Matematyka w6 7
Matematyka w6 7
i
ii 1. LEBESGUE MEASURE AND INTEGRATION
Remark 2. Here we will assume that φ is not the constant functions whose only value is +∞
or −∞, and that the range does not contain both +∞ and −∞, because if it did, the right side of
(1.1.4) could lose meaning.
Remark 3. Note that the left side of (1.1.5) is independent of the order in which the An ’s
are arranged. Hence, by the rearrangement theorem for series, if the right hand side of (1.1.5))
converges, it converges absolutely. Otherwise, the partial sums tend to +∞ or −∞.
Theorem 1. If φ is additive, then
(1) φ(∅)
k= 0.
S Pk
(2) φ An = n=1 φ(An ), if Ai ∩ Aj = ∅ for i 6= j.
n=1
(3) φ (A1 ∪ A2 ) + φ (A1 ∩ A2 ) = φ (A1 ) + φ (A2 ).
(4) If φ(A) ≥ 0 for all A, and A ⊂ B, then φ (A) ≤ φ (B).
added — PS (5) if A ⊂ B,and |φ (A)| < ∞, then φ (B − A) = φ (B) − φ (A).
and
∞
[ ∞
X
φ(A) = φ Bj = φ(Bj ).
j=1 j=1
iv 1. LEBESGUE MEASURE AND INTEGRATION
So
n
X ∞
X
lim φ(An ) = lim φ(Bj ) = φ(Bj ) = φ(A)
n→∞ n→∞
j=1 j=1
So
∞
X
(1.1.6) φ (A) = sup φn (A) ≤ φ (Aj )
n∈N j=1
k
S
On the other hand, Since Bk = Aj ⊂ A, then
j=1
Fix k ∈ N and ε > 0, for each j = 1, 2, . . . , k there exists nj so that φ (Aj ) < φnj (Aj ) + ε/k.
Since {φn (Aj ) : n ∈ N} is increasing, taking n = max {nj : j = 1, 2, . . . , k}, and using the countable
1.2. CONSTRUCTION OF THE LEBESGUE MEASURE v
aditivity of φn we have
Xk Xk k
[ k
[
φ (Aj ) ≤ φn (Aj ) + ε = φn Aj + ε ≤ φ Aj + ε ≤ φ (A) + ε.
j=1 j=1 j=1 j=1
k
X
(1.2.3) m (A) = m(Ij )
j=1
We denote by E the family of all elementary subsets of Rp . Note that E satisfies the following
properties.
k
S l
S
E 1 E is a ring, but not a σ-ring. Clearly the if A = In and B = Jm , then A ∪ B =
n=1 m=1
k
S l
S
In ∪ Jm and
n=1 m=1
k l
! k l
!
[ [ [ \
A−B = In − Jm = (In − Jm )
n=1 m=1 n=1 n=1
k+1
[ k
[ l
[
A= In = Ik+1 ∪ In = Ik+1 ∪ Jn
n=1 n=1 n=1
l
! l
[ [
= Ik+1 − Jn ∪ Jn
n=1 n=1
l
! l
\ [
= (Ik+1 − Jn ) ∪ Jn
n=1 n=1
where the Jn are disjoint intervals, since Ik+1 − Jn is the finite union of disjoint intervals
(see 5) and the intersection of intervals is an interval, then A is the union of a finite number
of disjoint intervals.
1.2. CONSTRUCTION OF THE LEBESGUE MEASURE vii
k
S l
S
E 4 m is additive on E. Indeed, If A = Ir and B = Jq , where the intervals Ir and Jq
r=1 q=1
are pairwise disjoint, and A ∩ B = ∅, then Ir ∩ Jq = ∅ for each r = 1, 2, . . . , k and q = 1,
2, . . . , l.
Sk Sl
Since A ∪ B = Ir ∪ Jq , then
r=1 q=1
k l
! k l
[ [ X X
m (A ∪ B) = m Ir ∪ Jq = m (Ir ) + m (Jq ) = m (A) + m (B)
r=1 q=1 r=1 q=1
Exercise 1. [Exercise 11.15] Let R be the ring of all elementary subsets of (0, 1]. If 0 < a <
b ≤ 1, define
φ ((a, b)) = φ ((a, b]) = φ ([a, b)) = φ ([a, b]) = b − a
but define
φ ((0, b)) = 1 + b
if 0 < b ≤ 1. Show that this gives an additive set function φ on R, which is not regular and which
cannot be extended to a countably additive set function on a σ-ring.
Solution 1. Here as in Definition 6 we define
Xk
(1.2.5) φ (A) = φ(Ij ),
j=1
k
S
if A = Ij , and the intervals Ij are pairwise disjoint.
j=1
First, if A is an elementary set, φ(A) is well defined by (1.2.5); that is, if two different decom-
positions of A into disjoint intervals are used, each gives rise to the same value of φ(A). Indeed, if
k
S Sl
A= In = Jm , where the intervals In and Jm are pairwise disjoint, then for each n = 1, 2,
n=1 m=1
. . . , k and m = 1, 2, . . . , l we have
l
[
In = A ∩ In = (Jm ∩ In )
m=1
k
[
Jm = A ∩ Jm = (Jm ∩ In )
n=1
since the family {Bmn = Jm ∩ In : n = 1, 2, . . . , k and m = 1, 2, . . . , l} is pairwise disjoint, then
k k l
! k l
!
X X X X X
φ(A) = φ(In ) = φ (Jm ∩ In ) = φ (Bmn )
n=1 n=1 m=1 n=1 m=1
l k
! l k
! l
X X X X X
= φ (Bmn ) = φ (Jm ∩ In ) = φ(Jm )
m=1 n=1 m=1 n=1 m=1
k
S
Recall that if A is an elementary set, then A = Ij is the union of a finite number of disjoint
j=1
intervals (see 1.2). So
( Pk
j=1 l(Ij ) if 0 is not the end point of any interval in A
φ (A) = P k
1 + j=1 l(Ij ) if 0 is the end point of any interval in A
where l(Ij ) is the length of the interval Ij . In particular, φ (A) < 1 if A is a closed set of (0, 1].
Indeed, note that 0 ∈ / A for any subset A ⊂ (0, 1].
Now, if 0 is the endpoint of any interval in A, since A is closed then 0 is limit point of A, and
0 ∈ A. which is clearly a contradiction.
If two elementary sets A and B are disjoint, at most one of them can have the point 0 as
the endpoint of one of its intervals. Then φ (A ∪ B) is the sum of the lengths of the intervals in
A ∪ B if neither set contains an interval having 0 as the endpoint, and 1 plus this sum if one of
1.2. CONSTRUCTION OF THE LEBESGUE MEASURE ix
them does contain an interval with 0 as endpoint. In either case φ (A ∪ B) = φ (A) + φ (B) when
A ∩ B = ∅.Thus, the function φ is additive.
The function φ is not regular, because by definition φ 0, 21 = 1 + 12 = 32 , but φ(A) < 1 if A
Example 1.
(b) Take p = 1, and let α be a monotonically increasing function defined for all real x. Put
µ ([a, b)) = α (b−) − α (a−) = sup α (t) − sup α (t)
t<b t<a
µ ([a, b]) = α (b+) − α (a−) = inf α (t) − sup α (t)
b<t t<a
µ ((a, b]) = α (b+) − α (a+) = inf α (t) − inf α (t)
b<t a<t
µ ((a, b)) = α (b−) − α (a+) = sup α (t) − inf α (t)
t<b a<t
Recall that if α is a monotonically increasing function then the set of points of discontinuity of α
is at most countable, and if α is continuous at x, then α (x−) = α (x) = α (x+). Also for each a,
b ∈ R with a < b, by the definition of infimum and supremum given ε > 0 there exists c, x, y, d
with c < a < x < y < b < d, so that α is continuous at c, x, y, d and
ε ε
α (a−) − < α (c) α (x) < α (a+) +
2 2
ε ε
α (b−) − < α (y) α (d) < α (b+) +
2 2
x 1. LEBESGUE MEASURE AND INTEGRATION
or equivalently
ε ε
−α (c) − < −α (a−) − α (a+) < −α (x) +
2 2
ε ε
α (b−) < α (y) + α (d) − < α (b+) .
2 2
The behavior of a monotonically increasing function around a discontinuity point x is sketched in
the figure below
Now we show that every regular set function on E can be extended to a countably additive set
function on a σ-ring which contains E.
Definition 8. Let µ be additive, regular, non-negative, and finite on E. Consider countable
coverings of any set E ⊂ Rp by open elementary sets An .
[∞
E⊂ An .
n=1
1.2. CONSTRUCTION OF THE LEBESGUE MEASURE xi
Define
∞
X
(1.2.6) µ∗ (E) = inf µ (An ) ,
n=1
where the infimum is taken over all countable coverings of E by open elementary sets.
µ∗ is called the outer measure of E, corresponding to µ.
It is clear that µ∗ (E) ≥ 0 for all E and that if E1 ⊂ E2 , then any countable coverings of E2
by open elementary sets is a countable coverings of E1 by open elementary sets and by properties
of infimum we have
(1.2.7) µ∗ (E1 ) ≤ µ∗ (E2 ) .
Theorem 4.
(1.2.11) S (A, B) = (A − B) ∪ (B − A)
Thus,
S (A, B) = (A − B) ∪ (B − A)
⊂ ((A − C) ∪ (C − B)) ∪ ((B − C) ∪ (C − A))
= ((A − C) ∪ (C − A)) ∪ ((C − B) ∪ (B − C))
= S (A, C) ∪ S (C, B)
S3
D1
d (A, B) = µ∗ (S (A, B)) = µ∗ (S (B, A)) = d (B, A)
d (A, A) = µ∗ (S (A, A)) = µ∗ (∅) = 0
D2 Since S (A, B) ⊂ S (A, C) ∪ S (C, B), by (1.2.7) and (1.2.8)
d (A, B) = µ∗ (S (A, B)) ≤ µ∗ (S (A, C) ∪ S (C, B))
≤ µ∗ (S (A, C)) + µ∗ (S (C, B)) = d (A, C) + d (C, B)
D3 As in the proof of D2, the inclusions in S3 and (1.2.7) and (1.2.8) imply
d (A1 ∪ A2 , B1 ∪ B2 )
d (A1 ∩ A2 , B1 ∩ B2 ) ≤ d (A1 , B1 ) ∪ d (A2 , B2 ) .
d (A1 − A2 , B1 − B2 )
The relations D1 and D2 show that d(A, B) satisfies the requirements of definition for a distance
except that d(A, B) = 0 does not imply A = B. For instance, if p = 1, µ = m, A = {an ∈ R : n ∈ N}
is countable, and B = ∅, then
d(A, B) = m∗ ((A − ∅) ∪ (∅ − A)) = m∗ (A)
ε ε ∞
S
If ε > 0, taken In = an − n+1 , an + n+1 , then In are elementary open sets and A ⊂ In
2 2 n=1
and
∞ ∞
X X ε
m∗ (A) ≤ m (In ) = n
=ε
n=1 n=1
2
Since ε is arbitrary, then m∗ (A) = 0.
If B = ∅, then D2 tells us that
µ∗ (A) = d(A, ∅) ≤ d(A, C) + d(C, ∅) = d(A, C) + µ∗ (C)
interchanging A and C we get
µ∗ (C) ≤ d(A, C) + µ∗ (A)
So if at least one of µ∗ (A) , µ∗ (C) is finite, then
(1.2.14) |µ∗ (A) − µ∗ (C)| ≤ d(A, C)
We write An → A, if
lim d(A, An ) = 0.
n→∞
If there is a sequence {An } of elementary sets such that An → A, we say that A is finitely
µ-measurable and write A ∈ MF (µ) .
1.2. CONSTRUCTION OF THE LEBESGUE MEASURE xv
So µ∗ (A) < ∞. By (1.2.15a) and (1.2.15c), MF (µ) is a ring. Since µ is additive by item 3 in
Theorem 1 we have µ (An ) + µ (Bn ) = µ (An ∪ Bn ) + µ (An ∩ Bn ) taking limit when n → ∞, by
(1.2.15d) and Theorem 2 (a) we obtain
µ∗ (A) + µ∗ (B) = µ∗ (A ∪ B) + µ∗ (A ∩ B) .
Since A ∩ B = ∅, implies µ∗ (A ∩ B) = 0, then if A ∩ B = ∅ we have
µ∗ (A) + µ∗ (B) = µ∗ (A ∪ B) .
And µ∗ is additive on MF (µ).
Now let
∞
[
A= An
n=1
with An ∈ MF (µ). for n = 1, 2, 3, . . . , Then Bn ∈ MF (µ) because Bn is a finite union of elements
∞
S
of MF (µ) , B1 ⊂ B1 ⊂ · · · ⊂ A, and A = Bn , if C1 = B1 and Cn = Bn − Bn−1 for n = 2, 3, · · · ,
n=1
then Cn ∈ MF (µ) because Cn is the difference of elements of MF (µ) , and as we see in the proof
∞
S
of Theorem 2 Ci ∩ Cj = ∅ for i 6= j and A = Ck .
k=1
By (1.2.8)
∞
X
(1.2.16) µ∗ (A) ≤ µ∗ (Ck ) .
k=1
n
Ck ⊂ A. the additivity of µ∗ implies
S
On the other hand, with the above notation Bn =
k=1
n n
!
X [
(1.2.17) µ∗ (Ck ) = µ∗ Ck = µ∗ (Bn ) ≤ µ∗ (A)
k=1 k=1
xvi 1. LEBESGUE MEASURE AND INTEGRATION
(a) If A is open, then A ∈ M (µ). Because every open set in Rp is the union of a countable
collection of intervals. To see this, using the density of Q in R, we can see that β =
{I = (a1 , b1 ) × (a2 , b2 ) × · · · × (ap , bp ) : ai , bi ∈ Q} is a countable base whose elements are
open intervals. Since Rp is an open set, taking complements we obtain that closed set is
in M (µ)
(b) If A ∈ M (µ) and ε > 0, there exist sets F and G that F ⊂ A ⊂ G, F is closed, G is open,
and
(1.2.20) µ (G − A) < ε, µ (A − F ) < ε.
Indeed, if µ (A) < ∞, i.e., A ∈ MF (µ) by (1.2.6), there exists a sequence {An } of
open elementary sets, so that
∞
[ ∞
X
A⊂ An and µ (An ) < µ (A) + ε,
n=1 n=1
∞
S
taking G = An , then
n=1
∞
X
µ (G − A) = µ (G) − µ (A) ≤ µ (An ) − (µ (A)) < ε.
n=1
For the second inequality, since M (µ) is a σ-ring, then Ac ∈ M (µ) , so there exists G
µ (G − Ac ) < ε.
xviii 1. LEBESGUE MEASURE AND INTEGRATION
(f) In case of the Lebesgue measure, every countable set has measure zero. In effect, for
each ε > 0, the interval centered on x and of volume 2ε is a covering of A = {x} by
elementary sets, with measure less than ε. As any countable set B is the countable union
of its elements, the remark 6 (e) shows that B has zero measurement. But there are
1.2. CONSTRUCTION OF THE LEBESGUE MEASURE xix
uncountable sets of measure zero. The Cantor set P is an example: Recall that P is
defined as
∞
\
P = En ,
n=0
where E0 = [0, 1]. Removing the the middle thirds ibterval of this intervals we obtain
E1 = E0 − ( 31 , 23 ), so E1 is [0, 13 ] ∪ [ 23 , 1]. Removing the middle thirds interval of these
intervals we obtain E2 and continuing in this way, we obtain a sequence of compact sets
En , such that
E0 ⊃ E1 ⊃ E2 · · ·
2 n
and En is the union of 2 intervals, each of length 3−n . So m (En ) =
n
3 .
Below we show the E1 , E2 , E3 and E4 .
Now P can be identified with the set of the sequences (a0 a1 . . . an . . .) where an = 0 or an = 2,
and using the same argument as the one that shows that [0, 1] is uncountable, we get that P is
uncountable. n
Since P ⊂ En for all n ∈ N, then m (P ) ≤ m (En ) = 32 , and taking limit when n → ∞ we
obtain m (P ) = 0.
Moreover, if we denote by c the cardinality of R, we obtain that cardinality of P is equal to c.
Because every subset of the set of measure 0 is the set of measure 0, we have at least 2c measurable
sets on R. Because the cardinality of the family of all subset of R is also 2c , the natural question
is: ‘are there unmeasurable sets?’. added — PS
(g) A Vitali set is a subset V of the interval [0, 1] of real numbers such that, for each real
number r, there is exactly one number v ∈ V such that v − r is a rational number. Vitali
sets are a set of representative of the group R/Q in [0, 1]. added — PS
Every Vitali set V is uncountable, and
(1.2.22) v − u is irrational for any u, v ∈ V, u 6= v.
A Vitali set is non-measurable. Indeed, assume that V is measurable and let q1 , q2 ,
. . . be an enumeration of the rational numbers in [−1, 1]. And let Vn be the translated
sets defined by
Vn = V + qn = {v + qn : v ∈ V },
Note that Vn ∩ Vm = ∅, because if y ∈ Vn ∩ Vm , then v + qn = y = u + qm , implies v − u
is rational in contradiction with (1.2.22).
∞
S
Also note that [0, 1] ⊆ Vn ⊆ [−1, 2].
n=1
xx 1. LEBESGUE MEASURE AND INTEGRATION
To see the first inclusion, consider any real number r ∈ [0, 1] and let v be the repre-
sentative in V for the equivalence class [r]; then r − v = qn for some rational number qn
∈ [−1, 1] which implies that r ∈ Vn
Since the Lebesgue measure is countably additive, then
∞
X
1≤ m(Vn ) ≤ 3.
n=1
Because the Lebesgue measure is translation invariant, we have m(Vn ) = m(V ) for all
n ∈ N, and therefore
X ∞
1≤ m(V ) ≤ 3.
n=1
But this is impossible. Summing infinitely many copies of the constant m(V ) yields either
zero or infinity, according to whether the constant is zero or positive. In neither case is the
sum in [1, 3]. So V cannot bee measurable. An adequate change of the above argument
shows that, for all measurable set A with m(A) > 0 there exists a non-measurable set B
with B ⊂ A.
Reciprocally, if f (x) > a − n1 for all n ∈ N, then f (x) is an upper bound of the set
a − n : n ∈ N , since a = sup a − n1 : n ∈ N we have f (x) ≥ a and
1
\ 1
x : f (x) > a − ⊂ {x : f (x) ≥ a}.
n
n∈N
Reciprocally, If f (x) < a + n1for all n ∈ N, then f (x) is lower bound of the set
1
a + n : n ∈ N , since a = inf a + n1 : n ∈ N we have f (x) ≥ a and
\ 1
x : f (x) < a + ⊂ {x : f (x) ≤ a}.
n
n∈N
Note that
{x : −f (x) > a} = {x : f (x) < −a}
Hence {x : −f (x) > a} is measurable for every real a if f is a measurable function.
Hence any of these conditions may be used instead of (1.2.21) to define measurability
Theorem 7. Let f be a measurable function on a measurable space X, then |f | is a measurable
function on X.
If a < 0, then {x : |f (x)| > a} = X, hence measurable.
Recall that |f (x)| > a, if and only if f (x) > a or −f (x) < −a. So
{x : |f (x)| > a} = {x : f (x) > a} ∪ {x : −f (x) < −a} ,
and by Theorem 6 and corollary 1 {x : |f (x)| > a} is measurable for every real a if f is a measurable
function.
Theorem 8. Let {fn } be a sequence of measurable functions on a measurable space X. Then
g(x) = sup {fn (x) : n ∈ N}
h(x) = inf {fn (x) : n ∈ N}
k(x) = lim sup {fn (x)}
n→∞
l(x) = lim inf {fn (x)}
n→∞
where 0(x) = 0.
(b) If f (x) = lim fn (x), then
n→∞
Theorem 9. Let f and g be measurable real valued functions defined on X, let F be a real
valued and continuous function on R2 , and put
h(x) = F (f (x), g(x))
then h is measurable.
In particular, f + g and f · g are measurable.
Let
Ga = {(u, v) : F (u, v) > a} = F −1 (a, +∞)
Since F is continuous, then Ga is an open subset of R2 , recall that every open set in R2 is the union
of a countable collection of intervals (because both open squares and open disks form a topological
∞
basis in R2 ), so we can write Ga =
S
In , where {In } is a sequence of open intervals: added — PS
n=1
In = (an , bn ) × (cn , dn ) .
Since f and g are measurable, then the sets
{x : an < f (x) < bn } = f −1 (an , bn ) ,
{x : cn < g(x) < dn } = g −1 (cn , dn ) ,
are measurable. So
{x : (f (x), g(x)) ∈ In } = {x : an < f (x) < bn } ∩ {x : cn < g(x) < dn }
is measurable, Hence the same is true for
∞
[
{x : (f (x), g(x)) ∈ In } = {x : (f (x), g(x)) ∈ Ga } = {x : h(x) = F (f (x), g(x)) > a} .
n=1
Above we see that all ordinary operations of analysis when we apply to measurable functions,
lead to measurable functions. but the next example shows that there exits continuous function g
and a measurable function h such that h ◦ g is non-measurable.
Exercise 2. [Exercise 11.3] If {fn } is a sequence of measurable functions, prove that the set
of points at which {fn (x)} converges is measurable.
Solution 2. Let A = {x : {fn (x)} converges}, if x ∈ A then {fn (x)} is a Cauchy sequence in
R. Thus, given n ∈ N, there existsn0 ∈ N such that |fp (x) − fq (x)| < n1 if p, q > n0 .
Note that for each n ∈ N. we have
\ 1
[ \ 1
x∈ z : |fp (z) − fq (z)| < ⊂ z : |fp (z) − fq (z)| <
p,q>n
n p,q>n
n
0 n0 ∈N 0
Clearly by construction if x ∈ C, then {fn (x)} is a Cauchy sequence in R, so {fn (x)} converges
and x ∈ A. in consequence A = C.
since fn is measurable for all n, then |fp − fq | is measurable for all p, q, in consequence
Finally,
the sets z : |fp (z) − fq (z)| < n1 are measurable for all p, q and n. Since the measurable sets form
a σ-ring then A = C is a measurable set.
Example 3. Consider the following sequence {fn } of functions defined by
0, 0 ≤ x < 31
1 1 2
f0 (x) =
2, 3 ≤x< 3
1, 23 ≤ x ≤ 1
1
0 ≤ x < 31
2 fn−1 (3x),
1 1 2
fn (x) =
2, 3 ≤x< 3
1 2
2 (1 + fn−1 (3x − 2)) , 3 ≤x≤1
This sequence of functions converges to the continuous function f know as the devil’s stair-
case function of Cantor, below we can see the graph of the first three functions fn and a sketch
of the graph of f .
Consider the function g(x) = f (x) + x, where f is the devil’s staircase function of Cantor. The
function g is strictly increasing homeomorphism from [0, 1] onto [0, 2] and note that the measure
1.5. SIMPLE FUNCTIONS xxv
of [0, 2] − g(C), where C is the Cantor set, is equal to two minus the sum of the measures of the
images of the middle thirds of the intervals in En , that we use to defined C, i.e.,
∞
2n
X
m([0, 2] − g(C)) = 2 − =1
n=0
3n+1
in the figure below we can see a sketch of the graph of g,
In the general case, consider f = f + − f − and apply the preceding construction to f + and f − .
Note that if f is measurable, then Fn , En,k are measurable sets, so sn are measurable functions.
Note that if f ≥ 0, then functions sn in the proof satisfy sn ≥ 0, and if f is bounded, i.e.,
|f (x)| < M for all x ∈ X and some M ∈ R, then equation (1.5.6) tells us that the sequence {sn }
converges uniformly to f .
1.6. INTEGRATION
Now we define integration on a measurable space X, in which M is the σ-ring of measurable
sets, and µ is the measure.
Definition 13. Suppose s ≥ 0
n
X
s(x) = ck χEk (x) x ∈ X
k=1
Pn
Suppose that s(x) = k=1 ck χEk (x) x ∈ X, then
∞
n n
!
X X [
IE (s) = ck µ (E ∩ Ek ) = ck µ Fj ∩ Ek
k=1 k=1 j=1
n
X ∞
[ Xn ∞
X
= ck µ (Fj ∩ Ek ) = ck µ (Fj ∩ Ek )
k=1 j=1 k=1 j=1
∞ ∞
n X X n
!
X X
= ck µ (Fj ∩ Ek ) = ck µ (Fj ∩ Ek )
k=1 j=1 j=1 k=1
∞
X
= IFj (s) .
j=1
Remark 9. The above theorem tells us that if φs (E) = IE (s) for each E ∈ M, then φs is
countably additive.
Theorem 12. Let s1 , s2 ≥ 0 be measurable simple functions on X,
(1) if s1 ≤ s2 , then for each E ∈ M we have
(1.6.3) IE (s1 ) ≤ IE (s2 )
(2) s1 + s2 ≥ 0 is a measurable simple function on X, and for each E ∈ M we have
(1.6.4) IE (s1 + s2 ) = IE (s1 ) + IE (s2 )
Suppose that
n
X k
X
s1 (x) = ck χEk , s2 (x) = dj χFj
k=1 j=1
where ck ≥ 0, dj ≥ 0.
Sn k
S
Since X = Ei = Fj , then for 1 ≤ i ≤ n, 1 ≤ j ≤ k, let Aij = Fj ∩ Ei ∩ E
i=1 j=1
k
[ n
[
Ei ∩ E = Aij , Fj ∩ E = Aij
j=1 i=1
n
[ n [
[ k
E= (Ei ∩ E) = Aij
i=1 i=1 j=1
Note that if 0 ≤ s1 ≤ s2 are measurable simple functions and Aij are as the above proof, then
m
X
s2 − s1 = (dj − ci ) χAij ≥ 0
j=1
and
(1.6.5) IE (s2 − s1 ) = IE (s2 ) − IE (s1 )
Definition 14. If f is measurable and non-negative, we define
Z
(1.6.6) f dµ = sup IE (s) ,
E
where the supremum is taken over all measurable simple functions s such that 0 ≤ s ≤ f .
The left member of (1.6.6) is called the Lebesgue integral of f , with respect to the measure
µ, over the set E. Note that the integral may have the value +∞.
Note that if 0 ≤ s is a measurable simple function, then s ≤ s, and by Theorem 12 (1)
IE (s1 ) ≤ IE (s) for all measurable simple functions s1 such that 0 ≤ s1 ≤ s so
Z
(1.6.7) sdµ = IE (s) .
E
where f = f + − f − and f + and f − are defined as Corollary 2. If at least one of the integrals (1.6.8)
is finite, we define
Z Z Z
(1.6.9) f dµ = f + dµ − f − dµ.
E E E
If both integrals in (1.6.8) are finite, then the integral in (1.6.9) is finite, and we say that f is
integrable (or summable) on E in the Lebesgue sense, with respect to µ and we write f ∈ L (µ)
on E. If µ = m the Lebesgue measure, we say that f is Lebesgue integrable on E and that
f ∈ L on E.
Note that if the integral in (1.6.9) is +∞ or −∞, then the integral of f over E is defined,
although f is not integrable on E in the above sense; f is integrable on E only if its integral over
E is finite.
All properties in Remark 11.23 of Rudin’s books are true, but not are evidently as he say, here
we only show those we need, and lead to later the rest
Remark 10. The integral satisfies the following properties:
xxx 1. LEBESGUE MEASURE AND INTEGRATION
In the case 0 ≤ f ;
1. If λ ≥ 0, then s1 is a simple function with 0 ≤ s1 ≤ λf , if and only if s1 = λs where s is
a simple function with 0 ≤ s ≤ f . So
Z Z Z Z Z
λf dµ = sup {IE (λs)} = sup λs dµ = sup λ s dµ = λ sup s dµ = λ f dµ
E E E E E
and consequently f dµ
to the end −
2. If λ < 0, then λf ≤ 0, so (λf ) = (−λ) f
Z Z Z Z Z
−
λf dµ = − (λf ) dµ = − (−λ) f dµ = − (−λ) f dµ = λ f dµ
E E E E E
So
Z
f + dµ ≤ M · µ (E) < ∞
E
Z
f + dµ ≤ M · µ (E) < ∞
E
and f ∈ L (µ) on E.
(c) If f, g ∈ L (µ) , and 0 ≤ f (x) ≤ g(x), then
Z Z
(1.6.11) f dµ ≤ gdµ.
E E
Thus, Z
f dµ = sup IE (s) = 0.
E 0≤s≤f
f + dµ = E f − dµ = 0, so
R R
In the general case, we have E
Z Z Z
f dµ = f + dµ − f − dµ = 0.
E E E
then
Z Z
(1.6.15) f dµ = lim fn dµ
E n→∞ E
First note that the sequence {fn (x)} is an increasing sequence of real functions so
f (x) = lim fn (x) = sup {fn (x)}
n→∞ n∈N
and fRis a measurable non-negative function with fn (x) ≤ f (x) for all x ∈ E and n ∈ N. Also we
end of the proof not here have E fn (x)dµ is an increasing sequence of real extended numbers, so
Z Z
lim fn dµ = sup fn dµ = α
n→∞ E n∈N E
for some α, and since fn (x) ≤ f (x) for all x ∈ E and n ∈ N, we have
Z
(1.6.16) α≤ f dµ.
E
Choose c such that 0 < c < 1, and let s be a simple measurable function such that 0 ≤ s ≤ f
and let
En = {x : fn (x) ≥ cs (x)} ,
for n ∈ N. Since
By (1.6.13), E1 ⊂ E2 ⊂ · · · ⊂ En ⊂ · · · . If x ∈ E, since 0 < c < 1 then
cs (x) < f (x) = sup {fn (x)} ,
n∈N
Since the φs is a countably additive set function (see Remark 9) by (1.6.17 ) we can apply
Theorem 11 and Remark 10 (a) to φs in (1.6.18), and taking limit when n → ∞ in(1.6.18), we
obtain
Z Z
(1.6.19) α = lim fn dµ ≥ lim cφs (En ) = cφs (E) = c sdµ
n→∞ E n→∞ E
1.6. INTEGRATION xxxiii
then write
Z Z Z
(1.6.23) φf (A) = f dµ = f + dµ − f − dµ = φf + (A) − φf − (A)
A A A
and apply (a) to φf + and φf − we obtain (b).
To prove (a), by Theorem (10) there exists a sequence {sn } of simple functions such that
(1.6.24) 0 ≤ s1 (x) ≤ s2 (x) ≤ ... ≤ f
for all x ∈ X. So that
(1.6.25) f = lim sn = sup {sn : n = 1, 2, 3, . . .}
n→∞
Remark 10 (c) shows that φsn ≤ φsn+1 for n = 1, 2, 3, . . . . So {φsn : n = 1, 2, 3, . . .} is a
sequence of countably additive functions on M. By Lebesgue’s monotone convergence theorem
Z Z Z
(1.6.26) φf (A) = f dµ = lim sn = sup sn dµ = sup {φsn (A)} ,
A n→∞ A n∈N A n∈N
and by Theorem 3, φf is a countably additive function on M.
Corollary 3. If A, B ∈ M, B ⊂ A and µ (B − A) = 0, then
Z Z
(1.6.27) f dµ = f dµ
A B
Remark 11. The preceding corollary shows that sets of measure zero are negligible in integra-
tion
xxxiv 1. LEBESGUE MEASURE AND INTEGRATION
Thus, µ ({x : f (x) 6= h(x)} ∩ E) = 0 if µ ({x : f (x) 6= g(x)} ∩ E) = 0 and µ ({x : g(x) 6= h(x)} ∩ E) =
0. Hence f ∼ g and g ∼ h implies f ∼ h. In conclusion, ∼ is an equivalence relation on E
If f ∼ g on E, then f ∼ g on A, for all A ⊂ E. let, B = {x : f (x) = h(x)} ∩ A and
C = {x : f (x) 6= h(x)} , then A = B ∪ C, B ∩ C = ∅. So if f and g are integrable on A then
Z Z Z Z
(1.6.30) f dµ = f dµ = gdµ = gdµ.
A B B A
If a property p holds for x ∈ E − A and if µ(A) = 0, we say that p holds for almost all x ∈ E,
or that p holds almost everywhere on E.
n
Note that if µ (A) > 0, where A = {x : f (x) = +∞} , then sn (x) = µ(A) χA ≤ f (x), so
Z Z
(1.6.31) n= sn dµ ≤ f dµ.
A A
R
Thus, A f dµ = +∞ and f ∈ / L(µ) on E, in consequence if f ∈ L(µ) on E, then f (x) must be finite
almost everywhere on E. In most cases we therefore do not lose generality if we assume the given
functions to be finite-valued from the outset.
R
Exercise 3. Let f be a measurable function on E. Then, if f ≥ 0 and E f dµ = 0, prove that
= 0 almost everywhere on E. Hint: Let En be the subset of E on which f (x) > n1 . Write
f (x) S
∞
A = n=1 En . Then µ(A) = 0 if and only if µ(En ) = 0 for every n.
Solution 3. Note that for every n, En ⊂ A , so if µ(A) = 0, then µ(En ) = 0 also. Conversely,
Not labelled if µ(En ) = 0 for every n, by Corollary ?? we have
∞
[ ∞
X
µ(A) = µ( En ) ≤ µ(En ) = 0
n=1 n=1
On the other hand, if µ(En ) > 0 for any n, since f ≥ n1 χEn on E,then
Z Z
1 1 1
0= f dµ ≥ χEn dµ = µ(E ∩ En ) = µ(En ) > 0
E E n n n
which is a clearly contradiction.
R
Exercise 4. [Exercise 11.2] Let f be a measurable function on E. Then, if A f dµ = 0 for
every measurable subset A of a measurable set E, then f (x) = 0 almost everywhere on E.
1.6. INTEGRATION xxxv
A+ ∩ A− = ∅, E = A+ ∪ A−
Note that
f + (x) x ∈ A+
f (x) = ,
−f − (x) x ∈ A−
f (x) x ∈ A+
f + (x) =
0 x ∈ A−
0 x ∈ A+
f − (x) =
−f (x) x ∈ A−
A = (A+ ∩ A) ∪ (A− ∩ A)
Since
Z Z Z Z Z
+ + + +
f dµ = f dµ + f dµ = f dµ + 0dµ = 0 + 0 = 0
A A+ ∩A A− ∩A A+ ∩A A− ∩A
Z Z Z Z Z
f − dµ = f − dµ + f − dµ = 0dµ + f − dµ = 0 + 0 = 0,
A A+ ∩A A− ∩A A+ ∩A A− ∩A
Since f + and f − are non-negative measurable functions, then exercise 11.1 implies that f + (x) =
0 and f − (x) = 0 almost everywhere on E, so
almost everywhere on E.
By Theorem 14 we have
Z Z Z Z
(1.6.33) f + dµ = φf + (E) = φf + (A) + φf + (B) = f + dµ + f + dµ = f + dµ
ZE ZA ZB ZA
− − −
(1.6.34) f dµ = φf − (E) = φf − (A) + φf − (B) = f dµ + f dµ = f . dµ
E A B B
and,
Z
|f | dµ = φ|f | (E) = φ|f | (A) + φ|f | (B)
E
Z Z Z Z
= |f | dµ + |f | dµ = +
f dµ + f − dµ
A B A B
Z Z
= f + dµ + f − dµ < ∞.
E E
Since the integrability of f implies that |f | s integrable, the Lebesgue integral is called an
absolutely convergent integral.
Theorem 16. Let f be a measurable function on E and g ∈ L (µ) on E. If |f | ≤ g, then
f ∈ L (µ) on E.
Note that 0 ≤ f + , f − ≤ |f | ≤ g so
Z Z
+
0≤ f dµ ≤ gdµ < ∞
ZE ZE
0≤ f − dµ ≤ gdµ < ∞
E E
and f ∈ L (µ) on E.
Exercise 5. [Exercise 11.4] If f ∈ L(µ) on E and g is bounded and measurable on E, then
f g ∈ L(µ) on E.
Solution 5. Since g is bounded, let M a positive constant such that |g (x)| ≤ M for all x ∈ E.
Since f ∈ L(µ) on E, by Remark 10 (a) and Theorem ?? we have M |f | ∈ L(µ) on E. Since f and
g are measurable, f g is also measurable. Now
|f g| (x) = |g (x)| |f (x)| ≤ M |f (x)| = M |f | (x)
so by Theorem ?? we have f g ∈ L(µ) on E.
Theorem 17. If f , g ∈ L (µ) on E, then
Z Z Z
(1.6.35) (f + g) dµ = f dµ + gdµ
E E E
1.6. INTEGRATION xxxvii
In the case 0 ≤ f , g let {sn }, {pn } be sequences of simple functions such that
0 ≤ s1 (x) ≤ s2 (x) ≤ . . . ≤ f,
0 ≤ p1 (x) ≤ p2 (x) ≤ . . . ≤ g,
for all x ∈ E. and
(1.6.36) f (x) = lim sn (x), g(x) = lim pn (x) x ∈ E.
n→∞ n→∞
Then
0 ≤ s1 (x) + p1 (x) ≤ s2 (x) + p2 (x) ≤ . . . ≤ f + g,
for all x ∈ E and
(1.6.37) (f + g) (x) = lim (sn + pn ) (x). x ∈ E.
n→∞
In the general case, since f, g ∈ L (µ) we can suppose that f (x) and g(x) are finite for all x ∈ E.
Let h = f + g, then
h+ − h− = h = f + g = f + − f − + g + − g −
or equivalently
(1.6.38) h+ + f − + g − = f + + g + + h−
Since both sides in equation (1.6.38) are the sum of three positive functions of L (µ) we conclude
Z Z Z Z
− −
+
h+ + f − + g − dµ
h dµ + f dµ + g dµ =
E E E
ZE
f + + g + + h− dµ
=
ZE Z Z
= f + dµ + g + dµ + h− dµ
E E E
Thus,
Z Z Z Z Z Z Z Z
(f + g) dµ = hdµ = h+ dµ − h− dµ = f + dµ − f − dµ + g + dµ − g − dµ
E E E E E E E E
Z Z
= f dµ + gdµ.
E E
ε
If f is bounded, i.e., |f (x)| ≤ M for some real positive M and all x ∈ E, then if µ (A) < δ = M,
since µ (A) < µ (E) we have
Z Z
φ|f | (A) = |f | dµ < M dµ = µ (A) M < ε.
A A
Otherwise, define
|f (x)| f (x) ≤ n
fn (x) =
n f (x) > n
Then {fn } is an increasing sequence of bounded, measurable functions with f = lim fn on E. By
n→∞
Lebesgue’s monotone convergence theorem, we can find N ∈ N so that
Z Z
ε
fN dµ > |f | dµ −
E E 2
and hence Z Z Z
ε
> |f | dµ − fN dµ = (|f | − fN ) dµ.
2 E E E
ε
If 0 < δ < 2N and µ (A) < δ, since 0 < fN (x) < N for all x ∈ E, then
Z Z Z Z
|f | dµ = (|f | − fn ) + fn dµ = (|f | − fn ) + fn dµ
A
ZA Z A A
≤ (|f | − fn ) + N dµ
E A
ε ε ε
< + N µ (A) < + = ε,
2 2 2
and the proof is complete.
Theorem 19. Suppose E ∈ M. Let {fn } be a sequence of non-negative measurable functions
on E. Let f be defined by
X∞
(1.6.39) f (x) = fk (x),
k=1
then
Z ∞ Z
X
(1.6.40) f dµ = fk dµ
E k=1 E
Pn
Note that the sequence {gn = k=1 fk } of the partial sums of (1.6.39) form a monotonically
increasing sequence such that
f = lim gn ,
n→∞
and by Theorem 17
Z n Z
X
(1.6.41) gn dµ = fk dµ
E k=1 E
Consider the sequence {gn = inf k≥n fk } , by Theorem 8 the functions gn are measurable func-
tions on E. On the other hand: or comma
Thus,
R1 R1
f (x)dm = 0 χ[ 1 ,1] (x)dm = m 12 , 1 = 21
Z 1 (
0 2k
n = 2k
fn (x)dm = R1 R1 2
,
(x)dm = 0 χ[0, 1 ) (x)dm = m 0, 12 = 21
0 f
0 2k+1
n = 2k + 1
2
This exercise show that we can obtain strict inequality in equation (1.6.43 ) in Fatou’s Lemma.
Theorem 21 (Lebesgue dominated convergence theorem). Suppose E ∈ M. Let {fn } be a
sequence of measurable functions on E and f a measurable function on E such that
(1.6.49) f = lim fn in E
n→∞
So
Z Z
(1.6.54) f dµ ≤ lim inf fn dµ
n→∞
ZE E
Z
(1.6.55) − f dµ ≤ −lim inf fn dµ
E n→∞ E
1.6. INTEGRATION xli
Thus,
Z Z Z Z
(1.6.57) lim sup fn dµ ≤ f dµ ≤ lim inf fn dµ ≤ lim sup fn dµ
n→∞ E E n→∞ E n→∞ E
So all inequalities in equation (1.6.57) are equalities which implies that the limit in (1.6.51)
exist and the equality in (1.6.51) is holds.
Corollary 4. If µ (E) < ∞, and {fn } is uniformly bounded on E, and f (x) = lim fn (x)
n→∞
on E, then
Z Z
(1.6.58) f dµ = lim fn dµ.
E n→∞ E
1 1
|fn (x) − 0| ≤ < < ε.
n N
So {fn } converges to 0 uniformly on R1 .
On the other hand, we have that fn = n1 χ[−n,n] . So for all n ∈ N
Z ∞ Z ∞
1 1 1
fn (x) dm = χ[−n,n] (x) dm = m ([−n, n]) = 2n = 2.
−∞ −∞ n n n
and the proof is complete.
xlii 1. LEBESGUE MEASURE AND INTEGRATION
Example 4. In this example we have a sequence {fk } of continuous function that converge to
a continuous function f , but
Z b Z b
lim fk dx 6= f dx
k→∞ a a
In this case consider 1
k2 x 0≤x≤
k
fk (x) = 2 1 2
k2 −x <x≤
k k k
2
0 k <x≤1
0, not o This sequence punctually converge to the function f (x) = 0 for all x ∈ [0, 1], below we can see the
graph of fk , this graph shows that the integral of fk is 1 equal to the area of the triangle in red on
the graph, so
Z b Z b
lim fk dx = 1 6= 0 = f dx.
k→∞ a a
The last example shows that the condition |fk (x)| ≤ g(x), in the dominated converge theorem
is necessary to obtain the conclusion.
Exercise 8. [Exercise 11.12] Suppose
(a) |f (x, y)| ≤ 1, for 0 ≤ x ≤ 1, 0 ≤ y ≤ 1
(b) for fixed x, f (x, y) is a continuous function of y,
(c) for fixed y, f (x, y) is a continuous function of x.
Put Z 1
g(x) = f (x, y)dy (0 ≤ x ≤ 1)
0
Is g continuous?
Solution 8. The answer is yes, g is continuous. We only need to show that if {xn } is a
sequence so that xn → x, then f (xn ) → f (x).
1.7. COMPARISON WITH THE RIEMANN INTEGRAL xliii
Indeed, first note that the continuity condition (b) implies that the right side on the definition
of g(x) is well defined (really we only need that for fixed x, fx (y) = f (x, y) is integrable on [0, 1]).
On the other hand, by (c), f (xn , y) → f (x, y) for each y ∈ [0, 1], in particular for almost every
y.
Since the set [0, 1] has finite measure, then the function h = χ[0,1] ∈ L (m)
Let fn (y) = f (xn , y) and fx (y) = f (x, y), then fn (y) → fx (y) and by (a)
|fn (y)| = |f (xn , y)| ≤ 1 = h (y)
for all n and y ∈ [0, 1], using the dominated convergence theorem we have
Z 1 Z 1 Z 1
lim g(xn ) = lim f (xn , y)dy = lim f (xn , y)dy = f (x, y)dy = g(x).
n→∞ n→∞ 0 0 n→∞ 0
This concludes the proof.