] | Ling Chen, Xiang Hu, Mi Chen . Optimal investment and reinsurance for the insurer and reinsurer with the joint exponential utility under the CEV model. AIMS Mathematics, 2023, 8(7): 15383-15410. doi: 10.3934/math.2023786 |
[4] | Wuyuan Jiang, Zechao Miao, Jun Liu . Optimal investment and reinsurance for the insurer and reinsurer with the joint exponential utility. AIMS Mathematics, 2024, 9(12): 35181-35217. doi: 10.3934/math.20241672 |
[5] | Emilio Gómez-Déniz . A quasi-conjugate bivariate prior distribution suitable for studying dependence in reinsurance and non reinsurance models with and without a layer. AIMS Mathematics, 2025, 10(5): 12055-12078. doi: 10.3934/math.2025546 |
[6] | Ying Fang, Guo Cheng, Zhongfeng Qu . Optimal reinsurance for both an insurer and a reinsurer under general premium principles. AIMS Mathematics, 2020, 5(4): 3231-3255. doi: 10.3934/math.2020208 |
[7] | Fuyun Sun, Yuelei Li . On the improved thinning risk model under a periodic dividend barrier strategy. AIMS Mathematics, 2021, 6(12): 13448-13463. doi: 10.3934/math.2021779 |
[8] | Wen Chao . Pricing catastrophe reinsurance under the standard deviation premium principle. AIMS Mathematics, 2022, 7(3): 4472-4484. doi: 10.3934/math.2022249 |
[9] | Riza Andrian Ibrahim, Sukono, Herlina Napitupulu, Rose Irnawaty Ibrahim . Modeling earthquake bond prices with correlated dual trigger indices and the approximate solution using the Monte Carlo algorithm. AIMS Mathematics, 2025, 10(2): 2223-2253. doi: 10.3934/math.2025103 |
[10] | Min Wang, Jiao Teng, Lei Wang, Junmei Wu . Application of ADMM to robust model predictive control problems for the turbofan aero-engine with external disturbances. AIMS Mathematics, 2022, 7(6): 10759-10777. doi: 10.3934/math.2022601 |
This paper investigates the robust optimal reinsurance strategy, which simultaneously takes into account the ambiguity aversion, the correlated claims and the joint interests of an insurer and a reinsurer. The correlated claims mean that future claims are correlated with historical claims, which are measured by an extrapolative bias. The joint interests of the insurer and the reinsurer are reflected by the competition between them. To better reflect competition, we assume that the insurer and the reinsurer are engaged in related insurance business. The insurer is allowed to purchase proportional reinsurance or acquire a new business. Under ambiguity aversion and the criterion of maximizing the expected utility of terminal wealth, we obtain explicit solutions for the robust optimal reinsurance strategy and the corresponding value function by using the stochastic dynamic programming approach. Furthermore, we obtain the optimal reinsurance strategy under four typical cases. A series of numerical experiments were carried out to illustrate how the robust optimal reinsurance strategy varies with model parameters, and the result analyses reveal some interesting phenomena and provide useful guidance for reinsurance in reality.
Nowadays, with the rapid development of economy and society, more and more individuals and institutions are beginning to buy insurance. This leads to a rapid increase in the premium income of the insurer; meanwhile, the claim risks faced by the insurer also increase. However, it is usually difficult to undertake claim risks for an insurer only depending upon his premium. Therefore, the insurer is likely to purchase reinsurance in order to transfer claim risks. The technique of stochastic control theory and the stochastic dynamic programming approach are widely used to cope with the optimal reinsurance problem. Different optimization criteria have been proposed in the literature, e.g., minimizing the probability of ruin criterion (see [1,2,3]), maximizing the expected utility criterion (see [4,5,6]) and maximizing or minimizing mean-variance (MV) criterion (see [7,8,9]).
The above studies only considered the interest of the insurer, and did not pay attention to the reinsurer's interest. In fact, the reinsurance problem involves the interests of both the insurer and the reinsurer. An insurer pays a reinsurance premium to a reinsurer in order to obtain compensation from the reinsurer in case a claim occurs, while a reinsurer is willing to bear the insurer's claims in exchange for the reinsurance premium. The authors of [10] considered the joint interests of an insurer and a reinsurer by maximizing the weighted sum of the insurer's and the reinsurer's MV utilities. The authors of [11] studied the optimal reinsurance problem that minimizes the convex combination of the value-at-risk measures of the insurer's loss and the reinsurer's loss under two types of constraints. The authors of [12] considered the joint interests of an insurer and a reinsurer by maximizing the product of the insurer's and the reinsurer's utilities. The authors of [13] and [14] considered the joint interests of an insurer and a reinsurer by maximizing the expected utility of the weighted sum of the insurer's wealth and the reinsurer's wealth. And the difference is that [13] considered the exponential utility, while [14] considered the hyberbolic absolute risk aversion utility. However, none of these studies considered competition.
With the development of economy and society, competition exists widely among investors, managers, insurers and reinsurers. However, it is a difficult issue to quantify competition. The authors of [15] demonstrated that relative performance can be used to quantify competition, which has recently been used to describe the competition between two insurers. The authors of [16] studied the competition between two insurers, where all of the surplus process parameters and asset price parameters were modeled by a common random factor and Markov chain. The authors of [17] investigated the competition between two insurers in the longevity risk transfer market. The authors of [18] studied the competition between two MV insurers, where the stock's price process is the constant elasticity of variance model. The authors of [19] considered the competition between two insurers under different classes of premium principles, and the price of the risky asset follows an Ornstein-Uhlenbeck process. In practice, there are often many, say n>2, insurers whose reinsurance strategies are related. It is more reasonable to study the reinsurance problem among n insurers. The authors of [20] first considered the competition among n insurers.
At present, few scholars study the competition between an insurer and a reinsurer. The authors of [21] recently studied the competition between an insurer and a reinsurer. Generally speaking, the insurer competes with the reinsurer in terms of insurance business or reinsurance business. As what was explained in [21], a Swiss reinsurance company engaged the property insurance business in China, and Ping An insurance company of China engaged the reinsurance business. Nevertheless, the authors of [21] did not consider this form of competition. This paper intends to extend the competition form of the results in [21].
In practice, insurers usually renew the terms of insurance contract based on policy-holders' past claims. The authors of [22,23,24] have shown that future claims are correlated with historical claims. Regarding discrete time models, the authors of [25] empirically studied the effect of firms' past loss on premium, and they demonstrated that the premium is paid according to the formula (1+ϖ)(θAL+(1−θ)EL). Here θ∈(0,1) is the weight factor, ϖ>0 is the safety factor and AL and EL respectively represent the firm's past yearly loss and the yearly expected loss. However, it is difficult to quantify the correlation between future claims and historical claims for a continuous time model. The authors of [26,27,28] established the correlation between future claims and historical claims by using an extrapolative bias model, which was pioneered in [29]. Furthermore, the discrete premium in [25] above-mentioned is extended to the continuous case. However, these studies did not consider the interest of the reinsurer. In practice, people expect to know the insurer's reinsurance decision when taking into account the joint interests of the insurer and the reinsurer.
Insurers often face an uncertain environment when making reinsurance decisions. It is well accepted that insurers are generally ambiguity-averse. This is because they must have enough funds to pay the claim of the policy-holders. The authors of [30] first proposed a robust control approach under the continuous-time framework to study ambiguity aversion. Nowadays, there are many studies on robust reinsurance and/or investment under ambiguity aversion. The authors of [31] investigated the robust optimal excess-of-loss reinsurance and investment problem with jumps. The authors of [32] considered the robust optimal dynamic reinsurance problem under the mean-RVaR premium principle. The authors of [33] investigated the robust reinsurance under a learning framework. The authors of [34] studied the robust optimal reinsurance problem for an ambiguity-averse insurer (AAI), where the insurer cannot obtain the perfect information of the claim. To the best of our knowledge, only the authors of [28] studied the robust optimal reinsurance problem under correlated claims; however, they did not consider the interest of the reinsurer.
