769
Views
1
CrossRef citations to date
0
Altmetric
Article

Properties of BLUEs in full versus small linear models

ORCID Icon, ORCID Icon & ORCID Icon
Pages 7684-7698 | Received 12 Aug 2021, Accepted 09 Mar 2022, Published online: 29 Mar 2022

Abstract

In this article we consider the partitioned linear model M12={y, X1β1+X2β2, V} and the corresponding small model M1={y, X1β1, V}. We focus on comparing the best linear unbiased estimators, BLUEs, of X1β1 under M12 and M1. In other words, we are interested in the effect of adding regressors on the BLUEs. Particular attention is paid on the consistency of the model, that is, whether the realized value of the response vector y belongs to the column space of (X1:V) or (X1:X2:V).

MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

In this article we consider the partitioned linear model y=X1β1+X2β2+ε and so-called small model (submodel) y=X1β1+ε, or shortly M12={y, Xβ, V}={y, X1β1+X2β2, V},M1={y, X1β1, V}. Here y is an n-dimensional observable response variable, and ε is an unobservable random error with a known covariance matrix cov(ε)=V=cov(y) and expectation E(ε)=0. The matrix X is a known n × p matrix, that is, XRn×p, partitioned columnwise as X=(X1:X2), XiRn×pi,i=1,2. Vector β=(β1,β2)Rp is a vector of fixed (but unknown) parameters; symbol stands for the transpose.

As for notation, r(A), A, A+, C(A), N(A), and C(A), denote, respectively, the rank, a generalized inverse, the (unique) Moore–Penrose inverse, the column space, the null space, and the orthogonal complement of the column space of the matrix A. By A we denote any matrix satisfying C(A)=C(A). Furthermore, we will write PA=PC(A)=AA+=A(AA)A to denote the orthogonal projector onto C(A). The orthogonal projector onto C(A) is denoted as QA=IaPA, where Ia is the a × a identity matrix and a is the number of rows of A. We write shortly M=InPX,Mi=InPXi,i=1, 2. One obvious choice for X is M.

When using generalized inverses it is essential to know whether the expressions are independent of the choice of the generalized inverses involved. The following lemma gives an important invariance condition; cf. Rao and Mitra (Citation1971, Lemma 2.2.4)

Lemma 1.1.

For nonnull matrices A and C the following holds: ABC=AB+Cfor all BC(C)C(B)  &  C(A)C(B).

For a given linear model M={y,Xβ,V}, let the set W(M) of nonnegative definite matrices be defined as (1.1) W(M)={WRn×n:W=V+XUUX,C(W)=C(X:V)}.(1.1) In Equation(1.1), U can be any matrix comprising p rows as long as C(W)=C(X:V) is satisfied. Lemma 1.2 collects together some important properties of the class W(M); see, for example, Puntanen, Styan, and Isotalo (Citation2011, Prop. 12.1 and 15.2).

Lemma 1.2.

Consider the model M={y,Xβ,V} and let W=V+XUUXW(M). Then(1.2) G12=X(XWX)XW+=PWVM(MVM)MPW=PWVM(MVM)+=PWVM(MVM)+M.(1.2) Moreover, the following statements are equivalent:

  1. C(X:V)=C(W),

  2. C(X)C(W),

  3. XWX is invariant for any choice of W,

  4. C(XWX)=C(X) for any choice of W,

  5. X(XWX)XWX=X for any choices of W and (XWX).

It is noteworthy that the matrix G12 in Equation(1.2) is invariant for the choice of the generalized inverses denoted as “”, and it is independent of any choice of WW(M). Notice also that the invariance properties in (d) and (e) in Lemma 1.2 are valid for all choices of WW(M). It is clear that VW(M) if and only if C(X)C(V).

In Lemma 1.2, the matrix W is nonnegative definite, denoted as WL0. A corresponding version of Lemma 1.2 can be presented for W=V+XTX which may not be symmetric but satisfies C(X:V)=C(W).

Corresponding to Equation(1.1), we will say that WiW(Mi) if there exist Ui such that (1.3) Wi=V+XiUiUiXi,C(Wi)=C(Xi:V),i=1,2.(1.3) For the partitioned linear model M12 we will say that WW(M12) if W=V+X1U1U1X1+X2U2U2X2, where U1 and U2 are defined as in Equation(1.3). For our considerations the actual choice of U1 and U2 does not matter as long as they satisfy Equation(1.3).

By the consistency of the model M it is meant that y lies in C(X:V) with probability 1. Hence we assume that under the consistent model M the observed numerical value of y satisfies yC(X:V)=C(X:VM)=C(X)C(VM)=C(X)C(MV), where “⊕” refers to the direct sum and “” refers to the direct sum of orthogonal subspaces. For the equality C(X:V)=C(X:VM), see Rao (Citation1974, Lemma 2.1).

