Вы находитесь на странице: 1из 29

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-denite) 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 sucient that A and B are positive and there is a contraction W (i.e. W 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 W 1 implies W W 1, hence B 1/2 W W B 1/2 B. Then T admits a factorization T = S S where S= A1/2 0 W B 1/2 B B 1/2 W W B 1/2
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 A1/2 C y
2

= sup
x

|(Cx, y)| B 1/2 y (Ax, x)

where A1/2 is dened 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 A1/2 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. Corollary I.1.2. If are positive. Proof. As in the proof of Theorem I.1 the positivity of This means that A1/2 C

A C

C B

is positive, then there is the minimum of all X for which

A C

C X

A C

C X

implies A1/2 C

A1/2 C X.

A1/2 C is the minimum in the assertion.


3

I. GEOMETRIC AND HARMONIC MEANS

Remark. If A has bounded inverse, the minimum in Corollary I.1.1 is just CA1 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 T = S . Further, among all T satisfying these conditions there are the minimum T and the maximum TM . Proof. It can be assumed that S = 1. Let S = K=H G. By assumption A is self-adjoint and Ax
2

A C

where A acts on G and C does from G to

+ Cx

for x G. A C C X for which

Therefore C C 1 A2 . T must be of the form T = C C 1X C 21 C 21

1+A
C

C 1+X

and

1A

are positive.

The inequalities C C

1 A2 and 21 (1 A) 1 imply that

1 A2 = 21 (1 A)(1 + A) 1 + A. By Remark after Corollary I.1.2 this implies 1 + A C 1 + A C the positivity of . Therefore there is the minimum B of all X for which C 1+1 C 1+X are positive. In fact, B is dened by B = (1 + A)1/2 C (1 + A)1/2 C 1. In order to conclude A C 1 A C that T = is the minimum in the assertion, it remains to show that is positive. C B C 1 B 1 A C 1 + A C Since is positive just as , (1 A)1/2 C is a well dened linear operator. By C 21 C 21
Corollary I.1.2 it reduces to prove the inequality (1 A)1/2 C y But (1 + A)1/2 C y
2 2

2 y

(1 + A)1/2 C y

+ (1 A)1/2 C y

= =2

1 A2

C y, C y = 2

1 A2

1/2

C y

Since C C 1 A2 implies

1 A2 C 0, C 1
2

1 A2

1/2

C y

2 y

A C with BM = 1 (1 A)1/2 C C BM be the maximum. This completes the proof. as expected. Analogously TM =

(1 A)1/2 C is shown to

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

Proof. Consider rst the case that A has bounded inverse. Let D = A1/2 BA1/2 . Then is positive if and only if of all Y for which Suppose that

A X

X B

1
A1/2 XA1/2 D

A1/2 XA1/2 D

is positive. We claim that D1/2 is the maximum

1 Y
Y

are positive. The positivity of

1
D1/2

D1/2 D

is obvious by Corollary I.1.2.

1 Y 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 x, y := (D1/2 x, y). Since Y is self-adjoint W x, y = x, W y , for every x, y H which means that W is self-adjoint with respect to this new scalar product. Since W 1, for any real number with || > 1 the operator W has bounded inverse. The operator ( W )1 is bounded with respect to the new scalar product, hence
| W x, x | x, x which implies D1/2 Y D1/2 . for all x H,

Thus D1/2 is the maximum as claimed. As a consequence, A X 1/2 1/2 A1/2 D1/2 A1/2 = A1/2 A1/2 BA1/2 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 C 0 whenever > > 0, the operator lim C 0+ X B is the maximum of all X for which A X X B are positive. This completes the proof.

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. (A)#(B) = (A#B) for 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

(vi) A#B = A1/2 A1/2 BA1/2 A1/2 if A has bounded inverse. 1 1 1 (vii) A #B = (A#B) if both A and B have bounded inverses. Proof. (i) to (iv) are immediate from denition. (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 = 1 (p1 + p2 ). Then since Ap = Ap1 #Ap2 and 2 B p = B p1 #B p2 by Corollary I.2.1, Ap B p follows also from the same Corollary.

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 CA1 C Bi i = 1, 2. Therefore by Corollary I.2.1 i C(A1 #A2 )1 C = C A1 #A1 C < CA1 C # CA1 C B1 #B2 , 1 2 1 2 which implies the positivity of A1 #A2 C C B1 #B2 by Corollary I.1.2.
1 2 (A

Corollary I.2.4. Arithmetic mean is greater than geometric mean, that is positive A and B.

+ B) A#B for

Proof. We may assume the bounded invertibility of A. Then by Corollary I.2.1 A#B = A1/2 A1/2 BA1/2 because X 1 2
1/2

A1/2 A1/2

1 1 1 1 + A1/2 BA1/2 A1/2 = (A + B), 2 2 2

1 + X 2 for any positive X.


