• No results found

Extended GMANOVA Model with a Linearly Structured Covariance Matrix

N/A
N/A
Protected

Academic year: 2021

Share "Extended GMANOVA Model with a Linearly Structured Covariance Matrix"

Copied!
17
0
0

Loading.... (view fulltext now)

Full text

(1)

' $

Department of Mathematics

Extended GMANOVA Model With a Linearly

Structured Covariance Matrix

Joseph Nzabanita, Dietrich von Rosen, Martin Singull

LiTH-MAT-R--2015/07--SE

(2)

Department of Mathematics Link¨oping University S-581 83 Link¨oping, Sweden.

(3)

Extended GMANOVA model with a linearly

structured covariance matrix

Joseph Nzabanita1,a, Dietrich von Rosen1,b, Martin Singull1

1Department of Mathematics,

Link¨oping University, SE–581 83 Link¨oping, Sweden. E-mail: joseph.nzabanita@liu.se E-mail: dietrich.von.rosen@liu.se

E-mail: martin.singull@liu.se

aDepartment of Mathematics,

University of Rwanda, PO.Box 3900 Kigali, Rwanda E-mail: j.nzabanita@ur.ac.rw

bDepartment of Energy and Technology,

Swedish University of Agricultural Sciences, SE–750 07 Uppsala, Sweden.

E-mail: dietrich.von.rosen@slu.se

Abstract

In this paper we consider the extended generalized multivariate analysis of variance (GMANOVA) with a linearly structured covariance matrix. The main theme is to find explicit estimators for the mean and for the linearly structured covariance matrix. We show how to decompose the residual space, the orthogonal complement to the mean space, into m + 1 orthogonal subspaces and how to derive explicit estimators of the covariance matrix from the sum of squared residuals obtained by projecting observations on those subspaces. Also an explicit estimator of the mean is derived and some properties of the proposed estimators are studied.

Keywords: estimation, extended growth curve model, GMANOVA, linearly struc-tured covariance matrix, residuals

1

Introduction

The growth curve problems are widely studied in several research areas such as medicine, econometrics, natural sciences, social sciences, etc. These problems are related to the analysis of short time series data where a characteristic of interest is measured on each unit over sev-eral time points. The classical growth curve model (GCM) by [1] also known in the statistical literature as the generalized multivariate analysis of variance (GMANOVA) has emerged as

(4)

a powerful tool to deal with such a kind of problems. The GMANOVA model was extended later on by [2] to handle different growth profiles. The resulting model was called sum of pro-files model and is known in different names such as extended growth curve model (EGCM), extended GMANOVA model, etc. In [2] an iterative algorithm to obtain the maximum likeli-hood estimators (MLEs), which could not be obtained explicitly, was proposed. von Rosen [3] studied the model and derived explicit MLEs under the additional nested subspaces condition on the between design matrices.

In this paper we consider the extended GMANOVA model as defined in [3]. Before we define it, we give some notations that we will use throughout this paper. C(A), r(A) and tr(A) denote the column space, the rank and the trace of a matrix A respectively. For a positive definite matrix S and any matrix A, PA,S = A(A′S−1A)−A′S−1 defines the projector onto

the space CS(A), where the subscript S in CS(A) indicates that the inner products are defined

via the positive definite matrix S−1. PoA,S = I − P′A,S = PAo,S−1 is a projector onto the

space C(A)⊥ = C(Ao), where Ao denotes any matrix of full rank spanning the orthogonal complement to the space C(A). If S = I, we simply write PA instead of PA,I. The symbol ⊗

denotes the Kronecker product of matrices or the tensor product of linear spaces. The symbol ⊕ denotes the direct sum of linear spaces while the symbol ⊞ denotes the direct sum of tensor spaces.

Definition 1.1 Let X : p × n, Ai : p × qi, Bi: qi× ki, Ci: ki× n, r1+ p ≤ n, i = 1, 2, . . . , m,

C(C′

i) ⊆ C(C′i−1), i = 2, 3, . . . , m, where ri = r(Ci). The Extended Growth Curve Model is

given by X = m X i=1 AiBiCi+ E,

where columns of E are assumed to be independently distributed as a p-variate normal distri-bution with mean zero and a positive definite dispersion matrix Σ; i.e. E ∼ Np,n(0, Σ, In).

The matrices Ai and Ci, often called design matrices, are known matrices whereas matrices

Bi and Σ are unknown parameter matrices.

