Skip to main content
Journal of Applied Statistics logoLink to Journal of Applied Statistics
. 2023 Jul 7;51(9):1729–1755. doi: 10.1080/02664763.2023.2232127

Family of bivariate distributions on the unit square: theoretical properties and applications

Roberto Vila a,b,CONTACT, Narayanaswamy Balakrishnan b, Helton Saulo a,c, Peter Zörnig a
PMCID: PMC11198150  PMID: 38933136

Abstract

We introduce the bivariate unit-log-symmetric model based on the bivariate log-symmetric distribution (BLS) defined in Vila et al. [25] as a flexible family of bivariate distributions over the unit square. We then study its mathematical properties such as stochastic representations, quantiles, conditional distributions, independence of the marginal distributions and marginal moments. Maximum likelihood estimation method is discussed and examined through Monte Carlo simulation. Finally, the proposed model is used to analyze some soccer data sets.

Keywords: Bivariate unit-log-symmetric distribution, bivariate log-symmetric distribution, MCMC, proportion data, soccer data, maximum likelihood estimation

2010 Mathematics Subject Classifications: 60E05, 62Exx, 62Fxx

1. Introduction

Bivariate distributions over the unit-square have been discussed in detail in the literature; see, e.g. Barreto-Souza and Lemonte [4] and Özbilen and Genç [20]. Many of them are based on the beta distribution and its generalizations; see Arnold and Ng [2] and Nadarajah et al. [18]. Models of this type have been studied since the 1980s. Some other distributions on the unit square are based on generalized arcsine and inverse Gaussian distributions. A recent model, the bivariate unit-sinh-normal distribution, is based on the bivariate Birnbaum-Saunders distribution; see Martinez-Flóres et al. [16]. Bivariate distributions over the unit square arise naturally in comparing indices, rates or proportions in the interval (0,1).

In this paper, we study the bivariate unit-log-symmetric (BULS) distribution defined over the unit-square, obtained as a modification of the bivariate log-symmetric (BLS) distribution introduced by Vila et al. [25]. The definitions of BLS and BULS distributions are given in Section 2, along with some special cases of the BULS. In Section 3, we discuss some properties of the new model, including a stochastic representation, marginal quantiles, and the conditional distributions of BULS. We derive compact formulas for the conditional densities, using the distribution functions of normal, Student-t, hyperbolic, Laplace and slash distributions. One of the uses of having closed formulas for the conditional densities (of the BULS model), for example, is in studying Heckman-type selection models (see Heckman [11]) when the selection variables have bounded support. In addition, we derive the distribution of the squared Mahalanobis distance of a random vector W=(W1,W2) with BULS distribution, and present a necessary condition for the independence of the components of W and formulas for the moments of W1 and W2. In Section 4, the log-likelihood function and the likelihood equations for the BULS distribution are presented. In Section 5, we carry out a Monte Carlo simulation study to evaluate the performance of the ML estimators by means of their bias, root mean square error and coverage probability. In Section 6, we present two applications to soccer data. Specifically, in Section 6.1, we model the vector W=(W1,W2), where W1 represents the time elapsed until a first kick goal (of any team) and the time elapsed until a goal of any type of the home team, and show that specific BULS distributions are suitable for modelling W. In Section 6.2, we consider the data of 2022 FIFA World Cup wherein the components of the vector W represent the pass completion proportions of medium passes (14 to 18 meters) and long passes (longer than 37 meters). We then demonstrate that these data can also be fitted well by BULS distributions.

2. Bivariate unit-log-symmetric model

In this section, we describe the bivariate unit-log-symmetric model (BULS). To define this model, we first need to describe the bivariate log-symmetric distribution (BLS) defined in Vila et al. [25].

2.1. BLS family of distributions

Following Vila et al. [25], a continuous random vector T=(T1,T2) is said to have a bivariate log-symmetric (BLS) distribution if its joint probability density function (PDF) is given by

fT1,T2(t1,t2;θ)=1t1t2σ1σ21ρ2Zgcgc(t1~22ρt1~t2~+t2~21ρ2),t1,t2>0, (1)

where ti~=log(ti/ηi)1/σi,ηi=exp(μi),i=1,2, with θ=(η1,η2,σ1,σ2,ρ) being the parameter vector, μiR, σi>0, i=1,2 and ρ(1,1). Furthermore, Zgc>0 is the partition function, that is,

Zgc=001t1t2σ1σ21ρ2gc(t1~22ρt1~t2~+t2~21ρ2)dt1dt2=π0gc(u)du, (2)

and gc is a scalar function referred to as the density generator (see Fang et al. [8]). The second integral in (2) is a consequence of a change of variables; for more details, see Proposition 3.1 of Vila et al. [25]. When a random vector T is BLS distributed, with parameter vector θ, we denote it by TBLS(θ,gc).

2.2. BULS family of distributions

We say that a continuous random vector W=(W1,W2) has a bivariate unit-log-symmetric (BULS) distribution with parameter vector θ=(η1,η2,σ1,σ2,ρ), denoted by WBULS(θ,gc), if its PDF is, for 0<w1,w2<1, given by

fW1,W2(w1,w2;θ)=1(1w1)t1σ1(1w2)t2σ21ρ2Zgcgc(w~122ρw~1w~2+w~221ρ2), (3)

where w~i=log(ti/ηi)1/σi,ti=log(1wi),ηi=exp(μi),i=1,2, with σi>0, i=1,2, ρ(1,1), and Zgc and gc are as given in (2). We shall prove later that the BULS PDF in (3) is obtained by taking Wi=1exp(Ti), i=1,2, with (T1,T2)BLS(θ,gc).

Table 1 presents some examples of bivariate unit-log-symmetric distributions.

Table 1.

Partition functions (Zgc) and density generators (gc) for some BULS distributions.

Distribution Zgc gc Parameter
Bivariate unit-log-normal 2π exp(x/2)
Bivariate unit-log-Student-t Γ(ν/2)νπΓ((ν+2)/2) (1+xν)(ν+2)/2 ν>0
Bivariate unit-log-hyperbolic 2π(ν+1)exp(ν)ν2 exp(ν1+x) ν>0
Bivariate unit-log-Laplace π K0(2x)
Bivariate unit-log-slash πq22q2 xq+22γ(q+22,x2) q>0

In Table 1, Γ(t)=0xt1exp(x)dx, t>0, is the complete gamma function, Kλ(u)=(1/2)(u/2)λ0tλ1exp(tu2/4t)dt, u>0, is the modified Bessel function of the third kind with index λ (see Appendix of Kotz et al. [14]), and γ(s,x)=0xts1exp(t)dt is the lower incomplete gamma function.

Let T=(T1,T2)BLS(θ,gc). From (3), it is clear that the random vector X=(X1,X2), with

Xi=log(Ti)=log[log(1Wi)],i=1,2, (4)

has a bivariate elliptically symmetric (BSY) distribution (see p. 592 in Balakrishnan and Lai [3]); that is, the PDF of X is

fX1,X2(x1,x2;θ)=1σ1σ21ρ2Zgcgc(x1~22ρx1~x2~+x2~21ρ2),<x1,x2<, (5)

where xi~=(xiμi)/σi,i=1,2, with θ=(μ1,μ2,σ1,σ2,ρ) being the parameter vector and Zgc is the partition function defined in (2). In this case, we shall use the notation XBSY(θ,gc).

It is a simple task to observe that the joint cumulative distribution function (CDF) of WBULS(θ,gc), denoted by FW1,W2(w1,w2;θ), is given by

FW1,W2(w1,w2;θ)=FT1,T2(log(1w1),log(1w2);θ)=FX1,X2(log[log(1w1)],log[log(1w2)];θ),

wherein FT1,T2(t1,t2;θ) and FX1,X2(x1,x2;θ) denote the CDFs of TBLS(θ,gc) and XBES(θ,gc), respectively.

