FA-Lecture_notes
FA-Lecture_notes
Contents
1
Acknowledgement. We are grateful to Xinfa Meng for his assistance in typing the notes.
1
2 DONGMENG XI AND JIN LI
Definition 1. Vector space (v.s.). A vector space V over a field F(R or C) is a set V with two
operations:
(1) Addition + : V × V → V . V equipped with “+” is an Abelian group (Commutative
and associative laws), with identity element (zero vector) and inverse elements (additive
inverse).
(2) Scalar multiplication · : F × V → V . The scalar multiplication satisfies the associative
and distributive laws; 1 · x = x, ∀x ∈ V , where 1 denotes the multiplicative identity.
Definition 2. Normed vector space (n.v.s.). A vector space E over a field F is said to be a
normed vector space if there is a function ∥ · ∥ : E → R satisfying
(1) ∥x + y∥ ≤ ∥x∥ + ∥y∥ for all x, y ∈ E
(2) ∥αx∥ = |α| · ∥x∥ for every x ∈ E, α ∈ F
(3) ∥x∥ > 0 if x ̸= 0.
The function ∥ · ∥ is called a norm, and we also denote a normed vector space E by (E, ∥ · ∥). For
simplicity, we will write n.v.s..
Definition 3. Metric space. A Metric space X is a set X equipped with a binary function
d : X × X → R satisfying for all x, y, z ∈ X
(1) d(x, y) ≤ d(x, z) + d(z, y) (Triangle inequality)
(2) d(x, y) = d(y, x)
(3) d(x, y) ≥ 0 with equality iff(if and only if) x = y
d(·, ·) is called the distance function.
Let (xk )k≥1 be a sequence in (X, d). We also simply write “a sequence (xk )k≥1 in X”, or “a
sequence (xk ) in X”.
Limit in (X, d). We say (xk )k≥1 converges to x̄, denoted by lim xk = x̄, if lim d(xk , x̄) = 0.
k→∞ k→∞
Cauchy sequence. A sequence (xn )n≥1 in X is said to be a Cauchy sequence if lim d(xi , xj ) = 0.
i,j→∞
Completeness. We say a metric space (X, d) is complete if every Cauchy sequence (xn )n≥1 in X
is associated with an x0 ∈ X s.t.(such that) lim d(xn , x0 ) = 0.
n→∞
Induced metric. Consider a n.v.s. (E, ∥ · ∥). Define d(x, y) = ∥x − y∥ on E, then (E, d) is a
LECTURE NOTES ON FUNCTIONAL ANALYSIS 3
Examples of n.v.s.
1◦ (Rn , ∥ · ∥K )
(1) If K is a convex, bounded and closed set with o ∈ intK and K = −K, then ∥x∥K =
inf {λ > 0 : x ∈ λK} is a norm on Rn .
(2) If ∥ · ∥ is a norm on Rn and let K = {x : ∥x∥ ≤ 1}, then K is a convex, bounded and
closed set with B2 (o, r) ⊂ K and K = −K. Here B2 (o, r) is a Euclidean ball with radius
r centered at the origin.
2◦ C([a, b])
(1) Define ∥f ∥∞ := sup |f (x)|, then(C([a.b]), ∥ · ∥∞ ) is a Banach space.
x∈[a,b]
Rb
(2) Define ∥f ∥1 := a |f (x)|dx, then (C([a.b]), ∥ · ∥1 ) is a n.v.s. but not complete.
One can confirm that (C(Ω̄), ∥ · ∥1 ) and (C(Ω̄), ∥ · ∥∞ ) are all normed vector space.
Pn α
where α = (α1 , · · · , αn ) is a multi-index with αi ≥ 0, i ≥ 1 and |α| = i=1 αi , D f (x) =
∂ |α|
α
∂x1 1 ···∂xα n f (x).
n
Define ∥f ∥1,1 = Ω |f | + nk=1 Ω |∂k f | for every f ∈ C 1 (Ω̄), where ∂k f = ∂x∂ k f . It is easily
R P R
seen that ∥ · ∥1,1 is a norm on the vector space C 1 (Ω̄), and we leave it as an exercise.
And we can also define ∥f ∥1,∞ = sup |f (x)| + nk=1 sup |∂k f (x)| for every f ∈ C 1 (Ω̄) as a norm
P
x∈Ω x∈Ω
on C 1 (Ω̄).
Definition 4. Open sets and Closed sets. Suppose (X, d) is a metric space.
Let x ∈ X and r > 0. Denote the open ball centered at x with radius r by Bx (r) = B(x, r) =
{z ∈ X : d(z, x) < r}, and the closed ball by B̄(x, r) = {z ∈ X : d(z, x) ≤ r}.
We say A ⊂ X is open if for every x ∈ A, there exists r > 0 s.t. B(x, r) ⊂ A. We also set the
empty set ∅ to be open. One can confirm that an open ball is open.
We say A ⊂ X is closed, if Ac = {x ∈ X : x ∈ / A} is open. One can confirm that a closed ball
is closed.
Ā = (int(Ac ))c is called the closure of A. One can also confirm that x ∈ Ā iff there exists
(xk )k≥1 ⊂ A s.t. xk → x as k → ∞.
4 DONGMENG XI AND JIN LI
Remark 1. We define intA := {x ∈ A : there exists r > 0 s.t. B(x, r) ⊂ A}, and x ∈ intA is
called an interior point of A.
Property 1. intA is open and A is open iff A = intA.
Property 2. For (X, d)
(1) Both X and ∅ are open.
(2) If Aα is open for every α ∈ I, then ∪ Aα is open.
α∈I
m
(3) If Aα is open for every 1 ≤ k ≤ m, m ∈ N+ , then ∩ Ak is open.
k=1
Examples:
1◦ . Rn .
2◦ . A discrete metric space X is a space such that d(x, y) = 1 for all x, y ∈ X. Each subset of
a discrete metric space is both open and closed.
3◦ . Any finite-dimensional linear subspace of a n.v.s is closed. But infinite-dimensional linear
subspace may not be closed. For instance, Cc∞ (Rn ) in L1 (Rn ) and polynomials in C([0, 1]).
Definition 6. Open cover. We say {Aα }α∈I is an open cover, if {Aα }α∈I is a class of open sets
s.t. K ⊂ ∪ Aα .
α∈I
Theorem 1.1. K is compact iff each open cover of K has a finite subcover.
Proof. Necessary part. Suppose any sequence (xk )k≥1 in K has a convergent subsequence con-
verging to a point in K. Let {Oi }i∈I (⊃ K) be an arbitrary open cover of K.
Step 1. We prove that there is an r > 0 such that for every x ∈ K, B(x, r) ⊆ Oi for some i ∈ I.
The proof is by contradiction. Assume that for any r > 0, there is an xr ∈ K, such that for each
i ∈ I, B(xr , r) ⊈ Oi . Now choose the sequence {xn }n≥1 in X so that
By the assumption, {xn }n≥1 has a convergent subsequence {xnk }k∈N , and xnk → x as k → ∞,
where x ∈ K. Then, there must be some i0 ∈ I such that x ∈ Oi0 , and since Oi0 is open, so there
exists r0 > 0 such that B(x, r0 ) ⊆ Oi0 . Choose N large enough such that d (x, xN ) < 21 r0 and
LECTURE NOTES ON FUNCTIONAL ANALYSIS 5
1
N
< 21 r0 . Now if y ∈ B (xN , 1/N ), then
d(x, y) ≤ d (x, xN ) + d (xN , y)
1 1
< r0 + r0 = r0
2 2
and hence y ∈ B(x, r0 ) ⊆ Oi0 . It follows that B (xN , 1/N ) ⊆ B(x, r0 ) ⊆ Oi0 , a contradiction!
Step 2. Next we prove that K is totally bounded i.e. for any ϵ > 0, there are finitely many balls
B(xi , ϵ) with xi ∈ K, such that K ⊂ m
S
i=1 B(xi , ϵ).
Otherwise, we can choose an arbitrary x1 ∈ K, then choose x2 ∈ K\B(x1 , ϵ), and step by step
we have
[k
xk+1 ∈ K \ B(xi , ϵ),
i=1
for k = 1, 2, · · · . The sequence (xk )k≥1 satisfies d(xk , xj ) ≥ ϵ for any k ̸= j, which contradicts to
the assumption that (xk )k≥1 has a convergent subsequence.
Step 3. Let ϵ < r where the r is chosen as in Step 1. Then by Step 2 there are finitely many balls
{B(xi , ϵ)}m
i=1 which covers K. Since Step 1 provides an Oki containing B(xi , ϵ), we have found
the finite open cover. Therefore, the necessary part is proved.
Sufficient part. Suppose each open cover of K has a finite subcover. We assume on the contrary
that there is a sequence {zk }k≥1 in K, such that any subsequence of {zk }k≥1 does not converge
to some point in K.
We notice that {zk }k≥1 must have infinitely many distinct points, since otherwise, {zk }k≥1
shall has a convergent subsequence converging to a point who appears infinitely many times in
this sequence.
Step 1. For an arbitrary x ∈ K, there is an ϵx > 0, such that B(x, ϵx ) ∩ {zk }k≥1 is a subset of
singleton {x}. Since otherwise, if there is an x such that for any ϵ > 0, B(x, ϵ) ∩ {zk }k≥1 has a
point different from x, then x will be a limit point to a subsequence, and this contradicts to the
original assumption.