1

1 1 A + B 1 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 dened as the limit of (A + ) : (B + ) as 0+ . The harmonic mean of two positive operators A and B must be dened by 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 (Xx, x) 2 (Bx, x) sup


y

|(Bx, y)| ((A + B)y, y)

Since by denition A : B = 2A(A + B)1 B = 2 B B(A + B)1 B , we have ((A : B)x, x) = 2 (Bx, x) sup
y

|(Bx, y)| ((A + B)y, y) X X

. X . X

This shows that A : B is the maximum of all X for which

2A 0 0 2B

Corollary I.3.1. Harmonic means have the following properties. (i) (ii) (iii) (iv) A : B = B : A. (A) : (B) = (A : B) for 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

(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) A1 : B 1 = 1 (A + B) if both A and B have bounded inverses. 2 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 A1 #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 CA1 C i
1 2 (B1

C . + 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 C 1 (A1 + A2 ) 2
1

C = C A1 : A1 C CA1 C : CA1 C B1 : B2 , 1 2 1 2
1 2 (A1

which implies the positivity of A1 : A2 C C + B2 )

+ A2 ) C C B 1 : B2

by Corollary I.1.2.

The positivity of

1 2 (B1

is proved analogously. The following identity holds for positive A and B; 1 (A + B) #(A : B) = A#B. 2

Corollary I.4.1.

A A#B A#B B is positive by Theorem I.4, which implies Proof. By denition

and

B A#B

A#B A

are positive, so that

1 2 (A

+ B) A#B A#B A:B

1 (A + B) #(A : B) A#B. 2 When A and B have bounded inverse, the above inequality, with A1 and B 1 instead of A and B respectively, leads, by taking inverse, to the inequality (A : B)# These two inequalities yield the assertion. 1 (A + B) 2 A#B.

CHAPTER II

Operator-Monotone Functions
Given a nite or innite open interval (, ) on the real line, let us denote by S(, ; H) the totality of all self-adjoint operators on H whose spectrum are included in the interval (, ). If there is no confusion, we shall write simply S(, ). A real-valued continuous function f on (, ) is said to be operator-monotone on (, ) if A, B S(, ; H) and A B implies f (A) f (B), where f (A) and f (B) are dened 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 () := a + b with b 0 is operator-monotone on (, ). 2. f () := p with 0 p 1 is operator-monotone on (0, ) by Corollary I.2.2. 1 1 3. For (, ) the function f () := is operator-monotone on (, ). In particular f () = is / operator-monotone on (0, ). In fact, if < , then A, B S(, ) A B implies 0 < ( A) ( B) so that ( A)1 ( B)1 hence f (A) f (B). If > , then 0 < B < A hence ( B)1 ( A)1 .

It is easy to see that a function f is operator-monotone on a nite interval (, ) if and only if g() := f ) + + ) is operator monotone on (1, 1). On this basis, we shall treat only the interval (1, 1).
1 2 ((

Lemma II.1. If a continuously dierentiable function f on (1, 1) is operator-monotone, then for any N choice i (1, 1) i = 1, . . . , N the matrix f [1] (i , j ) i,j=1 is positive, where f [1] (, ) is dened by f [1] (, ) =
f ()f ()

for = and f [1] (, ) = f ().

1 0 .. , considered as a self-adjoint operator on the N Proof. Observe the matrix A := . 0 N dimensional Hilbert space CN . Take any complex numbers 1 , . . . , N and consider the positive matrix N B := i j i,j=1 . Since A S 1, 1; CN , for suciently small > 0, A + B belongs to S 1, 1; CN . We claim that for any continuously dierentiable function g on (1, 1)

N 0

lim +

{g(A + B) g(A)} =
i,j=1 8

g [1] (i , j )Pi BPj

II. OPERATOR-MONOTONE FUNCTIONS

where Pi is the orthoprojection to the subspace spanned by the vector ei := (ij )N . In fact, this is true if j=1 g is polynomial, because for g() = n
n 1

{g(A + B) g(A)} =
k=1 N n

Ak1 BAnk + O( )

=
i,j=1 k=1 N

k1 nk Pi BPj + O( ) i j g [1] (i , j )Pi BPj + O( ).


i,j=1

Use approximation of g by polynomials on a suitable subinterval when g is merely continuously dierentiable. Now since f is operator-monotone and A + B A for all > 0, we have
N

f [1] (i , j )Pi BPj = lim +


i,j=1 0

{f (A + B) f (A)} 0.

Let x =

N i,j=1 ei ,

then the positivity of


N

N i,j=1

f [1] (i , j )Pi BPj implies


N

0
i,j=1

f [1] (i , j )(Pi BPj x, x) =


i,j=1

f [1] (i , j )i j .
N i,j=1

Since (i ) are arbitrary, this means that the matrix f [1] (i , j )

is positive.