Remark 2.1

As the components of W are defined as increasing functions of the components of X, by invariance under monotone transformations property of copulas, both will have the same associated copula (for a review of copula theory, see, e.g. Nelsen [19]). Therefore, the analysis of dependence measures for the variables W1 and W2 that depend only on the copula is equivalent to the analysis of those for X1 and X2. For the normal copula and Student's t copula, some dependency measures such as Kendall's tau, Spearman's rho, Blest's measure of rank correlation, and coefficients of tail dependence were all studied by Roncalli [21].

3. Some basic properties of the model

In this section, some mathematical properties of the bivariate unit-log-symmetric distribution are established.

3.1. Stochastic representation

Proposition 3.1

The random vector W=(W1,W2) has a BULS distribution if

W1=1exp[η1exp(σ1Z1)],W2=1exp[η2exp(σ2ρZ1+σ21ρ2Z2)],

where Z1=RDU1 and Z2=R1D2U2 with U1, U2, R, and D being mutually independent random variables, ρ(1,1), ηi=exp(μi), and P(Ui=1)=P(Ui=1)=1/2, i=1,2. The random variable D is positive and has PDF fD(d)=2/(π1d2),d(0,1). Further, the positive random variable R has its PDF as fR(r)=2rgc(r2)/0gc(u)du,r>0.

Proof.

It is well-known that (see Proposition 3.2 of Vila et al. [25]) the random vector T=(T1,T2) has a BLS distribution if

T1=η1exp(σ1Z1),T2=η2exp(σ2ρZ1+σ21ρ2Z2). (6)

Moreover, from (4), Wi=1exp(Ti), i=1,2. Hence, the result.

The following lemma provides a slight simplification in the representation of Proposition 3.1. This result plays a role in the next subsections, since all the probabilistic characteristics that depend on the distribution of ρZ1+1ρ2Z2 will be simplified since it has the same distribution as Z2.

Lemma 3.2

For a Borel subset B of (,), we have

P(ρZ1+1ρ2Z2B)=P(Z2B).

In other words, ρZ1+1ρ2Z2 and Z2 have the same distribution.

Proof.

It is clear that the density of ρZ1+1ρ2Z2 is related to the joint density fZ1,Z2 by

fρZ1+1ρ2Z2(s2)=11ρ2fZ1,Z2(z,s2ρz1ρ2)dz. (7)

From Equation (13) of Saulo et al. [23], the joint PDF of Z1 and Z2 is given by

fZ1,Z2(x,y)=1Zgcgc(x2+y2),<x,y<, (8)

and so the integral in (7) is

=11ρ2Zgcgc(z2+(s2ρz1ρ2)2)dz. (9)

Using the identity

z2+(s2ρz1ρ2)2=z22ρzs2+s221ρ2=(zρs21ρ2)2+s22

the integral in (9) can be expressed as

11ρ2Zgcgc((zρs21ρ2)2+s22)dz.

Making the change of variable s1=(zρs2)/1ρ2, the above integral becomes

1Zgcgc(s12+s22)ds1=fZ1,Z2(s1,s2)ds1,

where, in the last line, we have used (8). Hence,

fρZ1+1ρ2Z2(s2)=fZ1,Z2(s1,s2)ds1=fZ2(s2). (10)

Now, from (10), it is clear that ρZ1+1ρ2Z2 and Z2 are equal in distribution.

3.2. Marginal quantiles

Given p(0,1), let QWi(p) be the p-quantile of Wi, for i=1,2. By using the stochastic representation in Proposition 3.1, for W=(W1,W2)BULS(θ,gc), we have

p=P(W1QW1(p))=P(1exp[η1exp(σ1Z1)]QW1(p))=P(Z1log(log(1QW1(p))η1)1/σ1)

and

p=P(W2QW2(p))=P(1exp[η2exp(σ2ρZ1+σ21ρ2Z2)]QW2(p))=P(ρZ1+1ρ2Z2log(log(1QW2(p))η2)1/σ2).

Hence, the p-quantiles QZ1(p) and QZ2(p) of Z1 and Z2, respectively, are given by

log(log(1QW1(p))η1)1/σ1=QZ1(p)

and

log(log(1QW2(p))η2)1/σ2=QρZ1+1ρ2Z2(p)=QZ2(p),

where in the last equality we have used the fact that ρZ1+1ρ2Z2 and Z2 have the same distribution (see Lemma 3.2). Hence, the p-quantiles QW1(p) and QW2(p) are given by

QW1(p)=1exp[η1exp(σ1QZ1(p))],QW2(p)=1exp[η2exp(σ2QZ2(p))],

respectively.

3.3. Conditional distributions

Before enunciating and proving the main result (Theorem 3.4) of this subsection, we establish the following technical lemma which gets used subsequently.

Lemma 3.3

If W=(W1,W2)BULS(θ,gc), then the PDF of W2|(W1=w1) is given by

fW2(w2|W1=w1)=1(1w2)t2σ21ρ2fZ2(11ρ2(w~2ρw~1)|Z1=w~1), (11)

where w~i, i=1,2, and t2 are as defined in (3), and Z1 and Z2 are as given in Proposition 3.1.

Proof.

If W1=w1, then Z1=log[log(1w1)/η1]1/σ1=w~1. So, the conditional distribution of W2, given W1=w1, is the same as the distribution of

1exp[η2exp(σ2ρw~1+σ21ρ2Z2)]|W1=w1.

Consequently,

FW2(w2|W1=w1)=P(1exp[η2exp(σ2ρw~1+σ21ρ2Z2)]w2|W1=w1)=P(Z211ρ2(w~2ρw~1)|Z1=w~1).

Then, by differentiating FW2(w2|W1=w1) with respect to w2, (11) is readily obtained.

The following result provides a simple formula for determining the conditional distribution of W1, given W2B, whenever the marginal and conditional distributions of (Z1,Z2) are known. This result is essential for studying Heckman-type selection models (see Heckman [11]) when the selection variables have unitary support.

Theorem 3.4

For a Borel subset B of (0,1), let us define the following Borel set:

Br=11r2log(log(1B)η2)1/σ2r1r2w~1,1<r<1, (12)

where w~1 is as in (3). If WBULS(θ,gc), then the PDF of W1|(W2B) is given by

fW1(w1|W2B)=1(1w1)t1σ1fZ1(w~1)P(Z2Bρ|Z1=w~1)P(Z2B0),

in which t1 is as in (3), Br is as in (12), and Z1 and Z2 are as given in Proposition 3.1.

Proof.

Let B be a Borel subset of (0,1). Note that

fW1(w1|W2B)=fW1(w1)BfW2(w2|W1=w1)dw2P(W2B).

As fW1(w1)=fZ1(w~1)/[(1w1)t1σ1] and P(W2B)=P(ρZ1+1ρ2Z2B0), where B0 is as given in (12) with r=0, the term on the right-hand side of the above identity becomes

1(1w1)t1σ1fZ1(w~1)BfW2(w2|W1=w1)dw2P(ρZ1+1ρ2Z2B0).

By using the formula for fW2(w2|W1=w1) provided in Lemma 3.3, the above expression becomes

1(1w1)t1σ1σ21ρ2fZ1(w~1)×B1(1w2)t2fZ2(11ρ2w~2ρ1ρ2w~1|Z1=w~1)dw2P(ρZ1+1ρ2Z2B0),

where w~i and ti, i=1,2, are as in (3). Finally, by applying the change of variable z=(w~2ρw~1)/1ρ2, the above expression is

=1(1w1)t1σ1fZ1(w~1)BρfZ2(z|Z1=w~1)dzP(ρZ1+1ρ2Z2B0).

