Hostname: page-component-745bb68f8f-b95js Total loading time: 0 Render date: 2025-02-06T05:15:17.193Z Has data issue: false hasContentIssue false

On transform orders for largest claim amounts

Published online by Cambridge University Press:  22 November 2021

Yiying Zhang*
Affiliation:
Southern University of Science and Technology
*
*Postal address: Department of Mathematics, Southern University of Science and Technology, Shenzhen 518055, Guangdong, China. Email: zhangyy3@sustech.edu.cn
Rights & Permissions [Opens in a new window]

Abstract

This paper investigates the ordering properties of largest claim amounts in heterogeneous insurance portfolios in the sense of some transform orders, including the convex transform order and the star order. It is shown that the largest claim amount from a set of independent and heterogeneous exponential claims is more skewed than that from a set of independent and homogeneous exponential claims in the sense of the convex transform order. As a result, a lower bound for the coefficient of variation of the largest claim amount is established without any restrictions on the parameters of the distributions of claim severities. Furthermore, sufficient conditions are presented to compare the skewness of the largest claim amounts from two sets of independent multiple-outlier scaled claims according to the star order. Some comparison results are also developed for the multiple-outlier proportional hazard rates claims. Numerical examples are presented to illustrate these theoretical results.

Type
Original Article
Copyright
© The Author(s), 2021. Published by Cambridge University Press on behalf of Applied Probability Trust

1. Introduction

In the context of insurance, it is known that the annual premium paid by the policyholder on an annual basis to cover the cost of the insurance policy being purchased is the primary cost to the policyholder of transferring the risk to the insurer; see [Reference Balakrishnan, Zhang and Zhao3, Reference Frostig16, Reference Khaledi and Ahmadi20, Reference Zhang, Cai and Zhao41] and the related references therein. The smallest and largest claim amounts as well as their difference play a key role in determining the annual premium since they provide very useful information for insurance analysis. In the actuarial science context, the largest claim amount is involved within the largest claims reinsurance treaty and the ECOMOR reinsurance treaty, which are well known in charging premiums for insurers by studying these two important premium principles from insurance portfolios of policyholders; see, for example, [Reference Benktander9, Reference Kremer26Reference Ladoucette and Teugels28]. Also, the largest claim amount is sometimes termed the ‘probable maximum loss’ in the insurance industry, which is the maximum loss that an insurer would be expected to incur on a policy. This terminology is most often associated with insurance policies on property, such as fire or flood insurance; it represents the worst-case scenario for an insurer and can help to determine the premiums.

Stochastic orders, which have been used in various areas including management science, operations research, financial economics, reliability theory, and actuarial science, are important in studying the ordering properties of extreme and aggregate claim amounts from different viewpoints. For comprehensive discussions on different kinds of stochastic orders, interested readers may refer to the monographs [Reference Müller and Stoyan33, Reference Shaked and Shanthikumar35]. Let $X_{1},\ldots,X_{n}$ be a set of nonnegative random variables, with $X_{i}$ denoting the ith claim severity in an insurance portfolio for $i=1,\ldots,n$. Let $I_{1},\ldots,I_{n}$ be a group of Bernoulli random variables, independent of the $X_{i}$, with $I_{i}$ indicating whether claim $X_{i}$ occurs or not and such that $\mathbb{E}[I_{i}]=p_{i}$ for $i=1,\ldots,n$. A surge of research has sprung up on the study of the effects of dependence and heterogeneity among claim sizes and occurrence levels on ordering properties of the aggregate claim numbers $\sum_{i=1}^{n}I_{i}$ and the aggregate claim amount $\sum_{i=1}^{n}I_{i}X_{i}$ from different perspectives; see, for example, [Reference Barmalzan, Payandeh Najafabadi and Balakrishnan6, Reference Denuit and Frostig11Reference Dhaene and Goovaerts13, Reference Frostig16Reference Khaledi and Ahmadi20, Reference Ma29, Reference Wang37Reference Zhang and Zhao39, Reference Zhang, Zhao and Cheung42].

Let $Y_{1\;:\;n}$ ($Y_{1\;:\;n}^{\ast}$) and $Y_{n\;:\;n}$ ($Y_{n\;:\;n}^{\ast}$) be the smallest and largest claim amounts from the insurance portfolios $I_{1}X_{1},\ldots,I_{n}X_{n}$ ($I_{1}^{\ast}X_{1}^{\ast},\ldots,I_{n}^{\ast}X_{n}^{\ast}$). The ordering properties of $Y_{1\;:\;n}$ and $Y_{n\;:\;n}$ were discussed in [Reference Barmalzan, Payandeh Najafabadi and Balakrishnan8] in the sense of the usual stochastic and hazard rate orders by using the chain majorization orders imposed on the vectors of occurrence probabilities and distribution parameters. The likelihood ratio ordering between $Y_{1\;:\;n}$ and $Y_{1\;:\;n}^{\ast}$ were established in [Reference Barmalzan, Payandeh Najafabadi and Balakrishnan7] when the $X_{i}$ and $X_{i}^{\ast}$ are independent and heterogeneous Weibull distributed claims. Along these lines, [Reference Balakrishnan, Zhang and Zhao3] provided sufficient conditions to compare $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ according to the usual stochastic, the hazard rate, and the reversed hazard rate orders when the claims $X_{i}$ and $X_{i}^{\ast}$ have general distributions. They also presented sufficient conditions to stochastically compare claim ranges from two sets of heterogeneous (and dependent) claims by means of the usual stochastic order. Very recently, [Reference Zhang, Cai and Zhao41] showed that the weak supermajorization order between the transformed vectors of occurrence probabilities implies the usual stochastic ordering between the largest claim amounts when the claim severities are weakly stochastically arrangement increasing. For the case of independent multiple-outlier claims, they studied the effects of heterogeneity among sample sizes on the stochastic properties of the largest and smallest claim amounts in the sense of the hazard rate and likelihood ratio orders. Also, [Reference Zhang, Amini-Seresht and Zhao40] established sufficient (and necessary) conditions between $Y_{2\;:\;n}$ and $Y_{2\;:\;n}^{\ast}$ by means of the usual stochastic, the hazard rate, and the likelihood ratio orders.

On the other hand, simple and workable ways to facilitate comparisons of the skewness of the distributions of extreme claim amounts are in urgent need for both researchers and practitioners to extract important information, rather than woodenly adopting complex calculations and mathematical techniques. In this regard, stochastic comparisons on extreme claim amounts from different sets of insurance portfolios are strongly desired in many scenarios, which will be very helpful to insurers in evaluating the entire group of risks and charging premiums from policyholders. Stochastic comparisons between $Y_{1\;:\;n}$ and $Y_{1\;:\;n}^{\ast}$ were discussed in [Reference Barmalzan and Payandeh Najafabadi5] in the sense of the convex transform and right-spread orders under certain conditions when the claims are Weibull distributed. As a result, a lower/upper bound was established for the coefficient of variation of $Y_{1\;:\;n}$ when the claim risks are independent and heterogeneous Weibull distributed. The convex transform order and the right-spread order were investigated in [Reference Zhang, Cai and Zhao41] for the smallest claim amounts when the occurrence levels are dependent and the number of claims are different. However, establishing sufficient conditions for the convex transform and star orders of the largest claim amounts remains an open question.

The ordering results on maximum order statistics from different sets of continuous random variables (with zero probability taking values at the point 0) have been extensively studied in the sense of the transform orders; see, for example, [Reference Amini-Seresht, Qiao, Zhang and Zhao2, Reference Kochar and Xu23, Reference Kochar and Xu25, Reference Zhang, Ding and Zhao43]. Note that there exists a possibility that all the policyholders might have no claims during a specified time period, which means that the distribution of the largest claim amount has a (strictly) positive mass at 0 (cf. the proof of Theorem 1). Therefore, the traditional stochastic comparison results cannot be applied for ordering the largest claim amounts. To fill this gap, in this paper we investigate the ordering properties of the largest claim amounts in accordance with the convex transformation order and the star order. On one hand, if the $X_{i}$ ($X_{i}^{\ast}$) are independently exponentially distributed with hazard rates $\lambda_{i}$ ($\lambda$), and $\mathbb{E}[I_{i}]=p_{i}$ ($\mathbb{E}[I_{i}^{\ast}]=p$), $i=1,\ldots,n$, it is shown that, without any restrictions on the claim parameters,

\begin{equation*} p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}\Longrightarrow Y_{n\;:\;n}\geq_{\rm c}Y_{n\;:\;n}^{\ast},\end{equation*}

where $\geq_{\rm c}$ denotes the convex transformation order (see Section 2 for its detailed definition). As a result, a lower bound for the coefficient of variation of $Y_{n\;:\;n}$ is established. Furthermore, equivalent characterizations are presented for the dispersive and right-spread orders of the largest claim amounts arising from two sets of heterogeneous and homogeneous insurance claims. On the other hand, some insureds may have higher occurrence probabilities/claim sizes than the others (in the sense of some stochastic orders between their claim severity distributions) in some insurance portfolios. Therefore, it is natural to describe this phenomenon in terms of the multiple-outlier claims model. The second aim of the paper is to study the star ordering of the largest claim amounts from independent multiple-outlier scaled or proportional hazard rates (PHR) claims. Sufficient conditions will be established for comparing the skewness of the distributions of the largest claim amounts in terms of the star order. It is important to notice that our results can also be used in the area of reliability theory to stochastically compare the variability of the lifetimes of parallel systems subject to random shocks with different shocking probabilities during a fixed time period. Interested readers may refer to [Reference Zhang, Amini-Seresht and Zhao40] for more discussions.

The information we want to deliver from the present study is as follows: (i) The heterogeneity among the occurrence probabilities of claims and claim severities affects the skewness of the distribution of the largest claim amount in an insurance portfolio, which are quantitatively depicted by some transform orders. (ii) The lower bounds for the survival function, variance, and coefficient of variation of the largest claim amount from a set of insurance portfolios with heterogeneous claims can be established under certain conditions imposed on the occurrence probabilities and claim severity distribution parameters, which provide useful information in charging premiums for an insurance portfolio.

The remainder of the paper is organized as follows: Section 2 recalls some pertinent definitions and notions used in the subsequent sections. Section 3 presents ordering results between the largest claim amounts according to the convex transform order, the dispersive order, and the right-spread order when one set of insurance portfolios is heterogeneous while the set of insurance claims is homogeneous. Section 4 establishes sufficient conditions to compare the skewness of largest claim amounts from two sets of independent multiple-outlier scaled or PHR claims in the sense of the star ordering. Section 5 concludes the paper with some discussions.

2. Preliminaries

Throughout this paper, the term ‘increasing’ is used for ‘monotone non-decreasing’ and ‘decreasing’ is used for ‘monotone non-increasing’. Expectations are assumed to be finite whenever they appear, and inverse functions are assumed to be right continuous. Let $\mathbb{R}=(-\infty,+\infty)$, $\mathbb{R}_{+}=[0,\infty)$, and $\mathbb{R}_{++}=(0,\infty)$. Denote $\mathcal{I}^n_{+}=\{{{x}}\in\mathbb{R}_{+}^n\;:\;0\leq x_1\leq x_2\leq\cdots\leq x_n\}$ and $\mathcal{D}^n_{+}=\{\textbf{{x}}\in\mathbb{R}_{+}^n\;:\;x_1\geq x_2\geq\cdots\geq x_n\geq 0\}$. We shall use $\stackrel{\rm sgn}{=}$ to express that both sides of an equality have the same sign.

