This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Multimatricvariate and multimatrix variate distributions based on elliptically contoured laws under real normed division algebras

José A. Díaz-García
Universidad Autónoma de Chihuahua
Facultad de Zootecnia y Ecología
Periférico Francisco R. Almada Km 1, Zootecnia
33820 Chihuahua, Chihuahua, México
E-mail: jadiaz@uach.mx
Francisco J. Caro-Lopera
University of Medellin
Faculty of Basic Sciences
Carrera 87 No.30-65
Medellín, Colombia
E-mail: fjcaro@udemedellin.edu.co
Corresponding author
Key words. Multimatrix variate, real normed division algebras, matrix variate, multimatricvariate, random matrices, matrix variate elliptical distributions.
2000 Mathematical Subject Classification. 60E05; 62E15; 15A23; 15B52
Abstract

This paper proposes famillies of multimatricvariate and multimatrix variate distributions based on elliptically contoured laws in the context of real normed division algebras. The work allows to answer the following inference problems about random matrix variate distributions: 1) Modeling of two or more probabilistically dependent random variables in all possible combinations whether univariate, vector and matrix simultaneously. 2) Expected marginal distributions under independence and joint estimation of models under likelihood functions of dependent samples. 3) Definition of a likelihood function for dependent samples in the mentioned random dimensions and under real normed division algebras. The corresponding real distributions are alternative approaches to the existing univariate and vector variate copulas, with the additional advantages previously listed. An application for quaternionic algebra is illustrated by a computable dependent sample joint distribution for landmark data emerged from shape theory.

1 Introduction

A postmodern world of increasingly heuristic and interdisciplinary problems requires integrative solutions that do not always have the answer in the current frameworks of mathematics and statistics. Chaos theory as a paradigmatic element of phenomenal interrelation links more and more random variables. They are not only univariate and vector variables, but matrix variates, due to their great versatility in the description of variability and intrinsic multiple correlation. Likewise, the domain of the variables began to move from the real field to the remaining real normed division algebras, such as complex, quaternionic and octonionic. There the emerging models of multiple natural, exact and engineering sciences have their own scope and they are not restricted by the real field. Given multiple statistical models for marginal distributions in univariate and vector cases, the additional problem of parametric estimation arises from a set of samples provided by the scientific expert user of statistics. Then comes another challenge for statistics, which has historically been resolved by optimising likelihood functions as a joint law for sample distribution. The theory of copulas emerged as a possible solution to the problem of explaining a joint phenomenon of known interrelation. Countless link functions then appear for two variables, however, the intention of a dependent relationship is diluted when using the typical likelihood functions for estimating copula parameters from independent samples. The role of likelihood in the history of statistics is of such transcendence that it is a popular tool for estimating parameters; however, its profound simplicity based on sample independence is never debated. While copulas and similar theories represent the solution to the problem of dependence, the concept of likelihood function, defined as the product of the marginal densities, seems immutable and universal. Recently, the authors have begun the discussion of the role of classical likelihood functions in statistics, showing the differences in application in databases declared dependently probabilistic. Time series in particular are the source of the greatest discrepancy, given the historical effort in proposing models for variance and volatility. But maintaining the postulate of an estimate via likelihood functions on independent data is an obvious contradiction in the face of a temporal process that intrinsically is founded on dependency. A work that launched this new way of reframing likelihood functions appeared recently in Díaz-García et al. [7]. In a real data series, the discrepancy of likelihood function over independent time samples versus realistic likelihood function over dependent samples was observed. The study indicated that decisions diffused about the mean estimated from the database using independent likelihood were in the tails of the dependent expected likelihood distribution. The finding was possible thanks to the definition of the so-termed multivector variate distribution, a natural way of defining a likelihood function of dependent vector samples, further parameterised by a large class of elliptically contoured distributions that are isolated from the choices of popular models. Until then, the mutivector could be an approach to vector copulas, but the implemented theory is designed to address the matrix case, which is still an open problem for copula theory. Then, the matrix version appeared in the form of the so-termed classes of multimatricvariate distributions, which provided the likelihood function or joint matrix distributions with a family of elliptically contoured distributions with a variety of kurtosis and symmetries, see Díaz-García and Caro-Lopera [5]. The choice of distributions that are invariant under the class of elliptically contoured laws within the family of multimatricvariate and multimatrix variate distributions gives robustness to the analysis since it avoids the difficult paradigm of adjusting the link function, if we speak in parallel to the procedure. in vector copulas.

Multimatricvariate distributions specialise in distributions based on the determinant. Still to be defined a class of multiple distributions of random matrices that depended on the trace, the other function that has historically governed the theory of matrix distributions. Recently, the so-termeded multimatrix distributions appear in Díaz-García and Caro-Lopera [6]. Like the multimatricvariate distributions, they maintain the philosophy of probabilistic dependence, computability and a wide class of underlying distributions.

When we surpass the level of random vectors, where the most popular statistical techniques remain, such as copulas, we find immense difficulty for the calculation. We must observe the domain of the matrices, their transformations and Jacobians, and therefore their integration. It is generally difficult because it involves averaging over orthogonal groups, cones and hypercubes. The central, isotropic, non-central and non-isotropic cases also emerge. These integrations comprises the geometric filtering that comes for their definitions and factorisations.

At this point we find that the new multimatricvariate and multimatrix variate distributions on the real field satisfy the first 2 of the 3 conditions that we establish as a paradigm for a robust point of view of joint distributions, namely: 1) modeling of several dependent matrices capable of being combined with all univariate and/or vector variables governed by elliptically contoured models; 2) with strictly verifiable marginals under the trivial case of independence and testable from the dependent sample likelihood; and 3) founded on the same algebraic principles that allow them to be applied indistinctly from the real normed division algebras.

We now address the third desirable characteristic: its versatility in the four unique real normed division algebras.

Until a couple of decades ago it would be unthinkable to unify the theory of real random matrices with complex ones. History shows us that the theory of distributions of the central cases overflowed into an immense effort to characterize the indispensable elements of random matrices, namely: After its establishment, works on the complex case began to appear very slowly and completely conceptually distant from their real analogues. The so-termed statistical theory under real normed division algebras is to statistics what the expected theory of unification of forces is to physics. Its simplification is surprising in that all its results are reduced to modifying a beta parameter that goes from 1 (real) to 2 (complex) to 4 (quaternionic) to 8 (octonionic), see Díaz-García [4] were the transitions to complex, quaternionic and octonion are reached by changing the support group from orthogonal to unitary, compact symplectic or exceptional type.

If the real multimatricvariate distributions [5] and multimatrix variate distributions [6] articles are parallel compared with the results of the present work, an apparent repetition shall appear by its simplicity; Its notation makes the beta parameter open one of the four universes by simply changing the value, although understanding the profound difference that underpins them requires an extensive and difficult literature that began in other areas far from statistics. The non commutativity for quaternions and non associative for octonions promotes a deep research for some applications in those algebras.

We place the above discussion into the setting of two main problems that we shall address in this article.

The interest in multimatricvariate and multimatrix variate distributions has been motivated by the following two situations:

  1. 1.

    In different areas of knowledge (such as Finance and Hydrology, among others), people are interested in simultaneously modeling two random variables, say XX and YY, which are suspected of not being probabilistically independent. On the one hand, the marginal distributions of each variable are known, whether they are fX(x)f_{X}(x) and gY(y)g_{Y}(y). Typically this problem has been approached assuming that the random variables XX and YY are independent and, as a function of joint density of the two-dimensional vector (X,Y)(X,Y)^{\prime}, the product of the marginals, rX,Y(x,y)=fX(x)gY(y)r_{X,Y}(x,y)=f_{X}(x)g_{Y}(y), has been considered. Thus, for example, in this case the likelihood function, given the two-dimensional sample (x1,y1),,(xk,yk)(x_{1},y_{1}),\cdots,(x_{k},y_{k}), denoted as L(θ;(x1,y1),,(xk,yk))L(\boldmath{\theta};(x_{1},y_{1}),\cdots,(x_{k},y_{k})) is defined as

    L(θ;(x1,y1),,(xn,yn))\displaystyle L(\mathbf{\theta};(x_{1},y_{1}),\cdots,(x_{n},y_{n})) =\displaystyle= j=1krXj,Yj(xj,yj)\displaystyle\prod_{j=1}^{k}r_{X_{j},Y_{j}}(x_{j},y_{j})
    =\displaystyle= j=1kfXj(xj)gYj(yj)\displaystyle\prod_{j=1}^{k}f_{X_{j}}(x_{j})g_{Y_{j}}(y_{j})

    For some parameter vector θp\boldmath{\theta}\in\Re^{p} which is part of the density rX,Y(x,y)r_{X,Y}(x,y).

    Alternatively, based on variable changes on a set of independent random variables, the random vector (X,Y)(X,Y)^{\prime} was generated, where now the variables XX and YY are not independent and their density function is known joint tX,Y(x,y)fX(x)gY(y)t_{X,Y}(x,y)\neq f_{X}(x)g_{Y}(y), such that

    fX(x)=tX,Y(x,y)(dy) and gY(y)=tX,Y(x,y)(dx).f_{X}(x)=\int_{\mathfrak{\Re}}t_{X,Y}(x,y)(dy){\mbox{\hskip 14.22636pt and \hskip 14.22636pt}}g_{Y}(y)=\int_{\Re}t_{X,Y}(x,y)(dx).

    Under this approach we have that

    L(θ;(x1,y1),,(xk,yk))=j=1ktXj,Yj(xj,yj)L(\boldmath{\theta};(x_{1},y_{1}),\cdots,(x_{k},y_{k}))=\prod_{j=1}^{k}t_{X_{j},Y_{j}}(x_{j},y_{j})

    See Libby and Novick [24], Chen and Novick [3], Olkin and Liu [28], Nadarajah [26, 27] and Sarabia et al. [30] among many others.

    This situation also occurs in other multivariate problems. Then, parallel solutions were proposed in the vector and matrix cases, giving foothold to the study of bi-matrix variate distributions in the real and complex cases. In the last case joint distributions of random matrices, say 𝐗\mathbf{X} and 𝐘\mathbf{Y} dependent with joint density function, termed bimatrix variate distribution, are proposed such that the marginal densities of 𝐗\mathbf{X} and 𝐘\mathbf{Y} are the usual assumptions, see Olkin and Rubin [29], Díaz-García, and Gutiérrez-Jáimez [8, 9, 11], Bekker et al. [1], and [16] and references therein.

  2. 2.

    In another case, we are interested in defining the likelihood function by a joint function of the sample, but which is not defined as the product of the marginals, that is, the sample is not independent. In the univariate problem, one answer considers the elliptically contoured distribution of the vector (X1,,Xk)(X_{1},\dots,X_{k})^{\prime} as a likelihood function, noting that in reality the elliptically contoured distribution actually defines a distribution family, see Fang et al. [18], Fang and Zhang [17], Gupta and Varga [21] and the reference therein.

    Based on the family of matrix variate elliptically contoured distributions, the multimatrix variate and multimatricvariate distributions were proposed as a generalisation of the bi-matrix variate distributions, which are defined as the joint distribution of the dependent random matrices 𝐗1,,𝐗k\mathbf{X}_{1},\dots,\mathbf{X}_{k}, see Díaz-García et al. [7], and Díaz-García and Caro-Lopera [5, 6]. Thus, the multimatrix variate and multimatricvariate distributions can be used as likelihood functions for a sample of dependent random matrices with certain (usual) marginal distributions. Thus, the likelihood function of the sample 𝐗1,,𝐗k\mathbf{X}_{1},\dots,\mathbf{X}_{k} is defined as

    L(𝚯;𝐗1,,𝐗k)=f𝐗1,,𝐗k(𝐗1,,𝐗k).L(\mathbf{\Theta};\mathbf{X}_{1},\dots,\mathbf{X}_{k})=f_{\mathbf{X}_{1},\dots,\mathbf{X}_{k}}(\mathbf{X}_{1},\dots,\mathbf{X}_{k}).

Under the theory of multimatrix matrix or multimatricvariate distributions, each matrix 𝐗j\mathbf{X}_{j}, j=1,,kj=1,\dots,k into the density f𝐗1,,𝐗k(𝐗1,,𝐗k)f_{\mathbf{X}_{1},\dots,\mathbf{X}_{k}}(\mathbf{X}_{1},\dots,\mathbf{X}_{k}) can follow a different marginal distribution. This answers the following problem under an independent or dependent samples: Suppose that we have a matrix random sample as follows:


𝐗1\mathbf{X}_{1} 𝐗11\mathbf{X}_{11} 𝐗12\mathbf{X}_{12} \cdots 𝐗1r\mathbf{X}_{1r}
𝐗2\mathbf{X}_{2} 𝐗21\mathbf{X}_{21} 𝐗22\mathbf{X}_{22} \cdots 𝐗2r\mathbf{X}_{2r}
\vdots \vdots \vdots \ddots \vdots
𝐗k\mathbf{X}_{k} 𝐗k1\mathbf{X}_{k1} 𝐗k2\mathbf{X}_{k2} \cdots 𝐗kr\mathbf{X}_{kr}

And assume that the matrix 𝚯\mathbf{\Theta} contains the parameters of interest. Then the likelihood function L(𝚯;)L(\mathbf{\Theta};\cdot) can be defined as:

{j=1rf𝐗1j,,𝐗kj(𝐗1j,,𝐗kj),independence,f𝐗11,,𝐗1r,,𝐗k1,,𝐗kr(𝐗11,,𝐗1r,,𝐗k1,,𝐗kr),dependence.\left\{\begin{array}[]{ll}\displaystyle\prod_{j=1}^{r}f_{\mathbf{X}_{1j},\dots,\mathbf{X}_{kj}}(\mathbf{X}_{1j},\dots,\mathbf{X}_{kj}),&\hbox{independence,}\\ f_{\mathbf{X}_{11},\dots,\mathbf{X}_{1r},\dots,\mathbf{X}_{k1},\dots,\mathbf{X}_{kr}}(\mathbf{X}_{11},\dots,\mathbf{X}_{1r},\dots,\mathbf{X}_{k1},\dots,\mathbf{X}_{kr}),&\hbox{dependence.}\end{array}\right.

In the bi-matrix variate case, these problems have been studied in the real and complex cases, each giving rise to a series of non correlated publications. Fortunately, in terms of the theory of the real normed division algebras, a unification of the real and complex cases is possible. And an extension to the quaternionic and octonionic algebras is also feasible. It is worth mentioning that the octornionic case is still under research. At this time, they are valid for 2×22\times 2 octornionic matrices and in general it can only be conjectured that they may be valid.

In the present work, the multimatrix variate and multimatricvariate distributions are studied for matrix arguments which elements belong to the real normed division algebras. A brief description of the notation and some Jacobians for real normed division algebras is presented in Section 2. In addition, two more Jacobians are obtained and the definition of the matrix variate elliptically contoured distribution for real normed division algebras is presented. The main results on multimatrix variate and multimatricvariate distributions for real normed division algebras are obtained in Section 3. Some properties and extensions of multimatrix variate and multimatricvariate distribution with more than two different types of distributions in their arguments are studied in Section 4. An example in the quaternionic case is full derived in Section 5.

2 Notation and preliminary results

A detailed discussion of real normed division algebras may be found in Baez [2]. For convenience, we shall introduce some notations, although in general we adhere to standard notations.

A vector space is always a finite-dimensional module over the field of real numbers. An algebra 𝔉\mathfrak{F} is a vector space that is equipped with a bilinear map m:𝔉×𝔉𝔉m:\mathfrak{F}\times\mathfrak{F}\rightarrow\mathfrak{F} termed multiplication and a nonzero element 1𝔉1\in\mathfrak{F} termed the unit such that m(1,a)=m(a,1)=1m(1,a)=m(a,1)=1. As usual, we abbreviate m(a,b)=abm(a,b)=ab as abab. We do not assume 𝔉\mathfrak{F} associative. Given an algebra, we freely think of real numbers as elements of this algebra via the map ωω1\omega\mapsto\omega 1.

An algebra 𝔉\mathfrak{F} is a division algebra if given a,b𝔉a,b\in\mathfrak{F} with ab=0ab=0, then either a=0a=0 or b=0b=0. Equivalently, 𝔉\mathfrak{F} is a division algebra if the operation of left and right multiplications by any nonzero element is invertible. A normed division algebra is an algebra 𝔉\mathfrak{F} that is also a normed vector space with ab=ab||ab||=||a||||b||. This implies that 𝔉\mathfrak{F} is a division algebra and that 1=1||1||=1.

There are exactly four normed division algebras: real numbers (\Re), complex numbers (\mathfrak{C}), quaternions (\mathfrak{H}) and octonions (𝔒\mathfrak{O}), see Baez [2]. Taking into account that \Re, \mathfrak{C}, \mathfrak{H} and 𝔒\mathfrak{O} are the only normed division algebras; moreover, they are the only alternative division algebras, and all division algebras have a real dimension of 1,2,41,2,4 or 88, which is denoted by β\beta, see Baez [2, Theorems 1, 2 and 3]. In other branches of mathematics, the parameter α=2/β\alpha=2/\beta is used, see Edelman and Rao [15].

Let m,nβ{\mathcal{L}}^{\beta}_{m,n} be the linear space of all n×mn\times m matrices of rank mnm\leq n over 𝔉\mathfrak{F} with mm distinct positive singular values, where 𝔉\mathfrak{F} denotes a real finite-dimensional normed division algebra. Let 𝔉n×m\mathfrak{F}^{n\times m} be the set of all n×mn\times m matrices over 𝔉\mathfrak{F}. The dimension of 𝔉n×m\mathfrak{F}^{n\times m} over \Re is βmn\beta mn. Let 𝐀𝔉n×m\mathbf{A}\in\mathfrak{F}^{n\times m}, then 𝐀H=𝐀¯T\mathbf{A}^{H}=\overline{\mathbf{A}}^{T} denotes the usual conjugate transpose.

