dissertation
dissertation
net/publication/36283144
Article
Source: OAI
CITATIONS READS
101 8,278
1 author:
Dennis I. Merino
Southeastern Louisiana University
45 PUBLICATIONS 362 CITATIONS
SEE PROFILE
All content following this page was uploaded by Dennis I. Merino on 03 January 2016.
Baltimore, Maryland
i
Dedication
ii
Contents
Preface v
iii
2.7 A Sufficient Condition For Existence of A Square Root . . . . 55
2.8 A Canonical Form for
Complex Orthogonal Equivalence . . . . . . . . . . . . . . . . 60
Bibliography 120
Vita 123
iv
Preface
I came to America because it was a cool thing to do. And I was expected
to do so. It was the natural order of things back at the University of the
Philippines. You graduate with honors, you teach for a year or two, and
then you go to your Uncle Sam. I arrived at The Johns Hopkins University
without a clear objective (even though my application letter would tell you
otherwise). I did not know what to do. I guess I was lucky in that every
course I took was a palatable dish tasted for the first time. I was also given
choices I did not know I had.
In my second semester, I took Dr. Roger A. Horn’s class in Matrix Analy-
sis. One thing I noticed was that he was far from ordinary. He was (and
still is) a performer. An artist. And our classroom was his stage. I looked
forward to, and was swept away by, each and every performance. I knew
then what I wanted to do.
The notation in this work is standard, as in [HJ1] (except for the trans-
pose used in Chapter 4), and is explained at the beginning of each chapter.
Each chapter addresses a different topic and has its own set of notations and
definitions. Each chapter can be read independently of the others.
In Chapter 1, we study properties of coninvolutory matrices (EE = I),
and derive a canonical form under similarity as well as a canonical form un-
der unitary consimilarity for them. We show that any complex matrix has
a coninvolutory dilation and we characterize the minimum size of a conin-
volutory dilation of a square matrix. We characterize those A ∈ Mm,n (the
m-by-n complex matrices) that can be factored as A = RE with R real and
E coninvolutory, and we discuss the uniqueness of this factorization when A
is square and nonsingular.
Let A, C ∈ Mm,n and B, D ∈ Mn,m . The pair (A, B) is contragrediently
equivalent to the pair (C, D) if there are nonsingular complex matrices X ∈
v
Mm and Y ∈ Mn such that XAY −1 = C and Y BX −1 = D. In Chapter 2, we
develop a complete set of invariants and an explicit canonical form for this
equivalence relation. We show that (A, AT ) is contragrediently equivalent
to (C, C T ) if and only if there are complex orthogonal matrices P and Q
such that C = P AQ. Using this result, we show that the following are
equivalent for a given A ∈ Mn : (a) A = QS for some complex orthogonal Q
and some complex symmetric S; (b) AT A is similar to AAT ; (c) (A, AT ) is
contragrediently equivalent to (AT , A); (d) A = Q1 AT Q2 for some complex
orthogonal Q1 , Q2 ; (e) A = P AT P for some complex orthogonal P . We
then consider a generalization of the transpose operator to a general linear
operator φ : Mn → Mn that satisfies the following properties: for every
A, B ∈ Mn , (a) φ preserves the spectrum of A, (b) φ(φ(A)) = A, and (c)
φ(AB) = φ(B)φ(A). We show that (A, φ(A)) is contragrediently similar to
(B, φ(B)) if and only if A = X1 BX2 for some nonsingular X1 , X2 ∈ Mn that
satisfy X1−1 = φ(X1 ) and X2−1 = φ(X2 ). We also consider a factorization
of the form A = XY , where X ∈ Mn is nonsingular and X −1 = φ(X),
and Y ∈ Mn satisfies Y = φ(Y ). We also discuss the operators A → A∗
and A → A. We then study the nilpotent part of the Jordan canonical
forms of the products AB and BA. We use the canonical form for the
contragredient equivalence relation to give a new proof to a result due to
Flanders concerning the relative sizes of the nilpotent Jordan blocks of AB
and BA. We present a sufficient condition for the existence of square roots
of AB and BA and close the chapter with a canonical form for the complex
orthogonal equivalence relation (A → Q1 AQ2 , where Q1 ∈ Mm and Q2 ∈ Mn
are complex orthogonal matrices).
In Chapter 3, we characterize the linear operators on Mm,n that preserve
complex orthogonal equivalence, that is, T : Mm,n → Mm,n is linear and has
the property that T (A) is orthogonally equivalent to T (B) whenever A and
B are orthogonally equivalent.
In Chapter 4, we answer the question of Horn, Hong and Li in [HHL] con-
cerning characterization of linear operators in Mn (IF) that preserve unitary
t-congruence (A → U AU t , with unitary U ) in the cases IF = IR and IF = C.
The result in the real case is a rather pleasant surprise.
In the Appendix, we present a new proof for the Jordan canonical theorem
that incorporates ideas developed in Chapter 2.
I wish to thank Dr. Roger A. Horn and Dr. Chi-Kwong Li for their help
and guidance; and Dr. Irving Kaplansky for his inspiring basis-free methods.
vi
Chapter 1
A Real-Coninvolutory Analog
of the Polar Decomposition
1
2
Proposition 1.1.1 Let A ∈ Mn . Any two of the following implies the third:
(1) A is unitary.
(2) A is symmetric.
(3) A is coninvolutory.
Proof The forward implication is easily verified. For the converse, the condi-
−1
tions are equivalent to A = SBS −1 and A = SBS . Hence, AS = SB
and AS = SB. Let S = R1 + iR2 with R1 , R2 ∈ Mn (IR). Then
R1 = (S +S)/2 and R2 = (S −S)/2i, so A(αR1 +βR2 ) = (αR1 +βR2 )B
for all α, β ∈ IR. Let p(z) ≡ det(R1 + zR2 ). Then p(z) is a polynomial
of degree at most n and p(z) 6≡ 0 since p(i) 6= 0. Hence, p(β0 ) 6= 0
for some β0 ∈ IR. Thus, R ≡ R1 + β0 R2 ∈ Mn (IR) is nonsingular,
AR = RB, and A = RBR−1 .
Proof This follows directly from the theorem since q(A) = Sq(B)S −1 for
any polynomial q(t).
It is known that two symmetric matrices are similar if and only if they
are orthogonally similar and two Hermitian (even normal) matrices are sim-
ilar if and only if they are unitarily similar. An analogous result holds for
coninvolutory matrices.
Corollary 1.1.5 Two coninvolutory matrices are similar if and only if they
are real similar (that is, the similarity can be achieved via a nonsingular real
matrix).
Direct computations (and a simple induction for (c)) verify the following
facts and functional equations involving the matrices E2k (A, B).
by Lemma 1.1.7(g, c). Since E2k (λ) and eiB are similar coninvolu-
tory matrices, Corollary 1.1.5 ensures that they are real similar, so
there is some nonsingular C ∈ M2k (IR) such that E2k (λ) = CeiB C −1 =
−1
eiCBC = eiR with R ≡ CBC −1 ∈ M2k (IR). Since the spectrum
of iR ∼ Jk (α) ⊕ −Jk (α) is confined to the horizontal line {z ∈ C :
Im z = arg λ}, basic facts about primary matrix functions (see Ex-
ample (6.2.16) in [HJ2]) ensure that there is a polynomial p(t) (which
may depend on λ) such that iR = p(E2k (λ)).
Jl1 (eiθ1 ) ⊕ · · · ⊕ Jlr (eiθr ) ⊕ [Jk1 (λ1 ) ⊕ Jk1 (1/λ1 )] ⊕ · · · ⊕ [Jks (λs ) ⊕ Jks (1/λs )]
8
where each Rj ≡ 1i E2kj (Jkj (αj ), −Jkj (αj )) ∈ M2kj (IR) and αj ≡ ln λj (prin-
cipal branch) for j = 1, . . . , s.
For a coninvolutory E, the real matrix S in (b) and the coninvolutory matrix
F in (c) may be taken to be polynomials in E.
(d) If σ > 0 is given and if k is a given positive integer, then there exists
a coninvolutory matrix E such that σ is a singular value of E with
multiplicity k.
Proof The first three assertions follow directly from the singular value de-
−1
composition of E and the equation E = E . If σ > 0 and σ 6= 1,
then " #
σ 0
E2 (σ) = U U∗
0 1/σ
has singular values σ and 1/σ. Hence, we may take E = E2 (σ) ⊕ · · · ⊕
E2 (σ) (k copies). If σ = 1, then we may take E = Ik .
Let E ∈ Cn . Proposition 1.1.13(a) shows that the singular value decom-
position of E has the form E = U ΣV , where U and V are unitary and
1 1
Σ = [σ1 In1 ⊕ In1 ] ⊕ · · · ⊕ [σk Ink ⊕ Ink ] ⊕ In−2j (1.1.2)
σ1 σk
Pk
with σ1 > · · · > σk > 1 and j = i=1 ni . Partition the unitary matrix
X11 X12 · · · X1,l−2 X1,l−1 X1l
X21 X22 · · · X2,l−2 X2,l−1 X2l
.. .. . . . .. .. ..
. . . . .
