LectNotesAdvAnal
LectNotesAdvAnal
GEORGIOS PSARADAKIS
05/07/2024
Contents
1 Prerequisites 7
1.1 Subsets of the Euclidean space . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.1 Abstract sets and sequences of abstract sets . . . . . . . . . . . . . . 7
1.1.2 Subsets of the real line . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.3 Subsets of the n-dimensional Euclidean space . . . . . . . . . . . . . 9
1.2 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3 Volume of unit ball in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2 Measure 17
2.1 Measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Measurable sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3 Regularity of measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 Metric measures and Carathéodory’s criterion . . . . . . . . . . . . . . . . . 29
4 Integral 39
4.1 Measurable functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2 Lusin and Egoroff theorems . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3 Integration of measurable functions . . . . . . . . . . . . . . . . . . . . . . 45
4.4 L p spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3
4 CONTENTS
9 Elementary convexity 83
9.1 Convex functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
9.2 Lipschitz functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
9.3 A characterization of convex functions . . . . . . . . . . . . . . . . . . . . . 91
Preface
This set of notes grew from the graduate course “Advanced Analysis” I taught at the Mathe-
matics Institute of the University of Mannheim (Spring 2017, 2018, 2019 and Fall 2020).
The aim of the course was to present the analytic techniques required to prove various
functional inequalities. The lectures started with basic knowledge of real analysis (mea-
sure and integral) and then it went into some advanced topics in analysis such as L p spaces,
symmetric-decreasing rearrangement of functions, the Fourier transform, distributions and
Sobolev spaces. In parallel, the proofs of inequalities such as Riesz’s rearrangement in-
equality, Young’s inequality, Hardy-Littlewood-Sobolev inequality and logarithmic Sobolev
inequality, were given. The skeleton of the course was the first half of the second edition of
Analysis by E. H. Lieb and M. Loss. The first chapter of this book includes a very brief intro-
duction to measure theory. For instance, outer/inner regularity of the Lebesgue measure and
Lusin’s theorem are included in the exercises of the first chapter. Having taught only some
basic facts about the Lebesgue integral (sometimes without proofs), students found difficult
to study the book itself and were restricted to the classroom notes.
The aim of these notes is to quickly cover possible measure theoretic gaps of the students
knowledge so that to appreciate the remarkable book by Lieb and Loss. In order not to deviate
from the proving-inequalities-purpose of the course, we present: (i) a special proof of the
Euclidean isoperimetric inequality due to H. Hadwiger and D. Ohmann, that essentially uses
only the outer/inner regularity of Lebesgue measure, (ii) an elementary version of the coarea
formula which together with the isoperimetric inequality leads to the L1 -Sobolev inequality.
I start with a section of material students are expected to know from undergraduate anal-
ysis. The main part of sections 2, 4 and 5 is a more explanatory presentation of particular
subsections from the first two chapters of Measure theory and fine properties of functions by
L. C. Evans and R. F. Gariepy. For the presentation and proof of the isoperimetric inequality
through the Brunn-Minkowski inequality of section 3, I followed the book Geometric mea-
sure theory by H. Federer. The coarea formula of section 6 is from Sobolev spaces by V. G.
Maz’ya. Section 7 is a brief account on convex and Lipschitzian functions. For these notes
and throughout the course I have consulted several times the books found at the end of these
notes.
I want to thank Veniamin Gvozdik and Paul Nikolaev for spotting several misprints. Paul
5
6 CONTENTS
has provided me with the nice proof of Lemma 9.2.8. Any further misprints or comments are
very welcome at
psaradakis at outlook dot com
Chapter 1
Prerequisites
7
8 CHAPTER 1. PREREQUISITES
n x ≤ M ∀x ∈ A, (upper bound)
(i) R ∋ M = sup A ⇔
∀ ε > 0, ∃ xε ∈ A such that xε > M − ε. (least one)
Taking ε = 1/k, k ∈ N, we obtain a sequence {xk ∈ A}k∈N such that xk → M as k → ∞.
n x ≥ m ∀x ∈ A, (lower bound)
(ii) R ∋ m = inf A ⇔
∀ ε > 0, ∃ xε ∈ A such that xε < m + ε. (greatest one)
Taking ε = 1/k, k ∈ N, we obtain a sequence {xk ∈ A}k∈N such that xk → m as k → ∞.
Remark 1.1.1. Initially we will refer to this characterization when we use it. After some
point we will use it without any comment.
When finite, lim supk→∞ ak and lim infk→∞ ak correspond respectively to the largest and small-
est subsequential limit points of {ak }k∈N . Thus,
Exercise 1.1.3. Suppose there exists K ∈ N such that ak ≥ bk for all k ∈ N with k ≥ K. Prove
lim infk→∞ ak ≥ lim infk→∞ bk .
1.1. SUBSETS OF THE EUCLIDEAN SPACE 9
Definition 1.1.4. A point x ∈ Rn is called limit point of the sequence {xk ∈ Rn }k∈N , whenever
|xk − x| → 0 as k → ∞.
diam(A) := sup |x − y| x, y ∈ A .
dist(A1 , A2 ) := inf |x − y| x ∈ A1 , y ∈ A2 .
(iii) The open ball of radius r > 0 having centre at x ∈ Rn is denoted by Br (x); that is,
2. Interior of a set, open/closed sets, limit point of a set, closure and boundary of a set
Definition 1.1.6. Let A be a nonempty subset of Rn . Then
(ii) the set of all interior points of A is called interior of A and is denoted by A◦ ,
(iv) x ∈ Rn is called limit point of the set A if it is the limit point of a sequence in A,
10
/ is open by convention
10 CHAPTER 1. PREREQUISITES
(v) x ∈ A is called isolated point of A if it is not the limit of any nontrivial sequence in A,
(vi) the union of A and all its limit points is called the closure of A and is denoted by Ā,
(iv) x ∈ A is an isolated point of A if and only if Br (x) ∩ A = {x} for some r > 0.
Theorem 1.1.9. (i) The union (intersection) of any number of open (closed) sets is open
(closed).
(ii) The intersection (union) of a finite number of open (closed) sets is open (closed).
Exercise 1.1.11. (i) Prove that for any x = (x1 , ..., xn ) ∈ Rn , there holds
n
|x j | ≤ |x| ≤ ∑ |xi |, j = 1, ..., n.
i=1
(iv) Prove E ⊂ Rn is bounded if and only if for each i ∈ {1, ..., n}, the i-th coordinates of its
points form a bounded subset of R.
Definition 1.1.12. K ⊂ Rn is called compact if every open cover of K has a finite subcover.
Kε := {x ∈ Rn dist(x, K) ≤ ε}.
5. Intervals
Definition 1.1.21. For n ∈ N, an n-dimensional interval I is a subset of Rn of the form
I = x = (x1 , ..., xn ) αk ≤ xk ≤ βk , k = 1, ..., n
≡ [α1 , β1 ] × [α2 , β2 ] × ... × [αn , βn ],
where αk < βk for all k = 1, ..., n. If the edge lengths βk − αk are all equal, then I is called a
cube. For an open interval/cube take strict inequalities in the above form. Two intervals are
non-overlapping whenever their interiors are disjoint.
The volume of an interval equals the volume of its interior and is given by
n
Vol(I) = ∏ (βk − αk ).
k=1
12 CHAPTER 1. PREREQUISITES
Theorem 1.1.22. Every open set in R can be written as a countable union of disjoint open in-
tervals. Every open set in Rn , n ∈ N, can be written as a countable union of non-overlapping
intervals.
Exercise 1.1.23. For compact X ⊂ Rn , prove there exists a non-increasing sequence {Xk ⊂
Rn }k∈N with X = k∈N Xk and each Xk consists of a finite union of non-overlapping intervals.
T
Exercise 1.1.24. Show that an open set in Rn , n ∈ N\{1}, is not necessarily a countable union
of disjoint open sets by explaining why the open unit disk in the plane cannot be written as a
countable disjoint union of open squares.
Disjointness can be achieved in the following way:
Definition 1.1.25. Let k ∈ Z. The n-dimensional cubes of the form
h j j + 1 h j j + 1
1 1 n n
k
, k
× ... × k
, k , ( j1 , ..., jn ) ∈ Zn ,
2 2 2 2
are called (right) half-open cubes of side length 2−k , or simply dyadic cubes.
Theorem 1.1.26. Every open set in Rn , n ∈ N, can be written as a countable union of disjoint
dyadic cubes.
1.2 Functions
1. Pre-image properties
For any function f : X → Y and any A, {Ak }k∈N subsets of Y , there holds
c
f −1 (Ac ) = f −1 (A) and f −1 f −1 (Ak ) .
[ [
Ak =
k∈N k∈N
2. Continuity
We denote by E an arbitrary nonempty subset of Rn and by f : E → R̄ a given function.
For any limit point x0 of E we consider the punctured balls
B∗r (x0 ) := Br (x0 ) \ {x0 }, r > 0.
We set
Mr (x0 ) := sup f (x) and mr (x0 ) := inf f (x).
B∗r (x0 )∩E B∗r (x0 )∩E
As r decreases, {Mr (x0 )}r>0 is non-increasing, {mr (x0 )}r>0 is non-decreasing and we set
lim sup f (x) := lim Mr (x0 ) and lim inf f (x) := lim mr (x0 ).
E∋x→x0 r→0 E∋x→x0 r→0
1.2. FUNCTIONS 13
3. Differentiability
We denote by E an arbitrary nonempty subset of Rn and by f : E → Rm a given function.
where
Kε := {x ∈ Rn dist(x, K) ≤ ε}, Uε := {x ∈ Rn dist(x,U c ) > ε},
and ηε : Rn → [0, ∞) is the standard mollifier given by ηε (x) := ε −n η(x/ε), x ∈ Rn , where
c exp{(|x|2 − 1)−1 }
(
|x| < 1,
η(x) := (1.2.1)
0 |x| ≥ 1,
R
with the constant c > 0 being such that Rn η(x) dx = 1.
Theorem 1.2.9. The gradient vector of a C1 (Ω) scalar
function f at a point x0 of any non
empty level surface Sλ = {x ∈ Ω f (x) = λ }, λ ∈ R, is perpendicular to the tangent at x0 of
any C1 -curve that lies on Sλ and goes through x0 . In other words, ∇ f (x0 ) is perpendicular
to the surface Sλ at x0 ∈ Sλ .
proof. Suppose ⃗γ(t) ∈ Sλ for all t ∈ I := (α, β ) ⊆ R and t0 ∈ I is such that ⃗γ(t0 ) = x0 ∈ Sλ .
Then
f ⃗γ(t) = λ for all t ∈ I.
Differentiating with respect to t we readily get by the chain rule that
d⃗γ
∇ f ⃗γ(t) · (t) = 0 for all t ∈ I.
dt
For t = t0 this gives ∇ f (x0 ) · d⃗
γ
dt (t0 ) = 0, which implies the result since
d⃗γ
dt (t0 ) is the tangent
vector to the curve at its point ⃗γ(t0 ).
1.3. VOLUME OF UNIT BALL IN Rn 15
∥I ∥ := max diam(Ik ).
k∈{1,...,N}
We further define
N N
UI ( f ) := ∑ sup f (x) Vol(Ik ) and LI ( f ) := inf f (x) Vol(Ik ).
∑ x∈I
k=1 x∈Ik k=1 k
We say the Riemann integral of f on I exists whenever the limit lim∥I ∥→0 RI exists. If the
Riemann integral of f on I exists we write
Z
f = lim RI ( f ).
I ∥I ∥→0
inf UI ( f ) = sup UI ( f ).
I I
π s/2
ωs := , s ≥ 0.
