Skip to main content

The Likelihood Ratio Test of Equality of Mean Vectors with a Doubly Exchangeable Covariance Matrix

  • Chapter
  • First Online:
Methodology and Applications of Statistics

Part of the book series: Contributions to Statistics ((CONTRIB.STAT.))

Abstract

The authors derive the LRT statistic for the test of equality of mean vectors when the covariance matrix has what is called a double exchangeable structure. A second expression for this statistic, based on determinants of Wishart matrices with a block-diagonal parameter matrix, allowed for the expression of the distribution of this statistic as that of a product of independent Beta random variables. Moreover, the split of the LRT statistic into three independent components, induced by this second representation, will then allow for the expression of the exact distribution of the LRT statistic in a very manageable finite closed form for most cases and the obtention of very sharp near-exact distributions for the other cases. Numerical studies show that, as expected, due to the way they are built, these near-exact distributions are indeed asymptotic not only for increasing sample sizes but also for increasing values of all other parameters in the distribution, besides lying very close to the exact distribution even for extremely small samples.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 149.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 199.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 199.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Anderson, T.W.: An Introduction to Multivariate Statistical Analysis, 3rd edn. Wiley, Hoboken (2003)

    MATH  Google Scholar 

  • Arnold, B.C., Coelho, C.A., Marques, F.J.: The distribution of the product of powers of independent uniform random variables - a simple but useful tool to address and better understand the structure of some distributions. J. Multivariate. Anal. 113, 19–36 (2013)

    Article  MathSciNet  Google Scholar 

  • Coelho, C.A.: The Generalized Integer Gamma distribution - a basis for distributions in multivariate statistics. J. Multivariate Anal. 64, 86–102 (1998)

    Article  MathSciNet  Google Scholar 

  • Coelho, C.A.: The generalized near-integer gamma distribution: a basis for “near-exact’’ approximations to the distribution of statistics which are the product of an odd number of independent Beta random variables. J. Multivariate Anal. 89, 191–218 (2004)

    Article  MathSciNet  Google Scholar 

  • Coelho, C.A.: The Likelihood Ratio Test for Equality of Mean Vectors with Compound Symmetric Covariance Matrices. In: Gervasi, O., Murgante, B., Misra, S., Borruso, G., Torre, C.M., Rocha, A.M.A.C., Taniar, D., Apduhan, B.O., Stankova, E., Cuzzocrea, A. (eds.) Computational Science and Its Applications, Part V, pp. 20–32 (2017)

    Google Scholar 

  • Coelho, C.A., Alberto, R.P.: On the Distribution of the Product of Independent Beta Random Variables – Applications. In: Ghosh, I., Balakrishnan, N., Ng, H.K. (eds.) Advances in Statistics - Theory and Applications: Honoring the Contributions of Barry C. Arnold in Statistical Science, (in print) Springer (2021)

    Google Scholar 

  • Coelho, C.A., Arnold, B.C.: Finite Form Representations for Meijer G and Fox H Functions - Applied to Multivariate Likelihood Ratio Tests Using Mathematica\(^{\text{\textregistered} }\). MAXIMA and R. Lecture Notes in Statistics. Springer, Cham (2019)

    Book  Google Scholar 

  • Coelho, C.A., Marques, F.J.: The multi-sample block-scalar sphericity test: exact and near-exact distributions for its likelihood ratio test statistic. Commun. Statist. Theory Methods 42, 1153–1175 (2013)

    Article  MathSciNet  Google Scholar 

  • Coelho, C.A., Roy, A.: Testing the hypothesis of a doubly exchangeable covariance matrix. Metrika 83, 45–68 (2020)

    Article  MathSciNet  Google Scholar 

  • Coelho, C.A., Singull, M.: Testing for double complete symmetry. In: Holgersson, T., Singull, M. (eds.) Recent Developments in Multivariate and Random Matrix Analysis, pp. 17–39 (2020)

    Google Scholar 

  • Coelho, C.A., Arnold, B.C., Marques, F.J.: Near-exact distributions for certain likelihood ratio test statistics. J. Stat. Theory Pract. 4, 711–725 (2010)

    Article  MathSciNet  Google Scholar 

  • Coelho, C.A., Marques, F.J., Arnold, B.C.: The exact and near-exact distributions of the main likelihood ratio test statistics used in the complex multivariate normal setting. TEST 24, 386–416 (2015)

    Article  MathSciNet  Google Scholar 

  • Kshirsagar, A.M.: Multivariate Analysis. Marcel Dekker, New York (1972)

    MATH  Google Scholar 

  • Marques, F.J., Coelho, C.A., Arnold, B.C.: A general near-exact distribution theory for the most common likelihood ratio test statistics used in Multivariate Analysis. TEST 20, 180–203 (2011)

    Article  MathSciNet  Google Scholar 

  • Pavlenko, T., Roy, A.: Supervised classifiers for high-dimensional higher-order data with locally doubly exchangeable covariance structure. Comm. Statist. Theory Methods 46, 11612–11634 (2017)

