SciELO - Scientific Electronic Library Online

 
vol.46 número1A Note on the L1-Mean Ergodic TheoremManuel Sadosky (1914-2005) índice de autoresíndice de assuntospesquisa de artigos
Home Pagelista alfabética de periódicos  

Serviços Personalizados

Journal

Artigo

Indicadores

  • Não possue artigos citadosCitado por SciELO

Links relacionados

  • Não possue artigos similaresSimilares em SciELO

Compartilhar


Revista de la Unión Matemática Argentina

versão impressa ISSN 0041-6932versão On-line ISSN 1669-9637

Rev. Unión Mat. Argent. v.46 n.1 Bahía Blanca jan./jun. 2005

 

Some operator inequalities for unitarily invariant norms*†

Cristina Cano, Irene Mosconi and Demetrio Stojanoff§

Partially supported by Universidad Nac. del Comahue

§ Partially supported CONICET (PIP 4463/96), UNLP 11 X350 and ANPCYT (PICT03-09521)

Abstract:
Let L(H)  be the algebra of bounded operators on a complex separable Hilbert space H  . Let N  be a unitarily invariant norm defined on a norm ideal I ⊆ L(H)  . Given two positive invertible operators P,Q  ∈ L(H)  and k ∈ (- 2,2]  , we show that   (                       )  N  P T Q- 1 + P -1T Q + kT  ≥ (2 + k)N (T )  , T ∈ I  . This extends Zhang's inequality for matrices. We prove that this inequality is equivalent to two particular cases of itself, namely P =  Q  and        -1  Q  = P   . We also characterize those numbers k  such that the map Υ : L(H)  →  L(H)  given by               -1    -1  Υ(T  ) = P T Q  + P   T Q +  kT  is invertible, and we estimate the induced norm of Υ -1   acting on the norm ideal I  . We compute sharp constants for the involved inequalities in several particular cases.

Keywords and phrases: Positive matrices; Inequalities; Unitarily invariant norm.

AMS Subject Classification: Primary 47A30, 47B15.

1. INTRODUCTION

Let H  be a Hilbert space and denote by L(H)  the algebra of bounded linear operators on H  . In 1990, Corach-Porta-Recht [6] show that, for every invertible selfadjoint operator S ∈  L(H)  and for every T ∈ L(H)  , it holds that
∥∥     -1     -1   ∥∥    ∥∥  ∥∥   ST  S   + S   TS   ≥ 2  T .
(1)

Several authors have proved generalizations and alternative proofs of inequality (1). For example, Bhatia and Davis [3], Kittaneh (in two different ways, see [11] and [12]), and Andruchow, Corach and Stojanoff [2]. On the other hand, in 1993, Livshits and Ong [14] studied the invertibility of the map T  ↦-→  ST S -1 + S -1T S  for not necessarily selfadjoint S ∈  L(H)  . In 2001, Seddik, [15] proved that, for S ∈ L(H)  invertible and positive, T  ∈ L(H),  and k = 0, 1,2  ,
∥                      ∥          ∥  ∥  ∥kT + ST  S-1 + S- 1TS ∥ ≥ (k + 2)∥T ∥.
(2)

In 1999 Zhan [18] showed that, given two positive invertible matrices P,Q  ∈ Mn( ℂ)  , T  ∈ Mn( ℂ)  , and k ∈ (- 2,2]  , then
9P T Q -1 + P- 1T Q +  kT9  ≥ (2 + k) 9 T9,
(3)

for every unitarily invariant norm 9 ⋅9 on Mn( ℂ)  .

In this paper we work with unitarily invariant norms defined in some ideal I  of L(H)  (see Remark 2.1 or Simon's book [17]). We show that, for every unitarily invariant norm 9 ⋅9 , the following inequalities are equivalent, for every k ∈ (- 2,2]  :
9P T Q- 1 + P -1T Q + kT 9 ≥ (2 + k) 9 T 9   for   P, Q ∈  Gl(H)+,  T ∈ I .
(4)

9ST S +  S-1T S- 1 + kT 9 ≥ (2 + k) 9 T 9    for   S ∈ Gl(H)+,   T ∈ I .
(5)

9kT  + ST S- 1 + S -1T S9 ≥ (k + 2) 9 T 9    for   S ∈ Gl(H)+,   T ∈ I .
(6)

We give a proof of inequality (6), using a technical result about unitarily invariant norms, which allows us to obtain a reduction to the matricial case. In this case, we use a result of Bhatia and Parthasarathy [4], and some properties of the Hadamard product of matrices. This result was previously proved for k =  0  in [2], for not necessarily positive S  , P  and Q  . We study the operators associated to the three mentioned inequalities, and their restriction as operators on the norm ideal I  . We compute their spectra and, in some cases, their reduced minimum moduli (also called conorms). The rest of the paper deals with the estimation of sharp constants for inequality (5), with respect to the usual norm of L(H)  . We get the optimal constant, if one restricts to operators T ∈ L(H)+   . Using the notion of Hadamard index for positive matrices, studied in [7], we compute, for a fixed S ∈ Gl(H)+   , the constant

                                  -1   -1                                    +  M (S, k) = max{C  ≥  0 : ∥ST S + S  T S   + kT ∥ ≥ C ∥T ∥ for every T ∈ L(H)   },

for k ≥  0  (see Proposition 5.6). Finally, we give some partial results for T  ∈ Mn( ℂ)  , in lower dimensions, showing numerical estimates of sharp constants. For n = 2, 3  and 4  , we characterize the best intervals Jn  such that the inequality (6) holds in Mn( ℂ)  for every k ∈ Jn  .

In section 2, we fix several notations and state some preliminary results. We expose with some detail the theory of unitarily invariant norms defined on norm ideals of L(H)  , proving some technical results in this area. In section 3, we show the equivalence of the mentioned inequalities and we give the proof of (6). In section 4, we study the associated operators. In section 5, we describe the theory of Hadamard index, and we use it to obtain a description of the constant M (S, k)  . In section 6 we study the case of matrices of lower dimensions.

We wish to acknowledge Prof. G. Corach who shared with us fruitful discussions concerning these matters.

2. PRELIMINARIES

Let H  be a separable Hilbert space, and L(H)  be the algebra of bounded linear operators on H  . We denote L0(H)  the ideal of compact operators, Gl(H)  the group of invertible operators, L(H)h  the set of hermitian operators, L(H)+   the set of positive definite operators, U(H)  the unitary group, and Gl(H)+   the set of invertible positive definite operators.

