Home Mathematics A note on star partial order preservers on the set of all variance-covariance matrices
Article Open Access

A note on star partial order preservers on the set of all variance-covariance matrices

  • Gregor Dolinar , Bojan Kuzma , Janko Marovt EMAIL logo and Dijana Mosić
Published/Copyright: February 15, 2023
Become an author with De Gruyter Brill

Abstract

Let Hn+(R) be the cone of all positive semidefinite n × n real matrices. We describe the form of all surjective maps on Hn+(R) , n=3, that preserve the star partial order in both directions.

MSC 2010: 15B48; 47B49; 47L07; 54F05; 62J99

1 Introduction

Let F denote the field of all real numbers R or the field of all complex numbers C , and let Mm,n(F) be the set of all m × n matrices over F . When m=n, we write Mn(F) instead of Mn,n(F) . By A* we denote the conjugate transpose of AMm,n(F) (if A ∈ Mm, n(ℝ), then A*=At, the transpose of A). For AMm,n(F) , let ImA and rank(A) be the image and the rank of A, respectively. Let Hn(F) denote the set of all Hermitian (i.e., symmetric in the real case) matrices in Mn(F) . A matrix AHn(F) is said to be positive semidefinite if x*Ax⩾0 for every xFnMn,1(F) . We denote by Hn+(F) the cone of all positive semidefinite matrices in Hn(F) , and let Pn(F) be the set of all idempotent matrices in Hn+(F) (i.e., the set of all orthogonal projector (or self-adjoint idempotent) matrices in Mn(F)) . The study of positive semidefinite matrices is a flourishing area of mathematical investigation (see, e.g., the monograph [1] and the references therein). Positive semidefinite matrices have become fundamental computational objects in many areas of statistics, engineering, quantum information, and applied mathematics. They appear as variance-covariance matrices (also known as dispersion or covariance matrices) in statistics [5], as elements of the search space in convex and semidefinite programming [17], as kernels in machine learning [22], as density matrices in quantum information [16], and as diffusion tensors in medical imaging [4]. It is known (see, e.g., [5: page 4]) that every variance-covariance matrix is positive semidefinite, and that every real positive semidefinite matrix is a variance-covariance matrix of some multivariate distribution.

There are many partial orders that may be defined on various sets of matrices and a few of them have applications in statistics especially in the theory of linear models. Let A,BHn(F) . Then we say that A is below (or is dominated by) B with respect to the Löwner partial order and write

ALBifBAis positive semidefinite.

Let y=Xβ +ε be the vector form of a linear model. Here y is a real n × 1 random vector of observed quantities, which we try to explain with unknown parameters given by vector β, while the coefficients of the matrix X ∈ Mn, p(ℝ) are known (determined by the model), see, e.g., [5: Chapter 1]. It is assumed that mathematical expectation of ε is 0 and that σ 2D is the variance-covariance matrix of ε . The nonnegative parameter σ 2 and the vector of parameters (real numbers) β are unspecified, and D ∈ Hn(ℝ) is a known positive semidefinite matrix. We denote this linear model with the triplet (y, Xβ, σ 2D). Statistical analysis often focuses on answering questions about certain linear functions of the form Cβ for a specified real matrix C with p columns. We try to estimate Cβ by a linear function Ay of the response y (here A is a real matrix with n columns). We say that Ay is a linear unbiased estimator (LUE) of Cβ if the mathematical expectation of Ay equals Cβ for all possible values of β ∈ ℝp. The function Cβ is said to be estimable if it has LUE. The best linear unbiased estimator (BLUE) of an estimable Cβ is defined as LUE having the smallest variance-covariance matrix (“smallest” in terms of the Löwner partial order on Hn+(R)) .

Another well-known partial order is the star partial order.It was introduced in [8] and can be on Mm,n(F) defined as follows (see also [21: Chapter 5]). For A,BMm,n(F) , we write

(1) ABifAA=ABandAA=BA.

Two partial orders that are related to the star partial order are the left-star and the right-star partial orders [2]. For A,BMm,n(F) , we say that A is below B with respect to the left-star partial order and write

ABifAA=ABandImAImB.

Similarly, we define the right-star partial order. For A,BMm,n(F) , we write

ABifAA=ABandImAImB.

Note that for A,BMm,n(F) , A*B implies A* ⩽ B and A⩽ * B (see, e.g., [21: Theorem 6.5.14]), and that the star, the left-star, and the right-star partial orders are the same partial order on Hn(F) .

The left-star partial order has applications in the theory of linear models. Let us present two examples of such applications. Linear models (y, Xβ, σ 2D), where D=I is the identity matrix, are called the Gauss-Markov linear models. The following result gives an interpretation of the left-star order in such linear models (see [21: Theorem 15.3.7]).