The set of matrices 𝐇1𝔉n×m\mathbf{H}_{1}\in\mathfrak{F}^{n\times m} such that 𝐇1H𝐇1=𝐈m\mathbf{H}_{1}^{H}\mathbf{H}_{1}=\mathbf{I}_{m} is a manifold denoted 𝒱m,nβ{\mathcal{V}}_{m,n}^{\beta}, is termed the Stiefel manifold (𝐇1\mathbf{H}_{1} is also known as semi-orthogonal (β=1\beta=1), semi-unitary (β=2\beta=2), semi-symplectic (β=4\beta=4) and semi-exceptional type (β=8\beta=8) matrices, see Dray and Manogue [13]). The dimension of 𝒱m,nβ\mathcal{V}_{m,n}^{\beta} over \Re is [βmnm(m1)β/2m][\beta mn-m(m-1)\beta/2-m]. In particular, 𝒱m,mβ{\mathcal{V}}_{m,m}^{\beta} with dimension over \Re, [m(m+1)β/2m][m(m+1)\beta/2-m], is the maximal compact subgroup 𝔘β(m)\mathfrak{U}^{\beta}(m) of m,mβ{\mathcal{L}}^{\beta}_{m,m} and consists of all matrices 𝐇𝔉m×m\mathbf{H}\in\mathfrak{F}^{m\times m} such that 𝐇H𝐇=𝐈m\mathbf{H}^{H}\mathbf{H}=\mathbf{I}_{m}. Therefore, 𝔘β(m)\mathfrak{U}^{\beta}(m) is the real orthogonal group 𝒪(m)\mathcal{O}(m) (β=1\beta=1), the unitary group 𝒰(m)\mathcal{U}(m) (β=2\beta=2), compact symplectic group 𝒮p(m)\mathcal{S}p(m) (β=4\beta=4) or exceptional type matrices 𝒪o(m)\mathcal{O}o(m) (β=8\beta=8), for 𝔉=\mathfrak{F}=\Re, \mathfrak{C}, \mathfrak{H} or 𝔒\mathfrak{O}, respectively.

We denote by 𝔖mβ{\mathfrak{S}}_{m}^{\beta} the real vector space of all 𝐒𝔉m×m\mathbf{S}\in\mathfrak{F}^{m\times m} such that 𝐒=𝐒H\mathbf{S}=\mathbf{S}^{H}. Let 𝔓mβ\mathfrak{P}_{m}^{\beta} be the cone of positive definite matrices 𝐒𝔉m×m\mathbf{S}\in\mathfrak{F}^{m\times m}; then 𝔓mβ\mathfrak{P}_{m}^{\beta} is an open subset of 𝔖mβ{\mathfrak{S}}_{m}^{\beta}. Over \Re, 𝔖mβ{\mathfrak{S}}_{m}^{\beta} consist of symmetric matrices; over \mathfrak{C}, Hermitian matrices; over \mathfrak{H}, quaternionic Hermitian matrices (also termed self-dual matrices) and over 𝔒\mathfrak{O}, octonionic Hermitian matrices. Generically, the elements of 𝔖mβ\mathfrak{S}_{m}^{\beta} are termed Hermitian matrices, irrespective of the nature of 𝔉\mathfrak{F}. The dimension of 𝔖mβ\mathfrak{S}_{m}^{\beta} over \Re is [m(m1)β+2m]/2[m(m-1)\beta+2m]/2.

Let 𝔇mβ\mathfrak{D}_{m}^{\beta} be the diagonal subgroup of m,mβ\mathcal{L}_{m,m}^{\beta} consisting of all 𝐃𝔉m×m\mathbf{D}\in\mathfrak{F}^{m\times m}, 𝐃=diag(d1,,dm)\mathbf{D}=\mathop{\rm diag}\nolimits(d_{1},\dots,d_{m}).

For any matrix 𝐗𝔉n×m\mathbf{X}\in\mathfrak{F}^{n\times m}, d𝐗d\mathbf{X} denotes the matrix of differentials (dxij)(dx_{ij}). Finally, we define the measure or volume element (d𝐗)(d\mathbf{X}) when 𝐗𝔉m×n,𝔖mβ\mathbf{X}\in\mathfrak{F}^{m\times n},\mathfrak{S}_{m}^{\beta}, 𝔇mβ\mathfrak{D}_{m}^{\beta} or 𝒱m,nβ\mathcal{V}_{m,n}^{\beta}, see Dimitriu [12].

If 𝐗𝔉n×m\mathbf{X}\in\mathfrak{F}^{n\times m} then (d𝐗)(d\mathbf{X}) (the Lebesgue measure in 𝔉n×m\mathfrak{F}^{n\times m}) denotes the exterior product of the βmn\beta mn functionally independent variables

(d𝐗)=i=1nj=1mdxij where dxij=r=1βdxij(r).(d\mathbf{X})=\bigwedge_{i=1}^{n}\bigwedge_{j=1}^{m}dx_{ij}\quad\mbox{ where }\quad dx_{ij}=\bigwedge_{r=1}^{\beta}dx_{ij}^{(r)}.

If 𝐒𝔖mβ\mathbf{S}\in\mathfrak{S}_{m}^{\beta} (or 𝐒𝔗Lβ(m)\mathbf{S}\in\mathfrak{T}_{L}^{\beta}(m)) then (d𝐒)(d\mathbf{S}) (the Lebesgue measure in 𝔖mβ\mathfrak{S}_{m}^{\beta} or in 𝔗Lβ(m)\mathfrak{T}_{L}^{\beta}(m)) denotes the exterior product of the m(m+1)β/2m(m+1)\beta/2 functionally independent variables (or denotes the exterior product of the m(m1)β/2+mm(m-1)\beta/2+m functionally independent variables, if siis_{ii}\in\Re for all i=1,,mi=1,\dots,m)

(d𝐒)={ijmr=1βdsij(r),i=1mdsiii<jmr=1βdsij(r),if sii.(d\mathbf{S})=\left\{\begin{array}[]{ll}\displaystyle\bigwedge_{i\leq j}^{m}\bigwedge_{r=1}^{\beta}ds_{ij}^{(r)},&\\ \displaystyle\bigwedge_{i=1}^{m}ds_{ii}\bigwedge_{i<j}^{m}\bigwedge_{r=1}^{\beta}ds_{ij}^{(r)},&\hbox{if }s_{ii}\in\Re.\end{array}\right.

The context generally establishes the conditions on the elements of 𝐒\mathbf{S}, that is, if sijs_{ij}\in\Re, \in\mathfrak{C}, \in\mathfrak{H} or 𝔒\in\mathfrak{O}. It is considered that

(d𝐒)=ijmr=1βdsij(r)i=1mdsiii<jmr=1βdsij(r).(d\mathbf{S})=\bigwedge_{i\leq j}^{m}\bigwedge_{r=1}^{\beta}ds_{ij}^{(r)}\equiv\bigwedge_{i=1}^{m}ds_{ii}\bigwedge_{i<j}^{m}\bigwedge_{r=1}^{\beta}ds_{ij}^{(r)}.

Observe, too, that for the Lebesgue measure (d𝐒)(d\mathbf{S}) defined thus, it is required that 𝐒𝔓mβ\mathbf{S}\in\mathfrak{P}_{m}^{\beta}, that is, 𝐒\mathbf{S} must be a non singular Hermitian matrix (Hermitian positive definite matrix).

If 𝚲𝔇mβ\mathbf{\Lambda}\in\mathfrak{D}_{m}^{\beta} then (d𝚲)(d\mathbf{\Lambda}) (the Legesgue measure in 𝔇mβ\mathfrak{D}_{m}^{\beta}) denotes the exterior product of the βm\beta m functionally independent variables

(d𝚲)=i=1nr=1βdλi(r).(d\mathbf{\Lambda})=\bigwedge_{i=1}^{n}\bigwedge_{r=1}^{\beta}d\lambda_{i}^{(r)}.

If 𝐇1𝒱m,nβ\mathbf{H}_{1}\in\mathcal{V}_{m,n}^{\beta} then

(𝐇1Hd𝐇1)=i=1nj=i+1m𝐡jHd𝐡i.(\mathbf{H}^{H}_{1}d\mathbf{H}_{1})=\bigwedge_{i=1}^{n}\bigwedge_{j=i+1}^{m}\mathbf{h}_{j}^{H}d\mathbf{h}_{i}.

where 𝐇=(𝐇1|𝐇2)=(𝐡1,,𝐡m|𝐡m+1,,𝐡n)𝔘β(m)\mathbf{H}=(\mathbf{H}_{1}|\mathbf{H}_{2})=(\mathbf{h}_{1},\dots,\mathbf{h}_{m}|\mathbf{h}_{m+1},\dots,\mathbf{h}_{n})\in\mathfrak{U}^{\beta}(m). It can be proved that this differential form does not depend on the choice of the 𝐇2\mathbf{H}_{2} matrix. When m=1m=1; 𝒱1,nβ\mathcal{V}^{\beta}_{1,n} defines the unit sphere in 𝔉n\mathfrak{F}^{n}. This is, of course, an (n1)β(n-1)\beta- dimensional surface in 𝔉n\mathfrak{F}^{n}. When m=nm=n and denoting 𝐇1\mathbf{H}_{1} by 𝐇\mathbf{H}, (𝐇Hd𝐇)(\mathbf{H}^{H}d\mathbf{H}) is termed the Haar measure on 𝔘β(m)\mathfrak{U}^{\beta}(m).

The surface area or volume of the Stiefel manifold 𝒱m,nβ\mathcal{V}^{\beta}_{m,n} is

Vol(𝒱m,nβ)=𝐇1𝒱m,nβ(𝐇1Hd𝐇1)=2mπmnβ/2Γmβ[nβ/2],\mathop{\rm Vol}\nolimits(\mathcal{V}^{\beta}_{m,n})=\int_{\mathbf{H}_{1}\in\mathcal{V}^{\beta}_{m,n}}(\mathbf{H}^{H}_{1}d\mathbf{H}_{1})=\frac{2^{m}\pi^{mn\beta/2}}{\Gamma^{\beta}_{m}[n\beta/2]}, (1)

where Γmβ[a]\Gamma^{\beta}_{m}[a] denotes the multivariate Gamma function for the space 𝔖mβ\mathfrak{S}_{m}^{\beta}, and is defined by

Γmβ[a]\displaystyle\Gamma_{m}^{\beta}[a] =\displaystyle= 𝐀𝔓mβetr{𝐀}|𝐀|a(m1)β/21(d𝐀)\displaystyle\displaystyle\int_{\mathbf{A}\in\mathfrak{P}_{m}^{\beta}}\mathop{\rm etr}\nolimits\{-\mathbf{A}\}|\mathbf{A}|^{a-(m-1)\beta/2-1}(d\mathbf{A})
=\displaystyle= πm(m1)β/4i=1mΓ[a(i1)β/2],\displaystyle\pi^{m(m-1)\beta/4}\displaystyle\prod_{i=1}^{m}\Gamma[a-(i-1)\beta/2],

where etr()=exp(tr())\mathop{\rm etr}\nolimits(\cdot)=\exp(\mathop{\rm tr}\nolimits(\cdot)), |||\cdot| denotes the determinant and Re(a)>(m1)β/2\mathop{\rm Re}\nolimits(a)>(m-1)\beta/2, see Gross and Richards [20]. If 𝐀m,nβ\mathbf{A}\in\mathcal{L}_{m,n}^{\beta} then by vec(𝐀)\mathop{\rm vec}\nolimits(\mathbf{A}) we mean the mn×1mn\times 1 vector formed by stacking the columns of 𝐀\mathbf{A} under each other; that is, if 𝐀=[𝐚1𝐚2𝐚m]\mathbf{A}=[\mathbf{a}_{1}\mathbf{a}_{2}\dots\mathbf{a}_{m}], where 𝐚j1,nβ\mathbf{a}_{j}\in\mathcal{L}_{1,n}^{\beta} for j=1,2,,mj=1,2,\dots,m

vec(𝐀)=[𝐚1𝐚2𝐚m].\mathop{\rm vec}\nolimits(\mathbf{A})=\left[\begin{array}[]{c}\mathbf{a}_{1}\\ \mathbf{a}_{2}\\ \vdots\\ \mathbf{a}_{m}\end{array}\right].

Below are summarised some Jacobians in terms of the β\beta parameter. For a detailed discussion of this and related issues see Dimitriu [12], Edelman and Rao [15], Forrester [19] and Kabe [22].

Proposition 2.1.

Let 𝐀n,nβ\mathbf{A}\in{\mathcal{L}}_{n,n}^{\beta}, 𝐁m,mβ\mathbf{B}\in{\mathcal{L}}_{m,m}^{\beta} and 𝐂m,nβ\mathbf{C}\in{\mathcal{L}}_{m,n}^{\beta} be matrices of constants, 𝐘\mathbf{Y} and 𝐗m,nβ\mathbf{X}\in{\mathcal{L}}_{m,n}^{\beta} a matrices of functionally independent variables such that 𝐘=𝐀𝐗𝐁+𝐂\mathbf{Y}=\mathbf{AXB}+\mathbf{C}. Then

(d𝐘)=|𝐀H𝐀|βm/2|𝐁H𝐁|βn/2(d𝐗).(d\mathbf{Y})=|\mathbf{A}^{H}\mathbf{A}|^{\beta m/2}|\mathbf{B}^{H}\mathbf{B}|^{\beta n/2}(d\mathbf{X}). (2)
Proposition 2.2.

Let 𝐒𝔓mβ\mathbf{S}\in\mathfrak{P}_{m}^{\beta}. If 𝐘=𝐀𝐒𝐀H\mathbf{Y}=\mathbf{A}\mathbf{S}\mathbf{A}^{H}, 𝐀𝔓mβ.\mathbf{A}\in\mathfrak{P}_{m}^{\beta}.

(d𝐘)=|𝐀H𝐀|β(m1)/2+1(d𝐗).(d\mathbf{Y})=|\mathbf{A}^{H}\mathbf{A}|^{\beta(m-1)/2+1}(d\mathbf{X}). (3)
Proposition 2.3 (Singular value decomposition, SVDSVD).

Assume that 𝐗m,nβ\mathbf{X}\in{\mathcal{L}}_{m,n}^{\beta}, such that 𝐗=𝐕1𝐃𝐖H\mathbf{X}=\mathbf{V}_{1}\mathbf{DW}^{H} with 𝐕1𝒱m,nβ\mathbf{V}_{1}\in{\mathcal{V}}_{m,n}^{\beta}, 𝐖𝔘β(m)\mathbf{W}\in\mathfrak{U}^{\beta}(m) and 𝐃=diag(d1,,dm)𝔇m1\mathbf{D}=\mathop{\rm diag}\nolimits(d_{1},\cdots,d_{m})\in\mathfrak{D}_{m}^{1}, d1>>dm>0d_{1}>\cdots>d_{m}>0. Then

(d𝐗)=2mπτi=1mdiβ(nm+1)1i<jm(di2dj2)β(d𝐃)(𝐕1Hd𝐕1)(𝐖Hd𝐖),(d\mathbf{X})=2^{-m}\pi^{\tau}\prod_{i=1}^{m}d_{i}^{\beta(n-m+1)-1}\prod_{i<j}^{m}(d_{i}^{2}-d_{j}^{2})^{\beta}(d\mathbf{D})(\mathbf{V}_{1}^{H}d\mathbf{V}_{1})(\mathbf{W}^{H}d\mathbf{W}), (4)

where

τ={0,β=1;m,β=2;2m,β=4;4m,β=8.\tau=\left\{\begin{array}[]{rl}0,&\beta=1;\\ -m,&\beta=2;\\ -2m,&\beta=4;\\ -4m,&\beta=8.\end{array}\right.

As a consequence of this result, we have the following statement.

Proposition 2.4.

Let 𝐗m,nβ\mathbf{X}\in{\mathcal{L}}_{m,n}^{\beta}, and 𝐒=𝐗H𝐗𝔓mβ.\mathbf{S}=\mathbf{X}^{H}\mathbf{X}\in\mathfrak{P}_{m}^{\beta}. Then

(d𝐗)=2m|𝐒|β(nm+1)/21(d𝐒)(𝐕1Hd𝐕1),(d\mathbf{X})=2^{-m}|\mathbf{S}|^{\beta(n-m+1)/2-1}(d\mathbf{S})(\mathbf{V}_{1}^{H}d\mathbf{V}_{1}), (5)

with 𝐕1𝒱m,nβ\mathbf{V}_{1}\in{\mathcal{V}}_{m,n}^{\beta}.

Proposition 2.5.

Let 𝐒𝔓mβ.\mathbf{S}\in\mathfrak{P}_{m}^{\beta}. Then ignoring the sign, if 𝐘=𝐒1\mathbf{Y}=\mathbf{S}^{-1}

(d𝐘)=|𝐒|β(m1)2(d𝐒).(d\mathbf{Y})=|\mathbf{S}|^{-\beta(m-1)-2}(d\mathbf{S}). (6)
Theorem 2.1.

Assume that 𝐗m,nβ\mathbf{X}\in{\mathcal{L}}_{m,n}^{\beta} and 𝐘m,nβ\mathbf{Y}\in{\mathcal{L}}_{m,n}^{\beta} are matrices of functionally independent variables.

i)

Define 𝐘=𝐗(𝐈m𝐗H𝐗)1/2\mathbf{Y}=\mathbf{X}(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})^{-1/2}. Then

(d𝐘)=|𝐈m𝐗H𝐗|β(n+m+1)/21(d𝐗).(d\mathbf{Y})=\left|\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X}\right|^{-\beta(n+m+1)/2-1}(d\mathbf{X}). (7)
ii)

If 𝐗=𝐘(𝐈m+𝐘H𝐘)1/2\mathbf{X}=\mathbf{Y}(\mathbf{I}_{m}+\mathbf{Y}^{H}\mathbf{Y})^{-1/2}, we have

(d𝐗)=|𝐈m+𝐘H𝐘|β(n+m+1)/21(d𝐘),(d\mathbf{X})=\left|\mathbf{I}_{m}+\mathbf{Y}^{H}\mathbf{Y}\right|^{-\beta(n+m+1)/2-1}(d\mathbf{Y}), (8)

where nmn\geq m.

