Topics on Operator Inequalities T. Ando

Topics on Operator Inequalities
T. Ando
Division of Applied Mathematics
Research Institute of Applied Electricity
Hokkaido University, Sapporo, Japan
Research supported by Kakenhi 234004
CHAPTER I
Geometric and Harmonic Means
Throughout the lecture G,H,K denote Hilbert spaces. L(H) is the space of (bounded) linear operators
on H, while L+ (H) is the cone of positive (i.e. non-negative semi-definite) operators.
In this chapter we shall be concerned with simple binary operations in L+ (H), called geometric and
harmonic means.
Theorem!I.1. Suppose that H = G ⊕ K and a self-adjoint operator T on H is written in the form
A C∗
T =
where A and B act on G and K respectively and C acts from G to K. Then in order
C B
that T is positive it is necessary and sufficient that A and B are positive and there is a contraction W (i.e.
kW k ≤ 1) from G to K such that C = B 1/2 W A1/2 .
Proof. Suppose that A ≥ 0, B ≥ 0 and C = B 1/2 W A1/2 with a contraction W . The condition
kW k ≤ 1 implies W W ∗ ≤ 1, hence B 1/2 W W ∗ B 1/2 ≤ B. Then T admits a factorization T = S ∗ S where
!
A1/2
W ∗ B 1/2
S=
1/2 ,
0
B − B 1/2 W W ∗ B 1/2
which implies that T is positive.
Suppose conversely that T is positive. This means that
(Ax, x) + 2Re (Cx, y) + (By, y) ≥ 0
for
x ∈ G, y ∈ K.
A and B are obviously positive and the above inequality is easily seen to be equivalent to the following
(Ax, x)(By, y) ≥ |(Cx, y)|
2
for x ∈ G, y ∈ K.
Then for each y ∈ K the vector C ∗ y belongs to the range of A1/2 and
2
2
|(Cx, y)|
−1/2 ∗ 2
C y = sup
≤ B 1/2 y ,
A
(Ax, x)
x
where A−1/2 is defined to be the (unbounded) inverse of A1/2 restricted to the orthocomplement of the kernel
of A1/2 . Now there is a contraction U from K to G such that A−1/2 C ∗ = U B 1/2 . Finally W = U ∗ meets
the requirement.
Corollary I.1.1. If T is a positive operator on H and G is a closed subspace of H, then there is a
linear operator S such that T = S ∗ S and S(G) ⊆ G.
In fact, the operator S in the proof of Theorem I.1 meets the requirement.
!
!
A C∗
A C∗
Corollary I.1.2. If
is positive, then there is the minimum of all X for which
C B
C X
are positive.
!
∗ −1/2 ∗ A C∗
Proof. As in the proof of Theorem I.1 the positivity of
implies A−1/2 C ∗
A
C ≤ X.
C X
∗ −1/2 ∗ This means that A−1/2 C ∗
A
C is the minimum in the assertion.
3
4
I. GEOMETRIC AND HARMONIC MEANS
Remark. If A has bounded inverse, the minimum in Corollary I.1.1 is just CA−1 C ∗ .
Corollary I.1.3. If H ⊇ G and if S is a linear operator from G to H such that
(Sx, y) = (x, Sy)
for
x, y ∈ G
then there is a self-adjoint operator T on H such that T |G = S and kT k = kSk. Further, among all T
satisfying these conditions there are the minimum Tµ and the maximum TM .
!
A
Proof. It can be assumed that kSk = 1. Let S =
where A acts on G and C does from G to
C
K = H G. By assumption A is self-adjoint and
2
2
2
kAxk + kCxk ≤ kxk
for x ∈ G.
!
!
A C∗
1 + A C∗
∗
2
Therefore C C ≤ 1 − A . T must be of the form T =
for which
and
C X
C
1+X
!
1 − A −C ∗
are positive. The inequalities C ∗ C ≤ 1 − A2 and 2−1 (1 − A) ≤ 1 imply that
−C
1−X
C ∗ 2−1 C ≤ 2−1 1 − A2 = 2−1!(1 − A)(1 + A) ≤ 1 + A. By Remark after Corollary I.1.2 this implies
!
C∗
1 + A C∗
. Therefore there is the minimum Bµ of all X for which
C
1+1
C
1+X
−1/2 ∗ ∗
−1/2 ∗
are positive. In fact,
B
is
defined
by
B
=
(
1
+
A)
C
(
1
+
A)
C
−
1
.
In
order
to
conclude
µ
! µ
!
A C∗
1 − A −C ∗
that Tµ =
is the minimum in the assertion, it remains to show that
is positive.
C Bµ
−C
1 − Bµ
!
!
1 − A −C ∗
1 + A C∗
Since
is positive just as
, (1 − A)−1/2 C ∗ is a well defined linear operator. By
−C
21
C
21
Corollary I.1.2 it reduces to prove the inequality
2
2
2
(1 − A)−1/2 C ∗ y ≤ 2 kyk − (1 + A)−1/2 C ∗ y .
the positivity of
1+A
But
2 2
(1 + A)−1/2 C ∗ y + (1 − A)−1/2 C ∗ y =
=2
1 − A2
−1
C ∗ y, C ∗ y = 2 1 − A2
−1/2
2
C ∗ y .
1 − A2 C ∗
Since C ∗ C ≤ 1 − A2 implies
≥ 0,
C
1
!
2
1 − A2
−1/2
2
2
C ∗ y ≤ 2 kyk ,
!
∗ A C∗
as expected. Analogously TM =
with BM = 1 − (1 − A)−1/2 C ∗
(1 − A)−1/2 C ∗ is shown to
C BM
be the maximum. This completes the proof.
Theorem I.2. Let A and B be positive
operators on H. Then there is the maximum of all self-adjoint
!
A X
operators X on H for which
are positive.
X B
I. GEOMETRIC AND HARMONIC MEANS
5
−1/2
−1/2
A
X
X
B
!
Proof. Consider first the case that A has bounded inverse. Let D = A
BA
. Then
!
1
A−1/2 XA−1/2
is positive if and only if
is positive. We claim that D1/2 is the maximum
A−1/2 XA−1/2
D
!
!
1 Y
1 D1/2
of all Y for which
are positive. The positivity of
is obvious by Corollary I.1.2.
Y D
D1/2
D
!
1 Y
Suppose that
is positive. By Theorem I.1 there is a contraction W on H such that Y = D1/2 W .
Y D
Introduce a new scalar product on H by hx, yi := (D1/2 x, y). Since Y is self-adjoint hW x, yi = hx, W yi, for
every x, y ∈ H which means that W is self-adjoint with respect to this new scalar product. Since kW k ≤ 1,
for any real number λ with |λ| > 1 the operator λ − W has bounded inverse. The operator (λ − W )−1 is
bounded with respect to the new scalar product, hence
|hW x, xi| ≤ hx, xi
for all
x ∈ H,
which implies −D1/2 ≤ Y ≤ D1/2 . Thus D1/2 is the maximum as claimed. As!a consequence,
1/2 1/2
A X
A1/2 D1/2 A1/2 = A1/2 A−1/2 BA−1/2
A
is the maximum of all X for which
are positive.
X B
If A does not admit
bounded inverse, consider A = A + with > 0. Let C be the maximum of all X
!
A X
for which
are positive. Since obviously C ≥ C0 ≥ 0 whenever > 0 > 0, the operator lim+ C
→0
X B
!
A X
is the maximum of all X for which
are positive. This completes the proof.
X B
We shall call the maximum in Theorem I.2 the geometric mean of two positive operators A and B,
and denote it by A#B.
Corollary I.2.1. Geometric means have the following properties.
(i)
(ii)
(iii)
(iv)
(v)
A#B = B#A.
(αA)#(αB) = α(A#B) for α ≥ 0.
(A1 + A2 )#(B1 + B2 ) ≥ (A1 #B1 ) + (A2 #B2 ).
C(A#B)C ∗ ≤ (CAC ∗ )#(CBC ∗ ) for any linear operator C.
A#A = A, 1#A = A1/2 and 0#A = 0.