Proposition 1

Let L1=(y, X1β, σ2I) and L2=(y, X2β, σ2I) be any two (Gauss-Markov) linear models. Then X1* ⩽ X2 if and only if the following statements hold.

  1. The linear models L1 and L=(y, (X2X1)β, σ 2I) have no common estimable linear function of β;

  2. X1β is estimable under the model L2;

  3. The BLUE of X1β under the model L1 is also its BLUE under L2, and the variance-covariance matrix of the BLUE of X1β under the model L1 is the same as under the model L2.

Let X ∈ Mn, p(ℝ) and let sβ be an estimable linear parametric function (here s is a 1 × p real vector). A linear function of the response, ly (here l is a 1 × n real vector) is called a linear zero function (LZF) if E (ly)=0 for all possible values of β. Here E (ly) denotes the mathematical expectation of the random variable ly. Due to the additivity property of the mathematical expectation, one can easily see that by adding LZFs to LUE, we get other LUEs of sβ. Next we present another application in the theory of comparison of linear models when the model matrices are related via the left-star partial order (see [21: Corollary 15.3.8]).

Proposition 2

Let L1=(y, X1β, σ2I) and L2=(y, X2β, σ2I) be any two linear models such that X1* ⩽ X2. Then the BLUE of every estimable linear function of the parameters under L1 is a linear zero function under L=(y, (X2X1)β, σ2I) and vice versa.

Let A be some subset of Mn(F) and denote by ⩽ a partial order on A . We say that a map Φ:AA is a bi-preserver of the order ⩽ (or that it preserves the order ⩽ in both directions) when

ABif and only ifΦ(A)Φ(B)

for every A,BA . Instead of the linear model M=(y, Xβ, σ 2I) one might rather work with the transformed model Mˆ=(y,Xˆβ,σ2I) , where the new matrix XˆMn(R) is chosen so that its properties are more attractive than X ∈ Mn(ℝ) (e.g., elements of X that are very close to zero are transformed to zero), and thus it is natural (see [6]) to demand that the transformed model still retains most of the properties of the original model (e.g., has similar relations to other transformed models). Thus, in view of Propositions 1 and 2, it is interesting to know what transformations on Mn(ℝ) or perhaps on some subset of Mn(ℝ) (like Hn+(R)) preserve the left-star partial order. Linear bijective maps on Mn(F) that preserve either the star, or the left-star, or the right-star partial order in one direction (i.e., A ⩽ B implies Φ(A) ⩽ Φ(B), where ⩽ is the appropriate order) were investigated in [12] (see also [11]). In [6], the forms of surjective bi-preservers of the left-star partial order and the right-star partial order on Mn(F) , n=3, were described. Surjective bi-preservers of the star partial order on Mn(F) , n=3, were characterized in [18]. Since surjective bi-preservers of the star, the left-star, and the right-star partial orders on the full matrix algebra Mn(F) , n=3, have already been described (see also [13]) and because the set of all positive semidefinite real matrices has an important role in statistics, we focus our attention on the star partial order bi-preservers on Hn+(R) . In [10] (see also [9]), the following result was proved.

Theorem 3

Let n=3 be an integer. Then Φ:Hn+(R)Hn+(R) is a surjective, additive bi-preserver of the star partial order if and only if there exist an orthogonal matrix R ∈ Mn(ℝ) and λ>0 such that

Φ(A)=λRARt

for every AHn+(R) .

In this paper, we study bi-preservers of the star partial order on Hn+(R) that may be nonadditive. In Section 2, we give some preliminary results. The main result which describes the form of all surjective bi-preservers of the star partial order on Hn+(R) , n=3, is stated and proved in Section 3.

2 Preliminaries

The following properties of the star partial order are well-known and can be easily verified (see, e.g., [18: Proposition 2.1, Corollary 3.6]).

Proposition 4

Let A,BMn(F) and αF with α≠0. Let U,VMn(F) be unitary matrices (i.e., orthogonal when F=R) . The following statements are then equivalent.

  1. A* B.

  2. α A* α B.

  3. UAV*UBV.

Proposition 5

Let A,BMn(F) . If A*B and AB, then rank(A)<rank(B) .

The next lemma states that any Hermitian matrix A is idempotent if and only if the identity matrix I is its upper bound with respect to the star partial order (see [15: Theorem 3.4]).

Lemma 6

For AHn(F) , A*I if and only if A2=A.

