Chapter Five



Positive Definite Functions

Positive definite functions arise naturally in many areas of mathematics. In this chapter we study some of their basic properties, construct some examples, and use them to derive interesting results about positive matrices.


5.1  BASIC PROPERTIES

Positive definite sequences were introduced in Section 1.1.3. We repeat the definition. A (doubly infinite) sequence of complex numbers {an : nimages/nec-151-1.png} is said to be positive definite if for every positive integer N, we have

images/eq-151-1.png(5.1)

for every finite sequence of complex numbers ξ0, ξ1, . . . , ξN−1. This condition is equivalent to the requirement that for each N = 1, 2, . . . , the N × N matrix

images/eq-151-2.png(5.2)

is positive.

From this condition it is clear that we must have

images/eq-151-3.png(5.3)

A complex-valued function φ on images/nec-152-1.png is said to be positive definite if for every positive integer N, we have

images/eq-152-1.png(5.4)

for every choice of real numbers x0, x1, . . . , xN−1, and complex numbers ξ0, ξ1, . . . , ξN−1. In other words φ is positive definite if for each N = 1, 2, . . . the N × N matrix

images/eq-152-2.png(5.5)

is positive for every choice of real numbers x0, . . . , xN−1. It follows from this condition that

images/eq-152-3.png(5.6)

Thus every positive definite function is bounded, and its maximum absolute value is attained at 0.


5.1.1  Exercise

Let φ(x) be the characteristic function of the set images/nec-152-2.png; i.e., φ(x) = 1 if images/nec-152-3.png and φ(x) = 0 if images/nec-152-4.png. Show that φ is positive definite. This remains true when images/nec-152-5.png is replaced by any additive subgroup of images/nec-152-6.png.


5.1.2  Lemma

If φ is positive definite, then for all x1, x2

images/eq-152-4.png

Proof.   Assume, without loss of generality, that φ(0) = 1. Choose x0 = 0. The 3 × 3 matrix

images/eq-153-1.png

is positive. So for every vector u, the inner product images/nec-153-1.png. Choose u = (z, 1, −1) where z is any complex number. This gives the inequality

images/eq-153-2.png

Now choose images/nec-153-2.png. This gives

images/eq-153-3.png

Exercise 5.1.1 showed us that a positive definite function φ need not be continuous. Lemma 5.1.2 shows that if the real part of φ is continuous at 0, then φ is continuous everywhere on images/nec-153-3.png.


5.1.3  Exercise

Let φ(x) be positive definite. Then

(i)  images/nec-153-4.png is positive definite.

(ii)  For every real number t the function φ(tx) is positive definite.


5.1.4  Exercise

(i)   If φ1, φ2 are positive definite, then so is their product φ1φ2. (Schur’s theorem.)

(ii)  If φ is positive definite, then |φ|2 is positive definite. So is Re φ.


5.1.5  Exercise

(i)   If φ1, . . . , φn are positive definite, and a1, . . . , an are positive real numbers, then a1φ1 + · · · + anφn is positive definite.

(ii)  If {φn} is a sequence of positive definite functions and φn(x) → φ(x) for all x, then φ is positive definite.

(iii)  If φ is positive definite, then eϕ is positive definite, and so is eϕ+a for every images/nec-154-1.png.

(iv) If φ(x) is a measurable positive definite function and f(t) is a nonnegative integrable function, then images/nec-154-2.png is positive definite.

(v)  If µ is a finite positive Borel measure on images/nec-154-3.png and φ(x) a measurable positive definite function, then the function images/nec-154-4.png is positive definite. (The statement (iv) is a special case of (v).)


Let I be any interval and let K(x, y) be a bounded continuous complex-valued function on I × I. We say K is a positive definite kernel if

images/eq-154-1.png(5.7)

for every continuous integrable function f on the interval I.


5.1.6  Exercise

(i)   A bounded continuous function K(x, y) on I × I is a positive definite kernel if and only if for all choices of points x1, . . . , xN in I, the N × N matrix [[K(xi, xj)]] is positive.

(ii)  A bounded continuous function φ on images/nec-154-5.png is positive definite if and only if the kernel K(x, y) = φ(xy) is positive definite.


5.2  EXAMPLES

5.2.1

The function φ(x) = eix is positive definite since

images/eq-154-2.png

Exercise: Write the matrix (5.5) in this case as uu where u is a column vector.


This example is fundamental. It is a remarkable fact that all positive definite functions can be built from this one function by procedures outlined in Section 5.1.


5.2.2

The function φ(x) = cos x is positive definite.


Exercise: sin x is not positive definite. The matrix A in Exercise 1.6.6 has entries aij = cos(xixj) where xi = 0, π/4, π/2, 3π/4. It follows that | cos x| is not positive definite.


5.2.3

For each images/nec-155-1.png, φ(x) = eitx is a positive definite function.


5.2.4

Let images/nec-155-2.png and let f(t) ≥ 0. Then

images/eq-155-1.png(5.8)

is positive definite. More generally, if µ is a positive finite Borel measure on images/nec-155-3.png, then

images/eq-155-2.png(5.9)

is positive definite. The function images/nec-155-4.png is called the Fourier transform of f and images/nec-155-5.png is called the Fourier-Stieltjes transform of µ. Both of them are bounded uniformly continuous functions.


These transforms give us many interesting positive definite functions.


5.2.5

One of the first calculations in probability theory is the computation of an integral:

images/eq-156-1.png

The integral on the right hand side can be evaluated using Cauchy’s theorem. Let C be the rectangular contour with vertices −R, R, R + ix,R + ix. The integral of the analytic function f(z) = ez2/2 along this contour is zero. As R → ∞, the integral along the two vertical sides of this contour goes to zero. Hence

images/eq-156-2.png

So,

images/eq-156-3.png(5.10)

(This shows that, with a suitable normalization, the function ex2/2 is its own Fourier transform.) Thus for each a ≥ 0, the function φ(x) = eax2 is positive definite.


5.2.6

The function φ(x) = sin x/x is positive definite. To see this one can use the product formula

images/eq-156-4.png(5.11)

and observe that each of the factors in this infinite product is a positive definite function. Alternately, we can use the formula.

images/eq-156-5.png(5.12)

(This integral is the Fourier transform of the characteristic function images/nec-157-1.png.)

We have tacitly assumed here that φ(0) = 1. This is natural. If we assign φ(0) any value larger than 1, the resulting (discontinuous) function is also positive definite.


5.2.7

The integral

images/eq-157-1.png

shows that the function φ(x) = 1/(1 + ix) is positive definite. The functions 1/(1 − ix) and 1/(1 + x2) are positive definite.


5.2.8

The integral formulas

images/eq-157-2.png

and

images/eq-157-3.png

show that the functions 1/(1 + x2) and e−|x| are positive definite. (They are nonnegative and are, up to a constant factor, Fourier transforms of each other.)


5.2.9

From the product representations

images/eq-157-4.png(5.13)

and

images/eq-158-1.png(5.14)

one sees (using the fact that 1/(1 + a2x2) is positive definite) that the functions x/(sinh x) and 1/(cosh x) are positive definite. (Contrast this with 5.2.6 and 5.2.2.)


5.2.10

For 0 < α < 1, we have from (5.13)

images/eq-158-2.png(5.15)

Each factor in this product is of the form

images/eq-158-3.png

This shows that the function sinh αx/sinh x is positive definite for 0 ≤ α ≤ 1. In the same way using (5.14) one can see that cosh αx/cosh x is positive definite for −1 ≤ α ≤ 1. The function

images/eq-158-4.png

is positive definite for −1/2 ≤ α ≤ 1/2, as it is the product of two such functions.


5.2.11

The integral

images/eq-158-5.png

shows that tanh x/x is a positive definite function.

(Once again, it is natural to assign the functions sinh x/x, sinh αx/x and tanh x/x the values 1, α and 1, respectively, at x = 0. Then the functions under consideration are continuous and positive definite. Assigning them larger values at 0 destroys continuity but not positive definiteness.)


5.2.12

One more way of constructing positive definite functions is by convolutions of functions in images/nec-159-1.png. For any function f let images/nec-159-2.png be the function defined as images/nec-159-3.png. If images/nec-159-4.png then the function images/nec-159-5.png defined as

images/eq-159-1.png

is a continuous function vanishing at ∞. It is a positive definite function since

images/eq-159-2.png

5.2.13

Let R be a positive real number. The tent function (with base R) is defined as

images/eq-159-3.png(5.16)

A calculation shows that

images/eq-160-1.png

So, ∆R(x) is positive definite for all R > 0.


5.2.14

For R > 0, let δR be the continuous function defined as

images/eq-160-2.png(5.17)

From 5.2.6 it follows that δR is positive definite. The family {δR}R>0 is called the Fejér kernel on R. It has the following properties (required of any “summability kernel” in Fourier analysis):

(i)   δR(x) ≥ 0 for all x, and for all R > 0.

(ii)  For every a > 0, δR(x) → 0 uniformly outside [−a, a] as R → ∞.

(iii)  images/nec-160-1.png for every a > 0.

(iv)  images/nec-160-2.png for all R > 0.

Property (iv) may be proved by contour integration, for example.