Given an operator A ∈ L(H)  , R(A)  denotes the range of A  , kerA  the nullspace of A  , σ(A)  the spectrum of A  , A * the adjoint of A  , ∣A ∣ = (A *A)1∕2   the modulus of A  , ρ(A)  the spectral radius of A  , and ∥A ∥ the spectral norm of A  . Given a closed subspace S  of H  , we denote by P    S   the orthogonal projection onto S  .

When dim H  = n <  ∞ , we shall identify H  with  n  ℂ  , L(H)  with Mn( ℂ)  , and we use the following notations: H(n)  for L(H)h  ,        +  Mn( ℂ)   for      +  L(H)   , U(n)  for U(H)  , and Gl(n)  for Gl(H)  . A norm 9 ⋅9 in Mn( ℂ)  is called unitarily invariant if 9 U AV 9 =  9A9 for every A ∈  Mn( ℂ)  and U, V ∈ U(n)  .

Remark 2.1. The notion of unitarily invariant norms can be defined also for operators on Hilbert spaces. We give some basic definitions (see Simon's book [17]): Let A  ∈ L0(H)  . Then also ∣A ∣ ∈ L0(H)  . We denote by s(A) = (sk(A))k∈ℕ   , the sequence of eigenvalues of ∣A ∣ , taken in non increasing order and with multiplicity. If dim  R(A)  = n < ∞ , we take s (A)  = 0   k  for k > n  . The numbers s (A)   k  are called the singular values of A  .

Denote by C0   the set of complex sequences which converge to zero. Consider CF ⊆  C0   the set of sequences with finite non zero entries. For a ∈ C0   , denote ∣a∣ = (∣an∣)n∈ℕ ∈ C0   . A gauge symmetric function (or symmetric norm) is a map g : CF → ℝ  which satisfy the following properties:

g  is a norm on CF  ,
g(a) =  g(∣a∣)  for every a ∈ C        F  ,
•and g  is invariant under permutations.
We say that g  is normalized if g(e1) = 1  . For a ∈ C0   , define

g(a) = sup g (a1,...,an,0,...) ∈ ℝ ∪ {+ ∞}.         n∈ℕ

A unitarily invariant norm in L0(H)  is a map 9 ⋅9 : L0(H) →  ℝ+ ∪ { ∞} given by 9 A9  = g(s(A))  , A  ∈ L (H)        0  , where g  is a symmetric norm. The set

I = Ig = {A  ∈ L0(H)  : 9A9 <  ∞}

is a selfadjoint ideal of L(H)  , called the norm ideal associated to 9 ⋅9 . Then (I,9 ⋅ 9)  is a Banach space, and the following properties hold:
  1. If B ∈ I  , then 9 B9  = 9B  *9 , and 9 AB9   ≤ ∥A ∥ 9  B9 for every A  ∈ L(H)  .
  2. If A ∈ L(H)  has finite rank, then A  ∈ I  , because s(A) ∈ CF  .
  3. If dim R(A)  = 1  , then 9 A9  = s1(A)g(e1) =  g(e1)∥A ∥ .
  4. Given A ∈ L0(H)  and B  ∈ I  such that
              ∑k          ∑k  ∥A ∥(k) :=      sj(A) ≤      sj(B)  = ∥B ∥(k)   for every    k ∈ ℕ,            j=1         j=1
    then A  ∈ I  and 9 A9  ≤ 9B9 .
  5. For every ɛ > 0  and T  ∈ I  , there exists a finite rank operator S  such that 9 T - S9  <  ɛ  .

Some well known examples of unitarily invariant norms are the Schatten p  -norms                p1∕p  9 A9p  = tr(∣A ∣ )  , for 1 ≤ p ≤ ∞ , and the Ky-Fan norms ∥ ⋅ ∥(k)   , k ∈ ℕ  . The usual norm, which coincides with ∥ ⋅ ∥(1)   and 9 ⋅9 ∞ when restricted to L0(H)  , is also unitarily invariant. △

Proposition 2.2. Let N  be an unitarily invariant norm on an ideal I ⊆ L(H)  . Let {PF }F ∈F   be a increasing net of projections in      +  L(H)   which converges strongly to the identity (i.e., PF x --→  x       F ∈F  for every x ∈  H  ). Then

      N                        N  PFT  --→  T    and   PF T PF -- →  T   for every    T ∈ I .       F∈F                      F∈F

Proof. By Remark 2.1, for every ɛ > 0  , there exists a finite rank operator S  such that N (T - S) <  ɛ  . For every A ∈ I  and every projection P ∈  L(H)  , it holds that N (P AP ) ≤ N (P A) ≤  N (A)  . In particular, N (PF (T -  S)PF ) ≤ N (PF(T -  S)) < ɛ  for every F ∈  F  . Hence, we can assume that dim R(T ) = n <  ∞ . Given F ∈ F  , denote QF  =  I - PF  . Since N ( QF TQF  ) ≤ N (QF T )  and

T  - PF TPF  = QF T +  TQF  - QF T QF ,  F  ∈ F,

it suffices to prove that N (T - PF T) = N (QF T ) - -→ 0                           F∈F  . Fix x ∈ H  . Note that ∥(T *QF T )x ∥ ≤ ∥T ∥∥QF (T x)∥ --→  0                                 F ∈F  . Therefore

∥(T *QF T )1∕2x∥2 = 〈(T *QF T)x, x〉 ≤ ∥(T*QF  T)x∥ ∥x∥ --→  0 .                                                        F ∈F

This implies that   *      1∕2          SOT  (T  QF T )   = ∣QF T∣ -F-∈→F  0  , and all these operators act on the fixed finite dimensional subspace ker T⊥ , where the convergence of operators in every norm (included N  ) is equivalent to the SOT (or strong) convergence.

Remark 2.3. Let 9  ⋅9 be a unitarily invariant norm defined on a norm ideal I ⊆ L(H)  . The space L(H  ⊕ H)  can be identified with the algebra of block 2 × 2  matrices with entries in L(H)  , denoted by L(H)2 ×2   . Denote by I   2   the ideal of L(H  ⊕ H)  associated with the same norm 9 ⋅9 (i.e., by using the same symmetric norm g  ). Then, the following properties hold
  1. Let A  ∈ L0(H)  , and define A1 ∈ L0(H  ⊕ H)  as any of the following matrices
          (       )        (       )          0  A              0  0  A1 =    0   0     or      0  A    .

