Chapter Two



Positive Linear Maps

In this chapter we study linear maps on spaces of matrices. We use the symbol Φ for a linear map from images/nec-45-1.png into images/nec-45-2.png. When k = 1 such a map is called a linear functional, and we use the lower-case symbol φ for it. The norm of Φ is

images/eq-45-1.png

In general, it is not easy to calculate this. One of the principal results of this chapter is that if Φ carries positive elements of images/nec-45-3.png to positive elements of images/nec-45-4.png, then ||Φ|| = ||Φ(I)||.


2.1  REPRESENTATIONS

The interplay between algebraic properties of linear maps Φ and their metric properties is best illustrated by considering representations of images/nec-45-5.png in images/nec-45-6.png. These are linear maps that

(i)   preserve products; i.e., Φ(AB) = Φ(A)Φ(B);

(ii)   preserve adjoints; i.e., Φ(A) = Φ(A);

(iii)  preserve the identity; i.e., Φ(I) = I.

Let σ(A) denote the spectrum of A, and spr(A) its spectral radius.


2.1.1  Exercise

If Φ has properties (i) and (iii), then

images/eq-45-2.png(2.1)

Hence

images/eq-46-1.png(2.2)

Our norm || · || has two special properties related to the ∗ operation: ||A||2 = ||AA||; and ||A|| = spr(A) if A is Hermitian. So, if Φ is a representation we have

images/eq-46-2.png

Thus ||Φ(A)|| ≤ ||A|| for all A. Since Φ(I) = I, we have ||Φ|| = 1.

We have shown that every representation has norm one.

How does one get representations? For each unitary element of Mn, Φ(A) = UAU is a representation. Direct sums of such maps are representations; i.e., if U1, . . . , Ur are n × n unitary matrices, then images/nec-46-1.png is a representation.

Choosing Uj = I, 1 ≤ jr, we get the representation Φ(A) = images/nec-46-2.png. The operator images/nec-46-3.png is unitarily equivalent to images/nec-46-4.png, and images/nec-46-5.png is another representation.


2.1.2  Exercise

All representations of images/nec-46-6.png are obtained by composing unitary conjugations and tensor products with Ir, r = 1, 2, . . .. Thus we have exhausted the family of representations by the examples we saw above. [Hint: A representation carries orthogonal projections to orthogonal projections, unitaries to unitaries, and preserves unitary conjugation.]

Thus the fact that ||Φ|| = 1 for every representation Φ is not too impressive; we do know images/nec-46-7.png and ||UAU|| = ||A||.

We will see how we can replace the multiplicativity condition (i) by less restrictive conditions and get a richer theory.


2.2  POSITIVE MAPS

A linear map images/nec-46-8.png is called positive if Φ(A) ≥ O whenever AO. It is said to be unital if Φ(I) = I. We will say Φ is strictly positive if Φ(A) > O whenever A > O. It is easy to see that a positive linear map Φ is strictly positive if and only if Φ(I) > O.


2.2.1  Examples

(i)   φ(A) = trA is a positive linear functional; images/nec-47-1.png is positive and unital.

(ii)   Every linear functional on Mn has the form φ(A) = trAX for some images/nec-47-2.png. It is easy to see that φ is positive if and only if X is a positive matrix; φ is unital if trX = 1. (Positive matrices of trace one are called density matrices in the physics literature.)

(iii)  Let images/nec-47-3.png , the sum of all entries of A. If e is the vector with all of its entries equal to one, and E = ee, the matrix with all entries equal to one, then

images/eq-47-1.png

Thus φ is a positive linear functional.

(iv)  The map images/nec-47-4.png is a positive map of images/nec-47-5.png into itself. (Its range consists of scalar matrices.)

(v)   Let Atr denote the transpose of A. Then the map Φ(A) = Atr is positive.

(vi)  Let X be an n × k matrix. Then Φ(A) = XAX is a positive map from images/nec-47-6.png into images/nec-47-7.png.

(vii)  A special case of this is the compression map that takes an n× n matrix to a k × k block in its top left corner.

(viii)  Let P1, . . . , Pr be mutually orthogonal projections with images/nec-47-8.png Pr = I. The operator images/nec-47-9.png is called a pinching of A. In an appropriate coordinate system this can be described as

images/eq-47-2.png

Every pinching is positive. A special case of this is r = n and each Pj is the projection onto the linear span of the basis vector ej. Then C(A) is the diagonal part of A.

(ix)  Let B be any positive matrix. Then the map images/nec-48-1.png is positive. So is the map Φ(A) = AB.

(x)  Let A be a matrix whose spectrum is contained in the open right half plane. Let LA(X) = AX + XA. The operator LA on images/nec-48-2.png is invertible and its inverse images/nec-48-3.png is a positive linear map. (See the discussion in Exercise 1.2.10.)

(xi)  Any positive linear combination of positive maps is positive. Any convex combination of positive unital maps is positive and unital.


It is instructive to think of positive maps as noncommutative (matrix) averaging operations. Let C(X) be the space of continuous functions on a compact metric space. Let φ be a linear functional on C(X). By the Riesz representation theorem, there exists a signed measure µ on X such that

images/eq-48-1.png(2.3)

The linear functional φ is called positive if φ(f) ≥ 0 for every (pointwise) nonnegative function f. For such a φ, the measure µ representing it is a positive measure. If φ maps the function f ≡ 1 to the number 1, then φ is said to be unital, and then µ is a probability measure. The integral (2.3) is then written as

images/eq-48-2.png(2.4)

and called the expectation of f. Every positive, unital, linear functional on C(X) is an expectation (with respect to a probability measure µ). A positive, unital, linear map Φ may thus be thought of as a noncommutative analogue of an expectation map.


