Skip to main content
Log in

Transformation approaches of linear random-effects models

  • Original Paper
  • Published:
Statistical Methods & Applications Aims and scope Submit manuscript

Abstract

Assume that a linear random-effects model \(\mathbf{y}= \mathbf{X}\varvec{\beta }+ \varvec{\varepsilon }= \mathbf{X}(\mathbf{A}\varvec{\alpha }+ \varvec{\gamma }) + \varvec{\varepsilon }\) is transformed as \(\mathbf{T}\mathbf{y}= \mathbf{T}\mathbf{X}\varvec{\beta }+ \mathbf{T}\varvec{\varepsilon }= \mathbf{T}\mathbf{X}(\mathbf{A}\varvec{\alpha }+ \varvec{\gamma }) + \mathbf{T}\varvec{\varepsilon }\) by pre-multiplying a given matrix \(\mathbf{T}\) of arbitrary rank. These two models are not necessarily equivalent unless \(\mathbf{T}\) is of full column rank, and we have to work with this derived model in many situations. Because predictors/estimators of the parameter spaces under the two models are not necessarily the same, it is primary work to compare predictors/estimators in the two models and to establish possible links between the inference results obtained from two models. This paper presents a general algebraic approach to the problem of comparing best linear unbiased predictors (BLUPs) of parameter spaces in an original linear random-effects model and its transformations, and provides a group of fundamental and comprehensive results on mathematical and statistical properties of the BLUPs. In particular, we construct many equalities for the BLUPs under an original linear random-effects model and its transformations, and obtain necessary and sufficient conditions for the equalities to hold.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  • Arendacká B, Puntanen S (2015) Further remarks on the connection between fixed linear model and mixed linear model. Stat Pap 56:1235–1247

    Article  MATH  MathSciNet  Google Scholar 

  • Baksalary JK, Kala R (1981) Linear transformations preserving best linear unbiased estimators in a general Gauss-Markoff model. Ann Stat 9:913–916

    Article  MATH  MathSciNet  Google Scholar 

  • Baksalary JK, Kala R (1986) Linear sufficiency with respect to a given vector of parametric functions. J Stat Plann Inference 14:331–338

    Article  MATH  MathSciNet  Google Scholar 

  • Baksalary JK, Mathew T (1986) Linear sufficiency and completeness in an incorrectly specified general Gauss-Markov model. Sankhyā Ser A 48:169–180

    MATH  MathSciNet  Google Scholar 

  • Brumback BA (2010) On the built-in restrictions in linear mixed models, with application to smoothing spline analysis of variance. Commun Stat Theory Metheds 39:579–591

    Article  MATH  MathSciNet  Google Scholar 

  • Chaturvedi A, Kesarwani S, Chandra R (2008) Simultaneous prediction based on shrinkage estimator. In: Shalabh, Heumann C (eds) Recent advances in linear models and related areas. Essays in Honour of Helge Toutenburg. Springer, Heidelberg, pp 181–204

  • Demidenko E (2004) Mixed models: theory and applications. Wiley, New York

    Book  MATH  Google Scholar 

  • Dong B, Guo W, Tian Y (2014) On relations between BLUEs under two transformed linear models. J Multivar Anal 131:279–292

    Article  MATH  MathSciNet  Google Scholar 

  • Drygas H (1983) Sufficiency and completeness in the general Gauss-Markov model. Sankhyā Ser A 45:88–98

    MATH  MathSciNet  Google Scholar 

  • Drygas H (1985) Linear sufficiency and some applications in multilinear estimation. J Multivar Anal 16:71–84

    Article  MATH  MathSciNet  Google Scholar 

  • Gan S, Sun Y, Tian Y (2017) Equivalence of predictors under real and over-parameterized linear models. Commun Stat Theory Methods 46:5368–5383

  • Goldberger AS (1962) Best linear unbiased prediction in the generalized linear regression models. J Am Stat Assoc 57:369–375

    Article  MATH  MathSciNet  Google Scholar 

  • Goldstein H (1986) Multilevel mixed linear model analysis using iterative generalized least squares. Biometrika 73:43–56

    Article  MATH  MathSciNet  Google Scholar 

  • Goldstein H (2011) Multilevel statistical models, 4th edn. Wiley, West Sussex

    MATH  Google Scholar 

  • Groß J (1998) A note on the concepts of linear and quadratic sufficiency. J Stat Plann Inference 70:69–76

    Article  MATH  MathSciNet  Google Scholar 

  • Guttman L (1944) General theory and methods for matric factoring. Psychometrika 9:1–16

    Article  MATH  MathSciNet  Google Scholar 

  • Harville D (1976) Extension of the Gauss-Markov theorem to include the estimation of random effects. Ann Stat 4:384–395

    Article  MATH  MathSciNet  Google Scholar 

  • Haslett SJ, Puntanen S (2010a) Equality of BLUEs or BLUPs under two linear models using stochastic restrictions. Stat Pap 51:465–475

    Article  MATH  MathSciNet  Google Scholar 

  • Haslett SJ, Puntanen S (2010b) Note on the equality of the BLUPs for new observations under two linear models. Acta Comment Univ Tartu Math 14:27–33

    MATH  MathSciNet  Google Scholar 

  • Haslett SJ, Puntanen S (2011) On the equality of the BLUPs under two linear mixed models. Metrika 74:381–395

    Article  MATH  MathSciNet  Google Scholar 

  • Haslett SJ, Puntanen S, Arendacká B (2015) The link between the mixed and fixed linear models revisited. Stat Pap 56:849–861

    Article  MATH  MathSciNet  Google Scholar 

  • Heiligers B, Markiewicz A (1996) Linear sufficiency and admissibility in restricted linear models. Stat Prob Lett 30:105–111

