FAII Skript
FAII Skript
Functional Analysis II
April 3, 2013
1
Functional Analysis II
Definition 12.1.1. Let E be a Banach space over K and let T ∈ L (E) be a continuous
linear mapping on E.
i) λ ∈ σ(T ) ⇒ |λ| ≤ kT k;
ii) σ(T ) is closed, i.e. the resolvent set ̺(T ) := K \ σ(T ) is open;
We shall study linear operators on complex Hilbert spaces. Following definition recalls
the basic types of such operators.
ii) There is a unique operator T ∗ ∈ L (H), such that hT x, yi = hx, T ∗ yi for all x, y ∈ H.
This operator is called the adjoint operator to T ;
Few basic properties of these classes of operators may be found in Exercises. We give
one example here.
S((x1 , x2 , . . . )) = (0, x1 , x2 , . . . ).
We show that
where σp (S) is the set of eigenvalues of S (the so-called point spectrum) and σap (S) =
{λ ∈ C : inf{kSx − λxk : kxk = 1} = 0} is the approximative spectrum.
2
12 Spectral theory for bounded operators on complex Hilbert spaces
shows that λ 6= 0 and z = (−1/λ, −1/λ2 , −1/λ3 , . . . ), which does not lie in ℓ2 . We
conclude, that there is no z ∈ ℓ2 with (S − λI)z = (1, 0, 0, . . . ), i.e. (S − λI) does not map
ℓ2 onto ℓ2 , and is therefore not invertible. Therefore, σ(S) = {λ ∈ C : |λ| ≤ 1}.
If |λ| < 1 and x ∈ ℓ2 , we have kSx−λxk ≥ kSxk−kλxk = (1−|λ|)kxk, i.e. λ 6∈ σap (S).
Finally, if |λ| = 1, we put
1
xn = √ · (1, λ−1 , λ−2 , . . . , λ−(n−1) , 0, . . . ).
n
It follows, that kxn k = 1 for all n ∈ N and
Therefore, λI − T is injective and (λI − T )−1 : ran(λI − T ) → H with norm at most 1/d.
Hence, ran(λI − T ) and H are isomorphic and ran(λI − T ) is closed.
Let us assume, that there is x0 ∈ ran(λI − T )⊥ with kx0 k = 1. Then we obtain
0 = h(λI − T )x0 , x0 i = λ − hT x0 , x0 i
3
Functional Analysis II
Functional calculus is a way, how to define f (T ) for (at least some) functions f and
Xn
operators T . Surely, if T ∈ L (H) and p(z) = ak z k is a polynomial, we would like to
k=0
n
X
have p(T ) = ak T k .
k=0
Moreover, for H = Cn and T self-adjoint (i.e. a hermitian matrix), we may consider
the diagonalization
T = U −1 DU,
∞
X
where U is unitary4 and D is diagonal. Let f (z) = ak z k be holomorph on the whole
k=0
C, then (at least formally)
∞ ∞ ∞ ∞
!
X X X X
k −1 k −1 k −1 k
f (T ) = ak T = ak (U DU ) = ak U D U =U ak D U
k=0 k=0 k=0 k=0
= U −1 f (D)U,
where
f (d1 ) 0 0 ...
0 f (d2 ) 0 ...
f (D) = . . . ..
. . .
. . . .
0 0 . . . f (dn )
is again a diagonal matrix. We see, that to define f (T ), f needs to be defined only on
σ(T ) and “f (σ(T )) = σ(f (T ))”, the so-called Spectral Mapping Theorem.
i) Φ(z)5 = T, Φ(1) = I;
a) Φ is linear;
b) Φ is multiplicative: Φ(f g) = Φ(f ) ◦ Φ(g);
c) Φ is involutive: Φ(f )∗ = Φ(f );
iii) Φ is continuous.
4
12 Spectral theory for bounded operators on complex Hilbert spaces
n
X n
X
Existence: We set Φ(f ) = ak T k for f (z) = ak z k a polynomial. If we show continuity
k=0 k=0
of Φ on polynomials, then there would be a unique extension, which we denote by Φ again.
We shall use the Spectral Mapping Theorem for polynomials (cf. Exercise 1.4)
to obtain
We refer to Exercise 1.1.2 for the first equality, the second equality follows from a simple
calculation for polynomials, the third one is based on Exercise 2.3 and the fact that Φ(f f )
is self-adjoint.
All the required properties are then proven by approximation. Let us assume (for
example) that pn ⇉ f and qn ⇉ g for some polynomials pn , qn and f, g ∈ C(σ(T )). Then
we get
Φ(f g) ← Φ(pn qn ) = Φ(pn ) ◦ Φ(qn ) → Φ(f ) ◦ Φ(g).
The proof of the other properties is similar.
i) kf (T )k = kf k∞ := supλ∈σ(T ) |f (λ)|.
iv) σ(f (T )) = f (σ(T )), i.e. the Spectral Mapping Theorem holds for all f ∈ C(σ(T )).
Proof. (i) was proven for polynomials already before. For general pn ⇉ f we have
kΦ(f )k ← kΦ(pn )k = kpn k∞ → kf k∞ .
Let f ≥ 0 and take g ∈ C(σ(T )) with g 2 = f and g ≥ 0. Then
hf (T )x, xi = hg(T )x, g(T )∗ xi = hg(T )x, g(T )xi = hg(T )x, g(T )xi = kg(T )xk2 ≥ 0.
(iii) is again clear for polynomials, through approximation it follows for all f ∈ C(σ(T )).
Also (iv) was already discussed for polynomials. Let µ 6∈ f (σ(T )). Then g := (f − µ)−1 ∈
C(σ(T )) and g(f − µ) = (f − µ)g = 1. Hence, we get
5
Functional Analysis II
Let us now have a look on the spectral decomposition theorem for compact self-adjoint
operators and how it could be generalized to bounded operators. One obvious obstacle
would be that the spectrum does not have to be countable any more, i.e. we will have to
replace the sums by certain (appropriately interpreted) integrals. Furthermore, we have
now spent some time to define f (T ) for (at least some) functions f on σ(T ). We therefore
rewrite
X∞ ∞
X
Tx = λk hx, xk ixk = λk Ek x,
k=0 k=0
where Ek x = hx, xk ixk is the projection of x onto the linear span of xk . And if we put
fk (λj ) = δj,k , we get even
∞
X
fk (T )x = fk (λn )hx, xn ixn = hx, xk ixk = Ek x,
n=0
6
12 Spectral theory for bounded operators on complex Hilbert spaces
C(K)∗ ≈ M(K),
R
where µ(f ) = K f dµ. Furthermore, positive functionals (i.e. those with f ≥ 0 ⇒ ϕ(f ) ≥
0) correspond to non-negative measures.