We have thus proved that

fW1(w1|W2B)=1(1w1)t1σ1fZ1(w~1)BρfZ2(z|Z1=w~1)dzP(ρZ1+1ρ2Z2B0).

Finally, by combining the above identity with Lemma 3.2, the required result is obtained.

Using Theorem 3.4, for each generator (gc) in Table 1, we present closed formulas for the conditional densities of W1|(W2B) corresponding to bivariate unit-log-normal (Corollary 3.5), bivariate unit-log-Student-t (Corollary 3.6), bivariate unit-log-hyperbolic (Corollary 3.7), bivariate unit-log-Laplace (Corollary 3.8) and bivariate unit-log-slash (Corollary 3.9) distributions.

Corollary 3.5 Gaussian generator —

Let W=(W1,W2)BULS(θ,gc) and gc(x)=exp(x/2) be the generator of the bivariate unit-log-normal distribution. Then, for each Borel subset B of (0,1), the PDF of W1|(W2B) is given by (for 0<w1<1)

fW1(w1|W2B)=1(1w1)t1σ1ϕ(w~1)Φ(Bρ)Φ(B0),

where Φ(C)=Cϕ(x)dx and ϕ(x) is the standard normal PDF. Further, w~1 and t1 are as in (3), and Br is as in (12).

Proof.

It is well-known that the bivariate log-normal distribution has a stochastic representation as in (6), where Z1N(0,1) and Z2N(0,1), and Z2|(Z1=x)N(0,1) (see Abdous [1]). Hence, P(Z2B0)=Φ(B0) and P(Z2Bρ|Z1=w~1)=Φ(Bρ). Then, by applying Theorem 3.4, the required result follows.

Corollary 3.6 Student-t generator —

Let W=(W1,W2)BULS(θ,gc) and gc(x)=(1+(x/ν))(ν+2)/2, ν>0, be the generator of the bivariate unit-log-Student-t distribution with ν degrees of freedom. Then, for each Borel subset B of (0,1), the PDF of W1|(W2B) is given by (for 0<w1<1)

fW1(w1|W2B)=1(1w1)t1σ1fν(w~1)Fν+1(ν+1ν+w~12Bρ)Fν(B0),

where Fν(C)=Cfν(x)dx and fν(x) is the standard Student-t PDF with ν degrees of freedom.

Proof.

It is well-known that the bivariate log-Student-t distribution has a stochastic representation as in (6), where Z1tν and Z2tν (Student-t with ν degrees of freedom), and (see Corollary 3.7 of Vila et al. [25])

Z2|(Z1=x)ν+x2ν+1tν+1.

Hence, P(Z2B0)=Fν(B0) and

P(Z2Bρ|Z1=w~1)=Fν+1(ν+1ν+w~12Bρ).

By applying Theorem 3.4, the required result follows.

Corollary 3.7 Hyperbolic generator —

Let W=(W1,W2)BULS(θ,gc) and gc(x)=exp(ν1+x) be the generator of the bivariate unit-log-hyperbolic distribution. Then, for each Borel subset B of (0,1), the PDF of W1|(W2B) is given by (for 0<w1<1)

fW1(w1|W2B)=1(1w1)t1σ1fGH(w~1;3/2,ν,1)FGH(Bρ;1,ν,1+w~12)FGH(B0;3/2,ν,1),

where FGH(C;λ,α,δ)=CfGH(x;λ,α,δ)dx and fGH(x;λ,α,δ) is the generalized hyperbolic (GH) PDF (see Definition A.1 in the Appendix).

Proof.

It is well-known that the bivariate log-hyperbolic distribution has a stochastic representation as in (6), where Z1GH(3/2,ν,1) and Z2GH(3/2,ν,1) (see Subsection 2.1 of Deng and Yao [7]). Moreover, the distribution of Z2, given Z1=x, is GH(1/2,2,|x|) (Proposition A.2). Then, P(Z2B0)=FGH(B0;3/2,ν,1) and P(Z2Bρ|Z1=w~1)=FGH(Bρ;1,ν,1+w~12). By applying Theorem 3.4, the required result follows.

Corollary 3.8 Laplace generator —

Let W=(W1,W2)BULS(θ,gc) and gc(x)=K0(2x) be the generator of the bivariate unit-log-Laplace distribution. Then, for each Borel subset B of (0,1), the PDF of W1|(W2B) is given by (for 0<w1<1)

fW1(w1|W2B)=1(1w1)t1σ1fL(w~1)FGH(Bρ;12,2,|w~1|)FL(B0),

where FL(C)=CfL(x)dx and fL(x)=exp(2|x|)/2 is the Laplace PDF with scale parameter 1/2, and FGH is as defined in Corollary 3.7.

Proof.

It is well-known that the bivariate log-Laplace distribution has a stochastic representation as in (6), where Z1Laplace(0,1/2) and Z2Laplace(0,1/2) (see Subsection 5.1.4 of Kotz et al. [14]). Further, the distribution of Z2, given Z1=x, is GH(1/2,2,|x|) (Proposition A.3). Hence, P(Z2B0)=FL(B0) and P(Z2Bρ|Z1=w~1)=FGH(Bρ;1/2,2,|w~1|). By applying Theorem 3.4, the required result follows.

Corollary 3.9 Slash generator —

Let W=(W1,W2)BULS(θ,gc) and gc(x)=x(q+2)/2γ((q+2)/2,x/2), be the generator of the bivariate unit-log-slash distribution. Then, for each Borel subset B of (0,1), the PDF of W1|(W2B) is given by (for 0<w1<1)

fW1(w1|W2B)=1(1w1)t1σ1fSL(w~1;q)FESL(Bρ;w~1,q+1)FSL(B0;q),

where FSL(C;q)=CfSL(x;q)dx and fSL(x;q)=q01tqϕ(tx)dt is the classical slash PDF, and FESL(C;a,q)=CfESL(x;a,q)dx, where fESL(x;a,q) is the generalized hyperbolic (ESL) PDF (see Definition A.4 in the Appendix).

Proof.

It is well-known that the bivariate log-slash distribution has a stochastic representation as in (6), where Z1SL(q) and Z2SL(q) (see Section 2 of Wang and Genton [26]). Moreover, the distribution of Z2, given Z1=x, is ESL(x,q+1) (Proposition A.5). Hence, P(Z2B0)=FSL(B0;q) and P(Z2Bρ|Z1=w~1)=FESL(Bρ;w~1,q+1). By applying Theorem 3.4, the required result follows.

Table 2 below presents some examples of conditional PDFs corresponding to all the bivariate unit-log-symmetric distributions presented earlier in Table 1.

Table 2.

Conditional densities of W1|(W2B) and density generators (gc) for some BULS distributions.

Distribution gc fW1(w1|W2B)
Bivariate unit-log-normal exp(x/2) 1(1w1)t1σ1ϕ(w~1)Φ(Bρ)Φ(B0)
Bivariate unit-log-Student-t (1+xν)(ν+2)/2 1(1w1)t1σ1fν(w~1)Fν+1(ν+1ν+w~12Bρ)Fν(B0)
Bivariate unit-log-hyperbolic exp(ν1+x) 1(1w1)t1σ1fGH(w~1;3/2,ν,1)FGH(Bρ;1,ν,1+w~12)FGH(B0;3/2,ν,1)
Bivariate unit-log-Laplace K0(2x) 1(1w1)t1σ1fL(w~1)FGH(Bρ;12,2,|w~1|)FL(B0)
Bivariate unit-log-slash xq+22γ(q+22,x2) 1(1w1)t1σ1fSL(w~1;q)FESL(Bρ;w~1,q+1)FSL(B0;q)

3.4. Independence

Proposition 3.10

Let W=(W1,W2)BULS(θ,gc). If ρ=0 and the density generator gc in (3) is such that

gc(x2+y2)=gc1(x2)gc2(y2),(x,y)R2, (13)

for some density generators gc1 and gc2, then W1 and W2 are independent.

