Locally compact abelian groups

Jordan Bell
June 4, 2014

1 Introduction

These notes are a gloss on the first chapter of Walter Rudin’s Fourier Analysis on Groups, and may be helpful to someone reading Rudin. The results I do prove are proved in more detail than they are in Rudin. I caution that before reading the first chapter of that book it is know about the Gelfand transform on commutative Banach algebras because results from that are used without even stating them by Rudin. I at least state them.

2 Locally compact abelian groups

Let G be a locally compact abelian (LCA) group. There is a Haar measure m on G. It is a fact that if m and m are Haar measures on G, then there is a positive constant λ such that m=λm.11 1 Walter Rudin, Fourier Analysis on Groups, p. 2, §1.1.3. If G is compact then there is a unique Haar measure with m(G)=1, and if G is discrete then there is a unique Haar measure with m({x})=1 for each xG.

Because Haar measures on G are positive multiples of each other, the elements of Lp(G) do not depend on the Haar measure we use, but the value of fp will, where f:G is a Borel function.

If X is a normed vector space and f:GX is a function, we say that f is uniformly continuous if for every ϵ>0 there is a neighborhood V of 0 in G such that x-yV implies that f(x)-f(y)X<ϵ.

If f:G is a function and xG, we define fx:G by

fx(y)=f(y-x),yG.
Theorem 1.

If 1p< and fLp(G), then

xfx

is uniformly continuous GLp(G).

Proof.

Let ϵ>0. Because Cc(G) is dense in Lp(G), there is some gCc(G) such that g-fp<ϵ. Let K=suppg, and because K is compact, m(K)<. gCc(G) implies that g is uniformly continuous on G, so there is a neighborhood V of 0 in G such that if x-yV then |g(x)-g(y)|<ϵ(2m(K))-1/p. Then, for all xV and yG, as y-(y-x)V we have |g(y-x)-g(y)|<ϵ(2m(K))-1/p. That is, for all xV we have

gx-g<ϵ(2m(K))-1/p.

For xV, supp(gx-g)(K+x)K and m(K+x)=m(K), so for all xV,

gx-gp = (supp(gx-g)|gx-g|p𝑑m)1/p
(supp(gx-g)ϵp(2m(K))-1𝑑m)1/p
ϵ.

Because fx-gxp=f-gp, for all xV we have

fx-fpfx-gxp+gx-gp+g-fp<3ϵ.

Then, let x,yG with y-xV. The above inequality tells us

fy-x-fp<3ϵ.

But fx-fy=(f-fy-x)x and hxp=hp, so

fx-fyp=f-fy-xp<3ϵ,

showing that xfx is uniformly continuous. ∎

If f and g are Borel functions on G, for xG such that the integral exists we define

(f*g)(x)=Gf(x-y)g(y)𝑑m(y).

The operation of taking the convolution of functions is particularly suitable for functions that belong to L1(G), because if f,gL1(G), then (f*g)(x) is defined for almost all xG, and satisfies

f*g1f1g1;

this is proved in Rudin, together with other properties of the convolution.22 2 Walter Rudin, Fourier Analysis on Groups, p. 4, §1.1.6. From the above inequality, it follows that L1(G) with convolution as multiplication is a commutative Banach algebra. The map :*L1(G)L1(G) defined by f*(x)=f(-x)¯ for fL1(G), xG, is an isometric involution.

If G is discrete, define e on G by e(0)=1 and e(x)=0 for x0. Then e1=Ge(x)𝑑m(x)=xGe(x)=1, so eL1(G). For fL1(G) and xG,

(f*e)(x)=yGf(x-y)e(y)=f(x)e(0)=f(x),

showing that f*e=f. Hence e is unity in L1(G). It turns out that if G is not discrete then L1(G) does not have a unity.33 3 Walter Rudin, Fourier Analysis on Groups, p. 30, §1.7.3.

3 Dual groups

Let 𝕋={z:|z|=1} with the subspace topology inherited from . If G is a locally compact abelian group, we denote by Γ the set of continuous homomorphisms G𝕋. For γΓ and xG, we write

x,γ=γ(x).

We call a homomorphism G𝕋 a character of G, so elements of Γ are the continuous characters of G.

If A is a Banach algebra and Δ is the set of algebra homomorphisms h:A that are not identically zero, the Gelfand transform of xA is the map x^:Δ defined by

x^(h)=h(x),hΔ.

Δ is called the maximal ideal space of A.

If fL1(G), we define f^:Γ by

f^(γ)=Gf(x)-x,γ𝑑m(x),γΓ,

and call f^ the Fourier transform of f.

The following theorem is from Rudin.44 4 Walter Rudin, Fourier Analysis on Groups, p. 7, §1.2.2. It establishes a bijection between the dual group Γ of G and the maximal ideal space Δ of L1(G), and shows, if Γ and Δ are identified, that the Fourier transform is the same as the Gelfand transform.

Theorem 2.

For each γΓ, the map h:L1(G)C defined by h(f)=f^(γ) belongs to Δ. If hΔ, then there is a unique γΓ such that h(f)=f^(γ) for all fL1(G).

