The Gelfand transform, positive linear functionals, and positive-definite functions

Jordan Bell
June 24, 2015

1 Introduction

In this note, unless we say otherwise every vector space or algebra we speak about is over .

If A is a Banach algebra and eA satisfies xe=x and ex=x for all xA, and also e=1, we say that e is unity and that A is unital.

If A is a unital Banach algebra and xA, the spectrum of x is the set σ(x) of those λ for which λe-x is not invertible. It is a fact that if A is a unital Banach algebra and xA, then σ(x).11 1 Walter Rudin, Functional Analysis, second ed., p. 253, Theorem 10.13.

If A and B are Banach algebras and T:AB is a map, we say that T is an isomorphism of Banach algebras if T is an algebra isomorphism and an isometry.

Theorem 1 (Gelfand-Mazur).

If A is a Banach algebra and every nonzero element of A is invertible, then there is an isomorphism of Banach algebras A.

Proof.

Let xA. σ(x). If λ1,λ2σ(x), then neither λ1e-x nor λ2e-x is invertible, so they are both 0: x=λ1e and x=λ2e, whence λ1=λ2. Therefore σ(x) has precisely one element, which we denote by λ(x), and which satisfies

x=λ(x)e.

If x,yA, then x+y=λ(x)e+λ(y)e=(λ(x)+λ(y))e and also x+y=λ(x+y)e, so λ(x+y)=λ(x)+λ(y). If xA and α, then αx=αλ(x)e and also αx=λ(αx)e, so λ(αx)=αλ(x). Hence xλ(x) is linear. If λ0, then λ(λ0e)=λ0, showing that xλ(x) is onto. If λ(x)=λ(y) then x=λ(x)e=λ(y)e=y, showing that xλ(x) is one-to-one. Therefore xλ(x) is a linear isomorphism A.

If xA, then x=λ(x)e gives

x=λ(x)e=|λ(x)|e=|λ(x)|,

showing that the map xλ(x) is an isometry A. ∎

2 Complex homomorphisms

An ideal J of an algebra A is said to be proper if JA. An ideal is called maximal if it is a maximal element in the collection of proper ideals of A ordered by set inclusion.

The following theorem, which is proved using the fact that a maximal ideal is closed, the fact that a quotient of a Banach algebra with a closed ideal is a Banach algebra, and the Gelfand-Mazur theorem, states some basic facts about algebra homomorphisms from a Banach algebra to .22 2 Walter Rudin, Functional Analysis, second ed., p. 277, Theorem 11.5.

Theorem 2.

If A is a commutative unital Banach algebra and Δ is the set of all nonzero algebra homomorphisms A, then:

  1. 1.

    If M is a maximal ideal of A then there is some hΔ for which M=kerh.

  2. 2.

    If hΔ then kerh is a maximal ideal of A.

  3. 3.

    xA is invertible if and only if h(x)0 for all hΔ.

  4. 4.

    xA is invertible if and only if x does not belong to any proper ideal of A.

  5. 5.

    λσ(x) if and only if there is some hΔ for which h(x)=λ.

3 The Gelfand transform and maximal ideals

Suppose that A is a commutative unital Banach algebra and that Δ is the set of all nonzero algebra homomorphisms A. For each xA, we define x^:Δ by

x^(h)=h(x),hΔ.

We call x^ the Gelfand transform of x, and we call the map Γ:AΔ defined by Γ(x)=x^ the Gelfand transform.

We define A^={x^:xA}, and we call the set Δ with the initial topology for A^ the maximal ideal space of A. That is, the topology of Δ is the coarsest topology on Δ such that each x^:Δ is continuous. If X is a topological space, we denote by C(X) the set of all continuous functions X. C(X) is a commutative unital algebra, although it need not be a Banach algebra.

The radical of A, denoted radA, is the intersection of all maximal ideals of A. If radA={0}, we say that A is semisimple.

The following theorem establishes some basic facts about the Gelfand transform and the maximal ideal space.33 3 Walter Rudin, Functional Analysis, second ed., p. 280, Theorem 11.9.

Theorem 3.

If A is a commutative unital Banach algebra and Δ is the maximal ideal space of A, then:

  1. 1.

    Γ:AC(Δ) is an algebra homomorphism with kerΓ=radA.

  2. 2.

    If xA, then imx^=σ(x).

  3. 3.

    Δ is a compact Hausdorff space.

Proof.

Let x,yA and α. For hΔ,

Γ(αx+y)(h)=h(αx+y)=αh(x)+h(y)=αΓ(x)(h)+Γ(y)(h)=(Γ(x)+Γ(y)(h),

showing that Γ(αx+y)=αΓ(x)+Γ(y), and

Γ(xy)(h)=h(xy)=h(x)h(y)=Γ(x)(h)Γ(y)(h)=(Γ(x)Γ(y))(h),

showing that Γ(xy)=Γ(x)Γ(y). Therefore Γ:AC(Δ) is an algebra homomorphism. xkerΓ is equivalent to h(x)=0 for all hΔ, which is equivalent to xkerh for all hΔ. But by Theorem 2, {kerh:hΔ} is equal to the set of all maximal ideals of A, so xkerΓ is equivalent to xradA, i.e. kerΓ=radA.

Let xA. If λimx^ then there is some hΔ for which x^(h)=λ, and by Theorem 2, this yields λσ(x). Hence imx^σ(x). If λσ(x), then by Theorem 2 there is some hΔ for which h(x)=λ, i.e. there is some hΔ for which x^(h)=λ, i.e. λimx^. Hence σ(x)imx^. Therefore, imx^=σ(x).

It is straightforward to check that the topology of Δ is the subspace topology inherited from A* with the weak-* topology; in particular, the topology of Δ is Hausdorff. Therefore, to prove that Δ is compact it suffices to prove that Δ is a weak-* compact subset of A*. Let

K={λA*:λ1}.

By the Banach-Alaoglu theorem, K is a weak-* compact subset of A*. If hΔ, then because h is an algebra homomorphism A it follows that h1.44 4 Walter Rudin, Functional Analysis, second ed., p. 249, Theorem 10.7. Thus, ΔK. Therefore, to prove that Δ is compact it suffices to prove that Δ is a weak-* closed subset of A*.