Definition 1. Let X and Y be two nonnegative random variables with distribution functions F and G, and survival functions $\bar{F}$ and $\bar{G}$, respectively. Let $F^{-1}$ and $G^{-1}$ be the inverses of the distribution functions F and G, respectively. Then, X is said to be larger than Y in the

  1. (i) usual stochastic order (denoted by $X\geq_{\rm st} Y$) if $\bar{F}(x)\geq \bar{G}(x)$ for all $x\in\mathbb{R}_{+}$;

  2. (ii) convex transform order (denoted by $X\geq_{\rm c}Y$) if $F^{-1}[G(x)]$ is convex in $x\in\mathbb{R}_{+}$, or equivalently, $G^{-1}[F(x)]$ is concave in $x\in\mathbb{R}_{+}$;

  3. (iii) star order (denoted by $X\geq_{\star}Y$) if $F^{-1}[G(x)]$ is star-shaped in the sense that $F^{-1}[G(x)]/x$ is increasing in $x\in\mathbb{R}_{+}$;

  4. (iv) superadditive order (denoted by $X\geq_{\rm su}Y$) if $F^{-1}[G(x)]$ is superadditive in $x\in\mathbb{R}_{+}$, that is, $F^{-1}[G(x+y)]\geq F^{-1}[G(x)]+F^{-1}[G(y)]$ for all $x,y\in\mathbb{R}_{+}$;

  5. (v) new better than used in expectation (NBUE) order (denoted by $X\geq_{\rm NBUE}Y$) if

    \begin{equation*} \frac{1}{\mathbb{E}[X]}\int_{F^{-1}(u)}^{\infty}\bar{F}(x) \, \mathrm{d} x\geq \frac{1}{\mathbb{E}[Y]}\int_{G^{-1}(u)}^{\infty}\bar{G}(x) \, \mathrm{d} x \qquad\mbox{for all $u\in[0,1]$}; \end{equation*}
  6. (vi) Lorenz order (denoted by $X\geq_{\rm Lorenz}Y$) if $L_X(p)\leq L_Y(p)$ for all $p\in[0,1]$, where the Lorenz curve $L_X$ corresponding to X is defined as $L_X(p)=\int_0^{p}F^{-1}(u) \, \mathrm{d} u/\mathbb{E}[X]$;

  7. (vii) dispersive order (denoted by $X\geq_{\rm disp}Y$) if $F^{-1}(v)-F^{-1}(u)\geq G^{-1}(v)-G^{-1}(u)$, for $0< u\leq v<1$;

  8. (viii) right-spread order (denoted by $X\geq_{\rm RS}Y$) if

    \begin{equation*} \int_{F^{-1}(u)}^{\infty}\bar{F}(t) \, \mathrm{d} t\geq \int_{G^{-1}(u)}^{\infty}\bar{G}(t) \, \mathrm{d} t \qquad\mbox{for all $u\in(0,1)$.} \end{equation*}

The class of transform orders contains the convex transform order, the star order, and the superadditive order, which are commonly used in actuarial science to compare the severities of different risks/losses, and in reliability theory to compare the lifetimes of different coherent systems. These three types of stochastic orders are scale invariant, and usually adopted to compare the skewness of the distributions of interested random variables. The skewness of the distribution of the extreme claim amount plays a key role in many practical scenarios. For instance, in the context of actuarial science, the claims are usually right and positively skewed since they are bounded on the left but unbounded on the right, and the tail on the right side of the density function is longer/fatter than the left side. In order to compare the skewness of the distributions of the largest claim amounts, it is natural to establish sufficient conditions for some transform orders between them to analyze the effects of the heterogeneity among occurrence probabilities and claim severity parameters on the skewness of their distributions.

The convex transform order was proposed in [Reference Van Zwet36] to compare the skewness of probability distributions. It is also called the increasing failure rate order in reliability theory since, when f and g exist, the convexity of $G^{-1}[F(x)]$ means that

\begin{equation*}\frac{f(F^{-1}(u))}{g(G^{-1}(u))}=\frac{\tilde{r}_F(F^{-1}(u))}{\tilde{r}_G(G^{-1}(u))}\end{equation*}

is increasing in $u\in[0, 1]$, where $\tilde{r}_F$ and $\tilde{r}_G$ stand for the reversed hazard rate functions of F and G, respectively. Thus, $X\leq_{\rm c}Y$ can be understood as X aging faster than Y in some stochastic sense. The star order is called increasing failure rate in average in reliability theory. The Lorenz order, which is called harmonic new better than used in expectation in reliability theory, can be used in economics to measure the inequality of incomes. It is known that $X\geq_{\star} Y$ implies $X\geq_{\rm su} Y$, and

\begin{equation*} X\geq_{\rm c} Y\Longrightarrow X\geq_{\star} Y\Longrightarrow X\geq_{\rm NBUE}Y\Longrightarrow X\geq_{\rm Lorenz}Y \Longrightarrow {\rm CV}(X)\geq {\rm CV}(Y),\end{equation*}

where ${\rm CV}(X)=\sqrt{{\rm Var}(X)}/\mathbb{E}(X)$ and ${\rm CV}(Y)=\sqrt{{\rm Var}(Y)}/\mathbb{E}(Y)$ denote the coefficients of variation of X and Y, respectively. The dispersive order, which is stronger than the right-spread order, is a kind of partial order for comparing the variabilities of two probability distributions. It is known that both the usual stochastic order and the right-spread order (also called excess wealth order in economics) imply increasing convex order. In the following we shall sometimes use $F\geq_{\rm sym} G$ to denote that the two independent random variables X and Y with respective distributions F and G are such that $X\geq_{\rm sym} Y$, where ‘$\geq_{\rm sym}$ indicates some specified stochastic order. For comprehensive discussions on these useful orders, we refer interested readers to the monographs [Reference Denuit, Dhaene, Goovaerts and Kaas10, Reference Marshall and Olkin30, Reference Shaked and Shanthikumar35].

One useful tool in deriving various inequalities in statistics and probability is the notion of majorization, which characterizes the heterogeneity and dispersiveness among the coordinates of real-valued vectors.

Definition 2. Let $x_{1\;:\;n}\leq x_{2\;:\;n}\leq\cdots\leq x_{n\;:\;n}$ be the increasing arrangement of the components of the vector ${\textbf{{x}}}=(x_1,\ldots,x_n)$.

  1. (i) A vector ${\textbf{{x}}}\in\mathbb{R}^n$ is said to majorize another vector ${\textbf{{y}}}\in\mathbb{R}^n$ (written as ${\textbf{{x}}}\stackrel{\rm m}{\succeq} {\textbf{{y}}}$) if $\sum_{i=1}^jx_{i\;:\;n}\leq \sum_{i=1}^j y_{i\;:\;n}$ for $j=1,\ldots,n-1$, and $\sum_{i=1}^nx_{i\;:\;n}= \sum_{i=1}^n y_{i\;:\;n}$;

  2. (ii) A vector ${\textbf{{x}}}\in\mathbb{R}^n$ is said to weakly supermajorize another vector ${\textbf{{y}}}\in\mathbb{R}^n$ (written as ${\textbf{{x}}}\stackrel{\rm w}{\succeq} {\textbf{{y}}}$) if $\sum_{i=1}^jx_{i\;:\;n}\leq \sum_{i=1}^j y_{i\;:\;n}$ for $j=1,\ldots,n$;

  3. (iii) A vector ${\textbf{{x}}}\in\mathbb{R}^n$ is said to weakly submajorize another vector ${\textbf{{y}}}\in\mathbb{R}^n$ (written as ${\textbf{{x}}}\succeq_{\rm w} {\textbf{{y}}}$) if $\sum_{j=i}^nx_{j\;:\;n}\geq \sum_{j=i}^n y_{j\;:\;n}$ for $i=1,\ldots,n$;

  4. (iv) A vector ${\textbf{{x}}}\in\mathbb{R}^{n}_+$ is said to be p-larger than another vector ${\textbf{{y}}}\in\mathbb{R}^{n}_+$ (written as ${\textbf{{x}}}\stackrel{\rm p}{\succeq} {\textbf{{y}}}$) if $\prod_{i=1}^j x_{i\;:\;n}\leq \prod_{i=1}^j y_{i\;:\;n}$ for $j=1,\ldots,n$.

It is worth mentioning that the weak supermajorization order implies the p-larger order, and the majorization order implies both the weak supermajorization and submajorization orders. However, the reverse statement is not true in general. Interested readers are referred to [Reference Marshall, Olkin and Arnold31] for more detailed discussions on their properties and applications.

Next, we introduce the notion of Schur function and an important lemma that is useful to establish inequalities with respect to the majorization order.

Definition 3. A real-valued function $\varphi$ defined on a set $I\subseteq\mathbb{R}^{n}$ is said to be Schur-convex (Schur-concave) on I if ${{\textbf{{x}}}}\stackrel{\rm m}{\succeq}{{\textbf{{y}}}}$ implies $\varphi({{\textbf{{x}}}})\geq (\leq) \, \varphi({{\textbf{{y}}}})$ on I.

Lemma 1. ([Reference Marshall, Olkin and Arnold31]) Consider the real-valued continuously differentiable function $\phi$ on $J^{n}$, where $J\subseteq\mathbb{R}$ is an open interval. Then, $\phi$ is Schur-convex (Schur-concave) on $J^{n}$ if and only if $\phi$ is symmetric on $J^{n}$, and for all $i\neq j$ and all ${{\textbf{{u}}}}\in J^{n}$,

\begin{equation*}(u_i-u_j)\left(\frac{\displaystyle{\partial\phi({{\textbf{{u}}}})}}{\displaystyle{\partial u_{i}}}-\frac{\displaystyle{\partial\phi({{\textbf{{u}}}})}}{\displaystyle{\partial u_{j}}}\right)\geq0 \ (\leq0),\end{equation*}

where $\frac{\displaystyle{\partial\phi({{\textbf{{u}}}})}}{\displaystyle{\partial u_{i}}}$ denotes the partial derivative of $\phi$ with respect to its ith argument, $i=1,\ldots,n$.

3. Convex transform order

Assume that the nonnegative random variable $X_{i}$ has survival function $\bar{G}(t,\lambda_{i})$, where $\lambda_i\in\mathbb{R}_{++}$ is an index parameter, for $i=1,2,\ldots,n$. Let $I_{i}$ be a Bernoulli random variable such that $\mathbb{E}[I_{i}]=p_{i}$ for $i=1,2,\ldots,n$. Note that the random variable $Y_i\;:\!=\;I_{p_{i}}X_{i}$ is discrete-continuous, which admits zero with probability $1-p_i$, and $X_{i}$ with probability $p_i$, for $i=1,2,\ldots,n$. Then, the distribution function of $Y_{i}$ is given by $F_{Y_{i}}(t)=1-p_i\bar{G}(t,\lambda_{i})$ for $t\in\mathbb{R}_{+}$. Let $Y_{i}^{\ast}\;:\!=\;I_{i}^{\ast}X_{i}^{\ast}$ be the claim amounts arising from another insurance portfolio for $i=1,2,\ldots,n$. Denote by $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ the largest claim amounts from $Y_{1},\ldots,Y_{n}$ and $Y_{1}^{\ast},\ldots,Y_{n}^{\ast}$, respectively.

The well-known Cauchy–Schwarz inequality introduced in the following lemma plays a critical role in obtaining the main results.

Lemma 2. ([Reference Mitrinović32]) Let $(a_{1},a_{2},\ldots,a_{n})$ and $(b_{1},b_{2},\ldots,b_{n})$ be two sequences of real numbers. Then,

\begin{equation*}\left(\sum_{i=1}^{n}a_{i}^{2}\right)\left(\sum_{i=1}^{n}b_{i}^{2}\right)\geq \left(\sum_{i=1}^{n}a_{i}b_{i}\right)^{2},\end{equation*}

where the equality holds if and only if the sequences $(a_{1},a_{2},\ldots,a_{n})$ and $(b_{1},b_{2},\ldots,b_{n})$ are proportional, i.e. there is a constant $\lambda$ such that $a_{k}=\lambda b_{k}$ for each $k\in\{1,2,\ldots,n\}$.