The model in Definition 1.1 has been extensively studied by several authors and the book by Kollo and von Rosen [4] [Chapter 4] contains useful detailed information about uniqueness, estimability conditions, moments and approximative distributions of the maximum likelihood estimators. Recently other authors considered the model with slightly different conditions. For example in [5] the explicit MLEs are presented with the nested subspace conditions on the within design matrices instead. In [6, 7] the extended growth curve model without nested subspace conditions but with orthogonal design matrices is considered and generalized least-squares estimators and their properties are studied.

In most works on the extended GMANOVA model no particular attention is made on the structure of the covariance matrix. In fact there are few articles treating the problem of structured covariance matrix although it may be important in the growth curve analysis studies. For the classical growth curve model, the most studied structure are the uniform covariance structure and the serial covariance structure, see for example, [8, 9, 10]. The main

(5)

theme of this paper is to derive explicit estimators of parameters in the extended growth curve model with a linearly structured covariance matrix, which means that for Σ = (σij) the only

linear structure between the elements is given by |σij| = |σkl| 6= 0 and there exist at least one

(i, j) 6= (k, l) so that |σij| = |σkl| 6= 0. The examples of linear structures for the covariance

matrix are the uniform structure, the compound symmetry structure, the banded structure, the Toeplitz structure, etc.

The estimation procedure that we propose will rely on the decomposition of the residual space into m + 1 subspaces, see Theorem 2.3, and on the study of residuals obtained from projecting observations onto those subspaces. The paper [11] was the first to propose a residual based procedure to obtain explicit estimators for an arbitrary linear structured covariance matrix in the classical growth curve model as an alternative to iterative methods. The idea was later on applied to the sum of two profiles model in [12] and our aim here is to generalize results in [12] to the extended GMANOVA model with an arbitrary number of profiles.

2

Main idea and space decomposition

In this section we give some important results from which the main idea of our discussion is derived. To start with, we recall the theorem stated and proved in [3] about the estimation of the mean structure in the extended GMANOVA model.

Theorem 2.1 Let bBi’s be the maximum likelihood estimators of Bi’s in the model as in

Definition 1.1. Then Pr m X i=r AiBbiCi= m X i=r (I − Ti)XC′i(CiC′i)−Ci, where Pr = Tr−1Tr−2× · · · × T0, T0 = I, r = 1, 2, . . . , m + 1, Ti = I − PiAi(A′iP′iS−1i PiAi)−A′iP′iS−1i , i= 1, 2, . . . , m, Si = i X j=1 Kj, i= 1, 2, . . . , m, Kj = PjX(C′j−1(Cj−1C′j−1)−Cj−1− C′j(CjC′j)−Cj)X′P′j, C0 = I.

A useful result is the corollary of this theorem when r = 1, which gives the estimated mean structure, i.e., \ E[X] =Xm i=1AiBbiCi = m X i=1 (I − Ti)XC′i(CiC′i)−Ci. (1)

Replacing Ti in (1) by its expression given in Theorem 2.1 we get

\ E[X] = m X i=1 PiAi(A′iP′iS−1i PiAi)−A′iP′iS−1i XC′i(CiC′i)−Ci,

(6)

or equivalently \ E[X] = m X i=1 PPiAi,SiXPC′i. (2)

Noticing that the matrix PPiAi,Si and PC′

i are projector matrices, we see that estimators

of the mean structure is based on a projection of the observations on the space generated by the design matrices. Naturally, the estimators of the variance parameters are based on a projection of the observations on the residual space, that is the orthogonal complement to the design space.

If Σ would have been known, we would have from least squares theory the best linear estimator (BLUE) given by

^ E[X] = m X i=1 PPe iAi,ΣXPC′i, (3)

where Si in Pi is replaced with Σ to get ePi. Thus, we see that in the projections, if Σ is

unknown, the parameter has been replaced with Si’s, which according to their expressions are

not maximum likelihood estimators. However, Si’s define consistent estimators of Σ in the

sense that n−1Si → Σ in probability.

Applying the vec-operator on both sides of (3) we get vec(^E[X]) = m X i=1 (PC′ i⊗ PPeiAi,Σ)vecX.

