
The reduction of carbon emissions has attracted significant global attention. This paper empirically analyzes the dynamic nonlinear linkages among carbon markets, green bonds, clean energy, and electricity markets by constructing DCC-GARCH and TVP-VAR-SV models, and places the four markets under a unified framework to analyze the volatility risk from a time-varying perspective, thereby enriching the research on China's carbon market and renewable energy sector. We found that extreme events have a significant impact on the dynamic connectivity among the four markets. The analysis of the shock impact indicates that the carbon market has a positive effect on the power market in the short and medium terms, but has a mitigating impact in the long term. Especially, when the other markets are hit, the carbon market has evident fluctuation in 2020. The green bond market has a positive influence on the carbon market, whereas the power market demonstrates adverse effects in the short and medium terms. The New Energy Index negatively impacts the power market in the short and medium terms, but is expected to have a positive effect after 2020, highlighting the growing need for renewable energy in the power system transformation. According to the findings mentioned above, we put forward appropriate recommendations.
Citation: Jiamin Cheng, Yuanying Jiang. How can carbon markets drive the development of renewable energy sector? Empirical evidence from China[J]. Data Science in Finance and Economics, 2024, 4(2): 249-269. doi: 10.3934/DSFE.2024010
[1] | 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 |
[2] | 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 |
[3] | Wen Chao . Pricing catastrophe reinsurance under the standard deviation premium principle. AIMS Mathematics, 2022, 7(3): 4472-4484. doi: 10.3934/math.2022249 |
[4] | Jumanah Ahmed Darwish, Saman Hanif Shahbaz, Lutfiah Ismail Al-Turk, Muhammad Qaiser Shahbaz . Some bivariate and multivariate families of distributions: Theory, inference and application. AIMS Mathematics, 2022, 7(8): 15584-15611. doi: 10.3934/math.2022854 |
[5] | 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 |
[6] | Samah M. Ahmed, Abdelfattah Mustafa . Estimation of the coefficients of variation for inverse power Lomax distribution. AIMS Mathematics, 2024, 9(12): 33423-33441. doi: 10.3934/math.20241595 |
[7] | Ammar M. Sarhan, Rabab S. Gomaa, Alia M. Magar, Najwan Alsadat . Bivariate exponentiated generalized inverted exponential distribution with applications on dependent competing risks data. AIMS Mathematics, 2024, 9(10): 29439-29473. doi: 10.3934/math.20241427 |
[8] | G. M. Mansour, M. A. Abd Elgawad, A. S. Al-Moisheer, H. M. Barakat, M. A. Alawady, I. A. Husseiny, M. O. Mohamed . Bivariate Epanechnikov-Weibull distribution based on Sarmanov copula: properties, simulation, and uncertainty measures with applications. AIMS Mathematics, 2025, 10(5): 12689-12725. doi: 10.3934/math.2025572 |
[9] | H. M. Barakat, M. A. Alawady, I. A. Husseiny, M. Nagy, A. H. Mansi, M. O. Mohamed . Bivariate Epanechnikov-exponential distribution: statistical properties, reliability measures, and applications to computer science data. AIMS Mathematics, 2024, 9(11): 32299-32327. doi: 10.3934/math.20241550 |
[10] | Hassan Okasha, Mazen Nassar, Saeed A. Dobbah . E-Bayesian estimation of Burr Type XII model based on adaptive Type-Ⅱ progressive hybrid censored data. AIMS Mathematics, 2021, 6(4): 4173-4196. doi: 10.3934/math.2021247 |
The reduction of carbon emissions has attracted significant global attention. This paper empirically analyzes the dynamic nonlinear linkages among carbon markets, green bonds, clean energy, and electricity markets by constructing DCC-GARCH and TVP-VAR-SV models, and places the four markets under a unified framework to analyze the volatility risk from a time-varying perspective, thereby enriching the research on China's carbon market and renewable energy sector. We found that extreme events have a significant impact on the dynamic connectivity among the four markets. The analysis of the shock impact indicates that the carbon market has a positive effect on the power market in the short and medium terms, but has a mitigating impact in the long term. Especially, when the other markets are hit, the carbon market has evident fluctuation in 2020. The green bond market has a positive influence on the carbon market, whereas the power market demonstrates adverse effects in the short and medium terms. The New Energy Index negatively impacts the power market in the short and medium terms, but is expected to have a positive effect after 2020, highlighting the growing need for renewable energy in the power system transformation. According to the findings mentioned above, we put forward appropriate recommendations.
In the preface of the book of [1], it is stated that reinsurance is a fascinating field, where several of the challenges of classical insurance are amplified, particularly when it comes to dealing with extreme situations like large claims and rare events. However, leaving this enthusiastic phrase aside, the reinsurance field is also exciting because of its practicality and ability to solve many financial problems for insurance companies. As is well known by the actuarial community, a reinsurance contract establishes an agreement or contract between the insurer and the reinsurer, whereby the former (also called the ceding company) indemnifies the latter a part of the risk subscribed by the insurer, with a third party (usually the customer) paying, in turn, a reinsurance premium. As such, it is a contract similar to the one that an insurance company signs with a client to cover damages of any kind. The difference, however, is that the reinsurer covers only a part of the compensation, called priority. This type of reinsurance is called excess-of-loss reinsurance. For a detailed study of reinsurance, in all its forms, readers are referred to the works of [1,2,3,4,5] and the references therein. Recent works on reinsurance include contributions such as [6,7,8,9]. In these recent works, the reinsurance technique is also used in non-life insurance, but its scope is broadened, including, for example, non-life multiline insurers under the Solvency Ⅱ umbrella, risk transfers in the context of the 2007-08 global crisis, reinsurance using two layers and multiple reinsurers, and reinsurance in a game-theoretic setting, considering one insurer and two reinsurers.
As part of the collective risk model, in excess-of-loss reinsurance, it is usual to assume that the risk parameters associated with the random variables, the number of claims, and the claim size are independent. This is the main idea in [10], inspired by the ideas developed in [11]. For example, we can describe this situation by taking the Poisson frequency and the Pareto severity as appropriate distributional assumptions. The use of the Bayesian methodology facilitates the computation of predictive and posterior moments of the accumulated claims and the interpretation of the parameters involved. [12] also studied the Bayesian estimators in Paretian excess-of-loss reinsurance. Although these derivations have traditionally been built on the assumption of independence (see [13]), there has been great interest in modeling dependent risks due to the increasing complexity of the current insurance and reinsurance products.
In this paper, we will pay attention to the calculation of the prior and posterior moments of the size of the total claims, assuming some degree of dependence derived from the prior distributions of the risk parameters (profiles) involved in the Bayesian framework. This degree of dependence will be modeled using the Sarmanov-Lee family of distributions; specific details of this type of copula can be seen in [14,15]. This model allows us to study the impact of dependence on the next year's total claims costs, showing that a low degree of correlation produces highly sensitive results. Furthermore, the case where the moments are calculated separately without assuming dependence is also considered. In recent years, some actuarial works have considered the Sarmanov family of bivariate distributions: see, for example, [16,17,18]. Also, [19] considered, from a Bayesian point of view, the calculation of credibility premiums for the compound loss under a bivariate prior for the risk parameters of the frequency and severity distributions.
The impact of assuming the dependence feature is a matter of concern. At first sight, it would seem that the effect of including dependence has a low impact in the first and second moments; in fact, the reinsurer should be prudent, since ignoring the hypothesis of dependence might lead to budget imbalance, especially when there is a change of scale in the monetary units used. This work assumes the dependence hypothesis between the risk profiles instead of the claims number and the size of the corresponding claims. The former approach not only retains the main features of the latter case but also reduces the mathematical burden of the independence assumption between the number and the severity.
The rest of the paper has the following structure: The basic collective risk model considered in this paper is introduced in Section 2. Here, the bivariate prior distribution assumed for the risk profile (dependent on two parameters) is also presented and studied. In the original paper of [10], a classical gamma distribution was assumed, but this seems inadequate considering this risk parameter's support. To make the paper self-contained, we recall the main results of [10] in Section 3. Here, we also give closed-form expressions for the risk, collective, and Bayes net premiums in the excess-of-loss reinsurance model. An extension of the basic collective risk model and the reinsurance model, which includes some degree of dependence, is introduced in Section 4. Here, we also provide properties of this extended model together with the expressions of the collective and Bayesian net premiums, which are written in closed-form expressions. The cases with and without a layer are both studied. Numerical applications are provided in Section 5, and final comments are given in the last section.
Let N(t) be the random variable denoting the number of claims reported to the insurer during [0,t] and let Y1,Y2,…,YN(t) be the corresponding individual claim amounts up to time t. Assume the following: (i) {N(t),t≥0} is a time homogeneous Poisson process with intensity λ>0, i.e., pn(t)=exp(−λt)(λt)n/n! and (ii) {Yi}i=1,2,… are stochastically independent and independent of N(t) with common survival function ˉF(y|c,ψ)=(c/y)ψ, y≥c, i.e., the classical Pareto distribution with scale parameter c>0 and shape parameter ψ>0. Its mean is given by E(Y)=cψ/(ψ−1), ψ>1. The compound process described above has a cumulative distribution function (see, for example, [20]) given by
G(x)=∞∑n=0pn(t)Fn∗(x), | (2.1) |
where Fn∗(⋅) represents the nth-fold convolution of F(⋅|c,ψ). In practice, it is difficult to derive a closed-form expression for the distribution function given in (2.1). Therefore, numerical approximations have been proposed based on the normal and gamma distribution and fast Fourier transform. Other approximations that have been considered are based on Edgeworth expansions and the Gram-Charlier series. Furthermore, an exact recursive expression for (2.1) can be found by using Panjer's recursion formula (see [21] and [22, Chapter 4], among others). A closed-form expression for the Pareto type Ⅲ (Lomax) distribution has been derived for the convolution above in [23], and the asymptotic distribution of an infinite sum of i.i.d. classical Pareto type Ⅰ random variables is provided in [24].
Nevertheless, in the expression (2.1), it is not necessary to calculate premiums based on the moments of the distribution. We will consider that the parameter c is known, that the risk profiles λ and ψ are unknown, and that they take values from the random variables Λ and Ψ, respectively. As it is well-known, a premium calculation principle ([25,26]; among others) assigns to each vector of risk profiles Θ=(λ,ψ) a premium within the set P∈R+, the action space. Using the net premium principle, it is easy to see that, in this case, it gives the unknown risk premium
P(Θ)=E[N(t)]E[Y]=cλtψψ−1,ψ>1. |
We have followed the idea of [10] in order to define the net premium. That is, the net premium is defined as the sum of losses instead of the sum of discounted losses, which leads to the exclusion of the time value of the monetary losses.
In the absence of experience, the actuary computes the collective (prior) premium, Eπ[P(Θ)], where π is the prior distribution assigned to the risk profile Θ. On the other hand, if experience is available, the actuary takes a sample
ℵ={Nc(T)=n,Yc,t=yti,i=1,2,…,n}, | (2.2) |
and uses this information to estimate the unknown risk premium P(Θ) through the Bayes premium P∗, obtained by computing the posterior expectation Eπ∗[P(Θ)]. Here, π∗ is the posterior distribution of the risk parameter, Θ, given the sample information provided in (2.2). For that purpose, we need also the likelihood function, which is given by
ℓ(λ,ψ|ℵ=n,yt1,…,ytn)=(Tλ)nn!exp(−Tλ)n∏i=1ψc(ytic)−(ψ+1),∝λnexp(−Tλ)ψnexp(−zψ), | (2.3) |
where z=∑ni=1log(yt/c).