Now, we present a sufficient condition for the convex transform order between the largest claim amounts when one set of claims is heterogeneous and another set of claims is homogeneous.

Theorem 1. Let $X_{i}$ ($X_{i}^{\ast}$) be independent exponential random variables with hazard rate $\lambda_{i}$ ($\lambda_{i}^{\ast}=\lambda$), $i=1,2,\ldots,n$. Let $I_{1},\ldots,I_{n}$ ($I_{1}^{\ast},\ldots,I_{n}^{\ast}$) be a set of independent Bernoulli random variables such that $\mathbb{E}[I_{i}]=p_{i}$ ($\mathbb{E}[I_{i}^{\ast}]=p$), $i=1,2,\ldots,n$. Then,

\begin{equation*} p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}\Longrightarrow Y_{n\;:\;n}\geq_{\rm c}Y_{n\;:\;n}^{\ast}.\end{equation*}

Proof. Let $F_{n}$ ($f_{n}$) and $G_{n}$ ($g_{n}$) be the distribution (density) functions of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$, respectively. Then, for $x\in\mathbb{R}_{+}$, we have $F_{n}(x)=\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})$ and $G_{n}(x)=(1-p\mathrm{e}^{-\lambda x})^{n}$. In light of [Reference Marshall and Olkin30, Proposition 21.A.7], it is sufficient to show that $G_{n}^{-1}F_{n}(x)$ is concave on $x\in\mathbb{R}_{+}$. Since $F_{n}(x)\geq F_{n}(0)=\prod_{i=1}^{n}(1-p_{i})$ and $G_{n}(x)\geq G_{n}(0)=(1-p)^{n}$, we know that the inverse of $G_{n}$ exists by the assumption that $p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}$. Note that, for $x\in\mathbb{R}_{+}$,

\begin{equation*}G_{n}^{-1}F_{n}(x)=-\frac{1}{\lambda}\ln\left(\frac{1-\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{1/n}}{p}\right).\end{equation*}

Then, we have

\begin{equation*}g_{n}(G_{n}^{-1}F_{n}(x))=n\lambda\left[1-\prod_{i=1}^{n}\big(1-p_{i}\mathrm{e}^{-\lambda_{i}x}\big)^{1/n}\right]\prod_{i=1}^{n}\big(1-p_{i}\mathrm{e}^{-\lambda_{i}x}\big)^{(n-1)/n}.\end{equation*}

Upon differentiating $G_{n}^{-1}F_{n}(x)$ with respect to x, we obtain

\begin{eqnarray*}\frac{\mathrm{d} G_{n}^{-1}F_{n}(x)}{\mathrm{d} x}&=& \frac{f_{n}(x)}{g_{n}(G_{n}^{-1}F_{n}(x))}\\[5pt] &=& \frac{\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}\mathrm{e}^{-\lambda_{i}x}}{1-p_{i}\mathrm{e}^{-\lambda_{i}x}}\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})}{n\lambda\left[1-\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{1/n}\right]\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{(n-1)/n}}\\[5pt] &=& \frac{\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}\mathrm{e}^{-\lambda_{i}x}}{1-p_{i}\mathrm{e}^{-\lambda_{i}x}}\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{1/n}}{n\lambda\left[1-\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{1/n}\right]}.\end{eqnarray*}

Thus, it is enough to show that

\begin{equation*} h(x)=\frac{\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}}{\mathrm{e}^{\lambda_{i}x}-p_{i}}}{\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{-1/n}-1}\end{equation*}

is decreasing in $x\in\mathbb{R}_{+}$. By taking the derivative of h(x), this is equivalent to proving that

\begin{multline*} {\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}^2\mathrm{e}^{\lambda_{i}x}}{(\mathrm{e}^{\lambda_{i}x}-p_{i})^2}\left[\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{-1/n}-1\right]} \\ \geq \frac{1}{n}\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}}{\mathrm{e}^{\lambda_{i}x}-p_{i}}\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}\mathrm{e}^{-\lambda_{i}x}}{1-p_{i}\mathrm{e}^{-\lambda_{i}x}}\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{-1/n},\end{multline*}

i.e.

\begin{equation*} \sum_{i=1}^{n}\frac{p_{i}\lambda_{i}^2\mathrm{e}^{-\lambda_{i}x}}{(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^2}\left[1-\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{1/n}\right]\geq\frac{1}{n}\left(\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}\mathrm{e}^{-\lambda_{i}x}}{1-p_{i}\mathrm{e}^{-\lambda_{i}x}}\right)^2.\end{equation*}

By applying the Cauchy–Schwarz inequality in Lemma 3, it can be seen that

\begin{equation*} \sum_{i=1}^{n}\frac{p_{i}\lambda_{i}^2\mathrm{e}^{-\lambda_{i}x}}{(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^2}\sum_{i=1}^{n}p_{i}\mathrm{e}^{-\lambda_{i}x}\geq\left(\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}\mathrm{e}^{-\lambda_{i}x}}{1-p_{i}\mathrm{e}^{-\lambda_{i}x}}\right)^2.\end{equation*}

Thus, we only need to show that

\begin{equation*} \sum_{i=1}^{n}\frac{p_{i}\lambda_{i}^2\mathrm{e}^{-\lambda_{i}x}}{(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^2}\left[1-\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{1/n}\right]\geq\frac{1}{n}\sum_{i=1}^{n}\frac{p_{i}\lambda_{i}^2\mathrm{e}^{-\lambda_{i}x}}{(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^2}\sum_{i=1}^{n}p_{i}\mathrm{e}^{-\lambda_{i}x} ,\end{equation*}

i.e.

\begin{equation*} \frac{1}{n}\sum_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})\geq\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})^{1/n},\end{equation*}

which holds naturally by applying the arithmetic–geometric mean inequality.

Remark 1. In light of the proof of Theorem 3, the condition $p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}$ imposed on the occurrence probabilities is a technical condition to ensure that the function $G_{n}^{-1}[F_{n}(x)]$ is well defined. Under the setting when $p=p_{1}=\cdots=p_{n}=1$, Theorem 1 reduces to the result of [Reference Kochar and Xu23, Theorem 3.1]. It is worth mentioning that [Reference Zhang, Ding and Zhao43] generalized the result of [Reference Kochar and Xu23, Theorem 3.1] to the scale model. Moreover, since the exponential distribution is a special case of the PHR model, it is thus of natural interest to generalize the result of Theorem 1 when the claims have scale or PHR distributions. We leave these as open problems.

The following result can be readily obtained from Theorem 1 by applying the fact that the convex transform order implies the star order and the Lorenz order.

Corollary 1. Under the setup of Theorem 1, $p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}\Longrightarrow Y_{n\;:\;n}\geq_{\rm \star \, (Lorenz)}Y_{n\;:\;n}^{\ast}$.

The next result provides a lower bound on the coefficient of variation of the largest claim amount from a set of heterogeneous and independent insurance portfolios.

Corollary 2. Under the setup of Theorem 1, if $p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}$,

\begin{equation*} {\rm CV}(Y_{n\;:\;n})\geq \sqrt{\frac{\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\Big(\sum_{j=1}^{k}\frac{1}{j^2}+\big(\sum_{j=1}^{k}\frac{1}{j}\big)^2\Big)}{\Big(\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\sum_{j=1}^{k}\frac{1}{j}\Big)^2}-1}.\end{equation*}

Proof. Let $X_{k\;:\;k}^{\ast}$ be the largest order statistics from indpendent and identically distributed random variables $X_{i_{1}}^{\ast},\ldots,X_{i_{k}}^{\ast}$ with common hazard rate $\lambda$, for any $1\leq i_{1}<\cdots<i_{k}\leq n$. From [Reference Barlow and Proschan4, p. 60], it follows that

\begin{equation*} \mathbb{E}[X_{k\;:\;k}^{\ast}]=\frac{1}{\lambda}\sum_{j=1}^{k}\frac{1}{j} , \qquad \mbox{Var}[X_{k\;:\;k}^{\ast}]=\frac{1}{\lambda^2}\sum_{j=1}^{k}\frac{1}{j^2}.\end{equation*}

Note that the probability that the largest claim amount $Y_{n\;:\;n}^{\ast}$ takes value 0 is equal to $(1-p)^n$. Then, we have

\begin{equation*} \mathbb{E}[Y_{n\;:\;n}^{\ast}]=\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\mathbb{E}[X_{k\;:\;k}^{\ast}]=\frac{1}{\lambda}\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\sum_{j=1}^{k}\frac{1}{j}\end{equation*}

and

\begin{eqnarray*}\mathbb{E}[(Y_{n\;:\;n}^{\ast})^2]&=&\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\mathbb{E}[(X_{k\;:\;k}^{\ast})^2]\\&=&\frac{1}{\lambda^2}\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\left(\sum_{j=1}^{k}\frac{1}{j^2}+\left(\sum_{j=1}^{k}\frac{1}{j}\right)^2\right).\end{eqnarray*}

Then, by substituting $\mathbb{E}[Y_{n\;:\;n}^{\ast}]$ and $\mathbb{E}[(Y_{n\;:\;n}^{\ast})^2]$ into

\begin{equation*}{\rm CV}(Y_{n\;:\;n}^{\ast})=\frac{\sqrt{{\rm Var}[Y_{n\;:\;n}^{\ast}]}}{\mathbb{E}[Y_{n\;:\;n}^{\ast}]}=\sqrt{\frac{\mathbb{E}[(Y_{n\;:\;n}^{\ast})^2]}{\mathbb{E}^2[Y_{n\;:\;n}^{\ast}]}-1},\end{equation*}

the desired result can be obtained.

The next result gives equivalent characterizations of the dispersive order and the right-spread order between the largest claim amounts from two sets of independent heterogeneous and homogeneous insurance portfolios.

Theorem 2. Under the setup of Theorem 1, if $p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}$, then

  1. (i) $Y_{n\;:\;n}\geq_{\rm disp}Y_{n\;:\;n}^{\ast}$ if and only if $Y_{n\;:\;n}\geq_{\rm st}Y_{n\;:\;n}^{\ast}$;

  2. (ii) $Y_{n\;:\;n}\geq_{\rm RS}Y_{n\;:\;n}^{\ast}$ if and only if $\mathbb{E}[Y_{n\;:\;n}]\geq\mathbb{E}[Y_{n\;:\;n}^{\ast}]$.

Proof.

  1. (i) By using [Reference Ahmed, Alzaid, Bartoszewicz and Kochar1, Theorem 3], for two random variables X and Y, $X\geq_{\rm su}Y$ and $X\geq_{\rm st}Y$ imply that $X\geq_{\rm disp}Y$. According to Theorem 1, we have $Y_{n\;:\;n}\geq_{\rm c}Y_{n\;:\;n}^{\ast}$ and thus $Y_{n\;:\;n}\geq_{\rm su}Y_{n\;:\;n}^{\ast}$. Hence, the condition $Y_{n\;:\;n}\geq_{\rm st}Y_{n\;:\;n}^{\ast}$ implies that $Y_{n\;:\;n}\geq_{\rm disp}Y_{n\;:\;n}^{\ast}$. Conversely, observing that $Y_{n\;:\;n}\geq_{\rm disp}Y_{n\;:\;n}^{\ast}$ implies that $Y_{n\;:\;n}\geq_{\rm st}Y_{n\;:\;n}^{\ast}$, since their distributions have a common left-hand point of the support. Then, the required result follows immediately.

  2. (ii) It was shown in [Reference Fernández-Ponce, Kochar and Muñoz-Perez15, Theorem 4.3] that $X\geq_{\rm NBUE}Y$ and $\mathbb{E}[X]\geq\mathbb{E}[Y]$ imply $X\geq_{\rm RS}Y$. Note that $Y_{n\;:\;n}\geq_{\rm c}Y_{n\;:\;n}^{\ast}$ implies that $Y_{n\;:\;n}\geq_{\rm NBUE}Y_{n\;:\;n}^{\ast}$, and $Y_{n\;:\;n}\geq_{\rm RS}Y_{n\;:\;n}^{\ast}$ implies that $\mathbb{E}[Y_{n\;:\;n}]\geq\mathbb{E}[Y_{n\;:\;n}^{\ast}]$. Then, the required result can be obtained by applying Theorem 1.