Lemma II.2. If a continuously dierentiable function f on (1, 1) is operator-monotone, then there exists a nite positive measure m on the closed interval [1, 1] such that
1

f () = f (0) +
1

dm(t) 1 t

for

(1, 1).

Proof. Consider the linear subspace H of C(1, 1), spanned by the functions f (t) := f [1] (, t) where runs over (1, 1). Introduce a scalar product in H by i fi ,
i j

j fj

=
i,j

f [1] (i , j )i j .

This is positive semi-denite by Lemma II.1. Let H be the associated Hilbert space. Let us show that n implies fn f 0. In fact, fn f
2

= f [1] (n , n ) 2f [1] (n , ) + f [1] (, ) = f (n ) 2 f (n ) f () + f (). n

Now by continuous dierentiability of f f (n ) 2 f (n ) f () + f () f () 2f () + f () = 0. n

Let D be the linear subspace of H, spanned by all f ( = 0). Then D is dense in H, as shown above. Dene a linear map T from D to H by T f := 1 {f f0 } T is symmetric hence well-dened, because T f , f = f () f () f (0) = f , T f . ( ) ( = 0).

10

II. OPERATOR-MONOTONE FUNCTIONS

We claim that T admits a (not necessarily bounded) self-adjoint extension T . This follows from a theorem of von Neumann, (see [12] p.1231), because the anti-linear involution J, dened by J
i

i fi

=
i

i fi ,

makes D invariant and commutes with T . Since by denition ( )f = (1 T )(f f ), the kernel if 1 T is orthogonal to the linear subspace spanned by f ( = , and = 0). This subspace is dense in H as shown above, so that 1 T is injective, hence (1 T )1 is a (unbounded) self-adjoint )f implies f = (1 T )1 f0 . operator. Therefore f0 = (1 T Now let T = tdE(t) be the spectral representation of the self-adjoint operator T , and dm(t) = dE(t)f0 , f0 . Then for = 0 1 {f () f (0)} = f [1] (, 0) = f , f0 = (1 T )1 f0 , f0

1 dm(t). 1 t

To complete the proof, it remains to prove that the measure m is concentrated on the closed interval [1, 1]. Take > 1, and let us show that [, ) is an m-zero set
1 0

1 dm(t)d (1 t)2 =

1 0 1

1 dm(t)d (1 t)2
1

f , f d =
0 0

f ()d = f 1 f (0) < .

By Fubinis theorem we have


1 0

1 dm(t)d = (1 t)2

1 ddm(t) (1 t)2 s2 dsdm(t).

1 t

1 0

But this last expression is nite inly if [, ) is an m-zero set. Analogously (, ] is an m-zero set. Remark. In Lemma II.2
1 1

dm(t) = f0 , f0 = f (0).

Lemma II.3. If a continuously dierentiable function f on (1, 1) is operator-monotone, then for any choice 1 < 1 < 1 < 2 < 2 < 1 det f [1] (1 , 1 ) f [1] (1 , 2 ) f [1] (2 , 1 ) f [1] (2 , 2 ) 0.

Proof. We shall use the notations in the proof of the previous Lemma. First remark that
1

f [1] (, ) = f , f =
1

1 dm(t). (1 t)(1 t)

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 1 t0 )1 (1 1 t0 )1 (1 1 t0 )1 (1 2 t0 )1 det , (1 2 t0 )1 (1 1 t0 )1 (1 2 t0 )1 (1 2 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 1 = 1 and 2 = 2 has non-negative by Lemma II.1, by using continuity argument there are 1 and 2 such that 1 1 1 and 2 2 2 , and det f [1] (1 , 1 ) f [1] (1 , 2 ) f [1] (2 , 1 ) f [1] (2 , 2 ) = 0.

This implies that there are 1 and 2 such that |1 | + |2 | = 0 and fi , 1 f1 + 2 f2 = 0 (i = 1, 2).

Choose 1 and 2 so that 1 + 2 = 1 + 2 and 1 2 + 2 1 = 1 2 + 2 1 . Then it follows that 0 = 1 f1 + 2 f2 , 1 f1 + 2 f2


1

=
1

{1 + 2 (1 2 + 2 1 )t}2 dm(t). (1 1 t)(1 2 t)(1 1 t)(1 2 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 1 + 2 = 0 and 1 2 + 2 1 = 0, which contradicts |1 | + |2 | = 0. This contradiction completes the proof. Now let us return to general operator-monotone functions. Take an innitely many times dierentiable 1 function on (, ) such that is non-negative, vanishes outside of (1, 1) and 1 (t)dt = 1. Suppose that f is operator-monotone on (1, 1). For any 0 < < 1, dene a function f( ) on (1+ , 1 ) by 1 f( ) (t) := (s/ )f (t s)ds.

