Previous Up Next
Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License.

14 Functional Spaces

In this section we describe various Banach spaces of functions on sets with measure.

14.1 Integrable Functions

Let (X,L,µ) be a measure space. For 1≤ p<∞, we define Lp(µ) to be the space of measurable functions f:X→K such that

  
 


X

f
p  d µ < ∞. 

We define ||·||p : Lp(µ)→[0,∞) by

⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p = 




 


X

f
p  d µ 




1/p





 
    (f∈ Lp(µ)). 

Notice that if f=0 almost everywhere, then | f |p=0 almost everywhere, and so ||f||p=0. However, there can be non-zero functions such that f=0 almost everywhere. So ||·||p is not a norm on Lp(µ).

Exercise 1 Find a measure space (X,µ) such that lp=Lp(µ), that is the space of sequences lp is a particular case of function spaces considered in this section. It also explains why the following proofs are referencing to Section 11 so often.
Lemma 2 (Integral Hölder inequality) Let 1<p<∞, let q∈(1,∞) be such that 1/p + 1/q=1. For fLp(µ) and gLq(µ), we have that fg is summable, and
 


X

fg
  d µ ≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q.  (80)
Proof. Recall that we know from Lem. 2 that
    
ab
 ≤ 

a
p
p
 + 

b
q
q
    (a,b∈K).  
Now we follow the steps in proof of Prop. 4. Define measurable functions a,b:X→K by setting
    a(x) = 
f(x)
⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p
,   b(x) = 
g(x)
⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q
    (x∈ X). 
So we have that
    
a(x) b(x) 
 ≤ 

f(x) 
p
p⎪⎪
⎪⎪
f⎪⎪
⎪⎪
pp
 +

g(x) 
q
q⎪⎪
⎪⎪
g⎪⎪
⎪⎪
qq
    (x∈ X). 
By integrating, we see that
    
 


X

ab
  d µ ≤ 
1
p⎪⎪
⎪⎪
f⎪⎪
⎪⎪
pp
 


X

f
p  d µ + 
1
q⎪⎪
⎪⎪
g⎪⎪
⎪⎪
qq
 


X

g
q  d µ = 
1
p
 + 
1
q
 = 1. 
Hence, by the definition of a and b,
    
 


X

fg
 ≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q, 
as required. □
Lemma 3 Let f,gLp(µ) and let a∈K. Then:
  1. ||af||p = | a | ||f||p;
  2. || f+g ||p ≤ ||f||p + ||g||p.
In particular, Lp is a vector space.
Proof. Part 1 is easy. For 2, we need a version of Minkowski’s Inequality, which will follow from the previous lemma. We essentially repeat the proof of Prop. 5.

Notice that the p=1 case is easy, so suppose that 1<p<∞. We have that

     
 


X

f+g
p  d µ
= 
 


X

f+g
p−1
f+g
  d µ 
         
 
≤ 
 


X

f+g
p−1

f
 + 
g

d µ 
         
 
= 
 


X

f+g
p−1
f
  d µ + 
 


X

f+g
p−1
g
  d µ.
         

Applying the lemma, this is

    ≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p




 


X

f+g
q(p−1)  d µ 




1/q





 
+ ⎪⎪
⎪⎪
g⎪⎪
⎪⎪
p




 


X

f+g
q(p−1)  d µ 




1/q





 
. 

As q(p−1)=p, we see that

    ⎪⎪
⎪⎪
f+g⎪⎪
⎪⎪
pp ≤ 
⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p + ⎪⎪
⎪⎪
g⎪⎪
⎪⎪
p
⎪⎪
⎪⎪
f+g⎪⎪
⎪⎪
pp/q. 

As pp/q = 1, we conclude that

    ⎪⎪
⎪⎪
f+g⎪⎪
⎪⎪
p ≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p + ⎪⎪
⎪⎪
g⎪⎪
⎪⎪
p, 

as required.

In particular, if f,gLp(µ) then af+gLp(µ), showing that Lp(µ) is a vector space. □

We define an equivalence relation ∼ on the space of measurable functions by setting fg if and only if f=g almost everywhere. We can check that ∼ is an equivalence relation (the slightly non-trivial part is that ∼ is transitive).

