Chapter Three



Completely Positive Maps

For several reasons a special class of positive maps, called completely positive maps, is especially important. In Section 3.1 we study the basic properties of this class of maps. In Section 3.3 we derive some Schwarz type inequalities for this class; these are not always true for all positive maps. In Sections 3.4 and 3.5 we use general results on completely positive maps to study some important problems for matrix norms.

Let images/nec-75-1.png be the space of m× m block matrices [[Aij]] whose i, j entry is an element of images/nec-75-2.png. Each linear map images/nec-75-3.png induces a linear map images/nec-75-4.png defined as

images/eq-75-1.png(3.1)

We say that Φ is m-positive if the map Φm is positive, and Φ is completely positive if it is m-positive for all m = 1, 2, . . .. Thus positive maps are 1-positive.

The map Φ(A) = Atr on images/nec-75-5.png is positive but not 2-positive. To see this consider the 2 × 2 matrices Eij whose i, j entry is one and the remaining entries are zero. Then [[Eij]] is positive, but [[Φ(Eij)]] is not.

Let images/nec-75-6.png, the space of n × k matrices. Then the map Φ(A) = V AV from images/nec-75-7.png into images/nec-75-8.png is completely positive. To see this note that for each m

images/eq-75-2.png

If images/nec-75-9.png, then

images/eq-75-3.png(3.2)

is completely positive.

Let φ be any positive linear functional on images/nec-76-1.png. Then there exists a positive matrix X such that φ(A) = tr AX for all A. If uj, 1 ≤ jn, constitute an orthonormal basis for images/nec-76-2.png, then we have

images/eq-76-1.png

So, if we put vj = X1/2uj, we have

images/eq-76-2.png

This shows that in the special case k = 1, every positive linear map images/nec-76-3.png can be represented in the form (3.2) and thus is completely positive.


3.1  SOME BASIC THEOREMS

Let us fix some notations. The standard basis for images/nec-76-4.png will be written as ej, 1 ≤ jn. The matrix images/nec-76-5.png will be written as Eij. This is the matrix with its i, j entry equal to one and all other entries equal to zero. These matrices are called matrix units. The family {Eij : 1 ≤ i, jn} spans images/nec-76-6.png.

Our first theorem says all completely positive maps are of the form (3.2).


3.1.1  Theorem (Choi, Kraus)

Let images/nec-76-7.png be a completely positive linear map. Then there exist images/nec-76-8.png, such that

images/eq-76-3.png(3.3)

Proof.   We will find Vj such that the relation (3.3) holds for all matrix units Ers in images/nec-76-9.png. Since Φ is linear and the Ers span images/nec-76-10.png this is enough to prove the theorem.

We need a simple identification involving outer products of block vectors. Let images/nec-76-11.png. We think of v as a direct sum images/nec-76-12.png, where images/nec-76-13.png; or as a column vector

images/eq-77-1.png

Identify this with the k × n matrix

images/eq-77-2.png

whose columns are the vectors xj. Then note that

images/eq-77-3.png

So, if we think of vv as an element of images/nec-77-1.png we have

images/eq-77-4.png(3.4)

The matrix images/nec-77-2.png is a positive element of images/nec-77-3.png. So, if images/nec-77-4.png is an n-positive map, [[Φ(Ers)]] is a positive element of images/nec-77-5.png.

By the spectral theorem, there exist vectors images/nec-77-6.png such that

images/eq-77-5.png

Thus for all 1 ≤ r, sn

images/eq-77-6.png

as required.     ■

Note that in the course of the proof we have shown that if a linear map images/nec-77-7.png is n-positive, then it is completely positive. We have shown also that if Φn([[Ers]]) is positive, then Φ is completely positive.

The vectors vj occurring in the proof are not unique; and so the Vj in the representation are not unique. If we impose the condition that the family {vj} does not contain any zero vector and all vectors in it are mutually orthogonal, then the Vj in (3.3) are unique up to unitary conjugations. The proof of this statement is left as an exercise.

The map Φ is unital if and only if images/nec-78-1.png. Unital completely positive maps form a convex set. We state, without proof, two facts about its extreme points. The extreme points are those Φ for which the set {ViVj : 1 ≤ i, jnk} is linearly independent. For such Φ, the number of terms in the representation (3.3) is at most k.


3.1.2  Theorem (The Stinespring Dilation Theorem)

Let images/nec-78-2.png be a completely positive map. Then there exist a representation

images/eq-78-1.png

and an operator

images/eq-78-2.png

such that ||V ||2 = ||Φ(I)|| and

images/eq-78-3.png

Proof.   The equation (3.3) can be rewritten as

images/eq-78-4.png
images/eq-78-5.png

Note that if Φ is unital, then V V = I. Hence V is an isometric embedding of images/nec-79-1.png in images/nec-79-2.png and V a projection. The representation images/nec-79-3.png is a direct sum of nk copies of A. This number could be smaller in several cases. The representation with the minimal number of copies is unique upto unitary conjugation.


3.1.3  Corollary

Let images/nec-79-4.png be completely positive. Then ||Φ|| = ||Φ(I)||. (This is true, more generally, for all positive linear maps, as we saw in Chapter 2.)


Next we consider linear maps whose domain is a linear subspace images/nec-79-5.png and whose range is images/nec-79-6.png. To each element Φ of images/nec-79-7.png corresponds a unique element φ of images/nec-79-8.png. This correspondence is described as follows. Let Sij, 1 ≤ i, jk be elements of images/nec-79-9.png. Then

images/eq-79-1.png(3.5)

where we use the notation [T]i,j for the i, j entry of a matrix T.

If ej, 1 ≤ jk is the standard basis for images/nec-79-10.png, and x is the vector in images/nec-79-11.png given by images/nec-79-12.png, then (3.5) can be written as

images/eq-79-2.png(3.6)

In the reverse direction, suppose φ is a linear functional on images/nec-79-13.png. Given an images/nec-79-14.png let Φ(A) be the element of images/nec-79-15.png whose i, j entry is

images/eq-79-3.png(3.7)

where Eij, 1 ≤ i, jk, are the matrix units in images/nec-79-16.png.

It is easy to see that this sets up a bijective correspondence between the spaces images/nec-79-17.png and images/nec-79-18.png. The factor 1/k in (3.5) ensures that Φ is unital if and only if φ is unital.