For parts (a) and (b) of Lemma 1.3, see, for example, Puntanen, Styan, and Isotalo (Citation2011, Th. 8). and for part (c), see the rank rule of the matrix product of Marsaglia and Styan (Citation1974, Cor. 6.2). Claim (d) is straightforward to confirm.

Lemma 1.3.

Consider X=(X1:X2) and let M2=InPX2. Then

  1. C(X1:X2)=C(X1:M1X2),

  2. M=InP(X1:X2)=In(PX2+PM2X1)=M2QM2X1=QM2X1M2,

  3. r(M2X1)=r(X1)dimC(X1)C(X2),

  4. C(X2)C(X1:V)C(M1X2)C(M1V).

For Lemma 1.4, see, for example, Puntanen, Styan, and Isotalo (Citation2011, p. 152).

Lemma 1.4.

For conformable matrices A and B the following three statements are equivalent:(a)PAPBis an orth. projector,(b)PAPBL0,(c) C(B)C(A).If any of the above conditions holds thenPAPB=PC(A)C(B)=P(IPB)A.

Let A and B be arbitrary m × n matrices. Then, in the consistent linear model M, the estimators Ay and By are said to be equal (with probability 1) if (1.4) Ay=Byfor all yC(X:V)=C(X:VM)=C(W),(1.4) where WW(M). Thus, if A and B satisfy Equation(1.4), then AB=CQW for some matrix C. It is crucial to notice that in Equation(1.4) we are dealing with the “statistical” equality of the estimators Ay and By. In Equation(1.4) y refers to a vector in Rn. Thus we do not make any notational difference between a random vector and its observed value.

According to the well-known fundamental BLUE-equation, see Lemma 2.1 in Section 2, Ay is the BLUE of Xβ if and only if A(X:VM)=(X:0). Obviously (A+NQW)y is another representation of BLUE for any n × n matrix N. However, the equality Ay=(A+NQW)yfor all yC(W) holds when the model is consistent in the sense that yC(W). The properties of the BLUE deserve particular attention when C(X:V)=Rn does not hold: then there is an infinite number of multipliers B such that By is BLUE but for all such multipliers the vector By itself is unique once the response y has been observed. The case of two linear models, Bi={y,Xβ,Vi},i=1,2, is extensively studied by Mitra and Moore (Citation1973). They ask, for example, when is a specific linear representation of the BLUE of μ=Xβ under B1 also a BLUE under B2, and when is the BLUE of μ=Xβ under B1 irrespective of the linear representation used in its expression, also a BLUE under B2.

The purpose of this paper is to consider the models M1 and M12 in the spirit of Mitra and Moore (Citation1973). We pick up particular fixed representations for the BLUE s of μ1=X1β1 under these two models, say G1y and G1#y, and study the conditions under which they are equal for all values of yC(X1:X2:V) or yC(X1:V), that is, (1.5) G1W1=G1#W1,orG1W=G1#W.(1.5) Moreover, we review the conditions under which Equation(1.5) holds for all representations of the BLUE s, not only for fixed G1 and G1#. Some related considerations were made by Haslett, Markiewicz, and Puntanen (Citation2020) when these models are supplemented with the new unobservable random vector y*, coming from y*=Kβ1+ε*, where the covariance matrix of y* is known as well as the cross-covariance matrix between y* and y.

The well-known (or pretty well-known) results are given as Lemmas, while the new (or at least not so well-known) results are represented as Propositions.

2. The fundamental BLUE equations

A linear statistic By is said to be linear unbiased estimator, LUE, for the parametric function Kβ in M12 if its expectation is equal to Kβ, which happens if and only if K=XB; in this case Kβ is said to be estimable. The LUE By is the best linear unbiased estimator, BLUE, of estimable Kβ if By has the smallest covariance matrix in the Löwner sense among all LUEs of Kβ: cov(By)Lcov(B#y)for all B#:B#X=K. It is well known that μ1=X1β1 is estimable under M12 if and only if C(X1)C(X2)={0},i.e.,r(M2X1)=r(X1). For Lemma 2.1, characterizing the BLUE, see, for example, Rao (Citation1973, p. 282).

Lemma 2.1.

