<<

arXiv:1604.00243v1 [math.RA] 1 Apr 2016 Let Introduction 1 h ae,w eoetesto all of set the denote we paper, the arxwt h prpit ie For size. appropriate the with e ( M Let A inmatrices. nion ojgt rnps Hriinajit arxadtern of rank the and matrix adjoint) (Hermitian conjugate by matrix iey egvn For given. be tively, h olwn qain in equations following the A isrgc nttt o ple rbeso ehnc n Mathe and Mechanics of Problems Applied for Institute Pidstrygach ( = steuiu solution unique the is H h entoso h eeaie nes arcscnb extended be can matrices inverse generalized the of definitions The h or-ers nes of inverse Moore-Penrose The e emta oiiedfiiematrices definite positive Hermitian Let R m ouin flf n ih ytm fqaeno ierequ linear quaternion Moor of weighted systems the right for and rules n left Cramer the of the of solutions get an theory we obtained the the matrix, using of adjoint By framework . the column-row within mutative obtained qua a been of inverse have Moore-Penrose det weighted and the limit of matri representation, resentations this quaternion Using a derived. of been inverse have Moore-Penrose weighted the of a eemnna ersnain fteqaeno weighted quaternion the of representations Determinantal or-ers nes n orsodn rmrsrule. Cramer’s corresponding and inverse Moore-Penrose and × H ij X n, n ) egtdsnua au eopsto WV)adareprese a and (WSVD) decomposition value singular Weighted = n by and , ∈ H ∈ C { etern of ring the be ) H H a etera n ope ubrfils epciey Throughout respectively. fields, number complex and real the be (3 n 0 n × × + M m n A fUrie kan,[email protected] Ukraine, Ukraine, of NAS a ( ) sHriinif Hermitian is H aifigtefloigeutos[1], equations following the satisfying 1 i m r MAX A + × a n ∈ 2 X h e fall of set the j H X ) + m ∗ = n [2]: a × = 3 × A n vnKyrchei Ivan k ( ( MAX , AX XA M,N + n AXA XAX | A h egtdMoePnoeinverse Moore-Penrose weighted the i Abstract A m 2 utrinmtie and matrices quaternion ∗ ) ) = × = A ∗ ∗ ∈ ftemti qain 1 n 2 and (2) and (1) equations matrix the of 1 (4 ; j = = n m A = = ∈ H 2 arcsoe h utrinalgebra quaternion the over matrices AX XA . = m × H A X N M × n (1) ; (2) ; k ( ) n n × 2 (3) ; . eoe by denoted , and m arcsover matrices NXA = ednt by denote we , − N 1 a , ) forder of ∗ 0 = a , NXA 1 riatlrep- erminantal eno matrix ternion a , ations. A H I yWSVD by x 2 m lg fthe of alogs † a , eteidentity the be steunique the is , A A iharank a with e-Penrose . and 3 ∗ h matrix The . oncom- ntation rank , ∈ oquater- to R aisof matics n respec- , } A the (4) of r . In particular, when M = Im and N = In, the matrix X satisfying the equations (1), (2), (3M), (4N) is the Moore-Penrose inverse A†. It is known various representations of the weighted Moore-Penrose. In par- ticular, limit representations have been considered in [3,4]. Determinantal rep- resentations of the complex (real) weighted Moore-Penrose have been derived by full-rank factorization in [5], by limit representation in [6] using the method first introduced in [7], and by minors in [8]. A basic method for finding the Moore-Penrose inverse is based on the singular value decomposition (SVD). It is available for quaternion matrices, (see, e.g. [9, 10]). In [10, 11], using SVD of quaternion matrices, the limit and determinantal representations of the Moore- Penrose inverse over the quaternion skew field have been obtained within the framework of the theory of the noncommutative column-row determinants that have been introduced in [12]. † Cm×n The weighted Moore-Penrose inverse AM,N ∈ can be explicitly ex- pressed by the weighted singular value decomposition (WSVD) which at first has been obtained in [13] by Cholesky factorization. In [14] WSVD of real matri- ces with singular weights has been derived using weighted orthogonal matrices and weighted pseudoorthogonal matrices. Song at al. [15, 16] have studied the weighted Moore-Penrose inverse over the quaternion skew field and obtained its determinantal representation within the framework of the theory of the column-row determinants. But WSVD of quaternion matrices has not been considered and for obtaining a determinantal representation there was used auxiliary matrices which different from A, and weights M and N. The main goals of the paper are introducing WSVD of quaternion matrices and representation of the weighted Moore-Penrose inverse over the quaternion skew field by WSVD, and then by using this representation, obtaining its limit and determinantal representations. In this paper we shall adopt the following notation. Let α := {α1,...,αk}⊆{1,...,m} and β := {β1,...,βk}⊆{1,...,n} α be subsets of the order 1 ≤ k ≤ min {m,n}. By Aβ denote the subma- trix of A determined by the rows indexed by α, and the columns indexed α by β. Then, A α denotes the principal submatrix determined by the rows H α and columns indexed by α. If A ∈ M (n, ) is Hermitian, then by |Aα| de- note the corresponding principal of det A. For 1 ≤ k ≤ n, denote by Lk,n := { α : α = (α1,...,αk) , 1 ≤ α1 ≤ ... ≤ αk ≤ n} the collection of strictly increasing sequences of k integers chosen from {1,...,n}. For fixed i ∈ α and j ∈ β, let

Ir, m{i} := { α : α ∈ Lr,m,i ∈ α}, Jr, n{j} := { β : β ∈ Lr,n, j ∈ β}.

The paper is organized as follows. We start with some basic concepts and results from the theory of the row-column determinants and of Hermitian quater- nion matrices in Section 2. Weighted singular value decomposition and a rep- resentation of the weighted Moore-Penrose inverse of quaternion matrices by WSVD have been considered in Subsection 3.1, and its limit representations in

2 Subsection 3.2. In Section 4, we give the determinantal representations of the weighted Moore-Penrose inverse. In Subsection 4.1, if the matrices N−1A∗MA and AN−1A∗M are Hermitian, and if they are non-Hermitian in Subsection 4.2. In Section 5 we obtain explicit representation formulas of the weighted Moore-Penrose solutions (analogs of Cramer’s rule) of the left and right sys- tems of linear equations over the quaternion skew field. In Section 5, we give a numerical example to illustrate the main result.

2 Preliminaries

For a quadratic matrix A = (aij ) ∈ M (n, H) can be define n row determinants and n column determinants as follows. Suppose Sn is the symmetric group on the set In = {1,...,n}.

Definition 2.1 [11] The i-th row of A = (aij ) ∈ M (n, H) is defined for all i = 1,n by putting rdet A = (−1)n−r a a . . .a ...a ...a , i iik1 ik1 ik1 +1 ik1+l1 i ikr ikr +1 ikr +lr ikr σ S X∈ n σ = (iik1 ik1+1 ...ik1+l1 ) (ik2 ik2+1 ...ik2+l2 ) ... (ikr ikr +1 ...ikr+lr ) , with conditions ik2

Definition 2.2 [11] The j-th column determinant of A = (aij ) ∈ M (n, H) is defined for all j = 1,n by putting cdet A = (−1)n−r a ...a . . .a ...a a , j jkr jkr +lr jkr +1ikr j jk1+l1 jk1+1jk1 jk1 j τ S X∈ n τ = (jkr +lr ...jkr +1jkr ) ... (jk2+l2 ...jk2+1jk2 ) (jk1+l1 ...jk1+1jk1 j) , with conditions, jk2 < jk3 <...

Proposition 2.1 [11] If b ∈ H, then rdetiAi. (b · ai.)= b·rdetiA and cdeti A.i (a.ib)= cdeti Ab for all i = 1,n.