3.1.4  Theorem

Let images/nec-80-1.png be an operator system in images/nec-80-2.png, and let images/nec-80-3.png be a linear map. Then the following three conditions are equivalent:

(i)   Φ is completely positive.

(ii)  Φ is k-positive.

(iii)  The linear functional φ defined by (3.5) is positive.


Proof.   Obviously (i) images/nec-80-4.png (ii). It follows from (3.6) that (ii) images/nec-80-5.png (iii). The hard part of the proof consists of establishing the implication (iii) images/nec-80-6.png (i).

Since images/nec-80-7.png is an operator system in images/nec-80-8.png is an operator system in images/nec-80-9.png. By Krein’s extension theorem (Theorem 2.6.6), the positive linear functional φ on images/nec-80-10.png has an extension images/nec-80-11.png, a positive linear functional on images/nec-80-12.png. To this images/nec-80-13.png corresponds an element images/nec-80-14.png of images/nec-80-15.png defined via (3.7). This images/nec-80-16.png is an extension of Φ (since images/nec-80-17.png is an extension of φ). If we show images/nec-80-18.png is completely positive, it will follow that Φ is completely positive.

Let m be any positive integer. Every positive element of images/nec-80-19.png can be written as a sum of matrices of the type images/nec-80-20.png where Aj, 1 ≤ jm are elements of images/nec-80-21.png. To show that images/nec-80-22.png is m-positive, it suffices to show that images/nec-80-23.png is positive. This is an mk × mk matrix. Let x be any vector in images/nec-80-24.png. Write it as

images/eq-80-1.png

Then

images/eq-80-2.png(3.8)

using (3.7). For 1 ≤ im let Xi be the k × k matrix

images/eq-81-1.png

Then images/nec-81-1.png. In other words

images/eq-81-2.png

So (3.8) can be written as

images/eq-81-3.png

Since images/nec-81-2.png is positive, this expression is positive. That completes the proof.     ■


In the course of the proof we have also proved the following.


3.1.5  Theorem (Arveson’s Extension Theorem)

Let images/nec-81-3.png be an operator system in images/nec-81-4.png and let images/nec-81-5.png be a completely positive map. Then there exists a completely positive map images/nec-81-6.png that is an extension of Φ.

Let us also record the following fact that we have proved.


3.1.6  Theorem

Let images/nec-81-7.png be a linear map. Let m = min(n, k). If Φ is m-positive, then it is completely positive.


For l < m, there exists a map Φ that is l-positive but not (l + 1)-positive.

We have seen that completely positive maps have some desirable properties that positive maps did not have: they can be extended from an operator system images/nec-82-1.png to the whole of images/nec-82-2.png, and they attain their norm at I for this reason (even when they have been defined only on images/nec-82-3.png). Also, there is a good characterization of completely positive maps given by (3.3). No such simple representation seems possible for positive maps. For example, one may ask whether every positive map images/nec-82-4.png is of the form

images/eq-82-1.png

for some n × k matrices Vi, Wj. For n = k = 3, there exist positive maps Φ that can not be represented like this.

For these reasons the notion of complete positivity seems to be more useful than that of positivity.

We remark that many of the results of this section are true in the general setting of C-algebras. The proofs, naturally, are more intricate in the general setting.

In view of Theorem 3.1.6, one expects that if Φ is a positive linear map from a C-algebra a into a C-algebra b, and if either a or b is commutative, then Φ is completely positive. This is true.


3.2  EXERCISES

3.2.1

We have come across several positive linear maps in Chapter 2. Which of them are completely positive? What are (minimal) Stinespring dilations of these maps?


3.2.2

Every positive linear map Φ has a restricted 2-positive behaviour in the following sense:

images/eq-82-2.png

[Hint: Use Proposition 2.7.3 and Proposition 2.7.5.]


3.2.3

Let Φ be a strictly positive linear map. Then the following three conditions are equivalent:

(i)   Φ is 2-positive.

(ii)  If A, B are positive matrices and X any matrix such that BXA−1X, then Φ(B) ≥ Φ(X)Φ(A)−1Φ(X).

(iii)  For every matrix X and positive A we have Φ(XA−1X) ≥ Φ(X) Φ(A)−1 Φ(X).


[Compare this with Exercise 2.7.4 and Proposition 2.7.5.]


3.2.4

Let images/nec-83-1.png be the map defined as Φ(A) = 2 (tr A) IA. Then Φ is 2-positive but not 3-positive.


3.2.5

Let A and B be Hermitian matrices and suppose A = Φ(B) for some doubly stochastic map Φ on images/nec-83-2.png. Then there exists a completely positive doubly stochastic map Ψ such that A = Ψ(B). (See Exercise 2.7.13.)


3.2.6

Let images/nec-83-3.png be the collection of all 2×2 matrices A with a11 = a22. This is an operator system in images/nec-83-4.png. Show that the map Φ(A) = Atr is completely positive on images/nec-83-5.png. What is its completely positive extension on images/nec-83-6.png?


3.2.7

Suppose [[Aij]] is a positive element of images/nec-83-7.png. Then each of the m × m matrices images/nec-83-8.png, and images/nec-83-9.png is positive.


3.3  SCHWARZ INEQUALITIES

In this section we prove some operator versions of the Schwarz inequality. Some of them are extensions of the basic inequalities for positive linear maps proved in Chapter 2.

Let µ be a probability measure on a space X and consider the Hilbert space L2(X, µ). Let Ef = images/nec-84-1.png fdµ be the expectation of a function f. The covariance between two functions f and g in L2(X, µ) is the quantity

images/eq-84-1.png(3.9)

The variance of f is defined as

images/eq-84-2.png(3.10)

(We have come across this earlier in (2.23) where we restricted ourselves to real-valued functions.) The expression (3.9) is plainly an inner product in L2(X, µ) and the usual Schwarz inequality tells us

images/eq-84-3.png(3.11)

This is an important, much used, inequality in statistics.

