Download as pdf or txt
Download as pdf or txt
You are on page 1of 49

Partial Solutions to Folland’s Real Analysis: Part I

(Assigned Problems from MAT1000: Real Analysis I)

Jonathan Mostovoy - 1002142665


University of Toronto
January 20, 2018

Contents
1 Chapter 1 3
1.1 Folland 1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Folland 1.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Folland 1.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Boxes vs cylinder sets w.r.t. ‡-algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Folland 1.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.6 Folland 1.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.7 Folland 1.9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.8 Folland 1.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.9 Folland 1.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.10 Folland 1.17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.11 Folland 1.18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.12 Folland 1.26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.13 Folland 1.28 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.14 Folland 1.30 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.15 Folland 1.31 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.16 Folland 1.33 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2 Chapter 2 15
2.1 Folland 2.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Folland 2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Folland 2.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4 Folland 2.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Folland 2.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.6 Folland 2.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.7 Folland 2.9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.8 Folland 2.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.9 Folland 2.12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.10 Folland 2.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.11 Folland 2.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.12 Folland 2.16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.13 Folland 2.17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.14 Differentiable functions are Borel Measurable . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.15 Folland 2.20 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

1
2.16 Folland 2.21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.17 Folland 2.24 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.18 Folland 2.34 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.18.1 Folland 2.33 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.19 Folland 2.39 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.20 Folland 2.42 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.21 Folland 2.44: Lusin’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.22 Folland 2.46 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.23 Folland 2.48 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.24 Folland 2.49 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3 Chapter 3 30
3.1 Folland 3.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2 Folland 3.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.3 Folland 3.12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.4 Folland 3.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.5 Folland 3.17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.6 Folland 3.20 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.7 Folland 3.21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.8 Folland 3.24 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.9 Folland 3.25 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.10 Folland 3.26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4 Chapter 5 39
4.1 Folland 5.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2 Folland 5.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.3 Folland 5.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.4 Folland 5.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.5 Folland 5.9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

5 Chapter 6 44
5.1 Folland 6.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.2 Folland 6.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3 Folland 6.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.4 Folland 6.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.5 Folland 6.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.6 Folland 6.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

2
1 Chapter 1
1.1 Folland 1.2
Prove the following Proposition:
Proposition. 1.1:
BR is generated by each of the following:
(a) the open intervals: E1 = {(a, b) | a < b},
(b) the closed intervals: E2 = {[a, b] | a < b},

(c) the half-open intervals: E3 = {(a, b] | a < b} or E4 = {[a, b) | a < b},


(d) the open rays: E5 = {(a, Œ) | a œ R} or E6 = {(≠Œ, a) | a œ R},
(e) the closed rays: E7 = {[a, Œ) | a œ R} or E8 = {(≠Œ, a] | a œ R},

Proof. Most of the proof is already completed by Folland. What was shown is that M(Ej ) µ BR ’j =
1, . . . , 8. To finish the proof and show BR = M(Ej ) ’j, we can simply show that BR µ M(Ej ) ’j.
By invoking Lemma 1.1, if the family of open sets lie in M(Ej ), then it must be that BR µ M(Ej ).
Furthermore, it is actually sufficient to only show that all the open intervals lie in M(Ej ) since every
open set in R is a countable union of open intervals. Thus, we complete our proof by directly showing
the following:
1. (a, b) œ E1 ∆ (a, b) œ M(E2 ).

2. (a, b) = fiŒ
1 [a + n
≠1
, b ≠ n≠ 1] œ M(E2 )
3. (a, b) = fiŒ
1 (a, b ≠ n 1] œ M(E3 )

4. (a, b) = fiŒ
1 [a + n
≠1
, b) œ M(E4 )
! "c
5. (a, b) = (a, Œ) fl (≠Œ, b) = (a, Œ) fl [b, Œ)c = (a, Œ) fl flŒ 1 (b ≠ n ≠1
, Œ) œ M(E5 )
! ≠1 c
"
6. (a, b) = (a, Œ) fl (≠Œ, b) = (≠Œ, a]c fl (≠Œ, b) = flŒ 1 (≠Œ, a + n ) fl (≠Œ, b) œ M(E6 )
! Π"
7. (a, b) = (a, Œ) fl (≠Œ, b) = fi1 [a + n , Œ) fl [b, Œ)c œ M(E7 )
≠1

! "
8. (a, b) = (a, Œ) fl (≠Œ, b) = (≠Œ, a]c fl fiŒ 1 (≠Œ, b ≠ n ≠1
] œ M(E8 )

1.2 Folland 1.4


Prove the following proposition:
Proposition. 1.2:

An algebra A is a ‡-algebra ≈∆ A is closed under countable increasing unions (i.e., if {Ej }Œ


1 µA
and E1 µ E2 µ · · · , then fiŒ
1 Ej œ A).

3
Proof. The forward direction (‡-algebra ∆ closed under countable increasing unions) is by the definition
of ‡-algebra (closed under countable unions ). The backward direction (closed under countable increasing
unions ∆ closed under countable increasing unions ∆ ‡-algebra) is slightly more involved:
If {Fi }Œ
1 œ A, then let us define Ej := fi1 Fi . Since countable unions of countable unions is countable,
j

and since {Ej }Œ 1 has the property of E1 µ E2 µ · · · , then we know that fi1 Ej œ A. However, since
Œ

it is also the case that fiŒ


1 Fi = fi1 Ej , we can conclude that fi1 Fi œ A as well, and thus proving the
Œ Œ

backward direction.

1.3 Folland 1.5


Prove the following Proposition:
Proposition. 1.3:

If M(E) is the ‡-algebra generated by E, then M(E) is the union of the ‡-algebras generated by
F– as F– ranges over all countable subsets of E.

Proof. We use the notation F– to denote a countable subset of E, and we let F := {F– | – œ A} denote
the (likely uncountable) set of all countable subsets of E. Let us also define M̂ := fi–œA M(F– ). We
proceed now by first showing that M̂ is indeed a ‡-algebra by showing that M̂ is closed under countable
unions and compliments: 奇数
Suppose {Ei }Œ 1 œ M̂. Since M̂ is simply the union of a many ‡-algebras, we know immediately that ’Ei
÷ at least one Fi s.t. Ei œ M(Fi ). Since a countable union of countable elements is countable, if we define
H := fiŒ 1 Fi where Ei œ M(Fi ), we know that H is also countable subset of E. We can now look at the
properties of the following ‡-algebra: M(H).
(1) Since Fi µ H µ M(H) ∆ M(Fi ) µ M(H) (by Lemma 1.1), and since Ei œ M(Fi ), we can say that
{Ei }Œ
1 œ M(H).
(2) Since H is a countable subset of E, we know that ÷— s.t. H = F— , and hence M(H) µ M̂.
Therefore, since M(H) is by construction a ‡-algebra and from (1) ({Ei }Œ
1 œ M(H)) it ∆ fi1 Ei œ M(H),
Œ

and by (2) (M(H) µ M̂) ∆ fiŒ 1 Ei œ M̂.

To now show M̂ is closed under compliments, suppose E œ M̂. By the same argument already used,
there must exist a countable subset F– µ E s.t. E œ M(F– ), and obviously since M(F– ) is a ‡-algebra,
E c œ M(F– ). Therefore, since M(F– ) µ M̂ ∆ E c œ M̂. We have thus shown that M̂ is is closed under
countable unions and compliments, and hence a ‡-algebra.
万⼀
To neatly finish up our proof, let us first note that ’– œ A, F– µ E ∆ M(F– ) µ M(E), and thus we
不可 怎么办 数
can also say M̂ µ M(E). To show the opposite relation, let Á œ E, then Á is trivially countable, so ÷—
s.t. Á = F— ∆ Á œ M̂. Now since this is true ’Á œ E, we can say that E µ M̂, which therefore (again by
Lemma 1.1) ∆ M(E) µ M. By showing both opposite relations, we can thus conclude that M(E) = M̂.

4
1.4 Boxes vs cylinder sets w.r.t. ‡-algebras
Exercise. 1.1:
Let A be an index set, {X– }–œA a family of non-empty sets and for each – œ A, M– be a ‡-algebra
on X– . Consider the product space: Ÿ
X= X–
–œA

Let M be the ‡-algebra generated by the cylinder


r sets C := {fi ≠1
– (E– ) | E– œ M– , – œ A}, and
M be the one generated by boxes
ú
:= { –œA E– | E– œ M– }. Show that M µ Mú , but in
general M ”= Mú
Hint 1: Proposition 1.3 implies that if A is countable then M = Mú ; we should thus take A to be
not countable.)
Hint 2: You might find useful to first prove the following intermediate result. For any AÕ µ A, let
MAÕ = M({fi ≠1 (E– ) | E– œ M– , – œ AÕ }); let now

M̃ = M AÕ
AÕ µA countable

Then show that M = M̃. (Hint2 : show that M̃ is a ‡-algebra which contains the cylinders...) The
above can be loosely stated as “any set in M is determined by countably many coordinates”

**Please note the notation used for the box and cylinder sets above.
r
Answer: To show M µ Mú , note that fi–≠1 (E– ) = —œA E— , where E— = X— ’ — ”= –. In this form, it
is clear that C µ µ Mú ∆ M µ Mú (by Lemma 1.1).
Next, let us prove that M = M̃:
Proof. Suppose {Fi }Œ1 œ M̃. Then since M̃ is a union of ‡-algebras, it must be that Fi œ MAÕ for at least
one AÕ . Taking AÕÕ to be the union of one of the AÕ s which satisfies Fi œ MAÕ for each i. Thus, AÕÕ will
naturally also be a countable set. Since AÕÕ is a countable set, MAÕÕ µ M̃ ∆ fiŒ1 Fi œ M̃ by Lemma 1.1.
Next, suppose F œ M̃, then ÷AÕ s.t. F œ MAÕ , which implies F c œ MAÕ , and since MAÕ µ M̃, ∆ F c œ M̃,
and hence M̃ is indeed a ‡-algebra.
Next, since AÕ µ A, ∆ MAÕ µ M(= MA ) ’AÕ , and thus since MAÕ µ M ’AÕ ∆ M̃ µ M. To show the
opposite inclusion, we know that ’ – œ A ÷ a countable subset AÕ µ A s.t. – œ AÕ , namely {–}. In this
form, it is perfectly clear that fi ≠1 (E– ) µ M̃, since fi ≠1 (E– ) œ MAÕ ={–} ∆ M µ M̃. And thus M = M̃.

Let us now turn our attention


r to the form
r in which the generating family of sets for MAÕ takes. Each set
is in the form fi (E– ) = ( —œAÕ E— )◊( “œA\AÕ X“ ), where A is a countable set, and E— = X— ’— ”= –.
≠1 Õ

In this form, it is clear that after


r countably r many intersections, compliments and unions, ’E œ MAÕ ,
E will still be in the form of ( —œAÕ E— ) ◊ ( “œA\AÕ X“ ), where AÕ is a countable set, and E— œ M— .
r r
However, when looking at the boxes, it is clear that ÷E œ M(B) s.t. E = ( —œB E— ) ◊ ( “œA\B X“ ),
where B is an uncountable set, and E— œ M— .

1.5 Folland 1.7


Prove the following Proposition:

5
Proposition. 1.4:
qn
If µ1 , . . . , µn are measures on (X, M) and a1 , . . . , an œ [0, Œ), then 1 aj µj is a measure on
(X, M).

Proof.qSince µi i œ {1, . . . , n} are measures, we know that µi ( ) = 0 ’i = 1, . . . , n, and therefore


µ := 1 aj µj ( ) = 0. Next, suppose {Ej }Œ 1 œ M and {Ej }1 disjoint, then:
n Œ

3h
n 4 ÿn
A 3h
n 4B ÿn
A Π3 4 B n
A Œ 3 4 B Œ
ÿ ÿ ÿ ÿ
µ Ei = aj · µj Ei = aj · µj Ei = aj · µj Ei = µ(Ei )
i=1 j=1 i=1 j=1 i=1 j=1 i=1 i=1

1.6 Folland 1.8


Prove the following Proposition:
Proposition. 1.5:

If (X, M, µ) is a measure space and {Ej }Œ 1 µ M, then µ(lim inf Ej ) Æ lim inf µ(Ej ). Also,
µ(lim sup Ej ) Ø lim sup µ(Ej ) provided that µ(fiŒ
1 Ej ) < Œ.

Proof. We first recall the definitions of lim inf and lim sup for a sequence of sets as:
Œ
AŒ B Œ
AΠB
€ ‹ ‹ €
lim inf (Fn ) := Fn , and lim sup(Fn ) := Fn
næŒ næŒ
k=1 n=k k=1 n=k

We now quickly prove the following Lemma:


Lemma. 1.1: A Corollary of Monotonicity and Subadditivity - (Again!)

If {Bi }Œ
1 µ M, then:

(a) µ(flŒ
1 Bi ) Æ µ(B1 ) (or µ(Bk ) by switching B1 for Bk ).

(b) µ(fiŒ
1 Bi ) Ø µ(B1 ) (or µ(Bk ) by switching B1 for Bk ).

Note: I sorta forget these were either covered or corollaries from Thm 1.8 in Folland, hence why I included
it here - oh well (but I did give a slightly more concise proof for (b) :) ) 我有点忘了这些都是Folland的Thm 1.8所涵盖的内容或推
论,因此我把它放在这⾥--哦,好吧(但我确实给了(b)⼀
个稍微简洁的证明 :) )
Proof.
(a) Since (flŒ
1 Bi ) Û (B1 \ fl1 Bi ) = B1 ∆ µ(B1 ) = µ(fl1 Bi ) + µ(B1 \ fl1 Bi ) ∆ µ(fl1 Bi ) Æ µ(B1 ).
Œ Œ Œ Œ

(b) Since (B1 ) Û (fiŒ


2 Bi \B1 ) = fi1 Bi ∆ µ(fi1 Bi ) = µ(B1 ) + µ(fi2 Bi \B1 ) ∆ µ(fi1 Bi ) Ø µ(B1 ).
Œ Œ Œ Œ

6
We now have all the necessary tools to prove the proposition as follows:
Q Q RR Q R Q R
Œ
€ ‹ Œ ‹Œ Œ

! " ú ú
µ lim inf Ej = µ a a Ej b b = lim µ a Ej = lim inf µ
b a Ej Æ lim inf µ (Ej )
b
kæŒ kæŒ kæŒ
k=1 j=k j=k j=k

ú ú
Where = is by µ’s “Continuity from below” since flŒ
j=k Ej µ flj=k+1 Ej ’k œ N, and Æ is by Lem 2.1 (a).
Œ

Q Q RR Q R Q R

迎 迎
Œ
‹ € Œ Œ
€ Œ

! " ı
µ lim sup Ej = µ = lim µ Ej = lim inf µ Ej b Ø lim inf µ (Ej )
ı
a a Ej b b a b a
kæŒ kæŒ kæŒ
k=1 j=k j=k j=k

ı
Where = is by µ’s “Continuity from above” since fiŒ
j=k+1 Ej µ fij=k Ej ’k œ N, and Æ is by Lem 2.1 (b).
ı Œ

1.7 Folland 1.9


Prove the following Proposition:
Proposition. 1.6:

If (X, M, µ) is a measure space and E, F œ M, then µ(E) + µ(F ) = µ(E fi F ) + µ(E fl F ).

Proof. Firstly, let us make the following observations:


(E\F ) Û F = (E fi F ), and (E fl F ) Û (E\F ) = E
Therefore, since µ is countably additive and therefore finitely additive, we can now see that:
! "
µ(E) + µ(F ) = µ (E fl F ) Û (E\F ) + µ(F )
= µ(E fl F ) + µ(E\F ) + µ(F )
! "
= µ(E fl F ) + µ (E\F ) Û F
= µ(E fl F ) + µ(E fi F )

1.8 Folland 1.10


Prove the following Proposition:
Proposition. 1.7:

Given a measure space, (X, M, µ) and E œ M, define µE (A) = µ(A fl E) for A œ M. Then µE is
a measure.

Proof. We first confirm that µE ( ) = 0 since µE ( ) = µ( fl E) = µ( ) = 0. Next, let {Fi }Œ


1 µ M and
1 disjoint. Then:
{Fi }Œ
AŒ B A A Œ BB AA Œ BB Œ Œ
€ € € ú
ÿ ÿ
µE Fi = µ E fl Fi =µ E fl Fi = µ (E fl Fi ) = µE (Fi )
i=1 i=1 i=1 i=1 i=1
ú
Where = since if {Fi }Œ
1 is a disjoint family of sets, then {Fi fl E}1 will be as well. Thus, we have shown
Œ

µE is indeed a measure.

7
1.9 Folland 1.13
Prove the following Proposition:
Proposition. 1.8:
Every ‡-finite measure is semi-finite.