The next theorem is essential for the development of the sequel of this paper. Theorem 2.2 Let P =Pmi=1PC′

i⊗ PPeiAi,Σ and Vi = CΣ( ePiAi), i = 1, 2, . . . , m. Then,

(i) The subspaces Vi’s are mutually orthogonal and

V1⊕ V2⊕ · · · ⊕ Vi= CΣ(A1 : A2 : · · · : Ai), i = 1, 2, . . . , m;

(ii) The matrix P is a projection matrix;

(iii) C(P ) =Pmi=1C(C′i) ⊗ Vi;

Proof. (i) From the definition of ePi’s and using Theorem 1.2.16. and Proposition 1.2.2. in

[4], we have the following

V1 = CΣ( eP1A1) = CΣ(A1),

V2 = CΣ( eP2A2)

= CΣ((I − PA1)A2)

(7)

V3 = CΣ( eP3A3)

= CΣ((I − PPe

2A2,Σ) eP2A3)

= CΣ( eP2(A2 : A3)) ∩ CΣ( eP2A2)⊥,

= CΣ((I − PA1)(A2 : A3)) ∩ CΣ((I − PA1)A2)⊥,

= (CΣ(A1)⊥∩ CΣ(A1 : A2 : A3)) ∩ (CΣ(A1 : A2) ∩ CΣ(A1)⊥)⊥, = (CΣ(A1)⊥∩ CΣ(A1 : A2 : A3)) ∩ (CΣ(A1 : A2)⊥+ CΣ(A1)), = CΣ(A1 : A2 : A3) ∩ (CΣ(A1)⊥∩ CΣ(A1 : A2)⊥+ CΣ(A1)⊥∩ CΣ(A1)), = CΣ(A1 : A2 : A3) ∩ CΣ(A1 : A2)⊥, and in general Vi = CΣ(A1: A2: · · · : Ai) ∩ CΣ(A1 : A2 : · · · : Ai−1)⊥, i= 1, 2, . . . , m, A0= 0.

This shows that the subspaces Vi’s are mutually orthogonal. Now we prove the second

asser-tion. Clearly

V1⊕ V2⊕ · · · ⊕ Vi ⊆ CΣ(A1 : A2 : · · · : Ai).

Since CΣ(A1 : A2 : · · · : Ai−1) ⊂ CΣ(A1: A2: · · · : Ai),

dim(Vi) = r(A1 : · · · : Ai) − r(A1 : · · · : Ai−1),

so that

i

X

j=1

dim(Vj) = r(A1) + r(A1 : A2) − r(A1) + · · ·

+ r(A1 : · · · : Ai) − r(A1: · · · : Ai−1)

= r(A1 : · · · : Ai).

This proves that the dimension of V1⊕ V2⊕ · · · ⊕ Vi equals the dimension of CΣ(A1 : A2 : · · · :

Ai) and hence the result follows.

(ii) We need to show that P is an idempotent matrix. Let Hi = PPeiAi and Gi =

PC′ i⊗ Hi. Then, P = Pm i=1Gi and P P = m X i=1 G2i +X i6=j GiGj.

On one hand, Gi is obviously an idempotent matrix as PC′

i and Hi are. On the other hand,

from calculations in (i), Hi may be rewritten as

Hi = PAi,Σ− PAi −1,Σ,

where Ai= (A1 : · · · : Ai) and thus, m X i=1 Hi = PAm,Σ and m X i=1 r(Hi) = r(PAm,Σ) = r(Am).

(8)

So, from Lemma 3 in [13] it follows that HiHj = 0 for i 6= j, which in turn implies that

GiGj = 0 for i 6= j since GiGj = PC′

iPC′j⊗ HiHj. Hence P

2 = P and P is a projector.

(iii) With notations and calculations introduced in the proof of (ii) it is clear that C(P ) = Pm

i=1C(Gi). But,

C(Gi) = C(PC′

i ⊗ Hi) = C(PC′i) ⊗ C(Hi) = C(PC′i) ⊗ Vi,

which completes the proof of Theorem 2.2.

We refer to the space C(P ) as the mean space and it is used to estimate the mean parameters whereas C(P )⊥ is referred to as the residual space and it is used to create residuals.

When Σ is not known it should be estimated. The general idea is to use the variation in the residuals. For our purposes we decompose the residual space into m + 1 orthogonal subspaces and Theorem 2.3 shows how such a decomposition is made.

