Functional Analysis.
Functional Analysis.
MEAS Len
Department of Mathematics
Royal University of Phnom Penh
2022
Outline
1 Part I: Foundations
Metric Space
Banach Space
Dual Space
Hilbert Space
2 Part II: Principle of Functional Analysis
Baire’s Category Theorem
Uniform Boundedness Principle
Open Mapping Theorem
Closed Graph Theorem
Hahn-Banach Theorem
3 Part III: Lebesgue Integral
Measurable functions
Lebesgue Integral
The Lp Spaces
Part I: Foundations Metric Space
Metric Space
Definition
A metric space is a pair (X, d) consisting of a set X and a function
d : X × X → R that satisfies the following axioms:
M1 d(x, y) ≥ 0, ∀x, y ∈ X, with equality if and only if x = y.
M2 d(x, y) = d(y, x), ∀x, y ∈ X.
M3 d(x, z) ≤ d(x, y) + d(y, z), ∀x, y, z ∈ X.
Remark:
d : X × X → R : “distance function”.
(M3 ): triangle inequality.
Easy Consequences:
1 d(x, z) ≥ |d(x, y) − d(z, y)|
2 d(x1 , xn ) ≤ d(x1 , x2 ) + · · · + d(xn−1 , xn ).
Definition
A sequence (xn ) in a metric space (X, d) converges to x ∈ X if for every ε > 0,
there exists N ∈ N such that
Notation: lim xn = x or xn → x.
n→∞
A sequence (xn ) in (X, d) is a Cauchy sequence if, for every ε > 0, there exists
N ∈ N such that
d(xm , xn ) < ε, for all m, n ≥ N.
Theorem
Suppose that (xn ) is a convergent sequence in a metric space (X, d). Then
1 the limit x = lim xn is unique
n→∞
2 any subsequence of (xn ) also converges to x
3 (xn ) is a Cauchy sequence.
Proof:
1 suppose lim xn = l and lim xn = m. Let ε > 0. Then for n large enough
n→∞ n→∞
d(l, m) ≤ d(l, xn ) + d(xn , m) < 2ε + 2ε = ε.
Thus 0 ≤ d(l, m) ≤ ε holds for any ε > 0. It follows that d(l, m) = 0, i.e,
l = m.
Proposition
In a metric space, a sequence (xn ) can only converge to one limit.
∃N1 , n ≥ N1 ⇒ xn ∈ Br (x)
∃N2 , n ≥ N2 ⇒ xn ∈ Br (y).
For n ≥ max(N1 , N2 ) this would result in xn ∈ Br (x) ∩ Br (y) = ∅, a
contradiction.
Proposition
Distinct points of a metric space can be separated by disjoint balls,
Proof: If x , y then d(x, y) > 0. Let r := d(x, y)/2, then Br (x) is disjoint from
Br (y) else we get a contradiction: z ∈ Br (x) ∩ Br (y)
Examples:
1 In any metric space, xn → x ⇔ d(xn , x) → 0 as n → ∞ (because
xn ∈ Bε (x) ⇔ d(xn , x) < ε).
2 n
In R, n+1 → 1 as n → ∞, since for any ε, there is an N such that 1
N <ε
(Archimedean property of R), so
n 1 1
n≥N ⇒ 1− = < < ε.
n+1 n+1 N
Continuity
Definition
A function f : X → Y between metric spaces is continuous when it preserves
convergence,
xn → x in X ⇒ f (xn ) → f (x) in Y.
Theorem
A function f : X → Y between metric spaces. TFAE
i f is continuous
ii ∀ε > 0, ∃δ > 0, ∀x ∈ X, dX (x, x0 ) < δ ⇒ dY (f (x), f (x0 )) < ε,
iii For every open set V in Y, f −1 (V) is open in X.
iv For every closed subset F in Y, f −1 (F) is closed in X.
f (x) ∈ Bε (f (x)) ⊆ V,
and so
∃δ > 0, f (Bδ (x)) ⊆ Bε (f (x)) ⊆ V.
In other words, x is an interior point of U, ∃δ > 0, Bδ (x) ⊆ f −1 (V) ⊆ U.
(iii) ⇒ (i): Let (xn ) be a sequence converging to x. Consider any open
neighborhood Bε (f (x)) of f (x). Then f −1 (Bε (f (x)) contains x and is open set by
(iii), so
Remark:
1 Continuous functions preserve convergence, a central concept in metric
space.
2 A homeomorphism between metric spaces X and Y is a mapping
f : X → Y such that f is bijective, f is continuous and f −1 is continuous.
3 A metric space X is said to be embedded in another space Y, when there
is a subset Z ⊆ Y such that X is homeomorphic to Z.
When X and Y are homeomorphic, they are not only the same as a sets
(bijection) but also wrt convergence
xn → x ⇐⇒ f (xn ) → f (x),
and
A is open in X ⇐⇒ f (A) is open in Y.
The most vivid picture is that of “deforming” one space continuously and
reversibly form the other. The classic example is that a ‘teacup’ is
homeomorphic to a “doughnut”.
Definition
A metric space is complete when every Cauchy sequence in it converges.
Theorem
The real number R is complete.
The proof of this theorem follows the Cauchy criterion in real analysis states
that a sequence converges if and only if it is Cauchy.
Proposition
(Q, d), d(x, y) = |x − y| is NOT complete.
√
Why? Consider a sequence rationals xn converging to 2 in (R, d),
specifically, we may assume
√ 1
xn − 2 ≤
n
√
(This is possible since there are rationals number arbitrarily close to 2).
Indeed, we can use recurrence x1 = 1 and
xn−1 1
xn = + ,
2 xn−1
√
for n > 1. As (xn ) converges, then it is Cauchy. But its limits 2 < Q.
