In this paper, we investigate some multivariate integral stochastic orderings of skew-normal random vectors. We derive the results of the sufficient and/or necessary conditions by applying an identity for Ef(Y)−Ef(X), where X and Y are multivariate skew-normal random vectors, f satisfies some weak regularity condition. The integral orders considered here are the componentwise convex, copositive, completely-positive orderings and their corresponding increasing ones as well as linear forms of stochastic orderings, which play a vital role in transforming the unmanageable multivariate components into an easy-to-handle univariate variable.
Citation: Xueyan Li, Chuancun Yin. Some stochastic orderings of multivariate skew-normal random vectors[J]. AIMS Mathematics, 2023, 8(10): 23427-23441. doi: 10.3934/math.20231190
[1] | Xueying Yu, Chuancun Yin . Some results on multivariate measures of elliptical and skew-elliptical distributions: higher-order moments, skewness and kurtosis. AIMS Mathematics, 2023, 8(3): 7346-7376. doi: 10.3934/math.2023370 |
[2] | Lijie Zhou, Liucang Wu, Bin Yang . Estimation and diagnostic for single-index partially functional linear regression model with $ p $-order autoregressive skew-normal errors. AIMS Mathematics, 2025, 10(3): 7022-7066. doi: 10.3934/math.2025321 |
[3] | Baishuai Zuo, Chuancun Yin . Stein’s lemma for truncated generalized skew-elliptical random vectors. AIMS Mathematics, 2020, 5(4): 3423-3433. doi: 10.3934/math.2020221 |
[4] | Dongjian Bai, Feng Wang . New methods based $ \mathcal{H} $-tensors for identifying the positive definiteness of multivariate homogeneous forms. AIMS Mathematics, 2021, 6(9): 10281-10295. doi: 10.3934/math.2021595 |
[5] | Waqar Afzal, Sayed M. Eldin, Waqas Nazeer, Ahmed M. Galal . Some integral inequalities for harmonical $ cr $-$ h $-Godunova-Levin stochastic processes. AIMS Mathematics, 2023, 8(6): 13473-13491. doi: 10.3934/math.2023683 |
[6] | Enrique de Amo, José Juan Quesada-Molina, Manuel Úbeda-Flores . Total positivity and dependence of order statistics. AIMS Mathematics, 2023, 8(12): 30717-30730. doi: 10.3934/math.20231570 |
[7] | Li Zhang, Rongfang Yan . Stochastic comparisons of series and parallel systems with dependent and heterogeneous Topp-Leone generated components. AIMS Mathematics, 2021, 6(3): 2031-2047. doi: 10.3934/math.2021124 |
[8] | Guangshuai Zhou, Chuancun Yin . Family of extended mean mixtures of multivariate normal distributions: Properties, inference and applications. AIMS Mathematics, 2022, 7(7): 12390-12414. doi: 10.3934/math.2022688 |
[9] | Xiao Zhang, Rongfang Yan . Stochastic comparisons of extreme order statistic from dependent and heterogeneous lower-truncated Weibull variables under Archimedean copula. AIMS Mathematics, 2022, 7(4): 6852-6875. doi: 10.3934/math.2022381 |
[10] | Taekyun Kim, Dae San Kim, Hye Kyung Kim . Some identities involving degenerate Stirling numbers arising from normal ordering. AIMS Mathematics, 2022, 7(9): 17357-17368. doi: 10.3934/math.2022956 |
In this paper, we investigate some multivariate integral stochastic orderings of skew-normal random vectors. We derive the results of the sufficient and/or necessary conditions by applying an identity for Ef(Y)−Ef(X), where X and Y are multivariate skew-normal random vectors, f satisfies some weak regularity condition. The integral orders considered here are the componentwise convex, copositive, completely-positive orderings and their corresponding increasing ones as well as linear forms of stochastic orderings, which play a vital role in transforming the unmanageable multivariate components into an easy-to-handle univariate variable.
In recent years, more and more scholars have paid extensive attention to the research of the theory of stochastic orderings and have achieved a lot of research results. Stochastic orderings are the class of partial order relationships, which are defined on a family of random variables. It is a powerful tool to describe the size relationship between random variables and compare the degree of correlation of random variables. Nowadays, stochastic order relationships are widely used in many fields of probability and statistics, such as the statistical theory of economics and actuarial data, the comparison of stochastic processes in physics and other disciplines, etc. The comparison of two or more ordered experimental groups based on multivariate data is commonly used in research and in various applications in the medical field. For more detailed theoretical results, interested readers can refer to [1,2,3].
The problem of the stochastic order of the multivariate normal distribution has been fully researched and described in detail by Müller [4] and Arlotto and Scarsini [5] and so on. The stochastic orderings of the multivariate elliptical distribution were later introduced by [6,7,8] and later the results of the multivariate normal distribution were extended to the general multivariate elliptical distribution with the special cases of multivariate normal, multivariate logistic, multivariate-t and multivariate Laplace distributions by Yin [9]. Although these two distributions have good properties and characteristics, they are idealized symmetrical distributions. However, in practice, data often have skewness and heavy tails. In particular, it was fitted by using these two distributions but cannot achieve a more perfect effect. Therefore, researchers have promoted the elliptical distribution, either by mixing the two methods or by adding skewness to the distribution then obtained a series of asymmetric distributions, which can better fit the actual data.
Azzalini [10] proposed the skew-normal distribution and then Azzalini and Valle [11] extended it to the multivariate skew-normal families. There are some properties and applications of the multivariate skew-normal distribution were discussed in [12]. This distribution represents a mathematically tractable extension of the density of multivariate normal distribution, with adding parameters to adjust for skewness. The multivariate skew-normal distribution offers reasonable flexibility in fitting real data while maintaining some convenient formal properties of the normal density. A complete exposition of the theory of the skew-normal distribution can be found in [13].