The next result presents sufficient conditions for the usual stochastic order and the dispersive order between the largest claim amounts arising from two sets of independent heterogeneous and homogeneous claims.

Theorem 3. Let $\bar{p}=\frac{1}{n}\sum_{i=1}^{n}p_i$ and $\bar{\lambda}=\frac{1}{n}\sum_{i=1}^{n}\lambda_i$. Under the setup of Theorem 1, if ${\textbf{{p}}}\in\mathcal{I}^n_{+}$, $\boldsymbol{\lambda}\in\mathcal{D}^n_{+}$, $\bar{p}\geq p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}$, and $\lambda\geq\bar{\lambda}$, we have $Y_{n\;:\;n}\geq_{\rm st \, (disp)}Y_{n\;:\;n}^{\ast}$.

Proof. It suffices to show that $Y_{n\;:\;n}\geq_{\rm st}Y_{n\;:\;n}^{\ast}$ since the usual stochastic order implies the dispersive order according to the proof of Theorem 2(i). Note that the distribution function of $Y_{n\;:\;n}$ is given by $F_{n}(x)=\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})$. Let $\psi(\boldsymbol{\lambda})=-\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})$. Since $\bar{\lambda}=\frac{1}{n}\sum_{i=1}^{n}\lambda_i$, it holds that $(\lambda_1,\ldots,\lambda_n)\stackrel{\rm m}{\succeq}(\bar{\lambda},\ldots,\bar{\lambda})$. Observe that, for any $1\leq i<j\leq n$,

\begin{eqnarray*} \frac{\partial\psi({\boldsymbol\lambda})}{\partial\lambda_i}-\frac{\partial\psi({\boldsymbol\lambda})}{\partial\lambda_j}&=&\frac{p_ix\mathrm{e}^{-\lambda_ix}}{1-p_{i}\mathrm{e}^{-\lambda_{i}x}}\psi({\boldsymbol\lambda})-\frac{p_jx\mathrm{e}^{-\lambda_jx}}{1-p_{j}\mathrm{e}^{-\lambda_{j}x}}\psi({\boldsymbol\lambda})\\&\stackrel{\rm sgn}{=}&\frac{p_j\mathrm{e}^{-\lambda_jx}}{1-p_{j}\mathrm{e}^{-\lambda_{j}x}}-\frac{p_i\mathrm{e}^{-\lambda_ix}}{1-p_{i}\mathrm{e}^{-\lambda_{i}x}}\leq0,\end{eqnarray*}

where the inequality is due to $p_i\geq p_j$ and $\lambda_i\leq\lambda_j$. Then, in light of Lemma 1, we have $\psi({\boldsymbol\lambda})\geq\psi((\bar{\lambda},\ldots,\bar{\lambda}))$, which in turn implies that

(1) \begin{equation}\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda_{i}x})\leq\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\bar{\lambda}x})\leq\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda x}).\end{equation}

Next, we need to show that $\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda x})\leq(1-p\mathrm{e}^{-\lambda x})^n$. It is easy to see that $(p_1,\ldots,p_n)\stackrel{\rm m}{\succeq}(\bar{p},\ldots,\bar{p})$. Let $\phi({\textbf{{p}}})=-\prod_{i=1}^{n}(1-p_i\mathrm{e}^{-\lambda x})$. Observe that, for any $1\leq i<j\leq n$,

\begin{equation*} \frac{\partial\phi({\textbf{{p}}})}{\partial p_i}-\frac{\partial\phi({\textbf{{p}}})}{\partial p_j}=\frac{\mathrm{e}^{-\lambda x}}{1-p_j\mathrm{e}^{-\lambda x}}\phi({\textbf{{p}}})-\frac{\mathrm{e}^{-\lambda x}}{1-p_i\mathrm{e}^{-\lambda x}}\phi({\textbf{{p}}})\stackrel{\rm sgn}{=} p_i-p_j\geq0.\end{equation*}

Then, by applying Lemma 1, it follows that $\phi({\textbf{{p}}})\geq\phi((\bar{p},\ldots,\bar{p}))$, which further implies that $\prod_{i=1}^{n}(1-p_{i}\mathrm{e}^{-\lambda x})\leq(1-\bar{p}\mathrm{e}^{-\lambda x})^n\leq(1-p\mathrm{e}^{-\lambda x})^n$. Then, the desired result is proved by combining this with (1).

Under certain conditions restricted on the occurrence probabilities, Theorem 3 states that the largest claim amount from the heterogenous insurance portfolio not only has a greater tail function than the largest claim amount from the homogeneous insurance portfolio, but also is more dispersive in the distance between the corresponding quantiles for any two fixed confidence levels. It is of interest to investigate whether the condition $\lambda\geq\bar{\lambda}$ in Theorem 3 could be relaxed to $\lambda\geq\tilde{\lambda}\;:\!=\;\left(\prod_{i=1}^n\lambda_i\right)^{1/n}$; this is left as an open problem.

The next theorem establishes an equivalent characterization for the right-spread ordering of the largest claim amounts.

Theorem 4. Under the setup of Theorem 1, if $p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}$ then $Y_{n\;:\;n}\geq_{\rm RS}Y_{n\;:\;n}^{\ast}$ if and only if $\lambda\geq\lambda^{\ast}$, where

(2) \begin{equation} \lambda^{\ast}=\frac{\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\sum_{j=1}^{k}\frac{1}{j}}{\sum_{k=1}^{n}(-1)^{k+1}\sum\limits_{1\leq i_1\leq\cdots\leq i_k\leq n}\left(\frac{\prod_{j=1}^{k}p_{i_j}}{\sum_{j=1}^{k}\lambda_{i_j}}\right)}. \end{equation}

Proof. According to the proof of Corollary 2, we have

\begin{equation*} \mathbb{E}[Y_{n\;:\;n}^{\ast}]=\frac{1}{\lambda}\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\sum_{j=1}^{k}\frac{1}{j}.\end{equation*}

On the other hand, we can calculate that

\begin{eqnarray*}\mathbb{E}[Y_{n\;:\;n}]&=&\int_{0}^{\infty}\left(1-\prod_{i=1}^{n}\big(1-p_{i}\mathrm{e}^{-\lambda_{i}x}\big)\right)\mathrm{d} x\\[5pt] &=& \int_{0}^{\infty}\left(\sum_{k=1}^{n}(-1)^{k+1}\sum\limits_{1\leq i_1\leq\cdots\leq i_k\leq n}\prod_{j=1}^{k}p_{i_j}\mathrm{e}^{-\sum_{j=1}^{k}\lambda_{i_j}x}\right)\mathrm{d} x\\[5pt] &=&\sum_{k=1}^{n}(-1)^{k+1}\sum\limits_{1\leq i_1\leq\cdots\leq i_k\leq n}\left(\frac{\prod_{j=1}^{k}p_{i_j}}{\sum_{j=1}^{k}\lambda_{i_j}}\right).\end{eqnarray*}

Then the desired result is proved by applying Theorem 2(ii).

Since the right-spread ordering between two random variables implies the ordering of their variances, the following corollary is a direct result of Theorem 4.

Corollary 3. Under the setup of Theorem 1, suppose that $p\geq1-\prod_{i=1}^{n}(1-p_{i})^{1/n}$. Then, $\lambda\geq\lambda^{\ast}$, where $\lambda^{\ast}$ is defined in (2), implies

(3) \begin{align} {\rm Var}[Y_{n\;:\;n}] & \geq \frac{1}{\lambda^2}\left[\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\left(\sum_{j=1}^{k}\frac{1}{j^2}+\left(\sum_{j=1}^{k}\frac{1}{j}\right)^2\right)\right. \notag\\ & \qquad\qquad\qquad\qquad\qquad\qquad\left.-\left(\sum_{k=1}^{n}\binom{n}{k}p^{k}(1-p)^{n-k}\sum_{j=1}^{k}\frac{1}{j}\right)^2\right], \end{align}

The following example illustrates the lower bound given in Corollary 3.

Example 1. Under the setup of Corollary 3, we assume that $(p_1,p_2,p_3)=(0.4,0.6,0.9)$ and $(\lambda_1,\lambda_2,\lambda_3)=(0.1,0.3,1.2)$. It can be calculated that $1-\prod_{i=1}^{3}(1-p_{i})^{1/3}=0.7116$. Then, under the assumption that $\lambda=\lambda^{\ast}$, one can compute the lower bound of the variance of $Y_{3\;:\;3}$ in (3) for each $p\geq0.7116$, which is displayed as in Figure 3. It is clear that the lower bound becomes smaller when the occurrence probability for the homogeneous insurance portfolio gets larger.

Figure 1. Plot of the lower bound of ${\rm Var}[Y_{3\;:\;3}]$ in Example 1 with respect to $p\in[0.7116,1]$.

To conclude this section, we establish the convex transform and star orderings for the largest claim amounts from two sets of homogeneous insurance portfolios when the baseline distributions of claim severities are different.

Theorem 5. Let $X_{i}$ ($X_{i}^{\ast}$) be independent random variables with common distribution $F_1$ ($F_2$), $i=1,2,\ldots,n$. Let $I_{1},\ldots,I_{n}$ ($I_{1}^{\ast},\ldots,I_{n}^{\ast}$) be a set of independent Bernoulli random variables such that $\mathbb{E}[I_{i}]=p$ ($\mathbb{E}[I_{i}^{\ast}]=p$), $i=1,2,\ldots,n$. Let $Y_{i}=I_{i}X_{i}$ and $Y_{i}^{\ast}=I_{i}^{\ast}X_{i}^{\ast}$ be the claim amounts arising from two sets of insurance portfolios, $i=1,2,\ldots,n$. Then, $F_1\geq_{\rm c \, (\star)} F_2\Longleftrightarrow Y_{n\;:\;n}\geq_{\rm c \, (\star)}Y_{n\;:\;n}^{\ast}$.

Proof. Let $F_{n}$ and $G_{n}$ be the distribution functions of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$, respectively. Then, for $x\in\mathbb{R}_{+}$, we have $F_{n}(x)=(1-p\bar{F}_1(x))^n$ and $G_{n}(x)=(1-p\bar{F}_2(x))^n$. First, it is easy to see that $F_{n}^{-1}(G_{n}(x))$ is well defined on $x\in\mathbb{R}_{+}$. Then, the convex transform ordering result is equivalent to showing that $F_{n}^{-1}(G_{n}(x))=\bar{F}_1^{-1}(\bar{F}_2(x))$ is convex in $x\in\mathbb{R}_{+}$, and the star ordering result is equivalent to showing that $F_{n}^{-1}(G_{n}(x))/x=\bar{F}_1^{-1}(\bar{F}_2(x))/x$ is increasing in $x\in\mathbb{R}_{+}$. Note that $\bar{F}_1^{-1}(\bar{F}_2(x))=F_1^{-1}(1-\bar{F}_2(x))=F_1^{-1}(F_2(x))$ for all $x\in\mathbb{R}_{+}$. Thus, the convex transform order holds if and only if $F_1^{-1}(F_2(x))$ is convex in $x\in\mathbb{R}_{+}$, and star order holds if and only if $F_1^{-1}(F_2(x))$ is star-shaped in $x\in\mathbb{R}_{+}$, which are implied by (ii) and (iii) in Definition 1, respectively.

A natural interesting question is to study the effects of the occurrence probabilities and the number of claims on the variability of largest claim amounts from homogeneous insurance portfolios by means of the convex transform order and the star order; this remains open and needs further investigation.