2.3  SOME BASIC PROPERTIES OF POSITIVE MAPS

We prove three theorems due to Kadison, Choi, and Russo and Dye. Our proofs use 2 × 2 block matrix arguments.


2.3.1  Lemma

Every positive linear map is adjoint-preserving; i.e., Φ(T) = Φ(T) for all T.

Proof.   First we show that Φ(A) is Hermitian if A is Hermitian. Every Hermitian matrix A has a Jordan decomposition

images/eq-49-1.png

So,

images/eq-49-2.png

is the difference of two positive matrices, and is therefore Hermitian. Every matrix T has a Cartesian decomposition

images/eq-49-3.png

So,

images/eq-49-4.png

2.3.2  Theorem ( Kadison’s Inequality)

Let Φ be positive and unital. Then for every Hermitian A

images/eq-49-5.png(2.5)

Proof.   By the spectral theorem, images/nec-49-1.png, where λj are the eigenvalues of A and Pj the corresponding projections with images/nec-49-2.png. Then images/nec-49-3.png and

images/eq-49-6.png

Since Pj are positive, so are Φ(Pj). Therefore,

images/eq-49-7.png

Each summand in the last sum is positive and, hence, so is the sum. By Theorem 1.3.3, therefore,

images/eq-50-1.png

2.3.3  Exercise

The inequality (2.5) may not be true if Φ is not unital.

Recall that for real functions we have (Ef)2Ef2. The inequality (2.5) is a noncommutative version of this. It should be pointed out that not all inequalities for expectations of real functions have such noncommutative counterparts. For example, we do have (Ef)4Ef4, but the analogous inequality Φ(A)4 ≤ Φ(A4) is not always true. To see this, let Φ be the compression map from images/nec-50-1.png to images/nec-50-2.png, taking a 3 × 3 matrix to its top left 2 × 2 submatrix. Let

images/eq-50-2.png

Then images/nec-50-3.png and images/nec-50-4.png .

This difference can be attributed to the fact that while the function f(t) = t4 is convex on the real line, the matrix function f(A) = A4 is not convex on Hermitian matrices.

The following theorem due to Choi generalizes Kadison’s inequality to normal operators.


2.3.4  Theorem (Choi)

Let Φ be positive and unital. Then for every normal matrix A

images/eq-50-3.png(2.6)

Proof.   The proof is similar to the one for Theorem 2.3.2. We have

images/eq-50-4.png

So

images/eq-51-1.png

is positive.     ■

In Chapter 3, we will see that the condition that A be normal can be dropped if we impose a stronger condition (2-positivity) on Φ.


2.3.5  Exercise

If A is normal, then Φ(A) need not be normal. Thus the left-hand sides of the two inequalities (2.6) can be different.


2.3.6  Theorem (Choi’s Inequality)

Let Φ be strictly positive and unital. Then for every strictly positive matrix A

images/eq-51-2.png(2.7)

Proof.   The proof is again similar to that of Theorem 2.3.2. Now we have images/nec-51-1.png with λj > 0. Then images/nec-51-2.png, and

images/eq-51-3.png

is positive. Hence, by Theorem 1.3.3

images/eq-51-4.png

2.3.7  Theorem (The Russo-Dye Theorem)

If Φ is positive and unital, then ||Φ|| = 1.

Proof.   We show first that ||Φ(U)|| ≤ 1 when U is unitary. In this case the eigenvalues λj are complex numbers of modulus one. So, from the spectral resolution images/nec-51-3.png, we get

images/eq-51-5.png

Hence, by Proposition 1.3.1, ||Φ(U)|| ≤ 1. Now if A is any contraction, then we can write images/nec-52-1.png where U, V are unitary. (Use the singular value decomposition of A and observe that if 0 ≤ s ≤ 1, then we have images/nec-52-2.png for some θ.) So

images/eq-52-1.png

Thus ||Φ|| ≤ 1, and since Φ is unital ||Φ|| = 1.     ■

Second proof.  Let ||A|| ≤ 1. Then A has a unitary dilation Â

images/eq-52-2.png(2.8)

(Check that this is a unitary element of M2n.)

Now let Ψ be the compression map taking a 2n × 2n matrix to its top left n × n corner. Then Ψ is positive and unital. So, the composition Φ ◦ Ψ is positive and unital. Now Choi’s inequality (2.6) can be used to get

images/eq-52-3.png

But this says

images/eq-52-4.png

This shows that ||Φ(A)|| ≤ 1 whenever ||A|| ≤ 1. Hence, ||Φ|| = 1.     ■

We can extend the result to any positive linear map as follows.


2.3.8  Corollary

Let Φ be a positive linear map. Then ||Φ|| = ||Φ(I)||.


Proof.   Let P = Φ(I), and assume first that P is invertible. Let

images/eq-52-5.png

Then Ψ is a positive unital linear map. So, we have

images/eq-52-6.png

Thus ||Φ|| ≤ ||P||; and since Φ(I) = P, we have ||Φ|| = ||P||. This proves the assertion when Φ(I) is invertible. The general case follows from this by considering the family Φε(A) = Φ(A) + εI and letting ε ↓ 0.     ■

The assertion of (this Corollary to) the Russo-Dye theorem is some times phrased as: every positive linear map on images/nec-53-1.png attains its norm at the identity matrix.