Proof. Let µ be a ‡-finite measure on the measurable space (X, M). Firstly, if µ(X) < Œ, µ will trivially
be semi-finite. Therefore, suppose µ is ‡-finite, but not finite. Now, let us arbitrarily pick E œ M s.t.
µ(E) = Œ (we know at least one such element exists, namely X, since otherwise µ would be finite). From
the definition of µ being ‡-finite, we know that ÷ {Fi }Œ
1 µ M s.t. X = fi1 Fi and µ(Fi ) < Œ ’i œ N.
Œ

One can easily see the following:


AŒ B Œ
€ ÿ
µ(E) = µ(E fl X) = µ (E fl Fi ) Æ µ(E fl Fi )
i=1 i=1

And since µ(E) = Œ


Œ
ÿ Œ
ÿ
∆ŒÆ µ(E fl Fi ) ∆ µ(E fl Fi ) = Œ
i=1 i=1

Furthermore, since E ”= (since otherwise µ(E) = 0 < Œ) and µ(E) = µ ( i=1 (E fl Xi )), we know there
must exist at least one k œ N s.t. µ(E fl Fk ) > 0. On the other-hand, since µ(Fk ) < Œ by construction,
so too will µ(E fl Fk ) < Œ. Therefore, since trivially E fl Fk µ E, we have shown that for an arbitrary
E œ M s.t. µ(E) = Œ, ÷k œ N s.t. Fk fl E µ E and µ(Fk fl E) < Œ; I.e., all ‡-finite measures are
semi-finite.

1.10 Folland 1.17


Prove the following Proposition:
Proposition. 1.9:

If µú is an outer measure
q on X and {Aj }Œ 1 is a sequence of disjoint µ -measurable sets, then
ú

µú (E fl (fiŒ
1 Aj )) = 1 µ (E fl Aj ) for any E µ X.
Œ ú

Proof. Firstly, since µú is an outer measure, we know that:


Œ
ÿ
µ (E fl
ú
(fiŒ
1 Aj )) =µ ú
((fiŒ
1 E fl Aj )) Æ µú (E fl Aj )
j=1

Now, let us define Bn := fin1 Ej . Now, since Aj is µú -measurable ’j œ N, we know that ’n > 1:

µú (E fl Bn ) = µú ((E fl Bn ) fl An ) + µú ((E fl Bn ) fl Acn ) = µú (E fl An ) + µú (E fl Bn≠1 )

Therefore, iteratively using the above formula (by induction) for Bn , . . . , B2 , and countable additivity
being trivial for n = 1, we have shown that:
Q R
€n n
ÿ
µúa
Efl Aj b = µú (E fl Aj ) , ’n œ N
j=1 j=1

8
Now, by monotonicty, we can easily see that:
Q R Q R
Œ
€ n
€ n
ÿ
µú aE fl Aj b Ø µú aE fl Aj b = µú (E fl Aj ) , ’n œ N
j=1 j=1 j=1

1 tŒ 2 qŒ
And hence µú E fl j=1 Aj Ø j=1 µ (E fl Aj ).
ú
And thus since we shown both Ø and Æ, we can
qŒ ú
conclude that µú (E fl (fiŒ
1 Aj )) = 1 µ (E fl Aj ).

1.11 Folland 1.18


Prove the following Proposition:
Proposition. 1.10:

Let A µ P(X) be an algebra, A‡ the collection of countable unions of sets in A, and A‡” the
collection of countable intersections of sets in A‡ . Let µ0 be a premeasure on A and µú the induced
outer measure.

a) For any E µ X and ‘ > 0, there exists B œ A‡ with E µ B and µú (B) Æ µú (E) + ‘
b) If µú (E) < Œ, then E is µú -measurable ≈∆ there exists C œ A‡” with E µ C and
µú (C\E) = 0.

c) If µ0 is ‡-finite, the restriction µú (E) < Œ in (b) is superfluous.

Proof.
a) Let us recall the definition of µú (E) as:
IŒ - Œ
J
ÿ - €
µ (E) := inf
ú
µ0 (Ai ) - {Ai }Œ
- 1 µ A, E µ Ai
i=1 i=1

Therefore, by the definition of inf, ’‘ > 0 ÷{Bi }Œ
1 s.t. E µ fiŒ
1 B i and 1 µ0 (Bi ) Æ µ (E) + ‘.
ú

Therefore, if we define B := {Bi }Œ


1 (same seq. as before), we note that B œ A‡ , and also that:

Œ
ÿ
ú
µú (B) Æ µ0 (Ai ) Æ µú (E) + ‘
i=1

ú
Where Æ because µ0 (Bi ) = µú (Bi ), and B is µú -measruable.
b) Let us begin with the forward direction (µú (E) < Œ, and E is µú -measurable). From part (a),
we know ÷ {Ci } µ A‡ s.t. E µ Ck and µú (Ck ) Æ µú (E) + k1 ’k œ N. Let us now define
C := flŒ 1 Ci , to which we notice that C œ A‡” and E µ C since E µ Ck ’k œ N, and hence
µú (E) Æ µú (C). Furthermore, we note that since Ck is µú -measurable, so too will Ckc , and hence
1 Ci = (fl1 Ci ) = C is µ -measurable, and hence C is µ -measurable. Now, the following
fiŒ c Œ c c ú ú

observation becomes apparent:


AΠB A n B
‹ ‹
µú (C) = µú Ci = lim µú Ci Æ lim µú (Cn ) = µú (E)
næŒ næŒ
i=1 i=1

9
Moreover, using the fact that E µ C the above now actually implies that µú (E) = µú (C). We also
recall that since E c is µú -measurable, and since we already showed that C was µú -measurable, we
can now also say that C fl E c = B\E is µú -measurable, and also note that hence:

µú (C\E) = µú (C) ≠ µú (C fl E) = µú (C) ≠ µú (E) = 0

For the backward direction (there exists C œ A‡” with E µ C and µú (C\E) = 0), first note that
since E µ C), C = (B\E)fiE. Next, since µú is the Carathèodory extension, C\E is µú -measurable.
Therefore, we can easily conclude that E = B\(B\E) is also µú -measurable.

c) Firstly, since µ0 is ‡-finite, we know that ÷ a disjoint set {Xi }Œ


1 µ A s.t. X = ÛXi and µ0 (Xk ) <
Œ ’k œ N. Next, since E µ X is measurable, so too will Ek := E fl Xk ’k œ N, and by
above and since {E fl Ei }Œ 1 is disjoint, we know that E = Û1 (E fl Xi ) = Û1 Ei , and naturally
Œ Œ

µ0 (E fl Xk ) < Œ ’k œ N. Since we are able to write E in this construction, E is µú -measurable.


We can now figure out the following line of reasoning:

Eµú -measurable ≈∆ Ei µú -measurable


≈∆ ÷Ci œ A s.t. Ei µ Ci , µú (Ci \Ei ) = 0
Q AΠBR Q AΠBR
€Œ Œ
€ ‹ € Œ Œ Œ
‹ ‹ €
≈∆ E µ C = Ci = a Aijk b µ a Aijk b œ A‡”
i=1 i=1 j=1 k=1 i=1 j=1 k=1

qŒ qŒ
Where µú (C\E) = µú (fiŒ 1 Ci \Ei ) Æ 1 µú (Ci \Ei ) = 1 0 = 0. And hence µ ú (E) < Œ did not
matter if µ0 is ‡-finite.

1.12 Folland 1.26


Prove the following Proposition (by using Folland, Theorem 1.19):
Proposition. 1.11:

If E œ Mµ and µ(E) < Œ, then ’‘ > 0 ÷ a set A that is a finite union of open intervals such that
µ(E—A) < Œ.

Proof. We recall that by Theorem 1.18, ÷Uopen s.t. E µ U and µ(U) Æ µ(E) + 12 ‘. Furthermore, by the
inequality just stated, we know that µ(U), µ(E) < Œ, and hence:
1
µ(U\E) = µ(U) ≠ µ(E) < ‘
2
Now, by recalling that all open sets in R can be written as ی
1 Ui , we know that ÷ {Ui }1 s.t. Û1 Ui = U.
Œ Œ

We now prove that actually:


! N " 1
÷N œ N s.t. µ(U) = µ (ÛŒ
1 U = U) < µ Û1 Ui + ‘
2
i

To see this, since {Ui }Œ


1 is disjoint:

Œ
ÿ
µ (Ui ) = µ(U) < µ(E) < Œ
i=1

10

Therefore,
qŒ the series 1 µ (Ui ) must converge, and hence, by the definition of convergent series’, ÷N œ N
1
s.t. N +1 µ (µ(Ui )) < 2 ‘, and thus the inequality we sought to prove has now been shown.

Carrying on, let us define Ũ := {Ui }N


1 . Since Ũ µ U ∆ µ(Ũ) Æ µ(U) < Œ and also ∆ Ũ\E µ U\E,
hence:
1
µ(Ũ\E) Æ µ(U\E) < ‘
2
Now, also since µ(Ũ) < Œ, and since Ũ µ U ∆ E\Ũ µ U \Ũ , we can see that:

1
Œ
ÿ
µ(E\Ũ) Æ µ(U\Ũ) = µ(U) ≠ µ(Ũ) = µ(Ui ) < ‘
2
i=N +1

Therefore, by combining the last two main inequalities, we have found a set A = Ũ which is a finite union
of open intervals such that:
1 1
µ(E—Ũ) = µ(E\Ũ) + µ(Ũ\E) < ‘+ ‘=‘
2 2

1.13 Folland 1.28


Prove the following Proposition:
Proposition. 1.12:
Let F be increasing and right continuous, and let µF be the associated measure. Then:
a) µF ({a}) = F (a) ≠ F (a≠)
b) µF ([a, b)) = F (b≠) ≠ F (a≠)
c) µF ([a, b]) = F (b) ≠ F (a≠)

d) µF ((a, b)) = F (b≠) ≠ F (a)

Proof.
a) We first note that we may construct {a} from a countable intersection of h-intervals as follows:
Œ

{a} = (a ≠ 1/n, a]
n=1

Furthermore, since (a ≠ 1/n, a] ∏ (a ≠ 1/(n + 1), a] ’n œ N, we may invoke continuity from above
in that:
ú
µF ({a}) = lim µF ((a ≠ 1/n, a]) = lim (F (a) ≠ F (a ≠ 1/n)) = F (a) ≠ F (a≠)
næŒ næŒ

ú
Where = can be rigorously shown by noting that since F is an increasing function:

lim F (a ≠ 1/n) = sup{F (x) | x < a} = F (a≠)


næŒ

11
b) We first note that we may construct [a, b) from a union of countable intersections and unions of
h-intervals as follows:
AΠB A ΠB
‹ €
[a, b) = [a, (a + b)/2] fi (a, b) = (a ≠ 1/n, (a + b)/2] fi (a, b ≠ 1/m]
n=1 m=1

Like in part a):


A Œ
B

µF ([a, (a + b)/2]) = lim µF (a ≠ 1/n, (a + b)/2]
næŒ
n=1
= lim (F ((a + b)/2) ≠ F (a ≠ 1/n))
næŒ
ú
= F ((a + b)/2) ≠ F (a≠)
ú
Where = is reasoned exactly as in a). Furthermore, since (a, b ≠ 1/m) µ (a, b ≠ 1/(m + 1)) ’m œ N,
we may invoke continuity from below in that:
AΠB

µF ((a, b)) = lim µF (a, b ≠ 1/m]
næŒ
n=1
= lim (F (b ≠ 1/m) ≠ F (a))
næŒ

= F (b≠) ≠ F (a)
ı

Where = can be rigorously shown by noting that since F is an increasing function:


ı

lim F (b ≠ 1/m) = sup{F (x) | x < b} = F (b≠)


næŒ

Therefore, since all the sets we’ve been dealing with so far have been bounded, we can see now that:
! " ! " ! " ! "
µF [a, b) = µF [a, (a + b)/2] + µF (a, b) ≠ µF (a, b) fl [a, (a + b)/2]
! " ! " ! "
= µF [a, (a + b)/2] + µF (a, b) ≠ µF (a, (a + b)/2]
Ë ! " ! "È Ë ! " ! "È Ë ! " ! "È
= F (a + b)/2 ≠ F a ≠ + F b ≠ ≠ F a ≠ F (a + b)/2 ≠ F a
Ë È Ë ! " ! "È Ë ! " ! "È
= F (b≠) ≠ F (a≠) + F (a + b)/2 ≠ F (a + b)/2 + F a ≠ F a
= F (b≠) ≠ F (a≠)

c) We first note that we may construct [a, b] from countable intersection of h-intervals as follows:
Œ

[a, b] = (a ≠ 1/n, b]
n=1

Thus, by making the change of variables of (a + b)/2 æ b, from the first half of b), we have already
shown that µF ([a, b]) = F (b) ≠ F (a≠).
d) We first note that we may construct (a, b) from a countable union of h-intervals as follows:
Œ

(a, b ≠ 1/n]
n=1

Thus, from the second half of b), we have already shown that µF ((a, b)) = F (b≠) ≠ F (a).

12
1.14 Folland 1.30
Prove the following Proposition:
Proposition. 1.13:

If E œ L and m(E) > 0, for any – < 1 ÷ an open interval Iˆ such that m(E fl I) > –m(I).

Proof. If – Æ 0, since m(E) > 0 ∆ ÷F µ E s.t. m(F ) > 0, and F = (a, b], a < b. If we thus take:
3 4
1 3
Iˆ = (a + b), (a + b)
4 4

We have m(E fl I) = m(I) > 0 Ø –m(I).


Now suppose 0 < – < 1. Since m is semi-finite, if m(Ê) = Œ, we can simply take E µ Ê s.t. 0 <
m(E) < Œ, and hence we actually restrict our problem to that of all E’s s.t. E œ L and 0 < m(E) < Œ.
Let us also quickly note/recall that:
! " ! " ! " ! "
m({b}) = 0 ∆ m((a, b]) = m (a, b) Û {b} = m (a, b) + m {b} = m (a, b)

Now, for the sake of contradiction, assume ’I = (a, b), a < b, we have: m(E fl I) Æ –m(I). Let us choose
‘1 > 0 so that ‘1 < 1≠– – (and hence –(1 + ‘1 ) < 1). Moreover,
q !from (Folland)
" Theorem 1.18, we know
that ’‘2 > 0 ÷I = ÛŒ 1 (ai , bi ) s.t. E µ I and m(I) = 1 m (ai , bi ) < m(E) + ‘2 . Next, from our
Œ

discussion on (a, b) v.s. (a, b], we can actually write I = Û1 (ai , bi ], where I still satisfies everything that
Œ

it did beforehand. Now, if we let ‘2 = m(E)‘1 , (which we can certainly do since m(E) < Œ), we see that:
3 4
1≠– 1
Œ
ÿ
m(I) = (ai , bi ] < m(E) + m(E)‘1 = m(E)(1 + ‘1 ) < m(E) 1 + = m(E)
i=1
– –
∆ –m(I) < m(E)

Therefore, by combining the above inequality with our assumption in that m(E fl Ik ) Æ –m(Ik ) ’k œ N,
and that E µ I, we see that:
Œ
ÿ Œ
ÿ
m(E) = m(E fl I) = m(E fl Ii ) Æ –m(Ii ) = –m(I) < m(E)
i=1 i=1

Which is obviously a contradiction on the requirement of m(E) > 0, hence the converse must be true:
I.e. our Proposition is true.

1.15 Folland 1.31


Prove the following Proposition:
Proposition. 1.14:

If E œ L, and m(E) > 0, the set {E ≠ E} := {x ≠ y | x, y œ E} contains an interval centered at


0. (If I is as in (Folland) Exercise 1.30, with – > 34 , then {E ≠ E} contains (≠ 12 m(I), 12 m(I)).)

Proof. From (Folland) 1.30, we know that ÷ I s.t. 34 m(I) < m(E fl I). Let us now define F := E fl I µ E,
and naturally we will have {F ≠ F } µ {E ≠ E}, hence if ÷ an interval centered at 0 in {F ≠ F }, so too
will that interval be in {E ≠ E}.

13
We now claim that F fl {F + x0 } ”= ∆ x0 œ {F ≠ F }. To see this, let y œ F fl {F + x0 } ∆ y œ
F and ÷x œ F s.t. y = x + x0 ∆ x0 = y ≠ x, y, x œ F ∆ x0 œ {F ≠ F }.
Trivially 0 œ {F ≠ F } since F ”= ! . Let us now let "z0 œ R s.t. |z0 | < 12 m(I) < 34 m(I) < m(F ). If we can
show that F fl {F + z0 } ”= ∆ ≠ 12 m(I), 12 m(I) µ {E ≠ E}. Therefore, the remainder of this proof
will be dedicated to showing F fl {F + z0 } = ” where |z0 | < 12 m(I).
Firstly, we note that:
3 1
m(I\F ) = m(I) ≠ m(F ) = m(I) ≠ m(E fl I) Æ m(I) ≠ m(F ) = m(F )
4 4
Furthermore, by applying the useful fact that A fl B = ((A\C) fl B) fi ((C\A) fl B) twice, we find:
# $ # $ # $
I fl {I + z0 } = F fl {F + z0 } fi F fl ({I\F } + z0 ) fi (I\F ) fl {I + z0 }
! "
Our strategy now will be to show that m F fl {F + z0 } > 0, which therefore would imply I fl {I + z0 }
also has positive measure, and hence cannot be empty. To see this first note the following four properties:
# $ # $ # $
m(I fl {I + z0 }) Æ m F fl {F + z0 } + m F fl ({I\F } + z0 ) + m (I\F ) fl {I + z0 }
# $ # $ # 1
and: m F fl ({I\F } + z0 ) Æ m {(I\F ) + z0 } = m I\F ] Æ m(F ) from previously
4
# $ 1
and: m F fl ({I\F } + z0 ) Æ m(I\F ) Æ m(I) again
4
1 # $
and: m(I) < m(I) ≠ |z0 | = m I fl {I + z0 }
2
And hence combing all these we see that:
1 # $ # $ 1 # $
m(I) < m I fl {I + z)} Æ m F fl {F + z0 } + m(I) ∆ m F fl {F + z0 } > 0
2 2

1.16 Folland 1.33


Prove the following Proposition:
Proposition. 1.15:

There exists a Borel set A µ [0, 1] such that 0 < m(A fl I) < m(I) for every sub-interval I of [0, 1].
(Hint: Every sub-interval of [0, 1] contains Cantor-type sets of positive measure.)