Consider the model M12 where η=Kβ is estimable. Then

  1. Ay=BLUE(Xβ)A(X:VM)=(X:0), that is, A{Pμ|M12},

  2. By=BLUE(Kβ)B(X:VM)=(K:0),  that is, B{Pη|M12}.

    In particular, if μ1=X1β1 is estimable,

  3. Cy=BLUE(μ1)C(X1:X2:VM)=(X1:0:0), that is, C{Pμ1|M12}.

Of course, under the model M1 we have Dy=BLUE(μ1)D(X1:VM1)=(X1:0),i.e.,D{Pμ1|M1}. To indicate that A{Pμ|M12} we will also use notations Ay=μ˜(M12)=BLUE(Xβ|M12),Ay{BLUE(Xβ|M12)}. Using Lemma 1.2 we can obtain, for example, the following well-known solution to A in Lemma 2.1: X(XWX)XW{Pμ|M12}, where WW(M12) and we can freely choose the generalized inverses involved. Expression X(XWX)XW is not necessarily unique with respect to the choice of W but by Lemma 1.2, the matrix G12=X(XWX)XW+=PWVM(MVM)MPW is unique whatever choices of W and (XWX) we have and moreover, G12 does not depend on the choice of WW(M12). The general solution for A in Lemma 2.1, can be expressed, for example, as G0=G12+NQW, where NRn×n is free to vary, and QW=InPW. Thus the solution for A (as well as for B and C) in Lemma 2.1 is unique if and only if C(X:V)=Rn.

Consider then the estimation of μ1=X1β1 under M12 assuming that μ1 is estimable. Premultiplying the model M12 by M2 yields the reduced model M12·2={M2y,M2X1β1,M2VM2}. Now the well-known Frisch–Waugh–Lovell theorem, see, for example, Groß and Puntanen (Citation2000, Sec. 6), states that the BLUE s of μ1 under M12 and M12·2 coincide. To obtain an explicit expression for the BLUE of M2X1β1 under M12·2 we need a W-matrix in M12·2. Now any matrix of the form M2VM2+M2X1T1T1X1M2=M2(V+X1T1T1X1)M2 satisfying (2.1) C[M2(V:X1T1)]=C[M2(V:X1)]=C(M2W1),(2.1) is a W-matrix in M12·2. Choosing T1=U1 as in Equation(1.3) we have M2WM2=M2W1M2W(M12·2). Thus the BLUE of M2X1β1 under M12·2 can be expressed as BLUE(M2X1β1|M12·2)=M2X1(X1Ṁ2X1)X1Ṁ2y, where Ṁ2=M2(M2W1M2)M2.

We observe that Equation(2.1) holds for T1=0 if and only if C(M2X1)C(M2V), that is, see part (d) of Lemma 1.3, (2.2) C(X1)C(X2:V).(2.2) Our conclusion: If Equation(2.2) holds, then the BLUE of M2X1β1 under M12·2 can be expressed as (2.3) BLUE(M2X1β1|M12·2)=M2X1(X1Ṁ2VX1)X1Ṁ2Vy,(2.3) where Ṁ2V=M2(M2VM2)M2. Actually, it can be shown that Equation(2.2) is also a necessary condition for Equation(2.3). It is obvious that under the estimability of μ1 we have (2.4a) BLUE(μ1|M12·2)=BLUE(μ1|M12)=X1(X1Ṁ2X1)X1Ṁ2y,(2.4a) (2.4b) BLUE(μ2|M12·1)=BLUE(μ2|M12)=X2(X2Ṁ1X2)X2Ṁ1y,(2.4b) where Ṁi=Mi(MiWMi)Mi,i=1,2.

An alternative expression for the BLUE of μ1 can be obtained by premultiplying the fundamental BLUE-equation X(XWX)XW(X1:X2:VM)=(X1:X2:0) by M2, yielding (2.5) (M2X1:0)(XWX)XW(X1:X2:VM)=(M2X1:0:0).(2.5) Because r(M2X1)=r(X1), we can, by the rank cancelation rule of Marsaglia and Styan (Citation1974), cancel M2 in Equation(2.5) and thus an alternative expression for Equation(2.4a) is μ˜1(M12)=(X1:0)(XWX)XWy. Now we should pay attention to numerous generalized inverses appearing in the representations of the BLUEs. Namely, when the observable response y belongs to a “correct” subspace of Rn, then there is no problem with the generalized inverses. In the next section we will consider particular unique representations of the multipliers of y and study the equality of the relevant estimators taking the space where y belongs into account.