Proof.

For f,gL1(G) and k=f*g, then for any γΓ, using -x,γ=-x+y,γ-y,γ,

k^(γ) = G(f*g)(x)-x,γ𝑑m(x)
= G-x,γ(Gf(x-y)g(y)𝑑m(y))𝑑m(x)
= Gg(y)-y,γ(Gf(x-y)-x+y,γ𝑑m(x))𝑑m(y)
= Gg(y)-y,γf^(x)𝑑m(y)
= f^(γ)g^(γ),

showing that ff^(γ) is an algebra homomorphism. By Urysohn’s lemma, for every neighborhood V of 0G, there is some fCc(G) with f(0)=1, 0f1, and suppfV. As 0,γ=1 and because x-x,γ is continuous, it follows that there is some neighborhood V of 0 and some f as above such that f^(γ)=Gf(x)-x,γ𝑑m(x)0, showing that ff^(γ) is nonzero.

Now let h:L1(G) be a nonzero algebra homomorphism. A homomorphism from a Banach algebra to is linear functional with norm 1. It is a fact that for every bounded linear functional Λ on L1(G) there is some ϕL(G) such that Λ(f)=Gfϕ𝑑m for all fL1(G), and ϕ=Λ.55 5 Walter Rudin, Fourier Analysis on Groups, p. 268, E10. Therefore, there is some ϕL(G), ϕ=1, such that

h(f)=Gfϕ𝑑m,fL1(G).

Then for f,gL1(G) we have

h(f)Ggϕ𝑑m = h(f)h(g)
= h(f*g)
= G(f*g)ϕ𝑑m
= G(Gf(x-y)g(y)𝑑m(y))ϕ(x)𝑑m(x)
= Gg(y)(Gf(x-y)ϕ(x)𝑑m(x))𝑑m(y)
= Gg(y)h(fy)𝑑m(y),

i.e.

G(h(f)ϕ(y)-h(fy))g(y)𝑑m(y)=0.

Because this is true for all gL1(G), it follows that for all fL1(G) and for almost all yG,

h(f)ϕ(y)=h(fy).

Because h0, there is some gL1(G) with h(g)0. Then for almost all (x,y)G×G,

h(g)ϕ(x+y)=h(gx+y)=h((gx)y)=h(gx)ϕ(y)=h(g)ϕ(x)ϕ(y),

and as h(g)0, for almost all (x,y)G×G,

ϕ(x+y)=ϕ(x)ϕ(y).

We define ϕ1:G by ϕ1(y)=h(gy)h(g), which is continuous because ygy is continuous and h is continuous. The function ϕ1 satisfies ϕ(y)=ϕ1(y) for almost all yG, so for almost all (x,y)G×G,

ϕ1(x+y)=ϕ1(x)ϕ1(y).

(x,y)ϕ1(x+y) and (x,y)ϕ1(x)ϕ1(y) are continuous and the above equality holds for almost all (x,y)G×G, so the above equality in fact holds for all (x,y)G×G. Hence ϕ1Γ. Define γ:G by γ(x)=ϕ1(-x). Then γΓ, and γ(x)=ϕ(-x) for almost all xG, so for all fL1(G),

f^(γ) = Gf(x)-x,γ𝑑m(x)
= Gf(x)γ(-x)𝑑m(x)
= Gf(x)ϕ(x)𝑑m(x)
= h(f).

Suppose that γ1,γ2Γ satisfy f^(γ1)=f^(γ2) for all fL1(G). Then -x,γ1=-x,γ2 for almost all xG, and because these are continuous they are in fact equal for all xG, i.e. γ1=γ2. ∎

Let A(Γ)={f^:fL1(G)}. Elements of A(Γ) are functions Γ. So far Γ has not been given a topology. We shall be interested in the initial topology on Γ with respect to the family of functions A(Γ). That is, the topology on Γ is the coarsest topology so that each element of A(Γ) is continuous. Furthermore, it is a fact that the topology on Γ is equal to the subspace topology on Γ inherited from L1(G)* with the weak-* topology. Finally, let the maximal ideal space Δ of L1(G) have the subspace topology inherited from L1(G)* with the weak-* topology. In Theorem 2 we presented a bijection ΓΔ, and can be proved that this bijection is a homeomorphism.

It is proved in Rudin66 6 Walter Rudin, Fourier Analysis on Groups, p. 10, §1.2.6. that with this topology, Γ is a locally compact abelian group. The proof constructs a basis for the topology of Γ, and because we will use this basis later it will be useful to write out the proof.

Theorem 3.
  1. 1.

    (x,γ)x,γ is continuous G×Γ.

  2. 2.

    For r>0 let Ur={z:|1-z|<r}. If K is a compact subset of G then

    W(K,r)={γΓ:x,γUr for all xK}

    is an open subset of Γ, and if C is a compact subset of Γ then

    V(C,r)={xG:x,γUr for all γC}

    is an open subset of G.

  3. 3.