Suppose that hiΔ is a net that weak-* converges to λA*. Then hi(e)λ(e), i.e. 1λ(e), so λ(e)=1. Thus λ0. Let x,yA. On the one hand, hi(xy)λ(xy), and on the other hand, hi(x)λ(x) and hi(y)λ(y), so hi(x)hi(y)λ(x)λ(y) and hence hi(xy)=hi(x)hi(y)λ(x)λ(y). Therefore, λ(xy)=λ(x)λ(y), and because λA* is linear, this shows that λ:A is an algebra homomorphism, and hence that λΔ. Therefore Δ is a weak-* closed subset of A*. ∎

If A is a commutative unital Banach algebra, the above theorem shows that Γ:AA^ is an algebra isomorphism if and only if radA={0}, i.e., Γ is an algebra isomorphism if and only if A is semisimple.

The above theorem tells us that if A is a commutative unital Banach algebra and xA, then imx^=σ(x). This gives us

x^=ρ(x), (1)

where ρ(x) is the spectral radius of x, defined by

ρ(x)=sup{|λ|:λσ(x)}.

Therefore, x^=0 is equivalent to ρ(x)=0, and so by the above theorem, xradA is equivalent to ρ(x)=0. Moreover, it is a fact that ρ(x)x.55 5 Walter Rudin, Functional Analysis, second ed., p. 253, Theorem 10.13. Therefore,

x^x. (2)

In the proof of Theorem 3 we used the fact66 6 Walter Rudin, Functional Analysis, second ed., p. 249, Theorem 10.7. that the norm of any algebra homomorphism from a Banach algebra to is 1. In particular, this means that any algebra homomorphism from a Banach algebra to is continuous. The following theorem shows that any algebra homomorphism from a Banach algebra to a commutative unital semisimple Banach algebra is continuous.77 7 Walter Rudin, Functional Analysis, second ed., p. 281, Theorem 11.10.

Theorem 4.

Suppose that A is a Banach algebra and that B is a commutative unital semisimple Banach algebra. If ψ:AB is an algebra homomorphism, then ψ is continuous.

Proof.

Because ψ:AB is linear, to prove that ψ is continuous, by the closed graph theorem88 8 Walter Rudin, Functional Analysis, second ed., p. 51, Theorem 2.15. it suffices to prove that

G={(x,ψ(x)):xA}

is closed in A×B. To prove that G is closed in A×B, it suffices to prove that if (xn,yn)G converges to (x,y)A×B then (x,y)G.

Let hΔB. Then ϕ=hψ:A is an algebra homomorphism. Because h:B and ϕ:A are algebra homomorphisms with codomain , they are both continuous. Therefore, h(yn)h(y) and ϕ(xn)ϕ(x). Therefore,

h(y)=limh(yn)=limh(ψ(xn))=lim(hψ)(xn)=limϕ(xn)=ϕ(x)=h(ψ(x)),

so h(y-ψ(x))=0. This is true for all hΔB, hence y-ψ(x)radB. But B is semisimple, so y-ψ(x)=0, i.e. y=ψ(x), so (x,y)G. ∎

If A is a commutative unital Banach algebra and xA, we recorded in (2) that x^x. The following lemma99 9 Walter Rudin, Functional Analysis, second ed., p. 282, Lemma 11.11. shows that if x2=x2 and x0, then infx^x1, hence that x^=x. Therefore, if x2=x2 for all xA, then Γ:AC(Δ) is an isometry.

Lemma 5.

Let A be a commutative unital Banach algebra. If

r=infx0x2x2,s=infx0x^x,

then s2rs.

Theorem 3 shows that if A is a commutative unital Banach algebra, then Γ:AC(Δ) is an algebra homomorphism. Therefore Γ(A)=A^ is a subalgebra of C(Δ). Moreover, Theorem 3 also shows that Δ is a compact Hausdorff space. Therefore, C(Δ) is a unital Banach algebra with the supremum norm. (If X is a topological space then C(X) is an algebra, but need not be a Banach algebra.) For A^ to be a Banach subalgebra of C(Δ) it is necessary and sufficient that A^ be a closed subset of the Banach algebra C(Δ). The following theorem gives conditions under which this occurs.1010 10 Walter Rudin, Functional Analysis, second ed., p. 282, Theorem 11.12.

Theorem 6.

If A is a commutative unital Banach algebra, then A is semisimple and A^ is a closed subset of C(Δ) if and only if there exists some K< such that x2Kx2 for all xA.

Proof.

Suppose that there is some 0<K< such that xA implies that x2Kx2. Then

r=infx0x2x2infx0x2Kx2=1K.

By Lemma 5, with s=infx0x^x we have

1Ks,

hence x^1Kx. Thus, if xA then x^1Kx, from which it follows that Γ:AC(Δ) is one-to-one. Since Γ is one-to-one, by Theorem 3 we get that A is semisimple. Suppose that x^nA^ converges to x^A^, i.e. x^n-x^0, i.e. Γ(xn-x)0. But Γ(xn-x)1Kxn-x, so xn-x0, showing that Γ-1:A^A is bounded. Therefore Γ:AA^ is bilipschitz, and so A^ is a complete metric space, from which it follows that A^ is a closed subset of C(Δ).

Suppose that A is semisimple and that A^ is a closed subset of C(Δ). The fact that A is semisimple gives us by Theorem 3 that Γ:AA^ is a bijection. The fact that A^ is closed means that A^ is a Banach algebra. Because Γ:AA^ is continuous, linear, and a bijection, by the open mapping theorem1111 11 Walter Rudin, Functional Analysis, second ed., p. 49, Corollary 2.12. it follows that there are positive real numbers a,b such that if xA then

axΓxbx.

Then infx0x^xa. By Lemma 5, it follows that infx0x2x2a2. Hence, for all x0 we have x2Kx2, with K=1a2. ∎

4 L1

Let M(n) denote the set of all complex Borel measures on n, and let S:n×nn be defined by S(x,y)=x+y. For μ1,μ2M(n), we denote by μ1×μ2 the product measure on n×n, and we define the convolution of μ1 and μ2 to be μ1*μ2=S*(μ1×μ2), the pushforward of μ1×μ2 with respect to S. That is, if E is a Borel subset of n, then

(μ1*μ2)(E) = (S*(μ1×μ2))(E)
= (μ1×μ2)(S-1(E))
= nnχE(x+y)𝑑μ1(x)𝑑μ2(y).