Lemma 12.2.7. Let H be a complex Hilbert space and let Q : H → C be a function. Then
the following are equivalent.
i) There exists exactly one operator A ∈ L (H) such that Q(x) = hAx, xi for all x ∈ H.
ii) There is a constant C > 0 such that |Q(x)| ≤ Ckxk2 , Q(x + y) + Q(x − y) =
2Q(x) + 2Q(y) and Q(λx) = |λ|2 Q(x) for all x, y ∈ H and all λ ∈ C.
Proof. It is easy to show that (i) ⇒ (ii). If (ii) is satisfied, we put (motivated by polar-
ization identity)
and X
Ax = Ψ(x, eγ )eγ ,
γ∈Γ
where (eγ )γ∈Γ is some orthonormal basis of H. The rest follows by simple calculations.
Finally, let us observe that Ψ(x, y) = hAx, yi. Uniqueness of A follows from Exercises.
This mapping is linear, non-negative (i.e. f ≥ 0 implies hf (T )x, xi ≥ 0 for all x ∈ H).
Therefore, there exists a non-negative Radon measure E x on σ(T ), such that
Z
hf (T )x, xi = f dE x for all f ∈ C(σ(T )).
σ(T )
Z
If g is a bounded Borel-measurable function on σ(T ), the integral gdE x converges
σ(T )
and we shall show (using the Lemma above) that it is equal to some hGx, xi for all x ∈ H
and some G ∈ L (H). This G will then be defined to be g(T ).
Let us now elaborate on the program just stated.
Theorem 12.2.9. Let T ∈ L (H) be self-adjoint.
i) Let x ∈ H. Then there is a non-negative Radon measure E x , such that
Z
hf (T )x, xi = f dE x
σ(T )
7
Functional Analysis II
Proof. The proof is based on Lemma 12.2.7. Obviously, the operator is unique (hGx, xi =
hG̃x, xi for all x ∈ H implies G = G̃). Furthermore, g ≥ 0 implies hGx, xi ≥ 0, i.e.
positivity, and g real implies hGx, xi is real, i.e. G is self-adjoint.
So, we have to prove the existence. We put
Z
Q(x) = gdE x .
σ(T )
Then
Z Z
x
|Q(x)| ≤ kgk∞ dE = kgk∞ 1dE x = kgk∞ h1(T )x, xi = kgk∞ kxk2 . (12.1)
σ(T ) σ(T )
and Z Z
f dE λx = |λ|2 f dE x
σ(T ) σ(T )
| {z } | {z }
hf (T )(λx),λxi |λ|2 ·hf (T )x,xi
Due to the uniqueness of the Riesz Representation Theorem, we get E x+y + E x−y =
2E x + 2E y and E λx = |λ|2 E x . This verifies the assumptions of Lemma 12.2.7 and finishes
the proof.
Yet another theorem from the measure theory we shall use is that the set of bounded
Borel-measurable functions is the smallest set, which includes bounded continuous func-
tions and is closed under pointwise limits of such functions.
Theorem 12.2.10. Let K ⊂ C be compact. Let B d (K) be the Banach space of bounded
Borel-measurable functions on K equipped with the supremum norm. Let C(K) ⊂ U ⊂
B d (K) be a set of functions with the following property
(fn )n∈N ⊂ U with f (t) := lim fn (t) existing everywhere and sup kfn k∞ < ∞ ⇒ f ∈ U.
n→∞ n∈N
Then U = B d (K).
Proof. The proof is based on measure theory. We refer to Werner, Lemma VII.1.5 for
details.
i) Φ(z) = T, Φ(1) = I;
8
12 Spectral theory for bounded operators on complex Hilbert spaces
iii) Φ is continuous;
iv) fn ∈ B d (σ(T )) with supn kfn k∞ < ∞ and fn (t) → f (t) for every t ∈ σ(T ) implies
hΦ(fn )x, yi → hΦ(f )x, yi for every x, y ∈ H.
Proof. Uniqueness: (i), (ii) and (iii) imply uniqueness on C(σ(T )) and Theorem 12.2.10
on all B d (σ(T )).
Existence: We define Φ(g) = G, where G is the operator from Theorem 12.2.9. We have
to verify (i) − (iv). We know that
Z
hf (T )x, xi = f dE x
σ(T )
for all f ∈ C(σ(T )), i.e. Φ(f ) = f (T ) for all f ∈ C(σ(T )), where f (T ) is the continuous
functional calculus. This implies (i).
Let g be real valued. Then we obtain
where we have used (12.1), the fact, that G is self-adjoint and Corollary 11.6. If g is
complex-valued, we split it first into its real and imaginary part.
The proof of (iv) follows from
Z Z
x
hΦ(fn )x, xi = fn dE → f dE x = hΦ(f )x, xi,
σ(T ) σ(T )
w
and fn (T )x −
→ f (T )x.
9
Functional Analysis II
∞
X
Let X be a Banach space and let T ∈ L(X). Let f (z) = an z n be a potential series
n=0
with convergence radius bigger then r(T ). Then f is analytic on σ(T ) and we can define
∞
X
f (T ) := an T n .
n=0
where Γ ⊂ Ω is a suitable curve, which goes around σ(T ) exactly one times, and Rλ (T ) is
the resolvent mapping of T . Of course, this idea is heavily inspired by the Cauchy integral
known from complex analysis.
Lemma 12.3.1. Let T ∈ L (H) be self-adjoint and let A ⊂ σ(T ) be a Borel set. Then
EA := χA (T ) is an orthogonal projection.
i) E∅ = χ∅ (T ) = 0, Eσ(T ) := χσ(T ) (T ) = I;
10
12 Spectral theory for bounded operators on complex Hilbert spaces
Proof. The proof follows from the properties of the Borel-measurable functional calculus.
E : Σ → L (H), E(A) = EA
We have to check, that this definition is consistent. It is quite easy to see that the
integral of a simple function f does not depend on the form in which f was written. The
limit procedure is then justified by the following Lemma:
11
Functional Analysis II
n
X
Proof. Let kxk = 1 and let f = αi χAi with mutually disjoint A1 , . . . , An . Then
i=1
Z
2
Xn
2
n
X
f dE (x)
=
αi EAi (x)
= |αi |2 · kEAi (x)k2
i=1 i=1
n
X
≤ max |αi |2 · kEAi (x)k2 ≤ kf k2∞ · kxk2 = kf k2∞ .
i=1,...,n
i=1
Remark 12.3.7. On one hand, we have for each T ∈ L (H) self-adjoint the spectral
measure EA := χA (T ). On the otherR hand, we can define for every spectral measure with
compact support the operator T := λdEλ , i.e. we integrate the identity function λ → λ
with respect to E. We shall show that these two operations are actually inverse to each
other.