Theorem 2.3 Let C(P ) and Vi be given as in Theorem 2.2. Then

C(P )⊥= I 1⊞ I2⊞ · · · ⊞ Im+1, where Ir = Wm−r+2⊗ (⊕r−1i=1Vi)⊥, r= 1, 2, . . . , m + 1, Wr = C(C′m−r+1) ∩ C(C′m−r+2)⊥, r= 1, . . . , m + 1, in which by convenience (⊕0 i=1Vi)⊥= ∅⊥= V0 =Rp, C0= I and Cm+1 = 0.

Proof. On one hand, the conditions C(C′i) ⊆ C(C′i−1), i = 2, 3, . . . , m, imply that C(C′1) can be decomposed as a sum of orthogonal subspaces as follows:

C(C′1) = [C(C′1) ∩ C(C′2)⊥] ⊕ [C(C′2) ∩ C(C′3)⊥] ⊕ · · ·

⊕[C(C′

m−1) ∩ C(C′m)⊥] ⊕ C(C′m).

On the other hand, by Theorem 2.2, the subspaces Vi, i = 1, 2, . . . , m; are orthogonal. Hence,

the result follows by letting

W1 = C(C′m), W2 = C(C′m−1) ∩ C(C′m)⊥, . . . ,

Wm = C(C′1) ∩ C(C′2)⊥,Wm+1 = C(C′1)⊥,

Vm+1 = (⊕mi=1Vi)⊥,V0= ⊕mi=1Vi⊕ Vm+1,

which completes the proof.

The residuals obtained by projecting data to these subspaces are Hr = (I − r−1 X i=1 PPe iAi,Σ)X(PC′r−1− PC′r), (4)

(9)

To get more insight on what is going on, we are going to illustrate the space decomposition for m = 3. In this case the BLUE of the mean is

^ E[X] = fM1+ fM2+ fM3, where f M1 = PA1XPC′ 1, f M2 = PT1A2XPC′ 2,T1= I − PA1,Σ= T1, f M3 = PT2A3XPC′ 3,T2= I − PA1,Σ− PT1A2,Σ = T2T1.

From here we see that the estimated mean is obtained by projecting observations on some subspaces. The matrices PA1, PT1A2 and PT2A3 are projectors onto the subspaces

V1 = CΣ(A1), V2 = CΣ(A1 : A2) ∩ CΣ(A1)⊥ and V3 = CΣ(A1 : A2 : A3) ∩ CΣ(A1 : A2)⊥

respectively. Figure 1 shows the whole space decomposed into mean and residual subspaces.

W1 W2 W3 W4 V1 V2 V3 V4 f M3 f M2 f M1 H1 H2 H3 H4

Figure 1: Decomposition of the whole space according to the within and between individuals design matrices illustrating the mean and residual spaces: V1 = CΣ(A1), V2 = CΣ(A1 : A2) ∩

CΣ(A1)⊥, V3 = CΣ(A1 : A2 : A3) ∩ CΣ(A1 : A2)⊥, V4 = CΣ(A1 : A2 : A3)⊥, W1 = C(C′3),

W2= C(C′2) ∩ C(C′3)⊥, W3= C(C1′) ∩ C(C′2)⊥, W4 = C(C′1)⊥.

In practice Σ is not known and should be estimated. A natural way to get an estimator of Σ is to use the sum of squared estimated residuals. If Σ is not structured we estimate the residuals in (4) with Rr = (I − r−1 X i=1 PPiAi,Si)X(PC′ r−1− PC′r), r = 1, 2, 3, . . . , m + 1,

(10)

where Pi and Si are given as in Theorem 2.1. Thus a natural estimator of Σ is obtained from

the sum of squared residuals, i.e.,

n bΣ= R1R′1+ R2R′2+ · · · + Rm+1R′m+1,

which is the maximum likelihood estimator.

3

Estimators when the covariance matrix is linearly structured

In this section we consider the extended growth curve model as in Definition 1.1, but with a lin-early structured covariance matrix Σ. This Σ will be denoted Σ(s)so that E ∼ Np,n(0, Σ(s), In).

The estimation procedure that we propose will rely on the decomposition of the spaces done in Section 2. We will sequentially estimate the inner product in the spaces Vi (i = 1, 2, . . . , m),

the residuals in (4), and finally estimate the covariance matrix using all estimated residuals. To start with, it is natural to use Q1 = H1H′1, H1 = X(I − PC′

1), to estimate the

inner product in the space V1. We apply the general least squares approach and

mini-mize trn(Q1− (n − r1)Σ(s))( )′

o

with respect to Σ(s), where the notation (Y )( )′ stands for

(Y )(Y )′. This is done in Lemma 3.1.