    Article  MATH  MathSciNet  Google Scholar 

  • Henderson CR (1975) Best linear unbiased estimation and prediction under a selection model. Biometrics 31:423–447

    Article  MATH  Google Scholar 

  • Ip WC, Wong H, Liu JS (2005) Sufficient and admissible estimators in general multivariate linear model. J Stat Plann Inference 135:371–383

    Article  MATH  MathSciNet  Google Scholar 

  • Isotalo J, Puntanen S (2006a) Linear prediction sufficiency for new observations in the general Gauss-Markov model. Commun Stat Theory Methods 35:1011–1023

    Article  MATH  MathSciNet  Google Scholar 

  • Isotalo J, Puntanen S (2006b) Linear sufficiency and completeness in the partitioned linear model. Acta Comment Univ Tartu Math 10:53–67

    MATH  MathSciNet  Google Scholar 

  • Jiang J (2007) Linear and generalized linear mixed models and their applications. Springer, New York

    MATH  Google Scholar 

  • Kala R, Pordzik PR (2009) Estimation in singular partitioned, reduced or transformed linear models. Stat Pap 50:633–638

    Article  MATH  MathSciNet  Google Scholar 

  • Kala R, Puntanen S, Tian Y (2017) Some notes on linear sufficiency. Stat Pap 58:1–17

  • Kornacki A (1995) The linear sufficiency and invariant linear sufficiency in a multivariate Gauss-Markov model. Metrika 42:240–241

    Article  Google Scholar 

  • Kornacki A (1997) Stability of invariant linearly sufficient statistics in the general Gauss-Markov model. Appl Math 42:71–77

    Article  MATH  MathSciNet  Google Scholar 

  • Kornacki A (2007) Different kinds of sufficiency in the general Gauss-Markov model. Math Slov 57:389–392

    MATH  MathSciNet  Google Scholar 

  • Kreft GG, de Leeuw J (1998) Introduction to multilevel modelling. Sage, Thousand Oaks

    Book  Google Scholar 

  • Laird N, Ware J (1982) Random-effects models for longitudinal data. Biometrics 38:963–974

    Article  MATH  Google Scholar 

  • Lange K, Chi EC, Zhou H (2014) A brief survey of modern optimization for statisticians. Int Stat Rev 82:46–70

    Article  MathSciNet  Google Scholar 

  • Liski EP, Trenkler G, Groß J (1997) Estimation from transformed data under the linear regression model. Statistics 29:205–219

    Article  MATH  MathSciNet  Google Scholar 

  • Longford NT (1993) Random coefficient models. Oxford University Press, Oxford

    MATH  Google Scholar 

  • Lu C, Sun Y, Tian Y (2016) Two competing linear random-effects models and their connections. Stat Pap. doi:10.1007/s00362-016-0806-3

    Google Scholar 

  • Ma R, Jørgensen B (2007) Nested generalized linear mixed models: an orthodox best linear unbiased predictor approach. J R Stat Soc Ser B 69:625–641

    Article  MathSciNet  Google Scholar 

  • Markiewicz A (1998) Comparison of linear restricted models with respect to the validity of admissible and linearly sufficient estimators. Stat Prob Lett 38:347–354

    Article  MATH  MathSciNet  Google Scholar 

  • Markiewicz A, Puntanen S (2009) Admissibility and linear sufficiency in linear model with nuisance parameters. Stat Pap 50:847–854

    Article  MATH  MathSciNet  Google Scholar 

  • Marsaglia G, Styan GPH (1974) Equalities and inequalities for ranks of matrices. Linear Multilinear Algebra 2:269–292

    Article  MATH  MathSciNet  Google Scholar 

  • McLean RA, Sanders WL, Stroup WW (1991) A unified approach to mixed linear models. Am Stat 45:54–64

    Google Scholar 

  • Morrell CH, Pearson JD, Brant LJ (1997) Linear transformations of linear mixed-effects models. Am Stat 51:338–343

    Google Scholar 

  • Müller J (1987) Sufficiency and completeness in the linear model. J Multivar Anal 21:312–323

    Article  MathSciNet  Google Scholar 

  • Muller KE, Stewart PW (2006) Linear model theory: univariate, multivariate, and mixed models. Wiley, New York

    Book  MATH  Google Scholar 

  • Oktaba W, Kornacki A, Wawrzosek J (1988) Invariant linearly sufficient trasformations of the general Gauss-Markoff model. Estimation and testing. Scand J Stat 15:117–124

    MATH  Google Scholar 

  • Penrose R (1955) A generalized inverse for matrices. Proc Camb Philos Soc 51:406–413

    Article  MATH  Google Scholar 

  • Puntanen S, Styan GPH, Isotalo J (2011) Matrix tricks for linear statistical models: our personal top twenty. Springer, Heidelberg

    Book  MATH  Google Scholar 

  • Rao CR (1971) Unified theory of linear estimation. Sankhyā Ser A 33:371–394

    MATH  MathSciNet  Google Scholar 

  • Rao CR (1973) Representations of best linear unbiased estimators in the Gauss-Markoff model with a singular dispersion matrix. J Multivar Anal 3:276–292

    Article  MATH  MathSciNet  Google Scholar 

  • Rao CR (1989) A lemma on optimization of matrix function and a review of the unified theory of linear estimation. In: Dodge Y (ed) Statistical data analysis and inference. Elsevier, North-Holland, pp 397–417

    Google Scholar 

  • Rao CR, Toutenburg H, Shalabh HC (2008) Linear models and generalizations least squares and alternatives, 3rd edn. Springer, Heidelberg

    MATH  Google Scholar 

  • Raudenbush SW, Bryk AS (2002) Hierarchical linear models: applications and data analysis methods, 2nd edn. Sage, Thousand Oaks

    Google Scholar 

  • Schmidt KD (1998) Prediction in the linear model: a direct approach. Metrika 48:141–147