2.3.9  Exercise

There is a simpler proof of this theorem in the case of positive linear functionals. In this case φ(A) = trAX for some positive matrix X. Then

images/eq-53-1.png

Here ||T||1 is the trace norm of T defined as ||T||1 = s1(T)+· · ·+sn(T). The inequality above is a consequence of the fact that this norm is the dual of the norm || · ||.


2.4  SOME APPLICATIONS

We have seen several examples of positive maps. Using the Russo-Dye Theorem we can calculate their norms easily. Thus, for example,

images/eq-53-2.png(2.9)

for every pinching of A. (This can be proved in several ways. See MA pp. 50, 97.)

If A is positive, then the Schur multiplier SA is a positive map. So,

images/eq-53-3.png(2.10)

This too can be proved in many ways. We have seen this before in Theorem 1.4.1.

We have discussed the Lyapunov equation

images/eq-53-4.png(2.11)

where A is an operator whose spectrum is contained in the open right half plane. (Exercise 1.2.10, Example 2.2.1 (x)). Solving this equation means finding the inverse of the Lyapunov operator LA defined as LA(X) = AX + XA. We have seen that images/nec-54-1.png is a positive linear map. In some situations W is known with some imprecision, and we have the perturbed equation

images/eq-54-1.png(2.12)

If X and XX are the solutions to (2.11) and (2.12), respectively, one wants to find bounds for ||ΔX||. This is a very typical problem in numerical analysis. Clearly,

images/eq-54-2.png

Since images/nec-54-2.png is positive we have images/nec-54-3.png. This simplifies the problem considerably. The same considerations apply to the Stein equation (Exercise 1.2.11).

Let images/nec-54-4.png be the k-fold tensor product images/nec-54-5.png and let images/nec-54-6.png be the k-fold product images/nec-54-7.png of an operator A on images/nec-54-8.png. For 1 ≤ kn, let images/nec-54-9.png be the subspace of images/nec-54-10.png spanned by antisymmetric tensors. This is called the antisymmetric tensor product, exterior product, or Grassmann product. The operator images/nec-54-11.png leaves this space invariant and the restriction of images/nec-54-12.png to it is denoted as ∧kA. This is called the kth Grassmann power, or the exterior power of A.

Consider the map images/nec-54-13.png. The derivative of this map at A, denoted as images/nec-54-14.png, is a linear map from images/nec-54-15.png into images/nec-54-16.png. Its action is given as

images/eq-54-3.png

Hence,

images/eq-54-4.png(2.13)

It follows that

images/eq-54-5.png(2.14)

We want to find an expression for ||Dk (A)||.

Recall that ∧k is multiplicative, ∗ - preserving, and unital (but not linear!). Let A = USV be the singular value decomposition of A. Then

images/eq-55-1.png

Thus

images/eq-55-2.png

and hence

images/eq-55-3.png

Thus to calculate ||Dk (A)||, we may assume that A is positive and diagonal.

Now note that if A is positive, then for every positive B, the expression (2.13) is positive. So images/nec-55-1.png is a positive linear map from images/nec-55-2.png into images/nec-55-3.png. The operator Dk (A)(B) is the restriction of (2.13) to the invariant subspace images/nec-55-4.png. So ∧k(A) is also a positive linear map. Hence

images/eq-55-4.png

Let A = diag(s1, . . . , sn) with s1s2 ≥ · · · ≥ sn ≥ 0. Then ∧kA is a diagonal matrix of size images/nec-55-5.png whose diagonal entries are si1si2 · · · sik, 1 ≤ i1 < i2 < · · · < ikn. Use this to see that

images/eq-55-5.png(2.15)

the elementary symmetric polynomial of degree k − 1 with arguments s1, . . . , sk.

The effect of replacing Dk (A)(B) by Dk (A)(I) is to reduce a highly noncommutative problem to a simple commutative one. Another example of this situation is given in Section 2.7.


2.5  THREE QUESTIONS

Let images/nec-56-1.png be a linear map. We have seen that if Φ is positive, then

images/eq-56-1.png(2.16)

Clearly, this is a useful and important theorem. It is natural to explore how much, and in what directions, it can be extended.


Question 1   Are there linear maps other than positive ones for which (2.16) is true? In other words, if a linear map Φ attains its norm at the identity, then must Φ be positive?


Before attempting an answer, we should get a small irritant out of the way. If the condition (2.16) is met by Φ, then it is met by −Φ also. Clearly, both of them cannot be positive maps. So assume Φ satisfies (2.16) and

images/eq-56-2.png(2.17)

2.5.1  Exercise

If k = 1, the answer to our question is yes. In this case φ(A) = trAX for some X. Then ||φ|| = ||X||1 (see Exercise 2.3.9). So, if φ satisfies (2.16) and (2.17), then ||X||1 = trX. Show that this is true if and only if X is positive. Hence φ is positive.


If k ≥ 2, this is no longer true. For example, let Φ be the linear map on images/nec-56-2.png defined as

images/eq-56-3.png

Then ||Φ|| = ||Φ(I)|| = 1 and Φ(I) ≥ O, but Φ is not positive.

It is a remarkable fact that if Φ is unital and ||Φ|| = 1, then Φ is positive. Thus supplementing (2.16) with the condition Φ(I) = I ensures that Φ is positive. This is proved in the next section.


Question 2   Let S be a linear subspace of images/nec-57-1.png and let images/nec-57-2.png be a linear map. Do we still have a theorem like the Russo-Dye theorem? In other words how crucial is the fact that the domain of Φ is Mn (or possibly a subalgebra)?