We will assume that the distribution of Λ is gamma with shape parameter ν>0 and rate parameter τ>0 and that the distribution of Ψ is Erlang with shape parameter γ∈N+={1,2,…} and rate parameter ξ>0. That is, it is assumed that Λ∼G(ν,τ) and Ψ∼E(γ,ξ,k), a k>0 translated Erlang variate, with probability density functions (pdfs) given by, *
π1(λ|ν,τ)=τνΓ(ν)λν−1exp(−τλ),λ>0, | (2.4) |
π2(ψ|γ,ξ,k)=ξγΓ(γ)(ψ−k)γ−1exp[−ξ(ψ−k)],ψ>k, | (2.5) |
*[10] assumes a gamma prior for the parameter ψ that seems inappropriate since from Proposition 3, we have that ψ>k. A translated gamma or Erlang distribution should be more accurate in the case considered here.
where Γ(⋅) is the Euler gamma function and Γ(γ)=(γ−1)!. The choice of the prior distributions, in addition to mathematical convenience (calculations are provided), is done in this way following Hesselager's work [10]. The fundamental objective of this work is to study the variation of the premiums assuming a certain degree of dependence between the risk profiles. Mean and variance of these two distributions are ν/τ, ν/τ2, k+γ/ξ, and γ/ξ2, respectively.
Finally, the joint pdf of Θ=(Λ,Ψ) is assumed to be the product of these densities. Thus,
π(λ,ψ|ν,τ,γ,ξ,k)=π1(λ|ν,τ)π2(ψ|γ,ξ,k), | (2.6) |
i.e., both random variables are assumed to be independent. The next result provides the posterior distribution of (2.6) given the sample information provided in (2.3). As we will see, the posterior distribution is conjugate for λ and almost conjugate (quasi-conjugate) for the ψ risk profile.
Proposition 1. Let {X(t)}t≥0 be the compound Poisson process defined above. Suppose that (Λ,Ψ) follows the prior distribution with pdf given in (2.6). Then, the posterior distribution of (Λ,Ψ) given the sample information ℵ={Nc(T)=n,Yc,t=yti,i=1,2,…,n}, assumed to be independent and identically distributed, is given by,
π∗(λ,ψ|ν∗,τ∗,γ∗,ξ∗,k)=π1(λ|ν∗,τ∗)π∗2(ψ|γ∗,ξ∗,k), |
where
π∗2(ψ|γ∗,ξ∗,k)=κ−1ψn(ψ−k)γ−1exp(−ξ∗ψ),ψ>k, |
and the updated parameters are given by
ν∗=ν+n, | (2.7) |
τ∗=τ+T, | (2.8) |
γ∗=γ+n, | (2.9) |
ξ∗=ξ+z, | (2.10) |
being κ≡κ(γ,ξ∗,k) the normalization constant given by
κ≡κ(γ,ξ∗,k)=exp(−ξ∗k)n∑j=0(nj)kn−jΓ(γ+j)(ξ∗)γ+j. | (2.11) |
Proof. The posterior pdf π∗1(λ|ν∗,τ∗) is straightforwardly obtained from (2.3), (2.4), and Bayes' theorem. Now, π∗2(ψ|γ∗,ξ∗,k) is proportional to
π∗2(ψ|γ∗,ξ∗,k)∝ψnexp(−zψ)π2(ψ|γ,ξ,k), |
where the constant of proportionality is given by
κ(γ,ξ∗,k)=∫∞kψn(ψ−k)γ−1exp(−ξ∗ψ)dψ=exp(−ξ∗k)n∑j=0(nj)kn−j∫∞0ψj+γ−1exp(−ξ∗ψ)dψ=exp(−ξ∗k)n∑j=0(nj)kn−jΓ(γ+j)(ξ∗)γ+j. |
Hence the result.
As we can see in the next result, bivariate, prior, and posterior distributions can now be used to get the collective (prior) and posterior (Bayes) premiums of the aggregated model described above based on the first moment (the net premium).
Proposition 2. The collective (prior) and Bayes (posterior) net premiums, i.e., k=1, for the compound aggregate model described above are given by,
P(ν,τ,γ,ξ)=cντξ+γ−1γ−1,γ>1, | (2.12) |
P∗(ν∗,τ∗,γ∗,ξ∗)=cν∗τ∗κ∗(γ,ξ∗,1)κ(γ,ξ∗,1),γ≥2, | (2.13) |
respectively.
Proof. Expression (2.12) is obtained as follows,
P(ν,τ,γ,ξ)=∫∞1∫∞0P(λ,ψ)π1(λ|ν,τ)π2(ψ|γ,ξ,1)dλdψ=cντξγΓ(γ)∫∞1ψ(ψ−1)γ−2exp[−ξ(ψ−1)]dψ=cντξγΓ(γ)Γ(γ−1)ξγ−1(1+γ−1ξ), |
from which we get the result. Observe that we have used the fact that the mean of the random variable Ψ following the pdf π2(ψ|γ,ξ) is 1+(γ−1)/ξ.
To get (2.13), we need to compute
κ∗=∫∞1ψψ−1π∗2(ψ|γ∗,ξ∗,1)dψ=κ−1∫∞1ψn+1(ψ−1)γ−2exp(−ξ∗ψ)dψ=κ−1γ−2∑j=0(γ−1j)(−1)j∫∞1ψγ∗−j−1exp(−ξ∗ψ)dψ. |
Now, taking into account (2.7) and (2.8), the result is obtained directly.
[10] proposed the aggregate model described in the previous section to calculate reinsurance premiums in the collective risk model. For that, let Nu(t)=∑N(t)i=1I(Yi>u) be the number of claims exceeding the level u during [0,t]. The latter expression is again a Poisson process with intensity λu=λPr(Yi>u)=λ(σ/u)ψ, for u≥σ.
Consider an excess-of-loss cover for the layer b−a in excess of a. For a claim Ya, exceeding the priority or deductible a, the reinsurer will cover the amount Zi=min{Y,b}−a. Our interest now lies on the random variable, the reinsurer's total claim amount during period [0,t], X(t)=∑Na(t)i=1Zi. It is clear that {X(t)}t≥0 is a compound Poisson process with intensity λa and distribution function given by
H(z|ψ)={F(a+z|a,ψ),0≤z<b−a,1,z≥b−a. |
Now, at time T, the practitioner wishes to predict the reinsurer's total claims cost for the next year, i.e. X=X(T+1)−X(T). Therefore, we must first compute the so-called risk premium of order k, k=1,2,…, i.e., premiums based on non-central moments. Observe that the first moment is the so-called net risk premium (see, for instance, [25]). Then, we need
E[Λ(σa)Ψμk(ψ)], | (3.1) |
where Θ=(Λ,Ψ) is the vector of risk parameters and μk(ψ)=E(Zki|ψ).
Then, it is necessary to compute μk(ψ)=E(Zki|ψ). A closed-form expression for this kth moment is provided in [10] giving
μk(ψ)=kakk−1∑j=0(k−1j)(−1)j1−(a/b)ψ−k+jψ−k+j. | (3.2) |
The latter result can be written as a closed-form expression regarding the incomplete Beta function, as shown in the following proposition.
Proposition 3. It is verified that the non-central moments μk(ψ)=E(Zki|ψ) can be written as
μk(ψ)=(ab)ψ(b−a)k+ψakI1−a/b(k+1,ψ−k), | (3.3) |
with ψ>k, where Iz(r,s) is the incomplete Beta function, given by
Iz(r,s)=∫z0tr−1(1−t)s−1dt,0≤z≤1. |
Proof. We have that
μk(ψ)=E(Zk|ψ)=(ab)ψ(b−a)k+∫ba(y−a)kψaψy−(1+ψ)dy=(ab)ψ(b−a)k+ψaψ∫ba(1−ay)kyk−(1+ψ)dy. |
Let now the change of variable be z=1−a/y, and we get that the last integral can be rewritten as
ψak∫1−a/b0z(k+1)−1(1−z)ψ−k−1dz=ψakI1−a/b(k+1,ψ−k), |
from which we have the result.
The main idea of Hesselager's model (see [10]) is to consider that the parameters λ and ψ (often termed as risk profiles) are unknown and will be regarded as outcomes of random parameters Λ and Ψ, respectively, and to treat the problem in a Bayesian framework. For that reason, the prior distributions of Λ and Ψ are required, and the premium provided in (3.1) is called risk premium, net risk premium when k=1.
Proposition 4. The risk, collective (prior), and Bayes (posterior) net premiums, i.e., k=1, are given by,
P(λ,ψ)=λψ−1[a−b(ab)ψ](ca)ψ,ψ>1, | (3.4) |
P(ν,τ,γ,ξ)=ντcξγγ−1{1[ξ+log(a/c)]γ−1−1[ξ+log(b/c)]γ−1},γ>1, | (3.5) |
P∗(ν∗,τ∗,γ∗,ξ∗)=ν∗τ∗aκ(γ∗,ξ∗+log(a/c),1)−bκ(γ∗,ξ∗+log(b/c),1)κ(γ∗,ξ∗,1), | (3.6) |
respectively and where a{<=>}c.
Proof. By taking t=1, we have that P(λ,ψ)=E(N(t))E(Y)=λμ1(ψ). Now, μ1(ψ) is obtained directly from (3.2) or (3.3) by taking k=1, from which we easily get (3.1). To get (3.5), we have to compute P(ν,τ,γ,ξ)=Ep[P(Λ,Ψ)], where the expectation is taken with respect to p(λ,ψ|ν,τ,γ,ξ,1). Now, due to the hypothesis of independence between the risk profiles, we have that
P(ν,τ,γ,ξ)=Eπ1(Λ)Eπ2{1Ψ−1[a−b(ab)Ψ]}, |
where Eπ1(Λ)=ν/τ. On the other hand, we have that
Eπ2{1Ψ−1[a−b(ab)Ψ]}=cξγΓ(γ)∫∞1(ψ−1)γ−1−1exp[−(ξ+log(ac))(ψ−1)]dψ−cξγΓ(γ)∫∞1(ψ−1)γ−1−1exp[−(ξ+log(cb))(ψ−1)]dψ=cξγΓ(γ)[Γ(γ−1)[ξ+log(ac)]γ−1−Γ(γ−1)[ξ+log(bc)]γ−1], |
from which we get (3.5) after some simple algebra. Expression (3.6), which is the posterior expectation P∗(ν∗,τ∗,γ∗,ξ∗,k)=Eπ∗[P(Λ,Ψ)], is obtained from (3.1) using the updated parameters provided in (2.7)–(2.10) and arranging the constant of normalization.
As a special case of the model studied before, we now consider that b→∞. This is the basic excess-of-loss reinsurance model considered in [27, p. 11] and [1, p. 49]. In this case, an unlimited excess-of-loss treaty exists with retention a>0, and the mean excess amount gives the expected amount to be paid by the reinsurer. Thus, z>0 and the risk premium can be obtained from (3.1) as
limb→∞P(λ,ψ)=λaψ−1(ca)ψ. |
Furthermore, from (3.5) and (3.6) and assuming that b→∞, the collective and Bayes premiums are given by,
P(ν,τ,γ,ξ)=ντcξγ(γ−1)[ξ+log(a/c)]γ−1, | (3.7) |
P∗(ν∗,τ∗,γ∗,ξ∗)=ν∗τ∗aκ(γ∗,ξ∗+log(a/c),1)κ(γ∗,ξ∗,1). | (3.8) |
An anonymous reviewer has pointed out this important detail: suppose the deductible a is sufficiently large. The excess loss should be roughly distributed as the generalized Pareto distribution (Pickands-Balkema-de Haan theorem), which might be of more interest because the underlying distribution of individual severity can be assumed arbitrarily. See, for instance [28].