The function f( ) is operator-monotone on (1 + , 1 ), because each f (t ) is so. Obviously f( ) is innitely many times dierentiable, 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 (, ) of (1, 1) f satises Lipschitz condition, that is, sup<s<t< f [1] (t, s) < . Proof. Let us use the notations in the preceding comment. Let 1 = 1 ( 1) and 2 = 1 ( + 1), and 2 2 apply Lemma II.3 to the operator-monotone function f( ) with 2 = t and 1 = s. By taking limit, we can conclude that f [1] (1 , s) f [1] (1 , 2 ) det 0, f [1] (t, s) f [1] (t, 2 ) or equivalently f [1] (1 , 2 )f [1] (t, s) f [1] (1 , s)f [1] (t, 2 ). The right side of the above inequality is bounded when s < t run over [, ]. Since f is non-decreasing, both f [1] (1 , 2 ) and f [1] (t, s) are non-negative. Therefore if f [1] (1 , 2 ) = 0, then f [1] (t, s) are bounded when s < t run over [, ]. Finally f [1] (1 , 2 ) = 0 implies f [1] (t, s) = 0 for all s < t in [, ]. This completes the proof. Theorem II.1. In order that a continuous function f on (1, 1) is operator-monotone, it is necessary and sucient that there is a nite positive measure m on [1, 1] such that
1

f () = f (0) +
1

dm(t) 1 t

for (1, 1).

12

II. OPERATOR-MONOTONE FUNCTIONS

Proof. Suppose that f admits a representation of the above form. For each t [1, 1] the function ht () := 1t is operator-monotone on (1, 1). In fact, if t = 0, h0 () = is operator-monotone. If t = 0, ht () = t1 + weighted Suppose conversely that f is operator-monotone. Then with the notations in front of Lemma II.4, for each 0 < < 1 f( ) ((1 )) is operator-monotone on (1, 1). Then since f( ) ((1 )) is continuously dierentiable, by Lemma II.2 there is a measure m on [1, 1] such that
1 t2 t1 is operator-monotone as stated in one of the 1 average 1 ht ()dm(t) is operator-monotone, and so is f .

Examples II.1. As a consequence, the

f( ) ((1 )) = f( ) (0) +
1

dm (t). 1 t

We claim that

1 1

dm (t) are bounded when


1 1

runs over (0, 1/2). As remarked after Lemma II.2, we have dm (t) = f( ) (0).

Since Lemma II.4 f satises Lipshitz condition on any closed subinterval of (1, 1), it has derivative f (t) for almost all and ess. sup |f (t)| = r < . Further it is easy to see that
<t<

f( ) (0) =

(1 )

f (s)
1

ds (1 )r .

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
1 n 1

lim

g(t)dm n (t) =
1 1

g(t)dm(t)

for all continuous functions g on [1, 1]. Therefore for (1, 1)


1 n

lim

dm n (t) = 1 t

1 1

dm(t). 1 t

Finally the assertion follows from


n

lim f(

n)

((1

n ))

= f ().

The integral representation in Theorem II.1 shows that an operator-monotone function f on (1, 1) is necessarily innitely many times dierentiable. Further more it admits analytic continuation to the upper and lower open half planes by
1

f () = f (0) +
1

dm(t) 1 t

for

with Im () = 0.

The function f maps the upper half plane to itself. Indeed


1

Im

f () =
1

Im () |1 t|

2 dm(t).

This observation is completed in the following theorem. Theorem II.2. Let f be a real-valued continuous function on a nite or innite interval (, ). In order that f is operator-monotone it is necessary and sucient that it admits an analytic continuation f to the upper and lower half planes such that Im f () > 0 for Im () > 0.

II. OPERATOR-MONOTONE FUNCTIONS

13

Proof. Suppose that f admits an analytic continuation f 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 nite positive measure m on := (, ) \ (, ) such that f () = a + b +

1 + t dm(t). t

The function g() := a + b with b 0 is obviously operator-monotone on (, ). For each t (, ) the / function ht () := 1+t is operator monotone, because t ht () = t + 1 + t2 t

1 and the function t is operator-monotone on (, ). Therefore the weighted average f () is also operatormonotone on (, ). This completes the proof.

Corollary II.2.1. If f is operator-monotone on (, ), then f () = a + b 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 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 nite or innite interval (, ) is said to be operator-convex if f 1 (A + B) 2 1 {f (A) + f (B)} 2 forA, B S(, ).