Proposition 4 For 1≤ p<∞, the collection of equivalence classes Lp(µ) / ∼ is a vector space, and ||·||p is a well-defined norm on Lp(µ) / ∼.
Proof. We need to show that addition, and scalar multiplication, are well-defined on Lp(µ)/∼. Let a∈K and f1,f2,g1,g2Lp(µ) with f1f2 and g1g2. Then it’s easy to see that af1+g1af2+g2; but this is all that’s required!

If fg then | f |p = | g |p almost everywhere, and so ||f||p = ||g||p. So ||·||p is well-defined on equivalence classes. In particular, if f∼ 0 then ||f||p=0. Conversely, if ||f||p=0 then ∫X | f |pd µ=0, so as | f |p is a positive function, we must have that | f |p=0 almost everywhere. Hence f=0 almost everywhere, so f∼ 0. That is,

    

f∈ Lp(µ) : f∼ 0 

= 

f∈ Lp(µ) : ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
p=0 

. 

It follows from the above lemma that this is a subspace of Lp(µ).

The above lemma now immediately shows that ||·||p is a norm on Lp(µ)/∼. □

Definition 5 We write Lp(µ) for the normed space (Lp(µ)/∼ , ||·||p).

We will abuse notation and continue to write members of Lp(µ) as functions. Really they are equivalence classes, and so care must be taken when dealing with Lp(µ). For example, if fLp(µ), it does not make sense to talk about the value of f at a point.

Theorem 6 Let (fn) be a Cauchy sequence in Lp(µ). There exists fLp(µ) with ||fnf||p→ 0. In fact, we can find a subsequence (nk) such that fnkf pointwise, almost everywhere.
Proof. Consider first the case of a finite measure space X. We again follow the three steps scheme from Rem. 7. Let fn be a Cauchy sequence in Lp(µ). From the Hölder inequality (80) we see that ||fnfm||1≤ ||fnfm||p (µ(X))1/q. Thus, fn is also a Cauchy sequence in L1(µ). Thus by the Theorem 42 there is the limit function fL1(µ). Moreover, from the proof of that theorem we know that there is a subsequence fnk of fn convergent to f almost everywhere. Thus in the Cauchy sequence inequality
    
 


X

fnk −fnm  
pd µ <ε
we can pass to the limit m→ ∞ by the Fatou Lemma 39 and conclude:
    
 


X

fnk −f  
pd µ <ε.
So, fnk converges to f in Lp(µ), then fn converges to f in Lp(µ) as well.

For a σ-finite measure µ we represent X=⊔k Xk with µ(Xk)<+∞ for all k. The restriction (fn(k)) of a Cauchy sequence (fn)⊂Lp(X,µ) to every Xk is a Cauchy sequence in Lp(Xk,µ). By the previous paragraph there is the limit f(k)Lp(Xk,µ). Define a function fLp(X,µ) by the identities f(x)=f(k) if xXk. By the additivity of integral, the Cauchy condition on (fn) can be written as:

    
 


X

fnfm
pd µ=
k=1
 


Xk

fn(k)fm(k)
pd µ<ε. 

It implies for any M:

    
M
k=1
 


Xk

fn(k)fm(k)
pd µ<ε. 

In the last inequality we can pass to the limit m→ ∞:

    
M
k=1
 


Xk

fn(k)f(k)
pd µ<ε. 

Since the last inequality is independent of M we conclude:

    
 


X

fnf
pd µ=
k=1
 


Xk

fn(k)f(k)
pd µ<ε. 