We focus here on the collective and Bayes net premiums for the aggregate amount of claims under a compound model, assuming some degree of dependence between the random variables Λ and Ψ. In our case, it is challenging to implement this assumption in the first stage of the model, involving Poisson and Pareto random variables. As pointed out by [29], an attractive way to model the dependence between risks is via the dependence between the risk profiles. Thus, we translate the non-dependence to the second stage (the prior distributions of the risk profiles).
In the statistical literature, there exists a lot of research papers dealing with the construction of bivariate distributions (see, for example, [30,31,32]). Most of them are based on the idea of building a bivariate distribution with given marginal distributions, and the Sarmanov-Lee family of distributions [14,15] follows this idea. Specifically, assume f1(x1) and f2(x2) are univariate probability density functions for the continuous case or probability mass functions for the discrete one, with supports defined on subsets of R (which can be the total real line). Let ϕi(t), i=1,2, be bounded nonconstant functions such that ∫∞−∞ϕi(t)fi(t) dt=0 (they are usually named mixing functions). Then, provided that ω is a real number that satisfies the condition 1+ω ϕ1(x1)ϕ2(x2)≥0 for all x1 and x2, the function given by
h(x1,x2)=f1(x1)f2(x2)[1+ω ϕ1(x1)ϕ2(x2)] | (4.1) |
which is a genuine bivariate joint density (or a probability mass function in the discrete case) with specified marginals f1(x1) and f2(x2). It is well-known that the family given in (4.1) is a special case of the construction provided in [33] and includes some of the well-known Farlie-Gumbel-Morgenstern distributions as special cases. For details, see [34,35,36]. Observe that the dependence is captured through the mixing functions and the ω parameter.
Let us assume that a certain degree of dependence between the random variables Λ and Ψ exist, and specify a marginal prior gamma distribution for the first random variable and a translated Erlang distribution for the latter one, which, in both cases, are the conjugate a priori distribution of the Poisson and classical Pareto distributions. See the work of [37], where conjugate analysis for the Pareto distribution is studied.
We first need the following Lemma to obtain the prior distribution after assuming some degree of dependence.
Lemma 1. Let us consider the following functions
φ1(λ|ν,τ)=exp(−λ)−δ1(ν,τ),λ>0,ν>0,τ>0,φ2(ψ|γ,ξ,k)=exp(−ψ)−δ2(γ,ξ,k),ψ>k,γ>0,ξ>0,δ1(ν,τ)=(1+τ−1)−ν,δ2(γ,ξ,k)=(1+ξ−1)−γexp(−k). |
Then, it is verified that
H(ν,τ,γ,ξ,k,ω)=1+ωφ1(λ|ν,τ)φ2(ψ|γ,ξ,k)≥0, | (4.2) |
if ω1≤ω≤ω2, where
ω1=(−max{r1,r2})−1<0, | (4.3) |
ω2=(max{r3,r4})−1>0, | (4.4) |
being
r1=δ1(ν,τ)δ2(γ,ξ,k),r2=(1−δ1(ν,τ))(exp(−k)−δ2(γ,ξ,k)),r3=(1−δ1(ν,τ))δ2(γ,ξ,k),r4=δ1(ν,τ)(exp(−k)−δ2(γ,ξ,k)). |
Proof. To see this, observe that because d(φ1(λ|ν,τ))/dλ<0, we have that φ1(λ|ν,τ) is a decreasing function on λ; also, since the support of λ is in the interval (0,∞), we have that the range of variation of φ1(λ|ν,τ) is (−δ1(ν,τ),1−δ1(ν,τ)). Using the same argument, we get that the range of variation of φ2(ψ|γ,ξ,k) is given by (−δ2(γ,ξ,k),exp(−k)−δ2(γ,ξ,k)). Now we have that the expression H(ν,τ,γ,ξ,ω)≥0 if
ω≥−1φ1(λ|ν,τ)φ2(ψ|γ,ξ,k),forφ1(λ|ν,τ)φ2(ψ|γ,ξ,k)>0,ω≤−1φ1(λ|ν,τ)φ2(ψ|γ,ξ,k),forφ1(λ|ν,τ)φ2(ψ|γ,ξ,k)<0, |
from which, taking into account that 0<δ1(ν,τ)<1 and 0<δ2(γ,ξ,k)<1, it is a simple exercise to see that the range of ω is given by (ω1,ω2) provided above.
The following result provides the prior distribution under the Sarmanov-Lee family of distributions.
Proposition 5. The expression given by
π(λ,ψ|ν,τ,γ,ξ,k,ω)=π1(λ|ν,τ)π2(ψ|γ,ξ,k)H(ν,τ,γ,ξ,k,ω),λ>0,ψ>k, | (4.5) |
where H(ν,τ,γ,ξ,k,ω) is given in (4.2), defines a genuine prior bivariate distribution of (Λ,Ψ) with marginal distributions given by π1(λ|ν,τ) and π2(ψ|γ,ξ,k), which always satisfies that ω1≤ω≤ω2 and ω1 and ω2 are given by (4.3) and (4.4), respectively.
Proof. When ω1<ω<ω2, we have that π(λ,ψ|ν,τ,γ,ξ,ω)>0. Now, taking into account that
∫∞0exp(−λ)π1(λ|ν,τ)dλ=δ1(ν,τ),∫∞kexp(−ψ)π2(ψ|γ,ξ,k)dψ=δ2(γ,ξ,k), |
we get in a straightforward way that
∫∞0∫∞kπ(λ,ψ|ν,τ,γ,ξ,k,ω)dλdψ=∫∞0∫∞kπ1(λ|ν,τ)π2(ψ|γ,ξ,k)dλdψ=1. |
Hence the result.
As can be seen, the prior bivariate distribution (4.5) can be written as a linear combination of the product of gamma and Erlang distributions. Additionally, the special case ω=0 provides the prior bivariate distribution given by π(λ,ψ|ν,τ,γ,ξ,k)=π1(λ|ν,τ)π2(ψ|γ,ξ,k), i.e. it represents the case in which we assume independence between Λc and Ψ. Figure 1 shows the pdf given in (4.5) for parameters values given by (τ,ν,γ,ξ)=(5,10,10,4) and different values of ω moving in the range [−1.24,30.19] obtained from (4.3)-(4.4). The discrepancies between the different graphs are best appreciated in the contour plots given in Figure 1 below.
Some properties of the bivariate distribution proposed in this section are now studied.
Lemma 2. The following relations are verified,
E[Λexp(−Λ)]=ντδ1(ν+1,τ), | (4.6) |
E[Ψexp(−Ψ)]=(k+γξ+1)δ2(ξ,γ)exp(−k). | (4.7) |
Proof. See the Appendix.
The next result shows the cross moment and correlation coefficient between the random variables Λc and Ψ when the prior distribution (4.5) is used.
Proposition 6. For the prior bivariate distribution (4.5), the cross moment and correlation coefficient between Λc and Ψ are given by
E(ΛΨ)=ντξ(1+ξ)[(1+ξ)(γ+kξ)+ωγτδ1(ν+1,τ)δ2(γ,ξ,k)],ρ(Λ,Ψ)=ω√νγδ1(ν+1,τ)δ2(γ,ξ,k)τ(1+ξ). | (4.8) |
Proof. By using Theorems 1 and 2 in [15], we have that
E(ΛΨ)=μ1μ2+ωη1(ν,τ)η2(γ,ξ,k),ρ=ωη1(ν,τ)η2(γ,ξ,k)/(σ1σ2), |
where
η1(ν,τ)=∫∞0λφ1(λ|ν,τ)π1(λ|ν,τ)dλ,η2(γ,ξ,k)=∫∞kψφ2(ψ|γ,ξ,k)π2(ψ|γ,ξ,k)dψ, |
μ1, μ2, σ1 and σ2, the means and standard deviations of the gamma pdf (2.4) and Erlang pdf (2.5), respectively. These are given by μ1=ν/τ, σ1=√ν/τ, μ2=k+γ/ξ, and σ2=γ/ξ2.
Now, by using the expressions (4.6) and (4.7) provided in Lemma 2 and some straightforward algebra, we get the result.
Observe that the correlation coefficient in (4.8) is directly proportional to ω. The proportionality factor, given by √νγδ1(ν+1,τ)δ2(γ,ξ,k)/(τ(1+ξ)), is always positive, and the sign of the correlations will depend on the sign of ω, which allows values in the range provided by (4.3)-(4.4). A study of the linear correlation coefficient as a function of the values of the hyperparameters is carried out in the next proposition.
Proposition 7. Assume that ω>0. Then, the following is verified:
(1) ρ(Λ,Ψ) is an increasing (decreasing) function on ν if τ>(<)[exp(1/(2ν)−1]−1, for all ζ>0 and γ∈N.
(2) ρ(Λ,Ψ) is an increasing (decreasing) function on τ if ν>(<)τ, for all ζ>0 and γ∈N.
(3) ρ(Λ,Ψ) is an increasing (decreasing) function on γ if ζ>(<)[exp(1/(2γ)−1]−1, for all τ>0 and ν>0.
(4) ρ(Λ,Ψ) is an increasing (decreasing) function on ζ if γ>(<)ζ, for all τ>0 and ν>0.
Proof. See the Appendix.
Figure 2 plots the correlation given as a function of ω. The correlation coefficients seem limited to values between -0.15 and 0.15, coherent with the correlation in most empirical cases found in actuarial statistics.
The posterior distribution can be computed easily, and the next result shows that this posterior distribution can be written as a linear combination of the posterior distribution studied in Section 2.
Proposition 8. The posterior distribution of (Λc,Ψ) given the sample information ℵ is a linear combination of products of posterior univariate gamma and Erlang densities, given by
π∗(λ,ψ|ℵ)=κp{[1+ωδ1(ν,τ)δ2(γ,ξ,k)](τ∗)−n−νκπ∗1(λ|ν∗,τ∗)π∗2(ψ|γ∗,ξ∗,k)+ω(τ∗+1)−n−νκ∗π∗1(λ|ν∗,τ∗+1)π∗2(ψ|γ∗,ξ∗+1,k)−ωδ1(ν,τ)(τ∗)−n−νκ∗π∗1(λ|ν∗,τ∗)π∗2(ψ|γ∗,ξ∗+1,k)−ωδ2(γ,ξ,k)(τ∗+1)−n−νκπ∗1(λ|ν∗,τ∗+1)π∗2(ψ|γ∗,ξ∗,k)}, | (4.9) |
where κ∗≡κ(γ∗,ξ∗+1,k) is given in (2.11) and the normalization constant, κp≡κp(γ∗,ξ∗,k), satisfies
κ−1p=1+ωδ1(ν,τ)δ2(γ,ξ)κ(τ∗)−ν−n+ωκ∗(τ∗+1)−ν−n−ωκ∗δ1(ν,τ)(τ∗)−ν−n−ωκ(τ∗+1)−ν−nδ2(γ,ξ,k). |
Proof. The result follows after some intensive algebra by using (2.3), (4.5), Bayes' Theorem and arranging parameters.