By vecΣ(K) we mean the patterned vectorization of the linearly structured matrix Σ(s), that is the columnwise vectorization of Σ(s) where all 0’s and repeated elements (by modulus) have been disregarded. Then there exists a transformation matrix T such that

vecΣ(K) = T vecΣ(s) or vecΣ(s)= T+vecΣ(K), (5) where T+denotes the Moore-Penrose generalized inverse of T . Furthermore, from [4], we have

dΣ(s) dΣ(K) = (T

+). (6)

Lemma 3.1 Let Q1 = H1H′1 = X(I − PC′ 1)X

, bΥ

1 = (n − r1)I. Then, the minimizer of

f1(Σ(s)) = tr n (Q1− (n − r1)Σ(s))( )′ o is given by vec bΣ(s)1 = T+(T+)′Υb′1Υb1T+ − (T+)′Υb′1vecQ1. Proof. We may write

trnQb1− (n − r1)Σ(s)  ()′o = vecQb1− (n − r1)Σ(s) ′ () = vec bQ1− bΥ1vecΣ(s) ′ (), and thus f1(Σ(s)) =  vec bQ1− bΥ1vecΣ(s) ′ (). (7)

(11)

Differentiating the expression in the right hand side of (7) with respect to vecΣ(K) and equalizing to 0, we get −2 dΣ (s) dΣ(K)Υb ′ 1(vec bQ1− bΥ1vecΣ(s)) = 0. (8)

From (5), (6) and (8) we obtain the linear equation (T+)′Υb′1Υb1T+vecΣ(K) = (T+)′Υb

1vec bQ1,

which is consistent and a general solution is given by vecΣ(K) =(T+)′Υb′1Υb1T+

−

(T+)′Υb′1vec bQ1+ ((T+)′Υb′1Υb1T+)oz,

where z is an arbitrary vector. Hence, using (5) we obtain a unique minimizer of (7) given by vec bΣ(s)1 = T+(T+)′Υb′1Υb1T+

−

(T+)′Υb′1vec bQ1, (9) and the proof is complete.

Lemma 3.1 gives the first estimator of Σ(s). Assuming that bΣ(s)1 is positive definite (which always holds for large n), we can use bΣ(s)1 to define the inner product in the space V1, and

there-fore we consider Cb

Σ(s)1

(A1) instead of CΣ(s)(A1). At the same time an estimator of M1, and also

that of H2 are found by projecting observations on C(C′1) ⊗ V1 and C(C′1) ∩ C(C′2)⊥

 ⊗ V⊥ 1 respectively, i.e., c M1 = PA 1, bΣ(s)1 XPC′ 1, (10) c H2 = (I − PA 1, bΣ (s) 1 )X(PC′ 1 − PC′2).

A second estimator of Σ(s) is obtained using the sum of Q1 and cH2cH ′ 2. Notice that c H2cH ′ 2 = (I − PA1, bΣ(s)1 )X(PC′1 − PC′2)X ′(I − P A1, bΣ(s)1 ) ′. Put bT1 = I − PA 1, bΣ (s) 1 and W1 = X(PC′ 1 − PC′2)X ′. Then cH 2Hc ′ 2 = bT1W1Tb ′ 1 and Q1 is

independent of W1. Therefore it is natural to condition cH2Hc ′

2 with respect to Q1 and

c H2Hc ′ 2|Q1∼ Wp  b T1Σ(s)Tb ′ 1, r1− r2  , where Wp(·, ·) stands for the Wishart matrix.

The following lemma gives a second estimator of Σ(s)where again the general least squares approach is employed.

(12)

Lemma 3.2 Let bQ2 = Q1+ cH2Hc ′

2, bΥ2= (n − r1)I + (r1− r2) bT1⊗ bT1. Then, the minimizer

of f2(Σ(s)) = tr n ( bQ2− [(n − r1)Σ(s)+ (r1− r2) bT1Σ(s)Tb ′ 1)( )′ o is given by vec bΣ(s)2 = T+(T+)′Υb′2Υb2T+ − (T+)′Υb′2vec bQ2. Proof. Similar as in Lemma 3.1.