Γ(1 + s/2)
If s ∈ N then ωs is precisely the volume of the unit ball of Rs . Also the volume of a ball of
radius r > 0 is in this case given by
Vol(Br (x0 )) = ωs rs .
√
(ii) Γ(1/2) = π.
Measure
2.1 Measures
Definition 2.1.1 (measure). A mapping µ : 2X → [0, ∞] is called measure1 provided
/ = 0, and
(i) µ(0)
A ⊆ B =⇒ µ(A) ≤ µ(B),
It is easy to see that one can replace the second requirement in the above definition with
monotonicity plus subadditivity.
Exercise 2.1.3. Let µ be a measure on X and suppose {Ak ⊆ X}k∈N satisfies ∑k∈N µ(Ak ) < ∞.
Prove that µ(lim supk→∞ Ak ) = 0.
17
18 CHAPTER 2. MEASURE
(i) Dirac’s delta measure concentrated to a given x ∈ X; that is the measure on X defined
by
n 1 if x ∈ A,
δx (A) := = χA (x) , A ⊆ X.
0 if x ∈
/ A.
Proof. Let A ⊆ k∈N Ak . If x ∈ / k∈N Ak then δx (A) ≤ δx ( k∈N Ak ) = 0 = ∑k∈N δx (Ak ).
S S S
Proof. Let A ⊆ k∈N Ak with L n (Ak ) < ∞ for all k ∈ N and fix ε > 0. Given k ∈ N, by
S
definition of L n (Ak ) and the variational characterization of inf, we can choose intervals
(k) (k)
{I j } j∈N such that Ak ⊆ j∈N I j and
S
(k) ε
∑ Vol(I j ) < L n (Ak ) + .
j∈N 2k
(k)
Since A ⊆ the definition of L n (A) implies
S S
k∈N j∈N I j ,
(k)
L n (A) ≤ ∑ ∑ Vol(I j ).
k∈N j∈N
Combining the two estimates, L n (A) < ∑k∈N L n (Ak ) + ε, which implies the subaddi-
tivity since ε > 0 is arbitrary.
(iii) The s-dimensional Hausdorff measure on Rn (denoted throughout by H s ), defined for
any s ≥ 0 and any A ⊆ Rn by
H s (A) := lim Hδs (A),
δ →0+
where (see §1.3 for the definition of ωs )
n diamC s o
j
Hδ (A) := inf ∑ ωs
[
s
A⊆ C j , diam(C j ) ≤ δ .
j∈N 2 j∈N
Remark 2.1.5. Observe that for δ1 < δ2 , all coverings of A by sets with diameters no
more than δ1 are included in the set of coverings of A by sets with diameters no more
than δ2 . Thus Hδs increases as δ decreases. This justifies writing “lim” in the above
definition and also
H s (A) = sup Hδs (A).
δ >0
2.1. MEASURES 19
Remark 2.1.6. For s = 0 we deduce that H 0 is just the counting measure; that is,
H 0 (A) equals the number of elements in A when A is finite, and infinity otherwise.
This follows from the fact that ω0 = 1, hence H 0 {a} = 1 for any a ∈ Rn .
Proof. We show first that Hδs is a measure. Let A ⊆ k∈N Ak with Hδs (Ak ) < ∞ for all
S
k ∈ N and fix ε > 0. Given k ∈ N, by definition of Hδs (Ak ) and the variational char-
(k) (k) (k)
acterization of inf, we can choose {C j } j∈N such that Ak ⊆ j∈N C j , diam(C j ) ≤ δ
S
Proof. Let B ⊆ with µ|A (Bk ) < ∞ for all k ∈ N and note
S
k∈N Bk
[ [
Bk ∩ A = (Bk ∩ A). (2.1.1)
k∈N k∈N
Using the monotonicity of µ, then (2.1.1) and finally the subadditivity of µ,
[
µ|A (B) ≤ µ Bk ∩ A ≤ ∑ µ(Bk ∩ A) = ∑ µ|A (Bk ).
k∈N k∈N k∈N
Exercise 2.1.7. Prove for each one of the measures given in the above example, that its value
on the empty set is zero.
Exercise 2.1.8. Prove L n {a} = 0 for a ∈ Rn . Prove that for any countable set E in Rn we
have L n (E) = 0.
Exercise 2.1.9. Prove that the Lebesgue measure of an interval equals its volume.
Remark 2.2.2. The above inequality is trivial when µ(B) = ∞. Also, since B = (B ∩ A) ∪
(B \ A) we get by the subadditivity of µ that µ(B) ≤ µ(B ∩ A) + µ(B \ A). In particular,
µ-measurability of A ⊆ X is equivalent with
(ii) 0,
/ X and sets of µ-measure 0 are µ-measurable.
Proof. The first assertion readily follows from the definition of µ-measurability and so is
the µ-measurability of 0/ and X. Let µ(A) = 0 and B ⊆ X with µ(B) < ∞. Since B ∩ A ⊆ A
and B \ A ⊆ B, we get from the monotonicity of µ that µ(B ∩ A) = 0 and µ(B \ A) ≤ µ(B),
respectively. Add these to get µ(B ∩ A) + µ(B \ A) ≤ µ(B) as required. To prove the third
assertion let C ⊆ X be µ-measurable and B ⊆ X with µ|A (B) < ∞. It is enough to show
µ|A (B ∩C) + µ|A (B \C)≤ µ|A (B). Since (B \C) ∩ A = (B ∩ A) \C, this is written as µ (B ∩
A) ∩C + µ (B ∩ A) \C ≤ µ(B ∩ A), which is implied by the µ-measurability of C. For the
final assertion, let B ⊆ X with µ(B) < ∞ and pick A1 , A2 ⊆ X. If A1 is µ-measurable then
If A2 is µ-measurable then
µ(B \ A1 ) = µ (B \ A1 ) ∩ A2 + µ (B \ A1 ) \ A2 .
(ii) If {Ak }k∈N is non-decreasing, then limk→∞ µ(Ak ) = µ k∈N Ak .
S
(iii) If {Ak }k∈N is non-increasing and µ(A1 ) < ∞, then limk→∞ µ(Ak ) = µ
T
k∈N Ak .
S T
(iv) The sets k∈N Ak and k∈N Ak are µ-measurable.
22 CHAPTER 2. MEASURE
Proof. (i) Let {Ak }k∈N be a sequence of µ-measurable disjoint sets. By the µ-measurability
of A j+1 , j ∈ N,
j+1
[ j+1
[ j+1
[
µ Ak = µ Ak ∩ A j+1 + µ Ak \ A j+1 (2.2.1)
k=1 k=1 k=1
j
[
= µ(A j+1 ) + µ Ak .
k=1
Repeating this formula in the last summand, we arrive at
j+1
[ j−1
[ j+1
µ Ak = µ(A j+1 ) + µ(A j ) + µ Ak = ... = ∑ µ(Ak ) ∀ j ∈ N.
k=1 k=1 k=1
The proof is completed by taking the limit as j → ∞. Note µ( k∈N Ak ) ≤ ∑k∈N µ(Ak )
S
is also
true by the subadditivity of µ. □
(ii) Let {Ak }k∈N be a non-decreasing sequence of µ-measurable sets. For any k ∈ N write
Ak as a disjoint union of a finite number of µ-measurable sets (because of Theorem 2.2.4-(i)
and (iv)) as follows
k
[
Ak = A1 ∪ (A j+1 \ A j ) .
j=1
Using (i),
k
µ(Ak ) = µ(A1 ) + ∑ µ(A j+1 \ A j ),
j=1
and taking the limit as k → ∞,
∞
lim µ(Ak ) = µ(A1 ) + ∑ µ(A j+1 \ A j ).
k→∞ j=1
Because of (i), µ(A1 ) = µ(A1 \ Ak ) + µ(Ak ), k ∈ N, and since µ(A1 ) < ∞ we write µ(A1 \
Ak ) = µ(A1 ) − µ(Ak ), k ∈ N. Taking the limit,
lim µ(A1 \ Ak ) = µ(A1 ) − lim µ(Ak ). (2.2.3)
k→∞ k→∞
\ \ \ ∪
S T T
On the other
hand, noting k∈N (A1 A k ) = A1 k∈N A k and since A
1 = A 1 k∈N
2(A k )
k∈N Ak , the subadditivity of µ gives µ(A1 ) ≤ µ k∈N (A1 \ Ak ) + µ
T S T
k∈N Ak . Since
µ(A1 ) < ∞ we arrive at
[ \
µ (A1 \ Ak ) ≥ µ(A1 ) − µ Ak . (2.2.4)
k∈N k∈N
Since Bk ∩ Bck = 0,
/ we get from (i) that µ|B (Bk ) + µ|B (Bck ) = µ|B (X) = µ(B). Taking com-
plements we see the intersection of countably many µ-measurable sets is also µ-measurable.
2 wecannot use equality in place of inequality here (see (i)) because we don’t know yet that
T
is
k∈N Ak
µ-measurable
24 CHAPTER 2. MEASURE
/ X ∈A,
(i) 0,
(ii) A ∈ A ⇒ X \ A ∈ A ,
(ii) the Borel σ -algebra of Rn ; that is the smallest σ -algebra of Rn containing the open
subsets of Rn .3 A set in the Borel σ -algebra will be called a Borel set.
Definition 2.3.3 (Borel, Borel regular and Radon measures). Let µ be a measure on Rn .
(ii) µ is Borel regular if: (a) µ is Borel and (b) for each C ⊆ Rn there exists Borel set B ⊇ C
with µ(B) ≤ µ(C) (hence µ(C) = µ(B) by the monotonicity of measures),
(iii) µ is called Radon if µ is Borel regular and µ(K) < ∞ for all compact K ⊂ Rn .
Show next that for each C ⊆ Rn , there exists Borel set B ⊇ C with L n (B) ≤ L n (C) (this does
not imply L n is Borel regular because we don’t know yet it is Borel).
(ii) If µ is Borel regular and A ⊂ Rn is a Borel set, then µ|A is Borel regular.
(iii) If µ is Borel regular and A ⊂ Rn is a µ-measurable set with µ(A) < ∞, then µ|A is
Radon.
Proof. (i) By definition of µ being Borel, any Borel set is µ-measurable. From theorem
2.2.4-(iii) we know that any µ-measurable set is µ|A -measurable. Thus any Borel set is
µ|A -measurable and hence µ|A is Borel. □
n
(ii) From (i) we readily have that µ|A is a Borel measure. Now let C ⊆ R . We want to
find a Borel set D such that C ⊆ D and µ|A (D) ≤ µ|A (C). By definition of µ being Borel
regular, given C ⊆ Rn , there exists Borel set B such that C ∩ A ⊆ B and µ|A (C) = µ(B). Thus,
it suffices to find a Borel set D with C ⊆ D and µ|A (D) ≤ µ(B). The set D := B ∪ Ac is clearly
Borel and also C ⊆ D. Furthermore, D ∩ A = B ∩ A and we get µ|A (D) = µ(B ∩ A) ≤ µ(B).
□
(iii) By definition of µ being Borel regular, there exists Borel set A′ such that A ⊆ A′ and
µ(A′ ) = µ(A). Since A is µ-measurable we have µ(A′ ) = µ(A) + µ(A′ \ A). Since µ(A) < ∞
′ n ′
we conclude µ(A \ A) = 0. Now given C ⊆ R we have µ (C ∩ A ) \ A = 0 and using the
µ-measurability of A once more,
This shows µ|A′ and µ|A agree on any set and since A′ is a Borel set we conclude from (ii)
that µ|A is Borel regular. Finally note µ|A (Rn ) = µ(A) < ∞; that is µ is a finite measure. In
particular µ|A (K) < ∞ for any compact K ⊂ Rn .
Lemma 2.3.6. Let µ be a Borel measure on Rn and let B be a Borel set.