    Article  MathSciNet  Google Scholar 

  • Roy, A., Fonseca, M.: Linear models with doubly exchangeable distributed errors. Comm. Statist. Theory Methods 41, 2545–2569 (2012)

    Article  MathSciNet  Google Scholar 

  • Seely, J.: Quadratic subspaces and completeness. Ann. Math. Statist. 42, 710–721 (1971)

    Article  MathSciNet  Google Scholar 

  • Szatrowski, T.D.: Explicit solutions, one iteration convergence and averaging in the multivariate Normal estimation problem for patterned means and covariances. Ann. Inst. Statist. Math. 30, 81–88 (1978)

    Article  MathSciNet  Google Scholar 

  • Szatrowski, T.D.: Necessary and sufficient conditions for explicit solutions in the multivariate Normal estimation problem for patterned means and covariances. Ann. Statist. 8, 802–810 (1980)

    MathSciNet  MATH  Google Scholar 

  • Tricomi, F.G., Erdélyi, A.: The asymptotic expansion of a ratio of Gamma functions. Pacific J. Math. 1, 133–142 (1951)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work is funded by Portuguese National Funds through FCT - Fundação para a Ciência e a Tecnologia, I.P., under the scope of the project UIDB/00297/ 2020 (CMA-FCT/UNL – Center for Mathematics and Applications)

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Carlos A. Coelho .

Editor information

Editors and Affiliations

Appendices

Appendix A: Proof of Expression (16)

For ease of exposition let us split the proof of (16) in the proof of three consecutive results. First let us prove the following result.

Result 1: The MLE of \(\Sigma \) under \(H_1\), which is \(A^*\) into (8), may be equivalently written as

$$\begin{aligned} A^*=P_{vm}\otimes \widehat{\Psi }_{1|H_1}+Q_{vm}\otimes \widehat{\Psi }_{2|H_1}+R_{vm}\otimes \widehat{\Psi }_{3|H_1}, \end{aligned}$$
(A.1)

where, from (7),

$$\begin{aligned} \begin{array}{l} \widehat{\Psi }_{1|H_1}=\widehat{U}_{|H_1}+(v-1)\widehat{V}_{|H_1}+v(m-1)\widehat{M}_{|H_1}\\ \widehat{\Psi }_{2|H_1}=\widehat{U}_{|H_1}+(v-1)\widehat{V}_{H_1}-v\widehat{M}_{|H_1}\\ \widehat{\Psi }_{3|H_1}=\widehat{U}_{|H_1}-\widehat{V}_{|H_1}\,, \end{array} \end{aligned}$$
(A.2)

and

$$\begin{aligned} \begin{array}{l} \displaystyle P_{vm}=\frac{1}{mv}J_{vm}\\ \displaystyle Q_{vm}=\frac{1}{v}(I_{m}\otimes J_v)-P_{vm}\\ \displaystyle R_{vm}=I_{vm}-Q_{vm}-P_{vm}\,. \end{array} \end{aligned}$$
(A.3)

Proof

The proof is easier if we start from (A.1) and try to end with (8), by using (A.2) and (A.3) to write

$$ \begin{array}{l} \displaystyle P_{vm}\otimes \widehat{\Psi }_{1|H_1}+Q_{vm}\otimes \widehat{\Psi }_{2|H_1}+R_{vm}\otimes \widehat{\Psi }_{3|H_1}\\ \displaystyle = \frac{1}{mv}J_{vm} \otimes \left( \widehat{U}_{|H_1}+(v-1)\widehat{V}_{|H_1}+v(m-1)\widehat{M}_{|H_1}\right) ~~~~~~~~~~~~~~~~~~~~~~~~\\ \end{array}$$
$$ \begin{array}{l} \displaystyle +\left( \frac{1}{v}(I_{m}\otimes J_v)-\frac{1}{mv}J_{vm}\right) \otimes \left( \widehat{U}_{|H_1}+(v-1)\widehat{V}_{|H_1}-v\widehat{M}_{|H_1}\right) ~~~~~~\\ +\left( I_{vm}-\frac{1}{v}(I_{m}\otimes J_v)+P_{vm}-P_{vm}\right) \otimes (\widehat{U}_{|H_1}-\widehat{V}_{|H_1})\\ \displaystyle =\left( \frac{1}{mv}J_{vm} +\frac{1}{v}(I_{m}\otimes J_v)-\frac{1}{mv}J_{vm}+I_{vm}-\frac{1}{v}(I_{m}\otimes J_v)\right) \otimes \widehat{U}_{|H_1}\\ \displaystyle +\left( \frac{v\!-\!1}{mv}J_{vm}+\frac{v\!-\!1}{v}(I_{m}\!\otimes \! J_v)-\frac{v\!-\!1}{mv}J_{vm}-I_{vm}+\frac{1}{v}(I_{m}\!\otimes \! J_v)\!\right) \!\otimes \! \widehat{V}_{|H_1}\\ \displaystyle +\left( \frac{m-1}{m}J_{vm}-(I_{m}\otimes J_v)+\frac{1}{m}J_{vm}\right) \otimes \widehat{M}_{|H_1}\\ \displaystyle =I_{vm}\otimes \widehat{U}_{|H_1}+(I_m\otimes (J_v-I_v))\otimes \widehat{V}_{|H_1}+(J_{vm}-(I_m\otimes J_m))\otimes \widehat{M}_{|H_1} \end{array}$$

which is (8).    \(\square \)

We now prove that \(P_{vm}\), \(Q_{vm}\) and \(R_{vm}\) in (A.3) are indeed mutually orthogonal projectors.