S
Step 2. For each x ∈ K, we select corresponding ϵx as in Step 1. Since x∈K B(x, ϵxi ) covers K,
there are x1 , · · · , xN , such that N
S SN
i=1 B(x i , ϵx i
) covers K. However, i=1 B(xi , ϵxi ) has at most
N elements in {zk }k≥1 , which contradicts to the fact that {zk }k≥1 has infinitely many distinct
points. Therefore, the sufficient part is proved. □
Definition 7. Continuous map. For two metric spaces (X, d1 ) and (Y, d2 ). Let f : X → Y be a
map.
Remark 2. We say f is continuous at x0 ∈ X, provided for all ϵ > 0, there exists δ > 0 s.t.
if x ∈ X satisfies d1 (x, x0 ) < δ, then d2 (f (x), f (x0 )) < ϵ or equivalently lim f (x) exists and
x→x0
6 DONGMENG XI AND JIN LI
lim f (x) = f (x0 ) . In particular, we say f is continuous on X, if f is continuous at each point
x→x0
of X.
Proposition 1.2. f : X → Y is continuous iff f −1 (A) is open in X, for every open set A in Y .
We leave it as an exercise.
Definition 8. Dual space. Let (E, ∥ · ∥) be a n.v.s.. The dual space of E is the space of all
continuous linear functionals on E and denote the dual space of E by E ∗ . The dual norm on E ∗
is defined by
|f (x)|
∥f ∥E ∗ = sup = sup f (x) = sup f (x).
x∈E ∥x∥ x∈E, x∈E,
∥x∥≤1 ∥x∥=1
Prove it as an exercise.
Definition 11. Separable space. A metric space is called separable if it has a countable dense
subset.
Definition 12. Completion. Let (X, d) be a metric space. We say (X,e d) is a completion of X,
if there exists an isometry φ : X → X,
e such that φ(X) is dense in X
e and Xe is complete.
Theorem 1.5. Let (X, d) be a metric space. Denote C[X] to be the collection of all Cauchy
sequences in X. Define a relation ‘∼’ on C[X] by (xn ) ∼ (yn ) iff lim d(xn , yn ) = 0. Then ‘∼’ is
n→∞
an equivalence relation on C[X]. Let X e := C[X]/ ∼ be the quotient space, equipped with metric
de defined by
e n )], [(yn )]) = lim d(xn , yn ) for all [(xn )], [(yn )] ∈ X,
d([(x e
n→∞
In order to simplify the symbol, we sometimes denote by (xn ) an arbitrary sequence (xn )n≥1
in X.
LECTURE NOTES ON FUNCTIONAL ANALYSIS 7
Proof. In the following proof, we may use the notation x to denote a sequence (xn ) in C[X], and
use [x] to denote an element in X̃.
Step 1. It is easy to verify the following conditions,
So ∼ is an equivalence relation.
Step 2. Next, we prove that de is a well-defined distance function on X. e
e Let (xn ), (x′n ) ∈ [x] and (yn ), (yn′ ) ∈ [y]. Since (xn ), (yn ) are Cauthy sequences,
Let [x], [y] ∈ X.
Thus lim d(xn , yn ) exists. On the other hand, since d(xn , yn ) ≤ d(xn , x′n ) + d(x′n , yn′ ) + d(yn , yn′ ),
n→∞
we have lim d(xn , yn ) = lim d(x′n , yn′ ), which means de is well-defined.
n→∞ n→∞
Let [z] ∈ X
e and (zn ) ∈ [z]. Then,
= 0,
where the last equation follows from the first inequality in (EQ1). This shows the completeness
of (X,
e d).
e
Step 4. Define φ(z) = [(z, z, · · · )] for every z ∈ X. Then φ : X → X e is an isometry, since for
any x, y ∈ X, d(x, y) = lim d(xn , yn ) = d([(xe n )], [(yn )]) where xn = x, yn = y. And for every
n→∞
[(xn )] ∈ X,
e lim d(φ(x
e k ), [(xn )]) = lim lim d(xk , xn ) = 0, thus φ(X) is dense in X.
e □
k→∞ k→∞ n→∞
(x′n ) are equivalent Cauthy sequences. Since X f2 is complete and φ2 is isometry which implies
de2 (φ2 (xn ), φ2 (x′n )) = d(xn , x′n ), we deduce that lim φ2 (xn ), lim φ2 (x′n ) both exist and are the
n→∞ n→∞
same.
Step 2. By the Claim above, we can define f : X f1 → Xf2 as follows. For every ye ∈ X f1 , define
f (e
y ) = lim φ2 (xn ) where (xn ) is an arbitrary sequence in X converging to ye. It is clear that f
n→∞
is well-defined.
f1 and (x1 ), (x2 ) be sequences in X s.t. lim φ1 (xk ) = yek k = 1, 2. Then
Let ye1 , ye2 ∈ X n n n
n→∞
de2 (f (ye1 ), f (ye2 )) = lim de2 (φ2 (x1n ), φ2 (x2n )) = lim d(x1n , x2n ) = de1 (ye1 , ye2 ). Consequently, f is an
n→∞ n→∞
isometry form X
f1 to X
f2 .
f ◦ φ1 (x) = f (e
y)
= lim φ2 (xn ) (where xn = x, n ∈ N+ )
n→∞
= φ2 (x).
This implies f ◦ φ1 = φ2 .
Step 4. Suppose there is another isometry f ′ : X f1 → X f2 satisfying f ′ ◦ φ1 = φ2 on X. The
condition f ′ ◦φ1 = φ2 = f ◦φ1 implies that f (e
y ) = f ′ (e
y ) for every y ∈ φ1 (X). Since isometry must
be continuous (Prove it!), and φ1 (X) is dense in X, e we have f (e y ) = f ′ (e
y ) for every y ∈ X.
e □
10 DONGMENG XI AND JIN LI
Definition 13. Inner product space (i.p.s.). An inner product space is a vector space H together
with a symmetric positive definite bilinear function (·, ·) : H × H → R satisfying that for any
λi ∈ R and ui , u, v ∈ H, i = 1, 2
Examples.
1◦ Rn := (x1 , · · · , xn )t : xk ∈ R, 1 ≤ k ≤ n
Rigorously, when referring to the L2 (Ω) space, we are essentially discussing a quotient space
L2 (Ω)/ ∼. Here, for u, v ∈ L2 (Ω), we say u ∼ v iff u(x) = v(x) a.e. x ∈ Ω. Under this notation,
for [u], [v] ∈ L2 (Ω)\ ∼, the inner product ([u], [v]) is still defined by
Z
(u, v)L2 = u(x)v(x)dx.
Ω
Note that it is independent of the choices of the representatives u and v. The benefit is that
it makes the “zero vector” of this space to be the unique one satisfying ([u], [u]) = 0. One can
easily confirm that (L2 (Ω)\ ∼, (·, ·)) is a inner product space.
Usually, we abandon the above notation L2 (Ω)\ ∼. Instead, we simply write L2 (Ω), and we
say u and v are the “same point in L2 (Ω)”, if u(x) = v(x) a.e.
4◦ L2g (Ω)
LECTURE NOTES ON FUNCTIONAL ANALYSIS 11
R R
Suppose g ∈ C(Ω), g ≥ 0 and Ω
g > 0. Define (u, v)g = Ω
u(x)v(x)g(x)dx for measurable functions u
and v, and Z
L2g (Ω) := {u : uis measurable and u2 g < ∞}.
Ω
5◦ (Product space) Let H1 and H2 be two inner product spaces. Denote H1 × H2 := {x1 ⊕ x2 :
xi ∈ Hi , i = 1, 2}. Define
It is not hard to see that (·, ·)H1 ×H2 is an inner product space of H1 × H2 . We say (H1 ×
H2 , (·, ·)H1 ×H2 ) (briefly write it as H1 × H2 ) is the product (inner product) space of H1 and H2 .
If H1 and H2 are Hilbert spaces, so does H1 × H2 .
1 1
Remark 3. Cauchy-Schwarz inequality. Suppose H is an i.p.s., then (u, v) ≤ (u, u) 2 (v, v) 2 holds
for any u, v ∈ H.
Proof. ∀u, v ∈ H, by (u − v, u − v) ≥ 0 and (u + v, u + v) ≥ 0, we get (u, u) + (v, v) ≥ 2|(u, v)|.
Letting ū = u 1 , v̄ = v 1 , we have (ū, ū) = (v̄, v̄) = 1, and hence |(ū, v̄)| ≤ 1. Equivalently,
(u,u) 2 (v,v) 2
1 1
|(u, v)| ≤ (u, u) (v, v) .
2 2
1
Induced norm. Now let |u| = (u, u) 2 for every u ∈ H. By Cauchy-Schwarz inequality, | · | is a
norm. We call it the induced norm.
Definition 14. Hilbert space. If the i.p.s. (H, | · |) is complete, we say H is a Hilbert space.
From now on, we assume H to be Hilbert space.
Remark 5.
(1) The examples 1◦ , 2◦ , 3◦ are all Hilbert spaces.