As before, replace L2(X, µ) by images/nec-84-2.png and the expectation E by a positive unital linear map Φ on images/nec-84-3.png. The covariance between two elements A and B of images/nec-84-4.png (with respect to a given Φ) is defined as

images/eq-84-4.png(3.12)

and variance of A as

images/eq-84-5.png(3.13)

Kadison’s inequality (2.5) says that if A is Hermitian, then var(A) ≥ O. Choi’s generalization (2.6) says that this is true also when A is normal. However, with no restriction on A this is not always true. (Let Φ(A) = Atr, and let images/nec-84-5.png .)

If Φ is unital and 2-positive, then by Exercise 3.2.3(iii) we have

images/eq-84-6.png(3.14)

for all A. This says that var(A) ≥ O for all A if Φ is 2-positive and unital. The inequality (3.14) says that

images/eq-84-7.png(3.15)

The inequality |Φ(A)| ≤ Φ(|A|) is not always true even when Φ is completely positive. Let Φ be the pinching map on images/nec-84-6.png. If A = images/nec-84-7.png, then images/nec-84-8.png and images/nec-84-9.png.

An analogue of the variance-covariance inequality (3.11) is given by the following theorem.


3.3.1  Theorem

Let Φ be a unital completely positive linear map on images/nec-85-1.png. Then for all A, B

images/eq-85-1.png(3.16)

Proof.   Let V be an isometry of the space images/nec-85-2.png into any images/nec-85-3.png. Then V V = I and V V I. From the latter condition it follows that

images/eq-85-2.png

This is the same as saying

images/eq-85-3.png

This inequality is preserved when we multiply both sides by the matrix images/nec-85-4.png on the left and by images/nec-85-5.png on the right. Thus

images/eq-85-4.png

This is the inequality (3.16) for the special map Φ(T) = V TV. The general case follows from this using Theorem 3.1.2.     ■


3.3.2  Remark

It is natural to wonder whether complete positivity of Φ is necessary for the inequality (3.16). It turns out that 2-positivity is not enough but 3-positivity is. Indeed, if Φ is 3-positive and unital, then from the positivity of the matrix

images/eq-85-5.png

it follows that the matrix

images/eq-85-6.png

is positive. Hence by Theorem 1.3.3 (see Exercise 1.3.5)

images/eq-86-1.png

In other words,

images/eq-86-2.png(3.17)

This is the same inequality as (3.16).

To see that this inequality is not always true for 2-positive maps, choose the map Φ on images/nec-86-1.png as in Exercise 3.2.4. Let A = E13, and B = E12, where Eij stands for the matrix whose i, j entry is one and all other entries are zero. A calculation shows that the inequality (3.17) is not true in this case.


3.3.3  Remark

If Φ is 2-positive, then for all A and B we have

images/eq-86-3.png(3.18)

The inequality (3.17) is a considerable strengthening of this under the additional assumption that Φ is 3-positive and unital. The inequality (3.18) is equivalent to

images/eq-86-4.png(3.19)

(for 2-positive linear maps Φ). This is an operator version of the Schwarz inequality.


3.4  POSITIVE COMPLETIONS AND SCHUR PRODUCTS

A completion problem gives us a matrix some of whose entries are not specified, and asks us to fill in these entries in such a way that the matrix so obtained (called a completion) has a given property.

For example, we are given a 2 × 2 matrix images/nec-86-2.png with only three of its entries and are asked to choose the unknown (2,2) entry in such a way that the norm of the completed matrix is minimal among all completions. Such a completion is obtained by choosing the (2,2) entry to be −1. This is an example of a minimal norm completion problem.

A positive completion problem asks us to fill in the unspecified entries in such a way that the completed matrix is positive. Sometimes further restrictions may be placed on the completion. For example the incomplete matrix images/nec-87-1.png has several positive completions: we may choose any two diagonal entries a, b such that a, b are positive and ab ≥ 1. Among these the choice that minimises the norm of the completion is a = b = 1.

To facilitate further discussion, let us introduce some definitions.

A subset J of {1, 2, . . . , n} × {1, 2, . . . , n} is called a pattern. A pattern J is called symmetric if


(i, i) ∈ J for 1 ≤ in, and
(i, j) ∈ J if and only if (j, i) ∈ J.


We say T is a partially defined matrix with pattern J if the entries tij are specified for all (i, j) ∈ J. We call such a_ T symmetric if J is symmetric, tii is real for all 1 ≤ in, and images/nec-87-2.png for (i, j) ∈ J.

Given a pattern J, let

images/eq-87-1.png

This is a subspace of images/nec-87-3.png, and it is an operator system if the pattern J is symmetric.

For images/nec-87-4.png, we use the notation ST for the linear operator

images/eq-87-2.png

and sT for the linear functional

images/eq-87-3.png

3.4.1  Theorem

Let T be a partially defined symmetric matrix with pattern J. Then the following three conditions are equivalent:

(i)   T has a positive completion.

(ii)  The linear map images/nec-88-1.png is positive.

(iii)  The linear functional sT on images/nec-88-2.png is positive.


Proof.   If T has a positive completion images/nec-88-3.png, then by Schur’s theorem images/nec-88-4.png is a positive map on images/nec-88-5.png. For images/nec-88-6.png. So, ST is positive on images/nec-88-7.png . This proves the implication (i) images/nec-88-8.png (ii). The implication (ii) images/nec-88-9.png (iii) is obvious. (The sum of all entries of a positive matrix is a nonnegative number.)

(iii) images/nec-88-10.png (i): Suppose sT is positive. By Krein’s extension theorem there exists a positive linear functional s on images/nec-88-11.png that extends sT. Let images/nec-88-12.png. Then the matrix images/nec-88-13.png is a completion of T. We have for every vector x

images/eq-88-1.png

Thus images/nec-88-14.png is positive.     ■


For Timages/nec-88-15.png let T# be the element of images/nec-88-16.png defined as images/nec-88-17.png. We have seen that T is a contraction if and only if T# is positive.


3.4.2  Proposition

Let images/nec-88-18.png be the operator system in images/nec-88-19.png defined as

images/eq-88-2.png

Then for any Timages/nec-88-20.png, the Schur multiplier ST is contractive on images/nec-88-21.png if and only if ST# is a positive linear map on the operator system images/nec-88-22.png.

Proof.   Suppose ST# is positive on S. Then

images/eq-88-3.png

i.e., ||A|| ≤ 1 images/nec-88-23.png ||TA|| ≤ 1. In other words ST is contractive on images/nec-88-24.png.