Again, for the question to be meaningful, we have to demand of S a little more structure. If we want to talk of positive unital maps, then S must contain some positive elements and I.


2.5.2  Definition

A linear subspace S of images/nec-57-3.png is called an operator system if it is ∗ closed (i.e., if A ∈ S, then A ∈ S ) and contains I.


Let S be an operator system. We want to know whether a positive linear map images/nec-57-4.png attains its norm at I. The answer is yes if k = 1, and no if k ≥ 2. However, we do have ||Φ|| ≤ 2||Φ(I)|| for all k.

A related question is the following:


Question 3   By the Hahn-Banach theorem, every linear functional φ on (a linear subspace) S can be extended to a linear functional images/nec-57-5.png on images/nec-57-6.png in such a way that images/nec-57-7.png. Now we are considering positivity rather than norms. So we may ask whether a positive linear functional φ on an operator system S in images/nec-57-8.png can be extended to a positive linear functional images/nec-57-9.png on images/nec-57-10.png. The answer is yes. This is called the Krein extension theorem. Then since images/nec-57-11.png, we have ||φ|| = φ(I).


Next we may ask whether a positive linear map Φ from images/nec-57-12.png into images/nec-57-13.png can be extended to a positive linear map images/nec-57-14.png from images/nec-57-15.png into images/nec-57-16.png. If this were the case, then we would have ||Φ|| = ||Φ(I)||. But we have said that this is not always true when k ≥ 2. This is illustrated by the following example.


2.5.3  Example

Let n be any number bigger than 2 and let S be the n×n permutation matrix

images/eq-58-1.png

Let S be the collection of all matrices C of the form C = aI+bS+cS, a, b, cimages/nec-58-1.png. (The matrices C are circulant matrices.) Then images/nec-58-2.png is an operator system in images/nec-58-3.png. What are the positive elements of images/nec-58-4.png? First, we must have a ≥ 0 and images/nec-58-5.png. The eigenvalues of S are 1, ω, . . . , ωn−1, the n roots of 1. So, the eigenvalues of C are

images/eq-58-2.png

and C is positive if and only if all these numbers are nonnegative.

It is helpful to consider the special case n = 4. The fourth roots of unity are {1, i, −1,i} and one can see that a matrix C of the type above is positive if and only if

images/eq-58-3.png

Let images/nec-58-6.png be the map defined as

images/eq-58-4.png

Then Φ is linear, positive, and unital. Since

images/eq-58-5.png

images/nec-58-7.png. So, Φ cannot be extended to a positive, linear, unital map from images/nec-58-8.png into images/nec-58-9.png.


2.5.4  Exercise

Let n ≥ 3 and consider the operator system images/nec-58-10.png defined in the example above. For every t the map images/nec-58-11.png defined as

images/eq-58-6.png

is linear and unital. Show that for 1 < t < 2 there exists an n such that the map Φ is positive.

We should remark here that the elements of images/nec-59-1.png commute with each other (though, of course, images/nec-59-2.png is not a subalgebra of images/nec-59-3.png).

In the next section we prove the statements that we have made in answer to the three questions.


2.6  POSITIVE MAPS ON OPERATOR SYSTEMS

Let images/nec-59-4.png be an operator system in images/nec-59-5.png. the set of self-adjoint elements of images/nec-59-6.png, and images/nec-59-7.png the set of positive elements in it.

Some of the operations that we performed freely in images/nec-59-8.png may take us outside images/nec-59-9.png. Thus if images/nec-59-10.png, then Re images/nec-59-11.png and Im T = images/nec-59-12.png are in images/nec-59-13.png. However, if images/nec-59-14.png, then the positive and negative parts A± in the Jordan decomposition of A need not be in images/nec-59-15.png. For example, consider

images/eq-59-1.png

This is an operator system. The matrix images/nec-59-16.png is in images/nec-59-17.png. Its Jordan components are

images/eq-59-2.png

They do not belong to S.

However, it is possible still to write every Hermitian element A of images/nec-59-18.png as

images/eq-59-3.png(2.18)

Just choose

images/eq-59-4.png(2.19)

Thus we can write every images/nec-59-19.png as

images/eq-59-5.png
images/eq-60-1.png

Using this decomposition we can prove the following lemma.


2.6.1  Lemma

Let Φ be a positive linear map from an operator system images/nec-60-1.png into images/nec-60-2.png. Then Φ(T) = Φ(T) for all images/nec-60-3.png.


2.6.2  Exercise

If A = P1P2 where P1, P2 are positive, then

images/eq-60-2.png

2.6.3  Theorem

Let Φ be a positive linear map from an operator system images/nec-60-4.png into images/nec-60-5.png. Then

images/eq-60-3.png

and

images/eq-60-4.png

(Thus if Φ is also unital, then ||Φ|| = 1 on the space images/nec-60-6.png, and ||Φ|| ≤ 2 on images/nec-60-7.png.)


Proof.   If P is a positive element of images/nec-60-8.png, then OP ≤ ||P||I, and hence O ≤ Φ(P) ≤ ||P||Φ(I).

If A is a Hermitian element of images/nec-60-9.png, use the decomposition (2.18), Exercise 2.6.2, and the observation of the preceding sentence to see that

images/eq-60-5.png

This proves the first inequality of the theorem. The second is obtained from this by using the Cartesian decomposition of T.     ■

Exercise 2.5.4 shows that the factor 2 in the inequality (ii) of Theorem 2.6.3 is unavoidable in general. It can be dropped when k = 1:


2.6.4  Theorem

Let φ be a positive linear functional on an operator system images/nec-61-1.png. Then ||φ|| = φ(I).