Theorem 12.3.8. Let E be a spectral measure R on R with compact support. Let T ∈ L (H)
be the self-adjoint operator defined by T = λdEλ . Then the mapping
Z
Ψ:f → f dE, Ψ : B d (σ(T )) → L (H)
σ(T )
Proof. We sketch first the proof that Eσ(T ) = I. Let f (λ) := λ and let fδ ⇉ f be simple
n(δ)
X
functions fδ = αi,δ χAi,δ with |fδ (λ) − λ| ≤ δ on the support of E and A1,δ , . . . , An(δ),δ
i=1
Z
Z
mutually disjoint. Let µ ∈ ̺(T ). As ̺(T ) is open and
fδ dE − T
=
(fδ (λ) − λ)dE
≤
R R
R Pn(δ)
δ, we conclude, that R fδ dE − µI = i=1 αi,δ EAi,δ − µI is invertible for 0 < δ ≤ δ0 (for
δ0 > 0 small enough). On the other hand, the inverse of such an operator has a norm
max{|αi,δ − µ|−1 : 1 ≤ i ≤ n(δ), EAi,δ 6= 0} → k(T − µI)−1 k. Hence, EAi,δ = 0 for
12
12 Spectral theory for bounded operators on complex Hilbert spaces
R
and follows from the fact, that the mapping g →R h( gdE)x, yiRis in C(σ(T ))∗ and may be
therefore represented by a measure νx,y , i.e. h( gdE)x, yi = gdνx,y . The proof is then
finished by Lebesgue dominated convergence theorem.
We estimate all the three summands as follows. Let Ψ be the Borel-measurable functional
calculus of T . Then Ψ(λ) = T and Ψ(f ) = f (T ). Hence
13
Functional Analysis II
Hence, kf (T ) − f (S)k = 0.
Finally, we estimate kf (S)−Sk. We use again (12.2) combined with the norm estimate
from Theorem 12.3.6 and obtain
Z
kf (S) − Sk =
(f (λ) − λ)dEλ
≤ kf (λ) − λk∞ ≤ δ
Let us summarize the structure of this section into the following diagram.
T ∈L (H)
Weierstraß Theorem Riesz
self-adjoint −−−−−−−−−−−−−−→ Cont. F.C.(Thm. 12.2.4) −−−−→ Borel FC (Thm. 12.2.11)
density of polynomials
EA :=χA (T )y
R
T ∈L (H) T := λdEλ R Thm. 12.3.6
self-adjoint ←−−−−−−− f dE ←−−−−−−− Spectral measure
Thm. 12.3.8
(f ◦ g)(S) = f (g(S)).
χA ◦ g = χg−1 (A)
and we have to show that χg−1 (A) (S) = χA (g(S)). Let F be the spectral measure of S
and E the spectral measure of g(S); then we have to show that Fg−1 (A) = EA for all Borel
sets A, or, equivalently,
hFg−1 (A) x, yi = hEA x, yi
for all x, y ∈ H and all Borel sets A.
Let x, y ∈ H be fixed and let us consider the measures
1
νx,y : A → hFg−1 (A) x, yi,
2
νx,y : A → hEA x, yi,
µx,y : A → hFA x, yi.
1 (A) = µ
Hence, νx,y −1
x,y (g (A)). We use the transformation law for the (complex-valued)
1
measures νx,y and µx,y
Z Z
1
hdνx,y = (h ◦ g)dµx,y
14
12 Spectral theory for bounded operators on complex Hilbert spaces
in the form
Z Z Z
1
h(λ)dνx,y (λ) = (h ◦ g)dµx,y = h(g(λ))dhFλ x, yi
and obtain
Z Z Z Z
λ dνx,y (λ) = g(λ) dhFλ x, yi = hg(S) x, yi = λ dhEλ x, yi = λn dνx,y
n 1 n n n 2
(λ).
Remark 12.4.2. We recall the notion of the push-forward measure. Let µ be a measure
on X1 and f : X1 → X2 a mapping. Then f∗ (µ)(B) = µ(f −1 (B)) for B ⊂ X2 is
called the push-forward measure of µ under f . To develop the complete theory (which
includes the transformation law we have used just now) one has to take care also about
the corresponding σ-algebras.
Corollary 12.4.3. Let T ∈ L (H) be positive and let n ∈ N. Then there exists a unique
positive S ∈ L (H) with S n = T.
Proof. The functions fn : t → t1/n are continuous, bounded and non-negative on σ(T ) ⊂
[0, ∞). We set S := fn (T ). Then S ≥ 0 and from t1/n . . . t1/n = t, it follows S n = T. Let
gn (t) = tn , then (fn ◦ gn )(t) = t for t ∈ [0, ∞) ⊃ σ(S) and
T + T∗ T − T∗
T = +i ,
2 } | {z
| {z 2i }
=:T1 =:T2
we can decompose T = T1 +iT2 , where T1 and T2 are self-adjoint and commuting operators.
Let E be the spectral measure associated with T1 and F the spectral measure associated
with T2 . We shall show that EA FB = FB EA for all sets A, B ⊂ R. We show first that
EA T2 = T2 EA . The same argument repeated for the pair (EA , T2 ) instead of (T1 , T2 ) then
gives the general statement.
15
Functional Analysis II
We observe that T1n T2 = T2 T1n , i.e. p(T1 )T2 = T2 p(T1 ) for all polynomials and, by
limiting arguments, also for all ϕ ∈ C(σ(T1 )). Finally, we consider ϕn → χA , which
implies ϕn (T1 )x → EA (x) for every x ∈ H. This finally gives
T2 (EA x) = T2 ( lim ϕn (T1 )x) = lim T2 (ϕn (T1 )x) = lim ϕn (T1 )(T2 x) = EA (T2 x).
n→∞ n→∞ n→∞
we get Z
T = T1 + iT2 = zdGz .
C
This finishes a sketch of the proof of the following theorem:
16
13 Spectral theory for unbounded operators
Also in this section, H denotes a Hilbert space over C.
iv) We say that T is closable if G(T ) is a graph of some linear operator T0 . This operator
is then called closure of T . We shall denote this by T0 = T .
The domain of T , i.e. dom(T ), is an essential part of the definition of every operator
T . For example, if S and T are two (unbounded) operators on H, we set
Proof. Let xn → 0 and T xn → z. We show that z = 0 and apply the Closed graph
theorem. But this follows easily from
D E
hz, zi = lim T xn , z = lim hT xn , zi = lim hxn , T zi = 0.
n→∞ n→∞ n→∞
17
Functional Analysis II
Since C0∞ (0, 1) is dense in L2 (0, 1), T is densely defined. It is easy to check, that T is
symmetric:
Z 1 Z 1 Z 1
′ ′
hT f, gi = if (t)g(t)dt = −i f (t)g (t)dt = f (t)ig′ (t)dt = hf, T gi.
0 0 0
However, T is not self-adjoint - the same calculation goes through also for f, g ∈ C01 (0, 1),
which means, that f → hT f, gi is continuous also for g ∈ C01 (0, 1) ⊂ dom(T ∗ ).
Therefore, there comes a question: what is the closure of T and the adjoint of T ?
Remark 13.1.6. In the sequel, we shall encounter the space AC([0, 1]) of absolutely
continuous functions on [0, 1]. Let us recall their definition and basic facts about them.