3. Some useful matrix results

Let us denote G1#=X1(X1Ṁ2X1)X1Ṁ2,D1#=(X1:0)(XWX)XW+,G2#=X2(X2Ṁ1X2)X2Ṁ1,D2#=(0:X2)(XWX)XW+, where Ṁ1 and Ṁ2 are now unique (once W is given) matrices defined as Ṁ1=M1(M1WM1)+M1=M1(M1W2M1)+M1,Ṁ2=M2(M2WM2)+M2=M2(M2W1M2)+M2. It is noteworthy that the following types of equalities hold: M1(M1WM1)+M1=M1(M1WM1)+=(M1WM1)+. Now under the estimability of μ1=X1β1 we have μ˜1(M12)=X1(X1Ṁ2X1)X1Ṁ2y=(X1:0)(XWX)XW+y,μ˜2(M12)=X2(X2Ṁ1X2)X2Ṁ1y=(0:X2)(XWX)XW+y, and μ˜(M12)=(G1#+G2#)y=(D1#+D2#)y for all yC(W). Because G1# and D1# belong to {Pμ1|M12}, they satisfy the equation (3.3) G1#W=D1#W.(3.3) Next we show that we also have (3.4) G1#QW=D1#QW.(3.4) We immediately observe that D1#QW=0 and what remains is to show that G1#QW=0. Now the equation G1#QW=X1(X1Ṁ2X1)X1Ṁ2QW=0 holds if and only if (3.5) X1Ṁ2QW=0,i.e.,C(Ṁ2X1)C(W).(3.5) Clearly Equation(3.5) holds because C(Ṁ2X1)C(Ṁ2)=C[(M2W1M2)+]=C(M2W1)C(W), where the last inclusion follows from C(X1:X2:V)=C[X2:M2(X1:V)]=C(X2:M2W1). Combining Equation(3.3) and Equation(3.4) gives the following result.

Proposition 3.1.

Assume that μ1 is estimable under M12. Then(3.6) G1#=X1(X1Ṁ2X1)X1Ṁ2=(X1:0)(XWX)XW+=D1#,(3.6) where Ṁ2=M2(M2W1M2)+M2. Moreover, the expressions in Equation(3.6) are invariant for any choices of generalized inverses (X1Ṁ2X1), W, and (XWX) as well as for the choice of WW(M12). Corresponding equality holds between G2# and D2#. Moreover,G12=X(XWX)XW+=G1#+G2#=D1#+D2#.

We will also need the following proposition.

Proposition 3.2.

DenoteG1#=X1(X1Ṁ2X1)X1Ṁ2,where Ṁ2=M2(M2W1M2)+M2. Then

  1. C(X1Ṁ2W)=C(X1Ṁ2X1)=C(X1M2),

  2. r(W1Ṁ2X1)=r(WṀ2X1)=r(X1Ṁ2X1)=r(M2X1),

  3. C(WG1#)=C[WṀ2X1(X1Ṁ2X1)X1]=C(WṀ2X1),

  4. C(G1#W)=C(X1M2).

    In particular, when μ1 is estimable under M12, we have

  5. C(X1Ṁ2W)=C(X1Ṁ2X1)=C(G1#W)=C(X1).

Proof.

Property (b) comes from the following: (3.7) r(M2X1)r(W1Ṁ2X1)=r[W1M2(M2W1M2)M2X1]r[M2W1M2(M2W1M2)M2X1]=r(M2X1).(3.7) The last equality in Equation(3.7) follows from the fact that C(M2X1)C(M2W1). The other statements can be confirmed in the corresponding way. □

Proposition 3.3 appears to be useful for our BLUE-considerations and it also provides some interesting linear algebraic matrix results. By A1/2 we refer to the nonnegative definite square root of a nonnegative definite matrix A and A+1/2=(A1/2)+ so that A1/2A+1/2=PA.

Proposition 3.3.

The following five statements hold:

  1. C(W+X)=C(WM:QW)=C(VM:QW),

  2. C(W1+X1)=C(W1M1:QW1)=C(VM1:QW1),

  3. PW1/2M2=PWPW+1/2X2,

  4. PWṀ2PW=W+W+X2(X2W+X2)X2W+,

  5. WṀ2X1=W1Ṁ2X1=[InX2(X2W+X2)X2W+]X1.

    The following three statements are equivalent:

  6. r(X2)=dimC(W1)C(X2)+dimC(W1)C(X2),

  7. r(W1)=r(W1M2)+r(W1X2),

  8. PW11/2M2=PW1PW1+1/2X2.

    If any of the conditions (f)–(h) holds, then

  9. PW1Ṁ2PW1=W1+W1+X2(X2W1+X2)X2W1+,

  10. WṀ2X1=W1Ṁ2X1=[PW1PW1X2(X2W1+X2)X2W1+]X1.

    If C(X2)C(W1), then

  11. WṀ2X1=W1Ṁ2X1=[InX2(X2W1+X2)X2W1+]X1.

Proof.

The first five statements (a)–(e) appear in Markiewicz and Puntanen (Citation2019, Sec. 4). The claim (h), that is, PW11/2M2=PW1PW1+1/2X2, holds if and only if, see Lemma 1.4, (3.8) C(W11/2M2)=C(W1+1/2X2:QW1)=C(W1+1/2X2)C(W1).(3.8) Now Equation(3.8) holds if and only if r(W11/2M2)=nr(W1+1/2X2:QW1), that is, r(W1)=r(W1M2)+r(W1X2), which further is equivalent to (f). Clearly (f) holds, for example, when C(X2)C(W1).