Now assume that bΣ(s)2 is positive definite and use it to define the inner product in V2,

i.e., consider Cb

Σ(s)2

( bT1A2) instead of CΣ(s)(T1A2). This gives us an estimator of M2, and

also that of H3 by projecting observations on C(C′2) ⊗ CΣb(s) 2 ( bT1A2) and (C(C′2) ∩ C(C′3)⊥) ⊗  Cb Σ(s)1 (A1) + Cb Σ(s)2 ( bT1A2) ⊥ respectively, i.e., c M2 = PTb 1A2, bΣ (s) 2 XPC′ 2, c H3 = Tb2X(PC′ 2 − PC′3), b T2 = I − PA 1, bΣ(s)1 − PTb 1A2, bΣ(s)2 .

To derive a third estimator of Σ(s), the idea is to use the sum bQ3 = bQ2 + cH3Hc ′ 3 in a

similar way as in Lemma 3.1 or Lemma 3.2. We continue the same process until all residuals are estimated and then use the sum of squared estimated residuals to estimate the covariance matrix that can be understood as a dispersion matrix. After the (r − 1)th stage of the process

we have already r − 1 estimates of Σ(s). In Lemma 3.3, we show how to obtain the rthestimate of Σ(s). Before we state it, we notice that after the (r − 1)th stage we have also the following quantities: Wi = X(PC′ i− PC′i+1)X ′ ∼ W p(Σ(s), ri− ri+1), (11) i= 0, 1, 2, . . . , m, b Pj = Tbj−1Tbj−2× · · · × bT0, bT0= I, j = 1, 2, . . . , r − 1, (12) b Ti = I − PPb iAi, bΣ (s) i , i= 1, 2, . . . , r − 1, (13) b Ti = I − i X j=1 Pb PjAj, bΣ(s) j , i= 1, 2, . . . , r − 1, (14) c HiHc ′ i = Tbi−1Wi−1Tb ′ i−1, bT0 = I, i = 1, 2, . . . , r, (15) b Qi = i X j=1 c HjcH′j = i X j=1 b Tj−1Wj−1Tb ′ j−1, (16) i= 1, 2, . . . , r, b Υi = i X j=1 (rj−1− rj) bTj−1⊗ bTj−1, i= 1, 2, . . . , r. (17)

(13)

Lemma 3.3 Let bTi, bQr and bΥr be defined as in (14), (16) and (17), respectively. Then, the minimizers of fr(Σ(s)) = tr ( b Qr− r X i=1 (ri−1− ri) bTi−1Σ(s)Tb ′ i−1 ! ()′ ) , r = 1, 2, . . . , m + 1, are given by vec bΣ(s)r = T+(T+)′Υb′rΥbrT+ − (T+)′Υb′rvec bQr. Proof. Similar as in Lemma 3.1.

Lemma 3.3 gives the rth estimate of Σ(s) and the (m + 1)th estimate of Σ(s) can be

understood as a dispersion matrix as it uses all information that we have in residuals. Now the results at our disposal permit us to propose estimators of parameters in the extended growth curve model with a linearly structured covariance matrix.

Theorem 3.4 Let the extended growth curve model be given by (1.1). Then (i) A consistent estimator of the structured covariance matrix Σ(s) is given by

vec bΣ(s)m+1= T+  (T+)′Υb′ m+1Υbm+1T+ − (T+)′Υb′ m+1vec bQm+1. (18)

(ii) An unbiased estimator of the mean is given by

\ E[X] = m X i=1 (I − bTi)XC′i(CiC′i)−Ci. (19)

where bTi, bQr and bΥr be defined as in (13), (16) and (17), respectively.

Proof. (i) The consistency of the estimator in (18) is established through the following implications (the notation ‘−→’ means convergence in probability)p

1 n− r1 Q1 −→ Σp (s) ⇓ b Σ(s)1 −→ Σp (s) ⇓ b Σ(s)2 −→ Σp (s) ⇓ .. . ⇓ b Σ(s)m+1 −→ Σp (s).

(14)

The above implications can be easily checked. The first line follows from the well known fact that the (corrected) sample covariance matrix is a consistent estimator of the true covariance matrix. The rest is established using Cramer-Slutsky’s theorem [14]. (ii) Next we establish the unbiasedness of the estimator given by (19). From Theorem 2.2. (i) and using uniqueness property of projectors it is possible to rewrite the estimated mean as

\ E[X] = m X i=1 P Ai, bΣ(s)i X(PC′ i− PC′i+1).