(i) If µ(B) < ∞, then ∀ ε > 0, ∃ closed C ⊆ B with µ(B \C) < ε.
Indeed, that would imply the existence of a closed set C with C ⊆ B and µ|B (B \ C) < ε,
which in turn implies (i) because µ|B (B \C) = µ(B \C). We next prove B ∈ F by showing
G := {A ∈ F Ac ∈ F }
is a σ -algebra containing all open sets. This would imply it contains all Borel sets, hence B
in particular.
We start by listing the required properties of F .
26 CHAPTER 2. MEASURE
(a) F contains all closed sets (this follows readily from the definition of F ).
Thus A ∈ F . □
which gives
µ|B (A \C) ≤ ∑ µ|B (Ak \Ck ) < ε.
k∈N
Sm
The sequence A \ ( k=1 Ck ) is non-increasing and so the
above inequality together
Sm
with theorem 2.2.5-(iii) imply lim µ|
m→∞ B A \ C
k=1 k < ε. In particular µ|B A \
Sm0
< ε for some m0 ∈ N. By theorem 1.1.9 we get m
S 0
k=1 Ck k=1 Ck is also closed and
thus A ∈ F . □
We may now proceed to check the collection of subsets G is a σ -algebra containing all
open sets. Since F contains all open and all closed sets, G contains all open sets. Conse-
quently it contains 0/ and Rn and by it’s definition, it contains the complement of each of it’s
elements. Let {Ak ∈ G }k∈N . Then Ak ∈ F , k ∈ N, and also Ack ∈ F c, k ∈ N, which imply
because of (c) and (b), respectively, that k∈N Ak ∈ F and k∈N Ak = k∈N Ak ∈ F . □
c
S S T
(ii) The sequence {Bk (0) \ B}k∈N is comprised of Borel sets with
Applying (i) we get closed sets {Ck ⊆ Bk (0) \ B}k∈N such that µ (Bk (0) \ B) \ Ck < ε2−k ,
k ∈ N. Set [
U := (Bk (0) \Ck ).
k∈N
Since B ⊆ Ckc 6 for all k ∈ N, we have B = k∈N (Bk (0) ∩ B) ⊆ k∈N (Bk (0) ∩ Ckc ) = U. Fur-
S S
thermore, observing
[ [ [
U \B = (Bk (0) \Ck ) \ B = (Bk (0) \Ck ) \ B = (Bk (0) \ B) \Ck ,
k∈N k∈N k∈N
we have
µ(U \ B) ≤ ∑µ (Bk (0) \ B) \Ck < ε.
k∈N
Finally notice U is open by it’s definition.
Proof. Assume first µ(A) < ∞ and let ε > 0. By theorem 2.3.5-(iii) we know µ|A is a Radon
measure. Applying theorem 2.3.7 for µ|A on the set Ac (observe µ|A (Ac ) = 0) we find open
U ⊇ Ac with µ|A (U) < ε. Then C := U c is closed with C ⊆ A. Since also U ∩ A = A \C, there
holds
ε > µ(A \C) = µ(A) − µ(C),
6 Taking complements in Ck ⊆ Bk (0) \ B, we get Bck (0) ∪ B ⊆ Ckc . In particular, B ⊆ Ckc .
28 CHAPTER 2. MEASURE
Next we show (2.3.1) is valid also when µ(A) = ∞. Define Dk := Bk (0) \ Bk−1 (0), k ∈ N, and
write A as a disjoint union of sets: A = k∈N (Dk ∩ A). The subadditivity of µ implies
S
Since µ is Radon, each Dk ∩ A is of finite µ-measure7 and we may apply the above argument
to get closed Ck , k ∈ N, such that Ck ⊆ Dk ∩ A with µ(Ck ) ≥ µ(Dk ∩ A) − 2−k , k ∈ N. Clearly,
k∈N Ck ⊆ A. Since {Ck }k∈N is a sequence of disjoint µ-measurable sets, using theorem
S
2.2.5-(i) we have
the reverse inequality being trivially true. Let C be a closed subset of A. By theorem 2.2.5-(ii)
we get
µ(C) = lim µ C ∩ B̄m (0) ≤ sup{µ(K) compact K ⊆ A},
m→∞
where in the last inequality we used C ∩ B̄m (0), m ∈ N, are compact subsets of A.
Exercise 2.3.9. Let 0 < λ < 1 and suppose X ⊂ R satisfies 0 < L 1 (X) < ∞. Prove there
exists an open interval I such that L 1 (X ∩ I) ≥ λ L 1 (I).
Lemma 2.3.10. Let {Et ⊂ Rn }t∈I be a family of disjoint Borel sets and µ be a Radon
measure on Rn . Then µ(Et ) > 0 for at most countably many t ∈ I .
where n 1o
Ik := t ∈ I µ Bk (0) ∩ Et > .
k
Then for any finite set J ⊂ Ik , k ∈ N, we have
[
µ Bk (0) ≥ µ Bk (0) ∩ Et
t∈J
= ∑ µ Bk (0) ∩ Et
t∈J
H 0 (J)
≥ .
k
Hence, for any k ∈ N the finite number kµ Bk (0) bounds H 0 (J) independently of the choice
By possibly dividing Ik we may further assume diam(Ik ) < dist(A, B) for all k ∈ N.
Hence the covering {Ik }k∈N splits in two subsequences {IkA }k∈N , {IkB }k∈N , the first of
which covers A and the second covers B. We have
L n (A) + L n (B) ≤ ∑ Vol(IkA) + ∑ Vol(IkB) = ∑ Vol(Ik ).
k∈N k∈N k∈N
Coupling these two inequalities, L n (A) + L n (B) < L n (A ∪ B) + ε, which implies the
claim since ε > 0 is arbitrary.
30 CHAPTER 2. MEASURE
Let A be the set of all sets of {Ck }k∈N having nonempty intersection with A; that is,
A := {Ck Ck ∩ A ̸= 0}.
/
Correspondingly set
B := {Ck Ck ∩ B ̸= 0}.
/
Clearly, A ⊆ Ck ∈A Ck and B ⊆ Ck ∈B Ck , while the assumption δ < dist(A, B) gives
S S
Coupling these two inequalities, Hδs (A) + Hδs (B) < H s (A ∪ B) + ε. Letting δ → 0,
H s (A) + H s (B) < H s (A ∪ B) + ε and the claim follows since ε > 0 is arbitrary.
Remark 2.4.5. By exercise 2.3.4 and by example 2.4.3-(i) we obtain L n is Borel regular.
Since it is clearly finite on compact sets, we conclude L n is Radon.
Remark 2.4.7. H s is not a Radon measure for 0 ≤ s < n. Using the Brunn-Minkowski
inequality of the next section one can prove the isodiametric inequality; that is, among all
bounded sets of Rn having the same fixed diameter, it is the ball that maximizes the volume,
or what is the same, for all X ⊆ Rn there holds
L n (X) ≤ L n (BX ).
where BX is defined to be any ball with radius equal to half the diameter of X. Using the
isodiametric inequality one can show that H n = L n . These facts will be included in a
possible future addendum to this notes.
32 CHAPTER 2. MEASURE
Chapter 3
X ε := x ∈ Rn dist(x, X) < ε ,
L n (X ε ) ≥ L n (BεrX ).
Definition 3.0.2 (Minkowski sum). The Minkowski sum of nonempty X,Y ⊆ Rn is given by
X +Y := {x + y x ∈ X, y ∈ Y }.
X1 ⊆ Y1 and X2 ⊆ Y2 =⇒ X1 + X2 ⊆ Y1 +Y2 .
33
34 CHAPTER 3. BRUNN-MINKOWSKI AND ISOPERIMETRIC INEQUALITIES
Observing X ε ⊇ X + Bε (0)1 and BεrX = BrX +ε 2 , the isoperimetric inequality would be readily
implied by the following inequality
1/n 1/n
L n X + Bε (0) ≥ ωn (rX + ε) (3.0.2)
1/n n 1/n
= L n (X) + L Bε (0) .
This shows the isoperimetric inequality is a special case of the following inequality
L n X + Bε (0) − L n (X)
Per(X) := lim inf .
ε↓0 ε
If X ⊂ Rn is bounded, noting BrX +ε ⊇ Brx + Bε (0) (combine the last two footnotes taking
X = BrX ), we further obtain from (3.0.2) and (3.0.1)
which asserts
among all bounded sets of Rn having the same fixed volume, it is the ball that minimizes
the perimeter,
1 Given y ∈ X + Bε (0) we know there exists x ∈ X and z ∈ Bε (0) such that y = x + z. Hence, dist(y, X) ≤
|y − x| = |z| < ε ⇒ y ∈ X ε . As an exercise, show these two sets are in fact equal whenever X is closed.
2 It is enough to prove Bε (0) = B
r r+ε (0). Given x ∈ B ε (0) we know dist x, B (0) < ε. For any z ∈ B (0)
r r r
we have |x| ≤ |x − z| + |z| < |x − z| + r. Hence |x| ≤ dist x, Br (0) + r < ε + r ⇒ x ∈ Br+ε (0). Conversely,
given x ∈ Br+ε (0) we know |x| < r + ε, so if x ∈ Br+ε (0) \ Br (0) then dist(x, Br (0)) = |x| − r < ε. Hence
Br+ε (0) \ Br (0) ⊂ Bεr (0). But also Br (0) ⊂ Bεr (0) and so Br+ε (0) ⊂ Bεr (0).
3 The smoother the boundary of the set X is, the n − 1 dimensional lower Minkowski content coincides with
its perimeter (in a weak notion which is out of the scope of these notes), H n−1 (∂ X) or even the surface area of
∂ X from calculus. This is why here (by abuse of notation) we denote it by Per.
35
or, equivalently,
among all bounded sets of Rn having the same fixed perimeter, it is the ball that maxi-
mizes the volume.
This last formulation is the classical isoperimetric statement. Finally we notice the perimeter
of BrX can be explicitly computed. Indeed,
and using once more (3.0.1) on the right, we deduce after coupling with (3.0.4),
1/n 1−1/n
Per(X) ≥ nωn L n (X) .
This is the classical expression of the isoperimetric inequality involving the notion of perime-
ter.
Lemma 3.0.5 (arithmetic-geometric mean inequality). 4 For positive real numbers a1 , ..., an ,
n 1/n 1 n
∏ ai ≤ ∑ ai .
i=1 n i=1
4 it’s proof will be an easy exercise (see remark 9.1.4) as soon as we learn about convex functions
36 CHAPTER 3. BRUNN-MINKOWSKI AND ISOPERIMETRIC INEQUALITIES
Step 2 - Reductions. From remark 2.4.5 we know L n is Radon and this implies the inner
regularity (theorem 2.3.8) of L n . Suppose X and Y are L n -measurable and pick compact
sets KX , KY such that KX ⊆ X and KY ⊆ Y . Then KX + KY ⊆ X +Y and assuming the Brunn-
Minkowski inequality holds true for all compact sets, we get from the monotonicity of L n
1/n 1/n 1/n
L n (X +Y ) ≥ L n (KX ) + L n (KY ) .