In view of the aforementioned state-of-the-art work, we will study the robust optimal reinsurance problem and simultaneously take into account the ambiguity aversion, the joint interests of an insurer and a reinsurer and the correlated claims. The insurer and the reinsurer can also engage in insurance business, and their insurance businesses are correlated. The insurer can reduce his claim risk by purchasing proportional reinsurance or acquiring a new business, which includes playing the role of a reinsurer. The joint interests of the insurer and the reinsurer are reflected by their competition. We construct the competition model through a relative wealth process. The interests of the insurer and the reinsurer are reflected through their respective wealth in the relative wealth process. That is, the relative wealth process includes the joint interests of the insurer and the reinsurer. Furthermore, the insurer determines the optimal reinsurance strategy under the relative wealth process. Considering the uncertainty of model, we assume that the insurer is ambiguity-averse. By maximizing the expected exponential utility of terminal wealth, we obtain explicit solutions for the robust optimal reinsurance strategy and the corresponding value function under the worst-case market scenario. To analyze the effects of key model features on the robust optimal reinsurance strategy, we consider four special cases, i.e., the no-competition case, the uncorrelated insurance businesses case, the no-ambiguity-aversion case and the uncorrelated claims case. Finally, we conduct numerical experiments to examine the effects of key model features on the robust optimal reinsurance strategy.
Compared to the existing literature, the main contributions of this paper are as follows:
● We simultaneously take into account the ambiguity aversion, the joint interests of an insurer and a reinsurer, and the correlated claims.
● The insurer can reduce his claim risk by purchasing proportional reinsurance or acquiring a new business, and the reinsurer can also engage in insurance business; furthermore, we quantify the competition between the insurer and the reinsurer.
● We obtain the explicit robust optimal reinsurance strategy, which can guide the actual reinsurance activities more effectively.
● We systematically analyze the influences of time, the correlated claims, the dependence of insurance business, the competition degree and the ambiguity aversion coefficient on the robust optimal reinsurance strategy through numerical experiments. Furthermore, the result analyses reveal some interesting phenomena, extending the result of [28], and provide useful guidance for reinsurance in reality.
The rest of the paper is organized as follows. The model settings and assumptions are given in Section 2. In Section 3, we introduce the robust reinsurance problem with correlated claims and competition. In Section 4, we obtain explicit solutions for the robust optimal reinsurance strategy and the corresponding value function, and we also provide the verification theorem. Section 5 illustrates our theoretical results through numerical experiments. The final section summarizes the paper.
Throughout this paper, we assume that all random variables and processes are defined on a filtered probability space (Ω,F,{Ft}t∈[0,T],P) satisfying the usual conditions, i.e., Ft is right continuous and P-complete; Ft stands for the information available until time t; T is the terminal time of reinsurance; P is the reference measure.
Without losing generality, suppose that there are two kinds of insurance businesses in the insurance market, which are named as insurance businesses 1 and 2. The insurance business i's, i=1,2, cumulative claims in the time interval [0,t] is denoted by
Li(t)=Gi(t)∑k=1Yik. |
Here Yik is the size of the insurance business i's kth claim, and {Yik,k=1,2,…} forms a sequence of independent and nonnegative random variables following a common distribution. The common random variable of {Yik,k=1,2,…} is denoted as Yi. Yi has a finite mean μi1=E(Yi) and second-order moment μi2=E[(Yi)2]. Ni(t) and N(t) are mutually independent Poisson processes with intensities of λi≥0 and λ≥0, respectively. N(t) is the number of claims caused by the common shock of the two insurance businesses, which reflects the interdependence of the two insurance businesses. Ni(t) is the number of claims independent of the common shock. Gi(t)=N(t)+Ni(t) is the counting process, representing the number of claims of business i up to time t. According to the additivity of the Poisson process, Gi(t) is a Poisson process with an intensity of λi+λ.
It is known from [35] that the aggregate claims process ∑Gi(t)k=1Yik can be approximated by the following Brownian motion with drift
Gi(t)∑k=1Yik≈(λ+λi)μi1t−√(λ+λi)μi2Wi(t), | (2.1) |
where {Wi(t),t∈[0,T]} is the standard Brownian motion. The two Brownian motions W1(t) and W2(t) are correlated with the correlation coefficient of
ρ:=λμ11μ21√(λ+λ1)(λ+λ2)μ12μ22. |
For presentation convenience, we define notations pi=(λ+λi)μi1 and qi=√(λ+λi)μi2, i=1,2.
In reality, future claims are usually correlated with historical claims. Therefore, the insurers always take the occurred claims, especially the recent claims into consideration when renewing the insurance contracts. Considering this and inspired by [29], especially by [26] and [28], we define the exponential weighted average of historical losses as follows
ν(t)=∫t0δe−δ(t−s)dL1(s−ds), 0<δ<1, | (2.2) |
which is an exponential weighted average of historical claims. ν(0)=ν0=0, δ is the extrapolation intensity and dL1(s−ds)=L1(s)−L1(s−ds) represents the claim amount occurred from time s−ds to time s. Equation (2.2) means that the insurer uses the exponential weighted average of historical claims to forecast the expected future claim. By using model (2.2), we quantify the correlation between future claims and historical claims.
To ensure that the weighted average of historical claims is finite, we introduce the following assumption.
Assumption 2.1. We assume that
|ν(t)|≤M, for ∀t∈[0,T],P-a.s., |
where M is a positive constant.
Substituting (2.1) into (2.2) and according to the derivative of the definite integral with a variable upper limit, we derive the differential form of ν(t) as
dν(t)=δp1e−δt−δq1dW1(t). | (2.3) |
Suppose that there exists one insurer in the insurance market who is engaged in insurance business 1. His premium rate c1 can be determined according to the exponential weighted average of historical losses and the expected future claims. By the exponential weighted average of historical losses, we can easily obtain the total weight of the past claims is ∫t0δe−δ(t−s)ds=1−e−δt, which is always less than one. Hence, a weight e−δt is naturally given to the expectation of the future claims. Then, the premium rate c1 is given by c1=(1+η1)(v(t)+p1e−δt), where η1>0 is the safety loading of the insurer.
To transfer claim risks, the insurer adopts reinsurance. For each t∈[0,T], the reinsurance level is associated with the parameter a(t), where a(t)∈[0,+∞) is the retention level of the insurer. 0≤a(t)≤1 corresponds to a proportional reinsurance. a(t)>1 corresponds to a new business requirement; for example, the insurer plays the role of a reinsurer. By acquiring a new business, the insurer will increase his income to hedge the claim risks. For more information about this reinsurance pattern, please see [36]. For this reinsurance pattern, the insurer only needs to pay a(t) of each claim, and the reinsurer pays the rest (1−a(t)). Let (1+η2)(1−a(t))(v(t)+p1e−δt) be the reinsurance premium rate, where η2>0 is the safety loading of the reinsurer. To exclude the insurer's arbitrage behavior, we require η2>η1. The insurer can invest his surplus in a risk-free money account with a constant interest rate r>0. With the reinsurance and money account being incorporated, the wealth process X1(t) of the insurer satisfies the following stochastic differential equation
dX1(t)=[rX1(t)+c1−(1+η2)(1−a(t))(v(t)+p1e−δt)]dt−a(t)[(v(t)+p1e−δt)dt−q1dW1(t)]=[rX1(t)+c1−(1+η2)(v(t)+p1e−δt)+a(t)η2(v(t)+p1e−δt)]dt+a(t)q1dW1(t). | (2.4) |
Considering the insurance business in practice, we assume that the reinsurer not only accepts the reinsurance business of the insurer, but also engages in insurance business 2. The premium rate received by the reinsurer from insurance business 2 is c2. Similar to that of the insurer, the reinsurer can also invest his surplus in the risk-free money account with the constant interest rate r>0. With the reinsurance and money account being incorporated, the wealth process X2(t) of the reinsurer satisfies the following stochastic differential equation
dX2(t)=[rX2(t)+c2+(1+η2)(1−a(t))(v(t)+p1e−δt)]dt−(1−a(t))[(v(t)+p1e−δt)dt−q1dW1(t)]−[p2dt−q2dW2(t)]=[rX2(t)+c2−p2+(1−a(t))η2(v(t)+p1e−δt)]dt+(1−a(t))q1dW1(t)+q2dW2(t). | (2.5) |
From the above discussion, it is obvious that the interdependence of the insurer's insurance business and the reinsurer's insurance business is due to the common Poisson process N(t). When signing a reinsurance contract, it is clear that both the insurer and reinsurer are concerned about the reinsurance premium (1+η2)(1−a(t))(v(t)+p1e−δt). The insurer wants (1+η2)(1−a(t))(v(t)+p1e−δt) to be as small as possible, while the reinsurer wants (1+η2)(1−a(t))(v(t)+p1e−δt) to be as large as possible. In addition, the insurer and the reinsurer are engaged in dependent insurance business. Therefore, competition between the insurer and the reinsurer exists both in insurance business and in the formulation of the reinsurance premium. In insurance practice, after signing a reinsurance contract, the insurer not only cares about his own wealth X1(t), but he also cares about the wealth X2(t) obtained by the reinsurer from the reinsurance contract. In practice, since the reinsurer may also sign reinsurance contracts with other insurers, the wealth of the reinsurer is generally greater than X2(t). However, for this insurer, he may most care about the wealth X2(t) of the reinsurer. This is because the decision of this insurer will only affect X2(t), and will not affect other wealth assets of the reinsurer. If X1(t)>X2(t), the insurer will continue to perform the reinsurance contract. If X1(t)<X2(t), the insurer may consider amending the reinsurance contract, i.e., increasing the proportion of claims paid by himself, thus reducing the reinsurer's wealth obtained from the reinsurance contract. Therefore, after signing the reinsurance contract, the insurer will always simultaneously pay attention to X1(t) and X2(t). To this end, similar to [21], we introduce the following relative performance model
(1−τ)X1(t)+τ(X1(t)−X2(t)). |
Here τ∈[0,1] captures the competition intensity between the insurer and the reinsurer, and measures the insurer's sensitivities to the performance of the other party. A large τ means that the insurer puts more weight on the performance of the reinsurer and cares more about his relative wealth increase. When τ=0, the insurer only cares about his own wealth.