VU =
X
l−2,1 Xl−2,2 · · · Xl−2,l−2 Xl−2,l−1 Xl−2,l
Xl−1,1 Xl−1,2 · · · Xl−1,l−2 Xl−1,l−1 Xl−1,l
Xl1 Xl2 · · · Xl,l−2 Xl,l−1 Xll
10
Let k · kF denote the Frobenius norm (square root of the sum of squares
of moduli of entries). Since all the columns and rows of V U are unit vectors,
we have
k[ X11 X12 · · · X1l ]k2F = n1 = k[ X11
T T
X21 · · · Xl1T ]k2F . (1.1.4)
T T
Equating the first rows of (1.1.3) gives X21 = X12 and X11 = σ12 X11 , . . .,
T T T T
Xl−2,1 = σ1 σk X1,l−2 , Xl−1,1 = (σ1 /σk )X1,l−1 , Xl1 = σ1 X1l . Since all of the
coefficients σ12 , . . . , σ1 σk , σ1 /σk , and σ1 are greater than one, substituting
these identities in (1.1.4) shows that X1i = Xi1 = 0 for all i 6= 2, and hence
kX12 k2F = n1 = kX21
T 2
kF = kX21 k2F .
Again using the fact that the rows of V U are unit vectors, we also have
n1 = k[ X21 X22 · · · X2l ]k2F ,
and hence X2i = 0 for all i ≥ 2. Similarly, Xi2 = 0 for all i ≥ 2. Thus
0 X12 0
T
V U = X12 0 0
0 0 V1
11
(a) A and A−1 have the same singular values, that is, A and A−1 are
unitarily equivalent.
Proof If A and A−1 have the same singular values, then A = V1 ΣV2 , where
Σ has the form (1.1.6) and V1 , V2 ∈ Mn are unitary. The converse
assertion in Theorem 1.1.14 ensures that A = (V1 ΣV1T )(V 1 V2 ) is a
factorization of the form asserted in (b). If A = EW as in (b), then
A = W (W T EW ) is a factorization of the form asserted in (c). If
A = V E as in (c), then one may take W = I in (d). If A = V EW as
in (d), then A−1 = W ∗ EV ∗ . Thus, the singular values of A and A−1
are those of E and E, which by Proposition 1.1.13(b) are the same.
Proof Let P be the unique positive definite square root of A. Notice that
P −1 EP ∈ Cn , hence Theorem 1.1.14 ensures that there exists a unitary
U and a Σ of the form (1.1.6) such that P −1 EP = U ΣU T . Let S ≡
−1
U ∗ P −1 , so that S ∗ = P −1 U . Then SAS ∗ = I and SES = Σ, as
desired.
13
under consimilarity,
" # " −1
# " #" #" −1
#
X 0 X 0 X 0 JR (A) ∗ X 0
B =
0 I 0 I 0 I ∗ ∗ 0 I
" −1
#
XJR (A)X ∗
=
∗ ∗
" #
A ∗
=
∗ ∗
is a coninvolutory dilation of A of size n + k. Lemma 1.1.20 guar-
antees that Cnt (at , bt ) has a coninvolutory dilation of size 2nt . More-
over, Jmt (λt ) has a coninvolutory dilation of size 2mt , and if λt = ±1
then Jmt (±1) has a coninvolutory dilation of size 2mt − 1. Thus,
Lemma 1.1.21 ensures that JR (A) has a coninvolutory dilation of size
2n − l, where l is the number of Jordan blocks in JR (A) with λt = ±1.
Now, n−l = rank(I −[JR (A)]2 ) = rank(I −X[JR (A)]2 X −1 ) = rank(I −
AA) = k0 . Therefore, A has a coninvolutory dilation of size n + k0 .
Lemma 1.1.18 guarantees that A has a coninvolutory dilation of size
n + k if and only if k ≥ k0 .
In [TK] several power dilation problems are considered: when does a given
square matrix A have a dilation
" #
A ∗
B=
∗ ∗
in a given class such that
" #
k Ak ∗
B = for k = 1, . . . , N ?
∗ ∗
Theorem 1.1.22 solves the power dilation problem for N = 1 in the class of
coninvolutory matrices; the situation for N > 1 is an open problem.
Proof The first assertion has been dealt with. For the second, suppose
−1
EE = I and E 2 = A A, and compute
−1
AE = A(E 2 )E = A(A A)E = AE.
Proof The forward assertion has already been established. For the con-
−1
verse, let F be coninvolutory and satisfy F 2 = G = F F . Then
Theorem 1.2.1 ensures that R(F, E0 ) ≡ F E0 has real entries and F =
R(F, E0 )E0 . Since E0 is a polynomial in G = F 2 , both E0−1 = E0 and
R(F, E0 ) are polynomials in F . It follows that both F and R(F, E0 )
commute with E0 . Thus, R(F, E0 )2 = F E0 F E0 = F 2 (E0 )2 = GG = I,
as desired.
Proof Theorem 1.2.1 ensures that R(A, E0 ) is real and A = R(A, E0 )E0 .
Moreover, Lemma 1.2.2 ensures that RE0 is coninvolutory and A =
(R(A, E0 )R)(RE0 ) is an RE decomposition if R2 = I and R is real and
commutes with E0 . For the last assertions, apply Lemma 1.2.2 to the
−1
coninvolutory matrix E1 , for which E12 = A A = E02 , and conclude
that R(E1 , E0 ) ≡ E1 E0 is real, E1 = R(E1 , E0 )E0 , R(E1 , E0 ) commutes
18
R1 = AE1
= AR(E1 , E0 )E0
= AE0 R(E1 , E0 )
= R(A, E0 )R(E1 , E0 ).
−1
If E0 is a given coninvolutory polynomial square root of A A, we see
that the set of all pairs (R, E) with R real, E coninvolutory, and A = RE
is precisely {(AE 0 R, RE0 ) : R is real, R2 = I, and R commutes with E0 }.
Depending on the Jordan canonical form of Log E0 (see Theorem (6.4.20)
in [HJ2]), there can be many R’s that satisfy these conditions. If we use
Theorem 1.1.11(b) to write E0 = eiS with S real, and if S = T CT −1 is its
real Jordan form (see Theorem (3.4.5) in [HJ1]) with T and C both real and
C = Cn1 ⊕ · · · ⊕ Cnk , then E0 = T eiC T −1 ≡ T (En1 ⊕ · · · ⊕ Enk )T −1 . Any
matrix of the form R = T (±In1 ⊕ · · · ⊕ ±Ink )T −1 is real, commutes with E0 ,
and satisfies R2 = I.
T
is an RE decomposition. Notice that AT2 and A2 have different ranges, so
AT2 does not have an RE decomposition. Thus, it is possible for one, but not
both, of the matrices A and AT to have an RE decomposition. We now show
that equality of the ranges of A and A is sufficient as well as necessary for A
to have an RE decomposition.
Theorem 1.2.4 Let A ∈ Mm,n be given. Then there exist a real R ∈ Mm,n
and a coninvolutory E ∈ Mn such that A = RE if and only if A and A have
the same range, that is, if and only if there exists a nonsingular S ∈ Mn such
that A = AS.
Proof The necessity of the condition has already been shown, so we only
need to show that it is sufficient.
We first show that the asserted equivalent conditions are both invariant
under right multiplication by a nonsingular matrix.
Proof The singular value decomposition ensures that there exist nonsingu-
lar matrices X ∈ Mm and Y ∈ Mn so that A = XRY , where R ∈ Mm,n
has real entries. Write X = E1 R1 and Y = R2 E2 , with each Ei conin-
volutory and each Ri real; then A = E1 (R1 RR2 )E2 .
The Contragredient
Equivalence Relation
21
22
2.1 Introduction
We denote the set of m-by-n matrices by Mm,n and write Mn ≡ Mn,n . Given
a scalar λ ∈ C, the n-by-n upper triangular Jordan block corresponding to
λ is denoted by Jn (λ). We say that A, B ∈ Mm,n are equivalent if there are
nonsingular matrices X ∈ Mm and Y ∈ Mn such that A = XBY .
equivalence of (A, B) and (C, D). We obtain a canonical form for this rela-
tion and use it to study a variety of matrix factorizations involving complex
orthogonal factors.
Lemma 2.2.1 Let positive integers k, n be given with k < n. Let Y1 ∈ Mn,k
and P ∈ Mk,n be given with P Y1 nonsingular. Then there exists a Y2 ∈
Mn,n−k such that P Y2 = 0 and [Y1 Y2 ] ∈ Mn is nonsingular.
Proof Since P has full (row) rank, we may let {ξ1 , . . . , ξn−k } be a basis for
the orthogonal complement of the span of the columns of P ∗ in Cn ,
and set Y2 ≡ [ξ1 . . . ξn−k ]. Then P Y2 = 0 and Y2 has full (column)
rank. Let η ∈ Ck and ζ ∈ Cn−k and suppose
" #
η
[Y1 Y2 ] = Y1 η + Y2 ζ = 0.
ζ
and partition BX −1 as
Compute
" # " # " #
J(AB) 0 C1 C1 D1 C1 D2
= XABX −1 = [D1 D2 ] = .