    {γ0+W(K,r):γ0ΓK is a compact subset of Γr>0} is a basis for the topology of Γ.

  4. 4.

    Γ is a locally compact abelian group.

Proof.

Let fL1(G). For x0G, γ0Γ and ϵ>0, because xfx is continuous there is a neighborhood V of x0 such that fx-fx01<ϵ for all xV, and because f^x0 is continuous there is a neighborhood W of γ0 such that |f^x0(γ)-f^x0(γ0)|<ϵ for all γW. If (x,γ)V×W, then

|f^x(γ)-f^x0(γ0)| |f^x(γ)-f^x0(γ)|+|f^x0(γ)-f^x0(γ0)|
< fx-fx01+ϵ
< 2ϵ,

showing that (x,γ)f^x(γ) is continuous. But

x,γf^(γ)=f^x(γ),xG,γΓ,

and it follows that (x,γ)x,γ is continuous.

Let K be a compact subset of G, r>0, and γ0W(K,r). For x0K, |x0,γ0-1|=δ<r. Because (x,γ)x,γ is continuous there is a neighborhood Vx0 of x0 and a neighborhood Wx0 of γ0 such that |x,γ-x0,γ0|<r-δ for (x,γ)Vx0×Wx0, and then x,γUr for (x,γ)Vx0×Wx0. As Kx0KVx0 and K is compact, there are x1,,xnK such that KVxi. Let W=Wxi, which is a finite intersection of neighborhoods of γ0 and hence itself a neighborhood of γ0. It is apparent that WW(K,r), showing that W(K,r) is open.

Let C be a compact subset of G, r>0, and x0V(C,r). For γ0C, |x0,γ0-1|=δ<r, so there is a neighborhood Vγ0 of x0 and a neighborhood Wγ0 of γ0 such that |x,γ-x0,γ0|<r-δ for (x,γ)Vγ0×Wγ0, therefore x,γUr for (x,γ)Vγ0×Wγ0. There are γ1,,γnC such that CWγi, and V=Vγi is a neighborhood of x0 that is contained in V(C,r), showing that V(C,r) is open.

Let γ0Γ and let W be a neighborhood of γ0. Because Γ has the initial topology for A(Γ), a local subbasis at γ0 is given by sets of the form {γΓ:|f^(γ)-f^(γ0)|<ϵ}, fL1(G) and ϵ>0. Therefore, there are f1,,fnL1(G) and ϵ1,,ϵn>0 such that

{γΓ:|f^i(γ)-f^i(γ0)|<ϵi}W. (1)

Let ϵ be the minimum of the ϵi, let giC0(G) with fi-gi1<ϵ3, let K be the union of the supports of gi, and let M be the maximum of gi1. With r<ϵ3M, for γγ0+W(K,r) we have

|f^i(γ)-f^i(γ0)| |f^i(γ)-g^i(γ)|+|g^i(γ)-g^i(γ0)|+|g^i(γ0)-f^i(γ0)|
fi-gi1+|Kgi(x)(-x,γ--x,γ0)𝑑m(x)|+fi-gi1
< 23ϵ+K|gi(x)||1--x,γ-γ0|𝑑m(x)
23ϵ+rgi1
< ϵ.

Thus, if γγ0+W(K,r) then γ belongs to the intersection (1), and this shows that γ0+W(K,r)W. This establishes that the collection of those sets of the form γ0+W(K,r), for γ0Γ, K a compact subset of Γ, and r>0, is a basis for the topology of Γ.

Let γ1,γ2Γ, let K be a compact subset of Γ, and let r>0. Because

(γ1+W(K,r/2))-(γ2+W(K,r/2))γ1-γ2+W(K,r),

for (γ,γ′′)(γ1+W(K,r/2))×(γ2+W(K,r/2)) we have γ-γ′′γ1-γ2+W(K,r), and this shows that (γ,γ′′)γ-γ′′ is continuous, which shows that Γ is a topological group. Γ is locally compact because it is homeomorphic to the maximal ideal space Δ of L1(G), and it is a fact that the maximal ideal space of any commutative Banach algebra is a locally compact Hausdorff space. This completes the proof. ∎

A fact that Rudin uses but merely asserts is the following (he refers to it being true for the Gelfand transform, and then merely asserts its truth there).

Theorem 4.

A(Γ)C0(Γ).

Proof.

Let Γ be the weak-* closure of Γ in L1(G)*. For ΛΓ, there is a net γiΓ that weak-* converges to Λ. For x,yG,

Λ(ab)=limiγi(ab)=limiγi(a)γi(b)=(limiγi(a))(limiγi(b))=Λ(a)Λ(b).

Similarly, Λ is linear, so Λ is an algebra homomorphism G. Hence taking the closure of Γ in L1(G)* has added precisely the map that is identically 0:

Γ=Γ{0}.

With K={ΛL1(G)*:Λ1}, the Banach-Alaoglu theorem tells us that K is a weak-* compact subset of L1(G)*. It is apparent that ΓK, so Γ is a weak-* compact subset of L1(G)*.