The functions ∆R and δR are Fourier transforms of each other (up to constant factors). So the positive definiteness of one follows from the nonnegativity of the other.


5.2.15

In this section we consider functions like the tent functions of Section 5.2.13.

a  Let φ be any continuous, nonnegative, even function. Suppose φ(x) = 0 for |x| ≥ R, and φ is convex and monotonically decreasing in the interval [0, R). Then φ is a uniform limit of sums of the form images/nec-160-3.png , where ak ≥ 0 and RkR. It follows from 5.2.13 that φ is positive definite.

b  The condition that φ is supported in [−R, R] can be dropped. Let φ be any continuous, nonnegative, even function that is convex and monotonically decreasing in [0, ∞). Let images/nec-161-1.png. Then φ is a uniform limit of sums of the form images/nec-161-2.png , where ak ≥ 0 and Rk > 0. Hence φ is positive definite. This is Pólya’s Theorem.

c  Let φ be any function satisfying the conditions in Part a of this section, and extend it to all of images/nec-161-3.png as a periodic function with period 2R. Since φ is even, the Fourier expansion of φ does not contain any sin terms. It can be seen from the convexity of φ in (0, R) that the coefficients an in the Fourier expansion

images/eq-161-1.png

are nonnegative. Hence φ is positive definite.


5.2.16

Using 5.2.15 one can see that the following functions are positive definite:

images/eq-161-2.png

The special case a = 1 of (iii) was seen in 5.2.8. The next theorem provides a further extension.


5.2.17  Theorem

The function φ(x) = exp(−|x|a) is positive definite for 0 ≤ a ≤ 2.


Proof.   Let 0 < a < 2. A calculation shows that

images/eq-161-3.png

where

images/eq-162-1.png

(The assumption a < 2 is needed to ensure that this last integral is convergent. At 0 the numerator in the integrand has a zero of order 2.) Thus we have

images/eq-162-2.png

where (t) = Ca dt. Let φn be defined as

images/eq-162-3.png

The second integral in the last line is a number, say cn, while the first is a function, say gn(x). This function is positive definite since cos xt is positive definite for all t. So, for each n, the function exp φn(x) is positive definite by Exercise 5.1.5 (iii). Since limn→∞ exp φn(x) = exp(−|x|a), this function too is positive definite for 0 < a < 2. Again, by continuity, this is true for a = 2 as well.     ■

For a > 2 the function exp(−|x|a) is not positive definite. This is shown in Exercise 5.5.8.


5.2.18

The equality

images/eq-162-4.png

shows that the matrix on the left-hand side is positive. Thus the n × n matrix A with entries aij = min(i, j) is positive. This can be used to see that the kernel K(x, y) = min(x, y) is positive definite on [0, ∞) × [0, ∞).


5.2.19  Exercise

Let B be the n × n matrix with entries bij = 1/max(i, j). Show that this matrix is positive by an argument similar to the one in 5.2.18.

Note that if A is the matrix in 5.2.18, then B = DAD, where D = diag(1, 1/2, . . . , 1/n).


5.2.20  Exercise

Let λ1, . . . , λn be positive real numbers. Let A and B be the n × n matrices whose entries are aij = min(λi, λj) and bij = 1/ max(λi, λj), respectively. Show that A and B are positive definite.


5.2.21  Exercise

Show that the matrices A and B defined in Exercise 5.2.20 are infinitely divisible.


5.2.22  Exercise

Let 0 < λ1λ2 ≤ · · · ≤ λn and let A be the symmetric matrix whose entries aij are defined as aij = λij for 1 ≤ ijn. Show that A is infinitely divisible.


5.2.23  Exercise

Let λ1, λ2, . . . , λn be real numbers. Show that the matrix A whose entries are

images/eq-163-1.png

is infinitely divisible. [Hint: Use Pólya’s Theorem.]


5.3  LOEWNER MATRICES

In this section we resume, and expand upon, our discussion of operator monotone functions. Recall some of the notions introduced at the end of Chapter 2. Let C1(I) be the space of continuously differentiable real-valued functions on an open interval I. The first divided difference of a function f in C1(I) is the function f[1] defined on I × I as

images/eq-164-1.png

Let images/nec-164-1.png be the collection of all n × n Hermitian matrices whose eigenvalues are in I. This is an open subset in the real vector space images/nec-164-2.png consisting of all Hermitian matrices. The function f induces a map from images/nec-164-3.png into images/nec-164-4.png.

If f ∈ C1(I) and images/nec-164-5.png we define f[1](A) as the matrix whose i, j entry is f[1](λi, λj), where λ1, . . . , λn are the eigenvalues of A. This is called the Loewner matrix of f at A.

The function f on images/nec-164-6.png is differentiable. Its derivative at A, denoted as Df(A), is a linear map on images/nec-164-7.png characterized by the condition

images/eq-164-2.png(5.18)

for all images/nec-164-8.png. We have

images/eq-164-3.png(5.19)

An interesting expression for this derivative in terms of Loewner matrices is given in the following theorem.


5.3.1  Theorem

Let f ∈ C1(I) and images/nec-164-9.png. Then

images/eq-164-4.png(5.20)

wheredenotes the Schur product in a basis in which A is diagonal.


One proof of this theorem can be found in MA (Theorem V.3.3). Here we give another proof based on different ideas.

Let [X, Y ] stand for the Lie bracket: [X, Y ] = XYY X. If X is Hermitian and Y skew-Hermitian, then [X, Y ] is Hermitian.


5.3.2  Theorem

Let f ∈ C1(I) and images/nec-165-1.png. Then for every skew-Hermitian matrix K

images/eq-165-1.png(5.21)

Proof.   The exponential etK is a unitary matrix for all images/nec-165-2.png. From the series representation of etK one can see that

images/eq-165-2.png

Since f is in the class C1, this is equal to

images/eq-165-3.png

For each images/nec-165-3.png, the collection

images/eq-165-4.png

is a linear subspace of images/nec-165-4.png. On images/nec-165-5.png we have an inner product images/nec-165-6.png = tr XY . With respect to this inner product, the orthogonal complement of CA is the space

images/eq-165-5.png

(It is easy to prove this. If H commutes with A, then

images/eq-166-1.png

Proof of Theorem 5.3.1.    Choose an orthonormal basis in which A = diag(λ1, . . . , λn). Let images/nec-166-1.png; i.e., H = [A, K] for some skew-Hermitian matrix K. By (5.21), Df(A)(H) = [f(A), K]. The entries of this matrix are

images/eq-166-2.png

These are the entries of f[1](A) ◦ H also. Thus the two sides of (5.20) are equal when images/nec-166-2.png. Now let H belong to the complementary space images/nec-166-3.png. The theorem will be proved if we show that the equality (5.20) holds in this case too. But this is easy. Since H commutes with A, we may assume H too is diagonal, H = diag(h1, . . . , hn). In this case the two sides of (5.20) are equal to the diagonal matrix with entries f(λi)hi on the diagonal.     ■


The next theorem says that f is operator monotone on I if and only if for all n and for all images/nec-166-4.png the Loewner matrices f[1](A) are positive. (This is a striking analogue of the statement that a real function f is monotonically increasing if and only if f(t) ≥ 0.)


5.3.3  Theorem

Let f ∈ C1(I). Then f is operator monotone on I if and only if f[1](A) is positive for every Hermitian matrix A whose eigenvalues are contained in I.


Proof.   Suppose f is operator monotone. Let images/nec-166-5.png and let H be the positive matrix with all its entries equal to 1. For small positive t, A + tH is in images/nec-166-6.png. We have A + tHA, and hence f(A + tH) ≥ f(A). This implies Df(A)(H) ≥ O. For this H, the right-hand side of (5.20) is just f[1](A), and we have shown this is positive.

To prove the converse, let A0, A1 be matrices in images/nec-167-1.png with A1A0. Let A(t) = (1 − t)A0 + tA1, 0 ≤ t ≤ 1. Then A(t) is in images/nec-167-2.png. Our hypothesis says that f[1](A(t)) is positive. The derivative A(t) = A1A0 is positive, and hence the Schur product f[1](A(t)) ◦ A(t) is positive. By Theorem 5.3.1 this product is equal to Df(A(t))(A(t)). Since

images/eq-167-1.png

and the integrand is positive for all t, we have f(A1) ≥ f(A0).     ■


We have seen some examples of operator monotone functions in Section 4.2. Theorem 5.3.3 provides a direct way of proving operator monotonicity of these and other functions. The positivity of the Loewner matrices f[1](A) is proved by associating with them some positive definite functions. Some examples follow.



5.3.4

The function

images/eq-167-2.png

is operator monotone on any interval I that does not contain the point −d/c.

To see this write down the Loewner matrix f[1](A) for any Aimages/nec-167-3.png. If λ1, . . . , λn are the eigenvalues of A, this Loewner matrix has entries

images/eq-167-3.png

This matrix is congruent to the matrix with all entries 1, and is therefore positive.


5.3.5

The function f(t) = tr is operator monotone on (0, ∞) for 0 ≤ r ≤ 1.