To prove the converse, assume D1, D2 > O, and note that

images/eq-89-1.png

If ST is contractive on images/nec-89-1.png, then

images/eq-89-2.png

i.e.,

images/eq-89-3.png

We have seen above that the last matrix is congruent to

images/nec-89-2.png. This shows that ST# is positive on images/nec-89-3.png.     ■


We can prove now the main theorem of this section.


3.4.3  Theorem (Haagerup’s Theorem)

Let Timages/nec-89-4.png. Then the following four conditions are equivalent:

(i)   ST is contractive; i.e., ||TA|| ≤ ||A|| for all A.

(ii)  There exist vectors vj, wj, 1 ≤ jn, all with their norms ≤ 1, such that tij = vıwj.

(iii)  There exist positive matrices R1, R2 with diag R1I, diag R2I and such that images/nec-89-5.png is positive.

(iv)  T can be factored as T = V W with ||V ||c ≤ 1, ||W ||c ≤ 1. (The symbol ||Y ||c stands for the maximum of the Euclidean norms of the columns of Y.)


Proof.   Let ST be contractive. Then, by Proposition 3.4.2, ST# is a positive operator on the operator system images/nec-90-1.png. By Theorem 3.4.1, T# has a positive completion. (Think of the off-diagonal entries of the two diagonal blocks as unspecified.) Call this completion P. It has a Cholesky factoring P = ΔΔ where Δ is an upper triangular 2n × 2n matrix. Write images/nec-90-2.png. Then

images/eq-90-1.png

Let vj, wj, 1 ≤ jn be the columns of V, W, respectively. Since P is a completion of T#, we have T = V W; i.e., images/nec-90-3.png. Since diag(V V ) = I, we have ||vj|| = 1. Since diag(WW + XX) = I, we have ||wj|| ≤ 1. This proves the implication (i) images/nec-90-4.png (ii).

The condition (ii) can be expressed by saying T = V W, where diag(V V ) ≤ I and diag(WW) ≤ I. Since

images/eq-90-2.png

this shows that the statement (ii) implies (iii). Clearly (iv) is another way of stating (ii).

To complete the proof we show that (iii) images/nec-90-5.png (i). Let Aimages/nec-90-6.png, ||A|| ≤ 1. This implies images/nec-90-7.png. Then the condition (iii) leads to the inequality

images/eq-90-3.png

But this implies ||TA|| ≤ 1. In other words ST is contractive.     ■


3.4.4  Corollary

For every T in images/nec-90-8.png, we have ||ST || = min {||V ||c ||W||c : T = V *W} .


3.5  THE NUMERICAL RADIUS

The numerical range of an operator A is the set of complex numbers

images/eq-91-1.png

and the numerical radius is the number

images/eq-91-2.png

It is known that the set W(A) is convex, and w(·) defines a norm. We have

images/eq-91-3.png

Some properties of w are summarised below. It is not difficult to prove them.

(i)   w(UAU) = w(A) for all A, and unitary U.

(ii)  If A is diagonal, then w(A) = max |aii|.

(iii)  More generally,

images/eq-91-4.png

(iv)  w(A) = ||A|| if (but not only if) A is normal.

(v)  w is not submultiplicative: the inequality w(AB) ≤ w(A)w(B) is not always true for 2 × 2 matrices.

(vi)  Even the weaker inequality w(AB) ≤ ||A||w(B) is not always true for 2 × 2 matrices.

(vii)  The inequality images/nec-91-1.png is not always true for 2 × 2 matrices A, B.

(viii) However, we do have images/nec-91-2.png for square matrices A, B of any size.

(Proof: It is enough to prove this when ||A|| = 1. Then A = images/nec-91-3.png where U, V are unitary. So it is enough to prove that images/nec-91-4.png if U is unitary. Choose an orthonormal basis in which U is diagonal, and use (iii).)

(ix)  If w(A) ≤ 1, then I ± ReAO.
images/nec-92-1.png

(x)   The inequality w(AB) ≤ w(A)w(B) may not hold even when A, B commute. Let

images/eq-92-1.png

Then w(A) < 1, w(A2) = w(A3) = 1/2. So w(A3) > w(A)w(A2) in this case.


Proposition 1.3.1 characterizes operators A with ||A|| ≤ 1 in terms of positivity of certain 2 × 2 block matrices. A similar theorem for operators A with w(A) ≤ 1 is given below.


3.5.1  Theorem (Ando)

Let Aimages/nec-92-2.png. Then w(A) ≤ 1 if and only if there exists a Hermitian matrix H such that images/nec-92-3.png is positive.


Proof.   If images/nec-92-4.png, then there exists an operator K with ||K|| ≤ 1 such that A = (I + H)1/2K(IH)1/2. So, for every vector x

images/eq-92-2.png

This shows that w(A) ≤ 1.

The proof of the other half of the theorem is longer. Let A be an operator with w(A) ≤ 1. Let images/nec-92-5.png be the collection of 2 × 2 matrices images/nec-93-1.png where x, y, z are complex numbers. Then images/nec-93-2.png is an operator system. Let images/nec-93-3.png be the unital linear map defined as

images/eq-93-1.png

It follows from property (ix) listed at the beginning of the section that Φ is positive. We claim it is completely positive. Let m be any positive integer. We want to show that if the m× m block matrix with the 2 × 2 block images/nec-93-4.png as its i, j entry is positive, then the m × m block matrix with the n×n block xijI + 2 1 (yijA+zijA) as its i, j entry is also positive. Applying permutation similarity the first matrix can be converted to a matrix of the form images/nec-93-5.png where X, Y, Z are m× m matrices. If this is positive, then we have Z = Y , and our claim is that

images/eq-93-2.png

We can apply a congruence, and replace the matrices X by I and Y by X−1/2Y X−1/2, respectively. Thus we need to show that

images/eq-93-3.png

The hypothesis here is (equivalent to) ||Y || ≤ 1. By property (viii) this implies images/nec-93-6.png. So the conclusion follows from property (ix).

We have shown that Φ is completely positive on images/nec-93-7.png. By Arveson’s theorem Φ can be extended to a completely positive map images/nec-93-8.png images/nec-93-9.png.