Among the numerous stochastic orderings, Hessian, increasing Hessian orderings and linear orderings have been discussed quite extensively in the literature in recent years. For example, the results on skewness orderings on the multivariate skew-normal can be found in Arevalillo and Navarro [14]. Pu et al. [15] studied a class of multivariate generalized location-scale mixtures of elliptical distributions with respect to stochastic orderings. Amiri et al. [16], Amiri and Balakrishnan [17] considered the Hessian, increasing Hessian orderings and linear orderings, which is a practical tool for reducing dimension in multivariate stochastic comparisons. Amiri and Balakrishnan [17] established some stochastic comparison results for multivariate scale-shape mixture of skew-normal distributions by restricting the conditions of parameters. Thus, it was shown that there exists some equivalent correlation between stochastic orderings. Pertinent results of the comparisons of partial integral stochastic orderings of skew-normal distribution can be found in [18]. Similarly, related results can be found in [19], in which the orders for matrix variate skew-normal distribution were studied. Current research on integral stochastic orderings of skew-normal distributions is limited and some well-known stochastic orderings such as componentwise convex, copositive, completely-positive orderings and a variety of increasing orderings such as increasing supermodular, increasing directionally convex, increasing copositive, increasing completely-positive and increasing componentwise convex orderings as well as linear forms of stochastic orderings have not been well studied.
Whether the existence of sufficient and/or necessary conditions for those stochastic orderings is still an open question. This paper intends to solve these issues. The results can be also seen as supplements to the results in [18].
This paper is organized as follows: Section 2 contains an introduction to the skew-normal distribution and a review of knowledge about integral stochastic orderings. Section 3 derives necessary and/or sufficient conditions for stochastic orderings of the skew-normal distribution. In Section 4, necessary and sufficient conditions for the linear orderings of the skew-normal distribution are given. Section 5 gives some brief summaries.
We firstly review the concept and property of the multivariate skew-normal distribution which is introduced by [12]. A n-dimensional random vector Z has the multivariate skew-normal distribution, denoted as Z∼SNn(ξ,Ω,α), if its probability density function has the following form:
fZ(z)=2ϕn(z;ξ,Ω)Φ(α⊤ω−1(z−ξ)), |
where ξ is the mean vector, α is the skewness parameter, ξ,α∈Rn, Ω=[ωij] is a n×n covariance matrix which has full rank, denote ¯Ω=ω−1Ωω−1 is the corresponding correlation matrix, where ω=diag(ω11,...,ωnn)12. The following notations will be used throughout the paper: The cumulative distribution function (CDF) of the univariate standard normal distribution is denoted by Φ(⋅) and the probability density function (PDF) of the n-dimensional normal distribution is represented by ϕn(.;ξ,Ω).
The characteristic function of Z is (refer to [20]):
ΨZ(t)=2exp(iξt−12t⊤Ωt)Φ(iδ⊤t)=exp(iξt−12t⊤Ωt){1+iτ(δ⊤t)},t∈Rn, | (2.1) |
where
δ=(1+α⊤¯Ωα)−12ω¯Ωα,τ(u)=√2π∫u0exp(z22)dz. |
Then, Z has the multivariate skew-normal distribution, also denoted as Z∼SNn(ξ,Ω,δ). Specially, the standard skew-normal distribution Z0∼SNn(0,¯Ω,δ), which can be abbreviated as Z0∼SNn(¯Ω,δ). Its mean vector and covariance matrix can be expressed as
E(Z0)=√2πδ,Cov(Z0)=¯Ω−2πδδ⊤. | (2.2) |
Consider the univariate skew-normal distribution Z1∼SN1(ξ1,σ21,δ21). Then, Z1 has a stochastic representation of the form (see [13])
Z1d=ξ1+σ1X1∣{X2<αX1}, |
where the random variables X1 and X2 are independent standard normal random variables and
α=δ1σ1√1−(δ1σ1)2. |
Lemma 2.1. ([18]) Suppose that X and Y are n-dimensional skew-normal random vectors
X∼SNn(ξ,Ω,δ),Y∼SNn(ξ∗,Ω∗,δ∗). | (2.3) |
Define the density function φλ(⋅) be distributed as
SNn(λξ∗+(1−λ)ξ, λΩ∗+(1−λ)Ω, λδ∗+(1−λ)δ,0≤λ≤1, |
where
ξλ=λξ∗+(1−λ)ξ, Ωλ=λΩ∗+(1−λ)Ω, δλ=λδ∗+(1−λ)δ. |
Moreover, define φkλ(.), k=1,2 by
φ1λ(u)=φλ(u),φ2λ(u)=ϕn(u;ξλ,Ωλ−δλδ⊤λ). |
Assume that f: Rn→R is a twice continuously differentiable function satisfying
(1) limxj→±∞f(x)φkλ(x)=0,
(2) limxj→±∞f(x)∂∂xiφkλ(x)=0,
(3) limxj→±∞φkλ(x)∂∂xif(x)=0,
where 0≤λ≤1, x∈Rn, i,j=1,⋯,n, k=1,2. Then,
E(f(Y)−f(X))=∫10∫Rn{((ξ∗−ξ)⊤∇f(x)+12tr((Ω∗−Ω)Hf(x)))φ1λ(x)+2√2π(δ∗−δ)⊤∇f(x)φ2λ(x)}dxdλ. | (2.4) |
Corollary 2.1. Suppose that f, X and Y satisfy the conditions of Lemma 2.1, such that for x∈Rn,
(1) Σni=1(ξ∗i−ξi)∂∂xif(x)≥0.