1/2 1/2
(vi) A#B = A1/2 A−1/2 BA−1/2
A
if A has bounded inverse.
(vii) A−1 #B −1 = (A#B)−1 if both A and B have bounded inverses.
Proof. (i) to (iv) are immediate from definition. (v) and (vi) are proved as in the proof of Theorem
I.2. (vii) follows from (vi).
As a consequence if A commutes with B, then A#B = (AB)1/2 . This justify the terminology “geometric
mean”.
Corollary I.2.2. If 0 ≤ p ≤ 1, then A ≥ B ≥ 0 implies Ap ≥ B p .
Proof. The assertion is true if p = 1 or 0. Therefore it remains to show that the set ∆ of p for which
the assertion is true is convex. Take p1 , p2 ∈ ∆, and let p = 12 (p1 + p2 ). Then since Ap = Ap1 #Ap2 and
B p = B p1 #B p2 by Corollary I.2.1, Ap ≥ B p follows also from the same Corollary.
6
I. GEOMETRIC AND HARMONIC MEANS
Corollary I.2.3. Let A1 and A2 (resp. !
B1 and B2 ) be positive operators on G (reps. K) and
! let C
∗
∗
Ai C
A1 #A2
C
be a linear operator from G to K. If
are positive i = 1, 2, then so is
.
C Bi
C
B1 #B2
Proof. We may assume the bounded invertibility of A1 and A2 . Then by assumption and Corollary
∗
I.1.2 CA−1
i = 1, 2. Therefore by Corollary I.2.1
i C ≤ Bi
−1
∗
∗
C(A1 #A2 )−1 C ∗ = C A−1
C∗ < CA−1
# CA−1
≤ B1 #B2 ,
1 #A2
1 C
2 C
!
A1 #A2
C∗
which implies the positivity of
by Corollary I.1.2.
C
B1 #B2
Corollary I.2.4. Arithmetic mean is greater than geometric mean, that is
positive A and B.
1
2 (A
+ B) ≥ A#B for
Proof. We may assume the bounded invertibility of A. Then by Corollary I.2.1
1/2
1
1 −1/2
1
1/2
−1/2
−1/2
1/2
1/2
−1/2
A#B = A
A
BA
A
≤A
1+ A
BA
A1/2 = (A + B),
2
2
2
1
because X ≤
1 + X 2 for any positive X.
2
−1
1 −1
−1
The harmonic mean of two positive operators A and B must be defined by
A +B
when
2
both A and B have bounded inverses. We shall denote harmonic mean by A : B. If A and B do not have
bounded inverses, the harmonic mean A : B is defined as the limit of (A + ) : (B + ) as → 0+ .
Theorem I.3. Let A and B be positive operators on a Hilbert space H. Then harmonic mean A : B is
the maximum of all X for which
!
!
2A 0
X X
≥
.
0 2B
X X
Proof. We may assume that both A and B have bounded inverse. Then the above inequality is
equivalent to the condition that for every x ∈ H
(Xx, x) ≤ 2 inf {(Ay, y) + (B(x − y), x − y)}
y
hence to the condition
(
2
|(Bx, y)|
(Xx, x) ≤ 2 (Bx, x) − sup
((A
+ B)y, y)
y
)
.
Since by definition
A : B = 2A(A + B)−1 B = 2 B − B(A + B)−1 B ,
we have
(
)
2
|(Bx, y)|
((A : B)x, x) = 2 (Bx, x) − sup
.
y ((A + B)y, y)
!
!
2A 0
X X
This shows that A : B is the maximum of all X for which
≥
.
0 2B
X X
Corollary I.3.1. Harmonic means have the following properties.
(i)
(ii)
(iii)
(iv)
A : B = B : A.
(αA) : (αB) = α(A : B) for α ≥ 0.
(A1 + A2 ) : (B1 + B2 ) ≥ (A1 : B1 ) + (A2 : B2 ).
C(A : B)C ∗ ≤ (CAC ∗ ) : (CBC ∗ ) for any linear operator C.
I. GEOMETRIC AND HARMONIC MEANS
7
(v) A : A = A, 1 : A = 2A(1 + A)−1 and 0 : A = 0.
(vi) A : B = 2A(A + B)−1 B if A has bounded inverse.
−1
(vii) A−1 : B −1 = 12 (A + B)
if both A and B have bounded inverses.
Corollary I.3.2. Geometric mean is greater than harmonic mean, that is A#B ≥ A : B for positive
A and B.
Proof. We may assume the bounded invertibility of A and B. Then by Corollary I.2.3 A−1 #B −1 ≤
1 −1
A + B −1 . By taking inverse, this yields, by Corollaries I.2.1 and I.3.1, A#B ≥ A : B.
2
Theorem I.4. Let A1 and A2 (resp. !
B1 and B2 ) be positive operators of G (resp. K) and let C
! be a
1
∗
∗
Ai C
(A1 + A2 )
C
linear operator from G to K. If
are positive i = 1, 2, then so are 2
and
C Bi
C
B 1 : B2
!
A1 : A2
C∗
.
1
C
2 (B1 + B2 )
Proof. We may assume the bound invertibility of A1 and A2 . Then by assumption and Corollary I.1.2
≤ Bi i = 1, 2. Therefore by Corollary I.3.1
−1
∗
1
−1
∗
∗
C
(A1 + A2 )
C ∗ = C A−1
C ≤ CA−1
: CA−1
≤ B1 : B2 ,
1 : A2
1 C
2 C
2
!
1
∗
(A
+
A
)
C
1
2
2
which implies the positivity of
by Corollary I.1.2.
The positivity of
C
B 1 : B2
!
A1 : A2
C∗
is proved analogously.
1
C
2 (B1 + B2 )
∗
CA−1
i C
Corollary I.4.1.
The following identity holds for positive A and B;
1
(A + B) #(A : B) = A#B.
2
!
!
A
A#B
B
A#B
Proof. By definition
and
are positive, so that
A#B
B
A#B
A
is positive by Theorem I.4, which implies
1
(A + B) #(A : B) ≥ A#B.
2
1
2 (A
+ B) A#B
A#B
A:B
!
When A and B have bounded inverse, the above inequality, with A−1 and B −1 instead of A and B respectively, leads, by taking inverse, to the inequality
1
(A : B)#
(A + B) ≤ A#B.
2
These two inequalities yield the assertion.
CHAPTER II
Operator-Monotone Functions
Given a finite or infinite open interval (α, β) on the real line, let us denote by S(α, β; H) the totality of
all self-adjoint operators on H whose spectrum are included in the interval (α, β). If there is no confusion,
we shall write simply S(α, β).
A real-valued continuous function f on (α, β) is said to be operator-monotone on (α, β) if A, B ∈
S(α, β; H) and A ≤ B implies f (A) ≤ f (B), where f (A) and f (B) are defined by familiar functional calculi
for self-adjoint operators.
An operator-monotone function is non-decreasing in the usual sense, but the converse is not true. This
chapter is devoted to intrinsic characterization of operator-monotonousness.
Examples II.1. 1. f (λ) := a + bλ with b ≥ 0 is operator-monotone on (−∞, ∞).
2. f (λ) := λp with 0 ≤ p ≤ 1 is operator-monotone on (0, ∞) by Corollary I.2.2.
1
3. For µ ∈
/ (α, β) the function f (λ) := µ−λ
is operator-monotone on (α, β). In particular f (λ) = − λ1 is
operator-monotone on (0, ∞). In fact, if µ < α, then A, B ∈ S(α, β) A ≤ B implies 0 < −(µ − A) ≤
−(µ − B) so that −(µ − A)−1 ≥ −(µ − B)−1 hence f (A) ≤ f (B). If µ > β, then 0 < µ − B < µ − A
hence (µ − B)−1 ≥ (µ − A)−1 .
It is easy to see that a function f is operator-monotone on a finite interval (α, β) if and only if g(λ) :=
f
− α)λ + α + β) is operator monotone on (−1, 1). On this basis, we shall treat only the interval
(−1, 1).