Proof.

The proof follows the same steps as the proof of Proposition 3.11 of Vila et al. [25]. For the sake of completeness, however, we present it here.

Let ρ=0. From (13), the joint density (3) of (W1,W2) is such that

fW1,W2(w1,w2;θ)=Zgc1Zgc2Zgcf1(w1;μ1,σ1)f2(w2;μ2,σ2),(w1,w2)(0,1)×(0,1), (14)

where fi(wi;μi,σi)=gci(wi~2)/[(1wi)tiσiZgci],0<wi<1, Zgci=gci(zi2)dzi,i=1,2, and wi~ and ti are as in (3). Integrating (14) in terms of w1 and w2, we obtain

Zgc1Zgc2Zgc=1,

and consequently, Zgc=Zgc1Zgc2. Therefore,

fW1,W2(w1,w2;θ)=f1(w1;μ1,σ1)f2(w2;μ2,σ2),(w1,w2)(0,1)×(0,1).

Moreover, it is easy to verify that f1 and f2 are PDFs corresponding to univariate symmetric random variables (see Vanegas and Paula [24]). Then, W1 and W2 are statistically independent, and even more, fi=fWi, for i=1,2 (see Proposition 2.5 of James [13]).

Remark 3.11

In Table 1, the density generator of the bivariate unit-log-normal is the unique one that satisfies (13).

3.5. Marginal moments and correlation function

For W=(W1,W2)BULS(θ,gc), 0<Wi<1, it is clear that 0E(Wir)1, for any r>0 and i=1,2. Therefore, the positive moments of Wi always exist.

In general, for any rR, the moments of Wi, i=1,2, admit the following representations:

μ1(r)E(W1r)=E{1exp[η1exp(σ1Z1)]}r,μ2(r)E(W2r)=E(1exp{η2exp(σ2[ρZ1+1ρ2Z2])})r=E{1exp[η2exp(σ2Z2)]}r, (15)

where in the last equality we have used the fact that ρZ1+1ρ2Z2 and Z2 have the same distribution (see Lemma 3.2). Here, Z1 and Z2 are as given in Proposition 3.1.

Closed-forms expressions for the product moments can be derived as follows. Law of total expectation gives

μ1,2E(W1W2)=01w1[01w2fW2(w2|W1=w1)dw2]fW1(w1)dw1. (16)

Lemma 3.3 provides a formula for the PDF of W2|(W1=w1). Moreover, fW1(w1)=fZ1(w~1)/[(1w1)t1σ1]. By using these formulas in (16), the product moments μ1,2 is equal to

1σ1σ21ρ201w1(1w1)t1[01w2(1w2)t2fZ2(11ρ2(w~2ρw~1)|Z1=w~1)dw2]fZ1(w~1)dw1,

where w~i and ti, for i=1,2, are as defined in (3).

The conditional and unconditional distributions of vector (Z1,Z2) corresponding to bivariate models of Table 1 are well-known (see proofs of Corollaries 3.5–3.9). Therefore, at least numerically, the respective product moments can be determined.

For illustrative purposes, we now present the product moments and correlation function formula only for the bivariate unit-log-normal model. In this case, Z1N(0,1), Z2N(0,1) and Z2|(Z1=x)N(0,1) (see proof of Corollary 3.5). So, using the last formula above for μ1,2, we have

μ1,2=1σ1σ21ρ201w1(1w1)t1[01w2(1w2)t2ϕ(11ρ2(w~2ρw~1))dw2]ϕ(w~1)dw1.

Consequently, the correlation function between W1 and W2, denoted by ρW, can be expressed as

ρW=1σ1σ21ρ201w1(1w1)t1[01w2(1w2)t2ϕ(11ρ2(w~2ρw~1))dw2]ϕ(w~1)dw1μ1(1)μ2(1)[μ1(2)μ12(1)][μ2(2)μ22(1)], (17)

where μ1(r) and μ2(r) are as given in (15), and w~i and ti, for i=1,2, are as defined in (3).

Table 3 shows some values of the correlation function in (17) for different choices of ρ in the special case when μ1=μ2=0 and σ1=σ2=1.

Table 3.

Correlation values for different choices of ρ.

ρ −0.9000 −0.7500 −0.5000 −0.2500 −0.1000 0.0000 0.1000 0.2500 0.5000 0.7500 0.9000
ρW −0.9413 −0.8406 −0.6678 −0.5275 −0.4481 −0.3929 −0.2880 −0.1463 0.1683 0.5339 0.7648

Remark 3.12

Given the first two moments of BULS model, some useful bounds for ρW can be provided. For example,

  1. Theorem 1 of Hössjer and Sjölander [12] gives
    min{μ1(1)μ2(1),[1μ1(1)][1μ2(1)]}[μ1(2)μ12(1)][μ2(2)μ22(1)]ρWmin{μ1(1)[1μ2(1)],[1μ1(1)]μ2(1)}[μ1(2)μ12(1)][μ2(2)μ22(1)];
  2. Corollary 3 of Hössjer and Sjölander [12] provides
    14[μ1(2)μ12(1)][μ2(2)μ22(1)]ρW14[μ1(2)μ12(1)][μ2(2)μ22(1)].

3.6. Squared mahalanobis distance

Given a probability distribution F on R2 with mean μ=(μ1,μ2), μiR, and positive-definite covariance matrix Σ, and given two points x and y in R2, the squared Mahalanobis distance between them with respect to F is

d2(x,y)d2(x,y;F)=(xy)Σ1(xy).

Let W=(W1,W2)BULS(θ,gc) with location (mean) vector μW=(μ1(1),μ2(1)) and covariance matrix

ΣW=(σW12ρWσW1σW2ρWσW1σW2σW22),

where σW12μ1(2)μ12(1), σW22μ2(2)μ22(1), μ1(r) and μ2(r) are the moments given in (15), and ρW is as given in (17). The (random) squared Mahalanobis distance of W from μW is then

d2d2(W,μW)=(WμW)ΣW1(WμW)=11ρW2[(W1μ1(1)σW1)22ρW(W1μ1(1)σW1)(W2μ2(1)σW2)+(W2μ2(1)σW2)2].

The following two results provide formulas for the PDF and CDF of d2. The respective proofs can be found in the Appendix (Section 1).

Proposition 3.13

If W=(W1,W2)BULS(θ,gc), then the CDF of d2d2(W,μW), denoted by Fd2, is given by

Fd2(w)=kZgcww[ws12ws121(1w1,s)t1,s(1w2,s)t2,sgc(w~1,s2+11ρ2(w~2,sρw~1,s)2)ds2]ds1,

with Zgc being as in (5), k=σW1σW21ρW2/(σ1σ21ρ2), and

w1,s=μ1(1)+σW1s1,w2,s=μ2(1)+σW2[ρWs1+1ρW2s2],w~i,s=log(ti,s/ηi)1/σi,ηi=exp(μi),ti,s=log(1wi,s),i=1,2. (18)

Proposition 3.14

If W=(W1,W2)BULS(θ,gc), then the PDF of d2d2(W,μW), denoted by fd2, is given by

fd2(w)=k2Zgcww1/ws12(1w1,s)t1,s(1xs1)[log(1xs1)]gc(w~1,s2+11ρ2[log1/σ2(log(1xs1)η2)ρw~1,s]2)ds1,

where xs1=μ2(1)+σW2[ρWs1+1ρW2ws12], Zgc is as given in (5), k=σW1σW21ρW2/(σ1σ21ρ2), and w~1,s is as in (18).