Proof. Suppose A2=A, AHn(F) . Then A*A=A2=A=AI=A*I and similarly AA*=IA*. So, A*I. Conversely, suppose A*I for AHn(F) . By, e.g., [20: Theorem 8], A2=A. □

Let x,yFn be nonzero. We denote by xy* a rank-one matrix in Mn(F) . Note that it represents a rank-one linear operator on Fn defined with (xy*)z=〈z, yx for every zFn (here 〈z, y〉=y*z is an inner product in Fn) . It is known that every rank-one matrix in Mn(F) may be written in this form and that PPn(F) is of rank-one if and only if P=x x* for some xFn with ∣ x∣=1. The following result was proved in [18] (see also [10]).

Proposition 7

Let AMn(F) and let x,yFn be nonzero. Then

xyAif and only ifAx=x,xyandAy=y,yx.

Corollary 8

Let AHn(F) and let xFn be nonzero. Then

xxAif and only ifAx=x,xx.

Denote by Ei,jMn(F) the matrix with all entries equal to zero except the (i, j)-entry which is equal to one, and by e1,e2,,enFn the standard basis vectors.

Lemma 9

Let xx*, yyHn+(F) be rank-one matrices with different spectra. Then x x* and y y* have a common upper bound with respect to the star partial order if and only if vectors x and y are orthogonal.

Proof By using unitary similarity, we may by Proposition 4 assume that x=λ e1 for some nonzero λR and y=α e1+β e2, α,βF , where 0<α2+β2 . Note that λ 2α 2+β 2 since x x* and y y* have different spectra.

Suppose first that x and y are orthogonal. Then α=0 and hence A=xx+yy=λ2e1e1+β2e2e2 is a common upper bound of x x* and y y*.

Suppose now x and y are not orthogonal, and let us assume that there exists A=(ai,j)Hn+(F) which is a common upper bound of x x* and y y*. By Corollary 8, we have A(λ e1)=λ 3e1 and thus Ae1=λ 2e1. Also, A(α e1+β e2)=(α 2+β 2)(α e1+β e2). It follows that

αλ2e1+βAe2=(α2+β2)αe1+(α2+β2)βe2

and therefore,

(2) βAe2=(α2+β2λ2)αe1+(α2+β2)βe2.

If α=0, then y=β e2 and thus x and y are orthogonal, a contradiction. If β=0, then we get a contradiction by (2) since α≠ 0 and α 2+β 2λ 2≠ 0. By Ae1=λ 2e1, we obtain a2,1=0, and since β≠ 0, we get (see (2)) that a1,2≠ 0. So, AHn(F) , a contradiction. □

Let AMn(F) be nonzero. As a direct corollary of the singular value decomposition there is a unique decomposition

A=j=1ktjVj,

where t1>t2>>tk>0 and V1, V2, ⋅, Vk are pairwise orthogonal nonzero partial isometries. This decomposition is known as the Penrose decomposition. Here rank A=j=1k rank (Vj) . In [14: page 371] (see also [18: Theorem 3.3]), the following result was proved.

Proposition 10

Let A, BMn(F) have the Penrose decompositions

A=j=1ktjVj,B=i=1luiWi.

Then A*B if and only if there exists a function φ:{1, 2, ⋅, k} → {1, 2, ⋅, l} such that tj=uφ(j) and Vj*Wφ(j) for all j ∈ {1, 2, ⋅, k}.

Remark 11

Note that the function φ from Proposition 10 is injective. Indeed, if φ(i)=φ(j), then ti=uφ(i)=uφ(j)=tj which implies by the definition of the Penrose decomposition that i=j.

Let now AHn+(F) be nonzero. By the spectral theorem there exists a unique (eigenvalue) decomposition

A=j=1kλjPj,

where λ1>λ2>>λk>0 are (nonzero) eigenvalues of A and P1, P2, …, Pk are pairwise orthogonal nonzero projectors. Observe that this is in fact also the Penrose decomposition of A. We thus have the following direct corollary of Proposition 10.

Corollary 12

Let A,BHn+(F) have the eigenvalue decompositions

A=j=1kλjPj,B=i=1lμiQi.

Then A*B if and only if there exists a function φ:{1, 2, ⋅, k} → {1, 2, ⋅, l} such that λj=μφ(j) and Pj*Qφ(j) for all j ∈ {1, 2, …, k}.

Let ℝ+=[0, ∞) denote the set of all nonnegative real numbers. For AHn+(F) , let Sp(A) denote the spectrum of A, i.e., the set of all eigenvalues of A, and let #Sp(A) denote the number of (pairwise distinct) eigenvalues of A.

Definition 1

Let f:ℝ+ → ℝ+ be a function. For AHn+(F) , let r=#Sp(A) and let pA be the polynomial of the degree r − 1 such that the restriction of pA to Sp(A) equals the restriction of f to Sp(A) . For this A, we define

f(A):=pA(A).