1
2 ((β
Lemma II.1. If a continuously differentiable function f on (−1, 1) is operator-monotone, then for any
N
choice λi ∈ (−1, 1) i = 1, . . . , N the matrix f [1] (λi , λj ) i,j=1 is positive, where f [1] (λ, µ) is defined by
f [1] (λ, µ) =
f (λ)−f (µ)
λ−µ
for λ 6= µ and f [1] (λ, λ) = f 0 (λ).


λ1
0


..
, considered as a self-adjoint operator on the N Proof. Observe the matrix A := 
.


0
λN
dimensional Hilbert space CN . Take any complex numbers ξ1 , . . . , ξN and consider the positive matrix
N
B := ξi ξ¯j i,j=1 . Since A ∈ S −1, 1; CN , for sufficiently small > 0, A + B belongs to S −1, 1; CN . We
claim that for any continuously differentiable function g on (−1, 1)
lim+ −1 {g(A + B) − g(A)} =
→0
N
X
i,j=1
8
g [1] (λi , λj )Pi BPj
II. OPERATOR-MONOTONE FUNCTIONS
9
where Pi is the orthoprojection to the subspace spanned by the vector ei := (δij )N
j=1 . In fact, this is true if
n
g is polynomial, because for g(λ) = λ
−1 {g(A + B) − g(A)} =
n
X
Ak−1 BAn−k + O()
k=1
=
N
X
n
X
λk−1
λn−k
Pi BPj + O()
i
j
i,j=1 k=1
=
N
X
g [1] (λi , λj )Pi BPj + O().
i,j=1
0
Use approximation of g by polynomials on a suitable subinterval when g is merely continuously differentiable.
Now since f is operator-monotone and A + B ≥ A for all > 0, we have
N
X
f [1] (λi , λj )Pi BPj = lim+ −1 {f (A + B) − f (A)} ≥ 0.
→0
i,j=1
Let x =
PN
i,j=1 ei ,
then the positivity of
0≤
N
X
PN
i,j=1
f [1] (λi , λj )Pi BPj implies
f [1] (λi , λj )(Pi BPj x, x) =
i,j=1
N
X
f [1] (λi , λj )ξi ξ¯j .
i,j=1
N
Since (ξi ) are arbitrary, this means that the matrix f [1] (λi , λj ) i,j=1 is positive.
Lemma II.2. If a continuously differentiable function f on (−1, 1) is operator-monotone, then there
exists a finite positive measure m on the closed interval [−1, 1] such that
Z 1
λ
f (λ) = f (0) +
dm(t) for λ ∈ (−1, 1).
−1 1 − λt
Proof. Consider the linear subspace H of C(−1, 1), spanned by the functions fλ (t) := f [1] (λ, t) where
λ runs over (−1, 1). Introduce a scalar product in H by
*
+
X
X
X
αi fλi ,
βj fµj =
f [1] (λi , µj )αi βj .
i
j
i,j
ˆ be the associated Hilbert space. Let us show that
This is positive semi-definite by Lemma II.1. Let H
λn → λ implies kfλn − fλ k → 0. In fact,
2
kfλn − fλ k = f [1] (λn , λn ) − 2f [1] (λn , λ) + f [1] (λ, λ)
= f 0 (λn ) − 2
f (λn ) − f (λ)
+ f 0 (λ).
λn − λ
Now by continuous differentiability of f
f 0 (λn ) − 2
f (λn ) − f (λ)
+ f 0 (λ) → f 0 (λ) − 2f 0 (λ) + f 0 (λ) = 0.
λn − λ
ˆ spanned by all fλ (λ 6= 0). Then D is dense in H,
ˆ as shown above. Define
Let D be the linear subspace of H,
ˆ by
a linear map T from D to H
T fλ := λ−1 {fλ − f0 }
(λ 6= 0).
T is symmetric hence well-defined, because
hT fλ , fµ i =
µf (λ) − λf (λ) f (0)
−
= hfλ , T fµ i .
λµ(λ − µ)
λµ
10
II. OPERATOR-MONOTONE FUNCTIONS
We claim that T admits a (not necessarily bounded) self-adjoint extension Tˆ. This follows from a theorem
of von Neumann, (see [12] p.1231), because the anti-linear involution J, defined by
!
X
X
J
αi fλi =
αi fλi ,
i
i
makes D invariant and commutes with T . Since by definition
(µ − λ)fµ = (1 − λTˆ)(µfµ − λfλ ),
the kernel if 1 − λTˆ is orthogonal to the linear subspace spanned by fµ (µ 6= λ, and µ 6= 0). This subspace
ˆ as shown above, so that 1 − λTˆ is injective, hence (1 − λTˆ)−1 is a (unbounded) self-adjoint
is dense in H
operator. Therefore f0 = (1 − λTˆ)fλ implies fλ = (1 − λTˆ)−1 f0 .
R∞
Now let Tˆ = −∞ tdE(t) be the spectral representation of the self-adjoint operator Tˆ, and dm(t) =
hdE(t)f0 , f0 i. Then for λ 6= 0
λ−1 {f (λ) − f (0)} = f [1] (λ, 0)
D
E
= hfλ , f0 i = (1 − λTˆ)−1 f0 , f0
Z ∞
1
=
dm(t).
1
−
λt
−∞
To complete the proof, it remains to prove that the measure m is concentrated on the closed interval [−1, 1].
Take α > 1, and let us show that [α, ∞) is an m-zero set
Z α−1 Z ∞
Z α−1 Z ∞
1
1
dm(t)dλ ≤
dm(t)dλ
2
(1
−
λt)
(1
−
λt)2
0
α
0
−∞
Z α−1
Z α−1
=
hfλ , fλ i dλ =
f 0 (λ)dλ = f α−1 − f (0) < ∞.
0
By Fubini’s theorem we have
Z α−1 Z
0
∞
α
0
1
dm(t)dλ =
(1 − λt)2
Z
∞
α
∞
Z
≥
α
α−1
Z
0
1
t
Z
1
dλdm(t)
(1 − λt)2
1
s−2 dsdm(t).
0
But this last expression is finite inly if [α, ∞) is an m-zero set. Analogously (−∞, −α] is an m-zero set.
R1
Remark. In Lemma II.2 −1 dm(t) = hf0 , f0 i = f 0 (0).
Lemma II.3. If a continuously differentiable function f on (−1, 1) is operator-monotone, then for any
choice −1 < λ1 < µ1 < λ2 < µ2 < 1
!
f [1] (λ1 , µ1 ) f [1] (λ1 , µ2 )
det
≥ 0.
f [1] (λ2 , µ1 ) f [1] (λ2 , µ2 )
Proof. We shall use the notations in the proof of the previous Lemma. First remark that
Z 1
1
[1]
f (λ, µ) = hfλ , fµ i =
dm(t).
(1
−
λt)(1
− µt)
−1
If the measure m is concentrated on a single point, say t0 , then the determinant in question is equal to a
scalar multiple of
!
(1 − λ1 t0 )−1 (1 − µ1 t0 )−1 (1 − λ1 t0 )−1 (1 − µ2 t0 )−1
det
,
(1 − λ2 t0 )−1 (1 − µ1 t0 )−1 (1 − λ2 t0 )−1 (1 − µ2 t0 )−1
which is just equal to 0.
II. OPERATOR-MONOTONE FUNCTIONS
11
Suppose that m is not concentrated on a single point and that the determinant has negative value.
Since the corresponding determinant with λ1 = µ1 and λ2 = µ2 has non-negative by Lemma II.1, by using
continuity argument there are µ01 and µ02 such that λ1 ≤ µ01 ≤ µ1 and λ2 ≤ µ02 ≤ µ2 , and
!
f [1] (λ1 , µ01 ) f [1] (λ1 , µ02 )
det
= 0.
f [1] (λ2 , µ01 ) f [1] (λ2 , µ02 )
This implies that there are α1 and α2 such that |α1 | + |α2 | 6= 0 and
fλi , α1 fµ01 + α2 fµ02 = 0 (i = 1, 2).
Choose β1 and β2 so that β1 + β2 = α1 + α2 and β1 λ2 + β2 λ1 = α1 µ02 + α2 µ01 . Then it follows that
0 = β1 fλ1 + β2 fλ2 , α1 fµ01 + α2 fµ02
Z 1
{α1 + α2 − (α1 µ02 + α2 µ01 )t}2
=
dm(t).