4. Maximum likelihood estimation of model parameters

Let {(W1i,W2i):i=1,,n} be a bivariate random sample of size n from the BULS(θ,gc) distribution with PDF as in (3), and let (w1i,w2i) be the corresponding observations of (W1i,W2i). Then, the log-likelihood function for θ=(η1,η2,σ1,σ2,ρ), without the additive constant, is given by

(θ)=ni=12log(σi)n2log(1ρ2)+i=1nloggc(w~1i22ρw~1iw~2i+w~2i21ρ2),0<w1i,w2i<1,

where w~ki=log(tki/ηk)1/σk,tki=log(1wki)>0 and ηk=exp(μk),k=1,2;i=1,,n.

In the case when a supremum θ^=(η1^,η2^,σ1^,σ2^,ρ^) exists, it must satisfy the following likelihood equations:

(θ)η1|θ=θ^=0,(θ)η2=0,(θ)σ1|θ=θ^=0,∂ℓ(θ)σ2|θ=θ^=0,∂ℓ(θ)∂ρ|θ=θ^=0, (19)

with

(θ)η1=2σ1η1(1ρ2)i=1n(ρw~2iw~1i)G(w~1i,w~2i),∂ℓ(θ)η2=2σ2η2(1ρ2)i=1n(ρw~1iw~2i)G(w~1i,w~2i),∂ℓ(θ)σ1=nσ1+2σ1(1ρ2)i=1nw~1i(ρw~2iw~1i)G(w~1i,w~2i),∂ℓ(θ)σ2=nσ2+2σ2(1ρ2)i=1nw~2i(ρw~1iw~2i)G(w~1i,w~2i),∂ℓ(θ)∂ρ=1ρ22(1ρ2)2i=1n(ρw~1iw~2i)(ρw~2iw~1i)G(w~1i,w~2i), (20)

where we have used the notation

G(w~1i,w~2i)=gc(xρ,i)gc(xρ,i), (21)

with xρ,i=(w~1i22ρw~1iw~2i+w~2i2)/(1ρ2),i=1,,n.

Observe that the likelihood equations in (19) can be written as

i=1nw~1iG(w~1i,w~2i)|θ=θ^=0,i=1n(w~1i2w~2i2)G(w~1i,w~2i)|θ=θ^=0,i=1nw~2i[2ρw~2i(1+ρ2)w~1i]G(w~1i,w~2i)|θ=θ^=nρ^(1ρ^2)2.

Any nontrivial root θ^ of the above likelihood equations is an ML estimator in the loose sense. When the parameter value provides the absolute maximum of the log-likelihood function, it becomes the ML estimator in the strict sense.

In the following proposition, we discuss the existence of the ML estimator ρ^ when all other parameters are known.

Proposition 4.1

Let gc be a density generator such that

gc(x)=r(x)gc(x),<x<, (22)

for some real-valued function r(x) with limρ±1r(xρ,i)=c(,0), where xρ,i, i=1,,n, are as in (21). If the parameters η1,η2,σ1 and σ2 are all known, then (20) has at least one root in the interval (1,1).

Proof.

The proof of this result follows by a direct application of Intermediate value theorem. For more details, see Proposition 5.1 of Vila et al. [25].

By using Morse theory, Mäkeläinen et al. [15] established that, under some regularity conditions, there is a unique MLE for θ. For the BULS model, no closed-form solution to the maximization problem is available, and the MLE can only be found by means of numerical optimization. Under mild regularity conditions (Cox and Hinkley [5] and Davison [6]), the asymptotic distribution of the ML estimator θ^ of θ is as follows: n(θ^θ)DN(0,I1(θ)), where 0 is the zero mean vector and I1(θ) is the inverse expected Fisher information matrix. The main use of the last convergence is to construct confidence regions and to perform hypothesis testing for θ (see Davison [6]).

4.1. Residual analysis

To assess the goodness of fit and departures from the assumptions of the model, we consider the stochastic relation

Resid(W)=W~122ρW~1W~2+W~221ρ2, (23)

where the random vector W=(W1,W2) follows a BULS distribution, with W~i=log(Ti/ηi)1/σi,Ti=log(1Wi), and ηi=exp(μi),i=1,2. The corresponding CDF and PDF of (23) are given respectively by

FResid(W)(x)=4Zgc0x[0xz12gc(z12+z22)dz2]dz1,x>0,fResid(W)(x)=πZgcgc(x),x>0,

where Zgc is as in (2).

For example, upon taking gc(x)=exp(x/2) and Zgc=2π (see Table 1), we get Resid(W)χ22 (chi-square with 2 degrees of freedom). Next, upon taking gc(x)=(1+(x/ν))(ν+2)/2 and Zgc=Γ(ν/2)νπ/Γ((ν+2)/2) (see Table 1), we have Resid(W)2F2,ν, where F2,ν denotes the F-distribution with 2 and ν degrees of freedom. Therefore, we can use the relation in (23) to check the goodness of fit, contrasting the empirical distribution against the theoretical one. Specifically, quantile-quantile (QQ) plots and the Kolmogorov-Smirnov test can then be used to assess the fit.

5. Simulation study

In this section, we carry out a Monte Carlo simulation study for evaluating the performance of the ML estimators of the parameters of BULS distribution. For illustrative purposes, we only present the results for the bivariate unit-log-normal model. The simulation scenario considered is as follows: 1,000 Monte Carlo replications, sample size n(25,100,500,700), vector of true parameters (η1,η2,σ1,σ2)=(1,1,0.5,0.5), ρ{0,0.25,0.5,0.75,0.95} (negative values of ρ produce the same results and so are omitted). To study the performance of the ML estimators, we computed the bias, root mean square error (RMSE), and coverage probability (CP) as

Bias^(θ^)=1Ni=1Nθ^(i)θ,RMSE^(θ^)=1Ni=1N(θ^(i)θ)2,CP^(θ^)=1Ni=1NI(θ[Lθ^(i),Uθ^(i)]),

where θ and θ^(i) are the true parameter value and its i-th ML estimate, N is the number of Monte Carlo replications, I is an indicator function taking the value 1 if θ[Lθ^(i),Uθ^(i)], and 0 otherwise, where Lθ^(i) and Uθ^(i) are the i-th upper and lower limit estimates of the 95% confidence interval. We expect that, as the sample size increases, the bias and RMSE would decrease, and the CP would approach the 95% nominal level.

The obtained simulation results are presented in Figure 1. We observe that the results obtained for the chosen bivariate unit-log-normal distribution are as expected in that as the sample size increases, the bias and RMSE both decrease and that the CP approaches the 95% nominal level. Finally, in general, the results do not seem to depend on the parameter ρ.

Figure 1.

Figure 1.

Monte Carlo simulation results for the bivariate unit-log-normal model.

6. Application to soccer data

In this section, two real soccer data sets, corresponding to times elapsed until scored goals of UEFA Champions League and pass completions of 2022 FIFA World Cup, are analyzed. The UEFA Champions League data set was extracted from Meintanis [17], while the 2022 FIFA World Cup data set is new and is analyzed for the first time here.

6.1. UEFA champions league

We consider a bivariate data set on the group stage of the UEFA Champions League for the seasons 2004/05 and 2005/06. Only matches with at least one goal scored directly from a kick by any team, and with at least one goal scored by the home team, are considered here; see Meintanis [17]. The first variable (W1) is the time (in minutes) elapsed until a first kick goal is scored by any team, and the second one W2 is the time (in minutes) elapsed until a first goal of any type is scored by the home team. The times are divided by 90 minutes (full game time) to obtain data on the unit square (0,1)×(0,1); see Table A1.