Proposition 2.2 [11] If for A ∈ M (n, H) there exists t ∈ In such that atj = bj + cj for all j = 1,n, then

rdeti A = rdeti At. (b) + rdeti At. (c) , cdeti A = cdeti At. (b) + cdeti At. (c) , where b = (b1,...,bn), c = (c1,...,cn) and for all i = 1,n.

3 Proposition 2.3 [11] If for A ∈ M (n, H) there exists t ∈ Jn such that ai t = bi + ci for all i = 1,n, then

rdetj A = rdetj A . t (b) + rdetj A . t (c) , cdetj A = cdetj A . t (b) + cdetj A . t (c) ,

T T where b = (b1,...,bn) , c = (c1,...,cn) and for all j = 1,n.

n n Remark 2.1 Let rdeti A = aij · Rij and cdetj A = Lij · aij for all j=1 i=1 i, j = 1,n, where by Rij and LPij denote the right and left ij-thP cofactor of A ∈ M (n, H), respectively. It means that rdeti A can be expand by right cofactors along the i-th row and cdetj A can be expand by left cofactors along the j-th column, respectively, for all i, j = 1,n.

The following theorem has a key value in the theory of the column and row determinants.

Theorem 2.1 [11] If A = (aij ) ∈ M (n, H) is Hermitian, then rdet1A = ··· = rdetnA = cdet1A = ··· = cdetnA ∈ R. Since all column and row determinants of a over H are equal, we can define the determinant of a Hermitian matrix A ∈ M (n, H). By defini- tion, we put det A := rdeti A = cdeti A, (5) for all i = 1,n. By using its row and column determinants the determinant of a quaternion Hermitian matrix has properties similar to a usual determinant of a real matrix. These properties are completely explored in [11] and can be summarized in the following theorems. Theorem 2.2 If the i-th row of a Hermitian matrix A ∈ M (n, H) is replaced with a left linear combination of its other rows, i.e. ai. = c1ai1. + ... + ckaik ., where cl ∈ H for all l = 1, k and {i,il}⊂ In, then

rdeti Ai. (c1ai1. + ... + ckaik .) = cdeti Ai. (c1ai1 . + ... + ckaik .)=0.

Theorem 2.3 If the j-th column of a Hermitian matrix A ∈ M (n, H) is re- placed with a right linear combination of its other columns, i.e. a.j = a.j1 c1 + H ... + a.jk ck, where cl ∈ for all l = 1, k and {j, jl}⊂ Jn, then

cdetj A.j (a.j1 c1 + ... + a.jk ck) = rdetj A.j (a.j1 c1 + ... + a.jk ck)=0.

Theorem 2.4 [11] If an arbitrary column of A ∈ Hm×n is a right linear combination of its other columns, or an arbitrary row of A∗ is a left linear combination of its others, then det A∗A =0.

Moreover, we have the criterion of nonsingularity of a Hermitian matrix.

4 Theorem 2.5 [11] The right-linearly independence of columns of A ∈ Hm×n or the left-linearly independence of rows of A∗ is the necessary and sufficient condition for det A∗A 6=0. The following theorem about determinantal representation of an inverse matrix of Hermitian follows immediately from these properties. Theorem 2.6 [11] If a Hermitian matrix A ∈ M (n, H) is such that det A 6=0, then there exist a unique right inverse matrix (RA)−1 and a unique left inverse matrix (LA)−1, and (RA)−1 = (LA)−1 =: A−1, which possess the following determinantal representations:

R11 R21 ··· Rn1 −1 1 R R ··· Rn (RA) =  12 22 2  , (6) det A ············ R1n R2n ··· Rnn     L11 L21 ··· Ln1 −1 1 L L ··· Ln (LA) =  12 22 2 . (7) det A ············ L n L n ··· Lnn  1 2    Here Rij , Lij are right and left ij-th cofactors of A respectively for all i, j = 1,n. Due to the noncommutativity of quaternions, there are two types of eigen- values. Definition 2.3 Let A ∈ M (n, H). A quaternion λ is said to be a right eigen- value of A if A·x = x·λ for some nonzero quaternion column-vector x. Similarly λ is a left eigenvalue if A · x = λ · x. The theory on the left eigenvalues of quaternion matrices has been investigated, in particular, in [17–19]. The theory on the right eigenvalues of quaternion matrices is more developed, for further details one may refer to [9,20,21]. From this theory we cite the following propositions. Proposition 2.4 [9] Let A ∈ M (n, H) is Hermitian. Then A has exactly n real right eigenvalues. Definition 2.4 Suppose U ∈ M (n, H) and U∗U = UU∗ = I, then the matrix U is called unitary. Proposition 2.5 [9] Let A ∈ M (n, H) be given. Then, A is Hermitian if and only if there are a U ∈ M (n, H) and a real ∗ D = diag (λ 1, λ 2,...,λ n) such that A = UDU , where λ1, ..., λn are right eigenvalues of A. Suppose A ∈ M (n, H) is Hermitian and λ ∈ R is its right eigenvalue, then A·x = x·λ = λ·x. This means that all right eigenvalues of a Hermitian matrix are its left eigenvalues as well. For real left eigenvalues, λ ∈ R, the matrix λI − A is Hermitian.

5 Definition 2.5 If t ∈ R, then for a Hermitian matrix A the polynomial pA (t)= det (tI − A) is said to be the characteristic polynomial of A. The roots of the characteristic polynomial of a Hermitian matrix are its real left eigenvalues, which are its right eigenvalues as well. Definition 2.6 Let A ∈ Hn×n be a Hermitian matrix and π(A) = π be the number of positive eigenvalues of A, ν(A)= ν be the number of negative eigen- values of A, and δ(A) = δ be the number of zero eigenvalues of A. Then the ordered triple ω = (π,ν,δ) will be called the inertia of A. We shall write ω = In A. Definition 2.7 A Hermitian matrix A ∈ Hn×n, is called positive (semi)definite if x∗Ax > 0(≥ 0) for any nonzero vector x ∈ Hn. The following properties are equivalent to A being positive definite and they can been expanded obviously to quaternion matrices. Proposition 2.6 All its eigenvalues are positive. Proposition 2.7 Its leading principal minors are all positive. Since all leading principal submatrices of a Hermitian matrix are Hermitian, then we can define leading principal minors as determinants of Hermitian sub- matrices in terms of Eq.(5). Every positive definite matrix A ∈ Hn×n has a unique square root defined 1 ∗ 1 1 ∗ by A 2 . It means, if A = UDU then A 2 = UD 2 U . We have [9,10] the following theorem about the singular value decomposition (SVD) of quaternion matrices. Hm×n Theorem 2.7 (SVD) Let A ∈ r . There exist unitary quaternion matrices D 0 V ∈ Hm×m and W ∈ Hn×n such that A = VΣW∗, where Σ = ∈ 0 0 Hm×n  2  r , and Dr = diag (σ1, σ2,...,σr) , σ1 ≥ σ2 ≥ ... ≥ σr > 0, and σi is the nonzero eigenvalues of A∗A for all i = 1, ..., r. Then A† = WΣ†V∗, where D−1 0 Σ = . 0 0   In [11], using the singular value decomposition of quaternion matrices, the limit and determinantal representations of the Moore-Penrose inverse over the quater- nion skew field have been obtained as follows. Lemma 2.1 [11] If A ∈ Hm×n and A† is its Moore-Penrose inverse, then + ∗ ∗ −1 ∗ −1 ∗ A = lim A (AA + αI) = lim (A A + αI) A , where α ∈ R+. α→0 α→0 Hm×n † Theorem 2.8 [11] If A ∈ r , then the Moore-Penrose inverse A = † Hn×m aij ∈ possess the following determinantal representations:

  ∗ ∗ β cdeti (A A) . i a.j β † β∈Jr, n{i} aij =  , (8) P ∗ β (A A) β β∈J r, n P

6 or ∗ ∗ α rdetj ((AA )j. (ai. )) α α∈I {j} a† r,m . ij = P ∗ α (9) |(AA ) α| α∈Ir, m P Hm×n ∗ ∗ Lemma 2.2 [11] Let A ∈ r , then A A and AA are both positive (semi)definite, and r nonzero eigenvalues of A∗A and AA∗ coincide. Proof. The proof of the second part immediately follows from the singular Hm×n  value decomposition of A ∈ r . n n−1 Lemma 2.3 [11] If A ∈ M (n, H) is Hermitian, then pA (t) = t − d1t + n−2 n d2t − ... + (−1) dn, where dk is the sum of principle minors of A of order k, 1 ≤ k

3 Weighted singular value decomposition and representations of the weighted Moore-Penrose inverse of quaternion matrices 3.1 Representations of the weighted Moore-Penrose in- verse of quaternion matrices by WSVD Denote A♯ = N−1A∗M. Now, we prove the following theorem about the weighted singular value decomposition (WSVD) of quaternion matrices. We give the proof of the theorem that different from analogous for real matrices in [13], and this method has more similar manner to [14], where WSVD of A ∈ Rm×n with positive definite weights B and C has been described as A = UDVT C. Hm×n Theorem 3.1 Let A ∈ r , and M and N be positive definite matrices of order m and n, respectively. Then there exist U ∈ Hm×m, V ∈ Hn×n satisfying ∗ ∗ −1 U MU = Im and V N V = In such that A = UDV∗, (10)

7 Σ 0 where D = , Σ = diag(σ , σ , ..., σ ), σ ≥ σ ≥ ... ≥ σ > 0 and σ2 0 0 1 2 r 1 2 r i   is the nonzero eigenvalues of A♯A or AA♯, which coincide.

♯ −1 ∗ ∗ 1 ∗ 1 Proof. First, consider A A = N A MA. Since A MA = (M 2 A) AM 2 , then, by Lemma 2.2, A∗MA is Hermitian positive semidefinite, and by Lemma ♯ 2 2.4 all eigenvalues of A A are nonnegative. Denote them by σi , where σ1 ≥ ... ≥ σn ≥ 0. Denote L = A♯A. Since LN−1 = N−1A∗MAN−1 is Hermitian and there n×n ∗ −1 exists a nonsingular V ∈ H such that V N V = In, then by Lemma 2.5, V∗LN−1V = Λ, (11) where V is unitary with weight N−1, and Λ is a diagonal matrix. −1 ∗ ∗ −1 ∗ 2 2 It follows from L = N VΛV = (V ) ΛV that Λ ≡ Σ1, where Σ1 is ♯ 2 2 diagonal with eigenvalues of A A on the principal diagonal, σii = σi for all i = 1, ..., n. Since rank L = rank A = r, then the number of nonzero diagonal 2 elements of Σ1 is equal r. Also, we note that

V∗LN−1V = V∗N−1A∗MAN−1V = −1 ∗ ∗ −1 −1 ∗ −1 2 V A (U ) U A (V ) = Σ1. (12) Consider the following matrix,

1 −1 m×n P = M 2 AN V ∈ H . (13)

By virtue of (11),

1 1 ∗ ∗ −1 ∗ 2 2 −1 2 P P = V N A M M AN V = Σ1. (14)   Let us introduce the following m × n matrix D ∈ Hm×n, Σ 0 D = , (15) 0 0   r×r where Σ ∈ H is a diagonal matrix with σ1 ≥ σ2 ≥ ... ≥ σr > 0 on the principal diagonal. Then, 1 P = M 2 UD. (16) 1 −1 1 By (13) and (16), we have M 2 AN V = M 2 UD. Due to the equality −1 N−1V = V∗, it follows (10). 2 ♯ Now we shall prove (10), where σi is the nonzero eigenvalues of AA = AN−1A∗M. Since AN−1A∗ and M are respectively Hermitian positive semidef- inite and definite, then by by Lemma 2.4 all eigenvalues of AA♯ are nonneg- 2 ative. Primarily, denote them by τi , where τ1 ≥ ... ≥ τm ≥ 0, and denote Q = AA♯. Since MQ = MAN−1A∗M is Hermitian and there exists a nonsin- m×m ∗ gular U ∈ H such that U MU = Im, then by Lemma 2.5, U∗MQU = Ω, (17)

8 where U is unitary with weight M, and Ω is a diagonal matrix. ∗ −1 2 2 It follows from Q = UΩU M = UΩU that Ω ≡ Σ2, where Σ2 is diagonal ♯ 2 2 with eigenvalues of AA on the principal diagonal, τii = τi for all i =1, ..., m. Since rank Q = rank A = r, then the number of nonzero diagonal elements of 2 Σ2 is equal r. Also, we have

U∗MQU = U∗MAN−1A∗MU = −1 ∗ −1 −1 ∗ ∗ −1 2 U A (V ) V A (U ) = Σ2. (18)

Comparing (12) and (18), and due to Lemma 2.2, we have that r nonzero ♯ ♯ 2 2 eigenvalues of AA coincide with r nonzero eigenvalues of A A, i.e. σi = τi for all i =1, ..., r. Consider the following matrix,

∗ − 1 m×n S = U MAN 2 ∈ H . (19)

By virtue of (17),

1 1 ∗ ∗ − 2 − 2 ∗ 2 SS = U MAN N A MU = Σ2. (20)   Consider again the matrix D ∈ Hm×n from (15). Then,

∗ − 1 S = DV N 2 . (21)

∗ − 1 ∗ − 1 By (19) and (21), we have U MAN 2 = DV N 2 . From this, due to (U∗M)−1 = U, we again obtain (10). † Now, we prove the following theorem about a representation of AM,N by Hm×n WSVD of A ∈ r with weights M and N. Hm×n Theorem 3.2 Let A ∈ r , M and N be positive definite matrices of order m and n, respectively. There exist U ∈ Hm×m, V ∈ Hn×n satisfying U∗MU = Σ 0 I and V∗N−1V = I such that A = UDV∗, where D = . Then m n 0 0 †   the weighted Moore-Penrose inverse AM,N can be represented

Σ−1 0 A† = N−1V U∗M, (22) M,N 0 0   2 where Σ = diag(σ1, σ2, ..., σr), σ1 ≥ σ2 ≥ ... ≥ σr > 0 and σi is the nonzero eigenvalues of A♯A or AA♯, which coincide.

† Proof. To prove the theorem we shall show that X = AM,N expressed by (22) satisfies the equations (1), (2), (3N), and (4M).

Σ 0 Σ−1 0 Σ 0 1) AXA = U V∗N−1V U∗MU V∗ = A, 0 0 0 0 0 0      

9 Σ−1 0 Σ 0 2) XAX = N−1V U∗MU V∗× 0 0 0 0     Σ−1 0 N−1V U∗M = X, 0 0  

Σ 0 Σ−1 0 ∗ (3M) (MAX)∗ = MU V∗N−1V U∗M = 0 0 0 0       I 0 Σ 0 Σ−1 0 MU U∗M = MU V∗N−1V U∗M = 0 0 0 0 0 0  m×m     MAX,

Σ−1 0 Σ 0 ∗ (4N) (NXA)∗ = NN−1V U∗MU V∗ = 0 0 0 0       I 0 Σ−1 0 Σ 0 NN−1V V∗ = NN−1V U∗MU V∗ = 0 0 0 0 0 0  n×n     NXA. 