With convolution as multiplication, M(n) is an algebra.

If μM(n), the variation of μ is the measure |μ|M(n), where for a Borel subset E of n, we define |μ|(E) to be the supremum of Aπ|μ(A)| over all partitions π of E into finitely many disjoint Borel subsets. The total variation of μ is μ=|μ|(n). One proves that is a norm on M(n) and that with this norm, M(n) is a Banach algebra.1212 12 See Walter Rudin, Real and Complex Analysis, third ed., chapter 6.

Let mn be Lebesgue measure on n, let δ be the Dirac measure on n, and let A be the set of those μM(n) for which there is some fL1(n) and some α with which

dμ=fdmn+αdδ.

One proves that A is a Banach subalgebra of M(n). A is a unital Banach algebra, with unity δ. In particular, A is a unital Banach algebra that contains the Banach algebra L1(n).

If f+αδ,g+βδA (identifying fL1(n) with the complex Borel measure whose Radon-Nikodym derivative with respect to mn is f), then

(f+αδ)*(g+βδ)=(f*g+βf+αg)+αβδ, (3)

where

(f*g)(x)=nf(y)g(x-y)𝑑mn(y).

If tn, let et(x)=exp(itx), and if fL1(n), define f^:n, the Fourier transform of f, by

f^(t)=nfe-t𝑑mn,tn.

If tn, define ht:A by

ht(f+αδ)=f^(t)+α,f+αδA,

and define h:A by

h(f+αδ)=α,f+αδA.

By (3) it is apparent that for each tn{}, the map ht is a homomorphism of algebras. It can be proved that Δ={ht:tn}{h}.1313 13 Walter Rudin, Functional Analysis, second ed., p. 285. Let n{} be the one-point compactification of n, and define T:n{}Δ by T(t)=ht, which is a bijection.

Suppose that tkt in n. If f+αδA, then because f^:n is continuous, we have

T(tk)(f+αδ)=htk(f+αδ)=f^(tk)+αf^(t)+α=T(t)(f+αδ).

Suppose that tk. If f+αδA, then by the Riemann-Lebesgue lemma we have f^(tk)0, and hence

T(tk)(f+αδ)=f^(tk)+αα=h(f+αδ)=T()(f+αδ).

Therefore, T:n{}Δ is continuous.

Suppose that htkht in Δ, tk,tn. If f+αδA, then htk(f+αδ)ht(f+αδ). But htk(f+αδ)=f^(tk)+α and ht(f+αδ)=f^(t)+α, so f^(tk)f^(t). Because this is true for all fL1(n), it follows that tkt. Suppose that htkh in Δ, tkn. If f+αδA, then htk(f+αδ)h(f+αδ), i.e. f^(tk)+αα, i.e. f^(tk)0. Because this is true for all fL1(n), it follows that tk. Therefore, T-1:Δn{} is continuous, and so Δ is homeomorphic to the one-point compactification of n.

5 Involutions

If A is an algebra, an involution of A is a map :*AA satisfying

  1. 1.

    (x+y)*=x*+y*

  2. 2.

    (αx)*=α¯x*

  3. 3.

    (xy)*=y*x*

  4. 4.

    x**=x.

We say that x is self-adjoint if x*=x.

Following Rudin, if A is a Banach algebra with an involution :*AA satisfying

xx*=x2,xA,

we say that A is a B*-algebra.

The following theorem shows that a commutative unital B*-algebra with maximal ideal space Δ is isomorphic as a B*-algebra to C(Δ).1414 14 Walter Rudin, Functional Analysis, second ed., p. 289, Theorem 11.18. (An isomorphism of B*-algebras is an isomorphism of Banach algebras that preserves the involution; the involution on C(Δ) is (xf(x))(xf(x)¯).)

Theorem 7 (Gelfand-Naimark).

If A is a commutative unital B*-algebra, then Γ:AC(Δ) is an isomorphism of Banach algebras, and if xA then Γ(x*)=Γ(x)¯.

Proof.

Let uA be self-adjoint, let hΔ, and let h(u)=α+iβ. For t, put z=u+ite. We have

h(z)=h(u)+h(ite)=α+iβ+it=α+i(β+t),

and

zz*=(u+ite)(u-ite)=u2+t2e,

hence

α2+(β+t)2=|h(z)|2z2=zz*=u2+t2eu2+t2,

i.e.

α2+β2+2βtu2.

Because this is true for all t, it follows that β=0. Therefore, if uA is self-adjoint then h(u).

Furthermore, if xA then with 2u=x+x* and 2v=i(x*-x) we have x=u+iv with u and v self-adjoint. Then x*=u-iv, and so

h(x*)=h(u-iv)=h(u)-ih(v)=h(x)¯.

This shows that if xA then Γ(x*)=Γ(x)¯. In particular, A^ is closed under complex conjugation. If h1h2, then there is some xA for which h1(x)h2(x), i.e. x^(h1)x^(h2), so A^ separates points in Δ. Because A^ is a unital Banach algebra, it follows from the Stone-Weierstrass theorem that A^ is dense in C(Δ).

Let xA. With y=xx*, we have y*=(xx*)*=x**x*=xx*=y, from which it follows that y2=y2. Assume by induction that ym=ym, for m=2n. Then, as (ym)*=ym,

y2m=ymym=ym(ym)*=ym2=(ym)2=y2m.

The spectral radius formula1515 15 Walter Rudin, Functional Analysis, second ed., p. 253, Theorem 10.13. gives

ρ(y)=limyn1/n,

and because ym=ym for m=2n, we have limym1/m=y. Because the limit of this subsequence is y, the limit of yn1/n is also y, so we obtain

ρ(y)=y.

But (1) tells us y^=ρ(y), so we have y^=y. Because y=xx*, using Γ(x*)=Γ(x)¯ and the fact that Γ is an algebra homomorphism, we get

Γ(y)=Γ(xx*)=Γ(x)Γ(x*)=Γ(x)Γ(x)¯=|Γ(x)|2.