(2) If Ω ⊂ Rn is open and bounded, C(Ω̄), (·, ·)L2 is still an inner prodect space, but not
Hilbert.
Theorem 2.1 (Projection onto a closed convex set). Let K ⊂ H be nonempty, closed and convex.
Then for any f ∈ H, there exists a unique element u ∈ K, s.t.
Moreover, it is characterized by
u ∈ K and (f − u, v − u) ≤ 0 ∀v ∈ K. (2.2)
(2) Characterization.
Necessary part. Suppose u ∈ K satisfies |f − u| = min |f − v| = d(K, f ). For an arbitrary
v∈K
v ∈ K, let Φ(t) = (f − ((1 − t)u + tv), f − ((1 − t)u + tv)). Then Φ(0) takes minimum on [0, 1],
and it follows that
Φ(t) − Φ(0)
lim+ ≥ 0,
t→0 t
which implies (f − u, u − v) ≥ 0 by directly computing derivation. So (2.2) holds.
LECTURE NOTES ON FUNCTIONAL ANALYSIS 13
Sufficient part. Suppose there exists u ∈ K, s.t. (2.2) holds for any v ∈ K. It follows from a
direct computation that for every v ∈ K,
|f − u|2 − |f − v|2 = (f − u, f − u) − (f − v, f − v)
= (u, u) − (v, v) + 2(f, v − u)
= (u, u) − (v, v) + 2(u, v − u) + 2(f − u, v − u)
≤ −(u, u) − (v, v) + 2(u, v) = −(u − v, u − v) ≤ 0.
(3) Uniqueness.
If u1 , u2 are two minimizers, then (f − u1 , u2 − u1 ) ≤ 0 as well as (f − u2 , u1 − u2 ) ≤ 0. Adding
them together, we get (u1 − u2 , u1 − u2 ) ≤ 0 i.e. u1 − u2 = 0 and u1 = u2 .
□
Definition 15. Metric projection. The Theorem 2.1 defines a map PK , by for every f ∈ K,
PK f = u ∈ K s.t. |u − f | = d(K, f ).
It is called the projection of f onto K. And the map PK is sometimes called the metric projection.
In addition, let n = f − u and V := {x ∈ H : (n, x) = (n, u)}, then V is a Support hyperplane of
K, and V − := {x : (n, x) ≤ (n, u)} ⊃ K.
Remark 6. In the case that M is closed and linear, PM f is called the orthogonal projection of f
onto M , and the map PM is linear.
14 DONGMENG XI AND JIN LI
Theorem 2.4 (Riesz-Frechet representation theorem). Given any ϕ ∈ H ∗ there exists a unique
1
u ∈ H, s.t. ϕ(v) = ⟨ϕ, v⟩H ∗ ,H = (u, v) for every v ∈ H. Moreover, ∥ϕ∥H ∗ = |u|H = ⟨u, u⟩ 2 .
Proof. Assume ϕ ̸= 0, otherwise ϕ(v) = (0, v).
Let M = [ϕ = 0] ⊂ H be a closed linear subspace. Choose x ∈ H\M , i.e. ϕ(x) ̸= 0, then
x−Px
(x − PM x, y) = 0 for every y ∈ M . Denote Px = PM x, and let u = ϕ(x) |x−P |2
.
x
2
ϕ(x)
We claim that ϕ(v) = (u, v) for every v ∈ H. In fact, by direct computation, (u, u) = |x−P x|
2 =
ϕ(x) ϕ(v)
ϕ |x−P 2x = ϕ(u). Given v ∈ H, let v0 = v − ϕ(u) u, then ϕ(v0 ) = 0, i.e. v0 ∈ M . By the Cor
x|
ϕ(v) ϕ(v)
2.3, (u, v0 ) = 0, thus (u, v) = u, v0 + ϕ(u) u = ϕ(u) (u, u) = ϕ(v).
In the end,
∥ϕ∥H ∗ = sup ϕ(v) = sup (u, v) ≤ |u||v| ≤ |u|,
∥v∥≤1 ∥v∥≤1
′
and on the other hand, for v = u/|u|,
Definition 16. Bounded linear operator. Let E, F be two normed vector spaces. If A : E → F
is linear, and if sup ∥Au∥F < ∞, we say A is a bounded linear operator.
∥u∥E ≤1
We denote the space of bounded linear operators by L(E, F ).
Remark 7. Actually one can confirm that any linear operator ϕ ∈ L(E, F ) iff ϕ is continuous.
Examples.
1◦ Bounded but not continuous functional: many examples are known in the previous study.
2◦ Continuous but not bounded functional: let en = (0, . . . , 0, 1, 0, . . . ) ∈ l2 , where 1 only
appears at the n-th coordinate. Define fn (x) = max{0, 1/2 − ∥x − en ∥} and
∞
X
f (x) = 2nfn (x)
n=1
for any x ∈ l2 . Clearly fn is supported on disjoint closed balls B̄(en , 1/2). Hence f is a continuous
functional l2 → R but f (en ) = n tells us that f is unbounded.
Definition 17. Bidual space. Let E be a normed vector space. Recall the definition of its dual
space E ∗ and E ∗ is a Banach space. Then we call E ∗∗ = (E ∗ )∗ its bidual.
Proposition 2.5. Let x ∈ E. Define ξ : E ∗ → R by ⟨ξ, f ⟩E ∗∗ ,E ∗ = ⟨f, x⟩E ∗ ,E for every f ∈ E ∗ .
Then
LECTURE NOTES ON FUNCTIONAL ANALYSIS 15
(1) ξ ∈ E ∗∗ .
(2) The map J : E → E ∗∗ , x 7→ ξ for ∀x ∈ E is linear and bounded.
J is called the canonical injection.
Proof.
(1) Clearly, ξ is linear. Since ⟨f, x⟩E ∗ ,E ≤ ∥f ∥E ∗ ∥x∥, we have ⟨ξ, f ⟩E ∗∗ ,E ∗ ≤ ∥x∥∥f ∥E ∗ . That is
to say, sup ⟨ξ, f ⟩E ∗∗ ,E ∗ ≤ ∥x∥ < ∞.
∥f ∥E ∗ ≤1
Proof. Define L : H → H ∗ by (Lu)(v) = (u, v), ∀v ∈ H. By Thm 2.4, for every ϕ ∈ H ∗ , there
exists unique u ∈ H, s.t. ϕ = Lu, and ∥Lu∥H ∗ = |u|. Thus, L : H → H ∗ is linear, bijective and
an isometry.
Define A : H ∗∗ → H ∗ , by for every ξ ∈ H ∗∗
Since Aξ ∈ H ∗ (Prove it!), the definition of A make sense. Then for every ξ ∈ H ∗∗ , there exists
v ∈ H s.t. Aξ = Lv. Now for any Lu ∈ H ∗ ,
Theorem 2.7. (M ⊥ )⊥ = M̄ .
Proof. Step 1. Firstly, we show that for any linear subspace N ⊂ H, N ⊥ is closed. Actually,
from ui → u with ui ∈ N ⊥ , we know (ui , v) = 0, ∀v ∈ N , so (u, v) = 0, ∀v ∈ N and u ∈ N ⊥ .
Step 2. Next, we show that M̄ ⊂ (M ⊥ )⊥ . For u ∈ M , (u, v) = 0, ∀v ∈ M ⊥ , and u ∈ (M ⊥ )⊥ . So
M ⊂ (M ⊥ )⊥ , and by Step 1, (M ⊥ )⊥ is closed, hence M̄ ⊂ (M ⊥ )⊥ .
Step 3. Finally, we prove that (M ⊥ )⊥ ⊂ M̄ . Suppose by contradiction that f ∈ (M ⊥ )⊥ \M̄ .
Denote Pf = PM̄ f . By Cor 2.3, (f − Pf , v) = 0 for all v ∈ M̄ , then f − Pf ∈ M̄ ⊥ . Since
f ∈ (M ⊥ )⊥ and clearly (M̄ )⊥ ⊂ (M )⊥ , so (f − Pf , f ) = 0. Hence (f − Pf , f − Pf ) = 0, which
means f = Pf ∈ M̄ , a contradiction! □
Theorem 2.8 (Orthogonal decomposition). Let M ⊂ H be a closed linear subspace. For any
u ∈ H, u = PM u + PM ⊥ u is an orthogonal decomposition of u to M and M ⊥ . Moreover, the
orthogonal decomposition is unique in the sense that if u = uM + uM ⊥ for some uM ∈ M and
uM ⊥ ∈ M ⊥ , then uM = PM u and uM ⊥ = PM ⊥ u.
Theorem 2.9 (Lax-Milgram). Assume a(·, ·) is a continuous, coercive, bilinear form on a Hilbert
space H. Then given any ϕ ∈ H ∗ , there exists a unique u ∈ H, s.t.
(2) Next, we handle the situation without symmetry. Our aim is to find u ∈ H, s.t. ⟨ϕ, v⟩ =
a(u, v) for every v ∈ H.
By Riesz-Frechet representation theorem, there exists an f ∈ H, s.t. ⟨ϕ, v⟩ = (f, v) for every v ∈
H. And if we fix an u ∈ H, the map v 7→ a(u, v) is continuous and linear on H. So again
by Riesz-Frechet representation theorem, there exists an element in H denoted by Au, s.t.
a(u, v) = (Au, v) for every v ∈ H. Then this naturally induces a map A : H → H, u 7→ Au.
