fa1_normed_spaces
fa1_normed_spaces
Functional analysis is concerned with normed spaces and with operators between normed
spaces.
As an example we consider the initial value problem on I = [t0 , t1 ],
with the right side f : I × Rn → Rn . We transform (1.1) into the integral equation
Z t
y(t) = y0 + f (s, y(s)) ds . (1.2)
t0
We define an operator
Z t
n n
T : C(I; R ) → C(I; R ) , (T y)(t) = y0 + f (s, y(s)) ds . (1.3)
t0
y = Ty . (1.4)
In this manner we have transformed our original problem into an equation whose unknown
variable is a function (in this case the function y), not a number nor a vector with finitely
many components. In this equation there appears a mapping (here T ) between function
spaces (here C(I; Rn )). Such mappings are usually called “operators”. The function
spaces are typically infinite-dimensional Banach or Hilbert spaces.
In the following we write K for R or C.
kxk = 0 ⇔ x = 0, (1.5)
We repeat some basic notions from the first year analysis course. (See, for example, my
lecture notes.)
Let X be a normed space. A sequence (xn )n∈N in X is called convergent to the limit
x ∈ X, written as
lim xn = x , (1.8)
n→∞
if
lim kxn − xk = 0 . (1.9)
n→∞
2
The sequence is called a Cauchy sequence, if for every ε > 0 there exists an N ∈ N
such that
kxn − xm k ≤ ε , for all n, m ≥ N . (1.10)
X is called complete, if every Cauchy sequence in X has a limit (which by the definition
above has to be an element of X. If X is complete, X is called a Banach space. For
x ∈ X and ε > 0 we call
the open and closed ε-ball around x, respectively. A subset O of X is called open in
X, if for every x ∈ O there exists an ε > 0 such that B(x, ε) ⊂ O. A subset A of X is
called closed in X, if X \ A is open in X, or equivalently, if for every convergent sequence
(xn )n∈N in X whose elements xn all belong to A, its limit also belongs to A. The closure
Y , the interior int (Y ) and the boundary ∂Y of a subset Y of X are given by
\ [
Y = A , int (Y ) = O , ∂Y = Y \ int (Y ) . (1.13)
A⊃Y O⊂Y
A closed O open
A subspace U of the vector space X becomes itself a normed space, if we define the norm
on U to be the restriction of the given norm on X. If U is complete then U is closed in
X; if X itself is complete, the converse also holds. The closure U of a subspace U of X
is also a subspace of X.
In the basic analysis lectures we have already encountered some Banach spaces. The
spaces (Kn , k · kp ), 1 ≤ p ≤ ∞, are Banach spaces with
n
! p1
X
kxkp = |xi |p , 1 ≤ p < ∞, kxk∞ = max |xi | . (1.14)
1≤i≤n
i=1
becomes a Banach space for this norm k · kp , if we “identify” functions which are equal
almost everywhere. (“Identify” means that we pass to the quotient space formed by
3
equivalence classes of functions which are equal almost everywhere.) This is discussed in
detail in the course on measure and integration theory.
Sequence spaces. Let us set D = N. Then B(D; K) coincides with the space of all
bounded sequences, denoted as `∞ (K),
`∞ (K) = {x : x = (xk )k∈N , xk ∈ K, sup |xk | < ∞} , kxk∞ = sup |xk | . (1.18)
k∈N k∈N
Because this is a special case of (1.15), the space `∞ (K) is a Banach space. Let us consider
the subsets
Proposition 1.2 We have c0 (K) ⊂ c(K) ⊂ `∞ (K). Endowed with the supremum norm,
the spaces c0 (K) are c(K) Banach spaces.
Proof: Exercise. It suffices to show that c(K) is a closed subspace of `∞ (K), and that
c0 (K) is a closed subspace of c(K). 2
Moreover, we consider the space ce (K) of all finite sequences,
The space ce (K) is a subspace of c0 (K); it is not closed in c0 (K). Thus, it is not a Banach
space; we have (exercise)
ce (K) = c0 (K) . (1.22)
Let x = (xk )k∈N be a sequence in K. We define
∞
! p1
X
kxkp = |xk |p , 1 ≤ p < ∞, (1.23)
k=1
and
`p (K) = {x : x = (xk )k∈N , xk ∈ K, kxkp < ∞} , (1.24)
the space of sequences which are summable to the p-th power.