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 () := a + b is operator-convex on (, ). 2. f () := 2 is operator-convex on (, ). In fact, for self-adjoint A and B 1 A2 + B 2 2 1 (A + B) 2
2

1 (A B)2 0. 4
1

1 3. For < the function f () := is operator-convex on (, ). In particular, f () := convex on (0, ). In fact, for A, B S(, ), A and B belong to S(0, ) and

is operator-

1 (A )1 + (B )1 = {(A ) : (B )}1 . 2 By Corollary I.2.4 and I.3.2. {(A ) : (B )}1 = 1 1 (A ) + (B ) 2 2 1 (A + B) 2


1 1

Lemma III.1. Let f be a twice continuously dierentiable function on (1, 1). If f is operator-convex, then for each (1, 1) the function g() := f [1] (, ) is operator-monotone. Conversely if f [1] (0, ) is operator-monotone, then f is operator-convex. Proof. Let f be operator-convex. Obviously g is continuously dierentiable on (1, 1). Inspection of Lemmas II.1 and II.2 will show that for the operator-monotoneousness of g it suces to prove that for any choice i (1, 1) i = 1, 2, . . . , N the matrix g [1] (i , j ) is positive. Observe the matrix i,j=1 1 0 A := N with N +1 = , considered as a self-adjoint operator on the (N + 1)-dimensional 0 N +1 1 . . 0 . . Hilbert space CN +1 . Take any complex numbers 1 , . . . , N and consider the matrix B := N 1 , . . . , N 0 Since A S 1, 1; CN +1 , for suciently small > 0 A + B belongs to S 1, 1; CN +1 . Since f is twice
14 N

III. OPERATOR-CONVEX FUNCTIONS

15

continuously dierentiable, just as in the proof of Lemma II.1, it can be shown that the matrix-valued function f (A + B) is twice dierentiable and d2 f (A + B) d 2
N +1

=
=0 i,j,k=1

f [2] (i , j , k )Pi BPj BPk


N +1

where Pi is the orthoprojection to the subspace spanned by the vector ei := (ij )j=1 and f [2] (s, t, u) = d2 f (A + B) [1] hs (t, u) with hs (t) = f [1] (s, t). Now the operator-convexity of f implies 0, as in the d 2 =0 N case of scalar convex functions. Let x = i=1 ei . Then
N +1

0
i,j,k=1 N

f [2] (i , j , k )(Pi BPj BPk x, x) =


N

=
i,k=1

f [2] (i , N +1 , k )k i =
i,k=1

g [1] (i , k )k i .
N i,j=1

Since (i ) are arbitrary, this means that the matrix g [1] (i , j )


[1]

is positive.

Suppose conversely that f (0, ) is operator-monotone. Then by Theorem II.1 there exists a nite positive measure m on [1, 1] such that
1

f [1] (0, ) = f (0) +


1

dm(t), 1 t

hence f () = a + b +

1 1

2 dm(t), 1 t
2

where a = f (0) and b = f (0). For each t [1, 1] the function ht () := 1t is operator-convex. In fact, if 2 t = 0, ht () = , and if t = 0, t2 ht () = t2 t1 + . 1 t These functions are operator-convex, as shown in Example III.1. Since the function a+b 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 f1 () = 0 f (t)dt and f2 () := f () are operator-convex. Proof. Since f is continuously dierentiable by Theorem II.2, both f1 and f2 are twice continuously dierentiable. Now since f1 (0, ) =
0 [1] 1

f (s)ds and f2 (0, ) = f (),

[1]

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 sucient that there are real numbers a and b, and a nite positive measure m on [1, 1] such that
1

f () = a + b +
1

2 dm(t) 1 t

for

(1, 1).

Proof. Suciency 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 )) is operator-convex for each 0 < < 1. Theref( ) ((1 )) f( ) ((1 )) fore by Lemma III.1 for any (1, 1) the function is operator-monotone on (1, 1) so that by taking limit as 0 f [1] (, ) is operator-monotone on ( + , 1) as well as (1, )

16

III. OPERATOR-CONVEX FUNCTIONS

for any 0 < . By Theorem II.2 this implies that f is twice continuously dierentiable 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
0

f () = a + b +

1 + t dm(t) t

where b 0 and m is a positive measure. It suces to prove that for each t < 0 the function ht () := 1+t t is operator-concave. If t = 0, h0 () = 1/ is operator-concave on (0, ), as mentioned in Example III.1. If t < 0, 1 + t2 ht () = t t is also operator-concave on (0, ), as shown in Example III.1. Since the function a + b 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 () > 0 on (0, ), then g() := f ()1 is operator-convex. Proof. Take A, B S(0, ). Since f is operator-concave and the function 1/ is operator-convex on (0, ), 1 1 f (A + B) {f (A) + f (B)} 2 2 and g 1 (A + 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 + b. Theorem III.4. A continuous function f on (0, ) with f (0) := lim and only if f ()/ is operator-monotone.
0+

