Next Article in Journal
Cox-Based and Elliptical Telegraph Processes and Their Applications
Previous Article in Journal
Asymmetric Wealth Effect between US Stock Markets and US Housing Market and European Stock Markets: Evidences from TAR and MTAR
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimal Reinsurance under the Linear Combination of Risk Measures in the Presence of Reinsurance Loss Limit

1
School of Mathematics and Statistics, Southwest University, Chongqing 400715, China
2
Department of Mathematics, University of Manchester, Manchester M13 9PL, UK
*
Author to whom correspondence should be addressed.
Risks 2023, 11(7), 125; https://doi.org/10.3390/risks11070125
Submission received: 25 May 2023 / Revised: 30 June 2023 / Accepted: 5 July 2023 / Published: 10 July 2023

Abstract

:
Optimal reinsurance problems under the risk measures, such as Value-at-Risk ( VaR ) and Tail-Value-at-Risk ( TVaR ), have been studied in recent literature. However, losses based on VaR may be underestimated and TVaR allows us to account better for catastrophic losses. In this paper, we propose a new family of flexible risk measures denoted by LVaR , which is a weighted combination of VaR and TVaR . Based on the new risk measures, we deal with the optimal reinsurance problem by minimizing the LVaR of the total risks of an insurer when two types of constraints for reinsurer’s risk exposure are considered. The results indicate that the two-layer reinsurance is always an optimal reinsurance policy with both types of constraints. Also, we find that the optimal reinsurance policy depends on the confidence level, the weight coefficient, the safety loading, the tolerance level, as well as the relations between them. Finally, we illustrate the results by numerical examples and compare them with the results in Lu et al.

1. Introduction

Reinsurance can be seen as a risk transfer that may help to reduce the risk exposure of the insurer and, hence, to stabilize the business. More precisely, the insurer transfers some part of risk to a reinsurance company at the expense of paying the corresponding reinsurance premium, which reduces the potential risk. Thanks to this tool, catastrophic risk, such as climate risks, initially hardly insurable, may become insurable by transferring risks (see, e.g., Charpentier 2008). Naturally, the reinsurance premiums increase as the risk transferred to a reinsurer increases. In this context, how to deal with the trade-off between the risk retained and the premium paid to the reinsurer becomes a major issue for an insurer.
The study of the optimal reinsurance problems can be traced back to the seminal papers of Borch (1960) and Arrow (1963). Calculating the reinsurance premium by using the expected value principle, Borch (1960) showed that the stop-loss reinsurance is optimal when the objective is to minimize the variance of the insurer’s retained loss. Choosing the same premium principle as Borch (1960), Arrow (1963) considered the optimal reinsurance problem by maximizing expected utility of the terminal wealth of a risk-averse insurer. The results also showed that the stop-loss reinsurance is optimal. Kaluszka (2001) extended Borch’s result by applying the mean-variance premium principle. Arrow’s result has been considered under other premium principles, see Young (1999) and Kaluszka (2005). For different purposes, the optimal reinsurance problem has been studied under more intricate optimization criteria and/or more general premium principles, see Albrecher et al. (2017) for a survey, and some examples. When the reinsurance premium is calculated according to the maximal possible claims principle, Kaluszka and Okolewski (2008) showed that the limited stop-loss and the truncated stop-loss are the optimal reinsurance contracts by maximizing the expected utility. Guerra and Centeno (2010) considered the optimal forms of reinsurance when the insurer seeks to maximize the adjustment coefficient of the retained risk. Using the maximization of the expected utility of terminal wealth, Zhang and Siu (2012) and Liang and Bayraktar (2014) studied the optimal reinsurance and investment strategies with different market assumptions.
In the fields of finance and insurance, Value-at-Risk ( VaR ) and Tail-Value-at-Risk ( TVaR ) are the most popular risk measures due to their merits and good properties. For this reason, many optimal reinsurance problems under the VaR and/or the TVaR have been studied in the literature. For instance, under the expected value premium principle, Cai and Tan (2007) provided the optimal retention of a stop-loss reinsurance by minimizing the VaR and the TVaR of the insurer’s total risk exposure. Cai et al. (2008) derived the optimal reinsurance within a class of increasing convex loss functions. Cheung (2010) extended the VaR -minimization reinsurance model in Cai et al. (2008) by considering Wang’s premium principle. Lu et al. (2016) considered the optimal reinsurance problem under the optimality criteria of VaR and TVaR risk measures when constraints for the reinsurer’s risk exposure are presented. Further extensions of the optimal reinsurance problems can be found in Balbás et al. (2009), Tan et al. (2011), and Chi and Tan (2011, 2013).
Let X denote the amount of loss initially assumed by an insurer in a given time period. Usually, X is assumed to be a non-negative random variable defined on the probability space Ω , F , P with continuous distribution function F X ( x ) . We further assume that F X ( x ) is strictly increasing on ( 0 , + ) but with a possible jump at 0 with P ( X = 0 ) = p 0 . S X ( x ) = 1 F X ( x ) and f X ( x ) are used to denote the survival function and the density function of X, respectively. Given a confidence level α ( 0 , 1 ) , the VaR and the TVaR of the risk X are defined as
VaR α ( X ) = inf x R : F X ( x ) α
and
TVaR α ( X ) = 1 1 α α 1 VaR s ( X ) d s ,
respectively. For a continuous loss distribution, there are many alternative names of TVaR , such as Expected Shortfall ( ES ), Conditional Tail Expectation ( CTE ), Average VaR ( AVaR ), and Conditional VaR ( CVaR ). We refer to Acerbi and Tasche (2002) and Rockafellar and Uryasev (2013) for the relationships between the various notions. It is well known that VaR is simple and easier to interpret, but it is, in general, not subadditive and, hence, not coherent in the sense of Artzner et al. (1999). Another major drawback of VaR is that it only takes into account the probability of a big loss, not the size of the loss, i.e., it completely ignores the tail loss beyond the reference VaR . As a result, the same VaR s may occur when dealing with different extreme losses, and moreover, VaR may underestimate the losses in practice, especially when heavy-tailed losses are incorrectly modeled with light-tailed distributions, such as the normal distribution. By comparison with VaR , TVaR defines a more conservative risk measure that is always subadditive. Since it is interpreted as the arithmetic average of VaR α over all levels s α , capital reserves based on TVaR are always larger than those based on VaR . Further discussions about the comparative advantages of TVaR and VaR , we refer to Embrechts and Hofert (2014), McNeil et al. (2015), and references therein.
As a generalization of VaR , the spectral risk measure has received attention in recent years. In fact, the spectral risk measure is a weighted average of VaR α with the weight function depending on the user’s risk-aversion, so that it could help us to link the risk measure to the user’s attitude towards risk ((Acerbi 2002) and Dowd et al. (2008)). Another interesting generalization of VaR α is the Lambda Value-at-Risk, which considers the dependence between the level α and the amount of the loss (see, e.g., Frittelli et al. 2014; Bellini and Peri 2022). Recall that, in the risk-management community, there is an ongoing debate on the advantages and disadvantages of TVaR and VaR . However, there is no evidence for global advantage of one risk measure against the other, and the size of capital reserves may be significantly different depending on which risk measure is used. To provide a risk assessment between that offered by TVaR and VaR , and capture more information about various attitudes towards risk, in this paper, we extend VaR and TVaR to a more general family of risk measures, denoted by LVaR , which is a linear combination of VaR and TVaR . Obviously, LVaR includes VaR and TVaR as special cases and allows us to consider VaR and TVaR simultaneously.
As shown in the literature, once the optimization criteria and the premium principle are determined, the optimal reinsurance problem becomes a purely mathematical problem, which makes the analysis easier. The expected value principle is one of the most popular premium principles both due to its transparency and simplicity (see, e.g., Albrecher et al. 2017), so that we assume that the reinsurance premium is calculated by this premium principle in this paper. To determine the amount of risk retained (or the premium paid to the reinsurer), the insurer has to make a choice among all feasible reinsurance treaties, and a reasonable criterion for the insurer is naturally to choose the reinsurance form which makes its total risk to be as small as possible. To this end, we revisit the optimal reinsurance problem by using the new risk measure LVaR to quantify the total risk exposure of the insurer.
Recently, attention has been paid to controlling reinsurer’s risk since the insurer may be under a heavy financial burden with no limit on coverage. Throughout the paper, we will consider two types of constraints proposed by Cummins and Mahul (2004) and Zhou et al. (2010). The first type due to Cummins and Mahul (2004) has the ceded losses constrained to be less than the upper limit. The second type is that the reinsurer’s loss after the payments is limited to be less than a certain predetermined level. As mentioned above, the motivation behind these two classes of upper limits on coverage arises from providing the insurer with limited liability with respect to the indemnity schedule. Instead of constraining the risk of the reinsurer, Boonen et al. (2016) established lower and upper bounds of the reinsurance premium which ensure the benefits of reinsurance to both insurers and reinsurers. More recently, their work has been generalized by Boonen et al. (2021) to the case where an insurer could use more than one reinsurer to reinsure its risk, i.e., the case where there is competition among multiple reinsurers. With similar constraints, Balbás et al. (2022) showed that the optimal reinsurance problem may be very complex if the expected profits of both insurers and reinsurers are required to be non-decreasing when the reinsurance contract is signed. Here, we focus on the two types of constraints discussed in Cummins and Mahul (2004) and Zhou et al. (2010).
Many researchers have discussed the existence of optimal reinsurance contracts. However, closed-form expressions of the optimal risk transferred to the reinsurer and the resulting total risk of the insurer have not been provided. Furthermore, the study of the optimal reinsurance problems under the VaR and the TVaR is generally discussed separately, which ignores the situation where one may consider both VaR and TVaR simultaneously. Thus, inspired by Lu et al. (2016), we derive the optimal transferred risk in closed-form, but the most important distinction is that we seek to deal with the optimal reinsurance problem under the LVaR , while Lu et al. (2016) considered this problem under the VaR and the TVaR separately. To sum up, the main contributions of the present paper are as follows. First, we introduce a new family of risk measures LVaR to capture more information about various attitudes towards risk and to enable us to consider VaR and TVaR simultaneously. Second, by minimizing LVaR of the insurer’s total risk and using the expected value principle, we identify optimal reinsurance contracts when the reinsurer’s risk exposure is constrained. It appears that the two-layer reinsurance is always an optimal reinsurance policy which shows the stability of optimality results when switching from VaR and TVaR to LVaR . Moreover, we show that the solutions of optimal reinsurance model in Lu et al. (2016) can be unified and generalized by using the LVaR . In addition, the optimal quantity of ceded losses depends on the confidence level α , the weight coefficient ω , the safety loading θ , the tolerance level L (or K), as well as the relations between them. For the insurer, our results provide explicit expressions of the optimal reinsurance contract which may help to foster the intuition and comprehension of the consequences of setting up a reinsurance contract.
The rest of the paper is organized as follows. In Section 2, we give some preliminaries, define a new family of risk measures, and describe the setup of the proposed reinsurance models. Section 3 states the main results and is structured as follows. Section 3.1 studies the optimal reinsurance problem under the new risk measures optimality criterion by considering the first type of constraint. Section 3.2 studies the same problem in Section 3.1 with the second type of constraint. Section 4 illustrates the results in Section 3 by numerical examples and also compares them with the results in Lu et al. (2016). Section 5 concludes our study. All proofs are given in the Appendix A.