Proposition
Let X be a complete metric space. A subset F ⊆ X is complete ⇐⇒ F is
closed in X.
Proof: Let F ⊆ X be complete, i.e any Cauchy sequence in F converges to a
limit in F. Let x ∈ F̄, with a sequence xn → x, xn ∈ F. Since convergent
sequence are Cauchy and F is complete, x must be in F. Thus F = F̄ is closed.
Conversely, let F be a closed in X and let (xn ) be a Cauchy sequence in F.
Then (xn ) is a Cauchy sequence in X, which is complete.
Therefore, xn → x for some x ∈ X. In fact x ∈ F̄ = F. Thus any Cauchy
sequence of F converges in F.
Remark: In fact, a complete subspace of any metric space is closed.
Remark: Two metric spaces may be homeomorphic yet one space be
complete and the other not. For example: R is homeomorphic to ]0, 1[. But
]0, 1[ is NOT closed in R, hence NOT complete.
Uniformly Continuity
Remark:
1 The difference from continuity is that, here, δ is independent of x.
2 Uniformly continuity ⇒ continuity.
1
3 But not every continuously map is uniformly. For example, f (x) := x on
]0, ∞[.
Proposition
A uniformly continuous function maps any Cauchy sequence to a Cauchy
sequence.
Example: Lipschitz maps are uniformly continuous: since for any ε > 0, we
ε
can let δ = 2c independent of x to obtain d(x, x0 ) < δ ⇒ d(f (x), f (x0 )) < cδ < ε.
Remark: NOT every uniformly continuous function is Lipschitz. For
√
example, x on √ [0, 1] is uniformly continuous. If it were Lipschitz, it would
√ √
satisfy | x − 0| ≤ c|x − 0| which leads to x ≥ 1c .
Definition
A function f : X → Y is Lipschitz map when
Proof: Consider the iteration xn+1 := f (xn ) starting with any x0 in X. Note that
Hence, by induction on n,
d(xn+1 , xn ) ≤ cn d(x1 , x0 ).
cn
Moreover, the rate of convergence is given at least by d(x, xn ) ≤ 1−c d(x1 , x0 ).
Suppose there are two fixed points x = f (x) and y = f (y), then
Connectedness
Definition
A subset C of a metric space is disconnected when it can be divided into (at
least) two disjoint non-empty subsets C = A ∪ B such that each subset is
converd exclusively by an open set. i.e,
A ⊆ U, B ∩ U = ∅, U open
B ⊆ V, A ∩ V = ∅, V open.
Theorem
The connected subsets of R are intervals.
Proposition
Continuous functions map connected sets to connected sets.
f : X → Y continuous, C ⊆ X connected ⇒ f (C) is connected.
Compactness
Definition
A set K is compact if any open cover has a finite subcover, i.e.,
[ iN
[
K⊆ Ai ⇒ ∃i1 , · · · , iN , K = Ai
i i=i1
Proposition
Compact sets are closed.
Therefore, x ∈ Br (x) ⊆ X \ K.
Proposition
i A closed subset of compact set is compact.
ii A finite union of compact set is compact.
Proof: (i) Let F be a closed subset of a compact set K, and let the open set Ai
cover F, then [
K ⊆ F ∪ (X \ F) ⊆ Ai ∪ (X \ F).
i
But K is compact and therefore a finite number of these open sets are enough
to cover it
[N [N
K⊆ Ai ∪ (X \ F), so F = Ai .
i=1 i=1
(ii) Let the open sets Ai cover the finite union of compact sets K1 ∪ · · · ∪ KN .
Then they cover each individual Kn and a finite number will then suffice in
each case,
kn
[
Kn ⊆ Aik .
k=1
Proposition
Continuous functions map compact sets to compact sets. That is,
f : K ⊂ X → Y is continuous and K is compact, then f (K) is compact.
Proof
Let the sets Ai be an open cover for f (K)
[
f (K) ⊆ Ai .
i
Heine-Borel Theorem
A set K ⊆ Rn is compact if and only if K is closed and bounded.
Let (X, dx ) and (Y, dY ) be metric spaces and assume that X is compact. Then
the space
C(X, Y) := {f : X → Y | f is continuous}
of continuous function from X to Y is a metric space with distance
X→R
x 7→ dY (f (x), g(x))
ε
So given any ε > 0, ∃N such that dY (fn (x), fm (x)) < 2 for any n, m ≥ N and any
x ∈ X.
dY (fn (x), f (x)) ≤ dY (fn (x), fm (x)) + dY (fm (x), f (x)) < ε
Equicontinuity
A subset F ⊂ C(X, Y) is called equicontinuous if for every ε > 0, there exists
a constant δ > 0 such that for all x, x0 ∈ X and all f ∈ F ,
Arzéla-Ascoli Theorem
Let (X, d) be a compact metric space and let F ⊂ C(X, Rn ). Then F is
compact if and only if it is closed, bounded and equicontinuous.
Normed space
Consequences:
1 kx − yk ≥ |kxk − kyk| .
2 kx1 + · · · + xn k ≤ kx1 k + · · · + kxn k.
Banach space
Definition
X is a Banach space if the metric space (X, d) is complete; i.e, if every Cauchy
sequence in X converges, where
d(x, y) := kx − yk
defines a distance on X.
n 1/p
X
kxkp := |xi |p , x ∈ lp .
i=1
where f : M → R measurable.
For p = ∞, (L∞ (µ), k.k∞ ) is a Banach space of bounded measurable
functions with
Remark:
1 The case X = Rn and µ equal Lebesgue measure. Then we have
Z !1/p
kf kp = p
|f (x)| dx .