    Then s(A1) = s(A)  , 9 A19  =  9A9 , and A1 ∈ I2   if and only if A  ∈ I  .
  2. Under the mentioned identification, I2 = I 2×2   . △

Given A = [aij],B  = [bij] ∈ Mn( ℂ)  denote by A ∘ B  the Hadamard product [aijbij]  . Every A ∈ Mn( ℂ)  defines a linear map ΦA : Mn( ℂ) →  Mn( ℂ)  given by
ΦA(B)  = A ∘ B ,   B ∈ Mn(  ℂ).
(7)

Given a norm 9 ⋅9 on Mn( ℂ)  , it induces a norm of the linear map ΦA  by means of
9 ΦA9  = max  {9A  ∘ X9  : X  ∈ Mn( ℂ) ,  9X9  ≤  1 }.
(8)

The following result collects two classical results of Schur about Hadamard (or Schur) products of positive matrices (see [16]), and a generalization of the second one for unitarily invariant norms, proved by Ando in [1, Proposition 7.7] .

Proposition 2.4 (Schur). Let A ∈ Mn(  ℂ)+   and B  ∈ Mn( ℂ)  . Then
  1. If B ∈ Mn(  ℂ)+   then also A ∘ B ∈  Mn( ℂ)+   .
  2. Denote by d  =  max  {A   : 1 ≤ i ≤ n}   A           ii . Then
    9A  ∘ B9  ≤ dA  9 B 9    and     9 ΦA9  =  dA ,
    (9)
    for every unitarily invariant norm 9 ⋅9 on Mn(  ℂ)  . □
3. EQUIVALENT INEQUALITIES Theorem 3.1. Let k ∈ ℝ  and 9 ⋅9 be an unitarily invariant norm on an ideal I ⊆ L(H)  . Then the following inequalities are equivalent:
  1. 9 P TQ -1 + P -1T Q + kT 9 ≥  (k + 2) 9 T 9 , for every T ∈ I  and               +  P, Q ∈  Gl(H)   .
  2. 9 ST S -1 + S-1T S + kT 9 ≥  (k + 2) 9 T 9 , for every T ∈ I  and S  ∈ Gl(H)+   .
  3. 9 ST S + S -1T S- 1 + kT 9 ≥ (k + 2) 9 T 9 , for every T ∈ I  and            +  S  ∈ Gl(H)   .

Proof. It is clear that 1 implies 2 and 3. Suppose that 2 holds. Consider the space L(H  ⊕ H)  ~= L(H)2  ×2   . For P,Q  ∈ Gl(H)+   and T ∈ I  , define the operators

     [         ]              [        ]  S1 =    P   0      and   T1 =    0  T    ∈ I2 .          0   Q                    0   0

Then

                            [                             ]        -1    - 1                0  P T Q- 1 + P -1T Q + kT  S1T1S 1  + S1  T1S1 + kT1 =    0             0              .

Therefore, as 9 S1AS 1-1+ S -11AS1  + kA9  ≥ (k + 2) 9 A9 for every A ∈ I2   , then

9P  TQ -1 + P -1TQ  + kT 9 ≥ (2 + k) 9 T9  ,  T ∈ I.

This shows 2 → 1  . The same arguments using      [           ]  S1 =    P    0-1          0   Q show 3 →  1  . □

Remark 3.2. As said in the Introduction, the inequality 2 of Theorem 3.1 was proved, for the usual norm, by Corach-Porta-Recht in [6] with k = 0  (and S  not necessarily positive), and by Ameur Seddik in [15] with k =  1,2  . The inequality 1 of Theorem 3.1 was proved, in the finite dimensional case, by X. Zhan in [18], for k ∈ (- 2,2]  . In the rest of this section, we give a proof of inequality 2 of Theorem 3.1 for k ∈ (- 2,2]  in the general setting.

Lemma 3.3. Let n ∈ ℕ  , λ =  (λ ,...,λ ) ∈ ℝn         1      n      +   and k ∈  (- 2,2]  . Let Cn(k, λ) ∈ Mn( ℂ)  be given by

                  λ λ  Cn(k, λ)ij = -------i-j------,  1 ≤ i,j ≤  n .               kλiλj + λ2i + λ2j

Then Cn(k, λ) ∈ Mn( ℂ)+   for every n ∈ ℕ  .

Proof. Note that Cn(k, λ)ij = (λiλ-j1 + λ-i1λj + k)- 1   . These numbers are well defined because x + x-1 ≥ 2  for every x ∈ ℝ       +   and k >  - 2  . Bhatia and Parthasarathy [4] proved that, for - 2 < k ∈ ℝ  and       n  λ ∈ ℝ +   , the matrix Zn(k, λ) ∈ Mn( ℂ)  with entries
                    1  Zn(k, λ)ij = ---------2----2-,  1 ≤ i,j ≤ n,               kλiλj + λi + λj
(10)

satisfies Zn(k, λ) ∈ Mn( ℂ)+   for every n ∈ ℕ  and λ ∈ ℝn+   if and only if k ∈  (- 2,2].

On the other hand, if λ ∈ ℝn+   , then the matrix L =  λλ* = (λiλj)1≤i,j≤n ∈ Mn( ℂ)+  .  By Propposition 2.4, Cn(k, λ) = Zn(k, λ) ∘ L ∈ Mn( ℂ)+   for every k ∈ (- 2,2]  . □

Theorem 3.4. Let S  ∈ Gl(H)+   and k ∈  (- 2,2]  . Then, for every unitarily invariant norm 9  ⋅9 on an ideal I ⊆ L(H)  , and for every T ∈ I  ,

9kT  + ST S- 1 + S -1T S9 ≥ (k + 2) 9 T 9 .

Proof. We follow the same steps as in [6]. By the spectral theorem, we can suppose that σ(S)  is finite, since S  can be approximated in norm by operators Sn  such that each σ(Sn)  is finite. We can suppose also that dim H  < ∞ , by choosing an adequate net of finite rank projections {P  }    F  F∈F   which converges strongly to the identity and replacing S, T  by PF SPF , PF T PF  . Indeed, the net may be chosen in such a way that SPF  =  PFS  and σ(PF SPF ) = σ(S)  for every F ∈  F  . Note that, by Proposition 2.2, 9 PF APF 9 converges to 9 A9 for every A ∈  I  .