Let
Xa(t):=(1−τ)X1(t)+τ(X1(t)−X2(t)) |
be the relative wealth process of the insurer. Then Xa(t) can be described as
dXa(t)=[rXa(t)+c1−τc2+τp2−(1+η2)(v(t)+p1e−δt)+η2(v(t)+p1e−δt)((1+τ)a(t)−τ)]dt+[(1+τ)a(t)−τ]q1dW1(t)−τq2dW2(t). | (2.6) |
We quantify the competition between the insurer and the reinsurer through the relative wealth. In the relative wealth process (2.6), X1(t) embodies the interest of the insurer and X2(t) embodies the interest of the reinsurer. In what follows, the insurer will determine the optimal reinsurance strategy based on the relative wealth process (2.6). That is, the insurer determines the optimal reinsurance strategy by simultaneously takeing into account the joint interests of the insurer and the reinsurer.
The above-mentioned model corresponds to a traditional framework. It is assumed that the insurer is ambiguity-neutral, and that he has complete confidence in the model (2.6) provided by the probability distribution P. However, in many situations, the insurer cannot know exactly the true P; thus, any particular probability distribution used to describe the model would lead to potential model misspecification. For this reason, we shall incorporate the probability distribution uncertainty into the reinsurance problem for an AAI. We assume that the AAI uses the model (2.6) as his reference model for the wealth process. Nevertheless, he is sceptical about this reference model and takes alternative models into consideration to determine a robust optimal strategy. Parallel to [30], the alternative models are defined by a class of probability distributions which are equivalent to P:
Q:={Q|Q∼P}. |
In what follows, we transform model (2.6) from probability distribution P to probability distribution Q. First, we introduce a process {θ(t)=(θ1(t),θ2(t))|t∈[0,T]} satisfying the following:
● θ1(t) and θ2(t) are {Ft}t≥0-measurable, for each t∈[0,T];
● θ1(t) and θ2(t) satisfy the following Novikov's condition
E[exp{12∫T0[(θ1(t))2+(θ2(t))2]dt}]<+∞. |
We denote the space of all such processes θ(t) by Θ.
To change the probability distribution from P to Q, we define for each θ(t)∈Θ a real-valued process {Λθ(t)|t∈[0,T]} on (Ω,F,{Ft}t≥0,P) by
Λθ(t)=exp{−∫t0θ1(s)dW1(s)−∫t0(θ1(s))2ds−∫t0θ2(s)dW2(s)−∫t0(θ2(s))2ds}. |
Under the Novikov's condition, similar to that in [37], we can prove that Λθ(t) is a P-martingale. Hence, E[Λθ(t)]=1. For each θ(t)∈Θ, we apply the Radon-Nikodym derivative to define a new probability distribution Qθ that is absolutely continuous with respect to P on FT as
dQθdP|FT:=Λθ(T), |
and the set of all Qθ is denoted by Q.
According to Girsanov's theorem, under the alternative measure Qθ, the stochastic processes Wθ1(t) and Wθ2(t) are standard Brownian motions, where
dWθ1(t)=dW1(t)+θ1(t)dt, dWθ2(t)=dW2(t)+θ2(t)dt. |
Since Brownian motions W1(t) and W2(t) are correlated with the correlation coefficient of ρ, Brownian motions Wθ1(t) and Wθ2(t) are correlated with the correlation coefficient of ρ. Furthermore, the dynamics of the wealth process Xa(t) in (2.6) under Qθ becomes
dXa(t)=[rXa(t)+c1−τc2+τp2−(1+η2)(ν(t)+p1e−δt)+η2(ν(t)+p1e−δt)((1+τ)a(t)−τ)−θ1((1+τ)a(t)−τ)q1+θ2τq2]dt+[(1+τ)a(t)−τ]q1dWθ1(t)−τq2dWθ2(t), | (2.7) |
with Xa(0)=x0. Correspondingly, the measure of extrapolative bias ν(t) in (2.3) becomes
dν(t)=[δp1e−δt+θ1(t)δq1]ds−δq1dWθ1(t). | (2.8) |
In order to introduce the robust reinsurance problem, we define the admissible strategy under the probability distribution Qθ as follows.
Definition 2.1. (Admissible Strategy). For any fixed t∈T, a reinsurance strategy a(t) is said to be admissible if it satisfies the following conditions:
(i) a(t) is progressively measurable with respect to {Ft}t≥0 for each t∈[0,T];
(ii) ∀ t∈[0,T], a(t)∈[0,+∞);
(iii) Et,x,ν{∫T0a2(t)dt}<+∞, where Et,x,ν[⋅]=E[⋅|Xa(t)=x,ν(t)=ν].
The set of all admissible reinsurance strategies is denoted by A.
In this section, we have proposed a new reinsurance model with correlated claims and competition. Compared with [28], our model has the following characteristics.
● The reinsurer can engage in insurance business.
● We quantify the competition between the insurer and the reinsurer.
● We simultaneously take into account the ambiguity aversion, the joint interests of the insurer and the reinsurer, and the correlated claims.
Therefore, we extend the model of [28].
In this section, we construct the robust optimal reinsurance problem. We first introduce the following penalty function:
Φ(t,Xa(t),ν(t),θ(t))=(θ1(t))22ϕ1(t,Xa(t),ν(t))+(θ2(t))22ϕ2(t,Xa(t),ν(t)), | (3.1) |
where ϕ1(t,Xa(t),ν(t)) and ϕ2(t,Xa(t),ν(t)) are nonnegative and capture the AAI's ambiguity aversions.
This penalty function can effectively reflect the penalty due to the deviation from the reference model, i.e., the model under the probability distribution P. The first and second terms describe the penalty caused by W1(t) and W2(t), respectively. The larger the ϕ1(t,Xa(t),ν(t)) and ϕ2(t,Xa(t),ν(t)), the less the insurer will be penalized, the less faith in the reference model had by the insurer, i.e., the insurer is more likely to consider the alternative model, which is the model under the probability distribution Qθ.