Proof. The first observation we need to make is that since |Q| = ›0 ∆ |Q ◊ Q| = ›0 (›0 := “countably
infinite”). Therefore, we can actually write the set of all closed sets Ik inside [0, 1] where Ik ’s endpoints
are rational numbers as a countable list: Iˆ = {Ij }Œ1 . By the hint, we know that every sub-interval of
[0, 1] contains Cantor-type sets (which will certainly have rational endpoints). Our plan will therefore
be through induction, to explicitly describe a Borel set made up of necessary Cantor-like sets which will
satisfy the needed inequality.
Let Ak , Bk be strict subsets of Ik (which we can do because we’re assuming I ”= , and due to the density
of the rationals) s.t. Ai fl Bk = and m(Ai ), m(Bj ) > 0 ’i, j Æ N . We can therefore define:
N
h
CN := IN \ (Aj fi Bj )
j=1

14
And therefore, we can find a Cantor-type set DN and D̃N s.t. m(DN ), D(D̃N ) > 0 ’N œ N. If we let
D := fiŒ
N =1 DN , then ’ sub-intervals I µ [0, 1], ÷N s.t. IN µ I and we will have:

0 < m(DN ) Æ m(D fl IN ) Æ m(D fl I) Æ m(D fl I) + m(D̃N ) Æ m(I)

I.e., by seeing that A = D 0 < m(I fl D) < m(I).

2 Chapter 2
2.1 Folland 2.1
Prove the following Proposition:
Proposition. 2.1:

Let f : X æ R and Y = f ≠1 (R). Then f is measurable ≈∆ f ≠1 ({≠Œ}) œ M, f ≠1 ({Œ}) œ M,


and f is measurable on Y .

Proof. To be clear on notation, if X = {±Œ}, then either X = {Œ} or X = {≠Œ}, and naturally
{≠Œ, Œ} = ” X.
For the forward direction, since f is measurable and {±Œ} œ BR , it implies f ≠1 ({±Œ}) œ M. Fur-
thermore, again by f ’s is measurability and since R œ BR , it implies f ≠1 (R) œ M. Therefore, we may
conclude that if B œ BR , then f ≠1 (B) œ M and f ≠1 (B)flf ≠1 (R) = f ≠1 (B)flY œ M, I.e., f is measurable
on Y .
For the converse, if we let B œ BR , then we can see that:
! ≠1 " ! ≠1 "
f (B) = f (B) fl f (R) Û f (B) fl f (R\R)
≠1 ≠1 ≠1

And since f ≠1 (R) is measurable, naturally f ≠1 (B) fl f ≠1 (R) = f ≠1 (B fl R) is as well. Next, we note that:

f ≠1 (B) fl f ≠1 (R\R) = f ≠1 (B) fl f ≠1 ({≠Œ, Œ}) = f ≠1 (B fl {≠Œ, Œ})

Which naturally is either f ≠1 ( ) = , f ≠1 ({≠Œ}), f ≠1 ({Œ}), or f ≠1 ({≠Œ, Œ} = f ≠1 ({≠Œ}) fi


f ≠1 ({Œ}), all of which are measurable since f ≠1 ({≠Œ}) and f ≠1 ({Œ}) are by assumption measurable.
Combining these two implications of our assumptions, we can see f is measurable since:
! ≠1 " ! ≠1 "
f (B) = f (B) fl f (R) Û f (B) fl f (R\R) œ M
≠1 ≠1 ≠1

2.2 Folland 2.2


Prove the following Proposition:
Proposition. 2.2:

Suppose f, g : X æ R are measurable.


a) f g is measurable (where 0 · (±Œ) = 0).

b) Fix a œ R, and define h(x) = a if f (x) = ≠g(x) = ±Œ, and h(x) = f (x) + g(x) otherwise.
Then h is measurable.

15
Proof. We actually do this problem in reverse ordering.
b) We prove this fact by separating the problem into 2 lemmas, and one final main result:
For the first mini-lemma, we note that AŒ := {x œ X | f (x) = ≠g(x) = ±Œ} is measurable since
f and g are measurable.
For the second mini lemma, we make the observation that:
1 2 1 2
h ({Œ}) = (f + g) ({Œ}) = f ({Œ}) fl g ((≠Œ, Œ]) fi f ((≠Œ, Œ]) fl g ({Œ})
≠1 ≠1 ≠1 ≠1 ≠1 ≠1

Since h(x) = Œ ≈∆ either [f (x) = Œ and g(x) > ≠Œ] or [g(x) = Œ and f (x) > ≠Œ], or [
f (x) = g(x) = Œ]. Similarly for the {≠Œ} (sub-) case:
1 2 1 2
h ({≠Œ}) = (f +g) ({≠Œ}) = f ({≠Œ})flg ([≠Œ, Œ)) fi f ([≠Œ, Œ))flg ({≠Œ})
≠1 ≠1 ≠1 ≠1 ≠1 ≠1

Since h(x) = ≠Œ ≈∆ either [f (x) = ≠Œ and g(x) < Œ] or [g(x) = ≠Œ and f (x) < Œ], or [
f (x) = g(x) = ≠Œ]. We naturally recognize the above to certainly be measurable (again) since f
and g are measurable.
Now for the final main result. Let b œ R, then:
h≠1 ((b, Œ]) = h≠1 ((b, Œ)) fi h≠1 ({Œ)}
Since we already showed that h≠1 ({Œ}) is measurable, we now seek to show that h≠1 ((b, Œ)) is
measurable. This can be seen since:
I
(f + g)≠1 ((b, Œ)) if a Æ b
h ((b, Œ)) =
≠1
(f + g)≠1 ((b, a)) fi (h)≠1 ({a}) fi (f + g)≠1 ((a, Œ)) if a > b
I
(f + g)≠1 ((b, Œ)) if a Æ b
=
AŒ fi (f + g)≠1 ((b, Œ)) if a > b
Where we already showed that AŒ is measurable, and by f and g’s measurability, all the sets
above which make up h≠1 ((bŒ)) are measurable, and hence h≠1 ((b, Œ]) is measurable; therefore,
h is measurable.
a) Let us define Q+ := {r œ Q | r > 0} and Q≠ := {r œ Q | r < 0}, which is a subsets of Q and hence
countable. Suppose now that f, g Ø 0, if a Ø 0, then we will have:
(f g)≠1 ((a, Œ]) = {x œ X | f (x)g(x) > a}
€ 3 4
= {x œ X | f (x) > r} fl {x œ X | g(x) > a/r}
rœQ+

Furthermore, if a < 0, (since f, g Ø 0) we have:


(f g)≠1 ((a, Œ]) = {x œ X | f (x)g(x) > a} = X
Therefore, since irregardless of a, (f g)≠1 ((a, Œ]) is a countable union of measurable sets, f g is
measurable for f, g Ø 0. Our strategy henceforth will be to write f = f + ≠ fú≠ and g = g + ≠ gú≠ ,
where f + := max(0, f ), fú≠ := ≠ min(0, f ), and similarly for g. Therefore, we naturally have:
! " ! ! ""
f g = (f + ≠ fú≠ )(g + ≠ gú≠ ) = f + g + + fú≠ gú≠ + ≠ f + gú≠ + fú≠ g +
Now, by our previous work, since f + , g + , fú≠ , gú≠ Ø 0, it follows that the first half of the above ex-
pression is measurable (since by part b, we showed that the addition of two measurable functions as
defined in this question is measurable). And also recalling that f measurable ≈∆ ≠f measurable,
we can therefore conclude that f g is indeed measurable.

16
2.3 Folland 2.3
Prove the following Proposition:
Proposition. 2.3:

If {fn } is a sequnce of measurable functions on X, then {x | lim fn (x) exists} is a measurable set.

Proof. We first recall that by (Folland) Proposition 2.7, when {fn } is defined as in the question, g3 (x) =
lim supnæŒ fn (x) and g4 (x) = lim inf næŒ fn (x) are both measurable. If, as in Exercise 2.2, we let a = 1,
then function g3 ≠ g4 is measurable (and is equal to 1 when g3 = g4 = ±Œ). Finally, by noting that
lim fn (x) exists ≈∆ g3 = g4 , we can actually write:

{x œ X | lim fn (x) exists} = Kernel(g3 ≠ g4 ) = {x œ X | g3 (x) = g4 (x)} = (g3 ≠ g4 )≠1 (0)

Which is most certainly measurable since g3 and g4 are measurable, and the difference of such measurable
functions is also measurable (Corollary of Exercise 4.2 by combining the fact that f measurable ≈∆
≠f measurable, and taking f ≠ g = f + (≠g)).

2.4 Folland 2.4


Prove the following Proposition:
Proposition. 2.4:

If f : X æ R and f ≠1 ((r, Œ]) œ M for each r œ Q, then f is measurable.

Proof. Firstly, by the density of the rationals, (a, Œ] = firœQ+ r


(r, Œ], where a œ R, and Q+
a := {r œ
Q | r > a}. Naturally since Q+a is countable and BR is generated by the intervals in the form of (a, Œ],
and since: €
f ((a, Œ]) µ
≠1
f ≠1 ((r, Œ]) œ M
rœQ+
a

By (Folland) Proposition 2.1, it follows that f is measurable.

2.5 Folland 2.7


Prove the following Proposition:
Proposition. 2.5:
Suppose that for each – œ R we are given a set E– œ M such that E– µ E— whenever – < —,
fi–œR E– = X, and fl–œR E– = . Then there is a measurable function f : X æ R such that
f (x) Æ – on E– and f (x) Ø – on E–c for every –. (Use (Folland) Exercises 2.4).

Proof. We claim that f (x) := inf{– œ R | x œ E– }, where E– has the same construction as given in the
Proposition, will satisfy the requirements of being measurable and the stated inequalities. We begin first
by showing the latter.
Suppose x œ E– , then by the construction of f , we immediately have f (x) Æ –. Now, suppose – œ E–c ,
then ’— Æ –, E–c µ E—c since E— µ E– ; therefore, x œ E—c ∆ x ”œ E— ’— Æ – ∆ f (x) Ø – if x œ E–c .

17
Again by the construction of f , it is clear that fi–œR E– = X and fi–œR E–c = X. From this, given ’x œ X,
we know that ÷ –, — œ R such that x œ E– and x œ E— and most importantly since –, — œ R:

≠Œ < – Æ f (x) Æ — < Œ

And hence f (x) ”= ±Œ irregardless of x. It’ll now be a lot easier to conclude measurability since we no
longer have to worry about the possibility that f (x) = ±Œ.
Let us now take r œ Q, and note that by first set of inequalities established, if x œ X, then f (x) < r ≈∆
÷q œ R s.t. x œ Eq . Equivalently: f ≠1 ((≠Œ, r)) = fiq<r E– . By the density of Q, we can actually restrict
that q, r œ Q. We therefore have:

f ≠1 ((≠Œ, r)) = Eq , where q, r œ Q
q<r

And since Eq œ M ’q, and since {q œ Q | q < r} is a countable set, we naturally have f ≠1 ((≠Œ, r)) œ M.
Furthermore, by the inequalities established, we also have:

f ≠1 ([r, Œ)) = Erc œ M
q<r

And since we showed this to be true ’r œ Q, by Exercise 4.4, f is measurable.

2.6 Folland 2.8


Prove the following Proposition:
Proposition. 2.6:
If f : R æ R is monotone, then f is Borel measurable.

Proof. We first state our strategy: If we can show that ’a œ R, f ≠1 ([amŒ)) is an interval, then f must
be Borel measurable., let us note that as trivial corollary of (Folland) Proposition 2.3, f measurable
≈∆ ≠f measurable. Thus, without loss of generality, assume f is monotone increasing. Suppose now
that a œ R, x œ f ≠1 ([a, Œ), and y œ [x, Œ). Therefore, since f is monotone increasing:

a Æ f (x) Æ f (y) ∆ y œ f ≠1 ([a, Œ))

Since this is true ’x, y œ [a, Œ), it actually proves that f ≠1 ([a, Œ)) is indeed an interval, and therefore
Borel measurable, and hence f is Borel measurable since this is true ’a œ R.

2.7 Folland 2.9


Prove the following Proposition:

18
Proposition. 2.7:

Let f : [0, 1] æ [0, 1] be the Cantor Function (Folland Section 1.5), and let g(x) = f (x) + x.
a) g is a bijection from [0, 1] to [0, 2], and h = g ≠1 is continuous from [0, 2] to [0, 1].
b) If C is the Cantor set, m(g(C)) = 1.

c) By (Folland) Exercise 29 of Chapter 1, g(C) contains a Lebesgue non-measurable set A. Let


B = g ≠1 (A). Then B is Lebesgue measurable but not Borel.
d) There exist a Lebesgue measurable function F and a continuous function G on R such that
F ¶ G is not Lebesgue measurable.

Proof.
a) We first recall (from Folland) that the Cantor Function, f (x) is monotone increasing, and naturally
h(x) = x is a strictly increasing function, and hence g(x) = f (x) + x is also strictly increasing
and therefore injective. Next, to show surjectivity, note that g is a continuous function, and
g(0) = f (0) + 0 = 0, and g(1) = f (1) + 1 = 2; hence, by the intermediate value theorem, g is
surjective.

We now have all the necessary components to conclude that g is a bijection, and since g is a
continuous bijective function, and [0, 1] is compact, g’s inverse, g ≠1 , is continuous from [0, 2] to
[0, 1].
b) Firstly, by g’s surjectivity, and C being measurable, we see that:

g([0, 1]\C) Û g(C) = g([0, 1] fl C c ) Û g(C) = [0, 2] ∆ m(g(C)) + m(g([0, 1]\C)) = 2

Next, since C is a closed set ∆ [0, 1]\C is an open set. Therefore, since all open subsets of [0, 1] may
be written as a countable union of disjoint open sets, let us write [0, 1]\C = ی 1 Oj , Oj = (aj , bj ).
Now, since f is by construction constant on [0, 1]\C, and recalling that m(C) = 0 ∆ m([0, 1]\C) =
1 ∆ m(ÛŒ 1 Oj ) = 1, we see:
3 3h
Œ 44 ÿŒ
m(g([0, 1]\C)) = m g Oj = m(g(Oj ))
j=1 j=1
Π3
ÿ 4
= m(f (bj ) ≠ f (aj )) + m(bj ≠ aj )
j=1
ÿŒ
= m(Oj ) since f (bj ) = f (aj ) ’j œ N
j=1
3h
Π4
=m Oj
j=1

=1

And hence m(g(C)) = 1 by the the first part of this proof.

c) To show Lebesgue measurability, naturally B µ C, and since C is measurable with measure m(C) =
0, it implies m(B) Æ m(C) = 0, and hence Lebesgue measurable since null sets are measurable.

19
For the sake of contradiction, suppose B = g ≠1 (A) is Borel measurable. In part a), we showed that
g ≠1 is continuous and bijective; therefore g(B) = g(g ≠1 (A)) = A. However, by the continuity of g,
if g ≠1 (A) was Borel, so too would g(g ≠1 (A)) = A, hence a contradiction since A is not Lebesgue
measurable; therefore, B cannot be Borel measurable.
I
1 if x œ B
d) Let F = ‰B ; I.e., F (x) = , and also set G = g ≠1 . Naturally G is Lebesgue mea-
0 if x œ B c

surable since it is continuous, we now wish to prove that so too is F . This can be seen by notic-
ing F ≠1 ((a, Œ)) = or B or R, but all these possibilities are Lebesgue measurable, hence F is
Lebesgue measurable. We can now look at the following reasoning:

(F ¶ G)≠1 ((1/2, Œ)) = G≠1 ¶ F ≠1 ([1/2, Œ)) = {x œ [0, 2] | ‰B (g ≠1 (x)) œ [1/2, Œ)}
= {x œ [0, 2] | g ≠1 (x) œ B}
= G≠1 (B) = g(g ≠1 (A)) = A

Now since A is not Lebesgue measurable, F ¶ G also will not be Lebesgue measurable.

2.8 Folland 2.10


Prove the following Proposition:
Proposition. 2.8:
The following implications are valid ≈∆ the measure µ is complete:
a) If f is measurable and f = g µ-a.e., then g is measurable.

b) If fn is measurable for n œ N and fn æ f µ-a.e., then f is measurable.