A Loewner matrix for this function is a matrix V with entries

images/eq-168-1.png

The numbers λi are positive and can, therefore, be written as exi for some xi. We have then

images/eq-168-2.png

This matrix is congruent to the matrix with entries

images/eq-168-3.png

Since sinh rx/(sinh x) is a positive definite function for 0 ≤ r ≤ 1 (see 5.2.10), this matrix is positive.


5.3.6  Exercise

The function f(t) = tr is not operator monotone on (0, ∞) for any real number r outside [0, 1].


5.3.7

The function f(t) = log t is operator monotone on (0, ∞).

A Loewner matrix in this case has entries

images/eq-169-1.png

The substitution λi = exi reduces this to

images/eq-169-2.png

This matrix is positive since the function x/(sinh x) is positive definite. (See 5.2.9.)

Another proof of this is obtained from the equality

images/eq-169-3.png

For each t the matrix [[1/(λi + t)(λj + t)]] is positive. (One more proof of operator monotonicity of the log function was given in Exercise 4.2.5.)


5.3.8

The function f(t) = tan t is operator monotone on images/nec-169-1.png.

In this case a Loewner matrix has entries

images/eq-169-4.png

This matrix is positive since the function sin x/x is positive definite. (See 5.2.6.)


5.3.9  Exercise

For 0 ≤ r ≤ 1 let f be the map f(A) = Ar on the space of positive definite matrices. Show that

images/eq-169-5.png(5.22)

5.4  NORM INEQUALITIES FOR MEANS

The theme of this section is inequalities for norms of some expressions involving positive matrices. In the case of numbers they reduce to some of the most fundamental inequalities of analysis.

As a prototype consider the arithmetic-geometric mean inequality images/nec-170-1.png for positive numbers a, b. There are many different directions in which one could look for a generalization of this to positive matrices A, B. One version that involves the somewhat subtle concept of a matrix geometric mean is given in Section 4.1. Instead of matrices we could compare numbers associated with them. Thus, for example, we may ask whether

images/eq-170-1.png(5.23)

for every unitarily invariant norm. This is indeed true. There is a more general version of this inequality that is easier to prove: we have

images/eq-170-2.png(5.24)

for every X. What makes it easier is a lovely trick. It is enough to prove (5.24) in the special case A = B. (The inequality (5.23) is a vacuous statement in this case.) Suppose we have proved

images/eq-170-3.png(5.25)

for all matrices X and positive A. Then given X and positive A, B we may replace A and X in (5.25) by the 2 × 2 block matrices images/nec-170-2.png and images/nec-170-3.png. This gives the inequality (5.24).

Since the norms involved are unitarily invariant we may assume that A is diagonal, A = diag(λ1, . . . , λn). Then we have

images/eq-170-4.png(5.26)

where Y is the matrix with entries

images/eq-171-1.png(5.27)

This matrix is congruent to the Cauchy matrix—the one whose entries are 1/(λi + λj). Since that matrix is positive (Exercise 1.1.2) so is Y . All the diagonal entries of Y are equal to 1. So, using Exercise 2.7.12 we get the inequality (5.25) from (5.26).

The inequalities that follow are proved using similar arguments. Matrices that occur in the place of (5.27) are more complicated and their positivity is not as easy to establish. But in Section 5.2 we have done most of the work that is needed.


5.4.1  Theorem

Let A, B be positive and let X be any matrix. Then for 0 ≤ ν ≤ 1 we have

images/eq-171-2.png(5.28)

Proof.   Follow the arguments used above in proving (5.24). To prove the second inequality in (5.28) we need to prove that the matrix Y whose entries are

images/eq-171-3.png(5.29)

is positive for 0 < ν < 1. (When ν = 1/2 this reduces to (5.27).) Writing

images/eq-171-4.png

we see that Y is congruent to the matrix Z whose entries are

images/eq-172-1.png

This matrix is like the one in 5.3.5. The argument used there reduces the question of positivity of Z to that of positive definiteness of the function cosh αx/(cosh x) for −1 < α < 1. In 5.2.10 we have seen that this function is indeed positive definite. The proof of the first inequality in (5.28) is very similar to this, and is left to the reader.     ■


5.4.2  Exercise

Show that for 0 ≤ ν ≤ 1

images/eq-172-2.png(5.30)

5.4.3  Exercise

For the Hilbert-Schmidt norm we have

images/eq-172-3.png(5.31)

for positive matrices A and 0 < ν < 1. This is not always true for the operator norm || · ||.


5.4.4  Exercise

For any matrix Z let

images/eq-172-4.png

Let A be a positive matrix and let X be a Hermitian matrix. Let S = AνXA1−ν, T = νAX + (1 − ν)XA. Show that for 0 ≤ ν ≤ 1

images/eq-172-5.png

In Chapter 4 we defined the logarithmic mean of a and b. This is the quantity

images/eq-173-1.png(5.32)

A proof of the inequality (4.3) using the ideas of Section 5.3 is given below.


5.4.5  Lemma

images/eq-173-2.png(5.33)

Proof.   Put a = ex and b = ey. A small calculation reduces the job of proving the first inequality in (5.33) to showing that t ≤ sinh t for t > 0, and the second to showing that tanh tt for all t > 0. Both these inequalities can be proved very easily.     ■


5.4.6  Exercise

Show that for A, B positive and for every X

images/eq-173-3.png(5.34)

This matrix version of the arithmetic-logarithmic-geometric mean inequality can be generalized to all unitarily invariant norms.


5.4.7  Theorem

For every unitarily invariant norm we have

images/eq-173-4.png(5.35)

Proof.   The idea of the proof is very similar to that of Theorem 5.4.1. Assume B = A, and suppose A is diagonal with entries λ1, . . . , λn on the diagonal. The matrix A1/2XA1/2 is obtained from images/nec-174-1.png by entrywise multiplication with the matrix Y whose entries are

images/eq-174-1.png

This matrix is congruent to one with entries

images/eq-174-2.png

We have seen in 5.3.7 that this matrix is positive. That proves the first inequality in (5.35).

The matrix images/nec-174-2.png is the Schur product of images/nec-174-3.png with the matrix W whose entries are

images/eq-174-3.png

Making the substitution λi = exi, we have

images/eq-174-4.png

This matrix is positive since the function tanh x/x is positive definite. (See 5.2.11.) That proves the second inequality in (5.35).     ■


5.4.8  Exercise

A refinement of the inequalities (5.28) and (5.35) is provided by the assertion

images/eq-174-5.png

for 1/4 ≤ ν ≤ 3/4. Prove this using the fact that (x cosh αx)/ sinh x is a positive definite function for −1/2 ≤ α ≤ 1/2. (See 5.2.10.)


5.4.9  Exercise

Let H, K be Hermitian, and let X be any matrix. Show that

images/eq-175-1.png

This is a matrix version of the inequality |sin x| ≤ |x|.


5.4.10  Exercise

Let H, K and X be as above. Show that

images/eq-175-2.png

5.4.11  Exercise

Let A, B be positive matrices. Show that

images/eq-175-3.png

Hence, if H, K are Hermitian, then

images/eq-175-4.png

for every matrix X.


5.5  THEOREMS OF HERGLOTZ AND BOCHNER

These two theorems give complete characterizations of positive definite sequences and positive definite functions, respectively. They have important applications throughout analysis. For the sake of completeness we include proofs of these theorems here. Some basic facts from functional analysis and Fourier analysis are needed for the proofs. The reader is briefly reminded of these facts.

Let M[0, 1] be the space of complex finite Borel measures on the interval [0, 1]. This is equipped with a norm images/nec-176-1.png, and is the Banach space dual of the space C[0, 1]. If images/nec-176-2.png converges to images/nec-176-3.png for every f ∈ C[0, 1], we say that the sequence {µn} in M[0, 1] converges to µ in the weak topology.

A basic fact about this convergence is the following theorem called Helly’s Selection Principle.


5.5.1  Theorem

Let {µn} be a sequence of probability measures on [0, 1]. Then there exists a probability measure µ and a subsequence {µm} of {µn} such that µm converges in the weak topology to µ.


Proof.   The space C[0, 1] is a separable Banach space. Choose a sequence {fj} in C[0, 1] that includes the function 1 and whose linear combinations are dense in C[0, 1]. Since ||µn|| = 1, for each j we have images/nec-176-4.png for all n. Thus for each jimages/nec-176-5.png is a bounded sequence of positive numbers. By the diagonal procedure, we can extract a subsequence {µm} such that for each j, the sequence images/nec-176-6.png converges to a limit, say ξj, as m → ∞.

If images/nec-176-7.png is any (finite) linear combination of the fj, let

images/eq-176-1.png

This is a linear functional on the linear span of {fj}, and |Λ0(f)| ≤ ||f|| for every f in this span. By continuity Λ0 has an extension Λ to C[0, 1] that satisfies |Λ(f)| ≤ ||f|| for all f in C[0, 1]. This linear functional Λ is positive and unital. So, by the Riesz Representation Theorem, there exists a probability measure µ on [0, 1] such that images/nec-176-8.png for all f ∈ C[0, 1].

Finally, we know that images/nec-176-9.png converges to images/nec-176-10.png for every f in the span of {fj}. Since such f are dense and the µm are uniformly bounded, this convergence persists for every f in C[0, 1].     ■


Theorem 5.5.1 is also a corollary of the Banach Alaoglu theorem. This says that the closed unit ball in the dual space of a Banach space is compact in the weak topology. If a Banach space X is separable, then the weak topology on the closed unit ball of its dual X is metrizable.