(2) Σni=1(δ∗i−δi)∂∂xif(x)≥0.
(3) Σni,j=1(ω∗ij−ωij)∂2∂xi∂xjf(x)≥0.
Then, E(f(X))≤E(f(Y)).
In the sequel, we introduce the concept of stochastic orderings. Integral stochastic ordering is the class of stochastic orderings that can be characterized by comparing the expectations of the random vectors X and Y. Let F is the class of measurable function f: Rn→R. If f satisfies Ef(Y)≥Ef(X) for two random vectors X and Y whose expectations are assumed to exist and for ∀f∈F where F is a measurable mapping set, then it is called the integral stochastic ordering X≤FY.
A function is supermodular if and only if its Hessian matrix has non-negative off-diagonal elements. f is increasing supermodular if and only if for all x∈Rn, there is ∇f(x)≥0 and ∂2f(x)∂xi∂xj≥0, for i≠j; f is increasing directionally convex if and only if for all x∈Rn, there is ∇f(x)≥0 and ∂2f(x)∂xi∂xj≥0, for 1≤i,j≤n.
For the notions mentioned above, if f: Rn→R is a twice continuously differentiable function then we write
∇f(x)=(∂∂x1f(x),...,∂∂xnf(x))⊤, Hf(x)=(∂2f(x)∂xi∂xj)n×n |
as the gradient vector and the Hessian matrix of f, respectively. For the n-tuple vectors a and b, we use the notations a≤b when ai≤bi, and a≥ 0 when ai≥ 0, for i=1,2,⋯,n.
Definition 2.1. If a n×n matrix A has quadratic form x⊤Ax≥0 for all x≥0 then A is said to be copositive. If there is a non-negative matrix Bm×n such that A=B⊤B then A is said to be completely positive.
Let S be the space of n×n-dimensional symmetric matrices and H be a closed convex cone in S, which satisfies the inner product ⟨A,B⟩=tr(AB), for A,B∈S. Then, we can define the function class as
FH={f:Rn→R:Hf(x)∈H,∀x∈Rn} |
and the class of increasing functions as
L={f:Rn→R:∇f(x)≥0,∀x∈Rn}. |
Let LH=FH∩L.
Definition 2.2. If there is λx∈C for x∈C,∀λ≥0 then a subset C of vector space V is called a cone. The cone C is convex if and only if αx+βy∈C, for ∀x,y∈C,α,β≥0. Besides, if C is closed under the inner product ⟨⋅,⋅⟩ then
C∗={y∈V:⟨x,y⟩≥0,∀x∈C} |
is called the dual of C. If C=C∗ is satisfied, then C is called self-dual.
We use Ccop to denote the cone of a copositive matrix and Ccp to denote the cone of a completely positive matrix. Let C∗cop and C∗cp are the duals of Ccop and Ccp, respectively. Cpsd and C+ are denoted as the cones of positive semi-definite matrix and non-negative matrix, respectively. C+off and C+diag are ordered as the cones of non-negative off-diagonal elements matrix and non-negative main diagonal elements matrix, respectively.
Lemma 2.2. ([5,21,22]) The cones Ccop and Ccp are closed and convex, and Ccp=C∗cop, Ccop=C∗cp. Cpsd and C+ are also closed and convex and self-dual Cpsd=C∗psd, C+=C∗+. C+off and C+diag are closed and convex and their dual cones are
C∗+off={B∈S:bii=0,bij≥0,i≠j∈{1,...,n}}, |
C∗+diag={B∈S:bii≥0,bij=0,i≠j∈{1,...,n}}. |
Let f: Rn→R be a function with twice continuously derivative. Then, f is convex if and only if Hf∈Cpsd, f is directionally convex if and only if it satisfies Hf∈C+, f is supermodular if and only if it satisfies Hf∈C+off, f is componentwise convex if and only if it satisfies Hf∈C+diag.
We introduce some important stochastic orderings as following:
Definition 2.3. ([5]) If f∈F=FH then this kind of stochastic orderings is called Hessian orderings. If f∈LH then this kind of stochastic orderings is called increasing Hessian orderings.
If for a∈Rn and a scalar convex function ψ, the function f: Rn→R satisfies the condition f(x)=ψ(a⊤X) then f is said to be linear-convex. If for a∈Rn+ and a scalar convex function ψ, the function f: Rn→R satisfies the condition f(x)=ψ(a⊤X) then f is said to be positive-linear-convex.
Definition 2.4. Suppose two random variables X and Y.
(1) Usual random order: If Ef(Y)≥Ef(X) holds true for any increasing function f then X≤stY.
(2) Convex order: If Ef(Y)≥Ef(X) holds true for any convex function f then X≤cxY.
(3) Componentwise convex order: If F is a class of twice differentiable functions f: Rn→R satisfying ∂2∂x2if(x)≥0, where x∈Rn and 1≤i≤n then X≤ccxY.
(4) Completely positive order: If F is a class of functions f that satisfies the condition Hf(x)∈Ccp then X≤cpY.
(5) Copositive order: If F is a class of functions f satisfying the condition Hf(x)∈Ccop then X≤copY.
(6) Increasing convex order: If Ef(Y)≥Ef(X) holds true for any increasing convex function f then X≤icxY.
(7) Increasing supermodular order: If F is a class of twice differentiable functions f: Rn→R satisfying ∇f(x)≥0, for all x∈Rn, and ∂2f(x)∂xi∂xj≥0, for x∈Rn, 1≤i<j≤n then X≤ismY.