Assuming that (f) holds we can write (3.9) PW1Ṁ2PW1=PW1M2(M2W1M2)+M2PW1=W1+1/2PW11/2M2W1+1/2=W1+1/2(PW1PW1+1/2X2)W1+1/2=W1+W1+X2(X2W1+X2)X2W1+.(3.9) From Equation(3.9) it follows that W1Ṁ2X1=W1[W1+W1+X2(X2W1+X2)X2W1+]X1=[PW1PW1X2(X2W1+X2)X2W1+]X1, and hence, supposing that C(X2)C(W1), we obtain (k): W1Ṁ2X1=[InX2(X2W1+X2)X2W1+]X1. Thus the proof is completed. □

4. Difference of the BLUEs under the full and small model

Next we introduce a particular expression for the difference (G1G1#)y which is valid for all yC(W).

Proposition 4.1.

Consider the models M12 and M1 and suppose that μ1=X1β1 is estimable under M12. Using the earlier notation, we have for all yC(W):(4.1) (G1G1#)y=G1G2#y=X1(X1W1+X1)X1W1+·X2(X2Ṁ1X2)X2Ṁ1y.(4.1)

Proof.

It is clear that G1G1#=G1#. Premultiplying G12=PWVM(MVM)MPW by G1 we observe that G1G12=G1 as G1VM=0. Thus we have (4.2) G1G1#=G1(G12G1#)=G1G2#.(4.2) The claim Equation(4.1) follows from Equation(4.2). □

Proposition 4.1 was proved by Haslett and Puntanen (Citation2010, Lemma 3.1) in the situation when C(X2)C(X1:V)=C(W1). Using different formulation and proof, it appears also in Werner and Yapar (Citation1996, Th. 2.3). See also Sengupta and Jammalamadaka (Citation2003, Ch. 9) and Güler, Puntanen, and Özdemir (Citation2014). In the full rank model, that is, when X has full column rank and V is positive definite, it appears, for example, in Haslett (Citation1996).

Remark 4.1.

We might be tempted to express the equality G1y=G1#y as (4.3) μ˜1(M1)=μ˜1(M12),i.e.,BLUE(μ1|M1)=BLUE(μ1|M12).(4.3) However, the notation used in Equation(4.3) can be problematic when the possible values of the response vector y are taken into account. It is clear that G1y is the BLUE of μ1 under M1 and we may write shortly G1y=μ˜1(M1). Now, there might be another estimator Ay for which we can also write Ay=μ˜1(M1) but, however, Ay and G1y may have different numerical observed values. The numerical value of the BLUE under M1 is unique if and only if y lies in C(W1).

Notice that in above considerations all the matrices G1, G12 and so on. are fixed. Let us check whether Equation(4.1) holds for arbitrary H1{Pμ1|M1}, H12{Pμ|M12} and so on.

Corollary 4.1.

Let us denoteH1=G1+N1QW1,H12=G12+N2QW,H1#=G1#+N3QW,H2#=G2#+N4QW,where the matrices N1,,N4 are free to vary. Then

  1. (H1H1#)y=G1G2#y+N1QW1y  for all yC(W),

  2. (H1H1#)y=H1H2#y  for all  yC(W).

    Moreover, the following two statements are equivalent:

  3. C(X2)C(W1),

  4. (H1H1#)y=G1G2#y  for all  yC(W).

Proof.

In view of (H1H1#)W=(G1+N1QW1G1#N3QW)W=(G1+N1QW1G1#)W=(G1G1#)W+N1QW1W=G1G2#W+N1QW1W, the statement (a) holds. We observe that H1H2#W=(G1+N1QW1)G2#W=G1G2#W+N1QW1G2#W. Thus the statement (b), that is, the equality (H1H1#)W=H1H2#W holds if and only if (4.5) QW1W=QW1G2#W.(4.5) Replacing W with (X1:X2:VM) in Equation(4.5) we observe that Equation(4.5) indeed holds. The equivalence of (c) and (d) is obvious. □

Proposition 4.2.

Consider the models M12 and M1 and suppose that μ1=X1β1 is estimable under M12. Then the following statements are equivalent:

  1. G1y=G1#y for all yC(W), that is, G1W=G1#W,

  2. G1y=G1#y for all yC(X1:X2),

  3. G1y=G1#y for all yRn, that is, G1=G1#,

  4. G1{Pμ1|M12}, that is, G1y{BLUE(μ1|M12)},

  5. X1W1+X2=0,

  6. G1X2=0,

  7. C(X2)C(W1+X1)=C(W1M1:QW1)=C(VM1:QW1).

Proof.