3.2 Limit representations of the weighted Moore-Penrose inverse over the quaternion skew field Due to [3] the following limit representation can be extended to H. We give the proof of the following lemma that different from ( [3], Corollary 3.4.) and based on WSVD. Hm×n Lemma 3.1 Let A ∈ r , and M and N be positive definite matrices of order m and n, respectively. Then

† ♯ −1 ♯ AM,N = lim (λI + A A) A . (23) λ→0

♯ −1 ∗ where A = N A M, λ ∈ R+ and R+ is the set of all positive real numbers. Proof. By Theorems 3.1 and 3.2, respectively, we have

Σ 0 Σ−1 0 A = U V∗, A† = N−1V U∗M, 0 0 M,N 0 0     2 R where Σ = diag(σ1, σ2, ..., σr), σ1 ≥ σ2 ≥ ... ≥ σr > 0 and σi ∈ is the nonzero eigenvalues of N−1A∗MA. Consider the matrix

Σ 0 D := , 0 0  

10 Hm×n where D = (σij ) ∈ r is such that σ11 ≥ σ22 ≥ ... ≥ σrr > σr+1 r+1 = ... = σqq = 0, q = min {n,m}. Then Σ∗ 0 Σ−1 0 D∗ = , D+ = , 0 0 0 0     ∗ ♯ −1 ∗ ∗ † −1 † ∗ and A = UDV , A = N VD U M, AM,N = N VD U M. Since N−1V = (V∗)−1, then we have

λI + A♯A = λI + N−1VD∗U∗MUDV∗ = λI + (V∗)−1D∗DV∗ = (V∗)−1(λI + D2)V∗. Further,

(λI + A♯A)−1A♯ = (V∗)−1(λI + D2)−1V∗N−1VD∗U∗M = N−1V(λI + D2)−1D∗U∗M. Consider the matrix

σ1 2 ... 0 σ1 +λ ...... 0   −1 . λI D2 D ... σr . . + =  0 σ2+λ .   r   . .    . ..     0 0     −1  It is obviously that lim λI + D2 D = D†. Then, λ→0 −1 2 −1  ∗ ∗ −1 † ∗ † lim N V(λI + D ) D U M = N VD U M = AM,N . λ→0 The lemma is proofed. + In the following lemma we give another limit representation of AM,N . Hm×n Lemma 3.2 Let A ∈ r , and M and N be positive definite matrices of order m and n, respectively. Then + ♯ ♯ −1 AM,N = lim A (λI + AA ) , (24) λ→0 ♯ −1 ∗ where A = N A M, λ ∈ R+. Proof. The proof is similar to the proof of Lemma 3.1 by using the fact from Theorem 3.1 that the nonzero eigenvalues of A♯A and AA♯ coincide. It is evidently the following corollary. Corollary 3.1 If A ∈ Hm×n, then the following statements are true. + ♯ −1 ♯ i) If rank A = n, then AM,N = A A A .

+ ♯  ♯ −1 ii) If rank A = m, then AM,N = A AA . + −1 iii) If rank A =n=m, then AM,N =A .

11 4 Determinantal representations of the weighted Moore-Penrose inverse over the quaternion skew field

Even though the eigenvalues of A♯A and AA♯ are real and nonnegative, they are not Hermitian in general. Therefor, we consider two cases, when A♯A and AA♯ both or one of them are Hermitian, and when they are non-Hermitian.

4.1 The case of Hermitian A♯A and AA♯. Let (A♯A) ∈ Hn×n be Hermitian. It means that (A♯A)∗ = (A♯A). Since N−1 and M are Hermitian, then

(N−1A∗MA)∗ = A∗MAN−1 = N−1A∗MA.

So, to the matrix (A♯A) be Hermitian the matrices N−1 and (A∗MA) should be commutative. Similarly, to (AA♯) be Hermitian the matrices M and (AN−1A∗) should be commutative. ♯ ♯ ♯ Denote by a.j and ai. the j-th column and the i-th row of A respectively.

Hm×n ♯ ♯ Lemma 4.1 If A ∈ r , then rank A A . i a.j ≤ r.    ♯ ♯ Proof. Let’s lead elementary transformations of the matrix A A . i a.j   right-multiplying it by elementary unimodular matrices Pi k −ajk , k 6= j. The   matrix P i k −ajk has −aj k in the (i, k) entry, 1 in all diagonal entries, and 0 in others. It is the matrix of an elementary transformation. Right-multiplying a matrix by P i k −ajk , where k 6= j, means adding to k-th column its i-th column right-multiplying  on −ajk. Then we get

♯ ♯ ♯ a1kak1 ... a1j ... a1kakn k6=j k6=j ♯ ♯  ......  A A . i a. j · Pi k (−aj k)= P P . ♯ ♯ ♯ k6=i a a ... a ... a a   Y  nk k1 nj nk kn    k6=j k6=j     P i−th P  The obtained matrix has the following factorization.

♯ ♯ ♯ a1kak1 ... a1j ... a1kakn k6=j k6=j  ...... P P  = ♯ ♯ ♯  ankak1 ... anj ... ankakn   k6=j k6=j     P i−th P 

12 ♯ ♯ ♯ a ...... a a a ... a 11 0 n1 11 12 1m ...... a♯ a♯ ... a♯   =  21 22 2m  0 ... 1 ... 0 j − th......  ......   a♯ a♯ ... a♯     n1 n2 nm   am1 ... 0 ... amn       i−th 

a11 ... 0 ... an1 ......   Denote by A˜ := 0 ... 1 ... 0 j − th. The matrix A˜ is ob-  ......     am1 ... 0 ... amn     i−th  tained from A by replacing all entries of the j-th row and of the i-th col- umn with zeroes except that the (j, i) entry equals 1. Elementary transfor- mations of a matrix do not change its rank and the rank of a matrix product A♯A a♯ does not exceed a rank of each factors. It follows that rank . i .j ≤ min rank A♯, rank A˜ . It is obviously that rank A˜ ≥ rank A= rank A♯. This completesn the proof.o The following lemma has been proved in the same way.

Hm×n ♯ ♯ Lemma 4.2 If A ∈ r , then rank AA . i a.j ≤ r.   Analogues of the characteristic polynomial are considered in the following lem- mas. Lemma 4.3 If A ∈ Hm×n, t ∈ R, and (A♯A) is Hermitian, then

♯ ♯ (ij) n−1 (ij) n−2 (ij) cdeti tI + A A . i a.j = c1 t + c2 t + ... + cn , (25)   (ij) ♯  ♯ (ij) ♯ ♯ β where cn = cdeti A A . i a. j and ck = cdeti A A . i a. j β β∈J {i}   k,n    for all k = 1,n − 1, i = 1,n, and j = 1,m. P 

♯ Proof. Denote by b. i the i-th column of the Hermitian matrix A A =: ♯ Hn×n (bij )n×n. Consider the Hermitian matrix tI + A A . i (b. i) ∈ . It differs from tI + A♯A an entry b . Taking into account Lemma 2.3 we obtain ii   ♯ n−1 n−2 det tI + A A . i (b. i)= d1t + d2t + ... + dn, (26)

♯ β where dk = det A A β is the sum of all principal minors of order k β∈Jk,n{i} ♯ that contain thePi-th column for all k = 1,n − 1 and dn = det A A . Therefore, 

13 we have ♯ a1lali l  ♯  P a2lali l ♯ b. i =   = a.lali,  .P   .  l  ♯  X  a ali   nl   l  ♯  P ♯  where a.l is the lth column-vector of A for all l = 1,m. Taking into account Theorem 2.1, Remark 2.1 and Proposition 2.1 we obtain on the one hand