Let Eij, 1 ≤ i, j ≤ 2 be the matrix units in images/nec-93-10.png. Then the matrix images/nec-93-11.png is positive. Thus, in particular, images/nec-93-12.png and images/nec-93-13.png are positive, and their sum is I since images/nec-93-14.png is unital.

Put images/nec-93-15.png. Then H is Hermitian, and

images/eq-93-4.png

Since images/nec-94-1.png is an extension of Φ, we have

images/eq-94-1.png

Thus

images/eq-94-2.png

and this matrix is positive.     ■


3.5.2  Corollary

For every A and k = 1, 2, . . .

images/eq-94-3.png(3.20)

Proof.   It is enough to show that if w(A) ≤ 1, then w(Ak) ≤ 1. Let w(A) ≤ 1. By Ando’s theorem, there exists a Hermitian matrix H such that

images/eq-94-4.png

Hence, there exists a contraction K such that

images/eq-94-5.png

Then

images/eq-94-6.png

where L = K[(IH2)1/2K]k−1 is a contraction. But this implies that

images/eq-94-7.png

So, by Ando’s Theorem w(Ak) ≤ 1.     ■

The inequality (3.20) is called the power inequality for the numerical radius.

Ando and Okubo have proved an analogue of Haagerup’s theorem for the norm of the Schur product with respect to the numerical radius. We state it without proof.


3.5.3  Theorem (Ando-Okubo)

Let T be any matrix. Then the following statements are equivalent:

(i)   w(TA) ≤ 1 whenever w(A) ≤ 1.

(ii)  There exists a positive matrix R with diagRI such that

images/eq-95-1.png

3.6  SUPPLEMENTARY RESULTS AND EXERCISES

The Schwarz inequality, in its various forms, is the most important inequality in analysis. The first few remarks in this section supplement the discussion in Section 3.3.

Let A be an n × k matrix and B an n × l matrix of rank l. The matrix

images/eq-95-2.png

is positive. This is equivalent to the assertion

images/eq-95-3.png(3.21)

This is a matrix version of the Schwarz inequality. It can be proved in another way as follows. The matrix B(BB)−1B is idempotent and Hermitian. Hence IB(BB)−1B and (3.21) follows immediately. The inequality (3.19) is an extension of (3.21).

Let A be a positive operator and let x, y be any two vectors. From the Schwarz inequality we get

images/eq-95-4.png(3.22)

An operator version of this in the spirit of (3.19) can be obtained as follows. For any two operators X and Y we have

images/eq-96-1.png

So, if Φ is a 2-positive linear map, then

images/eq-96-2.png

or, equivalently,

images/eq-96-3.png(3.23)

This is an operator version of (3.22).

There is a considerable strengthening of the inequality (3.22) in the special case when x is orthogonal to y. This says that if A is a positive operator with mIAMI, and images/nec-96-1.png, then

images/eq-96-4.png(3.24)

This is called Wielandt’s inequality. The following theorem gives an operator version.


3.6.1  Theorem

Let A be a positive element of images/nec-96-2.png with mIAMI. Let X, Y be two mutually orthogonal projection operators in images/nec-96-3.png. Then for every 2-positive linear map Φ on images/nec-96-4.png we have

images/eq-96-5.png(3.25)

Proof.   First assume that images/nec-96-5.png. With respect to this decomposition, let A have the block form

images/eq-96-6.png

By Exercise 1.5.7

images/eq-96-7.png

Apply Proposition 2.7.8 with Φ as the pinching map. This shows

images/eq-97-1.png

Taking inverses changes the direction of this inequality, and then rearranging terms we get

images/eq-97-2.png

This is the inequality (3.25) in the special case when Φ is the identity map. A minor argument shows that the assumption images/nec-97-1.png can be dropped.

Let α = (Mm)/(M + m). The inequality we have just proved is equivalent to the statement

images/eq-97-3.png

This implies that the inequality (3.25) holds for every 2-positive linear map Φ.     ■

We say that a complex function f on images/nec-97-2.png is in the Lieb class images/nec-97-3.png if f(A) ≥ 0 whenever AO, and |f(X)|2f(A)f(B) whenever images/nec-97-4.png. Several examples of such functions are given in MA (pages 268–270). We have come across several interesting 2 × 2 block matrices that are positive. Many Schwarz type inequalities for functions in the class L can be obtained from these block matrices.

The next few results concern maps associated with pinchings and their norms.

Let images/nec-97-5.png be the diagonal part of a matrix:

images/eq-97-4.png(3.26)

where images/nec-97-6.png is the orthogonal projection onto the one-dimensional space spanned by the vector ej. This is a special case of the pinching operation C introduced in Example 2.2.1 (vii). Since images/nec-97-7.png and PjO, we think of the sum (3.26) as a noncommutative convex combination. There is another interesting way of expressing for images/nec-97-8.png. Let ω = e2πi/n and let U be the diagonal unitary matrix

images/eq-98-1.png(3.27)

Then

images/eq-98-2.png(3.28)

(The sum on the right-hand side is the Schur product of A by a matrix whose i, j entry is

images/eq-98-3.png

This idea can be generalized.


3.6.2  Exercise

Partition n × n matrices into an r × r block form in which the diagonal blocks are square matrices of dimension d1, . . . , dr. Let C be the pinching operation sending the block matrix A = [[Aij]] to the block diagonal matrix C(A) = diag(A11, . . . , Arr). Let ω = e2πi/r and let V be the diagonal unitary matrix

images/eq-98-4.png

where Ij is the identity matrix of size dj. Show that

images/eq-98-5.png(3.29)

3.6.3  Exercise

Let J be a pattern and let images/nec-98-1.png be the map on images/nec-98-2.png induced by J as follows. The i, j entry of images/nec-98-3.png is aij for all (i, j) ∈ J and is zero otherwise. Suppose J is an equivalence relation on {1, 2, . . . , n} and has r equivalence classes. Show that

images/eq-99-1.png(3.30)

where W is a diagonal unitary matrix. Conversely, show that if images/nec-99-1.png can be represented as

images/eq-99-2.png(3.31)

where Uj are unitary matrices and λj are positive numbers with images/nec-99-2.png, then J is an equivalence relation with r equivalence classes. It is not possible to represent images/nec-99-3.png as a convex combination of unitary transforms as in (3.31) with fewer than r terms.