4. Star order

In the context of insurance practice, some insureds may have higher occurrence probabilities/claim sizes than others (in the sense of some stochastic orders between their claim severity distributions) in an insurance portfolio. Therefore, a natural way for describing this phenomenon falls in the multiple-outlier claims model. This section carries out stochastic comparisons on the largest claim amounts arising from multiple-outlier scaled or PHR claims in the sense of the star and dispersive orderings.

4.1. Multiple-outlier scaled claims

In this subsection we investigate the effect of heterogeneity among scale parameters on the skewness of the largest claim amounts from independent multiple-outlier scaled claims. The following lemma is originally due to [Reference Saunders and Moran34] and plays a key role in proving the main results.

Lemma 3. Let $\{G_{\lambda} \mid \lambda \in \mathbb{R}_{+}\}$ be a class of distribution functions such that $G_{\lambda}$ is supported on some open interval $I \subseteq \mathbb{R}_{+}$ and has density $g_{\lambda}$ which does not vanish on any subinterval of I. Then, $G_{\lambda} \leq_{\star} G_{\lambda^{\ast}}$, for $\lambda \leq \lambda^{\ast}$ if and only if $\big({\frac{\partial G_{\lambda}(x)}{\partial\lambda}}\big)/({xg_{\lambda}(x)})$ is decreasing in x, where $\frac{\partial G_{\lambda}(\cdot)}{\partial\lambda}$ is the derivative of $G_{\lambda}$ with respect to $\lambda$.

To begin with, let $G(\cdot)$ be the baseline distribution function with its reversed hazard rate function given by $\tilde{r}(\cdot)$.

Theorem 6. Let $X_{1},\ldots,X_{n_{1}}$ ($X_{1}^{\ast},\ldots,X_{n_{1}}^{\ast}$) be independent scaled random variables with distributions $G(\lambda_{1}x)$ ($G(\mu_{1}x)$), and $X_{n_{1}+1},\ldots,X_{n}$ ($X_{n_{1}+1}^{\ast},\ldots,X_{n}^{\ast}$) be another set of independent scaled random variables with distributions $G(\lambda_{2}x)$ ($G(\mu_{2}x)$) for $x\in\mathbb{R}_+$. Let $I_{1},\ldots,I_{n_{1}}$ be independent Bernoulli random variables such that $\mathbb{E}[I_{i}]=p_{1}$ for $i=1,2,\ldots,n_{1}$, and let $I_{n_{1}+1},\ldots,I_{n}$ be another set of independent Bernoulli random variables such that $\mathbb{E}[I_{j}]=p_{2}$ for $j=n_{1}+1,\ldots,n$. Suppose that $(p_{1}-p_{2})(\lambda_{1}-\lambda_{2})\geq0$ and $(\lambda_{1}-\lambda_{2})(\mu_{1}-\mu_{2})\geq0$. If both $x\tilde{r}(x)$ and $x\tilde{r}'(x)/\tilde{r}(x)$ are decreasing in $x\in\mathbb{R}_{+}$, we have

\begin{equation*} \frac{\lambda_{2\;:\;2}}{\lambda_{1\;:\;2}}\geq\frac{\mu_{2\;:\;2}}{\mu_{1\;:\;2}}\Longrightarrow Y_{n\;:\;n}\geq_{\rm \star}Y_{n\;:\;n}^{\ast}.\end{equation*}

Proof. Let $n_{2}=n-n_{1}$. Without loss of generality, it is assumed that $p_{1}\geq p_{2}$, $\lambda_{1}\geq\lambda_{2}$, and $\mu_{1}\geq\mu_{2}$. Note that the the distributions of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ have the same probability at the value 0. The proof is completed by using the following two steps.

Case 1: $\lambda_{1}+\lambda_{2}=\mu_{1}+\mu_{2}=c$. Let $\lambda_{1}=\lambda\geq\lambda_{2}$ and $\mu_{1}=\mu\geq\mu_{2}$. Based on Lemma 3, it is enough to show that $\big({\frac{\partial G_{n,\lambda}(x)}{\partial \lambda}}\big)({xg_{n,\lambda}(x)})$ is decreasing in $x\in\mathbb{R}_{+}$ for $\lambda\in[\frac{c}{2},c]$. Note that the distribution of $Y_{n\;:\;n}$ can be denoted as $G_{n,\lambda}(x)=[1-p_{1}\bar{G}(\lambda x)]^{n_{1}}[1-p_{2}\bar{G}((c-\lambda)x)]^{n_{2}}$, with its density function defined as

\begin{align*} g_{n,\lambda}(x) & = [1-p_{1}\bar{G}(\lambda x)]^{n_{1}-1}[1-p_{2}\bar{G}((c-\lambda)x)]^{n_{2}-1} \\ & \quad \times \left\{n_{1}p_{1}\lambda g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]+n_{2}p_{2}(c-\lambda) g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]\right\}\end{align*}

and

\begin{align*} \frac{\partial G_{n,\lambda}(x)}{\partial \lambda} & = x[1-p_{1}\bar{G}(\lambda x)]^{n_{1}-1}[1-p_{2}\bar{G}((c-\lambda)x)]^{n_{2}-1} \\ & \quad \times \left\{n_{1}p_{1}g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]-n_{2}p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]\right\}.\end{align*}

Then, it suffices to prove that

\begin{align*} \frac{\frac{\partial G_{n,\lambda}(x)}{\partial \lambda}}{xg_{n,\lambda}(x)} & = \frac{n_{1}p_{1}g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]-n_{2}p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]}{n_{1}p_{1}\lambda g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]+n_{2}p_{2}(c-\lambda) g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]} \\ & = \left\{n_{1}p_{1}g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]-n_{2}p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]\right\} \\ & \quad \times \left\{\lambda \{n_{1}p_{1}g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]-n_{2}p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]\}\right. \\ & \qquad \quad \left. + \, c n_{2}p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]\right\}^{-1} \\ & = \left(\lambda+\frac{c n_{2}p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]}{n_{1}p_{1}g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]-n_{2}p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]}\right)^{-1} \\ & = \left[\lambda+c\left(\frac{n_{1}}{n_{2}}\times\frac{p_{1}g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]}{p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]}-1\right)^{-1}\right]^{-1}\end{align*}

is decreasing in $x\in\mathbb{R}_{+}$ for $\lambda\in[\frac{c}{2},c]$. Thus, it is sufficient to prove that

\begin{eqnarray*}\Delta(x)&=&\frac{p_{1}g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]}{p_{2}g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]}\\&=& \frac{\tilde{r}(\lambda x)}{\tilde{r}((c-\lambda)x)}\times\frac{p_{1}G(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]}{p_{2}G((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]}\\&=:&\Delta_{1}(x)\times\Delta_{2}(x)\end{eqnarray*}

is decreasing in $x\in\mathbb{R}_{+}$ for $\lambda\in[\frac{c}{2},c]$, where

\begin{equation*}\Delta_{1}(x)=\frac{\tilde{r}(\lambda x)}{\tilde{r}((c-\lambda)x)} , \qquad \Delta_{2}(x)=\frac{p_{1}G(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]}{p_{2}G((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]}.\end{equation*}

On one hand, by using the assumption that $x\tilde{r}'(x)/\tilde{r}(x)$ is decreasing in $x\in\mathbb{R}_{+}$, we have

\begin{equation*}\left[\frac{\tilde{r}(\lambda x)}{\tilde{r}((c-\lambda)x)}\right]'\stackrel{\rm sgn}{=}\frac{\lambda x\tilde{r}'(\lambda x)}{\tilde{r}(\lambda x)}-\frac{(c-\lambda)x\tilde{r}'((c-\lambda)x)}{\tilde{r}((c-\lambda)x)}\leq0.\end{equation*}

Next, we will prove that $\Delta_{2}(x)$ is also decreasing in $x\in\mathbb{R}_{+}$. Observe that

\begin{align*}\Delta^{\prime}_{2}(x) & \stackrel{\rm sgn}{=} \left\{p_{1}\lambda g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]+p_{1}p_{2}(c-\lambda)g((c-\lambda)x)G(\lambda x)\right\} \\&\quad\ \times p_{2}G((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]\\&\quad\ -\left\{p_{2}(c-\lambda)g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]+p_{1}p_{2}\lambda g(\lambda x)G((c-\lambda)x)\right\} \\&\quad\ \times p_{1}G(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]\\&=p_{1}p_{2}[1-p_{1}\bar{G}(\lambda x)][1-p_{2}\bar{G}((c-\lambda)x)]\\&\quad\times\left[\lambda g(\lambda x)G((c-\lambda)x)-(c-\lambda)g((c-\lambda)x)G(\lambda x)\right]\\&\quad+p_{1}p_{2}(G\lambda x)G((c-\lambda)x)\left\{p_{2}(c-\lambda)g((c-\lambda)x)[1-p_{1}\bar{G}(\lambda x)]\right.\\&\qquad\left.-p_{1}\lambda g(\lambda x)[1-p_{2}\bar{G}((c-\lambda)x)]\right\}\\&\stackrel{\rm sgn}{=} \lambda x\tilde{r}(\lambda x)-(c-\lambda)x\tilde{r}((c-\lambda)x)+\frac{p_{2}(c-\lambda)x g((c-\lambda)x)}{1-p_{2}\bar{G}((c-\lambda)x)}-\frac{p_{1}\lambda x g(\lambda x)}{1-p_{1}\bar{G}(\lambda x)}\\&= \lambda x\tilde{r}(\lambda x)-(c-\lambda)x\tilde{r}((c-\lambda)x)\\&\quad+(c-\lambda)x\tilde{r}((c-\lambda)x)\times\frac{p_{2} G((c-\lambda)x)}{1-p_{2}\bar{G}((c-\lambda)x)}-\lambda x\tilde{r}(\lambda x)\times\frac{p_{1}G(\lambda x)}{1-p_{1}\bar{G}(\lambda x)}\\&= \lambda x\tilde{r}(\lambda x)-(c-\lambda)x\tilde{r}((c-\lambda)x)\\&\quad+(c-\lambda)x\tilde{r}((c-\lambda)x)\times\left(1+\frac{\frac{1}{p_{2}}-1}{G((c-\lambda)x)}\right)^{-1}-\lambda x\tilde{r}(\lambda x)\times\left(1+\frac{\frac{1}{p_{1}}-1}{G(\lambda x)}\right)^{-1}\\&\leq\lambda x\tilde{r}(\lambda x)-(c-\lambda)x\tilde{r}((c-\lambda)x)\\&\quad+(c-\lambda)x\tilde{r}((c-\lambda)x)\times\left(1+\frac{\frac{1}{p_{1}}-1}{G(\lambda x)}\right)^{-1}-\lambda x\tilde{r}(\lambda x)\times\left(1+\frac{\frac{1}{p_{1}}-1}{G(\lambda x)}\right)^{-1}\\&=\left[\lambda x\tilde{r}(\lambda x)-(c-\lambda)x\tilde{r}((c-\lambda)x)\right]\times\left[1-\left(1+\frac{\frac{1}{p_{1}}-1}{G(\lambda x)}\right)^{-1}\right]\\&\leq0,\end{align*}

where the first inequality is based on $1\geq p_{1}\geq p_{2}\geq0$ and $G((c-\lambda)x)\leq G(\lambda x)$, and the last inequality is due to the decreasing property of $x\tilde{r}(x)$ in $x\in\mathbb{R}_+$. Thus, the proof is finished.