Taking the supremum over all compact KX ⊆ X and over all KY ⊆ Y , we deduce (3.0.3). Thus
(3.0.3) will be true for all L n -measurable X and Y , if it is true for all compact X and Y . Let
X and Y be compact. Recalling exercise 1.1.23, we construct two non-increasing sequences
{Xk ⊇ X}k∈N and {Yk ⊇ Y }k∈N with X = k∈N Xk , Y = k∈N Yk and such that each Xk and Yk ,
T T
Taking the limit k → ∞ we use theorem 2.2.5-(iii) to arrive to (3.0.3). Indeed, noting {Xk +
Yk ⊇ X +Y }k∈N is a non-increasing sequence of sets with X +Y = k∈N (Xk +Yk ),5 we deduce
T
Then X + and X − are non-overlapping sets comprised of non-overlapping intervals and the
number of intervals in each one is strictly less than m − 1 (it can be at most m − 2 in case Y is
5 If
α ∈ k∈N (Xk + Yk ), then ∀ k ∈ N, ∃ xk ∈ Xk and ∃ yk ∈ Yk such that xk + yk = α. But {Xk }k∈N is non-
T
increasing, hence xk ∈ X1 for all k ∈ N and since X1 is compact, a subsequence of xk converges to x ∈ X1 . Since
xk ∈ Xk for all k ∈ N and {Xk }k∈N is non-increasing to X, we deduce x ∈ Xk for all k ∈ N, that is x ∈ X. In the
same way, we get a a subsequence of yk converging to y ∈ Y . Taking the subsequential limit in xk + yk = α, we
get x + y = α. Thus α ∈ X +Y . The reverse inclusion is immediate since X +Y ⊆ Xk +Yk for all k ∈ N.
37
L n (X + )
λ := .
L n (X)
L n (X +Y ) ≥ L n (X + +Y + ) + L n (X − +Y − ).
6 Note that the quotient on the left is a continuous function of t ∈ R, that takes all values of [0, 1].
38 CHAPTER 3. BRUNN-MINKOWSKI AND ISOPERIMETRIC INEQUALITIES
Chapter 4
Integral
α ∈ R.
39
40 CHAPTER 4. INTEGRAL
A := {A ⊆ Rm f −1 (A) is µ-measurable},
contains all open sets of Rm . If we show it is a σ -algebra then by definition of the Borel
σ -algebra it will contain also all Borel sets. To this end notice first that 0,
/ R m ∈ A . Also,
c
if A ∈ A then by a property of the pre-image we have f −1 (Ac ) = f −1 (A) , which is µ-
A := {A ⊆ R̄ f −1 (A) is µ-measurable},
(i) µE, f is bounded, non-increasing, µE, f (−∞) = L n (E) and µE, f (∞) = 0.
Theorem 4.1.9. If f : X → [0, ∞] is µ-measurable, there exist µ-measurable sets {Ak ⊂ X}k∈N
such that
1
f (x) = ∑ χAk (x) ∀ x ∈ X.
k∈N k
Proof. We define the sets
n 1 k−1 1 o
A1 := {x ∈ X f (x) ≥ 1} and Ak := x ∈ X f (x) ≥ + ∑ χA j (x) , k ∈ N \ {1}.
k j=1 j
First we observe
f (x) ≥ χA1 (x) ∀ x ∈ X.
Indeed, this is self evident if x ∈ A1 and trivially true by the nonnegativity of f in case x ∈
X \ A1 . Next we claim
1
f (x) ≥ χA1 (x) + χA2 (x) ∀ x ∈ X. (4.1.1)
2
Indeed, noting
A2 = {x ∈ X f (x) ≥ 3/2 if x ∈ A1 and f (x) ≥ 1/2 if x ∈ X \ A1 },
we see again that (4.1.1) is self evident if x ∈ A1 ∩ A2 and trivially true by the nonnegativity
of f in case x ∈ X \ (A1 ∩ A2 ). If x ∈ A2 \ A1 then 1/2 ≤ f (x) < 1 and this gives f (x) ≥ 1/2
which is (4.1.1) in this case. Finally if x ∈ A1 \ A2 then 1 ≤ f (x) < 3/2 and this gives f (x) ≥ 1
which is (4.1.1) in this case. In the same fashion we can prove inductively that for any m ∈ N
m
1
f (x) ≥ ∑ χA j (x) ∀ x ∈ X.
j=1 j
Letting m → ∞ we obtain
1
f (x) ≥ ∑ k χAk (x) ∀ x ∈ X. (4.1.2)
k∈N
To show the equality we observe first that in case f (x) = ∞ we have x ∈ Ak for all k ∈ N.
Thus ∑k∈N (1/k)χAk (x) = ∑k∈N (1/k) = ∞ = f (x) in this case. If f (x) < ∞, then there exists
kx ∈ N such that x ∈
/ Ak for all k ≥ kx . This means
1 k−1 1
f (x) < + ∑ χA j (x) ∀ k ≥ kx ,
k j=1 j
42 CHAPTER 4. INTEGRAL
Ai, j := A ∩ f −1 (Bi, j ), i, j ∈ N.
(P1 ) {Ai, j }i, j∈N are µ-measurable. Indeed, since {Bi, j }i, j∈N are Borel and f is µ-measurable,
theorem 4.1.4-(i) implies { f −1 (Bi, j )}i, j∈N are µ-measurable. Since A is µ-measurable
we conclude that {Ai, j }i, j∈N are µ-measurable.
(P3 ) {Ai, j } j∈N are disjoint for any i ∈ N. Indeed, x ∈ Ai, j if and only if x ∈ A and there exists
y ∈ Bi, j such that y = f (x). Assuming x ∈ Ai, j ∩ Ai,k with j ̸= k, we get y ∈ Bi, j and
y′ ∈ Bi,k such that y = f (x) = y′ . This contradicts the disjointness of {Bi, j } j∈N .
Hence, out of a partition {Bi, j } j∈N of the target space Rm , we have constructed a partition
{Ai, j } j∈N of A ⊆ Rn comprised of µ-measurable sets.
Now let ε > 0. Theorem 2.3.5-(iii) assures that µ|A is Radon and because of (P1 ) we may
apply inner regularity (theorem 2.3.8) to get compact sets {Ki, j ⊆ Ai, j }i, j∈N with µ|A (Ai, j \
4.2. LUSIN AND EGOROFF THEOREMS 43
Being a finite union of compact sets, each Di is compact. For each i ∈ N we pick {bi, j ∈
Bi, j } j∈N and define functions gi : Di → Rm by
Ni
gi (x) := ∑ bi, j χKi, j (x), x ∈ Di .
j=1
(P3 ) together with {Ki, j ⊆ Ai, j }i, j∈N imply {Ki, j }i, j∈N are disjoint and thus, each being com-
pact, of positive distance apart (use exercise 1.1.14). This means each gi is continuous. Also,
if x ∈ Di , then x ∈ Ki, j ⊆ Ai, j for some j ∈ {1, ..., Ni }. Thus f (x) ∈ Bi, j for some j ∈ {1, ..., Ni }.
Since gi (x) = bi, j ∈ Bi, j for this j, we obtain
1
| f (x) − gi (x)| ≤ diam(Bi, j ) < ∀ x ∈ Di . (4.2.2)
i
Finally we define the set \
K := Di .
i∈N
44 CHAPTER 4. INTEGRAL
of measures gives
µ(A \ K) ≤ ∑ µ(A \ Di ) < ε,
i∈N
by the definition of Di and (4.2.1). Also, if x ∈ K then x ∈ Di for all i ∈ N and because of
(4.2.2)
1
sup | f (x) − gi (x)| < .
x∈K i
m
Thus {gi : Di → R }i∈N is a sequence of continuous functions which converges uniformly to
f on K = i∈N Di . Standard analysis implies the limit function f is continuous in K.3
T
Then {A ∩Ci, j } j∈N are µ-measurable (by theorem 4.1.6-(ii)) and non-increase to j∈N A ∩
T
Ci, j . But fk → f µ-a.e. in A implies j∈N A ∩Ci, j ⊆ N. Taking into account the fact that
T
Hence, for each i ∈ N there exists Ni ∈ N such that µ(A ∩Ci,Ni ) < ε2−i .
We define next the set [
B := A \ Ci,Ni .
i∈N
c
S
Observing that A \ B = A ∩ A ∪ i∈N Ci,Ni = A ∩ i∈N Ci,Ni = i∈N (A ∩Ci,Ni )
S S
we get
µ(A \ B) ≤ ∑ µ(A ∩Ci,Ni ) < ε.
i∈N
Moreover, since for each i ∈ N we have x ∈ B ⇒ x ∈ / A ∩Ci,Ni , we get by the definition of the
Ci, j ’s that for each i ∈ N, any x ∈ B and all k ≥ Ni , there holds | fk (x) − f (x)| ≤ 2−i .
3 | f (x) −
f (x0 )| ≤ | f (x) − gi (x)| + |gi (x) − gi (x0 )| + |gi (x0 ) − f (x0 )| ≤ |gi (x) − gi (x0 )| + 2 maxx∈K | f (x) −
gi (x)| ∀x, x0 ∈ K
4.3. INTEGRATION OF MEASURABLE FUNCTIONS 45
Observe these are nonnegative functions and that the following decompositions are valid
g = g+ − g− and |g| = g+ + g− .
Remark 4.3.3. The above definition implies that if g : X → R̄ is simple, then there exist
disjoint {Ak ⊆ X}k∈N and {αk ∈ R̄}k∈N such that
[
g= ∑ αk χAk and Ak = X.
k∈N k∈N
Consider now µ to be a measure on the set X. The sets {Ak }k∈N in the above expression can
be taken µ-measurable if g is known to be µ-measurable. Indeed, for each k ∈ N we have
and thus by theorem 4.1.4-(ii) Ak is µ-measurable. Consequently, we have the useful fact4
Proposition 4.3.4. Given a µ-measurable simple function g : X → [0, ∞], then there exist
disjoint µ-measurable sets {Ak }k∈N and {αk ∈ (0, ∞]}k∈N such that
[
g= ∑ αk χAk and Ak ⊆ X.
k∈N k∈N
g := ∑ t k χEk .
k∈Z
Then X \{ f = 0} = k∈Z Ek and assigning the value 0 to g in { f = 0}, we have g(x) ≤ f (x) ≤
S
tg(x) µ-a.e. x ∈ X. From the definitions of upper and lower integrals of f we deduce
Z Z Z Z
f dµ ≤ tg(x) dµ = t g(x) dµ ≤ t f (x) dµ,
X X X X
R
where
R
we have used 4.3.7-(iii) to get the equality. This holds for any t > 1 and so X f dµ ≤
X
f (x) dµ.
Proof. Given ε > 0 we employ the outer and inner regularity properties of Radon measures
(theorems 2.3.7 and 2.3.8), to get open U ⊃ A and compact K ⊂ A such that
Hence µ(U \ K) < ε. Further, from lemma 1.2.6 we get a continuous function g : Rn → [0, 1]
with g = 1 on K and sprt(g) ⊂ U. It follows that |χA − g| = 0 in (U \ K)c and |χA − g| ≤ 1 on
U \ K. Altogether,
Z Z
p
|χA − g| dµ = |χA − g| p dµ ≤ µ(U \ K) < ε.
Rn U\K
Definition 4.3.15
R
(µ-summable function). f : X → R̄ is called µ-summable if it is µ-
integrable and X | f | dµ < ∞.
1 Z
µ {x ∈ X | f (x)| ≥ s} ≤ | f | dµ whenever s > 0.
s X
R
Use this to prove that X |f| dµ = 0 if and only if f = 0 µ-a.e. in X.
4.4 L p spaces
Let µ be a measure on a set X ̸= 0.
/
Notation 4.4.1. From now on we write {g > α} for {x ∈ X | g(x) > α}, etc.
Remark
4.4.3. Suppose µ {g > α} > 0 for all α ∈ R. Then ∄ α ∈ R such that µ {g >
α} = 0; that is, ∄ α ∈ R such that g(x) ≤ α for µ-a.e. x ∈ X; in other words, g = ∞ on a
set of positive µ-measure. This justifies
the middle definition of ess sup. On the other hand,
if ∃ α ∈ R such that µ {g > α} = 0, then g(x) ≤ α for µ-a.e. x ∈ X; that is α is an upper
bound for g (except possibly on a subset of X of µ-measure 0). Hence it is natural to define
ess sup as the least upper bound in this case.