Proof.

i) Define 𝐀=𝐘H𝐘=(𝐈m𝐗H𝐗)1/2𝐗H𝐗(𝐈m𝐗H𝐗)1/2\mathbf{A}=\mathbf{Y}^{H}\mathbf{Y}=(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})^{-1/2}\mathbf{X}^{H}\mathbf{X}(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})^{-1/2} and 𝐁=𝐗H𝐗\mathbf{B}=\mathbf{X}^{H}\mathbf{X}. And observe that

𝐀\displaystyle\mathbf{A} =\displaystyle= (𝐈m𝐗H𝐗)1/2𝐗H𝐗(𝐈m𝐗H𝐗)1/2\displaystyle(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})^{-1/2}\mathbf{X}^{H}\mathbf{X}(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})^{-1/2}
=\displaystyle= (𝐈m𝐁)1/2𝐁(𝐈m𝐁)1/2\displaystyle(\mathbf{I}_{m}-\mathbf{B})^{-1/2}\mathbf{B}(\mathbf{I}_{m}-\mathbf{B})^{-1/2}
=\displaystyle= [𝐁1(𝐈m𝐁)]1/2[𝐁1(𝐈m𝐁)]1/2\displaystyle\left[\mathbf{B}^{-1}(\mathbf{I}_{m}-\mathbf{B})\right]^{-1/2}\left[\mathbf{B}^{-1}(\mathbf{I}_{m}-\mathbf{B})\right]^{-1/2}
=\displaystyle= (𝐁1𝐈m)1/2(𝐁1𝐈m)1/2\displaystyle\left(\mathbf{B}^{-1}-\mathbf{I}_{m}\right)^{-1/2}\left(\mathbf{B}^{-1}-\mathbf{I}_{m}\right)^{-1/2}
=\displaystyle= (𝐁1𝐈m)1=(𝐈m𝐁)1𝐁=(𝐈m𝐁)1𝐈m.\displaystyle\left(\mathbf{B}^{-1}-\mathbf{I}_{m}\right)^{-1}=(\mathbf{I}_{m}-\mathbf{B})^{-1}\mathbf{B}=(\mathbf{I}_{m}-\mathbf{B})^{-1}-\mathbf{I}_{m}.

Then by (5), for 𝐇1,𝐆1𝒱m,n\mathbf{H}_{1},\mathbf{G}_{1}\in\mathcal{V}_{m,n}, and writing these for (d𝐀)(d\mathbf{A}) and (d𝐁)(d\mathbf{B}) we obtain

(d𝐀)\displaystyle(d\mathbf{A}) =\displaystyle= 2m|𝐀|β(nm+1)/2+1(d𝐘)(𝐇1Hd𝐇1)1\displaystyle 2^{m}|\mathbf{A}|^{-\beta(n-m+1)/2+1}(d\mathbf{Y})(\mathbf{H}^{H}_{1}d\mathbf{H}_{1})^{-1} (9)
(d𝐁)\displaystyle(d\mathbf{B}) =\displaystyle= 2m|𝐁|β(nm+1)/2+1(d𝐗)(𝐆1Hd𝐆1)1.\displaystyle 2^{m}|\mathbf{B}|^{-\beta(n-m+1)/2+1}(d\mathbf{X})(\mathbf{G}^{H}_{1}d\mathbf{G}_{1})^{-1}. (10)

Since 𝐀=(𝐈m𝐁)1𝐈m\mathbf{A}=(\mathbf{I}_{m}-\mathbf{B})^{-1}-\mathbf{I}_{m}, from (6), we have that

(d𝐀)=|𝐈m𝐁|β(m1)2(d𝐁).(d\mathbf{A})=|\mathbf{I}_{m}-\mathbf{B}|^{-\beta(m-1)-2}(d\mathbf{B}). (11)

Substituting (10) and (9) into (11) we have

2m|𝐀|β(nm+1)/2+1(d𝐘)(𝐇1Hd𝐇1)12^{m}|\mathbf{A}|^{-\beta(n-m+1)/2+1}(d\mathbf{Y})(\mathbf{H}^{H}_{1}d\mathbf{H}_{1})^{-1}\hskip 170.71652pt
=2m|𝐈m𝐁|(m+1)|𝐁|β(nm+1)/2+1(d𝐗)(𝐆1Hd𝐆1)1.\hskip 113.81102pt=2^{m}|\mathbf{I}_{m}-\mathbf{B}|^{-(m+1)}|\mathbf{B}|^{-\beta(n-m+1)/2+1}(d\mathbf{X})(\mathbf{G}^{H}_{1}d\mathbf{G}_{1})^{-1}.

Then, by the uniqueness of the nonnormalised measure on Stiefel manifold, (𝐇1Hd𝐇1)=(𝐆1Hd𝐆1)(\mathbf{H}^{H}_{1}d\mathbf{H}_{1})=(\mathbf{G}^{H}_{1}d\mathbf{G}_{1}). Thus,

(d𝐘)=|𝐀|β(nm+1)/2+1|𝐈m𝐁|β(m1)2|𝐁|β(nm+1)/2+1(d𝐗).(d\mathbf{Y})=|\mathbf{A}|^{\beta(n-m+1)/2+1}|\mathbf{I}_{m}-\mathbf{B}|^{-\beta(m-1)-2}|\mathbf{B}|^{-\beta(n-m+1)/2+1}(d\mathbf{X}).

And using |𝐀|=|(𝐈m𝐗H𝐗)1/2𝐗H𝐗(𝐈m𝐗H𝐗)1/2|=|(𝐈m𝐗H𝐗)|1|𝐗H𝐗||\mathbf{A}|=|(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})^{-1/2}\mathbf{X}^{H}\mathbf{X}(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})^{-1/2}|=|(\mathbf{I}_{m}-\mathbf{X}^{H}\mathbf{X})|^{-1}|\mathbf{X}^{H}\mathbf{X}| and |𝐁|=|𝐗H𝐗||\mathbf{B}|=|\mathbf{X}^{H}\mathbf{X}|, the required result is obtained.

ii). The proof is similar to the preceding exposition given in i). ∎

Corollary 2.1.

Assume that 𝐗m,nβ\mathbf{X}\in{\mathcal{L}}_{m,n}^{\beta} and 𝐘m,nβ\mathbf{Y}\in{\mathcal{L}}_{m,n}^{\beta} are matrices of functionally independent variables.

i)

Let 𝐘=(1tr𝐗H𝐗)1/2𝐗\mathbf{Y}=(1-\mathop{\rm tr}\nolimits\mathbf{X}^{H}\mathbf{X})^{-1/2}\mathbf{X}. Then

(d𝐘)=(1tr𝐗H𝐗)(βnm/2+1)(d𝐗).(d\mathbf{Y})=(1-\mathop{\rm tr}\nolimits\mathbf{X}^{H}\mathbf{X})^{-(\beta nm/2+1)}(d\mathbf{X}). (12)
ii)

If 𝐗=(1+tr𝐘H𝐘)1/2𝐘\mathbf{X}=(1+\mathop{\rm tr}\nolimits\mathbf{Y}^{H}\mathbf{Y})^{-1/2}\mathbf{Y}, we have

(d𝐗)=(1+tr𝐘H𝐘)(βnm/2+1)(d𝐘).(d\mathbf{X})=(1+\mathop{\rm tr}\nolimits\mathbf{Y}^{H}\mathbf{Y})^{-(\beta nm/2+1)}(d\mathbf{Y}). (13)
Proof.

i). Observing that 𝐘=(1tr𝐗H𝐗)1/2𝐗\mathbf{Y}=(1-\mathop{\rm tr}\nolimits\mathbf{X}^{H}\mathbf{X})^{-1/2}\mathbf{X} can be write as 𝐲=(1𝐱H𝐱)1/2𝐱\mathbf{y}=(1-\mathbf{x}^{H}\mathbf{x})^{-1/2}\mathbf{x}, where 𝐱=vec𝐗1,nmβ\mathbf{x}=\mathop{\rm vec}\nolimits\mathbf{X}\in\mathcal{L}_{1,nm}^{\beta} and 𝐲=vec𝐘1,nmβ\mathbf{y}=\mathop{\rm vec}\nolimits\mathbf{Y}\in\mathcal{L}_{1,nm}^{\beta}. Its proof is obtained as a particular case to that given for the theorem. ii). Its proof is analogous to one given to i). ∎

Definition 2.1.

It is said that the random matrix 𝐘m,nβ\mathbf{Y}\in\mathcal{L}_{m,n}^{\beta} has a matrix variate elliptical distribution, denoted as 𝐘n×mβ(𝝁,𝚯,𝚺,h)\mathbf{Y}\sim\mathcal{E}_{n\times m}^{\beta}(\boldsymbol{\mu},\mathbf{\Theta},\mathbf{\Sigma},h), if its density is

1|𝚺|βn/2|𝚯|βm/2h{βtr[𝚺1(𝐘𝝁)H𝚯1(𝐘𝝁)]}(d𝐘).\frac{1}{|\mathbf{\Sigma}|^{\beta n/2}|\mathbf{\Theta}|^{\beta m/2}}h\left\{\beta\mathop{\rm tr}\nolimits\left[\mathbf{\Sigma}^{-1}(\mathbf{Y}-\boldsymbol{\mu})^{H}\mathbf{\Theta}^{-1}(\mathbf{Y}-\boldsymbol{\mu})\right]\right\}(d\mathbf{Y}). (14)

where

𝔓1βumnβ/21h(βu)𝑑u<,\int_{\mathfrak{P}_{1}^{\beta}}u^{mn\beta/2-1}h(\beta u)du<\infty, (15)

and where 𝚯𝔓nβ\mathbf{\Theta}\in\mathfrak{P}_{n}^{\beta}, 𝚺𝔓mβ\mathbf{\Sigma}\in\mathfrak{P}_{m}^{\beta} and 𝝁m,nβ\boldsymbol{\mu}\in\mathcal{L}_{m,n}^{\beta} are constant matrices.

Finally, note that for a1,1βa\in\mathcal{L}_{1,1}^{\beta} constant, making the change of variable v=u/av=u/a and (du)=aβ(dv)(du)=a^{\beta}(dv) in Fang et al. [18, Equation 2.21, p. 26] we have

v𝔓1βvnmβ/21h(βav)(dv)=anmβ/2Γ1β[nmβ/2]πnmβ/2\int_{v\in\mathfrak{P}_{1}^{\beta}}v^{nm\beta/2-1}h(\beta av)(dv)=\frac{a^{-nm\beta/2}\Gamma^{\beta}_{1}[nm\beta/2]}{\pi^{nm\beta/2}} (16)

Finally, from Díaz-García, and Gutiérrez-Jáimez [10]

𝔓mβ|𝐕|β(nm+1)/21h(βtr𝚺1𝐕)(d𝐕)=|𝚺|βn/2Γmβ[βn/2]πβnm/2,\int_{\mathfrak{P}^{\beta}_{m}}|\mathbf{V}|^{\beta(n-m+1)/2-1}h\left(\beta\mathop{\rm tr}\nolimits\mathbf{\Sigma}^{-1}\mathbf{V}\right)(d\mathbf{V})=\frac{|\mathbf{\Sigma}|^{\beta n/2}\Gamma^{\beta}_{m}[\beta n/2]}{\pi^{\beta nm/2}}, (17)

where 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}_{m}^{\beta} and 𝚺𝔓mβ\mathbf{\Sigma}\in\mathfrak{P}_{m}^{\beta}.

3 Multimatrix variate and multimatricvariate distributions

First, note that any new particular distribution indexed by the kernel h()h(\cdot) form part of its density function, then it defines a family of distributions itself in terms of each possible choice of h()h(\cdot).

Assume that 𝐗n×mβ(𝟎,𝐈n,𝐈m;h)\mathbf{X}\sim\mathcal{E}_{n\times m}^{\beta}(\mathbf{0},\mathbf{I}_{n},\mathbf{I}_{m};h), such that n0+n1++nk=nn_{0}+n_{1}+\cdots+n_{k}=n, and 𝐗=(𝐗0H,𝐗1H,,𝐗kH)H\mathbf{X}=\left(\mathbf{X}^{H}_{0},\mathbf{X}^{H}_{1},\dots,\mathbf{X}^{H}_{k}\right)^{H}. Then (14) can be written as

dF𝐗0,𝐗1,,𝐗k(𝐗0,𝐗1,,𝐗k)=h[βtr(𝐗0H𝐗0+𝐗1H𝐗1++𝐗kH𝐗k)]i=0k(d𝐗i),dF_{\mathbf{X}_{0},\mathbf{X}_{1},\dots,\mathbf{X}_{k}}(\mathbf{X}_{0},\mathbf{X}_{1},\dots,\mathbf{X}_{k})=h[\beta\mathop{\rm tr}\nolimits(\mathbf{X}^{H}_{0}\mathbf{X}_{0}+\mathbf{X}^{H}_{1}\mathbf{X}_{1}+\cdots+\mathbf{X}^{H}_{k}\mathbf{X}_{k})]\bigwedge_{i=0}^{k}(d\mathbf{X}_{i}), (18)

where 𝐗im,niβ\mathbf{X}_{i}\in\mathcal{L}_{m,n_{i}}^{\beta}, i=0,1,,ki=0,1,\dots,k. Take in account that, only under a matrix variate normal distribution the random matrices are independent, see Fang and Zhang [17], Gupta and Varga [21] and Fang et al. [18]. In general, the random matrices 𝐗0,𝐗1,,𝐗k\mathbf{X}_{0},\mathbf{X}_{1},\dots,\mathbf{X}_{k} are probabilistically dependent.

Theorem 3.1.

Suppose that 𝐗n×mβ(𝟎,𝐈n,𝐈m;h)\mathbf{X}\sim\mathcal{E}_{n\times m}^{\beta}(\mathbf{0},\mathbf{I}_{n},\mathbf{I}_{m};h), with 𝐗im,niβ\mathbf{X}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, (remember that nimn_{i}\geq m), i=0,1,,ki=0,1,\dots,k.

i)

Define V=tr𝐗0H𝐗0V=\mathop{\rm tr}\nolimits\mathbf{X}_{0}^{H}\mathbf{X}_{0}. Then, the joint density dFV,𝐗1,,𝐗k(v,𝐗1,,𝐗k)dF_{V,\mathbf{X}_{1},\dots,\mathbf{X}_{k}}(v,\mathbf{X}_{1},\dots,\mathbf{X}_{k}) is given by

πn0mβ/2Γ1β[n0mβ/2]h[β(v+tri=1k𝐗iH𝐗i)]vn0mβ/21(dv)i=1k(d𝐗i),\frac{\pi^{n_{0}m\beta/2}}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}h\left[\beta\left(v+\displaystyle\mathop{\rm tr}\nolimits\sum_{i=1}^{k}\mathbf{X}_{i}^{H}\mathbf{X}_{i}\right)\right]v^{n_{0}m\beta/2-1}(dv)\bigwedge_{i=1}^{k}\left(d\mathbf{X}_{i}\right), (19)

where V𝔓1βV\in\mathfrak{P}^{\beta}_{1}. This distribution shall be termed multimatrix variate generalised Gamma - Elliptical distribution.

ii)

Let 𝐕=𝐗0H𝐗0\mathbf{V}=\mathbf{X}_{0}^{H}\mathbf{X}_{0}. Hence, the joint density dF𝐕,𝐗1,,𝐗k(𝐕,𝐗1,,𝐗k)dF_{\mathbf{V},\mathbf{X}_{1},\dots,\mathbf{X}_{k}}(\mathbf{V},\mathbf{X}_{1},\dots,\mathbf{X}_{k}) is given by

πβn0m/2|𝐕|β(n0m+1)/21Γmβ[βn0/2]h[βtr(𝐕+i=1k𝐗iH𝐗i)](d𝐕)i=1k(d𝐗i),\frac{\pi^{\beta n_{0}m/2}|\mathbf{V}|^{\beta(n_{0}-m+1)/2-1}}{\Gamma_{m}^{\beta}[\beta n_{0}/2]}h\left[\beta\mathop{\rm tr}\nolimits\left(\mathbf{V}+\displaystyle\sum_{i=1}^{k}\mathbf{X}_{i}^{H}\mathbf{X}_{i}\right)\right](d\mathbf{V})\bigwedge_{i=1}^{k}\left(d\mathbf{X}_{i}\right), (20)

where 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}^{\beta}_{m}. This distribution shall be called multimatricvariate generalised Whishart - Elliptical distribution.

Proof.

We have that the joint density function of 𝐗0,𝐗1,,𝐗k\mathbf{X}_{0},\mathbf{X}_{1},\dots,\mathbf{X}_{k} is

h[βtr(𝐗0H𝐗0+𝐗1H𝐗1++𝐗kH𝐗k)]i=0k(d𝐗i).h[\beta\mathop{\rm tr}\nolimits(\mathbf{X}^{H}_{0}\mathbf{X}_{0}+\mathbf{X}^{H}_{1}\mathbf{X}_{1}+\cdots+\mathbf{X}^{H}_{k}\mathbf{X}_{k})]\bigwedge_{i=0}^{k}(d\mathbf{X}_{i}). (21)
i)

Let V=tr𝐗0H𝐗0V=\mathop{\rm tr}\nolimits\mathbf{X}_{0}^{H}\mathbf{X}_{0}, hence by (5)

(d𝐗0)=21vn0mβ/21(dv)(𝐡1Hd𝐡1),(d\mathbf{X}_{0})=2^{-1}v^{n_{0}m\beta/2-1}(dv)\wedge(\mathbf{h}^{H}_{1}d\mathbf{h}_{1}),

where 𝐡1𝒱1,n0mβ\mathbf{h}_{1}\in\mathcal{V}^{\beta}_{1,n_{0}m}. Thus, the multimatrix variate joint density function

dFV,𝐡1,𝐗1,,𝐗k(v,𝐡1,𝐗1,,𝐗k)dF_{V,\mathbf{h}_{1},\mathbf{X}_{1},\dots,\mathbf{X}_{k}}(v,\mathbf{h}_{1},\mathbf{X}_{1},\dots,\mathbf{X}_{k})

is

vn0mβ/212h[β(v+tri=1k𝐗iH𝐗i)](dv)(𝐡1Hd𝐡1)i=1k(d𝐗i).\frac{v^{n_{0}m\beta/2-1}}{2}h\left[\beta\left(v+\mathop{\rm tr}\nolimits\sum_{i=1}^{k}\mathbf{X}^{H}_{i}\mathbf{X}_{i}\right)\right](dv)\wedge(\mathbf{h}^{H}_{1}d\mathbf{h}_{1})\bigwedge_{i=1}^{k}(d\mathbf{X}_{i}).