♯ ♯ det tI + A A . i (b. i) = cdeti tI + A A . i (b. i)= tI A♯A a♯ a  tI A♯A a♯ a = cdeti + .l .l l i = cdeti + . i .l · li (27) l l X    X    On the other hand having changed the order of summation, we get for all k = 1,n − 1

♯ β ♯ β dk = det A A β = cdeti A A β = β∈J {i} β∈J {i} Xk,n  Xk,n  ♯ ♯ β cdeti A A . i a.lal i β = β∈J {i} l Xk,n X     ♯ ♯ β cdeti A A . i a.l β · al i. (28) l β∈J {i} X Xk, n     By substituting (27) and (28) in (26), and equating factors at al i when l = j, we obtain the equality (25). By analogy can be proved the following lemma. Lemma 4.4 If A ∈ Hm×n and t ∈ R, and AA♯ is Hermitian, then

♯ ♯ (ij) n−1 (ij) n−2 (ij) rdetj (tI + AA )j. (ai.)= r1 t + r2 t + ... + rn ,

(ij) ♯ ♯ (ij) ♯ ♯ α where rn = rdetj (AA )j. (ai. ) and rk = rdetj (AA )j. (ai. ) α α∈I {j} r,m   for all k = 1,n − 1, i = 1,n, and j = 1,m. P The following theorem introduce the determinantal representations of the weighted Moore-Penrose by analogs of the classical adjoint matrix. Hm×n ♯ ♯ Theorem 4.1 Let A ∈ r . If A A or AA are Hermitian, then the † † Hn×m weighted Moore-Penrose inverse AM,N = a˜ij ∈ possess the following determinantal representations, respectively, 

♯ ♯ β cdeti A A . i a.j β † β∈Jr, n{i} a˜ij =     , (29) P ♯ β (A A) β β∈J r, n P

14 or ♯ ♯ α rdetj (AA )j. (ai. ) α α∈I {j} a† r,m . ˜ij = P  ♯ α  (30) |(AA ) α| α∈Ir, m P −1 Proof. At first we prove (29). By Lemma 3.1, A† = lim αI + A♯A A♯. Let α→0 ♯ ♯ n×n A A is Hermitian. Then matrix αI + A A ∈ H is a full-rank Hermitian matrix. Taking into account Theorem 2.6 it has an inverse, which we represent  as a left inverse matrix

L11 L21 ...Ln1 −1 1 L L ...Ln αI + A♯A =  12 22 2  , det (αI + A∗A) ......   L1n L2n ...Lnn    ♯  where Lij is a left ij-th cofactor of αI + A A. Then we have

−1 αI + A♯A A♯ = n n n ♯ ♯ ♯  Lk1ak1 Lk1ak2 ... Lk1akm k=1 k=1 k=1  n n n  P L a♯ P L a♯ ... P L a♯ 1 k2 k1 k2 k2 k2 km . = det(αI+A♯A)  k=1 k=1 k=1     ...... P P P   n n n   L a♯ L a♯ ... L a♯   kn k1 kn k2 kn km   k=1 k=1 k=1   P P P  Using the definition of a left cofactor, we obtain

♯ ♯ ♯ ♯ cdet1 αI+A A a cdet1 αI+A A a ( ).1( .1) ( ).1( .m) det(αI+A♯A) ... det(αI+A♯A) † AM,N = lim  ......  . (31) α→0 ♯ ♯ ♯ ♯ cdetn αI+A A a cdetn αI+A A a ( ).n( .1) ( ).n( .m)  I A♯A ... I A♯A   det(α + ) det(α + )    ♯ n n−1 n−2 By Lemma 2.3, we have det αI + A A = α + d1α + d2α + ... + dn, ♯ β ♯ where dk = A A β is a sum of principal minors of A A of order k β∈J k, n for all k = 1,nP− 1 and d = det A♯A. Since rank A♯A = rank A = r and n dn = dn−1 = ... = dr+1 = 0, it follows that ♯ n n−1 n−2 n−r det αI + A A = α + d1α + d2α + ... + drα . By using (25), we have 

∗ ♯ (ij) n−1 (ij) n−2 (ij) cdeti (αI + A A).i a.j = c1 α + c2 α + ... + cn   (ij) ♯ ♯ β for all i = 1,n and j = 1,m, where ck = cdeti (A A). i a.j β for β∈J {i} k,n    (ij) ♯ ♯ P all k = 1,n − 1 and cn = cdeti A A .i a.j .    15 (ij) Now we prove that ck = 0, when k ≥ r + 1 for all i = 1,n, and j = 1,m. ♯ ♯ ♯ ♯ By Lemma 4.1 rank A A . i a.j ≤ r, then the matrix A A . i a.j has no r more right-linearly independent   columns.    ♯ ♯ β Consider (A A) . i a.j β, when β ∈ Jk,n{i}. It is a principal submatrix ♯ ♯   of A A . i a.j of order k ≥ r + 1. Deleting both its i-th row and column, we ♯ obtain a principal  submatrix of order k − 1 of A A. We denote it by M. The following cases are possible. Let k = r + 1 and det M 6= 0. In this case all columns of M are right- linearly independent. The addition of all of them on one coordinate to columns ♯ ♯ β of A A . i a.j β keeps their right-. Hence, they are    A♯A a♯ β i basis in a matrix . i .j β, and the -th column is the right lin- ear combination of its basic  columns. From this by Theorem 2.4, we get ♯ ♯ β cdeti A A . i a.j β = 0, when β ∈ Jk,n{i} and k ≥ r + 1. If k= r + 1 and  det M = 0, than p, (p < k), columns are basis in M and in A♯A a♯ β A♯A a♯ β . i .j β. Then due to Theorem 2.4, we obtain cdeti . i .j β = 0 as well.       If k > r+1, then by Theorem 2.5 it follows that det M = 0 and p, (p < k−1), M A♯A a♯ β columns are basis in the both matrices and . i .j β. Therefore, ♯ ♯  β   by Theorem 2.4, we obtain cdeti A A . i a.j β = 0.  ♯  ♯ β Thus in all cases, we have cdeti AA . i a.j β = 0, when β ∈ Jk,n{i} and r +1 ≤ k

(ij) ♯ ♯ β ck = cdeti A A . i a.j β =0, β∈J {i} Xk,n     (ij) ♯ ♯ cn = cdeti A A . i a.j =0.   ♯ ♯ (ij) n−1 (ij) n−2  (ij) n−r Hence, cdeti αI + A A . i a. j = c1 α + c2 α + ... + cr α for all i = 1,n and j = 1,m. By substituting these values in the matrix from (31), we obtain

(11) n−1 (11) n−r (1m) n−1 (1m) n−r c1 α +...+cr α c1 α +...+cr α n n−1 n−r ... n n−1 n−r α +d1α +...+drα α +d1α +...+drα + ...... AM,N = lim   = α→0 (n1) n−1 (n1) n−r (nm) n−1 (nm) n−r c1 α +...+cr α c1 α +...+cr α n n−1 n−r ... n n−1 n−r  α +d1α +...+drα α +d1α +...+drα  c(11) c(1m)  r ... r  dr dr  ......  . c(n1) c(nm) r ... r  dr dr    (ij) ♯ ♯ β ♯ β Here cr = cdeti A A . i a.j β and dr = A A β . β∈J {i} β∈Jr, n r, n    P  P 

16 + Thus, we have obtained the determinantal representation of AM,N by (29). +  Similarly can be proved the determinantal representation of AM,N by (30).