2. Preliminaries

2.1. A New Risk Measure

Recall that, from the definitions in (1) and (2), the VaR measures the minimum loss, so that a disadvantage when using VaR is that catastrophic losses can be underestimated. Since the TVaR measures average losses in the most adverse cases rather than just the minimum loss, naturally, losses based on TVaR are much larger than those based on VaR . Therefore, our objective is to propose a new family of risk measures which provides a risk assessment that lies somewhere between those offered by the VaR and the TVaR . The new family of risk measures, named LVaR , is defined as a linear combination of VaR and TVaR , and is helpful in providing risk managers with more flexible tools. We now formally define LVaR .
Definition 1.
The LVaR of a random variable X at a confidence level α ( 0 , 1 ) is defined as
LVaR α ( X ) = ω · TVaR α ( X ) + ( 1 ω ) · VaR α ( X ) ,
where ω [ 0 , 1 ] is the weight coefficient.
Remark 1.
Obviously, the LVaR risk measures include VaR and TVaR as special cases by setting ω = 0 and ω = 1 , respectively. Moreover, we can easily check that VaR α ( X ) LVaR α ( X ) TVaR α ( X ) .
Remark 2.
Note that different perspectives will lead to different weight coefficients. For example, a more pessimistic insurer may prefer the larger ω. As a consequence, the choice of the weight coefficient ω can be complicated in practice, and hence is not addressed here.

2.2. Optimal Reinsurance Model

Under a typical reinsurance arrangement, let I ( X ) be the part of loss transferred from the insurer to a reinsurer which we refer to as the ceded loss function. As usual in the reinsurance policy, we assume that the ceded loss function I ( X ) satisfies the following properties: (1) I ( 0 ) = 0 and I ( x ) is an increasing function, and (2) I x 2 I x 1 x 2 x 1 for any 0 x 1 x 2 . It is easy to check that these conditions imply that 0 I ( x ) x for any x 0 . In the sequel, we will denote by F the set of all ceded loss functions satisfying conditions (1) and (2). Associated with the ceded loss function I ( X ) , we denote R I ( X ) = X I ( X ) as the retained loss of the insurer. As an exchange of the undertaking risk, the insurer should pay a reinsurance premium to the reinsurer. Here, we assume that the reinsurance premium is determined by the common expectation premium principle and expressed as
Π I ( X ) = ( 1 + θ ) E I ( X ) ,
where θ > 0 is the safety loading factor. Consequently, the total risk exposure of the insurer in the presence of reinsurance, denoted by T I ( X ) , is the sum of the retained loss and the reinsurance premium, i.e.,
T I ( X ) = R I ( X ) + Π I ( X ) .
The objective of this paper is, by minimizing the new risk measure LVaR , to consider the optimal reinsurance problem when the constraints for the reinsurer’s risk exposure are presented. Two types of constraints due to Cummins and Mahul (2004) and Zhou et al. (2010) are considered. Specifically, we are interested in seeking an optimal reinsurance policy over the following two sets of ceded loss functions:
  • F 1 = I ( x ) F | I ( x ) L ;
  • F 2 = I ( x ) F | I ( x ) Π I ( X ) K .
For I F i , i = 1 , 2 , LVaR α T I ( X ) can be seen as a function of I ( X ) , and it suggests the amount of assets required to protect against adverse outcomes of the total risk T I ( X ) . For a fixed confidence level, a prudent risk management generally requires LVaR α T I ( X ) to be as small as possible. This allows us to determine the optimal ceded loss function by minimizing LVaR α T I ( X ) in the above classes of ceded loss functions. Our optimal reinsurance model can be formulated as follows:
LVaR optimization criterion
LVaR α T I ( X ) = min I F i LVaR α T I ( X ) ,
where i = 1 , 2 , and I is the resulting optimal ceded loss function.
We conclude this section by introducing the following notations:
δ = 1 + θ ω 1 α , θ = θ 1 + θ .

3. Main Results

3.1. Optimal Reinsurance under the Constraint of F 1

In this section, we will solve (3) under the constraint F 1 :
LVaR α T I ( X ) = min I F 1 LVaR α T I ( X ) .
First, note that both VaR and TVaR are translation invariant and
VaR α g ( X ) = g VaR α ( X )
for any increasing and continuous function g (see McNeil et al. (2015) and Theorem 1 in Dhaene et al. (2002)). Thus,
VaR α T I ( X ) = VaR α R I ( X ) + Π I ( X ) = VaR α R I ( X ) + Π I ( X ) = R I VaR α ( X ) + Π I ( X ) = VaR α ( X ) I VaR α ( X ) + ( 1 + θ ) E I ( X ) .
Moreover, by noticing that TVaR α ( X ) = VaR α ( X ) + VaR α ( X ) S X ( x ) d x 1 α , we can obtain that
TVaR α T I ( X ) = TVaR α R I ( X ) + Π I ( X ) = TVaR α R I ( X ) + Π I ( X ) = E R I ( X ) | R I ( X ) > VaR α R I ( X ) + Π I ( X ) = E R I ( X ) VaR α R I ( X ) | R I ( X ) > VaR α R I ( X ) + VaR α R I ( X ) + Π I ( X ) = VaR α R I ( X ) R I ( x ) VaR α R I ( X ) d F R I ( X ) ( x ) P R I ( X ) > VaR α R I ( X ) + VaR α R I ( X ) + Π I ( X ) = 1 1 α E R I ( x ) VaR α R I ( X ) + + VaR α ( X ) I VaR α ( X ) + ( 1 + θ ) E I ( X ) .
By applying (5) again, we have
E R I ( x ) VaR α R I ( X ) + = E X I ( X ) VaR α ( X ) + I VaR α ( X ) + = E X I ( X ) VaR α ( X ) + I VaR α ( X ) + I X VaR α ( X ) + I X < VaR α ( X ) = E X I ( X ) VaR α ( X ) + I VaR α ( X ) · I X VaR α ( X ) = E X VaR α ( X ) · I X VaR α ( X ) E I ( X ) I VaR α ( X ) · I X VaR α ( X ) = E X VaR α ( X ) + VaR α ( X ) I ( x ) I VaR α ( X ) d F X ( x ) = E X VaR α ( X ) + VaR α ( X ) I ( x ) d F X ( x ) + ( 1 α ) I VaR α ( X ) .
Substituting (8) into (7), we obtain
TVaR α T I ( X ) = 1 1 α E X VaR α ( X ) + + VaR α ( X ) 1 1 α VaR α ( X ) I ( x ) d F X ( x ) + ( 1 + θ ) E I ( X ) .
Therefore, we have from (6) and (9) that
LVaR α T I ( X ) = ω · TVaR α T I ( X ) + ( 1 ω ) · VaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) ( 1 ω ) I VaR α ( X ) + ( 1 + θ ) E I ( X ) ω 1 α VaR α ( X ) I ( x ) d F X ( x ) ,
or equivalently,
LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) ( 1 ω ) I VaR α ( X ) + ( 1 + θ ) 0 VaR α ( X ) I ( x ) d F X ( x ) + δ VaR α ( X ) I ( x ) d F X ( x ) ,
where Ψ ( α ) = Δ 1 1 α E X VaR α ( X ) + .
To proceed, define
D 1 = Δ ( a , b ) | 0 a VaR α ( X ) b , b a L ,
G 1 = Δ g ( x ; a , b ) | g ( x ; a , b ) = ( x a ) + ( x b ) + , x 0 , ( a , b ) D 1 .
We can easily verify that G 1 F 1 .
Lemma 2.
For any confidence level α p 0 , 1 and I ( x ) F 1 , there always exists a function g ( x ; a , b ) G 1 such that LVaR α T g ( X ) LVaR α T I ( X ) .
Remark 3.
By Lemma 2, (4) reduces to the following optimal problem:
LVaR α T g ( X ) = min g G 1 LVaR α T g ( X ) .
Theorem 1.
For any confidence level α p 0 , 1 , the following results hold according to the sign of δ.
(i) 
For the case of δ < 0 , let a 0 be the solution of the equation
1 ( 1 + θ ) S X ( a ) + δ S X ( a + L ) = 0
with respect to a.
(a) 
If VaR α ( X ) L a 0 , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x a 0 + x a 0 + L + and
min I F 1 LVaR α T I ( X ) = ω Ψ ( α ) + a 0 + ( 1 + θ ) a 0 VaR α ( X ) S X ( x ) d x + δ VaR α ( X ) a 0 + L S X ( x ) d x ;
(b) 
If VaR α ( X ) L > a 0 , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR α ( X ) L + x VaR α ( X ) + and
min I F 1 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) L + ( 1 + θ ) VaR α ( X ) L VaR α ( X ) S X ( x ) d x .
(ii) 
For the case of δ = 0 ,
(a) 
If VaR α ( X ) L VaR θ ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR θ ( X ) + ( x r ) + with r VaR α ( X ) , VaR θ ( X ) + L and
min I F 1 LVaR α T I ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x ;
(b) 
If VaR α ( X ) L > VaR θ ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR α ( X ) L + x VaR α ( X ) + and
min I F 1 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) L + ( 1 + θ ) VaR α ( X ) L VaR α ( X ) S X ( x ) d x .
(iii) 
For the case of δ > 0 ,
(a) 
If α θ , the minimum of LVaR α T I ( X ) is attained at I ( x ) = 0 and
min I F 1 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) ;
(b) 
If α > θ and VaR α ( X ) L VaR θ ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR θ ( X ) + x VaR α ( X ) + and
min I F 1 LVaR α T I ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x ;
(c) 
If α > θ and VaR α ( X ) L > VaR θ ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR α ( X ) L + x VaR α ( X ) + and
min I F 1 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) L + ( 1 + θ ) VaR α ( X ) L VaR α ( X ) S X ( x ) d x .
Remark 4.
For δ > 0 , the optimal ceded loss function of the three cases can be rewritten in an unified form
I ( x ) = ( x r ) + x VaR α ( X ) + ,
where r = max VaR α ( X ) L , min VaR α ( X ) , VaR θ ( X ) . If the safety loading factor θ is large enough, the optimal ceded loss function is I ( x ) = 0 , which means that the insurer purchases no reinsurance.