By integration over 𝐡1𝒱1,n0mβ\mathbf{h}_{1}\in\mathcal{V}^{\beta}_{1,n_{0}m} using (1), the desired result is obtained.

ii)

Define 𝐕=𝐗0H𝐗0\mathbf{V}=\mathbf{X}_{0}^{H}\mathbf{X}_{0}. Then by (5), we have that

(d𝐗0)=2m|𝐕|β(n0m+1)/21(d𝐕)(𝐇1Hd𝐇1),(d\mathbf{X}_{0})=2^{-m}|\mathbf{V}|^{\beta(n_{0}-m+1)/2-1}(d\mathbf{V})(\mathbf{H}_{1}^{H}d\mathbf{H}_{1}),

where 𝐇1𝒱m,n0β\mathbf{H}_{1}\in\mathcal{V}^{\beta}_{m,n_{0}}. The desired result is obtained make de change of variable in (21) and integrating over 𝐇1𝒱m,n0β\mathbf{H}_{1}\in\mathcal{V}^{\beta}_{m,n_{0}}, using (1).

Proceeding as Díaz-García et al. [7, Equation 4.2], defining Vi=tr𝐗iH𝐗iV_{i}=\mathop{\rm tr}\nolimits\mathbf{X}_{i}^{H}\mathbf{X}_{i}, i=0,1,,ki=0,1,\dots,k, and as in Díaz-García and Caro-Lopera [5, Equation (1), p. 216] the following result in general case, is obtained.

Theorem 3.2.

Suppose that 𝐗=(𝐗0H,,𝐗kH)H\mathbf{X}=\left(\mathbf{X}^{H}_{0},\dots,\mathbf{X}^{H}_{k}\right)^{H} has a matrix variate spherical distribution, with 𝐗im,niβ\mathbf{X}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, i=0,1,,ki=0,1,\dots,k. Then:

i)

If we define Vi=tr𝐗iH𝐗iV_{i}=\mathop{\rm tr}\nolimits\mathbf{X}_{i}^{H}\mathbf{X}_{i}, i=0,1,,ki=0,1,\dots,k, it is obtained

dFV0,,Vk(v0,,vk)=πnmβ/2i=0kvinimβ/21Γ1β[nimβ/2]h(βi=0kvi)i=0k(dvi),dF_{V_{0},\ldots,V_{k}}(v_{0},\ldots,v_{k})=\pi^{nm\beta/2}\prod_{i=0}^{k}\frac{v_{i}^{n_{i}m\beta/2-1}}{\Gamma_{1}^{\beta}\left[n_{i}m\beta/2\right]}h\left(\beta\sum_{i=0}^{k}v_{i}\right)\bigwedge_{i=0}^{k}(dv_{i}), (22)

where n=n0++nkn=n_{0}+\cdots+n_{k}, Vi𝔓1βV_{i}\in\mathfrak{P}^{\beta}_{1}, i=0,1,,ki=0,1,\dots,k, which distribution shall be termed multivariate generalised Gamma distribution.

ii)

The joint density dF𝐕0,𝐕1,,𝐕k(𝐕0,𝐕1,,𝐕k)dF_{\mathbf{V}_{0},\mathbf{V}_{1},\dots,\mathbf{V}_{k}}(\mathbf{V}_{0},\mathbf{V}_{1},\dots,\mathbf{V}_{k}) when 𝐕i=𝐗iH𝐗i\mathbf{V}_{i}=\mathbf{X}_{i}^{H}\mathbf{X}_{i}, i=0,1,,ki=0,1,\dots,k. is given by

πβnm/2i=0k|𝐕i|β(nim+1)/21Γmβ[βni/2]h(βtri=0k𝐕i)i=0k(d𝐕i),\pi^{\beta nm/2}\prod_{i=0}^{k}\frac{|\mathbf{V}_{i}|^{\beta(n_{i}-m+1)/2-1}}{\Gamma_{m}^{\beta}[\beta n_{i}/2]}h\left(\beta\displaystyle\mathop{\rm tr}\nolimits\sum_{i=0}^{k}\mathbf{V}_{i}\right)\bigwedge_{i=0}^{k}\left(d\mathbf{V}_{i}\right), (23)

where 𝐕i𝔓mβ\mathbf{V}_{i}\in\mathfrak{P}^{\beta}_{m}, i=0,1,,ki=0,1,\dots,k. This distribution shall be termed multimatricvariate generalised Whishart distribution.

Particular cases of these two distributions have been studied in the literature under a normal distribution in the real, complex and quaternionic cases, see Nadarajah [26], Fang and Zhang [17], Libby and Novick [24], Díaz-García and Caro-Lopera [5] and Li and Xue [23], among many other authors.

Theorem 3.3.

Assume that 𝐗n×mβ(𝟎,𝐈n,𝐈m;h)\mathbf{X}\sim\mathcal{E}_{n\times m}^{\beta}(\mathbf{0},\mathbf{I}_{n},\mathbf{I}_{m};h), with 𝐗im,niβ\mathbf{X}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, i=0,1,,ki=0,1,\dots,k.

i)

Define V=tr𝐗0H𝐗0V=\mathop{\rm tr}\nolimits\mathbf{X}^{H}_{0}\mathbf{X}_{0} and 𝐓i=V1/2𝐗i\mathbf{T}_{i}=V^{-1/2}\mathbf{X}_{i}, i=1,,ki=1,\dots,k.
The joint density dFV,𝐓1,,𝐓k(v,𝐓1,,𝐓k)dF_{V,\mathbf{T}_{1},\dots,\mathbf{T}_{k}}(v,\mathbf{T}_{1},\dots,\mathbf{T}_{k}) is given by

πn0mβ/2Γ1β[n0mβ/2]h[βv(1+i=1ktr𝐓iH𝐓i)]vnmβ/21(dv)i=1k(d𝐓i),\frac{\pi^{n_{0}m\beta/2}}{\Gamma_{1}^{\beta}[n_{0}m\beta/2]}h\left[\beta v\left(1+\displaystyle\sum_{i=1}^{k}\mathop{\rm tr}\nolimits\mathbf{T}^{H}_{i}\mathbf{T}_{i}\right)\right]v^{nm\beta/2-1}(dv)\bigwedge_{i=1}^{k}\left(d\mathbf{T}_{i}\right), (24)

where n=n0+n1++nkn=n_{0}+n_{1}+\cdots+n_{k}, V𝔓1βV\in\mathfrak{P}^{\beta}_{1} and 𝐓im,niβ\mathbf{T}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, i=1,,ki=1,\dots,k. This distribution shall be termed multimatrix variate generalised Gamma - Pearson type VII distribution.

ii)

Define 𝐕=𝐗0H𝐗0\mathbf{V}=\mathbf{X}^{H}_{0}\mathbf{X}_{0} and 𝐓i=𝐗i𝐕1/2\mathbf{T}_{i}=\mathbf{X}_{i}\mathbf{V}^{-1/2}, i=1,,ki=1,\dots,k.
Then, the joint density dF𝐕,𝐓1,,𝐓k(𝐕,𝐓1,,𝐓k)dF_{\mathbf{V},\mathbf{T}_{1},\dots,\mathbf{T}_{k}}(\mathbf{V},\mathbf{T}_{1},\dots,\mathbf{T}_{k}) is given by

πβn0m/2Γmβ[βn0/2]|𝐕|β(nm+1)/21h[βtr𝐕(𝐈m+i=1k𝐓iH𝐓i)](d𝐕)i=1k(d𝐓i),\frac{\pi^{\beta n_{0}m/2}}{\Gamma^{\beta}_{m}[\beta n_{0}/2]}|\mathbf{V}|^{\beta(n-m+1)/2-1}h\left[\beta\mathop{\rm tr}\nolimits\mathbf{V}\left(\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}\mathbf{T}^{H}_{i}\mathbf{T}_{i}\right)\right](d\mathbf{V})\bigwedge_{i=1}^{k}\left(d\mathbf{T}_{i}\right), (25)

where n=n0+n1++nkn=n_{0}+n_{1}+\cdots+n_{k}, 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}^{\beta}_{m} and 𝐓im,niβ\mathbf{T}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, i=1,,ki=1,\dots,k. This distribution shall be called multimatricvariate generalised Wishart-T distribution.

Proof.
i)

The density (24) is follow from (19) defining 𝐓i=V1/2𝐗i\mathbf{T}_{i}=V^{-1/2}\mathbf{X}_{i}, i=1,,ki=1,\dots,k. Hence by Proposition 2.1, we have

i=1k(d𝐗i)=v(nn0)mβ/2i=1k(d𝐓i).\bigwedge_{i=1}^{k}(d\mathbf{X}_{i})=v^{(n-n_{0})m\beta/2}\bigwedge_{i=1}^{k}(d\mathbf{T}_{i}).

Then the required result follows.

ii)

Now, from (20), making the change of variable 𝐓i=𝐗i𝐕1/2\mathbf{T}_{i}=\mathbf{X}_{i}\mathbf{V}^{-1/2}, i=1,,ki=1,\dots,k and considering that

i=1k(d𝐗i)=|𝐕|β(nn0)m/2i=1k(d𝐓i).\bigwedge_{i=1}^{k}(d\mathbf{X}_{i})=|\mathbf{V}|^{\beta(n-n_{0})m/2}\bigwedge_{i=1}^{k}(d\mathbf{T}_{i}).

by Proposition 2.1. The result is obtained.

Corollary 3.1.

Under the hypotheses of the theorem

i)

The marginal density dF𝐓1,,𝐓k(𝐓1,,𝐓k)dF_{\mathbf{T}_{1},\dots,\mathbf{T}_{k}}(\mathbf{T}_{1},\dots,\mathbf{T}_{k}) is termed multimatrix variate Pearson type VII and is given by

Γ1β[nmβ/2]π(nn0)mβ/2Γ1β[n0mβ/2](1+tri=1k𝐓iH𝐓i)nmβ/2i=1k(d𝐓i),\frac{\Gamma^{\beta}_{1}[nm\beta/2]}{\pi^{(n-n_{0})m\beta/2}\Gamma^{\beta}_{1}[n_{0}m\beta/2]}\left(1+\mathop{\rm tr}\nolimits\displaystyle\sum_{i=1}^{k}\mathbf{T}^{H}_{i}\mathbf{T}_{i}\right)^{-nm\beta/2}\bigwedge_{i=1}^{k}\left(d\mathbf{T}_{i}\right), (26)

where 𝐓im,niβ\mathbf{T}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}} and n=n0+n1++nkn=n_{0}+n_{1}+\cdots+n_{k}.

ii)

Similarly, the termed multimatricvariate Pearson VII distribution is the marginal density dF𝐓1,,𝐓k(𝐓1,,𝐓k)dF_{\mathbf{T}_{1},\dots,\mathbf{T}_{k}}(\mathbf{T}_{1},\dots,\mathbf{T}_{k}) of 𝐓1,,𝐓k\mathbf{T}_{1},\dots,\mathbf{T}_{k} and is given by

Γmβ[βn/2]πβ(nn0)m/2Γmβ[βn0/2]|𝐈m+i=1k𝐓iH𝐓i|βn/2i=1k(d𝐓i).\frac{\Gamma^{\beta}_{m}[\beta n/2]}{\pi^{\beta(n-n_{0})m/2}\Gamma^{\beta}_{m}[\beta n_{0}/2]}\left|\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}\mathbf{T}^{H}_{i}\mathbf{T}_{i}\right|^{-\beta n/2}\bigwedge_{i=1}^{k}\left(d\mathbf{T}_{i}\right). (27)

With 𝐓im,niβ\mathbf{T}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}} and n=n0+n1++nkn=n_{0}+n_{1}+\cdots+n_{k}.

Proof.
i)

Integrating (24) over V𝔓1βV\in\mathfrak{P}^{\beta}_{1} using (16) the density (26) is obtained.

ii)

Analogously, integrating (25) over 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}^{\beta}_{m}, using (17), is obtained

𝔓mβ|𝐕|β(nm+1)/21h[βtr𝐕(𝐈m+i=1k𝐓iH𝐓i)](d𝐕)\int_{\mathfrak{P}^{\beta}_{m}}|\mathbf{V}|^{\beta(n-m+1)/2-1}h\left[\beta\mathop{\rm tr}\nolimits\mathbf{V}\left(\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}\mathbf{T}^{H}_{i}\mathbf{T}_{i}\right)\right](d\mathbf{V})\hskip 85.35826pt
=Γmβ[βn/2]|𝐈m+i=1k𝐓iH𝐓i|βn/2πβnm/2,\hskip 142.26378pt=\frac{\Gamma^{\beta}_{m}[\beta n/2]\left|\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}\mathbf{T}^{H}_{i}\mathbf{T}_{i}\right|^{-\beta n/2}}{\pi^{\beta nm/2}},

and the desired result is archived.

Theorem 3.4.

Assume that 𝐗=(𝐗0H,,𝐗kH)H\mathbf{X}=\left(\mathbf{X}^{H}_{0},\dots,\mathbf{X}^{H}_{k}\right)^{H} has a matrix variate spherical distribution, with 𝐗im,niβ\mathbf{X}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, i=0,1,,ki=0,1,\dots,k.

i)

Define V=tr𝐗0H𝐗0V=\mathop{\rm tr}\nolimits\mathbf{X}^{H}_{0}\mathbf{X}_{0} and 𝐑i=(V+tr𝐗iH𝐗i)1/2𝐗i\mathbf{R}_{i}=\left(V+\mathop{\rm tr}\nolimits\mathbf{X}^{H}_{i}\mathbf{X}_{i}\right)^{-1/2}\mathbf{X}_{i}, i=1,,ki=1,\dots,k. Then the joint density of V,𝐑1,,𝐑kV,\mathbf{R}_{1},\dots,\mathbf{R}_{k}, denoted as dFV,𝐑1,,𝐑k(v,𝐑1,,𝐑k)dF_{V,\mathbf{R}_{1},\dots,\mathbf{R}_{k}}(v,\mathbf{R}_{1},\dots,\mathbf{R}_{k}), is given by

πn0mβ/2Γ1β[n0mβ/2]vnmβ/21h[βv(1+i=1ktr𝐑iH𝐑i(1tr𝐑iH𝐑i))]\frac{\pi^{n_{0}m\beta/2}}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}v^{nm\beta/2-1}h\left[\beta v\left(1+\displaystyle\sum_{i=1}^{k}\frac{\mathop{\rm tr}\nolimits\mathbf{R}^{H}_{i}\mathbf{R}_{i}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}_{i}\mathbf{R}_{i}\right)}\right)\right]\hskip 85.35826pt
×i=ik(1tr𝐑iH𝐑i)nimβ/21(dv)i=1k(d𝐑i),\hskip 56.9055pt\times\prod_{i=i}^{k}\left(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}_{i}\mathbf{R}_{i}\right)^{-n_{i}m\beta/2-1}(dv)\bigwedge_{i=1}^{k}\left(d\mathbf{R}_{i}\right), (28)

where n=n0+n1++nkn=n_{0}+n_{1}+\cdots+n_{k}, V𝔓1βV\in\mathfrak{P}^{\beta}_{1} and 𝐑im,niβ\mathbf{R}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, and tr𝐑iH𝐑i<1\mathop{\rm tr}\nolimits\mathbf{R}^{H}_{i}\mathbf{R}_{i}<1 i=1,,ki=1,\dots,k. This distribution shall be termed multimatrix variate generalised Gamma-Pearson type II distribution.

ii)

Define 𝐕=𝐗0H𝐗0\mathbf{V}=\mathbf{X}^{H}_{0}\mathbf{X}_{0} and 𝐑i=𝐗i(𝐕+𝐗iH𝐗i)1/2\mathbf{R}_{i}=\mathbf{X}_{i}(\mathbf{V}+\mathbf{X}_{i}^{H}\mathbf{X}_{i})^{-1/2}, i=1,,ki=1,\dots,k. Then the joint density of 𝐕,𝐑1,,𝐑k\mathbf{V},\mathbf{R}_{1},\dots,\mathbf{R}_{k}, denoted as dF𝐕,𝐑1,,𝐑k(𝐕,𝐑1,,𝐑k)dF_{\mathbf{V},\mathbf{R}_{1},\dots,\mathbf{R}_{k}}(\mathbf{V},\mathbf{R}_{1},\dots,\mathbf{R}_{k}), can be written as

πβn0m/2Γmβ[βn0/2]|𝐕|β(nm+1)/21h[βtr𝐕(𝐈m+i=1k(𝐈m𝐑iH𝐑i)1𝐑iH𝐑i)]\frac{\pi^{\beta n_{0}m/2}}{\Gamma^{\beta}_{m}[\beta n_{0}/2]}|\mathbf{V}|^{\beta(n-m+1)/2-1}h\left[\beta\mathop{\rm tr}\nolimits\mathbf{V}\left(\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}(\mathbf{I}_{m}-\mathbf{R}^{H}_{i}\mathbf{R}_{i})^{-1}\mathbf{R}^{H}_{i}\mathbf{R}_{i}\right)\right]
×i=ik|𝐈m𝐑iH𝐑i|β(nim+1)/21(d𝐕)i=1k(d𝐑i),\hskip 56.9055pt\times\prod_{i=i}^{k}|\mathbf{I}_{m}-\mathbf{R}^{H}_{i}\mathbf{R}_{i}|^{\beta(n_{i}-m+1)/2-1}(d\mathbf{V})\bigwedge_{i=1}^{k}\left(d\mathbf{R}_{i}\right), (29)

where n=n0+n1++nkn=n_{0}+n_{1}+\cdots+n_{k}, 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}^{\beta}_{m} and 𝐑im,niβ\mathbf{R}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, 𝐈m𝐑iH𝐑i𝔓mβ\mathbf{I}_{m}-\mathbf{R}^{H}_{i}\mathbf{R}_{i}\in\mathfrak{P}^{\beta}_{m}, i=1,,ki=1,\dots,k. This distribution shall be termed multimatricvariate generaliosed Wishart-Pearson type II distribution.