We can suppose that S  is diagonal by a unitary change of basis in ℂn  . Take S =  diag(λ1,λ2, ...,λn)  . Then            - 1    -1  k T +  ST S   + S   T S = Bn(k, λ) ∘ T  , where
            (             )   (           2    2 )  B (k, λ)  =   k + λi-+ λj-  =   k-λiλj-+-λi-+-λj-  ,   1 ≤ i,j ≤ n.   n      ij         λj   λi             λiλj
(11)

Since      - 1  x + x   ≥  2  for every x ∈ ℝ  , x > 0  , it follows that, if k ∈ (- 2,+ ∞)  , then Bn(k, λ)ij > 0  for every 1 ≤ i,j ≤  n  . Consider the matrix Cn(k, λ)  given by Cn(k, λ)ij = Bn(k, λ)-ij1  . Hence, in order to prove inequality (6) for every T ∈ Mn(  ℂ)  , it suffices to show that 9 Cn(k, λ) ∘ A9 ≤ (k + 2)-1 9 A9 for A ∈ Mn(  ℂ)  and k ∈ (- 2,2]  . By Lemma 3.3,                   +  Cn(k, λ) ∈ Mn( ℂ)   for every k ∈ (- 2,2]  . Finally, note that                    - 1  Cn(k, λ)ii = (k + 2)   , 1 ≤ i ≤ n.  Therefore, inequality (6) holds by Eq. (9) in Proposition 2.4. □

As a consequence of this result and Theorem 3.1, we get an infinite dimensional version, for every unitarily invariant norm, of Zhang inequality:

Corollary 3.5. Let              +  P, Q ∈ Gl(H)   and k ∈  (- 2,2]  . Then, for every unitarily invariant norm 9  ⋅9 on an ideal I ⊆ L(H)  , and for every T ∈ I  ,

9P  TQ -1 + P -1T Q + kT 9 ≥  (2 + k) 9 T 9 .

Corollary 3.6. Let S ∈ Gl(H)+   and k ∈  (- 2,2]  . Then, for every unitarily invariant norm 9  ⋅9 on an ideal I ⊆ L(H)  , and for every T ∈ I  ,

9ST  S + S- 1TS -1 + kT 9 ≥ (2 + k) 9 T 9 .

4. THE ASSOCIATED OPERATORS

In this section we study the operators associated with the inequalities proved in the previous section. Given P,Q  ∈ Gl(H)+   and k ∈ ℝ  , we consider the bounded operatorΥP,Q,k : L(H) →  L(H)  associated with inequality (4):
ΥP,Q,k(T ) = PT Q -1 + P- 1TQ  + kT ,  T ∈ L(H)  .
(12)

Hence, for every unitarily invariant norm 9 ⋅9 defined on an ideal I ⊆ L(H)  , inequality (4) means that 9  Υ     (T )9 ≥  (2 + k) 9 T 9      P,Q,k for T ∈ I  , - 2 < k ≤ 2  . Given          +  S ∈ L(H)   and k ∈ ℝ  , define the operators ΦS,k : L(H) →  L(H)  and ΨS,k : L(H) →  L(H)  associated with inequalities (6) and (5): ΦS,k = ΥS,S,k  and ΨS,k = ΥS,S -1,k  . In this section we characterize, for fixed P,Q  ∈ Gl(H)+   , those k ∈ ℝ  such that ΥP,Q,k  is invertible. In some cases we estimate, for a given norm on some ideal of L(H)  , the induced norms of their inverses.

Remark 4.1. Let u,v ∈ H  . Denote by u ⊗ v ∈ L(H)  the operator given byu ⊗ v(z) = 〈z,v 〉u  , z ∈ H  . It is clear that R(u  ⊗ v) = span {u} . Hence u ⊗ v ∈ I  for every norm ideal induced by a unitarily invariant norm 9 ⋅9 . The following properties are easy to see:
  1. 9 u ⊗ v9  = ∥u ⊗ v∥ = ∥u ∥∥v ∥ .
  2. The map (u,v) ↦→  u ⊗ v  is sesqui linear.
  3. If T ∈ L(H)  , then T (u ⊗ v) = (Tu) ⊗ v  and (u ⊗ v)T =  u ⊗ (T*v)  . △

Proposition 4.2. Let              +  P,Q  ∈ Gl(H)   . Denote by ΥP,Q =  ΥP,Q,0   . Then

               -1    -1  σ(ΥP,Q)  = {λμ   +  λ  μ  : λ ∈ σ (P) ,μ ∈ σ(Q)}.

Moreover, ΥP,Q   has the same spectrum, if it is considered as acting on any norm ideal I  associated with a unitarily invariant norm.

Proof. Fix the norm ideal I  and consider the restriction ΥP,Q : I → I  . Let AP,Q  : I → I  be given by AP,Q(T ) = P T Q- 1   , T  ∈ I  . Note that ΥP,Q =  AP,Q + A -1                   P,Q  . Therefore, by the known properties of the Riesz functional calculus for operators on Banach spaces (in this case, the Banach space is I  and the map is              -1  f (z) =  z + z   ), it suffices to show that               -1  σ(AP,Q) = {λ μ   : λ ∈ σ (P) , μ ∈ σ(Q)} .

Given C ∈ L(H)  , denote by LC  : I → I  (resp. RC  ) the operator given by LC (T ) = CT  (resp. RC (T ) = TC  ), T ∈ I  . By Remark 2.1, these operators are bounded. If C ∈ Gl(H)  , then LC-1 = (LC )-1   , and similarly for RC  . Hence σ(L  ) ⊆ σ(C)      C  and σ(R   ) ⊆ σ(C)      C  . Note that A     = L  R  -1 = R  -1L    P,Q     P  Q       Q   P  . Therefore

σ(AP,Q) ⊆  σ(LP )σ(RQ -1) ⊆ {λμ- 1 : λ ∈ σ (P ), μ ∈ σ(Q)}.

Given λ ∈ σ(P )  , μ ∈ σ(Q)  and ɛ > 0  , let x, y ∈ H  be unit vectors such that ∥P x - λx ∥ < ɛ  and    -1  ∥Q   y - μy ∥ < ɛ  . Such vectors exist because P -  λI  and   -1    -1  Q    - μ  I  are selfadjoint operators. Consider the rank one operator x ⊗ y ∈ I  . Then, by Remark 4.1, AP,Q(x  ⊗ y) = P x ⊗ (Q- 1)*y =  Px ⊗ Q -1y  . Hence

             -1                        - 1          - 1  9(AP,Q  - λμ   IdI)x ⊗ y9   = 9P  x ⊗ Q   y - λx-1⊗ μ   y9        -1     - 1                              ≤ 9(P  x - λx) ⊗ Q   y 9 + 9 λx ⊗ (Q   y - μ   y)9                              = ∥P x - λx ∥∥Q - 1y ∥ + ∥λx∥ ∥Q -1y - μ-1y ∥                              < (∥Q -1y∥ + ∥λx ∥) ɛ ≤ (∥Q -1∥ + ∥P ∥) ɛ .