If fL1(G) and ϵ>0, then because f^:Γ is continuous,

K0={γΓ:|f^(γ)|ϵ}

is a closed subset of Γ. The only way K0 would fail to be a weak-* closed subset of Γ is if 0 belonged to its weak-* closure, and it is straightforward to see that this is not the case. So K0 is in fact a weak-* closed subset of the weak-* compact set Γ, and hence is itself weak-* compact. It follows that K0 is a compact subset of Γ, and this shows that f^C0(Γ). ∎

Now that we know that A(Γ)C0(Γ), it does not take long to verify that the conditions of the Stone-Weierstrass theorem are satisfied (for distinct γ1,γ2Γ there is some fL1(G) with f^(γ1)f^(γ2); A(Γ) is self-adjoint; for each γΓ there is some fL1(G) with f^(γ)0), and hence that A(Γ) is dense in C0(Γ).

Theorem 5.

If G is discrete then Γ is compact, and if G is compact then Γ is discrete.

Proof.

We remarked earlier that the bijection ΓΔ is a homeomorphism, where Δ is the maximal ideal space of L1(G) and has the subspace topology inherited from L1(G)* with the weak-* topology. If G is discrete, then L1(G) is unital, and it is a fact that the maximal ideal space of a unital commutative Banach algebra is compact, so Γ is compact.

Suppose that G is compact, with Haar measure m satisfing m(G)=1. If γΓ and there is some x0G with γ(x0)1 (i.e. γ is not the 0 homomorphism), then

Gx,γ𝑑m(x)=x0,γGx-x0,γ𝑑m(x)=x0,γGx,γ𝑑m(x).

As x0,γ1, this means that Gx,γ𝑑m(x)=0. Therefore,

Gx,γ𝑑m(x)={1γ=0,0γ0.

As G is compact, χGL1(G), and

χ^G(γ)=GχG(x)-x,γ𝑑m(x)=G-x,γ𝑑m(x)={1γ=0,0γ0.

χ^G is continuous, so {γΓ:χ^G(γ)=0} is a closed subset of Γ, hence its complement {γΓ:χ^G(γ)0} is an open subset of Γ. But by the above this complement is {0}, and Γ is a topological group so this implies that each singleton is open, meaning that Γ is discrete. ∎

4 Regular complex Borel measures on G

Let M(G) be the set of regular complex Borel measures on G. If E is a Borel set in G, define En={(x1,,xn):x1++xnE}, which is a Borel set in Gn. For μ1,,μnM(G), we define

(μ1**μn)(E)=(μ1××μn)(En).

It is proved in Rudin77 7 Walter Rudin, Fourier Analysis on Groups, p. 13, §1.3.1. that μ1**μnM(G), and that with convolution as multiplication and norm μ=|μ|(G) (the total variation norm), M(G) is a commutative Banach algebra with unity δ0.

The Fourier transform of μM(G) is the function μ^:Γ defined by

μ^(γ)=G-x,γ𝑑μ(x),γΓ.

We write B(Γ)={μ^:μM(G)}. For fL1(G) we have proved that f^C0(G). We prove now that for μM(G), μ^ is bounded and uniformly continuous on Γ. However, δ0M(G), and for γΓ,

δ^0(γ)=G-x,γ𝑑δ0(x)=0,γ=γ(0)=1,

so δ^0C0(Γ). The proof is from Rudin.88 8 Walter Rudin, Fourier Analysis on Groups, p. 15, §1.3.3.

Theorem 6.

If μM(G), then μ^:ΓC is bounded and uniformly continuous.

Proof.

For any γΓ,

|μ^(γ)|G|-x,γ|d|μ|(x)=Gd|μ|(x)=|μ|(G)<,

where |μ| is the variation of μ. Hence μ^ is bounded.

|μ| is regular, so for any δ>0 there is some compact set K such that |μ|(K)<δ, where K=GK. For γ1,γ2Γ, x,γ1-γ2=x,γ1x,γ2-1 and hence

|x,γ1-x,γ2|=|1-x,γ1-γ2|,

with which we get

|μ^(γ1)-μ^(γ2)|G|1-x,γ1-γ2|d|μ|(x).

We know that W(K,δ) defined in Theorem 3 is an open neighborhood of 0. If γ1-γ2W(K,δ), then by the definition of W(K,δ), for all xK we have |1-x,γ1-γ2|<δ, giving

G|1-x,γ1-γ2|d|μ|(x)Kδd|μ|(x)+K2d|μ|(x)<δμ+2δ,

showing that μ^ is uniformly continuous.

If fL1(G), we define μfM(G) by

μf(E)=Ef(x)𝑑m(x),

for Borel subsets E of G, where m is Haar measure on G. Thus, μf is absolutely continuous with respect to m and has density f. Then, for γΓ,

μ^f(γ)=G-x,γ𝑑μf(x)=G-x,γf(x)𝑑m(x)=f^(γ),

showing that A(Γ)B(Γ). Furthermore, f1=μf, hence it makes sense to identify L1(G) with its image in M(G) under the map fμf. To talk about a function that belongs to M(G) is to speak about μf for some fL1(G). It is proved in Rudin that L1(G) is a closed ideal in the Banach algebra M(G).99 9 Walter Rudin, Fourier Analysis on Groups, p. 16, §1.3.4.