Result 2 : \(P_{vm}\), \(Q_{vm}\) and \(R_{vm}\) in (A.3) are mutually orthogonal projectors with \({\mathrm{rank}}(P_{vm})=1\), \({\mathrm{rank}}(Q_{vm})=m-1\), \({\mathrm{rank}}(R_{vm})=m(v-1)\).

Proof

\(P_{vm}\), \(Q_{vm}\) and \(R_{vm}\) in (A.3) are idempotent, since

$$\begin{aligned} P_{vm}P_{vm}= & {} \frac{1}{mv}J_{vm}\frac{1}{mv}J_{vm}=\frac{1}{(mv)^2}mvJ_{vm}=\frac{1}{mv}J_{vm}=P_{vm},\\[5pt] Q_{vm}Q_{vm}= & {} \left( \frac{1}{v}(I_{m}\otimes J_v)-P_{vm}\right) \left( \frac{1}{v}(I_{m}\otimes J_v)-P_{vm}\right) \\= & {} \frac{1}{v^2}(I_{m}\!\otimes \! J_v)(I_{m}\!\otimes \! J_v)-\frac{1}{v}(I_{m}\!\otimes \! J_v)P_{vm}-\frac{1}{v}P_{vm}(I_{m}\!\otimes \! J_v)+P_{vm} \\= & {} \frac{1}{v}(I_{m}\otimes J_v)-\frac{2}{vm}J_{vm}+\frac{1}{vm}J_{vm}=\frac{1}{v}(I_{m}\otimes J_v)-P_{vm}=Q_{vm}, \end{aligned}$$
$$\begin{aligned} R_{vm}R_{vm}= & {} (I_{vm}-Q_{vm}-P_{vm})(I_{vm}-Q_{vm}-P_{vm})=I_{vm}-Q_{vm}-P_{vm}\\&-Q_{vm}+Q_{vm}Q_{vm}+Q_{vm}P_{vm}-P_{vm}+P_{vm}Q_{vm}+P_{vm}P_{vm}\\ {}= & {} I_{vm}-Q_{vm}-P_{vm}-Q_{vm}+Q_{vm}-P_{vm}+P_{vm}\\= & {} I_{vm}-Q_{vm}-P_{vm}=R_{vm}, \end{aligned}$$

and they are mutually orthogonal since

$$ \begin{array}{rcl} \displaystyle P_{vm}Q_{vm}&{}=&{}P_{vm}\left( \frac{1}{v}(I_{m}\otimes J_v)-P_{vm}\right) =\frac{1}{v}P_{vm}(I_{m}\otimes J_v)-P_{vm}\\ &{}=&{}\displaystyle \frac{1}{vm}J_{vm}\left( \frac{1}{v}(I_{m}\otimes J_v)-I_{vm}\right) =0_{vm},\\ R_{vm}P_{vm}&{}=&{}(I_{vm}-Q_{vm}-P_{vm})P_{vm}=P_{vm}-P_{vm}=0_{vm},\\ R_{vm}Q_{vm}&{}=&{}(I_{vm}-Q_{vm}-P_{vm})Q_{vm}=Q_{vm}-Q_{vm}=0_{vm}. \end{array}$$

Since \(P_{vm}\), \(Q_{vm}\) and \(R_{vm}\) are idempotent, we have

$$ \begin{array}{l} \displaystyle {\mathrm{rank}}(P_{vm})=\mathrm{tr}(P_{vm})=\frac{1}{vm}\mathrm{tr}(I_{vm})=1\\ \displaystyle {\mathrm{rank}}(Q_{vm})=\mathrm{tr}(Q_{vm})=\frac{1}{v}\mathrm{tr}(I_{m}\otimes J_v)\!-\!\mathrm{tr}(P_{vm})=\frac{mv}{v}\!-\!1=m\!-\!1\\ \displaystyle {\mathrm{rank}}(R_{vm})=\mathrm{tr}(R_{vm})=\mathrm{tr}(I_{vm})\!-\!\mathrm{tr}(Q_{vm})\!-\!\mathrm{tr}(P_{vm})=vm\!-\!m=(v\!-\!1)m. \end{array}$$

   \(\square \)

Setting together the previous two results with the next one, we will finally be able to prove (16).