Rn
2 The case X = N and µ equal to the counting measure. Then Lp and lp are
coincide.
MEAS Len Functional Analysis
Part I: Foundations Banach Space
Theorem
Let (X, k.kX ) and (Y, k.kY ) be real normed spaces and let T : X → Y be a linear
operator. TFAE
1 T is bounded
2 T is continuous
3 T is continuous at x = 0.
Proof
(1) ⇒ (2). If T is bounded, then
kTxk ≤ kTkkxk.
Tn (x + y) = Tn x + Tn y, Tn (λx) = λTn x
then
T(x + y) = Tx + Ty, T(λx) = λTx
by continuity of addition and scalar multiplication.
MEAS Len Functional Analysis
Part I: Foundations Banach Space
kTn − Tk ≤ 2
Definition
Let X be a real vector space. Two norms k.k and k.k0 on X are equivalent if
there is a constant c ≥ 1 such that
1
kxk ≤ kxk0 ≤ ckxk for all x ∈ X.
c
Theorem
Let X be a finite dimensional real vector space. Then any two norms on X are
equivalent.
Proof
Step 2: There is a constant δ > 0 such that δkxk2 ≤ kxk for all x ∈ X.
The set S := {x ∈ X | kxk2 = 1} is compact wrt to k.k2 by the Heine-Borel
theorem and the function S → R : x 7→ kxk is continuous. Hence there exists
an element x0 ∈ S such that kx0 k ≤ kxk for all x ∈ S. Define δ := kx0 k > 0. The
2 x ∈ S, hence kxk2 x ≥ δ and so
every nonzero vector x ∈ X satisfies kxk−1 −1
kxk ≥ δkxk2 .
Corollary
Every finite dimensional normed space is complete.
Proof: This holds for Rn for Euclidean norm. By theorem it holds for every
norms on Rn . Thus it holds for every finite dimensional normed spaces.
Corollary
Every finite dimensional linear subspace of normed space is closed.
Corollary
Let (X, k.k) be a finite-dimensional normed space and let K ⊂ X. Then K is
compact if and only if K is closed and bounded.
Theorem
Let (X, k.k) be a normed space and denote the closed unit ball and the closed
unit sphere in X by
Proof
(1) ⇒ (2).
Let X be a normed space with dim X = n and B be its closed unit ball. Then X
is homeomorphic to Rn , so let T : X → Rn be the homeomorphism (by
definition T and T −1 are continuous). Since B is bounded set in X and T is
bounded linear operator, then T(B) is bounded. So T(B) is bounded and
closed since T −1 is continuous (by the fact that the inverse images of open sets
under a continuous function are open and similarly for closed sets). Hence
T(B) is a closed and bounded subset of Rn and so T(B) is compact. Since T −1
is continuous then T −1 (T(B)) = B is compact.
(2) ⇒ (3).
It follows from the fact that a closed subset of a compact set in a topological
space is compact.
(3) ⇒ (1).
We argue indirectly and show that if X is infinite-dimensional then S is not
compact. Thus assume X is infinite-dimensional. We claim that there exists a
sequence xi ∈ X such that kxi k = 1, kxi − xj k ≥ 12 for all i, j ∈ N with i , j (*).
This is then a sequence in S that does not have any convergent subsequence
and so it follows that S is not compact.
To prove the existence of a sequence in X satisfying (*) we argue by
induction. For i = 1, choose any element xi ∈ S. If x1 , . . . , xk ∈ S satisfy
kxi − xj k ≥ 21 for i , j, consider the subspace Y ⊂ X spanned by the vectors
x1 , . . . , xk . This is a closed subspace of X by corollary and is not equal to X as
dim X = ∞. Hence Riesz Lemma asserts that there exists a vector x = xk+1 ∈ S
such that kx − yk ≥ 21 for all y ∈ Y and hence in particular, kxk+1 − xi k ≥ 12 for
i = 1, . . . , k. This shows that there exists a sequence xi ∈ X that satisfies (*)
for i , j.
Riesz Lemma
Let (X, k.k) be a normed space and let Y ⊂ X be a closed linear subspace that
is not equal to X. Fix a constant 0 < δ < 1. Then there exists a vector x ∈ X
such that
kxk = 1, inf kx − yk ≥ 1 − δ.
y∈Y
x0 − y0 − kx0 − y0 ky d
kx − yk = ≥ ≥1−δ
kx0 − y0 k kx0 − y0 k
for all y ∈ Y.
Theorem
Let X be a normed space and let Y be a Banach space. Then L(X, Y) is a
Banach space with respect to the operator norm.
for all x ∈ X and all m, n ∈ N. Hence (Tn x)n∈N is a Cauchy sequence in Y for
every x ∈ X. Since Y is complete, this implies that the limit Tx := limn→∞ Tn x
exists for all x ∈ X. This defines a map T : X → Y.
T is linear since
and
This implies
lim kT − Tn k = 0.
n→∞
Dual Space
Definition
Let X be a normed space over R (or C). The space L(X, R) (or L(X, C)) is
called the dual space of X and is denoted by X ∗ .
Remark:
1 X ∗ = L(X, R) ≡ space of bounded (continuous) linear functionals on X.
2 By theorem, X ∗ = L(X, R) is a Banach spaces for every normed space X.
Let (M, F , µ) be a measure space and fix a constant 1 < p < ∞. Define the
number 1 < q < ∞ by
1 1
+ = 1.
p q
The Hölder inequality asserts that f ∈ Lp (µ), g ∈ Lq (µ) then fg ∈ L1 (µ) and
Z
fgdµ ≤ kf kp kgkq .
M
Lq (µ) → Lp (µ)∗
g 7→ Λg
is an isometric isomorphism.