That is, y^=|x^|2 and with y^=y we obtain

x^2=y^=y=xx*=x2,

i.e.

x^=x.

This shows that Γ:AC(Δ) is an isometry. In particular, Γ maps closed sets to closed sets, so A^=Γ(A) is a closed subset of C(Δ). We have already established that A^ is dense in C(Δ), so A^=C(Δ). The fact that Γ is an isometry yields that Γ is one-to-one, and the fact that A^=C(Δ) means that that Γ is onto, hence Γ is a bijection, and therefore it is an isomorphism of algebras. Because Γ is an isometry, it is an isomorphism of Banach algebras. ∎

The following theorem states conditions under which a self-adjoint element of a unital Banach algebra with an involution has a square root.1616 16 Walter Rudin, Functional Analysis, second ed., p. 294, Theorem 11.26.

Theorem 8.

Let A be a unital Banach algebra with an involution :*AA. If xA is self-adjoint and σ(x) contains no real λ with λ0, then there is some self-adjoint yA satisfying y2=x.

If A is a Banach algebra and xA, we say that xA is normal if xx*=x*x. If A is a Banach algebra with involution :*AA, by x0 we mean that x is self-adjoint and σ(x)[0,), and we say that x is positive. The following theorem states basic facts about the spectrum of elements of a unital B*-algebra.1717 17 Walter Rudin, Functional Analysis, second ed., p. 294, Theorem 11.28.

Theorem 9.

If A is a unital B*-algebra, then:

  1. 1.

    If x is self-adjoint, then σ(x).

  2. 2.

    If x is normal, then ρ(x)=x.

  3. 3.

    If xA, then ρ(xx*)=x2.

  4. 4.

    If x0 and y0, then x+y0.

  5. 5.

    If xA, then xx*0.

  6. 6.

    If xA, then e+xx* is invertible.

6 Positive linear functionals

Suppose that A is a Banach algebra with an involution :*AA. If F:A is a linear map such that F(xx*) is real and 0 for all xA, we say that F is a positive linear functional. In particular, if hΔ and xA, then from Theorem 7 we have h(x*)=h(x)¯, and so h(xx*)=h(x)h(x*)=h(x)h(x)¯=|h(x)|20. Thus, the elements of Δ are positive linear functionals.

We shall use the following theorem to prove the theorem after it.1818 18 Walter Rudin, Functional Analysis, second ed., p. 137, Theorem 5.20.

Theorem 10.

If X is a real or complex Banach space, X1 and X2 are closed subspaces of X, and X=X1+X2, then there is some γ< such that for every xX there are x1X1,x2X2 satisfying x=x1+x2 and

x1+x2γx.

The following theorem establishes some basic properties of positive linear functionals on a unital Banach algebra with an involution.1919 19 Walter Rudin, Functional Analysis, second ed., p. 296, Theorem 11.31.

Theorem 11.

Suppose that A is a unital Banach algebra with an involution :*AA. If F:A is a positive linear functional, then:

  1. 1.

    F(x*)=F(x)¯.

  2. 2.

    |F(xy*)|2F(xx*)F(yy*).

  3. 3.

    |F(x)|2F(e)F(xx*)F(e)2ρ(xx*).

  4. 4.

    If x is normal, then |F(x)|F(e)ρ(x).

  5. 5.

    If A is commutative, then F=F(e).

  6. 6.

    If there is some β such that x*βx for all xA, then Fβ1/2F(e).

  7. 7.

    F is a bounded linear map.

Proof.

Suppose that x,yA. For any α, we have on the one hand F((x+αy)(x+αy)*)0, and on the other hand

F((x+αy)(x+αy)*)=F((x+αy)(x*+α¯y*))=F(xx*+α¯xy*+αyx*+|α|2yy*).

Therefore,

F(xx*)+α¯F(xy*)+αF(yx*)+|α|2F(yy*)0. (4)

Applying (4) with α=1 gives

F(xx*)+F(xy*)+F(yx*)+F(yy*)0.

In particular, this expression is real, and because F(xx*) and F(yy*) are real we get that F(xy*)+F(yx*) is real, so ImF(yx*)=-ImF(xy*). Applying (4) with α=i gives

F(xx*)-iF(xy*)+iF(yx*)+F(yy*)0.

In particular, this expression is real, and so -iF(xy*)+iF(yx*) is real, i.e. F(xy*)-F(yx*) is imaginary, so ReF(yx*)=ReF(xy*). Therefore F(yx*)=F(xy*)¯. Using y=e yields

F(x*)=F(x)¯.

Suppose that x,yA and that F(xy*)0. For any t, using (4) with α=t|F(xy*)|F(xy*) gives

F(xx*)+t|F(xy*)|F(xy*)¯F(xy*)+t|F(xy*)|F(xy*)F(yx*)+t2F(yy*)0,

i.e.

F(xx*)+t|F(xy*)|+t|F(xy*)|F(xy*)F(yx*)+t2F(yy*)0,

and as F(yx*)=F((xy*)*)=F(xy*)¯, we have

F(xx*)+2t|F(xy*)|+t2F(yy*)0.

For t=-|F(xy*)|F(yy*) this is

F(xx*)-2|F(xy*)|2F(yy*)+|F(xy*)|2F(yy*)0,

i.e.

|F(xy*)|2F(xx*)F(yy*).

Suppose that xA. Because xe*=x and ee*=e, we have

|F(x)|2F(e)F(xx*).

We shall prove that F(xx*)F(e)ρ(xx*). Let t>ρ(xx*). It then follows that σ(te-xx*) is contained in the open right half-plane, and thus by Theorem 8 there is some self-adjoint uA satisfying u2=te-xx*. Then

F(te-xx*)=F(u2)=F(uu*)0,

so

F(xx*)tF(e).

Because this is true for all t>ρ(xx*), we obtain

F(xx*)F(e)ρ(xx*).

Suppose that x is normal. It is a fact that if x and y belong to a unital Banach algebra and xy=yx, then σ(xy)σ(x)σ(y).2020 20 Walter Rudin, Functional Analysis, second ed., p. 293, Theorem 11.23. Thus σ(xx*)σ(x)σ(x*), from which we get

ρ(xx*)ρ(x)ρ(x*).