Therefore   - 1  λμ   ∈ σ(AP,Q)  , because 9 x ⊗ y9  = ∥x∥ ∥y∥ = 1  . This shows that               -1  σ(AP,Q) = { λμ   : λ ∈ σ(P ), μ ∈ σ(Q)} , and the proof is complete. □

Corollary 4.3. Let S ∈ Gl(H)+   and k ∈ ℂ  . Then Υ    P,Q,k   is invertible if and only if           -1    - 1  - k ∕∈ {λ μ   + λ  μ  : λ ∈ σ (P ),μ ∈ σ(Q)} .

Proof. Just note that ΥP,Q,k = ΥP,Q + k  IdI   . Then apply Proposition 4.2. □

Remark 4.4. Let 9 ⋅9 be a unitarily invariant norm defined on a norm ideal I ⊆ L(H)  . By Remark 2.1, 9  AB9  ≤  ∥A ∥ 9 B9 for A ∈ L(H)  and B ∈ I  . Given a linear operator Υ  : I → I  , we denote by 9  Υ9 the induced norm:

9 Υ9 =  sup{9 Υ(T )9 : T ∈ I,9T 9 =  1}.

By a standard ɛ∕2  argument and using the continuity of "taking inverse", one can show that, for k ∈ (- 2,2]  fixed, the map Gl(H)+  × Gl(H)+  ∋ (P, Q) ↦→  9 Υ- 1  9                                   P,Q,k is continuous. △

Proposition 4.5. Let P, Q ∈ Gl(H)+   and k ∈ (- 2,2]  . Let 9  ⋅9 be a unitarily invariant norm defined on a norm ideal I ⊆ L(H)  . Then 9 Υ -P1,Q,k9 ≤ (k + 2)- 1   . Moreover, if σ(P ) ∩ σ (Q) ⁄= ∅ , then 9 Υ -1  9  = (k + 2)-1      P,Q,k   .

Proof. The inequality follows from Corollary 3.5. Suppose that λ ∈  σ(P ) ∩ σ (Q)  is an eigenvalue for both P  and Q  . Let x,y ∈ H  be unit vectors such that P x =  λx  and Qy  = λy  . Consider x ⊗  y ∈ I  . Then, since Q =  Q* and λ ∈  R  , it is easy to see that ΥP,Q,k(x ⊗ y) = (2 + k) (x ⊗ y)  . Hence, 9 Υ -S1,k9 =  (k + 2)- 1   in this case. An easy consequence of spectral theory is that every S ∈ Gl(H)+   such that λ ∈ σ(S)  can be arbitrarily approximated by positive invertible operators such that λ  is a isolated point of their spectra, hence an eigenvalue. Applying this, jointly for P  and Q  , for some λ ∈ σ (P ) ∩ σ (Q)  , and using the fact that the map (P,Q)  ↦→ 9 Υ -P1,Q,k9 is continuous, the proof is completed. □

Corollary 4.6. Let S ∈ Gl(H)+   and k ∈ (- 2,2]  . Let 9 ⋅9 be a unitarily invariant norm defined on a norm ideal I ⊆ L(H)  . Then 9 Φ -19  = (k + 2)-1      S,k   . If there exists λ ∈ σ(S)  such that also   -1  λ   ∈  σ(S)  , then also     -1            -1  9 Ψ S,k9 = (k + 2)   .

Proof. The first case follows applying Proposition 4.5 with P =  Q = S  . Note that the hypothesis σ(P ) ∩ σ(Q) ⁄=  ∅ becomes obvious. For the second, take P = S  and Q  = S -1   . Note that σ (S -1) = {λ- 1 : λ ∈ σ(S)} . □

Remark 4.7. The Forbenius norm 9 A922 = trA *A  works on the ideal of Hilbert Schmit operators, which is a Hilbert space with this norm. In this case, the operator Υ    P,Q,k  defined in Eq. (12) is positive, so that 9 Υ -1  9  =  ρ(Υ -1  )      P,Q,k  2       P,Q,k  . Therefore, Proposition 4.2 gives the sharp constant for inequality (4) for this norm. Observe that     -1              -1  9 Υ P,Q,k92 = (k + 2)   if and only if σ(P ) ∩ σ (Q) ⁄= ∅ . △

5. SHARP CONSTANTS AND HADAMARD INDEX

Preliminary results. In this subsection, we shall give a brief exposition of the definitions and results of the theory of Hadamard index, which we shall use in the rest of the section. All the results are taken from [7].

Denote by p = (1,...,1) ∈ ℝn  and P  = pp* ∈ Mn( ℂ)+   , the matrix with all its entries equal to 1  . Given A  ∈ M  (ℂ)+         n   and N  a norm on M   (ℂ)     n  , we define the N  -index of A  by

I(N, A) =  max{ λ ≥ 0 : N (A ∘ B) ≥ λN (B) ∀  B ∈ P (n)}

and the minimal index of A  by

I(A)  = max  { λ ≥  0 : A ∘ B ≥  λB   ∀ B  ∈ Mn( ℂ)+  }        = max  { λ ≥  0 : A -  λP ≥  0 }.

The index relative to the spectral norm on Mn(  ℂ)  will be denoted by I(sp, A)  , and the index relative to the 2-norm ∥B ∥22 =  tr B *B  , B  ∈ Mn( ℂ)  , is denoted by I2(A)  .

Proposition 5.1. Let            +  A ∈  Mn( ℂ)   . Then I(A)  ⁄= 0  if and only if p ∈ R(A)  . □

If      (  a  b )  A  =    ¯      ∈ P2          b  c   , then
0 ⁄= I(sp, A) = I(A)    ⇔    b ∈ ℝ  and  0 ≤ b ≤ min{a, c} ⁄= 0.
(13)

Theorem 5.2. Let A ∈ Mn( ℂ)+   with nonnegative entries such that all Aii ⁄= 0  . Then the following conditions are equivalent:
  1. There exist       n  u ∈ ℝ  with nonnegative entries such that Au =  p .
  2. I(sp, A) = I(A)  . □

Theorem 5.3. Let B  ∈ Mn( ℂ)+   such that bij ≥ 0  for all i,j  . Then there exists a subset J0   of 핀n = {1,2,...,n} such that I(sp,B) =  I(sp,BJ0) = I(BJ0)  . Therefore I(sp,B)  = min{  I(sp,B  ) : J ⊆  𝕀                          J         n  and I(sp,B  ) = I(B  ) }.         J        J  □

Let x =  (λ1, ...,λn)* ∈ ℝn+   , L  = {λ1,...,λn} and k ∈  ℝ  . We consider

      (            )  Λx =   λiλj + --1--   ∈  Mn( ℂ)+    and                λiλj  ij

           (        -1---    )                      +  En(x, k) =  λiλj +  λ λ +  k  ij = Λx + kP ∈ Mn(  ℂ)  .                       i j
(14)