Table 4 provides descriptive statistics for the variables W1 and W2, including minimum, median, mean, maximum, standard deviation (SD), coefficient of variation (CV), coefficient of skewness (CS), and coefficient of kurtosis (CK). We observe in the variable W1, the mean and median to be, respectively, 0.454 and 0.456, i.e. the mean is almost equal to the median, which indicates symmetry in the data. The CV is 49.274%, which means a moderate level of dispersion is present around the mean. Furthermore, the CS value also confirms the symmetry nature. The variable W2 has mean 0.365 and median 0.311, which indicates a small positively skewed feature in the distribution of the data. Moreover, the CV value is 69.475%, showing a moderate level of dispersion around the mean. The CS confirms the small skewed nature and the CK value indicates the small kurtosis feature in the data.

Table 4.

Summary statistics for the UEFA Champions League data set.

Variables n Minimum Median Mean Maximum SD CV CS CK
W1 37 0.022 0.456 0.454 0.911 0.224 49.274 0.164 −0.930
W2 37 0.022 0.311 0.365 0.944 0.254 69.475 0.522 −0.839

The ML estimates and the standard errors (in parentheses) for the bivariate unit-log-symmetric model parameters are presented in Table 5. The extra parameters, associated with the log-Student-t, log-hyperbolic and log-slash models, were estimated by using the profile log-likelihood; see Saulo et al. [22]. Table 5 also presents the log-likelihood value, and the values of the Akaike (AIC) and Bayesian (BIC) information criteria. We observe that the log-hyperbolic model provides better fit than all other models based on the values of log-likelihood, AIC and BIC.

Table 5.

ML estimates (with standard errors in parentheses), and the log-likelihood, AIC and BIC values for the indicated bivariate unit-log-symmetric models.

Distribuiton η^1 η^2 σ^1 σ^2 ρ^ ν^ Log-likelihood AIC BIC
Log-normal 0.5288* 0.3414* 0.8865* 1.1355* 0.4956* −36.693 83.386 91.441
  (0.0771) (0.0637) (0.1031) (0.1320) (0.1240)        
Log-Student-t 0.5541* 0.3783* 0.7431* 0.9734* 0.4723* 7 −35.487 80.974 89.029
  (0.0751) (0.0672) (0.1033) (0.1308) (0.1463)        
Log-hyperbolic 0.5458* 0.3816* 0.8456* 1.0950* 0.4893* 2 −35.470 80.940 88.996
  (0.0752) (0.0677) (0.1162) (0.1462) (0.1428)        
Log-Laplace 0.5680* 0.5679* 0.9928* 1.3231* 0.5281* −36.009 82.019 90.073
  (0.0020) (0.0021) (0.1692) (0.2164) (0.1639)        
Log-slash 0.5629* 0.3715* 0.6203* 0.8302* 0.4472* 5 −35.560 81.120 89.174
  (0.0749) (0.0666) (0.0847) (0.1096) (0.1472)        

Note: * significant at 5% level.

Figure 2 shows the QQ plots of Resid(W), defined in (23), for the bivariate unit-log-symmetric models considered in Table 5. The QQ plot is a plot of the empirical quantiles of Resid(W) against the theoretical quantiles of the respective reference distribution (see Section 4.1). Therefore, points falling along a straight line would indicate a good fit. From Figure 2, we see clearly that, with the exception of log-Student-t case, the empirical distributions of Resid(W) in the considered models conform relatively well with their reference distributions. We also see that, in all the cases, there is a point away from the reference line, which may be an outlier.

Figure 2.

Figure 2.

QQ plot of Resid(W) for the indicated models. (a) Log-normal (b) Log-Student-t (c) Log-hyperbolic (d) Log-Laplace (e) Log-slash

We can use the Kolmogorov-Smirnov test [9] to verify the assumption of the theoretical distribution of Resid(W). As the log-hyperbolic model provided the best fit in terms of log-likelihood, AIC and BIC, we only report the result for this case. The p-value is 0.9998, and therefore we can not reject the null hypothesis that the sample is drawn from the reference distribution.

6.2. 2022 FIFA world cup

We now use the data on the 2022 FIFA World Cup to illustrate the model developed in the preceding sections. These data are available at https://www.kaggle.com/. The first variable (W1) is the medium pass completion proportion, that is, successful passes between 14 and 18 meters. The second variable (W2) is the long pass completion proportion, namely, passes longer than 37 meters; see Table A1.

Table 6 provides descriptive statistics for the variables W1 and W2. We observe that the variable W1 has mean equal to the median, which indicates symmetry in the data. The CV is 4.376%, which means a low level of dispersion around the mean. Furthermore, the CS value is relatively small, which also confirms the symmetric nature. The variable W2 has mean equal to 0.550 and median equal to 0.556, which indicates a symmetric feature in the distribution of the data. Moreover, the CV value is 13.713%, showing a low level of dispersion around the mean. The CS confirms the symmetric nature and the CK value indicates the small kurtosis feature in the data.

Table 6.

Summary statistics for the 2022 FIFA World Cup data set.

Variables n Minimum Median Mean Maximum SD CV CS CK
W1 32 0.769 0.860 0.860 0.931 0.038 4.376 −0.373 −0.194
W2 32 0.427 0.556 0.550 0.751 0.075 13.713 0.308 −0.425

Table 7 presents the estimation results for the bivariate unit-log-symmetric models, and these reveal that the log-normal model provides better fit than all other models based on the values of log-likelihood, AIC and BIC.

Table 7.

ML estimates (with standard errors in parentheses), and the log-likelihood, AIC and BIC values for the indicated bivariate unit-log-symmetric models.

Distribuiton η^1 η^2 σ^1 σ^2 ρ^ ν^ Log-likelihood AIC BIC
Log-normal 1.9872* 0.7953* 0.1364* 0.2089* 0.7343* 20.791 −31.581 −24.252
  (0.0479) (0.0294) (0.0171) (0.0261) (0.0815)        
Log-Student-t 1.9954* 0.7936* −0.1257* −0.1949* 0.7423* 9 20.130 −30.260 −22.931
  (0.0485) (0.0299) (0.0178) (0.0271) (0.0868)        
Log-hyperbolic 1.9908* 0.7942* 0.3956* 0.6088* 0.7378* 10 20.618 −31.236 −23.907
  (0.0482) (0.0296) (0.0523) (0.0800) (0.0841)        
Log-Laplace 1.9908* 0.8089* 0.1563* 0.2425* 0.7471* 16.915 −23.830 −16.501
  (0.0023) (0.0021) (0.0278) (0.0415) (0.0938)        
Log-slash 1.9897* 0.7935* 0.1173* 0.1802* 0.7392* 8 20.613 −28.919 −23.898
  (0.0482) (0.0295) (0.0154) (0.0237) (0.0844)        

Note: * significant at 5% level.

Figure 3 shows the QQ plots of Resid(W) (see Section 4.1) for the bivariate unit-log-symmetric models considered in Table 7. We see clearly that the log-normal model provides better fit than all other bivariate unit-log-symmetric models. By applying the Kolmogorov-Smirnov test for the log-normal case, the corresponding p-value is found to be 0.9993, and therefore, the null hypothesis is not rejected.

Figure 3.

Figure 3.

QQ plot of Resid(W) for the indicated models. (a) Log-normal (b) Log-Student-t (c) Log-hyperbolic (d) Log-Laplace (e) Log-slash.

7. Conclusions

In this paper, we have proposed a family of bivariate distributions over the unit square. By suitably defining the density generator, we can transform any distribution over the real line into a bivariate distribution over the region (0,1)×(0,1). Such a model has several potential applications, since the simultaneous modeling of quantities like proportions, rates or indices frequently arises in applied sciences like economics, medicine, engineering and social sciences. We have discussed several theoretical properties such as stochastic representation, quantiles, conditional distributions, independence and moments. We have also carried out a Monte Carlo simulation study and have demonstrated some applications in the analyses soccer data. The present research can be extended in several possible directions. By changing the density generator, numerous special forms of the BULS distribution can be constructed. Furthermore, generalizations to higher dimensions can be studied. We are currently working in these directions and hope to report the findings in a future paper.