A function f : [0, 1] → R (or C) is called absolutely continuous, if for every ε > 0 there is
a δ > 0, such that
Xm
|f (bj ) − f (aj )| < ε
j=1
P
for every 0 ≤ a1 < b1 ≤ a2 < b2 ≤ · · · ≤ am < bm ≤ 1 with j (bj − aj ) < δ. Obviously, ev-
ery Lipschitz function on [0, 1] is absolutely continuous. Using the notion of total variation,
one shows that every absolutely continuous function is a difference of two nondecreasing
absolutely continuous functions. This, combined with the fact that nondecreasing func-
tions are differentiable almost everywhere, then gives that absolutely continuous functions
are also differentiable almost everywhere, i.e. it makes sense to ask if their derivative is
an element of L1 ([0, 1]).
We shall also need the following properties:
R1 R1 ′
(a) (partial integration): If f, g ∈ AC([0, 1]), then 0 f g′ = [f (x)g(x)]x=1x=0 − 0 f g.
18
13 Spectral theory for unbounded operators
Let us suppose first, that h ∈ AC([0, 1]) and h′ ∈ L2 ([0, 1]). Then
Z 1 Z 1 Z 1
′ 1 ′
hT f, hi = f h̄ = [f h̄]0 − f h̄ = − f h̄′ = hf, −h′ i,
0 0 0
We see, that Z
x
′
f (·), g(x) + h =0
0
Rx
for all f ∈ dom(T ). In other words, the function x → g(x) + R0 h is orthogonal to f ′ for all
x
f ∈ dom(T ). Therefore, this function is constant and (since 0 h is absolutely continuous)
g is also in AC([0, 1]) and g ′ = −h ∈ H. Furthermore, this shows that T ∗ g = h = −g ′ .
we get
dom(T ∗ ) = dom(T ) and T = T ∗ .
19
Functional Analysis II
i) T ∗ is closed.
Roughly speaking, T sends f to its Taylor series, and the domain contains only functions
for which this series converges uniformly and absolutely. We shall show that T is a densely
defined operator with dom(T ∗ ) = {0}.
We proceed in the following way
20
13 Spectral theory for unbounded operators
iii) Finally, ker(T ) and ran(T ) are dense for the operator above:
Obviously, ker T ⊃ C0∞ ((−1, 1) \ {0}), i.e. ker T contains all functions vanishing on
some neighborhood of zero and this set is dense in H. Furthermore, the range of
T includes all polynomials, as the constant Cf might be arbitrary large, hence also
ran(T ) is dense in H.
i) The set
ii) The mapping R : ̺(T ) → L (H), Rλ = R(λ) := (λI − T )−1 is called the resolvent
mapping.
i) Mf is densely defined:
Let An := {x ∈ Ω : |f (x)| ≤ n}. Then gχAn ∈ dom(Mf ) for all n ∈ N and all g ∈ H.
And gχAn → g in H finishes the argument.
21
Functional Analysis II
iii) σp (Mf ) = {z ∈ R : f −1 (z) has positive measure} are the eigenvalues (the so-called
point spectrum) of Mf and every eigenvalue of Mf has infinite multiplicity:
If the set Az := f −1 (z) has positive measure, then {g ∈ H : supp g ⊂ Az } is
an infinite-dimensional subspace of eigenvectors. If, on the other hand, Mf g(t) =
f (t)g(t) = zg(t) for almost every t ∈ (−1, 1), we get f (t) = z almost everywhere on
the support of g and the claim follows.
iv) z ∈ C is in σ(Mf ) if, and only if, for every ε > 0, the set {x ∈ Ω : |f (x) − z| < ε} =
f −1 ({w ∈ C : |w − z| < ε}) has positive measure:
Let z ∈ C be such that An := f −1 ({w ∈ C : |w − z| < 1/n}) has positive measure
for every n ∈ N and consider gn 6= 0 with supp gn ⊂ An . Then gn ∈ dom(Mf ) and
Z 1 Z
2 2 kgn k2
k(Mf − zI)gn k = |(f (t) − z)gn (t)| dt = |(f (t) − z)gn (t)|2 dt ≤
−1 An n2
If T is not closed, then (T − λI) is not closed for every λ ∈ C, hence G(T − λI) is
not closed. On the other hand, if (T − λI)−1 would be bounded (i.e. in L (H)), then
G((T − λI)−1 ) would be closed and, therefore, also G(T − λI) would be closed.
This can be summarized by saying that σ(T ) = C whenever T is not closed.
i) ̺(T ) is open.
Rλ − Rµ = (µ − λ)Rλ Rµ
holds.
Hence, λI − T is invertible.
and the similar identity for multiplication from the right side. So, only the (easy)
inspection of domains is missing.
22
13 Spectral theory for unbounded operators
Theorem 13.2.4. Let T be a densely defined operator. Then ker(T ∗ ) = ran(T )⊥ , where
Proof. y ∈ ker(T ∗ ) if, and only if, hT x, yi = 0 for all x ∈ dom(T ) and if, and only if
y ⊥ ran(T ).
Example 13.2.7. We present an example, where the change of the domain of the operator
has a big impact on its spectrum. This demonstrates once more, how important is the
right choice of the domain.
We consider again the operator f → if ′ on L2 (0, 1), on the following domains:
23
Functional Analysis II
It is quite easy to prove (similar to previous arguments) that both S and T are closed.
We show that
σ(S) = C, σ(T ) = ∅.
The claim on σ(S) is very easy to confirm. Just notice that ez (x) = e−izx ∈ dom(S)
for all z ∈ C and (S − z)ez = 0.
To find σ(T ), fix z ∈ C and let
Z x
(Rz f )(x) = −ie−izx eizt f (t)dt.
0
This is well defined for all f ∈ L2 (0, 1) and (Rz f )(x) is absolutely continuous function of
x ∈ [0, 1]. Hence Rz f ∈ L2 (0, 1) as well. An easy calculation shows that
Remark 13.3.2. i) Formally, it looks very similar to the case of bounded operators.
R
ii) σ(T ) might be an unbounded subset of R, therefore σ(T ) λdEλ integrates an un-
bounded function over an unbounded subset of R.
iii) Let E be a spectral measure on R and f be an unbounded (but everywhere finite)
Borel-measurable function. We let
Z
Df := {x ∈ H : |f (t)|2 dhEt x, xi < ∞}. (13.1)
R
24
13 Spectral theory for unbounded operators
Here, the integration with respect to d|hEt x, yi| means the integration with respect to
the variation of the measure µx,y : A → hEA x, yi, i.e. |µx,y |.
Proof. (Sketch)
First, we show that Df is a linear set. From
hEA (x + y), x + yi = kEA (x + y)k2 ≤ (kEA xk + kEA yk)2 ≤ 2(kEA xk2 + kEA yk2 )
we deduce that µx+y,x+y ≤ 2µx,x + 2µy,y , where again µx,x (A) = hEA x, xi. It follows, that
x, y ∈ Df implies x + y ∈ Df . Similar calculation shows also that cx ∈ Df for c ∈ C.