3.6.4  Exercise

Let V be the permutation matrix

images/eq-99-3.png(3.32)

Show that

images/eq-99-4.png(3.33)

Find n2 unitary matrices Wj such that

images/eq-99-5.png(3.34)

This gives a representation of the linear map images/nec-99-4.png from images/nec-99-5.png into scalar matrices.

It is of some interest to consider what is left of a matrix after the diagonal part is removed. Let

images/eq-100-1.png(3.35)

be the off-diagonal part of A. Using (3.28) we can write

images/eq-100-2.png

From this we get

images/eq-100-3.png(3.36)

This inequality is sharp. To see this choose images/nec-100-1.png, where E is the matrix all of whose entries are equal to one.


3.6.5  Exercise

Let B = EI. We have just seen that the Schur multiplier norm

images/eq-100-4.png(3.37)

Find an alternate proof of this using Theorem 3.4.3.


3.6.6  Exercise

Use Exercise 3.6.3 to show that

images/eq-100-5.png

This inequality can be improved:

(i)   Every matrix is unitarily similar to one with constant diagonal entries. [Prove this by induction, with the observation that images/nec-100-2.png for some unit vector x.]

(ii)  Thus, in some orthonormal basis, removing images/nec-100-3.png has the same effect as removing images/nec-100-4.png from A. Thus

images/eq-101-1.png(3.38)

and this inequality is sharp.


3.6.7  Exercise

The Schur multiplier norm is multiplicative over tensor products; i.e.,

images/eq-101-2.png

3.6.8  Exercise

Let images/nec-101-1.png. Show, using Theorem 3.4.3 and otherwise, that

images/eq-101-3.png

Let Δn be the triangular truncation operator taking every n×n matrix to its upper triangular part. Then we have images/nec-101-2.png. Try to find ||Δ3||.


3.6.9  Exercise

Fill in the details in the following proof of the power inequality (3.20).

(i)   If a is a complex number, then |a| ≤ 1 if and only if Re(1−za) ≥ 0 for all z with |z| < 1.

(ii)  w(A) ≤ 1 if and only if Re(IzA) ≥ O for |z| < 1.

(iii)  w(A) ≤ 1 if and only if Re((IzA)−1) ≥ O for |z| < 1.

(iv)  Let ω = e2πi/k. Prove the identity

images/eq-101-4.png

(v)  If w(A) ≤ 1, then

images/eq-102-1.png

(vi)  Assume w(A) ≤ 1. Use (v) and (iii) to conclude that w(Ak) ≤ 1.


By Exercise 3.2.7, if [[Aij]] is a positive element of images/nec-102-1.png, then the m × m matrices [[ tr Aij ]] and images/nec-102-2.png are positive. Matricial curiosity should make us wonder whether this remains true when tr is replaced by other matrix functions like det, and the norm || · ||2 is replaced by the norm || · ||.

For the sake of economy, in the following discussion we use (temporarily) the terms positive, m-positive, and completely positive to encompass nonlinear maps as well. Thus we say a map images/nec-102-3.png is positive if Φ(A) ≥ O whenever AO, and completely positive if [[Φ(Aij)]] is positive whenever a block matrix [[Aij]] is positive. For example, det(A) is a positive (nonlinear) function, and we have observed that images/nec-102-4.png is a completely positive (nonlinear) function. In Chapter 1 we noted that a function images/nec-102-5.png is completely positive if and only if it can be expressed in the form (1.40).


3.6.10  Proposition

Let φ(A) = ||A||2. Then φ is 2-positive but not 3-positive.


Proof.   The 2-positivity is an easy consequence of Proposition 1.3.2. The failure of φ to be 3-positive is illustrated by the following example in images/nec-102-6.png. Let

images/eq-102-2.png

Since X, Y and Z are positive, so is the matrix

images/eq-102-3.png

If we write A as [[Aij]] where Aij, 1 ≤ i, j ≤ 3 are 2 × 2 matrices, and replace each Aij by ||Aij||2 we obtain the matrix

images/eq-103-1.png

This matrix is not positive as its determinant is negative.     ■


3.6.11  Exercise

Let images/nec-103-1.png be the map defined as Φ(X) = |X|2 = XX. Use the example in Exercise 1.6.6 to show that Φ is not two-positive.


3.6.12  Exercise

Let images/nec-103-2.png be the k-fold tensor power of A. Let A = [[Aij]] be an element of images/nec-103-3.png. Then images/nec-103-4.png is a matrix of size (mn)k whereas images/nec-103-5.png is a matrix of size mnk. Show that the latter is a principal submatrix of the former. Use this observation to conclude that images/nec-103-6.png is a completely positive map from images/nec-103-7.png to images/nec-103-8.png.


3.6.13  Exercise

For 1 ≤ kn let ∧kA be the kth antisymmetric tensor power of an n × n matrix A. Show that ∧k is a completely positive map from images/nec-103-9.png into images/nec-103-10.png. If

images/eq-103-2.png

is the characteristic polynomial of A, then ck(A) = tr ∧kA. Hence each ck is a completely positive functional. In particular, det is completely positive.

Similar considerations apply to other “symmetry classes” of tensors and the associated “Schur functions.” Thus, for example, the permanent function is completely positive.


3.6.14  Exercise

Let images/nec-103-11.png be any 4-positive map. Let X, Y, Z be positive elements of images/nec-103-12.png and let

images/eq-103-3.png

Then A = [[Aij]] is positive. Let X = [I,I, I,I]. Consider the product X[[Φ(Aij)]]X and conclude that

images/eq-104-1.png(3.39)

Inequalities of the form (3.39) occur in other contexts. For example, if P, Q and R are (rectangular) matrices and the product P QR is defined, then the Frobenius inequality is the relation between ranks:

images/eq-104-2.png

The inequality (4.49) in Chapter 4 is another one with a similar structure.


3.7  NOTES AND REFERENCES

The theory of completely positive maps has been developed by operator algebraists and mathematical physicists over the last four decades.

The two major results of Section 3.1, the theorems of Stinespring and Arveson, hold in much more generality. We have given their baby versions by staying in finite dimensions.