Remark 13

Note that for every fixed AHn+(F) the polynomial pA, introduced in Definition 1, exists and is unique. Moreover, we could take any polynomial qA such that its restriction to Sp(A) equals the restriction of f to Sp(A) . Namely, for any λSp(A) , we have (qApA)(λ)=f(λ)−f(λ)=0 and thus (qApA)(A)=0.

Remark 14

Note that when f is analytic or even measurable, f(A) as defined in Definition 1 matches f(A) obtained via the functional calculus.

Remark 15

Let f:ℝ+ → ℝ+ be a function with f(0)=0, and let Φ:Hn+(F)Hn+(F) be defined as Φ(A)=f(A). Let A=j=1kλjPj be the eigenvalue decomposition of AHn+(F) . Observe, by the orthogonality of projectors involved, that then Φ(A)=j=1kf(λj)Pj and thus Φ(Hn+(F))Hn+(F) .

Lemma 16

Let f:ℝ+ → ℝ+ be a bijective function with f (0)=0 and let Φ:Hn+(F)Hn+(F) be defined as

Φ(A)=f(A).

Then Φ is a bijective star partial order bi-preserver.

Proof. Let BHn+(F) be nonzero and let B=i=1lμiQi be the eigenvalue decomposition of B. Let A=i=1lf1(μi)Qi . This is up to a permutation of indices the eigenvalue decomposition of A. Thus

Φ(A)=i=1lf(f1(μi))Qi=i=1lμiQi=B.

Since f (0)=0, we have Φ (0)=0 and thus Φ is surjective. Let now Φ(A)=Φ(B)≠0 and let j=1kλjPj be the eigenvalue decomposition of Φ(A). Then A=j=1kf1(λj)Pj=B . If Φ(A)=Φ(B)=0, then A=B=0 since f (0)=0 and f is injective. We may conclude that Φ is injective and thus bijective.

Suppose now A*B for some nonzero A, BHn+(F) . Let A=j=1kλjPj , B=i=1lμiQi be the eigenvalue decompositions of A and B, respectively. By Corollary 12, there exists a function φ:{1, 2, …, k} → {1, 2, …, l} such that λj=μφ(j) and Pj*Qφ(j) for all j ∈ {1, 2, …, k} . Also,

Φ(A)=j=1kf(λj)Pj,Φ(B)=i=1lf(μi)Qi

and therefore, since λj=μφ(j) implies f(λj)=f(μφ(j)), we may again by Corollary 12 conclude that Φ(A) ⩽ *Φ(B). If A=0, then Φ(A)=f (0)=0 and so (see (\refdef_star)) A*B and Φ(A) ⩽ *Φ(B) for every BHn+(F) . We proved that A*B implies Φ(A) ⩽ *Φ(B) for every A,BHn+(F) . Since f is bijective, we may similarly show that Φ(A) ⩽ *Φ(B) implies A*B. □

3 Statement and proof of the main result

We use the notation A<B if A*B and AB, and denote by diag(di)Mn(R) a diagonal matrix with real diagonal elements di, i=1, 2, …, n. For a subspace V of ℝn, we denote by PV ∈ Pn(ℝ) the orthogonal projector matrix with ImPV=V . Observe first (see [3]) that then for any pair of matrices P, Q ∈ Pn(ℝ), we have

PQif and only ifP=QP=PQif andonly ifImPImQ.

Our main result follows.

Theorem 17

Let n=3 be an integer. Then Φ:Hn+(R)Hn+(R) is a surjective bi-preserver of the star partial order if and only if there exist an orthogonal matrix Q ∈ Mn(ℝ), a positive λR , and a bijective function f:ℝ+↦ℝ+ with f (0)=0 such that

Φ(A)=λQf(A)Qt

for every AHn+(R) .

Proof. Let Φ:Hn+(R)Hn+(R) be of the form Φ(A)=λ Qf (A)Qt, AHn+(R) , where Q, λ and f are as above. From Proposition 4 and Lemma 16, it follows that Φ is a bijective bi-preserver of the star partial order.

Conversely, let Φ:Hn+(R)Hn+(R) be a surjective bi-preserver of the star partial order. We will split the proof into several steps. The first four steps are standard and we include them for the sake of completeness.

Step 1. Φ is bijective. Let Φ(A)=Φ(B) for A,BHn+(R) . Thus Φ(A) ⩽ *Φ(B) and Φ(B) ⩽ *Φ(A) and since Φ is a bi-preserver of the star partial order, we have A*B and B*A. So A=B, i.e., Φ is injective and hence bijective.

Step 2. Φ (0)=0. By (\refdef_star), 0 ⩽ *A for every AHn+(R) . So on the one hand 0 ⩽ *Φ (0) and on the other hand, since Φ −1 has the same properties as Φ, 0 ⩽ *Φ −1(0). Thus Φ (0) ⩽ *0 and therefore Φ (0)=0.