Hilbert Space
H × H → R : (x, y) 7→ hx, yi
Lemma
The inner product and norm satisfy the Cauchy-Schwarz inequality
kx + yk ≤ kxk + kyk
Hence
0 ≤ kη − hξ, ηiξk2 = hη, η − hξ, ηiξi = 1 − hξ, ηi2 .
This implies |hξ, ηi| ≤ 1 and hence
Definition
A Hilbert space is an inner product space which is complete as a metric space.
Proposition
The inner product is continuous.
Do all norms on vector spaces come from inner products and if not, which
property characterizes inner product spaces?
Proposition (Parallelogram Law)
A norm is induced from an inner product if and only if it satisfies, for all x, y,
The statement asserts that the sum of the lengths squared of diagonals of
parallelogram equals that of the sides.
Proof: It follows from adding the identities
Theorem
Let H be a Hilbert space and let K ⊂ H be a nonempty closed convex subset
of H. Then there exists a unique element x0 ∈ K such that kx0 k ≤ kxk for all
x ∈ K.
Proof: Define δ := inf{kxk | x ∈ K} ≥ 0.
Existence:
Chose a sequence (xi ) ⊂ K with limi→∞ kxi k = δ. We prove that (xi ) is a
Cauchy sequence. Let ε > 0. Then there exists an i0 ∈ N such that
i ∈ N, i ≥ i0 , then
ε2
kxi k2 < δ2 + .
4
Let i, j ∈ N such that i ≥ i0 and j ≥ i0 . Then 21 (xi + xj ) ∈ K since K is convex
and hence kxi + xj k ≥ 2δ.
This implies
and therefore x0 = x1 .
Theorem (Riesz)
Let H be a Hilbert space and let Λ : H → R be a bounded linear functional.
Then there exists a unique element y ∈ H such that
Proof:
Existence:
If Λ = 0 then the vector y = 0 satisfies (1). Hence assume Λ , 0 and define
K := {x ∈ H | Λ(x) = 1}.
x ∈ H, Λ(x) = 0 ⇒ hx0 , xi = 0.
This implies
hy, xi = kx0 k−2 hx0 , xi = λ = Λ(x).
Thus y satisfies (1).
We prove (2). Assume that y ∈ H satisfies (1). If y = 0 then Λ = 0 and so
kyk = 0 = kΛk. Hence assume y , 0. Then
Uniqueness:
Assume y, z ∈ H satisfy hy, xi = hz, xi = Λ(x) for all x ∈ H. Then
hy − z, xi = 0 for all x ∈ H.
Take x = y − z to obtain
ky − zk2 = hy − z, y − zi = 0
and so y − z = 0.
Corollary
Let H be a Hilbert space and let E ⊂ H be a closed subspace. Then
H = E ⊕ E⊥ .
The orthogonal space E of H is defined as
Proof: Assume that Åk = ∅ for all k. Then for U ⊂ Ak open, nonempty, U \ Ak
is open, nonempty for any k ∈ N. Then there exists a ball Bε (x) ⊂ U \ Ak with
ε ≤ 1k . Hence we can inductively choose balls Bεk (xk ) such that
1
Bεk (xk ) ⊂ Bεk−1 (xk−1 ) \ Ak and εk ≤ .
k
x = lim xl ∈ X
l→∞
and x ∈ Bεk (xk ) for all k. Since Bεk (xk ) ∩ Ak = ∅, we have that
x ∈ k∈N Ak = X, a contradiction.
S
being intersection of closed sets, are closed and it follows from (∗) that they
form a cover of X.Then Baire’s Category theorem yields that Åk , ∅ for some
k0 and hence there exists a Bε0 (x0 ) ⊂ Ak0 . Noting that
Banach-Steinhaus
Let X be a Banach space and let Y be a normed space. Suppose T ⊂ L(X, Y)
with
sup kTxkY < ∞ for every x ∈ X.
T∈T
i.e.,
{y ∈ Y | kykY < δ} ⊂ {Tx | x ∈ X, kxkX < 1}.
Indeed, let U be open and let x ∈ U. Choose an ε > 0 with Bε (x) ⊂ U.
Now, Bδ (0) ⊂ T(B1 (0)) implies Bεδ (Tx) ⊂ T(Bε (x)) ⊂ T(U) and hence
T(U) is open.
That is, every surjective bounded linear operator between two Banach spaces
is open.
It follows from the Baire’s Category theorem that there exists a k0 and a ball
Bε0 (y0 ) in Y with
Bε0 (y0 ) ⊂ T(Bk (0)).
This means that for y ∈ Bε0 (0) there exists points xi ∈ Bk0 (0) with Txi → y0 + y
as i → ∞. On choosing an x0 ∈ X with Tx0 = y0 , this implies that
!
xi − x0 y xi − x0
T → and < 1,
k0 + kx0 k k0 + kx0 k k0 + kx0 k
However, we want to show that such an inclusion holds without the closure of
the set on RHS, for a smaller δ if necessary. We note that (∗) implies that
y ∈ Bδ (0) implies there exists an x ∈ B1 (0) with y − Tx ∈ Bδ/2 (0). Then
2(y − Tx) ∈ Bδ (0). Hence for y ∈ Bδ (0) we can inductively choose points
yk ∈ Bδ (0) and xk ∈ B1 (0) such that
Then
2−k−1 yk+1 = 2−k yk − T(2−k xk ),
and so m
X
T 2 xk = y − 2−m−1 ym+1 → y as m → ∞.