A satisfies
1◦ (A(λu + λ′ u′ ), v) = a(λu + λ′ u′ , v) = λa(u, v) + λ′ a(u′ , v) = λ(Au, v) + λ′ (Au′ , v) =
(λAu + λ′ Au′ , v), so A is linear.
2◦ (Au, Au) = a(u, Au) ≤ c|Au||u|, so |Au| ≤ c|u| i.e. A is continuous.
3◦ |Au||u| ≥ (Au, u) = a(u, u) ≥ c′ |u|2 , so |Au| ≥ c′ |u|.
Claim 1. The condition 3◦ above implies that A is injective and the range R(A) = {Au : u ∈ H}
is closed. In fact, if Au1 = Au2 , then |u1 − u2 | ≤ c1′ |Au1 − Au2 | and u1 = u2 . If (fi := Aui )i≥1
converges to f¯ ∈ H, then |ui − uj | ≤ c1′ |Aui − Auj | → 0 as i, j → ∞, and there exists ū ∈ H, s.t.
ui → ū. By the continuous of A, we have Aū = f¯. Thus R(A) is closed.
Claim 2. R(A)⊥ = {o}. Actually, if w ∈ R(A)⊥ ⊂ H, then (Au, w) = 0 for all u ∈ H. Taking
u = Aw, we have a(w, w) = (Aw, w) = 0 ≥ c′ |w|2 , so w = 0.
18 DONGMENG XI AND JIN LI
Therefore, by Thm 2.7, R(A) = (R(A)⊥ )⊥ = H. Since R(A) is closed, we have R(A) = H.
As a result, A : H → H is surjective and injective, and hence there exists u ∈ H, s.t. Au = f .
That is to say, ⟨ϕ, v⟩ = (f, v) = (Au, v) = a(u, v) for every v ∈ H. □
We leave its proof as an exercise, and actually we can prove a more general result for Lp (Ω).
Definition 20. (1) For f, g ∈ Cc1 (Ω), (f, g)H01 := Ω f (x)g(x)dx + nk=1 Ω ∂k f (x)∂k g(x)dx
R P R
1
Cc (Ω).
defines an inner product on the linear space
1 1
(2) Define H0 (Ω) to be the completion of Cc (Ω), | · |H01 , where | · |H01 denotes the induced norm
of (·, ·)H01 .
to
be a linear subspace
of X. Denote the closure of X1 in Y by X̄1 . Then, X̄1 is a completion of
1
Cc (Ω), | · |H01 .
Definition 21. For each “point” in H01 (Ω), we identify it with a point (u, u1 , . . . , un ) ∈ X̄1 .
Usually, we write u ∈ H01 (Ω), instead of (u, u1 , ..., un ) ∈ X̄1 . Here u means a function in
H01 (Ω) as well as a function in L2 (Ω). For each j, we denote ∂j u := uj , and ∂j u is said to be the
weak partial derivative of u.
Remark 8. It is reasonable to write u ∈ H01 . It is easily seen that, for f ∈ Cc1 (Ω) ⊂ H01 , ∂j f is
determined by f . In fact, for (u, u1 , ..., un ) ∈ X̄1 , by the knowledge of real analysis, uj is also
uniquely determined a.e. by u.
Proof. Let (f )i≥1 be a Cauchy sequence in (Cc1 (Ω), (·, ·)H 1 ). Then ((f i , ∂1 f i , . . . , ∂n f i ))i≥1 is a
i
Cauchy sequence in X̄1 . Since Y is complete, there is a unique (u, u1 , . . . , un ) ∈ Y such that
|f i − u|L2 → 0 and |∂j f i − uj | → 0 as i → ∞ for each j = 1, . . . , n. For any v ∈ Cc1 (Ω), from
integration by parts, it follows that
Z Z
∂j f (x)v(x)dx = − f i (x)∂j v(x)dx,
i
Ω Ω
2.4.1. Weak Solution to the Dirichlet Problem. Now we consider the problem
(
−∆u(x) + u(x) = f (x) x ∈ Ω
(P 1)
u(x) = 0 x ∈ ∂Ω
where f ∈ C(Ω̄).
In functional analysis, we consider this problem from a dual viewpoint. Define T : L2 (Ω) →
(L2 (Ω))∗ , by Z
(T f )(v) = f (x)v(x)dx, ∀v ∈ L2 (Ω).
Ω
2 2 ∗
Define A : C (Ω̄) → (L (Ω)) by
Z
(Au)(v) = (−∆u + u)vdx, ∀v ∈ L2 (Ω).
Ω
20 DONGMENG XI AND JIN LI
Au = T f.
Bu = T f.
Such a u ∈ H01 (Ω) is said to be the weak solution of (P1). One benefit now is the completeness!
To find u ∈ H01 (Ω) s.t. Bu = T f holds, recalling the definitions of (u, v)H01 and (u, v)L2 , it is
equivalent to solving
(u, v)H01 = (f, v)L2 for every v ∈ H01 (Ω).
Since (f, v)L2 ≤ |f |2 |v|2 ≤ |f |2 |v|H01 , ϕ(v) = (f, v)L2 is a bounded linear operator on H01 (Ω).
Thus there exists unique u ∈ H01 (Ω), s.t.
n n n
!
X ∂ X ∂ X ∂
a1j (x) u(x), a2j (x) u(x), ..., anj (x) u(x) .
j=1
∂x j j=1
∂x j j=1
∂x j
Pn
Definition 22. Elliptic operator. We say L is elliptic, if there exists θ > 0, s.t. i,j=1 aij (x)ξ i ξ j ≥
θ|ξ|2 , for any x ∈ Ω and ξ = (ξ1 , ..., ξn ) ∈ Rn .
LECTURE NOTES ON FUNCTIONAL ANALYSIS 21
Proof.
Step 1. Define a(·, ·) : H01 (Ω) × H01 (Ω) → R by
n Z Z
X ∂ ∂
a(u, v) = aij (x) u(x) v(x)dx + a0 (x)u(x)v(x)dx for every u, v ∈ H01 (Ω)
i,j=1 Ω ∂x j ∂x i Ω
Therefore, to prove (1), it suffices to verify that a(·, ·) satisfies the assumptions of Lax-Milgram
theorem.
It is clear that a(·, ·) : H01 (Ω) × H01 (Ω) → R is bilinear. We will show that a(·, ·) is continuous
and coercive.
On one hand, since Ω̄ ⊂ Rn is compact, aij ∈ C 1 (Ω̄) and a0 ∈ C(Ω̄), there exists an M > 0
s.t. |aij (x)| ≤ M, |a0 (x)| ≤ M for any x ∈ Ω. Hence,
n Z Z
X ∂ ∂
|a(u, v)| = aij (x) u(x) v(x)dx + a0 (x)u(x)v(x)dx
i,j=1 Ω
∂xj ∂xi Ω
n
X ∂ ∂
≤M u(x), v(x) + M (u, v)L2
i,j=1
∂x j ∂x i L2
≤ M 2 (u, v)H01 .
22 DONGMENG XI AND JIN LI
≥ min{θ, θ0 }|u|H01 .
Definition 23. Algebraic basis. Let E be an n.v.s. and let {ei }i∈I be a family of vectors in E.
Notice that the index set I here may not be countable. We say {ei }i∈I is an algebraic basis, if
every x ∈ E can be uniquely written as
X
x= xi ei , for some finite subset J ⊂ I and xi ∈ R.
i∈J
Definition 24. Hilbert sum. Let {En }n≥1 be a sequence of closed subspaces of H. We say that
H is the Hilbert sum of En ’s and denote it by H = ⊕ En , if the following holds,
n≥1
(1) {En }n≥1 are mutually orthogonal, i.e. (u, v) = 0 for any u ∈ En , v ∈ Em with n ̸= m.
∞
S
(2) The linear space spanned by En is dense in H.
n=1
Theorem 3.1. Assume that H is the Hilbert sum of En ’s. Given u ∈ H, set un = PEn u and
sn = nk=1 uk . Then,
P
lim sn = u.
n→∞
Now it is reasonable to write ∞
P
k=1 uk and we have such following Bessel-Parseval’s identity,
∞
X
|uk |2 = |u|2 .
k=1
P∞
Lemma 3.2. Let (vn )n≥1 be any sequence in H, s.t. (vm , vn ) = 0 for any m ̸= n and k=1 |vk |2 <
Pn
∞. Then s := lim sn := lim vk exists and
n→∞ n→∞ k=1
∞
X
|s|2 = |vk |2 .
k=1
Pm
Proof of Lemma 3.2. Clearly, |sn − sm |2 = 2
k=n+1 |vk | → 0 as n, m → ∞, n < m. Thus
s = lim sn ∈ H exists, by the completeness of H. By the definition of induced norm and a
n→∞
direct computation, it follows that |s|2 = lim |sn |2 = ∞ 2
P
n→∞ k=1 |vk | . □
(u − un , v) = 0 for all v ∈ En ,
24 DONGMENG XI AND JIN LI
and hence (u, un ) = |un |2 , n ∈ N. Adding them, we get (u, sn ) = nk=1 |uk |2 . At the same time,
P
(sn , sn ) = nk=1 |uk |2 , hence (u, sn ) = |sn |2 . Therefore, |sn | ≤ |u| which implise that ∞ 2
P P
k=1 |uk | ≤
n
|u|2 . Now, by Lem 3.2, we have s = lim sn = lim uk exists and |s|2 = ∞ 2
P P
n→∞ n→∞ k=1 k=1 |uk | .