(8) Increasing directionally convex order: If F is a class of twice differentiable functions f: Rn→R, satisfying ∇f(x)≥0, for all x∈Rn and ∂2f(x)∂xi∂xj≥0, for x∈Rn, 1≤i,j≤n then X≤idcxY
(9) Increasing componentwise convex order: If F is a class of twice differentiable functions f: Rn→R, satisfying ∇f(x)≥0 and ∂2∂x2if(x)≥0, for x∈Rn, 1≤i≤n then X≤iccxY.
(10) Increasing copositive: If F is a class of increasing functions f satisfying condition Hf(x)∈Ccop then X≤icopY.
(11) Increasing completely-positive: If F is a class of increasing functions f satisfying condition Hf(x)∈Ccp then X≤icpY.
Then we introduce the definition of several linear stochastic orderings.
(a) If a⊤X≤sta⊤Y holds true for any a∈Rn then X is said to be less than Y in the sense of linear-usual stochastic order, which is denoted as X≤lstY.
(b) If a⊤X≤sta⊤Y holds true for any a∈Rn+ then X is said to be less than Y in the sense of positive-linear-usual stochastic order, which is denoted as X≤plstY.
(c) If a⊤X≤cxa⊤Y holds true for any a∈Rn then X is said to be less than Y in the sense of linear-convex order, which is denoted as X≤lcxY.
(d) If a⊤X≤cxa⊤Y holds true for any a∈Rn+ then X is said to be less than Y in the sense of positive-linear-convex order, which is denoted as X≤plcxY.
(e) If a⊤X≤icxa⊤Y holds true for any a∈Rn+ then X is said to be less than Y in the sense of increasing-positive-linear-convex order, which is denoted as X≤plcxY.
This section establishes the sufficient and/or necessary conditions for the stochastic comparison between two random variables that obey the skew-normal distributions. The proofs of sufficient conditions of the stochastic comparison are fully used the Lemma 2.1 and the identity (2.4). When proving the necessary conditions, various methods are used to compare the parameters under the premise of considering the property of the stochastic orderings.
Lemma 3.1. Suppose that a n-dimensional random vector X∼SNn(ξ,Ω,δ). Then,
E(X)=ξ+√2πδ. |
If the second order moment exists,
Cov(X)=Ω+ξξ⊤+√2π(ξδ⊤+δξ⊤). |
Theorem 3.1. Suppose that the random variables X and Y are defined as in (2.3).
(1) If ξ=ξ∗,δ=δ∗,ωii≤ω∗ii,1≤i≤n and ωij=ω∗ij,1≤i<j≤n then X≤ccxY.
(2) If ξ=ξ∗ then X≤ccxY if and only if δ=δ∗,ωii≤ω∗ii,1≤i≤n and ωij=ω∗ij,1≤i<j≤n.
(3) If δ=δ∗ then X≤ccxY if and only if ξ=ξ∗,ωii≤ω∗ii,1≤i≤n and ωij=ω∗ij,1≤i<j≤n.
Proof. (1) If a convex function f is twice derivable, and Hf(x)∈C+diag, then X≤ccxY by substituting the condition ξ=ξ∗,δ=δ∗,ωii≤ω∗ii,1≤i≤n, ωij=ω∗ij and 1≤i<j≤n into Corollary 2.1.
(2)-(3) The sufficiency can be directly obtained through Corollary 2.1. The necessity is proved below. Let X≤ccxY, considering a componentwise convex function which satisfies the Definition 2.4. Suppose that
f1(x)=xi, f2(x)=−xi, 1≤i≤n. |
Combining with E(f(X))≤E(f(Y)), it quickly yields E(f(X))=E(f(Y)). Considering the condition ξ=ξ∗ and using Lemma 3.1, we have δ=δ∗ immediately. The same is true for the proof of (3). Consider the functions
f3(x)=xixj, f4(x)=−xixj, f5(x)=x2i, |
where
1≤i≤n, 1≤i<j≤n. |
Obviously, they all satisfy the definition of componentwise convex functions in Definition 2.4. Therefore, we have
E(XiXj)≤E(YiYj), Cov(Xi)≤Cov(Yi). |
Then, by combining with the expression about the second moment in Lemma 3.1, we can get the conclusion ωii≤ω∗ii,1≤i≤n, ωij=ω∗ij and 1≤i<j≤n. This ends the proof of Theorem 3.1.
By considering the n-dimensional random variables X0 and Y0 which obey the standardized skew-normal distributions, we get the following theorem. The distributions of X0 and Y0 are respectively denoted as
X0∼SNn(¯Ω,δ),Y0∼SNn(¯Ω∗,δ∗). | (3.1) |
Theorem 3.2. Suppose that the random variables X0 and Y0 are defined as in (3.1). Then X0≤cpY0 if and only if δ=δ∗ and ¯Ω∗−¯Ω is copositive.
Proof. Sufficiency. Looking at the function f which satisfies the twice differentiable condition and Hf(x)∈Ccp, we take into account δ=δ∗ and ¯Ω∗−¯Ω in Corollary 2.1, then we derive E(f(X0))≤E(f(Y0)), which means X0≤cpY0.