0
0
−1 (1 − λ1 t)(1 − λ2 t)(1 − µ1 t)(1 − µ2 t)
Since the denominator is strictly positive on [−1, 1] and since the measure m is not concentrated on any single
point, the above relation implies that α1 + α2 = 0 and α1 µ02 + α2 µ01 = 0, which contradicts |α1 | + |α2 | 6= 0.
This contradiction completes the proof.
Now let us return to general operator-monotone functions. Take an infinitely many times differentiable
R1
function φ on (−∞, ∞) such that φ is non-negative, vanishes outside of (−1, 1) and −1 φ(t)dt = 1.
Suppose that f is operator-monotone on (−1, 1). For any 0 < < 1, define a function f() on (−1+, 1−)
by
Z
1 f() (t) :=
φ(s/)f (t − s)ds.
−
The function f() is operator-monotone on (−1 + , 1 − ), because each f (t − ) is so. Obviously f() is
infinitely many times differentiable, and as converges to 0, f() (t) converges to f (t) uniformly on any closed
subinterval of (−1, 1).
Lemma II.4. If a continuous function f on (−1, 1) is operator-monotone, then on any closed subinterval
(α, β) of (−1, 1) f satisfies Lipschitz condition, that is, supα<s<t<β f [1] (t, s) < ∞.
Proof. Let us use the notations in the preceding comment. Let λ1 = 12 (α − 1) and µ2 = 12 (β + 1), and
apply Lemma II.3 to the operator-monotone function f() with λ2 = t and µ1 = s. By taking limit, we can
conclude that
!
f [1] (λ1 , s) f [1] (λ1 , µ2 )
det
≥ 0,
f [1] (t, s)
f [1] (t, µ2 )
or equivalently
f [1] (λ1 , µ2 )f [1] (t, s) ≤ f [1] (λ1 , s)f [1] (t, µ2 ).
The right side of the above inequality is bounded when s < t run over [α, β]. Since f is non-decreasing, both
f [1] (λ1 , µ2 ) and f [1] (t, s) are non-negative. Therefore if f [1] (λ1 , µ2 ) 6= 0, then f [1] (t, s) are bounded when
s < t run over [α, β]. Finally f [1] (λ1 , µ2 ) = 0 implies f [1] (t, s) = 0 for all s < t in [α, β]. This completes the
proof.
Theorem II.1. In order that a continuous function f on (−1, 1) is operator-monotone, it is necessary
and sufficient that there is a finite positive measure m on [−1, 1] such that
Z 1
λ
f (λ) = f (0) +
dm(t) for λ ∈ (−1, 1).
1
−
λt
−1
12
II. OPERATOR-MONOTONE FUNCTIONS
Proof. Suppose that f admits a representation of the above form. For each t ∈ [−1, 1] the function
λ
ht (λ) := 1−λt
is operator-monotone on (−1, 1). In fact, if t = 0, h0 (λ) = λ is operator-monotone. If t 6= 0,
t−2
t−1 −λ is operator-monotone as stated in one of the
R1
average −1 ht (λ)dm(t) is operator-monotone, and so is f .
ht (λ) = −t−1 +
Examples II.1. As a consequence, the
weighted
Suppose conversely that f is operator-monotone. Then with the notations in front of Lemma II.4, for
each 0 < < 1 f() ((1 − )λ) is operator-monotone on (−1, 1). Then since f() ((1 − )λ) is continuously
differentiable, by Lemma II.2 there is a measure m on [−1, 1] such that
Z 1
λ
f() ((1 − )λ) = f() (0) +
dm (t).
−1 1 − λt
R1
We claim that −1 dm (t) are bounded when runs over (0, 1/2). As remarked after Lemma II.2, we have
Z 1
0
dm (t) = f()
(0).
−1
Since Lemma II.4 f satisfies Lipshitz condition on any closed subinterval of (−1, 1), it has derivative f 0 (t)
for almost all λ and ess. sup |f 0 (t)| = r < ∞. Further it is easy to see that
−<t<
0
f()
(0) =
(1 − )
Z
f 0 (−s)φ
−
s
ds ≤ (1 − )r .
R1
These considerations show the boundedness of −1 dm (t) when runs over (0, 1/2). Now by the Helly
theorem (see [11] p. 381) there is a sequence n → 0 and a measure m on [−1, 1] such that
Z 1
Z 1
lim
g(t)dmn (t) =
g(t)dm(t)
n→∞
−1
−1
for all continuous functions g on [−1, 1]. Therefore for λ ∈ (−1, 1)
Z 1
Z 1
λ
λ
lim
dmn (t) =
dm(t).
n→∞ −1 1 − λt
1
−
λt
−1
Finally the assertion follows from
lim f(n ) ((1 − n )λ) = f (λ).
n→∞
The integral representation in Theorem II.1 shows that an operator-monotone function f on (−1, 1) is
necessarily infinitely many times differentiable. Further more it admits analytic continuation to the upper
and lower open half planes by
Z 1
ζ
fˆ(ζ) = f (0) +
dm(t) for ζ with Im (ζ) 6= 0.
1
−
ζt
−1
The function fˆ maps the upper half plane to itself. Indeed
Z 1 Im (ζ)
ˆ
Im f (ζ) =
2 dm(t).
−1 |1 − ζt|
This observation is completed in the following theorem.
Theorem II.2. Let f be a real-valued continuous function on a finite or infinite interval (α, β). In order
ˆ
that f is operator-monotone it is necessary
andsufficient that it admits an analytic continuation f to the
upper and lower half planes such that Im fˆ(ζ) > 0 for Im (ζ) > 0.
II. OPERATOR-MONOTONE FUNCTIONS
13
Proof. Suppose that f admits an analytic continuation fˆ of the type mentioned above. Then by a
well-known theorem of Nevanlina (see [1] p. 7) there are a real number a, a non-negative number b and a
finite positive measure m
ˆ on Ω := (−∞, ∞) \ (α, β) such that
Z
1 + ζt
fˆ(ζ) = a + bζ +
dm(t).
ˆ
Ω t−ζ
The function g(λ) := a + bλ with b ≥ 0 is obviously operator-monotone on (α, β). For each t ∈
/ (α, β) the
function ht (λ) := 1+λt
is
operator
monotone,
because
t−λ
ht (λ) = −t +
1 + t2
t−λ
1
and the function t−λ
is operator-monotone on (α, β). Therefore the weighted average f (λ) is also operatormonotone on (α, β). This completes the proof.
Corollary II.2.1. If f is operator-monotone on (−∞, ∞), then f (λ) = a + bλ with some real a and
non-negative b.
This follows immediately from the integral representation in the proof of Theorem II.2, because the
measure m
ˆ must vanish.
CHAPTER III
Operator-Convex Functions
The notion next to monotoneousness seems convexity. In this respect, a real-valued continuous function
f on a finite or infinite interval (α, β) is said to be operator-convex if
1
1
f
(A + B) ≤ {f (A) + f (B)}
forA, B ∈ S(α, β).
2
2
The function f is said to be operator-concave if −f is operator-convex.
An operator-convex function is convex in the usual sense, but the converse is not true. This chapter is
devoted to intrinsic characterization of operator-convexity.
Examples III.1. 1. f (λ) := a + bλ is operator-convex on (−∞, ∞).
2. f (λ) := λ2 is operator-convex on (−∞, ∞). In fact, for self-adjoint A and B
2
1 2
1
1
A + B2 −
(A + B) = (A − B)2 ≥ 0.
2
2
4
1
3. For µ < α the function f (λ) := λ−µ
is operator-convex on (α, ∞). In particular, f (λ) :=
convex on (0, ∞). In fact, for A, B ∈ S(α, β), A − µ and B − µ belong to S(0, ∞) and
1
λ
is operator-
1
(A − µ)−1 + (B − µ)−1 = {(A − µ) : (B − µ)}−1 .
2
By Corollary I.2.4 and I.3.2.