Remark 4.1 If rank A = n, and (A♯A) is Hermitian, then by Corollary 3.1, + ♯ −1 ♯ ♯ −1 AM,N = A A A . Considering A A as a left inverse, we get the following representation of A+ ,  M,N 

♯ ♯ ♯ ♯ cdet1(A A). 1 a. 1 ... cdet1(A A). 1 a.m † 1 A =  ......    (32) M,N A  ddet ♯ ♯ ♯ ♯ cdetn(A A). n a. 1 ... cdetn(A A). n a.m .       +  If m>n, then by Theorem 4.2 for AM,N we have (29) as well.

Remark 4.2 If rank A = m, and (AA♯) is Hermitian, then by Corollary 3.1, † ♯ ♯ −1 ♯ −1 AM,N = A AA . Considering AA as a right inverse, we get the following representation of A+ ,  M,N 

rdet (AA♯) a♯ ... rdet (AA♯) a♯ 1 1 1. 1. m m. 1. A† = . (33) M,N ddetA  ......     rdet (AA♯) a♯ ... rdet (AA♯) a♯  1 1. n. m m. n.    +   If m

+ We obtain determinantal representations of the projection matrices AM,N A and † AAM,N in the following corollaries. Hm×n ♯ Corollary 4.1 If A ∈ r , where r < min {m,n} or r = m

♯ β cdeti A A . i (d.j ) β β∈Jr, n{i} pij =   , P ♯ β (A A) β β∈J r, n P ♯ n×n where d.j is the j-th column of A A ∈ H and for all i, j = 1,n. Proof. Representing A† by (29) and right-multiplying it by A, we obtain the † following presentation of an entry pij of AM,N A =: P = (pij )n×n.

17 ♯ ♯ β cdeti A A . i a. k β β∈Jr, n{i} pij =     · akj = P ♯ β k (A A) β X β∈J r, n P♯ ♯ β ♯ β cdeti A A . i a.k β · akj cdeti A A . i (d. j ) β β∈Jr, n{i} k β∈Jr, n{i} =     =   , P P ♯ β P ♯ β (A A) β (A A) β β∈J β∈J r, n r, n P P ♯ n×n where d.j is the j-th column of A A ∈ H and for all i, j = 1,n.  By analogy can be proved the following corollary. Hm×n Corollary 4.2 If A ∈ r , where r < min {m,n} or r = n

♯ α rdetj (A A )j. (gi. ) α α∈I {j} q r, m , ij = P ♯ α  |(AA ) α| α∈Ir, m P ♯ m×m where gi. is the i-th row of (AA ) ∈ H and for all i, j = 1,m.

4.2 The case of non-Hermitian A♯A and AA♯. In this subsection we derive determinantal representations of the weighted Moore- Penrose inverse of A ∈ Hm×n, when (AA♯) ∈ Hm×m and (A♯A) ∈ Hn×n are non-Hermitian. First, let (A♯A) ∈ Hn×n be non-Hermitian and rank(A♯A) < n. Consider (λI + A♯A)−1A♯. We have,

(λI + A♯A)−1A♯ = (λI + N−1A∗MA)−1A♯ = (N−1(λN + A∗MA))−1A♯ = ∗ −1 ∗ − 1 − 1 ∗ − 1 −1 − 1 ∗ (λN + A MA) A M = N 2 (λ + N 2 A MAN 2 ) N 2 A M = ∗ −1 − 1 1 − 1 1 − 1 − 1 ∗ 1 1 N 2 λ + M 2 AN 2 M 2 AN 2 N 2 A M 2 M 2 (34)       Since by Lemma 2.1

∗ −1 † 1 − 1 1 − 1 − 1 ∗ 1 1 − 1 lim λ + M 2 AN 2 M 2 AN 2 N 2 A M 2 = M 2 AN 2 , λ→0         then combining (23) and (34), we obtain

† † 1 1 1 1 − 2 2 − 2 2 AM,N = N M AN M . (35)   18 † † 1 − 1 Denote bya ˆij an ij-th entry of Aˆ := M 2 AN 2 . By determinantal representing (8) for Aˆ †, we obtain  

∗ ∗ 1 − 1 1 − 1 1 − 1 β cdeti M 2 AN 2 M 2 AN 2 m 2 an 2 β . i .j † β∈Jr, n{i}   aˆij = P    ∗    = 1 − 1 1 − 1 β M 2 AN 2 M 2 AN 2 β β∈J r, n     P ∗ 1 − 1 1 − 1 − 1 ∗ 1 β cdeti M 2 AN 2 M 2 AN 2 n 2 a m 2 β . i .j β∈Jr, n{i}   P    ∗    1 − 1 1 − 1 β M 2 AN 2 M 2 AN 2 β β∈J r, n     P ∗ ∗ 1 − 1 1 − 1 where m 2 an 2 denote the j-th column of M 2 AN 2 for all j = 1,m. .j −1  1   1 2 − 2 2 By nik denote an ik-th entry of N for all i, k = 1,n, and by mlj denote 1 an lj-th entry of M 2 for all l, j = 1,m, respectively. Then for the weighted + + Hn×m Moore-Penrose inverse AM,N = a˜ij ∈ , we have

n m  1 1 † − 2 † 2 a˜ij = nik aˆklmlj = k l X X 1 1 1 ∗ 1 1 1 − 2 − − − ∗ β nik · cdetk M 2 AN 2 M 2 AN 2 n 2 a m β . k .j k β∈Jr, n{i}   P P  ∗     = 1 − 1 1 − 1 β M 2 AN 2 M 2 AN 2 β β∈Jr, n P     − 1 1 1 1 2 − ∗ − − ∗ β nik cdetk N 2 A MAN 2 n 2 a m β k β∈J {i} . k .j r, n      , (36) P P − 1 ∗ − 1 β N 2 A MAN 2 β β∈J r, n   P for all i = 1,n, j = 1,m. ♯ † ♯ −1 ♯ If rank(A A)= n, then by Corollary 3.1, AM,N = (A A) A . So,

† −1 ∗ −1 −1 ∗ ∗ −1 ∗ AM,N = N A MA N A M = (A MA) A M. (37)

Since A∗MA is Hermitian, then we can use the determinantal representation ∗ n×m of a Hermitian inverse matrix (7). Denote A M =: (ˆa)ij ∈ H . So, we have

n L aˆ cdet (A∗MA) aˆ a˜† = k=1 ki kj = i .i .j . (38) ij det(A∗MA) det(A∗MA) P ∗ where aˆ.j is the j-th column of A M for all j = 1,m.

19 Now, let (AA♯) ∈ Hm×m be non-Hermitian and rank(AA♯) < m. By determinantal representing (9) for Aˆ †, we similarly obtain

∗ 1 − 1 1 − 1 − 1 ∗ 1 α rdetj M 2 AN 2 M 2 AN 2 n 2 a m 2 α j. i. † α∈Ir, m{j}   aˆij = P    ∗  , 1 − 1 1 − 1 α M 2 AN 2 M 2 AN 2 α α∈I r, m     P − 1 ∗ 1 − 1 ∗ 1 where n 2 a m 2 denote the i-th row of N 2 A M 2 for all i = 1,n. i. Finally, we get   

† a˜ij = ∗ α 1 − 1 1 − 1 −1 ∗ 1 rdetl M 2 AN 2 M 2 AN 2 n a m 2 l. i. α∈Ir, m{l} α 1 2 P     ∗   · mlj = 1 − 1 1 − 1 α l M 2 AN 2 M 2 AN 2 α X α∈Ir, m     α P 1 1 1 1 −1 −1 ∗ 2 rdetl M 2 AN AM 2 n a m 2 mlj l α∈I {l} l. i. α r, m      , (39) P P 1 −1 1 α M 2 AN AM 2 α α∈I r, m   P for all i = 1,n, j = 1,m. ♯ + ♯ ♯ −1 If rank(AA )= m, then by Corollary 3.1, AM,N = A (AA ) . So,