Expressions given in (4.5) and (4.9) can be used now to get the collective and Bayes net premiums under the non-independent model. These are shown in the following result.
Proposition 9. Under the bivariate model with dependence, the prior (collective) and Bayesian (posterior) premiums are given by,
Eπ[P(Λ,Ψ)]=[1+ωδ1(μ,τ)δ2(ξ,γ)]P(ν,τ,γ,ξ)+ωδ1(μ,τ)δ2(ξ,γ)[P(ν,τ+1,γ,ξ+1)−P(ν,τ,γ,ξ+1)−P(ν,τ+1,γ,ξ)], | (4.10) |
Eπ∗[P(Λ,Ψ)|ℵ]=κp{[1+ωδ1(ν,τ)δ2(γ,ξ,k)](τ∗)−n−νκP∗(ν∗,τ∗,γ∗,ξ∗,1),k)+ω(τ∗+1)−n−νκ∗P∗(ν∗,τ∗+1,γ∗,ξ∗+1,1)−ωδ1(ν,τ)(τ∗)−n−νκ∗P∗(ν∗,τ∗,γ∗,ξ∗+1,1)−ωδ2(γ,ξ,k)(τ∗+1)−n−νκP∗(ν∗,τ∗+1,γ∗,ξ∗,1)}, | (4.11) |
where P(ν,τ,γ,ξ) is the collective premium under the different models studied previously, the basic specific model (expression given in (2.12)), the reinsurance model with a layer (expression given in (3.5)), and the reinsurance model without a layer (expression given in (3.7)). Alternatively, P∗(ν∗,τ∗,γ∗,ξ∗,1) is the net Bayes premium under the models studied, expressions (2.13), (3.6), and (3.8).
Proof. See the Appendix.
In the following corollary, we give an alternative expression for (4.10).
Corollary 1. The collective premium given in (4.10) can be rewritten as
Eπ[P(Λ,Ψ)]=P(ν,τ,γ,ξ)+ωδ1(μ,τ)δ2(ξ,γ)τ+1[P(ν,τ,γ,ξ)−P(ν,τ,γ,ξ+1)]. |
Proof. See the Appendix.
We get the corresponding collective and Bayes premium under the independent model by taking ω=0 in (4.10) and (4.11).
Note that the Bayesian premium has been calculated sequentially rather than in a batch, which means that the count of the number of losses may not be a homogeneous Poisson over time.
In this section, two numerical examples will be considered. We first used the same dataset for comparison in [10,11]. The second dataset corresponds to catastrophe data regarding the total damage caused by a storm during 1956–1977, which is displayed in Table 1. The amounts and the year of occurrence are exhibited in this table. We have only considered losses below 6 billion dollars. A comprehensive version of this dataset is shown in [38, p. 128]. For this reason, in our analysis, we only consider 10 entries.
Number | Loss | Year |
1 | 2,000 | 1977 |
2 | 1,380 | 1971 |
3 | 2,000 | 1971 |
4 | 2,000 | 1964 |
5 | 2,580 | 1968 |
6 | 4,730 | 1971 |
7 | 3,700 | 1956 |
8 | 4,250 | 1961 |
9 | 4,500 | 1966 |
10 | 5,000 | 1958 |
The elicited hyperparameter values of the prior distributions are shown in Table 2, together with the bounds of the ω parameter and other statistics needed to obtain the collective and Bayesian net premiums. The maximum likelihood estimation method requires knowledge of the distribution obtained from (2.1), which is not simple; hence, the hyperparameters were estimated by the method of moments, where for the parameters (ν,τ), the sample mean and index of dispersion (ratio of variance to the mean assumed to be approximately 1.25) were used to calculate the estimates. Furthermore, for the parameters (ξ,γ), we used the sample mean with the coefficient of variation and the approximation of the value of γ to the next integer. The sample sizes, the values of the layers, and the corresponding time periods for each data set are also given.
Example 1 | Example 2 | |
ν | 2.56 | 0.3636 |
τ | 0.8 | 0.8 |
ξ | 0.972 | 1.197 |
γ | 2 | 2.65≈3 |
z | 6.48165 | 8.54057 |
c | 1.5 | 1.25 |
ω1 | -4.10 | -12.70 |
ω2 | 12.79 | 4.35 |
layer | 5 | 2 |
T | 5 | 21 |
n | 16 | 10 |
The collective and Bayesian net premiums computed using (4.10) and (4.11) are shown in Table 3 for the two examples considered and the basic collective model. Observe that the case ω=0 reduces to the model where independence between the risk profiles is assumed. The correlation coefficient obtained for each value of ω obtained via (4.8) is also shown in this table. It can be observed that incorporating dependence has a low impact on both premiums, as judged by the changes in ω when this dependence parameter increases or decreases (i.e., more separated from the hypotheses of independence).
Example 1 | Example 2 | |||||||
ω | Correlation | Collective | Bayes | ω | ρ | Collective | Bayes | |
-4 | -0.028 | 9.5854 | 8.5706 | -12 | -0.140 | 0.9920 | 1.8074 | |
-3 | -0.021 | 9.5555 | 8.5631 | -11 | -0.128 | 0.9850 | 1.7991 | |
-2 | -0.014 | 9.5256 | 8.5555 | -10 | -0.117 | 0.9780 | 1.7906 | |
-1 | -0.007 | 9.4957 | 8.5480 | -9 | -0.105 | 0.9710 | 1.7819 | |
0 | 0.000 | 9.4658 | 8.5404 | -8 | -0.093 | 0.9640 | 1.7729 | |
1 | 0.007 | 9.4359 | 8.5329 | -7 | -0.081 | 0.9570 | 1.7637 | |
2 | 0.014 | 9.4060 | 8.5253 | -6 | -0.070 | 0.9500 | 1.7542 | |
3 | 0.021 | 9.3761 | 8.5177 | -5 | -0.058 | 0.9431 | 1.7445 | |
4 | 0.028 | 9.3462 | 8.5102 | -4 | -0.046 | 0.9361 | 1.7345 | |
5 | 0.035 | 9.3163 | 8.5026 | -3 | -0.035 | 0.9291 | 1.7242 | |
6 | 0.042 | 9.2864 | 8.4950 | -2 | -0.023 | 0.9221 | 1.7136 | |
7 | 0.050 | 9.2565 | 8.4874 | -1 | -0.011 | 0.9151 | 1.7027 | |
8 | 0.057 | 9.2266 | 8.4798 | 0 | 0.000 | 0.9081 | 1.6914 | |
9 | 0.064 | 9.1967 | 8.4722 | 1 | 0.011 | 0.9011 | 1.6799 | |
10 | 0.071 | 9.1668 | 8.4647 | 2 | 0.023 | 0.8941 | 1.6679 | |
11 | 0.078 | 9.1369 | 8.4571 | 3 | 0.035 | 0.8871 | 1.6556 | |
12 | 0.085 | 9.1070 | 8.4495 | 4 | 0.046 | 0.8801 | 1.6428 |
The collective and Bayesian net premiums, including a layer of 1.5 million in excess of the priority a for Example 1 and a layer of 2 million in excess of the priority a for Example 2, are shown in Tables 4 and 5, respectively. Again, the case ω=0 reduces to the model where independence between the risk profiles is assumed, and the correlation coefficient is also included. It can be noticed that increasing the value of the deductible a decreases the value of both premiums. In cases where the deductible is 1.5 million, the Bayes premium is higher than the collective premiums in both examples.
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-4 | -0.028 | 11.1805 | 8.0838 | 2.8365 | 2.9943 | 1.5891 | 1.6574 |
-3 | -0.021 | 11.1990 | 8.0846 | 2.8288 | 2.9911 | 1.5823 | 1.6545 |
-2 | -0.014 | 11.2175 | 8.0855 | 2.8211 | 2.9879 | 1.5755 | 1.6517 |
-1 | -0.007 | 11.2361 | 8.0863 | 2.8135 | 2.9847 | 1.5687 | 1.6489 |
0 | 0.000 | 11.2546 | 8.0871 | 2.8058 | 2.9816 | 1.5619 | 1.6460 |
1 | 0.007 | 11.2731 | 8.0880 | 2.7981 | 2.9784 | 1.5551 | 1.6432 |
2 | 0.014 | 11.2917 | 8.0888 | 2.7905 | 2.9752 | 1.5483 | 1.6403 |
3 | 0.021 | 11.3102 | 8.0896 | 2.7828 | 2.9720 | 1.5415 | 1.6374 |
4 | 0.028 | 11.3287 | 8.0905 | 2.7751 | 2.9688 | 1.5347 | 1.6346 |
5 | 0.036 | 11.3472 | 8.0913 | 2.7675 | 2.9656 | 1.5279 | 1.6317 |
6 | 0.042 | 11.3658 | 8.0922 | 2.7598 | 2.9625 | 1.5211 | 1.6289 |
7 | 0.050 | 11.3843 | 8.0930 | 2.7521 | 2.9593 | 1.5143 | 1.6260 |
8 | 0.057 | 11.4028 | 8.0938 | 2.7445 | 2.9561 | 1.5075 | 1.6232 |
9 | 0.064 | 11.4214 | 8.0947 | 2.7368 | 2.9529 | 1.5007 | 1.6203 |
10 | 0.071 | 11.4399 | 8.0955 | 2.7291 | 2.9497 | 1.4939 | 1.6174 |
11 | 0.078 | 11.4584 | 8.0964 | 2.7214 | 2.9465 | 1.4871 | 1.6146 |
12 | 0.085 | 11.4769 | 8.0972 | 2.7138 | 2.9433 | 1.4803 | 1.6117 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-12 | -0.140 | 0.7151 | 0.7078 | 0.1829 | 0.3554 | 0.0914 | 0.2264 |
-11 | -0.128 | 0.7174 | 0.7059 | 0.1808 | 0.3542 | 0.0899 | 0.2255 |
-10 | -0.117 | 0.7197 | 0.7040 | 0.1787 | 0.3529 | 0.0883 | 0.2246 |
-9 | -0.105 | 0.7220 | 0.7020 | 0.1767 | 0.3516 | 0.0868 | 0.2236 |
-8 | -0.093 | 0.7243 | 0.7000 | 0.1746 | 0.3503 | 0.0852 | 0.222 |
-7 | -0.081 | 0.7266 | 0.6979 | 0.1725 | 0.3489 | 0.0837 | 0.2217 |
-6 | -0.070 | 0.7290 | 0.6957 | 0.1704 | 0.3475 | 0.0822 | 0.2207 |
-5 | -0.058 | 0.7313 | 0.6935 | 0.1683 | 0.3461 | 0.0806 | 0.2196 |
-4 | -0.046 | 0.7336 | 0.6913 | 0.1662 | 0.3446 | 0.0791 | 0.2185 |
-3 | -0.035 | 0.7359 | 0.6889 | 0.1641 | 0.3431 | 0.0775 | 0.2174 |
-2 | -0.023 | 0.7382 | 0.6865 | 0.1620 | 0.3415 | 0.0760 | 0.2163 |
-1 | -0.011 | 0.7405 | 0.6840 | 0.1599 | 0.3398 | 0.0744 | 0.2151 |
0 | 0.000 | 0.7420 | 0.6815 | 0.1578 | 0.3382 | 0.0729 | 0.2139 |
1 | 0.011 | 0.7452 | 0.6789 | 0.1557 | 0.3364 | 0.0714 | 0.2126 |
2 | 0.023 | 0.7475 | 0.6762 | 0.1536 | 0.3347 | 0.0698 | 0.2113 |
3 | 0.035 | 0.7498 | 0.6734 | 0.1516 | 0.3328 | 0.0683 | 0.2100 |
4 | 0.046 | 0.7521 | 0.6705 | 0.1495 | 0.3309 | 0.0667 | 0.2086 |
In addition, the collective and Bayesian net premiums without including a layer in excess of the priority a for both examples are shown in Tables 6 and 7, respectively. Again, the case ω=0 reduces to the model where independence between the risk profiles is assumed, and the correlation coefficient is also shown in these tables. It can be noticed that increasing the value of the deductible a decreases the value of both premiums. In cases where the deductible is 1.5 million, the Bayes premium is higher than the collective premiums in both examples. In Example 1, it can be observed that increasing the value of the deductible a decreases the value of both premiums. However, this is not verified in the second example, where the Bayes premiums increase for the deductible value a=1.5 million and decrease again for a=2.2. Also, the collective premiums increase for the deductible of a=2.2 million.