Proof.
i)

Consider in (24) the change of variable 𝐓i=(1tr𝐑iH𝐑i)1/2𝐑i\mathbf{T}_{i}=(1-tr\mathbf{R}^{H}_{i}\mathbf{R}_{i})^{-1/2}\mathbf{R}_{i}, i=1,,ki=1,\dots,k, hence by Theorem 2.1,

i=1k(d𝐓i)=i=ik(1tr𝐑iH𝐑i)(βnim/2+1)i=1k(d𝐑i),\bigwedge_{i=1}^{k}\left(d\mathbf{T}_{i}\right)=\prod_{i=i}^{k}\left(1-tr\mathbf{R}^{H}_{i}\mathbf{R}_{i}\right)^{-(\beta n_{i}m/2+1)}\bigwedge_{i=1}^{k}\left(d\mathbf{R}_{i}\right),

and the desired result follows.

ii)

Define 𝐓i=𝐑i(𝐈m𝐗iH𝐗i)1/2\mathbf{T}_{i}=\mathbf{R}_{i}(\mathbf{I}_{m}-\mathbf{X}_{i}^{H}\mathbf{X}_{i})^{-1/2}, i=1,,ki=1,\dots,k, with

i=1k(d𝐓i)=|𝐈m𝐑H𝐑|β(ni+m+1)/21i=1k(d𝐑i).\bigwedge_{i=1}^{k}\left(d\mathbf{T}_{i}\right)=\left|\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R}\right|^{-\beta(n_{i}+m+1)/2-1}\bigwedge_{i=1}^{k}\left(d\mathbf{R}_{i}\right).

The result is follow making this change of variable in (25).

Similarly to Corollary 3.1, from (28) and (29) we can obtain the multimatrix variate and multimatricvariate marginal densities dF𝐑1,,𝐑kdF_{\mathbf{R}_{1},\dots,\mathbf{R}_{k}}.

Corollary 3.2.

Consider the assumptions of the Theorem 3.4. Then

i)

The multimatrix variate marginal density dF𝐑1,,𝐑k(𝐑1,,𝐑k)dF_{\mathbf{R}_{1},\dots,\mathbf{R}_{k}}(\mathbf{R}_{1},\dots,\mathbf{R}_{k}) is

Γ1β[nmβ/2]π(nn0)mβ/2Γ1β[n0mβ/2][1+i=1ktr𝐑iH𝐑i(1tr𝐑iH𝐑i)]nmβ/2\frac{\Gamma^{\beta}_{1}[nm\beta/2]}{\pi^{(n-n_{0})m\beta/2}\Gamma^{\beta}_{1}[n_{0}m\beta/2]}\left[1+\displaystyle\sum_{i=1}^{k}\frac{\mathop{\rm tr}\nolimits\mathbf{R}_{i}^{H}\mathbf{R}_{i}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{R}_{i}^{H}\mathbf{R}_{i}\right)}\right]^{-nm\beta/2}\hskip 56.9055pt
×i=ik(1tr𝐑iH𝐑i)nimβ/21i=1k(d𝐑i),\hskip 85.35826pt\times\prod_{i=i}^{k}\left(1-\mathop{\rm tr}\nolimits\mathbf{R}_{i}^{H}\mathbf{R}_{i}\right)^{-n_{i}m\beta/2-1}\bigwedge_{i=1}^{k}\left(d\mathbf{R}_{i}\right), (30)

which shall be termed multimatrix variate Pearson type II distribution.

ii)

The multimaticvariate marginal density dF𝐑1,,𝐑k(𝐑1,,𝐑k)dF_{\mathbf{R}_{1},\dots,\mathbf{R}_{k}}(\mathbf{R}_{1},\dots,\mathbf{R}_{k}) is

Γmβ[βn/2]πβ(nn0)m/2Γmβ[n0/2]|𝐈mi=1k(𝐈m𝐑iH𝐑i)1𝐑iH𝐑i|βn/2\frac{\Gamma^{\beta}_{m}[\beta n/2]}{\pi^{\beta(n-n_{0})m/2}\Gamma^{\beta}_{m}[n_{0}/2]}\left|\mathbf{I}_{m}-\displaystyle\sum_{i=1}^{k}(\mathbf{I}_{m}-\mathbf{R}^{H}_{i}\mathbf{R}_{i})^{-1}\mathbf{R}^{H}_{i}\mathbf{R}_{i}\right|^{-\beta n/2}\hskip 142.26378pt
×i=1k|𝐈m𝐑iH𝐑i|β(nim+1)/21i=1k(d𝐑i),\hskip 56.9055pt\times\prod_{i=1}^{k}\left|\mathbf{I}_{m}-\mathbf{R}^{H}_{i}\mathbf{R}_{i}\right|^{\beta(n_{i}-m+1)/2-1}\bigwedge_{i=1}^{k}\left(d\mathbf{R}_{i}\right), (31)

which shall be termed multimatricvariate Pearson type II distribution.

Where n=n0+n1++nkn=n_{0}+n_{1}+\cdots+n_{k}, and 𝐑im,niβ\mathbf{R}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, i=1,,ki=1,\dots,k.

Proof.
i)

First, (30) is archived integrating (28) over V𝔓1βV\in\mathfrak{P}^{\beta}_{1} using (16).

ii)

Similarly, (31) is follows integrating (29) over V𝔓mβV\in\mathfrak{P}^{\beta}_{m} using (17).

Theorem 3.5.

Assuming the hypotheses of Theorem 3.3 and defining 𝐅i=𝐓iH𝐓i>𝟎\mathbf{F}_{i}=\mathbf{T}^{H}_{i}\mathbf{T}_{i}>\mathbf{0}, i=1,,ki=1,\dots,k.

i)

Then the joint density dFV,𝐅1,,𝐅k(v,𝐅1,,𝐅k)dF_{V,\mathbf{F}_{1},\dots,\mathbf{F}_{k}}(v,\mathbf{F}_{1},\dots,\mathbf{F}_{k}) is

πnmβ/2vnmβ/21Γ1β[n0mβ/2]i=1k(|𝐅i|β(nim+1)/21Γmβ[niβ/2])\frac{\pi^{nm\beta/2}v^{nm\beta/2-1}}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}\prod_{i=1}^{k}\left(\frac{|\mathbf{F}_{i}|^{\beta(n_{i}-m+1)/2-1}}{\Gamma^{\beta}_{m}[n_{i}\beta/2]}\right)\hskip 85.35826pt
×h[βv(1+i=1ktr𝐅i)](dv)i=1k(d𝐅i).\hskip 85.35826pt\times h\left[\beta v\left(1+\displaystyle\sum_{i=1}^{k}\mathop{\rm tr}\nolimits\mathbf{F}_{i}\right)\right](dv)\bigwedge_{i=1}^{k}\left(d\mathbf{F}_{i}\right). (32)

This distribution shall be termed multimatrix variate generalised Gamma-beta type II distribution.

ii)

Then the joint density dF𝐕0,𝐅1,,𝐅k(𝐕,𝐅1,,𝐅k)dF_{\mathbf{V}_{0},\mathbf{F}_{1},\dots,\mathbf{F}_{k}}(\mathbf{V},\mathbf{F}_{1},\dots,\mathbf{F}_{k}) is

πβmn/2i=0kΓmβ[βni/2]|𝐕|β(nm+1)/21i=1k|𝐅i|β(nim+1)/21\frac{\pi^{\beta mn/2}}{\displaystyle\prod_{i=0}^{k}\Gamma^{\beta}_{m}[\beta n_{i}/2]}|\mathbf{V}|^{\beta(n-m+1)/2-1}\prod_{i=1}^{k}|\mathbf{F}_{i}|^{\beta(n_{i}-m+1)/2-1}\hskip 85.35826pt
×h(βtr𝐕(𝐈m+i=1k𝐅i))(d𝐕)i=1k(d𝐅i).\hskip 56.9055pt\times h\left(\beta\mathop{\rm tr}\nolimits\mathbf{V}\left(\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}\mathbf{F}_{i}\right)\right)(d\mathbf{V})\bigwedge_{i=1}^{k}\left(d\mathbf{F}_{i}\right). (33)

This distribution can be termed multimatricvariate generalised Wishart-beta type II distribution.

Proof.

Multimatrix variate and multimatricvariate density functions (32) and (33) are obtained considering the change of variable 𝐅i=𝐓iH𝐓i\mathbf{F}_{i}=\mathbf{T}^{H}_{i}\mathbf{T}_{i}, i=1,,ki=1,\dots,k in expressions (24) and (25), respectively. Observing that by (2.4)

i=1k(d𝐓i)=2mki=1k|𝐅i|β(nim+1)/21i=1k(d𝐅i)i=1k(𝐇1iHd𝐇1i)\bigwedge_{i=1}^{k}\left(d\mathbf{T}_{i}\right)=2^{-mk}\prod_{i=1}^{k}|\mathbf{F}_{i}|^{\beta(n_{i}-m+1)/2-1}\bigwedge_{i=1}^{k}\left(d\mathbf{F}_{i}\right)\bigwedge_{i=1}^{k}\left(\mathbf{H}^{H}_{1_{i}}d\mathbf{H}_{1_{i}}\right)

where 𝐇1i𝒱ni,mβ\mathbf{H}_{1_{i}}\in\mathcal{V}^{\beta}_{n_{i},m}, i=1,,ki=1,\dots,k. Hence, integrating over 𝐇1i𝒱ni,mβ\mathbf{H}_{1_{i}}\in\mathcal{V}^{\beta}_{n_{i},m} i=1,,ki=1,\dots,k by (1) we have

𝐇11𝐇1ki=1k(𝐇1iHd𝐇1i)=2mkπβ(nn0)m/2i=1kΓmβ[βni/2]\int_{\mathbf{H}_{1_{1}}}\cdots\int_{\mathbf{H}_{1_{k}}}\bigwedge_{i=1}^{k}\left(\mathbf{H}^{H}_{1_{i}}d\mathbf{H}_{1_{i}}\right)=\frac{2^{mk}\pi^{\beta(n-n_{0})m/2}}{\displaystyle\prod_{i=1}^{k}\Gamma^{\beta}_{m}[\beta n_{i}/2]}

and the proof is complete. ∎

Corollary 3.3.

The corresponding marginal densities dF𝐅1,,𝐅k(𝐅1,,𝐅k)dF_{\mathbf{F}_{1},\dots,\mathbf{F}_{k}}(\mathbf{F}_{1},\dots,\mathbf{F}_{k}) of (32) is

i)
Γ1β[nmβ/2]Γ1β[n0mβ/2]i=1k(|𝐅i|β(nim+1)/21Γmβ[niβ/2])(1+i=1ktr𝐅i)nmβ/2i=1k(d𝐅i).\frac{\Gamma^{\beta}_{1}[nm\beta/2]}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}\prod_{i=1}^{k}\left(\frac{|\mathbf{F}_{i}|^{\beta(n_{i}-m+1)/2-1}}{\Gamma^{\beta}_{m}[n_{i}\beta/2]}\right)\left(1+\displaystyle\sum_{i=1}^{k}\mathop{\rm tr}\nolimits\mathbf{F}_{i}\right)^{-nm\beta/2}\bigwedge_{i=1}^{k}\left(d\mathbf{F}_{i}\right). (34)

Whose distribution can be termed multimatrix variate beta type II distribution.

ii)

And associated marginal density function dF𝐅1,,𝐅k(𝐅1,,𝐅k)dF_{\mathbf{F}_{1},\dots,\mathbf{F}_{k}}(\mathbf{F}_{1},\dots,\mathbf{F}_{k}) of (33) is given by

Γmβ[n/2]i=0kΓmβ[βni/2]i=1k|𝐅i|β(nim+1)/21|𝐈m+i=1k𝐅i|βn/2i=1k(d𝐅i).\frac{\Gamma^{\beta}_{m}[n/2]}{\displaystyle\prod_{i=0}^{k}\Gamma^{\beta}_{m}[\beta n_{i}/2]}\prod_{i=1}^{k}|\mathbf{F}_{i}|^{\beta(n_{i}-m+1)/2-1}\left|\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}\mathbf{F}_{i}\right|^{-\beta n/2}\bigwedge_{i=1}^{k}\left(d\mathbf{F}_{i}\right). (35)

This distribution shall be termed multimatricvariate beta type II distribution.

Proof.

The proof of both result are obtained integrating (32) and (33) with respect to V𝔓1βV\in\mathfrak{P}^{\beta}_{1} and 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}^{\beta}_{m} using (16) and (17), respectively. Or proceeding exactly as in the proof of Theorem 3.5

The density function (34) contain as particular case to the distribution in real case, proposed in Muirhead [25, Problem 3.18, p.118].

Theorem 3.6.

Assuming that 𝐁i=𝐑iH𝐑i𝔓𝔪β\mathbf{B}_{i}=\mathbf{R}^{H}_{i}\mathbf{R}_{i}\in\mathfrak{P_{m}^{\beta}}, i=1,,ki=1,\dots,k, in Theorem 3.4.

i)

Then the joint density dFV,𝐁1,,𝐁k(v,𝐁1,,𝐁k)dF_{V,\mathbf{B}_{1},\dots,\mathbf{B}_{k}}(v,\mathbf{B}_{1},\dots,\mathbf{B}_{k}), with tr𝐁i<1\mathop{\rm tr}\nolimits\mathbf{B}_{i}<1, i=1,,ki=1,\dots,k is

πnmβ/2vnmβ/21Γ1β[n0mβ/2]i=1k(|𝐁i|β(nim+1)/21Γmβ[niβ/2])h[βv(1+i=1ktr𝐁i(1tr𝐁i))]\frac{\pi^{nm\beta/2}v^{nm\beta/2-1}}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}\prod_{i=1}^{k}\left(\frac{|\mathbf{B}_{i}|^{\beta(ni-m+1)/2-1}}{\Gamma^{\beta}_{m}[n_{i}\beta/2]}\right)h\left[\beta v\left(1+\displaystyle\sum_{i=1}^{k}\frac{\mathop{\rm tr}\nolimits\mathbf{B}_{i}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{B}_{i}\right)}\right)\right]\hskip 85.35826pt
×i=ik(1tr𝐁i)nimβ/21(dv)i=1k(d𝐁i).\hskip 113.81102pt\times\prod_{i=i}^{k}\left(1-\mathop{\rm tr}\nolimits\mathbf{B}_{i}\right)^{-n_{i}m\beta/2-1}(dv)\bigwedge_{i=1}^{k}\left(d\mathbf{B}_{i}\right). (36)

This distribution shall be termed multimatrix variate generalised Gamma-beta type I distribution.

ii)

Then the joint density dF𝐕,𝐁1,,𝐁k(𝐕,𝐁1,,𝐁k)dF_{\mathbf{V},\mathbf{B}_{1},\dots,\mathbf{B}_{k}}(\mathbf{V},\mathbf{B}_{1},\dots,\mathbf{B}_{k}), where 𝐈m𝐁i𝔓mβ\mathbf{I}_{m}-\mathbf{B}_{i}\in\mathfrak{P}^{\beta}_{m}, i=1,2,,ki=1,2,\dots,k is

πβmn/2i=0kΓmβ[βni/2]|𝐕|β(nm+1)/21i=1k(|𝐁i||𝐈m𝐁i|)β(nim+1)/21\frac{\pi^{\beta mn/2}}{\displaystyle\prod_{i=0}^{k}\Gamma^{\beta}_{m}[\beta n_{i}/2]}|\mathbf{V}|^{\beta(n-m+1)/2-1}\prod_{i=1}^{k}\left(\frac{|\mathbf{B}_{i}|}{|\mathbf{I}_{m}-\mathbf{B}_{i}|}\right)^{\beta(n_{i}-m+1)/2-1}\hskip 56.9055pt
×h(βtr𝐕(𝐈m+i=1k(𝐈m𝐁i)1𝐁i))(d𝐕)i=1k(d𝐁i).\hskip 56.9055pt\times h\left(\beta\mathop{\rm tr}\nolimits\mathbf{V}\left(\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}(\mathbf{I}_{m}-\mathbf{B}_{i})^{-1}\mathbf{B}_{i}\right)\right)(d\mathbf{V})\bigwedge_{i=1}^{k}\left(d\mathbf{B}_{i}\right). (37)

This distribution can be termed multimatricvariate generalised Wishart-beta type I distribution.

Proof.