−k
k=0
Since
m
X m
X
2−k xk < 2−k ≤ 2 < ∞,
k=0 k=0
P
m −k
We have that k=0 2 xk m∈N is a Cauchy sequence in X. As X is complete,
there exists x = k=0 2 xk in X with kxk < 2. The continuity of T
∞ −k
P
then yields
that m
X
Tx = lim T 2−k xk = y.
m→∞
k=0
This shows that Bδ (0) ⊂ T(B2 (0)) or equivalently Bδ/2 ⊂ T(B1 (0)). Hence T is
open.
(⇐) The fact that Bδ (0) ⊂ T(B1 (0)) for some δ > 0 implies that
BR (0) ⊂ T(BR/δ (0)) for all R > 0.
That is, every bounded linear operator between two Banach spaces has a
bounded inverse.
Proof: T −1 is linear. It follows from above theorem that T is open and T −1 is
continuous.
When X and Y are Banach spaces, X × Y is a Banach space space with the
norm k(x, y)k. Indeed, let (zn ) be Cauchy in X × Y, where zn = (xn , yn ). Then
for every ε > 0 there is an N such that for all m, n ≥ N,
Hence (xn ) and (yn ) are Cauchy in X and Y, respectively and converge, say,
xn → x and yn → y since X and Y are complete. This implies that
zn → z = (x, y) since from (∗) with m → ∞ we have kzn − zk < ε for n > N.
Since the Cauchy sequence (zn ) was arbitrary, X × Y is complete.
P : G(T) → D(T)
(x, Tx) 7→ x.
Since G(T) and D(T) are complete, we can apply the inverse mapping
theorem and conclude that P−1 is bounded, say, k(x, Tx)k ≤ bkxk for some b
and all x ∈ D(T). Hence T is bounded because
Hahn-Banach Theorem
p(αx) = |α|p(x)
p(x + y) ≤ p(x) + p(y)
for all x, y ∈ V and for all scalars α. Let W be a subspace of V and let
f : W → R be a linear map such that
Then there exists a linear extension f̃ : V → R of f (i.e. f̃ (x) = f (x) for all
x ∈ W) such that
|f̃ (x)| ≤ p(x) for all x ∈ V.
Proof: Define p(x) = kf kW ∗ kxk. We see that p is defined for all x and satisfies
and
p(x + y) = kf kW ∗ kx + yk ≤ kf kW ∗ (kxk + kyk) = p(x) + p(y).
Hence we can apply the above theorem and conclude that there exists a linear
functional f̃ on V which is an extension of f and satisfies
Hence
kf̃ kV ∗ = sup |f̃ (x)| ≤ kf kW ∗ .
x∈V
kxk=1
Measurable functions
Definition
Let (X, M, µ) be a measure space and let f : X → [−∞, ∞]. We say that f is a
measurable function (or M-measurable) if for every a ∈ R, the set
{x : f (x) > a} is a measurable set (i.e., element of σ-algebra).
Example of measurable
1 Any continuous function f is measurable with measure space (R, L, m)
since the inverse image of any open set is open and L contains all open
subsets of R.
2 Discontinuous function f defined by
x,
1
x,0
f (x) =
0,
x = 0.
Definition
Let (X, M, µ) be a measure space and let E ⊆ X. The characteristic function of
E denoted by 1E , defined by
1, x∈E
1E =
0, x < E.
Theorem
Let (X, M, µ) be a measure space. The function 1E is measurable function
⇔ E is a measurable set. A simple function s is measurable ⇔ Cj is
measurable for j = 1, 2, · · · , n.
Suppose E is measurable.
If a ≤ 0, then {x : 1E (x) < a} = ∅ a measurable set.
If a > 1 then {x : 1E (x) < a} = X a measurable set.
If 0 < a ≤ 1, {x : 1E (x) < a} = X \ E a measurable set.
Suppose 1E is a measurable function. Then E = X \ {x : 1E (x) < 1/2} is a
measurable set.
Theorem
Let (X, M, µ) be a measure space and let f : X → [−∞, ∞]. The following are
equivalent:
a {x : f (x) > a} is a measurable set for every a ∈ R.
b {x : f (x) ≥ a} is a measurable set for every a ∈ R.
c {x : f (x) < a} is a measurable set for every a ∈ R.
d {x : f (x) ≤ a} is a measurable set for every a ∈ R.
∞
(a) ⇔ (b) : {x : f (x) ≥ a} = {x : f (x) > a − 1k } ∈ M.
T
k=1
∞
(b) ⇔ (a) : {x : f (x) > a} = {x : f (x) ≥ a + 1k } ∈ M.
S
k=1
(a) ⇔ (d)&(b) ⇔ (c): complement ∈ M.
Proposition
Let (X, M, µ) be a measure space and let f , g : X → R be measurable
functions, and let c ∈ R. Then |f |, cf , f + g, fg and f /g(g , 0) are measurable.
Proof
We have
X
if a < 0
{x : |f (x)| > a} =
{x : f (x) > a} ∪ {x : f (x) < −a} if a ≥ 0
Since f and g are measurable, the sets {x : f (x) > r} and {x : g(x) > a − r} are
rational for every rational r. Thus the intersection of these sets is measurable
and the countable union is also measurable.
MEAS Len Functional Analysis
Part III: Lebesgue Integral Measurable functions
It follows that
1
fg = [(f + g)2 − f 2 − g2 ]
2
is measurable.
Finally, if g , 0
{x : g(x) < 1/a} if a > 0
{x : 1/g(x) > a} = {x : g(x) > 0} if a = 0
{x : g(x) > 0} ∪ {x : g(x) < 1/a} if a < 0
Proposition
Let f , g be measurable functions. Then max{f , g}, min{f , g} are measurable
functions.
Proof
We have
We define
!
sup fn (x) : = sup{fn (x)}.
n∈N n∈N
inf fn (x) : = inf {fn (x)}.
n∈N n∈N
(lim sup fn )(x) = lim (sup{fk (x) : k ≥ n})
n→∞
= inf (sup{fk (x) : k ≥ n}).
n≥1
(lim inf fn )(x) = lim (inf{fk (x) : k ≥ n})
n→∞
= sup(inf{fk (x) : k ≥ n}).
n≥1
Theorem
Let (fn )n∈N be a sequence of measurable functions. Then
supn∈N fn , inf n∈N fn , lim sup fn , lim inf fn are measurable.
Proof
We have for every a ∈ R that
[ ∞
x : (sup fn )(x) > a = {x : fn (x) > a},
n=1
∞
\
{x : (inf fn )(x) > a} = {x : fn (x) > a},
n=1
Theorem
Let (fn )n∈N be a sequence of measurable functions and fn → f pointwise as
n → ∞. Then f is measurable.
Proof
Since fn → f pointwise, then
which is measurable.
Corollary
If f : X → [−∞, ∞] is a measurable function, then there exists a sequence of
measurable, simple functions (sn ) with lim sn (x) = f (x) for every x ∈ X.
n→∞
Definition
Let (X, M, µ) be a measure space and s is a measurable, nonnegative, simple
function
n
cj 1Cj , Cj = {x : s(x) = cj }.
X
s=
j=1
Remark: Convention: 0 · ∞ = 0.
Example:
if − 1 < x < 2
1,
if 2 ≤ x < 4
2,
f (x) =
3, if 4 ≤ x ≤ 8
0, otherwise.
Z
f (x)dx = 1 · µ((−1, 2)) + 2 · µ([2, 4)) + 3 · µ([4, 8])
R
= 1 × 3 + 2 × 2 + 3 × 4 = 19.
Example
Let µ be Lebesgue measure on R and let ϕ : R → R be given by
ϕ(x) = 2χ[0,2] (x) + 3χ(2,3] (x); that is
2, if 0 ≤ x ≤ 2
ϕ(x) = if 2 < x ≤ 3
3,
0,
otherwise
Then
Z
ϕ(x)dx = 2.µ([0, 2]) + 3.µ((2, 3]) = 2(2 − 0) + 3(3 − 2) = 7.
R
Example
Let µ be Lebesgue measure on R and let ϕ : R → R be given by
ϕ(x) = 2χ[−1,1] (x) + 3χ(3,7) (x) − 1χ[−4,−3) (x); that is
2, if − 1 ≤ x ≤ 1
if 3 < x < 7
3,
ϕ(x) =
−1, if − 4 ≤ x < −3
0, otherwise
Then
Z
ϕ(x)dx = 2.µ([−1, 1]) + 3.µ((3, 7)) − 1µ([−4, −3))
R
= 2.(1 − (−1)) + 3.(7 − 3) + (−1).((−3) − (−4)) = 15.
Properties
If ϕ, ψ are simple, nonnegative functions and if c ≥ 0, then
Z Z
cϕdµ = c ϕdµ,
Z Z Z
(ϕ + ψ)dµ = ϕdµ + ψdµ.
Theorem
Let (X, M, µ) be a measure space and let s be a measurable, nonnegative,
simple function. The set function ϕ : M → [0, ∞] defined by
Z
ϕ(E) = sdµ
E
is a positive measure on M.
Proof:
1 ϕ(∅) = 0 since µ(∅) = 0.
2 To show that ϕ is countably additive. Let (Ak )k≥1 ⊂ M pairwise disjoint
∞ n
sets in M and let A = Ak , s = cj 1Cj .
S P
k=1 j=1
Then
Z n
X
ϕ(A) = sdµ = cj µ(Cj ∩ A)
A j=1
n
∞
X [
= cj µ (Cj ∩ Ak )
j=1 k=1
n
X ∞
X
= cj µ(Cj ∩ Ak )
j=1 k=1
∞
XX n
= cj µ(Cj ∩ Ak )
k=1 j=1
∞ Z
X ∞
X
= sdµ = ϕ(Ak )
k=1 Ak k=1
Definition
Let (X, M, µ) be a measure space. If f : X → [0, ∞] is measurable and E ∈ M,
we
R define thenRLebesgue integral of f over E by o
E
fdµ = sup E
sdµ : s is measurable and simple with 0 ≤ s ≤ f on E .
If this value is finite, we say that f is Lebesgue integrable.
Definition
Let (X, M, µ) be a measure space and f a nonnegative, measurable function.
Let E ∈ M (i.e, E is measurable). We define the integral of f over E with
respect to µ to be Z Z
fdµ = f .1E dµ.
E
This is the Lebesgue integral of f .
Properties
If f and g are nonnegative, measurable functions and f ≤ g, then
Z Z
fdµ ≤ gdµ.
Proof.
If ϕ is a simple, nonnegative function such that 0 ≤ ϕ ≤ f , then it follows that
0 ≤ ϕ ≤ g. Thus the first holds.
Since f .1E ≤ f .1F , the second assertion follows from the first.
Definition
Let (X, M, µ) be a measure space. If f : X → [−∞, ∞] is measurable and
E ∈ M, then we define the Lebesgue integral of f over E by
Z Z Z
fdµ = f+ dµ − f− dµ
E E E
R R
provided either E f+ dµ or E f− dµ is finite.
If both integrals are finite we say that f is Lebesgue integrable.
Proposition
Let f and g be two functions that are integrable with respect to µ. If
f (x) = g(x) a.e., then Z Z
fdµ = gdµ.
X X
n
ai 1Ei (x),
X
ϕ(x) =
i=1
where ai ≥ 0. Set
n
ai 1Ei \Z (x).