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-4 | -0.028 | 13.2230 | 8.9452 | 4.7854 | 3.7701 | 3.4611 | 2.3652 |
-3 | -0.021 | 13.2187 | 8.9416 | 4.7555 | 3.7627 | 3.4326 | 2.3584 |
-2 | -0.014 | 13.2144 | 8.9379 | 4.7256 | 3.7553 | 3.4041 | 2.3515 |
-1 | -0.007 | 13.2101 | 8.9343 | 4.6957 | 3.7478 | 3.3756 | 2.3446 |
0 | 0.000 | 13.2058 | 8.9306 | 4.6658 | 3.7404 | 3.3470 | 2.3377 |
1 | 0.007 | 13.2015 | 8.9270 | 4.6359 | 3.7330 | 3.3185 | 2.3308 |
2 | 0.014 | 13.1972 | 8.9233 | 4.6060 | 3.7256 | 3.2900 | 2.3239 |
3 | 0.021 | 13.1929 | 8.9197 | 4.5761 | 3.7181 | 3.2615 | 2.3170 |
4 | 0.028 | 13.1886 | 8.9160 | 4.5462 | 3.7107 | 3.2330 | 2.3101 |
5 | 0.036 | 13.1843 | 8.9123 | 4.5163 | 3.7032 | 3.2044 | 2.3032 |
6 | 0.042 | 13.1800 | 8.9087 | 4.4864 | 3.6958 | 3.1759 | 2.2963 |
7 | 0.050 | 13.1758 | 8.9050 | 4.4565 | 3.6883 | 3.1474 | 2.2894 |
8 | 0.057 | 13.1715 | 8.9013 | 4.4266 | 3.6809 | 3.1189 | 2.2825 |
9 | 0.064 | 13.1672 | 8.8977 | 4.3967 | 3.6734 | 3.0903 | 2.2756 |
10 | 0.071 | 13.1629 | 8.8940 | 4.3668 | 3.6660 | 3.0618 | 2.2687 |
11 | 0.078 | 13.1586 | 8.8903 | 4.3369 | 3.6585 | 3.0333 | 2.2617 |
12 | 0.085 | 13.1543 | 8.8866 | 4.3070 | 3.6510 | 3.0048 | 2.2548 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-12 | -0.140 | 0.8994 | 0.3378 | 0.2273 | 1.5146 | 1.0860 | 0.9017 |
-11 | -0.128 | 0.8965 | 0.3310 | 0.2214 | 1.5072 | 1.0796 | 0.8959 |
-10 | -0.117 | 0.8935 | 0.3242 | 0.2156 | 1.4996 | 1.0729 | 0.8898 |
-9 | -0.105 | 0.8906 | 0.3174 | 0.2097 | 1.4918 | 1.0661 | 0.8836 |
-8 | -0.093 | 0.8877 | 0.3106 | 0.2038 | 1.4838 | 1.0592 | 0.8773 |
-7 | -0.081 | 0.8848 | 0.3038 | 0.1980 | 1.4755 | 1.0520 | 0.8707 |
-6 | -0.070 | 0.8819 | 0.2970 | 0.1921 | 1.4671 | 1.0446 | 0.8640 |
-5 | -0.058 | 0.8789 | 0.2902 | 0.1862 | 1.4584 | 1.0370 | 0.8571 |
-4 | -0.046 | 0.8760 | 0.2834 | 0.1804 | 1.4494 | 1.0292 | 0.8500 |
-3 | -0.035 | 0.8731 | 0.2766 | 0.1745 | 1.4402 | 1.0212 | 0.8427 |
-2 | -0.023 | 0.8702 | 0.2698 | 0.1687 | 1.4307 | 1.0129 | 0.8352 |
-1 | -0.011 | 0.8673 | 0.2630 | 0.1628 | 1.4210 | 1.0044 | 0.8275 |
0 | 0.000 | 0.8643 | 0.2562 | 0.1569 | 1.4109 | 0.9957 | 0.8195 |
1 | 0.011 | 0.8614 | 0.2494 | 0.1511 | 1.4006 | 0.9866 | 0.8113 |
2 | 0.023 | 0.8585 | 0.2426 | 0.1452 | 1.3899 | 0.9773 | 0.8028 |
3 | 0.035 | 0.8556 | 0.2358 | 0.1394 | 1.3789 | 0.9677 | 0.7940 |
4 | 0.046 | 0.8527 | 0.2290 | 0.1335 | 1.3675 | 0.9578 | 0.7850 |
Even though, for all the tables above, the predicted correlation coefficients seem to be minor, this is not a trivial issue since it could have a significant impact when changes of scales in the monetary units are considered; therefore, the analysis we have carried out in this work shows that although the correlation is usually low, the impact of assuming independence might have potential financial consequences. It is important to note that the dependence hypothesis is established between the risk profiles and not directly between the number of claims and the corresponding severity. However, given the mathematical difficulty that the latter scenario would entail, this can be replaced by transferring this hypothesis to the risk profiles, which retain characteristics of the former case. Even though when calculating bonus-malus premiums in the automobile insurance portfolio, independence between the number of claims and the amount of the claim is usually assumed, some recent works such as [39] have discussed the calculation of bonus-malus premiums for the compound loss by incorporating a certain degree of dependence between the frequency and severity. However, in practice, the correlation between both variables is usually low. For example, by considering the automobile insurance portfolio used in [40] and taken from [41] that corresponds to the years 2004-05 and contains data of 67,856 insureds, of whom 4,624 filed a claim, i.e., a positive claims amount, the correlation between the number of claims and the amount for the whole portfolio is 0.4818. Nevertheless, when it is restricted to the 4,624 policyholders that have declared at least one claim, the correlation drops to 0.0762, consistent with the values shown in Table 3.
We have presented an excess-of-loss reinsurance model derived from the compound Poisson distribution. The classical Pareto distribution was assumed for the severity of claims, and dependence between the number of claims and claims size was placed on the risk profiles. In this work, initially, we considered independence between the risk profiles whose parameters follow gamma and shifted Erlang distributions. Next, the hypothesis of independence was broken by assuming some degree of dependence between the random variables associated with these risk profiles via a prior bivariate distribution based on the Sarmanov-Lee probabilistic family. This model was used to study the effect of incorporating some degree of dependence of these variables on the collective and Bayesian net premiums. The results obtained lead us to conclude that even at moderate levels of correlation between the risk profiles, these aforementioned reinsurance premiums are more sensitive than those computed, assuming the independence of the risk profiles.
The methodology proposed in this paper is suitable for application in the collective model, in which the two relevant variables used in calculating premiums are used simultaneously: the number of claims and the amount associated with them. The difficulty of working simultaneously with these two variables is well-known, especially when obtaining the total claim amount distribution. Two scenarios that should be considered are, on the one hand, the study of the sensitivity of the obtained premiums to slight modifications in the choice of the prior distribution. To that end, we are discussing carrying out a Bayesian robustness analysis (local or global). On the other hand, choosing the Sarmanov-Lee family to construct the prior distribution that allows for a certain degree of dependence between the two parameters can be very rigid (it has the advantage of being almost conjugate) and could be replaced by a bivariate distribution based on another type of copula.
The author declares he has not used Artificial Intelligence (AI) tools in the creation of this article.
EGD was partially funded by grant PID2021-127989OB-I00 (Ministerio de Economía y Competitividad, Spain).
Prof. Emilio Gómez-Déniz is the Guest Editor of special issue "Advances in Probability Distributions and Social Science Statistics" for AIMS Mathematics. Prof. Emilio Gómez-Déniz was not involved in the editorial review and the decision to publish this article.
The author declares no conflict of interest.