    MATH  MathSciNet  Google Scholar 

  • Searle SR (1997) Built-in restrictions on best linear unbiased predictors (BLUP) of random effects in mixed models. Am Stat 51:19–21

    Google Scholar 

  • Shalabh (1995) Performance of Stein-rule procedure for simultaneous prediction of actual and average values of study variable in linear regression models. Bull Int Stat Instit 56:1375–1390

    Google Scholar 

  • Shao J, Zhang J (2015) A transformation approach in linear mixed-effects models with informative missing responses. Biometrika 102:107–119

    Article  MATH  MathSciNet  Google Scholar 

  • Tian Y (2004) More on maximal and minimal ranks of Schur complements with applications. Appl Math Comput 152:675–692

    MATH  MathSciNet  Google Scholar 

  • Tian Y (2015a) A new derivation of BLUPs under random-effects model. Metrika 78:905–918

    Article  MATH  MathSciNet  Google Scholar 

  • Tian Y (2015b) A matrix handling of predictions under a general linear random-effects model with new observations. Electron J Linear Algebra 29:30–45

    Article  MATH  MathSciNet  Google Scholar 

  • Tian Y, Jiang B (2016) An algebraic study of BLUPs under two linear random-effects models with correlated covariance matrices. Linear Multilinear Algebra 64:2351–2367

  • Tian Y, Jiang B (2017) A new analysis of the relationships between a general linear model and its mis-specified forms. J Korean Stat Soc 46:182–193

  • Tian Y, Puntanen S (2009) On the equivalence of estimations under a general linear model and its transformed models. Linear Algebra Appl 430:2622–2641

    Article  MATH  MathSciNet  Google Scholar 

  • Toutenburg H (1982) Prior information in linear models. Wiley, New York

    MATH  Google Scholar 

  • Verbeke G, Molenberghs G (2000) Linear mixed models for longitudinal data. Springer, New York

    MATH  Google Scholar 

Download references

Acknowledgements

The author is grateful to an anonymous referee for his/her helpful comments and constructive suggestions to an earlier version of this paper. This work was supported by the National Natural Science Foundation of China (Grant No. 11271384).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yongge Tian.

Appendix

Appendix

Statistical methods in many areas of application often involve mathematical computations with vectors and matrices. In particular, formulas and algebraic tricks for handling matrices in linear algebra and matrix theory play important roles in the derivation of predictors/estimators and characterizations of their properties under linear regression models. Recall that the rank of matrix is a conceptual foundation in linear algebra and matrix theory, which is the most significant finite nonnegative integer in reflecting intrinsic properties of matrices. The mathematical prerequisites for understanding the rank of matrix are minimal and do not go beyond elementary linear algebra. It has long history to establish rank formulas for block matrices and use the formulas in statistical inference, and a pioneer work in this aspect can be found in Guttman (1944). The intriguing connections between generalized inverses of matrices and rank formulas of matrices were recognized in 1970s, and a seminal work on rank formulas for matrices and their generalized inverses was presented in Marsaglia and Styan (1974). In order to establish and characterize various possible equalities for predictors/estimators under LRMs, and to simplify various matrix equalities composed by the Moore–Penrose inverses of matrices, we need a variety of matrix rank formulas for matrices and their Moore–Penrose generalized inverses as simple and effective tools. The results in the following lemma were established in Marsaglia and Styan (1974) and Tian (2004).

Lemma 2

Let \(\mathbf{A}\in \mathbb R^{m \times n},\) \( \mathbf{B}\in \mathbb R^{m \times k},\) and \(\mathbf{C}\in \mathbb R^{l \times n}.\) Then

$$\begin{aligned} r[\, \mathbf{A}, \, \mathbf{B}\, ]&= r(\mathbf{A})+ r(\mathbf{E}_{\mathbf{A}}\mathbf{B})= r(\mathbf{B}) + r(\mathbf{E}_{\mathbf{B}}\mathbf{A}), \end{aligned}$$
(A. 1)
$$\begin{aligned} r\begin{bmatrix} \mathbf{A}\\ \mathbf{C}\end{bmatrix}&= r(\mathbf{A}) + r(\mathbf{C}\mathbf{F}_{\mathbf{A}}) = r(\mathbf{C}) + r(\mathbf{A}\mathbf{F}_{\mathbf{C}}), \end{aligned}$$
(A. 2)
$$\begin{aligned} r\begin{bmatrix} \mathbf{A}\mathbf{A}^{\prime }&\mathbf{B}\\ \mathbf{B}^{\prime }&\mathbf{0}\end{bmatrix}&= r[\, \mathbf{A}, \, \mathbf{B}\,] + r(\mathbf{B}). \end{aligned}$$
(A. 3)

If \({\mathscr {R}}(\mathbf{A}_1^{\prime }) \subseteq {\mathscr {R}}(\mathbf{B}_1^{\prime })\), \({\mathscr {R}}(\mathbf{A}_2) \subseteq {\mathscr {R}}(\mathbf{B}_1)\), \({\mathscr {R}}(\mathbf{A}_2^{\prime }) \subseteq {\mathscr {R}}(\mathbf{B}_2^{\prime })\) and \({\mathscr {R}}(\mathbf{A}_3) \subseteq {\mathscr {R}}(\mathbf{B}_2),\) then

$$\begin{aligned} r(\mathbf{A}_1\mathbf{B}_1^{+}\mathbf{A}_2)&= r\begin{bmatrix} \mathbf{B}_1&\mathbf{A}_2\\ \mathbf{A}_1&\mathbf{0}\end{bmatrix} - r(\mathbf{B}_1), \end{aligned}$$
(A. 4)
$$\begin{aligned} r(\mathbf{A}_1\mathbf{B}_1^{+}\mathbf{A}_2\mathbf{B}_2^{+}\mathbf{A}_3)&= r\begin{bmatrix} \mathbf{0}&\mathbf{B}_2&\mathbf{A}_3\\ \mathbf{B}_1&\mathbf{A}_2&\mathbf{0}\\ \mathbf{A}_1&\mathbf{0}&\mathbf{0}\end{bmatrix} - r(\mathbf{B}_1) - r(\mathbf{B}_2). \end{aligned}$$
(A. 5)

In addition,  the following results hold.