On the basis of the penalty function Φ(t,Xa(t),ν(t),θ(t)), we formulate a robust optimal reinsurance problem which chooses the optimal strategy in the worst-case scenario as follows:
supa∈Ainfθ∈ΘEt,x,ν{∫TtΦ(s,Xa(s),ν(s),θ(s))ds+U(Xa(T))}, | (3.2) |
where the expectation is with respect to the alternative probability distribution Qθ, and U(⋅) is the utility function which is assumed to be continuous, increasing and strictly concave.
To solve problem (3.2), we define the optimal value function as
V(t,x,ν)=supa∈Ainfθ∈ΘEt,x,ν{∫TtΦ(s,Xa(s),ν(s),θ(s))ds+U(Xa(T))}. | (3.3) |
Let C1,2,2([0,T]×R×R+) denote the space of φ(t,x,ν) such that φ(t,x,ν) and its derivatives φt(t,x,ν), φx(t,x,ν), φxx(t,x,ν), φν(t,x,ν) and φνν(t,x,ν) are continuous on [0,T]×R×R+. For any function φ(t,x,ν)∈C1,2,2([0,T]×R×R+) and any fixed a(t)∈A, the usual infinitesimal generator Bθ,a for the wealth process Xa(t) in (2.7) is defined as
Bθ,aφ(t,x,ν)=φt(t,x,ν)+[rx+c1−τc2+τp2−(1+η2)(ν+p1e−δt)+η2(ν(t)+p1e−δt)((1+τ)a(t)−τ)−θ1((1+τ)a(t)−τ)q1+θ2τq2]φx(t,x,ν)+12[((1+τ)a(t)−τ)2q21+τ2q22−2((1+τ)a(t)−τ)τρq1q2]φxx(t,x,ν)+[δp1e−δt+θ1(t)δq1]φν(t,x,ν)+12δ2q21φνν(t,x,ν)+[−δq1((1+τ)a(t)−τ)q1+2ρδq1q2τ]φxν(t,x,ν). | (3.4) |
According to the optimality principle of dynamic programming, we can derive the Hamilton-Jacobi-Bellman-Isaacs equation for problem (3.3) as
supa∈Ainfθ∈Θ{Bθ,aV(t,x,ν)+Φ(t,x,ν,θ)}=0, | (3.5) |
with the boundary condition being V(T,x,ν)=U(x).
The following proposition is essential in solving the associated stochastic control problem.
Proposition 3.1. Suppose that there exist a function H(t,x,ν)∈C1,2,2([0,T]×R×R+) and a control strategy (θ∗,a∗)∈(Θ,A) such that
(1) Bθ,a∗H(t,x,ν)+Φ(t,x,ν,θ)≥0 for all θ∈Θ;
(2) Bθ∗,aH(t,x,ν)+Φ(t,x,ν,θ∗)≤0 for all a∈A;
(3) Bθ∗,a∗H(t,x,ν)+Φ(t,x,ν,θ∗)=0;
(4) for all (θ,a)∈(Θ,A), limt→T−H(t,Xa(t),ν(t))=U(Xa(T));
(5) {H(ϵ,Xa(ϵ),ν(ϵ))}ϵ∈ Ξ and {Φ(ϵ,Xa(ϵ),ν(ϵ),θ(ϵ))}ϵ∈ Ξ are uniformly integrable, where Ξ denotes the set of stopping times ϵ≤Ξ.
Then H(t,x,ν)=V(t,x,ν) and (θ∗,a∗) is an optimal control strategy.
The proof of this proposition is similar to that of Theorem 3.2 in [38], so we omit it here.
In the next section, we shall apply Proposition 3.1 to verify that the candidate reinsurance strategy and the corresponding value function are indeed optimal.
In this section, the aim is to derive the robust optimal reinsurance strategy for problem (3.3). We assume that the AAI has an exponential utility function, i.e.,
U(x)=−1me−mx, | (4.1) |
where m>0 is the absolute risk aversion parameter. Inspired by [39], we define the ambiguity preference functions as
ϕi(t,x,ν)=−βimV(t,x,ν), i=1,2. | (4.2) |
Here βi,i=1,2, are positive constants and stand for the ambiguity aversion parameters.
The following theorem provides the candidate robust optimal reinsurance strategy and the corresponding value function for problem (3.3).
Theorem 4.1. For problem (3.3) with the exponential utility function (4.1) and the ambiguity aversion coefficients specified in (4.2), the candidate robust optimal reinsurance strategy is given by
a∗(t)=11+τ[τ+η2(ν+p1e−δt)+(m+β1)δq21(2A(t)+B(t))+mq1q2τρer(T−t)(m+β1)q21er(T−t)]∨0, | (4.3) |
and the candidate worst-case measure is determined by
{θ∗1(t)=−β1q1[δ(2A(t)+B(t))−((1+τ)a∗(t)−τ)er(T−t)],θ∗2(t)=−β2τq2er(T−t), | (4.4) |
where A(t) and B(t) are defined by (4.7) and (4.8), respectively. For a∗(t)>0, the candidate optimal value function is given by
H(t,x,ν)=−1me−m[xer(T−t)+f(t,ν)], | (4.5) |
where
f(t,ν)=A(t)ν2+B(t)ν+D(t), | (4.6) |
and
A(t)=η24λ1μ12δ(m+β1)[e2δη2(T−t)−1], | (4.7) |
B(t)=eη2δ(T−t)∫Tt{2A(s)[δp1e−δs−2ρδq1q2τmer(T−s)+δmq1q2τρer(T−s)+η2δp1e−δs]−(1+η2)er(T−s)+η2p1e−δs(m+β1)q21+η2mq2τρer(T−s)(m+β1)q1}e−η2δ(T−s)ds, | (4.8) |
D(t)=∫Tt{[δp1e−δs−2mρδq1q2τer(T−s)+η2p1δe−δs+δmq1q2τρer(T−s)]B(s)+[c1−τc2−(1+η2)p1e−δs+τp2]er(T−s)−m+β22τ2q22e2r(T−s)+12q21(m+β1)×(η22p21e−2δs+m2q21q22τ2ρ2e2r(T−s)+2η2mq1q2τρp1e−δser(T−s))}ds. | (4.9) |
Proof. According to boundary condition V(T,x,ν)=U(x), we conjecture that the candidate value function has the following structure
H(t,x,ν)=−1me−m[xer(T−t)+f(t,ν)], | (4.10) |
where f(t,ν) is a function to be determined. It satisfies the boundary condition f(T,ν)=0. The partial derivatives of H(t,x,ν) are then given by
{Ht(t,x,ν)=[rxer(T−t)−ft(t,ν)]mH(t,x,ν),Hx(t,x,ν)=−mer(T−t)H(t,x,ν),Hν(t,x,ν)=−mfνH(t,x,ν),Hxx(t,x,ν)=m2e2r(T−t)H(t,x,ν),Hxν(t,x,ν)=m2fνer(T−t)H(t,x,ν),Hνν(t,x,ν)=m2(fν)2H(t,x,ν)−mfννH(t,x,ν). | (4.11) |
Substituting (4.11) into (3.5), we obtain after some simplifications that
supa∈Ainfθ∈Θ{θ21(t)2β1+θ22(t)2β2+ft+[rx+c1−τc2+τp2−(1+η2)(ν+p1e−δt)+η2(ν(t)+p1e−δt)((1+τ)a(t)−τ)−θ1((1+τ)a(t)−τ)q1+θ2τq2]er(T−t)−m2[((1+τ)a(t)−τ)2q21+τ2q22−2((1+τ)a(t)−τ)τρq1q2]e2r(T−t)+[δp1e−δt+θ1(t)δq1]fν+12δ2q21(−mf2ν+fνν)−[−δq1((1+τ)a(t)−τ)q1+2ρδq1q2τ]mfνer(T−t)}=0, | (4.12) |
where f is a short notation for f(t,ν).