Necessity. Take the functions
f1(x)=xi, f2(x)=−xi, 1≤i≤n, |
f1 and f2 satisfy obviously Hf(x)∈Ccp of the completely positive function in Definition 2.4. Considering the condition of X0≤cpY0 and combining with condition E(f(X0))≤E(f(Y0)), we deduce that their means are equal. Then, from the formula (2.2) we see that δ=δ∗. Let the function
f3(x)=12(x−E(X))⊤A(x−E(X)), |
where A is any n×n dimensional symmetric matrix and A∈Ccp, it is clear that
Hf3(x)=A∈Ccp, |
for ∀x∈Rn. In order to prove that ¯Ω∗−¯Ω is copositive, that is
E((X0−E(X0))⊤A(X0−E(X0)))≤E((Y0−E(Y0))⊤A(Y0−E(Y0))). |
Combining with the definition of standardized covariance in formula (2.2), it can be deduced that
tr[(¯Ω−2πδδ⊤)A]≤tr[(¯Ω∗−2πδ∗δ∗⊤)A]. |
Also, we conclude
tr[(¯Ω∗−¯Ω)A]≥0 |
by considering δ=δ∗. And because of A∈Ccp, ¯Ω∗−¯Ω∈C∗cp where C∗cp=Ccop. Consequently, we conclude that ¯Ω∗−¯Ω is copositive. This completes the proof of Theorem 3.2.
The following theorem gives the condition of copositive order for the multivariate skew-normal distribution. The result (1) proves the sufficient condition for the general case, and the necessary condition for the standard skew-normal distribution is proved in (2).
Theorem 3.3. Suppose that the random variables are defined as in (2.3) and (3.1).
(1) If ξ=ξ∗,δ=δ∗ and Ω∗−Ω is completely positive then X≤copY.
(2) If X0≤copY0 then δ=δ∗ and ¯Ω∗−¯Ω is completely positive.
Proof. (1) Combining known conditions with Corollary 2.1, it can be obtained immediately.
(2) Suppose that X0≤copY0 holds. We take copositive functions
f1(x)=xi, f2(x)=−xi, 1≤i≤n, |
where f1 and f2 satisfy obviously Hf(x)∈Ccop in Definition 2.4. Combining condition E(f(X0))≤E(f(Y0)), it can be deduced that their means are equal. Then, from (2.2) we can see that δ=δ∗. Take the function
f3(x)=12(x−E(X))⊤A(x−E(X)), |
where A is any n×n dimensional symmetric matrix and A∈Ccop, it is clear that
Hf3(x)=A∈Ccop |
for ∀x∈Rn. In order to prove that ¯Ω∗−¯Ω is completely positive, that is
E((X0−E(X0))⊤A(X0−E(X0)))≤E((Y0−E(Y0))⊤A(Y0−E(Y0))). |
Combining with the definition of standardized covariance in formula (2.2), it can be deduced that
tr[(¯Ω−2πδδ⊤)A]≤tr[(¯Ω∗−2πδ∗δ∗⊤)A] |
from which we conclude tr[(¯Ω∗−¯Ω)A]≥0 since δ=δ∗. And because of A∈Ccop, ¯Ω∗−¯Ω∈C∗cop where C∗cop=Ccp, Consequently, we conclude that ¯Ω∗−¯Ω is completely positive. This completes the proof of Theorem 3.3.
The following theorem introduces several stochastic orderings of univariate skew-normal distribution, which will be used in the some theorem proof. Suppose that the univariate random variables X1 and Y1 have univariate skew-normal distributions, denoted as
X1∼SN1(ξ1,σ21,δ1),Y1∼SN1(ξ2,σ22,δ2). | (3.2) |
Lemma 3.2. ([18]) Suppose that the random variables are defined as in (3.2). Then, X1≤stY1 if and only if ξ1≤ξ2,σ1=σ22,δ1≤δ2.
Lemma 3.3. Suppose that the random variables are defined as in (3.2).
(1) If ξ1≤ξ2,σ1≤σ2,δ1≤δ2 then X1≤icxY1.
(2) If X1≤icxY1 and ξ1=ξ2 then σ1≤σ2,δ1≤δ2.
Proof. (1) It is an immediate consequence of Corollary 2.1.
(2) Suppose that X1≤icxY1 and consider the increasing-convex function f(x)=xi, i=1,2. Then, we claim E(X1)≤E(Y1) because of E(f(Y)−f(X))≥0. By using Lemma 3.1 and the condition ξ1=ξ2, we have δ1≤δ2. Also, we know
X1∼SN1(ξ1,σ21,δ1), Y1∼SN1(ξ2,σ22,δ2). |
We claim that σ1≤σ2. Suppose that σ1>σ2. Then,
limt→+∞E(Y1−t)+E(X1−t)+=limt→+∞∫+∞t(1−FY1(x))dx∫+∞t(1−FX1(x))dx=limt→+∞FY1(t)−1FX1(t)−1=limt→+∞fY1(t)fX1(t)=0, | (3.3) |
where,
fX1(t)=2σ1ϕ((t−ξ1)/σ1)Φ(α1(t−ξ1)/σ1), |
fY1(t)=2σ2ϕ((t−ξ2)/σ2)Φ(α2(t−ξ2)/σ2). |
This contradicts X1≤icxY1. Therefore, σ1≤σ2.
Lemma 3.4. Suppose that the random variables are defined as in (3.2).
(1) If ξ1=ξ2,σ1≤σ2,δ1=δ2 then X1≤cxY1.
(2) If X1≤cxY1 and ξ1=ξ2 then σ1≤σ2,δ1=δ2.
Proof. (1) Take the convex function f and substitute the known conditions into Corollary 2.1, which we can get the result immediately.
(2) It is well known that X1≤cxY1 if and only if X1≤icxY1 and E(X1)=E(Y1), which combines with the mean formula in Lemma 3.1 and ξ1=ξ2, we get δ1=δ2. And we claim that σ1≤σ2 by using Lemma 3.3.
Theorem 3.4. Suppose that the random variables are defined as in (2.3) and (3.1).
(1) If ξ≤ξ∗, δ≤δ∗ and Ω∗−Ω is copositive then X≤icpY.
(2) If X0≤icpY0 then δ≤δ∗ and ¯Ω∗−¯Ω is copositive.