Result 3 : For the matrix \(A^{**}\) in (14), we may write

$$\begin{aligned} \widehat{\Psi }_{1|H_1}= & {} A_1^{**},\\ \widehat{\Psi }_{2|H_1}= & {} \frac{1}{m-1}\sum _{j=1}^{m-1}A_{jv+1}^{**},\\ \widehat{\Psi }_{3|H_1}= & {} \frac{1}{m(v-1)}\sum _{j=1}^m\sum _{k=1}^{v-1}A^{**}_{(j-1)v+k+1}\,. \end{aligned}$$

Proof

Let \({\mathrm{BTr}}_r(A)\) denote the sum of all \(r{\scriptstyle \times }r\) diagonal blocks of A. Then, for \(U_{|H_1}\), \(V_{|H_1}\) and \(W_{|H_1}\) in (9)–(11) we may write

$$\begin{aligned} \widehat{U}_{|H_1}= & {} \frac{1}{mv}{\mathrm{BTr}}_r(A),\\ \widehat{V}_{|H_1}= & {} \frac{1}{mv(v-1)}\left( {\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)-{\mathrm{BTr}}_r(A)\right) ,\\ \widehat{M}_{|H_1}= & {} \frac{1}{mv^2(m-1)}\left( {\mathrm{BTr}}_r((J_{vm}\otimes I_r)A)-{\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)\right) . \end{aligned}$$

Then, since from (14) we may write

$$\begin{aligned} A=\Gamma ' A^{**}\Gamma \,, \end{aligned}$$
(A.4)

from (A.2), and from the definition of the matrix \(\Gamma \) in (5), we have

$$\begin{aligned} \widehat{\Psi }_{1|H_1}= & {} \widehat{U}_{|H_1}+(v-1)\widehat{V}_{|H_1}+v(m-1)\widehat{M}_{|H_1}\\[5pt]= & {} \frac{1}{mv}{\mathrm{BTr}}_r(A)+\frac{1}{mv}\left( {\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)-{\mathrm{BTr}}_r(A)\right) \\&+\frac{1}{mv}\left( {\mathrm{BTr}}_r((J_{vm}\otimes I_r)A)-{\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)\right) \\[5pt]= & {} \frac{1}{mv}{\mathrm{BTr}}_r((J_{vm}\otimes I_r)A)={\mathrm{BTr}}_r((P_{vm}\otimes I_r)A)\\[5pt]= & {} {\mathrm{BTr}}_r((P_{vm}\otimes I_r)\Gamma ' A^{**}\Gamma ) ={\mathrm{BTr}}_r(\Gamma (P_{vm}\otimes I_r)\Gamma ' A^{**})\\= & {} {\mathrm{BTr}}_r((\,\underbrace{((\Gamma _m\otimes \Gamma _v)P_{vm}(\Gamma _m\otimes \Gamma _v)')}_{=diag(\underbrace{1,0,\ldots ,0}_{vm})}\otimes I_r) A^{**})=A_1^{**}\,, \end{aligned}$$

where \(A_1^{**}\) represents the first diagonal block of \(A^{**}\) of dimensions \(r{\scriptstyle \times }r\), since, given the definition of a Helmert matrix, \(\Gamma _m\otimes \Gamma _v\) is a \(vm{\scriptstyle \times }vm\) matrix whose first row is equal to \(\frac{1}{\sqrt{vm}}\underline{1}'_{vm}\), where \(\underline{1}_{vm}\) denotes a vector of 1’s of dimension vm, and as such \((\Gamma _m\otimes \Gamma _v)P_{vm}\) is a \(vm{\scriptstyle \times }vm\) matrix whose first row is equal to \(\frac{1}{\sqrt{vm}}\underline{1}'_{vm}\) and all other rows are null, so that

$$\begin{aligned} (\Gamma _m\otimes \Gamma _v)P_{vm}(\Gamma _m\otimes \Gamma _v)'=diag(\,\underbrace{1,0,\ldots ,0}_{vm}\,)\,, \end{aligned}$$
(A.5)

what implies

$$ ((\Gamma _m\otimes \Gamma _v)P_{vm}(\Gamma _m\otimes \Gamma _v)')\otimes I_r=bdiag(I_r,0_{v(m-1){\scriptstyle \times }v(m-1)})\,. $$

Using (A.4), and once again the definition of the matrix \(\Gamma \) in (5), we also have