Step 3. Φ preserves rank.Let AHn+(R) with rank(A)=k . There exists an orthogonal matrix U ∈ Mn(ℝ) such that

A=UDUt,

where D=diag(di) with di⩾0 for every i=1, 2, ⋅, n . Without loss of generality we may assume that d1d2dk>0=dk+1=dk+2==dn and write D=d1E1,1+d2E2,2+⋅+dkEk, k. Clearly,

0 < d 1 E 1 , 1 < d 1 E 1 , 1 + d 2 E 2 , 2 < < d 1 E 1 , 1 + d 2 E 2 , 2 + + d k E k , k = D < d 1 E 1 , 1 + d 2 E 2 , 2 + + d k E k , k + E k + 1 , k + 1 < < d 1 E 1 , 1 + d 2 E 2 , 2 + + d k E k , k + E k + 1 , k + 1 + + E n , n .

By Proposition 4 and since congruence preserves rank, we have

0 < Φ ( U ( d 1 E 1 , 1 ) U t ) < Φ ( U ( d 1 E 1 , 1 + d 2 E 2 , 2 ) U t ) < < Φ ( A ) < Φ ( U ( d 1 E 1 , 1 + d 2 E 2 , 2 + + d k E k , k + E k + 1 , k + 1 ) U t ) < < Φ ( U ( d 1 E 1 , 1 + d 2 E 2 , 2 + + d k E k , k + E k + 1 , k + 1 + + E n , n ) U t ) .

Since Φ (0)=0 and since Φ preserves the order ⩽ * and is injective, we obtain

0 < Φ ( U ( d 1 E 1 , 1 ) U t ) < Φ ( U ( d 1 E 1 , 1 + d 2 E 2 , 2 ) U t ) < < Φ ( A ) < Φ ( U ( d 1 E 1 , 1 + d 2 E 2 , 2 + + d k E k , k + E k + 1 , k + 1 ) U t ) < < Φ ( U ( d 1 E 1 , 1 + d 2 E 2 , 2 + + d k E k , k + E k + 1 , k + 1 + + E n , n ) U t ) .

By Proposition 5, every successor in the above chain of matrices is of rank strictly greater than its predecessor. Since rank of any matrix in Hn+(R) can not be greater than n, it follows that rankΦ(A)=k .

Step 4. We may without loss of generality assume that Φ(I)=I. By the previous step, Φ(I) is of rank n. It follows that there exist an orthogonal matrix U ∈ Mn(ℝ) and λi>0 , i=1, 2, ⋅, n, with Φ(I)=Udiag(λi)Ut . We may thus without loss of generality assume that Φ(I)=diag(λi) . Suppose there are λi, λj with λiλj, i, j ∈ {1, 2, ⋅, n} . By Corollary 12, there are exactly two rank-one matrices that are below λiEi, i+λjEj, j with respect to the star partial order, namely λiEi,i,λjEj,jHn+(R) . Since Φ preserves the order in both directions, we have Φ −1(λiEi, i+λjEj, j) ⩽ *I. By Lemma 6 and the previous step, Φ −1(λiEi, i+λjEj, j) is a rank-two self-adjoint idempotent matrix. This is a contradiction, since on the one hand Φ −1(λiEi, i) and Φ −1(λjEj, j) are the only two rank-one matrices that are below Φ −1(λiEi, i+λjEj, j) but on the other hand, since Φ −1(λiEi, i+λjEj, j) is a self-adjoint idempotent, for every rank-one self-adjoint idempotent P ∈ Mn(ℝ) with ImPImΦ1(λiEi,i+λjEj,j) , we have P*Φ −1(λiEi, i+λjEj, j), and there are infinitely many such matrices P (see (3)). It follows that Φ(I)=λ I for some λ>0 . Let now Ψ:Hn+(R)Hn+(R) be the map defined with

Ψ(A)=1λΦ(A),AHn+(R).

Then Ψ is a surjective map that by Proposition 4 preserves the order ⩽ * in both directions. Moreover, Ψ(I)=I.

We will thus from now on assume that

Φ(I)=I.

Step 5. There exists a bijective functionf:ℝ+ → ℝ+ such that f (0)=0 and Φ(λ x x*)=f(λ)Φ(x x*) for every x ∈ ℝn withx∣=1 and every λ>0 .