Thus we conclude that fnf in Lp(X,µ). □

Corollary 7Lp(µ) is a Banach space.
Example 8 If p=2 then Lp(µ)=L2(µ) can be equipped with the inner product:
⟨ f,g  ⟩ =
 


X
fḡ d µ. (81)
The previous Corollary implies that L2(µ) is a Hilbert space, see a preliminary discussion in Defn. 22.
Proposition 9 Let (X,L,µ) be a measure space, and let 1≤ p<∞. We can define a map Φ:Lq(µ) → Lp(µ)* by setting Φ(f)=F, for fLq(µ), 1/p+1/q=1, where
    F:Lp(µ)→K,   g ↦ 
 


X
fg  d µ    (gLp(µ)).  
Proof. This proof very similar to proof of Thm. 13. For fLq(µ) and gLp(µ), it follows by the Hölder’s Inequality (80), that fg is summable, and
    




 


X
fg  d µ  




 ≤ 
 


X

fg
  d µ ≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
q⎪⎪
⎪⎪
g⎪⎪
⎪⎪
p. 

Let f1,f2Lq(µ) and g1,g2Lp(µ) with f1f2 and g1g2. Then f1g1 = f2g1 almost everywhere and f2g1 = f2g2 almost everywhere, so f1g1 = f2g2 almost everywhere, and hence

    
 


X
f1g1  d µ =  
 


X
f2g2  d µ. 

So Φ is well-defined.

Clearly Φ is linear, and we have shown that ||Φ(f)|| ≤ ||f||q.

Let fLq(µ) and define g:X→K by

    g(x) = 




f(x) 

f(x) 
q−2
: f(x)≠0, 
      0: f(x)=0. 

Then | g(x) | = | f(x) |q−1 for all xX, and so

    
 


X

g
p  d µ = 
 


X

f
p(q−1)  d µ = 
 


X

f
q  d µ, 

so ||g||p = ||f||qq/p, and so, in particular, gLp(µ). Let F=Φ(f), so that

    F(g) = 
 


X
fg  d µ = 
 


X

f
q  d µ = ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
qq. 

Thus ||F|| ≥ ||f||qq / ||g||p = ||f||q. So we conclude that ||F|| = ||f||q, showing that Φ is an isometry. □

Proposition 10 Let (X,L,µ) be a finite measure space, let 1≤ p<∞, and let FLp(µ)*. Then there exists fLq(µ), 1/p+1/q=1 such that
    F(g) = 
 


X
fg  d µ    (gLp(µ)). 
Proof.[Sketch of the proof] As µ(X)<∞, for EL, we have that ||χE||p = µ(E)1/p < ∞. So χELp(µ), and hence we can define
    ν(E) = FE)    (EL). 
We proceed to show that ν is a signed (or complex) measure. Then we can apply the Radon-Nikodym Theorem 53 to find a function f:X→K such that
    FE) = ν(E) = 
 


E
f  d µ    (EL). 

There is then a long argument to show that fLq(µ), which we skip here. Finally, we need to show that

    
 


X
fg  d µ = F(g) 

for all gLp(µ), and not just for gE. That follows for simple functions with a finite set of values by linearity of the Lebesgue integral and F. Then, it can be extended by continuity to the entire space Lp(µ) in view in the following Prop. 14. □

Proposition 11 For 1<p<∞, we have that Lp(µ)* = Lq(µ) isometrically, under the identification of the above results.
Remark 12
  1. For p=q=2 we obtain a special case of the Riesz–Frechét theorem 11 about self-duality of the Hilbert space L2(µ).
  2. Note that L* is not isomorphic to L1, except finite-dimensional situation. Moreover if µ is not a point measure L1 is not a dual to any Banach space.
Exercise 13 Let µ be a measure on the real line.
  1. Show that the space L(ℝ,µ) is either finite-dimensional or non-separable.
  2. Show that for pq neither Lp(ℝ,µ) nor Lq(ℝ,µ) contains the other space.

14.2 Dense Subspaces in Lp

We note that fLp(X) if and only if | f |p is summable, thus we can use all results from Section 13 to investigate Lp(X).

Proposition 14 Let (X,L,µ) be a finite measure space, and let 1≤ p<∞. Then the collection of simple bounded functions attained only a finite number of values is dense in Lp(µ).
Proof.Let fLp(µ), and suppose for now that f≥0. For each n∈ℕ, let
    fn = min(n, 
1
n
 ⌊ nf ⌋). 
Then each fn is simple, fnf, and | fnf |p→0 pointwise. For each n, we have that
    0 ≤ fn ≤ f  0 ≤ ffn ≤ f, 
so that | ffn |p ≤ | f |p for all n. As ∫| f |pd µ<∞, we can apply the Dominated Convergence Theorem to see that
    
 
lim
n
 


X

fnf
p  d µ = 0, 
that is, ||fnf||p → 0.