Consider the statement (a) which is obviously equivalent to (d): (4.6) G1(X1:X2:VM)=G1#(X1:X2:VM).(4.6) Now G1VM=G1VM1QM1X2=0 and hence Equation(4.6) holds if and only if (4.7) (X1:G1X2:0)=(X1:0:0),(4.7) that is, G1X2=X1(X1W1X1)X1W1+X2=0, which is equivalent to X1W1+X2=0. The equivalence between (a) and (b) follows from the equivalence between Equation(4.6) and Equation(4.7).

To prove that (a) and (c) are equivalent we need to show that G1QW=G1#QW. It is clear that G1QW=0. Similarly, G1#QW=D1#QW=0. Thus (a) is equivalent to (c). The claim (g) follows from part (b) of Proposition 3.3. □

Remark 4.2.

Clearly (a) in Proposition 4.2 is equivalent to (i)G1(X1:X2)=G1#(X1:X2)=(X1:0)and(ii)G1V=G1#V, that is, (i) G1X2=0 and (ii) G1V=G1#V. Here is a question: where does the condition (ii) vanish in Proposition 4.2?

In view of Proposition 4.2, the condition (i) implies that G1=G1#, and hence trivially (ii) holds, that is, G1V=G1#V. However, (ii) does not imply (i). Moreover, the condition (ii) implies that cov(G1y)=cov(G1#y) which by Proposition 4.3 (see below) is equivalent to X1W+X2=0. Thus we can conclude that X1W1+X2=0X1W+X2=0.

In Propositions 4.34.5 we assume that μ1=X1β1 is estimable under M12.

Proposition 4.3.

The following statements are equivalent:

  1. G1y=G1#y for all yC(W1), that is, G1W1=G1#W1,

  2. G1#{Pμ1|M1}, that is, G1#y{BLUE(μ1|M1)},

  3. {BLUE(μ1 | M12)}{BLUE(μ1 | M1)}, that is, {Pμ1|M12}{Pμ1|M1},

  4. (H1H1#)W1=0  for all  H1{Pμ1|M1},H1#{Pμ1|M12},

  5. G1#VM1=0,

  6. C(WṀ2X1)=C(W1Ṁ2X1)=C(X1),

  7. WṀ2X1=W1Ṁ2X1=X1,

  8. X1W+X2=0,

  9. G1V=G1#V,

  10. cov(G1#yG1y)=0,

  11. cov(G1y)=cov(G1#y).

    Moreover, we always have

  12. cov(G1#yG1y)=cov(G1#y)cov(G1y),

  13. cov(G1y)Lcov(G1#y),

  14. X1W1+X2=0X1W+X2=0.

Proof.

It is clear that (b) is simply an alternative expression for (a) and similarly (d) for (c). The claim (a) holds if and only if G1(X1:VM1)=G1#(X1:VM1)=(X1:0), which gives (e): G1#VM1=0, that is, (4.8) X1(X1Ṁ2X1)X1Ṁ2VM1=0.(4.8) Premultiplying Equation(4.8) by X1Ṁ2 yields X1Ṁ2VM1=X1Ṁ2WM1=X1Ṁ2W1M1=0, that is, C(WṀ2X1)C(X1). In view of Proposition 3.2, we have r(WṀ2X1)=r(X1) and hence C(WṀ2X1)C(X1) becomes (4.9) C(WṀ2X1)=C(W1Ṁ2X1)=C(X1).(4.9) Thus we have shown that (e) and (f) are equivalent. Equality Equation(4.9) implies (4.10) X1W+X2=0,(4.10) that is, (f) implies (h). In view of part (e) of Proposition 3.3 we have (4.11) WṀ2X1=[InX2(X2W+X2)X2W+]X1.(4.11) Substituting Equation(4.10) into Equation(4.11) we observe that (h) implies (g), and so far we have confirmed the equivalence between (a) and any of (e)–(h).

The statement (c) holds if and only if (G1#+N2QW)(X1:VM1)=(X1:0)for all N2Rn×n, that is, (G1#+N2QW)VM1=0for all N2Rn×n, which holds if and only if G1#VM1=0. Thus (c) and (e) are equivalent.

The claim (a) holds if and only if G1(X1:V)=G1#(X1:V), which is precisely (l): G1V=G1#V. It is clear that (i) is equivalent to (j). Consider then cov(G1#G1)y=G1#VG1#+G1VG1G1#VG1G1VG1#. Notice that G1T1=G1#T1=T1, where T1=X1U1U1X1  and hence G1VG1=G1(W1T1)G1=X1(X1W1+X1)X1G1T1=X1(X1W1+X1)X1T1, and G1VG1#=G1(W1T1)G1#=G1W1G1#T1=X1(X1W1X1)X1G1#T1=X1(X1W1+X1)X1T1=G1VG1. Thus cov(G1#yG1y)=cov(G1#y)cov(G1y), and so (l) and (m) hold. Statement (l) obviously confirms the equivalence between (j) and (k). Property (n) is obvious. See also Remark 4.1. □

Next we consider the condition under which an arbitrary matrix from the set {Pμ1|M1} provides the BLUE for μ1 under M12.