Proof.   Let images/nec-61-2.png and ||T|| ≤ 1. We want to show |φ(T)| ≤ φ(I). If φ(T) is the complex number re, we may multiply T by e, and thus assume φ(T) is real and positive. So, if T = A+ iB in the Cartesian decomposition, then φ(T) = φ(A). Hence by part (i) of Theorem 2.6.3 φ(T) ≤ φ(I)||A|| ≤ φ(I)||T||.     ■


The converse is also true.


2.6.5  Theorem

Let φ be a linear functional on images/nec-61-3.png such that ||φ|| = φ(I). Then φ is positive.


Proof.   Assume, without loss of generality, that φ(I) = 1. Let A be a positive element of images/nec-61-4.png and let σ(A) be its spectrum. Let a = min σ(A) and b = max σ(A). We will show that the point φ(A) lies in the interval [a, b]. If this is not the case, then there exists a disk D = {z : |zz0| ≤ r} such that φ(A) is outside D but D contains [a, b], and hence also σ(A). From the latter condition it follows that σ(Az0I) is contained in the disk {z : |z| ≤ r} , and hence ||Az0I|| ≤ r. Using the conditions ||φ|| = φ(I) = 1, we get from this

images/eq-61-1.png

But then φ(A) could not have been outside D.

This shows that φ(A) is a nonnegative number, and the theorem is proved.     ■


2.6.6  Theorem (The Krein Extension Theorem)

Let images/nec-61-5.png be an operator system in images/nec-61-6.png. Then every positive linear functional on images/nec-61-7.png can be extended to a positive linear functional on images/nec-61-8.png.


Proof.   Let φ be a positive linear functional on images/nec-61-9.png. By Theorem 2.6.4, ||φ|| = φ(I). By the Hahn-Banach Theorem, φ can be extended to a linear functional images/nec-62-1.png on images/nec-62-2.png with images/nec-62-3.png. But then images/nec-62-4.png is positive by Theorem 2.6.5 (or by Exercise 2.5.1).     ■


Finally we have the following theorem that proves the assertion made at the end of the discussion of Question 1 in Section 2.5.


2.6.7  Theorem

Let images/nec-62-5.png be an operator system and let images/nec-62-6.pngk be a unital linear map such that ||Φ|| = 1. Then Φ is positive.


Proof.   For each unit vector x in images/nec-62-7.png, let

images/eq-62-1.png

This is a unital linear functional on images/nec-62-8.png. Since |φx(A)| ≤ ||Φ(A)|| ≤ ||A||, we have ||φx|| = 1. So, by Theorem 2.6.5, φx is positive. In other words, if A is positive, then for every unit vector x

images/eq-62-2.png

But that means Φ is positive.     ■


2.7  SUPPLEMENTARY RESULTS AND EXERCISES

Some of the theorems in Section 2.3 are extended in various directions in the following propositions.


2.7.1  Proposition

Let Φ be a positive unital linear map on images/nec-62-9.png and let A be a positive matrix. Then

images/eq-62-3.png

Proof.   Let 0 < r < 1. Using the integral representation (1.39) we have

images/eq-62-4.png

where µ is a positive measure on (0, ∞). So it suffices to show that

images/eq-62-5.png

for all λ > 0. We have the identity

images/eq-63-1.png

Apply Φ to both sides and use Theorem 2.3.6 to get

images/eq-63-2.png

The identity stated above shows that the last expression is equal to Φ(A)(λ + Φ(A))−1.     ■


2.7.2  Exercise

Let Φ be a positive unital linear map on images/nec-63-1.png and let A be a positive matrix. Show that

images/eq-63-3.png

if 1 ≤ r ≤ 2. If A is strictly positive, then this is true also when −1 ≤ r ≤ 0. [Hint: Use integral representations of Ar as in Theorem 1.5.8, Exercise 1.5.10, and the inequalities (2.5) and (2.7).]


2.7.3  Proposition

Let Φ be a strictly positive linear map on Mn. Then

images/eq-63-4.png(2.20)

whenever H is Hermitian and A > 0.


Proof.   Let

images/eq-63-5.png(2.21)

Then Ψ is positive and unital. By Kadison’s inequality we have Ψ(Y 2) ≥ Ψ(Y )2 for every Hermitian Y . Choose Y = A−1/2HA−1/2 to get

images/eq-63-6.png

Use (2.21) now to get (2.20).     ■


2.7.4  Exercise

Construct an example to show that a more general version of (2.20)

images/eq-64-1.png

where X is arbitrary and A positive, is not always true.


2.7.5  Proposition

Let Φ be a strictly positive linear map on images/nec-64-1.png and let A > O. Then

images/eq-64-2.png(2.22)

Proof.   Let Ψ be the linear map defined by (2.21). By the Russo-Dye theorem

images/eq-64-3.png

Let AXA−1X and put Y = A−1/2XA−1/2. Then Y Y = A−1/2 XA−1 XA−1/2I. Hence Ψ(A−1/2XA−1/2)Ψ(A−1/2XA−1/2) ≤ I. Use (2.21) again to get (2.22).     ■

In classical probability the quantity

images/eq-64-4.png(2.23)

is called the variance of the real function f. In analogy we consider

images/eq-64-5.png(2.24)

where A is Hermitian and Φ a positive unital linear map on images/nec-64-2.png. Kadison’s inequality says var(A) ≥ O. The following proposition gives an upper bound for var(A).