Proof.
a) For the forward direction, suppose a) holds. Then let N œ M be a measurable set s.t. µ(N ) = 0,
and N1 µ N . If we define f :© 0 and ‰N1 := 1 if x œ N1 , and 0 otherwise, then trivially f is
measurable and f = ‰N1 µ-a.e., so by our assumptions g is measurable. Now, by noting that
N1 ({1}) = N1 œ M by g’s measurability, and since this is true ’N1 µ N , we have arrived at the
‰≠1
definition of µ being complete.

For the backward direction, suppose µ is complete, and let f be measurable and f = g µ-a.e.
Explicitly, let N œ M be the measurable set s.t. µ(N ) = 0 and f (x) = g(x) ’x œ N c . Then if A is
measurable, we have:
# $ # $ # $ # $
g ≠1 (A) = g ≠1 (A) fl N Û g ≠1 (A) fl N c = g ≠1 (A) fl N Û f ≠1 (A)\N

Looking at the right hand side, we can see g ≠1 (A) fl N µ N is measurable by the definition of µ
being a complete measure since µ(N ) = 0. Furthermore, f ≠1 (A)\N µ f ≠1 (A) since f is measurable.
With these two facts, we may therefore conclude that g is indeed measurable.
b) For the forward direction, suppose b) holds. Then let N œ M be a measurable set s.t. µ(N ) = 0,
and N1 µ N . If we let fn = 0 and ‰N1 as before, then like in the forward direction of a), we have
fn æ ‰N1 µ-a.e., so ‰N1 is measurable. Therefore, ‰≠1N1 ({1}) œ M, and since this is true ’N1 µ N ,
we have arrived at the definition of µ being complete.

20
For the backward direction, suppose µ is complete, and fn is measurable ’n œ N, and fn æ f
µ-a.e. By (Folland) Proposition 2.7, g3 (x) = lim supjæŒ fj (x) is measurable since fn is measurable
’n œ N. Furthermore, since fn æ f µ-a.e., we have g3 = f µ-a.e., and thus by the backward
direction of part a) above, f is measurable.

2.9 Folland 2.12


Prove the following Proposition:
Proposition. 2.9:
s
If f œ L+ and f < Œ, then {x | f (x) = Œ} is a null set and {x | f (x) > 0} is ‡-finite.

s Let E := {x | f (x) = Œ}, F := {x | f (x) > 0}, Fn := {xŒ| f (x) > 1/n},
Proof. and f satisfy f œ L+
and f < Œ. Let us now define the two sets of functions {„n }1 and {Ïn }Œ 1 , where „n = n‰E and
Ïn = ‰Fn /n.
To prove E is a null set, we make the observation that since f (x) = Œ ’x œ E, and ‰n (x) < Œ ’n œ N,
we have:
⁄ ⁄
0 Æ „n (x) Æ f (x) ’x œ X ∆ nµ(E) = „n dµ Æ f dµ

1
∆ µ(E) Æ f dµ
n
s
Thus, since f ”µ < Œ, letting n æ Œ, we see that µ(E) = 0; I.e., E is a null set.
By the construction of {Fn }Œ
1 , we have fi1 Fn , so to conclude that F is ‡-finite, wessimply need to show
Œ

that µ(Fn ) < Œ ’n œ N. This is easily ascertained since f (x) > 1/n ’x œ Fn , and f < Œ, we have:
⁄ ⁄
1
0 Æ Ïn (x) Æ f (x) ’x œ Fn ∆ µ(Fn ) = Ïn dµ Æ f dµ
n

∆ µ(Fn ) Æ n f dµ < Œ

And hence µ(Fn ) < Œ ’n œ N, which implies F is ‡-finite.

2.10 Folland 2.13


Prove the following Proposition:
Proposition. 2.10:
+
s s s s
Suppose {fn }Œ
1 µ L , fn æ f pointwise,s and f = lim fn < Œ. Then E
f = lim f
E n
’E œ M. However, this need not be true if f = lim fn = Œ.

s s s
Proof. Let E œ M and f < Œ, and so we define ‰E s.t. E f = ‰E f , and so we have:
⁄ ⁄ ⁄ ⁄
f = ‰E f Æ f = lim fn < Œ
E

21
Furthermore, by (Folland) Theorem 2.15, we have:
⁄ ⁄ ⁄ ⁄
f = (‰E f + ‰E c f ) = ‰E f + ‰E c f

And similarly for substituting fn for f above. Now, since fn æ f ∆ ‰F fn æ ‰F f ’F œ M, we may


apply Fatou’s Lemma as follows:
⁄ ⁄ ⁄ 3⁄ ⁄ 4 ⁄ ⁄
ú úú
f = lim inf ‰E fn Æ lim inf fn = lim inf fn ≠ fn = f ≠ lim sup fn
E næŒ næŒ E næŒ Ec næŒ Ec

ú s s s úú s s s
Where we s have = sinces E f = ‰E f + ‰E c f , and = since lim inf fn = lim fn = f and
lim inf ≠ g = ≠ lim sup g. However, since all terms above are finite, we may gain by apply Fatou’s
Lemma (and in noticing the similarity to the steps made above) to see that:
⁄ 3⁄ ⁄ 4 ⁄ ⁄ ⁄ ⁄
lim sup fn = lim sup fn ≠ fn = f ≠ lim inf fn Æ f ≠ f
næŒ Ec næŒ E næŒ E E

And thus by substituting this in, we have:


⁄ ⁄ ⁄ ⁄
f Æ lim inf fn Æ lim sup fn Æ f
E næŒ E næŒ E E

And therefore all the inequalities in the equation(s) above are actually equalities, and so we have:
⁄ ⁄ ⁄ ⁄
lim inf fn = lim sup fn = lim f= f
næŒ E næŒ E næŒ E E
s s
We now turn our attention showing the above result need not hold if f = lim f = Πby means of a
counter-example. Let E = (0, 1], f = ‰[2,Œ) , and fn = ‰[2,Œ) + n‰(0,1/n] . Then fn æ f p.w., and:
⁄ ⁄
fn = nµ((0, 1/n]) = 1 ’n œ N ∆ lim fn = 1
(0,1] næŒ (0,1]
s s s s s
However, (0,1]
f = 0, thus E
f = lim E
fn need not be true if lim f= f = Œ.

2.11 Folland 2.14


Prove the following Proposition:
Proposition. 2.11:
+
s +
If
s f œ L s, let ⁄(E) = E
f dµ for E œ M. Then ⁄ is a measure on M, and for any g œ L ,
g d⁄ = f g dµ. (First Suppose that g is simple.)

Proof. Trivially, since f œ L+ , we have that ⁄(E) Ø 0 ’E œ M. Moreover, one can see that ⁄( ) = 0:
⁄ ⁄
⁄( ) = f dµ = ‰ f dµ = 0

22
To fully show
qŒthat ⁄ is a measure on M, we need that for any disjoint sequence of sets, {E j } Œ
1 œ M,
⁄(Û1 Ej ) = 1 ⁄(Ej ). We can deduce this fact from the following:
Œ

3h
Œ 4 ⁄ ⁄
⁄ Ej = f dµ = ‰(ÛŒ
1 Ej )
f dµ
j=1 ی
1 Ej
⁄ 3ÿ
Œ 4 Œ ⁄
ÿ
ú ú
= ‰Ej f dµ = ‰Ej f dµ = by (Folland) Theorem 2.15
j=1 j=1
Œ ⁄
ÿ Œ
ÿ
= f dµ = ⁄(Ej )
j=1 Ej j=1

We have thus shown all the necessary conditions for ⁄ to be a measure do indeed hold.
qn
Next, let g œ L+ , and assume that g is simple ∆ g = 1 aj ‰Ej . Therefore:
⁄ n
ÿ n ⁄
ÿ n ⁄
ÿ
g d⁄ = aj ⁄(Ej ) = f dµ = ‰Ej f dµ
j=1 j=1 Ej j=1
⁄ 3ÿ
n 4 ⁄
ú ú
= aj ‰Ej f dµ = gf dµ = by (Folland) Theorem 2.15
j=1

And so we get the required result when g is simple. However, by (Folland) Theorem 2.10, we know that
since f œ L+ , ÷{„n }Œ
1 s.t. 0 Æ „1 Æ „2 Æ · · · Æ f , „n æ f p.w., and „n æ f uniformly on any set on
which f is bounded. Therefore, we can apply the Monotone Convergence Theorem (used if = denoted)
ı

as follows:
⁄ ⁄ ⁄ ⁄
ú ú
g d⁄ = lim „n d⁄ = lim „n f dµ = gf dµ = since „n simple ’n œ N
ı ı
næŒ næŒ

2.12 Folland 2.16


Prove the following Proposition:
Proposition. 2.12:
s s !s "
If f œ L+ and f < Œ, ’‘ > 0 ÷E œ M s.t. µ(E) < Œ and E f > f ≠‘

Proof. Firstly, By (Folland) Exercise 2.12 (proved above - 5.2), we know that F := {x |f (x) > 0} is
‡-finite. In the proof of (Folland) 2.12, we showed that Fn := {x | f (x) > 1/n} has the nice properties
of µ(Fn ) < Œ and fiŒ 1 Fn = F . Furthermore, it is also apparent from the construction of Fn that
Fn µ Fn+1 ’n œ N - I.e., {Fn }Œ 1 is monotone increasing, and so {‰Fn }1 will be an increasing sequence
Œ

in L+ s.t. ‰Fn Æ ‰Fn+1 ’n œ N, and limnæŒ ‰Fn = ‰F .


Since {‰sFn }Œ
1 sand ‰F satisfy necessary conditions for the Monotone Convergence Theorem, and in
noticing f = ‰F f , we may apply it as follows:
⁄ ⁄ ⁄ ⁄
f = ‰F f = lim ‰Fn f = f
næŒ Fn

23
We also note that, since ‰Fn µ ‰F ’n œ N, we have:
⁄ ⁄ ⁄ ⁄
f = ‰F f Æ ‰Fn f = f ’n œ N
Fn
s s
Therefore, Fn is an increasing sequence with the limit of f . So by this convergence, we have ’‘ > 0,
÷N œ N such that: ⁄ 3⁄ 4
f> f ≠‘
FN
s !s "
I.e., we have proven the existence of an FN = E œ M which satisfies FN
f> f ≠ ‘.

2.13 Folland 2.17


Prove the following Proposition:
Proposition. 2.13:
Assume Fatou’s lemma and deduce the monotone convergence theorem from it.

+
Proof. Let {fn }Œ
1 be a sequence in L s.t. fj Æ fj+1 ’j œ N, and f = limnæŒ fn . If we’re assuming
Fatou’s Lemma, then: ⁄ ⁄ ⁄
f = lim inf fn Æ lim inf fn
næŒ næŒ
s
However,
s since {fn }Œ
1 is monotone increasing with the limit of f , we have fn Æ f ’n œ N ∆ fn Æ
f ’n œ N. And hence taking the lim sup on both sides, we get:
⁄ ⁄ ⁄
lim sup fn Æ lim sup f = f
næŒ næŒ

Therefore, in combining these two inequalities, we see:


⁄ ⁄ ⁄
lim sup fn Æ f Æ lim inf fn
næŒ næŒ

Which can be true ≈∆ all the inequalities above are actually equalities, hence we have:
⁄ ⁄ ⁄ ⁄
lim fn = lim sup fn = lim inf fn = f
næŒ næŒ næŒ

2.14 Differentiable functions are Borel Measurable


Exercise. 2.1:
Let f : R æ R be a differentiable function, show that its derivative f Õ is Borel Measurable.

Proof. Firstly, we note that by (Folland) Corollary 2.2, since f œ C 1 (R) ∆ f œ C(R), we have that f is
Borel measurable.

24
Next, we prove that gn := f (x + 1/n) is Borel measurable. This is actually quite easy since hn = x + 1/n
is naturally Borel measurable, and hence f ¶ hn = gn is Borel measurable since both f and gn are Borel
measurable.
Next, since f œ C 1 (R), we know that limhæ0 f (x+h)≠f
h
(x)
= f Õ (x), f Õ (x) œ R. Therefore, we can also say
that limnæŒ n(f (x + 1/n) ≠ f (x)) = f Õ (x) ’x œ R. Since we already showed f (x + 1/n) and f (x) are
Borel Measurable, by (Folland) Proposition 2.6, fnÕ := n(f (x + 1/n) ≠ f (x)) is Borel measurable ’n œ N.
Finally, by (Folland) Proposition 2.7, we can conclude that f Õ (x) = limnæŒ fnÕ (x) is Borel measurable
since f œ C 1 (R), fnÕ æ f , and {fn }Œ
1 is a sequence of Borel measurable functions.

2.15 Folland 2.20


Prove the following Proposition:
Proposition. 2.14:

(A generalizeds Dominated
s Convergence
s Theorem)
s If fn , gn , f, g œ L1 , fn æ f and gn æ g a.e.,
fn Æ gn and gn æ g, then fn æ f . (Rework the proof of the dominated convergence
theorem).

Proof. By the same reasoning as in Folland, WLOG we may assume fn and f are real-valued, and that
gn + fn Ø 0 a.e., and gn ≠ fn Ø 0 a.e. Now, we apply (Folland) Corollary (of Fatou’s Lemma) 2.19 to
both gn + fn and gn ≠ fn as follows (we can do so due to the convergent and L1 assumptions):
⁄ ⁄ ⁄ ⁄ ⁄
(g + f ) = lim(gn + fn ) Æ lim inf (gn + fn ) = g + lim inf fn
⁄ ⁄ ⁄ ⁄ ⁄
(g ≠ f ) = lim(gn ≠ fn ) Æ lim inf (gn ≠ fn ) = g ≠ lim sup fn

And so: ⁄ ⁄ ⁄ ⁄ ⁄ ⁄ ⁄ ⁄
lim sup fn ≠ gÆ≠ g+ f and g+ fÆ g + lim inf fn

And by combining these inequalities, we see that:


⁄ ⁄ ⁄
lim sup fn Æ f Æ lim inf fn
s
And since sf, fn œ sL1 , we know that the above inequalities imply equalities, everywhere, I.e., lim fn
exists and fn æ f .

2.16 Folland 2.21


Prove the following Proposition:
Proposition. 2.15:
1
s s s
Suppose fn , f œ L and fn æ f a.e. Then |fn ≠ f | æ 0 ≈∆ |fn | æ |f |, (Use (Folland)
Exercise 20).

25
s
Proof. For the forward direction, assume |fn ≠ f | æ 0, then since:
-⁄ ⁄ - -⁄ - ⁄ - - ⁄
- - - ! "- - ! " -
- |fn | ≠ |f |- = - |fn | ≠ |f | -- Æ - |fn | ≠ |f | - Æ |fn ≠ f |, since |fn | ≠ |f | Æ |fn ≠ f |
- - -

we
- s know that
s the- right hand
s sidesæ 0 as n æ Œ, and since the above holds ’n œ N (and since fn , f œ L1 ),
- |fn | ≠ |f |- æ 0 ∆ |fn | æ |f |.
s s
For the backward direction, assume s|fn | æs |f |. If we let gns := |fn | + |f |, then naturally |fn ≠ f | Æ gn ,
and since fn , f œ L1 , we know that gn = (|fn | + |f |) = 2 |f |. We may now invoke the generalized
dominated convergence theorem, (Folland) Exercise 2.20 above, which implies:
⁄ ⁄
lim |fn ≠ f | = lim |fn ≠ f |
s
And since fn æ f , we therefore have |fn ≠ f | æ 0.

2.17 Folland 2.24


2.18 Folland 2.34
Prove the following Proposition:
Proposition. 2.16:

Suppose |fn | Æ g œ L1 and fn æ f in measure.


s s
a) f = lim fn .
b) fn æ f in L1 .

Before we begin, we present Folland Exercise 33 as a necessary Lemma for part a):
2.18.1 Folland 2.33
Lemma. 2.1:
s s
If fn Ø 0 and fn æ f in measure, then f Æ lim inf fn .

Proof.
a) By (Folland) Theorem 2.30, ÷ a subsequence {fnk }Œ 1 s.t. fnk æ h, where f = h a.e. Furthermore,
by (Folland) Proposition 2.11, since f = h a.e., f is measurable. As is standard by this point, we
may assume fn and g are real-valued functions; therefore, g + fn Ø 0 a.e., and g ≠ fn Ø 0 a.e.
Moreover, we naturally have g + fn æ g + f , g ≠ fn æ g ≠ f in measure. We now make use of our
Lemma as follows:
⁄ ⁄ ⁄ ⁄ ⁄ ⁄
g+ f= (g + f ) Æ lim inf (g + fn ) = g + lim inf fn
⁄ ⁄ ⁄ ⁄ ⁄ ⁄
g≠ f= (g ≠ f ) Æ lim inf (g ≠ fn ) = g ≠ lim sup fn

And so in combining these inequalities, we have:


⁄ ⁄ ⁄
lim sup fn Æ f Æ lim inf fn

26
Which we recognizes as in previous
s exercises to be true ≈∆ all the above inequalities are actually
equalities; hence: f = lim fn .
b) From (Folland) Proposition 2.29, we know that since fn æ f in L1 , fn æ f in measure as well.
Thus, since:
1) - *2 1) - *2
µ x œ X - ||fn (x) ≠ f (x)| ≠ 0| Ø ‘ = µ x œ X - |fn (x) ≠ f (x)| Ø ‘ æ 0 as n æ Œ

We also have that |fn ≠ f | æ 0 in measure. Furthermore, since: |fn ≠ f | Æ |fn | + |f | Æ 2g œ L1 , we


may apply Part a) to see that:
⁄ ⁄
0 = 0 dµ = lim |fn ≠ f |dµ

Hence fn æ f in L1 .