0 N C2 C2 D1 C2 D2
25
Now compute
and
" # " #
−1 −1 C1 C1 D1 C1 D2
Y BX = Y (BX )= [D1 D2 ] =
C3 C3 D1 C3 D2
" # " #
J(AB) 0 J(AB) 0
= ≡ ,
0 C3 D2 0 B
26
Y F = C1 D1 = J(AB)
For a given A ∈ Mm,n and B ∈ Mn,m , the four cases (a)—(d) in Lemma
2.2.2 are exhaustive and mutually exclusive. The special forms achieved by
27
the contragredient equivalences in (c) and (d) are the canonical forms we
seek, but more work is required to achieve a canonical form in (a). Once
that is achieved, however, the special form of the reduction in (b) shows that
a canonical form in this case can be expressed as a direct sum of the canonical
forms for cases (a) and (d): If
" # " #
−1 Ik 0 −1 J(AB) 0
XAY = and Y BX = ,
0 A 0 B
and " #
−1 J(AB) 0
(Y1 Y )B(X1 X) = ,
0 G1 BF1−1
and (F1 AG−1 −1
1 )(G1 BF1 ) = F1 (AB)F1
−1
is nilpotent. Thus, if F1 and G1 are
chosen to put (A, B) into canonical form, we shall have achieved a canonical
form for (A, B).
In order to achieve a canonical form for (A, B) under contragredient equiv-
alence, we see that there are only two essential cases: AB is nonsingular or
AB is nilpotent. Before attacking the second case, it is convenient to sum-
marize what we have learned about the first.
Proof The forward implication is immediate (and does not require the rank
conditions). For the converse, the hypotheses ensure that the respective
Jordan canonical forms of AB and CD are J(AB)⊕0m−n and J(CD)⊕
0m−n and one may arrange the Jordan blocks in the nonsingular factor
J(CD) ∈ Mn so that J(AB) = J(CD). Inspection of the canonical
forms in Lemma 2.2.2 (c) and (d) now shows that (A, B) ∼ (C, D).
28
With the nonsingular case disposed of, we now begin a discussion of the
nilpotent case. Let m ≥ n, A ∈ Mm,n , and B ∈ Mn,m be given and suppose
AB (and hence also BA) is nilpotent. Then among all the nonzero finite
alternating products of the form
there is a longest one (possibly two, one of each type). Suppose that a longest
one is of Type I. There are two possibilities, depending on the parity of a
longest product.
Then
∗ y ∗ A(BA)k−1 x
·
P Y1 = QZ1 =
∈ Mk
·
·
y ∗ A(BA)k−1 x 0
Then " #
C1
AZ2 = Y C = [Y1 Y2 ] = Y1 C1 + Y2 A2 . (2.2.5)
A2
Now, use (2.2.5) and the identities (2.2.4b), (2.2.3b) and (2.2.4d) to compute
and
P AZ2 = (P A)Z2 = QZ2 = 0.
Since P Y1 is nonsingular, we conclude that C1 = 0. Thus, (2.2.5) simplifies
to
AZ2 = Y2 A2 (2.2.6)
30
so that " #
−1 Ik 0
Y AZ = . (2.2.7)
0 A2
Now, set D ≡ Z −1 BY2 and partition
" #
D1
D= with D1 ∈ Mk,m−k and B2 ∈ Mn−k,m−k .
B2
Now use (2.2.8) and the identities (2.2.4d), (2.2.3d) and (2.2.4b) to compute
and
QBY2 = (QB)Y2 = Jk (0)(P Y2 ) = 0.
Since QZ1 is nonsingular, we conclude that D1 = 0. Hence, (2.2.8) simplifies
to
BY2 = Z2 B2 , (2.2.9)
and we can use (2.2.3c) and (2.2.9) to compute
" #
JkT (0) 0
B[Y1 Y2 ] = [BY1 BY2 ] = [Z1 Z2 ] ,
0 B2
y∗B
y ∗ BAB
P ≡
.. ∈ Mk−1,m
.
y ∗ B(AB)k−2
and
y∗
∗
y BA
Q≡ ..
∈ Mk,n .
.
y ∗ (BA)k−2
y ∗ (BA)k−1
We define
(a) AZ1 = Y1 Hk
(b) P A = Kk Q
(2.2.15)
(c) BY1 = Z1 KkT
(d) QB = HkT P.
and
AZ2 = Y2 A2 . (2.2.17)
33
and " #
−1 JkT (0) 0
Y AZ = for some A2 ∈ Mm−k,n−k . (2.2.24)
0 A2
(b) if k = n ≤ m, then we may take B2 = 0 in (2.2.23) and A2 = 0 in
(2.2.24).
and " #
−1 KkT 0
Y AZ = for some A2 ∈ Mm−k,n−k+1 . (2.2.26)
0 A2
35
and
JB 0 ··· 0 0
0 B1 ··· 0 0
Y BX −1
= .. .. ... .. ..
(2.2.28)
. . . .
0 0 · · · Bp 0
0 0 ··· 0 0
where
The reason for introducing φ(A, B) is that the parts of the canonical form
in Theorem 2.2.4 that contribute to the nilpotent part of AB are completely
determined by the sequence φ(A, B).
Consider first what happens if A and B consist of only one canonical
block. For each choice of the possible pairs of canonical blocks, Table 1 gives
the ranks of the indicated products.
(Type, Case)
(I, 1) (I, 2) (II, 1) (II, 2)
A Ik Hk JkT (0) KkT
T T
B Jk (0) Kk Ik Hk
(AB)k 0 0 0 0
(BA)k 0 0 0 0
A(BA)k−1 1 0 0 0
B(AB)k−1 0 0 1 0
(AB)k−1 1 0 1 1
(BA)k−1 1 1 1 0
Table 1
Ranks of Products by Type and Case
Suppose AB is nilpotent and we are given the sequence φ(A, B). How can
we use φ(A, B) to determine the Ai in (2.2.27) and the Bi in (2.2.28)? Now
consider the general case of nilpotent AB. Let k be the smallest integer such
37
• Repeat the preceding three steps until the all the entries in the sequence
are 0.
The following result is immediate.
Lemma 2.2.6 Let integers m, n be given with m ≥ n, and suppose A, C ∈
Mm,n and B, D ∈ Mn,m . Suppose that AB is nilpotent. Then (A, B) ∼ (C, D)
if and only if φ(A, B) = φ(C, D).
Corollary 2.2.7 Let integers m, n be given and suppose A, C ∈ Mm,n and
B, D ∈ Mn,m . Then (A, B) ∼ (C, D) if and only if the following conditions
are satisfied:
(1) AB is similar to CD, and
(2) for all l = 0, 1, . . . , max {m, n},
(a) rank (BA)l = rank (DC)l ,
(b) rank A(BA)l = rank C(DC)l , and
(c) rank B(AB)l = rank D(CD)l .
Proof The forward implication is easily verified and was noted in the In-
troduction. For the converse, we may assume without loss of gener-
ality that m ≥ n. Theorem 2.2.4 guarantees that there exist non-
singular X1 , X2 ∈ Mm , Y1 , Y2 ∈ Mn , and a nonnegative integer k =
rank (AB)m = rank (CD)m such that
" # " #
A1 0 B1 0
X1 AY1−1 = and Y1 BX1−1 =
0 A 0 B
and " # " #
C1 0 D1 0
X2 CY2−1 = and Y2 DX2−1 = ,
0 C 0 D
where A1 , B1 , C1 , D1 ∈ Mk are nonsingular and AB and CD are nilpo-
tent. Since AB is similar to CD, the nonsingular parts of their respec-
tive Jordan canonical forms are the same. Therefore, A1 B1 is similar
to C1 D1 and hence Lemma 2.2.3 guarantees that (A1 , B1 ) ∼ (C1 , D1 ).
Moreover, (1) also ensures that rank (AB)l = rank (CD)l for all inte-
gers l = 0, 1, . . . , m. These identities and the rank identities (2) ensure
that φ(A, B) = φ(C, D), so (A, B) ∼ (C, D) by Lemma 2.2.6.
39
and
JB 0 ··· 0 0
0 Bσ(1) ··· 0 0
D≡ .. .. ... .. ..
, (2.2.30)
. . . .
0 0 · · · Bσ(p) 0
0 0 ··· 0 0
JA , JB , Ai , Bi , i = 1, . . . , p, are as in Theorem 2.2.4, and σ is any permuta-
tion of {1, . . . , p}.
and
rank ψ(A, B)2k+1 = rank A(BA)k + rank B(AB)k . (2.2.32)
If AB is similar to CD, ψ(A, B) is similar to ψ(C, D), and rank A(BA)l =
rank C(DC)l for all integers l ≥ 0, then the conditions of Corollary 2.2.7 are
satisfied, and hence (A, B) ∼ (C, D). Conversely, all of these conditions are
satisfied if (A, B) ∼ (C, D).
40
(3) rank A(BA)l = rank C(DC)l for all l = 0, 1, . . . , max {m, n}.
Lemma 2.3.1 Let A, C ∈ Mm,n with m ≥ n. Then the following are equiv-
alent:
(2) (A, AT ) ∼ (C, C T ) via orthogonal matrices, that is, there exist orthog-
onal Q1 ∈ Mm and Q2 ∈ Mn such that C = Q1 AQ2 .
" # " #
T T 0 A 0 C
(3) AA is similar to CC and is similar to .
AT 0 CT 0
41
for any B ∈ Mm,n and any integer k ≥ 0. Thus, we must also have
Definition 2.4.1 Let Sn+ denote the set of nonsingular symmetric (AT = A)
matrices in Mn , let Sn− denote the set of all nonsingular skew-symmetric
(AT = −A) matrices in Mn , and set Sn ≡ Sn+ ∪ Sn− . The spectrum of
A ∈ Mn , the set of eigenvalues of A, will be denoted by σ(A).
(3) p(φS (A)) = φS (p(A)) for all A ∈ Mn and all polynomials p(t).