It is a fact that σ(x*)=σ(x)¯,2121 21 Walter Rudin, Functional Analysis, second ed., p. 288, Theorem 11.15., so we have ρ(x)=ρ(x*), and thus

ρ(xx*)ρ(x)2.

But |F(x)|2F(e)2ρ(xx*), so we have |F(x)|2F(e)2ρ(x)2, i.e.

|F(x)|F(e)ρ(x).

Suppose that A is commutative, and let xA. Since A is commutative, x is normal and hence we have |F(x)|F(e)ρ(x), and as always we have ρ(x)x. Therefore, for every xA we have

|F(x)|F(e)x.

This implies that FF(e), and because the above inequality is an equality for x=e, we have F=F(e).

Suppose that there is some β such that x*βx for all xA. We have ρ(xx*)xx*xx*βx2. (We merely stipulated that A is a unital Banach algebra with an involution; if we had demanded that A be a B*-algebra, then we would have xx*=xx*=x2.) Using |F(x)|2F(e)2ρ(xx*) then gives us |F(x)|2βF(e)2x2, hence

|F(x)|β1/2F(e)x.

If F(e)=0, then |F(x)|20 for all xA, and hence F=0, which indeed is bounded. Otherwise, F(e)>0, and F is bounded if and only if 1F(e)F is bounded. Therefore, to prove that F is bounded it suffices to prove that F is bounded in the case where F(e)=1.

Let H be the set of all self-adjoint elements of A. H and iH are real vector spaces. For any xA, defining 2u=x+x* and 2v=i(x*-x), we have x=u+iv, and u,v are self-adjoint. It follows that

A=H+iH.

Because the elements of H are self-adjoint, the restriction of F to H is a real-linear map H. For uH, because u is self-adjoint it is in particular normal, and so |F(x)|F(e)ρ(x)F(e)x=x, because F(e)=1. Hence the restriction of F to H is a real-linear map H with norm 1, and therefore there is a unique bounded real-linear map Φ:H¯ whose restriction to H is equal to the restriction of F to H, and Φ=1.

Suppose that yH¯iH¯. There are unH with uny and there are vnH with ivny. Then un2y2 and -vn2y, or vn2-y2. Because |F(un)|2F(e)F(unun*)=F(un2), we have

|F(un)|2F(un2)F(un2+vn2).

Because un and vn are self-adjoint, un2+vn2 is normal and hence

|F(un2+vn2)|F(e)ρ(un2+vn2)=ρ(un2+vn2)un2+vn2,

and so we have

|F(un)|2un2+vn2.

But un2y and vn2-y, so un2+vn2y-y=0. Therefore, F(un)0, and so

Φ(y)=limF(un)0.

That is, if yH¯iH¯, then F(y)=0.

Because A=H+iH, certainly A=H¯+iH¯, so by Theorem 10 there is some γ< such that for all xA, there are x1H¯ and x2H¯ satisfying

x=x1+ix2,x1+x2γx.

Let xA and let x=x1+ix2, where x1,x2 satisfy the above, and let x=u+iv with u,vH, namely 2u=x+x* and 2v=i(x*-x). Supposing that x1-u and x2-vH¯iH¯, which Rudin asserts but whose truth is not apparent to me, we obtain F(x1-u)=0 and F(x2-v)=0, or F(x1)=F(u) and F(x2)=F(v). Then,

F(x)=F(u+iv)=F(u)+iF(v)=F(x1)+iF(x2)=Φ(x1)+iΦ(x2),

and therefore, because Φ=1 and because x1+x2γx,

|F(x)||Φ(x1)+iΦ(x2)||Φ(x1)|+|Φ(x2)|x1+x2γx,

showing that Fγ, and in particular that F is bounded. ∎

7 The Riesz-Markov theorem and extreme points

We say that a positive Borel measure μ on a compact Hausdorff space X is regular if for every Borel subset E of X we have

μ(E)=sup{μ(F):F is compact and FE}

and

μ(E)=inf{μ(G):G is open and EG}.

We say that a complex Borel measure μ on a compact Hausdorff space is regular if the positive Borel measure |μ| is regular, and we write μ=|μ|(X). The following is the Riesz-Markov theorem, stated for complex Borel measures on a compact Hausdorff space.2222 22 Walter Rudin, Real and Complex Analysis, third ed., p. 130, Theorem 6.19.

Theorem 12 (Riesz-Markov).

Suppose that X is a compact Hausdorff space. If Λ is a bounded linear functional on C(X), then there is one and only one regular complex Borel measure μ on X satisfying

Λf=Xf𝑑μ,fC(X).

This measure μ satisfies μ=Λ.

The following theorem uses the Riesz-Markov theorem to define a correspondence between positive linear functionals on a commutative unital Banach algebra with a symmetric involution and regular positive Borel measures on its maximal ideal space.2323 23 Walter Rudin, Functional Analysis, second ed., p. 299, Theorem 11.33.

Theorem 13.

Suppose that A is a commutative unital Banach algebra with an involution :*AA satisfying

h(x*)=h(x)¯,xA,hΔ. (5)

Let K be the set of all positive linear functionals F:A satisfying F(e)1, and let M be the set of all regular positive Borel measures μ on Δ satisfying μ(Δ)1. K and M are convex sets. If μM, then F:A defined by

Fμ(x)=Δx^𝑑μ,xA,

belongs to K, and this map μFμ is an isometric bijection MK.

Proof.

If F1,F2K and 0t1, then (1-t)F1+tF2 is linear, and it is straightforward to check that it is positive. Moreover, ((1-t)F1+tF2)(e)=(1-t)F1(e)+tF2(e)(1-t)+t=1, so (1-t)F1+tF2K. Therefore K is a convex set.