If x ∈ `p (K), α ∈ K, then
∞
! p1 ∞
! p1
X X
kαxkp = |αxk |p = |α| |xk |p = |α|kxkp .
k=1 k=1
4
In order to prove the triangle inequality, let x, y ∈ `p (K). For arbitrary N ∈ N we have
N
! p1 N
! p1 N
! p1
X X X
|xk + yk |p ≤ |xk |p + |yk |p ,
k=1 k=1 k=1
so
kx + ykp ≤ kxkp + kykp .
Thus, `p (K) is a normed space. In order to show that it is complete, let (xn )n∈N be a
Cauchy sequence in `p (K). Then for all k, n, m ∈ N we have
∞
X
|xnk − xm
k |
p
≤ |xnj − xm p n m p
j | = kx − x kp .
j=1
Therefore, (xnk )n∈N is a Cauchy sequence in K for all k. Since K is complete, there exists
x∞ n
k = lim xk , k ∈ N.
n→∞
N
! p1 N
! p1 N
! p1
X X X
|xnk − x∞
k |
p
≤ |xnk − xm
k |
p
+ |xm ∞ p
k − xk |
k=1 k=1 k=1
N
! p1
X
∞ p
≤ kxn − xm kp + |xm
k − xk | . (1.25)
k=1
N
! p1
X m(N )
|xk − x∞
k |
p
≤ ε. (1.27)
k=1
N
! p1
X
|xnk − x∞
k |
p
≤ 2ε , for all n ≥ M and all N ∈ N.
k=1
5
Passing to the limit N → ∞ we obtain
holds for all convergent sequences (xn )n∈N in X. From the analysis course we know that
the assertions
and
(i) T is continuous on X.
(ii) T is continuous in 0.
“(iv)⇒(i)⇒(ii)”: obvious.
“(ii)⇒(iii)”: Contraposition. Assume that (iii) does not hold. We choose a sequence
(xn )n∈N in X satisfying
kT (xn )kY > nkxn kX . (1.30)
We set
1
zn = xn ,
nkxn kX
6
this is possible, because xn 6= 0 due to (1.30). It follows that zn → 0, but kT (zn )kY > 1
and therefore T (zn ) does not converge to 0. Consequently, (ii) does not hold. 2
Not all linear mappings are continuous. Here is an example: the unit vectors {ek : k ∈ N}
form a basis of ce (K). We define a linear mapping T : ce (K) → K by T (ek ) = k. Then we
have kek k∞ = 1 and |T (ek )| = k, so (iii) in Proposition 1.4 is not satisfied.
In the following we will write kxk instead of kxkX if it is obvious which norm is meant.
We will also write T x instead of T (x).
Proposition 1.6 Let X, Y be finite dimensional normed spaces with dim(X) = dim(Y ).
Then X ' Y .
Proof: Let dim(X) = n, let {v1 , . . . , vn } be a basis of X, let e1 , . . . , en be the unit vectors
in Kn . We define
n
X n
X
n
T :K →X, Tx = xi vi if x= xi ei .
i=1 i=1
7
2
If X, Y are finite dimensional vector spaces with Y ⊂ X, Y 6= X, then dim(Y ) < dim(X),
and X und Y are not isomorphic, since bijective linear mappings leave the dimension
invariant. In the infinite dimensional case, the situation is not that simple. For example,
c0 (K) ⊂ c(K) and c0 (K) 6= c(K), but we have
is the inverse of T . One may compute directly that T ◦ S and S ◦ T are equal to the
identity on c0 (K) and c(K) resp., and that kT xk∞ ≤ 2kxk∞ as well as kSyk∞ ≤ 2kyk∞
hold for all x ∈ c(K) and all y ∈ c0 (K).
Proposition 1.7 Let (X1 , k·k1 ), . . . , (Xm , k·km ) be normed spaces. On the product space
m
Y
X= X i = X1 × · · · × Xm (1.37)
i=1
the expressions
m
! p1
X
kxk∞ = max kxi ki , kxkp = kxi kpi , (1.38)
1≤i≤m
i=1
Proof: Exercise. 2
Corollary 1.8 Let X be a normed space. The addition + : X × X → X and the scalar
multiplication · : K × X → X are continuous.