1 {f (A) + f (B)} 2 1 {g(A) + g(B)}. 2

f ( ) = 0 is operator-convex if

Proof. Suppose that f is operator-convex. By Theorem III.2 it is innitely many times dierentiable. Then by Lemma III.1 for each > 0 the function f ()f ( ) is operator-monotone hence, as the limit, the function f ()/ is operator-monotone. Suppose conversely that f ()/ is operator-monotone. Then as in the proof of Theorem III.3 there are a and b > 0 and a measure m such that
0

f ()/ = a + b +

1 + t dm(t), t

hence
0

f () = a + b2 +

(1 + t) dm(t). t

III. OPERATOR-CONVEX FUNCTIONS

17 (1+t) t

Since b2 is operator-convex, it suces to prove that for each t < 0 the function ht () := operator-convex on (0, ). For t = 0 this is obvious. If t = 0, ht () = {(1 + t2 ) t} + is operator-convex, as was shown in Example III.1. (1 + t2 ) |t| t

is

Let us investigate for what exponent < s < the function f () := s on (0, ) is operatormonotone, operator-convex or operator-concave. Examples III.2. 1. f () = s 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 () = s is operator-convex if and only if 1 s 2 or 1 s 0. In fact, by Theorem III.4 for s > 0 f () = s is operator-convex if and only if s1 is operator-monotone. Therefore for s > 0 f is operator-convex if and only if 1 s 2. By Corollary III.3.1 f () = s is operator-convex for s 1 s 0. For s < 1 the function 1 does not admit any analytic continuation f to the upper half 1 s plane such that Im (f ()) > 0 for Im () > 0, hence f () = is not operator-convex by Lemma III.1. Lemma III.2. Let f () > 0 on (0, ) and g() := f (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 f to the complement of the closed negative real semi-axis such that Im f () > 0 or < 0 according as 1 1 is an analytic continuation of g to the complement of Im () > 0 or Im () < 0. Then g () := f the closed negative real semi-axis such that Im (g()) > 0 or < 0 according as Im () > 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() := f ()/ is operator-monotone. By applying the rst part of this lemma to h, we can conclude that the 1 function h 1 = g()/ 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() := f 1 monotone, hence operator-concave by Theorem III.2. Therefore g hence f (A : B) f (A) : f (B). The other assertion can be proved quite analogously by using Lemma III.2. 1 1 A + B 1 2 1 g A1 + g A1 2
1

is operator-

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 from L(H) to L(K) preserves orderrelation, that is, A B implies (A) (B), and preserves adjoint operation, that is (A ) = (A) . It is said to be normalized if it transforms 1H to 1K . If is normalized, it maps S(, ; H) to S(, ; K). Lemma IV.1. A normalized positive linear map has the following properties. (i) A2 (A)2 for A S(, ; H). (ii) A1 (A)1 for A S(0, ; H). Proof. (i) By Remark after Corollary I.1.1 it suces to prove the positivity of

A2 (A)

(A) . (1)

Consider the spectral representation A =

tdE(t). Since and

A2 =

t2 dE(t)

1=
t2 t

dE(t),

we have, with tensor product notation, A2 (A) Since 2 2 matrices is positive. (ii) It suces to prove the positivity of (A) (1) . Since A is positive by assumption, it is written (1) A1 in the form A = 0+ tdE(t). Now the positivity in question follows from the positivity of matrices t 1 1 t
1

(A) (1)

t 1

dE(t).

t2 t

t 1

are positive, for all < t < the right hand of the above expression

for 0 < t < , as in the proof of (i).

Theorem IV.1. Let be a normalized positive linear map. If f is an operator-convex function on (, ), then f [(A)] [f (A)] for A S(, ; H). Proof. It suces to consider the case (, ) = (1, 1). By Theorem III.2 f admits a representation
1

f () = a + b +
1

2 dm(t) 1 t

with b 0 and a positive measure m. Since for A S(1, 1; H)


1

[f (A)] = a + b(A) +
1 18

A2 (1 tA)1 dm(t)

IV. POSITIVE MAPS

19

and f [(A)] = a + b(A) +

1 1

(A)2 {1 t(A)}1 dm(t), for 1 t 1.

it suces to show A2 (1 tA)1 (A)2 {1 t(A)}1 For t = 0, this follows from Lemma IV.1. For t = 0, again by Lemma IV.1 A2 (1 tA)1 = t2 t1 (A) + t2 (1 tA)1 t2 t1 (A) + t2 {1 t(A)}1 = (A)2 {1 t(A)}1 . This completes the proof. Corollary IV.1.1. Let be a normalized positive linear map. Then for a positive operator A (Ap ) (A)p (1 p 2) and (Ap ) (A)p (0 p 1).