Definition 25. Hilbert basis. A sequence (en )n≥1 in H is said to be a Hilbert basis of H, if it
satisfies
n n
(1) (en , em ) = δm , where δm is called Kronecker symbol defined by
(
n 1 n=m
δm := .
0 n ̸= m
(2) span {en : n ∈ N} is dense in H.
In some textbooks, it is also called complete orthonormal system or orthonormal basis. For
example, the Fourier series.
Corollary 3.3. Let (en )n≥1 be an orthonormal basis. Then for every u ∈ H, we have u =
P∞ n
(u, ek )ek and |u|2 = ∞ 2
P P
k=1 (u, ek )ek , i.e. u = n→∞
lim k=1 (u, ek ) . Conversely, given any sequence
P k=1
(αn )n≥1 ∈ l2 , the series ∞ k=1 αk ek converges to some element u ∈ H s.t. (u, ek ) = αk for all
2
P∞ 2
k ∈ N and |u| = k=1 αk .
Moreover, |u|2 = ∞ 2
P
k=1 (u, ek ) .
LECTURE NOTES ON FUNCTIONAL ANALYSIS 25
∞ n
Conversely, (αk )k≥1 ∈ l2 , αk 2 < +∞, and hence lim |αk ek |2 < +∞. Then by Lem
P P
k=1 n→∞ k=1
n
P
3.2 lim αk ek exists. Denote it by u and it is clear that (u, ek ) = αk for all k ∈ N and
n→∞ k=1
∞
|u|2 = αk 2 .
P
□
k=1
Proof. Let {vn : n ∈ N} be a countable dense subset of H. Let Fk denotes span{vi : 1 ≤ i ≤ k}.
∞
S
Clearly, Fk is dense in H. Now we construct an orthonormal basis as follows.
k=1
1◦ If F1 = {o}, reindex by letting Fk+1 = Fk , for all k ≥ 1. If F1 ̸= {o}, let e1 = v1 /|v1 |.
Then span{e1 } = F1 .
◦
2 If F2 = F1 , reindex by letting Fk+1 = Fk , for all k ≥ 2. If F2 ̸= F1 , let
.
e2 = v2 − (v2 , e1 )e1 v2 − (v2 , e1 )e1 .
Definition 26. Adjoint operator. Let L ∈ L(H1 , H2 ), where H1 , H2 are Hilbert spaces. Define
L∗ : H2 → H1 , by
(u, L∗ v)H1 = (Lu, v)H2 for every u ∈ H1 , v ∈ H2 .
And we call L∗ the adjoint operator of L.
26 DONGMENG XI AND JIN LI
Definition 27. Self-adjoint. If L ∈ L(H) and L = L∗ , i.e. (Lu, v) = (u, Lv) for all u, v ∈ H, we
say that L is self-adjoint.
Definition 28. Weak convergence. We say a sequence (xn )n≥1 in H converges weakly to x ∈ H,
written xn ⇀ x, if lim (y, xn ) = (y, x) for all y ∈ H.
n→∞
Remark 10. We say xn → x strongly, if |xn −x| → 0. And it is clear that xn → x strongly implies
xn ⇀ x. In fact, weak convergence is equivalent to strong convergence in finite-dimensional n.v.s.
However, a weak convergence may not be a strong convergence in infinite-dimensional Hilbert
√
space. E.g. sin nx ∈ L2 ([0, 1]) converges weakly to 0 but ∥ sin nx∥2 → 1/ 2 (as n → ∞).
Remark 11. It is clear that A is a compact operator is equivalent to that A(D) is compact in F
for any bounded set D ⊂ E.
Proof. Suppose (vn )n≥1 is a sequence in BH . We aim to show that there exists a subsequence
(vnk )k≥1 in BH such that (K ∗ (vnk ))k≥1 converges to a point w̄ ∈ H.
Step 1. Since K(BH ) is precompact, it must be separable. Let A0 = {Ku1 , · · · , Kuj , · · · } be a
countable dense subset
of K(BH ), where uj∈ BH . Since (vn , Ku1 ) ≤ ∥K∥|u1 ||vn | ≤ ∥K∥, there
(1) (1)
is a subsequence vn s.t. (vn , Ku1 ) converges. Do this step by step, we can extract
n≥1 n≥1
(k+1) (k) (k+1)
subsequence vn of vn s.t. (vn , Kuk+1 ) converges. Taking the diagonal
n≥1
n≥1 n≥1
(k) (k)
i.e. vk , we have (vk , Kuj )k≥1 converges for all j ∈ N.
n≥1
(k)
Step 2. We claim that (vk , w) converges uniformly for any w ∈ K(BH ).
k≥1
Since A0 is countable and dense in K(BH ), then for any ϵ > 0, there are finite elements
n0
S
Ku1 , · · · , Kun0 , s.t. K(BH ) ⊂ B(Kui , ϵ). For these elements, there exists N0 > 0, whenever
i=1
k, m ≥ N0 , there is
(k) (m)
(vk , Kuj ) − (vm , Kuj ) ≤ ϵ for all 1 ≤ j ≤ n0 .
LECTURE NOTES ON FUNCTIONAL ANALYSIS 27
Then, for any w ∈ K(BH ), there exists j0 ∈ {1, ..., n0 }, s.t. |w − Kuj0 | < ϵ and hence
(k) (m)
(vk , w) − (vm , w)
(k) (k) (m) (k) (m) (m)
≤ (vk , w) − (vk , Kuj0 ) + (vk , Kuj0 ) − (vm , Kuj0 ) + (vm , Kuj0 ) − (vm , w)
(k) (k) (m) (m)
≤ vk w − Kuj0 + (vk , Kuj0 ) − (vm , Kuj0 ) + vm w − Kuj0
≤3ϵ.
(k)
Thus it is proved that (vk , w) converges uniformly for any w ∈ K(BH ).
k≥1
(k) (k)
Step 3. Since (K ∗ vk , u) = (vk , Ku), it follows from step 2 that
(k) (k) (k)
K ∗ vk − K ∗ vm
(m)
= sup (K ∗ vk − K ∗ vm
(m) (m)
, u) = sup (vk − vm , Ku) → 0 as k, m → ∞.
|u|≤1 |u|≤1
(k)
Since H is complete, there exists w̄ ∈ H s.t. K ∗ vk → w̄ as k → ∞. □
Proof. Indeed, λ1 K1 (BH ) + λ2 K2 (BH ) is precompact. Suppose (Kn )n≥1 in K(H) converges to
K ∈ L(H), and (un )n≥1 isan arbitrary
sequence inBH . Then, by the proof of Thm 3.5, we
(k) (k)
can extract a subsequence uk of (un )n≥1 , s.t. Kn uk converges for all n ∈ N. Since
k≥1 k≥1
(k) (k)
∥Kuk − Kn uk ∥ ≤ ∥K − Kn ∥ → 0, we have
(k) (l) (k) (l)
∥Kuk − Kul ∥ ≤ 2∥K − Kn ∥ + Kn uk − ul
→ 2∥K − Kn ∥ as k, l → ∞
→0 as n → ∞.
Remark 12.
(a) This theorem is also true in the Banach setting.
(b) When we were talking about the weak solution of an elliptic PDE, we defined a continuous
and coercive bilinear form a(u, v) over H01 (Ω). By Lax-Milgram theorem, for f ∈ H01 (Ω)
28 DONGMENG XI AND JIN LI
Define map K by Kf = u. It can be proved by the compact injection H01 (Ω) ⊂ L2 (Ω)
that K is compact.
(c) Fredholm alternative for I − K studies the eigenvalues of K, and hence the elliptic oper-
ators in (b).
(d) Examples. (Not closed range). By Proposition 3.6, it is easy to verify that the mapping
K : l2 → l2 defined by
x xn
2
K(x1 , x2 , . . . , xn , . . . ) = x1 , , . . . , , . . . ,
2 n
is a compact operator. Moreover, R(K) is a dense set of l2 since it contains a set
{(x1 , x2 , . . . , xn , . . . ) ∈ l2 : only finitely many xn are not zero} which is dense in l2 . But
R(K) is not closed since 1, 21 , . . . , n1 , . . . ∈ l2 \ R(K).
Proof. (1) If dim N (I − K) = +∞, one can select an orthonormal set {uk }k∈N ⊂ N (I − K).
Then uk − Kuk = 0 for all k ∈ N. It follows that |uk − uj |2 = 2 = |Kuk − Kuj |2 for all k ̸= j.
However this contradicts to the compactness of K, as (Kuk )k≥1 ⊂ K(BH ) would not contain any
convergent subsequence. Thus (i) is proved.
(2) We will prove that there exists c0 > 0 s.t. |u − Ku| ≥ c0 |u| for all u ∈ N (I − K)⊥ . Otherwise,
there exists (uk )k≥1 ⊂ N (I − K)⊥ with |uk | = 1 and |uk − Kuk | → 0 as k → ∞. Since K is
compact, there exists v ∈ H and subsequence (ukj )j≥1 , s.t. Kukj → v. Thus ukj → v ∈ BH .