3.2. Optimal Reinsurance under the Constraint of F 2

In this section, we solve (3) under the constraint of F 2 :
LVaR α T I ( X ) = min I F 2 LVaR α T I ( X ) .
First, we introduce the following notations:
ψ ( a , b ) = Δ b a ( 1 + θ ) a b S X ( x ) d x ,
D 2 = Δ ( a , b ) | 0 a VaR α ( X ) , b a , ψ ( a , b ) K ,
G 2 = Δ g ( x ; a , b ) | g ( x ; a , b ) = ( x a ) + ( x b ) + , x 0 , ( a , b ) D 2 .
We can easily verify that G 2 F 2 .
Before solving (15), we discuss properties of ψ ( a , b ) needed for future analysis. For any a 1 R , taking the partial derivative of ψ a 1 , b with respect to b, we have
ψ a 1 , b b = 1 ( 1 + θ ) S X ( b )
so that
b VaR θ ( X ) ψ a 1 , b b 0 .
From (19), we see that ϕ a 1 , b is convex with respect to b.
For g X ; a 1 , b G 2 , we have g X ; a 1 , b X , which implies that E g X ; a 1 , b E [ X ] . Noting that E g X ; a 1 , b = a 1 b S X ( x ) d x , we have from (16) that
ψ a 1 , b b a 1 ( 1 + θ ) E [ X ] .
Moreover, as b , we have
b a 1 ( 1 + θ ) E [ X ]
and
ψ a 1 , b .
It is obvious that ψ a 1 , a 1 = 0 by the definition of ψ ( a , b ) . Thus, by the continuity and convexity of ψ a 1 , b with respect to b, there exists a unique b 1 > max a 1 , VaR θ ( X ) such that ψ a 1 , b 1 = K and ψ a 1 , b K for b a 1 , b 1 .
For any a R , let β ( a ) be the unique solution to the equation ψ ( a , b ) = K with b max a , VaR θ ( X ) , . For b > a , taking the derivatives with respect to a on both sides of the equation ψ ( a , b ) = K entails
b = β ( a ) = 1 ( 1 + θ ) S X ( a ) 1 ( 1 + θ ) S X ( b ) .
Furthermore, we can verify that 1 ( 1 + θ ) S X ( b ) > 0 when b = β ( a ) > VaR θ ( X ) . Thus,
a VaR θ ( X ) β ( a ) 0 ,
which implies that β ( a ) is convex with respect to a. Therefore, the set D 2 can be rewritten as
D 0 = Δ ( a , b ) | 0 a VaR α ( X ) , a b β ( a ) .
Lemma 3.
For any confidence level α p 0 , 1 and I ( x ) F 2 , there always exists a function g ( x ; a , b ) G 2 such that LVaR α T g ( X ) LVaR α T I ( X ) .
Remark 5.
By Lemma 3, we can solve (15) by considering the following optimal problem:
LVaR α T g ( X ) = min g G 2 LVaR α T g ( X ) .
To facilitate subsequent analysis, let b 0 = β VaR θ ( X ) , that is, let b 0 be the unique solution to
b VaR θ ( X ) ( 1 + θ ) VaR θ ( X ) b S X ( x ) d x = K
subjected to b > VaR θ ( X ) . Differentiating both sides with respect to K yields
b K = 1 1 ( 1 + θ ) S X ( b ) > 0 ,
which implies that b is increasing with respect to K. If b 0 < VaR α ( X ) , the equation β ( a ) = VaR α ( X ) must have two solutions in , VaR α ( X ) , denoted by a 0 and a 1 with a 0 < a 1 . Furthermore, let b 1 = β VaR α ( X ) , then b 1 > VaR α ( X ) .
Theorem 2.
For any confidence level α p 0 , 1 , the optimal decision in (15) for the insurer is given as follows according to the sign of δ.
(i) 
For the case of δ < 0 ,
(a) 
If b 0 VaR α ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR θ ( X ) + x b 0 + and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) K + h b 0 ,
where h ( b ) = Δ b ω 1 α VaR α ( X ) b S X ( x ) d x ;
(b) 
If b 0 < VaR α ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = ( x r ) + x β ( r ) + with r max 0 , a 0 , a 1 , and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) K .
(ii) 
For the case of δ = 0 ,
(a) 
If α = θ , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR α ( X ) + ( x r ) + for any r VaR α ( X ) , b 1 , and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) ;
(b) 
If α > θ and b 0 VaR α ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR θ ( X ) + ( x r ) + for any r VaR θ ( X ) , b 0 , and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x ;
(c) 
If α > θ and b 0 < VaR α ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = ( x r ) + x β ( r ) + with r max 0 , a 0 , a 1 , and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) K .
(iii) 
For the case of δ > 0 ,
(a) 
If α θ , the minimum of LVaR α T I ( X ) is attained at I ( x ) = 0 and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) ;
(b) 
If α > θ and b 0 VaR α ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR θ ( X ) + x VaR α ( X ) + , and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x ;
(c) 
If α > θ and b 0 < VaR α ( X ) , the minimum of LVaR α T I ( X ) is attained at I ( x ) = ( x r ) + x β ( r ) + with r max 0 , a 0 , a 1 , and
min I F 2 LVaR α T I ( X ) = ω Ψ ( α ) + VaR α ( X ) K .
Remark 6.
Theorem 2 shows that for δ > 0 larger safety loading factor θ leads to no reinsurance. Otherwise, the two layer reinsurance is optimal.

4. Numerical Illustrations

In this section, we present some examples to illustrate the results relying on Theorems 1 and 2. We also compare them with the results in Lu et al. (2016) by setting different confidence levels and different weight coefficients.
In the first two examples, we consider that X follows from an exponential distribution or a normal distribution. Both examples show that under the optimal reinsurance policy larger weight coefficients lead to less ceded losses (or fewer reinsurance premiums), and, hence, the total risks measured by the LVaR s lie somewhere between that measured by the VaR and the TVaR . As a new family of risk measures, LVaR provides flexible risk assessments. For example, the conservative insurer may prefer the larger weight coefficient in practice. Additionally, most risk managers would argue that the heavy-tailed distribution is riskier than the light-tailed distribution since larger losses are more likely to occur when the distribution is heavy-tailed. Therefore, we consider cases of heavy-tailed distributions presented in Examples 3–5. From tables given later, we can conclude that the absolute values of the difference between VaR α T I ( X ) and TVaR α T I ( X ) for heavy-tailed distributions are larger than those for light-tailed distributions. Moreover, the higher confidence level leads to the larger absolute value of the difference. In such cases, the new family of flexible risk measures LVaR permits risk managers to seek an equilibrium between different demands better. Also, we can see that for cases of heavy-tailed distributions the optimal ceded losses are larger, in other words, the insurer needs to hold less losses and pay more to the reinsurer for larger ceded losses.
Example 1.
We assume that X is an exponential random variable with survival function S X ( x ) = e 0.01 x . The optimal reinsurance strategy under risk measures with both types of constraints are given in Table 1 and Table 2 by setting θ = 3 , L = 120 , and K = 160 , where the risk measure ρ represents VaR , LVaR , and TVaR . ρ α T I ( X ) denotes the risk under the optimal reinsurance strategy I ( x ) and given confidence level α.
Example 2.
Suppose that X N ( 40 , 100 ) . The optimal reinsurance solutions under risk measures with both types of constraints are given in Table 3 and Table 4 by setting θ = 4 , L = 100 , and K = 120 , where the risk measure ρ represents VaR , LVaR , and TVaR . ρ α T I ( X ) denotes the risk under the optimal reinsurance strategy I ( x ) and given confidence level α.
Example 3.
Suppose that X follows Pareto distribution with F ( x ) = 1 x σ 1 / γ , 0 < σ x . For γ = 1 / 3 , σ = 120 , the optimal reinsurance solutions under risk measures with both types of constraints are given in Table 5 and Table 6 by setting θ = 4 , L = 100 , and K = 140 , where the risk measure ρ represents VaR , LVaR , and TVaR . ρ α T I ( X ) denotes the risk under the optimal reinsurance strategy I ( x ) and given confidence level α.
Example 4.
We suppose that X follows Fréchet distribution F ( x ) = exp x μ σ 1 / γ with μ = 5 , γ = 1 / 3 , and σ = 50 . The optimal reinsurance solutions under risk measures with both types of constraints are given in Table 7 and Table 8 by setting θ = 3 , L = 100 , and K = 120 , where the risk measure ρ represents VaR , LVaR , and TVaR . ρ α T I ( X ) denotes the risk under the optimal reinsurance strategy I ( x ) and given confidence level α.
Example 5.
We suppose that X has Burr distribution F ( x ) = 1 η + x ρ α η 1 / ρ with α = 1 , ρ = 3 , and η = 40 . The optimal reinsurance solutions under risk measures with both types of constraints are given in Table 9 and Table 10 by setting θ = 4 , L = 120 and K = 160 , where the risk measure ρ represents VaR , LVaR , and TVaR . ρ α T I ( X ) denotes the risk under the optimal reinsurance strategy I ( x ) and given confidence level α.

5. Conclusions

It is well known that VaR and TVaR are the most popular risk measures and both of them have been widely used in the literature (see, e.g., Cai and Tan 2007; Cai et al. 2008; Lu et al. 2016). However, these two risk measures were generally considered separately. By noticing that TVaR α VaR α and the size of capital reserves may be significantly different based on VaR and TVaR , we have proposed a new family of risk measures named LVaR which is a linear combination of VaR and TVaR . This new risk measure considers VaR and TVaR simultaneously and helps us to obtain a risk assessment by LVaR . A major issue in a reinsurance contract is to guarantee a balance between the ceded risk and the contract premium. From the perspective of the insurer, the problem is how to choose a ceded loss function such that the total losses are as small as possible (or the total benefits are as large as possible). To this end, we have revisited the optimal reinsurance problem by minimizing the LVaR of the total risk of the insurer when the reinsurer’s risk exposure has an upper limit.
In this paper, the optimal reinsurance problem has been studied between one insurer and one reinsurer. However, as introduced by Boonen et al. (2021), the setting with one insurer and multiple reinsurers may be more realistic in practice. This inspires our future work to consider the case with multiple reinsurers.
The most important results are presented in Theorems 1 and 2, which provide the solutions of our optimal reinsurance model. It is shown that the two-layer reinsurance is always the optimal reinsurance strategy under two types of constraints. Furthermore, we have found that the minimums of LVaR of an insurer’s total risks are larger than those of VaR and smaller than those of TVaR . The same holds for deductibles. In numerical illustrations, we have noted significant differences in the size of capital reserves depending on adopted weight coefficients, and moreover, higher capital reserves are obtained for larger weight coefficients.
By introducing the weighted coefficient ω , the new risk measure LVaR may help financial institutions, such as insurance companies, to quantify risk based on different situations since the weight coefficient ω reflects the user’s attitude towards risk. More precisely, the larger ω means the higher sensitivity to the severity of losses exceeding VaR . Most importantly, if a insurer intends to transfer some part of risk to one reinsurance, our work may provide a way to determine the size of the transferred risk.