Proof. As shown in Examples III.1, p is operator-convex on (0, ) for 1 p 2 while p is operator-convex for 0 p 1. Corollary IV.1.2. If is a normalized positive map and if A is a positive operator, then (Ap ) 1/q (Aq ) whenever 1 p q or 1 q p 1 q. 2
p/q 1/p

Proof. By Corollary IV.1.1 (Aq ) (Ap ) whenever p q. If p 1in addition, this implies 1/p q 1/q p 1/p (A ) (A ) by Corollary I.2.2. If q 1 p 1 q, (Ap ) (Aq )1/q , because 1/q is 2 operator-monotone. Corollary IV.1.3. If is a positive linear map, for any positive operators A and B (A : B) (A) : (B) and (A#B) (A)#(B). Proof. We may assume A S(0, ; H). Let G be the closure of the range of (A). Then there is uniquely a normalized positive linear map from L(H) to L(G) such that (A)1/2 (C)(A)1/2 = A1/2 CA1/2 Since the functions f () := for any positive C and (1#C) = (g(C)) g((C)) = 1#(C). With C = A1/2 BA1/2 this leads to the following (A : B) = (A)1/2 (1 : C)(A)1/2 = (A)1/2 {1 : (C)}(A)1/2 = (A) : (B) and analogously (A#B) (A)#(B).
2 1+

for C L(H).

and g() := 1/2 are operator-concave on (0, ), by Theorem IV.1 we have (1 : C) = (f (C)) f ((C)) = 1 : (C)

20

IV. POSITIVE MAPS

The converse of Theorem IV.1 is also true. Theorem IV.2. Let 0 and f a continuous function on (, ) with f (0) = 0. If (f (A)) f ((A)) for every normalized positive linear map and A S(, ) then f is operator-convex. Proof. Let K be the subspace of H H, consisting of all vectors x x. Dene the linear map 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 is positive and normalized. Take A, B S(, ; H). Then by denition we have f while f A 0 0 B 1 =f 2 1 = 2 = Now by assumption f which implies f This completes the proof. A (non-linear) map from a convex subset of L(H) to L(K) is said to be a convex map (resp. a concave map) if 1 (A + B) 1 {(A) + (B)} (resp. 1 (A + B) 1 {(A) + (B)}). 2 2 2 2 We shall be concerned with convexity or concavity of maps s,t (A) = As At dened on S(0, ; H) where < s, t < . Lemma IV.2. If and are concave maps with range in S(0, ; K) then the maps (A) := (A)#(A) and (A) := (A) : (A) are concave. Proof. By Corollary I.2.1 and concavity of and 1 (A + B) 2 = 1 1 (A + B) # (A + B) 2 2 1 1 1 1 (A) + (B) # (A) + (B) 2 2 2 2 1 (A + B) 2 1 {f (A) + f (B)}. 2 A 0 0 B f A 0 0 B 1 2 f f 1 1 1 1 1 1
1 2 (A 1 2 (A

A 0 0 B

f (A) 0 0 f (B)

1 4

f (A) + f (B) f (A) + f (B) , f (A) + f (B) f (A) + f (B)

1 1

0 0 f (0) 0 f f

1 2 (A

0 + B) 0

1 2

1 1 1 1 1 2 1 1 1 1

1 2 (A

+ B) .

+ B) + B)

1 2 (A 1 2 (A

+ B) + B)

1 {(A)#(A) + (B)#(B) 2 1 = {(A) + (B)}, 2 which proves the concavity of . The concavity of is proved analogously by using Corollary I.3.1. Theorem IV.3. The map p,q (A) := AP Aq is concave if 0 p, q and p + q 1.

IV. POSITIVE MAPS

21

Proof. Consider the set of (p, q) in R2 for which p,q are concave. We claim that is a convex set. + In fact, let (pi , qi ) and p = 1 (p1 + p2 ), q = 1 (q1 + q2 ). Then since 2 2 p,q (A) = Ap Aq = (Ap1 #Ap2 ) (Aq1 #Aq2 ) = (Ap1 Aq1 ) # (Ap2 Aq2 ) = p1 ,q1 (A)#p2 ,q2 (A), by Lemma IV.2 the map p,q is concave. Obviously (0, 0), (1, 0) and (0, 1) belong to , 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)
p q (A1 : B1 )p (A2 : B2 )q (Ap Aq ) : (B1 B2 ) 1 2

whenever 0 p, q and p + q 1. Proof. The concavity of p,q is seen to be equivalent to the inequality (A : B)p (A : B)q (Ap Aq ) : (B p B q ) . With A = A1 0 0 A2 and B = B1 0 0 B2 this inequality implies the inequality in the assertion.

Theorem IV.4. If f and g are positive, operator-monotone functions on (0, ), then the map (A) := f (A)1 g(A)1 is convex. Proof. Let h() := f 1 to that for A, B S(0, )
1

and k() := g 1

. Then the convexity of is seen to be equivalent

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 p,q (A) = Ap Aq is convex if 0 p, q 1. This follows from Theorem IV.3 and Corollary I.2.2. Theorem IV.5. The map p,q (A) = Ap Aq is convex if 0 p q 1 < 1. Proof. Let us consider rst the case q = p + 1 < 2. The well-known integral representation of p for 0 < p < 1 (see [11], [12] p. )

p = 1 sin(p)
0

tp1 ( + t)1 dt

22

IV. POSITIVE MAPS

is applied to get p,p+1 (A) = A1 A


p

(1 A)

= 1 sin(p)
0

t1p A1 A2

A1 A + t

dt.