Then, since K is continuous, we have Kv = v, and hence v ∈ N (I − K). However this would
imply (v, ukj ) = 0 for all j ∈ N, so letting j → ∞, (v, v) = 0 which is a contradiction!
(3) We claim that for any A ∈ L(H), v ∈ R(A), there exists u ∈ N (A)⊥ s.t. v = Au. Actually,
since for v ∈ R(A) there exists u0 ∈ H s.t. v = Au0 , let u = PN (A)⊥ u0 . Then, (u0 − u, w) = 0 for
all w ∈ N (A)⊥ . This implies u0 − u ∈ (N (A)⊥ )⊥ = N (A), and it follows that v = Au0 = Au.
Next let (vk )k≥1 ⊂ R(I − K) satisfying vk → v. There is (uk )k≥1 ⊂ N (I − K)⊥ s.t. uk − Kuk =
vk . By (2), we have |vn − vm | ≥ c0 |un − um | which implies that uk → ū for some ū. From the
continuity of I − K it follows that ū − K ū = v, i.e. v ∈ R(I − K). This means R(I − K) is
closed.
(4) Let A ∈ L(H). v ∈ N (A∗ ) is equivalent to (Au, v) = (u, A∗ v) = 0 for all u ∈ H. This is
equivalent to v ∈ R(A)⊥ . So take A = I − K, and note that (I − K)∗ = I − K ∗ , so together
with R(A) = (R(A)⊥ )⊥ , we obtain (ii).
we have already known that I − K is injective, so there exists u ∈ H\H1 , s.t. (I − K)u ∈ H1 ,
which cannot equal to (I − K)v for any v ∈ H1 . Therefore, letting Hk = (I − K)k H, we have
Hk ⊊ Hk−1 . Choose uk ∈ Hk ∩ Hk+1 ⊥ with |uk | = 1, k ∈ N. The existence of uk for all k ∈ N
holds by the projection theorem. In fact, there exists α ∈ Hk \Hk+1 , then we can set
uk = α − PHk+1 α α − PHk+1 α .
And one can easily confirm that such uk is the desired element. Then for any n > m, we have
Since −(un − Kun ) + (um − Kum ) + un is contained in Hm+1 , together with um ∈ Hm+1 ⊥ , we
get |Kun − Kum | ≥ |um | = 1 from parallelogram law. This contradicts to the compactness of K.
Thus we have proved that N (I − K) = {o} ⇒ R(I − K) = H.
Conversely, if R(I −K) = H, then by (ii) we have N (I −K ∗ ) = {o}, and hence R(I −K ∗ ) = H.
Thus finally, again by (ii), we get N (I − K) = {o}, and (iii) is proved.
(6) We claim that if A ∈ L(H) with dim R(A) < +∞, then A ∈ K(H). Actually, for all u ∈ BH ,
∥Au∥ ≤ ∥A∥, hence A(BH ) is bounded in R(A) which is finite-dimensional. Thus A(BH ) is
precompact.
(7) We will prove dim N (I −K) ≥ dim R(I −K)⊥ = dim N (I −K ∗ ). Otherwise, assume dim R(I −
K)⊥ > dim N (I − K). Then, we can define some map A : N (I − K) → R(I − K)⊥ to be a
one-to-one bounded linear map, and extend it to H by setting Au = 0 for all u ∈ N (I − K)⊥ .
Claim N (I − (A + K)) = {o}. Indeed, if u − (A + K)u = 0, then Au = (I − K)u ∈ R(I − K).
3.3. Spectrum.
Let T ∈ L(E), and E is a Banach space.
Definition 30. Resolvent set. The resolvent set of T denoted by ρ(T ) is defined by
Definition 31. Spectrum. The complement of resolvent set in R is called the spectrum, denoted
by σ(T ), i.e. σ(T ) = R\ρ(T ).
N (T − λI) ̸= {o}.
Remark 13.
Before studying the properties of spectrum, we invoke a result in the next section here. The
Corollary 4.7 tells us that for Banach spaces E, F , if L ∈ L(E, F ) and L is bijective, then
L−1 ∈ L(E, F ).
Proof of Lemma 3.9. Otherwise, there exists (en )n≥1 s.t. (ei , ej ) = δij , then |ei − ej |2 = 2 for any
i ̸= j, which implies (en )n≥1 does not have a convergent subsequence, a contradiction!. □
LECTURE NOTES ON FUNCTIONAL ANALYSIS 31
(3) Suppose (λk )k≥1 is a sequence of distinct numbers in σ(K) s.t. λk → λ. We will show λ = 0.
Indeed, since λk ∈ σ(K), there exists wk ̸= 0, s.t. Kwk = λk wk . Let Hk = span{w1 , · · · , wk }.
Then Hk ⊂ Hk+1 and Hk ̸= Hk+1 , for each k ≥ 1, since (wk )k≥1 are linear independent, and it
can be shown by induction. Observe that (K − λk I)Hk ⊂ Hk−1 for k ≥ 2. Choose an element
uk ∈ Hk for every k ≥ 2, with uk ∈ Hk−1 ⊥ , |uk | = 1 and choose u1 = |ww1
1|
. Now, if k > l, then
Hl−1 ⊊ Hl ⊂ Hk−1 ⊊ Hk . Thus,
Kuk Kul (K − λk I)uk (K − λl I)ul
− = − − ul + uk > 1,
λk λl λk λl
⊥
since
uk ∈ Hk−1 and (K − λk I)uk , (K − λl I)ul , ul ∈ Hk−1 . We get that if λk → λ ̸= 0, then
Kuk
λk
does not have a convergent subsequence, a contradiction! □
k≥1
(λu − T u, v) = (w, v) ∀v ∈ H.
(2) We will prove M ∈ σ(T ). Define [u, v] := (M u − T u, v). Then [·, ·] is a symmetric bilinear
form and [u, u] ≥ 0, ∀u ∈ H. Then, by a similar proof of the Cauchy-Schwarz inequality for inner
product, we have
[u, v]2 ≤ [u, u][v, v],
32 DONGMENG XI AND JIN LI
1 1
which means |(M u − T u, v)| ≤ (M u − T u, u) 2 (M v − T v, v) 2 holds for any u, v ∈ H. Letting
v = M u − T u, we get
1 1
|M u − T u|2 ≤ (M u − T u, u) 2 (M v − T v, M u − T u) 2
1 1
≤ ∥M I − T ∥ 2 |M u − T u|(M u − T u, u) 2 .
.
Then
1 1
|M u − T u| ≤ c(M u − T u, u) 2 where c = ∥M I − T ∥ 2 .
If M ∈ ρ(T ), by Cor 4.7, (M I − T )−1 ∈ L(H). Let (uk )k≥1 be a sequence in H satisfying |uk | = 1
and (T uk , uk ) → M as k → ∞. Then we get
1
|M uk − T uk | ≤ c(M − (T uk , uk )) 2 → 0.
Hence
|uk | = |(M I − T )−1 (M uk − T uk )| ≤ ∥(M I − T )−1 ∥ · |M uk − T uk | → 0,
which contradicts to the assumption that |uk | = 1. Therefore, M ∈ σ(T ), and likewise we have
m ∈ σ(T ). □
Theorem 3.11. Let H be a separable Hilbert space and T ∈ K(H) be self-adjoint. Then there
exists a countable orthonormal basis of H composed of eigenvectors of T .
Proof. By Thm 3.8, let (λk )k≥1 comprise the sequence of distinct eigenvalues of T , excepting 0.
Set λ0 = 0. Write H0 = N (T ) and Hk = N (T − λk I), k ≥ 1. Then, by Thm 3.7, 0 < dim Hk <
∞, 0 ≤ dim H0 ≤ ∞, k ≥ 1. If u ∈ Hk , v ∈ Hl for k ̸= l, then
inf (T u, u) = sup (T u, u) = 0.
u∈F ⊥ u∈F ⊥
2(T u, v) = (T (u + v), u + v) − (T u, u) − (T v, v) ∀v ∈ F ⊥ .
Theorem 4.1 (Baire category theorem). Let X be a complete metric space and let {Xn }n≥1 be
a sequence of closed subsets in X. Assume that
then
int (∪∞
n=1 Xn ) = ∅.
Remark 14.
(1) A set A is called no where dense, if int Ā = ∅.
(2) If On is open for all n ≥ 1, and Ōn = X, then ∩∞n=1 On = X.
∞
(3) If ∪n=1 Xn = X, then there must exist an n0 s.t. int Xn0 ̸= ∅.
Definition 34. linear operators space. Let E, F be two n.v.s.. Denote L(E, F ) to be the space
of continuous(=bounded) linear operators from E into F with norm
Remark. A brief statement is that one can derive a global estimate from pointwise estimates.
It follows that
1
∥Ti (z)∥ ≤ (n0 + ∥Ti (x0 )∥) for all z ∈ B(o, 1).
r
This together with (4.1), implies sup ∥Ti ∥ < ∞. □
i∈I
(2) By (1), ∥Tn x∥ ≤ c∥x∥ for some universal constant c > 0. It follows from this and ∥Tn x −
T x∥ → 0 that ∥T x∥ ≤ c∥x∥. Then T ∈ L(E, F ).