5.5.2  Herglotz’ Theorem

Let images/nec-177-1.png be a positive definite sequence and suppose a0 = 1. Then there exists a probability measure µ on [−π, π] such that

images/eq-177-1.png(5.36)

Proof.   The positive definiteness of {an} implies that for every real x we have

images/eq-177-2.png

This inequality can be expressed in another form

images/eq-177-3.png

Let fN(x) be the function given by the last sum. Then

images/eq-177-4.png

For any Borel set E in [−π, π], let

images/eq-177-5.png

Then µN is a probability measure on [−π, π]. Apply Helly’s selection principle to the sequence {µN }. There exists a probability measure µ to which (a subsequence of ) µN converges in the weak topology. Thus for every n

images/eq-177-6.png
images/eq-178-1.png

We remark that the sum

images/eq-178-2.png

is called the Fejér kernel and is much used in the study of Fourier series.

The condition a0 = 1 in the statement of Herglotz’ theorem is an inessential normalization. This can be dropped; then µ is a finite positive measure with ||µ|| = a0.

Bochner’s theorem, in the same spirit as Herglotz’, says that every continuous positive definite function on images/nec-178-1.png is the Fourier-Stieltjes transform of a finite positive measure on R. The proof needs some approximation arguments. For the convenience of the reader let us recall some basic facts.

For images/nec-178-2.png we write images/nec-178-3.png for its Fourier transform defined as

images/eq-178-3.png

This function is in images/nec-178-4.png, the class of continuous functions vanishing at ∞. We write

images/eq-178-4.png

for the inverse Fourier transform of f. If the function images/nec-178-5.png is in images/nec-178-6.png (and this is not always the case) then images/nec-178-7.png

The Fourier transform on the space images/nec-179-1.png is defined as follows. Let images/nec-179-2.png. Then images/nec-179-3.png is defined as above. One can see that images/nec-179-4.png and the map images/nec-179-5.png is an L2-isometry on the space images/nec-179-6.png. This space is dense in images/nec-179-7.png. So the isometry defined on it has a unique extension to all of images/nec-179-8.png. This unitary operator on images/nec-179-9.png is denoted again by images/nec-179-10.png. The inverse of the map images/nec-179-11.png is defined by inverting this unitary operator. The fact that the Fourier transform is a bijective map of images/nec-179-12.png onto itself makes some operations in this space simpler.

Let δR be the function defined in 5.2.14. The family {δN} is an approximate identity: as N → ∞, the convolution δNg converges to g in an appropriate sense. The “appropriate sense” for us is the following.

If g is either an element of images/nec-179-13.png, or a bounded measurable function, then

images/eq-179-1.png(5.37)

In the discussion that follows we ignore constant factors involving 2π. These do not affect our conclusions in any way.

The Fourier transform “converts convolution into multiplication;” i.e.,

images/eq-179-2.png

The Riesz representation theorem and Helly’s selection principle have generalizations to the real line. The space images/nec-179-14.png is a separable Banach space. Its dual is the space images/nec-179-15.png of finite Borel measures on images/nec-179-16.png with norm images/nec-179-17.png. Every bounded sequence {µn} in images/nec-179-18.png has a weak convergent subsequence {µm}; i.e., for every images/nec-179-19.png converges to images/nec-179-20.png. This too is a special case of the Banach-Alaoglu theorem.


5.5.3  Bochner’s Theorem

Let φ be any function on the real line that is positive definite and continuous at 0. Then there exists a finite positive measure µ such that

images/eq-180-1.png(5.38)

Proof.   By Lemma 5.1.2, φ is continuous everywhere. Suppose in addition that images/nec-180-1.png. Using (5.6) we see that

images/eq-180-2.png(0)

Thus φ is in the space images/nec-180-2.png also. Hence, there exists images/nec-180-3.png such that

images/eq-180-3.png(5.39)

Let ∆N(x) be the tent function defined in (5.16). Then

images/eq-180-4.png(5.40)

This integral (of a continuous function over a bounded interval) is a limit of Riemann sums. Let xj = j N/K,KjK. The last integral is the limit, as K → ∞, of sums

images/eq-180-5.png

These sums can be expressed in another way:

images/eq-180-6.png(5.41)

where c(K, N) is a positive number. (See the proof of Herglotz’ theorem where two sums of this type were involved.) Since φ is positive definite, the sum in (5.41) is nonnegative. Hence, the integral (5.41), being the limit of such sums, is nonnegative. As N → ∞ the integral in (5.40) tends to the one in (5.39). So, that integral is nonnegative too. Thus f(t) ≥ 0.

Now let φ be any continuous positive definite function and let

images/eq-181-1.png

Since φ is bounded, φn is integrable. Since φ(x) and ex2/n are positive definite, so is their product φn(x). Thus by what we have proved in the preceding paragraph, for each n

images/eq-181-2.png

We have the relation δNφn = (∆Nfn)^, i.e.,

images/eq-181-3.png(5.42)

At x = 0 this gives

images/eq-181-4.png

Let N → ∞. This shows

images/eq-181-5.png(0)

i.e., images/nec-181-1.png and ||fn||1φ(0). Let n(t) = fn(t)dt. Then {µn} are positive measures on images/nec-181-2.png and ||µn|| ≤ φ(0). So, by Helly’s selection principle, there exists a positive measure µ, with ||µ|| ≤ φ(0), to which (a subsequence of ) µn converges in the weak topology.

The equation (5.42) says

images/eq-182-1.png(5.43)

Keep N fixed and let n → ∞. For the right-hand side of (5.43) use the weak convergence of µn to µ, and for the left-hand side the Lebesgue-dominated convergence theorem. This gives

images/eq-182-2.png(5.44)

Now let N → ∞. Since φ is a bounded measurable function, by (5.37) the left-hand side of (5.44) goes to φ(x) a.e. The right-hand side converges by the bounded convergence theorem. This shows

images/eq-182-3.png

Since the two sides are continuous functions of x, this equality holds everywhere.     ■


Of the several examples of positive definite functions in Section 5.2 some were shown to be Fourier transforms of nonnegative integrable functions. (See 5.2.5 - 5.2.8.) One can do this for some of the other functions too.


5.5.4

The list below gives some functions φ and their Fourier transforms φimages/nec-182-1.png (ignoring constant factors).

images/eq-182-4.png
images/eq-183-1.png

Let φ be a continuous positive definite function. Then the measure µ associated with φ via the formula (5.38) is a probability measure if and only if φ(0) = 1. In this case φ is called a characteristic function.



5.5.5  Proposition

Let φ be a characteristic function. Then

images/eq-183-2.png

for all x and n = 1, 2, . . . .

Proof.   By elementary trigonometry

images/eq-183-3.png

An iteration leads to the inequality

images/eq-183-4.png

From (5.38) we have

images/eq-183-5.png

5.5.6  Corollary

Suppose φ is a positive definite function and φ(x) = φ(0) + o(x2); i.e.,

images/eq-184-1.png(0)

Then φ is a constant.

Proof.   We may assume that φ(0) = 1. Then using the Proposition above we have for all x and n

images/eq-184-2.png

The hypothesis on φ implies that the last expression goes to zero as n → ∞. Hence, Re φ(x) = 1 for all x. But then φ(x) ≡ 1.     ■


5.5.7  Exercise

Suppose φ is a characteristic function, and φ(x) = 1 + o(x) + o(x2) in a neighbourhood of 0, where o(x) is an odd function. Then φ ≡ 1. [Hint: consider φ(x)φ(−x).]


5.5.8  Exercise

The functions ex4 , 1/(1 + x4), and e−|x|a for a > 2, are not positive definite.

Bochner’s theorem can be used also to show that a certain function is not positive definite by showing that its Fourier transform is not everywhere nonnegative.


5.5.9  Exercise

Use the method of residues to show that for all t > 0

images/eq-184-3.png

It follows from Bochner’s theorem that the function f(x) = 1/(1 + x4) is not positive definite.


5.6  SUPPLEMENTARY RESULTS AND EXERCISES

5.6.1  Exercise

Let U be a unitary operator on any separable Hilbert space images/nec-185-1.png. Show that for each unit vector x in images/nec-185-2.png the sequence

images/eq-185-1.png(5.45)

is positive definite.

This observation is the first step on one of the several routes to the spectral theorem for operators in Hilbert space. We indicate this briefly.

Let U be a unitary operator on images/nec-185-3.png. By Exercise 5.6.1 and Herglotz’ theorem, for each unit vector x in images/nec-185-4.png, there exists a probability measure µx on the interval [−π, π] such that

images/eq-185-2.png(5.46)

Using a standard technique called polarisation, one can obtain from this, for each pair x, y of unit vectors a complex measure µx,y such that

images/eq-185-3.png(5.47)

Now for each Borel subset E ⊂ [−π, π] let P(E) be the operator on images/nec-185-5.png defined by the relation

images/eq-185-4.png(5.48)

It can be seen that P(E) is an orthogonal projection and that P(·) is countably additive on the Borel σ-algebra of [−π, π]. In other words P(·) is a projection-valued measure. We can then express U as an integral

images/eq-185-5.png(5.49)

This is the spectral theorem for unitary operators. The spectral theorem for self-adjoint operators can be obtained from this using the Cayley transform.