If αn → α and xn → x, then
8
Definition 1.9 (Space of operators, dual space)
Let X, Y be normed spaces. We define
The space L(X; K) is called the dual space of X, denoted X ∗ . The elements of X ∗ are
called functionals. 2
From linear algebra and analysis it is known that L(X; Y ) is a vector space.
kT xk
kT k = sup (1.40)
x∈X, x6=0 kxk
and
kT k = sup kT xk = sup kT xk , (1.42)
x∈X, kxk≤1 x∈X, kxk=1
as well as
kT k = inf{C : C > 0, kT xk ≤ Ckxk for all x ∈ X} . (1.43)
If Y is a Banach space, then so is L(X, Y ).
Proof: Let T ∈ L(X; Y ), let C > 0 with kT xk ≤ Ckxk for all x ∈ X. Dividing by kxk
we see that kT k ≤ C, so kT k ∈ R+ , and (1.41) as well as (1.43) hold. From (1.41) we
obtain that
sup kT xk ≤ sup kT xk ≤ kT k .
x∈X, kxk=1 x∈X, kxk≤1
Since
kT xk x
= T( ) ,
kxk kxk
(1.42) follows. We have
For α ∈ K we get
9
Thus, the properties of a norm are satisfied. We now prove that L(X, Y ) is complete. Let
(Tn )n∈N be a Cauchy sequence in L(X; Y ). Since
kTn x − Tm xk = k(Tn − Tm )(x)k ≤ kTn − Tm k kxk
the sequence (Tn x)n∈N is a Cauchy sequence in Y for every x ∈ X. Then (as Y is complete)
T x = lim Tn x
n→∞
Example 1.11
T is linear, Z b
|T x| = x(t) dt ≤ (b − a)kxk∞
a
with equality if x is constant. Therefore, T is continuous and kT k = b − a.
3. X = L1 ([a, b]; R) with the L1 norm, T as before, then
Z b Z b
|T x| = x(t) dt ≤ |x(t)| dt = kxk1
a a
10
4. X as before, f ∈ C([a, b]; R), T : X → R,
Z b
Tx = f (t)x(t) dt .
a
T is linear, and
Z b Z b
|T x| ≤ |f (t)| |x(t)| dt ≤ kf k∞ |x(t)| dt = kf k∞ kxk1 .
a a
5. Let D = (0, 1) × (0, 1), k ∈ L2 (D; R). On X = L2 ((0, 1); R) we want to define an
integral operator T : X → X by
Z 1
(T x)(s) = k(s, t)x(t) dt , s ∈ (0, 1) .
0
Once we have proved that this integral is well defined, it is clear that T is linear.
We have (the integrals are well defined as a consequence of Fubini’s and Tonelli’s
theorem, the second inequality follows from Hölder’s inequality)
Z 1 Z 1 2 Z 1 Z 1 2
kT xk2L2 ((0,1);R) = k(s, t)x(t) dt ds ≤ |k(s, t)| |x(t)| dt ds
0 0 0 0
Z 1 Z 1 Z 1
2 2
≤ |k(s, t)| dt · |x(t)| dt ds
0 0 0
Z 1Z 1
2
= kxkL2 ((0,1);R) |k(s, t)|2 dt ds .
0 0
11
6. Let X, Y be normed spaces with dim(X) < ∞. Then every linear mapping T : X →
Y is continuous: Let {v1 , . . . , vn } be a basis of X. Then for
n
X
x= xi vi ∈ X , xi ∈ K ,
i=1
we have that
n
X n
X n
X
kT xk = xi T vi ≤ |xi | kT vi k ≤ max kT vi k |xi | .
1≤i≤n
i=1 i=1 i=1
Therefore,
n
X
kT k ≤ max kT vi k , if kxk = |xi | .
1≤i≤n
i=1
7. Let T : Rn → Rm be linear, let T (x) = Ax for the matrix A ∈ R(m,n) . Via this
correspondence, we can identify the space L(Rn ; Rm ) with the space R(m,n) of all
m × n-matrices. The operator norm then becomes a so-called matrix norm whose
explicit form depends on the choice of the norms in Rn and Rm . Matrix norms
play an important role in the construction and analysis of numerical algorithms, in
particular in numerical linear algebra. 2
Proof: For all x ∈ X we have k(S ◦ T )xk ≤ kSk kT xk ≤ kSk kT k kxk. The assertion now
follows from Proposition 1.10. 2
Seminorms and quotient spaces.