4.4. L p SPACES 49
Also,
Remark 4.4.5. From Definition 4.3.15 we notice that L1 (X, µ) is the collection of all µ-
summable functions.
this proves (i). For (ii), let µ(X) > 0 and assume first that ∃ α ∈ [0, ∞) such that µ {| f | >
α} = 0. This implies (see the remark following the definition of ess sup) that
ess supX | f | = inf{α ∈ R | µ {| f | > α} = 0} =: M ∈ [0, ∞).
follows. Finally, the case where µ(X) = 0 is also clear since ess supX | f | = 0 by its definition,
p
R
and also X | f | dµ = 0.
Theorem 4.4.7 (Hölder inequality). Let p, q ∈ [1, ∞] satisfying 1/p + 1/q = 1. If f ∈ L p ,
g ∈ Lq then
R
ess supX |g| X | f | dµ if p = 1,
Z
1/p R 1/q
| f g| dµ ≤ p q
R
X | f | dµ X |g| dµ if 1 < p < ∞,
X
R
ess supX | f | X |g| dµ if p = ∞.
Proof. Let 1 < p < ∞ (the cases p = 1 and p = ∞ are trivial).R We know Rthe convexity
inequality |ab| ≤ 1p |a| p + q1 |b|q for all a, b ∈ R (prove it!). Hence if X | f | p dµ = X |g|q dµ = 1
we get
Z
1
Z
1
Z
1 1 Z 1/p Z 1/q
p q p q
| f g|dµ ≤ | f | dµ + |g| dµ = + = 1 = | f | dµ |g| dµ .
X p X q X p q X X
p q
R R
If X |f| dµ, X |g| dµ > 0, normalize f , g as follows
f g
f˜ := R 1/p , g̃ := R 1/q .
p q
X | f | dµ X |g| dµ
The proof follows by applying Hölder’s inequality with exponents p and p/(p − 1) on both
terms of the right hand side and then rearranging terms in the resulting inequality.
In particular, applying Fatou’s lemma to the sequence gk := | flk | p gives f ∈ L p . Applying then
the dominated convergence theorem for the sequence hk := | flk − f | p , we deduce ∥ flk − f ∥ p →
0, as k → ∞. This, together with the fact that { fk }k∈N is Cauchy in L p , imply ∥ fk − f ∥ p → 0,
as k → ∞.
52 CHAPTER 4. INTEGRAL
Chapter 5
where
A := g µ-integrable, simple, g ≤ lim inf fk µ-a.e. in X .
k→∞
Let g ∈ A . By proposition 4.3.4 we may assume g = ∑ j∈N α j χA j , for some {α j > 0} j∈N and
disjoint µ-measurable {A j ⊆ X} j∈N with j∈N A j ⊆ X.
S
Proof of claim. Let j ∈ N. We only prove A j ⊆ k∈N B j,k , the reverse inclusion being self
S
53
54 CHAPTER 5. THE BASIC THEOREMS OF ADVANCED ANALYSIS
This implies the existence of k0 ∈ N such that tα j < infk≥ℓ fk µ-a.e. in A j , for all ℓ ≥ k0 ,
which in turn gives tα j < fℓ µ-a.e. in A j , for all ℓ ≥ k0 . □
Using {A j } j∈N are disjoint with ⊆ X and then A j ⊇ B j,k for all k ∈ N, we get for
S
j∈N A j
any m ∈ N
Z m Z
fk dµ ≥ ∑ fk dµ
X j=1 A j
m Z m
≥ ∑ fk dµ ≥ t ∑ α j µ(B j,k ),
j=1 B j,k j=1
the last inequality coming straight from the definition of B j,k . Observing next B j,k+1 ⊇ B j,k
for all k ∈ N and taking the lim infk→∞ , we arrive at
Z m
lim inf fk dµ ≥ t lim µ(B j,k )
∑ α j k→∞
k→∞ X j=1
m m
=t ∑ α j µ(∪k∈NB j,k ) = t ∑ α j µ(A j ),
j=1 j=1
which implies Z Z
lim inf fk dµ ≥ sup g dµ.
k→∞ X g∈A X
and so Z Z
lim sup fk dµ ≤ lim fk dµ. (5.1.3)
k→∞ X X k→∞
(ii) gk → | f | as k → ∞, µ-a.e. in X,
Indeed, for (i) observe that gk ≤ k and gk ≤ | f | both µ-a.e. in X. (ii) is obvious. For
(iii) notice in case | f (x)| < k that gk (x) = | f (x)| = gk+1 (x), in case k ≤ | f (x)| < k + 1 that
gk (x) = k ≤ | f (x)| = gk+1 (x) and in case | f (x)| ≥ k + 1 that gk (x) = k < k + 1 = gk+1 (x).
Because of (ii) and (iii) we get through the monotone convergence theorem
Z Z
lim gk dµ = | f | dµ;
k→∞ X X
But (i) implies the difference in the absolute value is non-positive and so
Z Z
| f | dµ < gkε dµ + ε.
X X
Let δ := ε/kε . For any µ-measurable A ⊆ X with µ(A) < δ we rewrite the above inequality
Z Z Z
| f | dµ < gkε dµ + (gkε − | f |) dµ + ε.
A A Ac
Property (i) implies the second integral on the right is non-positive and also the first integral
does not exceed kε µ(A). Hence
Z
| f | dµ < kε δ + ε = 2ε.
A
Rescaling ε the proof is complete.
Theorem 5.1.5 (dominated convergence - Lebesgue). Let f , fk : X → R̄, k ∈ N, be µ-
measurable, g : X → [0, ∞] be µ-summable, satisfying
(i) fk → f as k → ∞, µ-a.e. in X,
(ii) | fk | ≤ g µ-a.e. in X, for all k ∈ N.
Then Z
lim | fk − f | dµ = 0. (5.1.4)
k→∞ X
Proof. First notice that (ii) and (i) imply | f | ≤ g µ-a.e. in X. This and (ii) again give through
the triangle inequality
2g − | fk − f | ≥ 2g − | fk | − | f | ≥ 0 µ-a.e. in X ∀ k ∈ N.
Applying Fatou’s lemma to the sequence defined by the left hand side gives
Z Z
lim inf 2g − | fk − f | dµ ≤ lim inf 2g − | fk − f | dµ,
X k→∞ k→∞ X
and using (i) once more together with the summability of g and exercise 1.1.2
Z Z Z
2g dµ ≤ 2g dµ − lim sup | fk − f | dµ.
X X k→∞ X
Indeed, let fk , f satisfy the hypotheses of the theorem and take gk := fk − f , g := f in the
above statement (the choice
p for g is eligible
because of remark 5.1.10). Then the elementary
1 p p
inequality |a| + |b| ≤ 2 |a| + |b| gives p
Z Z p
|gk | p dµ ≤ | fk | + | f | dµ
X X Z Z
p p
≤2 | fk | dµ + | f | dµ ≤ 2 p+1 M,
p
X X
where we used (II) to get to the last inequality. Taking the limit and using (5.1.8)
Z
lim sup |gk + g| p − |gk | p − |g| p dµ ≤ ε M̃.
k→∞ X
Definition 5.2.1 (product measure). The product measure of µ and ν is the measure µ × ν :
2X×Y → [0, ∞] given by
where the infimum is taken over all collections of µ-measurable sets Ak ⊆ X and ν-measurable
sets Bk ⊆ Y , k ∈ N, such that [
S⊆ (Ak × Bk ).
k∈N
60 CHAPTER 5. THE BASIC THEOREMS OF ADVANCED ANALYSIS
Definition 5.2.2 (σ -finite set). The set A ⊆ X is called σ -finite with respect to µ if
[
A= Ak , where Ak are µ-measurable and µ(Ak ) < ∞ ∀ k ∈ N.
k∈N
Exercise 5.2.3. Prove that if A ⊆ X is σ -finite with respect to µ, then there exist disjoint
µ-measurable sets Bk , k ∈ N, such that A = k∈N Bk and µ(Bk ) < ∞ for all k ∈ N.
S
Definition 5.2.4 (σ -finite function). The function f : X → R̄ is called σ -finite with respect
to µ if f is µ-measurable and the set {x ∈ X f (x) ̸= 0} is σ -finite with respect to µ.
Theorem 5.2.5 (Fubini). (i) µ × ν is a regular measure on X ×Y .
(ii) If A ⊆ X is µ-measurable and B ⊆ Y is ν-measurable, then A × B is µ × ν-measurable
and (µ × ν)(A × B) = µ(A)ν(B).
(iii) If S ⊆ X × Y is σ -finite with respect to µ × ν, then Sy := {x ∈ X (x, y) ∈ S} is µ-
measurable for ν-a.e. y ∈ Y , Sx := {y ∈ Y (x, y) ∈ S} is ν-measurable for µ-a.e.
x ∈ X, µ(Sy ) is ν-integrable and ν(Sx ) is µ-integrable. Moreover,
Z Z
(µ × ν)(S) = µ(Sy ) dν(y) = ν(Sx ) dµ(x).
Y X
Lemma 5.2.8 (continuous version of the Minkowski inequality). Let p ∈ [1, ∞). For any
f ∈ L p (Rn × Rn , L n ) we have
Z Z p 1/p Z Z 1/p
nn p n
| f (x, y)| dL (y) dL (x) ≤ | f (x, y)| dL (x) dL n (y).
Rn Rn Rn Rn
p p−1
dL n (y). Then, writing F(x) = F(x) F(x)
R
Proof. Set F(x) := Rn | f (x, y)| , we get
Z p Z Z p−1
n
I := F(x) dL (x) = | f (x, y)| dL n (y) F(x) dL n (x)
Rn R n
Z Z Rn
p−1
= | f (x, y)| F(x) dL n (x) dL n (y),
Rn Rn
from Fubini’s theorem. Applying now Hölder’s inequality on the inner integral we deduce
Z Z 1/p Z p (p−1)/p
p n n
I≤ | f (x, y)| dL (x) F(x) dL (x) dL n (y)
Rn Rn Rn
Z Z 1/p
= I(p−1)/p | f (x, y)| p dL n (x) dL n (y).
Rn Rn
1/p
Hence I1/p ≤ p dL n (x) dL n (y) which is the desired inequality.
R R
Rn Rn | f (x, y)|
...
62 CHAPTER 5. THE BASIC THEOREMS OF ADVANCED ANALYSIS
Chapter 6
Proof. Starting from the left hand side and using the definition of ν we have
Z Z
φ f (x) dµ(x) = ν 0, f (x) dµ(x) (6.0.4)
X ZX Z ∞
= χ (t) dν(t) dµ(x).
X 0 0, f (x)
63
64 CHAPTER 6. SOME REPRESENTATION FORMULAS
Since χ (t) = χ
{ f >t} (x) for all t ∈ [0, ∞), x ∈ X, we conclude
0, f (x)
Z Z ∞Z
φ f (x) dµ(x) = χ{ f >t} (x) dµ(x) dν(t)
X Z0∞ X
= µ { f > t} dν(t).
0
For the formula (6.0.2), the assumption ν [0,t) = ν [0,t] for any t ≥ 0 when applied to
(6.0.4) gives Z Z
φ f (x) dµ(x) = ν 0, f (x) dµ(x).
X X
We conclude with the proof by following the same steps as before. Finally, to get formula
(6.0.3) notice that
Z Z Z f (x)
φ ′ (t) dt dµ(x)
φ f (x) dµ(x) =
X ZX Z0 ∞
= χ (t)φ ′ (t) dL 1 (t) dµ(x).
X 0 0, f (x)
Now proceed as above from the application of Fubini’s theorem and on.
Remark 6.0.2. Under the assumptions for (6.0.2) to hold true, since { f > t} ⊆ { f ≥ t},
monotonicity of measures, formulas (6.0.1), (6.0.2) and exercise 4.3.16 readily imply µ({ f >
t}) = µ({ f ≥ t}) for ν-a.e. t ∈ [0, ∞); that is,
Observe that in case µ is a Radon measure on Rn and f is Borel measurable then this follows
from lemma 2.3.10. Indeed, the sets Et := { f = t}, t ∈ I := [0, ∞) form a family of disjoint
Borel sets. Hence lemma 2.3.10 asserts µ(Et ) > 0 for at most countable many t ∈ I ; that is,
Definition 6.0.4. Let p, q > 0. The Lorentz space L p,q is the collection of all measurable
functions f defined on Rn , such that [ f ]L p,q < ∞, where
!1/q
q
Z ∞ q
[ f ]L p,q := f ∗ (t) t q/p−1 dt .
p 0
(ii) If χω is the characteristic function of a set ω with finite Lebesgue measure, then
[χω ]L p,q = [L n (ω)]1/p for all p, q > 0.
We prove next an analogous to (6.0.6) formula for f = |∇u|, where u ∈ Cc1 (Rn ).2 To establish
it we will need the following proposition and its consequence, the divergence or Gauss-Green
theorem. The interested student can find the proof in [EG]-§3.3.
Proposition 6.0.6. The measure H n−1 of a sufficiently smooth bounded surface agrees with
the surface’s area as taught in vector calculus.
Theorem 6.0.7 (divergence theorem). Let Ω ⊂ Rn be a bounded domain with smooth bound-
ary and ⃗F = (F1 , ..., Fn ), Fi ∈ C1 (Ω̄) for any i = 1, ..., n, be a vector field. Then
Z Z
div ⃗F(x) dx = ⃗F(x) ·⃗ν∂ Ω (x) dH n−1 (x),
Ω ∂Ω
where ⃗ν∂ Ω (x) is the outwards pointing unit normal at the point x ∈ ∂ Ω.
Theorem 6.0.8 (representation formula for the integral of the modulus of the gradient).
For any u ∈ Cc1 (Rn ) we have
Z Z ∞
H n−1 {|u| = t} dL 1 (t),
|∇u(x)| dx = (6.0.7)
Rn 0
⃗ ε : Rn → Rn by
Proof. Let ε > 0 and u ∈ Cc1 (Rn ). Define also U
⃗ ε (x) := p ∇u(x)
U .
|∇u(x)|2 + ε
Noting that for any t ≥ 0 we have {u+ > t} = {u > t}, formula (6.0.5) implies
Z ∞
+
u (x) = χ{u>t} (x) dL 1 (t) L n -a.e. in Rn .
0
Also, since for any t ≥ 0 we have {u− ≥ t} = {u ≤ −t}, the same formula gives
Z ∞
−
u (x) = χ{u≤−t} (x) dL 1 (t)
0
Z 0
= χ{u≤t} (x) dL 1 (t)
−∞
Z 0
χ{u>t} (x) − 1 dL 1 (t) L n -a.e. in Rn .
=−
−∞
We then have
Z Z Z ∞
+ ⃗ ε (x) dL (x) =
u (x) div U n
χ{u>t} (x) dL 1 (t) div U
⃗ ε (x) dL n (x)
Rn n
ZR∞ Z 0
= ⃗ ε (x) dx dL 1 (t),
div U
0 {u>t}
67
where we have applied Fubini’s theorem in the second equality and the divergence theorem
⃗ ε (x)dx = 0). Adding the last two equalities
R
in the last one (to show that Rn div U
Z Z ∞Z
⃗ ε (x) dx =
u(x) div U ⃗ ε (x) dx dL 1 (t),
div U
Rn −∞ {u>t}
from which, after an integration by parts on the left hand side and applying the divergence
theorem on the right hand side4 , we conclude
Z Z ∞Z
− ⃗ ε (x) dx =
∇u(x) · U ⃗ ε (x) ·⃗ν(x) dH n−1 (x) dL 1 (t),
U
Rn −∞ {u=t}
where ⃗ν(x) is the unit normal at a point x of the surface {u = t} pointing in the direction
where u ≤ t. From Theorem 1.2.9 we know ∇u(x) ·⃗ν(x) = −|∇u(x)| on {u = t} and so
The results follows letting ε ↓ 0 through the monotone convergence theorem applied in both
sides.
Combining the above representation theorem with the isoperimetric inequality we get the
sharp L1 -Sobolev inequality, or sharp Gagliardo-Nirenberg inequality; that is,
4 we
need to know that the level set {u = t} is a smooth surface to apply it. This is a consequence of Sard’s
lemma which is going to be added in the next update of the notes (see the Appendix of [W])
68 CHAPTER 6. SOME REPRESENTATION FORMULAS
It suffices to prove
Z ∞ (n−1)/n
L n {|u| > t} dL 1 (t) ≥ I (n−1)/n , (6.0.9)
0
where I := Rn |u(x)|n/(n−1) dx. To this end using (6.0.5) with f = |u| and then Fubini’s
R
theorem we write
Z
I= |u(x)||u(x)|1/(n−1) dx
Z Rn Z ∞
= χ{|u|>t} (x) dL 1 (t)|u(x)|1/(n−1) dL n (x)
Rn
Z ∞ Z0
= χ{|u|>t} (x)|u(x)|1/(n−1) dL n (x) dL 1 (t).
0 Rn
Applying next Hölder’s inequality with the conjugate exponents n/(n − 1) and n,
Z ∞ Z n/(n−1) (n−1)/n
I≤ χ{|u|>t} (x) n
dL (x) I 1/n dL 1 (t)
0 n
Z ∞R (n−1)/n
= I 1/n L n {|u| > t} dL 1 (t),
0
Proposition 7.1.2 (properties of fˆ). The Fourier transform enjoys the following properties:
(i) For f ∈ L1 (Rn ) we have fˆ ∈ L∞ (Rn ) with ∥ fˆ∥∞ ≤ ∥ f ∥1 . In particular, since for non-
negative f ∈ L1 (Rn ) we get fˆ(0) = ∥ f ∥1 , we deduce ∥ fˆ∥∞ = ∥ f ∥1 in this case.
n ˆ 1 n
λ f (κ) = λ f (λ κ) for all λ ∈ (0, ∞), where δλ f (x) := f (x/λ ), f ∈ L (R ).
(vi) δd
(viii) fˆ ∈ C(Rn ).
(ix) lim|κ|→∞ fˆ(κ) = 0.
2 /λ
(x) e−πλ |x| (κ) = λ −n/2 e−π|κ|
\2
for all λ ∈ (0, ∞).
69
70 CHAPTER 7. THE FOURIER TRANSFORM
Proof. Properties (i) and (ii) follow from Theorem 4.3.12. Properties (iii) and (iv) are con-
sequences of the Fubini theorem. Each of the property (v) and (vi) follows from the obvious
change of variables. Property (vii) is elementary. For a given κ0 ∈ Rn we write
Z
ˆf (κ) − fˆ(κ0 ) = gκ (x) dL n (x) where gκ (x) := e−2πiκ·x − e−2πiκ0 ·x f (x).
Rn
Clearly, limκ→κ0 gκ (x) = 0 and |gκ (x)| ≤ 2| f (x)|, both for L n -a.e. x ∈ Rn . Since f ∈ L1 (Rn ),
the dominated convergence theorem applies to establish (viii). @@@@ To prove (x), we
compute
Z Z √ √
−2πik·x−πλ |x|2 −π|k|2 /λ 2
e−|i π/λ k+ πλ x| dx
\
−πλ |x|2
e (k) = e dx = e
Rn Rn
Z Z n
−n/2 −π|k|2 /λ −|y|2 −n/2 −π|k|2 /λ 2
= (πλ ) e e dy = (πλ ) e e−t dt .
Rn R
R −t 2
√
Finally notice that Re dt = π (see Exercise 1.3.1).
Lemma 7.1.3 (continuity of the translation operator in L p ). Let p ∈ [1, ∞). For any f ∈
L p (Rn , L n ) we have limh→0 ∥τh f − f ∥ p = 0.
Proposition 7.1.4 (invertibility of the Fourier transform). If f , fˆ ∈ L1 (Rn ) then
Z
e2πik·x fˆ(k) dk = fˆˆ(−x) , for a.e. x ∈ Rn .
f (x) =
Rn
Because of Fubini’s theorem and the continuity of the translation operator in L1 , we can see
that Jε (x) converges to f (x) in L1 (Rn ) (see below, in the proof of “Approximation in L p theo-
rem”, for the details). Hence, for some subsequence δ (ε) we know Jδ (ε) (x) → f (x) for L n -
a.e. x ∈ Rn , as ε → 0. On the other hand, the integrand of Jδ (ε) (x) converges to e2πik·x fˆ(k)
as ε → 0, for all x ∈ Rn . Moreover, this same integrand is dominated by | fˆ(k)| which is (by
assumption) an L1 (Rn , dL n (k)) function. So the dominated convergence theorem applies to
end the proof.
7.1. THE FOURIER TRANSFORM IN L1 71
Remark 7.1.5. It turns out from the regularity properties (i), (viii) and (ix) of Proposition
7.1.2 that, given f ∈ L1 (Rn ) we have f ∈ L∞ (Rn ) ∩ C0 (Rn ), but note carefully that it is not
necessarily an L1 (Rn ) function. Hence the invertibility formula is not valid in general (con-
sider for instance f = χ(α,β ) in R).
Using the last lemma (continuity of the translation operator in L p ), as well as the con-
tinuous version of Minkowski’s inequality (see Lemma 5.2.8), one has the following useful
approximation theorem, already used with p = 1 in the “Invertibility of the Fourier trans-
form”.
Proof. Let
Z
−n/2 2 /ε
Iε := ε e−π|k−x| f (k) dL n (k) − f (x) .
Rn L p Rn ,dL n (x)
We use the above theorem with p = 2 in the proof of the following fundamental result:
Theorem 7.1.7 (Plancherel’s formula). If f ∈ L1 (Rn )∩L2 (Rn ), then fˆ ∈ L2 (Rn ) with ∥ fˆ∥2 =
∥ f ∥2 .
...
72 CHAPTER 7. THE FOURIER TRANSFORM
∥ fˆj − fˆl ∥2 = ∥ f j − fl ∥2 ∀ j, l ∈ N;
that is, { fˆj } j∈N is a Cauchy sequence in L2 . But L2 is complete and thus { fˆj } j∈N converges
to a function of L2 (Rn ) which we call the Fourier transform of f and denote it by fˆ.
Remark 7.2.2. Given f ∈ L2 (Rn ) we can always find sequences { f j ∈ L1 (Rn ) ∩ L2 (Rn )} j∈N
such that f j → f in L2 . For example, taking f j := η1/ j ∗ f , where ηε for ε > 0 is the standard
mollifier, we have { f j ∈ Cc∞ (Rn )} j∈N such that f j → f in L2 . Another example is { f j :=
f χB j } j∈N . Since by Hölder’s inequality ∥ f j ∥1 ≤ ∥ f ∥2 [L n (B j )]1/2 for all j ∈ N and also
∥ f j ∥2 ≤ ∥ f ∥2 for all j ∈ N, we have f j ∈ L1 (Rn ) ∩ L2 (Rn ) for all j ∈ N. Moreover,
Z
∥fj − f ∥22 = g j dL n , where g j := (1 − χB j )| f |2 .