(6) Suppose that S ∈ Sn+ and let S1 ∈ Sn+ be such that S12 = S. Then
S1 AS1−1 is φS symmetric if and only if A is symmetric.
(c) A commutes with φS (A) if and only if there exist commuting X and Y ∈
Mn such that A = XY , X is φS orthogonal, and Y is φS symmetric.
Proof For the forward implication, let X1 ≡ Y Y0−1 . All the assertions
follow from Theorem 2.4.5 and the observation that Y0 is a polyno-
mial in φS (A)A = Y 2 , which ensures that Y (and hence X1 ) com-
mutes with Y0 . Conversely, under the stated assumptions we have
XY = X(A, Y0 )X12 Y0 = X(A, Y0 )Y0 = A and φS (Y ) = φS (X1 Y0 ) =
Y0 φS (X1 ) = Y0 X1−1 = Y0 X1 = X1 Y0 = Y , so Y is φS symmetric.
46
Moreover,
so X is φS orthogonal.
Proof The forward implication is easily verified. For the converse, we have
ZBZ −1 = A = φS (φS (A)) = φS (ZφS (B)Z −1 ) = φS (Z)−1 BφS (Z),
so (φS (Z)Z)B = B(φS (Z)Z). Theorem 2.4.5 guarantees that there
exists a φS symmetric Y ∈ Mn that is a polynomial in φS (Z)Z, as
well as a φS orthogonal X ∈ Mn such that Z = XY . Since φS (Z)Z
commutes with B, Y also commutes with B. Thus, A = ZBZ −1 =
(XY )B(Y −1 X −1 ) = XBX −1 .
(2) (A, φS (A)) ∼ (B, φS (B)) via φS orthogonal matrices, that is, there exist
φS orthogonal X1 , X2 ∈ Mn such that A = X1 BX2 .
and follow the argument in the proof of Theorem 2.3.2 to show that:
Lemmata 2.4.12, 2.4.10 and 2.4.13 now imply the following result about
φS polar decomposition in the symmetric case, which is a generalization of
Theorem 2.3.2.
Similarly, φS (A)A = 0. Hence, AφS (A) is similar (in fact, equal) to φS (A)A,
but rank A = 1, so A does not have a φS polar decomposition.
It is an open question to characterize the even-rank matrices with even
dimensions that have a φS polar decomposition for a given S ∈ Sn− .
2.5 A∗ and A
It is natural to ask what the analogs of Lemma 2.3.1 are for the mappings
A → A or A → A∗ . Although these mappings satisfy some of the conditions
2.4.2, they are not linear and do not preserve the spectrum. For a given
A, C ∈ Mm,n , when is (A, A∗ ) ∼ (C, C ∗ )? If m = n, when is (A, A) ∼ (C, C)?
The first case is immediate, and we present it without proof.
Theorem 2.5.1 Let A, C ∈ Mm,n be given. Then the following are equiva-
lent:
(1) (A, A∗ ) ∼ (C, C ∗ ).
51
(4) (A, A∗ ) ∼ (C, C ∗ ) via unitary matrices, that is, A = U CV ∗ for some
unitary U ∈ Mn and V ∈ Mm .
−1
Recall that A, C ∈ Mn are consimilar if A = SCS for some nonsingu-
lar S ∈ Mn . We now apply the analysis of Lemma 2.3.1 to the conjugate
mapping.
Corollary 2.2.7 and Theorem 2.5.2 now proves the following, which is
Theorem (4.1) in [HH1].
and
Y (BA)Y −1 = J(AB) ⊕ Jn1 (0) ⊕ · · · ⊕ Jnp (0) ⊕ 0,
where J(AB) ∈ Mk is the nonsingular part of the Jordan canonical form of
AB, and |mi − ni | ≤ 1 for each i = 1, . . . , p. Conversely, let integers m1 ≥
· · · ≥ mp ≥ 1 and n1 ≥ · · · ≥ np ≥ 1 be given, and set m ≡ m1 + · · · + mp
and n ≡ n1 + · · · + np . Suppose that |mi − ni | ≤ 1 for each i = 1, . . . , p. Then
there exist A ∈ Mm,n and B ∈ Mn,m such that
and
BA is similar to Jn1 (0) ⊕ · · · ⊕ Jnp (0).
and
JB 0 ··· 0 0
0 B1 ··· 0 0
Y BX −1
= .. .. ... .. ..
,
. . . .
0 0 · · · Bp 0
0 0 ··· 0 0
where JA , JB ∈ Mk are nonsingular; each
T T T T
(Ai , Bi ) ∈ {(Imi , Jm i
(0)), (Jm i
(0), Imi ), (Hmi , Km i
), (Km i
, Hmi )},
We can also use the results in Section 2.2 to give a different proof for
Theorem (3) in [F] (see also [PM]).
Notice that
" #
−1 −1 −1 N11 N12 A
0 = N A = XN AY = (XN X )(XAY )=
N21 N22 A
n − k − 1 = n − (k + 1) = rank (BA)k+1
= rank B(AB)k A
≤ rank B(AB)k
≤ rank (AB)k
= (n − 1) − k = n − k − 1
and
n−k−1 = rank (BA)k+1
= rank BA(BA)k
≤ rank A(BA)k
= rank (AB)k A
≤ rank (AB)k
= n − k − 1,
56
and
T T
B1 ≡ [Jm 1
(0) ⊕ Iα1 ] ⊕ · · · ⊕ [Jm t
(0) ⊕ Iαt ]
with m1 ≥ · · · ≥ mt ≥ 0 and α1 ≥ · · · ≥ αt ≥ 0 (if all mi = αi = 0, then A1
and B1 are absent); and A2 and B2 contain all the canonical blocks of the
T
form Hmi and Km i
, that is,
" # " #
Hn1 0 Hns 0
A2 ≡ ⊕ ··· ⊕
0 KβT1 0 KβTs
A1 (B1 A1 )k = [Jm
T
1
(0)k ⊕ JαT1 (0)k+1 ] ⊕ · · · ⊕ [Jm
T
t
(0)k ⊕ JαTt (0)k+1 ] = 0
and
B1 (A1 B1 )k = [JmT
1
(0)k ⊕ JαT1 (0)k ] ⊕ · · · ⊕ [Jm
T
t
(0)k ⊕ JαTt (0)k ]
= [0 ⊕ JαT1 (0)k ] ⊕ · · · ⊕ [0 ⊕ JαTt (0)k ] 6= 0.
" # " #
Hn1 0 Hns 0
A2 ≡ ⊕ ··· ⊕ ,
0 KβT1 0 KβTs
and " # " #
KnT1 0 KnTs 0
B2 ≡ ⊕ ··· ⊕
0 Hβ1 0 Hβs
if and only if rank A(BA)k = rank B(AB)k for all k = 0, 1, 2, . . ..
T T
B1 = [Jm 1
(0) ⊕ Iβ1 ] ⊕ · · · ⊕ [Jm t
(0) ⊕ Iβt ],
" # " #
Hn1 0 Hns 0
A2 = ⊕ ··· ⊕ , (2.7.37)
0 KnT1 0 KnTs
and " # " #
KnT1 0 KnTs 0
B2 = ⊕ ··· ⊕ (2.7.38)
0 Hn1 0 Hns
if and only if rank (AB)k = rank (BA)k for all k = 1, 2, 3, . . ..
Theorem 2.7.4 Let A ∈ Mm,n and B ∈ Mn,m be given. Suppose that for
every integer k ≥ 0, we have
(1) rank A(BA)k = rank B(AB)k , and
(2) rank (BA)k+1 = rank (AB)k+1 .
Then AB and BA have square roots.
Proof Suppose m ≥ n. Then AB is similar to J(AB) ⊕ A1 B1 ⊕ A2 B2 ⊕ 0,
in which J(AB) (if present) is nonsingular and hence has a square
T T
root. Lemma 2.7.2 guarantees that A1 B1 = [Jm 1
(0) ⊕ Jm 1
(0)] ⊕ · · · ⊕
T T
[Jm t
(0)⊕Jmt (0)], and hence has a square root. Lemma 2.7.3 guarantees
T T T T
that A2 B2 = [Jn1 −1 (0)⊕ Jn1 (0)] ⊕· · · ⊕ [Jns −1 (0) ⊕Jns (0)], so that A2 B2
also has a square root. It follows that AB has a square root. Similarly,
BA has a square root.
The conditions of Theorem 2.7.4, though sufficient, are not necessary for
the existence of a square root of AB. The example A1 ≡ I2k and B1 ≡
Jk (0) ⊕ Jk (0) shows that condition (1) need not be satisfied (although (2) is).
The example
" # 0 0
1 0 0 0 1 0
A2 ≡ and B2 ≡
0 0 1 0 0 0
0 1
shows that condition (2) need not be satisfied (although (1) is). Moreover,
the example " # " #
A1 0 B1 0
A≡ and B ≡
0 A2 0 B2
shows that neither condition need be satisfied.
60
Proof Since rank X = rank X = rank φS (X) for any X ∈ Mn and any
S ∈ Sn , one checks that the conditions of Theorem 2.7.4 are satisfied
in each case.
shows that AT A can have a square root without being similar to AAT .
and
T T
B1 = [Jm 1
(0) ⊕ Im1 ] ⊕ · · · ⊕ [Jm t
(0) ⊕ Im1 ],
m1 ≥ · · · ≥ mt ≥ 1, are paired in sizes. It may not be possible to write
" # " #
Hr1 0 Hrq 0
A2 = ⊕ ··· ⊕ (2.8.39)
0 KsT1 0 KsTq
Corollary 2.2.7 now guarantees that (Ak , Bk ) ∼ (S2k (0), S2k (0)).