$$\begin{aligned} \widehat{\Psi }_{2|H_1}= & {} \widehat{U}_{|H_1}+(v-1)\widehat{V}_{|H_1}-v\widehat{M}_{|H_1}\\= & {} \frac{1}{mv}{\mathrm{BTr}}_r(A)+\frac{1}{mv}\left( {\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)-{\mathrm{BTr}}_r(A)\right) \\&-\frac{1}{mv(m-1)}\left( {\mathrm{BTr}}_r((J_{vm}\otimes I_r)A)-{\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)\right) \\= & {} \frac{1}{mv}\left( 1+\frac{1}{m-1}\right) {\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)\\&-\frac{1}{mv(m-1)}{\mathrm{BTr}}_r((J_{vm}\otimes I_r)A)\\= & {} \frac{1}{m-1}{\mathrm{BTr}}_r\left( \left( \left( \frac{1}{v}(I_m\otimes J_v)-\frac{1}{mv}J_{vm}\right) \otimes I_r\right) A\right) \\ {}= & {} \frac{1}{m-1}{\mathrm{BTr}}_r\left( \left( Q_{vm}\otimes I_r\right) A\right) =\frac{1}{m-1}{\mathrm{BTr}}_r((Q_{vm}\otimes I_r)\Gamma ' A^{**}\Gamma )\\= & {} \frac{1}{m-1}{\mathrm{BTr}}_r(\Gamma (Q_{vm}\otimes I_r)\Gamma ' A^{**})\\= & {} \frac{1}{m-1}{\mathrm{BTr}}_r((((\Gamma _m\otimes \Gamma _v)Q_{vm}(\Gamma _m\otimes \Gamma _v))\otimes I_r) A^{**})\\= & {} \frac{1}{m-1}\sum _{j=1}^{m-1}A_{jv+1}^{**}\,, \end{aligned}$$

where \(A_j^{**}\) represents the j-th diagonal block of \(A^{**}\) of dimensions \(r{\scriptstyle \times }r\), since

$$\begin{aligned} (\Gamma _m\otimes \Gamma _v)Q_{vm}(\Gamma _m\otimes \Gamma _v)'=diag(\,\underbrace{0,\ldots ,0}_{v},\underbrace{\underbrace{1,0,\ldots ,0}_{v},\ldots ,\underbrace{1,0,\ldots ,0}_{v}}_{m-1 \hbox {~times}}\,) \end{aligned}$$
(A.6)

and as such,

$$ \begin{array}{l} \displaystyle ((\Gamma _m\otimes \Gamma _v)Q_{vm}(\Gamma _m\otimes \Gamma _v)')\otimes I_r\\ \displaystyle =bdiag(I_{vr{\scriptstyle \times }vr},\underbrace{\underbrace{I_r, 0_{(v-1)r{\scriptstyle \times }(v-1)r}},\dots ,\underbrace{I_r,0_{(v-1)r{\scriptstyle \times }(v-1)r}}}_{m-1})\,. \end{array}$$

This is so because, given the definition of a Helmert matrix, for a \(v{\scriptstyle \times }v\) Helmert matrix \(\Gamma _v\), we have that \(\Gamma _vJ_v\) is a matrix whose first row is equal to \(\frac{1}{\sqrt{v}}\underline{1}'_v\), and as such we have

$$ \Gamma _vJ_v\Gamma _v'=diag(\,\underbrace{v,0,\dots ,0}_{v}\,)\,, $$

so that we have

$$ \begin{array}{l} \displaystyle \frac{1}{v}\left( I_m\otimes \Gamma _v J_v\Gamma '_v\right) \\ \displaystyle = bdiag(\underbrace{diag(\,\underbrace{1,0,\dots ,0}_{v}\,),diag(\,\underbrace{1,0,\dots ,0}_{v}\,),\dots ,diag(\,\underbrace{1,0,\dots ,0}_{v}\,)}_{m})\\ \displaystyle = diag(\,\underbrace{\underbrace{1,0,\dots ,0}_{v},\underbrace{1,0,\dots ,0}_{v},\dots ,\underbrace{1,0,\dots ,0}_{v}}_{m}\,) \end{array}$$

and as such we may write, using the relation in (A.5),

$$\begin{aligned}&\displaystyle (\Gamma _m\otimes \Gamma _v)Q_{vm}(\Gamma _m\otimes \Gamma _v)'\\&\displaystyle = (\Gamma _m\otimes \Gamma _v)\left( \frac{1}{v}(I_m\otimes J_v)-p_{vm}\right) (\Gamma _m\otimes \Gamma _v)'\\&\displaystyle = \left( \frac{1}{v}(\Gamma _m\otimes \Gamma _v J_v)-(\Gamma _m\otimes \Gamma _v)P_{vm}\right) (\Gamma _m\otimes \Gamma _v)'\\&\displaystyle = \frac{1}{v}(\Gamma _m\otimes \Gamma _v J_v)(\Gamma _m\otimes \Gamma _v)'-(\Gamma _m\otimes \Gamma _v)P_{vm}(\Gamma _m\otimes \Gamma _v)'\\&\displaystyle = \frac{1}{v}(\Gamma _m\Gamma _m'\otimes \Gamma _v J_v\Gamma _v')-(\Gamma _m\otimes \Gamma _v)P_{vm}(\Gamma _m\otimes \Gamma _v)'~~~~~~~~~~~~~~~~~\\&\displaystyle = \frac{1}{v}(I_m\otimes \Gamma _v J_v\Gamma _v'))-diag(\,\underbrace{1,0,\dots ,0}_{vm}\,)\\&\displaystyle = diag(\,\underbrace{\underbrace{1,0,\dots ,0}_{v},\underbrace{1,0,\dots ,0}_{v},\dots ,\underbrace{1,0,\dots ,0}_{v}}_{m}\,)-diag(\,\underbrace{1,0,\dots ,0}_{vm}\,)\\&\displaystyle = diag(\,\underbrace{0,\ldots ,0}_{v},\underbrace{\underbrace{1,0,\ldots ,0}_{v},\ldots ,\underbrace{1,0,\ldots ,0}_{v}}_{m-1\hbox {~times}}\,)\,. \end{aligned}$$