2.7.6  Proposition

Let Φ be a positive unital linear map and let A be a Hermitian operator with mIAMI. Then

images/eq-65-1.png(2.25)

Proof.   The matrices MIA and AmI are positive and commute with each other. So, (MIA)(AmI) ≥ O; i.e.,

images/eq-65-2.png

Apply Φ to both sides and then subtract Φ(A)2 from both sides. This gives the first inequality in (2.25). To prove the second inequality note that if mxM, then images/nec-65-1.png.     ■


2.7.7  Exercise

Let images/nec-65-2.png. We say x ≥ 0 if all its coordinates xj are nonnegative. Let e = (1, . . . , 1).

A matrix S is called stochastic if sij ≥ 0 for all i, j, and images/nec-65-3.png for all i. Show that S is stochastic if and only if

images/eq-65-3.png(2.26)

and

images/eq-65-4.png(2.27)

The property (2.26) can be described by saying that the linear map defined by S on images/nec-65-4.png is positive, and (2.27) by saying that S is unital.

If x is a real vector, let images/nec-65-5.png. Show that if S is a stochastic matrix and mxjM, then

images/eq-65-5.png(2.28)

A special case of this is obtained by choosing images/nec-65-6.png for all i, j. If images/nec-65-7.png, this gives

images/eq-65-6.png(2.29)

An inequality complementary to (2.7) is given by the following proposition.


2.7.8  Proposition

Let Φ be strictly positive and unital. Let 0 < m < M. Then for every strictly positive matrix A with mIAMI, we have

images/eq-66-1.png(2.30)

Proof.   The matrices AmI and MA−1I are positive and commute with each other. So, O ≤ (AmI)(MA−1I). This gives

images/eq-66-2.png

and hence

images/eq-66-3.png

Now, if c and x are real numbers, then (c − 2x)2 ≥ 0 and therefore, for positive x we have images/nec-66-1.png. So, we get

images/eq-66-4.png

A very special corollary of this is the inequality

images/eq-66-5.png(2.31)

for every unit vector x. This is called the Kantorovich inequality.


2.7.9  Exercise

Let f be a convex function on an interval [m, M] and let L be the linear interpolant

images/eq-66-6.png

Show that if Φ is a unital positive linear map, then for every Hermitian matrix A whose spectrum is contained in [m, M], we have

images/eq-67-1.png

Use this to obtain Propositions 2.7.6 and 2.7.8.

The space images/nec-67-1.png has a natural inner product defined as

images/eq-67-2.png(2.32)

If Φ is a linear map on images/nec-67-2.png, we define its adjoint Φ as the linear map that satisfies the condition

images/eq-67-3.png(2.33)

2.7.10  Exercise

The linear map Φ is positive if and only if Φ is positive. Φ is unital if and only if Φ is trace preserving; i.e., tr Φ(A) = tr A for all A.


We say Φ is a doubly stochastic map on images/nec-67-3.png if it is positive,unital, and trace preserving (i.e., both Φ and Φ are positive and unital).


2.7.11  Exercise

(i)  Let Φ be the linear map on Mn defined as Φ(A) = XAX. Show that Φ(A) = XAX.

(ii)  For any A, let SA(X) = AX be the Schur product map. Show that (SA) = SA∗.

(iii)  Every pinching is a doubly stochastic map.

(iv)  Let LA(X) = AX + XA be the Lyapunov operator, where A is a matrix with its spectrum in the open right half plane. Show that (L A 1) = (LA∗ )−1.

A norm ||| · ||| on Mn is said to be unitarily invariant if |||UAV ||| = |||A||| for all A and unitary U, V . It is convenient to make a normalisation so that |||A||| = 1 whenever A is a rank-one orthogonal projection.

Special examples of such norms are the Ky Fan norms

images/eq-68-1.png

and the Schatten p-norms

images/eq-68-2.png

Note that the operator norm, in this notation, is

images/eq-68-3.png(1)

and the trace norm is the norm

images/eq-68-4.png

The norm ||A||2 is also called the Hilbert-Schmidt norm.

The following facts are well known:

images/eq-68-5.png(2.34)

If ||A||(k) ≤ ||B||(k) for 1 ≤ kn, then |||A||| ≤ |||B||| for all unitarily invariant norms. This is called the Fan dominance theorem. (See MA, p. 93.)

For any three matrices A, B, C we have

images/eq-68-6.png(2.35)

If Φ is a linear map on images/nec-68-1.png and ||| · ||| any unitarily invariant norm, then we use the notation |||Φ||| for

images/eq-68-7.png(2.36)

In the same way,

images/eq-68-8.png

etc.

The norm ||A||1 is the dual of the norm ||A|| on Mn. Hence

images/eq-69-1.png(2.37)

2.7.12  Exercise

Let ||| · ||| be any unitarily invariant norm on Mn.

(i)   Use the relations (2.34) and the Fan dominance theorem to show that if ||Φ|| ≤ 1 and ||Φ|| ≤ 1, then |||Φ||| ≤ 1.

(ii)  If Φ is a doubly stochastic map, then |||Φ||| ≤ 1.

(iii)  If AO, then |||AX||| ≤ max aii|||X||| for all X.

(iv)  Let LA be the Lyapunov operator associated with a positively stable matrix A. We know that images/nec-69-1.png. Show that in the special case when A is normal we have images/nec-69-2.png = [2 min Re λi]−1, where λi are the eigenvalues of A.