Author Contributions

S.N. wrote Section 4. Q.X. and Z.P. wrote the remaining sections. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the Editor and the three referees for careful reading and comments which greatly improved the paper.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Proof Lemma 2.
According to the sign of δ , we consider δ 0 and δ < 0 in turn.
In the case of δ 0 , for any I ( x ) F 1 , let
g ( x ; a , b ) = x e α + x VaR α ( X ) + ,
where e α = VaR α ( X ) I VaR α ( X ) . Obviously, e α , VaR α ( X ) D 1 , and hence g ( x ; a , b ) G 1 . Furthermore, we have g x ; e α , VaR α ( X ) = 0 I ( x ) for 0 x < e α and g x ; e α , VaR α ( X ) = I VaR α ( X ) I ( x ) for x VaR α ( X ) . For e α x < VaR α ( X ) , by using the property of Lipschitz continuity of I ( x ) , we have g x ; e α , VaR α ( X ) = x VaR α ( X ) + I VaR α ( X ) I ( x ) . Thus, we conclude that g x ; e α , VaR α ( X ) I ( x ) for x 0 . Noting that g VaR α ( X ) ; e α , VaR α ( X ) = I VaR α ( X ) , by (11), we have LVaR α T g ( X ) LVaR α T I ( X ) . Hence, the desired result follows for δ 0 .
In the case of δ < 0 , for any I ( x ) F 1 , let
g ( x ; a , b ) = x e α + x ( e α + L ) + ,
where e α is defined as above. We can easily check that VaR α ( X ) e α + L and, hence, g ( x ; a , b ) G 1 . We can prove that g x ; e α , e α + L = 0 I ( x ) for 0 x < e α , and g x ; e α , e α + L = L I ( x ) for x e α + L . If e α x < VaR α ( X ) , we have from Lipschitz continuity of I ( x ) that
g x ; e α , e α + L = x VaR α ( X ) + I VaR α ( X ) I ( x ) .
Similarly, if VaR α ( X ) x < e α + L , we have
g x ; e α , e α + L = x VaR α ( X ) + I VaR α ( X ) I ( x ) .
Hence, by applying (11), we have the desired result for δ < 0 . The proof is complete. □
Proof Theorem 1.
By Lemma 2, we can solve the optimal problem (14) instead of (4). Applying the definition of g ( x ) in (13), we have
g VaR α ( X ) = VaR α ( X ) a ,
0 VaR α ( X ) g ( x ) d F X ( x ) = a VaR α ( X ) ( x a ) d F X ( x ) = a VaR α ( X ) S X ( t ) d t VaR α ( X ) a S X VaR α ( X ) ,
and
VaR α ( X ) g ( x ) d F X ( x ) = VaR α ( X ) b ( x a ) d F X ( x ) + b ( b a ) d F X ( x ) = VaR α ( X ) b S X ( t ) d t + VaR α ( X ) a S X VaR α ( X ) .
Combining (A1)–(A3) with (11), we obtain
LVaR α T g ( X ) = ω Ψ ( α ) + VaR α ( X ) ( 1 ω ) VaR α ( X ) a + ( 1 + θ ) a VaR α ( X ) S X ( t ) d t ( 1 + θ ) VaR α ( X ) a S X VaR α ( X ) + δ VaR α ( X ) b S X ( t ) d t + δ VaR α ( X ) a S X VaR α ( X ) = ω Ψ ( α ) + VaR α ( X ) ( 1 ω ) VaR α ( X ) a + ( 1 + θ ) a VaR α ( X ) S X ( t ) d t + δ VaR α ( X ) b S X ( t ) d t ω VaR α ( X ) a = ω Ψ ( α ) + a + ( 1 + θ ) a VaR α ( X ) S X ( t ) d t + δ VaR α ( X ) b S X ( t ) d t = Δ ϕ ( a , b ) .
The rest of the work is to minimize (A4) according the sign of δ .
(i)
If δ < 0 (which implies α > θ ), taking the partial derivative with respect to b on ϕ ( a , b ) yields
ϕ ( a , b ) b = δ S X ( b ) < 0 .
From (A5) and (12), for any α p 0 , 1 and ( a , b ) D 1 , we always have ϕ ( a , b ) ϕ ( a , a + L ) . Note that ( a , b ) D 1 implies ( a , a + L ) D 1 , so that the minimum of LVaR α T g ( X ) must be attained at ( a , b ) with b = a + L . As a result, it suffices to solve the following optimal problem
min a max 0 , VaR α ( X ) L , VaR α ( X ) φ ( a ) ,
where
φ ( a ) = ϕ ( a , a + L ) = ω Ψ ( α ) + a + ( 1 + θ ) a VaR α ( X ) S X ( t ) d t + δ VaR α ( X ) a + L S X ( t ) d t .
By taking the first two derivatives of φ ( a ) , we have
φ ( a ) = 1 ( 1 + θ ) S X ( a ) + δ S X ( a + L )
and
φ ( a ) = ( 1 + θ ) f X ( a ) δ S X ( a + L ) > 0 ,
which implies that φ ( a ) is convex. In addition, we can obtain
φ VaR θ ( X ) = 1 ( 1 + θ ) 1 θ + δ S X VaR θ ( X ) + L = δ S X VaR θ ( X ) + L < 0 ,
and
φ VaR α ( X ) = 1 ( 1 + θ ) ( 1 α ) + δ S X VaR α ( X ) + L 1 ( 1 + θ ) ( 1 α ) + δ S X VaR α ( X ) = 1 ω 0 .
Let a 0 be the solution to φ ( a ) = 0 . We consider the following two cases.
(a)
If VaR α ( X ) L a 0 , we have φ VaR α ( X ) L 0 . Combining with (A7) and (A8), we conclude that the minimum of φ ( a ) must be attained at a 0 with
a 0 max VaR θ ( X ) , VaR α ( X ) L , VaR α ( X ) .
(b)
If VaR α ( X ) L > a 0 , we obtain φ VaR α ( X ) L > 0 . Then by the monotonicity of φ ( a ) in VaR α ( X ) L , VaR α ( X ) , we find that φ ( a ) attains its minimum at VaR α ( X ) L .
Hence, the proof of (i) is complete.
(ii)
If δ = 0 (which implies α θ ), we have
LVaR α T g ( X ) = ω Ψ ( α ) + a + ( 1 + θ ) a VaR α ( X ) S X ( t ) d t = Δ η ( a )
with a max 0 , VaR α ( X ) L , VaR α ( X ) . Taking the first derivatives of η ( a ) , we obtain
η ( a ) = 1 ( 1 + θ ) S X ( a ) .
From the monotonicity of η ( a ) , we find that the minimum of η ( a ) is attained at VaR θ ( X ) for VaR α ( X ) L VaR θ ( X ) , as well as at VaR α ( X ) L for VaR α ( X ) L > VaR θ ( X ) . Equivalently, if VaR α ( X ) L VaR θ ( X ) , then LVaR α T I ( X ) attains its minimum at I ( x ) = x VaR θ ( X ) + ( x r ) + , where r is any real number in VaR α ( X ) , VaR θ ( X ) + L . If VaR α ( X ) L > VaR θ ( X ) , LVaR α T I ( X ) attains its minimum at I ( x ) = x VaR α ( X ) L + x VaR α ( X ) + .
(iii)
If δ > 0 , taking the partial derivative of ϕ ( a , b ) with respect to b yields
ϕ ( a , b ) b = δ S X ( b ) > 0 .
From (A9) and the properties of D 1 , for any α p 0 , 1 and ( a , b ) D 1 , we have ϕ ( a , b ) ϕ a , VaR α ( X ) . Thus, the minimum of LVaR α T I ( X ) must be attained at a , VaR α ( X ) D 1 . Consequently, it suffices to solve the following optimal problem
min a max 0 , VaR α ( X ) L , VaR α ( X ) γ ( a ) ,
where
γ ( a ) = ω Ψ ( α ) + a + ( 1 + θ ) a VaR α ( X ) S X ( t ) d t .
Taking the first derivatives of γ ( a ) yields
γ ( a ) = 1 ( 1 + θ ) S X ( a ) .
We consider the following cases.
(a)
If α θ , then from (A10), the minimum of LVaR α T I ( X ) is attained at I ( x ) = 0 .
(b)
If α > θ and VaR α ( X ) L VaR θ ( X ) , from (A10), LVaR α T I ( X ) attains its minimum at I ( x ) = x VaR θ ( X ) + x VaR α ( X ) + .
(c)
If α > θ and VaR α ( X ) L > VaR θ ( X ) , then the minimum of LVaR α T I ( X ) is attained at I ( x ) = x VaR α ( X ) L + x VaR α ( X ) + .
The proof is complete. □
Proof Lemma 3.
For any I ( x ) F 2 , let g 1 ( x ) = x e α + x VaR α ( X ) + with e α = VaR α ( X ) I VaR α ( X ) , and g 2 ( x ) = x e α + x e α M + and g 3 ( x ) = x ( x M ) + with M = K + ( 1 + θ ) E I ( X ) . By similar arguments to Lemma 2, we can show that g 1 ( x ) I ( x ) g 3 ( x ) and E g 1 ( X ) E I ( X ) E g 3 ( X ) .
We consider the following two cases, E g 2 ( X ) E I ( X ) and E g 2 ( X ) > E I ( X ) . In the case of E g 2 ( X ) E I ( X ) , for any function g 4 ( x ; a ) = ( x a ) + x a M + , we have
E g 4 ( X ; a ) = a a + M S X ( x ) d x .
Thus,
E g 4 ( X ; a ) a = S X ( a + M ) S X ( a ) < 0 ,
which indicates that E g 4 ( X ; a ) is decreasing with respect to a. From the continuity of E g 4 ( X ; a ) with respect to a, there must exist g ( x ; c ) = ( x c ) + x c M + with 0 c e α , such that E g ( X ) = E I ( X ) and g 2 ( x ) g ( x ) g 3 ( x ) . We can easily verify that g ( x ) G 2 for 0 c e α and g ( x ) g 2 ( x ) I ( x ) for x VaR α ( X ) . Then, by (10), g ( x ) is the desired function, that is, LVaR α T g ( X ) LVaR α T I ( X ) .
Similarly, in the case of E g 2 ( X ) > E I ( X ) , for any function g 5 ( x ; b ) = x e α + ( x b ) + , we obtain that
E g 5 ( X ; a ) = e α b S X ( x ) d x
and
E g 5 ( X ; a ) b = S X ( b ) > 0 ,