−1
1
1
−1
{(A − µ) : (B − µ)} ≥
(A − µ) + (B − µ)
2
2
−1
1
=
(A + B) − µ
.
2
Lemma III.1. Let f be a twice continuously differentiable function on (−1, 1). If f is operator-convex,
then for each µ ∈ (−1, 1) the function g(λ) := f [1] (µ, λ) is operator-monotone. Conversely if f [1] (0, λ) is
operator-monotone, then f is operator-convex.
Proof. Let f be operator-convex. Obviously g is continuously differentiable on (−1, 1). Inspection
of Lemmas II.1 and II.2 will show that for the operator-monotoneousness of g it suffices to prove that
N
for any choice λi ∈ (−1, 1) i = 1, 2, . . . , N the matrix g [1] (λi , λj )
is positive. Observe the matrix
i,j=1


λ1
0


A := 
λN
 with λN +1 = µ, considered as a self-adjoint operator on the (N + 1)-dimensional
0
λN +1


ξ¯1

.. 

0
. 
.
Hilbert space CN +1 . Take any complex numbers ξ1 , . . . , ξN and consider the matrix B := 



ξ¯N 
ξ , . . . , ξN 0
1
N +1
N +1
Since A ∈ S −1, 1; C
, for sufficiently small > 0 A + B belongs to S −1, 1; C
. Since f is twice
14
III. OPERATOR-CONVEX FUNCTIONS
15
continuously differentiable, just as in the proof of Lemma II.1, it can be shown that the matrix-valued
function f (A + B) is twice differentiable and
N
+1
X
d2 f (A + B) =
f [2] (λi , λj , λk )Pi BPj BPk
d2
=0
i,j,k=1
N +1
where Pi is the orthoprojection to the subspace spanned by the vector ei := (δij )j=1 and f [2] (s, t, u) =
d2 f (A + B) [1]
[1]
hs (t, u) with hs (t) = f (s, t). Now the operator-convexity of f implies
≥ 0, as in the
d2
=0
PN
case of scalar convex functions. Let x = i=1 ei . Then
0≤
N
+1
X
f [2] (λi , λj , λk )(Pi BPj BPk x, x) =
i,j,k=1
=
N
X
f [2] (λi , λN +1 , λk )ξk ξ¯i =
i,k=1
N
X
g [1] (λi , λk )ξk ξ¯i .
i,k=1
N
Since (ξi ) are arbitrary, this means that the matrix g [1] (λi , λj ) i,j=1 is positive.
Suppose conversely that f [1] (0, λ) is operator-monotone. Then by Theorem II.1 there exists a finite
positive measure m on [−1, 1] such that
Z 1
λ
f [1] (0, λ) = f 0 (0) +
dm(t),
1
−
λt
−1
hence
Z
1
f (λ) = a + bλ +
−1
λ2
dm(t),
1 − λt
2
0
λ
where a = f (0) and b = f (0). For each t ∈ [−1, 1] the function ht (λ) := 1−λt
is operator-convex. In fact, if
2
t = 0, ht (λ) = λ , and if t 6= 0,
t−2
ht (λ) = −t−2 − t−1 λ +
.
1 − λt
These functions are operator-convex, as shown in Example III.1. Since the function a+bλ is operator-convex,
too, the weighted average f is operator-convex. This completes the proof.
Theorem III.1. If a continuous function f on (−1, 1) is operator-monotone, then both the function
Rλ
f1 (λ) = 0 f (t)dt and f2 (λ) := λf (λ) are operator-convex.
Proof. Since f is continuously differentiable by Theorem II.2, both f1 and f2 are twice continuously
differentiable. Now since
Z 1
[1]
[1]
f1 (0, λ) =
f (sλ)ds and f2 (0, λ) = f (λ),
0
the assertion follows from Lemma III.1.
Theorem III.2. In order that a continuous function f on (−1, 1) is operator-convex, it is necessary and
sufficient that there are real numbers a and b, and a finite positive measure m on [−1, 1] such that
Z 1
λ2
f (λ) = a + bλ +
dm(t)
for λ ∈ (−1, 1).
−1 1 − λt
Proof. Sufficiency was shown already in the proof of Lemma III.1. Suppose that f is operator-convex.
By using the notations in the proof of Lemma II.4, f() ((1−)λ) is operator-convex for each 0 < < 1. Theref() ((1 − )λ) − f() ((1 − )µ)
fore by Lemma III.1 for any µ ∈ (−1, 1) the function
is operator-monotone on
λ−µ
(−1, 1) so that by taking limit as → 0 f [1] (µ, λ) is operator-monotone on (µ + δ, 1) as well as (−1, µ − δ)
16
III. OPERATOR-CONVEX FUNCTIONS
for any 0 < δ. By Theorem II.2 this implies that f is twice continuously differentiable on (−1, 1). Now the
argument of the proof of Lemma III.1 can be applied.
Now let us consider functions on the half-line (0, ∞).
Theorem III.3. An operator-monotone function f on (0, ∞) is operator-concave.
Proof. It is seen from the proof of Theorem II.2 that f admits a representation
Z 0
1 + λt
f (λ) = a + bλ +
dm(t)
−∞ t − λ
where b ≥ 0 and m is a positive measure. It suffices to prove that for each t < 0 the function ht (λ) := 1+λt
t−λ
is operator-concave. If t = 0, h0 (λ) = −1/λ is operator-concave on (0, ∞), as mentioned in Example III.1.
If t < 0,
1 + t2
ht (λ) = −t −
λ−t
is also operator-concave on (0, ∞), as shown in Example III.1. Since the function a + bλ is obviously
operator-concave, the weighted average f is operator-concave, too.
Corollary III.3.1. If a function f is operator-monotone and f (λ) > 0 on (0, ∞), then g(λ) := f (λ)−1
is operator-convex.
Proof. Take A, B ∈ S(0, ∞). Since f is operator-concave and the function 1/λ is operator-convex on
(0, ∞),
1
1
f
(A + B) ≥ {f (A) + f (B)}
2
2
and
g
1
(A + B)
2
−1
1
≤
{f (A) + f (B)}
2
1
≤ {g(A) + g(B)}.
2
Thus g is operator-convex.
Corollary III.3.2. A function on (0, ∞) is operator-monotone and operator-convex at the same time,
only if it is of the form a + bλ.
Theorem III.4. A continuous function f on (0, ∞) with f (0) := lim→0+ f () = 0 is operator-convex if
and only if f (λ)/λ is operator-monotone.
Proof. Suppose that f is operator-convex. By Theorem III.2 it is infinitely many times differentiable.
()
Then by Lemma III.1 for each > 0 the function f (λ)−f
is operator-monotone hence, as the limit, the
λ−
function f (λ)/λ is operator-monotone. Suppose conversely that f (λ)/λ is operator-monotone. Then as in
the proof of Theorem III.3 there are a and b > 0 and a measure m such that
Z 0
1 + tλ
f (λ)/λ = a + bλ +
dm(t),
−∞ t − λ
hence
f (λ) = aλ + bλ2 +
Z
0
−∞
λ(1 + λt)
dm(t).
t−λ
III. OPERATOR-CONVEX FUNCTIONS
17
Since bλ2 is operator-convex, it suffices to prove that for each t < 0 the function ht (λ) :=
operator-convex on (0, ∞). For t = 0 this is obvious. If t 6= 0,
ht (λ) = {−(1 + t2 ) − tλ} +
λ(1+λt)
t−λ
is
(1 + t2 ) |t|
λ−t
is operator-convex, as was shown in Example III.1.
Let us investigate for what exponent −∞ < s < ∞ the function f (λ) := λs on (0, ∞) is operatormonotone, operator-convex or operator-concave.
Examples III.2. 1. f (λ) = λs is operator-monotone (or operator-concave) if and only if 0 ≤ s ≤ 1.
This follows from Corollary I.2.2 and Theorem III.3 and from the fact that for s > 1 or < 0 the function
f is not concave.