Proposition 5.4. Let x = (λ ,...,λ  )* ∈ ℝn         1      n      +   , L =  {λ ,...,λ }         1      n . Consider the matrix Λ   x  defined before. Then I(Λx) ⁄=  0  if and only #L  ≤  2  . More precisely,

  1. If L = {λ1} (i.e. #L = 1), then Λx  = (λ21 + λ-12) P  and I(Λx) =  λ21 + λ-12   .
  2. If #L  > 1  , then the image of Λx  is the subspace generated by the vectors x =  (λ ,...,λ )         1      n  and y = (λ-1,...,λ -1)        1        n  .
  3. If #L  = 2  , say L = { λ,μ} , denote                      +  Λ0  = Λ{λ,μ}∈ M2( ℂ)   . Then
             (λ + μ)2  I(Λx) =  ---------=  I(Λ0),           1 + λ2μ2
    (15)
  4. If #L  > 2  then I(Λx) = 0  , because p ∕∈ span {x,y} . □

Sharp constants for positive operators. In the rest of this section, we shall use the results above in order to get the sharp constants for inequality (5), if we consider only positive  operators T  . Remark 5.5. Proposition 5.4 says that, in most cases (i.e. #L   > 2  ), I(Λx) = 0  . On the other hand, in order to apply index theory for the matrix E  (x, k) = Λ  + kP   n          x  , we need that                   +  En(x, k) ∈ Mn( ℂ)   . Since Λx + kP  ≥ 0  if and only if k ≥ - I(Λx)  (by the definition of the minimal index), we can only consider the case k ≥ 0  , in general. In this case, it is easy to check that
I(En(x, k)) = I(Λx) + k .
(16)

Hence, applying Proposition 5.4, we get formulas for I(En(x, k))  in any case. △

We shall compute I(sp, En(x,k))  using Theorem 5.3. Hence, we shall use the principal minors of En(x, k)  , which are matrices of the same type. Let J ⊆  {1,2,...,n} , LJ  = {λj : j ∈ J} and pJ,xJ ∈  RJ  the induced vectors. Then

En(x, k)J = E ∣J∣(xJ,k) =  (Λx)J +  kPJ =  ΛxJ + kPJ .

Claim: If #LJ  > 2  then, I(sp,En(x, k)J ) ⁄= I(En(x, k)J )  .

Indeed, by Proposition 5.4, I(ΛxJ) = 0  , because pJ ∈∕ R(ΛxJ )  . Note that every u ∈ En(x, k)-J 1{pJ } must satisfy ΛxJu = 0  , because R(PJ ) =  span (pJ )  . By Proposition 5.4, ker(ΛxJ) = R( ΛxJ)⊥ = {xJ ,yJ}⊥ , where y = (λ-1 1,...,λ-n1)  . As all entries of xJ  are strictly positive, if u ≥ 0  and 〈u,xJ〉 = 0  , then u = 0  . Therefore, the Claim follows from Theorem 5.2.

Hence, if I(sp,En(x, k)J) = I(En(x, k)J)  , then #LJ  ≤ 2  . If #LJ  =  2  , let i1,i2 ∈ J  such that λi1 ⁄= λi2   . By Theorem 5.2 there exists a vector 0 ≤ u ∈ ℝJ  such that En(x, k)J u = pJ  . Let z1 = ∑{uk   : k ∈ J and λk = λi} ≥  0                                 1  and z2 = ∑{uj   : j ∈ J, and λj = λi } ≥ 0.                                 2  Easy computations show that, if we denote E  = E  ({λ  ,λ  },k) = E  (x, k)    0    2   i1  i2        n      {i1,i2} , then E0(z1, z2) = (1,1)  and, by Theorem 5.2, I(E0) =  I(sp,E0)  . Moreover, by equations (15) and (16),

                                  (λi + λi )2  I(sp, En(x, k)J) = I(En(x, k)J) = ---1--2-22- + k = I(E0) =  I(sp,E0).                                     1 + λi1λ i2

Therefore, in order to compute I(sp,En(x, k))  using Theorem 5.3, we have to consider only the diagonal entries of En(x, k)  and some of its principal minors of size 2 × 2  . If λ  ⁄= λ    i    j  and E  =  E  (x, k)    0    n      {i,j} then, by equations (13) and (16),

I(E0)  = I(sp,E0)   ⇐ ⇒    I(sp,Λ{λi,λj}) = I(Λ{λi,λj})                        ⇐ ⇒    λ λ +  -1--≤ min{ λ2 + -1 , λ2 + -1}.                              i j   λiλj          i  λ2i    j   λ2j

Easy computations show that, if λi < λj  , this is equivalent to λ2≤  -1--≤  λ2   i   λiλj    j  , which implies that λi < 1 < λj  . So, by Theorem 5.3,
I(sp,En(x, k)) = min{M1,  M2}
(17)

where                  -2  M1 =  mini λ2i + λi +  k = mini En(x, k)ii  and

          { (λ +  λ )2                                1       }  M2  = min   --i---2j2-+  k : λi < 1 < λj  and  λ2i ≤ -----≤  λ2j  .               1 + λiλj                               λiλj

Now we can characterize the sharp constant for inequality (5), if we consider only positive  operators T  .

Proposition 5.6. Given S  ∈ Gl(H)+   and k ≥ 0  , denote by M (S,k)  the greatest number such that ∥kT  + ST S + S -1T S-1∥ ≥  M (S,k)∥T ∥ for every T  ∈ L(H)+   . Then M  (S,k) = min { M1(S, k),M2(S, k)} ,  where                        2    - 2  M1(S,  k) = minλ∈ σ(S)λ  + λ   + k  and

               {        2                                           }  M2(S,  k) = inf   (λ-+-μ)--+ k : λ, μ ∈ σ(S),λ < μ and λ2 ≤  1--≤  μ2 .                   1 + λ2μ2                                   λμ

In particular, if ∥S ∥ ≤ 1  , then M (S, k) = ∥S∥2 + ∥S ∥-2 + k  .