  1. (a)

    \(r[\, \mathbf{A}, \, \mathbf{B}\,] = r(\mathbf{A}) \Leftrightarrow {\mathscr {R}}(\mathbf{B}) \subseteq {\mathscr {R}}(\mathbf{A}) \Leftrightarrow \mathbf{A}\mathbf{A}^{+}\mathbf{B}= \mathbf{B}\Leftrightarrow \mathbf{E}_{\mathbf{A}}\mathbf{B}= \mathbf{0}.\)

  2. (b)

    \(r\begin{bmatrix} \mathbf{A}\\ \mathbf{C}\end{bmatrix} = r(\mathbf{A}) \Leftrightarrow {\mathscr {R}}(\mathbf{C}^{\prime }) \subseteq {\mathscr {R}}(\mathbf{A}^{\prime }) \Leftrightarrow \mathbf{C}\mathbf{A}^{+}\mathbf{A}= \mathbf{C}\Leftrightarrow \mathbf{C}\mathbf{F}_{\mathbf{A}} = \mathbf{0}.\)

  3. (c)

    \( r[\, \mathbf{A}, \, \mathbf{B}\,] = r(\mathbf{A}) + r(\mathbf{B}) \Leftrightarrow {\mathscr {R}}(\mathbf{A}) \cap {\mathscr {R}}(\mathbf{B}) = \{ \mathbf{0}\} \Leftrightarrow {\mathscr {R}}[(\mathbf{E}_\mathbf{A}\mathbf{B})^{\prime }] = {\mathscr {R}}(\mathbf{B}^{\prime }) \Leftrightarrow {\mathscr {R}}[(\mathbf{E}_\mathbf{B}\mathbf{A})^{\prime }] = {\mathscr {R}}(\mathbf{A}^{\prime }).\)

  4. (d)

    \( r \begin{bmatrix} \mathbf{A}\\ \mathbf{C}\end{bmatrix} = r( \mathbf{A}) + r(\mathbf{C}) \Leftrightarrow {\mathscr {R}}(\mathbf{A}^{\prime }) \cap {\mathscr {R}}(\mathbf{C}^{\prime }) = \{ \mathbf{0}\} \Leftrightarrow {\mathscr {R}}(\mathbf{C}\mathbf{F}_{\mathbf{A}}) = {\mathscr {R}}(\mathbf{C}) \Leftrightarrow {\mathscr {R}}(\mathbf{A}\mathbf{F}_{\mathbf{C}}) = {\mathscr {R}}(\mathbf{A}). \)

Lemma 3

(Penrose 1955) The linear matrix equation \(\mathbf{A}\mathbf{X}= \mathbf{B}\) is consistent if and only if \(r[\mathbf{A}, \, \mathbf{B}\,] = r(\mathbf{A}),\) or equivalently,  \(\mathbf{A}\mathbf{A}^{+}\mathbf{B}= \mathbf{B}.\) In this case,  the general solution of the equation can be written in the following parametric form \(\mathbf{X}= \mathbf{A}^{+} \mathbf{B}+ ( \mathbf{I}- \mathbf{A}^{+}\mathbf{A}) \mathbf{U},\) where \(\mathbf{U}\) is an arbitrary matrix.

Once a statistical model is formulated, its parameters can be estimated and predicted by various optimization methods. A brief survey on modern optimization methods in statistical analysis can be found in Lange et al. (2014). In any case, we expect that the optimization problems occurred in predictions/stimations of parameter spaces in linear models have analytical solutions so that we can use the analytical solutions to establish precise theory in the corresponding statistical inferences of the models. In order to directly solve the matrix minimization problem in (2.4), we need the following result on constrained quadratic matrix-valued function minimization problem, which have been recently proved in Tian (2015a) and have been applied in Gan et al. (2017), Lu et al. (2016), Tian (2015a, b) and Tian and Jiang (2016, 2017) to establish exact expressions of BLUPs under linear models with fixed and random effects.

Lemma 4

Let

$$\begin{aligned} f(\mathbf{L}) = \left( \mathbf{L}\mathbf{C}+ \mathbf{D}\right) \mathbf{M}\left( \mathbf{L}\mathbf{C}+ \mathbf{D}\right) ^{\prime } \ \ s.t. \ \ \mathbf{L}\mathbf{A}= \mathbf{B}, \end{aligned}$$

where \(\mathbf{A}\in {\mathbb R}^{p \times q}\), \(\mathbf{B}\in {\mathbb R}^{n \times q}\), \(\mathbf{C}\in {\mathbb R}^{p\times m}\) and \(\mathbf{D}\in {\mathbb R}^{n\times m}\) are given,  \(\mathbf{M}\in {\mathbb R}^{m \times m}\) is positive semi-definite, and the matrix equation \(\mathbf{L}\mathbf{A}= \mathbf{B}\) is consistent. Then there always exists a solution \(\mathbf{L}_0\) of \(\mathbf{L}_0\mathbf{A}= \mathbf{B}\) such that

$$\begin{aligned} f(\mathbf{L}) \succcurlyeq f(\mathbf{L}_0) \end{aligned}$$

holds for all solutions of \(\mathbf{L}\mathbf{A}= \mathbf{B}.\) In this case,  the matrix \(\mathbf{L}_0\) satisfying the above inequality is determined by the following consistent matrix equation

$$\begin{aligned} \mathbf{L}_0\left[ \mathbf{A}, \, \mathbf{C}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp } \right] = \left[ \mathbf{B}, \, -\mathbf{D}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp } \right] . \end{aligned}$$