Stinespring’s theorem was proved in W. F. Stinespring, Positive functions on C-algebras, Proc. Amer. Math. Soc., 6 (1955) 211–216. To put it in context, it is helpful to recall an earlier theorem due to M. A. Naimark.

Let images/nec-104-1.png be a compact Hausdorff space with its Borel σ-algebra images/nec-104-2.png, and let images/nec-104-3.png) be the collection of orthogonal projections in a Hilbert space images/nec-104-4.png. A projection-valued measure is a map images/nec-104-5.png from images/nec-104-6.png into images/nec-104-7.png that is countably additive: if {Si} is a countable collection of disjoint sets, then

images/eq-104-3.png

for all x and y in images/nec-104-8.png. The spectral theorem says that if A is a bounded self-adjoint operator on images/nec-104-9.png, then there exists a projection-valued measure on [−||A||, ||A||] taking values in images/nec-104-10.png, and with respect to this measure A can be written as the integral images/nec-104-11.png.

Instead of projection-valued measures we may consider an operator-valued measure. This assigns to each set S an element E(S) of images/nec-104-12.png, the map is countably additive, and sup {||E(S)|| : Simages/nec-104-13.png } <. Such a measure gives rise to a complex measure

images/eq-104-4.png(3.40)

for each pair x, y in images/nec-105-1.png. This in turn gives a bounded linear map Φ from the space C(X) into images/nec-105-2.png via

images/eq-105-1.png(3.41)

This process can be reversed. Given a bounded linear map Φ : images/nec-105-3.png we can construct complex measures µx,y via (3.40) and then an operator-valued measure E via (3.39). If E(S) is a positive operator for all S, we say the measure E is positive.

Naimark’s theorem says that every positive operator-valued measure can be dilated to a projection-valued measure. More precisely, if E is a positive images/nec-105-4.png-valued measure on images/nec-105-5.png, then there exist a Hilbert space images/nec-105-6.png, a bounded linear map images/nec-105-7.png, and a images/nec-105-8.png- valued measure P such that

images/eq-105-2.png

The point of the theorem is that by dilating to the space images/nec-105-9.png we have replaced the operator-valued measure E by the projection-valued measure P which is nicer in two senses: it is more familiar because of its connections with the spectral theorem and the associated map Φ is now a ∗-homomorphism of C(X).

The Stinespring theorem is a generalization of Naimark’s theorem in which the commutative algebra C(X) is replaced by a unital Calgebra. The theorem in its full generality says the following. If Φ is a completely positive map from a unital C-algebra a into images/nec-105-10.png, then there exist a Hilbert space images/nec-105-11.png, a unital ∗-homomorphism (i.e., a representation) images/nec-105-12.png, and a bounded linear operator V : images/nec-105-13.png with ||V ||2 = ||Φ(I)|| such that

images/eq-105-3.png

A “minimal” Stinespring dilation (in which images/nec-105-14.png is a smallest possible space) is unique up to unitary equivalence.

The term completely positive was introduced in this paper of Stinespring. The theory of positive and completely positive maps was vastly expanded in the hugely influential papers by W. B. Arveson, Subalgebras of C-algebras, I, II, Acta Math. 123 (1969) 141–224 and 128 (1972) 271–308. In the general version of Theorem 3.1.5 the space images/nec-105-15.png is replaced by an arbitrary C-algebra a, and images/nec-105-16.png is replaced by the space images/nec-105-17.png of bounded operators in a Hilbert space images/nec-105-18.png. This theorem is the Hahn-Banach theorem of noncommutative analysis.

Theorem 3.1.1 is Stinespring’s theorem restricted to algebras of matrices. It was proved by M.-D. Choi, Completely positive linear maps on complex matrices, Linear Algebra Appl., 10 (1975) 285–290, and by K. Kraus, General state changes in quantum theory, Ann. of Phys., 64 (1971) 311–335. It seems that the first paper has been well known to operator theorists and the second to physicists. The recent developments in quantum computation and quantum information theory have led to a renewed interest in these papers.

The book M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press, 2000, is a popular introduction to this topic. An older book from the physics literature is K. Kraus, States, Effects, and Operations: Fundamental Notions of Quantum Theory, Lecture Notes in Physics Vol. 190, Springer, 1983.

A positive matrix of trace one is called a density matrix in quantum mechanics. It is the noncommutative analogue of a probability distribution (a vector whose coordinates are nonnegative and add up to one). The requirement that density matrices are mapped to density matrices leads to the notion of a trace-preserving positive map. That this should happen also when the original system is tensored with another system (put in a larger system) leads to trace-preserving completely positive maps. Such maps are called quantum channels. Thus quantum channels are maps of the form (3.3) with the additional requirement images/nec-106-1.png. The operators Vj are called the noise operators, or errors of the channel.

The representation (3.3) is one reason for the wide use of completely positive maps. Attempts to obtain some good representation theorem for positive maps were not very successful. See E. Størmer, Positive linear maps of operator algebras, Acta Math., 110 (1963) 233–278, S. L. Woronowicz, Positive maps of low dimensional matrix algebras, Reports Math. Phys., 10 (1976) 165–183, and M.-D. Choi, Some assorted inequalities for positive linear maps on C-algebras, J. Operator Theory, 4 (1980) 271–285. Let us say that a positive linear map images/nec-106-2.png is decomposable if it can be written as

images/eq-106-1.png

If every positive linear map were decomposable it would follow that every real polynomial in n variables that takes only nonnegative values is a sum of squares of real polynomials. That the latter statement is false was shown by David Hilbert. The existence of a counterexample to the question on positive linear maps gives an easy proof of this result of Hilbert. See M.-D. Choi, Positive linear maps, cited in Chapter 2, for a discussion.

The results of Exercises 3.2.2 and 3.2.3 are due to Choi and are given in his 1980 paper cited above. The idea that positive maps have a restricted 2-positive behavior seems to have first appeared in T. Ando, Concavity of certain maps ..., Linear Algebra Appl., 26 (1979) 203–241. Examples of maps on images/nec-107-1.png that are (n − 1)-positive but not n-positive were given in M.-D. Choi, Positive linear maps on C-algebras, Canadian J. Math., 24 (1972) 520–529. The simplest examples are of the type given in Exercise 3.2.4 (with n and (n − 1) in place of 3 and 2, respectively).