2. f (λ) = λs is operator-convex if and only if 1 ≤ s ≤ 2 or −1 ≤ s ≤ 0. In fact, by Theorem III.4
for s > 0 f (λ) = λs is operator-convex if and only if λs−1 is operator-monotone. Therefore for s > 0
f is operator-convex if and only if 1 ≤ s ≤ 2. By Corollary III.3.1 f (λ) = λs is operator-convex for
s
−1
−1 ≤ s ≤ 0. For s < −1 the function λλ−1
does not admit any analytic continuation fˆ to the upper half
plane such that Im (fˆ(ζ)) > 0 for Im (ζ) > 0, hence f (λ) = λs is not operator-convex by Lemma III.1.
Lemma III.2. Let f (λ) > 0 on (0, ∞) and g(λ) := f (λ−1 )−1 . If f is operator-monotone, so is g. If f is
operator-convex and f (0) = 0, the function g is operator-convex.
Proof. Suppose that f is operator-monotone. Then by Theorem II.2 it admits an analytic continuation
fˆ to the complement of the closed negative real semi-axis such that Im fˆ(ζ) > 0 or < 0 according as
−1
Im (ζ) > 0 or Im (ζ) < 0. Then gˆ(ζ) := fˆ ζ −1
is an analytic continuation of g to the complement of
the closed negative real semi-axis such that Im (g(ζ)) > 0 or < 0 according as Im (ζ) > 0 or < 0. Therefore
again by Theorem II.2 g is operator-monotone.
Suppose next that f is operator-convex and f (0) = 0. Then by Theorem III.4 the function h(λ) :=
f (λ)/λ is operator-monotone. By applying the first part of this lemma to h, we can conclude that the
−1
function h λ−1
= g(λ)/λ is operator-monotone, hence by Theorem III.4 g is operator-convex.
Theorem III.5. Let f be a continuous positive function on (0, ∞) and A, B positive operators. If f is
operator-monotone then
f (A : B) ≤ f (A) : f (B).
If f is operator-convex and f (0) = 0 then
f (A : B) ≥ f (A) : f (B).
Proof. Suppose that f is operator-monotone. Then by Lemma III.2 g(λ) := f λ−1
monotone, hence operator-concave by Theorem III.2. Therefore
1 −1
1
−1
g
A +B
≥
g A−1 + g A−1
2
2
−1
is operator-
hence
f (A : B) ≤ f (A) : f (B).
The other assertion can be proved quite analogously by using Lemma III.2.
CHAPTER IV
Positive Maps
A (non-linear) transformation which maps L+ (H), the set of positive operators on H, to L+ (K) will be
called positive. In this chapter we shall study some special classes of positive maps.
Let us start with positive linear maps. A positive linear map Φ from L(H) to L(K) preserves orderrelation, that is, A ≤ B implies Φ(A) ≤ Φ(B), and preserves adjoint operation, that is Φ(A∗ ) = Φ(A)∗ . It
is said to be normalized if it transforms 1H to 1K . If Φ is normalized, it maps S(α, β; H) to S(α, β; K).
Lemma IV.1. A normalized positive linear map φ has the following properties.
(i) Φ A2 ≥ Φ(A)2 for A ∈ S(−∞, ∞; H).
(ii) Φ A−1 ≥ Φ(A)−1 for A ∈ S(0, ∞; H).
(i) By Remark after Corollary I.1.1 it suffices to prove the positivity of
R∞
Consider the spectral representation A = −∞ tdE(t). Since
Z ∞
Z ∞
A2 =
t2 dE(t)
and 1 =
dE(t),
Proof.
−∞
Φ A2
Φ(A)
!
Φ(A)
.
Φ(1)
−∞
we have, with tensor product notation,
! Z
!
∞
Φ A2
Φ(A)
t2 t
=
⊗ dE(t).
Φ(A) Φ(1)
t 1
−∞
!
t2 t
Since 2 × 2 matrices
are positive, for all −∞ < t < ∞ the right hand of the above expression
t 1
is positive.
!
Φ(A)
Φ(1)
. Since A is positive by assumption, it is written
(ii) It suffices to prove the positivity of
Φ(1) Φ A−1
R∞
in the form
! A = 0+ tdE(t). Now the positivity in question follows from the positivity of matrices
t
1
1
t
−1
for 0 < t < ∞, as in the proof of (i).
Theorem IV.1. Let Φ be a normalized positive linear map. If f is an operator-convex function on
(α, β), then
f [Φ(A)] ≤ Φ[f (A)]
for A ∈ S(α, β; H).
Proof. It suffices to consider the case (α, β) = (−1, 1). By Theorem III.2 f admits a representation
Z 1
λ2
f (λ) = a + bλ +
dm(t)
−1 1 − λt
with b ≥ 0 and a positive measure m. Since for A ∈ S(−1, 1; H)
Z 1
Φ[f (A)] = a + bΦ(A) +
Φ A2 (1 − tA)−1 dm(t)
−1
18
IV. POSITIVE MAPS
and
Z
1
f [Φ(A)] = a + bΦ(A) +
−1
19
Φ(A)2 {1 − tΦ(A)}−1 dm(t),
it suffices to show
Φ A2 (1 − tA)−1 ≥ Φ(A)2 {1 − tΦ(A)}−1
for
− 1 ≤ t ≤ 1.
For t = 0, this follows from Lemma IV.1. For t 6= 0, again by Lemma IV.1
Φ A2 (1 − tA)−1 = −t2 − t−1 Φ(A) + t2 φ (1 − tA)−1
≥ −t−2 − t−1 Φ(A) + t−2 {1 − tΦ(A)}−1
= Φ(A)2 {1 − tΦ(A)}−1 .
This completes the proof.
Corollary IV.1.1. Let Φ be a normalized positive linear map. Then for a positive operator A
Φ (Ap ) ≥ Φ(A)p
(1 ≤ p ≤ 2)
and
φ (Ap ) ≤ Φ(A)p
(0 ≤ p ≤ 1).
Proof. As shown in Examples III.1, λp is operator-convex on (0, ∞) for 1 ≤ p ≤ 2 while −λp is
operator-convex for 0 ≤ p ≤ 1.
1/p
Corollary IV.1.2. If Φ is a normalized positive map and if A is a positive operator, then Φ (Ap )
1/q
Φ (Aq )
whenever 1 ≤ p ≤ q or 12 q ≤ p ≤ 1 ≤ q.
≤
p/q
Proof. By Corollary IV.1.1 Φ (Aq )
≥ Φ (Ap ) whenever p ≤ q. If p ≥ 1in addition, this implies
1/p
q 1/q
p 1/p
Φ (A )
≥ Φ (A )
by Corollary I.2.2. If q ≥ 1 ≥ p ≥ 12 q, Φ (Ap )
≤ Φ(Aq )1/q , because λ1/q is
operator-monotone.
Corollary IV.1.3. If Φ is a positive linear map, for any positive operators A and B
Φ(A : B) ≤ Φ(A) : Φ(B)
and
Φ(A#B) ≤ Φ(A)#Φ(B).
Proof. We may assume A ∈ S(0, ∞; H). Let G be the closure of the range of Φ(A). Then there is
uniquely a normalized positive linear map Ψ from L(H) to L(G) such that
Φ(A)1/2 Ψ(C)Φ(A)1/2 = Φ A1/2 CA1/2
for C ∈ L(H).
Since the functions f (λ) :=
for any positive C
2λ
1+λ
and g(λ) := λ1/2 are operator-concave on (0, ∞), by Theorem IV.1 we have
Ψ(1 : C) = Ψ(f (C)) ≤ f (Ψ(C)) = 1 : Ψ(C)
and
Ψ(1#C) = Ψ(g(C)) ≤ g(Ψ(C)) = 1#Ψ(C).
With C = A−1/2 BA−1/2 this leads to the following
Φ(A : B) = Φ(A)1/2 Ψ(1 : C)Φ(A)1/2
= Φ(A)1/2 {1 : Ψ(C)}Φ(A)1/2 = Φ(A) : Φ(B)
and analogously
Φ(A#B) ≤ Φ(A)#Φ(B).
20
IV. POSITIVE MAPS
The converse of Theorem IV.1 is also true.