63
Corollary 2.2.7 now guarantees that (Ak , Bk ) ∼ (S2k−1 (0), S2k−1 (0)).
We have now analysed all of the basic blocks except those in (2.8.41) and
(2.8.42). If A3 and B3 are present, there is no hope of finding a symmetric S
such that (A3 , B3 ) ∼ (S, S) because there must be at least one k for which
rank A3 (B3 A3 )k 6= rank B3 (A3 B3 )k . Notice that it is precisely the presence
of A3 and B3 that is the obstruction to writing A = QS when m = n. The
next two results will permit us to handle this remaining case.
(b1) Ci ∈ Mni −1,ni has rank ni − 1, CiT Ci = Sni (0), and Ci CiT is similar to
Sni −1 (0) for each i ,
(b2) Dj ∈ Mβj ,βj −1 has rank βj − 1, Dj DjT = Sβj (0), and DjT Dj is similar
to Sβj −1 (0) for each j , and
Proof One checks that (A3 , B3 ) ∼ (C, C T ), so that (A, AT ) ∼ (R, RT ), where
" #
S 0
R≡ .
0 C
(2) rank A = rank AT A if and only if the following three conditions hold:
(2a) S2 is absent,
(2b) S3 = 0, or is absent, and
(2c) C may have some Ci but C does not have Dj .
(3) rank (AT A)k = rank (AAT )k for all k = 1, 2, 3, . . . if and only if C = 0.
68
69
We have divided the proof of the theorem into three parts. In Section 3.2,
we establish that either T = 0 or T is nonsingular. In Section 3.3, we show
that T has the form asserted in the theorem, except that Q1 and Q2 are
nonsingular, but are not necessarily orthogonal. Finally, in Section 3.4, the
theorem is proved.
Similar problems have been studied in [HHL] and [HLT], and we use their
general approach. We analyse the orbits
O(A) = {X ∈ Mm,n : X ∼ A}
set Gij ≡ Eij − Eji . Notice that all Fij and Gij are skew-symmetric matrices.
We denote the standard basis of Cn by {e1 , . . . , en }. The n-by-n identity
matrix is denoted by I, or when necessary for clarity, by In . A vector x ∈ Cn
is said to be isotropic if xT x = 0.
dim span O(A) ≤ dim span O(T (A)) and dim TA ≤ dim TT (A)
for every A ∈ V.
Remark The argument used to prove the preceding lemma can be used to
obtain its conclusion under more general hypotheses: Let G be a given group
of nonsingular linear operators on Mm,n , and say A ∼ B if A = L(B) for some
L ∈ G. Let T be a given linear operator on Mm,n such that T (A) ∼ T (B)
whenever A ∼ B. If X ∈ TX and Y 6∈ TY , then T (X) 6= L(Y ) for all L ∈ G.
Proof Let A ∈ Mm,n be a given nonzero matrix. There exists orthogonal (in
fact permutation matrices) P ∈ Mm and Q ∈ Mn such that the (1,1)-
entry of B = P AQ, say b11 , is nonzero. For a given positive integer k,
define the diagonal orthogonal matrix
Note that if X1 , X2 ∈ Mn,k have full rank, Lemma 3.2.5 ensures that
there exists an orthogonal Q ∈ Mn such that X1 = QX2 if and only if
X1T X1 = X2T X2 . In particular, for n ≥ 2 and any given nonzero z ∈ Cn ,
there are two possibilities:
Lemma 3.2.6 Let E11 , E12 , E21 , E22 ∈ Mm,n be standard basis matrices, and
let E ≡ E11 − E22 + iE12 + iE21 . Then
(a) O(E11 ) = {q1 q2T : q1 ∈ Cm , q2 ∈ Cn , and q1T q1 = q2T q2 = 1}.
B ≡ AQ = [ b1 b2 b3 . . . bn ]
BF3j = [ 0 0 − bj 0 . . . 0 b3 0 . . . 0]
and the only matrices in SB with nonzero entries in the third column are
BF13 = [ − b3 0 b1 0 . . . 0 ] and BF23 = [ 0 − b3 b2 0 . . . 0 ]; thus, b3 = 0
and bj is a linear combination of b1 and b2 . Hence, rank A = rank B = 2 if
n > 3. Note that if n = 2 or m = 2 and if A ∈ Mm,n with rank A ≥ 2, then
rank A = 2.
TA = {XA + AY : X ∈ Mm , Y ∈ Mn and X + X T = 0, Y + Y T = 0}
and
(a) dim TA = m + n − 2 if A ∈ {E11 , E11 + iE12 , E11 + iE21 };
(b) dim TA = m + n − 3 if A = E11 − E22 + iE12 + iE21 ;
(c) dim TA ≥ 2n − 3 if rank A ≥ 2. Moreover, if n > 3, rank A ≥ 2,
and dim TA = 2n − 3, then rank A = 2 and there exists a permutation
matrix Q ∈ Mn such that TAQ = span SAQ ;
(d) dim TA ≥ 3n − 6 if rank A ≥ 3.
(e) If rank A = 2 and dim TA = 2n − 3, then there exists a permutation
matrix Q ∈ Mn such that dim span SAQ = 2n − 3.
74
Proof The asserted form of TA , as well as assertions (c) and (d), have been
verified. Assertions (a) and (b) follow from direct computations. We
consider in detail only the case in which A = E11 +iE12 ; the other cases
can be dealt with similarly. Let X = [xij ] ∈ Mm and Y = [yij ] ∈ Mn
be skew-symmetric. Then
0 0 0 ··· 0
x21 ix21 0 ··· 0
XA =
.. .. .. . . ..
. . . . .
xm1 ixm1 0 ··· 0
and
iy21 y12 y13 + iy23 · · · y1n + iy2n
0 0 0 ··· 0
AY =
.. .. .. ... ..
. . . .
0 0 0 ··· 0
Proof Lemma 3.2.1 shows that dim TT −1 (E) ≤ dim TE , while Lemma 3.2.6
and Lemma 3.2.8(a, b) give dim TE ≤ m + n − 2.
Case 1. aT1 q
a1 6= 0
Let α = aT1 a1 , so α 6= 0. Lemma 3.2.5 ensures that Qa1 = [ α 0 . . . 0 ]T
for some orthogonal Q ∈ Mm , so
" #
α ∗
QA = .
0 A1
h i
If we write A1 ≡ b2 b3 . . . bn , then b2 6= 0 since a1 and a2 are linearly
independent. There are two possibilities: b2 is isotropic or not.
(i) Suppose bT2 b2 6= 0. After applying the preceding argument to A1 , we
see that there exists an orthogonal Q1 ∈ Mm such that
α δ ∗
B ≡ Q1 A = 0 β ∗ , α, β, δ ∈ C, with αβ 6= 0,
0 0 A2
76
0 b2 ∗
Notice that the second column of QA is isotropic and independent of the first
column.
(i) If b2 = 0, then there is δ 6= 0 such that
α δ ∗
B ≡ QA = iα −iδ ∗ .
0 0 0
Let ψ = {(2, 2), (2, 3), (3, 3), (3, 4), (4, 4)}. If E ∈ Mm,n with n ≥ m and
(m, n) 6∈ ψ, we have shown that rank T −1 (E) = 1 whenever rank E = 1. We
treat the five special cases (m, n) ∈ ψ separately. We use the following two
results.
Lemma 3.3.7 Let A ∈ Mm,n . If dim TA = dim span SAQ for some orthog-
onal Q ∈ Mn , then AAT = αIm for some α ∈ C.
Proof Suppose dim TA = dim span SAQ . Since dim TA = dim TAQ , we have
TAQ = span SAQ . Hence, for each skew-symmetric Y ∈ Mm there
exists a skew-symmetric X ∈ Mn such that Y AQ = AQX. Thus
Y AAT = Y AQT (AQT )T = (AQ)X(AQ)T is skew-symmetric for every
skew-symmetric Y , and hence AAT = αIm by Lemma 3.3.6.
Proof Suppose E ∈ M2,3 has rank 1. Since rank T −1 (E) ≤ m = 2, there are
only two possibilities: rank T −1 (E) = 1 (which is the assertion of Proposi-
tion 3.3.1), or rank T −1 (E) = 2. We wish to exclude the latter possibility.
We look at
D ≡ {Y ∈ M2,3 : dim TY ≤ 3}.