To see this, let x ∈ ℝn with ∣ x∣=1 and let λ>0 . By unitary similarity we may without loss of generality assume that x=e1, i.e., x x*=E1,1. The matrices λ E1,1, (λ +2) E2,2, (λ +3) E3,3, …, (λ +n) En, n are pairwise orthogonal and have pairwise different spectra. Similarly, the matrices E1,1, (λ +2) E2,2, (λ +3) E3,3, …, (λ +n) En, n are pairwise orthogonal and have pairwise different spectra. By Lemma 9 and since Φ preserves the star partial order, it follows that matrices

Φ(λE1,1),Φ(λ+2E2,2),Φ(λ+3E3,3),,Φ(λ+nEn,n)

are pairwise orthogonal and similarly

Φ(E1,1),Φ(λ+2E2,2),Φ(λ+3E3,3),,Φ(λ+nEn,n)

are also pairwise orthogonal. Since Φ(λ E1,1) and Φ(E1,1) are pairwise orthogonal to the same set of n − 1 pairwise orthogonal rank-one matrices, they must be linearly dependent, i.e., Φ(λE1,1)=λe1Φ(E1,1) , λe1>0 . So, Φ(λ x x*)=λxΦ(x x*), λx>0 , for every x ∈ ℝn with ∣ x∣=1 and λ>0 .

Suppose now x, y ∈ ℝn with ∣ x∣=∣ y∣=1 are linearly independent and let λ>0 . We have Φ(λ x x*)=λxΦ(x x*) and Φ(λ y y*)=λyΦ(y y*). To show that λx=λy, let P ∈ Pn(ℝ) be of rank-two with x,yImP . By Corollary 12, λ x x*, λ y y**λ P. Note that Φ(λ P) is of rank-two and suppose Φ(λP)=μ1z1z1+μ2z2z2 is the eigenvalue decomposition of Φ(λ P) with μ1μ2. By Corollary 12, μ1z1z1 , μ2z2z2 are the only two rank-one matrices below Φ(λ P) with respect to the star partial order. This is a contradiction, since Φ preserves the order in both directions and there are infinitely many rank-one matrices below λ P. So, Φ(λ P)=μ Q, where μ>0 and Q ∈ Pn(ℝ) is of rank-two. It follows that Φ(λ x x*), Φ(λ y y*) ⩽ *μ Q and thus by Corollary 12, λx=μ=λy.

Recall that Φ (0)=0. We may conclude that there exists a function \f:ℝ+ → ℝ+ such that f (0)=0 and Φ(λ x x*)=f(λ)Φ(x x*) for every x ∈ ℝn with ∣ x∣=1 and every λ>0 . Function f is bijective since Φ is bijective.

Step 6. Letx x*, y y* ∈ Pn(ℝ) be pairwise orthogonal. ThenΦ(x x*), Φ(y y*) are pairwise orthogonal.

By unitary similarity, we may without loss of generality assume that x x*=E1,1 and y y*=E2,2. Lemma 9 implies Φ(E1,1) and Φ (2E2,2) are pairwise orthogonal, and Step 5 yields Φ (2E2,2)=f (2)Φ(E22). Thus, Φ(E2,2) is orthogonal to Φ(E1,1).

Step 7. We may without loss of generality assume thatΦ(P)=P for everyP ∈ Pn(ℝ). Let P ∈ Pn(ℝ) . It follows that P*I by Lemma 6. Therefore Φ(P) ⩽ *Φ(I)=I. Again, by Lemma 6, Φ(P) ∈ Pn(ℝ). Since Φ −1 has the same properties as Φ, we may conclude that

PPn(R)if and only ifΦ(P)Pn(R),

i.e., Φ preserves the set of all orthogonal projector matrices. Recall that we may identify subspaces of ℝn with elements of Pn(ℝ). Let C(Rn) be the lattice of all subspaces of ℝn. It follows that the map Φ induces a lattice automorphism, i.e., a bijective map τ:C(Rn)C(Rn) such that

MNif and only ifτ(M)τ(N)

for all M,NC(Rn) . It was proved in [19: page 246] (see also [7: pages 820 and 823]) that for n=3 every such a map is induced by an invertible linear operator, i.e., there exists an invertible linear operator T:ℝn → ℝn such that τ(M)=T(M) for every MC(Rn) . For the map Φ, we thus have

(4) Φ(P)=PT(ImP)

for every P=PImPPn(R) . Let x x*, y y* ∈ Pn(ℝ) be pairwise orthogonal. By Step 6, Φ(x x*), Φ(y y*) are pairwise orthogonal and hence by (4),

0=Tx,Ty

and therefore 〈TtTx, y〉=0. This equation holds for every y ∈ ℝn with ∣ y∣=1 and 〈x, y〉=0. Since TtTx,y=xyTtTxx,yy , we may conclude that for any fixed x ∈ ℝn, we have 〈TtTx, y〉=0 for every y ∈ ℝn with 〈x, y〉=0. So TtTx is a scalar multiple of x, i.e., TtT and I are locally linearly dependent. It is known that for linear operators of rank at least 2 local linear dependence implies (global) linear dependence (see, e.g., [23: page 1869]). Note that TtTHn+(R) . Therefore,