Rn
Remark 7.2.3. Although there are many sequences such that f j → f in L2 , fˆ is independent
of the one we choose. Indeed, suppose { f j ∈ L1 (Rn ) ∩ L2 (Rn )} j∈N satisfies lim j→∞ ∥ f j −
f ∥2 = 0. Then fˆ is defined as the L2 -limit of fˆj , hence lim j→∞ ∥ fˆj − fˆ∥2 = 0. Let {g j ∈
L1 (Rn ) ∩ L2 (Rn )} j∈N be one more sequence such that lim j→∞ ∥g j − f ∥2 = 0. Then
where we have used Plancherel’s formula to get the middle equality. Thus, lim j→∞ ∥ĝ j −
fˆ∥2 = 0 which says fˆ is the L2 -limit of {ĝ j } j∈N too. □
f + β g = α fˆ + β ĝ.
(ii) (linearity) If f , g ∈ L2 (Rn ) and α, β ∈ C, then α \
7.2. THE FOURIER TRANSFORM IN L2 73
By Hölder’s inequality,
Z rZ
− n2 −π|k−x|2 /ε n
f j (k) − f (k) dL n (k) ≤ ε − 2
2 /ε
e−2π|k−x|
ε e dk ∥ f j − f ∥2 → 0,
Rn Rn
as j → ∞. Also fˆ is defined as the L2 -limit of fˆj ; that is, lim j→∞ ∥ fˆj − fˆ∥2 = 0. Hence,
Z rZ
2πik·x−επ|k|2 ˆ ˆ n 2
e−2επ|k| dk ∥ fˆj − fˆ∥2 → 0,
e f j (k) − f (k) dL (k) ≤
Rn Rn
From Theorem - Approximation in L p with p = 2, we know the rhs of (∗) converges to f (x) in
L2 . Hence there is a subsequence (that we don’t rename) such that the rhs of (∗) converges to
f (x) for L n -a.e. x ∈ Rn . To see that the lhs of (∗) converges (up to a subsequence) to fˆˆ(−x)
2
for L n -a.e. x ∈ Rn , set first gε (k) := e−επ|k| fˆ(k) and observe that gε ∈ L1 (Rn ) ∩ L2 (Rn ) for
all ε > 0. This allows to write lhs(∗) = ĝε (−x). Next we claim that gε converges to fˆ in L2 .
Indeed, we have
Z
2 2
∥gε − fˆ∥22 = 1 − e−επ|k| | fˆ(k)|2 dL n (k), ε > 0.
Rn
the claim. Summarizing, we have {gε ∈ L1 (Rn )∩L2 (Rn )}ε>0 such that limε→0 ∥gε − fˆ∥2 = 0.
By definition of the Fourier transform in L2 we readily get fˆˆ is the L2 -limit of ĝε ; or, fˆˆ(−x)
is the L2 -limit of ĝε (−x) which equals the lhs of (∗). Passing to new subsequence we get that
lhs(∗) converges to fˆˆ(−x) for L n -a.e. x ∈ Rn .
Chapter 8
In this chapter, after proving in detail the Riesz representation theorem, we will combine
it with the Hahn-Banach theorem from functional analysis, to produce two more important
examples of Radon measures:
We assume all students are familiar with the notions of a linear (or vector) space and also
linear subspace of a linear space.
For the proof of the following fundamental theorem of functional analysis, we refer to the
first two pages from the book of H. Brezis; [Br].
75
76 CHAPTER 8. MORE EXAMPLES OF RADON MEASURES
F(λ ) ≤ p(λ ) ∀ λ ∈ Λ′ .
F (λ ) ≤ p(λ ) ∀ λ ∈ Λ,
Accepting for the moment that µ is a Radon measure and that λ is positively homogeneous,
nondecreasing and additive, we show firstly that
Z
λ(f) = f dµ for all f ∈ Cc+ (Rn ). (8.1.2)
Rn
1 equivalently: sup ℓ( f ) f ∈ Cc (Ω; Rm ), | f | ≤ 1, sprt( f ) ⊆ K < ∞, for all compact K ⊂ Ω
8.1. THE RIESZ REPRESENTATION THEOREM 77
So pick f ∈ Cc+ (Rn ) and let ε > 0. Consider a partition Nj=1 [t j−1 ,t j ] of the interval [0, 2M],
S
where
M := max | f |,
Rn
and such that 0 < t j − t j−1 < ε. Since f −1 {t} ≡ {x ∈ Rn f (x) = t}, t ∈ [0, ∞), is a
family of disjoint
Borel sets and µ is a Radon measure, by lemma 2.3.10 we may assume that
µ f −1 {t j } = 0 for all j = 1, ..., N. Then by the subadditivity of µ we have
N
f −1 {ti } .
[
µ(Z) = 0, where Z := (8.1.3)
j=1
Set
U j := f −1 (t j−1 ,t j ) ≡ {x ∈ Rn t j−1 < f (x) < t j },
j = 1, ..., N.
Obviously:
N
[
the sets Z,U1 , ...,UN are disjoint and { f > 0} = Z ∪ Uj . (8.1.4)
j=1
N
µ(A) = ∑ µ U j \ {h j = 1} ( by (8.1.4) and (8.1.3) )
j=1
N
≤ ∑µ Uj \ Kj ( because K j ⊆ {h j = 1} )
j=1
≤ε ( from (b) ).
Also, f 1 − ∑Nj=1 h j ≤ MχA and we claim that ℓ f (1 − ∑Nj=1 h j ) ≤ Mµ(A). Indeed,
N
≤ sup |ℓ(g)| g ∈ Cc (Rn ; Rm ), |g| ≤ MχA
ℓ f 1− ∑ hj
j=1
= M sup |ℓ(g)| g ∈ Cc (Rn ; Rm ), |g| ≤ χA ,
because ℓ(cg) = cℓ(g) for any c ≥ 0. Next note that since A is open there holds
{g ∈ Cc (Rn ; Rm ), |g| ≤ χA } = {g ∈ Cc (Rn ; Rm ), |g| ≤ 1, sprt(g) ⊂ A}.
We deduce
N
ℓ f 1− ∑ hj ≤ Mµ(A).
j=1
hence, taking into account all the underlined properties of λ and (8.1.5), we obtain
N N
λ f ∑ h j ≤ λ ( f ) ≤ εM + λ f ∑ h j ⇒
j=1 j=1
N N
∑ λ ( f h j ) ≤ λ ( f ) ≤ εM + ∑ λ ( f h j ) ⇒
j=1 j=1
N N
∑ t j−1 µ(U j ) − ε/N ≤ λ ( f ) ≤ εM + ∑ t j µ(U j ) ⇒
j=1 j=1
N N
∑ t j−1 µ(U j ) − 2εM ≤ λ ( f ) ≤ εM + ∑ t j µ(U j ), (8.1.6)
j=1 j=1
8.2. POSITIVE LINEAR FUNCTIONALS AND DISTRIBUTIONS 79
where in the last inequality we used t j−1 ≤ tN = 2M for all j = 1, ..., N. Next we estimate
R
Rn f dµ. Because of (8.1.4) we have
Z N Z
f dµ = ∑ f dµ ⇒
Rn j=1 U j
N Z N
∑ t j−1 µ(U j ) ≤ Rn
f dµ ≤ ∑ t j µ(U j ). (8.1.7)
j=1 j=1
from the fact that t j − t j−1 < ε and (8.1.4). This proves (8.1.2) since ε > 0 is arbitrary. □
...
Definition 8.2.1. We denote by D(U) the space consisting all functions in Cc∞ (U) accom-
panied with the following notion of convergence: A sequence {φm ∈ Cc∞ (U)}m∈N is said to
converge to φ ∈ Cc∞ (U) in D(U) (and write φm → φ in D(U)) whenever
F(λ ) = ℓ(λ ), λ ∈ Λ′ ,
to get that ℓ can be extended to a linear mapping (that we don’t rename) from Cc (U) into R,
so that
ℓ( f ) ≤ ℓ(g) max | f | ∀ f ∈ Cc (U), sprt( f ) ⊆ K.
U
In other words,
sup ℓ( f ) f ∈ Cc (U; R), | f | ≤ 1, sprt( f ) ⊆ K < ∞,
for any compact K ⊂ U. Hence, the Riesz representation theorem with m = 1 applies to get
a Radon measure µ on U and a µ-measurable function σ : U → R such that σ = ±1 µ-a.e.
8.3. WEAK DERIVATIVES OF FUNCTIONS OF BOUNDED VARIATION 81
of positive µ-measure and, given ε > 0, employ the outer and inner regularity properties of
Radon measures (theorems 2.3.7 and 2.3.8), to get open O ⊃ A and compact K ⊂ A such that
Hence µ(O\K) < ε. Further, from lemma 1.2.6 plus standard mollification (see remark 1.2.8)
we get a Cc∞R(U) function f : U →R
[0, 1] with f R= 1 on K and sprt( f ) ⊂ O. The hypothesis
that ℓ( f ) = U f σ dµ ≥ 0, gives O\K f σ dµ + K f σ dµ ≥ 0. By the properties of f and σ ,
this readily implies µ(K) < ε. Since ε is arbitrary, (8.2.1) gives µ(A) = 0 and the proof is
complete.
To this end, let us write S for the supremum on the right hand side of (8.3.1). Given g ∈
Cc1 (U; Rn ) we may integrate by parts to get
Z Z Z
f (x) div g(x) dx = − ∇ f (x) · g(x) dx ≤ |∇ f (x) · g(x)| dx.
U U U
Hence in particular Z
S≥ |∇ f (x)| dx.
K
We have shown nZ o
′
S ≥ sup |∇ f (x)| dx K ⊂ U , K compact .
K
It remains to apply Theorem 2.3.8 to get
nZ o Z Z
′
sup |∇ f (x)| dx K ⊂ U , K compact = |∇ f (x)| dx = |∇ f (x)| dx,
K U′ U
Since the right hand side of (8.3.1) makes sense if f is merely locally L n -summable on
U, defining the quantity
Z nZ o
n 1 n
|D f | := sup f div g dL g ∈ Cc (V ; R ), |g| ≤ 1 in V , V ⊆ U,
V V
we introduce:
Theorem 8.3.3. For any f ∈ BVloc (U), there exists a Radon measure µ on U and a µ-
measurable function σ : U → Rn such that
Proof.
...
Chapter 9
Elementary convexity
In case the reverse inequality holds true in (9.1.1), then f is called concave.
Remark 9.1.2. Let −∞ ≤ a < b ≤ ∞. The mean value theorem shows that if f : (a, b) → R
has nonnegative second derivative at any point of (a, b), then f is convex in (a, b).
Remark 9.1.3. Using induction we obtain for any k ∈ N the equivalent inequality
k k
f λ x ≤
∑ i i ∑ λi f (xi),
i=1 i=1
valid for all λi ∈ [0, 1], i = 1, ..., k, satisfying ∑ki=1 λi = 1, and all xi ∈ Ω, i = 1, ..., k. For
instance, for k = 3 we have
!
λ λ
2 3
f (λ1 x1 + λ2 x2 + λ3 x3 ) = f λ1 x1 + (λ2 + λ3 ) x2 + x3
λ2 + λ3 λ2 + λ3
λ λ3
2
≤ λ1 f (x1 ) + (λ2 + λ3 ) f x2 + x3
λ2 + λ3 λ2 + λ3
λ λ3
2
≤ λ1 f (x1 ) + (λ2 + λ3 ) f (x2 ) + f (x3 )
λ2 + λ3 λ2 + λ3
≤ λ1 f (x1 ) + λ2 f (x2 ) + λ3 f (x3 ).
83
84 CHAPTER 9. ELEMENTARY CONVEXITY
Remark 9.1.4. The arithmetic-geometric inequality of lemma 3.0.5 follows from the above
remark by taking k = n, λi = 1/n, xi = log ai and f (t) = et , t ∈ R.