Case 2: $\lambda_{1}+\lambda_{2}\neq\mu_{1}+\mu_{2}$. In this case, there exists some $\alpha>0$ such that $\lambda_{1}+\lambda_{2}=\alpha(\mu_{1}+\mu_{2})$. Now, let $Z_{n\;:\;n}$ be the largest claim amount from $I_{1}Z_{1}, \ldots, I_{n_{1}}Z_{n_{1}}$, $I_{n_{1}+1}Z_{n_{1}+1}, \ldots, I_{n}Z_{n}$, where $Z_{1},\ldots,Z_{n_{1}}$ have distribution $G(\alpha\mu_{1}x)$ and $Z_{n_{1}+1},\ldots,Z_{n}$ have distribution $G(\alpha\mu_{2}x)$. According to Case 1, $Y_{n\;:\;n}\geq_{\rm \star}Z_{n\;:\;n}$. Since the star order is scale invariant, it follows that $Y_{n\;:\;n}\geq_{\rm \star}Y_{n\;:\;n}^{\ast}$.

Remark 2. By considering the claim severity as the scale model $G(\lambda_i x)$ in Theorem 6, $i=1,2$, it holds that $G(\lambda_1 x)\leq G(\lambda_2 x)$ for $\lambda_1\leq\lambda_2$ and $x>0$. This means that the policyholders with claim severity distribution $G(\lambda_1 x)$ have larger claims than the policyholders with claim severity distribution $G(\lambda_2 x)$ in the sense of the usual stochastic order. The result of Theorem 6 implies that, under suitable conditions imposed on the baseline distribution, the occurrence probabilities, and the scale parameters, more heterogeneity among the scale parameters leads to a more skewed distribution function of the largest claim amount. For the case of $p_1=p_2=1$, the result of Theorem 6 reduces to [Reference Ding, Yang and Ling14, Theorem 8(a)], for which it is only required that $x\tilde{r}'(x)/\tilde{r}(x)$ is decreasing in $x\in\mathbb{R}_{+}$.

There are many distribution families that satisfy the conditions of Theorem 6. For example, a random variable X is said to have a generalized gamma distribution if its density function can be written as $g_{\alpha,\beta}(x)=\frac{\alpha}{\Gamma(\beta/\alpha)}x^{\beta-1}\mathrm{e}^{-x^{\alpha}}$, $x\in\mathbb{R}_{+}$, where $\alpha>0$ and $\beta>0$ are the shape parameters. It includes the widely used exponential $(\alpha=1,\beta=1)$, Weibull $(\alpha=\beta)$, and gamma $(\alpha=1)$ distributions as special cases. If we denote by $\tilde{r}(x;\alpha,\beta)$ the reversed hazard rate function of the generalized random variable, [Reference Khaledi, Farsinezhad and Kochar21] proved that $x\tilde{r}(x;\alpha,\beta)$ is decreasing in $x\in\mathbb{R}_{+}$ for $\alpha,\beta>0$, and [Reference Ding, Yang and Ling14] showed that $x\tilde{r}'(x;\alpha,\beta)/\tilde{r}(x;\alpha,\beta)$ is decreasing in $x\in\mathbb{R}_{+}$ for $\alpha,\beta>0$. Thus, the conditions of Theorem 6 always hold for the generalized gamma distribution.

Next, we give a numerical example to illustrate Theorem 6.

Figure 2. Plots of coefficients of variation of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ under different settings.

Example 2. Suppose the baseline distribution is exponential with parameter 1. Figure 2 displays the plots of the coefficients of variation of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ under the following six settings:

  1. (a) $(n_1,n_2)=(8,2)$, $(\lambda_1,\lambda_2)=(0.9,0.3)$, $(\mu_1,\mu_2)=(0.25,0.1)$;

  2. (b) $(n_1,n_2)=(8,2)$, $(\lambda_1,\lambda_2)=(0.9,0.3)$, $(\mu_1,\mu_2)=(1.6,0.6)$;

  3. (c) $(n_1,n_2)=(2,10)$, $(\lambda_1,\lambda_2)=(0.9,0.3)$, $(\mu_1,\mu_2)=(0.25,0.1)$;

  4. (d) $(n_1,n_2)=(2,10)$, $(\lambda_1,\lambda_2)=(0.9,0.3)$, $(\mu_1,\mu_2)=(1.6,0.6)$;

  5. (e) $(n_1,n_2)=(2,4)$, $(\lambda_1,\lambda_2)=(0.9,0.3)$, $(\mu_1,\mu_2)=(0.6,1.6)$;

  6. (f) $(n_1,n_2)=(4,2)$, $(\lambda_1,\lambda_2)=(0.9,0.3)$, $(\mu_1,\mu_2)=(0.6,1.6)$.

Under the settings of (a)–(d), these assumptions satisfy the conditions of Theorem 6 and thus the coefficient of variation of $Y_{n\;:\;n}$ is always larger than that of $Y_{n\;:\;n}^{\ast}$ for $p_1\geq p_2$. On the other hand, as observed in Figures 2(e) and 2(f), their coefficients of variation will cross each other if the condition $(\lambda_1-\lambda_2)(\mu_1-\mu_2)\geq0$ is not fulfilled.

4.2. Multiple-outlier PHR claims

Independent random variables $X_1,\ldots, X_n$ are said to follow a proportional hazard rates model if the survival function of $X_i$ can be written as $\bar{F}_i(x)=[\bar{F}(x)]^{\lambda_{i}}$ for $i = 1,\ldots,n$, where $\bar{F}(x)$ is the baseline survival function. Let $h(\cdot)$ be the hazard rate function of F. Then the survival function of $X_i$ can be written as $\bar{F}_i(x) = \mathrm{e}^{-\lambda_{i}R(x)}$ for $i = 1,\ldots,n$, where $R(x) =\int_{0}^{x}h(t) \, \mathrm{d} t$ is the cumulative hazard rate function of X. Many well-known distributions are special cases of the PHR model such as the exponential, Weibull, Pareto, and Lomax distributions.

This subsection studies the effects of heterogeneity among hazard rate parameters on the skewness of the largest claim amount from independent multiple-outlier PHR claims.

Theorem 7. Let $X_{1},\ldots,X_{n}$ ($X_{1}^{\ast},\ldots,X_{n}^{\ast}$) be independent random variables with survival functions $(\bar{F}^{\lambda_{1}}{\textbf{1}}_{n_{1}},\bar{F}^{\lambda_{2}}{\textbf{1}}_{n_{2}})$ ($(\bar{F}^{\mu_{1}}{\textbf{1}}_{n_{1}},\bar{F}^{\mu_{2}}{\textbf{1}}_{n_{2}})$), where $n_{1}+n_{2}=n$. Let $I_{1},\ldots,I_{n}$ be a set of independent Bernoulli random variables such that $\mathbb{E}[I_{i}]=p$ for $i=1,2,\ldots,n$. Suppose that $(\lambda_{1}-\lambda_{2})(\mu_{1}-\mu_{2})\geq0$. If $\frac{R(x)}{xh(x)}$ is increasing in $x\in\mathbb{R}_{+}$, we have $(\lambda_{1}{\textbf{1}}_{n_{1}},\lambda_{2}{\textbf{1}}_{n_{2}})\stackrel{\rm m}{\succeq}(\mu_{1}{\textbf{1}}_{n_{1}},\mu_{2}{\textbf{1}}_{n_{2}})\Longrightarrow Y_{n\;:\;n}\geq_{\rm \star}Y_{n\;:\;n}^{\ast}$,

Proof. Without loss of generality, it is assumed that $\lambda_{1}\leq\lambda_{2}$ and $\mu_{1}\leq\mu_{2}$. Let $\lambda=\lambda_{2}$, $\mu=\mu_{2}$, and $n_{1}\lambda_{1}+n_{2}\lambda_{2}=n_{1}\mu_{1}+n_{2}\mu_{2}=1$. Then we have $\lambda_{1}=(1-n_{2}\lambda_{2})/n_{1}$ and $\lambda\in[1/(n_{1}+n_{2}),1/n_{2})$. The distribution function of $Y_{n\;:\;n}$ is given by

\begin{equation*}F_{\lambda, n\;:\;n}(x)=\left[1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}\right]^{n_{1}}\left[1-p\mathrm{e}^{-\lambda R(x)}\right]^{n_{2}}.\end{equation*}

Taking the derivative of $F_{Y_{n\;:\;n}}(x)$ with respective to $\lambda$ gives

\begin{multline*}\frac{\partial F_{\lambda, n\;:\;n}(x)}{\partial \lambda} = -R(x)\left[\frac{n_{2}p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}{1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}-\frac{n_{2}p\mathrm{e}^{-\lambda R(x)}}{1-p\mathrm{e}^{-\lambda R(x)}}\right]\\\times\left[1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}\right]^{n_{1}}\left[1-p\mathrm{e}^{-\lambda R(x)}\right]^{n_{2}}.\end{multline*}

The density function of $Y_{n\;:\;n}$ is

\begin{multline*}f_{\lambda, n\;:\;n}(x) = h(x)\left[\frac{(1-n_{2}\lambda)p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}{1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}+\frac{\lambda n_{2}p\mathrm{e}^{-\lambda R(x)}}{1-p\mathrm{e}^{-\lambda R(x)}}\right] \\ \times\left[1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}\right]^{n_{1}}\left[1-p\mathrm{e}^{-\lambda R(x)}\right]^{n_{2}}.\end{multline*}

By making use of Lemma 3, it is equivalent to showing that

\begin{equation*}-\frac{\frac{\partial F_{\lambda, n\;:\;n}(x)}{\partial \lambda}}{xf_{\lambda, n\;:\;n}(x)}=\frac{R(x)}{xh(x)}\cdot\frac{\frac{n_{2}p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}{1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}-\frac{n_{2}p\mathrm{e}^{-\lambda R(x)}}{1-p\mathrm{e}^{-\lambda R(x)}}}{\frac{(1-n_{2}\lambda)p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}{1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}+\frac{\lambda n_{2}p\mathrm{e}^{-\lambda R(x)}}{1-p\mathrm{e}^{-\lambda R(x)}}}=:\frac{R(x)}{xh(x)}\cdot\frac{\Omega_{1}(x)}{\Omega_{2}(x)}\end{equation*}

is increasing in $x\in\mathbb{R}_{+}$. Note that

\begin{eqnarray*}\Omega_{1}(x)&=&\frac{n_{2}p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}{1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}-\frac{n_{2}p\mathrm{e}^{-\lambda R(x)}}{1-p\mathrm{e}^{-\lambda R(x)}}\\&\stackrel{\rm sgn}{=}&\left[\frac{1}{p}\exp\left\{\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}-1\right]^{-1}-\left[\frac{1}{p}\mathrm{e}^{\lambda R(x)}-1\right]^{-1}\geq0,\end{eqnarray*}

where the inequality is due to $(1-n_{2}\lambda)/n_{1}\leq \lambda$. Based on the assumption that $\frac{R(x)}{xh(x)}$ is increasing in $x\in\mathbb{R}_{+}$, it suffices to show that $\Omega_{3}(x)=\Omega_{1}(x)/\Omega_{2}(x)$ is also increasing in $x\in\mathbb{R}_{+}$. Observe that

\begin{equation*}\Omega_{3}(x)=\left[n_{2}^{-1}\left(1-\frac{p\mathrm{e}^{-\lambda R(x)}}{1-p\mathrm{e}^{-\lambda R(x)}}\times\left(\frac{p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}{1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}\right)^{-1}\right)^{-1}-\lambda\right]^{-1}.\end{equation*}

It is enough to prove that

\begin{equation*}\Omega_{4}(x)=\frac{p\mathrm{e}^{-\lambda R(x)}}{1-p\mathrm{e}^{-\lambda R(x)}} \!\left(\frac{p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}{1-p\exp\left\{-\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}\right)^{-1}\!\!=\frac{\exp\left\{\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}-p}{\mathrm{e}^{\lambda R(x)}-p}\end{equation*}

is decreasing in $x\in\mathbb{R}_{+}$, which can be obtained from the fact that $(1-n_{2}\lambda)/n_{1}\leq \lambda$ and $x/(1-p\mathrm{e}^{-x})$ is increasing in $x\in\mathbb{R}_{+}$ for any $p\in[0,1]$, and the observation that

\begin{align*} \Omega^{\prime}_{4}(x) & \stackrel{\rm sgn}{=} \frac{1-n_{2}\lambda}{n_{1}}h(x)\exp\left\{\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}\left[\mathrm{e}^{\lambda R(x)}-p\right] \\[5pt] & \quad \ - \lambda h(x)\mathrm{e}^{\lambda R(x)}\left[\exp\left\{\left(\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}-p\right] \\[5pt] & \stackrel{\rm sgn}{=} \frac{\frac{1-n_{2}\lambda}{n_{1}}R(x)}{1-p\exp\left\{\left(-\frac{1-n_{2}\lambda}{n_{1}}\right)R(x)\right\}}-\frac{\lambda R(x)}{1-p\mathrm{e}^{-\lambda R(x)}}\leq0.\end{align*}

Hence, the desired result follows.