(3) Since ∥Tn x∥ ≤ ∥Tn ∥∥x∥ for all n ∈ N, we have ∥T x∥ = lim ∥Tn x∥ ≤ lim ∥Tn ∥∥x∥ for all
n→∞ n→∞
x ∈ E. It follows that ∥T ∥ ≤ lim ∥Tn ∥. □
n→∞
36 DONGMENG XI AND JIN LI
Corollary 4.4. Let G be a Banach space and let B ∗ be a subset of G∗ . Assume that
Then B ∗ is bounded.
Proof. It is the direct corollary of Theorem 4.2 for E = G, F = R and the family {Ti : i ∈ I} =
B∗. □
Then B is bounded.
Proof. Recall that x defines a Jx ∈ E ∗∗ s.t. ⟨Jx, f ⟩ = ⟨f, x⟩ for all f ∈ E ∗ and ∥Jx∥ = ∥x∥
(follows from Corollary 5.5). Notice that ⟨JB, f ⟩ is bounded. By Corollary 4.4, JB is bounded
and hence B is bounded. □
4.3. The open mapping theorem and The closed graph theorem.
Theorem 4.6 (Open mapping theorem). Let E, F be two Banach spaces and let T ∈ L(E, F )
that is surjective. Then there exists a constant c > 0 such that
Remark 15.
(1) If we in addition assume that T is injective, then T −1 will also be continuous. Hence T
is a homeomorphism.
(2) Let U be open, and let y0 ∈ T (U ). Then there exists B(x0 , r) ⊂ U with T x0 = y0 and
r > 0. Then T (U ) ⊃ T (x0 ) + T (B(o, r)) ⊃ y0 + B(o, c) = B(y0 , c), which means T (U )
is open. That is, T maps open sets to open sets. Reversely, if a linear map T : E → F
maps open sets to open sets, then T must be surjective.
(3) Clearly, a bijective T maps open sets to open sets if and only if it maps closed sets to
closed sets. However, if T ∈ L(E, F ) is only surjective but not injective, then T may not
map closed sets to closed sets. E.g. Let T : R2 → R such that T (x1 , x2 ) = x1 . Then
C = {(x1 , x2 ) | x1 > 0, x2 = 1/x1 } is closed but T (C) is not closed.
Proof. Step 1. Firstly, we prove a weak version: If T is linear and onto, then there exists constant
c > 0 s.t.
T (B(o.1)) ⊃ B(o, 2c). (4.2)
LECTURE NOTES ON FUNCTIONAL ANALYSIS 37
and hence
x0 + rz1 + (−x0 ) + rz2 ∈ 2Xn0 for all z1 , z2 ∈ B(o, 1).
This implies that
rz ∈ Xn0 for all z ∈ B(o, 1).
r
Let c = 2n0
, we have B(o, 2c) ⊂ T (B(o, 1)).
Step 2. Next we show that: Assume T ∈ L(E, F ) that satisfies (4.2). Then
Now we show x̄ ∈ BE (o, 1), and by the arbitriness of y ∈ BF (o, c) our proof is finished.
Actually, we denote d = 21 − ∥x1 ∥ > 0, since ∥xk ∥ < 21k for all k ∈ N, 1 − ∥x̄∥ ≥ 12 − ∥x1 ∥ + 21 −
∥ ∞
P
k=2 xk ∥ ≥ d > 0. Thus x̄ ∈ BE (o, 1). □
Corollary 4.7. If T ∈ L(E, F ) in Thm. 4.6 is additionally bijective, then T is a homeomorphism
between E and F .
Proof. Since T has inverse T −1 which is linear, by Thm 4.6,
Corollary 4.8. Let E be a v.s. equipped with two norms ∥ · ∥1 and ∥ · ∥2 . If E is complete for
both norms, and if there exists c ≥ 0 s.t.
then the two norms are equivalent, i.e. there exists c′ > 0 s.t.
1
∥x∥2 ≤ ∥x∥1 ≤ c′ ∥x∥2 for all x ∈ E.
c′
(Equivalence will imply that the topolgies generated by the norms are the same.)
Proof. Let E = (E, ∥ · ∥1 ), F = (E, ∥ · ∥2 ) and T = I. Then there exists constant c̃ s.t. BE (o, 1) ⊃
BF (o, c̃), which means
c̃x
≤ 1.
∥x∥2 1
Thus∥x∥1 ≤ 1c̃ ∥x∥2 , and now we can set c′ = max c, 1c̃ .
□
Theorem 4.9 (Closed graph theorem). Let E, F be two Banach spaces and let T be a linear
operator from E into F . Assume that the graph of T , G(T ) := {(x, T x) ∈ E × F : x ∈ E} is
closed in E × F . Then T is continuous.
Remark 16.
(1) The graph of any continuous map is closed.
(2) The norm in E × F is defined by
Now from ∥x∥1 ≤ ∥x∥2 and Corollary 4.8, there exists c > 0 s.t. ∥x∥2 ≤ c∥x∥1 and c > 1
because ∥ · ∥F doesn’t vanish. Thus ∥T x∥F ≤ (c − 1)∥x∥E , which means T ∈ L(E, F ). □
N (A) × {o} = G ∩ L
E × R(A) = G + L
{o} × N (A∗ ) = G⊥ + L⊥
R(A∗ ) × F ∗ = G⊥ + L⊥
Under these assumptions, there exists a linear functional f defined on E that extends g, i.e.
g(x) = f (x) for all x ∈ G and f (x) ≤ p(x) for all x ∈ E.
we have
g(x) − p(x − x0 ) ≤ p(y + x0 ) − g(y).
Thus let sup g(x) − p(x − x0 ) ≤ α ≤ inf p(y + x0 ) − g(y) . Then for all x ∈ G,
x∈G y∈G
Hence
x
h(x + tx0 ) = g(x) + tα ≤ t p + x0 = p(x + tx0 ).
t
Thus we have extended g from G to span{G, x0 }. Since E is finite dimensional, through finitely
many such steps we can extend g to E.
Remark 19. One geometric meaning of Theorem 5.1 can be stated as below.
Suppose p(x) ≥ 0 for all x ∈ E. Let B = {x : p(x) ≤ 1}. Define Hg ⊂ G by Hg = {x ∈
G : g(x) = 1}. It follows that Hg ∩ B = ∅. Then the statement of Thm 5.1 means that one can
extend Hg to Hf ⊂ E s.t. Hf ∩ B = ∅.
Lemma 5.2 (Zorn). Every nonempty ordered set that is inductive has a maximal element.
It is well-defined since if x ∈ D(h) ∩ D(h′ ) for h, h′ ∈ Q, we must have h(x) = h′ (x) for that Q
is totally ordered. Then it is clear that h̄ is linear and h ≤ h̄ for all h ∈ Q, i.e. h̄ is an upper
bound of Q.
Step 2. By Zorn’s Lemma, there exists a maximal element in P , say f . We claim that D(f ) = E.
Otherwise, there exists x0 ∈ E\D(f ). Let D′ = span{D(f ), x0 } and let F (x + tx0 ) = f (x) + tα,
where α is chosen s.t.
Remark. Taking maximal element is a useful tool in dealing with infinite dimensional case.
Recalling the conception of Dual space, now we give several related properties based on Hahn-
Banach theorem.
Proof. Take p(x) = ∥g∥G∗ ∥x∥. Applying Thm 5.1, there exists f s.t. f ≤ p, f = g on G. □
Proof. Take G = Rx0 , then g(tx0 ) = t∥x0 ∥2 . Thus ∥g∥G∗ = ∥x0 ∥. Now by Cor 5.3, there exists
f0 ∈ E ∗ s.t. ∥f0 ∥ = ∥g∥G∗ = ∥x0 ∥ and ⟨f0 , x0 ⟩ = g(x0 ) = ∥x0 ∥2 . □
By Corollary 5.4, there exists f0 ∈ E ∗ , s.t. ∥f0 ∥ = ∥x∥ and ⟨f0 , x⟩ = ∥x∥2 . Let f1 = f0 /∥f0 ∥,
then |⟨f1 , x⟩| = ∥x∥ and ∥f1 ∥ = 1, which means
□
LECTURE NOTES ON FUNCTIONAL ANALYSIS 43
H = {x ∈ E : f (x) = α}
where f is some linear functional that does not vanish identically and α ∈ R. For simplicity,
write H = [f = α].
We call a set half space with form
For an arbitrary ϵ > 0, there exist α1 > 0, α2 > 0, s.t. for all x ∈ α1 A, y ∈ α2 A,
x + y ∈ α1 A + α2 A = (α1 + α2 )A.
Then p(x + y) < α1 + α2 < p(x) + p(y) + 2ϵ. By the arbitrariness of ϵ, we obtain p(x + y) ≤
p(x) + p(y).