so that E g 5 ( X ; a ) is increasing with respect to b. Because E g 5 ( X ; a ) is continuous with respect to a, there must exist g ( x ; k ) = x e α + ( x k ) + with k VaR α ( X ) , e α + M , such that E g ( X ) = E I ( X ) and g 1 ( x ) g ( x ) < g 2 ( x ) . As a result, we have g VaR α ( X ) g 1 VaR α ( X ) = I VaR α ( X ) . If x 0 , VaR α ( X ) , we can easily check that g ( x ) I ( x ) , and combining this with E g ( X ) = E I ( X ) , we obtain VaR α ( X ) g ( x ) d F X ( x ) VaR α ( X ) I ( x ) d F X ( x ) . Furthermore, for k VaR α ( X ) , e α + M , we can find that k e α < M = K + ( 1 + θ ) E I ( X ) = K + ( 1 + θ ) E g ( X ) , which implies that g ( x ) G 2 . Hence, by (10), we have LVaR α T g ( X ) LVaR α T I ( X ) and the proof is complete. □
Proof Theorem 2.
For any g ( x ; a , b ) G 2 , if b VaR α ( X ) , we have g VaR α ( X ) = b a and
VaR α ( X ) g ( x ) d F X ( x ) = VaR α ( X ) ( b a ) d F X ( x ) = ( b a ) ( 1 α ) .
Then, applying (10), we obtain that
LVaR α T g ( X ) = ω Ψ ( α ) + VaR α ( X ) ( 1 ω ) ( b a ) + ( 1 + θ ) E g ( X ) ω ( b a ) = ω Ψ ( α ) + VaR α ( X ) b a ( 1 + θ ) E g ( X ) = ω Ψ ( α ) + VaR α ( X ) ψ ( a , b ) = Δ η 1 ( a , b )
for b VaR α ( X ) , where ψ ( a , b ) = b a ( 1 + θ ) E g ( X ) . If b > VaR α ( X ) , we have that g VaR α ( X ) = VaR α ( X ) a and
VaR α ( X ) g ( x ) d F X ( x ) = VaR α ( X ) b ( x a ) d F X ( x ) + b ( b a ) d F X ( x ) = VaR α ( X ) b a x d t d F X ( x ) + ( b a ) S X ( x ) = a VaR α ( X ) VaR α ( X ) b d F X ( x ) d t + VaR α ( X ) b t b d F X ( x ) d + ( b a ) S X ( x ) = a VaR α ( X ) S X VaR α ( X ) S X ( b ) d t + VaR α ( X ) b S X ( t ) S X ( b ) d t + ( b a ) S X ( x ) = VaR α ( X ) b S X ( t ) d t + VaR α ( X ) a ( 1 α ) .
Let h ( b ) = Δ b ω 1 α VaR α ( X ) b S X ( t ) d t . By (10), we obtain that
LVaR α T g ( X ) = ω Ψ ( α ) + VaR α ( X ) ( 1 ω ) VaR α ( X ) a + ( 1 + θ ) E g ( X ) ω 1 α VaR α ( X ) b S X ( t ) d t ω VaR α ( X ) a = ω Ψ ( α ) + a + ( 1 + θ ) E g ( X ) ω 1 α VaR α ( X ) b S X ( t ) d t = ω Ψ ( α ) b a ( 1 + θ ) E g ( X ) + b ω 1 α VaR α ( X ) b S X ( t ) d t = ω Ψ ( α ) ψ ( a , b ) + h ( b ) = Δ η 2 ( a , b ) .
Thus,
LVaR α T g ( X ) = η 1 ( a , b ) , b VaR α ( X ) , η 2 ( a , b ) , b > VaR α ( X ) .
From (A11) and (A12), we have
η 1 ( a , b ) a = η 2 ( a , b ) a = 1 ( 1 + θ ) S X ( a ) ,
η 1 ( a , b ) b = 1 + ( 1 + θ ) S X ( b ) ,
η 2 ( a , b ) b = δ S X ( b )
and
h ( b ) = 1 ω 1 α S X ( b ) .
For convenience, define
D 3 = Δ ( a , b ) | ( a , b ) D 0 , b VaR θ ( X ) , D 4 = Δ ( a , b ) | ( a , b ) D 0 , b VaR α ( X ) .
The rest is to show the results, in turn, according to δ < 0 , δ = 0 and δ > 0 . First, we consider the case of δ < 0 . If b 0 VaR θ ( X ) , note that δ < 0 implies α > θ . We have from (A14)–(A17) that for any ( a , b ) D 3 ,
η 1 ( a , b ) η 1 ( a , a ) = η 1 VaR θ ( X ) , VaR θ ( X ) > η 1 VaR θ ( X ) , VaR α ( X ) = η 2 VaR θ ( X ) , VaR α ( X ) η 2 VaR θ ( X ) , b 0 = ω Ψ ( α ) K + h β b 0 .
For any ( a , b ) D 4 D 3 , we have
η 1 ( a , b ) η 1 a , VaR α ( X ) = η 2 a , VaR α ( X ) η 2 VaR θ ( X ) , VaR α ( X ) η 2 VaR θ ( X ) , b 0 = ω Ψ ( α ) K + h β b 0 .
For any ( a , b ) D 0 D 4 , we obtain
η 2 ( a , b ) η 2 a , β ( a ) = ω Ψ ( α ) K + h β ( a ) ω Ψ ( α ) K + h β b 0 .
Then, the minimum of LVaR α T I ( X ) is ω Ψ ( α ) K + h β b 0 which is attained at I ( x ) = x VaR θ ( X ) + x b 0 + .
Similarly, if b 0 < VaR θ ( X ) , for any ( a , b ) D 3 ,
η 1 ( a , b ) η 1 ( a , a ) = η 1 VaR θ ( X ) , VaR θ ( X ) η 1 VaR θ ( X ) , b 0 = ω Ψ ( α ) + VaR α ( X ) K .
For any ( a , b ) D 4 D 3 , we obtain
η 1 ( a , b ) η 1 a , min β ( a ) , VaR α ( X ) η 1 r , β ( r ) = ω Ψ ( α ) + VaR α ( X ) K ,
where r is any real number in max 0 , a 0 , a 1 . For any ( a , b ) D 0 D 4 , we have
η 2 ( a , b ) η 2 a , β ( a ) = ω Ψ ( α ) K + h β ( a ) > ω Ψ ( α ) K + h VaR α ( X ) = ω Ψ ( α ) + VaR α ( X ) K .
Hence, the minimum of LVaR α T I ( X ) is attained at I ( x ) = ( x r ) + x β ( r ) + with r max 0 , a 0 , a 1 as b 0 < VaR α ( X ) . The proof for the case of δ < 0 is complete.
For the case of δ = 0 , note that δ = 0 implies α θ . If α = θ , we have b 0 VaR θ ( X ) = VaR α ( X ) , then for any ( a , b ) D 4 ,
η 1 ( a , b ) η 1 ( a , a ) = η 1 VaR α ( X ) , VaR α ( X ) = η 2 VaR α ( X ) , VaR α ( X ) = ω Ψ ( α ) + VaR α ( X ) .
For ( a , b ) D 0 D 4 ,
η 2 ( a , b ) η 2 VaR θ ( X ) , b = η 2 VaR α , b = ω Ψ ( α ) + VaR α ( X ) .
Hence, the minimum of LVaR α T I ( X ) is ω Ψ ( α ) + VaR α ( X ) which is attained at I ( x ) = x VaR α ( X ) + ( x r ) + with r VaR α ( X ) , β VaR α ( X ) .
If α > θ and b 0 VaR α ( X ) , for any ( a , b ) D 3 ,
η 1 ( a , b ) η 1 ( a , a ) = η 1 VaR θ ( X ) , VaR θ ( X ) > η 1 VaR θ ( X ) , VaR α ( X ) = η 2 VaR θ ( X ) , VaR α ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x .
For any ( a , b ) D 4 D 3 ,
η 1 ( a , b ) η 1 a , VaR α ( X ) = η 2 a , VaR α ( X ) η 2 VaR θ ( X ) , VaR α ( X ) = η 2 VaR θ ( X ) , r = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x ,
where r is any real number in VaR θ ( X ) , VaR α ( X ) . For any ( a , b ) D 0 D 4 ,
η 2 ( a , b ) η 2 VaR θ ( X ) , b = η 2 VaR θ ( X ) , r = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x ,
where r is any real number in VaR α ( X ) , b 0 . Hence, the minimum of LVaR α T I ( X ) is
ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x
and is attained at I ( x ) = x VaR θ ( X ) + ( x r ) + , where r is any real number in VaR θ ( X ) , b 0 .
If α > θ and b 0 < VaR α ( X ) , for any ( a , b ) D 3 ,
η 1 ( a , b ) η 1 ( a , a ) = η 1 VaR θ ( X ) , VaR θ ( X ) > η 1 VaR θ ( X ) , b 0 = ω Ψ ( α ) + VaR α ( X ) K .
For any ( a , b ) D 4 D 3 ,
η 1 ( a , b ) η 1 a , min β ( a ) , VaR α ( X ) η 1 r , β ( r ) = ω Ψ ( α ) + VaR α ( X ) K ,
where r is any real number in max 0 , a 0 , a 1 . For any ( a , b ) D 0 D 4 ,
η 2 ( a , b ) = η 2 a , β ( a ) = ω Ψ ( α ) K + h β ( a ) > ω Ψ ( α ) K + h β VaR α ( X ) = ω Ψ ( α ) + VaR α ( X ) K .
Hence, the minimum of LVaR α T I ( X ) is ω Ψ ( α ) + VaR α ( X ) K which is attained at I ( x ) = ( x r ) + x β ( r ) + , where r is any real number in max 0 , a 0 , a 1 .
For the case of δ > 0 , if α θ , for any ( a , b ) D 4 ,
η 1 ( a , b ) η 1 ( a , a ) = ω Ψ ( α ) + VaR α ( X ) .
For any ( a , b ) D 0 D 4 ,
η 2 ( a , b ) η 2 a , VaR α ( X ) η 2 VaR α ( X ) , VaR α ( X ) = ω Ψ ( α ) + VaR α ( X ) .
Hence, the minimum of L VaR α T I ( X ) is ω Ψ ( α ) + VaR α ( X ) as α θ and is attained at I ( x ) = 0 .
If α > θ and b 0 VaR α ( X ) , for any ( a , b ) D 3 ,
η 1 ( a , b ) η 1 ( a , a ) = η 1 VaR θ ( X ) , VaR θ ( X ) > η 1 VaR θ ( X ) , VaR α ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x .
For any ( a , b ) D 4 D 3 ,
η 1 ( a , b ) η 1 a , VaR α ( X ) η 1 VaR θ ( X ) , VaR α ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x .
For any ( a , b ) D 0 D 4 ,
η 2 ( a , b ) η 2 VaR θ ( X ) , b η 2 VaR θ ( X ) , VaR α ( X ) = η 1 VaR θ ( X ) , VaR α ( X ) = ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x .
Hence, the minimum of LVaR α T I ( X ) is
ω Ψ ( α ) + VaR θ ( X ) + ( 1 + θ ) VaR θ ( X ) VaR α ( X ) S X ( x ) d x
and is attained at I ( x ) = x VaR θ ( X ) + x VaR α ( X ) + .
Finally, if α > θ and b 0 > VaR α ( X ) , for any ( a , b ) D 3 ,
η 1 ( a , b ) η 1 ( a , a ) = η 1 VaR θ ( X ) , VaR θ ( X ) > η 1 VaR θ ( X ) , b 0 = ω Ψ ( α ) + VaR α ( X ) K .
For any ( a , b ) D 4 D 3 ,
η 1 ( a , b ) η 1 a , min β ( a ) , VaR α ( X ) η 1 r , β ( r ) = ω Ψ ( α ) + VaR α ( X ) K ,
where r is any real number in max 0 , a 0 , a 1 . For any ( a , b ) D 0 D 4 ,
η 2 ( a , b ) > η 2 a , VaR α ( X ) = η 1 a , VaR α ( X ) η 1 VaR θ ( X ) , VaR α ( X ) = ω Ψ ( α ) + VaR α ( X ) ψ VaR θ ( X ) , VaR α ( X ) > ω Ψ ( α ) + VaR α ( X ) K .
Hence, the minimum of LVaR α T I ( X ) is ω Ψ ( α ) + VaR α ( X ) K and is attained at I ( x ) = ( x r ) + x β ( r ) + , where r is any real number in max 0 , a 0 , a 1 .
Therefore, by combining the above analyses, we complete the proof of Theorem 2. □