Using the linearity of the expectation operator and independence of P

Ai, bΣ(s)i and X(PC′ i− PC′ i+1), we have E[\E[X]] = m X i=1 EhP Ai, bΣ(s)i i EhX(PC′ i − PC′i+1) i = EhP Am, bΣ(s)m i EhXPC′ m i +E  P Am−1, bΣ(s)m−1  EhX(PC′ m−1− PC′m) i +E  P Am−2, bΣ(s)m−2  EhX(PC′ m−2− PC′m−1) i + · · · +EhP A1, bΣ(s)1 i E h X(PC′ 1 − PC′2) i . Since E[X] =Pmi=1AiBiCi, using the facts that C(C′i) ⊆ C(C′i−1) and

P An, bΣ(s)n n X i=1 AiBiCi ! = n X i=1 AiBiCi, we get E[\E[X]] = m X i=1 AiBiCi ! PC′ m+ m−1X i=1 AiBiCi ! (PC′ m−1− PC′m) + · · · + 2 X i=1 AiBiCi ! (PC′ 2 − PC′3) + A1B1C1(PC′1 − PC′2) = AmBmCm+ m−1X i=1 AiBiCi ! PC′ m+ Am−1Bm−1Cm−1 + m−2X i=1 AiBiCi ! PC′ m−1 − m−1X i=1 AiBiCi ! PC′ m+ · · · +A2B2C2+ A1B1C1PC′ 2 − 2 X i=1 AiBiCi ! PC′ 3 +A1B1C1− A1B1C1PC′ 2.

(15)

4

Simulation study

The aim of this section is to investigate the performance of the proposed algorithm for small sample sizes. In fact, for some structures of the covariance matrix, the proposed technique may produce non positive definite estimates for small sample sizes. In our simulations we will also study the same problem with respect to the number of profiles m.

4.1 Procedures and methods

We consider the following three scenarios:

Scenario S1: The covariance matrix has the uniform structure: Σ(s)= 3[(1 − 0.7)I4+ 0.7J4],

where I4 and J4 stand for the 4 × 4 identity matrix and the 4 × 4 matrix of ones respectively.

Scenario S2: The covariance matrix has the Toeplitz structure with different variances: Σ(s)=     4 1.2 0.8 0.9 1.2 5 1.2 0.8 0.8 1.2 6 1.2 0.9 0.8 1.2 7     . Scenario S3: The covariance matrix has the banded structure:

Σ(s)=     8 2 0 0 2 10 4 0 0 4 12 6 0 0 6 14     .

In all scenarios the number of simulation runs was set to 5000 for each sample size consid-ered (n = 12, 24, 36, 60, 72, 240, 360), and the percentage of non positive definite estimates of the covariance matrix was calculated. Data is generated from X ∼ Np,n(Pmi=1AiBiCi, Σ, I), m =

1, 2, 3, with p = 4. The within design matrices are A′1=  1 1 1 1 1 2 3 4  , A′2 = 12 22 32 42 , A′3= 13 23 33 43 .

The between design matrices are chosen so that we have the same number of individuals in each group (two group for m = 1, 2 and three groups for m = 3). For example, for m = 3, we have C1 =  1′ n/3⊗   1 0 0   : 1′ n/3⊗   0 1 0   : 1′ n/3⊗   0 0 1     , C2 =  1′n/3⊗  0 0  : 1′n/3⊗  1 0  : 1′n/3⊗  0 1  , C3 =  0′n/3 : 0′n/3 : 1′n/3,

(16)

Table 1: Percentage of non positive definite estimates of different linearly structured covariance matrices according to the number of profiles (m) and the sample size (n).

Number of profiles Sample size Linear structure

m n S1 S2 S3 1 12 0 5.48 7.82 24 0 0.04 0.64 36 0 0 0.26 60 0 0 0 72 0 0 0 240 0 0 0 360 0 0 0 2 12 0 12.56 15.12 24 0 0.38 1.54 36 0 0.02 0.24 60 0 0 0.02 72 0 0 0 240 0 0 0 360 0 0 0 3 12 0 86.98 63.84 24 0 39.64 14.84 36 0 12.98 4.08 60 0 1.60 0.22 72 0 0.50 0.10 240 0 0 0 360 0 0 0

where the notation 1k (respectively 0k) stands for a column vector of 1’s (respectively of 0’s)

of size k. The parameter matrices are B1 =  1 1 1 1 2 3  , B2 = 2 3 , B3= 4. 4.2 Results

Table 1 contains percentages of non positive definite estimates of the covariance matrix for each scenario considered. One can see that for the uniform structure our techniques perform well as we have 100% of positive definite estimates in all cases. Also we have good performance for the other two cases where we have small percentages of non definite estimates for n < 36 whether we have one profile or two profiles. However, for three profiles case we have high percentage of non positive definite estimates for n < 36. Hence, concerning the number of profiles m, we note that as m increases the percentages of non positive definite estimates also increase except for the case of uniform structure.