Theorem IV.2. Let α ≤ 0 ≤ β and f a continuous function on (α, β) with f (0) = 0. If
Φ(f (A)) ≥ f (Φ(A))
for every normalized positive linear map Φ and A ∈ S(α, β) then f is operator-convex.
Proof. Let K be the subspace of H ⊕ H, consisting
of all vectors x ⊕ x. Define the linear
!
! map Φ from
A11 A12
1 A11 + A22 A11 + A22
L(H ⊕ H) to L(K), that assigns to
the operator
, considered on
4 A11 + A22 A11 + A22
A21 A22
K. Then Φ is positive and normalized. Take A, B ∈ S(α, β; H). Then by definition we have
"
!#
!
!
A 0
f (A)
0
1 f (A) + f (B) f (A) + f (B)
Φ f
=Φ
=
,
4 f (A) + f (B) f (A) + f (B)
0 B
0
f (B)
while
"
A
f Φ
0
0
B
!#
"
1
1
!
!
!#
−1
0
0
1 1
1
√
1
0 12 (A + B)
2 −1 1
!
!
!
f (0)
0
1 1
1 1 −1
1
√
=√
0
f 12 (A + B)
2 1 1
2 −1 1
!
1 f 12 (A + B) f 12 (A + B)
.
=
2 f 12 (A + B) f 12 (A + B)
"
A 0
0 B
1
=f √
2
Now by assumption
Φ f
!#
"
A 0
≥f Φ
0 B
!#
which implies
f
1
1
(A + B) ≤ {f (A) + f (B)}.
2
2
This completes the proof.
A (non-linear) map Φ from a convex subset of L(H) to L(K) is said to be a convex map (resp. a
concave map) if Φ 12 (A + B) ≤ 12 {Φ(A) + Φ(B)} (resp. Φ 12 (A + B) ≥ 12 {Φ(A) + Φ(B)}).
We shall be concerned with convexity or concavity of maps Φs,t (A) = As ⊗ At defined on S(0, ∞; H)
where −∞ < s, t < ∞.
Lemma IV.2. If Φ and Ψ are concave maps with range in S(0, ∞; K) then the maps Θ(A) := Φ(A)#Ψ(A)
and Ξ(A) := Φ(A) : Ψ(A) are concave.
Proof. By Corollary I.2.1 and concavity of Φ and Ψ
1
1
1
Θ
(A + B) = Φ
(A + B) #Ψ
(A + B)
2
2
2
1
1
1
1
≥
Φ(A) + Φ(B) #
Ψ(A) + Ψ(B)
2
2
2
2
1
≥ {Φ(A)#Ψ(A) + Φ(B)#Ψ(B)
2
1
= {Θ(A) + Θ(B)},
2
which proves the concavity of Θ. The concavity of Ξ is proved analogously by using Corollary I.3.1.
Theorem IV.3. The map Φp,q (A) := AP ⊗ Aq is concave if 0 ≤ p, q and p + q ≤ 1.
IV. POSITIVE MAPS
21
Proof. Consider the set Ω of (p, q) in R2+ for which Φp,q are concave. We claim that Ω is a convex set.
In fact, let (pi , qi ) ∈ Ω and p = 12 (p1 + p2 ), q = 12 (q1 + q2 ). Then since
Φp,q (A) = Ap ⊗ Aq
= (Ap1 #Ap2 ) ⊗ (Aq1 #Aq2 )
= (Ap1 ⊗ Aq1 ) # (Ap2 ⊗ Aq2 )
= Φp1 ,q1 (A)#Φp2 ,q2 (A),
by Lemma IV.2 the map Φp,q is concave. Obviously (0, 0), (1, 0) and (0, 1) belong to Ω, hence so does (p, q)
for which 0 ≤ p, q and p + q ≤ 1.
Corollary IV.3.1. For positive Ai and Bi (i = 1, 2)
(A1 : B1 )p ⊗ (A2 : B2 )q ≤ (Ap1 ⊗ Aq2 ) : (B1p ⊗ B2q )
whenever 0 ≤ p, q and p + q ≤ 1.
Proof. The concavity of Φp,q is seen to be equivalent to the inequality
With A =
A1
0
(A : B)p ⊗ (A : B)q ≤ (Ap ⊗ Aq ) : (B p ⊗ B q ) .
!
!
0
B1 0
and B =
this inequality implies the inequality in the assertion.
A2
0 B2
Theorem IV.4. If f and g are positive, operator-monotone functions on (0, ∞), then the map Φ(A) :=
f (A)−1 ⊗ g(A)−1 is convex.
−1
−1
Proof. Let h(λ) := f λ−1
and k(λ) := g λ−1
. Then the convexity of Φ is seen to be equivalent
to that for A, B ∈ S(0, ∞)
1
{h(A) ⊗ k(A) + h(B) ⊗ k(B)}.
2
By Lemma III.2 both h and k are operator-monotone, so that by Theorem III.5
h(A : B) ⊗ k(A : B) ≤
{h(A : B) ⊗ k(A : B)} ≤ {h(A) : h(B)} ⊗ {k(A) : k(B)}
and further by Corollaries I.3.2 and I.2.4
{h(A) : h(B)} ⊗ {k(A) : k(B)} ≤ {h(A)#h(B)} ⊗ {k(A)#k(B)}
= {h(A) ⊗ k(A)}#{h(B) ⊗ k(B)}
≤
1
{h(A) ⊗ k(A) + h(B) ⊗ k(B)}.
2
Corollary IV.4.1. The map Φ−p,−q (A) = A−p ⊗ A−q is convex if 0 ≤ p, q ≤ 1.
This follows from Theorem IV.3 and Corollary I.2.2.
Theorem IV.5. The map Φ−p,q (A) = A−p ⊗ Aq is convex if 0 ≤ p ≤ q − 1 < 1.
Proof. Let us consider first the case q = p + 1 < 2. The well-known integral representation of λp for
0 < p < 1 (see [11], [12] p. )
Z ∞
p
−1
λ = π sin(pπ)
tp−1 λ(λ + t)−1 dt
0
22
IV. POSITIVE MAPS
is applied to get
(1 ⊗ A)
Z ∞
−1
= π −1 sin(pπ)
t1−p A−1 ⊗ A2 A−1 ⊗ A + t
dt.
Φ−p,p+1 (A) = A−1 ⊗ A
p
0
Therefore it suffices to prove that for each t > 0 the map
−1
Φt (A) := A−1 ⊗ A2 A−1 ⊗ A + t
is convex. To this end, remark, first of all
n
Φt (A) = 1 ⊗ A − t (A ⊗ 1)−1 +
1 ⊗ t−1 A
−1 o−1
,
the first term of which is a convex map. It remains to show that the map
n
−1 o−1
Θt (A) := (A ⊗ 1)−1 + 1 ⊗ t−1 A
is convex. But this follows from Lemma IV.2, because 2Θt (A) = Ψ(A) : Ξ(A) where Ψ(A) := A ⊗ 1 and
Ξ(A) := 1 ⊗ t−1 A are obviously concave.
Next let us consider the case p < q − 1. Let r = p(q − 1). Since 0 < r < 1, the function λr is
operator-concave by Theorem III.4, hence for positive A and B
r
1
1
1
1
A+ B
≥ Ar + B r ,
2
2
2
2
which implies
1
1
A+ B
2
2
−r
≥
1 r 1 r
A + B
2
2
−1
Further the operator-monotoneousness of λq−1 yields
−p −(q−1)
1
1
1 r 1 r
A+ B
≥
A + B
.
2
2
2
2
It follows from the first part of the proof, as Corollary IV.3.1 does from Theorem IV.3, that for positive
A, B, C and D
−(q−1) q
o
1
1
1
1
1 n −(q−1)
C+ D
⊗
A+ B
≤
C
⊗ Aq + D−(q−1) ⊗ B q .
2
2
2
2
2
Now use these inequalities with C = Ar and D = B r to get
−p q −(q−1) q
1
1
1
1
1 r 1 r
1
1
A+ B
⊗
A+ B
≤
A + B
⊗
A+ B
2
2
2
2
2
2
2
2
1 −p
≤
A ⊗ Aq + B −p ⊗ B q .