References

  1. Acerbi, Carlo. 2002. Spectral measures of risk: A coherent representation of subjective risk aversion. Journal of Banking and Finance 26: 1505–18. [Google Scholar] [CrossRef] [Green Version]
  2. Acerbi, Carlo, and Dirk Tasche. 2002. On the coherence of expected shortfall. Journal of Banking and Finance 26: 1487–503. [Google Scholar] [CrossRef] [Green Version]
  3. Albrecher, Hansjörg, Jan Beirlant, and Jozef L. Teugels. 2017. Reinsurance: Actuarial and Statistical Aspects. Chichester: John Wiley and Sons. [Google Scholar]
  4. Arrow, Kenneth J. 1963. Uncertainty and the welfare economics of medical care. American Economic Review 53: 941–73. [Google Scholar]
  5. Artzner, Philippe, Freddy Delbaen, Jean-Marc Eber, and David Heath. 1999. Coherent measures of risk. Mathematical Finance 9: 203–28. [Google Scholar] [CrossRef]
  6. Balbás, Alejandro, Beatriz Balbás, Raquel Balbás, and Antonio Heras. 2022. Risk transference constraints in optimal reinsurance. Insurance: Mathematics and Economics 103: 27–40. [Google Scholar] [CrossRef]
  7. Balbás, Alejandro, Beatriz Balbás, and Antonio Heras. 2009. Optimal reinsurance with general risk measures. Insurance: Mathematics and Economics 44: 374–84. [Google Scholar] [CrossRef] [Green Version]
  8. Bellini, Fabio, and Ilaria Peri. 2022. An axiomatization of Λ-quantiles. SIAM Journal on Financial Mathematics 13: SC26–SC38. [Google Scholar] [CrossRef]
  9. Boonen, Tim J., Ken Seng Tan, and Sheng Chao Zhuang. 2016. Pricing in reinsurance bargaining with comonotonic additive utility functions. ASTIN Bulletin 46: 507–30. [Google Scholar] [CrossRef] [Green Version]
  10. Boonen, Tim J., Ken Seng Tan, and Sheng Chao Zhuang. 2021. Optimal reinsurance with multiple reinsurers: Competitive pricing and coalition stability. Insurance: Mathematics and Economics 101: 302–19. [Google Scholar] [CrossRef]
  11. Borch, Karl. 1960. An attempt to determine the optimum amount of stop loss reinsurance. In Transactions of the 16th International Congress of Actuaries I. Brussels: Uden forlag, pp. 597–610. [Google Scholar]
  12. Cai, Jun, and Ken Seng Tan. 2007. Optimal retention for a stop-loss reinsurance under the VaR and CTE risk measure. Astin Bulletin 37: 93–112. [Google Scholar] [CrossRef] [Green Version]
  13. Cai, Jun, Ken Seng Tan, Chengguo Weng, and Yi Zhang. 2008. Optimal reinsurance under VaR and CTE risk measures. Insurance: Mathematics and Economics 43: 185–96. [Google Scholar] [CrossRef]
  14. Charpentier, Arthur. 2008. Insurability of climate risks. The Geneva Papers 33: 91–109. [Google Scholar] [CrossRef] [Green Version]
  15. Cheung, Ka Chun. 2010. Optimal reinsurance revisited—A geometric approach. Astin Bulletin 40: 221–39. [Google Scholar] [CrossRef]
  16. Chi, Yichun, and Ken Seng Tan. 2011. Optimal reinsurance under VaR and CVaR risk measures: A simplified approach. Astin Bulletin 41: 487–509. [Google Scholar]
  17. Chi, Yichun, and Ken Seng Tan. 2013. Optimal reinsurance with general premium principles. Insurance: Mathematics and Economics 52: 180–9. [Google Scholar] [CrossRef]
  18. Cummins, J. David, and Oliver Mahul. 2004. The demand for insurance with an upper limit on coverage. Journal of Risk and Insurance 71: 253–64. [Google Scholar] [CrossRef]
  19. Dhaene, J., M. Denuit, M.J. Goovaerts, R. Kaas, and D. Vyncke. 2002. The concept of comonotonicity in actuarial science and finance: Theory. Insurance: Mathematics and Economics 31: 3–33. [Google Scholar] [CrossRef]
  20. Dowd, Kevin, John Cotter, and Ghulam Sorwar. 2008. Spectral risk measures: Properties and limitations. Journal of Financial Services Research 34: 61–75. [Google Scholar] [CrossRef] [Green Version]
  21. Embrechts, Paul, and Marius Hofert. 2014. Statistics and quantitative risk management for banking and insurance. Annual Review of Statistics and Its Application 1: 493–514. [Google Scholar] [CrossRef] [Green Version]
  22. Frittelli, Marco, Marco Maggis, and Ilaria Peri. 2014. Risk measures on 𝒫( R ) and value at risk with probability/loss function. Mathematical Finance 24: 442–63. [Google Scholar] [CrossRef] [Green Version]
  23. Guerra, Manuel, and Maria de Lourdes Centeno. 2010. Optimal reinsurance for variance related premium calculation principles. Astin Bulletin 40: 97–121. [Google Scholar] [CrossRef]
  24. Kaluszka, Marek. 2001. Optimal reinsurance under mean-variance premium principles. Insurance: Mathematics and Economics 28: 61–67. [Google Scholar] [CrossRef]
  25. Kaluszka, Marek. 2005. Optimal reinsurance under convex principles of premium calculation. Insurance: Mathematics and Economics 36: 375–98. [Google Scholar] [CrossRef]
  26. Kaluszka, Marek, and Andrzej Okolewski. 2008. An extension of Arrow’s result on optimal reinsurance contract. Journal of Risk and Insurance 75: 275–88. [Google Scholar] [CrossRef]
  27. Liang, Zhibin, and Erhan Bayraktar. 2014. Optimal reinsurance and investment with unobservable claim size and intensity. Insurance: Mathematics and Economics 55: 156–66. [Google Scholar] [CrossRef]
  28. Lu, Zhi Yi, Li Li Meng, Yujin Wang, and Qingjie Shen. 2016. Optimal reinsurance under VaR and TVaR risk measures in the presence of reinsurer’s risk limit. Insurance: Mathematics and Economics 68: 92–100. [Google Scholar] [CrossRef]
  29. McNeil, Alexander J., Rüdiger Frey, and Paul Embrechts. 2015. Quantitative Risk Management: Concepts, Techniques and Tools. Princeton: Princeton University Press. [Google Scholar]
  30. Rockafellar, R. Tyrrell, and Stan Uryasev. 2013. The fundamental risk quadrangle in risk management, optimization and statistical estimation. Surveys in Operations Research and Management Science 18: 33–53. [Google Scholar] [CrossRef]
  31. Tan, Ken Seng, Chengguo Weng, and Yi Zhang. 2011. Optimality of general reinsurance contracts under CTE risk measure. Insurance: Mathematics and Economics 49: 175–87. [Google Scholar] [CrossRef]
  32. Young, Virginia R. 1999. Optimal insurance under Wang’s premium principle. Insurance: Mathematics and Economics 25: 109–22. [Google Scholar] [CrossRef]
  33. Zhang, Xin, and Tak Kuen Siu. 2012. On optimal proportional reinsurance and investment in a Markovian regime-switching economy. Acta Mathematica Sinica 28: 67–82. [Google Scholar] [CrossRef]
  34. Zhou, Chunyang, Wenfeng Wu, and Chongfeng Wu. 2010. Optimal insurance in the presence of insurer’s loss limit. Insurance: Mathematics and Economics 46: 300–7. [Google Scholar] [CrossRef]
Table 1. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X F ( x ) = 1 e 0.01 x , x > 0 , θ = 3 , and L = 120 .
Table 1. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X F ( x ) = 1 e 0.01 x , x > 0 , θ = 3 , and L = 120 .
α The Risk Measure ρ        ρ α T I ( X ) I ( x )
VaR       198.629 ( x 138.629 ) + ( x 230.259 ) +
LVaR ( ω = 0.2 )       218.629 ( x 138.629 ) + ( x 230.259 ) +
0.90 LVaR ( ω = 0.5 )       245.889 ( x 145.889 ) + ( x 265.889 ) +
LVaR ( ω = 0.8 )       264.958 ( x 164.958 ) + ( x 284.958 ) +
TVaR       275.909 ( x 175.909 ) + ( x 295.909 ) +
VaR       225.976 ( x 179.573 ) + ( x 299.573 ) +
LVaR ( ω = 0.2 )       245.976 ( x 179.573 ) + ( x 299.573 ) +
0.95 LVaR ( ω = 0.5 )       275.976 ( x 179.573 ) + ( x 299.573 ) +
LVaR ( ω = 0.8 )       303.003 ( x 203.003 ) + ( x 323.003 ) +
TVaR       317.692 ( x 217.692 ) + ( x 337.692 ) +
VaR       258.497 ( x 230.656 ) + ( x 350.656 ) +
LVaR ( ω = 0.2 )       278.497 ( x 230.656 ) + ( x 350.656 ) +
0.97 LVaR ( ω = 0.5 )       308.497 ( x 230.656 ) + ( x 350.656 ) +
LVaR ( ω = 0.8 )       338.205 ( x 238.205 ) + ( x 358.205 ) +
TVaR       355.218 ( x 255.218 ) + ( x 375.218 ) +
VaR       349.798 ( x 340.517 ) + ( x 460.517 ) +
LVaR ( ω = 0.2 )       369.798 ( x 340.517 ) + ( x 460.517 ) +
0.99 LVaR ( ω = 0.5 )       399.798 ( x 340.517 ) + ( x 460.517 ) +
LVaR ( ω = 0.8 )       429.798 ( x 340.517 ) + ( x 460.517 ) +
TVaR       449.392 ( x 349.392 ) + ( x 469.392 ) +
VaR       571.704 ( x 570.776 ) + ( x 690.776 ) +
LVaR ( ω = 0.2 )       591.703 ( x 570.776 ) + ( x 690.776 ) +
0.999 LVaR ( ω = 0.5 )       621.703 ( x 570.776 ) + ( x 690.776 ) +
LVaR ( ω = 0.8 )       651.703 ( x 570.776 ) + ( x 690.776 ) +
TVaR       671.698 ( x 571.699 ) + ( x 691.699 ) +
Table 2. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X F ( x ) = 1 e 0.01 x , x > 0 , θ = 3 , and K = 160 .