In this case,  the general expression of \(\mathbf{L}_0\) and the corresponding \(f(\mathbf{L}_0)\) and \(f(\mathbf{L})\) are given by

$$\begin{aligned} \mathbf{L}_0&= \mathop {{{\mathrm{argmin}}}}\limits _{\mathbf{L}\mathbf{A}= \mathbf{B}} f(\mathbf{L}) = \left[ \mathbf{B}, \, -\mathbf{D}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp }\right] \left[ \mathbf{A}, \, \mathbf{C}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp } \right] ^{+} + \mathbf{U}\left[ \mathbf{A}, \, \mathbf{C}\mathbf{M}\mathbf{C}^{\prime }\right] ^{\perp },\\ f(\mathbf{L}_0)&= \min _{\mathbf{L}\mathbf{A}= \mathbf{B}} f(\mathbf{L}) = \mathbf{K}\mathbf{M}\mathbf{K}^{\prime } - \mathbf{K}\mathbf{M}\mathbf{C}^{\prime }\mathbf{T}\mathbf{C}\mathbf{M}\mathbf{K}^{\prime },\\ f(\mathbf{L})&= f(\mathbf{L}_0) + \left( \mathbf{L}\mathbf{C}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp } + \mathbf{D}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp }\right) \mathbf{T}\left( \mathbf{L}\mathbf{C}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp } + \mathbf{D}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp }\right) ^{\prime }, \end{aligned}$$

where \(\mathbf{K}= \mathbf{B}\mathbf{A}^{+}\mathbf{C}+ \mathbf{D},\) \(\mathbf{T}= \left( \mathbf{A}^{\perp }\mathbf{C}\mathbf{M}\mathbf{C}^{\prime }\mathbf{A}^{\perp }\right) ^{+},\) and \(\mathbf{U}\in {\mathbb R}^{n\times p}\) is arbitrary.

With the supports of the formulas in Lemmas 24, we are capable of proving all the assertions in Sects. 2 and 3.

Proof of Lemma 1

It is obvious that \(\mathrm{E}\left( \mathbf{L}{Ty} - \varvec{\phi }\,\right) = \mathbf{0}\Leftrightarrow \mathbf{L}\widehat{\mathbf{X}}\varvec{\alpha }- \mathbf{K}\varvec{\alpha }=\mathbf{0}\ \hbox {for all } \varvec{\alpha }\Leftrightarrow \mathbf{L}\widehat{\mathbf{X}}= \mathbf{K}\). From Lemma 3, the matrix equation is consistent if and only if (2.10) holds. \(\square \)

Proof of Theorem 1

Under (2.9), (2.4) is equivalent to finding a solution \(\mathbf{L}_0\) of \(\mathbf{L}_0\mathbf{T}\widehat{\mathbf{X}} = \mathbf{K}\) such that

$$\begin{aligned} f(\mathbf{L}) \succcurlyeq f(\mathbf{L}_0) \ \ \hbox {s.t.} \ \ \mathbf{L}\mathbf{T}\widehat{\mathbf{X}} = \mathbf{K} \end{aligned}$$
(A. 6)

holds in the Löwner partial ordering. From Lemma 4, there always exists a solution \(\mathbf{L}_0\) of \(\mathbf{L}_0\mathbf{T}\widehat{\mathbf{X}} = \mathbf{K}\) such that \(f(\mathbf{L}) \succcurlyeq f(\mathbf{L}_0)\) holds for all solutions of \(\mathbf{L}\mathbf{T}\widehat{\mathbf{X}} = \mathbf{K}\), and the \(\mathbf{L}_0\) is determined by the matrix equation \(\mathbf{L}_0[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] = [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]\), establishing the matrix equation in (2.11). The general solution of the equation is given in (2.13) by Lemma 3.

Results (a)–(d) are routine consequences of (2.13).

Taking dispersion operation of (2.13) yields (2.14) and (2.16). Also from (1.33) and (2.13), the covariance matrix between \({\mathrm{BLUP}}_{{\mathscr {N}}}(\varvec{\phi })\) and \(\varvec{\phi }\) is

$$\begin{aligned} \mathrm{Cov}\{\,{\mathrm{BLUP}}_{{\mathscr {N}}}(\varvec{\phi }), \, \varvec{\phi }\,\}&= \mathbf{L}\mathbf{T}\mathrm{Cov}\{\,\mathbf{y}, \, \varvec{\phi }\}\\&= [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\mathbf{W}^{\prime }, \end{aligned}$$

thus establishing (2.15). Further by (2.13),

$$\begin{aligned} \mathbf{L}\mathbf{T}\widetilde{\mathbf{X}} = \left( [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+} + \mathbf{U}[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp } \right) \mathbf{T}\widetilde{\mathbf{X}}. \end{aligned}$$

Substituting it into (2.9) yields

$$\begin{aligned} \mathrm{Cov}(\varvec{\phi }- \mathbf{L}\mathbf{T}\mathbf{y})&= \left( [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\widetilde{\mathbf{X}} - \mathbf{J}\right) \varvec{\Sigma }\\&\ \ \ \times \left( [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\widetilde{\mathbf{X}} - \mathbf{J}\right) ^{\prime }, \end{aligned}$$

thus establishing (2.17).