Suppose that μ1,μ2M, that a1,a2 are nonnegative real numbers, and let μ=a1μ1+a2μ2. If E is a Borel subset of Δ, then for any ϵ>0 there are compact subsets F1,F2 of Δ such that μ1(E)<μ1(F1)-ϵ and μ2(E)<μ2(F2)-ϵ. With F=F1F2, we have

μ(F) = a1μ1(F)+a2μ2(F)
a1μ1(F1)+a2μ2(F2)
a1(μ1(E)+ϵ)+a2(μ2(E)+ϵ)
= μ(E)+(a1+a2)ϵ.

It follows that μ(E)=sup{μ(F):F is compact and FE}. If E is a Borel subset of Δ, then for any ϵ>0 there are open subsets G1,G2 of Δ such that μ1(E)>μ1(G1)-ϵ and μ2(E)>μ2(G2)-ϵ. With G=G1G2, we have

μ(G) = a1μ1(G)+a2μ2(G)
a1μ1(G1)+a2μ2(G2)
< a1(μ1(E)+ϵ)+a2(μ2(E)+ϵ)
= μ(E)+(a1+a2)ϵ.

It follows that μ(E)=inf{μ(G):G is open and EG}. Therefore, μ=a1μ1+a2μ2 is a regular positive Borel measure. In particular, if 0t1 and a1=1-t, a2=t, then μ is a regular positive Borel measure. Finally, for μ=(1-t)μ1+tμ2, 0t1, we have, because μ1(Δ)1 and μ2(Δ)1,

μ(Δ)=(1-t)μ1(Δ)+tμ2(Δ)(1-t)+t=1,

so μM, showing that M is a convex set.

Let μM. It is apparent that Fμ:A is linear. For xA, we have Γ(xx*)=Γ(x)Γ(x*), and as Γ(x*)=Γ(x)¯ by (5), we get Γ(xx*)=|Γ(x)|2. As |Γ(x)|2(h)0 for all hΔ, we have

Fμ(xx*)=ΔΓ(xx*)𝑑μ=Δ|Γ(x)|2𝑑μ0,

showing that Fμ is a positive linear functional. Furthermore, e^(h)=h(e)=1 for all hΔ, so

Fμ(e)=μ(Δ)1,

showing that FμK.

If xradA, then ρ(x)=0 by (1), and so F(x)=0 by Theorem 11. We define F^:A^

F^(x^)=F(x);

this makes sense because if x^=y^ then Γ(x-y)=0, and so by Theorem 3 we have x-yradA and hence F(x-y)=0, i.e. so F(x)=F(y). For xA, x is normal because A is commutative so we have by Theorem 11 that

|F^(x^)|=|F(x)|F(e)ρ(x)

and by (1) we have ρ(x)=x^, so

|F^(x^)|F(e)x^.

As F^(e^)=F(e), it follows that F^=F(e). By (5) and because A^ separates points in Δ, applying the Stone-Weierstrass we obtain that A^ is dense in C(Δ). Because F^ is a continuous linear functional on the dense subspace A^ of C(Δ), there is a unique continuous linear functional Λ on C(Δ) such that Λ=F^ on A^, and Λ=F^. Applying Theorem 12, there is one and only one regular complex Borel measure μ on X that satisfies

Λf=Δf𝑑μ,fC(Δ), (6)

and μ=Λ=F^=F(e). It follows that μFμ is one-to-one. Because e^(h)=1 for all hΔ,

μ(Δ)=ΔχΔ𝑑μ=Δe^𝑑μ=Λe^=F^(e^)=F(e)=μ=|μ|(Δ).

The fact that μ(Δ)=|μ|(Δ) implies that μ is a positive measure. The above equalities also state μ(Δ)=F(e), and since FK we have F(e)1, hence μ(Δ)1. Therefore, μM. For xA, as x^C(Δ) we have by (6) that

Fμ(x)=Δx^𝑑μ=Λx^=F^(x^)=F(x),

showing that F=Fμ. This shows that μFμ is onto, and therefore μFμ is a bijection MK. ∎

Because the map μFμ in the above theorem is an isometric bijection MK, it follows that that μ is an extreme point of M if and only if Fμ is an extreme point of K.

It is a fact that the set of extreme points of the set of regular Borel probability measures on a compact Hausdorff space X is {δx:xX}.2424 24 Barry Simon, Convexity: An Analytic Viewpoint, p. 128, Example 8.16. Given this, one proves that the set of extreme points of M is {0}{δh:hΔ}. For xA, F0(x)=0, i.e. F0=0. For hΔ and xA,

Fδh(x)=Δx^𝑑δh=x^(h)=h(x),

so Fδh=h. Therefore, the extreme points of K are {0}Δ, that is, the set of algebra homomorphisms A.

Corollary 14.

Suppose that A is a commutative unital Banach algebra with involution :*AA satisfying

h(x*)=h(x)¯,xA,hΔ.

If K is the set of all positive linear functionals F:A satisfying F(e)1, then

extK={0}Δ.

Moreover, it is straightforward to check that the set K in the above corollary is a weak-* closed subset of A*: if FiK is a net that weak-* converges to ΛA*, one checks that Λ(xx*)0 for all xA and that Λe1. By the Banach-Alaoglu theorem, the set B={ΛA*:Λ1} is weak-* compact, and if FK then F=F(e) by Theorem 11 and F(e)1, so KB. Hence, K is a weak-* compact subset of A*. Therefore, the Krein-Milman theorem2525 25 Walter Rudin, Functional Analysis, second ed., p. 75, Theorem 3.23. tells us that K is equal to the weak-* closure of the convex hull of the set of its extreme points, and by the above corollary this means that K is equal to the weak-* closure of the convex hull of {0}Δ.

8 Positive definite functions

A function ϕ:n is said to be positive-definite if r1, x1,,xrn,c1,,cr imply that

i,j=1rcicj¯ϕ(xi-xj)0;

in particular, for ϕ to be positive-definite demands that the left-hand side of this inequality is real.

A positive-definite function need not be measurable. For example, is a vector space over , and if ψ: is a vector space automorphism of over , one proves that xeiψ(x) is a positive-definite function , and that there are ψ for which xeiψ(x) is not measurable.

The following theorem states some basic facts about positive-definite functions. More material on positive-definite functions is presented in Bogachev; for example, if ϕ:n is a measurable positive-definite function, then there is a continuous positive-definite function n that is equal to ϕ almost everywhere.2626 26 Vladimir I. Bogachev, Measure Theory, vol. 1, p. 221, Theorem 3.10.20. See also Anthony W. Knapp, Basic Real Analysis, p. 406.