Table 2. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X F ( x ) = 1 e 0.01 x , x > 0 , θ = 3 , and K = 160 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 198.629 ( x 138.629 ) + ( x 230.259 ) +
LVaR ( ω = 0.2 ) 218.629 ( x 138.629 ) + ( x 230.259 ) +
0.90 LVaR ( ω = 0.5 ) 240.642 ( x 138.629 ) + ( x 390.580 ) +
LVaR ( ω = 0.8 ) 246.679 ( x 138.629 ) + ( x 390.580 ) +
TVaR 250.704 ( x 138.629 ) + ( x 390.580 ) +
VaR 218.629 ( x 138.629 ) + ( x 299.573 ) +
LVaR ( ω = 0.2 ) 238.629 ( x 138.629 ) + ( x r ) + with r [ 138.629 , 390.580 ]
0.95 LVaR ( ω = 0.5 ) 250.704 ( x 138.629 ) + ( x 390.580 ) +
LVaR ( ω = 0.8 ) 262.779 ( x 138.629 ) + ( x 390.580 ) +
TVaR 270.829 ( x 138.629 ) + ( x 390.580 ) +
VaR 226.629 ( x 138.629 ) + ( x 350.656 ) +
LVaR ( ω = 0.2 ) 243.996 ( x 138.629 ) + ( x 390.580 ) +
0.97 LVaR ( ω = 0.5 ) 264.121 ( x 138.629 ) + ( x 390.580 ) +
LVaR ( ω = 0.8 ) 284.246 ( x 138.629 ) + ( x 390.580 ) +
TVaR 297.663 ( x 138.629 ) + ( x 390.580 ) +
VaR 300.517 ( x r ) + x β ( r ) + with r [ 42.186 , 280.255 ] and β ( r ) [ 390.580 , 460.517 ]
LVaR ( ω = 0.2 ) 320.517 ( x r ) + x β ( r ) + with r [ 42.186 , 280.255 ] and β ( r ) [ 390.580 , 460.517 ]
0.99 LVaR ( ω = 0.5 ) 350.517 ( x r ) + x β ( r ) + with r [ 42.186 , 280.255 ] and β ( r ) [ 390.580 , 460.517 ]
LVaR ( ω = 0.8 ) 380.517 ( x r ) + x β ( r ) + with r [ 42.186 , 280.255 ] and β ( r ) [ 390.580 , 460.517 ]
TVaR 400.517 ( x r ) + x β ( r ) + with r [ 42.186 , 280.255 ] and β ( r ) [ 390.580 , 460.517 ]
VaR 530.776 ( x r ) + x β ( r ) + with r [ 0 , 529.162 ] and β ( r ) [ 390.580 , 690.776 ]
LVaR ( ω = 0.2 ) 550.775 ( x r ) + x β ( r ) + with r [ 0 , 529.162 ] and β ( r ) [ 390.580 , 690.776 ]
0.999 LVaR ( ω = 0.5 ) 580.775 ( x r ) + x β ( r ) + with r [ 0 , 529.162 ] and β ( r ) [ 390.580 , 690.776 ]
LVaR ( ω = 0.8 ) 610.775 ( x r ) + x β ( r ) + with r [ 0 , 529.162 ] and β ( r ) [ 390.580 , 690.776 ]
TVaR 630.774 ( x r ) + x β ( r ) + with r [ 0 , 529.162 ] and β ( r ) [ 390.580 , 690.776 ]
Table 3. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X N ( 40 , 100 ) , θ = 4 , and L = 100 .
Table 3. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X N ( 40 , 100 ) , θ = 4 , and L = 100 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 156.309 ( x 124.162 ) + ( x 168.155 ) +
LVaR ( ω = 0.2 ) 165.778 ( x 124.162 ) + ( x 168.155 ) +
0.90 LVaR ( ω = 0.5 ) 179.981 ( x 124.162 ) + ( x r ) + with r [ 168.155 , 224.162 ]
LVaR ( ω = 0.8 ) 183.531 ( x 130.423 ) + ( x 230.423 ) +
TVaR 185.618 ( x 133.936 ) + ( x 233.936 ) +
VaR 169.535 ( x 124.162 ) + ( x 204.485 ) +
LVaR ( ω = 0.2 ) 177.892 ( x 124.162 ) + ( x 204.485 ) +
0.95 LVaR ( ω = 0.5 ) 185.618 ( x 133.937 ) + ( x 233.937 ) +
LVaR ( ω = 0.8 ) 190.944 ( x 142.453 ) + ( x 242.453 ) +
TVaR 193.931 ( x 146.997 ) + ( x 246.997 ) +
VaR 174.278 ( x 128.079 ) + ( x 228.079 ) +
LVaR ( ω = 0.2 ) 182.024 ( x 128.079 ) + ( x 228.079 ) +
0.97 LVaR ( ω = 0.5 ) 191.469 ( x 143.261 ) + ( x 243.261 ) +
LVaR ( ω = 0.8 ) 198.218 ( x 153.286 ) + ( x 253.286 ) +
TVaR 201.874 ( x 158.466 ) + ( x 258.466 ) +
VaR 192.459 ( x 172.635 ) + ( x 272.635 ) +
LVaR ( ω = 0.2 ) 199.236 ( x 172.635 ) + ( x 272.635 ) +
0.99 LVaR ( ω = 0.5 ) 209.402 ( x 172.635 ) + ( x 272.635 ) +
LVaR ( ω = 0.8 ) 218.730 ( x 180.772 ) + ( x 280.772 ) +
TVaR 223.571 ( x 186.826 ) + ( x 286.826 ) +
VaR 252.207 ( x 249.023 ) + ( x 349.023 ) +
LVaR ( ω = 0.2 ) 257.744 ( x 249.023 ) + ( x 349.023 ) +
0.999 LVaR ( ω = 0.5 ) 266.050 ( x 249.023 ) + ( x 349.023 ) +
LVaR ( ω = 0.8 ) 274.356 ( x 249.023 ) + ( x 349.023 ) +
TVaR 279.786 ( x 251.536 ) + ( x 351.536 ) +
Table 4. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X N ( 40 , 100 ) , θ = 4 , and K = 120 .
Table 4. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X N ( 40 , 100 ) , θ = 4 , and K = 120 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 156.309 ( x 124.162 ) + ( x 168.155 ) +
LVaR ( ω = 0.2 ) 165.778 ( x 124.162 ) + ( x 168.155 ) +
0.90 LVaR ( ω = 0.5 ) 179.981 ( x 124.162 ) + ( x r ) + with r [ 124.162 , 299.231 ]
LVaR ( ω = 0.8 ) 180.431 ( x 124.162 ) + ( x 299.231 ) +
TVaR 180.731 ( x 124.162 ) + ( x 299.231 ) +
VaR 169.535 ( x 124.162 ) + ( x 204.485 ) +
LVaR ( ω = 0.2 ) 177.892 ( x 124.162 ) + ( x 204.485 ) +
0.95 LVaR ( ω = 0.5 ) 180.731 ( x 124.162 ) + ( x 299.231 ) +
LVaR ( ω = 0.8 ) 181.631 ( x 124.162 ) + ( x 299.231 ) +
TVaR 182.231 ( x 124.162 ) + ( x 299.231 ) +
VaR 174.172 ( x 124.162 ) + ( x 228.079 ) +
LVaR ( ω = 0.2 ) 180.231 ( x 124.162 ) + ( x 299.231 ) +
0.97 LVaR ( ω = 0.5 ) 181.731 ( x 124.162 ) + ( x 299.231 ) +
LVaR ( ω = 0.8 ) 183.231 ( x 124.162 ) + ( x 299.231 ) +
TVaR 184.231 ( x 124.162 ) + ( x 299.231 ) +
VaR 178.287 ( x 124.162 ) + ( x 272.635 ) +
LVaR ( ω = 0.2 ) 182.231 ( x 124.162 ) + ( x 299.231 ) +
0.99 LVaR ( ω = 0.5 ) 186.732 ( x 124.162 ) + ( x 299.231 ) +
LVaR ( ω = 0.8 ) 191.232 ( x 124.162 ) + ( x 299.231 ) +
TVaR 194.232 ( x 124.162 ) + ( x 299.231 ) +
VaR 229.023 ( x r ) + x β ( r ) + with r [ 47.219 , 222.453 ] and β ( r ) [ 299.231 , 349.023 ]
LVaR ( ω = 0.2 ) 234.560 ( x r ) + x β ( r ) + with r [ 47.219 , 222.453 ] and β ( r ) [ 299.231 , 349.023 ]
0.999 LVaR ( ω = 0.5 ) 242.866 ( x r ) + x β ( r ) + with r [ 47.219 , 222.453 ] and β ( r ) [ 299.231 , 349.023 ]
LVaR ( ω = 0.8 ) 251.172 ( x r ) + x β ( r ) + with r [ 47.219 , 222.453 ] and β ( r ) [ 299.231 , 349.023 ]
TVaR 256.709 ( x r ) + x β ( r ) + with r [ 47.219 , 222.453 ] and β ( r ) [ 299.231 , 349.023 ]
Table 5. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X F ( x ) = 1 x 120 3 , θ = 4 , and L = 100 .
Table 5. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X F ( x ) = 1 x 120 3 , θ = 4 , and L = 100 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 123.163 ( x 85.197 ) + ( x 138.532 ) +
LVaR ( ω = 0.2 ) 149.016 ( x 85.197 ) + ( x 138.532 ) +
0.90 LVaR ( ω = 0.5 ) 187.796 ( x 85.197 ) + ( x r ) + with r [ 138.532 , 235.197 ]
LVaR ( ω = 0.8 ) 207.890 ( x 93.182 ) + ( x 243.182 ) +
TVaR 220.799 ( x 98.571 ) + ( x 248.571 ) +
VaR 147.079 ( x 85.197 ) + ( x 205.730 ) +
LVaR ( ω = 0.2 ) 179.652 ( x 85.197 ) + ( x 205.730 ) +
0.95 LVaR ( ω = 0.5 ) 220.799 ( x 98.571 ) + ( x 248.571 ) +
LVaR ( ω = 0.8 ) 257.344 ( x 114.895 ) + ( x 264.895 ) +
TVaR 280.031 ( x 125.783 ) + ( x 275.783 ) +
VaR 164.667 ( x 116.196 ) + ( x 266.196 ) +
LVaR ( ω = 0.2 ) 203.286 ( x 116.196 ) + ( x 266.196 ) +
0.97 LVaR ( ω = 0.5 ) 261.214 ( x 116.196 ) + ( x 266.196 ) +
LVaR ( ω = 0.8 ) 315.204 ( x 143.696 ) + ( x 293.696 ) +
TVaR 347.493 ( x 161.109 ) + ( x 311.109 ) +
VaR 299.146 ( x 286.991 ) + ( x 436.991 ) +
LVaR ( ω = 0.2 ) 354.845 ( x 286.991 ) + ( x 436.991 ) +
0.99 LVaR ( ω = 0.5 ) 438.394 ( x 286.991 ) + ( x 436.991 ) +
LVaR ( ω = 0.8 ) 521.943 ( x 286.991 ) + ( x 436.991 ) +
TVaR 577.121 ( x 300.555 ) + ( x 450.555 ) +
VaR 930.918 ( x 930.000 ) + ( x 1080.000 ) +
LVaR ( ω = 0.2 ) 1050.918 ( x 930.000 ) + ( x 1080.000 ) +
0.999 LVaR ( ω = 0.5 ) 1230.918 ( x 930.000 ) + ( x 1080.000 ) +
LVaR ( ω = 0.8 ) 1410.918 ( x 930.000 ) + ( x 1080.000 ) +
TVaR 1530.917 ( x 930.984 ) + ( x 1080.983 ) +
Table 6. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X F ( x ) = 1 x 120 3 , θ = 4 , and K = 140 .
Table 6. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X F ( x ) = 1 x 120 3 , θ = 4 , and K = 140 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 123.163 ( x 85.197 ) + ( x 138.532 ) +
LVaR ( ω = 0.2 ) 149.016 ( x 85.197 ) + ( x 138.532 ) +
0.90 LVaR ( ω = 0.5 ) 187.796 ( x 85.197 ) + ( x r ) + with r [ 85.197 , 369.788 ]
LVaR ( ω = 0.8 ) 198.601 ( x 85.197 ) + ( x 369.788 ) +
TVaR 205.804 ( x 85.197 ) + ( x 369.788 ) +
VaR 147.079 ( x 85.197 ) + ( x 205.730 ) +
LVaR ( ω = 0.2 ) 179.652 ( x 85.197 ) + ( x 205.730 ) +
0.95 LVaR ( ω = 0.5 ) 205.831 ( x 85.197 ) + ( x 369.788 ) +
LVaR ( ω = 0.8 ) 227.414 ( x 85.197 ) + ( x 369.788 ) +
TVaR 241.820 ( x 85.197 ) + ( x 369.788 ) +
VaR 158.831 ( x 85.197 ) + ( x 266.196 ) +
LVaR ( ω = 0.2 ) 193.798 ( x 85.197 ) + ( x 369.788 ) +
0.97 LVaR ( ω = 0.5 ) 229.815 ( x 85.197 ) + ( x 369.788 ) +
LVaR ( ω = 0.8 ) 265.831 ( x 85.197 ) + ( x 369.788 ) +
TVaR 289.842 ( x 85.197 ) + ( x 369.788 ) +
VaR 236.991 ( x r ) + x β ( r ) + with r [ 15.450 , 211.637 ] and β ( r ) [ 369.788 , 436.991 ]
LVaR ( ω = 0.2 ) 292.690 ( x r ) + x β ( r ) + with r [ 15.450 , 211.637 ] and β ( r ) [ 369.788 , 436.991 ]
0.99 LVaR ( ω = 0.5 ) 376.238 ( x r ) + x β ( r ) + with r [ 15.450 , 211.637 ] and β ( r ) [ 369.788 , 436.991 ]