Note that the arbitrary matrix \(\mathbf{U}\) in (2.13) can be rewritten as \(\mathbf{U}= \mathbf{U}_1 + \mathbf{U}_2 + \mathbf{U}_3\), while \([\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\) in (2.13) can be decomposed as the sum of three terms

$$\begin{aligned}&[\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\\&= [\,\mathbf{K}, \, \mathbf{0}\,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\\&\quad + [\,\mathbf{0}, \, \mathbf{G}[\mathbf{I}_p, \, \mathbf{0}]\varvec{\Sigma }(\mathbf{T}\widetilde{\mathbf{X}})^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,][\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\\&\quad + [\,\mathbf{0}, \, \mathbf{H}[\mathbf{0}, \, \mathbf{I}_n]\varvec{\Sigma }(\mathbf{T}\widetilde{\mathbf{X}})^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,][\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\\&: = \mathbf{Q}_1 + \mathbf{Q}_2 + \mathbf{Q}_3. \end{aligned}$$

Hence

$$\begin{aligned} {\mathrm{BLUP}}_{{\mathscr {N}}}(\varvec{\phi })&= \left( \mathbf{Q}_1 + \mathbf{U}_1[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }\mathbf{T}\right) \mathbf{y}\\&\ \ \ \ + \left( \mathbf{Q}_2 + \mathbf{U}_2[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }\mathbf{T}\right) \mathbf{y}\\&\ \ \ \ + \left( \mathbf{Q}_3 + \mathbf{U}_3[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }\mathbf{T}\right) \mathbf{y}\\&= \mathrm{BLUE}_{{\mathscr {N}}}(\mathbf{K}\varvec{\alpha }) + {\mathrm{BLUP}}_{{\mathscr {N}}}(\mathbf{G}\varvec{\gamma }) + {\mathrm{BLUP}}_{{\mathscr {N}}}(\mathbf{H}\varvec{\varepsilon }), \end{aligned}$$

thus establishing (2.18).

From (2.13), the covariance matrix between \(\mathrm{BLUE}_{{\mathscr {N}}}(\mathbf{K}\varvec{\alpha })\) and \( {\mathrm{BLUP}}_{{\mathscr {N}}}(\mathbf{G}\varvec{\gamma }+ \mathbf{H}\varvec{\varepsilon })\) is

$$\begin{aligned}&{\mathrm{Cov}}\{\, {\mathrm{BLUE}}_{{\mathscr {N}}}(\mathbf{K}\varvec{\alpha }), \, {\mathrm{BLUP}}_{{\mathscr {N}}}(\mathbf{G}\varvec{\gamma }+ \mathbf{H}\varvec{\varepsilon }) \,\} \nonumber \\&= [\,\mathbf{K}, \, \mathbf{0}\,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}\mathbf{V}\mathbf{T}^{\prime }\left( [\,\mathbf{0}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+} \right) ^{\prime }. \end{aligned}$$
(A. 7)

Applying (A. 5) to (A. 7) and simplifying, we obtain

$$\begin{aligned}&r({\mathrm{Cov}}\{\,{\mathrm{BLUE}}_{{\mathscr {N}}}(\mathbf{K}\varvec{\alpha }), \, {\mathrm{BLUP}}_{{\mathscr {N}}}(\mathbf{G}\varvec{\gamma }+ \mathbf{H}\varvec{\varepsilon })\,\}) \\&\quad = r\left( [\,\mathbf{K}, \, \mathbf{0}\,][\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp } \,]^{+}\mathbf{T}\mathbf{V}\mathbf{T}^{\prime }\left( [\,\mathbf{0}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp } \,] [\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp } \,]^{+}\right) ^{\prime } \right) \\&\quad = r\begin{bmatrix} \mathbf{0}&\begin{bmatrix} (\mathbf{T}{\widehat{\mathbf{X}}})^{\prime }\\ (\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }\mathbf{T}\mathbf{V}\mathbf{T}^{\prime } \end{bmatrix}&\begin{bmatrix} \mathbf{0}\\ (\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }\mathbf{T}\mathbf{W}^{\prime } \end{bmatrix} \\ [\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }\,]&\mathbf{T}\mathbf{V}\mathbf{T}^{\prime }&\mathbf{0}\\ [\,\mathbf{K}, \, \mathbf{0}\,]&\mathbf{0}&\mathbf{0}\end{bmatrix} - 2r[\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }]\\&\quad = r\begin{bmatrix} \begin{bmatrix} \mathbf{0}&\mathbf{0}\\ \mathbf{0}&- (\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }\mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp } \end{bmatrix}&\begin{bmatrix} (\mathbf{T}{\widehat{\mathbf{X}}})^{\prime }\\ \mathbf{0}\end{bmatrix}&\begin{bmatrix} \mathbf{0}\\ (\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }\mathbf{T}\mathbf{W}^{\prime } \end{bmatrix} \\ [\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{0}\,]&\mathbf{T}\mathbf{V}\mathbf{T}^{\prime }&\mathbf{0}\\ [\,\mathbf{K}, \, \mathbf{0}\,]&\mathbf{0}&\mathbf{0}\end{bmatrix} - 2r[\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }] \\&\quad = r\begin{bmatrix} \mathbf{0}&(\mathbf{T}{\widehat{\mathbf{X}}})^{\prime } \\ \mathbf{T}{\widehat{\mathbf{X}}}&\mathbf{T}\mathbf{V}\mathbf{T}^{\prime } \\ \mathbf{K}&\mathbf{0}\end{bmatrix} + r\left[ \, (\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }\mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }, \, (\mathbf{T}{\widehat{\mathbf{X}}})^{\perp }\mathbf{T}\mathbf{W}^{\prime } \right] - 2r\left[ \,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}{\widetilde{\mathbf{X}}}\varvec{\Sigma }\,\right] \\&\quad = r\begin{bmatrix} \mathbf{T}{\widehat{\mathbf{X}}}\\ \mathbf{K}\end{bmatrix} + r\begin{bmatrix} (\mathbf{T}{\widehat{\mathbf{X}}})^{\prime }\\ \mathbf{T}\mathbf{V}\mathbf{T}^{\prime } \end{bmatrix} + r\left[ \,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}{\widetilde{\mathbf{X}}}\varvec{\Sigma }, \, \mathbf{T}\mathbf{W}^{\prime }\right] \\&\qquad - r(\mathbf{T}{\widehat{\mathbf{X}}}) - 2r[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}{\widetilde{\mathbf{X}}}\varvec{\Sigma }\,] \quad (by (A.1) \hbox { and } (A.3)) \\&\quad = r(\mathbf{T}{\widehat{\mathbf{X}}}) + r\begin{bmatrix} (\mathbf{T}{\widehat{\mathbf{X}}})^{\prime }\\ \varvec{\Sigma }(\mathbf{T}{\widetilde{\mathbf{X}}})^{\prime } \end{bmatrix} + r[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}{\widetilde{\mathbf{X}}}\varvec{\Sigma }\,] - r(\mathbf{T}{\widehat{\mathbf{X}}}) - 2r[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}{\widetilde{\mathbf{X}}}\varvec{\Sigma }\,] \\&\quad = 0, \end{aligned}$$