Making the change of variable 𝐁i=𝐑iH𝐑i\mathbf{B}_{i}=\mathbf{R}^{H}_{i}\mathbf{R}_{i}, i=1,,ki=1,\dots,k in the densities (28) and (29), respectively, and noting that

i=1k(d𝐑i)=2mki=1k|𝐁i|β(nim+1)/21i=1k(d𝐁i)i=1k(𝐇1iHd𝐇1i)\bigwedge_{i=1}^{k}\left(d\mathbf{R}_{i}\right)=2^{-mk}\prod_{i=1}^{k}|\mathbf{B}_{i}|^{\beta(n_{i}-m+1)/2-1}\bigwedge_{i=1}^{k}\left(d\mathbf{B}_{i}\right)\bigwedge_{i=1}^{k}\left(\mathbf{H}^{H}_{1_{i}}d\mathbf{H}_{1_{i}}\right)

where 𝐇1i𝒱ni,mβ\mathbf{H}_{1_{i}}\in\mathcal{V}^{\beta}_{n_{i},m}, i=1,,ki=1,\dots,k. The proof is archived, integrating over 𝐇1i𝒱ni,mβ\mathbf{H}_{1_{i}}\in\mathcal{V}^{\beta}_{n_{i},m} i=1,,ki=1,\dots,k using (1). In which case

𝐇11𝐇1ki=1k(𝐇1iHd𝐇1i)=2mkπβ(nn0)m/2i=1kΓmβ[βni/2].\int_{\mathbf{H}_{1_{1}}}\cdots\int_{\mathbf{H}_{1_{k}}}\bigwedge_{i=1}^{k}\left(\mathbf{H}^{H}_{1_{i}}d\mathbf{H}_{1_{i}}\right)=\frac{2^{mk}\pi^{\beta(n-n_{0})m/2}}{\displaystyle\prod_{i=1}^{k}\Gamma^{\beta}_{m}[\beta n_{i}/2]}.

Integrating (36) and (37) with respect to vv and 𝐕\mathbf{V} using (16) and (17), respectively; we obtain the marginal densities dF𝐁1,,𝐁k(𝐁1,,𝐁k)dF_{\mathbf{B}_{1},\dots,\mathbf{B}_{k}}(\mathbf{B}_{1},\dots,\mathbf{B}_{k}) of multimatrix variate and multimatricvariate beta type I distribution. Summarising:

Corollary 3.4.
i)

The density function dF𝐁1,,𝐁k(𝐁1,,𝐁k)dF_{\mathbf{B}_{1},\dots,\mathbf{B}_{k}}(\mathbf{B}_{1},\dots,\mathbf{B}_{k}) can written as

Γ1β[nmβ/2]Γ1β[n0mβ/2]i=1k(|𝐁i|β(nim+1)/21Γmβ[niβ/2](1tr𝐁i)nimβ/2+1)\frac{\Gamma^{\beta}_{1}[nm\beta/2]}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}\prod_{i=1}^{k}\left(\frac{|\mathbf{B}_{i}|^{\beta(n_{i}-m+1)/2-1}}{\Gamma^{\beta}_{m}[n_{i}\beta/2]\left(1-\mathop{\rm tr}\nolimits\mathbf{B}_{i}\right)^{n_{i}m\beta/2+1}}\right)\hskip 85.35826pt
×(1+i=1ktr𝐁i(1tr𝐁i))nmβ/2i=1k(d𝐁i).\hskip 113.81102pt\times\left(1+\displaystyle\sum_{i=1}^{k}\frac{\mathop{\rm tr}\nolimits\mathbf{B}_{i}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{B}_{i}\right)}\right)^{-nm\beta/2}\bigwedge_{i=1}^{k}\left(d\mathbf{B}_{i}\right). (38)

Whose marginal distribution shall be termed multimatrix variate beta type I distribution.

ii)

In this case, the density function dF𝐁1,,𝐁k(𝐁1,,𝐁k)dF_{\mathbf{B}_{1},\dots,\mathbf{B}_{k}}(\mathbf{B}_{1},\dots,\mathbf{B}_{k}) is

Γmβ[βn/2]i=0kΓmβ[βni/2]i=1k(|𝐁i||𝐈m𝐁i|)β(nim+1)/21\frac{\Gamma^{\beta}_{m}[\beta n/2]}{\displaystyle\prod_{i=0}^{k}\Gamma^{\beta}_{m}[\beta n_{i}/2]}\prod_{i=1}^{k}\left(\frac{|\mathbf{B}_{i}|}{|\mathbf{I}_{m}-\mathbf{B}_{i}|}\right)^{\beta(n_{i}-m+1)/2-1}\hskip 113.81102pt
×|𝐈m+i=1k(𝐈m𝐁i)1𝐁i|βn/2i=1k(d𝐁i).\hskip 56.9055pt\times\left|\mathbf{I}_{m}+\displaystyle\sum_{i=1}^{k}(\mathbf{I}_{m}-\mathbf{B}_{i})^{-1}\mathbf{B}_{i}\right|^{-\beta n/2}\bigwedge_{i=1}^{k}\left(d\mathbf{B}_{i}\right). (39)

This marginal distribution shall be named multimatricvariate beta type I distribution.

Finally

Theorem 3.7.

Assume that 𝐗=(𝐗0H,,𝐗kH)H\mathbf{X}=\left(\mathbf{X}^{H}_{0},\dots,\mathbf{X}^{H}_{k}\right)^{H} has a matrix variate spherical distribution, with 𝐗im,niβ\mathbf{X}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}, nimn_{i}\geq m, i=0,1,,ki=0,1,\dots,k. Define V=tr𝐗0H𝐗0V=\mathop{\rm tr}\nolimits\mathbf{X}^{H}_{0}\mathbf{X}_{0} and 𝐕i=𝐗iH𝐗i\mathbf{V}_{i}=\mathbf{X}^{H}_{i}\mathbf{X}_{i}, i=1,,ki=1,\dots,k.
Then, the joint density dFV,𝐕1,,𝐕k(v,𝐕1,,𝐕k)dF_{V,\mathbf{V}_{1},\dots,\mathbf{V}_{k}}(v,\mathbf{V}_{1},\dots,\mathbf{V}_{k}) is given by

πβnm/2vβnm/21Γ1β[βn0m/2]i=1k(|𝐕i|β(nim+1)/21Γmβ[βni/2])\frac{\pi^{\beta nm/2}v^{\beta nm/2-1}}{\Gamma^{\beta}_{1}[\beta n_{0}m/2]}\prod_{i=1}^{k}\left(\frac{|\mathbf{V}_{i}|^{\beta(ni-m+1)/2-1}}{\Gamma^{\beta}_{m}[\beta n_{i}/2]}\right)\hskip 142.26378pt
h[β(v+i=1ktr𝐕i)](dv)i=1k(d𝐕i),\hskip 85.35826pth\left[\beta\left(v+\displaystyle\sum_{i=1}^{k}\mathop{\rm tr}\nolimits\mathbf{V}_{i}\right)\right](dv)\bigwedge_{i=1}^{k}\left(d\mathbf{V}_{i}\right), (40)

where V𝔓1βV\in\mathfrak{P}^{\beta}_{1}, 𝐕i𝔓mβ\mathbf{V}_{i}\in\mathfrak{P}^{\beta}_{m}, i=1,,ki=1,\dots,k. This distribution shall be named multimatrix variate generalised Gamma - generalised Wishart distribution.

Proof.

From Theorem 3.1 we have

πn0mβ/2vn0mβ/21Γ1β[n0mβ/2]h[β(v+tri=1k𝐗iH𝐗i)](dv)i=1k(d𝐗i).\frac{\pi^{n_{0}m\beta/2}v^{n_{0}m\beta/2-1}}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}h\left[\beta\left(v+\mathop{\rm tr}\nolimits\displaystyle\sum_{i=1}^{k}\mathbf{X}^{H}_{i}\mathbf{X}_{i}\right)\right](dv)\bigwedge_{i=1}^{k}\left(d\mathbf{X}_{i}\right).

Defining 𝐕i=𝐗iH𝐗i\mathbf{V}_{i}=\mathbf{X}^{H}_{i}\mathbf{X}_{i} with i=1,,ki=1,\dots,k and proceeding as in the proof of Theorem 3.6, the result is immediate. ∎

4 Some properties and extensions

Now, multimatrix and multimatric variate distributions for two and three matrix arguments can be achived. Then we can obtain two or more different classes of marginal distributions. The metodology can be extended to more than three different marginal distributions. In addition, the inverse distributions of some multimatrix and multimatric variate distributions are also obtained.

Theorem 4.1.

Let be 𝐗=(𝐗0H,𝐗1H,𝐗2H)H\mathbf{X}=\left(\mathbf{X}^{H}_{0},\mathbf{X}^{H}_{1},\mathbf{X}^{H}_{2}\right)^{H} has a matrix variate spherical distribution, with 𝐗im,niβ\mathbf{X}_{i}\in\mathcal{L}^{\beta}_{m,n_{i}}.

i)

Define V0=tr𝐗0H𝐗0V_{0}=\mathop{\rm tr}\nolimits\mathbf{X}_{0}^{H}\mathbf{X}_{0}, 𝐓=V01/2𝐗1\mathbf{T}=V_{0}^{-1/2}\mathbf{X}_{1}, and 𝐑=𝐑=V1/2𝐗2\mathbf{R}=\mathbf{R}=V^{-1/2}\mathbf{X}_{2}, where V=(V0+tr𝐗2H𝐗2)V=(V_{0}+\mathop{\rm tr}\nolimits\mathbf{X}_{2}^{H}\mathbf{X}_{2}). The joint density dFV,𝐓,𝐑(v,𝐓,𝐑)dF_{V,\mathbf{T},\mathbf{R}}(v,\mathbf{T},\mathbf{R}) is given by

πn0mβ/2vnmβ/21Γβ1[n0mβ/2]h{βv[1+(1tr𝐑H𝐑)tr𝐓H𝐓]}\frac{\pi^{n_{0}m\beta/2}v^{nm\beta/2-1}}{\Gamma\mathbf{\beta}_{1}[n_{0}m\beta/2]}h\left\{\beta v\left[1+\left(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R}\right)\mathop{\rm tr}\nolimits\mathbf{T}^{H}\mathbf{T}\right]\right\}\hskip 142.26378pt
×(1tr𝐑H𝐑)(n0+n1)mβ/21(dv)(d𝐓)d𝐑).\hskip 85.35826pt\times\left(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R}\right)^{(n_{0}+n_{1})m\beta/2-1}(dv)\wedge(d\mathbf{T})\wedge d\mathbf{R}). (41)

where n=n0+n1+n2n=n_{0}+n_{1}+n_{2}, V𝔓1βV\in\mathfrak{P}^{\beta}_{1}, 𝐓m,n1β\mathbf{T}\in\mathcal{L}^{\beta}_{m,n_{1}}, 𝐑m,n2β\mathbf{R}\in\mathcal{L}^{\beta}_{m,n_{2}} such that tr𝐑H𝐑1\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R}\leq 1. This distribution shall be termed threematrix variate generalised Gamma - Pearson type VII - Pearson type II distribution.

ii)

Let be 𝐕0=𝐗0H𝐗0\mathbf{V}_{0}=\mathbf{X}^{H}_{0}\mathbf{X}_{0}, 𝐓=𝐗1𝐕01/2\mathbf{T}=\mathbf{X}_{1}\mathbf{V}_{0}^{-1/2}, 𝐕=𝐕0+𝐗2H𝐗2\mathbf{V}=\mathbf{V}_{0}+\mathbf{X}^{H}_{2}\mathbf{X}_{2}, and 𝐑=𝐗2𝐕1/2\mathbf{R}=\mathbf{X}_{2}\mathbf{V}^{-1/2}. Then the joint density dF𝐕,𝐓,𝐑(𝐕,𝐓,𝐑)dF_{\mathbf{V},\mathbf{T},\mathbf{R}}(\mathbf{V},\mathbf{T},\mathbf{R}) is given by

πβn0m/2Γmβ[βn0/2]|𝐕|β(nm+1)/21h{βtr[𝐕+(𝐈m𝐑H𝐑)𝐕1/2𝐓H𝐓𝐕1/2]}\frac{\pi^{\beta n_{0}m/2}}{\Gamma^{\beta}_{m}[\beta n_{0}/2]}|\mathbf{V}|^{\beta(n-m+1)/2-1}h\left\{\beta\mathop{\rm tr}\nolimits\left[\mathbf{V}+(\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R})\mathbf{V}^{1/2}\mathbf{T}^{H}\mathbf{TV}^{1/2}\right]\right\}
×|𝐈m𝐑H𝐑|β(n0+n1m+1)/21(d𝐕)(d𝐓)(d𝐑),\hskip 56.9055pt\times|\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R}|^{\beta(n_{0}+n_{1}-m+1)/2-1}(d\mathbf{V})\wedge(d\mathbf{T})\wedge\left(d\mathbf{R}\right), (42)

where n=n0+n1+n2n=n_{0}+n_{1}+n_{2}, 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}^{\beta}_{m}, 𝐈m𝐑H𝐑𝔓mβ\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R}\in\mathfrak{P}^{\beta}_{m}, 𝐓m,n1β\mathbf{T}\in\mathcal{L}^{\beta}_{m,n_{1}} and 𝐑m,n1β\mathbf{R}\in\mathcal{L}^{\beta}_{m,n_{1}}. The distribution of 𝐕,𝐓,𝐑\mathbf{V},\mathbf{T},\mathbf{R} shall be termed trimatricvariate generalised Wishart-Pearson VII-Pearson type II distribution.

Proof.
i)

From Theorem 3.1, dFV0,𝐗1,𝐗2(v0,𝐗1,𝐗2)dF_{V_{0},\mathbf{X}_{1},\mathbf{X}_{2}}(v_{0},\mathbf{X}_{1},\mathbf{X}_{2}) is

πn0mβ/2Γ1β[n0mβ/2]h[β(v0+tr𝐗1H𝐗1+tr𝐗2H𝐗2)]v0n0mβ/21\frac{\pi^{n_{0}m\beta/2}}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]}h\left[\beta\left(v_{0}+\mathop{\rm tr}\nolimits\mathbf{X}_{1}^{H}\mathbf{X}_{1}+\mathop{\rm tr}\nolimits\mathbf{X}_{2}^{H}\mathbf{X}_{2}\right)\right]v_{0}^{n_{0}m\beta/2-1}\hskip 56.9055pt
(dv0)(d𝐗1)(d𝐗2).\hskip 142.26378pt(dv_{0})\wedge\left(d\mathbf{X}_{1}\right)\wedge\left(d\mathbf{X}_{2}\right). (43)

Let V=V0+tr𝐗2H𝐗2V=V_{0}+\mathop{\rm tr}\nolimits\mathbf{X}_{2}^{H}\mathbf{X}_{2}, 𝐓=V01/2𝐗1\mathbf{T}=V_{0}^{-1/2}\mathbf{X}_{1}, and 𝐑=V1/2𝐗2\mathbf{R}=V^{-1/2}\mathbf{X}_{2}. Thus, 𝐗1=V01/2𝐓\mathbf{X}_{1}=V_{0}^{1/2}\mathbf{T}, 𝐗2=V1/2𝐑\mathbf{X}_{2}=V^{1/2}\mathbf{R}, and

V0=Vtr𝐗2H𝐗2=VVtr𝐑H𝐑=V(1tr𝐑H𝐑).V_{0}=V-\mathop{\rm tr}\nolimits\mathbf{X}_{2}^{H}\mathbf{X}_{2}=V-V\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R}=V(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R}).

Thus 𝐓=[V(1tr𝐑H𝐑)]1/2𝐗1\mathbf{T}=[V(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R})]^{-1/2}\mathbf{X}_{1} and (dv)=(dv0)(dv)=(dv_{0}). Then, the volume element (dv0)(d𝐗1)(d𝐗2)(dv_{0})\wedge\left(d\mathbf{X}_{1}\right)\wedge\left(d\mathbf{X}_{2}\right) is

v(n1+n2)mβ/2(1tr𝐑H𝐑)n1mβ/2(dv)(d𝐓)(d𝐑).v^{(n_{1}+n_{2})m\beta/2}(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R})^{n_{1}m\beta/2}(dv)\wedge\left(d\mathbf{T}\right)\wedge\left(d\mathbf{R}\right). (44)

From (43), substituting V=V0+tr𝐗2H𝐗2V=V_{0}+\mathop{\rm tr}\nolimits\mathbf{X}_{2}^{H}\mathbf{X}_{2}, 𝐗1=[V(1tr𝐑H𝐑)]1/2𝐓\mathbf{X}_{1}=[V(1-\mathop{\rm tr}\nolimits\mathbf{R}^{H}\mathbf{R})]^{-1/2}\mathbf{T} and by (44), the desired result is obtained.

ii)

From (20), dF𝐕0,𝐗1,𝐗2(𝐕0,𝐗1,𝐗2)dF_{\mathbf{V}_{0},\mathbf{X}_{1},\mathbf{X}_{2}}(\mathbf{V}_{0},\mathbf{X}_{1},\mathbf{X}_{2}) is

πβn0m/2|𝐕0|β(n0m+1)/21Γmβ[βn0/2]\frac{\pi^{\beta n_{0}m/2}|\mathbf{V}_{0}|^{\beta(n_{0}-m+1)/2-1}}{\Gamma_{m}^{\beta}[\beta n_{0}/2]}\hskip 199.16928pt
×h[βtr(𝐕0+𝐗1H𝐗1+𝐗2H𝐗2)](d𝐕0)(d𝐗1)(d𝐗2).\hskip 28.45274pt\times h\left[\beta\mathop{\rm tr}\nolimits\left(\mathbf{V}_{0}+\mathbf{X}_{1}^{H}\mathbf{X}_{1}+\mathbf{X}_{2}^{H}\mathbf{X}_{2}\right)\right](d\mathbf{V}_{0})\wedge\left(d\mathbf{X}_{1}\right)\wedge\left(d\mathbf{X}_{2}\right).

Define 𝐕=𝐕0+𝐗2H𝐗2\mathbf{V}=\mathbf{V}_{0}+\mathbf{X}_{2}^{H}\mathbf{X}_{2}, 𝐓=𝐗1𝐕01/2\mathbf{T}=\mathbf{X}_{1}\mathbf{V}_{0}^{-1/2}, and 𝐑=𝐗2𝐕1/2\mathbf{R}=\mathbf{X}_{2}\mathbf{V}^{-1/2}. Hence by Proposition 2.1,

(d𝐕0)(d𝐗1)(d𝐗2)=|𝐕0|βn1/2|𝐕0+𝐗2H𝐗2|βn2/2(d𝐕)(d𝐓)(d𝐑).(d\mathbf{V}_{0})\wedge(d\mathbf{X}_{1})\wedge(d\mathbf{X}_{2})=|\mathbf{V}_{0}|^{\beta n_{1}/2}|\mathbf{V}_{0}+\mathbf{X}_{2}^{H}\mathbf{X}_{2}|^{\beta n_{2}/2}(d\mathbf{V})\wedge(d\mathbf{T})\wedge(d\mathbf{R}).