Proposition 4.4.

The following statements are equivalent:

  1. {BLUE(μ1 | M1)}{BLUE(μ1 | M12)}, that is, {Pμ1|M1}{Pμ1|M12},

  2. (H1H1#)W=0  for all  H1{Pμ1|M1},H1#{Pμ1|M12},

  3. C(X2)C(X1:VM1), that is, C(W1)=C(W), and X1W1+X2=0,

  4. C(X2)C(VM1),

  5. C(X2:VM)C(VM1),

  6. {BLUE(μ1 | M1)}={BLUE(μ1 | M12)}, that is, {Pμ1|M1}={Pμ1|M12},

  7. C(X2:VM)=C(VM1).

Proof.

Notice first that (b) is simply an alternative way to express (a). The statement (a) holds if and only if (G1+N1QW1)(X1:X2:VM)=(X1:0:0)for all N1Rn×n, that is, (G1+N1QW1)X2=0for all N1Rn×n, which holds if and only if QW1X2=0 and G1X2=0, which is precisely (c). Moreover, (c) implies that (4.12) X2=X1A+VM1B(4.12) for some A and B and (4.13) X1W1+(X1A+VM1B)=X1W1+X1A=0.(4.13) Now Equation(4.13) implies that W1+X1A=0, which further implies that X1A=0, so that by Equation(4.12) we get (d). The claim (d) obviously implies (c). The equivalence between (d) and (e) is obvious because C(VM)C(VM1).

It is clear that (f) implies (b). Thus to confirm the equivalence of (b) and (f) we have to show that (4.14) (b){BLUE(μ1|M12)}{BLUE(μ1|M1)}.(4.14) This follows at once from Proposition 4.3 by noting that the right-hand side of Equation(4.14) means that (H1H1#)W1=0. The equivalence between (f) and (g) follows by combining part (d) of Proposition 4.4 and (k) of Proposition 4.3. □

Our next task is to find necessary and sufficient conditions for G1y=G1#y for all yC(W) when the inclusion C(X2)C(X1:V) holds.

Proposition 4.5.

Consider the models M12 and M1 and suppose that(4.15) C(X2)C(X1:V)=C(W1),i.e.,C(W1)=C(W).(4.15) Then the following statements are equivalent:

  1. G1W1=G1#W1,

  2. H1W1=H1#W1 for all H1 and H1#,

  3. {BLUE(μ1|M12)}{BLUE(μ1|M1)}, that is, {Pμ1|M12}{Pμ1|M1},

  4. {BLUE(μ1|M1)}{BLUE(μ1|M12)}, that is, {Pμ1|M1}{Pμ1|M12},

  5. BLUE(μ1|M1)=BLUE(μ1|M12) with probability 1,

  6. X1W1+X2=0,

  7. C(X2)C(VM1),

  8. X1C12X2=0, where C12 is defined as(4.16) (XW1+X)+=(X1W1+X1X1W1+X2X2W1+X1X2W1+X2)+=(C11C12C21C22).(4.16)

Proof.

The equivalence between (a)–(g) is obvious. Consider then part (h). Now we have (4.17) D1#=(X1:0)(XWX)XW+=(X1:0)(XW1X)XW1+.(4.17) Hence (a) holds, under Equation(4.15), if and only if G1W1=D1#W1, that is, G1(X1:VM1)=D1#(X1:VM1)=(X1:0), that is, (4.18) D1#VM1=D1#W1M1=0.(4.18) Using Equation(4.17) the equality Equation(4.18) becomes (4.19) D1#VM1=(X1:0)(XW1X)XW1+W1M1=(X1:0)(XW1X)XPW1M1=(X1:0)(XW1X)XM1=(X1:0)(XW1X)(0:M1X2)=X1C12X2M1=0,(4.19) where C12 is defined in Equation(4.16). In light of r(X2M1)=r(X2), we can cancel M1 in the last expression in Equation(4.19). This proves the equivalence between (a) and (h). □

5. Conclusions

In this article we consider the partitioned linear model M12={y, X1β1+X2β2, V} and the corresponding small model M1={y, X1β1, V}. We focus on comparing the BLUEs of μ1=X1β1 under M12 and M1. The observed numerical value of the BLUE is unique under the model M1 if the M1 is consistent in the sense that yC(X1:V) and the same uniqueness concerns the full model in the respective way. But now there may be some problems if we write (5.1) BLUE(X1β1|M1)=BLUE(X1β1|M12).(5.1) What is the meaning of the above equality? It is not fully clear because we know that under M1 the values of y vary over C(X1:V) but under M12 the values of y vary over C(X1:X2:V) and these column spaces may be different. However, if C(X1:V)=C(X1:X2:V) there is no difficulties to interpret the equality Equation(5.1), which means that Ay=Byfor all yC(X1:V), where Ay{BLUE(μ1|M1)} and By{BLUE(μ1|M12)}.