TtT=αI

for some α>0 . Let now Q=1αT . It follows that QtQ=1αTtT=I . So, Q is a linear isometry and since it is also invertible (and thus surjective), it is also a coisometry (QQt=I). For any P ∈ Pn(ℝ), we thus have Φ(P)=PQ(ImP) , where Q is an orthogonal operator,\i.e., it may be represented with an orthogonal matrix Q, where QQt=QtQ=I. Therefore for every P ∈ Pn(ℝ)

ImΦ(P)=Q(ImP)=QP(Rn)=QPQt(Rn)=ImQPQt.

Since clearly QPQt ∈ Pn(ℝ), we conclude that

Φ(P)=QPQt

for every P ∈ Pn(ℝ).

Without loss of generality, we assume from now on that

Φ(P)=P

for every P ∈ Pn(ℝ).

Step 8. Conclusion of the proof.Let AHn+(R) . Then we may write

A=i=1nλixixi,

where λi⩾0, i=1, 2, ⋅, n, are the eigenvalues of A and {x1, x2, ⋅, xn} is an orthonormal basis of ℝn. Then λixixi *A for every i=1, 2, ⋅, n and therefore

f(λi)xixi=f(λi)Φ(xixi)=Φ(λixixi)Φ(A),

where f:ℝ+ → ℝ+ is the function obtained in Step 5. By Corollary 12 and since Φ preserves rank, it follows that

Φ(A)=i=1nf(λi)Φ(xixi)=i=1nf(λi)xixi.

Recall Definition 1 to conclude that Φ(A)=f(A) for every AHn+(R) . Taking into account the assumptions from Steps 4 and 7, we may conclude that if Φ:Hn+(R)Hn+(R) is a surjective bi-preserver of the star partial order, there exist an orthogonal matrix Q ∈ Mn(ℝ), a positive λ, and a bijective function f:ℝ+↦ℝ+ with f (0)=0 such that

Φ(A)=λQf(A)Qt

for every AHn+(R) . □


The first, the second, and the third author acknowledge the financial support from the Slovene Research Agency, ARRS (research programs No. P1-0222, No. P1-0285, and No. P1-0288, and research project No. N1-0210). The fourth author is supported by the Ministry of Education, Science and Technological Development, Republic of Serbia, grant no. 451-03-68/2022-14/200124. The authors also acknowledge the bilateral project between Serbia and Slovenia (Generalized inverses, operator equations and applications) was financially supported by the Ministry of Education, Science and Technological Development, Republic of Serbia (grant no. 337-00-21/2020-09/32) and by the Slovene Research Agency, ARRS (grant BI-RS/20-21-039).


  1. (Communicated by Marek Balcerzak)

Acknowledgement

The authors wish to thank the anonymous reviewers for helpful and constructive comments that improved the presentation of this paper.

References

[1] Bahtia, R.: Positive Definite Matrices. Princeton Series in Applied Mathematics, Princeton University Press, Princeton, New Jersey, 2007.Search in Google Scholar

[2] Baksalary, J. K.—Mitra, S. K.: Left-star and right-star partial orderings, Linear Algebra Appl. 149 (1991), 73-89.Search in Google Scholar

[3] Baksalary, O. M.—Trenkler, G.: A partial ordering approach to characterize properties of a pair of orthogonal projectors, Indian J. Pure Appl. Math. 52 (2021), 323-334.Search in Google Scholar

[4] Chen, Y.—Dai, Y.—Han, D.—Sun, W.: Positive semidefinite generalized diffusion tensor imaging via quadratic semidefinite programming, SIAM J. Imaging Sci. 6(3) (2013), 1531-1552.Search in Google Scholar

[5] Christensen, R.: Plane Answers to Complex Questions: The Theory of Linear Models, Springer-Verlag, New York, 1996.Search in Google Scholar

[6] Dolinar, G.—Halicioglu, S.—Harmanci, A.—Kuzma, B.—Marovt, J.—Ungor, B.: Preservers of the left-star and right-star partial orders, Linear Algebra Appl. 587 (2020), 70-91.Search in Google Scholar

[7] Filmore, P. A.—Longstaff, W. E.: On isomorphisms of lattices of closed subspaces, Canad. J. Math. 5 (1984), 820-829.Search in Google Scholar

[8] Drazin, M. P.: Natural structures on semigroups with involution, Bull. Amer. Math. Soc. 84 (1978), 139-141.Search in Google Scholar