Obviously,(1.45) implies p(0) = 0, but p(x) = 0 does not imply x = 0. Every norm is
a seminorm. If (Y, q) is a seminormed space, then every linear mapping T : X → Y
generates a seminorm on X by
p(x) = q(T x) . (1.47)
Here are some examples of seminorms:
X = Rn , p(x) = |x1 | , (1.48)
X = B(D; K) , a ∈ D , p(x) = |x(a)| , (1.49)
Z 1
1
X = L ((0, 1); R) , p(x) = x(t) dt , (1.50)
0
X = C 1 ([0, 1]; R) , p(x) = kẋk∞ . (1.51)
12
From linear algebra, we recall the concept of a quotient space. Let X be a vector space,
let U be a subspace of X. Then
x ∼U z ⇔ x−z ∈U (1.52)
[x] = {z : z ∈ X, x ∼U z} (1.53)
(iii) If moreover X is a Banach space and U is closed, then (X/U, p) is a Banach space.
0 = p([x]) = inf kx − zk .
z∈U
ky − x − zk ≤ 2p([y − x])
13
and then setting ỹ = y − z. Let now ([xn ])n∈N be a Cauchy sequence in X/U . Passing to
a subsequence if necessary we may assume that
We now choose according to (1.59) an x̃1 ∈ [x1 ] and for n > 1 an x̃n ∈ [xn ] mit
x ∼U y ⇔ x(0) = y(0) .
A=X (1.61)
holds. If A is dense in (X, d) and B is dense in (A, dA ), then B is dense in (X, d).
Proof: By the approximation theorem of Weierstra”s, the set P of all polynomials is dense
in C([a, b]; K). Moreover, the set of all polynomials with rational coefficients is countable.
It is dense in P , and therefore in C([a, b]; K), too. 2
Proposition 1.17 Let X be a normed space, let (xn )n∈N be a sequence in X with X =
span {xn : n ∈ N}. Then X is separable.
14
Proof: Exercise. 2
Proposition 1.18 The space `p (K) is separable for 1 ≤ p < ∞. The space `∞ (K) is not
separable.
Proof: For p < ∞ we have `p (K) = span {en : n ∈ N}. The assertion then follows from
1.17. Let now p = ∞. For arbitrary M ⊂ N we define xM ∈ `∞ (K) by
(
M 1, k ∈ M ,
xk =
0, k ∈ /M.
15
Proof: We want to define an isometric isomorphism T : `q (K) → (`p (K))∗ by
∞
X
(T x)(y) = xk y k , x ∈ `q (K) , y ∈ `p (K) . (1.64)
k=1
Thus, (1.64) defines for any given x ∈ `q (K) a linear continuous mapping T x : `p (K) → K
which satisfies
kT xk ≤ kxkq . (1.65)
Therefore T : `q (K) → (`p (K))∗ is well-defined. It follows from (1.64) that T is linear
and from (1.65) that T is continuous. T is injective, as (T x)(ek ) = xk , and consequently
T x = 0 implies that xk = 0 for all k. We now prove that T is surjective. Let y ∗ ∈ (`p (K))∗
be arbitrary. We want to find an x = (xk )k∈N in `q (K) such that T x = y ∗ . Since
(T x)(ek ) = xk for such an x, we have to define
xk = y ∗ (ek ) . (1.66)
Let y ∈ ce (K),
N
X
y= yk ek , yk ∈ K , N ∈ N, (1.67)
k=1
be given. We have
N
! N N
X X X
∗ ∗ ∗
y (y) = y yk ek = yk y (ek ) = y k xk . (1.68)
k=1 k=1 k=1
We choose
yk = |xk |q−1 sign (xk ) , 1≤k≤N. (1.69)
Then
|yk |p = |xk |p(q−1) = |xk |q = xk yk ,
and it follows from (1.68) that
N N N
! p1 N
! p1
X X X X
|xk |q = xk yk = y ∗ (y) ≤ ky ∗ k kykp = ky ∗ k |yk |p = ky ∗ k |xk |q .