Since dim TT −1 (Y ) ≤ dim TY , it must be the case that T −1 (Y ) ∈ D whenever
Y ∈ D. If E has rank 1, then Lemma 3.2.8(a, b) shows that E ∈ D. Suppose
79
Thus, {E12 , E13 , E21 } ⊂ O(E11 ) ∩ TE11 so that O(E11 ) ∩ TE11 contains at
least three vectors. Moreover, χ ≡ {T1 (E12 ), T1 (E13 ), T1 (E21 )} contains
80
Proof Let E ∈ M3,4 have rank 1 and let A ≡ T −1 (E). Lemma 3.3.4 shows
that rank A is either 1 or 2. Suppose rank A = 2. We apply the analysis of
the proof of Lemma 3.3.5 to A. Notice that Cases 1(i, ii) and 2(i, iii) are
not possible here. Thus, A has the form considered in Case 2(ii):
α δ x1 x4
A = iα β x2 x5
0 λ x3 x6
with λ 6= 0 and all the columns of A are isotropic vectors. It now follows
from Lemma 3.2.8(a, c) and Lemma 3.2.1 that dim TA = 5. Moreover,
Lemma 3.2.8(e) ensures that dim span SAP = 5 for some permutation matrix
P ∈ M4 . Hence, Lemma 3.3.7 guarantees that there exists a ∈ C such that
AAT = aI3 . Since it is always the case that rank A ≥ rank AAT , we must
have a = 0, that is, the rows of A are also isotropic vectors. Thus, there exists
an orthogonal Q ∈ M4 such that the first row of AQ is [ α iα 0 0 ]. Since
AAT = 0, the third row of AQ has the form [ 0 0 c d ], where c2 + d2 = 0
and d 6= 0. It follows that
α iα 0 0
AQ = iα −α ebd ed
0 0 bd d
that G12 AQ = AQ(a1 F12 + a2 F13 + a3 F14 + a4 F23 + a5 F24 ). Examining the
(1, 1), (2, 1), (3, 1) and (2, 4) entries, we get a1 = −1, a2 = a3 = a5 = 0. This
is a contradiction since for any a4 , G12 AQ 6= −AQF12 + a4 AQF23 . Hence,
dim TA = dim TAQ ≥ 6, again a contradiction. Since we can exclude all the
possibilities associated with rank A = 2, it follows that rank A = 1.
Proof First note that Xn ≡ E11 − E22 + iE12 + iE21 = Xn (iF12 ) ∈ TXn for
all n = 2, 3, . . ., where Eij , Xn ∈ Mn . Note also that, TIn = {X ∈ Mn :
82
xT AT Ax = xT P T AT AP x
Proof Under the stated assumptions, Proposition 3.3.12 ensures that there
exist nonsingular M ∈ Mm and N ∈ Mn such that either T (A) = MAN
for all A ∈ Mm,n , or m = n and T (A) = M AT N for all A ∈ Mm,n . We
consider only the case T (A) = M AN ; the case T (A) = MAT N can
be dealt with similarly. Let an orthogonal P ∈ Mn be given. Since
T preserves orthogonal equivalence, for each A ∈ Mm,n there exist
orthogonal matrices Q ∈ Mm and Z ∈ Mn (which depend on A and P )
such that T (A) = Q T (AP ) Z. Hence,
85
86
Notice that the singular values of any A = [aij ] ∈ K4 (IF) are uniquely
P
determined by tr A∗ A = ni,j=1 |aij |2 and |det A|2 , and hence these two
invariants determine A ∈ K4 (IF) uniquely up to unitary t-congruence.
Thus, X, Y ∈ K4 (IF) are unitarily t-congruent if and only if X + is
unitarily t-congruent to Y + . In addition, X and X + are unitarily t-
congruent whenever X ∈ K4 (IF).
87
In [HHL], the authors studied the linear operators T on Mn (IF) that pre-
serve t-congruence, that is, T (A) is t-congruent to T (B) whenever A and B
are t-congruent. At the end of the paper they raised the question of charac-
terizing the linear operators on Mn (IF) that preserve unitary t-congruence.
The purpose of this chapter is to answer their question with the following
results.
T (X) = µU (X + X t )U t + νU (X − X t )U t
As we shall see, even though the result in the complex case (Theo-
rem 4.1.1) is very similar to Theorem (4.1) of [HHL], it requires a refinement
of their proof and several additional ideas to obtain our result. The real case
is even more interesting. Our Theorems 4.1.2 and 4.1.3 are quite different
from the result in [HHL]. The complication is due to the fact that Mn (IR)
can be decomposed as a direct sum of the subspaces
Λn ≡ {λI : λ ∈ IR},
Sn (IR) ≡ {A ∈ Mn (IR) : A = At and tr A = 0}, and
0
4.2 Preliminaries
The following three results can be found in [HHL]. By private correspon-
dence, the author has learned that Professor M. H. Lim has obtained different
90
(b) There exist U ∈ Un (C) and scalars µ, ν ∈ IR with (µ, ν) 6= (0, 0) such
that
T (X) = µU XU t + ν(tr X)I for all X ∈ Sn (IR).
Lemma 4.2.6 span O(A) = Sn (C) for every nonzero A ∈ Sn (C). Conse-
quently, if T is a linear operator on Mn (C) that preserves unitary t- congru-
ence, then ker T ∩ Sn (C) 6= {0} if and only if Sn (C) ⊂ ker T .
Lemma 4.2.7 span O(A) = Sn0 (IR) for every nonzero A ∈ Sn0 (IR). Conse-
quently, if T is a linear operator on Mn (IR) that preserves orthogonal simi-
larity, then ker T ∩ Sn0 (IR) 6= {0} if and only if Sn0 (IR) ⊂ ker T .
Proof Suppose A ∈ Sn0 (IR). We have already shown that span O(A) ⊂
Sn0 (IR). We now prove that Sn0 (IR) ⊂ span O(A). There exist Q ∈
Un (IR) and α1 , α2 , . . . , αn−1 , αn ∈ IR such that
so E11 − Enn ∈ span O(A). Since Eij + Eji ∈ O(E11 − Enn ), we have
Eij + Eji ∈ span O(A) for all 1 ≤ i < j ≤ n. The conclusion follows.
Lemma 4.2.8 Let A ∈ Sn (IR) be given. Suppose that A 6∈ Λn ∪Sn0 (IR). Then
span O(A) = Sn (IR). Consequently, if T is a linear operator on Mn (IR) that
preserves orthogonal similarity, then ker T ∩ Sn (IR) 6= {0} if and only if
Sn (IR) ⊂ ker T .
92
Proof We follow and refine slightly the proof of Lemma (4.2) of [HHL]. Let
{Pi : 1 ≤ i ≤ 2n } denote the group of all real diagonal orthogonal
matrices in Mn (IF) and define
2n
X
P(X) ≡ Pi XPit for X ∈ Mn (IF).
i=1
T (X) + T (X)t
L(X) ≡ for all X ∈ Kn (C).
2
One checks that L is a linear operator that preserves unitary t-congruence.
Moreover, if for some nonzero B ∈ Kn (C), L(B) = 0, then Lemma 4.2.6
guarantees that L(Kn (C)) = 0. Notice that T (Kn (C)) ⊂ Kn (C)) if and
only if L = 0. Suppose then that L(A) 6= 0 for some (necessarily) nonzero
A ∈ Kn (C), that is, suppose L is nonsingular. There exist U ∈ Un (C) and a
scalar a1 > 0 such that
U AU t = a1 F12 + C (4.3.1)
where C ≡ 02 ⊕ C1 ∈ Kn (C) for some C1 ∈ Kn−2 (C). Notice that for any
µ ∈ C with |µ| = 1, ZAZ t = a1 F12 +µ2 C is unitarily t-congruent to A, where
Z ≡ (I2 ⊕ µIn−2 )U ∈ Un (C). Hence,
rank L(A) = rank L(a1 F12 + µC) = rank (a1 L(F12 ) + µL(C))
94
The analysis (for the complex case) of the proof of Lemma (4.4) of [HHL]
can be modified to show that this is not possible.
Hence, for this case, we may regard T as a nonzero linear operator on Kn (C).
Lemma 4.2.1 guarantees that Theorem 4.1.1 holds for this case.
Suppose now that n = 2 and let 0 6= A ∈ ker T ∩ Sn (C). Again,
Lemma 4.2.6 guarantees that T (Sn (C)) = 0. Let
T (F12 )
A0 ≡ − .
2
Since X − X t = −(tr XF12 )F12 for all X ∈ M2 (C), we have
T (X) = 12 [T (X + X t ) + T (X − X t )]
= 12 T (X − X t )
= − 12 (tr XF12 )T (F12 )
= (tr XF12 )A0 .
Let us now consider the nonsingular T . Lemmata 4.2.1, 4.2.2, and 4.3.5
imply that there exist nonzero scalars µ, ν ∈ C and V1 , V2 ∈ Un (C) such that
and either
By considering various choices for X and Y , we will show that we may take
V2 = βV1 in (A1), for some nonzero β ∈ C. Moreover, we will show that
(A2) cannot happen.
96
and if n = 2k we let
Y ≡ B ⊕ 2B ⊕ · · · ⊕ kB ∈ Kn (C),
Y ≡ B ⊕ 2B ⊕ · · · ⊕ kB ⊕ 0 ∈ Kn (C),
X + Y = W (U V XV t U t + V U Y U t V t )W t (4.3.2)
X = W U V XV t U t W t = (W U V )X(W U V )t (4.3.3)
and
Y = W V U Y U t V t W t = (W V U )Y (W V U )t . (4.3.4)
97
X(W U V ) = (W U V )X (4.3.5)
Y (W V U ) = (W V U )Y. (4.3.6)
or
W V U = W11 ⊕ W22 ⊕ · · · ⊕ Wkk ⊕ Wk+1,k+1 if n = 2k + 1,
where Wii ∈ U2 (C) for each i = 1, 2, . . . , k and Wk+1,k+1 ∈ U1 (C).
Since U was arbitrary, we may choose U so that U ∗ V ∗ U is diagonal.
