course-Exercises-M2206
course-Exercises-M2206
Lecture Notes
c This document can be reproduced in its entirely for non commercial use.
Contents
3
CHAPTER 1
1. Vector spaces
It is assumed that the reader has already encountered vector spaces before but a brief
summery of the theory is provided here for reference purposes. In this section, X denotes
a vector space over K, (K = R or C). In this context the real or the complex numbers
will be called scalars.
Definition 1.1. A vector space X is a non empty set on which are defined an
operation of vectors + : X × X 7→ X satisfying associativity, commutativity, zero and
inverse axioms, and an operation of scalar multiplication × : K × X → X that satisfies
the following laws, for every x, y, z ∈ K and λ, µ ∈ K,
(1) x + (y + z) = (x + y) + z,
(2) x + y = y + x,
(3) 0X + x = x,
(4) x + (−x) = 0X ,
(5) λ(x + y) = λx + λy,
(6) (λ + µ)x = λx + µx,
(7) (λµ)x = λ(µx),
(8) 1x = x.
2. Normed spaces
In this section, we assume that K denotes one of the fields R or C, and that X is a
vector space over K.
2.1. Norms.
Definition 2.1. We call a norm on a vector space X any mapping N : X 7−→ R+
(in general we write N (x) = kxk) verifying, for every x, y ∈ X and λ ∈ K, the following
three properties:
(N1 ) N (x) = 0 ⇔ x = 0 .
(N2 ) N (λx) = |λ|N (x).
(N3 ) N (x + y) ≤ N (x) + N (y).
A vector space X endowed with a norm N is called a normed vector space (n.v.s.) and
we denote it by the pair (E, N ). If necessary, norms on different spaces are distinguished
by subscript such that as k · kX . A positive function that satisfies (N 2) and (N 3) is called
semi-norm.
Example 2.2.
(1) [Usual norm in R or C] The absolute value function x 7→ |x| defines a norm
on R and the modulus function z 7→ |z| defines a norm on C. These are called
the usual norms.
(2) The mappings defined, for every x = (x1 , · · ·, xn ) ∈ Rn :
n n
!1/2
X X
kxk1 = |xi | , kxk2 = |xi |2 and kxk∞ = max |xi | .
1≤i≤n
i=1 i=1
are norms on Rn .
(3) Let a, b ∈ R (a < b) and X = C([a, b], R) the vector space of real valued contin-
uous functions over [a, b]. The mappings defined, for every f ∈ X, by
Z b Z b 1/2
2
kf k1 = |f (x)| dx , kf k2 = |f (x)| dx and kf k∞ = max |f (x)| .
a a x∈[a,b]
are norms on X.
Remark 2.3.
(1) We denote by C([a, b], K) = {f : [a, b] → K : f is continuous } with K = R or C.
In the case of K = R, we write C[a, b] instead of C([a, b], K) for simplicity.
(2) [Induced norm] Let F be a vector subspace of a normed space (X, N ). The
restriction of N to F is a norm on F called the induced norm of (X, N ) on F .
2. NORMED SPACES 7
2.2. Topology induced by the norm. Using the properties of the norm in Defini-
tion 2.1, it is easy to verify that (the proof is left as an exercise)
Proposition 2.4. Normed spaces are metric spaces: Let (X, k · k) be a normed
space. The function d : X × X → [0, ∞) defined by d(x, y) = kx − yk defines a distance
on X. This distance is called the distance induced by the norm k · k.
The previous proposition guaranties that every normed vector space is a metric space.
However, the converse is not necessary true. To clarify this, we make the following
definition.
Definition 2.5. Distance induced by a norm: Suppose that (X, d) is a metric
space and X is a vector space on K. The distance d is said to be induced by a norm if
there exists a norm k · k in X such that, for all (x, y) ∈ X × X, dk·k (x, y) = kx − yk.
It is worth to mention that under some conditions, a distance can provide a norm.
Proposition 2.6. Let E be a vector space endowed with a distance d verifying two
additional properties:
(1) d is homogeneous i.e. ∀x, y ∈ E, ∀λ ∈ K, d(λx, λy) = |λ|d(x, y).
(2) d is invariant by translation i.e. ∀x, y, a ∈ E, d(x + a, y + a) = d(x, y).
Then the mapping defined by
N : E −→ R+
x 7−→ N (x) = d(x, 0)
is a norm over E.
Proof. Left as an exercise.
Example 2.7.
(1) [Usual norm in R] The distance induced by the the usual norm | · | on R is the
usual distance on R, that is (x, y) 7−→ d(x, y) = |x − y|.
(2) The discrete distance in R is not induced by a norm.
Since every normed space is a metric space, we get a notion of open sets in normed
spaces, as well as many other topological notions.
Definition 2.8. [Topology induced by the norm] Let (X, k · k) be a normed
space and dk·k the distance induced by the norm k · k. We introduce the following class of
subsets of X:
Tk·k = {A ⊂ X : A is open in (X, dk·k )} .
The class Tk·k is called the topology induced by k · k.
Remark 2.9. [Topological notions in normed spaces] Let (X, k · k) be a normed
vector space and dk·k be the distance induced by k · k. Since (X, dk·k ) is a metric space,
then we can talk about various topological notions in X, for example:
(1) [Convergence] A sequence (un ) is convergent in (X, k · k) if ∃ u ∈ X such that
kun − uk → 0 ;
(2) [Cauchy sequences] A sequence (un ) is Cauchy in (X, k · k) if, for all ε > 0,
∃ n0 ∈ N, ∀ p ≥ q ≥ n0 =⇒ kup − uq k < ε.
(3) [Continuity] A function f : (X, k · k) → (R, | · |) is continuous at u0 ∈ X if
∀ ε > 0, ∃ δ > 0, ku − u0 k < δ =⇒ |f (u) − f (u0 )| < ε ;
8 1. NORMED VECTOR SPACES
(2) It is worth noting that if the sequence (fn ) ⊂ C([a, b], R) converges uniformly to
a function f on [a, b], then the then f ∈ C([a, b], R) (see Theorem 7.4). Con-
sequently, the uniform convergence means convergence in C([a, b], R), k · k∞ ,
where k · k∞ is the norm introduced in Example 2.2.
(3) The norm k · k∞ is also called the norm of uniform convergence.
Remark 3.2.
(1) In many situations, it is customary to skip the reference to the norm k · k from
the notation of the balls and simply write B(a, r) and B[a, r].
(2) [The unit balls] For a = 0 and r = 1, the balls B(0, 1) and B[0, 1] are called
the unit open and closed balls respectively.
The algebraic structure in a vector space induce a geometric property which is the
convexity. Recall that X is assumed a vector space (on the field K ∈ {R, C}).
Definition 3.3. [Convex set]
Proof. Let (X, k · k) be a normed space, a ∈ X and r > 0. We will prove that B(a, r)
and B[a, r] are convex. Let x ∈ B(a, r), y ∈ B(a, r) and t ∈ [0, 1]. We have,
tx+(1−t)y −a = t(x−a)+(1−t)(y−a) ≤ tkx−ak+(1−t)ky−ak < tr+(1−t)r = r ,
i.e. tx + (1 − t)y ∈ B(a, r). This proves that B(a, r) is convex. The proof of the convexity
of B[a, r] is similar.
Remark 3.5. A potential application of Theorem 3.4 can be as follows. If A ⊂ X is
not convex, then A can not be a (closed or open) ball in X relative to any norm. Now if
d is a distance in X such that Bd (a, r) is not convex for some a ∈ X and r > 0, then the
distance d is not induced by a norm.
In a normed space, the property of boundedness takes the pleasant form described in:
Theorem 3.6. [Bounded subsets] Let (X, k · k) be a normed space. The following
two statements are equivalent:
(1) A ⊂ X is bounded in (X, k · k) ;
(2) ∃ C > 0, ∀ x ∈ A, kxk ≤ C .
Proof. Suppose that A ⊂ X. If A is bounded in (X, k · k), then A ⊂ B[a, M ] for
some a ∈ X and M > 0. Let C = kak+M . We have, for all x ∈ A, kxk ≤ kx−ak+kak ≤
M + kak = C. Conversely, suppose that, for all x ∈ A, kxk ≤ C. We have A ⊂ B[0, C].
Hence A is bounded in (X, k · k).
Exercise 1.1. Let (X, k · k) be a normed space.
(1) Let S ⊂ X and λ ∈ C. Show that, if S is bounded in (X, k · k), then the set
λS = {λx : x ∈ S} is bounded in (X, k · k).
(2) Let S ⊂ X and T ∈ X. Show that, if S and T are bounded in (X, k · k), then
the set S + T = {x + y : x ∈ S and y ∈ T } is bounded in (X, k · k).
In the next theorem, we will see that the algebraic operations in a normed space will
force it to be unbounded.
Theorem 3.7. [A normed space is unbounded]
Let (X, k · k) be a normed space and X 6= {0}.
(1) For every α > 0, there exists x ∈ X such that kxk = α.
(2) X is not bounded in (X, k · k).
(3) Every non-null vector subspace of X is not bounded in (X, k · k).
Proof. The proof involves the trick of normalization of a non-zero vector. Let α > 0,
α
y ∈ X and y 6= 0. Let x = y (notice the use of the algebraic operations in X). Clearly,
kyk
x ∈ X and kxk = α.
If X is bounded, then there exists M > 0 such that, for all x ∈ X, kxk ≤ M . Select
x ∈ X such that kxk = M + 1. Consequently, M + 1 = kxk ≤ M , i.e. 1 ≤ 0 which is a
contradiction.
Every non-null subspace is a normed space, hence not bounded by the above consid-
erations.
Remark 3.8. Since compact metric spaces are bounded, we get in light of Theorem 3.7
that every normed space (X, k · k) is not compact.
5. EQUIVALENT NORMS 11
are norms on X.
The proof of Theorem 4.1 is left as an exercise.
5. Equivalent Norms
In this section, we will study the notion of equivalent norms. These are norms having
distinct expressions but defined on the same vector space. We shall see that this notion
is highly non-trivial in infinite dimensional spaces.
5.1. Generalities.
Definition 5.1. [Equivalent norms]
Let N1 and N2 be two norms in a vector space X. The norms N1 and N2 are said to
be equivalent if there exist two constants 0 < c1 < c2 such that,
∀ x∈X, c1 N1 (x) ≤ N2 (x) ≤ c2 N1 (x) . (5.1)
12 1. NORMED VECTOR SPACES
Example 5.2. The norms N∞ , N1 and N2 , defined in Theorem 4.1, are equivalent on
the product space X. More precisely we have
√
∀ x ∈ X , N∞ (x) ≤ N1 (x) ≤ ` N2 (x) ≤ ` N∞ (x) .
Remark 5.3.
(1) In (5.1), the constants c1 and c2 are fixed while the point x is allowed to vary in
X. The strength of (5.1) is that the inequality remains true when x varies. In
other words, the constants c1 and c2 do not depend on the variable x.
(2) If N1 and N2 are norms on X, then they induce two distances dN1 and dN2
respectively. The norms N1 and N2 are equivalent if and only if the distances
dN1 and dN2 are equivalent too.
(3) Consequently, when N1 and N2 are equivalent norms, then they induce the same
topology, i.e. TN1 = TN2 . The converse is also true. We will present an elegant
proof of this in Chapter 2. Another elementary proof will be given in Exercise 3.
5.2. Case of finite dimension. In this section we discuss finite dimensional normed
vector spaces. These spaces share many properties with the Euclidean spaces, in particular
Theorems 5.5 and 6.1.
Our discussion of finite dimensional spaces relies on:
Lemma 5.4. Let (X, k·k) be a normed vector space of finite dimension d. Suppose that
{e1 , · · · , ed } is a basis of X. For all u = (u1 , · · · , ud ) ∈ Rd , define f (u) = u1 e1 +· · ·+ud ed .
(1) The function N (u) = kf (u)k is a norm on Rd ;
(2) ∃ c1 , c2 > 0 such that, ∀ u ∈ Rd , c1 kuk2 ≤ N (u) ≤ c2 kuk2 , where k · k2 is the
Euclidean norm in Rd . ;
(3) The function f : (Rd , k · k2 ) → (X, k · k) is a homeomorphism, i.e. f and f −1 are
continuous.
Proof. That N is a norm on Rd is left as an exercise to the reader. The existence of
the constants c1 and c2 follows from Theorem 5.5 which states that the norms N (·) and
k · k2 are equivalent.
Since {e1 , · · · , ed } is a basis of X, then f is bijective. The continuity of f and f −1
follows from the inequality in Item (2) above; actually, f and f −1 are Lipschitz functions.
Theorem 5.5. [Equivalent norms in finite dimension spaces] Let N be a norm
in Rd . The norm N and the Euclidean norm
q
x = (x1 , x2 , · · · , xd ) 7→ kxk2 = x21 + x22 + · · · + x2d
are equivalent norms.
Proof. Step 1. Let
e1 = (1, 0, · · · , 0) , e2 = (0, 1, 0, · · · , 0) , ed = (0, · · · , 0, 1) ,
be the canonical basis of R . For all x = (x1 , x2 , · · · , xd ) ∈ Rd , we have
d
d
X
x= xk ek ,
k=1
and by triangle’s inequality we have,
X
N (x) ≤ |xk |N (ek ) .
k=1
5. EQUIVALENT NORMS 13
6.2. Case of infinite dimension. In the Euclidean space Rd , the Hein-Borel theo-
rem tells us that the unit ball B[0, 1] is compact. This fact will not be true in an infinite
dimensional space. The Riesz compactness theorem will clarify this, as it will state that
the unit ball in a normed space is compact if and only if the space is finite dimensional.
The lengthy proof of this theorem will occupy the whole of this section.
Theorem 6.5. [Riesz’s theorem]
Suppose that (X, k · k) is a normed space. The unit closed ball B[0, 1] in (X, k · k) is
compact if and only if the vector space X has a finite dimension (i.e. dim(X) is finite).
The proof of Theorem 6.5 is split into several lemmas. In the rest of this section,
(X, k · k) denotes a normed space.
Lemma 6.6. Suppose that M is a closed proper subspace of X. For all ∈]0, 1[, there
exists y ∈ X such that
kyk = 1 and dist(y, M ) ≥ 1 − .
Proof. Since M 6= X, there exists x ∈ X \ M . Let α = dist(x, M ). Since M is closed
and x 6∈ M , M = M and α = dist(x, M ) > 0.
α
Let ∈]0, 1[ be fixed. Note that α = dist(x, M ) = inf{kx−mk : m ∈ M } and α ≤ 1− .
By the definition of the infimum, we get m0 ∈ M such that α ≤ kx − m0 k ≤ α/(1 − ).
Now, we define y = kx − m0 k−1 (x − m0 ). Clearly kyk = 1. Let m ∈ M . We will prove
that ky − mk ≥ 1 − . Since M is a subspace, m0 + kx − m0 km ∈ M . Consequently, by
definition of α = dist(x, M ),
ky − mk = kx − m0 k−1 x − m0 − kx − m0 km ≥ kx − m0 k−1 α .
Since kx−m0 k ≤ α/(1−), kx−m0 k−1 ≥ α−1 (1−) and consequently, ky−mk ≥ 1−.
Lemma 6.7. Suppose that 0 < < 1. If dim(X) = +∞, there exist a sequence (un )
in X and a sequence of subspaces (Mn ) such that, for all n ≥ 2,
• kun k = 1 ;
• un ∈ Mn ;
• dist(un , Mn−1 ) ≥ 1 − ;
• Mn ⊂ Mn+1 and dimMn = n .
Proof. Choose u1 ∈ X such that ku1 k = 1. Let M1 = Span(u1 ). M1 is a subspace
of X, dim(M1 ) = 1 and M1 6= X since dim(X) = +∞. Furthermore, M1 is closed
by Theorem 6.3. Applying Lemma 6.6, we obtain u2 ∈ X such that ku2 k = 1 and
dist(u2 , M1 ) ≥ 1 − ε.
Next, we define M2 = Span(u1 , u2 ). Clearly, M1 ⊂ M2 , dim(M2 ) = 2 and M2 is closed
by Theorem 6.3. Furthermore, M2 6= X since dim(X) = +∞. Applying Lemma 6.6,
16 1. NORMED VECTOR SPACES
7. Banach spaces
In this section, we discuss the completeness in the case of infinite dimensional normed
vector spaces. More precisely, we study the convergence of Cauchy sequences in (C[0, 1], k · kp )
for p ∈ {1, 2, ∞}.
We begin by the following definition:
Definition 7.1. [Banach spaces] Let (X, k · k) be a normed space and dk·k be the
distance induced by k · k. The space (X, k · k) is said to be a Banach space if the metric
space (X, dk·k ) is complete, i.e. Cauchy sequences in (X, k · k) are convergent.
Now we recall the following theorem:
Theorem 7.2. In any normed vector space, we have
(1) Every convergent sequence is a Cauchy sequence. The opposite is not necessary
true.
(2) Every Cauchy sequence is bounded.
(3) Every Cauchy sequence that has an adherent value is convergent 1.
Remark 7.3. In Rn any bounded sequence has an adherent value (this is asserted
by the Bolzano-Weirstrass theorem). Consequently, the above theorem 7.2 ensures that
every Cauchy sequence in Rn is convergent i.e. Rn is a Banach space. On other hand,
due to the non-compactness of the closed ball in every infinite dimensional vector space
X (see Theorem 6.5), it can not be guaranteed that any bounded sequence in X has an
adherent value. Hence the question of convergence of Cauchy sequences in (the infinite
dimensional space) X is far from trivial.
1Apoint a is an adherent (or cluster) value of a sequence (xn ) if there exists a subsequence (xnk )
such that xnk → a.
7. BANACH SPACES 17
In the following two theorems we will find that the space (C[0, 1], k · kp ) is a Banach
space for p = ∞ unlike the cases of p ∈ {1, 2} (see Example 2.2 for the definition of k · kp ,
for p = 1, 2). The main reason for this is that the space C([0, 1]) is an infinite dimensional
vector space.
Theorem 7.4. The space (C([0, 1]), k · k∞ ) is a Banach space, i.e. Cauchy sequences
are convergent.
Proof. Let (fn ) be a Cauchy sequence in C([0, 1]), k · k∞ ). To prove Theorem 7.4,
we must find f ∈ C([0, 1]) such that fn → f in C([0, 1]), k · k∞ .
Step 1. Definition of the function f : [0, 1] → R :
Let ε > 0. Since (fn ) is a Cauchy sequence in (C([0, 1]), k · k∞ ), then there exists
n0 ∈ N such that,
ε
∀ n ≥ n0 , ∀ p ∈ N , kfn+p − fn k∞ < . (7.1)
6
By definition of k · k∞ , we infer from (7.1),
ε
∀ x ∈ [0, 1] , ∀ n ≥ n0 , ∀ p ∈ N , |fn+p (x) − fn (x)| ≤ kfn+p − fn k∞ < < ε . (7.2)
6
Consequently, for all x ∈ [0, 1], the sequence (fn (x)) is a Cauchy sequence in R, hence
convergent. Let us denote by f (x) the limit of the sequence fn (x) in R. That way, we get
a function f : [0, 1] → R.
Step 2. The function f is in the space C([0, 1]) :
Let x0 ∈ [0, 1]. We will prove that the function f defined is Step 1 is continuous at x0 .
Let ε > 0 and n0 ∈ N be as in Step 1. Since the function fn0 is continuous at x0 , then,
there exists δ > 0 such that,
ε
x ∈ [0, 1] and |x − x0 | < δ =⇒ |fn0 (x) − fn0 (x0 )| < .
6
In the sequel, we suppose that x ∈ [0, 1] and |x − x0 | < δ. By (7.2), we have, for all p ∈ N,
|fn0 +p (x) − fn0 +p (x0 )| ≤ |fn0 +p (x) − fn0 (x)| + |fn0 (x) − fn0 (x0 )| + |fn0 (x0 ) − fn0 +p (x0 )|
ε ε ε ε
< + + = .
6 6 6 2
Since the inequality mentioned previously is true for all p ∈ N, then
ε
lim |fn0 +p (x) − fn0 +p (x0 )| ≤ < ε .
p→∞ 2
By Step 1, we deduce that,
|f (x) − f (x0 )| < ε .
Step 3. fn → f in C([0, 1]), k · k∞ :
By (7.2) ,
ε
∀ n ≥ n0 , ∀ x ∈ [0, 1] , lim |fn+p (x) − fn (x)| ≤ .
p→∞ 6
By Step 1, we get,
ε
∀ x ∈ [0, 1] , ∀ n ≥ n0 , |f (x) − fn (x)| ≤ .
6
By definition of the distance d, we deduce that,
ε
∀ n ≥ n0 , kf − fn k∞ ≤ < ε .
6
This proves that lim kf − fn k∞ = 0.
n→∞
18 1. NORMED VECTOR SPACES
The reader is asked to verify that the sequence (fn ) is Cauchy in (C([0, 1]), k · kp ). We
will prove that (fn ) is not convergent in (C([0, 1]), k · kp ). If (fn ) where convergent, then
there exists f ∈ C([0, 1]) such that kfn − f kp → 0. However,
Z 1/2 Z 1
p
∀ n ≥ 3, 0 ≤ |1 − f (x)| dx ≤ |fn (x) − f (x)|p dx .
0 0
Z 1
Sending n to infinity, we get that |1 − f (x)|p dx = 0. Consequently, the continuity of
0
f yields that f (x) = 1 for all x ∈ [0, 12 ].
Now, let a ∈] 21 , 1]. We will prove that f (a) = 0. Select n0 ≥ 3 such that 21 + n10 < a ≤ 1.
Consequently,
Z 1 Z 1
p
∀ n ≥ n0 , 0 ≤ |f (x)| dx ≤ |fn (x) − f (x)|p dx .
a 0
By sending n to infinity, we get that f (a) = 0. This is true for all a ∈] 12 , 1]. By continuity
of f , we get f ( 12 ) = lima→ 1 + f (a) = 0. On the other hand, we proved that f ( 12 ) = 0. This
2
yields the desired contradiction.
The same proof can be applied on every subsequence of (fn ). That is, if (fnk ) is a
convergent subsequence of (fn ), then it will converge to a function f ∈ C([0, 1]) that
+
satisfies f ( 12 ) = 1 and f ( 12 ) = 0.
Remark 7.6. (The space L2 ([0, 1])
According to Theorem 7.5, the space (C([0, 1], k · k2 ) is not a Banach space. However,
this issue can be resolved as follows. One can construct a larger space of functions,
denoted by L2 ([0, 1]) and called the space of square integrable functions, such that, for
every function f ∈ L2 ([0, 1]), the integral
Z 1
|f (x)|p dx exists .
0
This space is needed in Chapter 4 later. Obviously, the space L2 ([0, 1]) contains C([0, 1]).
All we need to know about L2 ([0, 1]) is the following property. If f ∈ L2 ([0, 1]), then there
exists a sequence of continuous functions (fn )n≥1 ⊂ C([0, 1]) such that
• (fn ) Z
is a Cauchy sequence in (C([0, 1]), k · k2 ) ;
1
• lim |fn (x) − f (x)|2 dx → 0 .
n→∞ 0
7. BANACH SPACES 19
A deep theorem asserts that (L2 ([0, 1]), k·k2 ) is a Banach space. However, the construction
of the space L2 is a subtle problem that is beyond the scope of this course. One way to
define L2 is through a completion process by collecting all Cauchy sequences in C([0, 1])
in equivalent classes.
20 1. NORMED VECTOR SPACES
Exercises
Exercise 1. In each of the following cases, check if the function N defines a norm in the
space X.
(1) X = R and N (x) = sin(x) ;
(2) X = R and N (x) = x2 + |x| ;
(3) X = R2 and N (x, y) = 2|x| + 3|y| ;
(4) X = R2 and N (x, y) = sup |x + ty| .
t∈[0,1]
Let N1 and N2 be two norms in a vector space X, and 0 < c1 < c2 . Show that the
following three statements are equivalent:
(a) ∀ v ∈ X, c1 N1 (v) ≤ N2 (v) ≤ c2 N1 (v)
.
1 1
(b) BN1 0, c2 ⊂ BN2 (0, 1) ⊂ BN1 0, c1 .
(c) BN1 0, c12 ⊂ BN2 [0, 1] ⊂ BN1 0, c11 .
Exercise 4. Suppose that (X, k · k) is a normed space, p and q the two functions defined
by,
kxk
p(x) = , q(x) = ln(1 + kxk) , (x ∈ X) .
1 + kxk
(1) Are p and q norms in X ?
(2) Are the functions dp (x, y) = p(x − y) and dq (x, y) = q(x − y) metrics in X ?
Z 1 1/2
2
Exercise 5. Let X = C([0, 1]) be normed by kf k2 = |f (x)| dx . Consider the
√ 0
two subsets Y = {ax : a ∈ R} and B = {ax + x : 0 ≤ a ≤ 1} .
(1) Show that Y is not bounded.
(2) Let f0 (x) = x2 . Find dist(f0 , Y ) and a function g ∈ Y such that kf0 − gk2 =
dist(f0 , Y ) .
EXERCISES 21
Exercise 7. Suppose that (X, k · k) is a normed space and V a subspace of X such that
V 6= X. Show that V ◦ = ∅. (Hint: if x ∈ V ◦ , y ∈ V and y 6= 0, then w = x + 2kyk
r
y∈
B(x, r)).
kf k∞ = sup |f (x)|.
x∈[0,1]
Exercise 10. Let X = C[−1, 1] endowed with the norm of uniform convergence. Study
22 1. NORMED VECTOR SPACES
(1) Show that F is closed in (X, k · k∞ ). What can we say about (X, k · k∞ )?
(2) Let (fn )n≥1 the sequence of real functions defined on [0, 1] by:
nt + 1 if t ∈ [0, n1 ],
fn (t) =
2 if t ∈] n1 , 1].
Exercise 13. Let E = C([0, 1], R) the vector space of continuous functions over [0, 1].
Consider the sequence of functions (fn ) defined by:
1 − nx if 0 ≤ x ≤ n1
fn (x) = .
0 if n1 ≤ x ≤ 1
Exercise 14. Let X = C([0, 1]) be the space of continuous functions from [0, 1] to R.
kf k∞ = sup |f (x)| ,
x∈[0,1]
EXERCISES 23
and
Z 1 1/2
2
kf k2 = |f (x)| dx .
0
(1) Show that k · k1 and k · k∞ are norms in X.
(2) Show that k · k2 is a norm in X. (Hint: Prove the following Cauchy-Schwarz
R1
inequality, 0 f (x)g(x) dx ≤ kf k2 kgk2 ).
(3) Give the distances assigned to each of the norms k · k1 , k · k∞ , k · k2 .
(4) Let n ∈ N∗ and define the function fn : [0, 1] → R as follows,
1 − nx if 0 ≤ x ≤ n1 ,
fn (x) =
0 if x ≥ n1 .
(a) Show that fn ∈ X ;
(b) Compute kfn k1 and kfn k∞ .
(c) Deduce that k · k1 and k · k∞ are not equivalent.
Exercise 15. Let X = C 1 [0, 1] be the set of real valued continuously differentiable
functions defined on the closed bounded interval [0, 1]. Consider the following mappings
on X Z 1
kf k1 = |f (t)|dt, kf k∞ = sup |f (t)|
0 t∈[0,1]
and
0
kf kX = kf k∞ + kf k1
(1) Show that k · kX is a norm on X.
(2) Show that for every f ∈ X we have
kf k∞ ≤ kf kX
(3) Show that k · k∞ is not equivalent to k · kX . (Hint: you may consider fn (x) =
sin(2πnx), n ∈ N)
be defined by
N (f ) = sup |f 00 (x)| + |f (x)| .
[0,1]
Deduce that
0
kf k∞ ≤ (e − 1)kf kX and kf k∞ ≤ ekf kX .
(3) Deduce that the two norms k · k1 and k · kX are equivalent over X.
Z 2
Exercise 18. Let X = C[0, 2] endowed with the norm kf k1 = |f (x)|dx. Consider
0
the sequence
0 if 0 ≤ t ≤ 1 − n1 ,
fn (t) = nt − n + 1 if 1 − n1 < t < 1,
1 if 1 ≤ t ≤ 2.
(1) Show that (fn )n is a Cauchy sequence in (X, k · k1 ).
(2) Suppose that (fn )n converges to some function f in (X, k · k1 ).
(a) Show that f (x) = 1 for every x ∈ [1, 2].
Z 1− 1
n
(b) Show that |f (x)|dx −→ 0 as n −→ +∞. Deduce that f (x) = 0 for all
0
x ∈ [0, 1[.
(c) Deduce that (fn )n does not converge in (X, k · k1 ).
(3) What can you say about the space (X, k · k1 )?
Exercise 19. Let X = C([0, 1]). Show that (X, k · k∞ ) is a Banach space.
( ∞
)
X
`1 (R) = x = (xk ) : xk ∈ R, |xk | < +∞ ,
k=1
EXERCISES 25
∞
X
equipped with the norm kxk1 = |xk |. Let B = {x = (xk ) ∈ `1 (R), kxk1 ≤ 1} and we
k=1
define the sequence (xn ) ⊂ `1 by
1 if i = n
xni =
0 6 n.
if i =
(1) Show that B is closed and bounded in (`1 (R), k · k1 ).
(2) Show that for every n, we have xn ∈ B.
(3) Show that kxn − xm k1 = 2 for n 6= m. Deduce that (xn ) is not a Cauchy sequence
in B.
(4) Deduce that B is not compact in (`1 (R), k · k1 ).
(5) What can you say about the dimension of `1 (R)?
Exercise 22.
This chapter is devoted to the study of continuous linear functions between normed
spaces and the spaces of such functions. Recall that a normed space carries an algebraic
and an analytic structure. These two structures are compatible in the following sense.
The analytic structure is defined by the distance function, the distance function is induced
by a norm, and the norm respects the vector space operations.
In the same way, when one considers functions between two normed spaces, a special
attention is given to those functions that are compatible with the algebraic structure,
i.e. linear functions. For these functions, it is natural to study their analytic properties,
i.e. continuity. This is a rich material in many aspects: There are examples of linear
functions that fail to be continuous; the study of spaces of continuous linear functions
is interesting in its own; spaces of continuous linear functions lead us to dual spaces; by
inspecting the dual of a given space, one may extract an additional information about the
space. In summary, the theory of bounded linear transformations plays a central role in
the development of functional analysis and its applications.
(b) ∀ G ∈ TY , F −1 (G) ∈ TX .
Using the above property, we can verify the result summarized in the following exercise:
Exercise 2.1. Let (X, k · kX ) and (Y, k · kY ) be two n.v.s. and let F be a linear
surjective application from X into Y . Show that the following statements are equivalent:
(1) F is a homeomorphism from X into Y ;
(2) There exist α, β > 0 such that ∀ x ∈ X, αkxkX ≤ kF (x)kY ≤ βkxkX .
Consequently, in the case when N1 and N2 are two distinct norms in X, we get
TN1 = TN2 ⇐⇒ I : (X, N1 ) → (X, N2 ) is a homeomorphism ,
where I : u ∈ X 7→ I(u) = u ∈ X is the identity function, TN1 and TN2 denote the classes
of open subsets of (X, N1 ) and (X, N2 ) respectively.
This theorem have important applications to the theory of differential and integral
equations, namely on existence of solutions. Some exercises will be devoted to applications
of this sort.
2. Linear continuous functions
In this section, we specialize in studying continuity for linear functions between two
normed spaces. We will see that the algebraic property of a function being linear is not
enough to guarantee continuity.
Let us start by recalling the notion of a linear function:
Definition 2.1. Let X and Y be two vector spaces on the some field K = (R or C).
A function L : X → Y is said to be linear if,
∀ x, y ∈ X , ∀ α, β ∈ K, L(αx + βy) = αL(x) + βL(y) .
Remark 2.2.
(1) If L is a linear function, then L(0) = 0.
(2) If L1 : X → Y and L2 : X → Y are two linear functions, then the function
L1 + L2 : X → Y
x 7→ L1 (x) + L2 (x)
is a linear function.
(3) If L : X → Y is a linear function and α ∈ C, then the function
αL : X → Y
f 7→ α L(f )
is a linear function.
(4) The space of all linear functions from X to Y is a vector space.
(5) The identity function
I: X → X
f 7→ I(x) = x
is a linear function.
2. LINEAR CONTINUOUS FUNCTIONS 29
Linear functions enjoy the property that continuity at one point guarantees continuity
in the space. That is a consequence of the algebraic property of linearity.
Theorem 2.3. Let (X, k · kX ) and (Y, k · kY ) be normed spaces and L : X → Y be a
linear function. The following four statements are equivalent:
(a) The function L is continuous at 0 (from (X, k · kX ) to (Y, k · kY )) ;
(b) The function L is continuous everywhere (from (X, k · kX ) to (Y, k · kY )) .
(c) The function L is uniformly continuous (from (X, k · kX ) to (Y, k · kY )) .
(d) ∃ M > 0, ∀ x ∈ X, kL(x)kY ≤ M kxkX .
Proof. First we prove (a)⇐⇒(b)⇐⇒(c). The implication (c) =⇒ (b) =⇒ (a) is
obviously true. So we will prove that (a) =⇒ (b) and (b) =⇒ (c).
Suppose that (a) is true. Let a ∈ X and ε > 0. Since L is continuous at 0 and
L(0) = 0, there exists δ > 0 such that
∀u ∈ X, kukX < δ =⇒ kL(u)kY < ε .
In particular, for u = x − a and x ∈ X, we have
kx − akX < δ =⇒ kL(x − a)kY < ε =⇒ kL(x) − L(a)kY < ε .
Here we used the linearity of L to write L(x − a) = L(x) − L(a). This proves that (b)
is true and consequently (a)⇐⇒(b). The implication (b) =⇒ (c) follows similarly as
(a) =⇒ (b).
Now it remains to show that (c)⇐⇒(d). Indeed, suppose that (c) is true. We will
prove (d) by contradiction. Suppose that (d) is false. Consequently, the negation of (d)
is true, i.e.
∀ M > 0 , ∃ a(M ) ∈ X , kL(a(M ))kY > M ka(M )kX . (2.1)
For all n ∈ N, define xn = a(n). Clearly, xn 6= 0 since, if xn = 0, then by (2.1),
1
0 = kL(xn )kY > M kxn kX = 0, i.e. 0 > 0. Consider the sequence, un = xn .
nkxn kX
1
It is clear that, for all n ∈ N, kun kX = n → 0, i.e. un → 0 in (X, k · kX ). By the
continuity of L, L(un ) → L(0) = 0 in (Y, k · kY ). By (2.1), for all n ∈ N, kL(un )kY > 1.
By taking n → ∞ on both sides of this inequality, we get 0 = lim kL(un )kY ≥ 1. This is
n→∞
a contradiction.
Now, suppose that (d) is true. By the linearity of L, for all x ∈ X and y ∈ X, (d)
asserts that kL(x) − L(y)kY = kL(x − y)kY ≤ M kx − ykX , i.e. L is M -Lipschitz. This
proves that L is uniformly continuous. Thus (a) is true.
Remark 2.4. The above theorem gives a characterization of linear continuous func-
tions. In addition to the algebraic property of linearity, continuity needs an extra condi-
tion, namely, L : X → Y is continuous if kL(·)kY and k · kX are comparable i.e.
L : (X, k · kX ) 7−→ (Y, k · kY ) is continuous ⇐⇒ ∃ M > 0, ∀ x ∈ X, kL(x)kY ≤ M kxkX
Theorem 2.3 validates the following definition:
Definition 2.5. Let (X, k · kX ) and (Y, k · kY ) be normed spaces.
(1) L : X → Y be a linear function.
30 2. BOUNDED LINEAR TRANSFORMATIONS
(a) If L is continuous (from (X, k · kX ) to (Y, k · kY )), then we define the norm
of L to be
kL(x)kY
kLk = sup .
x∈X kxkX
x6=0
Remark 2.6.
(1) Theorem 2.3 yields that L(X, Y ) is a vector space.
(2) The norm of a linear continuous function L from (X, k · kX ) to (Y, k · kY ) is also
denoted by
kLkL(X,Y ) or kLk .
L (X,k·kX ),(Y,k·kY )
This notation is helpful to remind the reader of the spaces and the norms.
(3) The notion of a bounded linear transformation/operator is not to be con-
fused with bounded functions. In the former, bounded-ness signifies that the norm
of the linear function is finite (hence meaning continuity).
(4) For bounded linear transformations, Theorems 2.3 states the following
continuity at a point ⇔ continuity everywhere ⇔ uniform continuity ⇔ Lipschitz function .
(5) One can extract useful information about the norm of a continuous linear function
L : X → Y as follows (these simply come from the definition of the sup and kLk).
Let L : X → Y be a linear function. It holds:
(a) If
∃ M > 0 , ∀ x ∈ X, kL(x)kY ≤ M kxkX ,
then we deduce that L is continuous and kLk ≤ M . (This remark is helpful
to extract an upper bound on kLk).
(b) For all x ∈ X \ {0}, we have
kL(x)kY
kLk ≥ .
kxkX
If x is selected properly, then this remark is helpful to extract a lower bound
on kLk.
(c) Let ε > 0 (no matter how small it is). There exists xε ∈ X such that,
kL(xε )kY
xε 6= 0 and ≥ kLk − ε .
kxε k
(d) The inequality kL(x)kY ≤ kLk kxkX is true for all x ∈ X.
(6) The norm of a linear continuous function can be expressed in the equivalent forms
kLk = sup kL(x)kY = sup kL(x)kY .
x∈X x∈X
kxkX ≤1 kxkX =1
3. THE SPACE OF BOUNDED LINEAR TRANSFORMATIONS 31
Let (Tn ) be a Cauchy sequence in L(X, Y ). To prove Theorem 3.2, we must find a
continuous linear function T : X → Y such that Tn → T in L(X, Y ). We will do this in
three steps.
Step 1. Definition of the function T :
Let ε > 0. Since (Tn ) is a Cauchy sequence in L(X, Y ), then there exists n0 ∈ N such
that,
ε
∀ n ≥ n0 , ∀ p ∈ N , kTn+p − Tn k < . (3.3)
2
Let x ∈ X. As a consequence of (3.3) and the definition of the norm in L(X, Y ), we have,
ε
∀ n ≥ n0 , ∀ p ∈ N , kTn+p (x) − Tn (x)kY < kf kX . (3.4)
2
By Remark ??, the sequence (T (xn )) is a Cauchy sequence in (Y, k · kY ). Since (Y, k · kY )
is a Banach space, we deduce that (T (xn )) is convergent in (Y, k · kY ). Let us denote by
T (x) the limit of T (xn ) in (Y, k · kY ). That way, to every x ∈ X, we assigned T (x) ∈ Y ;
this defines a function T : X → Y .
Step 2. The function T is in the space L(X, Y ):
Here we prove that T ∈ L(X, Y ). Let x ∈ X, y ∈ X and k ∈ R or C. We will prove
that T (x + y) = T (x) + T (y) and T (kx) = kT (x). Since Tn is a linear function, then
Tn (x + y) = Tn (x) + Tn (y) and Tn (kx) = kTn (y) . (3.5)
Now, by Step 1,
Tn (x + y) → T (x + y) , Tn (x) → T (x) , Tn (y) → T (y) in (Y, k · k) .
Furthermore, we get by Theorem 2.10
Tn (x) + Tn (y) → T (x) + T (y) and kTn (x) → kT (x) in (Y, k · k) .
We take n → ∞ in (3.5) and use that the limit is unique to write
T (x + y) = T (x) + T (y) and T (kx) = kT (y) .
Next we prove that T is continuous. By (3.4) used with p ∈ N and n = n0 , we have,
ε
∀ x ∈ X , ∀ p ∈ N , kTn0 +p (x) − Tn0 (x)kY < kxkX .
2
By sending p → ∞ and Step 1, we get,
ε
∀ x ∈ X , kT (x) − Tn0 (x)kY ≤ kxkX .
2
3. THE SPACE OF BOUNDED LINEAR TRANSFORMATIONS 33
By the triangle inequality and the fact that Tn0 ∈ L(X, Y ), we deduce further,
ε
∀ x ∈ X , kT (x)kY ≤ kTn0 (x)kY + kT (x) − Tn0 (x)kY ≤ (kTn0 + )kxkX .
2
By Theorem 2.3, we deduce that T is continuous.
Step 3. The convergence Tn → T holds in L(X, Y ):
By (3.4) and Step 2, we have,
ε
∀ x ∈ X , ∀ n ≥ n0 , kT (x) − Tn (x)k = lim kTn+p (x) − Tn (x)kY ≤ kxkX .
p→∞ 2
By definition of the norm in L(X, Y ) (cf. Remark 2.6, Item (5)-(a)), we get,
ε
∀ n ≥ n0 , kT − Tn k ≤ < ε .
2
This proves that lim kT − Tn k = 0.
n→∞
34 2. BOUNDED LINEAR TRANSFORMATIONS
Exercises
(1) Show that for every k > 0, Hk is closed in (E, k · k∞ ). What can we say about
Hk ?
(2) We define the sequence of functions (fn )n≥1 by
r
1
fn (x) = x +
n
(a) Show that the sequence (fn ) converges uniformly on [0, 1] to a limit f to be
determined. √
n
(b) Show that for all n ∈ N, fn is -Lipschitz.
2
(c) Show that f 6∈ H.
(3) Is H closed in (E, k · k∞ )? Is (H, k · k∞ ) a Banach space? Justify your answers.
Exercise 3. Let (X, N1 ) and (Y, N2 ) be normed spaces and f a mapping from X into Y .
Exercise 7. Let (X, k.k) be a normed space. Show that the mapping f : X 7−→ B(0, 1)
x
defined by f (x) = 1+kxk is a homeomorphism.
1
Exercise 8. Let E = Q ∩ [0, 1] and f : E 7−→ E the mapping defined by f (x) = .
x+2
Show that f is a contraction and that f does not have a fixed point in X. What can you
say about X?
Exercise 9. Consider the space E = C[0, 1] equipped with the norm of uniform conver-
gence k · k∞ .
We define the application T : (E, k · k∞ ) → (E, k · k∞ ) by
Z 1
1 xt
T (f )(x) = 1+ xe f (t)dt .
2 0
Exercise 10.
(1) Let α ≥ 0 and consider the space C[0, 1] equipped with the two norms:
kf k∞ = sup |f (t)| and kf kα = ke−αt f (t)k∞ , α ≥ 0.
t∈[0,1]
(a) Show that the norms k · k∞ and k · kα are equivalent over C[0, 1].
(b) Deduce that (C[0, 1], k · kα ) is a Banach space.
(2) Let f : [0, 1] → R be a α-Lipschitz function; that is, for all y, z ∈ R:
|f (y) − f (z)| ≤ α|y − z|.
The aim of this problem is to show that the following differential equation
( 0
y (t) = f (y(t)),
(∗)
y(0) = y0 ∈ R,
has a unique solution y ∈ C 1 [0, 1] = {f ∈ C[0, 1], f 0 ∈ C[0, 1]}.
(a) Show that if y ∈ C[0, 1] is a solution of (∗) then y ∈ C 1 [0, 1].
(b) Show that y ∈ C 1 [0, 1] is a solution of (∗) Zif and only if y ∈ C[0, 1] is a
t
solution of the integral equation y(t) = y0 + f (y(s))ds.
0
(c) Let T : (C[0, 1], k · kα ) 7−→ (C[0, 1], k · kα ) be the mapping defined by
Z t
T (y)(t) = y0 + f (y(s))ds, (t ∈ [0, 1]).
0
Show that T is a contraction.
(d) Deduce that there exists a unique solution y ∈ C 1 [0, 1] of (∗).
Exercise 11. Let E = {f ∈ C(R) : (1 + x2 )|f (x)| is bounded}. Consider the mappings
N : E 7−→ R+ and L : E 7−→ R defined by:
Z
2
N (f ) = sup(1 + x )|f (x)|, L(f ) = f (x)dx
x∈R R
Exercise 12. Let X = C[0, 2]. Using linear form show that the following spaces are
closed in (X, k · k∞ ):
Z 2
E = {f ∈ X : f (0) = 0}, and F = f ∈X: f (t) dt = 0 .
0
Exercise 13. Let E = C 1 [0, 1] and F = C[0, 1] endowed with the norm
Z 1
kf k1 = |f (x)|dx.
0
Consider the mappings I : E 7−→ E and D : E 7−→ F defined by
Z t
I(f )(t) = f (x)dx, (t ∈ [0, 1]); D(f ) = f 0 .
0
(1) Show that I is linear and continuous for the norm k · k1 .
(2) Show that D is linear but not continuous for the norm k · k1 .
(Hint: consider the sequence of functions fn (t) = (n + 1)tn , (t ∈ [0, 1]) .
(3) We define on E the mapping N by
N (f ) = kf k1 + kf 0 k1 .
Show that N is a norm on E and if E is endowed with this norm then D is
continuous.
(4) Let (gn ) the sequence of functions defined by:
∀t ∈ [0, 1], gn (t) = (n + 1)(1 − t)n
(a) Calculate kgn k1 et kI(gn )k1 .
(b) Deduce that kIk = 1.
(5) Let (hn ) be the sequence of functions defined by:
n n−1
∀t ∈ [0, 1], hn (t) = t
n+1
(a) Calculate N (hn ) et kD(hn )k1 .
(b) Deduce that kDk = 1.
Exercise 14. For f ∈ C[0, 1], we define the norms kf k∞ = sup |f (t)| and kf k1 =
t∈[0,1]
Z 1
|f (t)|dt. Are the following linear mappings continuous? Eventually, compute their
0
norms.
(1) φ : (C([0, 1]), k · k∞ ) −→ R, φ(f ) = f (0);
38 2. BOUNDED LINEAR TRANSFORMATIONS
equipped with the norm kxk∞ = sup |xk |. Consider the mapping f : `∞ 7−→ R defined by
k∈N
f (x) = x0 , ∀ x = (xk ) ∈ `∞ .
Show that f is a linear form and continuous. Calculate kf k.
Exercise 16. Consider E = C[−1, 1] endowed with the norm kf k∞ = sup |f (t)|. Let
t∈[−1,1]
F the space of 2π-periodic continuous functions from R into R, endowed with either the
Z π 1/2
norm N2 such that N2 (f ) = |f (t)|2 dt , or the norm N∞ : N∞ (f ) = sup |f (t)|.
−π t∈R
Let L : E 7−→ F be the mapping defined by
(L(f ))(t) = f (cos(t)).
(1) Show that L is well defined, linear and injective.
(2) Show that L is continuous for each of the norms N2 and N∞ of F , and calculate
their corresponding norms kLk2 and kLk∞ .
Exercise 17. Let E and F be two n.v.s. and let T ∈ L(E, F ) such that kLk ∈ [0, 1[.
Show that T is a contraction.
Exercise 18. Let X = C([0, 1], R) endowed with the norm kf k∞ = sup |f (x)| and let
x∈[0,1]
I : X 7−→ X be the mapping defined, for any f ∈ X, by
Z x
I(f )(x) = sin(x − t)f (t)dt, ∀ x ∈ [0, 1].
0
(1) Show that is well defined.
(2) Show that L is linear and continuous. Calculate kIk.
(3) Deduce that I has a unique fixed point.
(4) Determine the fixed point of I.
Exercise 19.
A. Let E be a Banach space and f : E 7−→ E a continuous mapping. Show that if
f 2 = f ◦ f is a contraction then f has a unique fixed point.
EXERCISES 39
Exercise 20. Let (E, k · k) be a normed space and C a convex compact of E. Consider
f : C 7−→ C an application verifying
kf (x) − f (y)k ≤ kx − yk, ∀ x, y ∈ C.
∗
Let x0 ∈ C. For every n ∈ N , consider the application fn : C 7−→ C defined by
f (x0 ) 1
fn (x) = + 1− f (x).
n n
(1) Show that fn is a contraction.
(2) Show that there exists a unique xn ∈ C such thatfn (xn ) = xn .
(3) Show that kfn (xn ) − f (x)k ≤ n1 kx − x0 k + 1 − n1 kxn − xk.
(4) Deduce that f has a fixed point.
CHAPTER 3
HILBERT SPACES
The space H endowed with an inner product h·, ·i is called an inner product space
(or a pre-Hilbert space). This is usually denoted by the pair (H, h·, ·i). If furthermore
K = R, then H is called a real inner product space.
Remark 1.2. Assume that (H, h·, ·i) is an inner product space. Then the following
consequences are satisfied:
(1) For all λ, µ ∈ K and h, g, h ∈ H, we have
hh, λf + µgi = λhh, f i + µhh, gi .
(2)
hf, f i = 0 ⇔ f = 0 .
(3) Polarization identity: For all f, g ∈ H, we have
1 i
hf, gi = kf + gk2 − kf − gk2 + kf + igk2 − kf − igk2 .
4 4
If H is a real inner product space, the polarization identity becomes
1
hf, gi = kf + gk2 − kf − gk2 .
4
(4) Parallelogram identity: For all f, g ∈ H, we have
1
kf k2 + kgk2 = kf + gk2 + kf − gk2 .
2
Next we point two typical examples of inner product spaces.
Example 1.3.
(1) [Complex Euclidean Spaces] Let d ≥ 1. The Euclidean space Cd is an inner
product space with the usual inner product defined as follows. For all u =
(u1 , · · · , ud ) ∈ Cd and v = (v1 , · · · , vd ) ∈ Cd ,
d
X
hu, vi = u · v = ui vi . (1.1)
i=1
(2) [Euclidean Spaces] Let d ≥ 1. The Euclidean space Rd is a real inner product
space with the usual inner product defined as follows. For all u = (u1 , · · · , ud ) ∈
Rd and v = (v1 , · · · , vd ) ∈ Rd ,
d
X
hu, vi = u · v = ui vi . (1.2)
i=1
(3) [Space of continuous functions] The space C([0, 1]) of continuous real-valued
functions on the interval [0, 1] is a real inner product space with the inner product
defined as follows
Z 1
hf, gi = f (x)g(x) dx f, g ∈ C([0, 1]) . (1.3)
0
Proof of (2):
Clearly, kf k ≥ 0 for all f ∈ H. Moreover, if kf k = 0, then hf, f i = 0 and by (IP1),
f = 0. p
Using (IP3), we get for all λ ∈ C and f ∈ H, kλf k = |λ|2 hf, f i = |λ| kf k.
Finally, we prove the triangle inequality. Let f, g ∈ H. We have
kf + gk2 = hf + g, f + gi
= kf k2 + kgk2 + 2 Re hf, gi
Example 1.7.
(1) The Euclidean space Rd with the usual inner product in (1.2) is a Hilbert space.
The complex Euclidean space Cd is a Hilbert space too.
(2) The space of continuous functions C([0, 1]) with the inner product in (1.3) is
not a Hilbert space. In fact, the norm induced by the inner product is kf k2 =
Z 1 1/2
2
|f (x)| dx , and (C([0, 1]), k · k2 ) is not a Banach space as pointed in
0
Theorem 7.5 (Chapter 1).
44 3. HILBERT SPACES
(3) The space L2 ([0, 1]) introduced in Remark 7.6 (Chapter 1) is a Hilbert space with
the inner product defined by the formula in (1.3). Note that L2 ([0, 1]) is loosely
defined in the following manner:
• L2 ([0, 1]) is a space of functions larger than
R 1 C([0,2 1]) ;
2
• If f ∈ L ([0, 1]), then we may compute 0 |f (x)| dx ;
• (L2 ([0, 1]), k · k2 ) is a Banach space ;
• C([0, 1]) is dense in L2 ([0, 1]) .
Theorem 1.8. [Continuity of the inner product]
Assume that (H, h·, ·i) is an inner product space. Let k · k be the norm in H induced
by the inner product.
Suppose that:
• (fn ) and (gn ) are two sequences in H ;
• fn → f and gn → g in (H, k · k) .
Then the sequence (hfn , gn i) ⊂ C is convergent and
hfn , gn i → hf, gi in (C, usual distance) .
Proof. For all n ≥ 1, we write
hfn , gn i − hf, gi = hfn , gn − gi + hfn − f, gi . (1.4)
We will prove that hfn , gn − gi → 0 and hfn − f, gi → 0 in C. Since fn → f in (H, k · k),
Remark 1.5 yields that
hfn − f, gi = Lg (fn − f ) → Lg (0) = 0 .
For the term hfn , gn − gi, we note that (fn ) is a convergent sequence, hence bounded.
Consequently,
∃ M > 0, ∀ n ≥ 1, kfn k ≤ M .
Next, we use the Cauchy-Schwarz inequality (Theorem 1.4) to write, for all n ≥ 1,
|hfn , gn − gi| ≤ kfn k kgn − gk ≤ M kgn − gk .
Since gn → g in (H, k · k), kgn − gk → 0 and we deduce that hfn , gn − gi → 0 in C.
Returning back to (1.4), we find
lim hfn , gn i − hf, gi = lim hfn , gn − gi + lim hfn − f, gi = 0 .
n→∞ n→∞ n→∞
Remark 1.9.
(1) We record the following special case of Theorem 1.4. If g ∈ H and fn → f in H,
then lim hfn , gi = hf, gi and lim hg, fn i = hg, f i.
n→∞ n→∞
(2) The inner product in H induces a norm k·k, hence a notion of convergence/continuity
in H. This in turn induces a norm k·kprod in the product space H ×H, hence a no-
tion of convergence/continuity in H × H (Theorem 4.1; Chapter 1). Theorem 1.4
states that the inner product
h·, ·i : (H × H, k · kprod ) → (H, k · k)
is a continuous function.
2. ORTHOGONALITY 45
2. Orthogonality
In this section, (H, h·, ·i) denotes an inner product space.
Definition 2.1. [Orthogonal vectors]
Two vectors f ∈ H and g ∈ H are said to orthogonal in (H, h·, ·i) if hf, gi = 0. We
will express this notion by f ⊥g.
Remark 2.2.
(1) For all f ∈ H, f ⊥0 since hf, 0i = 0.
(2) If f ⊥f , then hf, f i = 0 and f = 0.
(3) [Pythagorean identity] If f ⊥g, then kf + gk2 = kf k2 + kgk2 .
Definition 2.3. [Orthogonal complements]
Let M ⊂ H. The orthogonal complement of M in (H, h·, ·i) is the new subset
M ⊥ = {f ∈ H : ∀ g ∈ M , hf, gi = 0} .
Remark 2.4.
(1) Let f ∈ H. By definition 2.3, f ∈ M ⊥ if and only if f is orthogonal to every
vector in M .
(2) The orthogonal complement of {0} is {0}⊥ = H.
(3) H ⊥ = {0}. Actually, if f ∈ H ⊥ , then f ∈ H and hf, f i = 0, hence f = 0.
(4) For every nonempty subset M of H we have 0 ∈ M ⊥ , hence M ⊥ 6= ∅.
(5) M ⊥ is a subspace of H, since M ⊥ 6= ∅ and
∀ f1 , f2 ∈ M ⊥ , ∀ k ∈ R, kf1 ∈ M ⊥ and f1 + f2 ∈ M ⊥ .
(6) If N ⊂ M ⊂ H, then M ⊥ ⊂ N ⊥ .
(7) M ⊥ ∩ M = {0}.
Theorem 2.5. [The orthogonal complement is closed]
Suppose that M ⊂ H. The orthogonal complement M ⊥ is a closed set.
Proof. If M = ∅ then M ⊥ = ∅ is a closed set. Suppose M 6= ∅ and let N = M ⊥ .
We should prove that N ⊂ N . Let f ∈ N . There exists a sequence (fn ) ⊂ N such that
fn → f . Let g ∈ M . By Remark 1.9 (2), we get
lim hfn , gi = hf, gi .
n→∞
Hence,
2 2 1 1 2 1 1
∀ m, n ≥ 1, kgn − gm )k < 4λ + 2 + − 4λ = 2 + .
n m n m
It follows that (gn ) is a Cauchy sequence, and so the completeness of H yields that (gn )
is convergent. Thus there exists g ∈ H such that gn → g in H. Since M is closed and
(gn ) ⊂ M , we get that g ∈ M .
We set PM (f ) = g. As for the uniqueness of PM (f ), if h ∈ M satisfies kf − hk = λ,
we get by the parallelogram identity and (3.1)
1 1
2λ2 ≥ kf − PM (f )k2 + kf − hk2 = k2f − PM (f ) − hk2 + kPM (f ) − hk2
2 2
1 2 1
= 2 f − (PM (f ) + h) + kPM (f ) − hk2
2 2
2 1 2 1
≥ 2λ + kPM (f ) − hk since (PM (f ) + h) ∈ M .
2 2
This proves that kPM (f ) − hk = 0, i.e. PM (f ) = h.
kf − Pm (f )k2 ≤ kf − ut k2
2 2
= f − (1 − t)PM (f ) + tg = (f − PM (f )) + t PM (f ) − g
h i
= kf − PM (f )k2 + t2 kPM (f ) − gk2 + 2t Re hf − PM (f ), PM (f ) − gi .
h i
Thus, for all t ∈ ]0, 1], 2Re hf − PM (f ), g − PM (f )i ≤ tkPM (f ) − gk2 . Sending t to 0+
h i
in this inequality, we get that Re hf − PM (f ), g − PM (f )i ≤ 0.
That PM is linear is a consequence of the result in Lemma 3.8. This proves statement
(1) in Theorem 3.1. Statement (2) follows by Remark 3.4. Statement (3) is the content
of Lemma 3.6.
Finally, for statement (5), we pick u ∈ M such that kuk > 0 (this is possible since
M 6= {0}). Consequently, PM (u) = u (by Statement (2)) and the operator norm satisfies
(see Definition 2.5)
kPM (u)k kuk
kPM kL(H) ≥ = = 1.
kuk kuk
On the other hand, for all f ∈ H, PM (f ) ∈ M and f −PM (f )⊥PM (f ). By the Pythagorean
identity (Remark 2.2), kf k2 = kPM (f )k2 + kf − PM (f )k2 ≥ kPM (f )k2 , i.e. kPM (f )k ≤
kf k. This yields that kPM kL(H) ≤ 1.
4.2. The Riesz representation theorem. The projection theorem has another
important application, namely the representation of linear forms on a Hilbert space that
will allow us to identify a Hilbert space with its dual. This is the Riesz representation
theorem that has a fundamental role in the development of functional analysis and its
applications to other fields, notably the theory of linear partial differential equations.
In this section, H is a Hilbert space, h·, ·i denotes the inner product in H and k · k
denotes the norm induced by the inner product.
Definition 4.3. [Linear forms and topological dual]
(1) A linear form on H is a function L : H → K such that
• L is linear ;
• L is continuous .
(2) The (topological) dual of H, denoted by H ∗ , is the space of all linear forms in
H, i.e. H ∗ = L(H; K).
Remark 4.4.
(1) H ∗ is endowed with the operator norm: kLkH ∗ = sup kL(f )k. Since R is
f ∈H,kf k≤1
complete, it follows from Theorem 3.2 that (H ∗ , k · kH ∗ ) is a Banach space.
(2) [The fundamental example] We can embed H into H ∗ in the following simple
manner. For all g ∈ H, we define the linear function Lg : H → K by Lg (f ) =
hf, gi. It is easy to check that Lg ∈ H ∗ (see Remark 1.5) and kLg kH ∗ = kgk.
Now we can state the Riesz representation theorem, which explains how we can identity
the spaces H and H ∗ .
Theorem 4.5. [The Risez representation theorem]
7 Lg (·) := h·, gi ∈ H ∗ is an invertible linear isometry. More
The function g ∈ H →
precisely:
(1) For all g ∈ H, kLg kH ∗ = kgk ;
(2) For all L ∈ H ∗ , there exists a unique g ∈ H such that L = Lg ;
(3) For all k ∈ R and g ∈ H, Lkg = kLg ;
(4) For all g1 , g2 ∈ H, Lg1 +g2 = Lg1 + Lg2 .
Analogy with the Euclidean case. Before moving to the proof of Theorem 4.5, we
discuss an analogy between its content and the simple case of Euclidean spaces (or more
generally, finite dimensional inner product spaces). Actually, the content of Theorem 4.5
is of algebraic character in the case where H = Rd , d ≥ 1.
Consider Rd with the usual inner product in (1.2). Let L : Rd → R be a linear
function. The function L admits a canonical representation relative to the canonical basis
B = {e1 = (1, 0 · · · , 0), e2 = (0, 1, 0·, 0), · · · , ed = (0, · · · , 0, 1)} ,
52 3. HILBERT SPACES
namely
u1
u2
L(u) = [L]B
· ,
u = (u1 , u2 , · · · , ud ) ∈ Rd , (4.2)
·
ud
where [L]B is the matrix of L relative to the basis B (this is a 1 × d matrix). We have
[L]B = L(e1 ) L(e2 ) · · · L(ed ) .
Using the inner product in (1.2), we express the formula in (4.2) as follows
L(u) = u · g = hu, gi
where g = (L(e1 ), L(e2 ), · · · , L(ed )) ∈ Rd . This is the content of Theorem 4.5 for H = Rd .
Also, we actually proved that every linear function L : Rd → R is continuous (hence a
linear form).
The proof of Theorem 4.5 is of an analytic nature and valid for general Hilbert spaces
(typically infinite dimensional ones). It is one out of many cases where analysis is the
right tool for the passage from finite dimensional settings to infinite dimensional ones.
In proving Theorem 4.5, we will need the following remarkable conclusion:
5. Hilbert basis
5.1. Motivation. Consider the Hilbert space Rd with the usual inner product intro-
duced in (1.2) (this inner product induces the Euclidean norm k · k). The space Rd has a
distinguished basis (called the canonical basis)
B = {e1 = (1, 0, · · · , 0), e2 = (0, 1, 0, · · · , 0), · · · , ed = (0, · · · , 0, 1)} .
Let us list some remarkable properties of the canonical basis:
(1) The basis B is orthonormal in the sense:
• For all n ∈ {1, 2, · · · , d}, ken k = 1 ;
• For all n 6= m, hen , em i = 0 .
(2) Every u ∈ Rd can be expressed in the pleasant form u = dn=1 hu, en ien .
P
(3) (Parseval’s identity) The norm of every u ∈ Rd can be expressed in the form
kuk2 = dn=1 (hu, en i)2 .
P
The canonical basis B is an example of a Hilbert basis. We will discuss the existence of
such a basis in a general Hilbert space.
5.2. Orthonormal subsets. Here we introduce many notions (from linear algebra)
that will lead us to the remarkable Bessel’s inequality in inner product spaces.
In this section, we assume that H is an inner product space whose inner product is
h·, ·i. The norm in H induced by the inner product is denoted by k · k.
Definition 5.1. Let S ⊂ H.
(1) [Orthogonal subsets] The set S is said to be orthogonal if, for all f, g ∈ H, it
holds
f 6= g =⇒ hf, gi = 0 .
54 3. HILBERT SPACES
n
X n
X n
X n
X
kf k2 = krn + hf, fk ifk k2 = krn k2 + k hf, fk ifk k2 ≥ k hf, fk ifk k2 = |hf, fk i|2
k=1 k=1 k=1 k=1
so n
X
2
kf k ≥ |hf, fk i|2 .
k=1
∞
X
That is, the sequence of partial sums of the series |hf, fn i|2 is bounded. Since this is
n=1
a series of positive terms we conclude that it is convergent. Sending n to ∞, in the above
inequality, yields the Bessel’s inequality.
X∞
|hf, fn i|2 ≤ kf k2 .
n=1
6. Separable spaces
Here we introduce the notion of separable Hilbert spaces. In such spaces, we will
see that Paseval’s identity holds (this is where equality occurs in Bessel’s inequality -
Corollary ??). Bessel’s inequality is valid for any choice of an orthonormal subset S = (fn ).
However, for equality to occur, we need that S is in addition a Hilbert basis, a term
that we will define below.
Definition 6.1. [Hilbert basis]
Suppose that (H, h·, ·i) is an inner product space and S ⊂ H. The set S is said to be
a Hilbert basis of H if
• S is orthonormal ;
• span(S) is dense in H.
Note that we deviated from the standard algebraic definition of a basis. Actually, S
is a basis of H if span(S) = H. To be a Hilbert basis, we need only that span(S) = H,
that is equality is relaxed and replaced by density.
56 3. HILBERT SPACES
Of special importance are separable Hilbert spaces since they resemble very much
the Euclidean space Rd . The precise definition is:
Definition 6.2. [Separable spaces]
An inner product space (H, h·, ·i) is said to be separable if it has a countable Hilbert
basis.
Next, we establish how and in what sense do separable Hilbert spaces resemble the
Euclidean spaces. But let us first clarify the following notion: If (un ) is a sequence in an
inner product space H and u ∈ H, by writing
X∞
u= un (in H)
n=1
n
X
we mean that the sequence sn = uk is convergent to u in H.
k=1
Now we are ready to state:
Theorem 6.3. Suppose that (H, h·, ·i) is a separable Hilbert space. Let S = {f1 , f2 , · · · }
be a Hilbert basis of H. Then it holds the following:
(1) (Hilbertian decomposition) For all f ∈ H,
X∞
f= hf, fn ifn in H .
n=1
∞
X
By Bessel’s inequality, the series |hf, fk i|2 is convergent. Thus,
k=n+1
∞
X
∀ > 0, ∃ n0 ≥ 1 , ∀n ≥ n0 , ∀ p ≥ 1 , |hf, fk i|2 < ,
k=n+1
and consequently
n+p ∞
X X
2 2
ksn+p − sn k = |hf, fk i| ≤ |hf, fk i|2 < .
k=n+1 k=n+1
Thus, (sn ) is a Cauchy sequence in the Hilbert space H, hence convergent to some limit
u ∈ H.
Proof of (3):
Suppose that u ∈ span(S) and kf − u k < 3 . Since sn → u, there exists n0 ≥ 1 such
that, for all n ≥ n0 , ksn − uk < 3 .
Since u ∈ span(S), there exist N ≥ n0 and λ1 , λ2 , · · · , λN ∈ R such that
X N
u = λk fk . Using the Pythagorean identities in (5.1) and (5.2), we get
k=1
N
X 2
ksN − u k2 = hf, fk i − λk . (6.1)
k=1
Using Bessel’s inequality,
N
X 2
|hf − u , fk i|2 ≤ kf − u k2 < . (6.2)
k=1
9
But, by the orthogonality of S = (fn )n≥1 ,
N
X
∀ k ≥ 1, hf − u i = hf, fk i − hλm fm , fk i = hf, fk i − λk .
m=1
N
X 2 2
Thus, (6.2) becomes hf, fk i − λ2k < . Inserting this into (6.1), we get
k=1
9
ksN − u k <
.
3
But, since N ≥ n0 , we know that ksN − uk < 3 . Thus, by the triangle inequality,
kf − uk ≤ kf − u k + ku − sN k + ksN − uk < + + = .
3 3 3
With Lemma 6.4 in hand, we can present the
Let > 0 be arbitrary but fixed. Since S = (fn )n≥1 is a Hilbert basis of H, then,
span(S) is dense in H. Consequently, there exists u ∈ span(S) such that kf − u k < 3 .
By Lemma 6.4, kf − uk < . Since > 0 is arbitrary, we get that u = f .
X∞
This proves that hf, fn ifn = f .
n=1
Proof of (2):
Let f, g ∈ H be fixed. We know that
X n n
X
sn = hf, fk ifk → f and tn = hg, fk ifk → g in H .
k=1 k=1
Now by continuity of the norm and the inner product (see Theorem 1.8) we have
n
X ∞
X ∞
X
2
hf, gi = lim hsn , tn i = lim hf, fk ihg, fk i = hf, fk ihfk , gi and kf k = |hf, fk i|2 .
n→∞ n→∞
k=1 k=1 k=1
We will end this chapter by TheoremX 6.5. In its statement, `2 (K) is the space of all
sequences (an ) ⊂ K such that the series |an |2 is convergent. This is a Hilbert space
n≥1
with the following inner product
X
h(an ), (bn )i`2 = an b n .
n≥1
Two Hilbert spaces H1 and H2 are said to be unitary isomorphic if there exists an
invertible function U : H1 → H2 such that
∀ f, g ∈ H1 , hU (f ), U (g)iH2 = hf, giH1 .
Theorem 6.5. Let H be a separable Hilbert space. Then H is unitarily isomorphic
either to Kn , for some n ∈ N, or to the space `2 (K).
Proof. Suppose H is a separable Hilbert space. Assume first that dim(H) = +∞.
Then H has a dense countable (and infinite) orthonormal basis S = {f1 , f2 , · · · }. Consider
the map
U : H → `2 (K)
defined by
U (f ) = αn (f ) = hf, fn i .
n≥1
∞
X
For all f ∈ H, Bessel’s inequality yields that the sequence (αn ) ⊂ ` (K) since 2
|αn |2 =
n=1
∞
X
|hf, fn i|2 < +∞. And also Parseval’s identity gives
n=1
kU (f )k`2 = kf k. (6.3)
6. SEPARABLE SPACES 59
That U is linear follows from the properties of the inner product h·, ·i. That U is injective
follows from the identity (6.3).
= (αn )n≥1 ∈ `2 (K). An easy adaptation
It remains to check that U is surjective. Let αP
of the proof of Lemma 6.4 yields that the series ∞ n=1 αn fn is convergent in H. Hence it
converges to some f ∈ H. Consequently, we find that U (f ) = α.
If the Hilbert space H is finite dimensional and dim(H) = n, then H has an or-
thonormal basis formed by n vectors so H is unitarily isomorphic to KN , see Lemma 5.4
(Chapter 1).
Remark 6.6. We can just think about separable Hilbert spaces as `2 (K) or Kn when
we wish. To be more precise, the calculus in a general separable Hilbert space resembles
that in `2 (K) or Kn .
60 3. HILBERT SPACES
Exercises
Exercise 1.
Tell whether the following maps, are inner product or not on the vector space E :
n
X
(1) hx, yi = xi yi3 , E = Rn ;
i=1
∞
X ∞
X
2
(2) hx, yi = xn y n , E = ` (R) = {x = (xn )n : xn ∈ R and |xn |2 < +∞};
Zn=1 Z n=1
2 2
(3) hf, gi = f (x)g(x)e−x dx, E = {f ∈ C(R, R) : |f (x)|2 e−x dx < ∞};
ZR1 R
0 0 1
(4) hf, gi = f (x)g (x)dx, E = C ([0; 1], R);
0
Z 1
(5) hf, gi = f 0 (x)g 0 (x)dx, E = {C 1 ([0; 1], R) : f(0)=0}
Z0 1
(6) hf, gi = f (x)g(x) + f 0 (x)g 0 (x) dx, E = C 1 ([0; 1], R).
0
Exercise 2.
Let x and y be two nonnull vectors of a real pre-Hilbert space (E, h·, ·i) . Show that
x y kx − yk
2
− 2
= .
kxk kyk kxkkyk
Exercise 3. Let (E, h·, ·i) be a real pre-Hilbert space and u : E 7−→ E be a map. We
designate by k · k the associated norm to h·, ·i.
(1) Suppose that u verify the following conditions:
Exercise 4. Let (E, h·, ·i) be a real pre-Hilbert space and u : E 7−→ E be an application
verifing
hf (x), yi = hx, f (y)i
for all x, y ∈ E. Show that f is linear.
EXERCISES 61
Exercise 5. Let (E, h·, ·i) be a real pre-Hilbert space. Show that , for all x, y ∈ E,
h i h
|hx, yi| = kxk · kyk ⇐⇒ x and y are two linearly dependent vectors
i
(i.e. there exist α ∈ R such that y = αx)
Exercise 6. Let E = C([0, 1], R) equipped with norm k · k∞ . Put f (x) = x and g(x) = 1.
(1) Show that kf − gk2∞ + kf + gk2∞ 6= 2(kf k2∞ + kgk2∞ ).
(2) what can be said about the space (E, k · k∞ )?
Exercise 8. Let (X, h·, ·i) be a real Hilbet space, a ∈ X and f : X 7−→ X be a mapping
such that:
Exercise 9. Let (H, h·, ·i) be a real Hilbert space and f : H 7−→ R be a linear continueous
form (i.e f ∈ H ∗ ). We define the continueous map φ : H 7−→ R by:
1
φ(x) = kxk2 − f (x),
2
where k · k is the associated norm to the scalar product h·, ·i.
62 3. HILBERT SPACES
Exercise 10. Let E = C 1 ([0, 1], R) equipped with the inner product
Z 1
hf, gi = f (x)g(x) + f 0 (x)g 0 (x) dx.
0
Let u(x) = sin x et v(x) = cos x. Prove that u and v are orthogonal with respect to h·, ·i.
Deduce then the value of ku + vk.
Exercise 11. Let (E, h·, ·i) be a real pre-Hilbert space. Let u and v be two vectors such
that, for all λ ∈ R, ku + λvk ≥ kuk. Show that u and v are orthogonal.
Z 1
Exercise 12. Let E = C([0, 1], R) equipped by the inner product hf, gi = f (x)g(x)dx.
0
We define the sub-vector space F by
F = {f ∈ E : f (0) = 0}.
Calculate hf, xf i for f ∈ F ⊥ . Deduce that F ⊥ = {0}.
Z b
Exercise 13. Let E = C([a, b], R) equipped with the inner product hf, gi = f (x)g(x)dx.
a
We designate by F = R[x] the sub-vector space of E formed by polynomials.
(1) Prove that, for f ∈ F ⊥ and for P ∈ R[x], we have
kf k2 ≤ (b − a)kf k∞ kf − P k∞
(2) Deduce, by using the theorem of approximation of Weierstrass, that (F )⊥ = {0}.
∞
X
equipped with the inner product hx, yi = xn yn . For N ∈ N∗ fixed, we define the map
n=1
fN : `2 → R by
N
X
fN (x) = xn ∀ x = (xn )n≥1 .
n=1
(1) Show that (
fN is linear and continueous. )
XN
(2) Let AN = x = (xn )n≥1 ∈ `2 (R) : xn = 0 .
n=1
(a) Show that AN is closed in `2 (R) and that `2 (R) = AN (AN )⊥ .
L
(b) What c an be said about (AN , h·, ·i)?
(c) Let
n o
E = y = (yn )n≥1 ∈ `2 (R) : (yn = y1 , for 1 ≤ n ≤ N ) and (yn = 0 for n > N ) .
Exercise 15. We consider the Hilbert space E = R4 equipped with the inner product
4
X
hx, yi = xi yi . Let E1 = {(x1 , x2 , x3 , x4 ) ∈ E : x1 = x3 , x2 = x4 }.
i=1
(1) Prove that E1 is a closed sub-vector space of E.
(2) Prove that the two vectors a1 = (1, 0, −1, 0) and a2 = (0, 1, 0, −1) are orthogonal
to E1 , and deduce that E1⊥ .
(3) Let A = {x ∈ E1 : kxk = 2} and B = {x ∈ E1⊥ : kxk = 2}.
Z +∞
hf, gi = e−x f (x)g(x)dx, ∀ f, g ∈ E.
0
Exercise 17. Let H = L2 (] − π, π[) be the Hilbert space equipped with the inner product
Z π
1
hf, gi = f (x)g(x)dx.
2π −π
We designate by k · k the induced norm. √ √
We consider the functions f0 (x) = 1, f1 (x) = 2 cos x and f2 (x) = 2 sin x.
(1) Show thatn the family f0 , f1 , f2 is orthonormal in H. o
(2) Let K = f :]−π, π[7−→ R : f (x) = af0 (x)+bf1 (x)+cf2 (x), with a, b, c ∈ R .
Say why K is a closed a sub-vector space of H?
(3) Let f (x) = x2 . Determine hf, f0 i, hf, f1 i and hf, f2 i.
(4) Deduce, from the previous parts, the values of the real a, b and c such that the
integral
Z π
1 2
J(a, b, c) = x2 − af0 (x) − bf1 (x) − cf2 (x) dx
2π −π
is minimal.
Exercise 18. Find the real numbers a and b so that the following integrals are minimal:
Z 1 Z 1
2 2
(1) (x − a − bx) dx (2) (ex − a − bx)2 dx.
0 0
Exercise 19. Let E = C 1 ([0, 1], R) equipped with the inner product
Z 1
hf, gi = f (x)g(x) + f 0 (x)g 0 (x) dx.
0
Put
00
V = {f ∈ E : f (0) = f (1) = 0} and W = {f ∈ E : f is C 2 and f = f }.
(1) Prove that V and W are orthogonal.
f (1) − f (0)ch(1)
(2) For f ∈ E, we define g(x) = f (0)ch(x) + sh(x) and h = f − g.
sh(1)
(a) Verify that g ∈ W and h ∈ V .
(b) Deduce that V and W are complementary. Express the orthogonal projec-
tion on W .
EXERCISES 65
(c) Let Eα,β = {f ∈ E : f (0) = α, f (1) = β}. By using the function g calculate
Z 1
inf (|f (t)|2 + |f 0 (t)|2 )dt.
f ∈Eα,β 0
Exercise 20. Let H be Hilbert space, and F be a closed sub-vector space of H, not
reduce to zero. We denote by p the orthogonal projection of H on F . Show that:
(1) p ◦ p = p.
(2) For all x, y ∈ H, we have hp(x), yi = hx, p(y)i.
(3) kpk = 1.
Exercise 21. Let E = L2 (]0, 1[) equipped with the inner product h·, ·i and let
Z 1
F = {f ∈ E : f (x)dx = 0}.
0
(1) Prove that F is a closed sub-space of E.
(2) Prove that F ⊥ = span{h}, where h ≡ 1.
(3) Let g(x) = x2 . Calculate the orthogonal projection of g on F .
Exercise 22. Let E = C[0, 1] equipped with the norm k · k∞ and let K be a sub-space
of E defined by
n Z 1 o
K = f ∈ E; f (0) = 0 and f (x)dx = 0 .
0
(1) Prove that K is a closed convex of (E, k · k∞ ).
(2) Take f0 (x) = x and the sequence of functions (fn ) defined, for all n ≥ 1, by
(n − 1)2
− x if 0 ≤ x ≤ n1
2n − 1
fn (x) =
n
x− if n1 < x ≤ 1.
2n − 1
(a) Prove that fn ∈ K andZ f0 6∈ K.
1
(b) For f ∈ K, calculate (f (x) − f0 (x))dx . Deduce that
0
1
kf − f0 k∞ ≥ , ∀ f ∈ K.
2
n 1
(c) Prove that kfn − f0 k∞ = . Deduce that inf kf − f0 k∞ = .
2n − 1 f ∈K 2
(d) Suppose there exist a function f1 ∈ K such that kf1 −f0 k∞ = inf kf −f0 k∞ .
f ∈K
In this case, we say that f1 is the orthogonal projection of f0 on K.
(i) Put g(x) = kf1 − f0 k∞ − (f0 (x) − f1 (x)). Verify that
Z 1
g(x)dx = 0 et g ≥ 0.
0
(ii) Deduce that f1 = f0 and that f1 6∈ K. Give your conclusion.
(iii) Is (E, k · k∞ ) a Hilbert space? Justify.
66 3. HILBERT SPACES
Exercise 23.[Projection on a ball] Let (H, h·, ·i) be a real Hilbert space. Let B =
B(0, 1) be the unit closed ball of H. We denote by PB the orthogonal projection of H on
B.
(1) Le but de cette partie est de trouver l’expression de PB .
(a) Let x 6∈ B. Verify that, for all z ∈ B, we have
x x
hz − ,x − i≤0
kxk kxk
(b) Deduce that
x if x ∈ B
PB (x) = x
if x 6∈ B.
kxk
Exercise 25.
( ∞
)
X
`2 (R) = x = (xn )n≥1 : xn ∈ R, |xn |2 < +∞ ,
n=1
∞
X
equipped with the inner product hx, yi = xn yn . We designate by k · k`2 the induced
n=1
norm la norm. Let N ∈ N∗ fixed and defined the map fN : `2 → R by
N
X √
fN (x) = n xn ∀ x = (xn )n≥1 .
n=1
Exercise 28.
(1) Let (H, h·, ·i) be Hilbert space and (ei )i∈I be an orthonormal family. Show that
(ei )i∈I is a hilbert basis of H if and only if
hx, ei i = 0, ∀ i ∈ I =⇒ x = 0
(2) For all i ≥ 1, we on consider the sequence ei = (ein ) ∈ `2 (R) defined by
1 if i = n
i
en =
0 if i 6= n.
Show that (ei )i≥1 is a Hilbert basis of `2 (R).
CHAPTER 4
FOURIER SERIES
In this chapter, we will apply the theory of Hilbert spaces in the specific setting of the
spaces C([0, 2π], R) and L2 ([0, 2π], R), both endowed with the following inner product
Z 2π
hf, gi = f (x)g(x) dx .
0
We remind the reader that C([0, 2π]) is the space of real-valued continuous functions
on the interval [0, 2π]. The space L2 ([0, 2π]) is the space of square integrable functions:
It is a Hilbert space containing C([0, 2π]) as a dense subset (relative to the L2 -norm).
More details regarding this loosely defined space are given in Remark 7.6.
We will prove that the space L2 ([0, 2π]) is separable (see Chapter 3 - Definition 6.2),
i.e. it has a countable Hilbert basis. We will do this by discussing a special Hilbert basis
in L2 ([0, 2π]), namely the trigonometric basis defined by orthogonal sine and cosine
functions. The Hilbertian decomposition of functions relative to the trigonometric basis
(in the sense described in Chapter 3 - Theorem 6.3) will lead us to the notion of Fourier
series. Our approach to the subject will show us immediately that the Fourier series
are convergent relative to the L2 -norm, but also we will elaborate the discussion of their
convergence in other directions (uniform and point-wise convergence).
Both norms have their own advantage. One useful remark is the inequality
√
kf k2 ≤ 2πkf k∞ , (1.3)
which simply says that, if a sequence (fn ) converges (to f ) relative to the norm of uniform
convergence, then it converges (to the same limit f ) relative to the L2 -norm.
As opposed to Euclidean spaces, it is a quite delicate task to answer the following
question in the space of continuous functions: Prove that a subset D of C([0, 2π]) is dense
in C([0, 2π]) (relative to one of the norms k · k∞ or k · k2 ). The Stone-Weirstrass theorem
provides us with a wide class of dense subsets in C([0, 2π]). We state this theorem below.
69
70 4. FOURIER SERIES
hence span(S) is unitary. Finally, if a, b ∈ [0, 2π] and a 6= b, then we can find a function
f ∈ D such that f (a) 6= f (b). Indeed, we choose f as follows:
• If a, b ∈ [0, π] (or a, b ∈ [π, 2π]), we choose f (x) = cos x ;
• If a ∈ [0, π] and b ∈ [π, 2π] (or a ∈ [π, 2π] and b ∈ [0, π]), we choose f (x) = sin x .
Hence, span(S) separates points. Therefore, Theorem 1.1 yields that D = span(S) is
dense in (C([0, 2π]), k · k∞ ).
Consequently, we get the following two density results relative to the L2 -norm:
Corollary 1.3. The set span(S) introduced in (1.5) is dense in (C([0, 2π]), k · k2 ).
1There are many many proofs of the theorem, one simple proof is by B. Brosowski and F.
Deutsch ‘An elementary proof of the Stone-Weirstrass theorem, Proceedings of AMS, Vol. 81,
No. 1, 1981 (http://www.ams.org/journals/proc/1981-081-01/S0002-9939-1981-0589143-8/S0002-9939-
1981-0589143-8.pdf).
2An algebra is a vector space which is closed under multiplication and satisfying the usual algebraic
rules.
2. THE TRIGONOMETRIC SYSTEM 71
Proof. Let > 0 and f ∈ C([0, 2π]). By Theorem 1.2, there exists f ∈ span(S)
such that kf − f k∞ < √2π . Using the inequality in (1.3), we get that kf − f k2 < .
Corollary 1.4. The set span(S) introduced in (1.5) is dense in (L2 ([0, 2π]), k · k2 ).
Proof. Let > 0 and f ∈ L2 ([0, 2π]). Since C([0, 2π]) is dense in L2 ([0, 2π]), there
exists u ∈ L2 ([0, 2π]) such that kf − u k2 < 2 .
By Theorem 1.3, there exists f ∈ span(S) such that ku −f k2 < 2 . Using the triangle
inequality, we get
kf − f k2 ≤ kf − u k2 + ku − f k2 < + = .
2 2
holds).
Proof. Straightforward computations yield that,
Z 2π
|φn (x)|2 dx = 1 (n = 0, 1, 2 · · · )
0
Z 2π
|ψn (x)|2 dx = 1 (n = 1, 2, · · · )
0
Z 2π Z 2π Z 2π
φn (x)φm (x) dx = ψn (x)ψm (x) dx = φn (x)ψm (x) = 0 (if n 6= m)
0 0 0
Z 2π
φn (x)ψn (x) dx = 0 n = 0, 1, 2, · · · ) .
0
Remark 2.3.
(1) To stress their dependence on the function, the Fourier coefficients of the function
f are sometimes denoted by an (f ) and bn (f ).
(2) [Even and odd functions]
If the function f is even, a simple symmetry argument yields the following
formulas for the Fourier coefficients of f :
1 π 2 π
Z Z
a0 = f (x) dx , an = f (x) cos(nx) dx , bn = 0 (n ≥ 1) .
π 0 π 0
Similarly, if f is an odd function, the Fourier coefficients of f are given by
2 π
Z
a0 = an = 0 , bn = f (x) sin(nx) dx (n ≥ 1) .
π 0
(3) Using the functions {φn , ψn } introduced in (1.4), we may express the Fourier
series of a function f as follows
1 1 1
a0 = √ hf, φ0 i , an = √ hf, φn i , bn = √ hf, ψn i ,
2π π π
where n ∈ {1, 2, · · · }. Furthermore, the Fourier series of f can be expressed as
follows (compare with Theorem 6.3 - Chapter 3):
∞
X
hf, φ0 iφ0 + hf, φn iφn + hf, ψn iψn .
n=1
(4) The othonormal set S = {φn , ψn } introduced in (1.4) can be expressed using a
single sequence of functions (fn )n≥0 , where f0 = φ0 and for all n ≥ 1,
(
φ(n+1)/2 if n is odd
fn = ,
ψn/2 if n is even
so that f1 = φ1 , f2 = ψ2 , f3 = φ2 , f4 = ψ2 , etc.
The Fourier series of f becomes
∞
X
hf, fn ifn .
n=0
Using Remark 2.3 and Theorem 2.1, we infer from Chapter 3 - Theorem 6.3:
Theorem 2.4. [Convergence of Fourier series in L2 -norm]
If f ∈ L2 ([0, 2π]), then
(1) [Fourier series decomposition]
The Fourier series of f is convergent in (L2 ([0, 2π]), k · k2 ) and
∞
X
f (x) = a0 + an cos(nx) + bn sin(nx) in L2 ([0, 2π]) .
n=0
(4) The functions f and f 0 are Riemann integrable on [a, b] (hence square integrable,
i.e. in L2 ([a, b])) and
Z b n−1 Z xi+1
X Z b n−1 Z xi+1
X
0
f (x) dx = fi (x) dx , f (x) dx = fi0 (x) dx .
a i=0 xi a i=0 xi
Next, we will write a criterion for the convergence of a series relative to the norm of
uniform convergence.
Lemma 3.3. [Weirstrass M -test] Let (un ) be a sequence in C([0, 2π]). Suppose that
• There exists a sequence (an ) ⊂ R+ such that, ∀ n ≥ 1, kun k∞ ≤ an ;
X∞
• The series an is convergent .
n=1
∞
X
Then, the series un is convergent in (C([0, 2π]), k · k∞ ), i.e. the sequence of partial
n=1
sums
n
X
sn = uk
k=1
is a convergent sequence in (C([0, 2π]), k · k∞ ).
The proof of the lemma is left as an exercise. All we have to do is prove that the
sequence of partial sums (sn ) is a Cauchy sequence. The convergence then follows by the
completeness of the space (C([0, 2π]), k · k∞ ) (see Chapter 1 - Theorem 7.4).
Now, we can prove the main theorem of this section (compare with Theorem 2.4):
Theorem 3.4. [Uniform convergence of Fourier series]
Let f : [0, 2π] → R be a continuous function such that
• f (0) = f (2π) ;
• f is piecewise C 1 on [0, 2π] .
X∞
Then, the Fourier series of f , a0 + an cos(nx) + bn sin(nx) , uniformly converges to
n=1
f on [0, 2π] (i.e. it is a convergent series with limit f in (C([0, 2π]), k · k∞ )).
Proof. The proof is quite simple. We will denote by (an (g), bn (g)) the Fourier coef-
ficients of a Riemann integrable function g. With this notation in hand, we check by a
straightforward calculation, using in particular integration by parts and the assumption
that f (0) = f (2π), that, for all n ≥ 1,
1 1
an (f ) = − bn (f 0 ) and bn (f ) = an (f 0 ) .
n n
Next we use Cauchy’s inequality, for all s, t ∈ R, |st| ≤ 21 (s2 + t2 ), which follows by
expanding the square in (s ± t)2 ≥ 0. That way we obtain,
1 1 0 2 1 1 0 2
|an (f )| ≤ + |bn (f )| and |bn (f )| ≤ + |an (f )| .
2 n2 2 n2
Thus, the general term of the Fourier series of f satisfies
1 1
|an (f 0 )|2 + |bn (f 0 )|2 ,
|an cos(nx) + bn sin(nx)| ≤ 2 +
n 2
4. THE DIRICHELT THEOREM 75
1
P
where n≥1 n2 is a convergent sequence in R, and the same holds for the series
0 2 0 2
P
n≥1 (|an (f )| + |bn (f )| ), in light of Theorem
2.4 (Parseval’s identity). Now, Lemma 3.3
yields that the Fourier series of f is a uniformly convergent series.
That the limit of the Fourier series is equal to f follows from Theorem 2.4 (the limit
is equal to f relative to the L2 -norm) and the inequality in (1.2).
Proof. Define the function h(θ) = sin θ − θ. Clearly, h0 (θ) = cos θ − 1 ≤ 0, hence h
is decreasing. Consequently, for all θ ∈ [0, π/2], h(0) ≤ h(θ) ≤ h(π/2) with h(0) = 0. In
particular, this yields that sin θ ≤ θ.
Consider the function g(θ) = sin θ − 2θ
π
. It is easy to check that max0≤θ≤ π2 g(θ) = g(θ0 )
where θ0 = arccos π , and min0≤θ≤ π2 g(θ) = 0. Thus, for all θ ∈ [0, π2 ], we have g(θ) ≥ 0,
2
i.e. sin θ ≥ π2 θ.
Lemma 4.6. If g is a 2π-periodic function and Riemann integrable on [−π, π], then,
for all a ∈ R,
Z a+2π Z π
g(x) dx = g(x) dx .
a −π
Proof. Note that, for all u ∈ R, g(u) = g (u − 2π) + 2π = g(u − 2π), since g is
2π-periodic. Thus, the change of variable u = x + 2π yields
Z −π Z π Z a+2π Z a+2π
g(x) dx = g(u − 2π) du = − g(u) du = − g(x) dx .
a a+2π π π
Thus, we get
Z a+2π Z −π Z π Z a+2π Z π
g(x) dx = g(x) dx + g(x) dx + g(x) dx = g(x) dx .
a a −π π −π
Lemma 4.7. If g : [a, b] → R is a piecewise C 1 function, then
Z b
1
lim g(x) sin n + x dx = 0 .
n→∞ a 2
Proof. There exist points x0 < x1 < · · · < xm such that x0 = a, xm = b and
for all i ∈ {1, · · · , m}, g can be viewed as a C 1 function on the interval [xi−1 , xi ] (see
Remark 3.2). Moreover,
Z b m Z xi
1 X 1
g(x) sin n + x dx = g(x) sin n + x dx .
a 2 k=1 x i−1
2
Z xi
1
Let i ∈ {1, · · · , m} be fixed. We will prove that lim g(x) sin n + x dx = 0.
n→∞ x
i−1
2
Indeed, performing an integration by parts, we have
Z xi
1 Z xi cos n + 1
2
x g(x) cos n + 1
2
x xi
0
g(x) sin n + x dx = g (x) dx − .
xi−1 2 xi−1 n + 12 n + 12
xi−1
The definition of the limit yields the existence of δ > 0 such that,
f (x + t) − f (x) f (x − t) − f (x)
∀ t ∈ (0, δ) , − f 0 (x+ ) ≤ 1 and − f 0 (x− ) ≤ 1 .
t −t
Consequently, we get the conclusion of Lemma 4.8 with M = 1 + max |f 0 (x+ )|, |f 0 (x− )| .
n
1 X
Dn (t) = + cos(kt) .
2 k=1
Z π Z π n Z π
1 X π
Dn (t) dt = dt + cos(kt) dt = .
0 0 2 k=1 0
2
Proof of (2):
n
1 X 1
For t ∈ {0, ±2π, ±4π, ±6π · · · }, Dn (t) = + 1 = + n.
2 k=1 2
Proof of (3):
We will use the Euler formula: For all θ ∈ R, eiθ = cos θ + i sin θ (consequently,
cos θ = Re(eiθ )).
78 4. FOURIER SERIES
Proof of (4):
Let δ ∈ ]0, π/2[ be fixed. Since the function t 7→ tDn (t) is continuous, the Fundamental
Theorem of Calculus yields
Z δ Z δ
|tDn (t)| = lim |tDn (t)| dt .
0 →0+
We use the Jordan inequality (Lemma 4.5) and the identity in (3) above to write,
1
∀ ∈ ]0, π/2[ , ∀ t ∈ ], π/2[ , |tDn (t)| ≤ π sin n+ t ≤ π.
2
Consequently,
Z δ Z δ
lim |tDn (t)| dt ≤ lim π dt = πδ .
→0+ →0+
Lemma 4.10. Let f be a 2π-periodic function. Assume that f is Riemann integrable
on the interval [−π, π]. Then, for all x ∈ R,
Z π Z π
Dn (x − u)f (u) du = Dn (t) f (x − t) + f (x + t) dt ,
−π 0