According to the first-order optimality conditions, the processes θ1(t) and θ2(t) solving the inner infimum problem in (4.12) are given by
{θ∗1(t)=−β1q1[δfν−((1+τ)a(t)−τ)er(T−t)],θ∗2(t)=−β2τq2er(T−t). | (4.13) |
Inserting (4.13) into (4.12), we obtain
supa∈A{−q21β12[δfν−((1+τ)a(t)−τ)er(T−t)]2−β22τ2q22e2r(T−t)+ft+[c1−τc2+τp2−(1+η2)(ν+p1e−δt)+η2(ν(t)+p1e−δt)((1+τ)a(t)−τ)]er(T−t)−m2[((1+τ)a(t)−τ)2q21+τ2q22−2((1+τ)a(t)−τ)τρq1q2]e2r(T−t)+δp1e−δtfν+12δ2q21(−mf2ν+fνν)−[−δq1((1+τ)a(t)−τ)q1+2ρδq1q2τ]mfνer(T−t)}=0. | (4.14) |
According to the first-order optimality condition for a(t) with respect to the superemum problem in (4.14), we have
ˆa(t)=11+τ[τ+η2(ν+p1e−δt)+(m+β1)δq21fν+mq1q2τρer(T−t)(m+β1)q21er(T−t)]. | (4.15) |
Substituting (4.15) into (4.14), we have
ft+[δp1e−δt−2ρδq1q2τmer(T−t)+η2δ(ν+p1e−δt)+δmq1q2τρer(T−t)]fν+12δ2q21fνν+ν2η222(m+β1)q21+ν[η2p1e−δt(m+β)q21+η2mq2τρer(T−t)(m+β)q1−(1+η2)er(T−t)]+[c1−τc2−(1+η2)p1e−δt+τp2]er(T−t)−m+β22τ2q22e2r(T−t)+12q21(m+β1)×(η22p21e−2δt+m2q21q22τ2ρ2e2r(T−t)+2η2mq1q2τρp1e−δter(T−t))=0,f(T,ν)=0. | (4.16) |
We assume that the solution of (4.16) is
f(t,ν)=A(t)ν2+B(t)ν+D(t). |
Then, we obtain
ft=A′(t)ν2+B′(t)ν+D′(t), fν=2A(t)ν+B(t), fνν=2A(t). | (4.17) |
Substituting (4.17) into (4.16), we have after some simplifications that
ν2{A′(t)+2η2δA(t)+η222(m+β1)q21}+ν{B′(t)+η2δB(t)+2A(s)[δp1e−δt−2ρδq1q2τmer(T−t)+δmq1q2τρer(T−t)+η2δp1e−δt]−(1+η2)er(T−t)+η2p1e−δt(m+β1)q21+η2mq2τρer(T−t)(m+β1)q1}+{D′(t)+[δp1e−δt−2mρδq1q2τer(T−t)+η2p1δe−δt+δmq1q2τρer(T−t)]B(t)+[c1−τc2−(1+η2)p1e−δt+τp2]er(T−t)−m+β22τ2q22e2r(T−t)+12q21(m+β1)×(η22p21e−2δt+m2q21q22τ2ρ2e2r(T−t)+2η2mq1q2τρp1e−δter(T−t))}. | (4.18) |
To solve (4.18), we separate the terms with ν2, ν and the constant term, respectively. It can then be seen that the following system of ordinary differential equations should hold
A′(t)+2η2δA(t)+η222(m+β1)q21, A(T)=0, | (4.19) |
B′(t)+η2δB(t)+2A(s)[δp1e−δt−2ρδq1q2τmer(T−t)+δmq1q2τρer(T−t)+η2δp1e−δt]−(1+η2)er(T−t)+η2p1e−δt(m+β1)q21+η2mq2τρer(T−t)(m+β1)q1, B(T)=0, | (4.20) |
D′(t)+[δp1e−δt−2mρδq1q2τer(T−t)+η2p1δe−δt+δmq1q2τρer(T−t)]B(t)+[c1−τc2−(1+η2)p1e−δt+τp2]er(T−t)−m+β22τ2q22e2r(T−t)+12q21(m+β1)×(η22p21e−2δt+m2q21q22τ2ρ2e2r(T−t)+2η2mq1q2τρp1e−δter(T−t))=0, D(T)=0. | (4.21) |
Solving these ordinary differential equations gives us A(t), B(t) and D(t) as those shown in (4.7), (4.8) and (4.9).
Combining (4.6) and (4.10), we can obtain the candidate optimal value function in (4.5).
Inserting (4.6) into (4.15), we have
ˆa(t)=11+τ[τ+η2(ν+p1e−δt)+(m+β1)δq21(2A(t)ν+B(t))+mq1q2τρer(T−t)(m+β1)q21er(T−t)]. | (4.22) |
If the reinsurance strategy ˆa(t)>0, then the candidate robust optimal reinsurance strategy is a∗(t)=ˆa(t), and the corresponding optimal value function is given by (4.5). If ˆa(t)≤0, the supa{⋯} in (4.14) is achieved at the point 0 since the function in the interior of supa{⋯} is a decreasing function with respect to a(t) in the interval [0,+∞). This implies that the candidate robust optimal reinsurance strategy must be a∗(t)=0. Then, inserting a∗(t)=0 into (4.14), we can obtain the optimal value function as we did for the case a∗(t)=ˆa(t). This case means that the insurer transfers all of the claims to the reinsurer, which hardly occurs in practice. Therefore, we do not calculate the value function in this case.
In what follows, we shall apply Proposition 3.1 to verify that the candidate robust optimal reinsurance strategy and the corresponding candidate value function given in Theorem 4.1 are indeed optimal.
Theorem 4.2. (Verification theorem). The candidate robust optimal strategies a∗(t) and θ∗(t)=(θ∗1(t),θ∗2(t)) given by (4.3) and (4.4) respectively are indeed robust optimal strategies, and the candidate value function H(t,x,ν) given by (4.5) is the optimal value function V(t,x,ν) defined by (3.3).
Proof. From Theorem 4.1, it is easy to see that the conditions (2)–(4) in Proposition 3.1 hold for H(t,x,ν). Because H(t,x,ν) is a smooth function, it is generally assumed in the literature (for example, [39] and [37]) that the optimization problem
infθ∈Θsupu∈UEt,x,ν{∫TtΦ(s,Xus,ν(s),θ(s))ds+U(XuT)}, | (4.23) |
is equivalent to problem (3.5). Hence, the condition (1) in Proposition 3.1 holds. To verify that the condition (5) in Proposition 3.1 holds, the candidate strategy (a∗(t),θ∗(t)) and the candidate value function H(t,x,ν) under the probability distribution Qθ∗ need to satisfy the following three properties:
(i) E(supt∈[0,T]|H(t,Xa∗(t),ν(t))|4)<∞;
(ii) a∗(t) is an admissible strategy;
(iii) E(supt∈[0,T]|(θ∗1(t))22ϕ1(t,Xa∗(t),ν(t))+(θ∗2(t))22ϕ2(t,Xa∗(t),ν(t))|2)<∞.
Similar to inequality (91) in [28], we can find that property (i) holds. From the solution process of the Hamilton-Jacobi-Bellman-Isaacs equation (3.5), we know that conditions (i) and (ii) in Definition 2.1 are satisfied, and that the strategy a∗(t) is deterministic and state-independent; thus, condition (iii) in Definition 2.1 holds. Thus, a∗(t) is an admissible strategy. To prove property (iii), let
ˉΓ(t)=m(θ∗1(t))22β1+m(θ∗2(t))22β2. |
Obviously, ˉΓ(t) is bounded. According to (4.2) with H(t,x,ν) instead of V(t,x,ν), we obtain
E(supt∈[0,T]|(θ∗1(t))22ϕ1(t,Xa∗(t),ν(t))+(θ∗2(t))22ϕ2(t,Xa∗(t),ν(t))|2)=E(supt∈[0,T]|ˉΓ(t)|2|H(t,Xa∗(t),ν(t))|2)≤E(supt∈[0,T]|ˉΓ(t)|4)12(supt∈[0,T]|H(t,Xa∗(t),ν(t))|4)12<∞. |
The first inequality follows from the Cauchy-Schwarz inequality, and the second inequality follows from property (i). Thus, Theorem 4.2 holds.
To more intuitively analyze the effects of key model features on the robust optimal reinsurance strategy, we consider four special cases: the no-competition case, the uncorrelated insurance businesses case, the no-ambiguity-aversion case and the uncorrelated claims case. The resulting robust optimal reinsurance strategies for these four special cases can be directly derived from Theorem 4.1. The corresponding value functions can also be similarly obtained. However, their expressions are rather complicated and are thus omitted here.