Now, since x0 ∈ / A, we have p(x0 ) ≥ 1. Let g(tx0 ) = t. Then, g is linear on G = Rx0 and
g(tx0 ) = t ≤ p(tx0 ) = tp(x0 ) for any t > 0. By Theorem 5.1, there exists f extending g s.t.
is open. Since A ∩ B = ∅, o ∈
/ A − B. Let z0 ∈ A − B, by Step 1 there exists f that is continuous
and s.t. H = [f = 1] separates A − B − {z0 } and {−z0 }. Precisely, f (−z0 ) = 1 and for all
x − y − z0 ∈ A − B − {z0 }, there is
f (x − y − z0 ) ≤ p(x − y − z0 ) < 1.
This implies f (x) < f (y) for all x ∈ A, y ∈ B. Find an α ∈ R s.t. sup f (x) ≤ α ≤ inf f (y), then
x∈A y∈B
we complete the proof with the hyperplane [f = α]. □
Remark. If A, B are only assumed to be closed, their distance can tend to zero, and thus we
can’t strictly separates them.
LECTURE NOTES ON FUNCTIONAL ANALYSIS 45
f (x − y) ≥ α for all x ∈ A, y ∈ B,
f (rz) ≤ α holds for all z ∈ B(o, 1), and then holds for all ∥z∥ ≤ 1.
ϵ
Then there exists β s.t. inf x∈A f (x) − 2
≥ β ≥ supy∈B f (y) + 2ϵ . This means that Hβ = [f = β]
strictly separates A and B. □
Corollary 5.9. Let F ⊂ E be a linear subspace s.t. F̄ ̸= E, then there exists some f ∈ E ∗ , f ̸≡ 0,
s.t.
⟨f, x⟩ = 0 for all x ∈ F.
Remark. Corollayr 5.9 extends F to a closed space, which is actually a hyperplane H = [f = 0].
Remark 20. By Corollary 5.9, we find that if one can show that every continuous linear functional
on E that vanishes on F must vanish on E, then one can deduce that F is dense in E.
Proof. Let x0 ∈ E\F. By Thm 5.8 with A = F̄ and B = {x0 }, there is a closed hyperplane
[f = α] that strictly separates F̄ and {x0 }. Then for all x ∈ F ,
.
Remark 21. As Proposition 2.5 shows, J is linear and bounded, and furthermore J is an isometry,
i.e. ∥Jx∥E ∗∗ = ∥x∥E . Actually, by Cor 5.5,
J may not be surjective from E onto E ∗∗ (see Chapter 3 and 4). We can identify E with a
subspace of E ∗∗ .
Definition 40. Reflexive. We say E is reflexive if J(E) = E ∗∗ . In this case, we also write
E = E ∗∗ .
Definition 41. Perpendicularity. If M ⊂ E is a subspace,
If N ⊂ E ∗ is a subspace,
(M ⊥ )⊥ = M̄ .
(N ⊥ )⊥ ⊃ N̄ . (5.1)
Proof. (1) Suppose x ∈ M . Then ⟨f, x⟩ = 0 for all f ∈ M ⊥ , which means x ∈ (M ⊥ )⊥ . Since
(M ⊥ )⊥ is closed, we have M̄ ⊂ (M ⊥ )⊥ .
Suppose x ∈ (M ⊥ )⊥ . Then
/ M̄ , there exists H = [f0 = a0 ] with f0 ∈ E ∗ that strictly separates {x} and M̄ . It follows
If x ∈
that
⟨f0 , x⟩ > α0 > ⟨f0 , y⟩ for all y ∈ M̄ . (5.3)
Since ⟨f0 , λy⟩ < α0 for all λ ∈ R, we get ⟨f0 , y⟩ = 0 for all y ∈ M . This means that f0 ∈ M ⊥ .
However, by(5.2), ⟨f0 , x⟩ must also be zero. This contradict to (5.3), hence x must be in M̄ .
6. Lp spaces
1◦ When we talking about these function spaces, we automatically have assumed that we are
working on (Rn , M, µ), where µ is Lebesgue measure on Rn and M is the σ-algebra consisting
of Lebesgue measurable sets.
2◦ Sometimes we write dx instead of dµ. We use this notation because most results hold in
general measure space (X, Σ, µ).
Definition 42. Lp spaces. Let p ∈ [1, ∞), Ω ∈ M. Set
( Z p1 )
Lp (Ω) = f : Ω → R f is measurable and ∥f ∥p := |f |p dµ <∞ .
Ω
Here ∥ · ∥p is a norm follows from the Minkowski’s inequality which will be showed below.
Set
( )
f is measurable and there is a constant c ∈ R
L∞ (Ω) = f : Ω → R
such that |f (x)| ≤ c a.e. on Ω.
and ∥f ∥L∞ = ∥f ∥∞ =: inf{c : |f (x)| ≤ c a.e. on Ω}. For example, we define f : Ω → R by
1 x∈/ N+
f (x) =
x x ∈ N+
Then there isn’t C ∈ R s.t. |f (x)| ≤ C for all x ∈ R, but |f (x)| ≤ 1 a.e. x ∈ Ω.
Exercise : Assume µ(Ω) < ∞ and f ∈ Lp (Ω) for all p ≥ 1.
1◦ Prove that limp→∞ ∥f ∥p = ∥f ∥∞ .
2◦ Prove that ∥ · ∥∞ is a norm on L∞ (Ω).
Definition 43. Conjugate number. Let p ∈ [1, ∞], the number p′ satisfying
1 1
+ ′ =1
p p
will be called the conjugate exponent, especially when p = 1, p′ = ∞ and when p = ∞, p′ = 1.
′
Theorem 6.1 (Hölder’s inequality). Assume f ∈ Lp , g ∈ Lp with p ∈ [1, ∞]. Then f · g ∈ L1
and Z
|f · g|dx ≤ ∥f ∥p ∥g∥p′ .
′
Remark 22. When p ∈ (1, ∞), equality holds in Hölder’s inequality iff |f |p = c|g|p a.e. for some
constant c ≥ 0.
Remark 23. By Hölder’s inequality, one can confirm that Lp ⊂ Lq for any 1 ≤ q ≤ p.
(1) (J ◦ f ) is in L1 (Ω);
1 1
R R
(2) µ(Ω) Ω
J ◦ f (x)dµ(x) ≥ J µ(Ω) Ω
f (x)dµ(x) .
∥f + g∥p ≤ ∥f ∥p + ∥g∥p .
≤ |f + g|p−1 p′
∥f ∥p + |f + g|p−1 p′
∥g∥p (6.1)
∥f + g∥p ≤ ∥f ∥p + ∥g∥p .
Theorem 6.4 (Riesz’s respresentation theorem of Lp version). Suppose p ∈ (1, ∞). Let ϕ ∈
(Lp )∗ . Then there exists a unique function u ∈ Lp s.t.
′
Z
⟨ϕ, v⟩ = u · vdµ, for all v ∈ Lp .
We omit its proof here. And by this theorem, it is clear that when p ∈ (1, ∞), Lp is reflexive
and (Lp )∗ coincides with Lp .
′
Letting χB(o,k) (x) · fN (x) =: hk (x), we have |hk − fN |p → 0 a.e. and |hk − fN |p ≤ |2fN |p . Thus
there exists h s.t.
∥f − h∥p < 2ϵ, ∥h∥∞ < ∞ and supp h is compact.
For such function h, by Proposition 6.5 there exists g ∈ Cc (Rn ), s.t. ∥g − h∥1 ≤ ϵ and
∥g∥∞ ≤ ∥h∥∞ . Then
Z Z Z
p p−1
|h − g| dµ = |h − g||h − g| dµ ≤ |h − g|dµ ∥2h∥p−1∞ ,
and hence 1
1− 1 1
∥h − g∥p ≤ ∥h − g∥1p ∥2h∥∞ p ≤ cϵ p .
1
In the end, for any ϵ > 0, there exists g ∈ Cc (Rn ) s.t. ∥f − g∥p ≤ 2ϵ + cϵ p , which is exactly
our desired result. □
R
Theorem 6.7. Suppose j ∈ Cc (Rn ) with j ≥ 0 and Rn jdx = 1. For k ∈ N, define jk (x) =
R
k n j(kx), so that Rn jk dx = 1, and ∥jk ∥1 = ∥j∥1 . Let f ∈ Lp (Rn ) for some p ∈ [1, ∞), and define
Z
fk := jk ∗ f = jk (x − y)f (y)dy.
Rn
Then
Remark 25. If f is supported on a compact set K ⊂ Ω, and supp j = B, then supp jk = k1 B and
supp f ∗ jk ⊂ K + k1 B.
Step 2. If f ∈ Cc (Rn ), then there exists N > 0 s.t. supp f, supp j ⊂ B(o, N ). It follows that
Z
f (x) − fk (x) = (f (x)jk (y) − f (x − y)jk (y))dy
B(o, N
k
)
Z
≤ f (x) − f (x − y) jk (y)dy.
B(o, N
k
)
Since |fk ∗ h − h| → 0 uniformly by Step 2, then by Step 3 there exists N > 0 s.t. when k ≥ N
Z Z
|jk ∗ h − h|dx = |jk ∗ h − h|dx ≤ ϵ.
supp h+B(o, N
k
)
Definition 44. Locally p-th integrable function spaces. We define locally p-th integrable function
spaces by
Corollary 6.8. Let Ω ⊂ Rn be open and let u ∈ L1loc (Ω) be such that
Z
u · f dµ = 0 for all f ∈ Cc∞ (Ω).
Then u = 0 a.e. on Ω.