5.6.2  Exercise

Let B be an n × n Hermitian matrix. Show that for each unit vector u the function

images/eq-186-1.png

is a positive definite function on images/nec-186-1.png. Use this to show that the functions tr eitB and det eitB are positive definite.


5.6.3  Exercise

Let A, B be n × n Hermitian matrices and let

images/eq-186-2.png(5.50)

Is φ a positive definite function? Show that this is so if A and B commute.

The general case of the question raised above is a long-standing open problem in quantum statistical mechanics. The Bessis-Moussa-Villani conjecture says that the function φ in (5.50) is positive definite for all Hermitian matrices A and B.

The purpose of the next three exercises is to calculate Fourier transforms of some functions that arose in our discussion.


5.6.4  Exercise

Let φ(x) = 1/cosh x. Its Fourier transform is

images/eq-186-3.png

This integral may be evaluated by the method of residues. Let f be the function

images/eq-186-4.png

Then

images/eq-186-5.png

For any R > 0 the rectangular contour with vertices −R, R, R+ and −R + contains one simple pole, z = iπ/2, of f inside it. Integrate f along this contour and then let |R| → ∞. The contribution of the two vertical sides goes to zero. So

images/eq-187-1.png

where Resz=z0 f(z) is the residue of f at a pole z0.

A calculation shows that

images/eq-187-2.png

5.6.5  Exercise

More generally consider the function

images/eq-187-3.png(5.51)

Integrate the function

images/eq-187-4.png

along the rectangular contour with vertices −R, R, R + i2π and −R + i2π. The function f has two simple poles z = i(π ± arccos a) inside this rectangle. Proceed as in Exercise 5.6.4 to show

images/eq-187-5.png(5.52)

It is plain that images/nec-187-1.png. Hence by Bochner’s theorem φ(x) is positive definite for −1 < a < 1. By a continuity argument it is positive definite for a = 1 as well.


5.6.6  Exercise

Now consider the function

images/eq-187-6.png(5.53)

Use the function f and the rectangular contour of Exercise 5.6.5. Now f has two simple poles z = ± arccosh t+ inside this rectangle. Show that

images/eq-187-7.png(5.54)

It is plain that images/nec-188-1.png is negative for some values of t. So the function φ in (5.53) is not positive definite for any a > 1.


5.6.7  Exercise

Let λ1, . . . , λn be positive numbers and let Z be the n× n matrix with entries

images/eq-188-1.png

Show that if −2 < t ≤ 2, then Z is positive definite; and if t > 2 then there exists an n > 2 for which this matrix is not positive definite. (See Exercise 1.6.4.)


5.6.8  Exercise

For 0 < a < 1, let fa be the piecewise linear function defined as

images/eq-188-2.png

Show that fa is not positive definite. Compare this with 5.2.13 and 5.2.15. Express fa as the convolution of two characteristic functions.

The technique introduced in Section 4 is a source of several interesting inequalities. The next two exercises illustrate this further.


5.6.9  Exercise

(i)  Let A be a Hermitian matrix. Use the positive definiteness of the function sech x to show that for every matrix X

images/eq-188-3.png

(ii)  Now let A be any matrix. Apply the result of (i) to the matrices images/nec-188-2.png and show that

images/eq-188-4.png

for every matrix X. Replacing A by iA, one gets

images/eq-188-5.png

5.6.10  Exercise

Let A, B be normal matrices with ||A|| ≤ 1 and ||B|| ≤ 1. Show that for every X we have

images/eq-189-1.png

The inequalities proved in Section 5.4 have a leitmotiv. Let M(a, b) be any mean of positive numbers a and b satisfying the conditions laid down at the beginning of Chapter 4. Let A be a positive definite matrix with eigenvalues λ1 ≥ · · · ≥ λn. Let M(A, A) be the matrix with entries

images/eq-189-2.png

Many of the inequalites in Section 5.4 say that for certain means M1 and M2

images/eq-189-3.png(5.55)

for all X. We have proved such inequalites by showing that the matrix Y with entries

images/eq-189-4.png(5.56)

is positive definite. This condition is also necessary for (5.55) to be true for all X.


5.6.11  Proposition

Let M1(a, b) and M2(a, b) be two means. Then the inequality (5.55) is true for all X if and only if the matrix Y defined by (5.56) is positive definite.

Proof.   The Schur product by Y is a linear map on images/nec-189-1.png. The inequality (5.55) says that this linear map on the space images/nec-189-2.png equipped with the norm || · || is contractive. Hence it is contractive also with respect to the dual norm || · ||1; i.e.,

images/eq-189-5.png

Choose X to be the matrix with all entries equal to 1. This gives ||Y ||1n. Since Y is Hermitian

images/eq-189-6.png

where λi(Y ) are the eigenvalues of Y. Since yii = 1 for all i, we have

images/eq-190-1.png

Thus images/nec-190-1.png. But this is possible only if λi(Y ) ≥ 0 for all i. In other words Y is positive.     ■

Let us say that M1M2 if M1(a, b) ≤ M2(a, b) for all positive numbers a and b; and M1 << M2 if for every n and every choice of n positive numbers λ1, . . . , λn the matrix (5.56) is positive definite. If M1 << M2 the inequality (5.55) is true for all positive matrices A and all matrices X. Clearly M1M2 if M1 << M2. The converse is not always true.


5.6.12  Exercise

Let A(a, b) and G(a, b) be the arithmetic and the geometric means of a and b. For 0 ≤ α ≤ 1 let

images/eq-190-2.png

Clearly we have FαFβ whenever αβ. Use Exercise 5.6.6 to show that Fα << Fβ if and only if images/nec-190-2.png.

Using Exercise 2.7.12 one can see that if M1 << M2, then the inequality (5.55) is true for all unitarily invariant norms. The weaker condition M1M2 gives this inequality only for the Hilbert-Schmidt norm || · ||2.

In Exercises 1.6.3, 1.6.4, 5.2.21, 5.2.22 and 5.2.23 we have outlined simple proofs of the infinite divisibility of some special matrices. These proofs rely on arguments specifically tailored to suit the matrices at hand. In the next few exercises we sketch proofs of some general theorems that are useful in this context.

An n × n Hermitian matrix A is said to be conditionally positive definite if images/nec-190-3.png for all images/nec-190-4.png such that x1 + · · · + xn = 0. (The term almost positive definite is also used sometimes.)


5.6.13  Proposition

Let A = [[aij]] be an n × n conditionally positive definite matrix. Then there exist a positive definite matrix B = [[bij]] and a vector y = (y1, . . . , yn) in images/nec-191-1.png such that

images/eq-191-1.png(5.57)

Proof.   Let J be the n × n matrix all of whose entries are equal to 1/n. For any vector images/nec-191-2.png let x# = Jx and images/nec-191-3.png. Since images/nec-191-4.png we have

images/eq-191-2.png

Let B = AAJJA + JAJ. The inequality above says that

images/eq-191-3.png

In other words B is positive definite.

If C1, . . . , Cn are the columns of the matrix A, then the jth column of the matrix JA has all its entries equal to images/nec-191-5.png , the number obtained by averaging the entries of the column Cj. Likewise, if R1, . . . , Rn are the rows of A, then the ith row of AJ has all its entries equal to images/nec-191-6.png. Since A is Hermitian images/nec-191-7.png is the complex conjugate of images/nec-191-8.png. The matrix JAJ has all its entries equal to images/nec-191-9.png. Thus the i, j entry of the matrix AJ + JAJAJ is equal to images/nec-191-10.png. Let y be the vector

images/eq-191-4.png(5.57)

Then the equation (5.57) is satisfied.     ■


5.6.14  Exercise

Let A = [[aij]] be a conditionally positive definite matrix. Show that the matrix [[eaij ]] is positive definite. [Hint: If B = [[bij]] is positive definite, then [[ebij ]] is positive definite. Use Proposition 5.6.13.]


5.6.15  Exercise

Let A = [[aij]] be an n × n matrix with positive entries and let L = [[log aij]]. Let E be the matrix all whose entries are equal to 1. Note that Ex = 0 if x1 + · · · + xn = 0.

(i)  Suppose A is infinitely divisible. Let x be any vector with x1 + · · · + xn = 0, and for r > 0 let A(r) be the matrix with entries arij. Then

images/eq-191-5.png

Let r ↓ 0. This gives

images/eq-192-1.png

Thus L is conditionally positive definite.

(ii)  Conversely, if L is conditionally positive definite, then so is rL for every r ≥ 0. Use Exercise 5.6.14 to show that this implies A is infinitely divisible.

Thus a Hermitian matrix A with entries aij > 0 is infinitely divisible if and only if the matrix L = [[log aij]] is conditionally positive definite. The next exercise gives a criterion for conditional positive definiteness.


5.6.16  Exercise

Given an n×n Hermitian matrix B = [[bij]] let D be the (n−1)×(n−1) matrix with entries

images/eq-192-2.png(5.58)

Show that B is conditionally positive definite if and only if D is positive definite.