Corollary 4.1. If we do not consider the competition between the insurer and reinsurer (that is, only consider the interest of the insurer), i.e., τ=0, the robust optimal reinsurance strategy is given by
a∗(t)=η2(ν+p1e−δt)+(m+β1)δq21(2A(t)+˜B(t))(m+β1)q21er(T−t)∨0, | (4.24) |
where
˜B(t)=eη2δ(T−t)∫Tt{2A(s)[δp1e−δs+η2δp1e−δs]−(1+η2)er(T−s)+η2p1e−δs(m+β1)q21}e−η2δ(T−s)ds, | (4.25) |
and A(t) is given by (4.7).
The robust optimal reinsurance strategy (4.24) is similar to (28) in [28]. In other words, our model extends the robust optimal reinsurance strategy in [28] to the competition case, i.e., the case of simultaneously considering the joint interests of the insurer and the reinsurer.
Corollary 4.2. If we do not consider the interdependence between insurance businesses, i.e., ρ=0, the robust optimal reinsurance strategy is given by
a∗(t)=11+τ[τ+η2(ν+p1e−δt)+(m+β1)δq21(2A(t)+˜B(t))(m+β1)q21er(T−t)]∨0, | (4.26) |
where A(t) and ˜B(t) are given by (4.7) and (4.25), respectively. Furthermore, if we do not consider correlated claims, i.e., δ=ν(t)=0, then (4.26) reduces to
a∗(t)=11+τ[τ+η2p1(m+β1)q21er(T−t)]. | (4.27) |
We can see that the robust optimal reinsurance strategy (4.27) is similar to (3.5) in [21]. This shows that our model extends the robust optimal reinsurance strategy in [21] to a more general case.
Corollary 4.3. If we do not consider ambiguity aversion, i.e., β1=β2=0, the robust optimal reinsurance strategy is given by
a∗(t)=11+τ[τ+η2(ν+p1e−δt)+mδq21(2ˆA(t)+ˆB(t))+mq1q2τρer(T−t)mq21er(T−t)]∨0, | (4.28) |
where
ˆA(t)=η24λ1μ12δm[e2δη2(T−t)−1], | (4.29) |
ˆB(t)=eη2δ(T−t)∫Tt{2A(s)[δp1e−δs−2ρδq1q2τmer(T−s)+δmq1q2τρer(T−s)+η2δp1e−δs]−(1+η2)er(T−s)+η2p1e−δsmq21+η2mq2τρer(T−s)mq1}e−η2δ(T−s)ds. | (4.30) |
We find that (4.3) and (4.28) have very similar forms. However, we cannot arbitrarily say that ambiguity aversion has no significant influence on the robust optimal reinsurance strategy. In the next section, we will concretely analyze the influence of ambiguity aversion on the robust optimal reinsurance strategy through numerical experiments.
Corollary 4.4. If we do not consider correlation among claims, i.e., ν(t)=δ=0, the robust optimal reinsurance strategy is given by
a∗(t)=11+τ[τ+η2p1+mq1q2τρer(T−t)(m+β1)q21er(T−t)]. | (4.31) |
Comparing (4.3) and (4.31), we find that (4.31) becomes simpler. This shows that the correlated claims have a significant impact on the robust optimal reinsurance strategy. However, the specific forms of influence will be further discussed in the next section.
In this section, we will conduct a series of numerical experiments to illustrate the effects of time, the correlated claims, the dependence of insurance business, the competition degree and the ambiguity-aversion coefficient on the robust optimal reinsurance strategy. We also analyze the similarities and differences between the robust optimal reinsurance strategy obtained when the correlated claims, the dependence of insurance business and the competition are considered and are not considered, respectively.
By referring to the relevant studies such as [6,21,27,28], we set the basic parameters as those in Table 1.
t | T | λ1 | λ2 | λ | μ11 | μ12 | μ21 | μ22 | m | β1 | β2 | r | δ | ν | η1 | η2 | τ |
0 | 10 | 0.8 | 0.5 | 0.04 | 2.2 | 2.7 | 1.6 | 2.3 | 0.2 | 0.2 | 0.4 | 0.02 | 0.25 | 0.4 | 0.15 | 0.5 | 0.5 |
In this subsection, we examine the effects of time, the correlated claims, the competition degree and the ambiguity aversion coefficient on the robust optimal reinsurance strategy when the dependence between the insurer's insurance business and the reinsurer's insurance business is considered and is not considered, respectively.
Figure 1 illustrates the effect of time t on a∗(t) when the dependence of the insurance business is considered (λ>0) and is not considered (λ=0). From Figure 1, we can see that a∗(t) is increasing with respect to t, regardless of whether the dependence of the insurance business is considered or not. Since we consider the correlated claims, the insurer can obtain more claim information as time goes on. Therefore, as time passes, the insurer would reduce his reinsurance willingness.
For given time t=0, Figure 2 shows the influence of the extrapolation intensity δ on the robust optimal reinsurance strategy a∗(0) when the dependence of the insurance business is considered (λ>0) and is not considered (λ=0). The larger the δ, the larger the weighted average ν(t). The weighted average ν(t) is negatively correlated with the risk of the wealth process; hence, ν(t) can offset more risk of the wealth process as δ increases. And then, the insurer decreases his reinsurance demand as δ increases. Therefore, in Figure 2, we see that a∗(0) is an increasing function of δ, regardless of whether the dependence of insurance businesses is considered or not.
For given time t=0, Figure 3 illustrates the effect of τ on the robust optimal reinsurance strategy a∗(0) when the dependence of the insurance business is considered (λ>0) and is not considered (λ=0). From Figure 3, we can see that a∗(0) is increasing with respect to τ, regardless of whether the dependence of the insurance business is considered or not. A larger τ means that the insurer becomes more concerned about his relative wealth. This makes the insurer become more eager to surpass the reinsurer's wealth. Hence, the insurer increases the value of a∗(0) in order to reduce the reinsurance premium.
For given time t=0, Figure 4 discloses the effect of β1 on the robust optimal reinsurance strategy a∗(0) when the dependence of the insurance business is considered (λ>0) and is not considered (λ=0). From Figure 4, we can see that a∗(0) decreases with respect to β1, regardless of whether the dependence of the insurance business is considered or not. This is intuitive, as β1 stands for the ambiguity-aversion coefficient. With the increase of β1, the insurer would purchase more reinsurance to disperse the underlying distribution uncertainty risk.
From Figures 1–4, we can also see that when considering the dependence of an insurance business, the insurer keeps a larger a∗(t) and a∗(0). This is because, when considering the dependence of an insurance business, the insurer can simultaneously grasp the claims of multiple insurance businesses. Therefore, the insurer can more efficiently grasp the potential claims' risk. This directly leads to the increase of his retention level and the decrease of his reinsurance premium expenditure.
In this subsection, we examine the effect of time, the competition degree and the ambiguity aversion coefficient on the robust optimal reinsurance strategy when the correlated claims are considered and not considered, respectively.
Figure 5 shows the effect of t on a∗(t) when the correlated claims are considered (δ>0,ν(t)>0) and not considered (δ=ν(t)=0), where we set η2=0.3. For given time t=0, Figures 6 and 7 respectively show the effects of τ and β1 on a∗(0) when the correlated claims are considered (δ>0,ν(t)>0) and not considered (δ=ν(t)=0). Similar to Figure 1, we see in Figure 5 that a∗(t) is increasing with respect to t, regardless of whether the correlated claims are considered or not. Similar to Figures 3 and 4, we see in Figures 6 and 7 that a∗(0) is increasing and decreasing with respect to τ and β1, respectively, regardless of whether the correlated claims are considered or not. More importantly, we find from Figures 5–7, that, when the correlated claims are considered, the insurer keeps a larger a∗(t) and a∗(0). This is because, when the correlated claims are considered, the insurer can infer future claim information based on historical claim information. Therefore, the insurer can better grasp the claim risk. Hence, at this time, the insurer reduces his reinsurance willingness. Thus, this directly causes him to keep a large a∗(t) and a∗(0) to reduce the expenditure of the reinsurance premium.