The general case follows by taking positive and negative parts, and if K=ℂ, by taking real and imaginary parts first. □

Corollary 15 Let µ be the Lebesgue measure on the real line. The collection of simple bounded functions with compact supports attained only a finite number of values is dense in Lp(ℝ,µ).
Proof. Let fLp(ℝ,µ), since ∫ | f |pd µ = ∑k=−∞[k,k+1) | f |pd µ there exists N such that ∑k=−∞N + ∑Nk=∞[k,k+1) | f |pd µ < ε . By the previous Proposition, the restriction of f to [−N,N] can be ε-approximated by a simple bounded function f1 with support in [−N,N] attained only a finite number of values. Therefore f1 will be also (2ε)-approximation to f as well. □
Definition 16 A function f:ℝ→ ℂ is called step function if it a linear combination of a finite number of indicator functions of half-open disjoint intervals: f=∑k ck χ[ak,bk).

The regularity of the Lebesgue measure allows to make a stronger version of Prop. 14.

Lemma 17 The space of step functions is dense in Lp(ℝ).
Proof. By Prop. 14, for a given fLp(ℝ) and ε>0 there exists a simple function f0=∑k=1n ck χAk such that ||ff0||p<ε/2. Let M=||f0|| < ∞. By measurability of the set Ak there is Ck=⊔jmk [ajk,bjk) a disjoint finite union of half-open intervals such that µ(CkAk)<ε/2n3 M. Since Ak and Aj are disjoint for kj we also obtain by the triangle inequality: µ(CjAk)<ε/2n3 M and µ(CjCk)<2ε/2n3 M. We define a step function
  f1=
n
k=1
ck χCk=
n
k=1
mk
j
ck χ[ajk,bjk).
Clearly
   f1(x)=ck    for all  x ∈ Ak∖ ((Ck▵ Ak) ⋃(⋃j≠ kCj)).
Thus:
   µ({x ∈ ℝ  ∣  f0(x)≠ f1(x)}) ≤ n· n· 
ε
2n3M
 = 
ε
2 nM
.
Then ||f0f1||pnM· ε/2 n M=ε/2 because ||f1||< nM. Thus ||ff1||p<ε. □
Corollary 18 The collection of continuous function belonging to Lp(ℝ) is dense in Lp(ℝ).
Proof. In view of Rem. 29 and the previous Lemma it is enough to show that the characteristic function of an interval [a,b] can be approximated by a continuous function in Lp(ℝ). The idea of such approximation is illustrated by Fig. 4 and we skip the technical details. □

We will establish denseness of the subspace of smooth function in § 15.4.

Exercise 19 Show that every fL1(ℝ) is continuous on average, that is for any ε>0 there is δ>0 such that for all t such that | t |<δ we have:
 



f(x)−f(x+t) 
dx < ε . (82)

Here is an alternative demonstration of a similar result, it essentially encapsulate all the above separate statements. Let ([0,1],L,µ) be the restriction of Lebesgue measure to [0,1]. We often write Lp([0,1]) instead of Lp(µ).

Proposition 20 For 1≤ p<∞, we have that CK([0,1]) is dense in Lp([0,1]).
Proof. As [0,1] is a finite measure space, and each member of CK([0,1]) is bounded, it is easy to see that each fCK([0,1]) is such that ||f||p<∞. So it makes sense to regard CK([0,1]) as a subspace of Lp(µ). If CK([0,1]) is not dense in Lp(µ), then we can find a non-zero FLp([0,1])* with F(f)=0 for each fCK([0,1]). This was a corollary of the Hahn-Banach theorem 15.

So there exists a non-zero gLq([0,1]) with

    
 


[0,1]
fg  d µ = 0    (f∈ CK([0,1])). 

Let a<b in [0,1]. By approximating χ(a,b) by a continuous function, we can show that ∫(a,b) gd µ = ∫ g χ(a,b)d µ = 0.