We consider the resulting problems by picking up particular fixed expressions for the BLUE s of μ1=X1β1 under these two models, and study the conditions under which they are equal for all values of yC(X1:X2:V) or yC(X1:V). Moreover, we review the conditions under which all representations of the BLUE s in one model continue to be valid in the other model. Some related considerations, using different approach, have been made by Lu et al. (Citation2015), Tian (Citation2013), and Tian and Zhang (Citation2016).

Acknowledgements

Part of this research was done during meeting of an International Research Group on Multivariate and Mixed Linear Models in the Mathematical Research and Conference Center, Bȩdlewo, Poland, in November 2019 and February 2020. Thanks go to the anonymous referee for constructive remarks.

References

  • Groß, J., and S. Puntanen. 2000. Estimation under a general partitioned linear model. Linear Algebra and Its Applications 321, 131–44. doi:10.1016/S0024-3795(00)00028-8.
  • Güler, N., S. Puntanen, and H. Özdemir. 2014. On the BLUEs in two linear models via C.R. Rao’s Pandora’s Box. Communications in Statistics - Theory and Methods, 5, 43, 921–31. doi:10.1080/03610926.2013.826366.
  • Haslett, S. J. 1996. Updating linear models with dependent errors to include additional data and/or parameters. Linear Algebra and Its Applications 237(238):329–49.
  • Haslett, S. J., A. Markiewicz, and S. Puntanen. 2020. Properties of BLUEs and BLUPs in full vs. small linear models with new observations. In Recent developments in multivariate and random matrix analysis: Festschrift in honour of Dietrich von Rosen, eds. T. Holgersson and M. Singull, 123–46. Cham: Springer.
  • Haslett, S. J., and S. Puntanen. 2010. Effect of adding regressors on the equality of the BLUEs under two linear models. Journal of Statistical Planning and Inference 140, 104–10. doi:10.1016/j.jspi.2009.06.010.
  • Lu, C., S. Gan, and Y. Tian. 2015. Some remarks on general linear model with new regressors. Statistics & Probability Letters, 97, 16–24. doi:10.1016/j.spl.2014.10.015.
  • Markiewicz, A, and S. Puntanen. 2019. Further properties of the linear sufficiency in the partitioned linear model. In Matrices, statistics and big data, eds. S. E. Ahmed, F. Carvalho and S. Puntanen, 1–22. Cham: Springer.
  • Marsaglia, G, and G. P. H. Styan. 1974. Equalities and inequalities for ranks of matrices. Linear Multilinear Algebra 2:269–92.
  • Mitra, S. K, and B. J. Moore. 1973. Gauss–Markov estimation with an incorrect dispersion matrix. Sankhyā Series A 35:139–52.
  • Puntanen, S., G. P. H. Styan, and J. Isotalo. 2011. Matrix tricks for linear statistical models: our personal top twenty. Heidelberg: Springer.
  • Rao, C. R. 1973. Representations of best linear estimators in the Gauss–Markoff model with a singular dispersion matrix. Journal of Multivariate Analysis 3, 276–92. doi:10.1016/0047-259X(73)90042-0.
  • Rao, C. R. 1974. Projectors, generalized inverses and the BLUE’s. Journal of the Royal Statistical Society: Series B 36:442–8.
  • Rao, C. R, and S. K. Mitra. 1971. Generalized inverse of matrices and its applications. New York: Wiley.
  • Sengupta, D, and S. R. Jammalamadaka. 2003. Linear models: An integrated approach. River Edge: World Scientific.
  • Tian, Y. 2013. On properties of BLUEs under general linear regression models. Journal of Statistical Planning and Inference 43, 771–82. doi:10.1016/j.jspi.2012.10.005.
  • Tian, Y., and X. Zhang. 2016. On connections among OLSEs and BLUEs of whole and partial parameters under a general linear model. Statistics & Probability Letters 112, 105–12. doi:10.1016/j.spl.2016.01.019.
  • Werner, H.J., and C. Yapar. 1996. A BLUE decomposition in the general linear regression model. Linear Algebra and its Applications 237-238, 395–404. doi:10.1016/0024-3795(95)00542-0.