In this subsection, we examine the effects of time, the correlated claims and the ambiguity aversion coefficient on the robust optimal reinsurance strategy when the competition between the insurer and the reinsurer is considered and is not considered, respectively.
Figure 8 shows the effect of t on a∗(t) when the competition between the insurer and the reinsurer is considered (τ>0) and is not considered (τ=0). For given time t=0, Figures 9 and 10 respectively illustrate the effects of δ and β1 on a∗(0) when the competition between the insurer and the reinsurer is considered (τ>0) and is not considered (τ=0). Similar to Figure 1, we see in Figure 8 that a∗(t) is increasing with respect to t, regardless of whether the competition is considered or not. Similar to Figures 2 and 4, we see in Figures 9 and 10 that a∗(0) is increasing and decreasing with respect to δ and β1, respectively, regardless of whether the competition is considered or not. More importantly, we find from Figures 8–10, that, when the competition is considered, the insurer keeps a larger a∗(t) and a∗(0). The reason is similar to that for Figure 3. That is, when considering competition, the insurer hopes to surpass the reinsurer's wealth. While purchasing reinsurance can reduce his risk, it is nonetheless costly because the insurer needs to pay the reinsurance premium to the reinsurer for the reinsurance protection. This would decrease his wealth value relative to that of his competitor. Hence, the insurer keeps a larger a∗(t) and a∗(0) when the competition is considered.
Above, we have analyzed the effects of time, the correlated claims, the dependence of the insurance business, the competition degree and the ambiguity aversion coefficient on the robust optimal reinsurance strategy. To further analyze which of these model parameters has a more significant influence on the robust optimal reinsurance strategy, for given time t=0, we examine some joint effects of these model parameters on a∗(0) under the general situation, i.e., all the of above-mentioned model features are considered. The results are shown in Figures 11–14.
From Figures 11–14, we can see that the influence of the extrapolation intensity δ has the most significant influence on a∗(0). This means that the correlated claims have the most significant influence on the insurer's reinsurance decision. This may be because the insurer can better grasp the possible claim risk when considering the correlated claims. This shows the necessity and importance of considering the correlated claims.
A robust optimal reinsurance strategy selection problem was studied in this paper. To reduce claim risk, the insurer can purchase proportional reinsurance or acquire a new business; to increase wealth, the reinsurer can engage in insurance business. The interdependence between the insurance businesses of both the insurer and the reinsurer is considered. Meanwhile, the correlation between the insurer's future claims and historical claims is also considered. Furthermore, we have quantified the competition between the insurer and the reinsurer through relative wealth. These model settings extend the model in [28]. Under ambiguity aversion and the criterion of maximizing the expected utility of terminal wealth, we have obtained explicit solutions for the robust optimal reinsurance strategy and the corresponding value function by using the stochastic dynamic programming approach. Finally, a series of numerical experiments were carried out to examine the influences of the model parameters on the robust optimal reinsurance strategy.
There are still some issues worthy of investigation in the future. First, we considered the joint interests of the insurer and the reinsurer through competition, but we did not consider the decision of the reinsurer. It is interesting to consider the decision of the reinsurer. Second, it is also interesting to consider other reinsurance patterns, e.g., the excess-of-loss reinsurance and per-loss reinsurance. Third, it is worthwhile to consider other forms of competition between the insurer and the reinsurer.
This research was supported by the Natural Science Basic Research Plan in Shaanxi Province of China under Grant Number 2023-JC-YB-002 and the Humanities and Social Sciences Project of the Ministry of Education of China under Grant Number 21XJC910001.
The author declares that he has no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
[1] |
X. Han, Z. Liang, V. R. Young, Optimal reinsurance to minimize the probability of drawdown under the mean-variance premium principle, Scand. Actuar. J., 2020 (2020), 879–903. http://doi.org/10.1080/03461238.2020.1788136 doi: 10.1080/03461238.2020.1788136
![]() |
[2] |
H. Yener, Proportional reinsurance and investment in multiple risky assets under borrowing constraint, Scand. Actuar. J., 2020 (2020), 396–418. http://doi.org/10.1080/03461238.2019.1676301 doi: 10.1080/03461238.2019.1676301
![]() |
[3] |
Y. Yuan, Z. Liang, X. Han, Optimal investment and reinsurance to minimize the probability of drawdown with borrowing costs, J. Ind. Manag. Optim., 18 (2022), 933–967. http://doi.org/10.3934/jimo.2021003 doi: 10.3934/jimo.2021003
![]() |
[4] |
X. Lin, P. Yang, Optimal investment and reinsurance in a jump diffusion risk model, The ANZIAM Journal, 52 (2011), 250–262. http://doi.org/10.1017/S144618111100068X doi: 10.1017/S144618111100068X
![]() |
[5] |
X. Rong, Y. Yan, H. Zhao, Asymptotic solution of optimal reinsurance and investment problem with correlation risk for an insurer under the CEV model, Int. J. Control, 96 (2023), 839–852. http://doi.org/10.1080/00207179.2021.2015627 doi: 10.1080/00207179.2021.2015627
![]() |
[6] |
C. Ceci, K. Colaneri, A. Cretarola, Optimal reinsurance and investment under common shock dependence between financial and actuarial markets, Insur. Math. Econ., 105 (2022), 252–278. https://doi.org/10.1016/j.insmatheco.2022.04.011 doi: 10.1016/j.insmatheco.2022.04.011
![]() |
[7] |
Z. Liang, J. Bi, K. C. Yuen, C. Zhang, Optimal mean Cvariance reinsurance and investment in a jump-diffusion financial market with common shock dependence, Math. Meth. Oper. Res., 84 (2016), 155–181. http://doi.org/10.1007/s00186-016-0538-0 doi: 10.1007/s00186-016-0538-0
![]() |
[8] |
P. Yang, Z. Chen, L. Wang, Time-consistent reinsurance and investment strategy combining quota-share and excess of loss for mean-variance insurers with jump-diffusion price process, Commun. Stat.-Theor. M., 50 (2021), 2546–2568. https://doi.org/10.1080/03610926.2019.1670849 doi: 10.1080/03610926.2019.1670849
![]() |
[9] | W. Wang, D. Muravey, Y. Shen, Y. Zeng, Optimal investment and reinsurance strategies under 4/2 stochastic volatility model, Scand. Actuar. J., in press. http://doi.org/10.1080/03461238.2022.2108335 |
[10] |
D. Li, X. Rong, H. Zhao, Time-consistent reinsurance-investment strategy for an insurer and a reinsurer with mean-variance criterion under the CEV model, J. Comput. Appl. Math., 283 (2015), 142–162. http://doi.org/10.1016/j.cam.2015.01.038 doi: 10.1016/j.cam.2015.01.038
![]() |
[11] |
J. Cai, C. Lemieux, F. Liu, Optimal reinsurance from the perspectives of both an insurer and a reinsurer, Astin Bull., 46 (2016), 815–849. https://doi.org/10.1017/asb.2015.23 doi: 10.1017/asb.2015.23