But, 𝐗1=𝐓𝐕01/2\mathbf{X}_{1}=\mathbf{T}\mathbf{V}_{0}^{1/2}, 𝐗2=𝐑𝐕1/2\mathbf{X}_{2}=\mathbf{R}\mathbf{V}^{1/2}, and

𝐕0=𝐕𝐗2H𝐗2=𝐕𝐕1/2𝐑H𝐑𝐕1/2=𝐕1/2(𝐈m𝐑H𝐑)𝐕1/2.\mathbf{V}_{0}=\mathbf{V}-\mathbf{X}_{2}^{H}\mathbf{X}_{2}=\mathbf{V}-\mathbf{V}^{1/2}\mathbf{R}^{H}\mathbf{R}\mathbf{V}^{1/2}=\mathbf{V}^{1/2}(\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R})\mathbf{V}^{1/2}.

This way, 𝐗1=𝐓(𝐕1/2(𝐈m𝐑H𝐑)𝐕1/2)1/2\mathbf{X}_{1}=\mathbf{T}\left(\mathbf{V}^{1/2}(\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R})\mathbf{V}^{1/2}\right)^{1/2} and (d𝐕)=(d𝐕0)(d\mathbf{V})=(d\mathbf{V}_{0}). Therefore

(d𝐕0)(d𝐗1)(d𝐗2)=|𝐕|β(n1+n2)/2|𝐈m𝐑H𝐑|βn1/2(d𝐕)(d𝐓)(d𝐑)(d\mathbf{V}_{0})\wedge\left(d\mathbf{X}_{1}\right)\wedge\left(d\mathbf{X}_{2}\right)=|\mathbf{V}|^{\beta(n_{1}+n_{2})/2}|\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R}|^{\beta n_{1}/2}(d\mathbf{V})\wedge\left(d\mathbf{T}\right)\wedge\left(d\mathbf{R}\right)

Given that

|𝐕0+𝐗2H𝐗2|=|𝐕| and |𝐕0|=|𝐕||𝐈m𝐑H𝐑|.|\mathbf{V}_{0}+\mathbf{X}_{2}^{H}\mathbf{X}_{2}|=|\mathbf{V}|\mbox{ and }|\mathbf{V}_{0}|=|\mathbf{V}||\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R}|.

Finally, observe that tr𝐗1H𝐗1=tr𝐕1/2(𝐈m𝐑H𝐑)𝐕1/2𝐓H𝐓\mathop{\rm tr}\nolimits\mathbf{X}^{H}_{1}\mathbf{X}_{1}=\mathop{\rm tr}\nolimits\mathbf{V}^{1/2}(\mathbf{I}_{m}-\mathbf{R}^{H}\mathbf{R})\mathbf{V}^{1/2}\mathbf{T}^{H}\mathbf{T}. Then, the required result is obtained.

Corollary 4.1.

Under the Hypotheses of Theorem 4.1, define 𝐅=𝐓H𝐓\mathbf{F}=\mathbf{T}^{H}\mathbf{T} and 𝐁=𝐑H𝐑\mathbf{B}=\mathbf{R}^{H}\mathbf{R}.

i)

The density function of the termed trimatrix variate generalised Gamma - Pearson type VII - Pearson type II distribution, dF𝐓,𝐑(𝐓,𝐑)dF_{\mathbf{T},\mathbf{R}}(\mathbf{T},\mathbf{R}) is given by

πnmβ/2vnmβ/21|𝐅|β(n1m+1)/21|𝐁|β(n2m+1)/21Γ1β[n0mβ/2]Γmβ[n1β/2]Γmβ[n2mβ/2](1tr𝐁)(n0+n1)mβ/21\frac{\pi^{nm\beta/2}v^{nm\beta/2-1}|\mathbf{F}|^{\beta(n_{1}-m+1)/2-1}|\mathbf{B}|^{\beta(n_{2}-m+1)/2-1}}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]\Gamma^{\beta}_{m}[n_{1}\beta/2]\Gamma^{\beta}_{m}[n_{2}m\beta/2]}\left(1-\mathop{\rm tr}\nolimits\mathbf{B}\right)^{(n_{0}+n_{1})m\beta/2-1}
×h{βv[1+(1tr𝐁)tr𝐅]}(dv)d𝐅)d𝐁).\times h\left\{\beta v\left[1+\left(1-\mathop{\rm tr}\nolimits\mathbf{B}\right)\mathop{\rm tr}\nolimits\mathbf{F}\right]\right\}(dv)\wedge d\mathbf{F})\wedge d\mathbf{B}). (45)
ii)

Similarly, the density dF𝐕,𝐅,𝐔(𝐕,𝐅,𝐔)dF_{\mathbf{V},\mathbf{F},\mathbf{U}}(\mathbf{V},\mathbf{F},\mathbf{U}) is

πβnm/2|𝐕|β(nm+1)/21Γmβ[βn0/2]Γmβ[βn1/2]Γmβ[βn2/2]h[βtr(𝐕+(𝐈m𝐁)𝐕1/2𝐅𝐕1/2)]\frac{\pi^{\beta nm/2}|\mathbf{V}|^{\beta(n-m+1)/2-1}}{\Gamma^{\beta}_{m}[\beta n_{0}/2]\Gamma^{\beta}_{m}[\beta n_{1}/2]\Gamma^{\beta}_{m}[\beta n_{2}/2]}h\left[\beta\mathop{\rm tr}\nolimits\left(\mathbf{V}+(\mathbf{I}_{m}-\mathbf{B})\mathbf{V}^{1/2}\mathbf{FV}^{1/2}\right)\right]\hskip 56.9055pt
×|𝐈m𝐁|β(n0+n1m+1)/21|𝐅|β(n1m1)/21|𝐁|β(n2m+1)/21\hskip 56.9055pt\times|\mathbf{I}_{m}-\mathbf{B}|^{\beta(n_{0}+n_{1}-m+1)/2-1}|\mathbf{F}|^{\beta(n_{1}-m1)/2-1}|\mathbf{B}|^{\beta(n_{2}-m+1)/2-1}
(d𝐕)(d𝐅)(d𝐁),\hskip 227.62204pt(d\mathbf{V})\wedge(d\mathbf{F})\wedge\left(d\mathbf{B}\right), (46)

This density shall be termed trimatricvariate generalised Wishart-beta type II-beta type I distribution.

Where n=n0+n1+n2n=n_{0}+n_{1}+n_{2},V𝔓1βV\in\mathfrak{P}^{\beta}_{1}, 𝐅𝔓mβ\mathbf{F}\in\mathfrak{P}^{\beta}_{m}, 𝐁𝔓mβ\mathbf{B}\in\mathfrak{P}^{\beta}_{m}, 𝐕𝔓mβ\mathbf{V}\in\mathfrak{P}^{\beta}_{m}, tr𝐁1\mathop{\rm tr}\nolimits\mathbf{B}\leq 1 and 𝐈m𝐁𝔓mβ\mathbf{I}_{m}-\mathbf{B}\in\mathfrak{P}^{\beta}_{m}.

Proof.

The proofs of (45) and (46) are follows from (41) and (42), respectively; making the change of variables 𝐅=𝐓H𝐓\mathbf{F}=\mathbf{T}^{H}\mathbf{T}, 𝐁=𝐑H𝐑\mathbf{B}=\mathbf{R}^{H}\mathbf{R}, and by Proposition 2.4,

(d𝐓)d𝐑)=22m|𝐅|β(n1m+1)/21|𝐁|β(n2m+1)/21(d𝐅)d𝐁)i=12(𝐇1iHd𝐇1i).(d\mathbf{T})\wedge d\mathbf{R})=2^{-2m}|\mathbf{F}|^{\beta(n_{1}-m+1)/2-1}|\mathbf{B}|^{\beta(n_{2}-m+1)/2-1}(d\mathbf{F})\wedge d\mathbf{B})\bigwedge_{i=1}^{2}\left(\mathbf{H}^{H}_{1_{i}}d\mathbf{H}_{1_{i}}\right).

The desired results are archived, integrating over 𝐇1i𝒱ni,mβ\mathbf{H}_{1_{i}}\in\mathcal{V}^{\beta}_{n_{i},m} i=1,2i=1,2 using (1). Moreover

𝐇11𝐇12i=12(𝐇1iHd𝐇1i)=22mπβ(nn0)m/2i=12Γmβ[βni/2].\int_{\mathbf{H}_{1_{1}}}\int_{\mathbf{H}_{1_{2}}}\bigwedge_{i=1}^{2}\left(\mathbf{H}^{H}_{1_{i}}d\mathbf{H}_{1_{i}}\right)=\frac{2^{2m}\pi^{\beta(n-n_{0})m/2}}{\displaystyle\prod_{i=1}^{2}\Gamma^{\beta}_{m}[\beta n_{i}/2]}.

Additionally, we are interested in the distributions of the inverse of one or more of the arguments in the multimatrix variate or multimatricvariate distributions, which shall be termed inverse multimatrix variate or inverse multimatricvariate distributions.

Theorem 4.2.

Define 𝐀i=𝐁i1\mathbf{A}_{i}=\mathbf{B}^{-1}_{i}, i=1,,ri=1,\dots,r.

i)

Assume that 𝐁1,,𝐁r,𝐁r+1,,𝐁k\mathbf{B}_{1},\cdots,\mathbf{B}_{r},\mathbf{B}_{r+1},\cdots,\mathbf{B}_{k} have a multimatrix variate beta type I distribution. The join density function

dF𝐀1,,𝐀r,𝐁r+1,,𝐁k(𝐀1,,𝐀r,𝐁r+1,,𝐁k),dF_{\mathbf{A}_{1},\cdots,\mathbf{A}_{r},\mathbf{B}_{r+1},\cdots,\mathbf{B}_{k}}(\mathbf{A}_{1},\cdots,\mathbf{A}_{r},\mathbf{B}_{r+1},\cdots,\mathbf{B}_{k}),

is

Γ1β[nmβ/2]Γ1β[n0mβ/2]i=1kΓmβ[niβ/2]i=1r(|𝐀i|β(ni+m1)/21(1tr𝐀i1)nimβ/2+1)\frac{\Gamma^{\beta}_{1}[nm\beta/2]}{\Gamma^{\beta}_{1}[n_{0}m\beta/2]\prod_{i=1}^{k}\Gamma^{\beta}_{m}[n_{i}\beta/2]}\prod_{i=1}^{r}\left(\frac{|\mathbf{A}_{i}|^{-\beta(n_{i}+m-1)/2-1}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{A}^{-1}_{i}\right)^{n_{i}m\beta/2+1}}\right)\hskip 56.9055pt
×(1+i=1rtr𝐀i1(1tr𝐀i1)+i=r+1ktr𝐁i(1tr𝐁i))nmβ/2\hskip 56.9055pt\times\left(1+\displaystyle\sum_{i=1}^{r}\frac{\mathop{\rm tr}\nolimits\mathbf{A}^{-1}_{i}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{A}^{-1}_{i}\right)}+\sum_{i=r+1}^{k}\frac{\mathop{\rm tr}\nolimits\mathbf{B}_{i}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{B}_{i}\right)}\right)^{-nm\beta/2}
×i=r+1k(|𝐁i|β(nim+1)/21(1tr𝐁i)nimβ/2+1)i=1r(d𝐀i)i=r+1k(d𝐁i),\hskip 56.9055pt\times\prod_{i=r+1}^{k}\left(\frac{|\mathbf{B}_{i}|^{\beta(n_{i}-m+1)/2-1}}{\left(1-\mathop{\rm tr}\nolimits\mathbf{B}_{i}\right)^{n_{i}m\beta/2+1}}\right)\bigwedge_{i=1}^{r}\left(d\mathbf{A}_{i}\right)\bigwedge_{i=r+1}^{k}\left(d\mathbf{B}_{i}\right), (47)

where 𝐀i𝔓mβ\mathbf{A}_{i}\in\mathfrak{P}^{\beta}_{m}, 𝐁i𝔓mβ\mathbf{B}_{i}\in\mathfrak{P}^{\beta}_{m}, tr𝐀i<1\mathop{\rm tr}\nolimits\mathbf{A}_{i}<1, tr𝐁i<1\mathop{\rm tr}\nolimits\mathbf{B}_{i}<1.

ii)

Consider that 𝐁1,,𝐁r,𝐁r+1,,𝐁k\mathbf{B}_{1},\cdots,\mathbf{B}_{r},\mathbf{B}_{r+1},\cdots,\mathbf{B}_{k} have a multimatricvariate beta type I distribution. Then, the join density function

dF𝐀1,,𝐀r,𝐁r+1,,𝐁k(𝐀1,,𝐀r,𝐁r+1,,𝐁k)dF_{\mathbf{A}_{1},\cdots,\mathbf{A}_{r},\mathbf{B}_{r+1},\cdots,\mathbf{B}_{k}}(\mathbf{A}_{1},\cdots,\mathbf{A}_{r},\mathbf{B}_{r+1},\cdots,\mathbf{B}_{k})

is given by

Γmβ[βn/2]i=0kΓmβ[βni/2]i=1r(|𝐀i|β(m1)2|𝐀i𝐈m|β(nim+1)/21)\frac{\Gamma^{\beta}_{m}[\beta n/2]}{\displaystyle\prod_{i=0}^{k}\Gamma^{\beta}_{m}[\beta n_{i}/2]}\prod_{i=1}^{r}\left(\frac{|\mathbf{A}_{i}|^{-\beta(m-1)-2}}{|\mathbf{A}_{i}-\mathbf{I}_{m}|^{\beta(n_{i}-m+1)/2-1}}\right)\hskip 142.26378pt
×|𝐈m+i=1r(𝐀i𝐈m)1+i=r+1k(𝐈m𝐁i)1𝐁i|βn/2\hskip 56.9055pt\times\left|\mathbf{I}_{m}+\sum_{i=1}^{r}(\mathbf{A}_{i}-\mathbf{I}_{m})^{-1}+\sum_{i=r+1}^{k}(\mathbf{I}_{m}-\mathbf{B}_{i})^{-1}\mathbf{B}_{i}\right|^{-\beta n/2}
×i=r+1k(|𝐁i||𝐈m𝐁i|)β(nim+1)/21i=1r(d𝐀i)i=r+1k(d𝐁i),\hskip 36.98866pt\times\prod_{i=r+1}^{k}\left(\frac{|\mathbf{B}_{i}|}{|\mathbf{I}_{m}-\mathbf{B}_{i}|}\right)^{\beta(n_{i}-m+1)/2-1}\bigwedge_{i=1}^{r}\left(d\mathbf{A}_{i}\right)\bigwedge_{i=r+1}^{k}\left(d\mathbf{B}_{i}\right), (48)

where 𝐀i𝔓mβ\mathbf{A}_{i}\in\mathfrak{P}^{\beta}_{m}, 𝐁i𝔓mβ\mathbf{B}_{i}\in\mathfrak{P}^{\beta}_{m}, 𝐀i𝐈m𝔓mβ\mathbf{A}_{i}-\mathbf{I}_{m}\in\mathfrak{P}^{\beta}_{m}, 𝐈m𝐁i𝔓mβ\mathbf{I}_{m}-\mathbf{B}_{i}\in\mathfrak{P}^{\beta}_{m}

Proof.

The density functions (47) and (48) are archived from (38) and (39), respectively, defining 𝐀i=𝐁𝐢1\mathbf{A}_{i}=\mathbf{B_{i}}^{-1}, i=1,,ri=1,\cdots,r, and using the Proposition 2.5. ∎

The parameter domain of the real normed division algebra in the multimatrix and multimatric variate distributions can be extended as in the real and complex cases. However, the statistical and/or geometrical interpretation, perhaps can be lost. In any case, these distributions are valid if we replace ni/2n_{i}/2 by aia_{i}, n0m/2n_{0}m/2 by a0a_{0} and nm/2nm/2 by aa. Where the asa^{\prime s} are complex number with positive real part. From practical point of view, this parameter domain extension its allow to use nonlinear optimisation rather integer nonlinear optimisation in the procedure of estimation, among other possibilities.

Each distribution can be reparametrised in order to obtain a general expression for its density function. As in the normal case, the expressions obtained in this article appear in their standard form.

If (𝐕0,𝐕1,,𝐕k)(\mathbf{V}_{0},\mathbf{V}_{1},\dots,\mathbf{V}_{k}) follows a multimatricvariate or multimatrix variate generalised Whishart distribution, with density function (23). Define 𝐖i=𝚺i1/2𝐕i𝚺i1/2\mathbf{W}_{i}=\mathbf{\Sigma}_{i}^{1/2}\mathbf{V}_{i}\mathbf{\Sigma}_{i}^{1/2}, 𝚺i𝔓mβ\mathbf{\Sigma}_{i}\in\mathfrak{P}^{\beta}_{m}, i=0,1,,ki=0,1,\dots,k, then by Proposition 2.2, we have that

i=0k(d𝐕i)=i=0k|𝚺i|β(m1)/21i=0k(d𝐖i).\bigwedge_{i=0}^{k}\left(d\mathbf{V}_{i}\right)=\prod_{i=0}^{k}|\mathbf{\Sigma}_{i}|^{-\beta(m-1)/2-1}\bigwedge_{i=0}^{k}\left(d\mathbf{W}_{i}\right).