Proof. (1) For any f∈F where F is the class of increasing functions f which satisfy Hf(x)∈Ccp. Considering ξ≤ξ∗, δ≤δ∗ and Ω∗−Ω is copositive with Corollary 2.1, it is quickly implies that Ef(Y)≥Ef(X). Then, we draw the conclusion X≤icpY.
(2) Suppose that X0≤icpY0. We take the twice differentiable increasingly function f(x)=xi, satisfying Hf(x)∈Ccp. Then, from the formula
E(f(Y)−f(X))≥0 |
we can know that
E(f(X0))≤E(f(Y0)). |
Therefore, we get δ≤δ∗ by combining with the (2.2). Let
f(x)=g(a⊤x), a∈Rn+, |
where g is an increasing-convex function. Thus, f∈LCcp. To sum up, we imply
E(g(a⊤X0))≤E(g(a⊤Y0)), |
i.e., a⊤X0≤icxa⊤Y0, where
a⊤X0∼SN1(a⊤¯Ωa, a⊤δ), a⊤Y0∼SN1(a⊤¯Ω∗a, a⊤δ∗). |
Then, from the conclusion of Lemma 3.2, we get that
a⊤(¯Ω∗−¯Ω)a≥0, |
i.e., ¯Ω∗−¯Ω is copositive.
Theorem 3.5. Suppose that the random variables are defined as in (2.3) and (3.1).
(1) If ξ≤ξ∗, δ≤δ∗ and Ω∗−Ω is completely positive then X≤icopY.
(2) If X0≤icopY0 then δ≤δ∗ and ¯Ω∗−¯Ω is completely positive.
Proof. (1) Obviously, considering any increasing function f∈F, Hf(x)∈Ccop of f and Corollary 2.1 then Ef(Y) ≥Ef(X) is immediately available from ξ≤ξ∗, δ≤δ∗ and Ω∗−Ω is completely positive. Therefore, X≤icopY.
(2) According to the Definition 2.4, we take the twice differentiable increasingly function f(x)=xi satisfying Hf(x)∈Ccop. Then, the formula
E(f(Y)−f(X))≥0 |
shows
E(f(X0))≤E(f(Y0)). |
Combining with formula (2.2), we can derive that δ≤δ∗. Consider
f(x)=g(a⊤x), a∈Rn+, |
where g is an increasing-convex function, such that f is also an increasing-convex function, it yields that
Hf(x)=a⊤ag(2)(a⊤x). |
Note that g(2)(a⊤x)≥0, because of the convexity of g, so Hf(x)≥0, i.e., f∈LCcop. Assume that X0≤icopY0 then
E(g(a⊤X0))≤E(g(a⊤Y0)), |
i.e., a⊤X0≤icxa⊤Y0, where
a⊤X0∼SN1(a⊤¯Ωa, a⊤δ), a⊤Y0∼SN1(a⊤¯Ω∗a, a⊤δ∗). |
The conclusion of Lemma 3.2 shows that ¯Ω∗−¯Ω can be expressed by the product of any two non-negative matrices, that is to say ¯Ω∗−¯Ω is completely positive.
Theorem 3.6. Suppose that the random variables are defined as in (3.1), then X0≤ismY0 if and only if δ≤δ∗,ωii=ω∗ii,1≤i≤n, ωij≤ω∗ij and 1≤i<j≤n.
Proof. Sufficiency. According to Lemma 1 in [17], we conclude that X0≤LHY0 when δ≤δ∗, where random vectors X0 and Y0 are defined as (3.1). And H=C+off, its dual cone is
C∗+off={B∈S:bii=0,bij≥0,i≠j∈{1,...,n}}. |
Then, by the supermodular ordering in Definition 2.4, X0≤ismY0 can be obtained.
Necessity. Suppose that X0≤ismY0, then X0i≤stY0i, 1≤i≤n and X0≤iplcxY0 can be known from Müller and Stoyan [1]. First, X0i≤stY0i, 1≤i≤n guarantees the results of Lemma 3.2. Then, it is shown that δi≤δ∗i, ωii=ω∗ii and 1≤i≤n. Moreover, X0≤iplcxY0, which means
a⊤X0≤icxa⊤Y0, a∈Rn+. |
According to the conclusion given in Lemma 3.3, we can imply a⊤(¯Ω∗−¯Ω)a≥0, ¯Ω∗−¯Ω is copositive according to Definition 2.1. Combining the above derivation with the definition of the copositive matrix, it can be known that ¯Ω∗−¯Ω is a matrix whose diagonal elements are 0 and off-diagonal elements ≥0, which means that ωij≤ω∗ij and 1≤i<j≤n.
Theorem 3.7. Suppose that the random variables are defined as in (3.1). Then, X0−E(X0)≤idcxY0−E(Y0) if and only if ωij≤ω∗ij, 1≤i and j≤n.
Proof. Sufficiency. Let f∈F, where F be a class of twice differentiable and directionally convex functions. From Definition 2.4 we know that ∂2f(x)∂xi∂xj≥0. Then, according to the condition ωij≤ω∗ij,1≤i,j≤n and Corollary 2.1, we can get
X0−E(X0)≤idcxY0−E(Y0). |
Necessity. The condition
X0−E(X0)≤idcxY0−E(Y0) |
means
X0−E(X0)≤dcxY0−E(Y0). |
Then, according to the conclusion derived from Proposition 4.7 in [18], it can be known that ωij≤ω∗ij, 1≤i and j≤n.
Theorem 4.1. Suppose that the random variables are defined as in (2.3). Then, X≤plstY if and only if ξ≤ξ∗, δ≤δ∗ and Ω∗=Ω.