Suppose for now that K=ℝ. Let A = { x∈[0,1] : g(x)≥0 } ∈ L. By the definition of the Lebesgue (outer) measure, for є>0, there exist sequences (an) and (bn) with A ⊆ ∪n (an,bn), and ∑n (bnan) ≤ µ(A) + є.

For each N, consider ∪n=1N (an,bn). If some (ai,bi) overlaps (aj,bj), then we could just consider the larger interval (min(ai,aj), max(bi,bj)). Formally by an induction argument, we see that we can write ∪n=1N (an,bn) as a finite union of some disjoint open intervals, which we abusing notations still denote by (an,bn). By linearity, it hence follows that for N∈ℕ, if we set BN = ⊔n=1N (an,bn), then

    g χBN  d µ = g χ(a1,b1)⊔⋯⊔(aN,bN)  d µ = 0. 

Let B=∪n (an,bn), so AB and µ(B) ≤ ∑n (bnan) ≤ µ(A)+є. We then have that

    
g χBN  d µ − g χB  d µ  
 = 
g χB∖ (a1,b1)⊔⋯⊔(aN,bN)  d µ  
. 

We now apply Hölder’s inequality to get

     
    
χB∖ (a1,b1)⋃⋯⋃(aN,bN)  d µ 
1/p ⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q
= µ(B∖ (a1,b1)⊔⋯⊔(aN,bN))1/p⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q
         
 
≤ 


n=N+1
(bnan) 


1/p



 
⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q. 
         

We can make this arbitrarily small by making N large. Hence we conclude that

    g χB  d µ=0. 

Then we apply Hölder’s inequality again to see that

    
gχA  d µ  
 = 
gχA  d µ − gχB  d µ  
  = 
g χB∖ A  d µ  
 ≤ ⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q µ(B∖ A)1/p ≤ ⎪⎪
⎪⎪
g⎪⎪
⎪⎪
q є1/p. 

As є>0 was arbitrary, we see that ∫A gd µ=0. As g is positive on A, we conclude that g=0 almost everywhere on A.

A similar argument applied to the set {x∈[0,1] : g(x)≤0} allows us to conclude that g=0 almost everywhere. If K=ℂ, then take real and imaginary parts. □

14.3 Continuous functions

Let K be a compact (always assumed Hausdorff) topological space.

Definition 21 The Borel σ-algebra, B(K), on K, is the σ-algebra generated by the open sets in K (recall what this means from Section 11.5). A member of B(K) is a Borel set.

Notice that if f:K→K is a continuous function, then clearly f is B(K)-measurable (the inverse image of an open set will be open, and hence certainly Borel). So if µ:B(K)→K is a finite real or complex charge (for K=ℝ or K=ℂ respectively), then f will be µ-summable (as f is bounded) and so we can define

  φµ:CK(K) → K,   φµ(f) = 
 


K
f  d µ    (f∈ CK(K)). 

Clearly φµ is linear. Suppose for now that µ is positive, so that

  
φµ(f) 
 ≤ 
 


K

f
  d µ ≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
 µ(K)    (f∈ CK(K)). 

So φµCK(K)* with ||φµ||≤ µ(K).

The aim of this section is to show that all of CK(K)* arises in this way. First we need to define a class of measures which are in a good agreement with the topological structure.

Definition 22 A measure µ:B(K)→[0,∞) is regular if for each AB(K), we have
     
    µ(A)
= sup

µ(E) : E⊆ A  and E is compact 

         
 
= inf

µ(U) : A⊆ U  and U is open 

.
         
A charge ν=ν+−ν is regular if ν+ and ν are regular measures. A complex measure is regular if its real and imaginary parts are regular.

Note the similarity between this notion and definition of outer measure.

Example 23
  1. Many common measures on the real line, e.g. the Lebesgue measure, point measures, etc., are regular.
  2. An example of the measure µ on [0,1] which is not regular:
          µ(∅)=0,   µ(
    {
    1
    2
    }
    )=1,    µ(A)=+∞,
    for any other subset A⊂[0,1].
  3. Another example of a σ-additive measure µ on [0,1] which is not regular:
          µ(A)=

              0, if A is at most countable;
              +∞otherwise.