Appendices.

Appendix 1. Some proofs and additional results

For the convenience of readers, we present here some complementary results relating to Section 3.3 and the proofs of Propositions 3.13 and 3.14.

Definition A.1

We say that a random variable X follows a univariate generalized hyperbolic (GH) distribution, denoted by XGH(λ,α,δ), if its PDF is given by

fGH(x;λ,α,δ)=(α2/δ)λ2πKλ(δα2)Kλ1/2(αδ2+x2)(δ2+x2/α)1/2λ,<x<.

Here, Kr is the modified Bessel function of the third kind with index r, λR,αR and δ>0 is a scale parameter.

The following result has appeared in a multivariate version in Proposition 3 of Deng and Yao [7].

Proposition A.2 Hyperbolic generator —

Let Z=(Z1,Z2) be a random vector as in Proposition 3.1. If gc(x)=exp(ν1+x), then the conditional distribution of Z2, given Z1=x, is GH(1,ν,1+x2) and both of its unconditional distributions are GH(3/2,ν,1).

Proof.

By using (8), the joint PDF of Z1 and Z2 is (see Table 1)

fZ1,Z2(x,y)=ν2exp(ν)2π(ν+1)exp(ν1+x2+y2). (A1)

So, the marginal PDF of Z1 is given by

fZ1(x)=fZ1,Z2(x,y)dy=ν2exp(ν)2π(ν+1)exp(ν1+x2+y2)dy=2ν2exp(ν)2π(ν+1)0exp(ν1+x2+y2)dy.

By using Formula 6 of Section 3.46–3.48 of (see p. 364 of Gradshteyn and Ryzhik [10]) that 0exp(ab2+x2)dx=bK1(ab), the above integral becomes

2ν2exp(ν)2π(ν+1)1+x2K1(ν1+x2). (A2)

Now, as K3/2(ν)=2πexp(ν)(ν+1)(ν/2)3/2/ν3, the above espression becomes

ν3/22πK3/2(ν)K1(ν1+x2)(1+x2/ν)1=fGH(x;3/2,ν,1),

which proves that Z1GH(3/2,ν,1). Similarly, we can show that Z2GH(3/2,ν,1), as well.

On the other hand, from (A1) and (A2), and by using the well-known identity K1/2(z)=π/(2z)exp(z), the conditional PDF of Z2, given Z1=x, is obtained as

fZ2|Z1(y|x)=exp(ν1+x2+y2)21+x2K1(ν1+x2)=ν/1+x22πK1(ν1+x2)K1/2(ν1+x2+y2)(1+x2+y2/ν)1/2=fGH(x;1,ν,1+x2),

which completes the proof.

The following result has also appeared in a multivariate version in Theorem 6.7.1 of Kotz et al. [14].

Proposition A.3 Laplace generator —

Let Z=(Z1,Z2) be a random vector as in Proposition 3.1. If gc(x)=K0(2x), then the conditional distribution of Z2, given Z1=x, is GH(1/2,2,|x|) and both of its unconditional distributions are Laplace(0,1/2).

Proof.

By (8) and using the definitions of K0 and Zgc in Table 1, we have

fZ1,Z2(x,y)=1πK0(2(x2+y2))=12π01texp(tx2+y22t)dt. (A3)

We then find the marginal density of Z1 to be

fZ1(x)=fZ1,Z2(x,y)dy=12exp(2|x|)=fL(x), (A4)

where fL(x)=exp(2|x|)/2 is the Laplace PDF with scale parameter 1/2; that is, Z1Laplace(0,1/2). Similarly, we can show that Z2Laplace(0,1/2), as well.

On the other hand, by using (A3), (A4) and the well-known identity K1/2(z)=π/(2z)exp(z), the conditional PDF of Z2, given Z1=x, is obtained as

fZ2|Z1(y|x)=1πK0(2(x2+y2))12exp(2|x|)=(2/|x|)1/22πK1/2(2|x|)K0(2x2+y2)=fGH(x;1/2,2,|x|).

Then, from Definition A.1, we simply have Z2|(Z1=x)GH(1/2,2,|x|), as required.

Definition A.4

We say that a random variable X follows a univariate extended slash (ESL) distribution, denoted by XESL(a,q), if its PDF is given by

fESL(x;a,q)=01tqϕ(ta)ϕ(tx)dt01uq1ϕ(ua)du,<x<,

where ϕ denotes the PDF of the standard normal distribution.

If we now choose a=0, the classical slash (SL) PDF is obtained, given by

fSL(x;q)=q01tqϕ(tx)dt,<x<=q2q21π|x|(q+1)γ(q+12,x22).

In this case, we denote it by XSL(q).

Proposition A.5 Slash generator —

Let Z=(Z1,Z2) be a random vector as in Proposition 3.1. If gc(x)=x(q+2)/2γ((q+2)/2,x/2), then the conditional distribution of Z2, given Z1=x, is ESL(x,q+1) and both of its unconditional distributions are SL(q).

Proof.

By (8) and using the definition of Zgc in Table 1, the joint PDF of Z1 and Z2 is given by

fZ1,Z2(x,y)=q2q21π(x2+y2)q+22γ(q+22,x2+y22)=q01tq+1ϕ(tx)ϕ(ty)dt. (A5)

So, the marginal PDF of Z1 is

fZ1(x)=fZ1,Z2(x,y)dy=q01tq+1ϕ(tx)[ϕ(ty)dy]dt=q01tqϕ(tx)dt=fSL(x;q), (A6)

which proves that (see Definition A.4), Z1SL(q). Similarly, we can show that Z2SL(q).

On the other hand, by (A5) and (A6), the conditional PDF of Z2, given Z1=x, is obtained as

fZ2|Z1(y|x)=01tq+1ϕ(tx)ϕ(ty)dt01uqϕ(ux)du=fESL(y;x,q+1).

From Definition A.4, we then find that Z2|(Z1=x)ESL(x,q+1), as required.

Proof of Proposition 3.13.

Writing S1=[W1μ1(1)]/σW1 and ρWS1+1ρW2S2=[W2μ2(1)]/σW2, a simple algebraic manipulation shows that

d2=S12+S22.

Upon using the law of total expectation, the corresponding CDF of d2 is given by (for 0<w<1)

Fd2(w)=E[E(1{S12+S22w})|S1]=E[E(1{|S2|wS12})|S11{|S1|w}]=ww[ws12ws12fS2(s2|S1=s1)ds2]fS1(s1)ds1. (A7)

If S1=s1, then W1=μ1(1)+σW1s1. So, the conditional distribution of S2, given S1=s1, is the same as the distribution of

11ρW2[W2μ2(1)σW2ρWs1]|W1=σW1s1+μ1(1).

Hence, the PDF of S2, given S1=s1, is given by

fS2(s2|S1=s1)=σW21ρW2fW2(μ2(1)+σW2[ρWs1+1ρW2s2]|W1=μ1(1)+σW1s1).

By using Lemma 3.3, the above density can be expressed as

fS2(s2|S1=s1)=σW21ρW2fW2(w2,s|W1=w1,s)=k1(1w1,s)t1,s(1w2,s)t2,sfZ2(11ρ2(w~2,sρw~1,s)|Z1=w~1,s), (A8)

where kσW1σW21ρW2/(σ1σ21ρ2), and wi,s, w~i,s and ti,s are as given in (18).