LVaR ( ω = 0.8 ) 459.787 ( x r ) + x β ( r ) + with r [ 15.450 , 211.637 ] and β ( r ) [ 369.788 , 436.991 ]
TVaR 515.486 ( x r ) + x β ( r ) + with r [ 15.450 , 211.637 ] and β ( r ) [ 369.788 , 436.991 ]
VaR 880.000 ( x r ) + x β ( r ) + with r [ 0 , 878.669 ] and β ( r ) [ 369.788 , 1080 ]
LVaR ( ω = 0.2 ) 1000.000 ( x r ) + x β ( r ) + with r [ 0 , 878.669 ] and β ( r ) [ 369.788 , 1080 ]
0.999 LVaR ( ω = 0.5 ) 1180.000 ( x r ) + x β ( r ) + with r [ 0 , 878.669 ] and β ( r ) [ 369.788 , 1080 ]
LVaR ( ω = 0.8 ) 1360.000 ( x r ) + x β ( r ) + with r [ 0 , 878.669 ] and β ( r ) [ 369.788 , 1080 ]
TVaR 1480.000 ( x r ) + x β ( r ) + with r [ 0 , 878.669 ] and β ( r ) [ 369.788 , 1080 ]
Table 7. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X F ( x ) = exp x 5 50 3 , θ = 3 , and L = 100 .
Table 7. Optimal solutions under different risk measures with the constraint of F 1 under the settings that X F ( x ) = exp x 5 50 3 , θ = 3 , and L = 100 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 100.108 ( x 80.741 ) + ( x 110.863 ) +
LVaR ( ω = 0.2 ) 111.032 ( x 80.741 ) + ( x 110.863 ) +
0.90 LVaR ( ω = 0.5 ) 123.962 ( x 81.409 ) + ( x 181.409 ) +
LVaR ( ω = 0.8 ) 129.893 ( x 83.432 ) + ( x 183.432 ) +
TVaR 133.772 ( x 84.797 ) + ( x 184.797 ) +
VaR 108.290 ( x 80.741 ) + ( x 139.571 ) +
LVaR ( ω = 0.2 ) 121.955 ( x 80.741 ) + ( x r ) + with r [ 139.571 , 180.741 ]
0.95 LVaR ( ω = 0.5 ) 133.772 ( x 84.797 ) + ( x 184.797 ) +
LVaR ( ω = 0.8 ) 145.064 ( x 88.963 ) + ( x 188.963 ) +
TVaR 152.311 ( x 91.794 ) + ( x 191.794 ) +
VaR 112.261 ( x 80.741 ) + ( x 165.102 ) +
LVaR ( ω = 0.2 ) 127.273 ( x 82.529 ) + ( x 182.529 ) +
0.97 LVaR ( ω = 0.5 ) 146.287 ( x 89.432 ) + ( x 189.432 ) +
LVaR ( ω = 0.8 ) 163.912 ( x 96.586 ) + ( x 196.586 ) +
TVaR 174.945 ( x 101.444 ) + ( x 201.444 ) +
VaR 146.303 ( x 136.691 ) + ( x 236.691 ) +
LVaR ( ω = 0.2 ) 169.542 ( x 136.691 ) + ( x 236.691 ) +
0.99 LVaR ( ω = 0.5 ) 204.401 ( x 136.691 ) + ( x 236.691 ) +
LVaR ( ω = 0.8 ) 239.259 ( x 136.691 ) + ( x 236.691 ) +
TVaR 261.556 ( x 148.004 ) + ( x 248.004 ) +
VaR 405.479 ( x 404.917 ) + ( x 504.917 ) +
LVaR ( ω = 0.2 ) 455.486 ( x 404.917 ) + ( x 504.917 ) +
0.999 LVaR ( ω = 0.5 ) 530.496 ( x 404.917 ) + ( x 504.917 ) +
LVaR ( ω = 0.8 ) 605.506 ( x 404.917 ) + ( x 504.917 ) +
TVaR 655.511 ( x 405.550 ) + ( x 505.550 ) +
Table 8. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X F ( x ) = exp x 5 50 3 , θ = 3 , and K = 120 .
Table 8. Optimal solutions under different risk measures with the constraint of F 2 under the settings that X F ( x ) = exp x 5 50 3 , θ = 3 , and K = 120 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 100.108 ( x 80.741 ) + ( x 110.863 ) +
LVaR ( ω = 0.2 ) 111.032 ( x 80.741 ) + ( x 110.863 ) +
0.90 LVaR ( ω = 0.5 ) 123.111 ( x 80.741 ) + ( x 237.333 ) +
LVaR ( ω = 0.8 ) 126.578 ( x 80.741 ) + ( x 237.333 ) +
TVaR 128.889 ( x 80.741 ) + ( x 237.333 ) +
VaR 108.290 ( x 80.741 ) + ( x 139.571 ) +
LVaR ( ω = 0.2 ) 121.955 ( x 80.741 ) + ( x r ) + with r [ 80.741 , 237.333 ]
0.95 LVaR ( ω = 0.5 ) 128.889 ( x 80.741 ) + ( x 237.333 ) +
LVaR ( ω = 0.8 ) 135.822 ( x 80.741 ) + ( x 237.333 ) +
TVaR 140.444 ( x 80.741 ) + ( x 237.333 ) +
VaR 112.261 ( x 80.741 ) + ( x 165.102 ) +
LVaR ( ω = 0.2 ) 125.037 ( x 80.741 ) + ( x 237.333 ) +
0.97 LVaR ( ω = 0.5 ) 136.593 ( x 80.741 ) + ( x 237.333 ) +
LVaR ( ω = 0.8 ) 148.148 ( x 80.741 ) + ( x 237.333 ) +
TVaR 155.852 ( x 80.741 ) + ( x 237.333 ) +
VaR 117.308 ( x 80.741 ) + ( x 236.691 ) +
LVaR ( ω = 0.2 ) 140.444 ( x 80.741 ) + ( x 237.333 ) +
0.99 LVaR ( ω = 0.5 ) 175.111 ( x 80.741 ) + ( x 237.333 ) +
LVaR ( ω = 0.8 ) 209.778 ( x 80.741 ) + ( x 237.333 ) +
TVaR 232.889 ( x 80.741 ) + ( x 237.333 ) +
VaR 384.917 ( x r ) + x β ( r ) + with r [ 0 , 384.179 ] and β ( r ) [ 237.333 , 504.917 ]
LVaR ( ω = 0.2 ) 434.923 ( x r ) + x β ( r ) + with r [ 0 , 384.179 ] and β ( r ) [ 237.333 , 504.917 ]
0.999 LVaR ( ω = 0.5 ) 509.933 ( x r ) + x β ( r ) + with r [ 0 , 384.179 ] and β ( r ) [ 237.333 , 504.917 ]
LVaR ( ω = 0.8 ) 584.943 ( x r ) + x β ( r ) + with r [ 0 , 384.179 ] and β ( r ) [ 237.333 , 504.917 ]
TVaR 634.950 ( x r ) + x β ( r ) + with r [ 0 , 384.179 ] and β ( r ) [ 237.333 , 504.917 ]
Table 9. Optimal solutions under different risk measures with the constraint of F 1 .
Table 9. Optimal solutions under different risk measures with the constraint of F 1 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 77.587 ( x 63.496 ) + ( x 83.203 ) +
LVaR ( ω = 0.2 ) 86.448 ( x 63.496 ) + ( x 83.203 ) +
0.90 LVaR ( ω = 0.5 ) 99.738 ( x 63.496 ) + ( x r ) + with r [ 83.203 , 183.496 ]
LVaR ( ω = 0.8 ) 102.565 ( x 64.314 ) + ( x 184.314 ) +
TVaR 104.436 ( x 64.863 ) + ( x 184.863 ) +
VaR 85.980 ( x 63.496 ) + ( x 106.736 ) +
LVaR ( ω = 0.2 ) 96.987 ( x 63.496 ) + ( x 106.736 ) +
0.95 LVaR ( ω = 0.5 ) 104.436 ( x 64.863 ) + ( x 184.863 ) +
LVaR ( ω = 0.8 ) 109.982 ( x 66.532 ) + ( x 186.532 ) +
TVaR 113.624 ( x 67.662 ) + ( x 187.663 ) +
VaR 90.005 ( x 63.496 ) + ( x 127.432 ) +
LVaR ( ω = 0.2 ) 101.312 ( x 63.949 ) + ( x 183.949 ) +
0.97 LVaR ( ω = 0.5 ) 110.592 ( x 66.720 ) + ( x 186.720 ) +
LVaR ( ω = 0.8 ) 119.598 ( x 69.577 ) + ( x 189.577 ) +
TVaR 125.452 ( x 71.529 ) + ( x 191.529 ) +
VaR 95.128 ( x 65.043 ) + ( x 185.043 ) +
LVaR ( ω = 0.2 ) 113.624 ( x 67.662 ) + ( x 187.662 ) +
0.99 LVaR ( ω = 0.5 ) 139.572 ( x 76.565 ) + ( x 196.566 ) +
LVaR ( ω = 0.8 ) 163.204 ( x 86.098 ) + ( x 206.099 ) +
TVaR 177.766 ( x 92.674 ) + ( x 212.675 ) +
VaR 280.907 ( x 279.867 ) + ( x 399.867 ) +
LVaR ( ω = 0.2 ) 320.917 ( x 279.867 ) + ( x 399.867 ) +
0.999 LVaR ( ω = 0.5 ) 380.933 ( x 279.867 ) + ( x 399.867 ) +
LVaR ( ω = 0.8 ) 440.949 ( x 279.867 ) + ( x 399.867 ) +
TVaR 480.954 ( x 281.130 ) + ( x 401.130 ) +
Table 10. Optimal solutions under different risk measures with the constraint of F 2 .
Table 10. Optimal solutions under different risk measures with the constraint of F 2 .
α The Risk Measure ρ ρ α T I ( X ) I ( x )
VaR 77.587 ( x 63.496 ) + ( x 83.203 ) +
LVaR ( ω = 0.2 ) 86.448 ( x 63.496 ) + ( x 83.203 ) +
0.90 LVaR ( ω = 0.5 ) 99.738 ( x 63.496 ) + ( x r ) + with r [ 63.496 , 257.326 ]
LVaR ( ω = 0.8 ) 101.186 ( x 63.496 ) + ( x 257.326 ) +
TVaR 102.151 ( x 63.496 ) + ( x 257.326 ) +
VaR 85.980 ( x 63.496 ) + ( x 106.736 ) +
LVaR ( ω = 0.2 ) 96.987 ( x 63.496 ) + ( x 257.326 ) +
0.95 LVaR ( ω = 0.5 ) 102.151 ( x 63.496 ) + ( x 257.326 ) +
LVaR ( ω = 0.8 ) 105.046 ( x 63.496 ) + ( x 257.326 ) +
TVaR 106.976 ( x 63.496 ) + ( x 257.326 ) +
VaR 90.005 ( x 63.496 ) + ( x 127.432 ) +
LVaR ( ω = 0.2 ) 100.543 ( x 63.496 ) + ( x 257.326 ) +
0.97 LVaR ( ω = 0.5 ) 105.368 ( x 63.496 ) + ( x 257.326 ) +
LVaR ( ω = 0.8 ) 110.193 ( x 63.496 ) + ( x 257.326 ) +
TVaR 113.410 ( x 63.496 ) + ( x 257.326 ) +
VaR 95.084 ( x 63.496 ) + ( x 185.043 ) +
LVaR ( ω = 0.2 ) 106.976 ( x 63.496 ) + ( x 257.326 ) +
0.99 LVaR ( ω = 0.5 ) 121.453 ( x 63.496 ) + ( x 257.326 ) +
LVaR ( ω = 0.8 ) 135.929 ( x 63.496 ) + ( x 257.326 ) +
TVaR 145.580 ( x 63.496 ) + ( x 257.326 ) +
VaR 239.867 ( x r ) + x β ( r ) + with r [ 0.243 , 238.049 ] and β ( r ) [ 257.326 , 399.867 ]
LVaR ( ω = 0.2 ) 279.877 ( x r ) + x β ( r ) + with r [ 0.243 , 238.049 ] and β ( r ) [ 257.326 , 399.867 ]
0.999 LVaR ( ω = 0.5 ) 339.893 ( x r ) + x β ( r ) + with r [ 0.243 , 238.049 ] and β ( r ) [ 257.326 , 399.867 ]
LVaR ( ω = 0.8 ) 399.909 ( x r ) + x β ( r ) + with r [ 0.243 , 238.049 ] and β ( r ) [ 257.326 , 399.867 ]
TVaR 439.920 ( x r ) + x β ( r ) + with r [ 0.243 , 238.049 ] and β ( r ) [ 257.326 , 399.867 ]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xiong, Q.; Peng, Z.; Nadarajah, S. Optimal Reinsurance under the Linear Combination of Risk Measures in the Presence of Reinsurance Loss Limit. Risks 2023, 11, 125. https://doi.org/10.3390/risks11070125

AMA Style

Xiong Q, Peng Z, Nadarajah S. Optimal Reinsurance under the Linear Combination of Risk Measures in the Presence of Reinsurance Loss Limit. Risks. 2023; 11(7):125. https://doi.org/10.3390/risks11070125

Chicago/Turabian Style

Xiong, Qian, Zuoxiang Peng, and Saralees Nadarajah. 2023. "Optimal Reinsurance under the Linear Combination of Risk Measures in the Presence of Reinsurance Loss Limit" Risks 11, no. 7: 125. https://doi.org/10.3390/risks11070125

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop