10.john - K.Hunter - Measure Theory 14pp 2007 - TM
10.john - K.Hunter - Measure Theory 14pp 2007 - TM
Measure Spaces
• ∅, X ∈ A;
• if E ∈ A, then E c ∈ A;
• if En ∈ A for n ∈ N, then
∞
[ ∞
\
En ∈ A, En ∈ A.
n=1 n=1
1
3 Remark It is surprisingly complicated to obtain B(R) by starting from
the open sets and taking successively complements, countable unions, and
countable intersections. This process requires uncountably many iterations
before it gives B(R).
If (X, A) is a measurable space, then a map
µ : A → [0, ∞]
is called a measure if:
• µ(∅) = 0;
• if {En | n ∈ N} is a collection of disjoint sets in A, meaning that
Em ∩ En = ∅ for m 6= n, then
∞
! ∞
[ X
µ En = µ (En ) .
n=1 n=1
Here, we use the obvious conventions for measures and sums that are
equal to ∞.
The second property in the definition of a measure is called ‘countable addi-
tivity’, and it expresses the idea that the ‘volume’ of a disjoint union should
be the sum of the ‘volumes’ of its individual parts. We call (X, A, µ) a
measure space.
It follows from the definition that: (a) if E ⊂ F are measurable sets,
then µ(E) ≤ µ(F ); (b) if E1 ⊂ E2 ⊂ E3 ⊂ . . . is an increasing sequence of
measurable sets, then
∞
!
[
µ En = lim µ(En );
n→∞
n=1
2
Then ν is a measure on P(X) called counting measure. If X = N and
En = {1, 2, 3, . . . , n} ,
En = {n, n + 1, n + 2, . . .} ,
m (I) = Length(I)
for any interval I ⊂ R. Here, the length of an interval is defined in the usual
way; for example, Length ((a, b]) = b − a. More generally, there is a unique
Lebesgue measure m : B(Rd ) → [0, ∞] on Rd such that
m (I) = Volume(I)
3
measure is zero; for example, the standard Cantor set has Lebesgue measure
zero. In general, the Lebesgue measure of a subset N ⊂ Rd is zero if and
only if for every > 0 there is a countable collection {In | n ∈ N} of, not
necessarily disjoint, d-dimensional rectangles such that
∞
[ ∞
X
N⊂ In , Volume(In ) < .
n=1 n=1
4
finitely-additive translation and rotation invariant measures on R2 , but they
still have some counter-intuitive properties. For example, Laczkovich (1990)
proved that it is possible to decompose a circular disc into a finite number of
pieces (in Laczkovich’s proof, approximately 1050 non-Lebesgue measurable
sets) and reassemble them into a square of the same area.
Measurable Functions
f : X → R,
we equip R with its Borel σ-algebra B(R). Then f is measurable if and only
if f −1 ((a, ∞)) ∈ A for every a ∈ R. A function
f :X→C
is measurable if and only if its real and imaginary parts <f and =f are
measurable.
It is often convenient to consider extended real-valued functions
f : X → R,
5
A function φ : X → R is a simple function if
N
X
φ(x) = cn χEn (x)
n=1
Integration
6
13 Remark In this definition, we approximate the function f from below by
simple functions. In contrast with the definition of the Riemann integral, it
is not necessary to approximate a measurable function from both above and
below in order to define its integral.
If f : X → R, we write f = f+ − f− in terms of its positive and negative
parts
f+ = max{f, 0}, f− = max{−f, 0},
and define Z Z Z
f dµ = f+ dµ − f− dµ,
R R
provided that at least one of the integrals f+ dµ, f− dµ is finite.
We also have Z Z Z
|f | dµ = f+ dµ + f− dµ.
7
and if f, g : X → C, λ ∈ C then
Z Z
f dµ ≤ |f | dµ,
Z Z
λf dµ = λ f dµ,
Z Z Z
(f + g) dµ = f dµ + g dµ.
where the integral converges if and only if the series is absolutely conver-
gent. Thus, the theory of absolutely convergent series is a special case of
the Lebesgue integral. Note that conditionally convergent series, such as the
alternating harmonic series, do not correspond to a Lebesgue integral, since
both their positive and negative parts diverge.
15 Example Any Riemann integrable function f : [a, b] → R is integrable
with respect to Lebesgue measure, and the Riemann integral is equal to the
Lebesgue integral, Z b Z
f (x) dx = f dm.
a [a,b]
Thus, all of the usual results from elementary calculus remain valid for the
Lebesgue integral on R. A Lebesgue integrable function, however, need not
be Riemann integrable. In fact, a bounded function f : [a, b] → R is Riemann
integrable if and only if it is Lebesgue measurable and the set of disconti-
nuities {x ∈ [a, b] | f is discontinuous at x} has Lebesgue measure zero. For
example, the characteristic function of a Cantor set with non-zero measure
is Lebesgue integrable, but it is not Riemann integrable, nor is any modifi-
cation of the function on a set of measure zero Riemann integrable. We will
often write an integral with respect to Lebesgue measure on R or Rd as
Z
f dx.
8
Convergence Theorems
9
18 Theorem [Dominated Convergence] Suppose that (fn ) is an sequence of
measurable functions fn : X → C such that fn → f pointwise and |fn | ≤ g
where g : X → [0, ∞] is an integrable function, meaning that
Z
g dµ < ∞.
Then Z Z
fn dµ → f dµ as n → ∞.
Product Measures
Fubini’s theorem provides a simple and general condition for the equality
of multiple and iterated integrals of a function: namely, the function should
be integrable (see Theorem 23 for multiple integrals on Rd ).
Suppose that (X, A), (Y, B) are measurable spaces. The product σ-algebra
on the Cartesian product X × Y is the sigma algebra A ⊗ B generated by
the collection of all measurable rectangles {A × B | A ∈ A, B ∈ B}. Note
that this collection is not itself a σ-algebra; for example, the union of two
rectangles is not in general another rectangle.
19 Example Consider Rm , Rn equipped with the Borel σ-algebras B(Rm ),
B(Rn ). Then Rm × Rn = Rm+n , and one can show that
10
19 Theorem Suppose that (X, A, µ), (Y, B, ν) are σ-finite measure spaces.
There is a unique measure µ ⊗ ν : A ⊗ B → [0, ∞] on X × Y such that
are measurable.
21 Theorem [Fubini] A measurable function f : X × Y → C is integrable if
and only if either one of the iterated integrals
Z Z Z Z
y
|f | dµ dν, |fx | dν dµ
then
∞ ∞
! ∞ ∞
!
X X X X
amn = amn .
m=1 n=1 n=1 m=1
11
Finally, we state a version of Fubini’s theorem for functions on Rd .
23 Theorem A Lebesgue measurable function f : Rm+n → C is integrable,
meaning that Z
|f (x, y)| dxdy < ∞,
Rm+n
if and only if either one of the iterated integrals
Z Z Z Z
|f (x, y)| dx dy, |f (x, y)| dy dx
Rn Rm Rm Rn
Lp -spaces
One can also define the space L∞ (X) of essentially bounded measurable
functions, with norm given by the essential supremum. (Here ‘essential’
means ‘up to sets of measure zero’, but we omit the details.)
12
24 Theorem If 1 ≤ p < ∞, then Lp (X) equipped with the norm k · kp is a
Banach space.
The theorem includes the statement that the Lp -norm satisfies the trian-
gle inequality,
kf + gkp ≤ kf kp + kgkp .
This is called Minkowski’s inequality and it does not hold for p < 1, which
explains the restriction on p. It also includes the statement that Lp is com-
plete, meaning that if (fn )∞ p
n=1 is any Cauchy sequence in L , then there exists
f ∈ Lp such that kfn − f kp → 0 as n → ∞. The sequence need not converge
pointwise a.e., but there is a subsequence (fnk )∞
k=1 such that fnk → f point-
wise a.e. as k → ∞. Any function in Lp is the pointwise a.e. and Lp -limit
of a sequence of simple functions.
Finally, we give a theorem which states that Lp -functions on Rd can be
approximated by continuous functions. We consider functions f : Ω → C
where Ω ⊂ Rd is an arbitrary nonempty open set. We denote by Cc (Ω) the
set of continuous functions f : Ω → C with compact support in Ω, meaning
that the closure in Ω of the set {x ∈ Ω | f (x) 6= 0} is a compact set. A
function f ∈ Cc (Ω) is bounded and nonzero on a bounded set, so f ∈ Lp (Ω).
25 Theorem Cc (Ω) is a dense subset of Lp (Ω).
More explicitly, this theorem states that if f ∈ Lp (Ω) then, given any
> 0, there exists g ∈ Cc (Ω) such that
kf − gkp < .
It follows that Lp (Ω) is the completion of Cc (Ω) with respect to the Lp -norm.
13
References
14