Remark 9.1.5. To show the claim made in the proof of theorem 5.1.9; that is,
we observe first the function (0, ∞) ∋ t 7→ t p is convex in case p > 1. Hence, for A, B ∈ R and
λ ∈ (0, 1),
|A| |B| p
p p
|A + B| ≤ (|A| + |B|) = λ + (1 − λ ) ≤ λ 1−p |A| p + (1 − λ )1−p |B| p .
λ 1−λ
Since (9.1.2) is also true if ε ≥ 1, the claim follows. For 0 < p ≤ 1 observe instead
p
|a| + |b| ≤ |a| p + |b| p .
0 if x ∈ A,
IA (x) :=
∞ if else.
1−λ
(ii) Prove L n λ X + (1 − λ )Y ≥ L n (X) L n (Y )
λ
.
(iii) Prove that if the inequality of (ii) is known to be true for all λ ∈ [0, 1] and all X,Y ⊆ Rn ,
then it implies the Brunn-Minkowski inequality.
9.1. CONVEX FUNCTIONS 85
where we have used the triangle inequality to get to the middle line.
Then zx ∈ Brx (x), zy ∈ Bry (y) and also λ zx + (1 − λ )zy = z. Thus z is in the Minkowski
sum of the sets λ Brx (x) and (1 − λ )Bry (y). Since Brx (x), Bry (y) ⊂ Ω and Ω is convex,
we conclude z ∈ Ω. □
86 CHAPTER 9. ELEMENTARY CONVEXITY
Proof. Taking
1
Z
x := u dµ, y := u(z), z ∈ X,
µ(X) X
The proof now follows by taking mean values on X with respect to µ on both sides.
Exercise 9.1.13. Let µ be a measure on the set X and µ(X) = 1. Let u : X → [0, ∞] be
µ-summable.
R
(i) If Xu dµ = 1, prove
Z
u log u dµ ≥ 0.
X
1 Z p
R 1
≤ u dµ .
X up dµ X
(iii) Prove r Z 2 Z p Z
1+ u dµ ≤ 1 + u2 dµ ≤ 1 + u dµ.
X X X
88 CHAPTER 9. ELEMENTARY CONVEXITY
(ii) real functions in C1 (Ω) having uniformly bounded derivative in Ω, with Ω being any
open and convex subset of Rn .
Proof. Let f : Ω → R, f ∈ C1 (Ω), for which there exists M > 0 such that
|∇ f (x)| ≤ M ∀ x ∈ Ω.
Let x, y ∈ Ω and consider the straight line γ(t) := x + t(y − x), t ∈ [0, 1]. Applying
the
mean value theorem to the function u : [0, 1] → R defined by u(t) := f γ(t) , we get
t0 ∈ (0, 1) such that u′ (t0 ) = u(1) − u(0). By the chain rule we deduce
D f γ(t0 ) · (y − x) = f (y) − f (x).
Coupling this with the bound on the derivative of f we conclude that f is Lipschitz
with L = M.
Exercise 9.2.4. Let Ω ⊊ Rn be open and suppose that f : Ω̄ → R is Lipschitz with f = 0 on
∂ Ω. Prove that there exists a positive constant C such that
| f (x) − f (y)| ≤ CK |x − y| ∀ x, y ∈ K.
Example 9.2.7. Any convex function f : Ω → R, where Ω ⊆ Rn is open and convex, is locally
Lipschitz.
| f (y) − f (x)|
≤ max{|C1 |, |C2 |} for all x, y ∈ K, x ̸= y.
|y − x|
The right hand side is always finite under the above construction. Indeed, the function
| f (y) − f (x)|
F(x, y) := x ∈ ∂ X ′ , y ∈ ∂Y ′ ,
|y − x|
90 CHAPTER 9. ELEMENTARY CONVEXITY
For instructive reasons we provide a second proof for the statement in the above example.
It is based on the following fact1
Lemma 9.2.8. Let Q be an interval in Rn and y1 , . . . , y2n the vertices. Then every point x ∈ Q
can be written as a convex linear combination of the vertices, i.e. there exist λi ∈ [0, 1],
i = 1, . . . , 2n , such that
2n 2n
x = ∑ λi yi , ∑ λi = 1.
i=1 i=1
Proof. The proof is done by induction on the space dimension n ∈ N:
For n = 1 the statement is trivial. Suppose now the claim holds true for n − 1. Let
x = (x1 , ..., xn ) ∈ Q = [l11 , l12 ] × [l21 , l22 ] × ... × [ln1 , ln2 ]. Further, we denote by A the set of all
vertices having their last coordinate equal to ln1 and by B the set of all vertices having their
last coordinate equal to ln2 . Observe that A and B are disjoint, card(A) = card(B) = 2n−1 and
their union is the whole set of vertices of Q. Hence we can define a bijection
(
A→B
ϕ:
(z1 , z2 , ..., zn−1 , ln1 ) → (z1 , z2 , ..., zn−1 , ln2 ).
Naturally, we obtain for each ai ∈ A, i = 1, ..., 2n−1 , a curve γi (t) = tai + (1 − t)ϕ(ai ) for
t ∈ [0, 1]. Furthermore, let H = {z = (z1 , ..., zn ) ∈ Rn zn = xn } be a hyperplane. Since γi is
continuous for each i = 1, ..., 2n−1 and all curves are identical in the last coordinate, we obtain
a t ∗ such that the last coordinate of γi (t ∗ ) is equal to xn for all i. Next we observe the points
{γ1 (t ∗ ), ..., γ2n−1 (t ∗ )} lie in the hyperplane H and generate an interval in this hyperplane. By
our induction hypotheses we can write the point (x1 , ..., xn−1 ) as a convex linear combination
of the points {γ1 (t ∗ ), ..., γ2n−1 (t ∗ )} restricted to the hyperplane H. But since the linear com-
bination is convex and {γ1 (t ∗ ), ..., γ2n−1 (t ∗ )} lie in the hyperplane H, we are allowed to write
(x1 , ..., xn ) as a linear combination of {γ1 (t ∗ ), ..., γ2n−1 (t ∗ )} in Rn . Finally, observe that the
points {γ1 (t ∗ ), ..., γ2n−1 (t ∗ )} are convex linear combinations of our 2n vertices.
Second proof for the statement in example 9.2.7. Let Ql (y0 ) ⊆ Ω be a cube of side length
l > 0, center at y0 and vertices y1 , y2 , ..., y2n . Set V := {y1 , y2 , ..., y2n }. If x ∈ Ql (y0 ) then
1 this is a primitive version of the Krein-Milman theorem in Functional Analysis; see [Br, Theorem 1.13]
9.3. A CHARACTERIZATION OF CONVEX FUNCTIONS 91
n
from the previous lemma we know there exists λi ∈ [0, 1], i = 1, ..., 2n , with ∑2k=1 λk = 1 and
n
x = ∑2k=1 λk yk . The convexity of f implies through exercise 9.1.3 that
2n
f (x) ≤ ∑ λk f (yk ) ≤ max | f | < ∞.
V
k=1
Thus M := supQl f < ∞. Since y0 = (1/2)x + (1/2)(2y0 − x) and 2y0 − x ∈ Ql (y0 ) (why?),
applying the convexity property of f gives
1 1 1 1
f (y0 ) ≤ f (x) + f (2y0 − x) ≤ f (x) + M.
2 2 2 2
This gives infQl f ≥ 2 f (y0 ) − M. Hence f is bounded in Ql . It follows that f is locally
bounded in Ω (why?). Next let B̄3r (x0 ) ⊂ Ω and x, y be two distinct points in B̄r (x0 ). Select
µ > 0 and z ∈ ∂ B2r (x) so that
z − x = µ(y − x).
Then µ = 2r/|y − x| > 1 and y = (1/µ)z + (1 − 1/µ)x. Hence, the convexity of f gives
Proof. If f is convex then (9.3.1) is true. Also, by the above example and remark we get f is
continuous. To prove the other direction let x, y ∈ Ω and define γ : [0, 1] → Ω by
γ(λ ) := λ x + (1 − λ )y, λ ∈ [0, 1].
Define the sequence of sets
Dk := γ(λ ) λ ∈ {0/2k , 1/2k , ..., 2k /2k } , k ∈ N.
Claim. There holds f γ(λ ) ≤ λ f (x) + (1 − λ ) f (y) for all γ(λ ) ∈ Dk .
Accepting the claim for a moment, we can use k∈N Dk is dense in {γ(λ ) : λ ∈
S
[0, 1]}
and the continuity of u to deduce (9.1.1). Indeed, for any λ ∈ [0, 1], there exists λk ∈
{0/2k , 1/2k , ..., 2k /2k } k∈N such that limk→∞ λk = λ . Then limk→∞ γ(λk ) = γ(λ ) and the
continuity of f implies through theorem 1.2.3 that f γ(λ ) = limk→∞ f γ(λk ) . From the
claim we further obtain
f γ(λ ) ≤ lim λk f (x) + (1 − λk ) f (y) ,
k→∞
Example 9.3.2. The function f : Rn → R given by f (x) = |x|2 − dF2 (x), where dF is the
distance function to a nonempty closed subset F ⊊ Rn , is convex.
Proof. From example 9.2.3-(i) we know f is continuous, hence because of the above theorem
we only have to prove f (x/2 + y/2) ≤ f (x)/2 + f (y)/2. To this end pick z ∈ F such that
x + y x+y
dF = z− .
2 2
Then by expanding the square
x + y
f = −|z|2 + z · (x + y).
2
On the other hand
f (x) + f (y) |x|2 + |y|2 − dF2 (x) − dF2 (y)
=
2 2
|x|2 + |y|2 − |x − z|2 − |y − z|2
≥
2
2
= −|z| + z · (x + y),
as required.
94 CHAPTER 9. ELEMENTARY CONVEXITY
Bibliography
[Br] Brezis, H. Functional analysis, Sobolev spaces and partial differential equations. Univer-
sitext. Springer 2011.
[DB] DiBenedetto, E. Real analysis. 2nd ed. Birkhäuser Adv. Texts. Basler Lehrbücher.
Birkhäuser 2016.
[EvG] Evans, L. C., Gariepy, R. F. Measure theory and fine properties of functions. Stud. Adv.
Math. CRC Press 1992.
[F] Federer, H. Geometric measure theory. Grundlehren Math. Wiss. 153, Springer 1969.
[Fr] Friedlander, F.G. Introduction to the theory of distributions. 2nd ed. with additional material
by M. Joshi. Cambridge Univ. Press 1998.
[KrP] Krantz, S., Parks H. The geometry of domains in space. Birkhäuser Adv. Texts. Basler
Lehrbücher. Birkhäuser 1999.
[LL] Lieb, E. H., Loss, M. Analysis. 2nd ed. Grad. Stud. Math. 14. Amer. Math. Soc. 2001.
[Mg] Maggi, F. Sets of finite perimeter and geometric variational problems. An introduction to
geometric measure theory. Cambridge Stud. Adv. Math. 135. Cambridge Univ. Press 2012.
[Mz] Maz’ya, V. G. Sobolev spaces. Translated from Russian by T. Shaposhnikova. Springer Ser.
Soviet Math., Springer 1985.
[Schn] Schneider, R. Convex Bodies: The Brunn-Minkowski theory. 2nd expanded ed. Encyclope-
dia Math. Appl. 151. Cambridge Univ. Press 2014.
[V] Villani, C. Topics in Optimal Transportation. 2nd ed. Grad. Stud. Math. 58. Amer. Math.
Soc. 2016.
[WhZ] Wheeden, R. L.; Zygmund, A. Measure and integral. An introduction to real analysis. Pure
Appl. Math. 43. Marcel Dekker 1977.
95
96 BIBLIOGRAPHY
[Z] Ziemer, W. P. Modern real analysis. 2nd ed. with contributions by Monica Torres. Grad.
Texts in Math. 278. Springer 2017.