Now, W U V = W V U (U ∗ V ∗ U )V . Hence V = (U ∗ V ∗ U )∗ (W V U )∗ (W U V ) is
block-diagonal and has the same form as W V U , that is, either
V = V1 ⊕ V2 ⊕ · · · ⊕ Vk if n = 2k, (4.3.7)
or
V = V1 ⊕ V2 ⊕ · · · ⊕ Vk ⊕ Vk+1 if n = 2k + 1, (4.3.8)
where Vi ∈ U2 (C) for each i = 1, 2, . . . k and Vk+1 ∈ U1 (C).
We apply the same analysis to the same choices for X ∈ Sn (C) and
U ∈ Un (C), but this time we take
and
Y1 (W1 V U ) = (W1 V U )Y1 . (4.3.10)
The identity (4.3.9) implies that D1 ≡ W1 U V is diagonal and the identity
(4.3.10) implies that W1 V U has the form
A11 0 A13
W1 V U = 0 A22 0
,
A31 0 A33
The proof of Lemma (4.9) of [HHL] can be used (again, verbatim) to prove
the following result.
and
Y (W1 V U1 ) = (W1 V U1 )Y. (4.3.20)
Again, W1 U1 V is diagonal and the identity (4.3.20) shows that W1 V U1 has
the form
z11 0 0 z14
0 z 0
22 z23
. (4.3.21)
0 z32 z33 0
z41 0 0 z44
Now,
U1∗ V = V (W1 U1 V )∗ (W1 V U1 )U1∗ . (4.3.22)
Since W1 U1 V and U1 are diagonal, D ≡ (W1 U1 V )∗ (W1 V U1 )U1∗ has the form
(4.3.21). Since V has the form (4.3.17) or (4.3.18), the identity (4.3.22) shows
that D is diagonal, that is, z14 = z23 = z32 = z14 = 0. Now choose U1 having
distinct (diagonal) elements. Then the identity (4.3.22) implies that each of
the Vij in (4.3.17) and (4.3.18) has the form either
" #
ν1 0
(4.3.23)
0 ν2
or " #
0 ν4
. (4.3.24)
ν3 0
We will show that neither of these choices is possible.
To reiterate, we are assuming
1
T (X) ≡ (X + X t ) + αV (X − X t )+ V t for all X ∈ M4 (C), α 6= 0
2
where V ∈ U4 (C) has either the form (4.3.17) or (4.3.18), and each Vij has
the form either (4.3.23) or (4.3.24). Notice that if T (A1 ) = W T (A2 )W t , then
1 1
(A1 + At1 ) = W (A2 + At2 )W t (4.3.25)
2 2
and
1 1
V (A1 − At1 )+ V t = W V (A2 − At2 )+ V t W t . (4.3.26)
2 2
102
" #
0 V12
Case 1 V ≡ .
V21 0
Let
1 0 1 0 1 1 0 0
0 0 0 0 −1 0 0 0
A1 ≡ and A2 ≡ .
−1 0 0 0 0 0 0 0
0 0 0 2 0 0 0 2
With " #
0 1
U ≡1⊕ ⊕ 1 ∈ U4 (C), (4.3.27)
1 0
one checks that A1 = U A2 U t . Hence, there exists a W ∈ U4 (C) such that
T (A1 ) = W T (A2 )W t . Now the identity (4.3.25) shows that W has the form
w11 0 0 0
0 w22 w23 0
W ≡ .
0 w32 w33 0
0 0 0 w44
The identity (4.3.26) shows that
" # " #
0 −V21 BV12t 0 0
t =W W t, (4.3.28)
V12 BV21 0 0 V21 CV21t
where " # " #
0 0 0 1
B≡ and C ≡ .
0 1 −1 0
Since V21 CV21t = βC, where β ≡ det V21 6= 0, direct computation shows that
" # 0 0 0 0
0 0 0 0 0 βw23 w44
W Wt =
. (4.3.29)
0 V21 CV21t 0 0 0 βw33 w44
0 −βw23 w44 −βw33 w44 0
Now, use (4.3.28) and (4.3.29) to get
βw33 w44 = 0 (4.3.30)
and " # " #
0 0 0 0
V12 BV21t = V12 V21t = . (4.3.31)
0 1 0 βw23 w44
103
where β ≡ u11 u32 ν1 ν2 and δ ≡ u23 u44 ν3 ν4 . Since W1 and V are unitary,
|β| = |δ| = |νi νj | = 1 for all 1 ≤ i, j ≤ 2, and hence (4.3.34) is impossible.
One checks that A1 and A2 are unitarily t-congruent. Hence, there exists
a W ∈ U4 (C) such that T (A1 ) = W T (A2 )W t . As in Case 1, the identities
(4.3.25) and (4.3.26) hold. The identity (4.3.25) now shows that W has the
form
" # " #
w11 0 w33 w43
W ≡ W1 ⊕ W2 , with W1 = and W2 = .
0 w22 w34 w44
Since V1 has the form either (4.3.23) or (4.3.24) and W1 is diagonal, V1∗ W1 V1
is diagonal. Thus, if we write (V2∗ W2∗ V2 )t = [uij ], (4.3.36) shows that u21 =
u22 = 0. Hence, V2∗ W2∗ V2 is singular, a contradiction to the fact that both V2
and W2 are unitary.
n(n + 1) n(n − 1)
dim Sn0 (IR) = −1> = dim Kn (IR),
2 2
there exists a nonzero B ∈ Sn (IR) such that T1 (A) = B for some nonzero
A ∈ Sn0 (IR). It follows that T1 (Sn0 (IR)) ⊂ Sn (IR). Since Λn ⊕Sn0 (IR) = Sn (IR),
we may regard T1 as a mapping from Sn (IR) to Mn (IR). Define
T1 (X) + T1 (X)t
TS (X) ≡ for all X ∈ Sn (IR)
2
and
T1 (X) − T1 (X)t
TK (X) ≡ for all X ∈ Sn (IR).
2
Notice that TS (Sn (IR)) ⊂ Sn (IR). Lemma 4.2.4 guarantees that TS satisfies
one of the following conditions:
(1a) There exists B1 ∈ Sn (IR) such that TS (X) = (tr X)B1 for all X ∈
Sn (IR); or
(1b) There exist U ∈ Un (IR) and α, β ∈ IR such that TS (X) = α(tr X)I +
βU XU t for all X ∈ Sn (IR).
For every nonzero A ∈ Sn0 (IR), T1 (A) is nonzero and symmetric. It follows
that TS (A) = T1 (A) 6= 0 for all A ∈ Sn0 (IR). Hence (1a) cannot happen.
Notice also that TK (Sn (IR)) ⊂ Kn (IR), so that Sn0 ⊂ ker TK . Thus,
U XU t ≡ diag (1, . . . , k, . . . , h, . . . , n)
and
U Y1 U t ≡ diag (1, . . . , h, . . . , k, . . . , n).
Notice that X and Y1 are orthogonally similar. Hence, there exists W ∈
Un (IR) such that T1 (X) = W T1 (Y1 )W t . It follows that
U XU t = W (U Y1 U t )W t (4.4.37)
and
A1 = W A1 W t . (4.4.38)
The identity (4.4.37) and the fact that W ∈ Un (IR) show that W ≡ [wij ] has
entries wkh , whk , wii = ±1 where i = 1, . . . , n, i 6= h and i 6= k, and wij = 0
otherwise. The k th and hth rows of the identity (4.4.38) show that
Since ahk = −akh 6= 0, the identity (4.4.39a) shows that wkh whk = −1. The
identities (4.4.39b) and (4.4.39c) now show that ahi = aki = 0 for all integers
h 6= i 6= k.
Since n ≥ 3, there is an integer t such that 1 ≤ t ≤ n, t 6= k, and t 6= h.
Suppose t < k and this time consider
Notice that X and Y2 are also orthogonally similar. Repeat the previous
analysis and conclude that akh = 0. Hence, A1 = 0, as desired.
108
One checks that if a given linear operator L on Mn (IR) satisfies any one
of the three conditions (A1) - (A3), then L preserves orthogonal similarity.
We now consider T2 . Recall that either T2 = 0, or T2 is nonsingular on
Kn (IR). Hence, T2 (A) 6= 0 for some (necessarily) nonzero A ∈ Kn (IR) if and
only if Kn (IR) ∩ ker T2 = {0}.
T2 (X) + T2 (X)t
TS (X) ≡ for all X ∈ Kn (IR),
2
and
T2 (X) − T2 (X)t
TK (X) ≡ for all X ∈ Kn (IR).
2
Notice that for any X ∈ Kn (IR), we have T2 (X) = TS (X) + TK (X).
Moreover, TS (Kn (IR)) ⊂ Sn (IR) and TK (Kn (IR)) ⊂ Kn (IR). We will show
that if n ≥ 3, then TS = 0 so that Lemma 4.2.3 ensures that either (B1) or
(B2) holds. If n = 2, we will show that T2 satisfies (B3).
Suppose TS 6= 0. Since TS also preserves orthogonal similarity, TS is
nonsingular. Since −A ∈ O(A) for any A ∈ Kn (IR), we have −TS (A) ∈
O(TS (A)). Hence, if λ is an eigenvalue of TS (A) with multiplicity k, then −λ
is also an eigenvalue of TS (A) with multiplicity k. Moreover, if A 6= 0, then
TS (A) has a positive eigenvalue.