Second, we show that Df is dense. Let ωn = {t ∈ R : |f (t)| < n} ⊂ R and let
y ∈ ran(Eωn ). Then
Z Z
2
hEωnc y, yi = 0 and |f (t)| dhEt y, yi = |f (t)|2 dhEt y, yi ≤ n2 kyk2 < ∞.
R ωn
Therefore, y ∈ Df . Furthermore, Eωn x → x due to kx − Eωn xk2 = kEωnc xk2 = µx,x (ωnc ) →
0.
Finally, let us prove the inequality (13.2). First, we observe that (for x 6= 0)
As ( n )
X n
[
|µx,y |(A) = sup |µx,y (Ak )| : Ak = A, A1 , A2 , . . . , Ak disjoint ,
k=1 k=1
we get
( n n
)
X [
|µx,y |(A) ≤ sup kEAk xk · kEAk yk : Ak = A, A1 , A2 , . . . , Ak disjoint
k=1 k=1
!1/2 !1/2
X
n n
X n
[
≤ sup kEAk xk2 · kEAk yk2 : Ak = A, A1 , A2 , . . . , Ak disjoint
k=1 k=1 k=1
For a general function f , we consider a sequence of simple functions |fn | ր |f | and apply
Levi’s Theorem (sometimes also called Lebesgue monotone convergence theorem).
25
Functional Analysis II
Theorem 13.3.5. Let f : R → C be measurable. Then dom(Tf ) = H if, and only if, f is
essentially bounded w.r.t. to E.
Proof. Let x ∈ H, then hER x, xi ≤ kxk2 . If f is bounded, we integrate in (13.1) a bounded
function over a finite measure space, hence x ∈ Df .
If, on the other hand, dom(Tf ) = Df = H, then Tf ∈ L (H) by Closed Graph Theorem.
Furthermore, if ωn := {t ∈ R : |f (t)| ≥ n}, then Theorem 13.3.4 (ii) implies that kTf xk ≥
nkxk for x ∈ ran(Eωn ). Therefore Eωn = 0 for n large.
Proof. (Sketch of the proof of Theorem 13.3.1).
Let R := (T − iI)−1 ∈ L (H). Then RR∗ = R∗ R, i.e. R is normal (and R∗ = (T + iI)−1 ∈
L (H)). According to the spectral theorem for bounded normal operators,
Z
R = (T − iI)−1 = zdFz ,
σ(R)
26
14 Distributions and Fourier transform
14.1 The space S (Rn ) and the Fourier transform on S (Rn )
We shall use the usual notation from vector analysis, i.e. let n ∈ N, α ∈ Nn0 be a multiindex
and x = (x1 , . . . , xn ) ∈ Rn , then we write
xα := xα1 1 . . . xαnn ,
|α| := α1 + · · · + αn ,
∂ |α|
D α := ∂1α1 . . . ∂nαn = α1 ,
∂x1 . . . ∂xαnn
q
|x| := x21 + · · · + x2n ,
α! := α1 ! . . . αn !.
ii) A sequence (ϕj )j∈N ⊂ S (Rn ) is said to converge in S (Rn ) to ϕ ∈ S (Rn ) if, and
only if,
lim kϕj − ϕk(k,l) = 0 for all k, l ∈ N0 .
j→∞
S
This will be written as ϕj → ϕ.
Remark 14.1.2. The space S (Rn ) is a vector space (in the algebraic sense), but it is
not a Banach space (or normed space). It can be shown, that it is actually impossible to
introduce a norm S (Rn ), such that the convergence in this norm would be equivalent to
convergence, which we have just described. Nevertheless, it is possible to equip S (Rn )
with a topology, such that the convergence in this topology and S -convergence are equiv-
alent. Together with this topology, S (Rn ) becomes a topological vector space. We refer to
the Book of Rudin [4], which contains a lot of details on such spaces. On the other hand,
the space is metrizable, cf. Exercises.
S
Theorem 14.1.3. i) Let (ϕk )k∈N ⊂ S (Rn ) with ϕk → ϕ ∈ S (Rn ). Then ϕk → ϕ
also in Lp (Rn ) for every 0 < p ≤ ∞.
S S
ii) Let (ϕk )k∈N ⊂ S (Rn ) with ϕk → ϕ ∈ S (Rn ). Then D α ϕk → D α ϕ for every
α ∈ Nn0 .
S S
iii) Let (ϕk )k∈N ⊂ S (Rn ) with ϕk → ϕ ∈ S (Rn ). Then xα ϕk → xα ϕ for every α ∈ Nn0 .
S
iv) Let ϕ ∈ S (Rn ) and let τh ϕ(x) = ϕ(x − h). Then τh ϕ → ϕ for h → 0.
R
v) Let ϕ, ψ ∈ S (Rn ). Then ϕψ and (ϕ ∗ ψ)(x) = Rn ϕ(x − y)ψ(y)dy belong both to
S (Rn ) and it holds
D α (ϕ ∗ ψ) = (D α ϕ) ∗ ψ = ϕ ∗ (D α ψ).
27
Functional Analysis II
Proof. We have
|ϕj (x) − ϕ(x)| ≤ (1 + |x|2 )−k/2 kϕj − ϕk(k,0) .
Choosing k large enough, integrating over x ∈ Rn and using kϕj − ϕk(k,0) → 0 gives (i).
Similarly, (ii) and (iii) follows by using only the definition of S (Rn ).
To prove (iv), we have to estimate kϕ(x + h) − ϕ(x)k(k,l) . We estimate using Taylor
theorem
|D α ϕ(x + h) − D α ϕ(x)| ≤ cα |h| · sup sup |D α+β ϕ(y)|
|β|=1 y∈[x,x+h]
and obtain
kτh ϕ − ϕk(k,l) ≤ ck,l |h| · kϕk(k,l+1) , |h| ≤ 1.
Finally, we give the proof of (v). Let ϕ, ψ ∈ S (Rn ). Then (ϕ ∗ ψ) is well defined, i.e.
the integral converges for every x ∈ Rn . The identity for derivatives of (ϕ ∗ ψ) follows by
Lebesgue’s dominated convergence theorem. Indeed, let ej = (0, . . . , 0, 1, 0, . . . , 0) with 1
in the jth entry and zeros everywhere else.
(ϕ ∗ ψ)(x + hej ) − (ϕ ∗ ψ)(x)
D ej (ϕ ∗ ψ)(x) = lim
h→0
Z h
1
= lim [ϕ(x + hej − y) − ϕ(x − y)]ψ(y)dy
h→0 h Rn
Z
ϕ(z + hej ) − ϕ(z)
= lim · ψ(x − z)dz
h→0 Rn h
Then
ϕ(z + hej ) − ϕ(z)
− D ej ϕ(z) → 0
h
for every z ∈ Rn as h → 0 and since this expression is uniformly bounded by a constant
depending only on ϕ, its integral with respect to the measure ψ(x − z)dz converges to 0
as h → 0. The general case follows by repeating this argument by induction.