[1] |
Arif M, Naeem MA, Farid S, et al. (2021) Diversifier or more? Hedge and safe haven properties of green bonds during COVID-19. Energ Policy 168: 113102. https://doi.org/10.1016/j.enpol.2022.113102 doi: 10.1016/j.enpol.2022.113102
![]() |
[2] |
Avkiran NK, Cai L (2014) Identifying distress among banks prior to a major crisis using non-oriented super-SBM. Ann Oper Res 217: 31–53. https://doi.org/10.1007/s10479-014-1568-8 doi: 10.1007/s10479-014-1568-8
![]() |
[3] |
Broock WA, Scheinkman JA, Dechert WD, et al. (1996) A test for independence based on the correlation dimension. Economet Rev 15: 197–235. https://doi.org/10.1080/07474939608800353 doi: 10.1080/07474939608800353
![]() |
[4] | Bloomberg NEF (2020) Power sector to spend $5 billion on software by 2025. |
[5] |
Chang K, Ge F, Zhang C, et al. (2018) The dynamic linkage effect between energy and emissions allowances price for regional emissions trading scheme pilots in China. Renew Sust Energy Rev 98: 415–425. https://doi.org/10.1016/j.rser.2018.09.023 doi: 10.1016/j.rser.2018.09.023
![]() |
[6] |
Chai S, Chu W, Zhang Z, et al. (2022) Dynamic nonlinear connectedness between the green bonds, clean energy, and stock price: the impact of the COVID-19 pandemic. Ann Oper Res 2022: 1–28. https://doi.org/10.1007/s10479-021-04452-y doi: 10.1007/s10479-021-04452-y
![]() |
[7] |
Chan HY, Merdekawati M, Suryadi B (2022) Bank climate actions and their implications for the coal power sector. Energy Strateg Rev 39: 100799. https://doi.org/10.1016/j.esr.2021.100799 doi: 10.1016/j.esr.2021.100799
![]() |
[8] |
Chen Y, Jiang Y (2023) Integration of green energy equity and fossil energy markets in different time scales: evidence from the US, Europe and China. Int J Environ Pollut 72: 198–221. https://doi.org/10.1504/IJEP.2023.10060227 doi: 10.1504/IJEP.2023.10060227
![]() |
[9] |
Esmaeili P, Rafei M (2021) Dynamics analysis of factors affecting electricity consumption fluctuations based on economic conditions: Application of SVAR and TVP-VAR models. Energy 226: 120340. https://doi.org/10.1016/j.energy.2021.120340 doi: 10.1016/j.energy.2021.120340
![]() |
[10] | Engle Ⅲ RF, Sheppard K (2001) Theoretical and empirical properties of dynamic conditional correlation multivariate GARCH. https://doi.org/10.3386/w8554 |
[11] | Engle RF (2002) A simple class of multivariate generalized autoregressive conditional heteroskedasticity models. J Bus Econ Stat 20: 339–350. |
[12] |
Gong X, Shi R, Xu J, et al. (2021) Analyzing spillover effects between carbon and fossil energy markets from a time-varying perspective. Appl Energ 285: 116384. https://doi.org/10.1016/j.apenergy.2020.116384 doi: 10.1016/j.apenergy.2020.116384
![]() |
[13] |
Hanif W, Hernandez JA, Mensi W, et al. (2021) Nonlinear dependence and connectedness between clean renewable energy sector equity and European emission allowance prices. Energ Econ 101: 105409. https://doi.org/10.1016/j.eneco.2021.105409 doi: 10.1016/j.eneco.2021.105409
![]() |
[14] |
He L, Zhang L, Zhong Z, et al. (2019) Green credit, renewable energy investment and green economy development: Empirical analysis based on 150 listed companies of China. J Clean Prod 208: 363–372. https://doi.org/10.1016/j.jclepro.2018.10.119 doi: 10.1016/j.jclepro.2018.10.119
![]() |
[15] |
He Z (2020) Dynamic impacts of crude oil price on Chinese investor sentiment: Nonlinear causality and time-varying effect. Int Rev Econ Financ 66: 131–153. https://doi.org/10.1016/j.irej.iref.2019.11.004 doi: 10.1016/j.irej.iref.2019.11.004
![]() |
[16] |
Hammoudeh S, Ajmi AN, Mokni K (2021) Relationship between green bonds and financial and environmental variables: A novel time-varying causality. Energ Econ 92: 104941. https://doi.org/10.1016/j.eneco.2020.104941 doi: 10.1016/j.eneco.2020.104941
![]() |
[17] |
Li H, Li J, Jiang Y (2023) Exploring the Dynamic Impact between the Industries in China: New Perspective Based on Pattern Causality and Time-Varying Effect. Systems 11: 318. https://doi.org/10.3390/systems11070318 doi: 10.3390/systems11070318
![]() |
[18] | IRENA (2019) Innovation landscape for a renewable-powered future: Solutions to integrate variable renewables. Int Renew Energ Agency, Abu Dhabi. |
[19] |
Ji Q, Zhang D, Geng J (2018) Information linkage, dynamic spillovers in prices and volatility between the carbon energy markets. J Clean Prod 198: 972–978. https://doi.org/10.1016/j.jclepro.2018.07.126 doi: 10.1016/j.jclepro.2018.07.126
![]() |
[20] |
Ji Q, Xia T, Liu F, et al (2019) The information spillover between carbon price and power Sector returns: Evidence from the major European electricity companies. J Clean Prod 208: 1178–1187. https://doi.org/10.1016/j.jclepro.2018.10.167 doi: 10.1016/j.jclepro.2018.10.167
![]() |
[21] |
Li P, Zhang H, Yuan Y, et al. (2021) Time-varying impacts of carbon price drives in the EU ETS: A TETS: A TVP-VAR Analysis. Front Env Sci 9: 651791. https://doi.org/10.3389/fenvs.2021.651791 doi: 10.3389/fenvs.2021.651791
![]() |
[22] |
Li Y, Nie D, Li B, et al. (2020) The Spillover Effect between Carbon Emission Trading Price and Power Company Stock Price in China. Sustainability 12: 6573. https://doi.org/10.3390/su12166573 doi: 10.3390/su12166573
![]() |
[23] |
Li F, Cao X, Ou R (2021) A network-based evolutionary analysis of the diffusion of cleaner energy substitution in enterprises: the roles of PEST factors. Energ Policy 156: 112385. https://doi.org/10.1016/j.enpol.2021.112385 doi: 10.1016/j.enpol.2021.112385
![]() |
[24] |
Lin Z, Liao X, Jia H (2023) Could green finance facilitate low-carbon transformation of power generation? Some evidence from China. Int J Clim Chang Strategies Manage 15: 141–158. https://doi.org/10.1108/IJCCSM-03-2022-0039 doi: 10.1108/IJCCSM-03-2022-0039
![]() |
[25] |
Nong H, Guan Y, Jiang Y (2022) Identifying the volatility spillover risks between crude oil price and China's clean energy market. Electro Res Arch 30: 4593–4618. https://doi.org/10.3934/era.2022233 doi: 10.3934/era.2022233
![]() |
[26] |
Nguyen TTH, Naeem MA, Balli F, et al. (2021) Time-frequency co-movement among green bonds, stocks, commodities, clean energy, and conventional bonds. Financ Res Lett 40: 101739. https://doi.org/10.1016/j.frl.2020.101739 doi: 10.1016/j.frl.2020.101739
![]() |
[27] | Nakajima J (2011) Time-varying Parameter VAR Model with Stochastic Volatility: An Overview of Methodology and Empirical Applications. Inst Monetary Econ Stud, Bank of Japan. |
[28] |
Pham L (2021) Frequency connectedness and cross-quantile dependence between green bond and green equity markets. Energ Econ 98: 105257. https://doi.org/10.1016/j.eneco.2021.105257 doi: 10.1016/j.eneco.2021.105257
![]() |
[29] |
Piotr F, Witold O (2018) Nonlinear granger causality between grains and livestock. Agr Econ 64: 328–336. https://doi.org/10.17221/376/2016-AGRICECON doi: 10.17221/376/2016-AGRICECON
![]() |
[30] | Primiceri GE (2005) Time-Varying Structural Vector Autoregressions and Monetary Policy. Rev Econ Stud 72: 821–852. |
[31] |
Ren X, Cheng C, Wang Z, et al. (2021) Spillover and dynamic effects of energy transition and economic growth on carbon dioxide emissions for the European Union: a dynamic spatial panel model. Sustain Devt 29: 228–242. https://doi.org/10.1002/sd.2144 doi: 10.1002/sd.2144
![]() |
[32] |
Ren X, Li Y, Wen F, et al. (2022) The interrelationship between the carbon market and the green bonds market: Evidence from wavelet quantile-on-quantile method. Technol Forecast Soc Chang 179: 121611. https://doi.org/10.1016/j.techfore.2022.121611 doi: 10.1016/j.techfore.2022.121611
![]() |
[33] |
Reboredo JC, Ugolini A (2020) Price connectedness between green bond and financial markets. Econ Model 88: 25–38. https://doi.org/10.1016/j.econmod.2019.09.004 doi: 10.1016/j.econmod.2019.09.004
![]() |
[34] |
Samuel Asante Gyamerah, Clement Asare (2024) A critical review of the impact of uncertainties on green bonds. Green Financ 6: 78–91. https://doi: 10.3934/GF.2024004 doi: 10.3934/GF.2024004