Therefore it suces to prove that for each t > 0 the map t (A) := A1 A2 is convex. To this end, remark, rst of all t (A) = 1 A t (A 1)1 + A1 A + t
1

1 t1 A
1

the rst term of which is a convex map. It remains to show that the map t (A) := (A 1)1 +

1 t1 A

is convex. But this follows from Lemma IV.2, because 2t (A) = (A) : (A) where (A) := A 1 and (A) := 1 t1 A are obviously concave. Next let us consider the case p < q 1. Let r = p(q 1). Since 0 < r < 1, the function r is operator-concave by Theorem III.4, hence for positive A and B 1 1 A+ B 2 2 which implies 1 1 A+ B 2 2
p r r

1 r 1 r A + B , 2 2 1 r 1 r A + B 2 2
1

Further the operator-monotoneousness of q1 yields 1 1 A+ B 2 2 1 r 1 r A + B 2 2


(q1)

It follows from the rst part of the proof, as Corollary IV.3.1 does from Theorem IV.3, that for positive A, B, C and D 1 1 C+ D 2 2
(q1)

1 1 A+ B 2 2

1 C (q1) Aq + D(q1) B q . 2

Now use these inequalities with C = Ar and D = B r to get 1 1 A+ B 2 2


p

1 1 A+ B 2 2

1 r 1 r A + B 2 2

(q1)

1 1 A+ B 2 2

1 Ap Aq + B p B q . 2

This shows the convexity of p,q . It remains to consider the case p = 1 and q = 2. The convexity of 1,2 means that (A + B)1 (A + B)2 A1 A2 + B 1 B 2 . With C = A1/2 BA1/2 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

23

Considering the spectral representation A = A is an orthoprojection, i.e. A2 = A. Then

dE(), it suces 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 As , A At and R+ 0 s+t are concave, hence as mentioned in Example III.2 we have 0 s, t and s + t 1. If s,t is convex with s t, just as above, the maps A As and A 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 {, } s t is convex on the positive cone of R2 . Therefore, for arbitrarily xed , > 0, the function () := ( + )s ( )t is a convex function of in a neighborhood of 0. By dierentiation this implies If s,t is concave, using s(s 1)s2 t 2sts1 t1 + t(t 1)s t2 0. If 1 s 0 1 t 2 or 1 s t 2, by arbitrariness of and 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 s,t is concave while Theorem IV.4 and IV.5 exhaust all the case s,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 & Dun [2] dened A1 + 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 Lwner [19] and Kraus [15]. o 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 (A : B) (A) : (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 dierent 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. Dun. Series and parallel addition of matrices. J. Math. Anal. Appl., 26:576594, 1969. W. N. Anderson, Jr. and G. E. Trapp. Shorted operators. II. SIAM J. Appl. Math., 28:6071, 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):5367, 1976. [5] T. Ando. Concavity of certain maps on positive denite matrices and applications to Hadamard products. Linear Algebra Appl., 26:203241, 1979. [6] Julius Bendat and Seymour Sherman. Monotone and convex operator functions. Trans. Amer. Math. Soc., 79:5871, 1955. [7] Man Duen Choi. A Schwarz inequality for positive linear maps on C -algebras. Illinois J. Math., 18:565574, 1974. [8] Chandler Davis. A Schwarz inequality for convex operator functions. Proc. Amer. Math. Soc., 8:4244, 1957. [9] Chandler Davis. Notions generalizing convexity for functions dened on spaces of matrices. In Proc. Sympos. Pure Math., Vol. VII, pages 187201. 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 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., 31:317325, 1973. A. Kornyi. On a theorem of Lwner and its connections with resolvents of selfadjoint transformations. Acta Sci. Math. a o Szeged, 17:6370, 1956. Fritz Kraus. Uber konvexe Matrixfunktionen. Math. Z., 41(1):1842, 1936. M. Krein. The theory of self-adjoint extensions of semi-bounded Hermitian transformations and its applications. I. Rec.

[13] [14] [15] [16]

Math. [Mat. Sbornik] N.S., 20(62):431495, 1947. [17] Elliott H. Lieb. Convex trace functions and the Wigner-Yanase-Dyson conjecture. Advances in Math., 11:267288, 1973. [18] Elliott H. Lieb and Mary Beth Ruskai. Some operator inequalities of the Schwarz type. Advances in Math., 12:269273, 1974. [19] Karl Lwner. Uber monotone Matrixfunktionen. Math. Z., 38(1):177216, 1934. o [20] W. Pusz and S. L. Woronowicz. Functional calculus for sesquilinear forms and the purication map. Rep. Mathematical Phys., 8(2):159170, 1975. [21] A. Uhlmann. Relative entropy and the Wigner-Yanase-Dyson-Lieb concavity in an interpolation theory. Comm. Math. Phys., 54(1):2132, 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(, ), 8 S(, ; H), 8 s,t (A), 20

29

Вам также может понравиться