2.7.13  Exercise

Let A and B be Hermitian matrices. Suppose A = Φ(B) for some doubly stochastic map Φ on Mn. Show that A is a convex combination of unitary conjugates of B; i.e., there exist unitary matrices U1, . . . , Uk and positive numbers p1, . . . , pk with images/nec-69-3.png such that

images/eq-69-2.png

[Hints: There exist diagonal matrices D1 and D2, and unitary matrices W and V such that A = WD1W and B = V D2V . Use this to show that D1 = Ψ(D2) where Ψ is a doubly stochastic map. By Birkhoff’s theorem there exist permutation matrices S1, . . . , Sk and positive numbers p1, . . . , pk with images/nec-69-4.png such that

images/eq-69-3.png

Choose Uj = V SjW. (Note that the matrices Uj and the numbers pj depend on Φ, A and B.)]

Let images/nec-70-1.png be the set of all n × n Hermitian matrices. This is a real vector space. Let I be an open interval and let C1(I) be the space of continuously differentiable real functions on I. Let images/nec-70-2.png be the set of all Hermitian matrices whose eigenvalues belong to I. This is an open subset of images/nec-70-3.png. Every function f in C1(I) induces a map images/nec-70-4.png from images/nec-70-5.png into images/nec-70-6.png. This induced map is differentiable and its derivative is given by an interesting formula known as the Daleckii-Krein formula.

For each images/nec-70-7.png the derivative Df(A) at A is a linear map from images/nec-70-8.png into itself. If images/nec-70-9.png is the spectral decomposition of A, then the formula is

images/eq-70-1.png(2.38)

for every images/nec-70-10.png. For i = j, the quotient in (2.38) is to be interpreted as f(λi).

This formula can be expressed in another way. Let f[1] be the function on I × I defined as

images/eq-70-2.png

This is called the first divided difference of f. For images/nec-70-11.png, let f[1](A) be the n × n matrix

images/eq-70-3.png(2.39)

where λ1, . . . , λn are the eigenvalues of A. The formula (2.38) says

images/eq-70-4.png(2.40)

where ◦ denotes the Schur product taken in a basis in which A is diagonal. A proof of this is given in Section 5.3.

Suppose a real function f on an interval I has the following property: if A and B are two elements of images/nec-70-12.png and AB, then f(A) ≥ f(B). We say that such a function f is matrix monotone of order n on I. If f is matrix monotone of order n for all n = 1, 2, . . . , then we say f is operator monotone.

Matrix convexity of order n and operator convexity can be defined in a similar way. In Chapter 1 we have seen that the function f(t) = t2 on the interval [0, ∞) is not matrix monotone of order 2, and the function f(t) = t3 is not matrix convex of order 2. We have seen also that the function f(t) = tr on the interval [0, ∞) is operator monotone for 0 ≤ r ≤ 1, and it is operator convex for 1 ≤ r ≤ 2 and for −1 ≤ r ≤ 0. More properties of operator monotone and convex functions are studied in Chapters 4 and 5.

It is not difficult to prove the following, using the formula (2.40).


2.7.14  Exercise

If a function f ∈ C1(I) is matrix monotone of order n, then for each images/nec-71-1.png, the matrix f[1](A) defined in (2.39) is positive.


The converse of this statement is also true. A proof of this is given in Section 5.3. At the moment we note the following interesting consequence of the positivity of f[1](A).


2.7.15  Exercise

Let f ∈ C1(I) and let f be the derivative of f. Show that if f is matrix monotone of order n, then for each images/nec-71-2.png

images/eq-71-1.png(2.41)

By definition

images/eq-71-2.png(2.42)

and

images/eq-71-3.png

This expression is difficult to calculate for functions such as f(t) = tr, 0 < r < 1. The formula (2.41) gives an easy way to calculate its norm. Its effect is to reduce the supremum in (2.42) to the class of matrices B that commute with A.


2.8  NOTES AND REFERENCES

Since positivity is a useful and interesting property, it is natural to ask what linear transformations preserve it. The variety of interesting examples, and their interpretation as “expectation,” make positive linear maps especially interesting. Their characterization, however, has turned out to be slippery, and for various reasons the special class of completely positive linear maps has gained in importance.

Among the early major works on positive linear maps is the paper by E. Størmer, Positive linear maps of operator algebras, Acta Math., 110 (1963) 233–278. Research expository articles that explain several subtleties include E. Størmer, Positive linear maps of C-algebras, in Foundations of Quantum Mechanics and Ordered Linear Spaces, Lecture Notes in Physics, Vol. 29, Springer, 1974, pp.85–106, and M.-D. Choi, Positive linear maps, in Operator Algebras and Applications, Part 2, R. Kadison ed., American Math. Soc., 1982. Closer to our concerns are Chapter 2 of V. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge University Press, 2002, and sections of the two reports by T. Ando, Topics on Operator Inequalities, Sapporo, 1978 and Operator-Theoretic Methods for Matrix Inequalities, Sapporo, 1998.

The inequality (2.5) was proved in the paper R. Kadison, A generalized Schwarz inequality and algebraic invariants for operator algebras, Ann. Math., 56 (1952) 494–503. This was generalized by C. Davis, A Schwarz inequality for convex operator functions, Proc. Am. Math. Soc., 8 (1957) 42–44, and by M.-D. Choi, A Schwarz inequality for positive linear maps on C-algebras, Illinois J. Math., 18 (1974) 565–574. The generalizations say that if Φ is a positive unital linear map and f is an operator convex function, then we have a Jensen-type inequality

images/eq-72-1.png(2.43)

The inequality (2.7) and the result of Exercise 2.7.2 are special cases of this. Using the integral representation of an operator convex function given in Problem V.5.5 of MA, one can prove the general inequality by the same argument as used in Exercise 2.7.2. The inequality (2.43) characterises operator convex functions, as was noted by C. Davis, Notions generalizing convexity for functions defined on spaces of matrices, in Proc. Symposia Pure Math., Vol. VII, Convexity, American Math. Soc., 1963.