X
ψ(x) =
i=1
Then 0 ≤ ψ ≤ g and
Z n
X n
X Z
ψdµ = ai µ(Ei \ Z) = ai µ(Ei ) = ϕdµ.
X i=1 i=1 X
Thus, Z Z
ϕdµ ≤ gdµ.
X X
The reverse inequality follows the same manner. Therefore, in this case
Z Z
fdµ = gdµ.
X X
Finally, the general result follows by considering the positive and negative
parts of f and g.
Properties
Let f , g Lebesgue integrable
R R
i 0 ≤ f (x) ≤ g(x), ⇒ A fdµ ≤ A gdµ.
R R
ii A ⊆ B, 0 ≤ f (x) on B, ⇒ A fdµ ≤ B fdµ.
f + g is integrable and A f + gdµ = A fdµ + A gdµ.
R R R
iii
Proof
Note that the limit function f is measurable. The monotonicity of the integrals
yields Z Z Z
fn dµ ≤ fn+1 dµ ≤ fdµ
X X X
for every n. By monotone convergence of sequences, the limit of the integrals
exists as a value in [0, ∞], call it α. We have that
Z
α≤ fdµ.
X
Proof
Define a sequence of functions (gn ) by gn (x) = inf k≥n {fk (x)}. Note that the
functions in this sequence is measurable and increasing by construction. The
function lim inf fn is measurable and we can write lim inf fn = limn→∞ gn .
Applying Lebesgue’s monotone convergence theorem,
Z Z Z
lim inf fn dµ = lim gn dµ = lim gn dµ.
X X n→∞ n→∞ X
By the properties of infimum it follows that gn (x) ≤ fk (x) for every x ∈ X and
all k ≥ n, from which the monotonicity of the integral implies
Z Z
gn dµ ≤ inf fk dµ.
X k≥n X
Hence Z Z ! Z
lim inf fn dµ ≤ lim inf fk dµ = lim inf fn dµ.
X n→∞ k≥n X X
Example
For n = 1, 2, 3, . . . define
n
if 0 < x ≤ 1
fn (x) =
n
0
otherwise
Proof
For (a), Note that the function f is measurable. Since (fn ) converges pointwise
to f and |fn (x)| ≤ g(x) on A, it must also be the case that |f (x)| ≤ g(x) on A.
Then we have Z Z
|f |dµ ≤ gdµ
A A
and f is integrable.
For (b), We first observe that |fn − f | ≤ 2g. Then the sequence of functions
2g − |fn − f | is a nonnegative sequence of measurable functions and we may
apply Fatou’s Lemma. Notice that
limn→∞ |fn − f | = lim sup |fn − f | = lim inf |fn − f | = 0 on A from which it
follows that
Hence Z
lim sup |fn − f |dµ ≤ 0
A
However, |fn − f | ≥ 0 and thus
Z
lim inf |fn − f |dµ ≥ 0.
A
Therefore, we get
Z Z
0 ≤ lim inf |fn − f |dµ ≤ lim sup |fn − f |dµ ≤ 0
A A
Example
For each positive integer n and x ∈ [0, 1] define fn (x) to be
0
if x = 0
fn (x) =
x2
(1 − e− n ) √1x if 0 < x ≤ 1
if x = 0
0
g(x) =
√1 if 0 < x ≤ 1
x
Then g is integrable on [0, 1] and |fn (x)| ≤ g(x) for all x ∈ [0, 1]. Therefore, by
Lebesgue Dominated Convergence Theorem
Z 1 Z 1
lim fn (x)dx = 0dx = 0.
n→∞ 0 0
Remark
R1
This is a case where it is not easy to compute 0 fn for each n. Without
Lebesgue Dominated Convergence Theorem, it would be difficult to compute
R1
the limit of 0 fn (x)dx.
The Lp Spaces
Definition
Let (X, M, µ) be a measure space and let 1 ≤ p < ∞. We define Lp (X, M, µ)
to be the set of all measurable functions f : X → [−∞, ∞] such that
Z
|f |p dµ < ∞
X
Remark
Lp is a vector space with the usual rules of function addition and scalar
multiplication, but the rule k.kp does not define a norm. The problem here is
the fact that there are nonzero functions which integrate to zero; take a
function which is zero except on a set of measure zero. However such a
function is equal to zero a.e., if we assume the function is measurable.
Equality a.e. with respect to the measure µ defines an equivalence relation on
the collection of all functions which are measurable with respect to µ. This
equivalence relation then partitions the set of all measurable functions into
equivalent classes. In other words, for a measurable function
f : X → [−∞, ∞], the set of all functions which are a.e. equal to f can be
identified as
Example
With respect to the measure space (R, L, m), the Dirichlet function is in the
same class as zero function.
Moreover, if f and g are in the same equivalence class and are integrable, then
Z Z
fdµ = gdµ.
X X
Definition
Let (X, M, µ) be a measure space and let 1 ≤ p < ∞. We define Lp (X, M, µ) to
be the collection of equivalence classes [f ] of measurable functions which
satisfy Z Z
[|f |p ]dµ = |g|p dµ < ∞
X X
for g ∈ [f ] and define a rule k.kp : Lp → [0, ∞) by
Z !1/p
k[f ]kp = p
|g| dµ
X
Remark
Notation Lp for Lp (X, M, µ).
L1 is the spaces of all functions which are Lebesgue integrable with
repsect to (X, M, µ).
We generally write kf kp for k[f ]kp since when working with an
equivalence class, it is often sufficient to work with a single element of
the class.
Lp is a vector space with the standard definitions of function addition and
scalar multiplication within the equivalence class structure
Theorem
The space Lp equipped with the norm k.kp is a normed linear space.