Proof. We shall use the same steps as in the proof of Theorem 3.4 (and [6]). By the spectral theorem, we can suppose that σ(S)  is finite, since S  can be approximated in norm by operators             +  Sn ∈ Gl(H)   such that σ(Sn)  is a finite subset of σ(S)  , σ(Sn) ⊂  σ(Sn+1)  for all n ∈ ℕ  and ∪n∈ℕ σ(Sn)  is dense in σ(S)  . So M  (Sn,k)  (and Mi(Sn,  k), i = 1,2  ) converge to M  (S,k)  (resp. Mi(S, k), i = 1, 2  ).
We can also suppose that dim  H <  ∞ , by choosing an adequate net of finite rank projections {PF }F∈F   which converges strongly to the identity and replacing S, T  by PF SPF , PF T PF  . Indeed, the net may be chosen in such a way that SPF  = PF S  and σ(PF SPF ) = σ(S)  for all F ∈ F  . Note that for every A ∈ L(H)  , ∥PF APF ∥ converges to ∥A ∥ .
Finally, we can suppose that S  is diagonal by a unitary change of basis in ℂn  . In this case, if λ1, ...,λn  are the eigenvalues of S  (with multiplicity) and x = (λ1,...,λn)  , then          -1   -1  ST S + S   T S   + kT =  En(x, k) ∘ T  . Note that all our reductions (unitary equivalences and compressions) preserve the fact that T  ≥ 0  . Now the statement follows from formula (17). If ∥S ∥ ≤ 1  then M  (S) = M1(S)  , since M2(S)  is the infimum of the empty set. Note that M1(S)  is attained at λ = ∥S ∥ , because the map f (x) =  x + x-1   is decreasing on (0, 1]  . □

6. NUMERICAL RESULTS

Let S ∈  Gl(n)+   and k ∈ (- 2,+∞)  . Denote by                    {                                                     }  Nn(S, k)  =   max   c ≥ 0 : ∥kT + ST S + S -1T S-1∥ ≥  c∥T∥,  T ∈ Mn(  ℂ)  .  Corollary 3.6 says that, if k ≤ 2  , then Nn(S,  k) ≥ k + 2  for every           +  S  ∈ Gl(n)   . On the other hand, Corollary 4.6 says that, if there exists                 -1  λ ∈ σ (S) ∩ σ (S   )  , thenNn(S, k) = k + 2  . In this section we search conditions for S  which assure thatNn(S, k) = min λ∈ σ(S)λ2 + λ-2 + k  . As in the proof of Theorem 3.4, we can assume that S =  diag(λ)  for some λ = (λ1,λ2,...,λn) ∈ ℝn>0   . In this case, we have thatk T +  ST S + S -1TS -1 = E  (λ,k) ∘ T                              n  , where E  (λ,k) ∈ M  (ℂ)+    n          n   is the matrix defined in Eq. (14). Consider the matrix G  ∈ Mn( ℂ)h  with entries
g  = (e (λ,k)  )-1 = ---------1-------- = -------λiλj------ ,   ij     n      ij      k + λiλj + λ-i 1λ-j1   1 + k λiλj + λ2iλ2j
(18)

1 ≤ i,j ≤ n  . Then, for T ∈ Mn( ℂ)  ,                   - 1    -1  T = G  ∘ (T + ST S   + S   TS)  . Denote by ΦG  : Mn( ℂ) → Mn(  ℂ)  the map given by ΦG(B)  =  G ∘ B  , for B  ∈ Mn( ℂ)  . We conclude that Nn(S, k)- 1 = ∥ΦG ∥ .

Remark 6.1. There exists an extensive bibliography concerning methods for computing the norm of a Hadamard multiplier like ΦG  . The oldest result in this direction is Schur Theorem (Proposition 2.4) for the positive case. We have applied this result in the proof of Theorem 3.4, but it is not useful in this case, because G  ∕∈ Mn( ℂ)+   . The most general result is 1983's Haagerup theorem [10], which gives a complete characterization, but it is not effective. There exist also several fast algorithms (see, for example, [9]) which allow to make numerical experimentation for this problem. For example, we have observed that the behavior of the map ℝ ∋ t ↦-→  Nn(tS, k)  , for any fixed S  , is chaotic. But, as a great number of examples suggest, it seems that Nn(tS, k) = (2 + k)-1   if and only if tσ(S) ∩ t-1σ (S-1) ⁄= ∅ . Note that these cases are exactly those considered in Corollary 4.6. △

Cowen and others [8] and [9] proved the following result for hermitian matrices:

Theorem 6.2. Let B =  (bij) ∈ Mn( ℂ)h  such that 0 < b11 =  max  bjj            1≤j≤n  . Suppose that B  has rank two, or that B  has one positive eigenvalue and n - 1  non positive eigenvalues. Then the following conditions are equivalent:
  1. ∥ Φ  ∥ = b     B      11   .
  2. If 1 < j ≤ n  and J = {1, j} , then ∥ΦBJ ∥ = b11   .
  3. For 1 ≤  k ≤ n  , it holds that b211 + b11bkk - 2∣b1k∣2 ≥ 0  . □

Suppose that λ = (λ ,λ  ,...,λ  ) ∈ ℝn        1   2     n     >0   satisfies 1 ≥ λ  ≥ λ  ≥  ⋅⋅⋅ ≥ λ >  0       1     2          n  or 1 ≤ λ1 ≤  λ2 ≤ ⋅⋅⋅ ≤ λn.  Since that map             -1  f (t) = t + t   is increasing on (1,+ ∞)  and it is decreasing on (0,1)  , it follows that the matrix G  of Eq. (18) satisfies 0 < g11 = m1≤ajx≤n gjj  . This suggests that we could apply Theorem 6.2 for our problem. Unfortunately, for n > 2  , G  has rank greater than two, and it can have more that one positive eigenvalue. We prove the following result:

Proposition 6.3. Let S ∈ Gl(2)+   . Suppose that S ≤ I  , or I ≤ S. Then

Nn(S,  0) =  min  λ2 + λ- 2 .              λ∈ σ(S)

Proof. Suppose that σ(S) =  {λ ,λ }            1  2 with 1 ≤ λ  ≤ λ       1     2   or 1 ≥  λ  ≥ λ  > 0        1    2  . Let G  ∈ M2( ℂ)  as in equation (18), for k = 0  . Then det(G) ⁄=  0  . Hence, since g11 ≥ g22   , can apply Theorem 6.2. Then, in order to prove that                        -1  g11 = ∥ΦG ∥ = Nn(S,  0)   , it suffices to verify the inequality C = g211 + g11g22 - 2g212 ≥ 0 .  Note that

            λ2          λ2       λ2          λ2λ1  C   =    (--4-1--)2 + --4-1-----4-2----  2(--------)2            λ 1 + 1     (λ 1 + 1) (λ2 + 1)    λ1λ2 + 1           2                          2 2     6  2     4    4 4    6 6      =   λ1(λ2---λ1)(λ1-+-λ2)(--1 +-λ1λ2-+-2λ1λ-2 --2λ-2 --λ-1λ2 +-λ1λ2) .                           (λ41 + 1)2(λ42 + 1)(1 + λ21λ22)2