Rudin calls the following theorem the uniqueness theorem.1010 10 Walter Rudin, Fourier Analysis on Groups, p. 17, §1.3.6.

Theorem 7.

If μM(Γ) and

Γx,γ𝑑μ(γ)=0

for all xG, then μ=0.

Proof.

Let fL1(G).

Γf^(γ)𝑑μ(γ) = ΓGf(x)-x,γ𝑑m(x)𝑑μ(γ)
= Gf(x)Γ-x,γ𝑑μ(γ)𝑑m(x)
= 0.

Because A(Γ) is dense in C0(Γ), it follows that for all ϕC0(Γ),

Γϕ𝑑μ=0,

and this implies that μ=0. ∎

5 Positive-definite functions

A function ϕ:G is called positive-definite if for every N and every x1,,xNG, c1,,cN, we have

n,m=1Ncncm¯ϕ(xn-xm)0. (2)

In particular, the left-hand side of the above inequality is real.

Let ϕ be a character of G, which we do not assume to be continuous. Then,

n,m=1Ncncm¯ϕ(xn-xm)=n,m=1Ncncm¯ϕ(xn)ϕ(xm)¯=|n=1Ncnϕ(xn)|20,

so any character of G, whether or not it is continuous, is positive-definite.

Lemma 8.

If ϕ:GC is positive-definite, then

ϕ(0)0,
ϕ(-x)=ϕ(x)¯,|ϕ(x)|ϕ(0),xG,

and

|ϕ(x)-ϕ(y)|22ϕ(0)Re(ϕ(0)-ϕ(x-y)),x,yG.
Proof.

Take N=1, c1=1, x1=0. Then (2) is

ϕ(0)0.

Take N=2, x1=0,x2=x,c1=1,c2=c. Then (2) is

ϕ(0)+c¯ϕ(-x)+cϕ(x)+|c|2ϕ(0)0. (3)

Because ϕ(0) is real, this means that cϕ(x)+c¯ϕ(-x) is real, hence is equal to its own complex conjugate. Writing ϕ(x)=A+iB and ϕ(-x)=C+iD, for c=1 this implies that

A+iB+C+iD=A-iB+C-iD

and for c=i this is

iA-B-iC+D=-iA-B+iC+D.

The first equation tells us B+D=0, and the second equation tells us A-C=0. Thus

ϕ(-x)=C+iD=A-iB=ϕ(x)¯.

Use (3) with c chosen so that cϕ(x)=-|ϕ(x)|. |c|=1, and using ϕ(-x)=ϕ(x)¯,

2ϕ(0)+2|ϕ(x)|0.

Take N=3, x1=0,x2=x,x3=y, c1=1, λ,

c2=λ|ϕ(x)-ϕ(y)|ϕ(x)-ϕ(y),

c3=-c2; since ϕ(0)0, the claim is obviously true for the case x=y, which we discard. Then (2) is

(1+2|c2|2)ϕ(0)+c2¯(ϕ(-x)-ϕ(-y))+c2(ϕ(x)-ϕ(y))-|c2|2(ϕ(x-y)+ϕ(y-x))0.

Using the definition of c2 and the fact that ϕ(z)¯=ϕ(-z),

(1+2λ2)ϕ(0)+2λ|ϕ(x)-ϕ(y)|-λ2(ϕ(x-y)+ϕ(x-y)¯)0,

or

λ2(2ϕ(0)-2Reϕ(x-y))+2λ|ϕ(x)-ϕ(y)|+ϕ(0)0.

The fact that this quadratic polynomial does not take negative values implies that it has either 0 or 1 real roots, and hence that its discriminant is 0:

4|ϕ(x)-ϕ(y)|2-4(2ϕ(0)-2Reϕ(x-y))ϕ(0)0,

which is the claim. ∎

We remind ourselves that f*(x)=f(-x)¯.

Theorem 9.

If fL2(G), then ϕ=f*f* is positive-definite and belongs to C0(G).

Proof.
n,m=1Ncncm¯ϕ(xn-xm) = n,m=1Ncncm¯Gf(xn-xm-y)f(-y)¯𝑑y
= n,m=1Ncncm¯Gf(xn-y)f(xm-y)¯𝑑y
= G|n=1Ncnf(xn-y)|2𝑑y
0.

It is a fact that if 1<p<, 1p+1q=1 and fLp(G), gLq(G), then f*gC0(G).1111 11 Walter Rudin, Fourier Analysis on Groups, p. 4, §1.1.6.

The following theorem is from Rudin.1212 12 Walter Rudin, Fourier Analysis on Groups, p. 19, §1.4.2.

Theorem 10.

If μM(Γ), μ0 (i.e. μ=|μ|), and

ϕ(x)=Γx,γ𝑑μ(γ),xG,

then ϕ is uniformly continuous and positive-definite.