2
This shows the convexity of Φ−p,q .
It remains to consider the case p = 1 and q = 2. The convexity of Φ−1,2 means that
(A + B)−1 ⊗ (A + B)2 ≤ A−1 ⊗ A2 + B −1 ⊗ B 2 .
With C = A−1/2 BA−1/2 this inequality is equivalent to the inequality
(1 + C)−1 ⊗ (1 + C)A(1 + C) ≤ 1 ⊗ A + C −1 ⊗ CAC,
and further to the inequality
C ⊗ (CA + AC) ≤ C 2 ⊗ A + 1 ⊗ CAC.
IV. POSITIVE MAPS
Considering the spectral representation A =
A is an orthoprojection, i.e. A2 = A. Then
R∞
0
23
λdE(λ), it suffices to prove the above inequality for the case
C 2 ⊗ A + 1 ⊗ CAC − C ⊗ (CA + AC) = (C ⊗ A − 1 ⊗ AC)∗ (C ⊗ A − 1 ⊗ AC) ≥ 0.
This completes the proof.
!
A 0
instead of A, it is seen that the maps A 7→ As , A 7→ At and R+ 3
0 α
α → αs+t are concave, hence as mentioned in Example III.2 we have 0 ≤ s, t and s + t ≤ 1.
If Φs,t is convex with s ≤ t, just as above, the maps A 7→ As and A 7→ At are convex, so that
1 ≤ s ≤ t ≤ 2 or −1 ≤ s ≤ t ≤ 0 or −1 ≤ s ≤ 0 ≤ 1 ≤ t ≤ 2. Replacing A by scalar, we see that the map
{α, β} → αs β t is convex on the positive cone of R2 . Therefore, for arbitrarily fixed α, β > 0, the function
φ(λ) := (α + λ)s (β − λ)t is a convex function of λ in a neighborhood of 0. By differentiation this implies
If Φs,t is concave, using
s(s − 1)αs−2 β t − 2stαs−1 β t−1 + t(t − 1)αs β t−2 ≥ 0.
If −1 ≤ s ≤ 0 ≤ 1 ≤ t ≤ 2 or 1 ≤ s ≤ t ≤ 2, by arbitrariness of α and β the above inequality is equivalent to
s2 t2 ≤ st(s − 1)(t − 1).
If −1 ≤ s ≤ 0 < 1 ≤ t ≤ 2, this is possible only if s + t ≥ 1. But the case 1 ≤ s ≤ t ≤ 2 is not consistent
with the inequality. Thus we can conclude that Theorem IV.3 exhausts all the case Φs,t is concave while
Theorem IV.4 and IV.5 exhaust all the case Φs,t (s ≤ t) is convex.
Note
Chapter I. Corollary I.1.3 is due to Krein [16]. Geometric mean was introduced by Pusz and Woronowicz
[20], who proved Theorem I.2. Corollary I.2.3 can be considered as a non-commutative version of the result of
−1
Lieb & Ruskai [18]. Anderson & Duffin [2] defined A−1 + B −1
as parallel sum of two positive matrices.
Hilbert space operator case was treated in Anderson & Trapp [3], who proved Theorem I.3.
Chapter II and III. The content of these chapters is the famous theory of L¨owner [19] and Kraus [15].
The full account of the theory can be found in Donoghue [10] and Davis [9]. Theorem II.1 and II.2 are due
to Bendat & Sherman [6]. The Hilbert space method in Chapter II is due to Koranyi [14].
Chapter IV. Theorem IV.1 is due to Davis [8] and Choi [7], while Theorem IV.2 is pointed out in Davis
[9]. The inequality Φ(A : B) ≤ Φ(A) : Φ(B) was proved for a special case by Anderson & Trapp [4].
Theorem IV.3 and Corollary IV.4.1 were proved by Lieb [17] by a different method. Epstein [13] gave a
simpler proof. Uhlmann [21] also used geometric means to prove Theorem IV.3. The idea in the proof of
Theorem IV.5 will be developed in a forthcoming paper [5].
24
Bibliography
[1]
[2]
[3]
[4]
N. I. Akhiezer and I.M. Glazman. Theory of linear operators in Hilbert space. Pitman Pub., Boston :, 1981.
W. N. Anderson, Jr. and R. J. Duffin. Series and parallel addition of matrices. J. Math. Anal. Appl., 26:576–594, 1969.
W. N. Anderson, Jr. and G. E. Trapp. Shorted operators. II. SIAM J. Appl. Math., 28:60–71, 1975.
William N. Anderson, Jr. and George E. Trapp. A class of monotone operator functions related to electrical network theory.
Linear Algebra and Appl., 15(1):53–67, 1976.
[5] T. Ando. Concavity of certain maps on positive definite matrices and applications to Hadamard products. Linear Algebra
Appl., 26:203–241, 1979.
[6] Julius Bendat and Seymour Sherman. Monotone and convex operator functions. Trans. Amer. Math. Soc., 79:58–71, 1955.
[7] Man Duen Choi. A Schwarz inequality for positive linear maps on C ∗ -algebras. Illinois J. Math., 18:565–574, 1974.
[8] Chandler Davis. A Schwarz inequality for convex operator functions. Proc. Amer. Math. Soc., 8:42–44, 1957.
[9] Chandler Davis. Notions generalizing convexity for functions defined on spaces of matrices. In Proc. Sympos. Pure Math.,
Vol. VII, pages 187–201. Amer. Math. Soc., Providence, R.I., 1963.
[10] William F. Donoghue, Jr. Monotone matrix functions and analytic continuation. Springer-Verlag, New York, 1974. Die
Grundlehren der mathematischen Wissenschaften, Band 207.
[11] Nelson Dunford and Jacob T. Schwartz. Linear operators. Part I. Wiley Classics Library. John Wiley & Sons Inc., New
York, 1988. General theory, With the assistance of William G. Bade and Robert G. Bartle, Reprint of the 1958 original,
A Wiley-Interscience Publication.
[12] Nelson Dunford and Jacob T. Schwartz. Linear operators. Part II. Wiley Classics Library. John Wiley & Sons Inc., New
[13]
[14]
[15]
[16]
York, 1988. Spectral theory. Selfadjoint operators in Hilbert space, With the assistance of William G. Bade and Robert G.
Bartle, Reprint of the 1963 original, A Wiley-Interscience Publication.
H. Epstein. Remarks on two theorems of E. Lieb. Comm. Math. Phys., 31:317–325, 1973.
A. Kor´
anyi. On a theorem of L¨
owner and its connections with resolvents of selfadjoint transformations. Acta Sci. Math.
Szeged, 17:63–70, 1956.
¨
Fritz Kraus. Uber
konvexe Matrixfunktionen. Math. Z., 41(1):18–42, 1936.
M. Krein. The theory of self-adjoint extensions of semi-bounded Hermitian transformations and its applications. I. Rec.
Math. [Mat. Sbornik] N.S., 20(62):431–495, 1947.
[17] Elliott H. Lieb. Convex trace functions and the Wigner-Yanase-Dyson conjecture. Advances in Math., 11:267–288, 1973.
[18] Elliott H. Lieb and Mary Beth Ruskai. Some operator inequalities of the Schwarz type. Advances in Math., 12:269–273,
1974.
¨
[19] Karl L¨
owner. Uber
monotone Matrixfunktionen. Math. Z., 38(1):177–216, 1934.
[20] W. Pusz and S. L. Woronowicz. Functional calculus for sesquilinear forms and the purification map. Rep. Mathematical
Phys., 8(2):159–170, 1975.
[21] A. Uhlmann. Relative entropy and the Wigner-Yanase-Dyson-Lieb concavity in an interpolation theory. Comm. Math.
Phys., 54(1):21–32, 1977.
25
Index
concave map, 20
convex map, 20
geometric mean, 5
harmonic mean, 6
normalized map, 18
operator-concave function, 14
operator-convex function, 14
operator-monotone function, 8
positive operator, 3
27
Symbols
A : B, 6
A#B, 5
G,H,K, 3
L(H), 3
L+ (H), 3
S(α, β), 8
S(α, β; H), 8
Φs,t (A), 20
29