The Schwarz inequality is one of the most important and useful inequalities in mathematics. It is natural to seek its extensions in all directions and to expect that they will be useful. The reader should see the book J. M. Steele, The Cauchy-Schwarz Master Class, Math. Association of America, 2004, for various facets of the Schwarz inequality. (Noncommutative or matrix versions are not included.) Section IX.5 of MA is devoted to certain Schwarz inequalities for matrices. The operator inequality (3.19) was first proved for special types of positive maps (including completely positive ones) by E. H. Lieb and M. B. Ruskai, Some operator inequalities of the Schwarz type, Adv. Math., 12 (1974) 269–273. That 2-positivity is an adequate assumption was noted by Choi in his 1980 paper. Theorem 3.3.1 was proved in R. Bhatia and C. Davis, More operator versions of the Schwarz inequality, Commun. Math. Phys., 215 (2000) 239–244. It was noted there (observation due to a referee) that 4-positivity of Φ is adequate to ensure the validity of (3.16). That 3-positivity suffices but 2-positivity does not was observed by R. Mathias, A note on: “More operator versions of the Schwarz inequality,” Positivity, 8 (2004) 85–87. The inequalities (3.23) and (3.25) are proved in the paper of Bhatia and Davis cited above, and in a slightly different form in S.-G. Wang and W.-C. Ip, A matrix version of the Wielandt inequality and its applications, Linear Algebra Appl., 296 (1999) 171–181.

Section 3.4 is based on material in the paper V. I. Paulsen, S. C. Power, and R. R. Smith, Schur products and matrix completions, J. Funct. Anal., 85 (1989) 151–178, and on Paulsen’s two books cited earlier. Theorem 3.4.3 is attributed to U. Haagerup, Decomposition of completely bounded maps on operator algebras, unpublished report. Calculating the exact value of the norm of a linear operator on a Hilbert space is generally a difficult problem. Calculating its norm as a Schur multiplier is even more difficult. Haagerup’s Theorem gives some methods for such calculations.

Completion problems of various kinds have been studied by several authors with diverse motivations coming from operator theory, electrical engineering, and optimization. A helpful introduction may be obtained from C. R. Johnson, Matrix completion problems: a survey, Proc. Symposia in Applied Math. Vol. 40, American Math. Soc., 1990.

Theorem 3.5.1 was proved by T. Ando, Structure of operators with numerical radius one, Acta Sci. Math. (Szeged), 34 (1973) 11–15. The proof given here is different from the original one, and is from T. Ando, Operator Theoretic Methods for Matrix Inequalities, Sapporo, 1998. Theorem 3.5.3 is proved in T. Ando and K. Okubo, Induced norms of the Schur multiplier operator, Linear Algebra Appl., 147 (1991) 181–199. This and Haagerup’s theorem are included in Ando’s 1998 report from which we have freely borrowed. A lot more information about inequalities for Schur products may be obtained from this report.

The inequality (3.20) is called Berger’s theorem. The lack of sub-multiplicativity and of its weaker substitutes has been a subject of much investigation in the theory of the numerical radius.

We have seen that even under the stringent assumption AB = BA we need not have w(AB) ≤ w(A)w(B). Even the weaker assertion w(AB) ≤ ||A||w(B) is not always true in this case. A 12 × 12 counterexample, in which w(AB) > (1.01)||A||w(B) was found by V. Müller, The numerical radius of a commuting product, Michigan Math. J., 35 (1988) 255–260. This was soon followed by K. R. Davidson and J.A.R. Holbrook, Numerical radii of zero-one matrices, ibid., 35 (1988) 261–267, who gave a simpler 9 × 9 example in which w(AB) > C||A||w(B) where C = 1/ cos(π/9) > 1.064. The reader will find in this paper a comprehensive discussion of the problem and its relation to other questions in dilation theory.

The formula (3.28) occurs in R. Bhatia, M.-D. Choi, and C. Davis, Comparing a matrix to its off-diagonal part, Oper. Theory: Adv. and Appl., 40 (1989) 151–164. The results of Exercises 3.6.2–3.6.6 are also taken from this paper. The ideas of this paper are taken further in R. Bhatia, Pinching, trimming, truncating and averaging of matrices, Am. Math. Monthly, 107 (2000) 602–608. Finding the exact norm of the operator Δn of Exercise 3.6.8 is hard. It is a well-known and important result of operator theory that for large n, the norm ||Δn|| is close to log n. See the paper by R. Bhatia (2000) cited above.

The operation of replacing the matrix entries Aij of a block matrix [[Aij]] by f(Aij) for various functions f has been studied by several linear algebraists. See, for example, J. De Pillis, Transformations on partitioned matrices, Duke Math. J., 36 (1969) 511–515, R. Merris, Trace functions I, ibid., 38 (1971) 527–530, and M. Marcus and W. Watkins, Partitioned Hermitian matrices, ibid., 38(1971) 237–249. Results of Exercises 3.6.11-3.6.13 are noted in this paper of Marcus and Watkins. Two foundational papers on this topic that develop a general theory are T. Ando and M.-D. Choi, Non-linear completely positive maps, in Aspects of Positivity in Functional Analysis, North-Holland Mathematical Studies Vol. 122, 1986, pp.3–13, and W. Arveson, Nonlinear states on C-algebras, in Operator Algebras and Mathematical Physics, Contemporary Mathematics Vol. 62, American Math. Society, 1987, pp. 283–343. Characterisations of nonlinear completely positive maps and Stinespring-type representation theorems are proved in these papers. These are substantial extensions of the representation (1.40). Exercise 3.6.14 is borrowed from the paper of Ando and Choi.

Finally, we mention that the theory of completely positive maps is now accompanied by the study of completely bounded maps, just as the study of positive measures is followed by that of bounded measures. The two books by Paulsen are an excellent introduction to the major themes of this subject. The books K. R. Parthasarathy, An Introduction to Quantum Stochastic Calculus, Birkhäuser, 1992, and P. A. Meyer, Quantum Probability for Probabilists, Lecture Notes in Mathematics Vol. 1538, Springer, 1993, are authoritative introductions to noncommutative probability, a subject in which completely positive maps play an important role.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.17.174.204