Proof.
n,m=1Ncncm¯ϕ(xn-xm) = Γn,m=1Ncncm¯xn-xm,γdμ(γ)
= Γn,m=1Ncncm¯xn,γxm,γ¯dμ(γ)
= Γ|n=1Ncnxn,γ|2𝑑μ(γ)
0,

showing that ϕ is positive-definite.

μ is regular, so for any δ>0 there is a compact set C such that μ(C)<δ, where C=ΓC. We know V(C,δ) defined in Theorem 3 is an open neighborhood of 0. If x1-x2V(C,δ), then by the definition of V(C,δ), for all γC we have |1-x1-x2,γ|<δ, and so

|ϕ(x1)-ϕ(x2)| Γ|1-x1-x2,γ|𝑑μ(γ)
= C|1-x1-x2,γ|𝑑μ(γ)+C|1-x1-x2,γ|𝑑μ(γ)
Cδ𝑑μ(γ)+C2𝑑μ(γ)
< δμ+2δ.

The above theorem shows that the inverse Fourier transform of a nonnegative measure on Γ is a uniformly continuous positive-definite function on G. The following theorem shows that any continuous positive-definite function on G has this form. We remind ourselves that if a positive-definite function is continuous then it is uniformly continuous, by Lemma 8. We are following the proof given in Rudin.1313 13 Walter Rudin, Fourier Analysis on Groups, p. 19, §1.4.3.

Theorem 11 (Bochner’s theorem).

If ϕ:GC is uniformly continuous and positive-definite, then there is some nonnegative measure μM(Γ) such that

ϕ(x)=Γx,γ𝑑μ(γ),xG.
Proof.

As ϕ is positive-definite, ϕ(0) is a nonnegative real number, and |ϕ(x)|ϕ(0) for all xG. If ϕ(0)=0 then use μ=0. Otherwise, it makes sense to divide ϕ by ϕ(0) and the resulting function is also continuous and positive-definite. Thus without loss of generality we suppose that ϕ(0)=1.

Using the fact that ϕ is positive-definite one shows that for any fCc(G), GGf(x)f(y)¯ϕ(x-y)𝑑m(x)𝑑m(y) is nonnegative by approximating this integral with finite sums. Then as Cc(G) is dense in L1(G), the previous integral is nonnegative for any fL1(G). We define Tϕ:L1(G) by

Tϕ(f)=Gfϕ𝑑m,fL1(G),

and define, for f,gL1(G),

[f,g] = Tϕ(f*g*)
= G(f*g*)(x)ϕ(x)𝑑m(x)
= GGf(x-y)g(-y)¯ϕ(x)𝑑m(y)𝑑m(x)
= GGf(x)g(y)¯ϕ(x-y)𝑑m(x)𝑑m(y).

Therefore, [f,f]0, and thus [,] is an inner product on L1(G) and hence satisfies the Cauchy-Schwarz inequality:

|[f,g]|2[f,f][g,g],f,gL1(G).

Suppose that V is a symmetric neighborhood of 0 in G and define g=χVm(V).

[f,g]-Tϕ(f) = GGf(x)g(y)¯ϕ(x-y)𝑑m(x)𝑑m(y)-Gf(x)ϕ(x)𝑑m(x)
= G1m(V)Vf(x)ϕ(x-y)𝑑m(y)𝑑m(x)
-G1m(V)Vf(x)ϕ(x)𝑑m(y)𝑑m(x)
= Gf(x)1m(V)V(ϕ(x-y)-ϕ(x))𝑑m(y)𝑑m(x)

and

[g,g]-1=1m(V)2VV(ϕ(x-y)-1)𝑑m(x)𝑑m(y).

Because ϕ is uniformly continuous, for any δ>0 there is some V such that both these integrals have absolute value <δ, and then using the Cauchy-Schwarz inequality we get

|Tϕ(f)|2[f,f],fL1(G). (4)

Let fL1(G) and define h=f*f* and hn=hn-1*h for n2. |ϕ(x)|ϕ(0)=1 tells us ϕ=1 and so Tϕ1. In fact, one checks that Tϕ=1. Applying (4) to h,h2,h4, we obtain

|Tϕ(f)|2Tϕ(f*f*)=Tϕ(h)(Tϕ(h*h*))1/2=(Tϕ(h2))1/2,

so for any n1, because Tϕ=1,

|Tϕ(f)|2(Tϕ(h2n))2-nh2n12-n.

The spectral radius formula tells us that

limnh2n12-n=h^.

But h^=f^f*^=|f^|2, so |Tϕ(f)|2f^2, i.e.

|Tϕ(f)|f^,fL1(G).

We define Sϕ on A(Γ) by Sϕ(f^)=Tϕ(f); this makes sense because if f1^=f2^ then |Tϕ(f1-f2)|f1^-f2^=0, so Tϕ(f1)=Tϕ(f2). The above inequality means that

|Sϕ(f^)|f^,f^A(Γ).