In our proof of Theorem 2.3.7 we used the fact that any contraction is an average of two unitaries. The infinite-dimensional analogue says that the unit ball of a C-algebra is the closed convex hull of the unitary elements. (Unitaries, however, do not constitute the full set of extreme points of the unit ball. See P. R. Halmos, A Hilbert Space Problem Book, Second Edition, Springer, 1982.) This theorem about the closed convex hull is also called the Russo-Dye theorem and was proved in B. Russo and H. A. Dye, A note on unitary operators in C-algebras, Duke Math. J., 33 (1966) 413–416.

Applications given in Section 2.4 make effective use of Theorem 2.3.7 in calculating norms of complicated operators. Our discussion of the Lyapunov equation follows the one in R. Bhatia and L. Elsner, Positive linear maps and the Lyapunov equation, Oper. Theory: Adv. Appl., 130 (2001) 107–120. As pointed out in this paper, the use of positivity leads to much more economical proofs than those found earlier by engineers. The equality (2.15) was first proved by R. Bhatia and S. Friedland, Variation of Grassman powers and spectra, Linear Algebra Appl., 40 (1981) 1–18. The alternate proof using positivity is due to V. S. Sunder, A noncommutative analogue of |DXk| = |kXk−1|, ibid., 44 (1982) 87-95. The analogue of the formula (2.15) when the antisymmetric tensor product is replaced by the symmetric one was worked out in R. Bhatia, Variation of symmetric tensor powers and permanents, ibid., 62 (1984) 269–276. The harder problem embracing all symmetry classes of tensors was solved in R. Bhatia and J. A. Dias da Silva, Variation of induced linear operators, ibid., 341 (2002) 391–402.

Because of our interest in certain kinds of matrix problems involving calculation or estimation of norms we have based our discussion in Section 2.5 on the relation (2.16). There are far more compelling reasons to introduce operator systems. There is a rapidly developing and increasingly important theory of operator spaces (closed linear subspaces of C-algebras) and operator systems. See the book by V. Paulsen cited earlier, E. G. Effros and Z.-J. Ruan, Operator Spaces, Oxford University Press, 2000, and G. Pisier, Introduction to Operator Space Theory, Cambridge University Press, 2003. This is being called the noncommutative or quantized version of Banach space theory. One of the corollaries of the Hahn-Banach theorem is that every separable Banach space is isometrically isomorphic to a subspace of l; and every Banach space is isometrically isomorphic to a subspace of l(X) for some set X. In the quantized version the commutative space l is replaced by the noncommutative space images/nec-73-1.png where images/nec-73-2.png is a Hilbert space. Of course, it is not adequate functional analysis to study just the space l and its subspaces. Likewise subspaces of images/nec-73-3.png are called concrete operator spaces, and then subsumed in a theory of abstract operator spaces.

Our discussion in Section 2.6 borrows much from V. Paulsen’s book. Some of our proofs are simpler because we are in finite dimensions.

Propositions 2.7.3 and 2.7.5 are due to M.-D. Choi, Some assorted inequalities for positive linear maps on C-algebras, J. Operator Theory, 4 (1980) 271–285. Propositions 2.7.6 and 2.7.8 are taken from R. Bhatia and C. Davis, A better bound on the variance, Am. Math. Monthly, 107 (2000) 602–608. Inequalities (2.29), (2.31) and their generalizations are important in statistics, and have been proved by many authors, often without knowledge of previous work. See the article S. W. Drury, S. Liu, C.-Y. Lu, S. Puntanen, and G. P. H. Styan, Some comments on several matrix inequalities with applications to canonical correlations: historical background and recent developments, Sankhyā, Series A, 64 (2002) 453–507.

The Daleckii-Krein formula was presented in Ju. L. Daleckii and S. G. Krein, Formulas of differentiation according to a parameter of functions of Hermitian operators, Dokl. Akad. Nauk SSSR, 76 (1951) 13–16. Infinite dimensional analogues in which the double sum in (2.38) is replaced by a double integral were proved by M. Sh. Birman and M. Z. Solomyak, Double Stieltjes operator integrals (English translation), Topics in Mathematical Physics Vol. 1, Consultant Bureau, New York, 1967.

The formula (2.41) was noted in R. Bhatia, First and second order perturbation bounds for the operator absolute value, Linear Algebra Appl., 208/209 (1994) 367–376. It was observed there that this equality of norms holds for several other functions that are not operator monotone. If A is positive and f(A) = Ar, then the equality (2.41) is true for all real numbers r other than those in images/nec-74-1.png. This, somewhat mysterious, result was proved in two papers: R. Bhatia and K. B. Sinha, Variation of real powers of positive operators, Indiana Univ. Math. J., 43 (1994) 913–925, and R. Bhatia and J. A. R. Holbrook, Fréchet derivatives of the power function, ibid., 49(2000) 1155–1173. Similar equalities involving higher-order derivatives have been proved in R. Bhatia, D. Singh, and K. B. Sinha, Differentiation of operator functions and perturbation bounds, Commun. Math. Phys., 191 (1998) 603–611.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.221.163.13