We now show how the results of Exercises 5.6.14–5.6.16 may be used to prove the infinite divisibility of an interesting matrix.

For any n, the n × n Pascal matrix A is the matrix with entries

images/eq-192-3.png(5.59)

The entries of the Pascal triangle occupy the antidiagonals of A. Thus the 4 × 4 Pascal matrix is

images/eq-192-4.png

5.6.17  Exercise

Prove the combinatorial identity

images/eq-192-5.png

[Hint: Separate i + j objects into two groups, the first containing i objects and the second j objects. If we choose ik objects from the first group and k from the second, we have chosen i objects out of i + j.]


5.6.18  Exercise

Show that

images/eq-193-1.png

Use this to conclude that the Pascal matrix is a Gram matrix and is thus positive definite.


5.6.19  Exercise

Let L be the n × n lower triangular matrix whose rows are the rows of the Pascal triangle. Thus for n = 4

images/eq-193-2.png

Show that A = LL. This gives another proof of the positive definiteness of the Pascal matrix A.


5.6.20  Exercise

For every n, the n × n Pascal matrix is infinitely divisible. Prove this statement following the steps given below.

(i)  Use the results of Exercises 5.6.15 and 5.6.16. If B has entries bij = log aij, where aij are defined by (5.59), then the entries dij defined by (5.58) are given by

images/eq-193-3.png

We have to show that the matrix D = [[dij]] is positive definite.

(ii)  For x > 0 we have

images/eq-194-1.png

where µ is the probability measure on [0, ∞) defined as (t) = dt/t2. Use this to show that

images/eq-194-2.png

(iii)  Thus the matrix D can be expressed as

images/eq-194-3.png

where C(t) = [[cij(t)]] is a Cauchy matrix for all t ≥ 1. This shows that D is positive definite.


5.6.21  Exercise

The infinite divisibility of the Pascal matrix can be proved in another way as follows. Let λ1, . . . , λn be positive numbers, and let K be the n × n matrix with entries

images/eq-194-4.png

When λj = j, 1 ≤ jn, this is the Pascal matrix. Use Gauss’s product formula for the gamma function

images/eq-194-5.png

to see that

images/eq-194-6.png

Each of the factors in this product is the i, j entry of a matrix that is congruent to a Cauchy matrix. Hence K is infinitely divisible.

Let f be a nonnegative function on images/nec-194-1.png. If for each r > 0 the function (f(x))r is positive definite, then we say that f is an infinitely divisible function. By Schur’s theorem, the product of two infinitely divisible functions is infinitely divisible. If f is a nonnegative function and for each m = 1, 2, . . . the function (f(x))1/m is positive definite, then f is infinitely divisible.

Some examples of infinitely divisible functions are given in the next few exercises.


5.6.22  Exercise

(i)  The function f(x) = 1/(cosh x) is infinitely divisible.

(ii)  The function f(x) = 1/(cosh x + a) is infinitely divisible for −1 < a ≤ 1. [Hint: Use Exercises 1.6.4 and 1.6.5.]


5.6.23  Exercise

In Section 5.2.10 we saw that the function

images/eq-195-1.png

is positive definite. Using this information and Schur’s theorem one can prove that f is in fact infinitely divisible. The steps of the argument are outlined.

(i)  Let a and b be any two nonnegative real numbers. Then either images/nec-195-1.png is positive definite.

Hence

images/eq-195-2.png

is positive definite.

(ii)  Use the identity

images/eq-195-3.png

to obtain

images/eq-195-4.png

(iii)  For 0 < r < 1 we have the expansion

images/eq-196-1.png

where the coefficients an are nonnegative. Use Part (i) of this exercise and of Exercise 5.6.22 to prove that the series above represents a positive definite function. This establishes the assertion that (cosh αx)/(cosh x) is infinitely divisible for 0 < α < 1.


5.6.24  Exercise

The aim of this exercise is to show that the function

images/eq-196-2.png

is infinitely divisible. Its positive definiteness has been established in Section 5.2.10.

(i)   Imitate the arguments in Exercise 5.6.23. Use the identity

images/eq-196-3.png

to show that the function

images/eq-196-4.png

is infinitely divisible for 0 ≤ ba. (This restriction is needed to handle the term sinh(ab)x occurring in the series expansion.) This shows that the function (sinh αx)/(sinh x) is infinitely divisible for 1/2 ≤ α ≤ 1.

(ii)  Let α be any number in (0, 1) and choose a sequence

images/eq-196-5.png

where αii+1 ≥ 1/2. Then

images/eq-196-6.png

Each factor in this product is infinitely divisible, and hence so is the product.


5.6.25  Exercise

(i)   Use Exercise 5.6.24 to show that the function x/(sinh x) is infinitely divisible. [Hint: Take limit α ↓ 0.]

(ii)  Use this and the result of Exercise 5.6.23 to show that the function

images/eq-197-1.png

is infinitely divisible.


5.6.26  Exercise

Let λ1, . . . , λn be any real numbers. Use the result of Exercise 5.2.22 to show that the matrix

images/eq-197-2.png

is infinitely divisible. Thus the function f(x) = e−|x| is infinitely divisible. Use the integral formula

images/eq-197-3.png

to obtain another proof of this fact.


5.6.27  Exercise

Using the gamma function, as in Exercise 1.6.4, show that for every r > 0

images/eq-197-4.png

Thus the functions 1/(1 + ix)r, 1/(1 − ix)r, and 1/(1 + x2)r are positive definite for every r > 0. This shows that 1/(1 + x2) is infinitely divisible.


5.6.28  Exercise

Let a and b be nonnegative numbers with ab. Let 0 < r < 1. Use the integral formula (1.39) to show that

images/eq-197-5.png

where µ is a positive measure. This is equal to

images/eq-198-1.png

Show that this is positive definite as a function of x. Note that it suffices to show that for each λ > 0,

images/eq-198-2.png

is positive definite. This, in turn, follows from the integral representation

images/eq-198-3.png

where γ = [( + b)/(1 + λ)]1/2. Thus, for ab the function f(x) = (1 + bx2)/(1 + ax2) is infinitely divisible.


5.6.29  Exercise

Show that the function f(x) = (tanh x)/x is infinitely divisible. [Hint: Use the infinite product expansion for f(x).]


5.6.30  Exercise

Let t > −1 and consider the function

images/eq-198-4.png

Use the identity

images/eq-198-5.png

to obtain the equality

images/eq-198-6.png

Use the binomial theorem and Exercise 5.6.29 to prove that f is infinitely divisible for −1 < t ≤ 1.

Thus many of the positive definite functions from Section 5.2 are infinitely divisible. Consequently the associated positive definite matrices are infinitely divisible. In particular, for any positive numbers λ1, . . . , λn the n × n matrices V, W and Y whose entries are, respectively,

images/eq-199-1.png

are infinitely divisible.


5.6.31  Another proof of Bochner’s Theorem

The reader who has worked her way through the theory of Pick functions (as given in Chapter V of MA) may enjoy the proof outlined below.

(i)  Let φ be a positive definite function on images/nec-199-1.png, continuous at 0. Let z = x + iy be a complex number and put

images/eq-199-2.png(5.60)

Since φ is bounded, this integral is convergent for y > 0. Thus f is an analytic function on the open upper half-plane H+.

(ii)  Observe that

images/eq-199-3.png

and so from (5.60) we have

images/eq-199-4.png

First substitute u = t + v in both the integrals, and then interchange u and v in the second integral to obtain

images/eq-199-5.png
images/eq-200-1.png

Observe that these two double integrals are over the quarter-planes {(u, v) : uv ≥ 0} and {(u, v) : vu ≥ 0}, respectively. Hence

images/eq-200-2.png

Since φ is a positive definite function, this integral is nonnegative. (Write it as a limit of Riemann sums each of which is nonnegative.)

Thus f maps the upper half-plane into the right half-plane. So i f(z) is a Pick function.

(iii)  For η > 0

images/eq-200-3.png(0)

Hence, by Problem V.5.9 of MA, there exists a finite positive measure µ on images/nec-200-1.png such that

images/eq-200-4.png

(iv)  Thus we have

images/eq-200-5.png

(v)  Compare the expression for f in (5.60) with the one obtained in (iv) and conclude

images/eq-200-6.png

This is the assertion of Bochner’s theorem.


5.7  NOTES AND REFERENCES

Positive definite functions have applications in almost every area of modern analysis. In 1907 C. Carathéodory studied functions with power series

images/eq-201-1.png

and found necessary and sufficient conditions on the sequence {an} in order that f maps the unit disk into the right half-plane. In 1911 O. Toeplitz observed that Carathéodory’s condition is equivalent to (5.1). The connection with Fourier series and transforms has been pointed out in this chapter. In probability theory positive definite functions arise as characteristic functions of various distributions. See E. Lukacs, Characteristic Functions, Griffin, 1960, and R. Cuppens, Decomposition of Multivariate Probabilities, Academic Press, 1975. We mention just one more very important area of their application: the theory of group representations.