Next we show, that ϕ ∗ ψ belongs also to S (Rn ). We have (for N > n arbitrary)
Z
|(ϕ ∗ ψ)(x)| ≤ CN (1 + |x − y|2 )−N/2 (1 + |y|2 )−N/2 dy, (14.1)
Rn
where CN depends only on ϕ, ψ and N > n. The part of the integral in (14.1) over the set
{y ∈ Rn : |x − y| > |x|/2} is bounded by
Z
(1 + (|x|/2)2 )−N/2 (1 + |y|2 )−N/2 dy ≤ BN (1 + |x|2 )−N/2 ,
y:|x−y|>|x|/2
where BN depends only on N and the dimension. On the other hand, if |x|/2 > |y − x|,
then |y| > |x|/2 and the part of the integral in (14.1) over the set {z ∈ Rn : |y −x| < |x|/2}
may be estimated from above by
Z
(1 + |x − y|2 )−N/2 (1 + (|x|/2)2 )−N/2 dy ≤ BN (1 + |x|2 )−N/2 .
y:|x−y|<|x|/2
Therefore, (ϕ ∗ ψ)(x) decays at infinity as (1 + |x|2 )−N/2 - and this holds for every N > n.
Using D α (ϕ ∗ ψ) = (D α ϕ ∗ ψ) and replacing ϕ by D α ϕ, we obtain by the same argument,
that also D α (ϕ∗ψ) decays faster then the reciprocal of any polynomial, and ϕ∗ψ ∈ S (Rn )
follows.
Finally, that ϕψ belongs to S (Rn ) follows directly by the Leibniz rule.
28
14 Distributions and Fourier transform
We will prove later on, that the terminology used is correct, especially, that F −1 really
is an inverse of F.
Furthermore, let us mention that many authors use another normalization of Fourier
transform, i.e. Z
f (x)e−2πihx,ξi dx.
Rn
S
iii) Let (ϕj )∞ n
j=1 ⊂ S (R ) with ϕj → ϕ. Then
S S
Fϕj → Fϕ and F −1 ϕj → F −1 ϕ.
Proof. If ϕ ∈ S (Rn ) and α ∈ Nn0 , then one gets from Theorem 14.1.3 that xα ϕ ∈ S (Rn )
and D α ϕ ∈ S (Rn ). Furthermore, we get by Lebesgue’s dominated convergence theorem
Z
∂ 1
(Fϕ)(ξ) = (−i)xl e−ihx,ξi ϕ(x)dx = (−i)F(xl ϕ(x))(ξ).
∂ξl (2π)n/2 Rn
By iteration, we get (14.2). To prove (14.3), we use partial integration in xl -direction with
intervals tending to R to obtain
Z Z
i ∂ −ihx,ξi −i ∂ϕ
ξl (Fϕ)(ξ) = n/2
(e )ϕ(x)dx = n/2
(x)e−ihx,ξi dx
(2π) Rn ∂xl (2π) Rn ∂xl
∂ϕ
= (−i)F (ξ).
∂xl
∂2 ∂
(Fϕ)(ξ) = (−i) (F(xl ϕ(x))(ξ)) = (−i)2 F(xm xl ϕ(x))(ξ).
∂ξm ∂ξl ∂ξm
29
Functional Analysis II
ii) Let h ∈ Rn and ϕ ∈ S (Rn ). Then we denote by (τh ϕ)(x) = ϕ(x − h) the translation
operator. And the formula F(τh ϕ) = e−ihh,ξi Fϕ holds for all ϕ ∈ S (Rn ).
iii) Let h ∈ Rn and ϕ ∈ S (Rn ). Then we denote by (Mh ϕ)(x) = eihh,xi ϕ(x) the modu-
lation of ϕ. And the formula F(Mh ϕ) = τh (Fϕ) holds for every ϕ ∈ S (Rn ).
iv) Let ϕ, ψ ∈ S (Rn ). Then the convolution of ϕ and ψ satisfies the formula
F(ϕ ∗ ψ)(ξ) = (2π)n/2 (Fϕ)(ξ) · (Fψ)(ξ), ξ ∈ Rn .
30
14 Distributions and Fourier transform
2 2
Solving this equation, we observe that h(s) = h(0)e−s /2 = e−s /2 , which finishes the
proof.
We leave the proof of (vi) open in the moment and come back to that once we establish
the properties of the inverse Fourier transform.
Obviously, similar statements can be proven for F −1 in the same way. So far we know
that FS (Rn ) ⊂ S (Rn ) and F −1 S (Rn ) ⊂ S (Rn ).
Proof. When trying to evaluate F −1 (Fϕ) using Fubini’s theorem, one runs into non-
convergent integrals. Therefore, we use the following trick. We evaluate
Z Z
1 2 |ξ|2 /2 1
(Fϕ)(ξ)eihx,ξi e−ε dξ → (Fϕ)(ξ)eihx,ξi dξ = F −1 (Fϕ)(x)
(2π)n/2 Rn (2π)n/2 Rn
as ε → 0+ .
2 2
Let ψ(x) = e−ε |x| /2 . Then
2
|y|2
Fψ(y) = ε−n F e−|x| /2 (y/ε) = ε−n e− 2ε2 .
2 |x|2 /2
This, applied to ψ(x) = e−ε then gives
Z Z
1 ihx,ξi −ε2 |ξ|2 /2 ε−n |y|2
− 2
(Fϕ)(ξ)e e dξ = ϕ(x + y)e 2ε dy
(2π)n/2 Rn (2π)n/2 Rn
Z
1 2
= n/2
ϕ(x + εz)e−|z| /2 dz.
(2π) Rn
Finally, we come to the proof of (vi) in Theorem 14.1.6. Let us observe that in (14.4),
we have shown actually that
(2π)n/2 F −1 (Fϕ · ψ) = ϕ ∗ F −1 ψ.
31
Functional Analysis II
is also continuous, hence T ∈ S ′ (Rn ). With a slight modification of the arguments, the
same holds for also for f ∈ Lp (Rn ) for all 1 ≤ p ≤ ∞. However, it does not hold for p < 1,
and not for Lloc n
1 (R ).
Every distribution, which is equal to Tf for some f ∈ L1 (Rn ), is called regular dis-
tribution. Every distribution, which is not regular is called singular distribution. In this
sense, we may identify L1 (Rn ) with the set of regular distributions, which is a subset of
S ′ (Rn ). The following lemma shows, that this identification is also one-to-one.
Lemma 14.2.4. Let f, g ∈ L1 (Rn ) and let Tf (ϕ) = Tg (ϕ) for all ϕ ∈ S (Rn ). Then f = g
a.e.
Proof. There is a lot of different proofs in the literature. We shall present one which uses
the technique of mollification, which we shall encounter also later on. Obviously, it is
enough to show that if Th (ϕ) = 0 for some h ∈ L1 (Rn ) and all ϕ ∈ S (Rn ), then h = 0
a.e.