![]() |
[12] |
Y. Huang, Y. Ouyang, L. Tang, J. Zhou, Robust optimal investment and reinsurance problem for the product of the insurer's and the reinsurer's utilities, J. Comput. Appl. Math., 344 (2018), 532–552. http://doi.org/10.1016/j.cam.2018.05.060 doi: 10.1016/j.cam.2018.05.060
![]() |
[13] |
Y. Zhang, H. Zhao, X. Rong, K. Han, Optimal investment and reinsurance problem toward joint interests of the insurer and the reinsurer under default risk, Commun. Stat.-Theor. M., 51 (2022), 6535–6558. http://doi.org/10.1080/03610926.2020.1862872 doi: 10.1080/03610926.2020.1862872
![]() |
[14] |
Y. Zhang, P. Zhao, H. Zhou, The optimal reinsurance-investment problem considering the joint interests of an insurer and a reinsurer under HARA utility, Acta Math. Sci., 43 (2023), 97–124. http://doi.org/10.1007/s10473-023-0107-6 doi: 10.1007/s10473-023-0107-6
![]() |
[15] |
P. M. DeMarzo, R. Kaniel, I. Kremer, Relative wealth concerns and financial bubbles, Rev. Financ. Stud., 21 (2008), 19–50. https://doi.org/10.1093/rfs/hhm032 doi: 10.1093/rfs/hhm032
![]() |
[16] |
A. Bensoussan, C. C. Siu, S. C. P. Yam, H. Yang, A class of non-zero-sum stochastic differential investment and reinsurance games, Automatica, 50 (2014), 2025–2037. https://doi.org/10.1016/j.automatica.2014.05.033 doi: 10.1016/j.automatica.2014.05.033
![]() |
[17] |
K. Y. Kwok, M. C. Chiu, H. Y. Wong, Demand for longevity securities under relative performance concerns: stochastic differential games with cointegration, Insur. Math. Econ., 71 (2016), 353–366. http://doi.org/10.1016/j.insmatheco.2016.10.005 doi: 10.1016/j.insmatheco.2016.10.005
![]() |
[18] |
J. Zhu, G. Guan, S. Li, Time-consistent non-zero-sum stochastic differential reinsurance and investment game under default and volatility risks, J. Comput. Appl. Math., 374 (2020), 112737. http://doi.org/10.1016/j.cam.2020.112737 doi: 10.1016/j.cam.2020.112737
![]() |
[19] | X. Dong, X. Rong, H. Zhao, Non-zero-sum reinsurance and investment game with non-trivial curved strategy structure under Ornstein-Uhlenbeck process, Scand. Actuar. J., in press. https://doi.org/10.1080/03461238.2022.2139631 |
[20] |
P. Yang, Z. Chen, Y. Xu, Time-consistent equilibrium reinsurance-investment strategy for n competitive insurers under a new interaction mechanism and a general investment framework, J. Comput. Appl. Math., 374 (2020), 112769. http://doi.org/10.1016/j.cam.2020.112769 doi: 10.1016/j.cam.2020.112769
![]() |
[21] | P. Yang, Z. Chen, Optimal reinsurance pricing, risk sharing and investment strategies in a joint reinsurer-insurer framework, IMA J. Manag. Math., in press. http://doi.org/10.1093/imaman/dpac002 |
[22] |
G. Niehaus, A. Terry, Evidence on the time series properties of insurance premiums and causes of the underwriting cycle: new support for the capital market imperfection hypothesis, J. Risk Insur., 60 (1993), 466–479. http://doi.org/10.2307/253038 doi: 10.2307/253038
![]() |
[23] |
M. J. Browne, R. E. Hoyt, The demand for flood insurance: empirical evidence, J. Risk Uncertainty, 20 (2000), 291–306. https://doi.org/10.1023/A:1007823631497 doi: 10.1023/A:1007823631497
![]() |
[24] |
R. Ranyard, S. McHugh, Defusing the risk of borrowing: the psychology of payment protection insurance decisions, J. Econ. Psychol., 33 (2012), 738–748. https://doi.org/10.1016/j.joep.2012.02.002 doi: 10.1016/j.joep.2012.02.002
![]() |
[25] |
J. W. Ruser, Workers' compensation insurance, experience-rating, and occupational injuries, Rand J. Econ., 16 (1985), 487–503. http://doi.org/10.2307/2555508 doi: 10.2307/2555508
![]() |
[26] |
S. Chen, D. Hu, H. Wang, Optimal reinsurance problem with extrapolative claim expectation, Optim. Contr. Appl. Met., 39 (2018), 78–94. https://doi.org/10.1002/oca.2335 doi: 10.1002/oca.2335
![]() |
[27] |
D. Hu, H. Wang, Optimal proportional reinsurance with a loss dependent premium principle, Scand. Actuar. J., 2019 (2019), 752–767. http://doi.org/10.1080/03461238.2019.1604426 doi: 10.1080/03461238.2019.1604426
![]() |
[28] |
Z. Chen, P. Yang, Robust optimal reinsurance-investment strategy with price jumps and correlated claims, Insur. Math. Econ., 92 (2020), 27–46. https://doi.org/10.1016/j.insmatheco.2020.03.001 doi: 10.1016/j.insmatheco.2020.03.001
![]() |
[29] |
N. Barberis, R. Greenwood, L. Jin, A. Shleifer, X-CAPM: an extrapolative capital asset pricing model, J. Financ. Econ., 115 (2015), 1–24. https://doi.org/10.1016/j.jfineco.2014.08.007 doi: 10.1016/j.jfineco.2014.08.007
![]() |
[30] |
E. W. Anderson, L. P. Hansen, T. J. Sargent, A quartet of semi-groups for model specification, robustness, prices of risk, and model detection, J. Eur. Econ. Assoc., 1 (2023), 68–123. https://doi.org/10.1162/154247603322256774 doi: 10.1162/154247603322256774
![]() |
[31] |
D. Li, Y. Zeng, H. Yang, Robust optimal excess-of-loss reinsurance and investment strategy for an insurer in a model with jumps, Scand. Actuar. J., 2018 (2018), 145–171. http://doi.org/10.1080/03461238.2017.1309679 doi: 10.1080/03461238.2017.1309679
![]() |
[32] | W. Zhang, H. Meng, Robust optimal dynamic reinsurance policies under the mean-RVaR premium principle, Commun. Stat.-Theor. M., in press. http://doi.org/10.1080/03610926.2022.2076121 |
[33] |
D. Hu, H. Wang, Robust reinsurance contract with learning and ambiguity aversion, Scand. Actuar. J., 2022 (2022), 794–815. http://doi.org/10.1080/03461238.2022.2030398 doi: 10.1080/03461238.2022.2030398
![]() |
[34] |
Y. Yuan, Z. Liang, X. Han, Robust optimal reinsurance in minimizing the penalized expected time to reach a goal, J. Comput. Appl. Math., 420 (2023), 114816. http://doi.org/10.1016/j.cam.2022.114816 doi: 10.1016/j.cam.2022.114816
![]() |
[35] | J. Grandell, Aspects of risk theory, New York: Springer, 1991. https://doi.org/10.1007/978-1-4613-9058-9 |
[36] |
N. Bäuerle, Benchmark and mean-variance problems for insurers, Math. Meth. Oper. Res., 62 (2005), 159–165. https://doi.org/10.1007/s00186-005-0446-1 doi: 10.1007/s00186-005-0446-1
![]() |
[37] |
N. Branger, L. S. Larsen, Robust portfolio choice with uncertainty about jump and diffusion risk, J. Bank. Financ., 37 (2013), 5036–5047. https://doi.org/10.1016/j.jbankfin.2013.08.023 doi: 10.1016/j.jbankfin.2013.08.023
![]() |
[38] |
S. Mataramvura, B. Øksendal, Risk minimizing and HJBI equations for stochastic differential games, Stochastics, 80 (2008), 317–337. http://doi.org/10.1080/17442500701655408 doi: 10.1080/17442500701655408
![]() |
[39] |
P. J. Maenhout, Robust portfolio rules and asset pricing, Rev. Financ. Stud., 17 (2004), 951–983. http://doi.org/10.1093/rfs/hhh003 doi: 10.1093/rfs/hhh003
![]() |
1. | Peng Yang, Robust optimal reinsurance-investment problem for $ n $ competitive and cooperative insurers under ambiguity aversion, 2023, 8, 2473-6988, 25131, 10.3934/math.20231283 | |
2. | Yiming Su, Haiyan Liu, Mi Chen, Robust equilibrium reinsurance and investment strategy for the insurer and reinsurer under weighted mean-variance criterion, 2023, 31, 2688-1594, 6384, 10.3934/era.2023323 |
t | T | λ1 | λ2 | λ | μ11 | μ12 | μ21 | μ22 | m | β1 | β2 | r | δ | ν | η1 | η2 | τ |
0 | 10 | 0.8 | 0.5 | 0.04 | 2.2 | 2.7 | 1.6 | 2.3 | 0.2 | 0.2 | 0.4 | 0.02 | 0.25 | 0.4 | 0.15 | 0.5 | 0.5 |