And finally, using once again (A.4), and the definition of \(R_{vm}\) in (A.3), we have

$$\begin{aligned} \widehat{\Psi }_{3|H_1}= & {} \widehat{U}_{|H_1}-\widehat{V}_{|H_1}\\= & {} \frac{1}{mv}{\mathrm{BTr}}_r(A)-\frac{1}{mv(v-1)}\left( {\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)-{\mathrm{BTr}}_r(A)\right) \\= & {} \frac{1}{mv}\left( 1+\frac{1}{v-1}\right) {\mathrm{BTr}}_r(A)-\frac{1}{mv(v-1)}{\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)\\= & {} \frac{1}{m(v-1)}\left( {\mathrm{BTr}}_r(A)-\frac{1}{v}{\mathrm{BTr}}_r((I_m\otimes J_v\otimes I_r)A)\right) \\= & {} \frac{1}{m(v-1)}{\mathrm{BTr}}_r\left( (I_{mvr}-\frac{1}{v}(I_m\otimes J_v\otimes I_r))A\right) \\= & {} \frac{1}{m(v-1)}{\mathrm{BTr}}_r\left( ((I_{mv}-\frac{1}{v}(I_m\otimes J_v))\otimes I_r)A\right) \\= & {} \frac{1}{m(v-1)}{\mathrm{BTr}}_r\left( (R_{vm}\otimes I_r)A\right) \\= & {} \frac{1}{m(v-1)}{\mathrm{BTr}}_r\left( (R_{vm}\otimes I_r)\Gamma ' A^{**} \Gamma \right) \\= & {} \frac{1}{m(v-1)}{\mathrm{BTr}}_r\left( \Gamma (R_{vm}\otimes I_r)\Gamma ' A^{**} \right) \\= & {} \frac{1}{m(v-1)}{\mathrm{BTr}}_r((((\Gamma _m\otimes \Gamma _v) R_{vm}(\Gamma _m\otimes \Gamma _v)')\otimes I_r) A^{**} )\\= & {} \frac{1}{m(v-1)}\sum _{j=1}^m\sum _{k=1}^{v-1}A^{**}_{(j-1)v+k+1}\,,\\[-15pt] \end{aligned}$$

where, once again, \(A^{**}_j\) represents the j-th diagonal block of \(A^{**}\) of dimensions \(r{\scriptstyle \times }r\).

This is so given that from the definition of \(R_{vm}\) in (A.3) and from (A.5) and (A.6) it is clear that we have

$$ \begin{array}{l} \displaystyle (\Gamma _m\otimes \Gamma _v)R_{vm}(\Gamma _m\otimes \Gamma _v)'\\ \displaystyle =(\Gamma _m\otimes \Gamma _v)(\Gamma _m\otimes \Gamma _v)'-(\Gamma _m\otimes \Gamma _v)Q_{vm}(\Gamma _m\otimes \Gamma _v)'\\ \displaystyle -(\Gamma _m\otimes \Gamma _v)P_{vm}(\Gamma _m\otimes \Gamma _v)'~~~\\ \displaystyle =I_{vm}-diag(\,\underbrace{0,\ldots ,0}_{v},\underbrace{\underbrace{1,0,\ldots ,0}_{v},\ldots ,\underbrace{1,0,\ldots ,0}_{v}}_{m-1\hbox {~times}}\,)-diag(\,\underbrace{1,0,\ldots ,0}_{vm}\,)\\ \displaystyle =diag(0,\underbrace{\underbrace{1,\ldots ,1}_{v-1},0,\underbrace{1,\ldots ,1}_{v-1},\ldots ,0, \underbrace{1,\ldots ,1}_{v-1}}_{m\hbox {~times}}\,)\,, \end{array}$$

so that

$$ \begin{array}{l} \displaystyle ((\Gamma _m\otimes \Gamma _v)R_{vm}(\Gamma _m\otimes \Gamma _v)')\otimes I_r\\ \displaystyle =bdiag(\,\underbrace{0_{r{\scriptstyle \times }r},I_{(v-1)r},0_{r{\scriptstyle \times }r},I_{(v-1)r},\dots ,0_{r{\scriptstyle \times }r},I_{(v-1)r}}_{m\hbox {~times}}\,). \end{array}$$

   \(\square \)

Finally, expression (16) holds since from Result 1 and the mutual orthogonality of \(P_{vm}\), \(Q_{vm}\) and \(R_{vm}\) proven in Result 2 and yet the rank results therein, we may write

$$ |A^*|=|\widehat{\Psi }_{1|H_1}|\,|\widehat{\Psi }_{2|H_1}|^{m-1}\,|\widehat{\Psi }_{3|H_1}|^{m(v-1)}, $$

which given the expressions obtained in Result 3 for \(\widehat{\Psi }_{1|H_1}\), \(\widehat{\Psi }_{2|H_1}\) and \(\widehat{\Psi }_{3|H_1}\) yields (16).