Let G be a locally compact topological group. A (continuous) complex-valued function φ on G is positive definite if for each N = 1, 2, . . . , the N × N matrix images/nec-201-1.png is positive for every choice of elements g0, . . . , gN−1 from G. A unitary representation of G is a homomorphism images/nec-201-2.png from G into the group of unitary operators on a Hilbert space H such that for every fixed images/nec-201-3.png the map images/nec-201-4.png from G into images/nec-201-5.png is continuous. (This is called strong continuity.) It is easy to see that if Ug is a unitary representation of G in the Hilbert space images/nec-201-6.png, then for every images/nec-201-7.png the function

images/eq-201-2.png(5.61)

is positive definite on G. (This is a generalization of Exercise 5.6.1.) The converse is an important theorem of Gelfand and Raikov proved in 1943. This says that for every positive definite function φ on G there exist a Hilbert space images/nec-201-8.png, a unitary representation Ug of G in images/nec-201-9.png, and a vector images/nec-201-10.png such that the equation (5.61) is valid. This is one of the first theorems in the representation theory of infinite groups. One of its corollaries is that every locally compact group has sufficiently many irreducible unitary representations. More precisely, for every element g of G different from the identity, there exists an irreducible unitary representation of G for which Ug is not the identity operator.

An excellent survey of positive definite functions is given in J. Stewart, Positive definite functions and generalizations, an historical survey, Rocky Mountain J. Math., 6 (1976) 409–434. Among books, we recommend C. Berg, J.P.R. Christensen, and P. Ressel, Harmonic Analysis on Semigroups, Springer, 1984, and Z. Sasvári, Positive Definite and Definitizable Functions Akademie-Verlag, Berlin, 1994.

In Section 5.2 we have constructed a variety of examples using rather elementary arguments. These, in turn, are useful in proving that certain matrices are positive. The criterion in 5.2.15 is due to G. Pólya, Remarks on characteristic functions, Proc. Berkeley Symp. Math. Statist. & Probability, 1949, pp.115-123. This criterion is very useful as its conditions can be easily verified.

The ideas and results of Sections 5.2 and 5.3 are taken from the papers R. Bhatia and K. R. Parthasarathy, Positive definite functions and operator inequalities, Bull. London Math. Soc. 32 (2000) 214–228, H. Kosaki, Arithmetic-geometric mean and related inequalities for operators, J. Funct. Anal., 15 (1998) 429–451, F. Hiai and H. Kosaki, Comparison of various means for operators, ibid., 163 (1999) 300–323, and F. Hiai and H. Kosaki, Means for matrices and comparison of their norms, Indiana Univ. Math. J., 48 (1999) 899–936.

The proof of Theorem 5.3.1 given here is from R. Bhatia and K. B. Sinha, Derivations, derivatives and chain rules, Linear Algebra Appl., 302/303 (1999) 231–244. Theorem 5.3.3 was proved by K. Löwner (C. Loewner) in Über monotone Matrixfunctionen, Math. Z., 38 (1934) 177–216. Loewner then used this theorem to show that a function is operator monotone on the positive half-line if and only if it has an analytic continuation mapping the upper half-plane into itself. Such functions are characterized by certain integral representations, namely, f is operator monotone if and only if

images/eq-202-1.png(5.62)

for some real numbers α and β with β ≥ 0, and a positive measure µ that makes the integral above convergent. The connection between positivity of Loewner matrices and complex functions is made via Carathéodory’s theorem (mentioned at the beginning of this section) and its successors. Following Loewner’s work operator monotonicity of particular examples such as 5.3.5–5.3.8 was generally proved by invoking the latter two criteria (analytic continuation or integral representation). The more direct proofs based on the positivity of Loewner matrices given here are from the 2000 paper of Bhatia and Parthasarathy.

The inequality (5.24) and the more general (5.28) were proved in R. Bhatia and C. Davis, More matrix forms of the arithmetic-geometric mean inequality, SIAM J. Matrix Anal. Appl., 14 (1993) 132–136. For the operator norm alone, the inequality (5.28) was proved by E. Heinz, Beiträge zur Störungstheorie der Spektralzerlegung, Math. Ann., 123 (1951) 415–438. The inequality (5.24) aroused considerable interest and several different proofs of it were given by various authors. Two of them, R. A. Horn, Norm bounds for Hadamard products and the arithmetic-geometric mean inequality for unitarily invariant norms, Linear Algebra Appl., 223/224 (1995) 355–361, and R. Mathias, An arithmetic-geometric mean inequality involving Hadamard products, ibid., 184 (1993) 71–78, observed that the inequality follows from the positivity of the matrix in (5.27). The papers by Bhatia-Parthasarathy and Kosaki cited above were motivated by extending this idea further. The two papers used rather similar arguments and obtained similar results. The program was carried much further in the two papers of Hiai and Kosaki cited above to obtain an impressive variety of results on means. The interested reader should consult these papers as well as the monograph F. Hiai and H. Kosaki, Means of Hilbert Space Operators, Lecture Notes in Mathematics Vol. 1820, Springer, 2003.

The theorems of Herglotz and Bochner concern the groups images/nec-203-1.png and images/nec-203-2.png. They were generalized to locally compact abelian groups by A. Weil, by D. A. Raikov, and by A. Powzner, in independent papers appearing almost together. Further generalizations (non-abelian or non-locally compact groups) exist. The original proof of Bochner’s theorem appears in S. Bochner, Vorlesungen über Fouriersche Integrale, Akademie-Verlag, Berlin, 1932. Several other proofs have been published. The one given in Section 5.5 is taken from R. Goldberg, Fourier Transforms, Cambridge University Press, 1961, and that in Section 5.6 from N. I. Akhiezer and I. M. Glazman, Theory of Linear Operators in Hilbert Space, Dover, 1993 (reprint of original editions). A generalization to distributions is given in L. Schwartz, Théorie des Distributions, Hermann, 1954.

Integral representations such as the one given by Bochner’s theorem are often viewed as a part of “Choquet Theory.” Continuous positive definite functions φ(x) such that φ(0) = 1 form a compact convex set; the family images/nec-203-3.png is the set of extreme points of this convex set.

Exercise 5.6.1 is an adumbration of the connections between positive definite functions and spectral theory of operators. A basic theorem of M. H. Stone in the latter subject says that every unitary representation images/nec-203-4.png of R in a Hilbert space H is of the form Ut = eitA for some (possibly unbounded) self-adjoint operator A. (The operator A is bounded if and only if ||UtI|| → 0 as t → 0.) The theorems of Stone and Bochner can be derived from each other. See M. Reed and B. Simon, Methods of Modern Mathematical Physics, Vols. I, II, Academic Press, 1972, 1975, Chapters VIII, IX.

A sequence images/nec-204-1.png is of positive type if for every positive integer N, we have

images/eq-204-1.png(5.63)

for every finite sequence of complex numbers ξ0, ξ1, . . . , ξN−1. This is equivalent to the requirement that for each N = 1, 2, . . . , the N × N matrix

images/eq-204-2.png(5.64)

is positive. Compare these conditions with (5.1) and (5.2). (Matrices of the form (5.64) are called Hankel matrices while those of the form (5.2) are Toeplitz matrices.) A complex valued function φ on the positive half-line [0, ∞) is of positive type if for each N the N × N matrix

images/eq-204-3.png(5.65)

is positive for every choice of x0, . . . , xN−1 in [0, ∞). A theorem of Bernstein and Widder says that φ is of positive type if and only if there exists a positive measure µ on [0, ∞) such that

images/eq-204-4.png(5.66)

i.e., φ is the Laplace transform of a positive measure µ. Such functions are characterized also by being completely monotone, which, by definition, means that

images/eq-204-5.png

See MA p.148 for the connection such functions have with operator monotone functions. The book of Berg, Christensen, and Ressel cited above is a good reference for the theory of these functions.

Our purpose behind this discussion is to raise a question. Suppose f is a function mapping [0, ∞) into itself. Say that f is in the class images/nec-205-1.png if for each N the matrix

images/eq-205-1.png

is positive for every choice λ1, . . . , λN in [0, ∞). The class images/nec-205-2.png is precisely the operator monotone functions. Is there a good characterisation of functions in images/nec-205-3.png? One can easily see that if images/nec-205-4.png, then so does 1/f. It is known that images/nec-205-5.png is contained in images/nec-205-6.png; see, e.g., M. K. Kwong, Some results on matrix monotone functions, Linear Algebra Appl., 118 (1989) 129–153. (It is easy to see, using the positivity of the Cauchy matrix, that for every λ > 0 the function g(t) = λt/(λ + t) is in images/nec-205-7.png. The integral representation (5.62) then shows that every function in images/nec-205-8.png is in images/nec-205-9.png.)

The conjecture stated after Exercise 5.6.3 goes back to D. Bessis, P. Moussa, and M. Villani, Monotonic converging variational approximations to the functional integrals in quantum statistical mechanics, J. Math. Phys., 16 (1975) 2318–2325. A more recent report on the known partial results may be found in P. Moussa, On the representation of Tr images/nec-205-10.png as a Laplace transform, Rev. Math. Phys., 12 (2000) 621–655. E. H. Lieb and R. Seiringer, Equivalent forms of the Bessis-Moussa-Villani conjecture, J. Stat. Phys., 115 (2004) 185–190, point out that the statement of this conjecture is equivalent to the following: for all A and B positive, and all natural numbers p, the polynomial images/nec-205-11.png has only positive coefficients. When this polynomial is multiplied out, the co-efficient of λr is a sum of terms each of which is the trace of a word in A and B. It has been shown by C. R. Johnson and C. J. Hillar, Eigenvalues of words in two positive definite letters, SIAM J. Matrix Anal. Appl., 23 (2002) 916-928, that some of the individual terms in this sum can be negative. For example, tr A2B2AB can be negative even when A and B are positive.