Definition
If 1 < p < ∞, we define the conjugate exponent of p to be the unique real
number q such that
1 1
+ = 1.
p q
This can be written as q = p/(p − 1) from which it is obvious that 1 < q < ∞.
If p = 1, we define q = ∞ to be the conjugate exponent and vice versa if
p = ∞.
Lemma
If x, y ≥ 0 and p, q are conjugate exponents, then
xp yq
xy ≤ + .
p q
Proof
Fix y ≥ 0. If y = 0, then the inequality holds. For y > 0, define a function
f (x) = xy − xp /p on [0, ∞). We will show that f (x) ≤ yq /q for every x ∈ [0, ∞).
First observe that f 0 (x) = y − xp−1 and this function has exactly one zero in the
interval (0, ∞) at c = y1/(p−1) . Furthermore, if x < y1/(p−1) then y − xp−1 > 0
and if x > y1/(p−1) then y − xp−1 < 0 so c is a maximum of f on [0, ∞) and
(y1/(p−1) )p
f (c) = yy1/(p−1) −
p
yp/(p−1)
= yp/(p−1) −
p
=y p/(p−1)
(1 − 1/p)
y q
= .
q
Proof
If f = 0 or g = 0, then the inequality holds true. Therefore we may assume
that kf kp and kgkq are both nonzero. Applying the Lemma with x = |f |/kf kp
and y = |g|/kgkq ,
! ! !p !q
|f | |g| 1 |f | 1 |g|
≤ +
kf kp kgkq p kf kp q kgkq
|f | p |g| q
= p + q.
pkf kp qkgkq
We now integrate this inequality and use the linearity of the integral, we have
Z Z Z
1 1 1
|fg|dµ ≤ p |f | p
dµ + q |g|q dµ
kf kp kgkq X pkf kp X qkgkq X
1 1
= +
p q
= 1.
Multiplying on the left and right by kf kp kgkq gives the desired result.
kf + gkp ≤ kf kp + kgkp .
Proof
If p = 1, then the triangle inequality for absolute value and linearity of the
integral produce the inequality we seek. Now suppose 1 < p < ∞. If
kf + gkp = 0, then the inequality must hold by definition of k.kp . Now suppose
kf + gkp > 0 and let q be the conjugate exponent of p. We may then write
p = q(p − 1) and 1/q = (p − 1)/p, and thus
Z !1/q
k(f + g)p−1 kq = |f + g|(p−1)q dµ
X
Z !(p−1)/p
p−1
= |f + g| dµ
p
= kf + gkp .
X
Now,
Z Z
p
kf + gkp = |f + g| dµ =
p
|f + g||f + g|p−1 dµ
ZX X Z
Hence
kf + gkp ≤ kf kp + kgkp .
Riesz-Fischer Theorem
If 1 ≤ p < ∞ and (X, M, µ) is a measure space, then the space Lp (X, M, µ) is
complete.
Proof
Step 1. The Cauchy sequence: Let (fn ) be a Cauchy sequence in Lp . It
suffices to show that (fn ) has a convergence subsequence with limit f in Lp . By
definition of Cauchy sequence, for each k ∈ N, we can find Nk ∈ N so that
kfm − fn kp < 21k whenever m, n ≥ Nk . In addition, construct an increasing
sequence of natural numbers by choosing n1 = N1 and for k ≥ 2, set
nk = max{nk−1 , Nk } + 1. Then we have a subsequence (fnk ) with
kfnk+1 − fnk k < 21k for every k ∈ N. We want to show that (fnk ) converges to a
function f ∈ Lp . We simply refer to fnk as fk .
gk (x) = |f1 (x)| + |f2 (x) − f1 (x)| + |f3 (x) − f2 (x)| + · · · + |fk (x) − fk−1 (x)|, k ≥ 2.
It follows that
0 ≤ g1 (x) ≤ g2 (x) ≤ g3 (x) ≤ · · ·
and
< kf1 kp + 1
where we used
k ∞ ∞
X X X 1
kfj − fj−1 kp < kfj − fj−1 kp < = 1.
j=2 j=1 j=1
2j
This shows that gk ∈ Lp for every k ∈ N. Set g(x) = limk→∞ gk (x) and
gp (x) = limk→∞ gk (x)p . Moreover, the sequence (gk (x))p is increasing for each
x ∈ X and an application of monotone convergence theorem yields
Z !1/p Z !1/p Z !1/p
p p
kgkp = |g|p dµ = lim gk dµ = lim gk dµ
X X k→∞ k→∞ X
Z !1/p
p
= lim gk dµ = lim kgk kp ≤ kf1 kp + 1
k→∞ X k→∞
Fact
Let (X, d) be a metric space. Then X is complete if and only if every Cauchy
sequence has a convergent subsequence (with limit x ∈ X).
Let (xn ) be a Cauchy sequence in the metric space (X, d) and (xnk ) be its
convergence subsequence, say xnk → x. We claim that (xn ) is convergent with
the limit x. Then ∀ > 0, ∃N ∈ N such that for all n, m, k > N,
d(xn , xm ) < and d(xnk , x) <
2 2
Therefore, note that nk ≥ k, we have
d(xn , x) ≤ d(xn , xnk ) + d(xnk , x) < + = .
2 2
which implies that xn → x.
L∞ space
We define
where
kf k∞ = ess supx∈X |f (x)| = inf{M : |f (x)| ≤ M a.e.}
Theorem
The space (L∞ (X, M, µ), k.k∞ ) is a Banach space.
Summary
Theorem
The space (Lp (X, M, µ), k.kp ), 1 ≤ p < ∞ is a Banach space.
The space (L2 (X, M, µ), k.k2 ) is a Hilbert space.
The space (L∞ (X, M, µ), k.k∞ ) is a Banach space.