![]() |
[35] |
Strantzali E, Aravossis K (2016) Decision making in renewable energy investments: A review. Renew Sust Energ Rev 55: 885–898. https://doi.org/10.1016/j.rser.2015.11.021 doi: 10.1016/j.rser.2015.11.021
![]() |
[36] |
Wen F, Zhao H, Zhao L, et al. (2022) What drive carbon price dynamics in China. Int Rev Financ Anal 79: 101999. https://doi.org/10.1016/j.irfa.2021.101999 doi: 10.1016/j.irfa.2021.101999
![]() |
[37] |
Wu Y, Wang J, Ji S, et al. (2020) Renewable energy investment risk assessment for nations along China's Belt & Road Initiative: An ANP-cloud model method. Energy 190: 116381. https://doi.org/10.1016/j.energy.2019.116381 doi: 10.1016/j.energy.2019.116381
![]() |
[38] | Xiao H, Zhang Z, Wang A, et al. (2021) Evaluating energy security in China: a subnational analysis. In China's Energy Security: Analysis, Assessment and Improvement, 119–137. https://doi.org/10.1142/97817863492240005 |
[39] |
Yang L (2022) Idiosyncratic information spillover and connectedness network between the electricity and carbon markets in Europe. J Commod Mark 25: 100185. https://doi.org/10.1016/j.jcomm.2021.100185 doi: 10.1016/j.jcomm.2021.100185
![]() |
[40] |
Yin G, Li B, Fedorova N, et al. (2021) Orderly retire China's coal-fired power capacity via capacity payments to support renewable energy expansion. Iscience 24. https://doi.org/10.1016/j.isci.2021.103287 doi: 10.1016/j.isci.2021.103287
![]() |
[41] |
Zhao L, Liu W, Zhou M, et al. (2022) Extreme event shocks and dynamic volatility interactions: The stock, commodity, and carbon markets in China. Financ Res Lett 47: 102645.https://doi.org/10.1016/j.frl.2021.102645 doi: 10.1016/j.frl.2021.102645
![]() |
[42] |
Zhao X, Li Q, Xue W, et al. (2022) Research on ultra-short-term load forecasting based on real-time electricity price and window-based XGBoost model. Energies 15: 7367. https://doi.org/10.3390/en15197367 doi: 10.3390/en15197367
![]() |
[43] |
Zhao Y, Zhou Z, Zhang K, et al. (2023) Research on spillover effect between carbon market and electricity market: Evidence from Northern Europe. Energy 263: 126107. https://doi.org/10.1016/j.energy.2022.126107 doi: 10.1016/j.energy.2022.126107
![]() |
[44] |
Zhou K, Li Y (2019) Influencing factors and fluctuation characteristics of China's carbon emission trading price. Physica A 524: 459–474. https://doi.org/10.1016/j.physa.2019.04.249 doi: 10.1016/j.physa.2019.04.249
![]() |
[45] |
Zhou D, Chen B, Li J, et al. (2021) China's economic growth, energy efficiency, and industrial development: Nonlinear effects on carbon dioxide emissions. Discrete Dyn Nat Soc 2021: 1–17. https://doi.org/10.1155/2021/5547092 doi: 10.1155/2021/5547092
![]() |
[46] |
Zhou Y, Wu S, Liu Z, et al. (2023) The asymmetric effects of climate risk on higher-moment connectedness among carbon, energy and metals markets. Nat Commun 14: 7157. https://doi.org/10.1038/s41467-023-42925-9 doi: 10.1038/s41467-023-42925-9
![]() |
Number | Loss | Year |
1 | 2,000 | 1977 |
2 | 1,380 | 1971 |
3 | 2,000 | 1971 |
4 | 2,000 | 1964 |
5 | 2,580 | 1968 |
6 | 4,730 | 1971 |
7 | 3,700 | 1956 |
8 | 4,250 | 1961 |
9 | 4,500 | 1966 |
10 | 5,000 | 1958 |
Example 1 | Example 2 | |
ν | 2.56 | 0.3636 |
τ | 0.8 | 0.8 |
ξ | 0.972 | 1.197 |
γ | 2 | 2.65≈3 |
z | 6.48165 | 8.54057 |
c | 1.5 | 1.25 |
ω1 | -4.10 | -12.70 |
ω2 | 12.79 | 4.35 |
layer | 5 | 2 |
T | 5 | 21 |
n | 16 | 10 |
Example 1 | Example 2 | |||||||
ω | Correlation | Collective | Bayes | ω | ρ | Collective | Bayes | |
-4 | -0.028 | 9.5854 | 8.5706 | -12 | -0.140 | 0.9920 | 1.8074 | |
-3 | -0.021 | 9.5555 | 8.5631 | -11 | -0.128 | 0.9850 | 1.7991 | |
-2 | -0.014 | 9.5256 | 8.5555 | -10 | -0.117 | 0.9780 | 1.7906 | |
-1 | -0.007 | 9.4957 | 8.5480 | -9 | -0.105 | 0.9710 | 1.7819 | |
0 | 0.000 | 9.4658 | 8.5404 | -8 | -0.093 | 0.9640 | 1.7729 | |
1 | 0.007 | 9.4359 | 8.5329 | -7 | -0.081 | 0.9570 | 1.7637 | |
2 | 0.014 | 9.4060 | 8.5253 | -6 | -0.070 | 0.9500 | 1.7542 | |
3 | 0.021 | 9.3761 | 8.5177 | -5 | -0.058 | 0.9431 | 1.7445 | |
4 | 0.028 | 9.3462 | 8.5102 | -4 | -0.046 | 0.9361 | 1.7345 | |
5 | 0.035 | 9.3163 | 8.5026 | -3 | -0.035 | 0.9291 | 1.7242 | |
6 | 0.042 | 9.2864 | 8.4950 | -2 | -0.023 | 0.9221 | 1.7136 | |
7 | 0.050 | 9.2565 | 8.4874 | -1 | -0.011 | 0.9151 | 1.7027 | |
8 | 0.057 | 9.2266 | 8.4798 | 0 | 0.000 | 0.9081 | 1.6914 | |
9 | 0.064 | 9.1967 | 8.4722 | 1 | 0.011 | 0.9011 | 1.6799 | |
10 | 0.071 | 9.1668 | 8.4647 | 2 | 0.023 | 0.8941 | 1.6679 | |
11 | 0.078 | 9.1369 | 8.4571 | 3 | 0.035 | 0.8871 | 1.6556 | |
12 | 0.085 | 9.1070 | 8.4495 | 4 | 0.046 | 0.8801 | 1.6428 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-4 | -0.028 | 11.1805 | 8.0838 | 2.8365 | 2.9943 | 1.5891 | 1.6574 |
-3 | -0.021 | 11.1990 | 8.0846 | 2.8288 | 2.9911 | 1.5823 | 1.6545 |
-2 | -0.014 | 11.2175 | 8.0855 | 2.8211 | 2.9879 | 1.5755 | 1.6517 |
-1 | -0.007 | 11.2361 | 8.0863 | 2.8135 | 2.9847 | 1.5687 | 1.6489 |
0 | 0.000 | 11.2546 | 8.0871 | 2.8058 | 2.9816 | 1.5619 | 1.6460 |
1 | 0.007 | 11.2731 | 8.0880 | 2.7981 | 2.9784 | 1.5551 | 1.6432 |
2 | 0.014 | 11.2917 | 8.0888 | 2.7905 | 2.9752 | 1.5483 | 1.6403 |
3 | 0.021 | 11.3102 | 8.0896 | 2.7828 | 2.9720 | 1.5415 | 1.6374 |
4 | 0.028 | 11.3287 | 8.0905 | 2.7751 | 2.9688 | 1.5347 | 1.6346 |
5 | 0.036 | 11.3472 | 8.0913 | 2.7675 | 2.9656 | 1.5279 | 1.6317 |
6 | 0.042 | 11.3658 | 8.0922 | 2.7598 | 2.9625 | 1.5211 | 1.6289 |
7 | 0.050 | 11.3843 | 8.0930 | 2.7521 | 2.9593 | 1.5143 | 1.6260 |
8 | 0.057 | 11.4028 | 8.0938 | 2.7445 | 2.9561 | 1.5075 | 1.6232 |
9 | 0.064 | 11.4214 | 8.0947 | 2.7368 | 2.9529 | 1.5007 | 1.6203 |
10 | 0.071 | 11.4399 | 8.0955 | 2.7291 | 2.9497 | 1.4939 | 1.6174 |
11 | 0.078 | 11.4584 | 8.0964 | 2.7214 | 2.9465 | 1.4871 | 1.6146 |
12 | 0.085 | 11.4769 | 8.0972 | 2.7138 | 2.9433 | 1.4803 | 1.6117 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-12 | -0.140 | 0.7151 | 0.7078 | 0.1829 | 0.3554 | 0.0914 | 0.2264 |
-11 | -0.128 | 0.7174 | 0.7059 | 0.1808 | 0.3542 | 0.0899 | 0.2255 |
-10 | -0.117 | 0.7197 | 0.7040 | 0.1787 | 0.3529 | 0.0883 | 0.2246 |
-9 | -0.105 | 0.7220 | 0.7020 | 0.1767 | 0.3516 | 0.0868 | 0.2236 |
-8 | -0.093 | 0.7243 | 0.7000 | 0.1746 | 0.3503 | 0.0852 | 0.222 |
-7 | -0.081 | 0.7266 | 0.6979 | 0.1725 | 0.3489 | 0.0837 | 0.2217 |
-6 | -0.070 | 0.7290 | 0.6957 | 0.1704 | 0.3475 | 0.0822 | 0.2207 |
-5 | -0.058 | 0.7313 | 0.6935 | 0.1683 | 0.3461 | 0.0806 | 0.2196 |
-4 | -0.046 | 0.7336 | 0.6913 | 0.1662 | 0.3446 | 0.0791 | 0.2185 |
-3 | -0.035 | 0.7359 | 0.6889 | 0.1641 | 0.3431 | 0.0775 | 0.2174 |
-2 | -0.023 | 0.7382 | 0.6865 | 0.1620 | 0.3415 | 0.0760 | 0.2163 |
-1 | -0.011 | 0.7405 | 0.6840 | 0.1599 | 0.3398 | 0.0744 | 0.2151 |
0 | 0.000 | 0.7420 | 0.6815 | 0.1578 | 0.3382 | 0.0729 | 0.2139 |
1 | 0.011 | 0.7452 | 0.6789 | 0.1557 | 0.3364 | 0.0714 | 0.2126 |
2 | 0.023 | 0.7475 | 0.6762 | 0.1536 | 0.3347 | 0.0698 | 0.2113 |
3 | 0.035 | 0.7498 | 0.6734 | 0.1516 | 0.3328 | 0.0683 | 0.2100 |
4 | 0.046 | 0.7521 | 0.6705 | 0.1495 | 0.3309 | 0.0667 | 0.2086 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-4 | -0.028 | 13.2230 | 8.9452 | 4.7854 | 3.7701 | 3.4611 | 2.3652 |
-3 | -0.021 | 13.2187 | 8.9416 | 4.7555 | 3.7627 | 3.4326 | 2.3584 |
-2 | -0.014 | 13.2144 | 8.9379 | 4.7256 | 3.7553 | 3.4041 | 2.3515 |
-1 | -0.007 | 13.2101 | 8.9343 | 4.6957 | 3.7478 | 3.3756 | 2.3446 |
0 | 0.000 | 13.2058 | 8.9306 | 4.6658 | 3.7404 | 3.3470 | 2.3377 |
1 | 0.007 | 13.2015 | 8.9270 | 4.6359 | 3.7330 | 3.3185 | 2.3308 |
2 | 0.014 | 13.1972 | 8.9233 | 4.6060 | 3.7256 | 3.2900 | 2.3239 |
3 | 0.021 | 13.1929 | 8.9197 | 4.5761 | 3.7181 | 3.2615 | 2.3170 |
4 | 0.028 | 13.1886 | 8.9160 | 4.5462 | 3.7107 | 3.2330 | 2.3101 |
5 | 0.036 | 13.1843 | 8.9123 | 4.5163 | 3.7032 | 3.2044 | 2.3032 |
6 | 0.042 | 13.1800 | 8.9087 | 4.4864 | 3.6958 | 3.1759 | 2.2963 |
7 | 0.050 | 13.1758 | 8.9050 | 4.4565 | 3.6883 | 3.1474 | 2.2894 |
8 | 0.057 | 13.1715 | 8.9013 | 4.4266 | 3.6809 | 3.1189 | 2.2825 |