2.19 Folland 2.39


Prove the following Proposition:
Proposition. 2.17:
If fn æ f almost uniformly, then fn æ f a.e. and in measure.

1
Proof. We first recall the fn æ f almost uniformly means that ÷{En }Œ
1 µ M s.t. µ(En ) < n and fn æ f
c

uniformly on En . If we define E := fiŒ
1 En , then µ(E ) Æ lim inf µ(En ) = 0; hence, fn æ f a.e.
c c

Now to show fn æ f in measure, we proceed as follows. ’‘, ” > 0, since fn æ f almost uniformly,
÷E œ M and an N œ N s.t. ’n Ø N, |fn (x) ≠ f (x)| < ‘ ’x œ E and µ(F c ) < ”. An immediate result of
this set up is therefore:
1) *2
lim inf µ x œ X | |fn (x) ≠ f (x)| Ø ‘ Æ µ(F c ) < ”
næŒ

And since our result works ’” > 0, letting ” æ 0 proves fn æ f in measure.

2.20 Folland 2.42


Prove the following Proposition:
Proposition. 2.18:
Let µ be a counting measure on N. Then fn æ f in measure ≈∆ fn æ f uniformly.

Proof. For the forward direction, suppose fn æ f in measure (µ a counting measure on N). Then ’‘ > 0
÷N œ N s.t. ’n Ø N :
1) - *2 1 ) - *
µ x œ N |fn (x) ≠ f (x)| Ø ‘ <
- ∆ x œ N - |fn (x) ≠ f (x)| Ø ‘ =
2
I.e., |fn (x) ≠ f (x)| < ‘ ’x œ N (and n Ø N ), which is by definition uniform convergence.

27
For the converse, suppose fn æ f uniformly (again µ a counting measure on N). Then ’‘ > 0, ÷N œ N
s.t. ’n Ø N :
1) - *2
|fn (x) ≠ f (x)| < ‘ ’x œ N ∆ µ x œ N |fn (x) ≠ f (x)| Ø ‘ = µ( ) = 0
-

For which the latter equality vacuously satisfies our definition of convergence in measure.

2.21 Folland 2.44: Lusin’s Theorem


Prove the following Theorem:
Theorem. 2.1: Lusin’s Theorem
If f : [a.b] æ C is Lebesgue measurable and ‘ > 0, there is a compact set E µ [a, b] such that
µ(E c ) < ‘ and f |E is continuous. (Use Egoroff’s Theorem and (Folland) Theorem 2.26.)

Proof. Following the hint, by (Folland) Theorem 2.26, ÷ {fn }Œ


1 s.t. fn : [a, b] æ C and fn æ f a.e. Also,
by Egoroff’s Theorem, ÷F µ [a, b] s.t. µ(F ) < ‘/2 and fn æ f uniformly on F .
c

Naturally, since fn æ f uniformly on F c , f |F will be continuous. We now make use of (Folland) Theorem
1.18 which states that ’F œ Mµ :
µ(F ) = sup{µ(K) | K cpt µ F }
And so by definition of sup, ÷ E cpt µ F s.t. µ(F \E) < ‘/2. We thus have:
‘ ‘
µ(E c ) = µ(F c ) + µ(F \E) < + = ‘
2 2
And since E µ F , we know that f |E is also continuous, thereby proving the Theorem.

2.22 Folland 2.46


Prove the following Proposition:
Proposition. 2.19:

Let X = Y = [0, 1], M = N = B[0,1] , µ = Lebesgues smeasure, and s s ‹ = counting smeasure. If


D = {(x, x) | x œ [0, 1]} is the diagonal
s in X ◊Y , then ‰D dµd‹, ‰D d‹dµ, and ‰D d(µ◊‹)
are all unequal. (To compute ‰D d(µ ◊ ‹) = µ ◊ ‹(D), go back to the definition of µ ◊ ‹.)

s s
Proof. We begin by making the following two observations: ’x œ [0, 1], ‰D d‹(y) = {x} d‹(y) =
s s
‹({x}) = 1 and ’y œ [0, 1], ‰D dµ(x) = {y} d‹(x) = µ({y}) = 0. Therefore, we’ll now be able to
ss ss
compute ‰D dµd‹ and ‰D d‹dµ as follows:
⁄ ⁄ ⁄ 3⁄ 4 ⁄
‰D dµd‹ = ‰D (x, y)dµ(x) d‹(y) = 0d‹(y) = 0
⁄ ⁄ ⁄ 3⁄ 4 ⁄
‰D d‹dµ = ‰D (x, y)d‹(y) dµ(x) = dµ(x) = 1
[0,1]
s
We now claim that ‰D d(µ ◊ ‹) = Œ. To see this, suppose {An ◊ Bn }Œ
1 s.t. An , Bn µ [0, 1] (measurable
subsets) s.t. D µ fi1 (An ◊ Bn ). Therefore, [0, 1] µ fi1 (An ◊ Bn ). Because
Œ Œ
qŒ of this, ÷N œ N s.t.
s (AN ◊ BN ) > 0, and explicitly µ(AN ) > 0, and ‹(BN ) = Œ. Therefore, 1 µ(An )‹(Bn ) = Œ ∆
ú
µ
‰D d(µ ◊ ‹) = Œ.

28
2.23 Folland 2.48
Prove the following Proposition:
Proposition. 2.20:

Let X = Y = N, M = N = P(N), µ = ‹ = counting measure. Define:


Y
]1
_ if m = n
f (m, n) = ≠1 if m = n + 1
_
[
0 otherwise
s ss ss
Then |f |d(µ ◊ ‹) = Œ, and f dµd‹ and f d‹dµ exist and are unequal.

Proof. We first claim that µ ◊ µ is also a counting measure. We may actually note that fundamentally,
a counting measure · on N ◊ N will satisfy · (A ◊ B) = |A||B| = µ(A)‹(B). Therefore, since rectangles
generate the product ‡-algebra, the ‡-finitness of · implies µ ◊ ‹ = · .
We now proceed to computing each of the quantities of interest. For the first, if we let E := fiŒ
1 {{(n, n)}fi
{(n, n + 1)}}, then clearly |E| = Œ, and |f | = ‰E . Therefore:

|f |d(µ ◊ ‹) = |E| = Œ

Furthermore, the other two calculations are nearly immediate:


⁄ ⁄ ⁄ ⁄ ÿÿ ÿ
f dµd‹ = f (m, n)dµ(m)d‹(n) = f (m, n) = 0=0
n m n
⁄ ⁄ ⁄ ⁄ ÿÿ ÿ
f dµd‹ = f (m, n)d‹(n)d‹(m) = f (m, n) = ‰{m=1} = 1
m n m

2.24 Folland 2.49


Prove the following Proposition:
Proposition. 2.21:

Prove (Folland) Theorem 2.38 by using Theorem 2.37 and Proposition 2.12 together with the
following lemmas:
a) If E œ M ◊ N and µ ◊ ‹(E) = 0, then ‹(Ex ) = µ(E y ) = 0 for a.e. x and y.

b) If
s f is L-measurable
s y and f = 0 ⁄-a.e., then fx and f y are integrable for a.e. x and y, and
fx d‹ = f dµ = 0 for a.e. x and y. (Here the completeness of µ and ‹ is needed.)

Proof.
a) Immediately by (Folland) Theorem 2.36, since (X, M, µ) and (Y, N, ‹) are ‡-finite measure spaces,
we have: ⁄ ⁄
0 = µ ◊ ‹(E) = ‹(Ex )dµ(x) = µ(E y )d‹(y)

29
b) Let us define F := {(x, y) œ M ◊ N | f (x, y) ”= 0}. Thus, ÷E œ M ¢ N where µ ◊ ‹(E) = 0 and
F µ E. From a), µ(Ex ) = ‹(E y ) = 0 for x, y a.e. Furthermore, by the fact that Fx µ Ex and
F y µ E y , by linearity of measures we have µ(Fx ) = ‹(F y ) = 0 as well. We may now conclude thus
that: ⁄ ⁄ ⁄ ⁄
|fx |d‹ = ‰Fx |fx |d‹ = 0 = ‰F y |f y |dµ = |f y |dµ = 0 for a.e. x and y

For which the intended result trivially follows.


Please see the next (attached) page for our proof of (Folland) Theorem 2.38.

3 Chapter 3
3.1 Folland 3.2
Prove the following Proposition:
Proposition. 3.1:

a) If ‹ is a signed measure, E is ‹-null ≈∆ |‹|(E) = 0.


b) If ‹ ane µ are signed measures, ‹ ‹ µ ≈∆ |‹| ‹ µ ≈∆ ‹ + ‹ µ and ‹ ≠ ‹ µ.

Proof.

a) For the forward direction, suppose ‹ is a signed measure and that E is ‹-null. Suppose for the sake
of contradiction that |‹|(E) = ‹ + (E) + ‹ ≠ (E) > 0, where ‹ = ‹ + ≠ ‹ ≠ is the Jordon Decomposition
of ‹. By the Hahn Decomposition Theorem, ÷ P, N s.t. ‹(X) = ‹(P fi N ) = ‹ + (P ) ≠ ‹ ≠ (N ) and
‹ + (N ) = 0 = ‹ ≠ (P ).
We thus can thus make the following observations:

|‹|(E) = ‹ + (E) + ‹ ≠ (E) = 2‹ + (E) > 0 since ‹(E) = 0 ∆ ‹ + (E) = ‹ ≠ (E)

‹ + (E fl P ) = ‹ + (E fl N ) + ‹ + (E fl P ) = ‹ + (E fl X) = ‹ + (E) > 0 since 2‹ + (E) > 0


‹ ≠ (E fl P ) Æ ‹ ≠ (P ) = 0
And so:
‹(E fl P ) = ‹ + (E fl P ) + ‹ ≠ (E fl P ) = ‹ + (E fl P ) > 0
However, since E fl P µ E, and we are assuming that ‹(E) = 0, we arrive at a contradiction with
the last inequality. Thus, actually if E is ‹-null, |‹|(E) = 0.

For the converse, suppose |‹|(E) = 0, hence |‹|(E Õ ) = 0 ’E Õ µ E (E Õ measurable). Since |‹|(E Õ ) =
‹ + (E Õ ) + ‹ ≠ (E Õ ) = 0 ≈∆ ‹ + (E Õ ) = 0 = ‹ ≠ (E Õ ), we thus trivially satisfy ‹(E Õ ) = ‹ + (E Õ ) ≠
‹ ≠ (E Õ ) = 0 since both are already zero.
b) Let us recall that, explicitly, if ÷ E, F œ M s.t. E flF = , E ÛF = X and µ(E Õ ) = 0 = ‹(F Õ ) ’E Õ µ
E, F Õ µ F (E Õ , F Õ measurable), we denote this property as ‹ ‹ µ.

We begin by showing ‹ ‹ µ ∆ |‹| ‹ µ. To see this, since F is ‹-null, by Part a), we know that
|‹|(F ) = 0. Since |‹| is a positive (regular) measure, by monotonicity we have that |‹| is F -null.
Thus, by definition, |‹| ‹ µ.

30
We now show |‹| ‹ µ ∆ ‹ + ‹ µ and ‹ ≠ ‹ µ. Since ‹ = ‹ + + ‹ ≠ , we have ‹ + Æ ‹ and ‹ ≠ Æ ‹,
and so ‹ + (F Õ ) = ‹ ≠ (F Õ ) = 0, I.e., F is both ‹ + -null and ‹ ≠ -null; hence, ‹ + ‹ µ and ‹ ≠ ‹ µ.

We may now complete Part b) by showing [‹ + ‹ µ and ‹ ≠ ‹ µ] ∆ ‹ ‹ µ. Let us make explicit the
properties associated with ‹ + ‹ µ and ‹ ≠ ‹ µ by replacing the roles of E, F (from the beginning
of this proof) with A1 , A2 for ‹ + ‹ µ and B1 , B2 for ‹ ≠ ‹ µ. We first note that since A1 , B1
are both µ-null, so too is A1 fi B1 . This is true since by looking at the following representation:
A1 fi B1 © A1 Û (B1 \A1 ), and in noting B1 \A1 µ B1 , ’E Õ µ A1 fi B1 , ÷AÕ1 µ A1 , B1Õ µ B1 s.t.
E Õ = B1Õ Û A1 and both B1Õ and AÕ1 are µ-null. Furthermore, since A1 fi A2 = X = B1 fi B2 , we have
X\(A1 fi B1 ) = A2 fl B2 , which is both ‹ + -null and ‹ ≠ -null since A2 fl B2 µ A2 and A2 fl B2 µ B2 .
Thus, by setting E = A1 fi B1 , and F = A2 fl B2 , we see that indeed ‹ ‹ µ.

3.2 Folland 3.7


Prove the following Proposition:
Proposition. 3.2:

Suppose that ‹ is a signed measure on (X, M) and E œ M.


) - * ) - *
a) ‹ + (E) = sup ‹(F ) - E œ M, F µ E and ‹ ≠ (E) = ≠ inf ‹(F ) - F œ M, F µ E .
) qn - *
b) |‹|(E) = sup 1 |‹(E j )| - n œ N, E1 , . . . , E n are disjoint, and Û n
1 E j = E .

Proof.
a) For the first equality, ’F µ E we have:
‹(F ) = ‹ + (F ) ≠ ‹ ≠ (F ) Æ ‹ + (F ) Æ ‹ + (E)
And so ‹ + (E) Ø sup{‹(F ) | E œ M, F µ E}. To see the reverse inequality, if P, N are our
Hahn Decomposition of ‹, we naturally have ‹ + (E) = ‹(E fl P ), and since E fl P µ E, ‹ + (E) Æ
sup{‹(F ) | E œ M, F µ E}, and so:
) - *
‹ + (E) = sup ‹(F ) - E œ M, F µ E
For the second inequality, this follows very similarly. Explicitly, ’F µ E, we have:
≠‹(F ) = ‹ ≠ (F ) ≠ ‹ + (F ) Æ ‹ ≠ (F ) Æ ‹ ≠ (E)
And so ‹ ≠ (F ) Ø sup{≠‹(F ) | F œ M, F µ E} = ≠ inf{‹(F ) | F œ M, F µ E}. For the reverse
inequality, since ‹ ≠ (E) = ≠‹(E fl N ), and since E fl N µ E, ‹ ≠ (E) Æ sup{≠‹(F ) | F œ M, F µ
E} = ≠ inf{‹(F ) | F œ M, F µ E}. Combining our two inequalities, we see:
) - *
‹ ≠ (E) = ≠ inf ‹(F ) - F œ M, F µ E

b) Firstly, if P, N are again the Hahn Decomposition of ‹, then E = (E fl N ) Û (E fl P ), and so:

|‹|(E) = ‹ + (E) + ‹ ≠ (E) = ‹ + (E fl P ) + ‹ ≠ (E fl N )


= ‹ + (E fl P ) + ‹ ≠ (E fl N ) + ‹ + (E fl N ) + ‹ ≠ (E fl P )
¸ ˚˙ ˝
=0
= |‹(E fl P )| + |‹(E fl N )|

31
And so:
;ÿ
n - n
h <
-
|‹|(E) Æ sup |‹(Ej )| - n œ N, E1 , . . . , En are disjoint, and Ej = E
1 1

To see the reverse inequality, we note that ’E = Ûn1 Ej , we have:


n
ÿ n
ÿ
|‹(Ej )| = |‹ + (Ej ) ≠ ‹ ≠ (Ej )|
j=1 j=1
ÿn
! "
Æ ‹ + (Ej ) + ‹ ≠ (Ej )
j=1
ÿn
= |‹|(Ej )
j=1
3h
n 4
= |‹| Ej = |‹|(E)
j=1

And so:
;ÿ
n - n
h <
-
|‹|(E) Ø sup |‹(Ej )| - n œ N, E1 , . . . , En are disjoint, and Ej = E
1 1

And hence combining the two inequalities, we have:


;ÿn - n
h <
-
|‹|(E) = sup |‹(Ej )| - n œ N, E1 , . . . , En are disjoint, and Ej = E
1 1

3.3 Folland 3.12


Prove the following Proposition:
Proposition. 3.3:

For j = 1, 2, let µj , ‹j be ‡-finite measures on (Xj , Mj ) s.t. ‹j << µj . Then ‹1 ◊ ‹2 << µ1 ◊ µ2


and:
d(‹1 ◊ ‹2 ) d‹1 d‹2
(x1 , x2 ) = (x1 ) (x2 )
d(µ1 ◊ µ2 ) dµ1 dµ2

Proof. Let us begin by defining fj := dµjj for j = 1, 2. Thus, if A1 ◊ A2 is measurable, by the definition
d‹

of product measure and Radon-Nikodym derivative, we have:


⁄ ⁄
‹1 ◊ ‹2 (A1 ◊ A2 ) = ‹1 (A1 )‹2 (A2 ) = f1 dµ1 f2 dµ2
⁄A1 A2

= f1 ‰A1 dµ1 f2 ‰A2 dµ2
⁄ ⁄
= f1 f2 ‰A1 ‰A2 dµ1 dµ2
⁄ ⁄
ú
= f1 f2 d(µ1 ◊ µ2 )
A1 ◊A2

32
ú
Where we have = by Tonelli’s Theorem. Therefore, on A1 ◊ A2 measurable, (f1 f2 )(µ1 µ2 ) = ‹1 ‹2 ; and
thus we also have equality on the algebra of finite unions of A1 ◊ A2 ’s. Furhermore, by the uniqueness
of the extension from premeasure to measure, (f1 f2 )(µ1 µ2 ) = ‹1 ‹2 on M1 ¢ M2 . We thus immediately
have that if (µ1 ◊ µ2 )(E) = 0 ∆ (‹1 ◊ ‹2 )(E) = 0, and so ‹1 ◊ ‹2 << µ1 ◊ µ1 . Finally, since the
Radon-Nikodym derivative is unique, we have:
d(‹1 ◊ ‹2 ) d‹1 d‹2
(x1 , x2 ) = f1 (x1 )f2 (x2 ) = (x1 ) (x2 )
d(µ1 ◊ µ2 ) dµ1 dµ2

3.4 Folland 3.13


Prove the following Proposition:
Proposition. 3.4:

Let X = [0, 1], M = B[0,1] , m = Lebesgue measure, and µ = counting measure on M, then:
a) m << µ but dm ”= f dµ for any f .
b) µ has no Lebesgue decomposition with respect to m.