Expression (17) would be proven in a completely similar way.

Appendix B: The GIG, EGIG and GNIG distributions

In this appendix we define the GIG (Generalized Integer Gamma), EGIG (Exponentiated Generalized Integer Gamma) and GNIG (Generalized Near-Integer Gamma) distributions and establish the notation used concerning their p.d.f.’s and c.d.f.’s.

We say that the r.v. (random variable) X has a Gamma distribution with shape parameter \({r\,(>0)}\) and rate parameter \({\lambda \,(>0)}\), and we will denote this fact by \(X\sim \Gamma (r,\lambda )\), if the p.d.f. of X is

$$ f^{}_X(x)=\frac{\lambda ^r}{\Gamma (r)}\,e^{-\lambda x}\,x^{r-1}~~~~(x>0)\,. $$

Let \(X_j\sim \Gamma (r_j,\lambda _j)\) \({(j=1,\dots ,p)}\) be a set of p independent r.v.’s and consider the r.v.

$$ W=\sum ^p_{j=1}X_j\,. $$

In case all the \(r_j\in \mathbb {N}\), the distribution of W is what we call a GIG distribution (Coelho 1998).   If all the \(\lambda _j\) are different, W has a GIG distribution of depth p, with shape parameters \(r_j\) and rate parameters \(\lambda _j\), with p.d.f.

$$ f^{}_W(w)=f^{{ \,GIG}}\Bigl (w\,\bigl |\,\{r_j\}_{j=1:p};\{\lambda _j\}_{j=1:p};p\Bigr )=K\sum ^p_{j=1} P_j(w)\,e^{-\lambda _j w}\,, $$

and c.d.f.

$$ F^{}_W(w)=F^{{ \,GIG}}\Bigl (w\,\bigl |\,\{r_j\}_{j=1:p};\{\lambda _j\}_{j=1:p};p\Bigr )=1-K\sum ^p_{j=1}P^*_j(w)\,e^{-\lambda _jw}\,, $$

for \({w>0}\), where

$$\begin{aligned} K=\prod ^p_{j=1}\lambda _j^{r_j}\,,~~~~~~P_j(w)=\sum ^{r_j}_{k=1} c_{j,k}\,w^{k-1} \end{aligned}$$
(B.1)

and

$$\begin{aligned} P^*_j(w)=\sum ^{r_j}_{k=1}c_{j,k}(k-1)!\sum ^{k-1}_{i=0}\frac{w^i}{i!\,\lambda _j^{k-i}}\,, \end{aligned}$$
(B.2)

with

$$\begin{aligned} c_{j,r_j}=\frac{1}{(r_j-1)!}\mathop {\prod ^p_{i=1}}_{i\ne j}(\lambda _i-\lambda _j)^{-r_i}\,,~~~~j=1,\dots ,p\,, \end{aligned}$$
(B.3)

and, for \(k=1,\dots ,r_j-1\) and \(j=1,\dots ,p\),

$$\begin{aligned} c_{j,r_j-k}=\frac{1}{k}\sum _{i=1}^k\frac{(r_j-k+i-1)!}{(r_j-k-1)!}\,R(i,j,p)c_{j,r_j-(k-i)}\,, \end{aligned}$$
(B.4)

where

$$\begin{aligned} R(i,j,p)=\mathop {\sum ^p_{k=1}}_{k\ne j}r_k(\lambda _j-\lambda _k)^{-i}~~~~(i=1,\dots ,r_j-1)\,. \end{aligned}$$
(B.5)

In case some of the \(\lambda _j\) assume the same value as other \(\lambda _j\)’s, the distribution of W still is a GIG distribution, but in this case with a reduced depth. In this more general case, let \(\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g(\le p)}\) be the set of different \(\lambda _j\)’s and let \({\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g(\le p)}\) be the set of the corresponding shape parameters, with \(r_\ell \) \((\ell =1,\dots ,g)\) being the sum of all \(r_j\) \({(j\in \{1,\dots ,p\})}\) which correspond to the \(\lambda _j\) assuming the value \(\lambda _\ell \). In this case W will have a GIG distribution of depth g, with shape parameters \(r_\ell \) and rate parameters \(\lambda _\ell \) \({(\ell =1,\dots ,g)}\).