[9] Golubić, I.—Marovt, J.: Preservers of partial orders on the set of all variance-covariance matrices, Filomat 34(9) (2020), 3015-3030.Search in Google Scholar

[10] Golubić, I.—Marovt, J.: Monotone transformations on the cone of all positive semidefinite real matrices, Math. Slovaca 70(3) (2020), 733-744.Search in Google Scholar

[11] Guterman, A. E.: Linear preservers for matrix inequalities and partial orderings, Linear Algebra Appl. 331 (2001), 75-87.Search in Google Scholar

[12] Guterman, A. E.: Linear preservers for Drazin partial order, Comm. Algebra 29(9) (2001), 3905-3917.Search in Google Scholar

[13] Guterman, A. E.: Monotone additive transformations onmatrices, Mat. Zametki 81 (2007), 681-692.Search in Google Scholar

[14] Hartwig, R. E.—Drazin, M. P.: Lattice properties of the *-order for complex matrices, J. Math. Anal. Appl. 86 (1982), 359-378.Search in Google Scholar

[15] Hartwig, R. E.—Styan, G. P. H.: Partially ordered idempotent matrices. In: Proceedings of the Second International Tampere Conference in Statistics (T. Pukkila, S. Puntanen, eds.), Univ. of Tampere, Tampere, Finland, 1987, 361-383.Search in Google Scholar

[16] Holevo, A. S.: Statistical Structure of Quantum Theory. Lecture Notes in Phys., Springer-Verlag, Berlin, Heidelberg, New York, 2001.Search in Google Scholar

[17] de Klerk, E.: Aspects of Semidefinite Programming: Interior Point Algorithms and Selected Applications, Kluwer Academic Publishers, Dordrecht, 2002.Search in Google Scholar

[18] Legiša, P.: Automorphisms of Mn, partially ordered by the star order, Linear Multilinear Algebra 54(3) (2006), 157-188.Search in Google Scholar

[19] Mackey, G. W.: Isomorphisms of normed linear spaces, Ann. Math. 43 (1942), 244-260.Search in Google Scholar

[20] Marovt, J.—Rakić, D. S.—Djordjević, D. S.: Star, left-star, and right-star partial orders in Rickart *-rings, Linear Multilinear Algebra 63(2) (2015), 343-365.Search in Google Scholar

[21] Mitra, S. K.—Bhimasankaram, P.—Malik, S. B.: Matrix partial Orders, Shorted Operators and Applications, Word Scientific, London, 2010.Search in Google Scholar

[22] Mohri, M.—Rostamizadeh, A.—Talwalkar, A.: Foundations of Machine Learning, The MIT Press, Cambridge, Massachusetts, London, 2012.Search in Google Scholar

[23] Molnár, L.: Local automorphisms of operator algebras on Banach spaces, Proc. Amer. Math. Soc. 131(6) (2003), 1867-1874.Search in Google Scholar

Received: 2021-10-20
Accepted: 2022-02-01
Published Online: 2023-02-15
Published in Print: 2023-02-23

© 2023 Mathematical Institute Slovak Academy of Sciences

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. RNDr. Stanislav Jakubec, DrSc. passed away
  2. Schur m-power convexity for general geometric Bonferroni mean of multiple parameters and comparison inequalities between several means
  3. Conditions forcing the existence of relative complements in lattices and posets
  4. Radically principal MV-algebras
  5. A topological duality for dcpos
  6. Padovan or Perrin numbers that are concatenations of two distinct base b repdigits
  7. Addendum to “A generalization of a result on the sum of element orders of a finite group”
  8. Remarks on w-distances and metric-preserving functions
  9. Solution of logarithmic coefficients conjectures for some classes of convex functions
  10. Multiple periodic solutions of nonautonomous second-order differential systems with (q, p)-Laplacian and partially periodic potentials
  11. Asymptotic stability of nonlinear neutral delay integro-differential equations
  12. On Catalan ideal convergent sequence spaces via fuzzy norm
  13. Fourier transform inversion: Bounded variation, polynomial growth, Henstock–Stieltjes integration
  14. (ε, A)-approximate numerical radius orthogonality and numerical radius derivative
  15. Wg-Drazin-star operator and its dual
  16. On the Lupaş q-transform of unbounded functions
  17. K-contact and (k, μ)-contact metric as a generalized η-Ricci soliton
  18. Compactness with ideals
  19. Number of cells containing a given number of particles in a generalized allocation scheme
  20. A new generalization of Nadarajah-Haghighi distribution with application to cancer and COVID-19 deaths data
  21. Compressive sensing using extropy measures of ranked set sampling
  22. A note on star partial order preservers on the set of all variance-covariance matrices
Downloaded on 15.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ms-2023-0022/html
Scroll to top button