† −1 ∗ −1 ∗ −1 −1 ∗ −1 ∗ −1 AM,N = N A M AN A M = N A AN A . (40) Since AN−1A∗ is Hermitian and full-rank, then we can use the determinantal −1 ∗ representation of a Hermitian inverse matrix (6). Denote N A =: (ˇa)ij ∈ Hn×m. So, we have n aˆ R rdet (AN−1A∗) aˆ a˜+ = k=1 ik jk = j j. i. , (41) ij det(AN−1A∗) det(AN−1A∗) P −1 ∗ where aˆi. is the i-th row of N A for all i = 1,n. Thus, we have proved the following theorem. Hm×n ♯ Theorem 4.2 Let A ∈ r . If A A is non-Hermitian, then the weighted † † Hn×m Moore-Penrose inverse AM,N = a˜ij ∈ possess the determinantal rep- resentation (36) if r

20 5 Determinantal representation of the weighted Moore-Penrose solution of system linear equa- tions

Consider a right system linear equation over the quaternion skew field,

Ax = b (42) where A ∈ Hm×n is the coefficient matrix, b ∈ Hm×1 is a column of constants, and x ∈ Hn×1 is a unknown column. Due to [15] we have the following theorem that characterizes the weighted Moore-Penrose solution of (42).

♯ Theorem 5.1 The right system linear equation (42) with restriction x ∈ Rr(A ) + has the unique solution x˜ = AM,N b.

Theorem 5.2 Let A ∈ Hm×n and A♯A ∈ Hn×n be Hermitian. (i) If rank A = k ≤ m

♯ β cdeti A A . i (f) β β∈Jr, n{i} x˜i =   , (43) P ♯ β (A A) β β∈J r, n P where f = A♯ b, for all i = 1,n. (ii) If rank A = n, then for all i = 1,n we have

A♯A f cdeti .i ( ) x˜i = ♯ . (44) det A A + Proof. i) If rank A = k ≤ m

♯ ♯ β m cdeti A A . i a. j β β∈Jr, n{i} x˜i =     · bj = P ♯ β i=1 (A A) β X β∈J r, n P A♯A a♯ β b A♯A f β cdeti . i .j β · j cdeti . i ( ) β β∈Jr, n{i} j β∈Jr, n{i} =     =   , P P ♯ β P ♯ β (A A) β (A A) β β∈J β∈J r, n r, n P P where f = A♯b and for all i = 1 ,n. + + ii) If rank A = n, then AM,N can be represented by (32). Representing A b by component-wise directly gives (44). 

21 Theorem 5.3 Let A ∈ Hm×n and A♯A ∈ Hn×n be non-Hermitian. (i) If rank A = k ≤ m

− 1 1 1 β 2 − 2 ∗ − 2 ˆ nik cdetk N A MAN f β . k k β∈Jr, n{i} x˜i =    , P P − 1 ∗ − 1 β N 2 A MAN 2 β β∈J r, n   P 1 − 1 1 ˆ − ∗ 2 − where f = N 2 A M b and nik is an ik-th entry of N 2 for all i, k = 1,n.   (ii) If rank A = n, then for all i = 1,n we have

cdet (A∗MA) ˇf x˜ = i .i . i det(A∗MA)

where ˇf = A∗Mb for all j = 1,m.

Proof. The proof is similar to the proof of Theorem 5.2 using component-wise + representations of AM,N by (36) in the (i) point, and (38) in the (ii) point, respectively.  Consider a left system linear equation over the quaternion skew field,

xA = b (45) where A ∈ Hm×n is the coefficient matrix, b ∈ H1×n is a row of constants, and x ∈ H1×m is a unknown row. The following theorem characterizes the weighted Moore-Penrose solution of (45). ♯ Theorem 5.4 The left system linear equation (45) with restriction x ∈ Rl(A ) + has the unique solution x˜ = bAM,N .

Theorem 5.5 Let A ∈ Hm×n and AA♯ ∈ Hm×m be Hermitian. (i) If rank A = k ≤ n

♯ α rdeti A A j. (g) α α∈I {j} x r, m , ˜j = P  ♯ α  |(A A) α| α∈Ir, m P where g = b A♯, for all j = 1,m. (ii) If rank A = m, then for all j = 1,m we have AA♯ g rdetj j. ( ) x˜j = ♯ . det AA

22 Proof. The proof is similar to the proof of Theorem 5.2 using component-wise + representations of AM,N by (30) in the (i) point, and (33) in the (ii) point, respectively. 

Theorem 5.6 Let A ∈ Hm×n and AA♯ ∈ Hm×m be non-Hermitian. (i) If rank A = k ≤ n

1 1 α 1 −1 2 rdetl M 2 AN AM 2 (gˆ) mlj l. l α∈Ir, m{l} α x˜j =    , P P 1 −1 1 α M 2 AN AM 2 α α∈I r, m   P 1 1 1 −1 ∗ 2 where gˆ = b N A M 2 , mlj is lj-th entry of M 2 for all l, j = 1,m.   (ii) If rank A = m, then for all j = 1,m we have

rdet (AN−1A∗) gˇ x˜ = j j. . j det(AN−1A∗)

where gˇ = bN−1A∗.

Proof. The proof is similar to the proof of Theorem 5.2 using component-wise + representations of AM,N by (39) in the (i) point, and (41) in the (ii) point, respectively. 

6 Examples

In this section, we give examples to illustrate our results. 1. Let us consider the matrices 1 i j −k i 1 A = , (46)  k j −i  j −1 i     

23 16 − 2i − 2j + 10k −16+10i − 2j − 2k N−1 = 16+2i +2j − 10k 29 −19 − i − 13j − k , −16 − 10i +2j +2k −19+ i + 13j + k 29   2 k i 0  −k 2 0 j M = . (47) −i 0 2 k  0 −j −k 2    

23 By direct calculation we get that leading principal minors of M and N−1 are all positive. Therefore, due to Proposition 2.7, M and N−1 are positive definite matrices. Similarly, by direct calculation of leading principal minors of A∗A, we obtain rank A∗A = rank A = 2. Further,

A♯ = MA∗N−1 = 51 − 12i + 25j − 24k −43 − 18i + 39k −18 + 26i − 30j − 38k 19 − i − 50j − 42k −32i + 17j − 37k −24 − 50i + 26j + 24k −5 − 24i − 56j + k −38 − 25i − 18j − 67k . 5 − 6i − 50j + 11k 44 + 23i − 12j + 7k 30 + 38i + 5j + 37k 18 − 44i + 6j + 54k   Since,

A♯A = 178 41 + 47i + 47j + 43k −41 + 43i + 47j + 47k 41 − 47i − 47j − 43k 176 −40 − 46i − 42j − 46k −41 − 43i − 47j − 47k −40 + 46i + 42j + 46k 176 !

+ + H3×4 are Hermitian, then we shall be obtain AM,N = a˜ij ∈ due to Theorem 4.2 by Eq. (29). ♯ β  We have, A A β = 23380+ 23380+ 23380 = 70140, and β∈J 2, 3 P 

A♯A a♯ β i j k cdet1 . 1 .1 β = 6680+ 1670 + 3340 − 5010 + β∈J {1} X2, 3     6680 − 5010i + 3340j − 1640k = 13360 − 3340i + 6680j − 6680k.