Let ω ∈ S (Rn ) Rbe a smooth non-negative symmetric function with support in B :=
{y −n
R : |y| ≤ 1} and Rn ω(x)dx = 1. We define ωε (x) = ε ω(x/ε). Observe that also
Rn ωε (x)dx = 1. We observe the following facts
32
14 Distributions and Fourier transform
ii) Using Fubini’s theorem (and the symmetry of ω) we get quickly (as ωε ∗ ϕ ∈ S (Rn ))
Z Z
0= h(x)(ωε ∗ ϕ)(x)dx = ϕ(y)(h ∗ ωε )(y)dy.
Rn Rn
iv) Finally, we use that, for every t > 0, h ∈ L1 (Rn ) may be written as h = h1 + h2 ,
where h1 is continuous with compact support and kh2 k1 ≤ t.
where
Z
(∗) = |h1 (x) − ωε ∗ h1 (x)|dx
ZR Z
n
= [h1 (x) − h1 (x − y)]ωε (y)dy dx
ZR Rn
n
Z
≤ ωε (y) |h1 (x) − h1 (x − y)|dxdy
Rn Rn
≤ sup kh1 (·) − h1 (· − y)k1 .
y:|y|≤ε
Due to the bounded support of h1 and its uniform continuity, the last expression
goes to zero as ε → 0. Hence, choosing ε, t > 0 small enough, khk1 is arbitrary small
and, therefore, equal to zero, i.e. h = 0 a.e.
We show how one can extend the convolution also to measures. Let first ϕ, ψ ∈ S (Rn )
and f ∈ S (Rn ) as well. We denote by µ the measure on Rn , which has density ϕ with
respect to the Lebesgue measure. Furthermore, λ corresponds to ψ in the same way. Then
Z Z Z
(ϕ ∗ ψ)(f ) = (ϕ ∗ ψ)(x)f (x)dx = f (x) ϕ(x − y)ψ(y)dydx
ZR n R n
ZR n
33
Functional Analysis II
We observe that the last expression makes sense for all Borel measures µ and λ and
f ∈ C0 (Rn ), i.e. continuous functions which tend to zero at infinity.
Hence, for two Borel measures µ and λ on Rn , we define µ ∗ λ to be the unique measure
on Rn , such that (using the Riesz representation theorem)
Z Z
f d(µ ∗ λ) = f (x + y)dµ(x)dλ(y) (14.6)
Rn Rn ×Rn
for all f ∈ C0 (Rn ). Using standard arguments, this holds also for all bounded Borel-
measurable functions, i.e. fξ (x) = eihx,ξi . Therefore, we obtain immediately the formula
Furthermore, kµ ∗ λk ≤ kµk · kλk, where the norms denote total variation of measures.
Finally, if µ is absolutely continuous with respect to the Lebesgue measure, then the same
is true for µ ∗ λ. To see this, consider f = χE , where the Lebesgue measure of E is zero.
Then Z
0= f (x + y)dµ(x)
Rn
for all y ∈Rn , and (µ ∗ λ)(E) = 0 follows.
Finally, (14.6) implies that
Remark 14.2.6. Let us sketch the proof (given by Schwartz) that it is impossible to
introduce a continuous multiplication of distributions.
R Let ψ ∈ S (Rn ) be compactly
supported non-negative function with ψ(0) = 1, ψ = 1. Then ψk (x) := kn ψ(kx) converge
to δ in S ′ (Rn ) as k → ∞. By the definition of multiplication of distribution δ and test
function ψk , we obtain
for all ϕ ∈ S (R) with ϕ(0) > 0. On the other hand, assuming that it is possible to define
continuous multiplication in S ′ (Rn ) implies that (δψk )(ϕ) → (δ · δ)(ϕ) = δ2 (ϕ), i.e. a
contradiction.
34
14 Distributions and Fourier transform
iv) Let h ∈ Rn and T ∈ S ′ (Rn ). Then we denote by (τh T )(ϕ) = T (ϕ(· + h)), ϕ ∈
S (Rn ), the translation operator. The formula F(τh T ) = e−ihh,ξi FT holds for all
T ∈ S ′ (Rn ).
Proof. We obtain
for every ϕ ∈ S (Rn ), i.e. F(F −1 T ) = T. The other identities follow in the same manner.
vi) If g(x) = −ixj f (x), and g ∈ L1 (Rn ), then Ff is differentiable at ξ and ∂j (Ff )(ξ) =
Fg(ξ).
35
Functional Analysis II
Proof. The proof is done by direct substitutions to the Definition, the last statement
follows by Lebesgue dominated convergence theorem.
Let us mention, that the properties of F and F are very similar. Nevertheless, the
identities above are understood in pointwise sense, the properties of F were proven in the
distributional sense.
We show that F = F on L1 (Rn ) ֒→ S ′ (Rn ) (and we shall after that use only the letter
F for any of the Fourier transforms).
Theorem 14.3.3. Let f ∈ L1 (Rn ). Then Ff is a regular distribution and Ff = Ff in
the distributional sense.
Proof. If f ∈ L1 (Rn ), then Ff is obviously bounded, cf.
Z
|Ff (ξ)| =
1 f (x)e−ihx,ξi
dx ≤ kf k1 , ξ ∈ Rn .
(2π) n/2 (2π)n/2
Rn
i.e. (Ff )(ϕ) = (Ff )(ϕ) for all ϕ ∈ S (Rn ) are equal in the distributional sense.
Theorem 14.3.4. Let f ∈ L1 (Rn ). Then Ff ∈ C0 (Rn ), i.e. Ff ∈ C(Rn ) and lim Ff (x) =
|x|→∞
0.
Proof. The boundedness of Ff was already discussed above. The continuity follows from
Lebesgue convergence theorem (with |f (x)| as an integrable majorant):
Z Z
1 −ihξ+h,xi 1
f (x)e dx → f (x)e−ihξ,xi dx as h → 0.
(2π)n/2 Rn (2π)n/2 Rn
Finally, if n = 1 and f = χ[a,b] with −∞ < a < b < ∞, we get
Z b
1 −iξt 1 e−iξa − e−iξb
Fχ[a,b] (ξ) = e dt = →0
(2π)1/2 a (2π)1/2 iξ
Q
if ξ → ±∞. In the same way, if n > 1 and g(x) = nj=1 χ[aj ,bj ] (xj ) on Rn , then
Yn
1 e−iξj aj − e−iξj bj
Fg(ξ) = →0
(2π)n/2 j=1 iξj
if |ξ| → ∞. The same conclusion therefore holds also for finite sums of step functions of
intervals. Finally, we use that every function f ∈ L1 (Rn ) might be approximated by such
a finite sum function h to arbitrary precision in the L1 (Rn )-norm. We obtain
kf − hk1
|Ff (ξ)| ≤ |F(f − h)(ξ)| + |Fh(ξ)| ≤ + |Fh(ξ)|.