Therefore Sϕ is a bounded linear functional on A(Γ), and because A(Γ) is dense in C0(Γ), Sϕ can be extended to a bounded linear functional on C0(Γ) with norm Tϕ=1. But Γ is a locally compact Hausdorff space, so by the Riesz representation theorem there is a unique measure μM(Γ) such that

Sϕ(g)=Γg(-γ)𝑑μ(γ),gC0(Γ),

and μ=Sϕ=1; we state the above with g(-γ) rather than g(γ) for later convenience. For fL1(G),

Tϕ(f) = Sϕ(f^)
= Γf^(-γ)𝑑μ(γ)
= ΓGf(x)-x,-γ𝑑m(x)𝑑μ(γ)
= Gf(x)(Γx,γ𝑑μ(γ))𝑑m(x).

But the definition of Tϕ states

Tϕ(f)=Gf(x)ϕ(x)𝑑m(x).

Since these two expressions for Tϕ(f) are equal for all fL1(G), we get that

Γx,γ𝑑μ(γ)=ϕ(x)

for almost all xG. Since both sides of the above equality are continuous, they are equal for all xG. For x=0,

1=ϕ(0)=Γ0,γ𝑑μ(γ)=Γ𝑑μ(γ)Γd|μ|(γ)μ=1.

Hence Γ𝑑μ(γ)=Γd|μ|(γ), from which it follows that μ=|μ|, and therefore μ is a nonnegative measure. ∎

6 The inversion theorem

Define B(G) to be the set of those f:G for which there is some μfM(Γ) such that

f(x)=Γx,γ𝑑μf(γ),xG.

It is apparent from Theorem 7 that there is at most one μfM(Γ) such that the above holds.

The following proof is from Rudin.1414 14 Walter Rudin, Fourier Analysis on Groups, p. 22, §1.5.1.

Theorem 12 (Inversion theorem).

If fL1(G)B(G), then f^L1(Γ).

If the Haar measure m on G is fixed, then there is a Haar measure mΓ on Γ such that for all fL1(G)B(G),

f(x)=Γf^(γ)x,γ𝑑mΓ(γ),xG.
Proof.

Write B1=L1(G)B(G). For fB1 and hL1(G),

(h*f)(0) = Gh(x)f(-x)𝑑m(x)
= Gh(x)Γ-x,γ𝑑μf(γ)𝑑m(x)
= Γh^(γ)𝑑μf(γ).

For gB1, we have h*gL1(G) and h*fL1(G), and so using the above equality,

Γh*g^𝑑μf=((h*g)*f)(0)=((h*f)*g)(0)=Γh*f^𝑑μg,

hence

Γh^g^𝑑μf=Γh^f^𝑑μg,f,gB1,hL1(G).

Because A(Γ) is dense in C0(Γ) and the above holds for all hL1(G), it follows that

g^dμf=f^dμg,f,gB1. (5)

We define T:Cc(Γ) as follows. Let ψCc(Γ), K=suppψ. For γ0K, there is some ϕC0(Γ) such that ϕ(γ0)0, and because A(Γ) is dense in C0(Γ) there is therefore some fL1(G) such that f^(γ0)0. With δ=|f^(γ0)|, there is some uCc(G) with u-f1<δ, and

|u^(γ0)-f^(γ0)|u-f1<δ,

which shows that u^(γ0)0. u*u*^=|u^|20, so there is some open neighborhood U of γ0 on which u*u*^ is positive, as it is a continuous function. Since K is compact, it is covered by finitely many of these open neighborhoods. Call the corresponding functions u1,,unCc(G), and write

g=u1*u1*++un*un*.

g satisfies g^(γ)>0 for all γK. Because each un belongs to Cc(G), un*un* belongs to Cc(G) and so gCc(G). Moreover, by Theorem 9, each un*un* is positive-definite, and one checks that as g is a linear combination of positive-definite functions with nonnegative coefficients it is itself positive-definite. Because g is positive-definite, by Bochner’s theorem it belongs to B(G), and because gCc(G), g belongs to L1(G). Hence gB1. We have now proved that there is at least one element of B1 whose Fourier transform does not vanish on K. Suppose that f is any such function. Then using (5),

Γψf^𝑑μf=Γψf^g^g^𝑑μf=Γψf^g^f^𝑑μg=Gψg^𝑑μg.

Thus, it makes sense to define

Tψ=Γψg^𝑑μg. (6)

One checks that T is linear. Because g is positive-definite, the measure μg supplied by Bochner’s theorem is nonnegative, and hence if ψ0 then Tψ0, namely, T is positive. There are fB1 and ψCc(G) such that Γψ𝑑μf0, and ψf^Cc(G), so there is some gB1 satisfying

T(ψf^)=Γψf^g^𝑑μg=Γψ𝑑μf0,

showing that T0.