The following subspace of the space of all simple functions is helpful.

As we are working only with compact spaces, for us, “compact” is the same as “closed”. Regular measures somehow interact “well” with the underlying topology on K.

We let M(K) and M(K) be the collection of all finite, regular real or complex charges (that is, signed or complex measures) on B(K).

Exercise 24 Check that, M(K) and M(K) are real or complex, respectively, vector spaces for the obvious definition of addition and scalar multiplication.

Recall, Defn. 31, that for µ∈ MK(K) we define the variation of µ

  ⎪⎪
⎪⎪
µ⎪⎪
⎪⎪
 = sup



n=1

µ(An) 




, 

where the supremum is taken over all sequences (An) of pairwise disjoint members of B(K), with ⊔n An=K. Such (An) are called partitions.

Proposition 25 The variation ||·|| is a norm on MK(K).
Proof. If µ=0 then clearly ||µ||=0. If ||µ||=0, then for AB(K), let A1=A, A2=KA and A3=A4=⋯=∅. Then (An) is a partition, and so
    0 = 
n=1

µ(An) 
 = 
µ(A) 
 + 
µ(K∖ A) 
. 
Hence µ(A)=0, and so as A was arbitrary, we have that µ=0.

Clearly ||aµ|| = | a |||µ|| for a∈K and µ∈ MK(K).

For µ,λ∈ MK(K) and a partition (An), we have that

    
 
n

(µ+λ)(An) 
 = 
 
n

µ(An)+λ(An) 
≤ 
 
n

µ(An) 
+
 
n

λ(An) 
 ≤ ⎪⎪
⎪⎪
µ⎪⎪
⎪⎪
 + ⎪⎪
⎪⎪
λ⎪⎪
⎪⎪
. 

As (An) was arbitrary, we see that ||µ+λ|| ≤ ||µ|| + ||λ||. □

To get a handle on the “regular” condition, we need to know a little more about CK(K).

Theorem 26 (Urysohn’s Lemma) Let K be a compact space, and let E,F be closed subsets of K with EF=∅. There exists f:K→[0,1] continuous with f(x)=1 for xE and f(x)=0 for xF (written f(E)={1} and f(F)={0}).
Proof. See a book on (point set) topology. □
Lemma 27 Let µ:B(K)→[0,∞) be a regular measure. Then for UK open, we have
    µ(U) = sup





 


K
f  d µ  : f∈ C(K), 0≤ f≤χU





. 
Proof. If 0≤ f≤χU, then
    0 = 
 


K
 0  d µ ≤ 
 


K
f  d µ ≤ 
 


K
 χU  d µ = µ(U). 
Conversely, let F=KU, a closed set. Let EU be closed. By Urysohn Lemma 26, there exists f:K→[0,1] continuous with f(E)={1} and f(F)={0}. So χEf ≤ χU, and hence
    µ(E) ≤ 
 


K
f  d µ ≤ µ(U). 
As µ is regular,
    µ(U) = sup

µ(E) : E⊆ U closed 

≤ sup





 


K
f  d µ : 0≤ f≤χU





≤ µ(U). 
Hence we have equality throughout. □

The next result tells that the variation coincides with the norm on real charges viewed as linear functionals on C(K).

Lemma 28 Let µ∈ M(K). Then
    ⎪⎪
⎪⎪
µ⎪⎪
⎪⎪
 = ⎪⎪
⎪⎪
φµ⎪⎪
⎪⎪
 := sup










 


K
f  d µ 




 : f∈ C(K), ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
≤ 1 





. 
Proof. Let (A,B) be a Hahn decomposition (Thm. 36) for µ. For fC(K) with ||f||≤ 1, we have that
     





 


K
f  d µ 




≤ 




 


A
f  d µ 




+




 


B
f  d µ 




= 




 


A
f  d µ+




+




 


B
f  d µ




         
 
≤ 
 


A

f
  d µ+ + 
 


B

f
  d µ ≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪

µ(A) − µ(B)
≤ ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
⎪⎪
⎪⎪
µ⎪⎪
⎪⎪
, 
         
using the fact that µ(B)≤0 and that (A,B) is a partition of K.