Proof.
a) Firstly, if E œ M and µ(E) = 0, then it must be that E = , and so m(E) = m( ) = 0; I.e.,
m << µ. Suppose for the sake of contradiction that dm = f dµ, then ’x œ [0, 1] and E = {x}, we
have: ⁄ ⁄
0 = m(E) = f dµ = dm = m(E) = 0
E E
Thus we must have that f © 0 on [0, 1]. However:
⁄ ⁄
1 = m([0, 1]) = f dµ = 0dµ = 0
[0,1] [0,1]

I.e. we’ve reached a contradiction and hence dm ”= f dµ for any f .


b) Suppose, for the sake of contradiction, that µ has a Lebesgue decomposition w.r.t. m; namely:
µ = ⁄ + fl where ⁄ ‹ m and fl << m. Since ⁄ ‹ m, by definition we know that ÷E, F s.t X = E Û F
where E is ⁄-null and F is m-null (or just m(F ) = 0 since m is a positive measure). Suppose
x œ F , then µ({x}) = 1, but ⁄({x}) = 0 and m({x}) = 0 ∆ fl({x}) = 0, which would be a
contradiction unless we have F = . Thus, since X = E Û F = E Û ∆ E = X. However, since
m(E) = m([0, 1]) = 1, yet we are requiring E to be m-null, we arrive at a contradiction. Thus, @ a
Lebesgue decomposition of µ w.r.t. m.

3.5 Folland 3.17


Prove the following Proposition:
Proposition. 3.5:

Let (X, M, µ) be a ‡-finite measure space,s N a sub-‡-algebra


s of M, and ‹ = µ|N. If f œ L1 (µ),
÷g œ L1 (‹) (thus g is N-measurable) s.t. E f dµ= E gd‹ ’E œ N; if g Õ is another such function,
then g = g Õ ‹-a.e. (In Probability Theory, g is call the conditional expectation of f on N.)

33
Proof. Let us begin by definings the measure ⁄ s.t. d⁄ = f dµ and its integration is restricted to E œ N;
I.e., ’E sœ N, we have ⁄(E) = E f dµ. To easily see that ⁄ << ‹, note that if ‹(E) = 0 ∆ µ(E) = 0 ∆
⁄(E) = E ddµ = 0. We thus have shown the necessary conditions for us to invoke The Lebesgue-Radon-
Nikodym theorem. Explicitly, the Radon-Nikodym derivative, g = d⁄ d‹ exists and is ‹-integrable where
f dµ = d⁄ = gd‹; I.e.; ⁄ ⁄ ⁄
f dµ = d⁄ = gd‹ ’E œ N
E E E
s s
Finally, if g Õ satisfies E f dµ = E g Õ d‹, then naturally d⁄ = g Õ d‹, and since the Radon-Nikodym deriva-
tive is unique, we must also have g = g Õ ‹-a.e.

3.6 Folland 3.20


Prove the following Proposition:
Proposition. 3.6:

If ‹ is a complex measure on (X, M) and ‹(X) = |‹|(X), then ‹ = |‹|.

Proof. Suppose that d|‹| = f dµ as in the definition of |‹|. Then if E œ M, then we will have:
ú ú
‹(E) + ‹(E c ) = ‹(X) = |‹|(X) = |‹|(E) + |‹|(E c ), where we have = by assumption

And so:
‹(E c ) ≠ |‹|(E c ) = |‹|(E) ≠ ‹(E)
Taking the real part of the LHS, and using (Folland) Proposition 3.13a (|‹(E)| Æ |‹|(E)), we see that:
! "
Re ‹(E c ) ≠ |‹|(E c ) Æ Re(‹(E c ) ≠ |‹(E c )|)
= ‹r (E c ) ≠ |‹(E c )|
Ò
= ‹r (E ) ≠ ‹r2 (E c ) + ‹i2 (E c )
c


Æ ‹r (E ) ≠ ‹r2 (E c ) = 0
c

And similarly for the RHS:


! "
Re |‹|(E) ≠ ‹(E) Ø Re(|‹(E)| ≠ ‹(E))
= |‹(E)| ≠ ‹r (E)
Ò
= ‹r2 (E) + ‹i2 (E) ≠ ‹r (E c )

Æ ‹r2 (E) ≠ ‹r (E) = 0

And so combining the fact that Re(LHS) Æ 0 Æ Re(RHS), but obviously since Re(LHS) = Re(RHS),
we must have that:
Re(|‹|(E) ≠ ‹(E)) = 0 ∆ |‹|(E) = ‹r (E)
But since, again by (Folland) Proposition 3.13a, we have |‹(E)| Æ |‹|(E), we see that this must be true
≈∆ ‹i (E) = 0, and so:

|‹|(E) = ‹r (E) = ‹r (E) + ‹i (E) = ‹(E) ’E œ M (I.e., ‹ = |‹|)

34
3.7 Folland 3.21
Prove the following Proposition:
Proposition. 3.7:

Let ‹ be a complex measure on (X, M). If E œ M, define:


;ÿ
n <
-
µ1 (E) = sup |‹(Ej )| - n œ N, E1 , . . . , En disjoint, E = Ûnj=1 Ej
j=1
;ÿŒ <
-
µ2 (E) = sup |‹(Ej )| - n œ N, E1 , E2 , . . . disjoint, E = ÛŒ
j=1 Ej
j=1
;- ⁄ - <
- --
µ3 (E) = sup -- f d‹ -- - |f | Æ 1
E

Then µ1 = µ2 = µ3 = |‹|. (First show that µ1 Æ µ2 Æ µ3 . To see that µ3 = |‹|, let f = d‹


d|‹| and
apply (Folland) Prop 3.13. To see that µ3 Æ µ1 , approximate f by a simple function.)

) *
Proof.
) We proceed * as in the hint. Trivially, µ1 Æ µ2 since given E œ M, 1 | E = Û1 Ej
{Ej }Œ Œ

{Ej }n1 | E = Ûn1 Ej (set Ej = ’j > n).
To see µ2 Æ |‹| Æ µ3 , let f := d|‹| ,
d‹
and so if E = ی
1 Ej , we have:

5ÿ
Œ 6 ÿŒ
|‹(Ej )| Æ |‹|(Ej ) by (Folland) Prop. 3.13a
j=1 j=1
5 6 ⁄
= |‹|(E) = d|‹|
E

= |f |2 d|‹| by (Folland) Prop. 3.13b
⁄E
= f f d|‹|
⁄E
= f d‹ by (Folland) Prop. 3.9a
E
-⁄ - 5;Ó ⁄ Ô- <6
- - -
Æ -- f d‹ -- œ f d‹ - |f | Æ 1
E E

And so µ2 Æ |‹| Æ µ3 by the steps with square brackets around them.


To show now that µ3 Æ µ1 , let D := {z œ C | |z| Æ 1}. Trivially D is compact, and thus ÷{zj }n1 s.t.
’‘ > 0:
€n
B‘ (zj ) ∏ D
1

Moreover, by definition of supremum, ’‘ > 0, ÷f s.t. |f | Æ 1 and:


-⁄ -
- -
µ3 (E) Æ -- f d‹ -- + ‘
E

If we are assuming |f | Æ 1, then f ≠1 (D) = X, and so we will have X = fin1 f ≠1 (B‘ (zj )) as well. By
defining Bj := f ≠1 (B‘ (zj )), we now perform the standard “shuffle” to make a disjoint sequence {Aj }n1

35
out of {Bj }n1 , namely we let A1 = B1 , and Aj = Bq j \ fii=1 Bi , so that X = Û1 Aj . Now, in following the
j n

hint, we explicitly define the simple function „ := 1 zj ‰Aj .


n

Naturally |„| Æ 1 and |f (x) ≠ „(x)| < ‘. Thus:


-⁄ - -⁄ - -⁄ - -⁄ -
- - - - - - - -
µ3 (E) Æ -- f d‹ -- + ‘ Æ -- f d‹ -- ≠ -- „d‹ -- + -- „d‹ -- + ‘
E
- ⁄E ⁄
E
- -⁄ E -
- - - -
Æ- - f d‹ ≠ „d‹ - + -
- - „d‹ -- + ‘ by the Reverse Triangle Inequality
E E E
-⁄ - -⁄ -
- - - -
Æ -- |f ≠ „|d‹ -- + -- „d‹ -- + ‘
E
⁄ -⁄ -E
- -
Æ ‘d|‹| + - - „d‹ -- + ‘
E E
-⁄ -
- -
Æ ‘|‹|(E) + ‘ + -- „d‹ --
E
s
So, by letting ‘ æ 0, we have µ3 (E) Æ | E „d|‹|. Now, let us define {Ej }n1 by Ej = Aj fl E so that
E = Ûn1 Ej ; thus:
-⁄ - -⁄ ÿ n - -ÿ ⁄ -
- - - - - n -
µ3 (E) Æ -- „d‹ -- = -- zj ‰Aj ‰E d‹ -- = -- zj d‹ --
E j=1 j=1 Ej
-ÿn - ÿ n
- -
= -- ‹(Ej )-- Æ |zj ||‹(Ej )|
j=1 j=1
n
ÿ ; <
-
Æ |‹(Ej )| œ {Ej }1 E = Ûn1 Ej
n -

j=1

And so µ3 Æ µ1 . Thus since we were able to show µ1 Æ µ2 Æ |‹| Æ µ3 Æ µ1 , every inequality above is
actually an equality and in fact: µ1 = µ2 = µ3 = |‹|.

3.8 Folland 3.24


Prove the following Proposition:
Proposition. 3.8:

If f œ L1loc and f is continuous at x, then x is in the Lebesgue set of f .

Proof. To show that x is in the Lebesgue set of f , we need to show that:



1
lim |f (y) ≠ f (x)|dy = 0
r√0 m(Br (x)) B (x)
r

To see this, suppose that ‘ > 0. By the definition of continuity of f at x, we know that ÷” > 0 s.t. if
||x ≠ y|| < ”, I.e., y œ B” (x), we have |f (x) ≠ f (y)| < ‘. We therefore yield the following inequality for
0 < r < ”: ⁄ ⁄
1 1
|f (y) ≠ f (x)| Æ ‘dy = Ar (‘) = ‘
m(Br (x)) Br (x) m(Br (x)) Br (x)
Thus, since ‘ was arbitrary, we may conclude that our limit is indeed = 0; I.e., x is in the Lebesgue set
of f .

36
3.9 Folland 3.25
Prove the following Proposition:
Proposition. 3.9:

If E is a Borel set in Rn , the density, DE (x), of x is defined as:

m(E fl Br (x))
DE (x) = lim
r√0 m(Br (x))

whenever the limit exists.


a) Show that: I
1 for a.e. x œ E
DE (x) =
0 for a.e. x œ E c

b) Find examples of E and x s.t. DE (x) is a given number – œ (0, 1), or such that DE (x) does
not exist.

Proof.

a) Let us begin by defining ‹(A) := m(E fl A) ’A œ BRn . Then, by construction we have ‹ << m and
dm = ‰E . Furthermore, since {Br (x)}r>0 vacuously satisfies the requirements for a set to shrink
d‹

nicely to x œ Rn , we may make use of (Folland) Theorem 3.22. Explicitly, by 3.22, we have:

m(E fl B(r, x)) ‹(Br (x))


DE (x) = lim = lim = ‰E for m ≠ almost everyx œ Rn
r√0 m(B(r, x)) r√0 m(Br (x))

Which is precisely what we wanted to show.


b) For the first example, we are looking for an E and an x s.t. DE (x) = –, where – œ (0, 1).
Suppose we are dealing in BR2 and we set x = (0, 0) and let E = {(x, y) | x = t cos(◊), y =
t sin(◊), t > 0, ◊ œ (0, 2fi–)}. Intuitively, E is the interior of the 2-dimensional, infinitely-extending,
cone whose vertex is the point x, and walls are defined as the positive x-axis and the line starting
at the origin and passing thought the point (cos(2fi–), sin(2fi–)) on the unit circle. Therefore,
E fl Br (x) will be the interior of the same cone as defined before, but now bounded by the curve
beginning at (r, 0), which traverses c.c.w. along Cr (0) and stops at (r cos(2fi–), r sin(2fi–)). With
this geometrical understanding, we can now easily recognize that since m(Br (x)) = 2fir2 , m(E fl
Br (x)) = –m(Br (x)) = –2fir2 . Since our results are true ’r > 0, we have:

m(E fl B(r, x)) –2fir2


DE (x) = lim = lim =–
r√0 m(B(r, x)) r√0 2fir 2

For the second example, we are looking for an E and an x s.t. DE (x) does not exist. Suppose for
this example we turn our thoughts to BR1 (so that Br (x) = (x ≠ r, x + r)). Let us now set x = 0,
and define E as follows:
Π3 4 3 4 3 4 3 4
h 1 1 1 1 1 1 1 1
E= , = , Û , Û , Û ···
n=1
2 2n+1 22n 8 4 32 16 128 64

Our strategy henceforth will be to compose a countable subsequence, rk , s.t. rk √ 0 and where
m(EflB k (x))
limkæŒ m(Br r(x)) will be undefined, therefore also rendering DE (x) undefined. To do this, we

37
set rk = 21k . Naturally, we have m(Brk (x)) = 22k = 2k≠1
1
. We now decompose (with a slight abuse
of notation) {rk }1 = {r2l }1 + {r2l+1 }1 , I.e. separate rk into two subsequences, one where k is
Œ Œ Õ Œ

even, the other when k is odd. For the former (k is even), we have:
ÿ 3 1 1
4 ÿ
1 1 ÿ 1
Π3 43 4
1 2 1
m(E fl Brk (x) ) = m 2n+1 , 2n = = = =
2 2 22n+1 2k n=0 22n+1 2k 3 3 · 2k≠1
nØl nØl

And (quite) similarly for k being odd we have:


3 4 3 43 4
1 1 1 1 ÿ 1 1 2 1
ÿ ÿ Œ
m(E fl Brk (x) ) = m 2n+1 , 2n = = = =
2 2 22n+1 2k+1 n=0 22n+1 2k+1 3 3 · 2k
nØl+1 nØl+1
1
And so, in recalling again that m(Brk ) = 2k≠1 we have:
Y
1/(3·2k≠1 )
m(E fl Brk (x)) ]
1/2k≠1
= 13 if k = 2l, l œ N
= 1/(3·2k )
m(Brk (x)) [ k≠1 =
1
if k = 2l + 1, l œ N
1/2 6

m(EflBrk (x))
And so limkæŒ Brk (x) is undefined, and therefore so too is DE (x) by our previous reasoning.

3.10 Folland 3.26


Prove the following Proposition:
Proposition. 3.10:
If ⁄ and µ are positive, mutually singular Borel measures on Rn and ⁄ + µ is regular, then so are
⁄ and µ.