Then the density dF𝐖0,𝐖1,,𝐖k(𝐖0,𝐖1,,𝐖k)dF_{\mathbf{W}_{0},\mathbf{W}_{1},\dots,\mathbf{W}_{k}}(\mathbf{W}_{0},\mathbf{W}_{1},\dots,\mathbf{W}_{k}) is

πβnm/2i=0k(|𝚺i1/2𝐖i𝚺i1/2|β(nim+1)/21Γmβ[βni/2])h(βtri=0k𝚺i1/2𝐖i𝚺i1/2)\pi^{\beta nm/2}\prod_{i=0}^{k}\left(\frac{|\mathbf{\Sigma}_{i}^{-1/2}\mathbf{W}_{i}\mathbf{\Sigma}_{i}^{-1/2}|^{\beta(n_{i}-m+1)/2-1}}{\Gamma_{m}^{\beta}[\beta n_{i}/2]}\right)h\left(\beta\displaystyle\mathop{\rm tr}\nolimits\sum_{i=0}^{k}\mathbf{\Sigma}_{i}^{-1/2}\mathbf{W}_{i}\mathbf{\Sigma}_{i}^{-1/2}\right)
i=0k|𝚺i|β(m1)/21i=0k(d𝐖i).\hskip 227.62204pt\prod_{i=0}^{k}|\mathbf{\Sigma}_{i}|^{-\beta(m-1)/2-1}\bigwedge_{i=0}^{k}\left(d\mathbf{W}_{i}\right).

Hence

πβnm/2i=0k(|𝐖i|β(nim+1)/21Γmβ[βni/2]|𝚺i|βni/2)h(βtri=0k𝚺i1𝐖i)i=0k(d𝐖i),\pi^{\beta nm/2}\prod_{i=0}^{k}\left(\frac{|\mathbf{W}_{i}|^{\beta(n_{i}-m+1)/2-1}}{\Gamma_{m}^{\beta}[\beta n_{i}/2]|\mathbf{\Sigma}_{i}|^{\beta n_{i}/2}}\right)h\left(\beta\displaystyle\mathop{\rm tr}\nolimits\sum_{i=0}^{k}\mathbf{\Sigma}_{i}^{-1}\mathbf{W}_{i}\right)\bigwedge_{i=0}^{k}\left(d\mathbf{W}_{i}\right),

where 𝐖i𝔓mβ\mathbf{W}_{i}\in\mathfrak{P}^{\beta}_{m}, i=0,1,,ki=0,1,\dots,k.

5 Example

In this Section we provide an example in quaternions. Finding a random suitable data base for this algebra is difficult, then we try first to explain a way of generating a number of applications by using data bases of the literature of shape theory.

We start with a known representation of a quaternion number in terms of 2×22\times 2 complex matrices. Let q=a+b𝐢+c𝐣+d𝐤q=a+b\mathbf{i}+c\mathbf{j}+d\mathbf{k} be a quaternion, then qq can be written in terms of a the following 2×22\times 2 matrix of complex entries:

𝐙=[a+b𝐢c+d𝐢c+d𝐢ab𝐢]\mathbf{Z}=\begin{bmatrix}a+b\mathbf{i}&c+d\mathbf{i}\\ -c+d\mathbf{i}&a-b\mathbf{i}\end{bmatrix} (49)

Thus 𝐙\mathbf{Z} can be seen as an array of 4 complex points, with a double symmetry: a+b𝐢a+b\mathbf{i} and ab𝐢a-b\mathbf{i} are symmetric respect the \Re axis; meanwhile c+d𝐢c+d\mathbf{i} and c+d𝐢-c+d\mathbf{i} are symmetric about the \Im axis.

Now, shape theory deals, for example, with sets of planar figures summarised by corresponding landmarks between populations in order to obtain means, variability and discrimination via statistics on certain quotients spaces. Instead of the noisy Euclidean space, the statistics is performed with equivalent classes after filtering out some non meaning geometrical information, such as scaling, translation, rotation, reflection, etc.. Thus, a landmark data with the referred symmetries can be set into a vector variate quaternion sample for estimation of the extended real parameters ai=ni/2,i=1,,ka_{i}=n_{i}/2,i=1,\ldots,k of the distributions here derived. We focus on the mouse vertebra landmark data given for example in Dryden and Mardia [14]. The sample consists of 23 small, 23 large and 30 control second thoracic vertebrae with 60 landmarks. After transforming the data for a suitable application of the complex matrix representation, we have in figure 1 an example of an small vertebra.

Refer to caption
Figure 1: A 60 landmark small vertebra with high symmetry, divided in four parts A,B,C,DA,B,C,D for suitable application of a quaternionic complex matrix representation.

The required sample most follows the symmetric suggestions of 𝐙\mathbf{Z}. For getting this end, just cut the bones on landmarks 3030 and 4545 and two sectors are obtained: ABCABC, from landmarks 11 to 4545; and, DD, from landmarks 4646 to 6060. Now the free part DD can be placed symmetrically to sector BB (landmarks 1515 to 3030) as a reflection on the imaginary axis \Im. Finally, for each bone we have 14 pairs of landmarks (au+bu𝐢,cu+14+du+14𝐢)(a_{u}+b_{u}\mathbf{i},c_{u+14}+d_{u+14}\mathbf{i}), each one representing the quaternion qu=au+bu𝐢+cu+14𝐣+du+14𝐤q_{u}=a_{u}+b_{u}\mathbf{i}+c_{u+14}\mathbf{j}+d_{u+14}\mathbf{k}, where u=2,,15u=2,\ldots,15. Namely, for each u=2,,15u=2,\ldots,15, the first landmark au+bu𝐢a_{u}+b_{u}\mathbf{i} belongs to the sector AA (symmetric to sector DD, respect \Re axis) and it is paired with the second landmark cu+15+du+15𝐢c_{u+15}+d_{u+15}\mathbf{i} (symmetric to the translated sector CC). Summarising, the sample for the three classes of bones consists of upper landmarks 2 to 29, distributed by the pairs (2,16),(3,17),,(14,28),(15,29)(2,16),(3,17),\cdots,(14,28),(15,29). With each pair providing a quaternion, we have the following three dependent samples: 23 quaternion vectors of size 14 for the small group (Figure 2(a)), 23 quaternion vectors of size 14 for the large class (Figure 2(b)), and 30 quaternion vectors of size 14 for the control set (Figure 2(c)).

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 2: (a) Dependent sample of 23 quaternion vectors of size 14 for the small group. (b) Dependent sample of 23 quaternion vectors of size 14 for the large class. (c) Dependent sample of 30 quaternion vectors of size 14 for the control set. The sector AA (green) with complex au+bu𝐢a_{u}+b_{u}\mathbf{i} constitutes the first two components of the quaternion qu=au+bu𝐢+cu+14𝐣+du+14𝐤q_{u}=a_{u}+b_{u}\mathbf{i}+c_{u+14}\mathbf{j}+d_{u+14}\mathbf{k}, and the sector BB (red) with complex cu+15+du+15𝐢c_{u+15}+d_{u+15}\mathbf{i} indexes the last two components of qu,u=2,,15q_{u},u=2,\ldots,15.

The mouse vertebra landmark data has been studied in several works based on the classical assumptions of normality and independent probabilistic sample, see for example Dryden and Mardia [14] and the references therein. Gaussian restriction can not deal properly with the outlier shapes, meanwhile assuming a sample of independent small and large bones just facilitates the estimations via likelihood function, but it seems to be out of underlying sample extraction and population description given in the original source of the experiments back to the earlies 70s.

This work provides two solutions of the previous problems. First, within the multimatricvariate and multimatrix variate distributions, the choice of distributions that are invariant under the family of β\beta-elliptically contoured distributions. A notorious advantage facing the lack of knowledge about the random matrix law for the landmark data. And second, the distributions here derived give join density functions of dependent matrices, a fact eliminates the historical controversy between the probabilistic independence of the sample data. Another important fact, hidden for the classical studies based on Gaussianity (independently equivalent) resides on the real normed division algebra supporting the landmark data. In particular, the use of quaternions via the complex matrix representation integrates in the study the high symmetry of the bones. Finally, for this data, the distributions under quaternions are extremely simple, they are reduced to vectors instead of the real matrix setting given in the referred studies.

For the sake of illustration and simple computation, we consider the multimatrix variate beta type II distribution (34) as the likelihood function dependent sample invariant under the quaternionic elliptically contoured distribution.

For the small and large samples, k=23k=23, β=4\beta=4 and m=1m=1, meanwhile in the control group k=30k=30. In the three samples we use (34) for the maximum likelihood estimates of a0=n0/2a_{0}=n_{0}/2 and a=ni/2,i=1,,ka=n_{i}/2,i=1,\ldots,k. Here, 𝐓i\mathbf{T}_{i} is a 14×114\times 1 quaternion vector, then Fi=𝐓iH𝐓iF_{i}=\mathbf{T}_{i}^{H}\mathbf{T}_{i} is a real value for all i=1,,ki=1,\ldots,k. Thus the likelihood in (34) takes the form:

Γ1β[(a0+ka)mβ]Γ1β[a0mβ](Γmβ[aβ])ki=1kFiβ(am2+12)1(1+i=1kFi)(a0+ka)mβ.\frac{\Gamma_{1}^{\beta}\left[(a_{0}+ka)m\beta\right]}{\Gamma_{1}^{\beta}\left[a_{0}m\beta\right]\left(\Gamma_{m}^{\beta}\left[a\beta\right]\right)^{k}}\prod_{i=1}^{k}F_{i}^{\beta\left(a-\frac{m}{2}+\frac{1}{2}\right)-1}\left(1+\sum_{i=1}^{k}F_{i}\right)^{-(a_{0}+ka)m\beta}.

The computations are performed in the Optimx package of R under several methods of optimisations and a wide range of seeds for a consistent estimation. The results are given in the following table

Sample a^0\hat{a}_{0} a^\hat{a}
Small 0.040714 45.194923
Large 0.03294941 12.82131179
Control 0.03765324 32.99296063

Finally, we can use again the symmetry of the modified mouse vertebra data in order to define 14×214\times 2 quaternion matrices 𝐓i,i=1,,k\mathbf{T}_{i},i=1,\ldots,k. In this case the first column is the same formed by sectors AA and BB, and the second column corresponds to sector DD (the reflection of sector AA) and the translated and reflected sector CC, which is symmetric respect sector BB. Then we obtain the 2×22\times 2 quaternion matrix 𝐅i=𝐓iH𝐓i\mathbf{F}_{i}=\mathbf{T}^{H}_{i}\mathbf{T}_{i}, i=1,,ki=1,\dots,k, and the likelihood function (34) can be computed in terms of the latent roots of the quaternionic Hermitian matrices 𝐅i\mathbf{F}_{i}.

The use of high symmetric planar landmark data for characterising quatenion applications opens an interesting perspective in shape theory, in particular the computation of probabilities are tractable expressions which can be implemented easily. Finally, under these symmetries, (49) shows a way of avoiding the quaternions by performing a study based on block 2×22\times 2 complex matrices, a real normed division algebra more easily understood and handled because its commutative property. These aspects are taking part of a future work.

6 Conclusions

This work has set the multimatrix and multimatric variate distributions in a unified approach for the real normed division algebras. The distributions are also indexed by the class of elliptical contoured models. The main advantages of the proposed theory are in agreement with the current paradigms of the distribution theory: 1) The distributions are computable in a simple PC. 2) After integrations, the results can be seen as joint distributions of several combinations of scalars, vectors and matrix variates, some of them invariant under the family of matrix variate elliptically contoured distributions. An ideal property for situations where the marginals and joint distributions are completely unknown. 3) The multimatrix variate and multimatricvariate distributions share the philosophy of copula theory, but without the restriction to reals, vectors and likelihood copula parameter estimation based on independent distributions. 4) The join distributions can be seen as likelihood functions of probabilistically dependent matrices, as a more real alternative a likelihood function of independent sample variate. 5) The multimatrix variate and multimatricvariate distributions emerged into as unified point of view for all the real normed division algebras, just modulated by a parameter β=1,2,4,8\beta=1,2,4,8. 6) The properties presented here are valid for all real normed division algebras, then several applications can be switched according to the sample dependent origin. Finally, a application of symmetric landmark data popularised in real shape theory is translated into the quaternion setting. Current a research about multiple computation of probabilities on symmetric cones is considered.

References

  • Bekker et al. [2011] Bekker, A., Roux, J. J. J., Ehlers, E., Arashi, M. (2001). Bimatrix variate beta type IV distribution: relation to Wilks’s statistics and bimatrix variate Kummer-beta type IV distribution. Communications in Statistics - Theory and Methods, 40, 4165-4178.
  • Baez [2002] J. C. Baez, J. C. 2002. The octonions, Bulletin of the American Mathematical Society, 39, 145–205.
  • Chen and Novick [1984] Chen, J. J., Novick, M. R. (1984). Bayesian analysis for binomial models with generalized beta prior distributions. Journal of Educational Statistics, 9, 163–175.
  • Díaz-García [2014] Díaz-García, J. A., (2014). Integral properties of zonal spherical functions, hypergeometric functions and invariant polynomials. Journal of the Iranian Statistical Society, 13 (1), 83-124.
  • Díaz-García and Caro-Lopera [2022] Díaz-García, J. A., Caro-Lopera, F. J. (2022). Multimatricvariate distribution under elliptical models. Journal of Statistical Planning and Inference, 2016, 109-117.
  • Díaz-García and Caro-Lopera [2024] Díaz-García, J. A., Caro-Lopera, F. J. (2024). Multimatrix variate distribution. https://arxiv.org/abs/2405.02498.
  • Díaz-García et al. [2022] Díaz-García, J. A., Caro-Lopera, F. J., Pérez Ramírez, F. O. (2022). Multivector variate distributions: An application in Finance. Sankhyā, 84-A,Part 2, 534-555.
  • Díaz-García, and Gutiérrez-Jáimez [2010a] Díaz-García, J. A., Gutiérrez-Jáimez, R. (2010a). Bimatrix variate generalised beta distributions. South African Statistical Journal, 44, 193-208.
  • Díaz-García, and Gutiérrez-Jáimez [2010b] Díaz-García, J. A., Gutiérrez-Jáimez, R. (2010b). Complex bimatrix variate generalised beta distributions. Linear Algebra and its Applications, 432 (2-3), 571-582.
  • Díaz-García, and Gutiérrez-Jáimez [2011a] Díaz-García, J. A., Gutiérrez-Jáimez, R. (2011a). On Wishart distribution: Some extensions. Linear Algebra and its Applications, 435, 1296-1310.
  • Díaz-García, and Gutiérrez-Jáimez [2011b] Díaz-García, J. A., Gutiérrez-Jáimez, R. (2011b). Noncentral bimatrix variate generalised beta distributions. Metrika, 73(3), 317-333.
  • Dimitriu [2002] Dimitriu, I. (2002). Eigenvalue statistics for beta-ensembles. PhD thesis, Department of Mathematics. Massachusetts Institute of Technology, Cambridge, MA.
  • Dray and Manogue [1999] Dray, T., Manogue, C. A. (1999). The exceptional Jordan eigenvalue problem, International Journal of Theoretical Physics, 38(11), 2901–2916.
  • Dryden and Mardia [1998] Dryden, I. L., Mardia, K. V. (1998). Statistical Shape Analysis, Wiley, Chichester.
  • Edelman and Rao [2005] Edelman, A. Rao, R. R. (2005). Random matrix theory. Acta Numerica, 14, 233–297.
  • Ehlers [2011] Ehlers, R. (2011). Bimatrix variate distributions of Wishart ratios with application. Doctoral dissertation, Faculty of Natural & Agricultural Sciences University of Pretoria, Pretoria. http://hdl.handle.net/2263/31284.
  • Fang and Zhang [1990] Fang, K. T., Zhang, Y. T. (1990). Generalized Multivariate Analysis, Science Press, Springer-Verlag, Beijing.
  • Fang et al. [1990] Fang, K. T., Zhang, Y. T., Ng, K. W. (1990). Symmetric Multivariate and realted distributions. Springer-Science+Business Media, B. V., New Delhi.
  • Forrester [2009] Forrester, P. J. (2009). Log-gases and random matrices. To appear. Available in: http://www.ms.unimelb.edu.au/~matpjf/matpjf.html.
  • Gross and Richards [1987] Gross, K. I., Richards, D. ST, P. (1987). Special functions of matrix argument I: Algebraic induction zonal polynomials and hypergeometric functions. Transactions of the American Mathematical Society, 301(2), 475–501.
  • Gupta and Varga [1993] Gupta, A. K., Varga, T. (1993). Elliptically Contoured Models in Statistics. Kluwer Academic Publishers, Dordrecht.
  • Kabe [1984] Kabe, D. G. (1984). Classical statistical analysis based on a certain hypercomplex multivariate normal distribution. Metrika, 31, 63–76.
  • Li and Xue [2009] Li, F., Xue, Y. (2009). Zonal polynomials and hypergeometric functions of quaternion matrix argument, Communications in Statistics - Theory and Methods, 38(8), 1184-1206.
  • Libby and Novick [1982] Libby, D. L., Novick, M. R. (1982). Multivariate Generalized beta distributions with applications to utility assessment. Journal of Educational Statistics, 7, 271–294.
  • Muirhead [2005] Muirhead, R. J. (2005). Aspects of Multivariate Statistical Theory. John Wiley & Sons, New York.
  • Nadarajah [2007] Nadarajah, S. (2007). A bivariate gamma model for drought. Water Resources Research, 43, W08501, doi:10.1029/2006WR005641.
  • Nadarajah [2013] Nadarajah, S. (2013). A bivariate distribution with gamma and beta marginals with application to drought data. Journal of Applied Statistics, 36(3), 277-301.
  • Olkin and Liu [2003] Olkin, I., Liu, R. (2003). A bivariate beta distribution. Statistics and Probability Letters, 62, 407–412.
  • Olkin and Rubin [1964] Olkin, I., Rubin, H. (1964). Multivariate beta distributions and independence properties of Wishart distribution. Annals of Mathematical Statistics, 35, 261–269. Correction 1966, 37(1), 297.
  • Sarabia et al. [2014] Sarabia, J. M., Prieto, F., Jordá, V. (2014). Bivariate beta-generated distributions with application to well-being data. Journal of Statistical Distributions and Applications, 1:15. http://www.jsdajournal.com/content/1/1/15.