The r.v. \({Z=e^{-W}}\) has then what Arnold et al. (2013) call an Exponentiated Generalized Integer Gamma (EGIG) distribution of depth g, with p.d.f.

$$\begin{aligned} \begin{array}{rcl} \displaystyle f^{}_Z(z) &{} = &{} \displaystyle f^{{ EGIG}}\biggl (z~\bigl |\,{\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g};\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g};g\biggr )\\ &{} = &{} \displaystyle f^{{ \,GIG}}\biggl (-\log \,z~\bigl |\,{\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{j=1:g};\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g};g\biggr )\,\frac{1}{z}\\ &{} = &{} \displaystyle K^*\sum _{\ell =1}^g P_\ell (-\log \,z)\,z^{\lambda _\ell -1}~~~~~~~~~~(0<z<1) \end{array} \end{aligned}$$
(B.6)

and c.d.f.

$$\begin{aligned} \begin{array}{rcl} \displaystyle F^{}_Z(z) &{} = &{} \displaystyle F^{{ EGIG}}\biggl (z~\bigl |\,{\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g};\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g};g\biggr )\\ &{} = &{} \displaystyle 1-F^{{ \,GIG}}\biggl (-\log \,z~\bigl |\,{\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g};\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g};g\biggr )\\ &{} = &{} \displaystyle K^*\sum _{\ell =1}^g P^*_\ell (-\log \,z)\,z^{\lambda _\ell }~~~~~~~~~~(0<z<1)\,, \end{array} \end{aligned}$$
(B.7)

for

$$\begin{aligned} K^*=\prod _{\ell =1}^g \lambda _\ell ^{r_\ell } \end{aligned}$$
(B.8)

and \(P_\ell (\,\cdot \,)\) and \(P^*_\ell (\,\cdot \,)\) given by (B.1) and (B.2), with j replaced by \(\ell \).

Then let W be a r.v. with a GIG distribution of depth g, with rate parameters \(\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g}\) and shape parameters \({\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g}\) \((\in \mathbb N)\) and let \(W^*\sim \Gamma (r,\lambda )\), with \(r\in {\mathbb R}^+\backslash {\mathbb N}\). Let further W and \(W^*\) be two independent r.v.’s. Then the r.v.

$$ Y=W+W^* $$

has a Generalized Near-Integer Gamma (GNIG) distribution (Coelho 2004) of depth \({g+1}\), with rate parameters \(\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g}\) and \(\lambda \) and corresponding shape parameters \({\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g}\) \((\in \mathbb N)\) and r, with p.d.f.

$$ \begin{array}{rcl} f_{Y}^{}(y) &{} = &{} \displaystyle f^{{ \,GNIG}}\Bigl (y\,|\,{\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g},r;\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g},\lambda ;g+1\Bigr )\\ &{} = &{} \displaystyle K^*\lambda ^r\sum _{\ell =1}^g e^{-\lambda _\ell y}\sum _{k=1}^{r_\ell }\left\{ c_{\ell ,k}\frac{\Gamma (k)}{\Gamma (k+r)}y^{\,k+r-1}\,\!_1F_1\bigl (r,k\!+\!r,-(\lambda \!-\!\lambda _\ell )y\bigr )\right\} \end{array}$$

and c.d.f.

$$ \begin{array}{rcl} F_{Y}^{}(y) &{} = &{} \displaystyle F^{{ \,GNIG}}\Bigl (y\,|\,{\mathop {\{}\limits ^{\approx }}\!\!r_\ell \!\!{\mathop {\}}\limits ^{\approx }}_{\ell =1:g},r;\widetilde{\{}\lambda _\ell \widetilde{\}}_{\ell =1:g},\lambda ;g+1\Bigr )\\ &{} = &{} \displaystyle \frac{\lambda ^r\,y^r}{\Gamma (r+1)}\,_1F_1\bigl (r,r+1,-\lambda y\bigr )\\ &{} &{} \displaystyle -K^{**}\! \sum _{\ell =1}^g e^{-\lambda _\ell y}\sum _{k=1}^{r_\ell }c^*_{\ell ,k}\sum _{i=0}^{k-1}\left\{ \!\frac{y^{\,r+i}\,\lambda _\ell ^i}{\Gamma (r^*\!+\!i)}\,_1F_1\bigl (r,r^*\!+\!i,-(\lambda \!-\!\lambda _\ell )y\bigr )\!\right\} \end{array}$$

where \({z>0}\), \(K^{**}=K^*\lambda ^r\), for \(K^*\) in (B.8), \({r^*=r+1}\), and

$$ c^*_{\ell ,k}=\frac{c_{\ell ,k}}{\lambda _\ell ^k}\,\Gamma (k) $$

for \(c_{\ell ,k}\) given by (B.3)–(B.5) with j replaced by \(\ell \), and

$$ _1F_1(a,b,y)=\sum _{i=0}^\infty \frac{\Gamma (a+i)}{\Gamma (b+i)}\frac{\Gamma (b)}{\Gamma (a)}\,\frac{z^i}{i!} $$

is the Kummer confluent hypergeometric function.

Rights and permissions

Reprints and permissions

Copyright information

© 2021 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Coelho, C.A., Pielaszkiewicz, J. (2021). The Likelihood Ratio Test of Equality of Mean Vectors with a Doubly Exchangeable Covariance Matrix. In: Arnold, B.C., Balakrishnan, N., Coelho, C.A. (eds) Methodology and Applications of Statistics. Contributions to Statistics. Springer, Cham. https://doi.org/10.1007/978-3-030-83670-2_8

Download citation

Publish with us

Policies and ethics