Proof. If K cpt µ BRn , then ⁄(K), µ(K) < (⁄ + ‹)(K) < Œ. Furthermore, suppose that E, F form the
singular decomposition of ⁄, µ; I.e., Rn = E Û F and ’F1 µ F, F1 œ BRn , µ(F1 ) = 0, and similarly for E
w.r.t. ⁄.
Suppose now that A œ BRn . By definition of ⁄ + µ’s regularity, we know that:
) *
(⁄ + µ)(A) = inf (⁄ + µ)(U open
)|U∏E
Therefore, ’‘ = 2≠k , k œ N, ÷ Uopen s.t. (⁄ + µ)(Uk ) < (⁄ + µ)(A) + ‘. Thus, we may construct a
countable sequence of these such Uk ’s, namely {Uk }Œ
1 , for which when letting k æ Œ, we have:

lim (⁄ + µ)(Uk ) = (⁄ + µ)(A), where (⁄ + µ)(Uk ) Ø (⁄ + µ)(A) ’k œ N by positivity of measures


kæŒ

By our set up of the singular decomposition of ⁄, µ we also note that we may express (⁄ + µ)(Uk ) =
(⁄ + µ)(Uk fl E) + (⁄ + µ)(Uk fl F ) = µ(Uk fl E) + ⁄(Uk fl F ), and similarly for A, namely: (⁄ + µ)(A) =
µ(A fl E) + ⁄(A fl F ). Furthermore, since µ, ⁄ are positive measures, and by construction Uk ∏ A ∆
Uk fl E ∏ A fl E (and similarly Uk fl F ∏ A fl F ), we have µ(Uk fl E) Ø µ(A fl E) and ⁄(Uk fl F ) Ø ⁄(A fl F ).
By applying the last result twice, we can reach the following result:
(⁄ + µ)(Uk ) ≠ (⁄ + µ)(A) = ⁄(Uk fl F ) + µ(Uk fl E) ≠ ⁄(A fl F ) ≠ µ(A fl E)
Ø ⁄(Uk fl F ) ≠ ⁄(A fl F )
= ⁄(Uk fl F ) + ⁄(Uk fl E) ≠⁄(A fl F ) ≠ ⁄(A fl E)
¸ ˚˙ ˝ ¸ ˚˙ ˝
=0 =0
= ⁄(Uk ) ≠ ⁄(A) Ø 0

38
For which we already showed that the LHS has limit = 0, and thus taking limits on every equation in
the above reasoning shows that limkæŒ ⁄(Uk ) = ⁄(A). Furthermore, by the exact same steps but in
swapping ⁄ ¡ µ, we see that limkæŒ µ(Uk ) = µ(A) as well. Therefore, the same approximation by open
sets from above for the definition of (⁄ + µ)’s regularity also works as an approximation by open sets
from above for all sets A œ BRn for ⁄ and µ, hence we have arrived at the definition of ⁄ and µ being
regular measures.

4 Chapter 5
4.1 Folland 5.1
Prove the following Proposition:
Proposition. 4.1:

If X is a normed vector space over K (= R or C), then addition and scalar multiplication are
continuous
- from -X ◊ X and K ◊ X to X. Moreover, the norm is continuous from X to [0, Œ); in
fact, -||x|| ≠ ||y||- Æ ||x ≠ y||.

Proof. Let us define A : X ◊ X æ X as the addition map (I.e., defined as A(x, y) = x + y). The by
construction, A is a linear map from the NVS X ◊ X to the NVS X. We thus will have (for (x, y) œ X ◊ X):

||A(x, y)|| = ||x + y|| Æ ||x|| + ||y|| Æ 2 max{||x||, ||y||} = 2||(x, y)||

And therefore by (Folland) Proposition 5.2, since the above shows A is bounded, A must also be contin-
uous.
Let now define M : X ◊ X æ X as the scalar multiplication map (I.e., defined as M (–, x) = –x). Suppose
now that ‘ > 0 and we choose ” = min{1, ‘}. Then if (–, x) œ K ◊ X such that ||(–, x)|| < ”, we have:

max{|–|, ||x||} < ” Æ ‘ ∆ ||M (–, x)|| = ||–x|| = |–|||x|| < ” 2 Æ ” Æ ‘

And so is continuous at (0, 0), and hence again by (Folland) Proposition 5.2, M is continuous.
Lastly, let again ‘ > 0, but now set ” = ‘. If x, y œ X such that ||x ≠ y|| < ”, then:

||x|| = ||x ≠ y + y|| Æ ||x ≠ y|| + ||y|| ∆ ||x|| ≠ ||y|| Æ ||x ≠ y||

And similarly for ||y|| ≠ ||x|| Æ ||y ≠ x|| = ||x ≠ y||. Therefore:
- -
-||x|| ≠ ||y||- Æ ||x ≠ y|| < ” = ‘

And so || · || is uniformly continuous and therefore continuous from X to [0, Œ).

4.2 Folland 5.2


Prove the following Proposition:
Proposition. 4.2:

L(X, Y) is a vector space and the function || · || defined by (Folland, Equation 5.3) is a norm on it.
In particular, the three expressions on the right of (5.3) are always equal.

39
Proof. We begin by defining:
) *
||T ||1 := sup ||T x|| | ||x|| = 1
; <
||T x||
||T ||2 := sup | x ”= 0
||x||
) *
||T ||3 := sup C | ||T x|| Æ C||x|| ’x œ X

As
- in (Folland)
- Equation 5.3. We thus begin by showing ||·||1 = ||·||2 = ||·||3 . Firstly, if x œ X, x ”= 0, then
-x/||x||- = 1 and so T (x)/||x|| = T (x/||x||) Æ || · ||1 . Since this is true ’x œ X, we may take the supremum
and hence ||T ||2 Æ ||T ||1 . Next, again if x œ X and ||x|| = 1, then ||T x|| Æ ||T ||3 , and again taking the
supremum implies ||T ||1 Æ ||T ||3 . Lastly, again supposing x œ X, we simply have ||T x|| Æ ||T ||2 , therefore
||T ||3 Æ ||T ||2 . Summarizing we have: ||T ||1 Æ ||T ||3 Æ ||T ||2 Æ ||T ||1 , and hence all our inequalities
above are actually equalities, and proving the equivalence of the above forms of (5.3).
To prove that || · || does indeed define a norm, suppose S, T œ L(X, Y), and x œ X. We thus have:

||(S + T )x|| = ||Sx + T x|| Æ ||Sx|| + ||T x|| Æ (||S|| + ||T ||)||x|| ∆ ||S + T || Æ ||S|| + ||T ||

If now – œ K(= C or R), then:

||–T x|| = |–|||T x|| Æ |–||T ||||x|| ∆ ||–T || Æ |–|||T ||


∆||–≠1 (–T )|| Æ |–≠1 |||–T || ∆ |–|||T || Æ ||–T ||
∆||–T || = |–|||T ||

And finally, ||T || = 0 ≈∆ ||T x|| = 0 ’x œ X and T © 0. Hence we’ve shown all the conditions for || · ||
to be a norm.

4.3 Folland 5.5


Prove the following Proposition:
Proposition. 4.3:
If X is a normed vector space, the closure of any subspace of X is a subspace.

Proof. Let X be a subspace of X and X denote its closure. Firstly, by definition, 0 œ X. The other
property that we need to show is that if that if x, y œ X, and a, b œ K, then ax + by œ X as well. Since
x, y œ X, we know that ÷ {xj }Œ 1 µ X and {yj }1 µ X s.t. xn æ x and yn æ y with respect to the norm,
n

|| · || on X. So, ’‘/2 > 0, ÷N1 , N2 œ N s.t. ||xn ≠ x|| < ‘/2 and ||yn ≠ y|| < ‘/2 ’n Ø N1 , N2 respectively.
So, ’n Ø N = max(N1 , N2 ), we have:
-- -- 1‘2
--(axn + byn ) ≠ (ax + by)-- Æ ||axn ≠ ax|| + ||byn ≠ by|| = |a|||xn ≠ x|| + |b|||yn ≠ y|| < 2 =‘
2
And so since axn + byn æ ax + by, and axn + byn œ X it implies ax + by œ X by the definition of X.
Therefore, X is indeed a subspace of X.

4.4 Folland 5.6


Prove the following Proposition:

40
Proposition. 4.4:
Suppose
qn that Xqis a finite-dimensional vector space. Let e1 , . . . , en be a basis for X and define
|| 1 aj ej ||1 = 1 |aj |.
n

a) || · ||1 is a norm on X.
qn
b) The map (a1 , . . . , an ) æ 1 aj ej is a continuous form K n with the usual Euclidean topology
to X with the topology defined by || · ||1 .
c) {x œ X | ||x||1 = 1} is compact in the topology defined by || · ||1 .
d) All norms on X are equivalent. (Compare any norm to || · ||1 .)

Proof.
qn
a) We can first see that ||x||1 = 0 ≈∆ x = 0 since 1 |aj | = 0 ≈∆ aj = 0 ’j = 1, . . . , n, and
0 := 0e1 + · · · + 0en .

Next, to see the triangle inequality,


qwe first noteq that the triangle inequality naturally holds ’x, y œ
K. Therefore, if x, y œ X ∆ x = 1 –j ej , y = 1 —j ej , and hence:
n n

-- ÿ -- -- ÿ --
-- n n
ÿ -- -- n -- n
ÿ
||x + y||1 = ---- –j ej + —j ej ---- = ---- (–j + —j )ej ---- = |–j + —j |
j=1 j=1 1 j=1 1 j=1
n
ÿ n
ÿ
Æ |–j | + |—j | = ||x||1 + ||y||1
j=1 j=1

So the triangle inequality holds. Now suppose ⁄ œ K, we therefore have:


-- ÿ -- -- ÿ --
-- n -- -- n -- n
ÿ n
ÿ
||⁄x||1 = ----⁄ –j ej ---- = ---- (⁄–j )ej ---- = |⁄–j | = |⁄| |–j | = |⁄|||x||1
j=1 1 j=1 1 j=1 j=1

And hence we have shown the three conditions for || · ||1 to be a norm on X.
qn
b) q
From Part a), by dropping the absolute values in expressions of the form 1 |aj |, and replacing it by
1 aj ej , the one inequality
qn now becomes an equality, and hence the rest proves that T : K æ X,
n n

where T (a1 , · · · , an ) = 1 aj ej , is a linear map. We may now invoke (Folland) Proposition 5.2,
which states T is continuous ≈∆ T is continuous at 0.

Let ‘ > 0, and ” = ‘/n. Then if:

||x ≠ 0|| = ||x|| = (a21 + · · · a2n )1/2 < ” ∆ a2i Æ (a21 + · · · + a2n ) < ” 2 ’i = 1, . . . , n

And so |ai | < |”| = ‘/n. Therefore, we have:


-- ÿ -- 3 4
-- n -- ‘
||T x||1 = ---- aj ej ---- = |a1 | + · · · + |an | < n =‘
j=1 1 n

qn
c) We begin by showing := {(a1 , . . . , an ) œ K n | 1 |aj | = 1} µ K is compact. To see this, we
n

can simply show that is closed and boundedq since µ K n = Cn or Rn . The boundness of is
easy to see since: ||(a1 , . . . , an )||2 = ||x||2 := ( 1 a2j )1/2 ∆ |aj | Æ 1 ’j = 1, . . . , n ∆ B2 (0) ∏ ,
n

hence is bounded.

41
To see is closed, we show that c is open. If x œ c
, then:
; n
ÿ <
x œ (a1 , . . . , an ) | |aj | =
” 1
1
; n
ÿ < ; n
ÿ <
© (a1 , . . . , an ) | |aj | < 1 Û (a1 , . . . , an ) | |aj | > 1 := 1 Û 2
1 1
qn qn
I.e., if x q
= (x1 , . . . , xn ), then 1 |xj | < 1 or 1 |xj | > 1. Assume x œ 1, and y œ K n . Letting
‘1 = 1 ≠ 1 |xj | > 0, then in taking ”1 = ‘1 /n, we have:
n

‘1
||x ≠ y||2 < ”1 ∆ |xi ≠ yi | Æ ||x ≠ y||2 < ”1 = ’i œ {1, . . . , n}
n
ÿn 1‘2 ÿ n
∆ ||x ≠ y||1 = |xj ≠ yj | < n =1≠ |xj |
j=1
n j=1
n
ÿ n
ÿ n
ÿ
∆ |xj | ≠ |yj | < 1 ≠ |xj | since |a| ≠ |b| < |b ≠ a|
j=1 j=1 j=1
ÿn
∆ |yj | < 1
j=1
qn
And so B‘1 /n (x) µ 1 , so 1 is open. Now suppose x œ 2. Letting ‘2 = 1 |xj | ≠ 1 and y œ K n
as before. Then letting ”2 = ‘2 /2, we have:
‘2
||x ≠ y||2 < ”2 ∆ |xi ≠ yi | Æ ||x ≠ y||2 < ”2 = ’i œ {1, . . . , n}
n
ÿn 1‘2 ÿ n
∆ ||x ≠ y||1 = |xj ≠ yj | < n = |xj | ≠ 1
j=1
n j=1
n
ÿ n
ÿ n
ÿ
∆ |xj | ≠ |yj | < |xj | ≠ 1 since |b| ≠ |a| < |b ≠ a|
j=1 j=1 j=1
ÿn
∆ |yj | > 1
j=1

And so B‘2 /n (x) µ 2 , and hence 2 is open. Now since c = 1 Û 2 , we can now conclude that
c
is open, and hence is closed, and hence compact. Furthermore, in Part b), we showed that T
(as defined in Part b) is continuous. Therefore, since:
T ( ) = {x œ X | ||x||1 = 1}
We may now conclude that since is compact, so too is {x œ X | ||x||1 = 1} in the topology defined
by || · ||1 .
d) Suppose || · || : ‰ æ RØ0 is an arbitrary norm on X. We recall that to show || · || and || · ||1 are
equivalent, we need to find C1 , C2 > 0 s.t. C1 ||x||1 Æ ||x|| Æ C2 ||x||1 ’x œ X. If x = 0, then
||x||1 = ||x|| since both are norms, selecting any C1 Æ C2 where C1 , C2 > 0 proves the equivalence
of these norms for x = 0; therefore, assume x ”= 0.
qn
If we let C2 = max({||ej ||n1 ), then if x œ X ∆ x = 1 xj ej , then:
n
ÿ n
ÿ
ú ú
||x|| Æ |xj |||ej || Æ C2 = C2 ||x||1 where we have Æ from the -inequality
j=1 j=1

42
So we have found an appropriate C2 .
We now claim that || · || is continuous in the topology defined by || · ||1 . To see this, let ‘ > 0, and
” = ‘/n. If x, y œ X and ||x ≠ y||1 < ”, then by what we found above:
3 4

||x ≠ y|| Æ C2 ||x ≠ y||1 < C2 =‘
C2

Which tells us that || · || is indeed continuous on X in the topology defined by || · ||1 .


By Part c), we recall that A := {x œ X | ||x||1 = 1} is a compact set in the topology defined by ||·||1 .
Therefore, by the continuity of || · ||, and since we are assuming x ”= 0, we know that minxœA ||x||
exists, so let’s call this min C1 . Explicitly now:
-- --
-- x --
C1 Æ ---- -- ∆ C1 ||x||1 Æ ||x|| ’x œ X
||x||1 --

Hence completing our proof since we found both C1 , C2 which satisfy the necessary inequality.

4.5 Folland 5.9


Prove the following Proposition:
Proposition. 4.5:

Let C k ([0, 1]) be the space of functions on [0, 1] possessing continuous derivatives up to order k
on [0, 1], including one-sided derivatives at the endpoints.

a) If f œ C([0, 1]), then f œ C k ([0, 1]) ≈∆ f is k times continuously differentiable on (0, 1)


and limx√n f (j) (x) and limx¬1 f (j) (x) exist for j Æ k. (The mean value theorem is useful.)
qk
b) ||f || = 0 ||f (j) ||u is a norm on C k ([0, 1]) that makes C k ([0, 1]) into a Banach space. (Use
induction on k. The essential point is that if {fn } µ C 1 ([0, 1]), fn æ f uniformly, and
1
fnÕ æ g uniformly, thens x f œ C ([0, 1]) and f = g. The easy way to prove this is to show
Õ

that f (x) ≠ f (0) = 0 g(t)dt.)

Proof.

a) We’ll proceed to prove this claim through induction. Suppose k = 0, then the forward case of
f œ C([0, 1]) implying f is differentiable on (0, 1) and and limx√n f (x) and limx¬1 f (x) existing is
by the definition of C([0, 1]).

Now, for the backward direction (k = 0), suppose f œ C((0, 1)), limx√n f (x), and limx¬1 f (x) exist
- this, however, is simply the definition of f œ C([0, 1]).
(j) (j) (j)
Let L0 := limx√n f (x) and L1 := limx¬n f (j) (x). Now assume the property above holds for
k = n ≠ 1. The forward direction is simply by definition. For the backward direction, if we wish to
(j)
show that f being k times differentiable on (0, 1) and limx√n f (x) and limx¬n f (j) (x) existing for
j Æ n implies f œ C k ([0, 1]), we may proceed as follows. Firstly, by the existence of the one sided
(j)
derivatives, we know that ’‘ > 0, ÷” > 0 such that if 0 < x < ”, then |f (j) (x) ≠ L0 | < ‘, ’j Æ n.
(j)
Furthermore, WLOG, we may omit the L1 case since all we need to chance in the argument is

43
that ” < x < 1 instead of 0 < x < ”. Moreover, by the mean value theorem, ÷x̂ œ (0, ”) s.t.
f (j≠1) (x) ≠ f (j≠1) (0) = (x ≠ 0)f (j) (x̂) = xf (j) (x̂) Therefore:
- (j≠1) (j≠1)
-
-f (x) ≠ f (x) -
-
- ≠ L0 -- = |f (j) (x̂) ≠ L1 | < ‘ since x̂ œ (0, ”)
x

f (j≠1) (x)≠f (j≠1) (x)


And so limx√0 x = L0 ’j Æ n, and by the exact same argument for 1, we see that
(j≠1) (j≠1)
(x)≠f (x)
limx¬0 f x = L1 ’j Æ n, and so f œ C (n) ([0, 1]), completing our inductive step and
proving this proposition ’k œ N.
1 0
b) We proceed, as hinted, by induction. Suppose that {fn }Œ 1 is Cauchy in C . Then fn æ f in C
and fnÕ æ g in C. Therefore: ⁄ x
fn (x) = fn (0) + fnÕ (y)dy
0

However, since fnÕ æ g, by the dominated convergence theorem we have:


⁄ x ⁄ x
f (x) = lim f (x) = lim fn (0) + lim fnÕ (y)dy = f (0) + g(y)dy
næŒ næŒ næŒ 0 0

And therefore by the fundamental theorem of calculus we may conclude that g = f Õ , and so fn æ f
in C 1 .