Theorem 15.

If ϕ:n is positive-definite, then

ϕ(0)0, (7)

and for all xn we have

ϕ(x)¯=ϕ(-x) (8)

and

|ϕ(x)|ϕ(0). (9)
Proof.

Using r=1 and c1=1, we have for all x1n that ϕ(x1-x1)0, i.e. ϕ(0)0.

Using r=2, for x1,x2n and c1,c2 we have

c1c1¯ϕ(x1-x1)+c1c2¯ϕ(x1-x2)+c2c1¯ϕ(x2-x1)+c2c2¯ϕ(x2-x2)0.

Take x1=x and x2=0, with which

|c1|2ϕ(0)+c1c2¯ϕ(x)+c2c1¯ϕ(-x)+|c2|2ϕ(0)0;

in particular, the left-hand side is real, and because ϕ(0) is real by (7), this implies that c1c2¯ϕ(x)+c2c1¯ϕ(-x) is real. That is, it is equal to its complex conjugate:

c1c2¯ϕ(x)+c2c1¯ϕ(-x)=c1¯c2ϕ(x)¯+c2¯c1ϕ(-x)¯.

The fact that this holds every c1,c2 implies that ϕ(x)¯=ϕ(-x).

Again using that

c1c1¯ϕ(x1-x1)+c1c2¯ϕ(x1-x2)+c2c1¯ϕ(x2-x1)+c2c2¯ϕ(x2-x2)0,

with x1=x,x2=0 and c2=1 we get

|c1|2ϕ(0)+c1ϕ(x)+c1¯ϕ(-x)+ϕ(0)0.

Applying (8) gives

|c1|2ϕ(0)+c1ϕ(x)+c1ϕ(x)¯+ϕ(0)0.

For c1 such that |c1|=1,

2ϕ(0)+2Re(c1ϕ(x))0,

or

-Re(c1ϕ(x))ϕ(0).

Thus, taking c1 such that |c1|=1 and for which -Re(c1ϕ(x))=|ϕ(x)|, we get |ϕ(x)|ϕ(0). ∎

The following lemma about positive-definite functions follows a proof in Bogachev.2727 27 Vladimir I. Bogachev, Measure Theory, vol. 1, p. 221, Lemma 3.10.19.

Lemma 16.

If ϕ:n is a measurable positive-definite function and fL1(n) is nonnegative, then

nnϕ(x-y)f(x)f(y)𝑑mn(x)𝑑mn(y)0.
Proof.

For r2 and for any x1,,xr and c1=1,,cr=1, we have

j,k=1rϕ(xj-xk)0,

or

rϕ(0)+jkϕ(xj-xk)0.

By (9), ϕ is bounded. It follows that we can integrate both sides of the above inequality over (n)r with respect to the positive measure

f(x1)f(xr)dmn(x1)dmn(xr).

Writing

I=nf(x)𝑑mn(x),

we obtain

rϕ(0)Ir+jknnϕ(xj-xk)f(x1)f(xr)𝑑mn(x1)𝑑mn(xr)0,

and so, writing

J=nnϕ(x-y)f(x)f(y)𝑑mn(x)𝑑mn(y),

we have

rϕ(0)Ir+jkJIr-20,

or

rϕ(0)Ir+r(r-1)JIr-20.

If I=0, then because f is nonnegative it follows that f is 0 almost everywhere, in which case J=0, so the claim is true. If I>0, then dividing by r(r-1)Ir-2 we obtain

1r-1ϕ(0)I2+J0.

This inequality holds for all r2, so taking r yields

J0,

which is the claim. ∎

For f,gL1(n),

(f*g)(x)=nf(y)g(x-y)𝑑mn(y).

The support of a function f:n, denoted suppf, is the closure of the set {xn:f(x)0}. We denote by Cc(n) the set of all continuous functions f:n such that suppf is a compact set. It is straightforward to check that an element of Cc(n) is uniformly continuous on n. The following theorem is similar to the previous lemma, but applies to functions that need not be nonnegative.2828 28 Gerald B. Folland, A Course in Abstract Harmonic Analysis, p. 85, Proposition 3.35.

Theorem 17.

For f:n, define f~:n by f~(x)=f(-x)¯. If ϕ:n is a continuous positive-definite function, then for all fCc(n), we have

n(f*f~)ψ0.

If μ is a complex Borel measure on n, the Fourier transform of μ is the function μ^:n defined by

μ^(ξ)=ne-ξ𝑑μ,ξn.

One proves using the dominated convergence theorem that μ^ is continuous.

Theorem 18.

If μ is a finite positive Borel measure on n, then μ^:n is positive-definite.

Proof.

For ξ1,,ξrn and c1,,cr, we have

j,k=1rcjck¯μ^(ξj-ξk) = j,k=1rcjck¯ne-i(ξj-ξk)x𝑑μ(x)
= nj,k=1rcje-iξjxcke-iξkx¯dμ(x)
= n(j=1rcje-iξjx)(k=1rcke-iξkx)¯𝑑μ(x)
= n|j=1rcje-iξjx|2𝑑μ(x)
0.

The following proof of Bochner’s theorem follows an exercise in Rudin.2929 29 Walter Rudin, Functional Analysis, second ed., p. 303, Exercise 14. Other references on Bochner’s theorem are the following: Barry Simon, Convexity: An Analytic Viewpoint, p. 153, Theorem 9.17; Edwin Hewitt and Kenneth A. Ross, Abstract Harmonic Analysis, vol. II, p. 293, Theorem 33.3; Mark A. Pinsky, Introduction to Fourier Analysis and Wavelets, p. 220, Theorem 3.9.16; Walter Rudin, Fourier Analysis on Groups, p. 19, Theorem 1.4.3; Yitzhak Katznelson, An Introduction to Harmonic Analysis, third ed., p. 170; Vladimir I. Bogachev, Measure Theory, vol. II, p. 121, Theorem 7.13.1; Gerald B. Folland, A Course in Abstract Harmonic Analysis, p. 95, Theorem 4.18.