Conversely, as µ is regular, for є>0, there exist closed sets E and F with EA, FB, and with µ+(E)> µ+(A)−є and µ(F)>µ(B)−є. By Urysohn Lemma 26, there exists f:K→[0,1] continuous with f(E)={1} and f(F)={0}. Let g=2f−1, so g is continuous, g takes values in [−1,1], and g(E)={1}, g(F)={−1}. Then

     
    
 


K
g  d µ
= 
 


E
 1  d µ + 
 


F
 −1  d µ + 
 


K∖ (E⋃ F)
g  d µ 
         
 
= µ(E) − µ(F) + 
 


A∖ E
g  d µ + 
 


B∖ F
g  d µ 
         
           

As EA, we have µ(E) = µ+(E), and as FB, we have −µ(F)=µ(F). So

     
 


K
g  d µ
> µ+(A)−є + µ(B) − є + 
 


A∖ E
g  d µ + 
 


B∖ F
g  d µ 
         
 
≥ 
µ(A) 
 + 
µ(B) 
 − 2є − 
µ(A∖ E) 
 − 
µ(B∖ F) 
         
 
≥ 
µ(A) 
 + 
µ(B) 
 − 4є. 
         

As є>0 was arbitrary, we see that ||φµ|| ≥ | µ(A) |+| µ(B) |=||µ||. □

Thus, we know that M(K) is isometrically embedded in C(K)*.

14.4 Riesz Representation Theorem

To facilitate an approach to the key point of this Subsection we will require some more definitions.

Definition 29 A functional F on C(K) is positive if for any non-negative function f≥ 0 we have F(f)≥0.
Lemma 30 Any positive linear functional F on C(X) is continuous and ||F||=F(1), where 1 is the function identically equal to 1 on X.
Proof. For any function f such that ||f||≤ 1 the function 1−f is non negative thus: F(1)−F(f)=F(1−f)>0, Thus F(1)>F(f), that is F is bounded and its norm is F(1). □

So for a positive functional you know the exact place where to spot its norm, while a linear functional can attain its norm in an generic point (if any) of the unit ball in C(X). It is also remarkable that any bounded linear functional can be represented by a pair of positive ones.

Lemma 31 Let λ be a continuous linear functional on C(X). Then there are positive functionals λ+ and λ on C(X), such that λ=λ+−λ.
Proof. First, for fC(K) with f≥0, we define
     
    λ+(f)
= sup

λ(g) : g∈ C(K), 0≤ g≤ f

≥0, 
         
λ(f)
= λ+(f) − λ(f) = sup

λ(g)−λ(f): g∈ C(K), 0≤ g≤ f

         
 
= sup

λ(h): h∈ C(K), 0≤ h+f≤ f

         
 
= sup

λ(h): h∈ C(K), −f ≤ h ≤ 0 

≥ 0. 
         
In a sense, this is similar to the Hahn decomposition (Thm. 36).

We can check that

    λ+(tf) = tλ+(f),   λ(tf) = tλ(f)    (t≥0, f≥0). 

For f1,f2≥ 0, we have that

     
λ+(f1+f2)
= sup

λ(g): 0≤ g ≤ f1+f2

         
 
= sup

λ(g1+g2): 0≤ g1+g2 ≤ f1+f2

         
 
≥ sup

λ(g1) + λ(g2): 0≤ g1f1, 0 ≤ g2 ≤ f2

         
 = λ+(f1) + λ+(f2).           

Conversely, if 0≤ gf1+f2, then set g1 = min(g,f1), so 0≤ g1f1. Let g2 = gg1 so g1g implies that 0≤ g2. For xK, if g1(x)=g(x) then g2(x) = 0 ≤ f2(x); if g1(x)=f1(x) then f1(x)≤ g(x) and so g2(x) = g(x)−f1(x) ≤ f2(x). So 0 ≤ g2f2, and g = g1 + g2. So in the above displayed equation, we really have equality throughout, and so λ+(f1+f2) = λ+(f1) + λ+(f2). As λ is additive, it is now immediate that λ(f1+f2) = λ(f1) + λ(f2)

For fC(K) we put f+(x)=max(f(x),0) and f(x)=−min(f(x),0). Then f±≥ 0 and f=f+f. We define:

    λ+(f) = λ+(f+) − λ+(f),   λ(f) = λ(f+) − λ(f). 