The next example provides an illustration for the condition $\frac{R(x)}{xh(x)}$ is increasing in $x\in\mathbb{R}_{+}$ employed in Theorem 7 when the baseline distribution is taken as the power-generalized Weibull distribution.

Example 3. A random variable X is said to have the power-generalized Weibull distribution, denoted by $X\sim PGW(\nu,\gamma)$, if its density function is given as $f(x;\nu,\gamma)=\frac{\nu}{\gamma}x^{\nu-1}(1+x^\nu)^{\frac{1}{\gamma}-1}\exp\big\{1-(1+x^\nu)^{\frac{1}{\gamma}}\big\}$, $x>0$, $\nu,\gamma>0$. Then its survival function can be derived as $\bar{F}(x;\nu,\gamma)=\exp\big\{1-(1+x^\nu)^{\frac{1}{\gamma}}\big\}$ for $x>0$ and $\nu,\gamma>0$. The hazard rate function is $h(x;\nu,\gamma)=\frac{\nu}{\gamma}x^{\nu-1}(1+x^\nu)^{\frac{1}{\gamma}-1}$ for $x>0$ and $\nu,\gamma>0$. Then, we have

\begin{equation*} \psi_1(x)\;:\!=\;\frac{R(x;\nu,\gamma)}{xh(x;\nu,\gamma)}=\frac{\int_0^x t^{\nu-1}(1+t^\nu)^{\frac{1}{\gamma}-1} \, \mathrm{d} t}{x^{\nu}(1+x^\nu)^{\frac{1}{\gamma}-1}},\qquad x>0,\quad \nu,\gamma>0.\end{equation*}

So,

\begin{eqnarray*} \psi^{\prime}_1(x) &\stackrel{\rm sgn}{=}& x^{\nu-1}(1+x^\nu)^{\frac{1}{\gamma}-1}\times x^{\nu}(1+x^\nu)^{\frac{1}{\gamma}-1}-\int_0^x t^{\nu-1}(1+t^\nu)^{\frac{1}{\gamma}-1} \, \mathrm{d} t \\ &&\quad\times\left[\nu x^{\nu-1}(1+x^\nu)^{\frac{1}{\gamma}-1}+\left(\frac{1}{\gamma}-1\right)\nu x^{2\nu-1}(1+x^\nu)^{\frac{1}{\gamma}-2}\right] \\ &\stackrel{\rm sgn}{=}& x^{\nu}(1+x^\nu)^{\frac{1}{\gamma}} -\nu\left(1+\frac{x^\nu}{\gamma}\right)\int_0^x t^{\nu-1}(1+t^\nu)^{\frac{1}{\gamma}-1} \, \mathrm{d} t.\end{eqnarray*}

Note that it is not easy to judge the sign of $\psi^{\prime}_1(x)$ on $x>0$ for arbitrary $\nu,\gamma>0$. Here, we consider the following two special cases:

Case 1: $\nu=1$ and $\gamma\geq1$. For this case, one can note that

\begin{equation*} \psi^{\prime}_1(x) \stackrel{\rm sgn}{=} x(1+x)^{\frac{1}{\gamma}}-\left(1+\frac{x}{\gamma}\right)\int_0^x (1+t^\nu)^{\frac{1}{\gamma}-1} \, \mathrm{d} t = x+r-r(1+x)^{\frac{1}{\gamma}}=:\psi_2(x).\end{equation*}

Since $\psi^{\prime}_2(x)=1-(1+x)^{\frac{1}{\gamma}-1}\geq0$ for all $x>0$ and $\gamma\geq1$, it follows that $\psi_2(x)$ is increasing in $x>0$. Thus, $\psi_2(x)\geq\lim_{x\to0_+}\psi_2(x)=0$, which implies that $\psi^{\prime}_1(x)\geq0$, i.e. $\psi_1(x)$ is increasing in $x>0$ when $\nu=1$ and $\gamma\geq1$.

Case 2: $\gamma=1$. It is easy to check that $\psi^{\prime}_1(x)=0$, which means that $\psi_1(x)$ is a constant for $x>0$. Simple calculations can be implemented to derive that $\psi_1(x)=\frac{1}{\nu}$. This case corresponds to the traditional Weibull distribution, and has been also verified in some recent works; see [Reference Amini-Seresht, Qiao, Zhang and Zhao2, Reference Kochar and Xu25], for example.

Remark 3. The result of Theorem 7 was proved in [Reference Ahmed, Alzaid, Bartoszewicz and Kochar2, Theorem 3.10] for the case of $p=1$. One may wonder whether the result of Theorem 7 could be generalized to the setting of different occurrence probabilities; however, we cannot prove it with a similar proof method to Theorem 7. We leave it as an open problem.

The next example states that the condition $(\lambda_1-\lambda_2)(\mu_1-\mu_2)\geq0$ in Theorem 7 cannot be removed.

Example 4. Let $\bar{F}(x)=\mathrm{e}^{-x}$ for $x\in\mathbb{R}_+$. Assume that $(n_1,n_2)=(3,4)$, $(\lambda_1,\lambda_2)=(1.9,0.3)$, and $(\mu_1,\lambda_2)=(13/30,1.4)$. It is plain that $(\lambda_1-\lambda_2)(\mu_1-\mu_2)<0$ and $(\lambda_{1}{\textbf{1}}_{n_{1}},\lambda_{2}{\textbf{1}}_{n_{2}})\stackrel{\rm m}{\succeq}(\mu_{1}{\textbf{1}}_{n_{1}},\mu_{2}{\textbf{1}}_{n_{2}})$. As shown in Figure 3, the coefficients of variation of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ cross each other for $p\in[0,1]$. Hence, $(\lambda_1-\lambda_2)(\mu_1-\mu_2)\geq0$ is necessary in Theorem 7.

The following result can be proved by using a similar proof method to Theorem 7; the details are omitted for brevity.

Figure 3. Plots of coefficients of variation of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$.

Theorem 8. Let $X_{1},\ldots,X_{n}$ ($X_{1}^{\ast},\ldots,X_{n}^{\ast}$) be independent random variables with survival functions $(\bar{F}^{\lambda_{1}}{\textbf{1}}_{n_{1}},\bar{F}^{\lambda}{\textbf{1}}_{n_{2}})$ ($(\bar{F}^{\lambda_{2}}{\textbf{1}}_{n_{1}},\bar{F}^{\lambda}{\textbf{1}}_{n_{2}})$), where $n_{1}+n_{2}=n$. Let $I_{1},\ldots,I_{n}$ be a set of independent Bernoulli random variables such that $\mathbb{E}[I_{i}]=p$ for $i=1,\ldots,n$. Suppose that $\lambda_{1}\leq\lambda_{2}\leq\lambda$. If $\frac{R(x)}{xh(x)}$ is increasing in $x\in\mathbb{R}_{+}$, we have $Y_{n\;:\;n}\geq_{\rm \star}Y_{n\;:\;n}^{\ast}$.

A general result follows stating that the largest claim amount becomes more skewed if the hazard rate parameters have more heterogeneity.

Theorem 9. Under the same setup as Theorem 7, we assume that $\lambda_{1}\leq\mu_{1}\leq\mu_{2}\leq\lambda_{2}$. If $\frac{R(x)}{xh(x)}$ is increasing in $x\in\mathbb{R}_{+}$, we have $(\lambda_{1}{\textbf{1}}_{n_{1}},\lambda_{2}{\textbf{1}}_{n_{2}})\stackrel{\rm w}{\succeq}(\mu_{1}{\textbf{1}}_{n_{1}},\mu_{2}{\textbf{1}}_{n_{2}})\Longrightarrow Y_{n\;:\;n}\geq_{\rm \star}Y_{n\;:\;n}^{\ast}$.

Proof. There is some $\lambda^{\prime}_{1}$ such that $\lambda_{1}<\lambda^{\prime}_{1}\leq\mu_{1}$, $(\lambda^{\prime}_{1}{\textbf{1}}_{n_{1}},\lambda_{2}{\textbf{1}}_{n_{2}})\stackrel{\rm m}{\succeq}(\mu_{1}{\textbf{1}}_{n_{1}},\mu_{2}{\textbf{1}}_{n_{2}})$. Let $Z_{1},\ldots,Z_{n}$ be a set of independent random variables with tail functions $(\bar{F}^{\lambda^{\prime}_{1}}{\textbf{1}}_{n_{1}}$, $\bar{F}^{\lambda_{2}}{\textbf{1}}_{n_{2}})$. Denote $Z_{n\;:\;n}^{\ast}=\max\{I_{p}Z_{1},\ldots,I_{p}Z_{n_{1}},I_{p}Z_{n_{1}+1},\ldots,I_{p}Z_{n}\}$. From Theorem 8, we know that $Y_{n\;:\;n}\geq_{\rm \star}Z_{n\;:\;n}^{\ast}$. According to Theorem 7, $Z_{n\;:\;n}^{\ast}\geq_{\rm \star}Y_{n\;:\;n}^{\ast}$ holds. Hence, the theorem is proved.

It is interesting to note that the result of Theorem 9 is contained in [Reference Amini-Seresht, Qiao, Zhang and Zhao2, Theorem 3.11] under the setting of $p=1$. Indeed, it was proved in [Reference Amini-Seresht, Qiao, Zhang and Zhao2, Theorem 3.15] that the conclusion still holds if the weak supermajorization order is replaced by the p-larger order. We conjecture that the result of Theorem 9 might also hold under the p-larger order; this is left as an open problem.

The next result shows that more heterogeneity among the hazard rate parameters leads to more variability of the largest claim amount according to the dispersive order.

Theorem 10. Under the same setup as Theorem 7, we assume that $\lambda_{1}\leq\mu_{1}\leq\mu_{2}\leq\lambda_{2}$. If $\frac{R(x)}{xh(x)}$ is increasing in $x\in\mathbb{R}_{+}$, we have $(\lambda_{1}{\textbf{1}}_{n_{1}},\lambda_{2}{\textbf{1}}_{n_{2}})\stackrel{\rm w}{\succeq}(\mu_{1}{\textbf{1}}_{n_{1}},\mu_{2}{\textbf{1}}_{n_{2}})\Longrightarrow Y_{n\;:\;n}\geq_{\rm disp}Y_{n\;:\;n}^{\ast}$.

Proof. It was proved in [Reference Ahmed, Alzaid, Bartoszewicz and Kochar1] that, for two random variables X and Y, if $X \leq_{\star}Y$ then $X \leq_{\rm st} Y$ implies that $X \leq_{\rm disp}Y$. According to [Reference Balakrishnan, Zhang and Zhao3, Theorem 3.7], $Y_{n\;:\;n}\geq_{\rm st}Y_{n\;:\;n}^{\ast}$ holds. Then, the desired result can be obtained from Theorem 9.