which implies that \({\mathrm{Cov}}\{\,{\mathrm{BLUE}}_{{\mathscr {N}}}(\mathbf{K}\varvec{\alpha }), \, {\mathrm{BLUP}}_{{\mathscr {N}}}(\mathbf{G}\varvec{\gamma }+ \mathbf{H}\varvec{\varepsilon })\,\}\) is a zero matrix, establishing (2.19). Equation (2.20) follows from (2.18) and (2.19). Result (g) follows from (2.13). \(\square \)

Proof of Theorem 2

From Definition 3(a), \({\mathrm{BLUP}}_{{\mathscr {M}}}(\varvec{\phi }) = \mathbf{G}\mathbf{T}\mathbf{y}\) holds for some matrix \(\mathbf{G}\) if and only if the coefficient matrix \(\mathbf{G}\mathbf{T}\) of \(\mathbf{y}\) satisfies the matrix equation in (2.22), i.e.,

$$\begin{aligned} \mathbf{G}\mathbf{T}[\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } ] = [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } ]. \end{aligned}$$
(A. 8)

From Lemma 3, the equation in (A. 8) is solvable for \(\mathbf{G}\) if and only if

$$\begin{aligned} r\begin{bmatrix} \mathbf{T}{\widehat{\mathbf{X}}}&\quad \mathbf{T}\mathbf{V}{\widehat{\mathbf{X}}}^{\perp } \\ \mathbf{K}&\quad \mathbf{W}{\widehat{\mathbf{X}}}^{\perp } \end{bmatrix} = r[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}{\widehat{\mathbf{X}}}^{\perp } \,], \end{aligned}$$
(A. 9)

where by (1.5), (1.33), and (A. 2),

$$\begin{aligned}&r\begin{bmatrix} \mathbf{T}\widehat{\mathbf{X}}&\quad \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \\ \mathbf{K}&\quad \mathbf{W}\widehat{\mathbf{X}}^{\perp } \end{bmatrix} = r\begin{bmatrix} \mathbf{T}\widehat{\mathbf{X}}&\quad {\mathrm{Cov}}\{\mathbf{T}\mathbf{y},\, \widehat{\mathbf{X}}^{\perp }\mathbf{y}\}\\ \mathbf{K}&\quad {\mathrm{Cov}}\{\varvec{\phi },\, \widehat{\mathbf{X}}^{\perp }\mathbf{y}\} \end{bmatrix} = r\begin{bmatrix} \mathbf{T}{\widehat{\mathbf{X}}}&\quad {\mathrm{Cov}}\{\mathbf{T}\mathbf{y},\, \mathbf{y}\}\\ \mathbf{0}&\quad \widehat{\mathbf{X}}^{\prime }\\ \mathbf{K}&\quad {\mathrm{Cov}}\{\varvec{\phi },\, \mathbf{y}\} \end{bmatrix} - r(\widehat{\mathbf{X}}),\\&r[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}{\widehat{\mathbf{X}}}^{\perp } \,] = r[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathrm{Cov}\{\mathbf{T}\mathbf{y},\, {\widehat{\mathbf{X}}}^{\perp }\mathbf{y}\}\,] = r\begin{bmatrix} \mathbf{T}{\widehat{\mathbf{X}}}&\quad {\mathrm{Cov}}\{\mathbf{T}\mathbf{y},\, \mathbf{y}\}\\ \mathbf{0}&\quad {\widehat{\mathbf{X}}}^{\prime } \end{bmatrix} - r({\widehat{\mathbf{X}}}). \end{aligned}$$

Hence (A. 9) is equivalent to (d) and (e), respectively. Thus, (a), (d), and (e) are equivalent. The equivalence of (e) and (f) follows from Lemma 2(b).

From Definition 3(a), \({\mathrm{BLUP}}_{{\mathscr {M}}}(\varvec{\phi })= {\mathrm{BLUP}}_{{\mathscr {N}}}(\varvec{\phi })\) holds definitely if and only if the coefficient matrix of \(\mathbf{y}\) in (2.13) satisfies the matrix equation in (2.22), i.e.,

$$\begin{aligned}&\left( [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}+ \mathbf{U}[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }\mathbf{T}\right) [\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]\nonumber \\&\quad = [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,]. \end{aligned}$$
(A. 10)

From Lemma 3, this equation is solvable for \(\mathbf{U}\) if and only if

$$\begin{aligned}&r\begin{bmatrix} [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,] - [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,] \\ [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,] \end{bmatrix} \nonumber \\&\quad = r([\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]). \end{aligned}$$
(A. 11)