Then, 8 − 2i +4j − 4k a˜+ = . 11 42 Similarly, we obtain

+ −7−3i+6k + −3+4i−5j−6k + 3−8j−7k a˜12 = 42 , a˜13 = 42 , a˜14 = 42 ,

+ −5i+3j−6k + −4−8i+2j+2k + −1−4i−9j + −6−4i−3j−11k a˜21 = 42 , a˜22 = 42 , a˜23 = 42 , a˜24 = 42 ,

+ −1−i−8j+2k + 7+4i−2j+k + 5+6i+j+6k + 3−7i+j+9k a˜31 = 42 , a˜32 = 42 , a˜33 = 42 , a˜34 = 42 . Finally, we obtain

+ AM,N = − − − − − − − − − 1 8 2i + 4j 4k 7 3i + 6k 3 + 4i 5j 6k 3 8j 7k −5i + 3j − 6k −4 − 8i + 2j + 2k −1 − 4i − 9j −6 − 4i − 3j − 11k . 42 −1 − i − 8j + 2k 7 + 4i − 2j + k 5 + 6i + j + 6k 3 − 7i + j + 9k ! (48)

24 2. Consider the right system of linear equations,

Ax = b, (49) where the coefficient matrix A is (46) and the column b = (10 i k)T . Using (48), by the matrix method we have for the weighted Moore-Penrose solution + ˜x = AM,N b of (49) with weights M and N from (47),

11 − 13i − 2j +4k 15 − 9i +7j − 3k −16+5i +5j +4k x˜ = , x˜ = , x˜ = . 1 42 2 42 3 42 (50) Now, we shall find the weighted Moore-Penrose solution of (49) by Cramer’s rule (43). Since

67 − 80i − 12j + 25k f = A♯b = 91 − 55i + 43j − 19k , −97+30i + 31j + 24k   then we have

♯ β cdeti A A . i (f) β β∈J2, 3{i} 18370 − 21710i − 3340j + 6680k x˜1 =   = = P ♯ β 70140 (A A) β β∈J r, n P 11 − 13i − 2j +4k , 42

25050 − 15030i + 11690j − 5010k 15 − 9i +7j − 3k x˜ = = , 2 70140 42 −26720 + 8350i + 8350j + 6680k −16+5i +5j +4k x˜ = = . 3 70140 42 As we expected, the weighted Moore-Penrose solutions by Cramer’s rule and be the matrix method coincide.

7 Conclusion

In this paper, we derive determinantal representations of the weighted Moore- Penrose by WSVD within the framework of the theory of the noncommutative column-row determinants. Recently, within the framework of the theory of the noncommutative column-row determinants we have been obtain the deter- minantal representations of the Drazin inverse [24] and the weighted Drazin inverse [25] and corresponding determinantal representations of generallized in- verse solutions of some matrix equations in [26–30].

25 References

[1] R. Penrose, A generalized inverse for matrices, Proc. Camb. Philos. Soc. 51 (1955) 406–413. [2] K.M. Prasad, R.B. Bapat, A note of the Khatri inverse, Sankhya: Indian J. Stat. 54 (1992) 291-295. [3] Y. Wei, H. Wu, The representation and approximation for the weighted MoorePenrose inverse, Appl. Math. Comput. 121 (2001) 17-28. [4] I. V. Sergienko, E. F. Galba, V. S. Deineka, Limiting representations of weighted pseudoinverse matrices with positive definite weights. Problem regularization, Cybernetics and Systems Analysis 39(6) (2003) 816-830. [5] P. S. Stanimirovic’, M. Stankovic’, Determinantal representation of weighted Moore-Penrose inverse, Mat. Vesnik 46 (1994) 41-50. [6] X. Liu, G. Zhu, G. Zhou, Y. Yu, An analog of the adjugate matrix for the (2) outer inverse AT,S, Mathematical Problems in Engineering, Volume 2012, Article ID 591256, 14 pages. [7] I.I. Kyrchei, Analogs of the adjoint matrix for generalized inverses and corresponding Cramer rules, Linear and Multilinear Algebra 56(4) (2008) 453-469. [8] X. Liu, Y. Yu, H. Wang, Determinantal representation of weighted gener- alized inverses, Appl. Math. Comput. 218(7) (2011) 3110-3121. [9] F. Zhang, Quaternions and matrices of quaternions, Appl. 251 (1997) 21-57. [10] I.I. Kyrchei, Determinantal representation of the MoorePenrose inverse ma- trix over the quaternion skew field, Journal of Mathematical Sciences 180(1) (2012) 23-33. [11] I. I. Kyrchei, The theory of the column and row determinants in a quater- nion linear algebra, Advances in Mathematics Research 15, pp. 301-359. Nova Sci. Publ., New York, (2012).

[12] I.I. Kyrchei Determinantal representations of the Moore-Penrose inverse over the quaternion skew field and corresponding Cramer’s rules, Linear and Multilinear Algebra 59(4) (2011) 413-431. [13] C.F. Van Loan, Generalizing the singular value decomposition, SIAMJ. Numer. Anal. 13 (1976) 76–83. [14] E. F. Galba, Weighted singular decomposition and weighted pseudoinver- sion of matrices, Ukr. Math. J. 48(10)(1996) 1618-1622.

26 [15] G. Song, Q. Wang, H. Chang, Cramer rule for the unique solution of re- stricted matrix equations over the quaternion skew field, Comput Math. Appl. 61 (2011) 1576-1589. [16] G.J. Song, Q.W. Wang, Condensed Cramer rule for some restricted quater- nion linear equations, Appl. Math. Comput. 208(2) (2009) 556-563. [17] L. Huang, Wasin So, On left eigenvalues of a , Linear Algebra Appl. 323 (2001) 105-116. [18] W. So, Quaternionic left eigenvalue problem, Southeast Asian Bulletin of Mathematics 29 (2005) 555-565. [19] R. M. W. Wood, Quaternionic eigenvalues, Bull. Lond. Math. Soc. 17 (1985) 137-138. [20] A. Baker, Right eigenvalues for quaternionic matrices: a topological ap- proach, Linear Algebra and its Applications 286 (1999) 303-309. [21] T. Dray, C. A. Manogue, The octonionic eigenvalue problem, Advances in Applied Clifford Algebras 8(2) (1998) 341-364. [22] R. A. Horn, C. R. Johnson, Matrix Analysis. Cambridge etc., Cambridge University Press, (1985). [23] A. Ben-Israel and T.N.E. Grenville. Generalized Inverses: Theory and Ap- plications. Springer- Verlag, Berlin, (2002). [24] I. Kyrchei, Determinantal representations of the Drazin inverse over the quaternion skew field with applications to some matrix equations, Appl. Math. Comput. 238 (2014), pp. 193-207. [25] I. Kyrchei, Determinantal representations of the W-weighted Drazin inverse over the quaternion skew field. Appl. Math. Comput. 264 (2015) 453–465. [26] I. Kyrchei, Cramer’s rule for quaternion systems of linear equations, J. Math. Sci. 155 (6) (2008) 839–858. [27] I. Kyrchei, Cramer’s rule for some quaternion matrix equations, Appl. Math. Comput. 217(5) (2010) 2024–2030. [28] I. Kyrchei, Explicit formulas for determinantal representations of the Drazin inverse solutions of some matrix and differential matrix equations, Appl. Math. Comput. 219 (2013) 1576–1589. [29] I. Kyrchei, Analogs of Cramer’s rule for the minimum norm least squares solutions of some matrix equations, Appl. Math. Comput. 218 (2012) 6375– 6384. [30] I. Kyrchei, Explicit representation formulas for the minimum norm least squares solutions of some quaternion matrix equations, Linear Algebra and Its Applications, 438 no. 1, (2013), pp. 136152.

27