Remark 4. Since both $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ are discrete-continuous random variables, the dispersive order in Theorem 10 should be understood in the sense that $F^{-1}_{Y_{n\;:\;n}}(0)=F^{-1}_{Y_{n\;:\;n}^{\ast}}(0)=(1-p)^{n}$ and $F^{-1}_{Y_{n\;:\;n}}(u)-F^{-1}_{Y_{n\;:\;n}}(v)\geq F^{-1}_{Y_{n\;:\;n}^{\ast}}(u)-F^{-1}_{Y_{n\;:\;n}^{\ast}}(v)$ for any $(1-p)^{n}<v\leq u<1$.

5. Conclusions

The largest claim amount plays an important role in insurance analysis for insurance companies to determine the annual premiums paid by policyholders. We study the ordering properties of largest claim amounts according to some transform orderings. It is shown that, without any restrictions on the claim severity parameters, the largest claim amount from a set of independent and heterogeneous exponential claims is more skewed than that from a set of independent and homogeneous exponential claims in the sense of the convex transform ordering. Equivalent characterizations for the dispersive order and the right-spread order are also presented. Also, we establish sufficient conditions to compare the skewness of largest claim amounts from two sets of independent multiple-outlier scaled or PHR claims by means of the star ordering. Some useful lower bounds are provided for the coefficient of variation, the variance, and the survival function of the largest claim amount from a set of heterogeneous insurance claims.

It should be noted that our results can also be used in other areas such as reliability theory. More specifically, more heterogeneity among the two types of components subject to random shocks in a parallel system may lead to the system lifetime ageing faster in the sense of the star ordering (cf. Theorems 6 and 7) under appropriate conditions. Further investigations are needed to generalize the results in Section 3 to the case when the claim severities have more general distributions (e.g. PHR and scale models). On the other hand, from the viewpoint of statistical inference, it might be of interest to implement some parametric estimation methods in the inference issues of the parameters in the occurrence probabilities and claim severity distributions, and present some significant connections with the class of transform orders studied here.

Acknowledgements

The author acknowledges the insightful comments and suggestions from two anonymous reviewers, which have improved the presentation of the paper. The author also acknowledges the National Natural Science Foundation of China (no. 12101336) and the Natural Science Foundation of Tianjin (no. 20JCQNJC01740).

References

Ahmed, A. N., Alzaid, A., Bartoszewicz, J. and Kochar, S. C. (1986). Dispersive and superadditive ordering. Adv. Appl. Prob. 18, 10191022.10.2307/1427262CrossRefGoogle Scholar
Amini-Seresht, E., Qiao, J., Zhang, Y. and Zhao, P. (2016). On the skewness of order statistics in multiple-outlier PHR models. Metrika 79, 817836.10.1007/s00184-016-0579-7CrossRefGoogle Scholar
Balakrishnan, N., Zhang, Y. and Zhao, P. (2018). Ordering the largest claim amounts and ranges from two sets of heterogeneous portfolios. Scand. Actuarial J. 2018, 2341.10.1080/03461238.2017.1278717CrossRefGoogle Scholar
Barlow, R. and Proschan, R. (1981). Statistical Theory of Reliability and Life Testing. Silver Spring, MD: Madison.Google Scholar
Barmalzan, G. and Payandeh Najafabadi, A. T. (2015). On the convex transform and right-spread orders of smallest claim amounts. Insurance Math. Econom. 64, 380384.10.1016/j.insmatheco.2015.07.001CrossRefGoogle Scholar
Barmalzan, G., Payandeh Najafabadi, A. T. and Balakrishnan, N. (2015). Stochastic comparison of aggregate claim amounts between two heterogeneous portfolios and its applications. Insurance Math. Econom. 61, 235241.10.1016/j.insmatheco.2015.01.010CrossRefGoogle Scholar
Barmalzan, G., Payandeh Najafabadi, A. T. and Balakrishnan, N. (2016). Likelihood ratio and dispersive orders for smallest order statistics and smallest claim amounts from heterogeneous Weibull sample. Statist. Prob. Lett. 110, 17.10.1016/j.spl.2015.11.009CrossRefGoogle Scholar
Barmalzan, G., Payandeh Najafabadi, A. T. and Balakrishnan, N. (2017). Ordering properties of the smallest and largest claim amounts in a general scale model. Scand. Actuarial J. 2017, 105124.10.1080/03461238.2015.1090476CrossRefGoogle Scholar
Benktander, G. (1978). Largest claims reinsurance (LCR). A quick method to calculate LCR-risk rates from excess of loss risk rates. ASTIN Bull. 10, 5458.10.1017/S0515036100006346CrossRefGoogle Scholar
Denuit, M., Dhaene, J., Goovaerts, M. J. and Kaas, R. (2006). Actuarial Theory for Dependent Risks: Measures, Orders and Models. John Wiley, Chichester.Google Scholar
Denuit, M. and Frostig, E. (2006). Heterogeneity and the need for capital in the individual model. Scand. Actuarial J. 2006, 4266.10.1080/03461230500518690CrossRefGoogle Scholar
Dhaene, J. and Goovaerts, M. J. (1996). Dependency of risks and stop-loss order. ASTIN Bull. 26, 201212.10.2143/AST.26.2.563219CrossRefGoogle Scholar
Dhaene, J. and Goovaerts, M. J. (1997). On the dependency of risks in the individual life model. Insurance Math. Econom. 19, 243253.10.1016/S0167-6687(96)00015-7CrossRefGoogle Scholar
Ding, W., Yang, J. and Ling, X. (2017). On the skewness of extreme order statistics from heterogeneous samples. Commun. Statist. Theory Meth. 46, 23152331.10.1080/03610926.2015.1041984CrossRefGoogle Scholar
Fernández-Ponce, J. M., Kochar, S. C. and Muñoz-Perez, J. (1998). Partial orderings of distributions based on right-spread functions. J. Appl. Prob. 35, 221228.10.1239/jap/1032192565CrossRefGoogle Scholar
Frostig, E. (2001). A comparison between homogeneous and heterogeneous portfolios. Insurance Math. Econom. 29, 5971.10.1016/S0167-6687(01)00073-7CrossRefGoogle Scholar
Frostig, E. (2006). On risk dependence and MRL ordering. Statist. Prob. Lett. 76, 231243.10.1016/j.spl.2005.07.012CrossRefGoogle Scholar
Hu, T. and Ruan, L. (2004). A note on multivariate stochastic comparisons of Bernoulli random variable. J. Statist. Planning Infer. 126, 281288.10.1016/j.jspi.2003.07.012CrossRefGoogle Scholar
Hu, T. and Wu, Z. (1999). On dependence of risks and stop-loss premiums. Insurance Math. Econom. 24, 323332.10.1016/S0167-6687(99)00007-4CrossRefGoogle Scholar
Khaledi, B. E. and Ahmadi, S. S. (2008). On stochastic comparison between aggregate claim amounts. J. Statist. Planning Infer. 138, 22432251.10.1016/j.jspi.2007.10.018CrossRefGoogle Scholar
Khaledi, B. E., Farsinezhad, S. and Kochar, S. C. (2011). Stochastic comparisons of order statistics in the scale models. J. Statist. Planning Infer. 141, 276286.10.1016/j.jspi.2010.06.006CrossRefGoogle Scholar
Kochar, S. C., Li, X. and Shaked, M. (2002). The total time on test transform and the excess wealth stochastic orders of distributions. Adv. Appl. Prob. 34, 826845.10.1239/aap/1037990955CrossRefGoogle Scholar
Kochar, S. C. and Xu, M. (2009). Comparisons of parallel systems according to the convex transform order. J. Appl. Prob. 46, 342352.10.1017/S0021900200005490CrossRefGoogle Scholar
Kochar, S. C. and Xu, M. (2010). On the right spread order of convolutions of heterogeneous exponential random variables. J. Multivar. Anal. 101, 165176.10.1016/j.jmva.2009.07.001CrossRefGoogle Scholar
Kochar, S. C. and Xu, M. (2014). On the skewness of order statistics with applications. Ann. Operat. Res. 212, 127138.10.1007/s10479-012-1212-4CrossRefGoogle Scholar
Kremer, E. (1983). Distribution-free upper bounds on the premiums of the LCR and ECOMOR treaties. Insurance Math. Econom. 2, 209213.10.1016/0167-6687(83)90013-6CrossRefGoogle Scholar
Kremer, E. (1998). Largest claims reinsurance premiums under possible claims dependence. ASTIN Bull. 28, 257267.10.2143/AST.28.2.519069CrossRefGoogle Scholar
Ladoucette, S. A. and Teugels, J. L. (2006). Reinsurance of large claims. J. Comput. Appl. Math. 186, 163190.10.1016/j.cam.2005.03.069CrossRefGoogle Scholar
Ma, C. (2000). Convex orders for linear combinations of random variables. J. Statist. Planning Infer. 84, 1125.10.1016/S0378-3758(99)00143-3CrossRefGoogle Scholar
Marshall, A. W. and Olkin, I. (2007). Life Distributions. Springer, New York.Google Scholar
Marshall, A. W., Olkin, I. and Arnold, B. C. (2011). Inequalities: Theory of Majorization and Its Applications, 2nd edn. Springer, New York.10.1007/978-0-387-68276-1CrossRefGoogle Scholar
Mitrinović, D. S. (1970). Analytic Inequalities. Springer, Berlin.10.1007/978-3-642-99970-3CrossRefGoogle Scholar
Müller, A. and Stoyan, D. (2002). Comparison Methods for Stochastic Models and Risks. John Wiley, New York.Google Scholar
Saunders, I. W. and Moran, P. A. (1978). On the quantiles of the gamma and F distributions. J. Appl. Prob. 15, 426432.10.2307/3213414CrossRefGoogle Scholar
Shaked, M. and Shanthikumar, J. G. (2007). Stochastic Orders. Springer, New York.10.1007/978-0-387-34675-5CrossRefGoogle Scholar
Van Zwet, W. R. (1970). Convex Transformations of Random Variables (Math. Centre Tracts 7), 2nd edn. Mathematical Centre, Amsterdam.Google Scholar
Wang, Y. (1993). On the number of successes in independent trials. Statistica Sinica 3, 295312.Google Scholar
Xu, M. and Balakrishnan, N. (2011). On the convolution of heterogeneous Bernoulli random variables. J. Appl. Prob. 48, 877884.10.1239/jap/1316796922CrossRefGoogle Scholar
Zhang, Y. and Zhao, P. (2015). Comparisons on aggregate risks from two sets of heterogeneous portfolios. Insurance Math. Econom. 65, 124135.10.1016/j.insmatheco.2015.09.004CrossRefGoogle Scholar
Zhang, Y., Amini-Seresht, E. and Zhao, P. (2019a). On fail-safe systems under random shocks. Appl. Stoch. Models Business Industry 35, 591602.10.1002/asmb.2349CrossRefGoogle Scholar
Zhang, Y., Cai, X. and Zhao, P. (2019b). Ordering properties of extreme claim amounts from heterogeneous portfolios. ASTIN Bull. 49, 525554.10.1017/asb.2019.7CrossRefGoogle Scholar
Zhang, Y., Zhao, P. and Cheung, K. C. (2019c). Comparisons on aggregate claim numbers and amount: A study of heterogeneity. Scand. Actuarial J. 2019, 273290.10.1080/03461238.2018.1557738CrossRefGoogle Scholar
Zhang, Y., Ding, W. and Zhao, P. (2020). On variability of series and parallel systems with heterogeneous components. Prob. Eng. Inf. Sci. 34, 626645.10.1017/S0269964819000263CrossRefGoogle Scholar
Figure 0

Figure 1. Plot of the lower bound of ${\rm Var}[Y_{3\;:\;3}]$ in Example 1 with respect to $p\in[0.7116,1]$.

Figure 1

Figure 2. Plots of coefficients of variation of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$ under different settings.

Figure 2

Figure 3. Plots of coefficients of variation of $Y_{n\;:\;n}$ and $Y_{n\;:\;n}^{\ast}$.