We now make our inductive step. Assume the statement is true up until j = k. Suppose then that
(k+1) (k) (k+1)
1 is Cauchy in C
{fn }Œ and fn æ f in C k and fn æ g in C k+1 . Therefore, fn æ fn
k+1
(k+1) (k+1) (k+1)
in C, and fn æ g in C. We therefore may conclude that fn æ fn in C, and ultimately
(k+1)
fn æ f in C .

To finish our proof, we need to prove that || · || is indeed a norm. Firstly, if f ”© 0, then naturally
||f || =
” 0, so ||f || = 0 ≈∆ f © 0. Since || · || is simply a sum of other norms, the triangle inequality
and absolutely scalability are both trivially immediate like definiteness.

5 Chapter 6
5.1 Folland 6.3
Prove the following Proposition:
Proposition. 5.1:

If 1 Æ p < r Æ Œ, Lp fl Lr is a Banach space with norm ||f || = ||f ||p + ||f ||r , and if p < q < r, the
inclusion map Lp fl Lr æ Lq is continuous.

Proof. We begin by first showing that Lp fl Lr is a Banach Space w.r.t. ||f || = ||f ||p + ||f ||r (I.e., show
Lp fl Lr a normed vector space and complete w.r.t. ||f ||).
The fact that || · ||r and || · ||p are norms implies || · || is a norm. Firstly, || · || Ø 0 since || · ||p , || · ||r Ø 0.
Now, suppose f, g œ Lr fl Lp , and ⁄ œ K, then we have:

||f + g|| = ||f + g||p + ||f + g||r Æ ||f ||p + ||g||p + ||f ||r + ||g||r = ||f || + ||g||

44
||⁄f || = ||⁄f ||p + ||⁄f ||r = |⁄| ||f ||p + |⁄| ||f ||r = |⁄| ||f ||
||f || = 0 ≈∆ ||f ||p = ||f ||r = 0 ≈∆ f © 0 µ-a.e.
We can also immediately see that Lp fl Lr is a vector space since if u, v œ Lp fl Lr , then u, v œ Lp and
Lr , and so all our conditions for being a vector subspace are satisfied since both Lp and Lr are vector
subspaces.
Suppose now that {fn }Œ 1 be a Cauchy sequence in L fl L . By noting that ’n, m œ N, we have
p r

||fn ≠ fm ||p Æ ||fn ≠ fm || and ||fn ≠ fm ||r Æ ||fn ≠ fm ||, and hence {fn }Œ
1 are also Cauchy in L and L .
p r

We can thus define g and h as lim fn in Lp and Lr respectively. Let ‘ > 0, then ÷ N œ N s.t. if we take
” = ‘(p+1)/p , then letting ||fn ≠ g||p < ”, and in setting E := {x œ X | ‘ Æ |fn (x) ≠ g(x)|}, we have:
⁄ ⁄ ⁄
1 1 1 1! "p 1 ! "p
µ(E) = p p
‘ dµ Æ p p
|fn ≠ g| dµ Æ p |fn ≠ g| dµ = p ||fn ≠ g||p < p ” = ‘
p
‘ E ‘ E ‘ ‘ ‘

I.e., µ(E) < ‘ ∆ {fn }Œ


1 converges in measure to g. If r < Œ, the argument holds for interchanging p
for r. If r = Œ, then ÷ a subsequence fnk of {fn }Œ
1 s.t. fnk æ h µ-a.e. We have therefore shown that
g = h, and so g œ L fl L . Therefore, since fn æ g in Lp and Lr , we have fn æ g in Lp fl Lr - and hence
p r

Lp fl Lr is a Banach space with norm || · ||.


Let now p < q < r. By (Folland) Proposition 6.10, we know that ÷⁄ œ (0, 1) s.t.||f ||⁄p ||f ||1≠⁄
r where
1 1≠⁄
q = p + r . Thus, since ||f ||p Æ ||f || and ||f ||r Æ ||f ||, we have:

||f ||q Æ ||f ||⁄p ||f ||1≠⁄


r Æ ||f ||⁄ ||f ||1≠⁄ = ||f ||

Suppose now that ‘ > 0 and f, g œ Lp fl Lr , then if ||f ≠ g|| < ” = ‘, we have ||f ≠ g||q Æ ||f ≠ g|| < ‘ by
the above inequality. Hence ÿ : Lp fl Lr æ Lq is uniformly continuous (and naturally continuous as well).

5.2 Folland 6.4


Prove the following Proposition:
Proposition. 5.2:

If 1 Æ p < r Æ Œ, Lp + Lr is a Banach Space with norm ||f || = inf{||g||p + ||h||r | f = g + h},


and if p < q < r, the inclusion map Lq æ Lp + Lr is continuous.

Proof. We begin by showing || · ||, as defined, is a norm. Firstly, || · || Ø 0 since || · ||p , || · ||r Ø 0. Now,
suppose f1 , f2 œ Lr + Lp , and ⁄ œ K, then we have:
Ó - Ô
-
||f1 + f2 || = inf ||g||p + ||h||r - f1 + f2 = g + h
Ó - Ô
-
= inf ||g1 + g2 ||p + ||h1 + h2 ||r - f1 + f2 = g + h = (g1 + g2 ) + (h1 + h2 )
Ó! " ! " -- Ô
Æ inf ||g1 ||p + ||g2 ||p + ||h1 ||r + ||h2 ||r - f1 + f2 = g + h = (g1 + g2 ) + (h1 + h2 )
Ó - Ô Ó - Ô
- -
Æ inf ||g1 ||p + ||h1 ||r - f1 = g1 + h1 + inf ||g2 ||p + ||h2 ||r - f2 = g2 + h2
= ||f1 || + ||f2 ||

45
Ó - Ô
-
||⁄f || = inf ||⁄g||p + ||⁄h||r - ⁄f = ⁄(g + h)
Ó - Ô
-
= inf |⁄| ||g||p + |⁄| ||h||r - ⁄f = ⁄(g + h)
Ó - Ô
-
= |⁄| inf ||g||p + ||h||r - f = g + h
= |⁄| ||f ||
||f || = 0 ≈∆ ||f ||p = ||f ||r = 0 ’g, h s.t. f = g + h ≈∆ f © 0 µ-a.e.
We can also immediately see that Lp + Lr is a vector space since if u, v œ Lp + Lr , then u = u1 + u2 , v =
v1 + v2 where u1 , v1 œ Lp and u2 , v2 œ Lr , and so all our conditions for being a vector subspace are
satisfied since both Lp and Lr are vector subspaces.
To show completeness, we make use of (Folland) Theorem 5.1 which states that a q normed vector space, X,
is complete ≈∆ every absolutely convergent series in X converges. So, suppose 1 fn be an absolutely
Œ

convergent series in Lp + Lr . By the definition of inf and || · ||, we know that ’n œ N, ÷ gn œ Lp , hn œ Lr


s.t.
qŒfn = gn q + hn where ||gn ||p + ||hn ||r < ||fn || + 2≠n . Therefore,
qΠfrom this
qΠinequality, and since both
1 fn and 2 are absolutely convergent, so too will 1 gn and 1 hn . Since Lp and Lr are
Œ ≠n
1q qN
Banach spaces, 1 gn æ g œ Lp and 1 hn æ h œ Lr . Furthermore, byqdefinition || · || Æ || · ||p and
N

||·|| Æ ||·||r , so combining these two reverse inequalities, we have 1 fn = 1 (gn +hn ), which therefore
Œ

has a limit in Lp + Lr , explicitly g + h œ Lp + Lr . We have thus show all the necessary conditions for
Lp + Lr to be a Banach Space w.r.t. || · ||.
Suppose p < q < r and f œ Lq . Let E := {x œ X | 1 < |f (x)|}. Thus, by the construction of E, we
therefore have: |f ‰E |p Æ |f ‰E |q and |f ‰E c |p Æ |f ‰E c |q (I.e., f ‰E œ Lp , f ‰E œ Lr ), and hence:
||f || = ||f ‰E + f ‰E c || Æ ||f ‰E ||p + ||f ‰E ||r Æ ||f ‰E ||q + ||f ‰E c ||q = ||f ||q
Suppose now that ‘ > 0 and f, g œ Lq , then if ||f ≠ g||q < ” = ‘, we have ||f ≠ g|| Æ ||f ≠ g||q < ‘ by the
above inequality. Hence ÿ : Lq æ Lp + Lr is uniformly continuous (and naturally continuous as well).

5.3 Folland 6.5


Prove the following Proposition:
Proposition. 5.3:
Suppose 0 < p < q < Œ. Then:
a) Lp ”µ Lq ≈∆ X contains sets of arbitrarily small positive measure.
b) Lq ”µ Lp ≈∆ X contains sets of arbitrarily large finite measure.

c) What about the case of q = Œ?

Proof.
a) We first prove the following Lemma:
Lemma. 5.1: Chebyshev’s Inequality
3 4p
||f ||p
µ(Et ) Æ
t
Where Et = {x œ X | |f (x)| Ø t} and p œ (0, Œ).

46
Proof. Let g(x) = xp if x Ø t, and 0 otherwise. We thus have 0 Æ tp ‰Et Æ |f |p ‰Et , and hence:
⁄ ⁄ ! "p
1 1 ||f ||
µ(Et ) = p
p
tp ‰Et dµ Æ p |f |p dµ Æ
t t Et tp

Now back to the problem at hand. For the forward direction, we proceed via the contrapositive,
I.e., suppose ÷‘ > 0 s.t. ’E µ M(X), µ(E) ”œ (0, ‘). From Chebyshev’s Inequality, we know that
! ||f ||p "p
÷T s.t. ’t Ø T , µ(Et ) = 0 since µ(Et ) Æ t æ 0, and so |f | Æ T a.e. So:
⁄ ⁄ ⁄ ⁄
|f |q dµ = |f |q dµ + |f |q dµ Æ T q µ(Et ) + |f |p dµ < Œ
Et Etc Etc

And so f œ Lq .

For the converse, suppose ’‘ > 0, ÷E œ M(X) s.t. µ(E) œ (0, ‘). Let us define {Fn }Œ
1 where
0 < µ(Fn ) < 1/n so that µ(Fn ) æ 0. By defining Gn := Fn \ fiŒ n+1 Fm , we must have 0 <
µ(Fn ) Æ µ(fiŒn Gm ). Furthermore, by taking subsequences, we may actually assume now that
0 < µ(Gm ) Æ 2≠m . Now if we define:
Œ
ÿ ! "≠1/q
f := µ(Gn ) ‰Gn n≠2/p (Ø 0)
n=1

Then we have:
⁄ ⁄ ⁄ ÿ
1 fi2
Œ ÿŒ
! "≠p/q
|f |p dµ = f p dµ = µ(Gn ) ‰Gn n dµ =
≠2
= <Œ
n=1 n=1
n 2 6

And so f œ Lp ; however, one can see that f ”œ Lq since:


⁄ ⁄ ÿŒ Œ
! "1≠p/q ≠2q/p ÿ
|f | dµ = f dµ =
q q
µ(Gn ) n Ø 2p/q≠1 n≠2q/p = Œ
n=1 n=1

b) For the forward direction, the proof here is completely analogous to that in a). For the converse,
! "≠1/q ! "≠1/(p+1)
by substituting µ(Gn ) for µ(Gn ) , and noting that now we have 2m Æ µ(Gm ) < Œ
instead of 0 < µ(Gm ) Æ 2≠m , the same results as in a) still hold.
c) For the case of q = Œ, we have LŒ ”µ Lp ≈∆ µ(X) = Œ, since if |f |p < C œ RØ0 , we have:
⁄ ⁄
|f |p dµ Æ C dµ Æ Cµ(X) < Œ

5.4 Folland 6.7


Prove the following Proposition:
Proposition. 5.4:

If f œ Lp fl LŒ for some p < Œ, so that f œ Lq ’q > p, then ||f ||Œ = limqæŒ ||f ||q .

47
Proof. We may first assume f ”© 0 a.e. by the triviality of this case. From (the proof of Folland)
Proposition 6.10, we know that:
! "1≠p/q ! "p/q
||f ||q Æ ||f ||Œ ||f ||p
And so: 1! "1≠p/q ! "p/q 2
lim sup ||f ||q Æ lim sup ||f ||Œ ||f ||p = ||f ||Œ
qæŒ qæŒ
Furthermore, by our initial assumption, we have ||f ||Π> 0. Suppose now that 0 < a < ||f ||Πand
Ea := {x œ X | |f (x)| Ø a}. We thus have:

! " p ! "q ! q "1/q 1! "q 21/q
a µ(Ea ) Æ ||f ||p Æ
q q
|f | dµ Æ ||f ||q ∆ a µ(Ea ) Æ ||f ||q
Ea
! "1/q
∆ lim inf a µ(Ea ) Æ lim inf ||f ||q
qæŒ qæŒ

∆ a Æ lim inf ||f ||q


qæŒ

And so letting a æ ||f ||Œ , we thus have:


lim sup ||f ||q Æ ||f ||Œ Æ lim inf ||f ||q
qæŒ qæŒ

And so we must have all our inequalities become equalities: hence limqæŒ ||f ||q = ||f ||Œ .

5.5 Folland 6.10


Prove the following Proposition:
Proposition. 5.5:

Suppose 1 Æ p < Œ. If fn , f œ Lp and fn æ f a.e., then ||fn ≠ f ||p æ 0 ≈∆ ||fn ||p æ ||f ||p .
[Use Exercise 20 in (Folland) 2.3.]

Proof. For the forward direction, if ||fn ≠ f ||p æ 0, by the triangle inequality we have:
- -
-||fn ||p ≠ ||f ||p - Æ ||fn ≠ f ||p æ 0
And we therefore have ||fn ||p æ ||f ||p .
For the converse, suppose ||fn ||p æ ||f ||p . We now quickly prove the following result:
If z, w œ C, then |z ≠ w|p Æ 2p≠1 (|z|p + |w|p ) ’p Ø 1
By the second derivative test, g(z) = |z|p is convex (I.e., g(tz + (1 ≠ t)w) Æ tg(z) + (1 ≠ t)g(w)). So, if
we set t = 1/2, and move the 2p over to the other side, we have:
- -p
- z ≠ w -- 1 p 1 p
|z ≠ w|p Æ 2p≠1 (|z|p + |w|p ) ≈∆ -- Æ |z| + |w|
2 - 2 2
For which the latter is recognizably true due to the convexity of | · |p for p Ø 1 (and in making a change
of variables wÕ = ≠w)
Carrying on, let us define gn := 2p≠1 (|f |p + |fn |p ) ≠ |f ≠ fn |p . By the above inequality, we know that
gn Ø 0, and so we may apply Fatou’s Lemma:
⁄ ⁄
! "p ! "p
2 ||f ||p Æ lim inf gn = 2 ||f ||p ≠ lim sup |f ≠ fn |p dµ
p p
næŒ næŒ
s
And so lim sup |f ≠ fn |p dµ Æ 0 ∆ ||f ≠ fn ||p æ 0.

48
5.6 Folland 6.14
Prove the following Proposition:
Proposition. 5.6:
If g œ LŒ , the operator T defined by T f = f g is bounded on Lp for 1 Æ p Æ Œ. Its operator
norm is at most ||g||Œ with equality if µ is semi-finite.

Proof. Firstly, we may assume g ”© 0 due to the triviality of this case. We now proceed to see that:
⁄ ⁄
! "p ! " p ! "p ! "p # $
||T f ||p = |f g|p dµ = |f |p |g|p dµ Æ ||g||Œ |f |p dµ = ||g|Œ | ||f ||p Æ since |g| Æ ||g||Œ

∆ ||T || Æ ||g||Œ
To see equality if µ is semi-finite, suppose 0 < ‘ < ||g||Œ , By µ’s semi-finitness, ÷ E s.t. ||g||Œ ≠ ‘ <
|g| ’x œ E. Thus, we have:
! "
||T ‰E ||p = ||g‰E || > ||g||Œ ≠ ‘ ||‰E ||p ∆ ||T || > ||g||Œ ≠ ‘ ∆ ||T || Ø ||g||Œ

Where we have the last implication by ‘’s arbitrarily, and to satisfy both equalities, we must have
||g||Π= ||T ||.

49

You might also like