Proof. Sufficiency. According to the conclusion derived from Proposition 4.3 in [18], we can see that X≤stY from the known conditions ξ≤ξ∗, δ≤δ∗ and Ω∗=Ω. Then, we conclude X≤plstY.
Necessity. X≤plstY implies a⊤X≤sta⊤Y for a∈Rn+. From the necessary and sufficient conditions of the usual stochastic ordering of the skew-normal distribution, it can be known that a⊤ξ∗≥a⊤ξ, a⊤δ∗≥a⊤δ and a⊤(Ω∗−Ω)a=0,a≥0. The conclusion is obviously available.
Theorem 4.2. Suppose that the random variables are defined as in (2.3) and (3.1).
(1) If ξ=ξ∗,δ=δ∗,Ω∗−Ω is PSD then X≤lcxY.
(2) X0≤lcxY0 if and only if δ=δ∗,¯Ω∗−¯Ω≥0.
Proof. (1) From ξ=ξ∗ and δ=δ∗, Ω∗−Ω is positive semi-definite, we can conclude X≤cxY and then X≤lcxY.
(2) Because of X0≤lcxY0, we have a⊤X0≤cxa⊤Y0,a∈Rn+, where
a⊤X0∼SN1(a⊤¯Ωa,a⊤δ), a⊤Y0∼SN1(a⊤¯Ω∗a,a⊤δ∗). |
Combining with the known necessary and sufficient conditions of convex order, we directly get a⊤δ∗=a⊤δ and a⊤(¯Ω∗−¯Ω)a≥0,a∈Rn. That is, δ=δ∗,¯Ω∗−¯Ω≥0, which means ¯Ω∗−¯Ω is positive semi-definite.
Theorem 4.3. Suppose that the random variables are defined as in (2.3) and (3.1).
(1) If ξ=ξ∗,Ω∗−Ω is copositive and δ=δ∗ then X≤plcxY.
(2) If X0≤plcxY0 then δ=δ∗ and ¯Ω∗−¯Ω is copositive.
Proof. (1) By using the given conditions and Theorem 3.2, it is obvious that X≤cpY can be obtained. Consider a convex function g: Rn→R and let
f(x)=g(a⊤x), a∈Rn+, |
then f: Rn→R is also a convex function. According to completely positive ordering in Definition 2.4, the condition Ef(Y)≥Ef(X) is satisfied, that is,
E(g(a⊤X))≤E(g(a⊤Y)), |
so X≤plcxY is launched.
(2) Suppose that X0≤plcxY0, then a⊤X0≤cxa⊤Y0,a∈Rn+. Combining with the known conclusion of univariate convex order in Lemma 3.4, it is shown that a⊤X0≤cxa⊤Y0 is equivalent with a⊤δ∗=a⊤δ and a⊤(¯Ω∗−¯Ω)a is positive semi-definite. Then, we immediately conclude that δ=δ∗ and ¯Ω∗−¯Ω is copositive.
Theorem 4.4. Suppose that the random variables are defined as in (2.3) and (3.1).
(1) If ξ≤ξ∗, δ≤δ∗ and Ω∗−Ω is copositive then X≤iplcxY.
(2) If X0≤iplcxY0 then δ≤δ∗ and ¯Ω∗−¯Ω is copositive.
Proof. (1) Apply the conclusion derived in Theorem 3.4, if ξ≤ξ∗, δ≤δ∗ and Ω∗−Ω is copositive, then X≤icpY can be deduced. At this time, consider an increasing-convex function g: Rn→R and let f(x)=g(a⊤x),a∈Rn+. Then, f: Rn→R is also an increasing-convex function on Rn and f satisfies the expression of copositive ordering in Definition 2.4, such that ∂∂xif(x)≥0 and
Hf(x)=a⊤ag(2)(a⊤x)∈Ccp,a∈Rn+. |
Thus,
E(g(a⊤X))≤E(g(a⊤Y)), a∈Rn+, |
which means X≤iplcxY.
(2) When X0≤iplcxY0, there is a⊤X0≤icxa⊤Y0,a∈Rn+ where
a⊤X0∼SN1(a⊤¯Ωa,a⊤δ), a⊤Y0∼SN1(a⊤δ∗,a⊤¯Ω∗a). |
According to the conclusion of increasing-convex order in Lemma 3.3, we can get a⊤δ≤a⊤δ∗ and a⊤(¯Ω∗−¯Ω)a≥0,a≥0, i.e., δ≤δ∗, ¯Ω∗−¯Ω is copositive.
In this paper we consider the multivariate integral stochastic ordering of skew-normal distribution, including componentwise convex, copositive, completely positive orderings and increasing componentwise convex, increasing copositive, increasing completely positive, increasing directionally convex, increasing supermodular orderings, etc, as well as some important linear stochastic orderings. We obtain some necessary and/or sufficient conditions. As for future research directions, we will draw on the idea of a projection pursuit, which is a multivariate statistical method aimed at finding interesting data projections [23], in the studying of linear random orders. Furthermore, the results might be extended to other skew-symmetric distributions, such as the (generalized) skew-elliptical distributions, which are natural generalizations of skew-normal distributions [24].
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors express their sincere thanks to the editors and the anonymous reviewers for all their useful comments and suggestions on an earlier version of this manuscript which led to this improved version. The research was supported by the National Natural Science Foundation of China (No. 12071251).
All authors declare no conflicts of interest in this paper.