The matrix Z in Exercise 5.6.7 was studied by M. K. Kwong, On the definiteness of the solutions of certain matrix equations, Linear Algebra Appl., 108 (1988) 177–197. It was shown here that for each n ≥ 2, there exists a number tn such that Z is positive for all t in (−2, tn], and further tn > 2 for all n, tn = ∞, 8, 4 for n = 2, 3, 4, respectively. The complete solution (given in Exercise 5.6.7) appears in the 2000 paper of Bhatia-Parthasarathy cited earlier. The idea and the method are carried further in R. Bhatia and D. Drissi, Generalised Lyapunov equations and positive definite functions, SIAM J. Matrix Anal. Appl., 27 (2005) 103-295–114. Using a Fourier transforms argument D. Drissi, Sharp inequalities for some operator means, preprint 2006, has shown that the function f(x) = (x cosh α x)/ sinh x is not positive definite when |α| > 1/2. The result of Exercise 5.6.9 is due to E. Andruchow, G. Corach, and D. Stojanoff, Geometric operator inequalities, Linear Algebra Appl., 258 (1997) 295-310, where other related inequalities are also discussed. The result of Exercise 5.6.10 was proved by D. K. Jocić, Cauchy-Schwarz and means inequalities for elementary operators into norm ideals, Proc. Am. Math. Soc., 126 (1998) 2705–2711. Cognate results are proved in D. K. Jocić, Cauchy-Schwarz norm inequalities for weak-integrals of operator valued functions, J. Funct. Anal., 218 (2005) 318–346.

Proposition 5.6.11 is proved in the Hiai-Kosaki papers cited earlier. They also give an example of two means where M1M2, but M1 << M2 is not true. The simple example in Exercise 5.6.12 is from R. Bhatia, Interpolating the arithmetic-geometric mean inequality and its operator version, Linear Algebra Appl., 413 (2006) 355–363.

Conditionally positive definite matrices are discussed in Chapter 4 of the book R. B. Bapat and T. E. S. Raghavan, Nonnegative Matrices and Applications, Cambridge University Press, 1997, and more briefly in Section 6.3 of R. A. Horn and C. R. Johnson, Topics in Matrix Analysis, Cambridge University Press, 1991. This section also contains a succinct discussion of infinitely divisible matrices and references to original papers. The results of Exercises 5.6.20 and 5.6.21 are taken from R. Bhatia, Infinitely divisible matrices, Am. Math. Monthly, 113 (2006) 221–235, and those of Exercises 5.6.23, 5.6.24 and 5.6.25 from R. Bhatia and H. Kosaki, Mean matrices and infinite divisibility, preprint 2006. In this paper it is shown that for several classes of means m(a, b), matrices of the form [[m(λi, λj)]] are infinitely divisible if images/nec-206-1.png for all a and b; and if images/nec-206-2.png, then matrices of the form [[1/m(λi, λj)]] are infinitely divisible. The contents of Exercise 5.6.28 are taken from H. Kosaki, On infinite divisibility of positive definite functions, preprint 2006. In this paper Kosaki uses very interesting ideas from complex analysis to obtain criteria for infinite divisibility.

We have spoken of Loewner’s theorems that say that the Loewner matrices associated with a function f on [0, ∞) are positive if and only if f has an analytic continuation mapping the upper half-plane into itself. R. A. Horn, On boundary values of a schlicht mapping, Proc. Am. Math. Soc., 18 (1967) 782–787, showed that this analytic continuation is a one-to-one (schlicht) mapping if and only if the Loewner matrices associated with f are infinitely divisible. This criterion gives another proof of the infinite divisibility of the matrices in Sections 5.3.5 and 5.3.7.

Infinitely divisible distribution functions play an important role in probability theory. These are exactly the limit distributions for sums of independent random variables. See, for example, L. Breiman, Probability, Addison-Wesley, 1968, pp. 190–196, or M. Loeve, Probability Theory, Van Nostrand, 1963, Section 22. The two foundational texts on this subject are B. V. Gnedenko and A. N. Kolmogorov, Limit Distributions for Sums of Independent Random Variables, Addison-Wesley, 1954, and P. Lévy, Théorie de l’ Addition des Variables Aléatoires, Gauthier-Villars, 1937. The famous LévyKhintchine Formula says that a continuous positive definite function φ, with φ(0) = 1, is infinitely divisible if and only if it can be represented as

images/eq-207-1.png

where γ is a real number, G is a nondecreasing function of bounded variation, and the integrand at u = 0 is to be interpreted as the limiting value −t2/2. This representation of log f(t) is unique. See the book by Gnedenko and Kolmogorov, p. 76. A list of such representations for some of the examples we have considered is given in Lukacs, p. 93.

A positive definite function is said to be self-decomposable if for every real number ν in (0, 1) there exists a positive definite function φν such that

images/eq-207-2.png

Such functions are of importance in probability theory; see Section 23.3 of Loeve’s book. It is shown there that if φ is self-decomposable, then it is infinitely divisible, and so are the functions φν, 0 < ν < 1. Our discussion in Section 5.2.10 shows that the functions 1/(1 + x2), 1/cosh x and x/sinh x are self-decomposable. So, the infinite divisibility of some of the functions in Exercises 5.6.22–5.6.28 is a consequence also of Loeve’s theorem.

In Exercise 5.6.28 we observed that if ab ≥ 0, then the function f(x) = (1 + bx2)/(1 + ax2) is infinitely divisible, and in Exercise 5.6.29 asked the reader to use this to show that g(x) = (tanh x)/x is infinitely divisible. In the paper Z. J. Jurek and M. Yor, Self decomposable laws associated with hyperbolic functions, Probability and Mathematical Statistics, 24 (2004), 181–190, it is observed that the function f is not self-decomposable but the function g is. Some of the results in H. Kosaki, On infinite divisibility of positive definite functions, preprint 2006, can be rephrased to say that certain functions are self-decomposable. His list includes, for example, 1/(cosh x + a), −1 ≤ a ≤ 1. The relevance and importance of these functions in probability theory is explained in the paper of Jurek and Yor, and in the references therein.

Finally, we make a few comments on operator inequalities with special reference to the arithmetic-geometric mean inequality. Operator inequalities are sought and found in three different versions. If A and B are positive, we may have sometimes an inequality AB. By Weyl’s monotonicity principle, this implies that λj(A) ≥ λj(B), 1 ≤ jn, where λj(A) is the jth eigenvalue of A counted in decreasing order. (This is equivalent to the existence of a unitary matrix U such that AUBU.) This, in turn implies that |||A||| ≥ |||B||| for all unitarily invariant norms. If A and B are not positive, we may ask whether |A| ≥ |B|. This implies the set of inequalities sj(A) ≥ sj(B) for all the singular values, which in turn implies |||A||| ≥ |||B||| for all unitarily invariant norms.

In Chapter 4 we saw an arithmetic-geometric mean inequality of the first kind; viz., images/nec-208-1.png for any two positive matrices. The inequality (5.23) and its stronger version (5.24) are of the third kind. An inequality of the second kind was proved by R. Bhatia and F. Kittaneh, On the singular values of a product of operators, SIAM J. Matrix Anal. Appl., 11 (1990) 272–277. This says that images/nec-208-2.png, for 1 ≤ jn. This inequality for singular values implies the inequality (5.23) for norms. A stronger version sj(A1/2XB1/2) ≤ sj(AX + XB)/2 is not always true. So, there is no second level inequality generalising (5.24).

For positive numbers a and b, the arithmetic-geometric mean inequality may be written in three different ways:

(i)  images/nec-208-3.png,

(ii)  images/nec-208-4.png,

(iii)  images/nec-208-5.png .

While each of these three may be obtained from the other, the matrix versions suggested by them are different. For example (i) leads to the question whether

images/eq-208-1.png

This is different from the inequality of Bhatia and Kittaneh stated above. It is not known whether this is true when n > 2. Weaker than this is the third-level inequality

images/eq-209-1.png

This too is known to be true for a large class of unitarily invariant norms (including Schatten p-norms for p = 1 and for p ≥ 2). It is not known whether it is true for all unitarily invariant norms. From properties of the matrix square function, one can see that this last (unproven) inequality is stronger than the assertion

images/eq-209-2.png

This version of the arithmetic-geometric mean inequality is known to be true. Thus there are quite a few subtleties involved in noncommutative versions of simple inequalities. A discussion of some of these matters may be found in R. Bhatia and F. Kittaneh, Notes on matrix arithmetic-geometric mean inequalities, Linear Algebra Appl., 308 (2000) 203–211, where the results just mentioned are proved. We recommend the monograph X. Zhan, Matrix Inequalities, Lecture Notes in Mathematics Vol. 1790, Springer, 2002 for a discussion of several topics related to these themes.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.147.79.11