As when we were dealing with integration, we can check that λ+ and λ become linear functionals; by the previous Lemma they are bounded. □

Finally, we need a technical definition.

Definition 32 For fC(K), we define the support of f, written supp(f), to be the closure of the set {xK : f(x)≠0}.
Theorem 33 (Riesz Representation) Let K be a compact (Hausdorff) space, and let λ∈ CK(K)*. There exists a unique µ∈ MK(K) such that
    λ(f) = 
 


K
f  d µ    ( f∈ CK(K) ). 
Furthermore, ||λ|| = ||µ||.
Proof. Let us show uniqueness. If µ12MK(K) both induce λ then µ = µ1−µ2 induces the zero functional on CK(K). So for fC(K),
     
0
= ℜ 
 


K
f  d µ = 
 


K
f  d µr
         
 
= ℑ
 


K
f  d µ = 
 


K
f  d µi. 
         
So µr and µi both induce the zero functional on C(K). By Lemma 28, this means that ||µr|| = ||µi||=0, showing that µ = µr + iµi = 0, as required.

Existence is harder, and we shall only sketch it here. Firstly, we shall suppose that K=ℝ and that λ is positive.

Motivated by the above Lemmas 27 and 28, for UK open, we define

    µ*(U) = sup

λ(f):  f∈ C(K),       0≤ f≤χU,  supp(f)⊆ U

. 

For AK general, we define

    µ*(A) = inf

µ*(U):  U⊆ K  is open,       A⊆ U

. 

We then proceed to show that

If λ is not positive, then by Lemma 31 represent it as λ=λ+−λ for positive λ±. As λ+ and λ are positive functionals, we can find µ+ and µ positive measures in M(K) such that

    λ+(f) = 
 


K
f  d µ+,   λ(f) = 
 


K
f  d µ    (f∈ C(K)). 

Then if µ = µ+ − µ, we see that

    λ(f) = λ+(f) − λ(f) = 
 


K
f  d µ    (f∈ C(K)). 

Finally, if K=ℂ, then we use the same “complexification” trick from the proof of the Hahn-Banach Theorem 15. Namely, let λ∈ C(K)*, and define λr, λiC(K)* by

    λr(f) = ℜ  λ(f),   λi(f) = ℑ λ(f)    ( f∈ C(K) ). 

These are both clearly ℝ-linear. Notice also that | λr(f) | = | ℜ λ(f) | ≤ | λ(f) | ≤ ||λ|| ||f||, so λr is bounded; similarly λi.

By the real version of the Riesz Representation Theorem, there exist charges µr and µi such that

    ℜ λ(f) = λr(f) = 
 


K
f  d µr,   ℑλ(f) = λi(f) = 
 


K
f  d µi    (f∈ C(K) ). 

Then let µ=µr+iµi, so for fC(K),

     
 


K
f  d µ
= 
 


K
f  d µr + i
 


K
f  d µi
         
 
= 
 


K
 ℜ (f)  d µr + i
 


K
 ℑ(f)  d µr  + i
 


K
 ℜ (f)  d µi − 
 


K
 ℑ(f)  d µi
         
 = λr(ℜ (f)) + iλr(ℑ(f)) + iλi(ℜ (f)) − λi(ℑ(f))          
 = ℜ  λ(ℜ (f)) + iℜ λ(ℑ(f)) +iℑλ(ℜ (f)) − ℑλ(ℑ(f))          
 = λ(ℜ (f) + iℑ(f)) = λ(f),          

as required. □

Notice that we have not currently proved that ||µ|| = ||λ|| in the case K=ℂ. See a textbook for this.

site search by freefind advanced

Last modified: November 6, 2024.
Previous Up Next