Replacing (A8) in (A7), we get the following formula for the CDF of the squared Mahalanobis distance:

Fd2(w)=kww[ws12ws121(1w1,s)t1,s(1w2,s)t2,sfZ2(11ρ2(w~2,sρw~1,s)|Z1=w~1,s)ds2]fZ1(w~1,s)ds1.

Moreover, the vector (Z1,Z2) has a standard elliptical distribution (spherical) (see Item 2.11 of Saulo et al. [23]), denoted by (Z1,Z2)BSY(μ1=0,μ2=0,σ1=1,σ2=1,ρ=0,gc), where gc is the density generator in (5). Hence, the required result.

Proof of Proposition 3.14.

By differentiating Fd2(w) (in Proposition 3.13) with respect to w and then by using the following well-known Leibniz integral rule

ddxa(x)b(x)h(x,y)dy=h(x,b(x))b(x)h(x,a(x))a(x)+a(x)b(x)h(x,y)xdy,

the required result follows.

Appendix 2. Data sets

Table A1.

UEFA Champions League and 2022 FIFA World Cup data sets.

  UEFA FIFA
  W1 W2 W1 W2
1 0.289 0.222 0.888 0.541
2 0.700 0.200 0.815 0.474
3 0.211 0.211 0.907 0.624
4 0.733 0.944 0.891 0.606
5 0.444 0.444 0.827 0.517
6 0.544 0.544 0.898 0.557
7 0.089 0.089 0.856 0.462
8 0.767 0.789 0.861 0.618
9 0.433 0.433 0.890 0.603
10 0.911 0.533 0.860 0.477
11 0.800 0.800 0.920 0.646
12 0.733 0.689 0.894 0.587
13 0.278 0.100 0.913 0.648
14 0.456 0.033 0.849 0.471
15 0.178 0.833 0.781 0.427
16 0.200 0.200 0.828 0.442
17 0.244 0.156 0.864 0.581
18 0.467 0.467 0.820 0.527
19 0.022 0.022 0.846 0.526
20 0.400 0.578 0.879 0.601
21 0.378 0.378 0.860 0.481
22 0.589 0.433 0.885 0.616
23 0.600 0.078 0.862 0.592
24 0.567 0.311 0.769 0.463
25 0.844 0.711 0.845 0.495
26 0.711 0.167 0.846 0.489
27 0.289 0.533 0.931 0.751
28 0.178 0.178 0.863 0.555
29 0.489 0.144 0.856 0.447
30 0.278 0.156 0.879 0.569
31 0.611 0.122 0.812 0.613
32 0.544 0.544 0.841 0.594
33 0.267 0.267
34 0.489 0.333
35 0.467 0.033
36 0.300 0.522
37 0.311 0.311

Funding Statement

Roberto Vila and Helton Saulo gratefully acknowledge financial support from CNPq, CAPES and FAPDF, Brazil.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • 1.Abdous B., Fougères A.-L., and Ghoudi K., Extreme behaviour for bivariate elliptical distributions, Canad. J. Statist. 33 (2005), pp. 317–334. [Google Scholar]
  • 2.Arnold B.C. and Ng H.K.T., Flexible bivariate beta distributions, J. Multivariate Anal. 102 (2011), pp. 1194–1202. [Google Scholar]
  • 3.Balakrishnan N. and Lai C-D., Continuous Bivariate Distributions, Springer-Verlag, New York, 2009. [Google Scholar]
  • 4.Barreto-Souza W. and Lemonte A., Bivariate Kumaraswamy distribution: properties and a new method to generate bivariate classes, Statistics 47 (2013), pp. 1321–1342. [Google Scholar]
  • 5.Cox D.R. and Hinkley D.V., Theoretical Statistics, Chapman and Hall, London, England, 1974. [Google Scholar]
  • 6.Davison A.C., Statistical Models, Cambridge University Press, Cambridge, England, 2008. [Google Scholar]
  • 7.Deng X. and Yao J., On the property of multivariate generalized hyperbolic distribution and the stein-type inequality, Comm. Statist. Theory Methods 47 (2018), pp. 5346–5356. [Google Scholar]
  • 8.Fang K.T., Kotz S., and Ng K.W., Symmetric Multivariate and Related Distributions, Chapman and Hall, London, England, 1990. [Google Scholar]
  • 9.Govindarajulu Z., Nonparametric Inference, World Scientific, Publishers, Singapore, 2007. [Google Scholar]
  • 10.Gradshteyn I.S. and Ryzhik I.M., Table of Integrals, Series and Products, Academic Press, San Diego, 2000. [Google Scholar]
  • 11.Heckman J.J., Sample selection bias as a specification error, Econometrica 47 (1979), pp. 153–161. [Google Scholar]
  • 12.Hössjer O. and Sjölander A., Sharp lower and upper bounds for the covariance of bounded random variables, Statist. Probab. Lett. 182 (2022), p. 109323. [Google Scholar]
  • 13.James B.R., Probabilidade: Um Curso Em Nível Intermediário, Projeto Euclides, Brazil, 2004. [Google Scholar]
  • 14.Kotz S., Kozubowski T.J., and Podgórski K., The Laplace Distribution and Generalizations, John Wiley & Sons, New York, 2001. [Google Scholar]
  • 15.Mäkeläinen T., Schmidt K., and Styan G.P.H., On the existence and uniqueness of the maximum likelihood estimate of a vector-valued parameter in fixed-size samples, Ann. Statist. 9 (1981), pp. 758–767. [Google Scholar]
  • 16.Martínez-Flórez G., Lemonte A.J., Moreno-Arenas G., and Tovar-Falón R., The bivariate unit-sinh-normal distribution and its related regression model, Mathematics 10 (2022), pp. 3125. [Google Scholar]
  • 17.Meintanis S.G., Test of fit for Marshall-Olkin distributions with applications, J. Statist. Plann. Inference 137 (2007), pp. 3954–3963. [Google Scholar]
  • 18.Nadarajah S., Shih S.H., and Nagar D.K., A new bivariate beta distribution, Statistics 51 (2017), pp. 455–474. [Google Scholar]
  • 19.Nelsen R.B., An Introduction to Copulas, 2nd ed., Springer-Verlag, New York, 2006. [Google Scholar]
  • 20.Özbilen Ö. and Genç A.İ., A bivariate extension of the Omega distribution for two-dimensional proportional data, Math. Slovaca 72 (2022), pp. 1605–1622. [Google Scholar]
  • 21.Roncalli T., Copulas and dependence modeling, in Handbook of Financial Risk Management, 1st ed., Chapman and Hall/CRC, Press, Boca Raton, FL, 2020.
  • 22.Saulo H., Dasilva A., Leiva V., Sánchez L., and Fuente-Mella H.L., Log-symmetric quantile regression models, Stat. Neerl. 76 (2022), pp. 124–163. [Google Scholar]
  • 23.Saulo H., Vila R., Cordeiro S.S., and Leiva V., Bivariate symmetric Heckman models and their characterization, J. Multivariate Anal. 193 (2023), p. 105097. [Google Scholar]
  • 24.Vanegas L.H. and Paula G.A., Log-symmetric distributions: statistical properties and parameter estimation, Braz. J. Probab. Stat. 30 (2016), pp. 196–220. [Google Scholar]
  • 25.Vila R., Balakrishnan N., Saulo H., and Protazio A., Bivariate log-symmetric models: theoretical properties and parameter estimation, preprint (2022). Available at arXiv:2211.13839, pp. 1–23.
  • 26.Wang J. and Genton M., The multivariate skew-slash distribution, J. Statist. Plann. Inference 136 (2006), pp. 209–220. [Google Scholar]

Articles from Journal of Applied Statistics are provided here courtesy of Taylor & Francis

RESOURCES