9 | 0.064 | 13.1672 | 8.8977 | 4.3967 | 3.6734 | 3.0903 | 2.2756 |
10 | 0.071 | 13.1629 | 8.8940 | 4.3668 | 3.6660 | 3.0618 | 2.2687 |
11 | 0.078 | 13.1586 | 8.8903 | 4.3369 | 3.6585 | 3.0333 | 2.2617 |
12 | 0.085 | 13.1543 | 8.8866 | 4.3070 | 3.6510 | 3.0048 | 2.2548 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-12 | -0.140 | 0.8994 | 0.3378 | 0.2273 | 1.5146 | 1.0860 | 0.9017 |
-11 | -0.128 | 0.8965 | 0.3310 | 0.2214 | 1.5072 | 1.0796 | 0.8959 |
-10 | -0.117 | 0.8935 | 0.3242 | 0.2156 | 1.4996 | 1.0729 | 0.8898 |
-9 | -0.105 | 0.8906 | 0.3174 | 0.2097 | 1.4918 | 1.0661 | 0.8836 |
-8 | -0.093 | 0.8877 | 0.3106 | 0.2038 | 1.4838 | 1.0592 | 0.8773 |
-7 | -0.081 | 0.8848 | 0.3038 | 0.1980 | 1.4755 | 1.0520 | 0.8707 |
-6 | -0.070 | 0.8819 | 0.2970 | 0.1921 | 1.4671 | 1.0446 | 0.8640 |
-5 | -0.058 | 0.8789 | 0.2902 | 0.1862 | 1.4584 | 1.0370 | 0.8571 |
-4 | -0.046 | 0.8760 | 0.2834 | 0.1804 | 1.4494 | 1.0292 | 0.8500 |
-3 | -0.035 | 0.8731 | 0.2766 | 0.1745 | 1.4402 | 1.0212 | 0.8427 |
-2 | -0.023 | 0.8702 | 0.2698 | 0.1687 | 1.4307 | 1.0129 | 0.8352 |
-1 | -0.011 | 0.8673 | 0.2630 | 0.1628 | 1.4210 | 1.0044 | 0.8275 |
0 | 0.000 | 0.8643 | 0.2562 | 0.1569 | 1.4109 | 0.9957 | 0.8195 |
1 | 0.011 | 0.8614 | 0.2494 | 0.1511 | 1.4006 | 0.9866 | 0.8113 |
2 | 0.023 | 0.8585 | 0.2426 | 0.1452 | 1.3899 | 0.9773 | 0.8028 |
3 | 0.035 | 0.8556 | 0.2358 | 0.1394 | 1.3789 | 0.9677 | 0.7940 |
4 | 0.046 | 0.8527 | 0.2290 | 0.1335 | 1.3675 | 0.9578 | 0.7850 |
Number | Loss | Year |
1 | 2,000 | 1977 |
2 | 1,380 | 1971 |
3 | 2,000 | 1971 |
4 | 2,000 | 1964 |
5 | 2,580 | 1968 |
6 | 4,730 | 1971 |
7 | 3,700 | 1956 |
8 | 4,250 | 1961 |
9 | 4,500 | 1966 |
10 | 5,000 | 1958 |
Example 1 | Example 2 | |
ν | 2.56 | 0.3636 |
τ | 0.8 | 0.8 |
ξ | 0.972 | 1.197 |
γ | 2 | 2.65≈3 |
z | 6.48165 | 8.54057 |
c | 1.5 | 1.25 |
ω1 | -4.10 | -12.70 |
ω2 | 12.79 | 4.35 |
layer | 5 | 2 |
T | 5 | 21 |
n | 16 | 10 |
Example 1 | Example 2 | |||||||
ω | Correlation | Collective | Bayes | ω | ρ | Collective | Bayes | |
-4 | -0.028 | 9.5854 | 8.5706 | -12 | -0.140 | 0.9920 | 1.8074 | |
-3 | -0.021 | 9.5555 | 8.5631 | -11 | -0.128 | 0.9850 | 1.7991 | |
-2 | -0.014 | 9.5256 | 8.5555 | -10 | -0.117 | 0.9780 | 1.7906 | |
-1 | -0.007 | 9.4957 | 8.5480 | -9 | -0.105 | 0.9710 | 1.7819 | |
0 | 0.000 | 9.4658 | 8.5404 | -8 | -0.093 | 0.9640 | 1.7729 | |
1 | 0.007 | 9.4359 | 8.5329 | -7 | -0.081 | 0.9570 | 1.7637 | |
2 | 0.014 | 9.4060 | 8.5253 | -6 | -0.070 | 0.9500 | 1.7542 | |
3 | 0.021 | 9.3761 | 8.5177 | -5 | -0.058 | 0.9431 | 1.7445 | |
4 | 0.028 | 9.3462 | 8.5102 | -4 | -0.046 | 0.9361 | 1.7345 | |
5 | 0.035 | 9.3163 | 8.5026 | -3 | -0.035 | 0.9291 | 1.7242 | |
6 | 0.042 | 9.2864 | 8.4950 | -2 | -0.023 | 0.9221 | 1.7136 | |
7 | 0.050 | 9.2565 | 8.4874 | -1 | -0.011 | 0.9151 | 1.7027 | |
8 | 0.057 | 9.2266 | 8.4798 | 0 | 0.000 | 0.9081 | 1.6914 | |
9 | 0.064 | 9.1967 | 8.4722 | 1 | 0.011 | 0.9011 | 1.6799 | |
10 | 0.071 | 9.1668 | 8.4647 | 2 | 0.023 | 0.8941 | 1.6679 | |
11 | 0.078 | 9.1369 | 8.4571 | 3 | 0.035 | 0.8871 | 1.6556 | |
12 | 0.085 | 9.1070 | 8.4495 | 4 | 0.046 | 0.8801 | 1.6428 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-4 | -0.028 | 11.1805 | 8.0838 | 2.8365 | 2.9943 | 1.5891 | 1.6574 |
-3 | -0.021 | 11.1990 | 8.0846 | 2.8288 | 2.9911 | 1.5823 | 1.6545 |
-2 | -0.014 | 11.2175 | 8.0855 | 2.8211 | 2.9879 | 1.5755 | 1.6517 |
-1 | -0.007 | 11.2361 | 8.0863 | 2.8135 | 2.9847 | 1.5687 | 1.6489 |
0 | 0.000 | 11.2546 | 8.0871 | 2.8058 | 2.9816 | 1.5619 | 1.6460 |
1 | 0.007 | 11.2731 | 8.0880 | 2.7981 | 2.9784 | 1.5551 | 1.6432 |
2 | 0.014 | 11.2917 | 8.0888 | 2.7905 | 2.9752 | 1.5483 | 1.6403 |
3 | 0.021 | 11.3102 | 8.0896 | 2.7828 | 2.9720 | 1.5415 | 1.6374 |
4 | 0.028 | 11.3287 | 8.0905 | 2.7751 | 2.9688 | 1.5347 | 1.6346 |
5 | 0.036 | 11.3472 | 8.0913 | 2.7675 | 2.9656 | 1.5279 | 1.6317 |
6 | 0.042 | 11.3658 | 8.0922 | 2.7598 | 2.9625 | 1.5211 | 1.6289 |
7 | 0.050 | 11.3843 | 8.0930 | 2.7521 | 2.9593 | 1.5143 | 1.6260 |
8 | 0.057 | 11.4028 | 8.0938 | 2.7445 | 2.9561 | 1.5075 | 1.6232 |
9 | 0.064 | 11.4214 | 8.0947 | 2.7368 | 2.9529 | 1.5007 | 1.6203 |
10 | 0.071 | 11.4399 | 8.0955 | 2.7291 | 2.9497 | 1.4939 | 1.6174 |
11 | 0.078 | 11.4584 | 8.0964 | 2.7214 | 2.9465 | 1.4871 | 1.6146 |
12 | 0.085 | 11.4769 | 8.0972 | 2.7138 | 2.9433 | 1.4803 | 1.6117 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-12 | -0.140 | 0.7151 | 0.7078 | 0.1829 | 0.3554 | 0.0914 | 0.2264 |
-11 | -0.128 | 0.7174 | 0.7059 | 0.1808 | 0.3542 | 0.0899 | 0.2255 |
-10 | -0.117 | 0.7197 | 0.7040 | 0.1787 | 0.3529 | 0.0883 | 0.2246 |
-9 | -0.105 | 0.7220 | 0.7020 | 0.1767 | 0.3516 | 0.0868 | 0.2236 |
-8 | -0.093 | 0.7243 | 0.7000 | 0.1746 | 0.3503 | 0.0852 | 0.222 |
-7 | -0.081 | 0.7266 | 0.6979 | 0.1725 | 0.3489 | 0.0837 | 0.2217 |
-6 | -0.070 | 0.7290 | 0.6957 | 0.1704 | 0.3475 | 0.0822 | 0.2207 |
-5 | -0.058 | 0.7313 | 0.6935 | 0.1683 | 0.3461 | 0.0806 | 0.2196 |
-4 | -0.046 | 0.7336 | 0.6913 | 0.1662 | 0.3446 | 0.0791 | 0.2185 |
-3 | -0.035 | 0.7359 | 0.6889 | 0.1641 | 0.3431 | 0.0775 | 0.2174 |
-2 | -0.023 | 0.7382 | 0.6865 | 0.1620 | 0.3415 | 0.0760 | 0.2163 |
-1 | -0.011 | 0.7405 | 0.6840 | 0.1599 | 0.3398 | 0.0744 | 0.2151 |
0 | 0.000 | 0.7420 | 0.6815 | 0.1578 | 0.3382 | 0.0729 | 0.2139 |
1 | 0.011 | 0.7452 | 0.6789 | 0.1557 | 0.3364 | 0.0714 | 0.2126 |
2 | 0.023 | 0.7475 | 0.6762 | 0.1536 | 0.3347 | 0.0698 | 0.2113 |
3 | 0.035 | 0.7498 | 0.6734 | 0.1516 | 0.3328 | 0.0683 | 0.2100 |
4 | 0.046 | 0.7521 | 0.6705 | 0.1495 | 0.3309 | 0.0667 | 0.2086 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-4 | -0.028 | 13.2230 | 8.9452 | 4.7854 | 3.7701 | 3.4611 | 2.3652 |
-3 | -0.021 | 13.2187 | 8.9416 | 4.7555 | 3.7627 | 3.4326 | 2.3584 |
-2 | -0.014 | 13.2144 | 8.9379 | 4.7256 | 3.7553 | 3.4041 | 2.3515 |
-1 | -0.007 | 13.2101 | 8.9343 | 4.6957 | 3.7478 | 3.3756 | 2.3446 |
0 | 0.000 | 13.2058 | 8.9306 | 4.6658 | 3.7404 | 3.3470 | 2.3377 |
1 | 0.007 | 13.2015 | 8.9270 | 4.6359 | 3.7330 | 3.3185 | 2.3308 |
2 | 0.014 | 13.1972 | 8.9233 | 4.6060 | 3.7256 | 3.2900 | 2.3239 |
3 | 0.021 | 13.1929 | 8.9197 | 4.5761 | 3.7181 | 3.2615 | 2.3170 |
4 | 0.028 | 13.1886 | 8.9160 | 4.5462 | 3.7107 | 3.2330 | 2.3101 |
5 | 0.036 | 13.1843 | 8.9123 | 4.5163 | 3.7032 | 3.2044 | 2.3032 |
6 | 0.042 | 13.1800 | 8.9087 | 4.4864 | 3.6958 | 3.1759 | 2.2963 |
7 | 0.050 | 13.1758 | 8.9050 | 4.4565 | 3.6883 | 3.1474 | 2.2894 |
8 | 0.057 | 13.1715 | 8.9013 | 4.4266 | 3.6809 | 3.1189 | 2.2825 |
9 | 0.064 | 13.1672 | 8.8977 | 4.3967 | 3.6734 | 3.0903 | 2.2756 |
10 | 0.071 | 13.1629 | 8.8940 | 4.3668 | 3.6660 | 3.0618 | 2.2687 |
11 | 0.078 | 13.1586 | 8.8903 | 4.3369 | 3.6585 | 3.0333 | 2.2617 |
12 | 0.085 | 13.1543 | 8.8866 | 4.3070 | 3.6510 | 3.0048 | 2.2548 |
a=0.8 | a=1.5 | a=2.2 | |||||
ω | Correlation | Collective | Bayes | Collective | Bayes | Collective | Bayes |
-12 | -0.140 | 0.8994 | 0.3378 | 0.2273 | 1.5146 | 1.0860 | 0.9017 |
-11 | -0.128 | 0.8965 | 0.3310 | 0.2214 | 1.5072 | 1.0796 | 0.8959 |
-10 | -0.117 | 0.8935 | 0.3242 | 0.2156 | 1.4996 | 1.0729 | 0.8898 |
-9 | -0.105 | 0.8906 | 0.3174 | 0.2097 | 1.4918 | 1.0661 | 0.8836 |
-8 | -0.093 | 0.8877 | 0.3106 | 0.2038 | 1.4838 | 1.0592 | 0.8773 |
-7 | -0.081 | 0.8848 | 0.3038 | 0.1980 | 1.4755 | 1.0520 | 0.8707 |
-6 | -0.070 | 0.8819 | 0.2970 | 0.1921 | 1.4671 | 1.0446 | 0.8640 |
-5 | -0.058 | 0.8789 | 0.2902 | 0.1862 | 1.4584 | 1.0370 | 0.8571 |
-4 | -0.046 | 0.8760 | 0.2834 | 0.1804 | 1.4494 | 1.0292 | 0.8500 |
-3 | -0.035 | 0.8731 | 0.2766 | 0.1745 | 1.4402 | 1.0212 | 0.8427 |
-2 | -0.023 | 0.8702 | 0.2698 | 0.1687 | 1.4307 | 1.0129 | 0.8352 |
-1 | -0.011 | 0.8673 | 0.2630 | 0.1628 | 1.4210 | 1.0044 | 0.8275 |
0 | 0.000 | 0.8643 | 0.2562 | 0.1569 | 1.4109 | 0.9957 | 0.8195 |
1 | 0.011 | 0.8614 | 0.2494 | 0.1511 | 1.4006 | 0.9866 | 0.8113 |
2 | 0.023 | 0.8585 | 0.2426 | 0.1452 | 1.3899 | 0.9773 | 0.8028 |
3 | 0.035 | 0.8556 | 0.2358 | 0.1394 | 1.3789 | 0.9677 | 0.7940 |
4 | 0.046 | 0.8527 | 0.2290 | 0.1335 | 1.3675 | 0.9578 | 0.7850 |