Straightforward computations with 1 ≤ λ1 ≤  λ2   show that C > 0  , since the polynomial P (λ1,λ2) = - 1 + λ21λ22 + 2λ61λ22 - 2λ42 - λ41λ42 + λ61λ62 > 0  in this case. A similar analysis shows that still C >  0  for 1 ≥ λ1 ≥  λ2 > 0  . The result follows by applying Theorem 6.2. □

It was proved by Kwong (see [13]) that, if λ ∈ ℝn+   , then the matrix Zn(k, λ) ∈ Mn( ℂ)  defined in Eq. (10), is positive semidefinite in the following cases: n = 2  and k ∈ (- 2,∞)  , n =  3  and k ∈ (- 2,8)  , and n =  4  and k ∈ (- 2,4)  . Therefore, by the proof of Theorem 3.4, inequality (6) holds in Mn( ℂ)  in these cases, for every unitarily invariant norm. Note that the proof Theorem 3.1 does not give similar estimates for the inequalities (4) and (5), because one needs to duplicate dimensions.

A numerical approach suggests that these intervals are optimal, both for the positivity of the matrix Bn(k,λ)  , defined in Eq. (11), and for inequality (6). In the case of M3( ℂ)  , if - 2 < k ∈ ℝ  and                     3  λ =  (λ1,λ2,λ3) ∈ ℝ >0   then, using symbolic computation with the software Mathematica, one obtains a kind of "proof" of the fact that detB3(k, λ) ≥ 0  for every λ ∈ ℝ3>0   if and only if -  2 < k ≤ 8.  The 2 × 2  principal sub matrices of B  (k,λ)    3  have the form B (k,(λ ,λ ) )   2      i  j  , and they live in M   (ℂ)+     2   for every k > - 2  . Therefore,

                  +                    3  B3(k, λ) ∈ M3( ℂ)     for every  λ ∈  ℝ>0  ⇐ ⇒  - 2 < k ≤ 8.

Likewise, for the 4 × 4  matrix case, it suffices to study detB4(k, λ)  , for λ ∈ ℝ4>0   , and one obtains similar results.

Denote by kn  the maximum number k ∈ ℝ  such that inequality (6) holds in Mn( ℂ)  for the spectral norm. By the preceding comments, and the proof of Theorem 3.4,

k2 = + ∞,   k3 ≥  8,  k4 ≥ 4,   and     kn ≥ 2    for    n ≥ 5.

Computer experimentation using the softwares Mathematica and Matlab suggests that, also in this case, k3 =  8  , k4 = 4  , and kn = 2  , for n ≥  5  . In other words, inequality (6) holds for Mn( ℂ)  in the same intervals as it holds that Bn(k, λ) ∈ Mn( ℂ)+   for every λ ∈ ℝn>0   .

References

[1]    T. Ando, Majorization and inequalities in matrix theory, Linear Algebra Appl. 199: 17-67, 1994.         [ Links ]

[2]    E. Andruchow, G. Corach and D. Stojanoff, Geometric operator inequalities, Linear Algebra Appl. 258: 295-310, 1997.         [ Links ]

[3]    R. Bhatia, and CH. Davis, More matrix forms of the Arithmetic-Geometric mean inequality, SIAM J. Matrix Anal. Appl. 14: 132-136, 1993.         [ Links ]

[4]    R. Bhatia and K. R. Parthasarathy, Positive definite functions and operator inequalities, Bull. London Math. Soc. 32: 214-228, 2000.         [ Links ]

[5]    R. Bhatia, Matrix Analysis, Springer-Verlag, New York, 1997.         [ Links ]

[6]    G. Corach, H. Porta and L. Recht, An operator inequality, Linear Algebra Appl. 142: 153-158, 1990.         [ Links ]

[7]    G. Corach and D. Stojanoff, Index of Hadamard multiplication by positive matrices II, Linear Algebra Appl. 332-334 (2001), 503-517.         [ Links ]

[8]    C. C. Cowen, K. E. Debro and P. D. Sepansky, Geometry and the norms of Hadamard multipliers, Linear Algebra Appl. 218: 239-249, 1995.         [ Links ]

[9]    C. C. Cowen, P. A. Ferguson, D. K. Jackman, E. A. Sexauer, C. Vogt y H. J. Woolf, Finding Norms of hadamard multipliers, Linear Algebra Appl. 247, 217-235 (1996).         [ Links ]

[10]    U. Haagerup, Decompositions of completely bounded maps on operator algebras, unpublished manuscript.         [ Links ]

[11]    F. Kittaneh, A note on the arithmetic-geometric inequality for matrices, Linear Algebra Appl. 171 (1992), 1-8.         [ Links ]

[12]    F. Kittaneh, On some operator inequalities, Linear Algebra Appl. 208/209:19-28, 1994.         [ Links ]

[13]    M.K. Kwong, On the definiteness of the solutions of certain matrix equations, Linear Algebra Appl. 108: 177-197, 1988.         [ Links ]

[14]    L. Livshits and S. C. Ong, On the invertibility properties of the map         -1    -1  T ↦→ ST S   +S   TS  and operator-norm inequalities, Linear Algebra Appl. 183: 117-129, 1993.         [ Links ]

[15]    A. Seddik, Some results related to the Corach-Porta-Recht inequality, Proceedings of the American Mathematical Society, 129: 3009-3015, 2001.         [ Links ]

[16]    I. Schur, Bemerkungen zur theorie de beschränkten bilineareformen mit unendlich vielen veränderlichen, J. Reine Angew. Math. 140: 1-28, 1911.         [ Links ]

[17]    B. Simon, Trace ideals and their applications, London Mathematical Society Lecture Note Series, 35, Cambridge University Press, Cambridge-New York, 1979.         [ Links ]

[18]    X. Zhan, Inequalities for unitarily invariant norms, SIAM J. Matrix Anal. Appl. 20:466-470, 1999.         [ Links ]

Cristina Cano
Depto. de Matemática,
FaEA-UNC,
Neuquén, Argentina.
cbcano@uncoma.edu.ar

Irene Mosconi
Depto. de Matemática,
FaEA-UNC,
Neuquén, Argentina.
imosconi@uncoma.edu.ar

Demetrio Stojanoff
Depto. de Matemática,
FCE-UNLP, La Plata, Argentina
and IAM-CONICET.
demetrio@mate.unlp.edu.ar

Recibido: 3 de junio de 2005
Aceptado: 21 de noviembre de 2005

Creative Commons License Todo o conteúdo deste periódico, exceto onde está identificado, está licenciado sob uma Licença Creative Commons