k=1 k=1 k=1 k=1
Thus,
N
! 1q
X
|xk |q ≤ ky ∗ k . (1.70)
k=1
16
Passing to the limit N → ∞ yields x ∈ `q (K) and
kxkq ≤ ky ∗ k . (1.71)
We recall that ce (K) is dense in (`p (K), k · kp ). It follows from Proposition 1.19 that
T x = y∗ . (1.72)
Therefore T is surjective, and from (1.65) and (1.71) we obtain that kT xk = kxkq . There-
fore, T is an isometric isomorphism. 2
We mention some other results concerning the representation of dual spaces. We have
c0 (K)∗ ∼
= `1 (K) , `1 (K)∗ ∼
= `∞ (K) . (1.73)
For the case p = 2, (1.74) is a consequence of a general result for Hilbert spaces, which
will be treated in the next chapter. We sketch the proof for arbitrary p ∈ (1, ∞); its
general structure is the same as that for the sequence space `p (K). Setting
Z
(T x)(y) = x(t)y(t) dt , x ∈ Lq (D; K), y ∈ Lp (D; K) , (1.75)
D
one obtains a linear continuous mapping T : Lq (D; K) → Lp (D; K)∗ with kT xk = kxkq .
Indeed, by virtue of Hölder’s inequality,
Z Z 1q Z p1
q p
|(T x)(y)| ≤ |x(t)| |y(t)| dt ≤ |x(t)| dt |y(t)| dt = kxkq kykp , (1.76)
D D D
and for
y(t) = sign (x(t))|x(t)|q−1
we have |y(t)|p = |x(t)|q and therefore
Z Z
q−1
(T x)(y) = x(t)sign (x(t))|x(t)| dt = |x(t)|q dt
D D
Z 1q Z p1
q q
= |x(t)| dt |x(t)| dt = kxkq kykp .
D D
In order to prove that T is surjective, for a given y ∗ ∈ Lp (D; K)∗ one constructs an
x ∈ Lq (D; K) satisfying T x = y ∗ by employing the Radon-Nikodym theorem from measure
and integration theory.
17
One also has the result that
L1 (D; K)∗ ∼
= L∞ (D; K) . (1.77)
The representation theorem of Riesz states that, for compact sets D ⊂ Rn , the space
C(D; R)∗ is isometrically isomorphic to the space of all signed regular measures on the
Borel σ-algebra on D. In particular, for every y ∗ ∈ C(D; R)∗ there exists a measure µ
such that Z
∗
y (y) = y dµ . (1.78)
D
of the partial sums converges to an element s ∈ X, we say that the corresponding series
P ∞
k=1 xk converges, and we define
X∞
xk = s . (1.80)
k=1
P∞
The series k=1 xk is called absolutely convergent if
∞
X
kxk k < ∞ . (1.81)
k=1
Because addition and scalar multiplication in normed spaces are continuous operations,
we have the rules
∞
X ∞
X ∞
X ∞
X ∞
X
(xk + yk ) = xk + yk , αxk = α xk , (1.82)
k=1 k=1 k=0 k=1 k=1
α ∈ K. They are valid when the limits on the right side exist.
P∞
Proposition 1.22 Let X be a Banach space, assume that the series k=1 xk converges
absolutely. Then it also converges, and
∞
X ∞
X
xk ≤ kxk k . (1.83)
k=1 k=1
Moreover, every reordering of the series converges, and the limits are identical.
Proof: Let n
X
σn = kxk k .
k=1
18
For the partial sums defined in (1.79) we have, if n > m,
n
X
ksn − sm k ≤ kxk k = |σn − σm | .
k=m+1
Since (σn ) is a Cauchy sequence in K, also (sn ) is a Cauchy sequence in X, hence conver-
gent to some s ∈ X. Due to ksn k ≤ |σn |, (1.83) follows from
∞
X
ksk = lim ksn k ≤ lim |σn | = kxk k .
n→∞ n→∞
k=1
Let ∞
X
x̃k , x̃k = xπ(k) , π : N → N bijective,
k=1
P
be a reordering of xk with the partial sums
n
X
s̃n = x̃k .
k=1
therefore ks̃n − sn k → 0. 2
19