[1] | A. Müller, D. Stoyan, Comparison methods for stochastic models and risks, John Wiley & Sons, Inc., 2002. |
[2] | M. Denuit, J. Dhaene, M. Goovaerts, R. Kaas, Actuarial theory for dependent risks: measures, orders and models, John Wiley & Sons, Inc., 2005. https://doi.org/10.1002/0470016450 |
[3] | M. Shaked, J. G. Shanthikumar, Stochastic orders, Springer, 2007. https://doi.org/10.1007/978-0-387-34675-5 |
[4] |
A. Müller, Stochastic ordering of multivariate normal distributions, Ann. Inst. Stat. Math., 53 (2001), 567–575. http://doi.org/10.1023/A:1014629416504 doi: 10.1023/A:1014629416504
![]() |
[5] |
A. Arlotto, M. Scarsini, Hessian orders and multinormal distributions, J. Multivar. Anal., 100 (2009), 2324–2330. https://doi.org/10.1016/j.jmva.2009.03.009 doi: 10.1016/j.jmva.2009.03.009
![]() |
[6] |
Z. Landsman, A. Tsanakas, Stochastic ordering of bivariate elliptical distributions, Stat. Probab. Lett., 76 (2006), 488–494. https://doi.org/10.1016/j.spl.2005.08.016 doi: 10.1016/j.spl.2005.08.016
![]() |
[7] |
O. Davidov, S. Peddada, The linear stochastic order and directed inference for multivariate ordered distributions, Ann. Stat., 41 (2013), 1–40. https://doi.org/10.1214/12-AOS1062 doi: 10.1214/12-AOS1062
![]() |
[8] |
X. Pan, G. Qiu, T. Hu, Stochastic orderings for elliptical random vectors, J. Multivar. Anal., 148 (2016), 83–88. http://doi.org/10.1016/j.jmva.2016.02.016 doi: 10.1016/j.jmva.2016.02.016
![]() |
[9] |
C. C. Yin, Stochastic orderings of multivariate elliptical distributions, J. Appl. Probab., 58 (2021), 551–568. http://doi.org/10.1017/JPR.2020.104 doi: 10.1017/JPR.2020.104
![]() |
[10] |
A. Azzalini, A class of distributions which includes the normal ones, Scand. J. Stat., 46 (1986), 171–178. http://doi.org/10.6092/ISSN.1973-2201/711 doi: 10.6092/ISSN.1973-2201/711
![]() |
[11] |
A. Azzalini, A. D. Valle, The multivariate skew-normal distribution, Biometrika, 83 (1996), 715–726. https://doi.org/10.1093/biomet/83.4.715 doi: 10.1093/biomet/83.4.715
![]() |
[12] |
A. Azzalini, A. Capitanio, Statistical applications of the multivariate skew normal distribution, J. R. Stat. Soc. Ser. B, 61 (1999), 579–602. https://doi.org/10.1111/1467-9868.00194 doi: 10.1111/1467-9868.00194
![]() |
[13] | A. Azzalini, The skew-normal and related families, Cambridge University Press, 2014. https://doi.org/10.1017/CBO9781139248891 |
[14] |
J. M. Arevalillo, H. Navarro, A stochastic ordering based on the canonical transformation of skew-normal vectors, Test, 28 (2019), 475–498. https://doi.org/10.1007/s11749-018-0583-5 doi: 10.1007/s11749-018-0583-5
![]() |
[15] | T. Pu, Y. Y. Zhang, C. C. Yin, Generalized location-scale mixtures of elliptical distributions: definitions and stochastic comparisons, Commun. Stat., 2023. https://doi.org/10.1080/03610926.2023.2165407 |
[16] |
M. Amiri, S. Izadkhah, A. Jamalizadeh, Linear orderings of the scale mixtures of the multivariate skew-normal distribution, J. Multivar. Anal., 179 (2020), 104647. https://doi.org/10.1016/j.jmva.2020.104647 doi: 10.1016/j.jmva.2020.104647
![]() |
[17] |
M. Amiri, N. Balakrishnan, Hessian and increasing-Hessian orderings of scale-shape mixtures of multivariate skew-normal distributions and applications, J. Comput. Appl. Math., 402 (2022), 113801. https://doi.org/10.1016/J.CAM.2021.113801 doi: 10.1016/J.CAM.2021.113801
![]() |
[18] |
D. Jamali, M. Amiri, A. Jamalizadeh, Comparison of the multivariate skew-normal random vectors based on the integral stochastic ordering, Commun. Stat., 50 (2021), 5215–5227. http://doi.org/10.1080/03610926.2020.1740934 doi: 10.1080/03610926.2020.1740934
![]() |
[19] | T. Pu, N. Balakrishnan, C. C. Yin, An identity for expectations and characteristic function of matrix variate skew-normal distribution with applications to associated stochastic orderings, Commun. Math. Stat., 2022. https://doi.org/10.1007/s40304-021-00267-2 |
[20] |
H. M. Kim, M. G. Genton, Characteristic functions of scale mixtures of multivariate skew-normal distributions, J. Multivar. Anal., 102 (2011), 1105–1117. http://doi.org/10.1016/j.jmva.2011.03.004 doi: 10.1016/j.jmva.2011.03.004
![]() |
[21] | M. Hall, Combinatorial theory, John Wiley & Sons, Inc., 1988. https://doi.org/10.1002/9781118032862 |
[22] | Y. L. Tong, Stochastic orders and their applications, Academic Press, 1994. https://doi.org/10.1137/1037117 |
[23] | P. J. Huber, Projection pursuit, Ann. Stat., 13 (1985), 435–475. |
[24] |
N. Loperfido, A note on skew-elliptical distributions and linear functions of order statistics, Stat. Probab. Lett., 78 (2008), 3184–3186. http://doi.org/10.1016/j.spl.2008.06.004 doi: 10.1016/j.spl.2008.06.004
![]() |