Let ψCc(Γ) and γ0Γ. There is some gB1 such that g^ is positive on both K and K+γ0. For f(x)=-x,γ0g(x), xG, we have f^(γ)=g^(γ+γ0), γΓ, and μf(E)=μg(E-γ0). For ψ0Cc(Γ) defined by ψ0(γ)=ψ(γ-γ0),

Tψ0=Γψ(γ-γ0)g^(γ)𝑑μg(γ)=Γψ(γ)f^(γ)𝑑μf(γ)=Tψ,

showing that T is translation invariant. Then by the Riesz representation theorem, there is some nonnegative regular measure mΓ on Γ satisfying

Tψ=Γψ𝑑mΓ,ψCc(Γ).

This measure mΓ is translation invariant and not the zero measure because T has these properties, and this means that it is a Haar measure on Γ.

For fB1,

Γψ𝑑μf=T(ψf^)=Γψf^𝑑mΓ,ψCc(Γ),

which implies that

dμf=f^dmΓ,fB1.

Because μf< (as μfM(Γ)), the above equality implies that f^L1(Γ). Moreover, by the definition of μf, for any xG we have

f(x)=Γx,γ𝑑μf(γ)=Γx,γf^𝑑mΓ(γ).

Using the inversion theorem, we prove the following lemma.1515 15 Walter Rudin, Fourier Analysis on Groups, p. 23, §1.5.2.

Lemma 13.

{x0+V(C,r):x0GC is a compact subset of Gr>0} is a basis for the topology of G.

Γ separates points in G.

Proof.

Let mΓ be the Haar measure on Γ specified in the inversion theorem. Suppose that V is a neighborhood of 0 in G. Let W be a compact neighborhood of 0 in G satisfying W-WV. (One proves that there are such W.) Define f=χWm(W) and g=f*f*. g is continuous and positive-definite, and suppgW-W. Because g is continuous and positive-definite, by Bochner’s theorem it belongs to B(G), and because suppgW-W it belongs to L1(G), so we can apply the inversion theorem to get g^L1(Γ) and

Γg^𝑑mΓ(γ)=g(0)=Gf(-y)f(-y)¯𝑑m(y)=G|f(y)|2𝑑m(y)=1.

Because g^=|f^|20, there is a compact set C in Γ such that

Cg^(γ)𝑑mΓ(γ)>23.

To say that xV(C,1/3) means |1-x,γ|<13 for all γC and hence Rex,γ>23, and satisfies

g(x)=ReCg^(γ)x,γ𝑑mΓ(γ)+ReCg^(γ)x,γ𝑑mΓ(γ).

The first term is >49 and the second term has absolute value <13, so g(x)>19 for xV(C,1/3). But g(x)>19 means that xsuppg=W-WV, so

V(C,1/3)V,

from which it follows that V(C,r), C compact and r>0, is a local basis at 0.

For x0G, x00, let V be a neighborhood of 0 that does not include x0, and the above gives xV(C,1/3), i.e., there is some γΓ such that |1-x,γ|13, and hence x,γ1. Therefore, if x1x2, there is some γΓ such that x1-x2,γ1, and hence x1,γx2,γ, which is what it means to say that Γ separates points in G. ∎

7 Pontryagin duality theorem

The dual group Γ of G is itself a locally compact abelian group and so has a dual group Γ^. We proved in Theorem 3 that (x,γ)x,γ is continuous, and therefore for any xG, the function α(x):Γ defined by

γ,α(x)=x,γ,γΓ,

belongs to Γ^. For x,yG, α(xy)Γ^ satisfies

γ,α(xy)=xy,γ=x,γy,γ=γ,α(x)γ,α(y),

showing that α:GΓ^ is a homomorphism. The following theorem, proved in Rudin,1616 16 Walter Rudin, Fourier Analysis on Groups, p. 28, §1.7.2. shows that α is an isomorphism of topological groups. That is, it states that a locally compact abelian group is isomorphic as a topological group to its double dual. Let LCA denote the category of locally compact abelian groups, where morphisms are continuous group homomorphisms. Taking the double dual of an element of LCA is a functor, and it can be proved that there is a natural isomorphism between the identity functor in LCA and the double dual functor.1717 17 For more, see the nLab page: http://ncatlab.org/nlab/show/Pontrjagin+dual

Theorem 14 (Pontryagin duality theorem).

α:GΓ^ defined by

γ,α(x)=x,γ,γΓ,

is an isomorphism of topological groups.

We proved earlier that if G is discrete then Γ is compact and that if G is compact then Γ is discrete, but had not established that if Γ is compact then G is discrete or if Γ is discrete then G is compact, but we obtain these conclusions from the Pontryagin duality theorem: if Γ is compact then Γ^ is discrete, and G is isomorphic as a topological group to Γ^ so G is discrete, and likewise if Γ is discrete then G is compact.

8 Further reading

Keith Conrad, http://www.math.uconn.edu/~kconrad/blurbs/gradnumthy/characterQ.pdf works out explicitly the form of all characters of , and shows that the group of all characters of (namely, the dual group of when has the discrete topology), is isomorphic as a group to the quotient group 𝔸/. This gives a satisfying reason for caring about the p-adic numbers p and the adeles 𝔸.