Theorem 19 (Bochner).

If ϕ:n is continuous and positive-definite, then there is some finite positive Borel measure ν on n for which ϕ=ν^.

Proof.

Let A be the Banach algebra defined in §4, whose elements are those complex Borel measures μ on n for which there is some fL1(n) and some α such that

dμ=fdmn+αdδ,

where mn is Lebesgue measure on n. For f+αδ,g+βδA, we have

(f+αδ)*(g+βδ)=(f*g+βf+αg)+αβδ;

we are identifying fL1(n) with the complex Borel measure whose Radon-Nikodym derivative with respect to mn is f. The norm on A is the total variation norm of a complex measure; one checks that for f+αδ this is

f+αδ=f+|α|,

where f=n|f(x)|𝑑mn(x).

For fL1(n), we define f~L1(n) by f~(x)=f(-x)¯, and we define :*AA by

(f+αδ)*=f~+α¯δ,f+αδA.

On the one hand,

((f+αδ)*(g+βδ))* = ((f*g+βf+αg)+αβδ)*
= f*g~+β¯f~+α¯g~+αβ¯δ
= f*g~+β¯f~+α¯g~+αβ¯δ,

and

f*g~(x)=nf(y)g(-x-y)𝑑mn(y)¯.

On the other hand,

(g+βδ)**(f+αδ)* = (g~+β¯δ)*(f~+α¯δ)
= (g~*f~+α¯g~+β¯f~)+βα¯δ,

and

(g~*f~)(x) = ng~(y)f~(x-y)𝑑mn(y)
= ng(-y)¯f(-x+y)¯𝑑mn(y)
= ng(-y-x)f(y)𝑑mn(y)¯.

Therefore we have

((f+αδ)*(g+βδ))*=(g+βδ)**(f+αδ)*.

Thus :*AA is an involution (the other properties demanded of an involution are immediate).

We define F:A by

F(f+αδ)=nfϕ𝑑mn+αϕ(0),f+αδA.

It is apparent that F is linear, and because |ϕ(x)|ϕ(0) for all x,

|F(f+αδ)| |nfϕ𝑑mn|+|α|ϕ(0)
n|f||ϕ|𝑑mn+|α|ϕ(0)
ϕ(0)n|f|𝑑mn+|α|ϕ(0)
= ϕ(0)f+αδ,

from which it follows that F=ϕ(0), and in particular that F is bounded. Let A0={f+αδA:fCc(n)}. Because F:A is bounded and A0 is a dense subset of A, to prove that F is a positive linear functional it suffices to prove that for all f+αδA0 we have F((f+αδ)*(f+αδ)*)0.

For gCc(n), by Theorem 17 we obtain

F(g*g*)=F(g*g~)=n(g*g~)ϕ𝑑mn0. (10)

Define η:n by

η(x)={exp(-11-|x|2)|x|<10|x|1,

and for ϵ>0, define ηϵ:n by ηϵ(x)=ϵ-nη(xϵ). Let f+αδA0 and define gϵ=f+αηϵCc(n). From (10) we have F(gϵ*gϵ*)0 for any ϵ>0. On the other hand,

F(gϵ*gϵ*) = F((f+αηϵ)*(f~+α¯ηϵ))
= F(f*f~+α¯f*ηϵ+αηϵ*f~+|α|2ηϵ*ηϵ)
= F(f*f~)+α¯n(f*ηϵ)ϕ𝑑mn+αn(ηϵ*f~)ϕ𝑑mn
+|α|2n(ηϵ*ηϵ)ϕ𝑑mn.

We take as granted that

n(f*ηϵ)ϕ𝑑mnnfϕ𝑑mn

as ϵ0, that

n(ηϵ*f~)ϕ𝑑mnnf~ϕ𝑑mn

as ϵ0, and that

n(ηϵ*ηϵ)ϕ𝑑mnϕ(0)

as ϵ0. Furthermore,

F((f+αδ)*(f+αδ)*) = F((f+αδ)*(f~+α¯δ))
= F(f*f~+α¯f+αf~+|α|2)

Thus

F(gϵ*gϵ*)F((f+αδ)*(f+αδ)*)

as ϵ0. Since F(gϵ*gϵ*)0 for all ϵ>0, it follows that

F((f+αδ)*(f+αδ)*)0.

Therefore, F:A is a positive linear functional.

Because F is a positive linear functional and F(e)=F(δ)=1, we can apply Theorem 12, according to which there is a regular positive Borel measure μ on Δ satisfying

F(f+αδ)=ΔΓ(f+αδ)𝑑μ,f+αδA,

and hence, from the definition of F,

nfϕ𝑑mn+αϕ(0)=ΔΓ(f+αδ)𝑑μ,f+αA.

We state the following again from §4 for easy access. If tn, define ht:A by

ht(f+αδ)=f^(t)+α,f+αδA,

and also define h:A by

h(f+αδ)=α,f+αδA.

Let n{} be the one-point compactification of n. We proved in §4 that the map T:n{}Δ defined by T(t)=ht is a homeomorphism. With ν=(T-1)*μ we have μ=T*ν, and then

ΔΓ(f+αδ)𝑑μ = ΔΓ(f)𝑑μ+αΔΓ(δ)𝑑μ
= ΔΓ(f)d(T*ν)+αΔχΔ𝑑μ
= n{}Γ(f)T𝑑ν+αμ(Δ)
= n{}Γ(f)T(t)𝑑ν(t)+αF(δ)
= n{}ht(f)𝑑ν(t)+αϕ(0)
= nf^(t)𝑑ν(t)+αϕ(0).

Therefore

nfϕ𝑑mn+αϕ(0)=nf^(t)𝑑ν(t)+αϕ(0),

i.e.

nfϕ𝑑mn=nf^(t)𝑑ν(t).

As

nf^(t)𝑑ν(t) = n(ne-itxf(x)𝑑mn(x))𝑑ν(t)
= nf(x)ne-ixt𝑑ν(t)𝑑mn(x)
= nf(x)ν^(x)𝑑mn(x),

we have

nfϕ𝑑mn=nfν^𝑑mn.

This is true for all fL1(n), from which it follows that ϕ=ν^. ∎