Simplifying both sides by elementary block matrix operations, we obtain

$$\begin{aligned}&r\begin{bmatrix} [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,] \quad -[\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,] \\ [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,] \end{bmatrix}\\&\quad =r\begin{bmatrix} [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,] - [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]&\mathbf{0}\\ [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]&[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] \end{bmatrix}\\&\qquad \,- \ r[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]\\&\quad =r\begin{bmatrix} [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,]&[\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] \\ [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]&[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] \end{bmatrix} - r[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\widetilde{\mathbf{X}}\varvec{\Sigma }\,]\\&\quad =r\begin{bmatrix} \mathbf{K}&\quad \mathbf{W}\widehat{\mathbf{X}}^{\perp } \\ \mathbf{T}\widehat{\mathbf{X}}&\quad \mathbf{T}\mathbf{V}\widehat{\mathbf{X}}^{\perp } \end{bmatrix} - r[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\widetilde{\mathbf{X}}\varvec{\Sigma }\,]\\&\quad = r\begin{bmatrix} \mathbf{T}\widehat{\mathbf{X}}&\quad {\mathrm{Cov}}\{\mathbf{T}\mathbf{y},\, \widehat{\mathbf{X}}^{\perp }\mathbf{y}\}\\ \mathbf{K}&\quad {\mathrm{Cov}}\{\varvec{\phi },\, \widehat{\mathbf{X}}^{\perp }\mathbf{y}\} \end{bmatrix} - r[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\widetilde{\mathbf{X}}\varvec{\Sigma }\,]\\&\quad = r\begin{bmatrix} \mathbf{T}\widehat{\mathbf{X}}&\quad {\mathrm{Cov}}\{\mathbf{T}\mathbf{y},\, \mathbf{y}\}\\ \mathbf{0}&\quad \widehat{\mathbf{X}}^{\prime }\\ \mathbf{K}&\quad {\mathrm{Cov}}\{\varvec{\phi },\, \mathbf{y}\} \end{bmatrix} - r(\widehat{\mathbf{X}}) - r[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\widetilde{\mathbf{X}}\varvec{\Sigma }\,], \end{aligned}$$

and

$$\begin{aligned}&r([\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}{\widehat{\mathbf{X}}})^{\perp } \,]^{\perp }[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\mathbf{V}{\widehat{\mathbf{X}}}^{\perp } \,]) = r[\, \mathbf{T}\widehat{\mathbf{X}}, \, {\mathrm{Cov}}\{\mathbf{T}\mathbf{y},\, \widehat{\mathbf{X}}^{\perp }\mathbf{y}\}\,] - r[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\widetilde{\mathbf{X}}\varvec{\Sigma }\,]\\&\quad = r\begin{bmatrix} \mathbf{T}{\widehat{\mathbf{X}}}&\quad {\mathrm{Cov}}\{\mathbf{T}\mathbf{y},\, \mathbf{y}\}\\ \mathbf{0}&\quad {\widehat{\mathbf{X}}}^{\prime } \end{bmatrix} - r({\widehat{\mathbf{X}}}) - r[\,\mathbf{T}{\widehat{\mathbf{X}}}, \, \mathbf{T}\widetilde{\mathbf{X}}\varvec{\Sigma }\,]. \end{aligned}$$

Substituting the two equalities into (A. 11) leads to the equivalences of (b), (d), and (e).

From Definition 3(b), \({\mathrm{BLUP}}_{{\mathscr {M}}}(\varvec{\phi })= {\mathrm{BLUP}}_{{\mathscr {N}}}(\varvec{\phi })\) holds with probability 1 if and only if the coefficient matrices in (2.13) and (2.23) satisfy

$$\begin{aligned}&\left( [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,] [\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]^{+} + \mathbf{U}_1[\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]^{\perp } \right) [\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]\\&\quad = \left( [\,\mathbf{K}, \, \mathbf{W}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,] [\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{+}\mathbf{T}+ \mathbf{U}[\,\mathbf{T}\widehat{\mathbf{X}}, \, \mathbf{T}\mathbf{V}\mathbf{T}^{\prime }(\mathbf{T}\widehat{\mathbf{X}})^{\perp } \,]^{\perp }\mathbf{T}\right) \\&\qquad \quad \times \,[\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,], \end{aligned}$$

which can further reduce to (A. 10) from

$$\begin{aligned} {[}\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,][\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]^{+}[\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,] = [\,\mathbf{K}, \, \mathbf{W}\widehat{\mathbf{X}}^{\perp } \,] \ \ {\mathrm{and}} \ \ [\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,]^{\perp }[\,\widehat{\mathbf{X}}, \, \mathbf{V}\widehat{\mathbf{X}}^{\perp } \,] = \mathbf{0}. \end{aligned}$$

Hence, (b) and (c) are equivalent. \(\square \)

Proof of Theorem 3

Equations (3.1) and (3.2) follow from (2.18) and (2.28) by setting \(\varvec{\phi }= \mathbf{T}\mathbf{y}= \mathbf{T}\widehat{\mathbf{X}}\varvec{\alpha }+ \mathbf{T}\mathbf{X}\varvec{\gamma }+ \mathbf{T}\varvec{\varepsilon }\). \(\square \)

Proof of Theorem 4

Results (a)–(c) follow from Theorem 2 by setting \(\varvec{\phi }=\mathbf{T}\mathbf{X}\varvec{\beta }\), \(\mathbf{T}\widehat{\mathbf{X}}\varvec{\alpha }\), \(\mathbf{T}\mathbf{X}\varvec{\gamma }\), and \(\mathbf{T}\varvec{\varepsilon }\), respectively. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tian, Y. Transformation approaches of linear random-effects models. Stat Methods Appl 26, 583–608 (2017). https://doi.org/10.1007/s10260-017-0381-3

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10260-017-0381-3

Keywords

Mathematics Subject Classification

Navigation