109
(r − 1)(2n − r + 2)
k ≥ κ(r) ≡ , (4.4.40)
2
then Theorem (1) of [FL] guarantees that Z contains a nonzero matrix B
whose largest eigenvalue has multiplicity at least r. This result, together with
the fact that the eigenvalues of TS (A) are paired, will give us a contradiction.
Let n = 2t + 1, where t = 1, 2, . . .. Then 2 ≤ t + 1 ≤ n − 1 and
n(n − 1) 3t(t + 1)
k= = t(2t + 1) ≥ = κ(t + 1).
2 2
Hence, there exists a nonzero B ∈ Im T whose largest eigenvalue λ has
multiplicity at least t + 1. Since B ∈ Im T and B 6= 0, we must have λ > 0.
But since −λ is also an eigenvalue of B with multiplicity at least t + 1, this
is a contradiction.
Now let n = 2t, where t = 3, 4, . . .. Then 2 ≤ t + 1 ≤ n − 1 and
n(n − 1) t(3t + 1)
k= = t(2t − 1) ≥ = κ(t + 1).
2 2
Hence, there exists a nonzero B ∈ Im T whose largest eigenvalue is positive
and has multiplicity at least t + 1. Again, this is a contradiction.
Suppose now that n = 4. Notice that dim Im TS = 6 > 4 = κ(2).
Hence, Im TS contains a matrix of the form Y0 ≡ Qdiag (a, a, −a, −a)Qt
where Q ∈ U4 (IR) and a > 0. We will show that this cannot happen.
One checks that the tangent space at any A ∈ M4 (IR) is given by
TA ≡ {XA + At X t : X ∈ K4 (IR)}.
Moreover,
dim TY0 = dim TQt Y0 Q = 2
and
dim TA ≥ 4 for any nonzero A ∈ K4 (IR).
Since TS is nonsingular on K4 (IR), it must be that
where we put A0 ≡ − 12 T (F12 ). Since F12 and −F12 are orthogonally similar,
A0 is orthogonally similar to −A0 . Hence, tr A0 = tr (−A0 ) = −tr A0 , so
that tr A0 = 0. Notice that if A0 ∈ K2 (IR), then condition (B1) also holds.
and
0 0 0
1 1 t
X4 ≡ I + V Q( 0 0 1 t
⊕ 0n−3 )Q V ∈ Mn (IR).
n 2c
0 −1 0
One checks that X3 and X4 are orthogonally similar. Hence T (X3 ) is orthog-
onally similar to T (X4 ). It follows that
a1 0 1 a1 0 0
B1 ≡
0 a2 0 and B2 ≡ 0 a2 1
−1 0 a3 0 −1 a3
Proof The forward implication is easily verified. For the converse, notice
that the conditions imply that A and B have the same eigenvalues and
singular values. Hence, the matrices Aα ≡ A + αI and Bα ≡ B + αI
also satisfy the three conditions for any α ∈ IR. Notice that A and B
are orthogonally similar if and only if there is some α ∈ IR such that Aα
and Bα are orthogonally similar. Thus, we may assume that tr A 6= 0
and det A 6= 0. Since A and B have the same singular values, there
exist X1 , X2 , V1 , V2 ∈ U2 (IR) such that
X1 AX1t = ΣX2 and V1 BV1t = ΣV2 , (4.4.41)
where Σ ≡ diag (σ1 , σ2 ) and σ1 ≥ σ2 > 0. Since the determinant is
invariant under orthogonal similarity, it follows from condition (a) and
(4.4.41) that det (X2 ) = det (V2 ). Now, any U ∈ U2 (IR) can be written
as " #
cos θ sin θ
U≡ , for some θ ∈ IR.
∓ sin θ ± cos θ
113
Hence, (b) and the fact that tr A 6= 0 show that X2 and V2 have
the same diagonal elements. Direct verification now shows that X2 =
W V2 W t , where W ≡ diag (1, ±1) ∈ U2 (IR). Hence,
ΣX2 = ΣW V2 W t = W ΣV2 W t .
Suppose K2 (IR) 6⊂ ker T . Assume further that S20 (IR) ⊂ ker T . Write
1 1
X = (tr X)I + XS 0 − (tr XF12 )F12 . (4.4.42)
2 2
Then XS 0 ∈ S20 (IR) and hence
Then we have det T (X1 ) = det T (X2 ). Direct computation shows that this
implies tr A0 B0 = 0. Moreover, we have
For this case, it follows that tr At0 B0 = 0. Hence, T satisfies condition (1) of
Theorem 4.1.3.
Suppose S20 (IR) 6⊂ ker T and write X as in (4.4.42). Since the restriction
of T = T1 + T2 to S20 (IR) is nonsingular, it follows from Lemma 4.4.2 that
115
116
and
Y ≡ [y Ay · · · Ak−1 y] ∈ Mn,k .
Since x∗ Ak−1 y 6= 0, the columns of
∗ x∗ Ak−1 y
·
X1 Y =
∈ Mk
·
·
x∗ Ak−1 y 0
are independent, so X1 Y is nonsingular (if k = n, this means X1 is nonsin-
gular). Moreover,
X1 A = Jk (0)X1 (A.0.1)
and
AY = Y JkT (0). (A.0.2)
If k = n we have X1 A = Jn (0)X1 and X1 is nonsingular, so Jn (0) = X1 AX1−1 .
If k < n, Lemma A.1 ensures that there is an X2 ∈ Mn−k,n such that
X2 Y = 0 (A.0.3)
and " #
X1
X≡ ∈ Mn is nonsingular.
X2
Set B ≡ (X2 A)X −1 ∈ Mn−k,n and partition B = [B1 A2 ] with B1 ∈ Mn−k,k
and A2 ∈ Mn−k . Notice that
" #
X1
X2 A = BX = [B1 A2 ] = B1 X1 + A2 X2 . (A.0.4)
X2
Now use (A.0.4) and the identities (A.0.3) and (A.0.2) to compute
and
X2 AY = X2 (AY ) = X2 (Y JkT (0)) = (X2 Y )JkT (0) = 0.
Since X1 Y is nonsingular, we conclude that B1 = 0. Thus, (A.0.4) simplifies
to
X2 A = A2 X2 (A.0.5)
118
where each Ti ∈ Mni is upper triangular with all diagonal entries equal to λi ,
i = 1, . . . , r.
For a block matrix of the form (A.0.6), notice that each Ti − λi Ini is
nilpotent for i = 1, . . . , r. Lemma A.2 guarantees that there exists a non-
singular Si ∈ Mni and positive integers mi , i1 ≥ · · · ≥ imi ≥ 1 such
that Ji1 (0) ⊕ · · · ⊕ Jimi (0) = Si (Ti − λi Ini )Si−1 = Si Ti Si−1 − λi Ini . Hence,
Si Ti Si−1 = Ji1 (λi )⊕· · ·⊕Jimi (λi ). If we now consider the similarity of (A.0.6)
obtained with the direct sum S1 ⊕ · · · ⊕ Sr , we obtain the desired result:
Once one has the Jordan form, its uniqueness (up to permutation of the
diagonal blocks) can be established by showing that the sizes and multiplici-
ties of the blocks are determined by the finitely many integers rank (A−λI)k ,
λ an eigenvalue of A, k = 1, . . . , n; see the proof of Theorem (3.1.11) in [HJ1].
Bibliography
120
121
[Ho] Y.P. Hong. A Canonical Form Under φ—Equivalence. Linear Alg. Appl.
147 (1991), 501—549.
[HH1] Y.P. Hong and R. A. Horn. A Canonical Form for Matrices Under
Consimilarity. Linear Algebra Appl. 102 (1988), 143—168.
[HHL] Y.P. Hong, R. A. Horn, and C.K. Li. Linear Operators Preserving
t—Congruence on Matrices. Linear Algebra Appl. to appear.
[HLT] R. A. Horn, C.K. Li, and N.K. Tsing. Linear Operators Preserving
Certain Equivalence Relations on Matrices. SIAM J. Matrix Analysis
Appl. 12 (1991), 195—204.
[LRT] C.K. Li, L. Rodman, and N.K. Tsing. Linear Operators Preserving
Certain Equivalence Relations Originating in System Theory. Linear
Algebra Appl. 161 (1992), 165—225.
[LT1] C.K. Li and N.K. Tsing. G—Invariant Norms and G(c)—Radii. Linear
Algebra Appl. 150 (1991), 150—179.
[LT2] C.K. Li and N.K. Tsing. Linear Preserver Problems: A Brief Intro-
duction and Some Special Techniques. Linear Algebra Appl. 162—164
(1992), 217—236.
123
124
DENNIS I. MERINO
817 Scarlett Drive
Baltimore, MD 21204
(301) 296 3620
e-mail: [email protected]
Personal
Education / Awards
Employment Experience
1986-1988 Instructor
Department of Mathematics
The University of the Philippines, Quezon City, Philippines
Taught elementary courses in Analysis and Algebra.
Chairman of the Student-Faculty Relations Commitee, 1987.
Trainer for the participants in the International
Mathematical Olympiad, 1988
Publications
R. A. Horn and D. I. Merino. A Real-Coninvolutory Analog of the
Polar Decomposition. (preprint)
References
Dr. Roger A. Horn
Department of Mathematical Sciences
The Johns Hopkins University
Baltimore, MD 21218
Dr. Chi-Kwong Li
Department of Mathematics
The College of William and Mary
Williamsburg, VA 23185
Dr. Alan J. Goldman
Department of Mathematical Sciences
The Johns Hopkins University
Baltimore, MD 21218