(17)

References

[1] R. F. Pothoff and S. N. Roy, “A generalized multivariate analysis of variance model useful especially for growth curve problems,” Biometrika, vol. 51, pp. 313–326, 1964.

[2] A. P. Verbyla and W. N. Venables, “An extension of the growth curve model,” Biometrika, vol. 75, no. 1, pp. 129–138, 1988.

[3] D. von Rosen, “Maximum likelihood estimators in multivariate linear normal models,” Journal of multivariate analysis, vol. 31, no. 2, pp. 187–200, 1989.

[4] T. Kollo and D. von Rosen, Advanced Multivariate Statistics with Matrices. Dordrecht, The Netherlands: Springer, 2005.

[5] K. Filipiak and D. von Rosen, “On MLEs in an extended multivariate linear growth curve model,” Metrika, vol. 75, no. 8, pp. 1069–1092, 2012.

[6] J. Hu, “Properties of the explicit estimators in the extended growth curve model,” Statis-tics, vol. 44, no. 5, pp. 477–492, 2010.

[7] J. Hu, G. Yan, and J. You, “Estimation for an additive growth curve model with orthog-onal design matrices,” Bernoulli, vol. 17, no. 4, pp. 1400–1419, 2011.

[8] J. C.-S. Lee and S. Geisser, “Applications of growth curve prediction,” Sankhy¯a: The Indian Journal of Statistics, Series A, vol. 37, no. 2, pp. 239–256, 1975.

[9] J. C. Lee, “Prediction and estimation of growth curves with special covariance structures,” Journal of the American Statistical Association, vol. 83, no. 402, pp. 432–440, 1988.

[10] C. G. Khatri, “Testing some covariance structures under a growth curve model,” Journal of Multivariate Analysis, vol. 3, no. 1, pp. 102–116, 1973.

[11] M. Ohlson and D. von Rosen, “Explicit estimators of parameters in the growth curve model with linearly structured covariance matrices,” Journal of Multivariate Analysis, vol. 101, no. 5, pp. 1284–1295, 2010.

[12] J. Nzabanita, M. Singull, and D. von Rosen, “Estimation of parameters in the extended growth curve model with a linearly structured covariance matrix,” Acta et Commenta-tiones Universitatis Tartuensis de Mathematica, vol. 16, no. 1, pp. 13–32, 2012.

[13] C. G. Khatri, “Some results for the singular normal multivariate regression models,” Sankhy¯a: The Indian Journal of Statistics, Series A, pp. 267–280, 1968.

[14] H. Cram´er, Mathematical Methods in Statistics. Princeton: Princeton University Press, 1946.

References

Related documents

The Standard 2 is about specifying the expected learning outcomes, in terms of knowledge and skills, of an education to prepare the students for the intended professional role,

Flera familjehemsföräldrar beskriver sin relation till fosterbarn som att de tycker om fosterbarnet som det vore sitt eget men att det saknas något för att det exakt

• Byta till mindre kärl för tyngre avfallsfraktioner, t ex matavfall (se tipsblad). Är märkningen av kärlens olika fraktioner

When we looked at relative gene expression value of cyp1a from PCB, PFOS, PCB with PFOS, PCB with PFHxA, except PFHxA alone, we could see differences in average from 13 times up to

• Byta till mindre kärl för tyngre avfallsfraktioner, t ex matavfall (se tipsblad). Är märkningen av kärlens olika fraktioner

Vår studie uppmärksammar hur elever i läs- och skrivsvårigheter och dyslexi upplever motivation som en del i det egna lärandet och ambitionen är att kunskapen ska leda till

Specific objectives are (i) to derive explicit estimators of parameters in the extended growth curve model when the covariance matrix is linearly structured, (ii) to propose