(2π)n/2
36
14 Distributions and Fourier transform
As the last summand goes to zero and the first can be made arbitrary small, we obtain
the conclusion.
i) The inversion theorem is proven for f ∈ L1 (Rn ) and Ff ∈ L1 (Rn ), and holds
pointwise a.e.
iii) One shows that for f ∈ L1 (Rn ) ∩ L2 (Rn ), the following Plancherel identity holds:
kFf k2 = kf k2 .
Proof. We know from Lemma 14.2.4 that if Th (ϕ) = 0 for h ∈ L1 (Rn ) and all ϕ ∈ S (Rn ),
then h = 0 a.e. This can be easily generalised. Let u ∈ L1 (R n n
−|x| 2 /2 n
R ) and v ∈ L∞ (R ).
Furthermore, let ψ(x) =Re . Then (u + v)ψ ∈ L1 (R ) and Rn (u + v)ϕ = 0 for all
ϕ ∈ S (R ) implies also Rn [(u + v)ψ]ϕ = 0 for all ϕ ∈ S (Rn ). Hence (u + v)ψ = 0 a.e.
n
for f ∈ L1 (Rn ) and ϕ ∈ S (Rn ) (actually, even ϕ ∈ L1 (Rn ) can be allowed). The same
formula holds also for F −1 . Finally, we obtain
Z Z Z Z
−1 −1 −1
[F (Ff )]ϕ = (Ff )(F ϕ) = f (F (Fϕ)) = f ϕ.
kf k2 = kFf k2 .
Proof. For f ∈ L1 (Rn ) ∩ L2 (Rn ), let h = f ∗ f¯˜. Then we have the following
i) h ∈ L1 (Rn );
37
Functional Analysis II
= f (y)[f (y − x) − f (y)]dy.
Rn
The first integral tends to zero (using again the compact support of f1 and its uniform
continuity), the second is bounded by 2tkf k2 < ∞.
iii) Fh = (2π)n/2 |Ff |2 ≥ 0;
iv) Finally, we obtain
Z Z
2 |ξ|2 /2
kFf k22 = (2π) −n/2
(Fh)(ξ)dξ = lim (2π) −n/2
(Fh)(ξ)e−ε dξ
Rn ε→0 Rn
Z
2 2
= lim ε−n (2π)−n/2 h(x)e−|x| /(2ε ) dx
ε→0 R n
Z
= h(0) = f (x)f^(−x)dx = kf k22 ,
Rn
where we used the continuity of h at zero and
Z
−n −|x|2 /(2ε2 )
lim ε (2π) −n/2
h(x)e dx − h(0)
ε→0 Rn
Z
−n −n/2 2 2
= lim ε (2π) |h(x) − h(0)|e−|x| /(2ε ) dx
ε→0
ZR
n
2 2
≤ lim ε−n (2π)−n/2 |h(x) − h(0)|e−|x| /(2ε ) dx
ε→0 x:|x|≥t
Z
−n −n/2 2 2
+ lim ε (2π) |h(x) − h(0)|e−|x| /(2ε ) dx.
ε→0 x:|x|≤t
The first integral goes (for every t > 0) to zero11 , the second can be made arbitrary
small by choosing t > 0 small.
38
14 Distributions and Fourier transform
Z X N
1
= cj ck e−ihxj −xk ,ξi dν(ξ)
(2π)n/2 Rn j,k=1
Z X N
1
= cj e−ihxj ,ξi ck e−ihxk ,ξi dν(ξ)
(2π)n/2 Rn j,k=1
2
Z X N
1
−ihxj ,ξi
=
(2π)n/2 Rn j=1
cj e dν(ξ) ≥ 0.
39
Functional Analysis II
One may now conjecture, that every closed translation invariant subspace is obtained
exactly in this manner. So, to every closed translation invariant subspace M we have to
construct a set E ⊂ Rn , such that f ∈ M if, and only if, Ff (x) = 0 a.e. on E. The
obvious construction
\ \
E= Ef = {x ∈ Rn : Ff (x) = 0}
f ∈M f ∈M
runs into serious difficulties, when we realize that each Ef is defined only up to set of
measure zero and that there are uncountably many f ∈ M . Hence, we lose every control
about E.
So, let M be a closed translation invariant subspace of L2 (Rn ) and let M̂ be its image
under Fourier transform. Let P denote the orthogonal projection onto M̂ . Hence, to each
f ∈ L2 (Rn ) there is P f ∈ M̂ , such that f − P f ⊥ M̂ , i.e.
(f − P f ) ⊥ P g, f, g ∈ L2 (Rn )
and also
(f − P f ) ⊥ P g(x)e−ihα,xi , f, g ∈ L2 (Rn ), α ∈ Rn .
This is equivalent to
Z
(f − P f )(x)P g(x)e−ihα,xi dx = 0, f, g ∈ L2 (Rn ), α ∈ Rn ,
Rn
f · P g = (P f ) · (P g), f, g ∈ L2 (Rn ).
f · P g = P f · g, f, g ∈ L2 (Rn ).
This may be (roughly speaking) interpreted as that (P f )/f is constant for all f ∈ L2 (Rn ).
To avoid the devision by zero here, we consider some strictly positive function in L2 (Rn ),
2
for example g(x) = e−|x| will do, and put
(P g)(x)
ϕ(x) := , x ∈ Rn .
g(x)
ϕ2 · g = ϕ · P g = P 2 g = P g = ϕ · g,
40
14 Distributions and Fourier transform
41
Functional Analysis II
#(2M ) ≤ 2#(M ) + 8M
aN
0 (F ) = 1/2(F (1) + F (−1)), aN
1 (F ) = 1/2(F (1) − F (−1)).
42
14 Distributions and Fourier transform
We shall use the first formula as the definition and afterwards, we show that it coincides
with the second one.
43
Functional Analysis II
is equal to Z
T (τy ϕ̃(·))ψ(y)dy,
Rn
i.e. that we may interchange the integration and the application of T . To do this, it is
enough to prove the convergence of the Riemann sums of the integral under discussion in
S (Rn ). We sketch the arguments but leave out the details.
For each N ∈ N, we partition [−N, N ]n into (2N 2 )n cubes Qm , m = 1, . . . , (2N 2 )n
with centers ym . One has to show that the functions
1 X
x→ (τym ϕ̃)(x)ψ(ym )
Nn m
converge in S (Rn ) to Z
x→ (τy ϕ̃)(x)ψ(y)dy
Rn
as N → ∞. Although not difficult, we leave out the technical details.
Definition 14.5.4. Let T ∈ S ′ (Rn ). Then the support of T is the intersection of all
closed sets K ⊂ Rn , such that T (ϕ) = 0 for all ϕ ∈ S (Rn ) with supp ϕ ⊂ Rn \ K.
The support of the distribution δx0 is exactly the set {x0 }.
Lemma 14.5.5. Let T ∈ S ′ (Rn ). Then T has compact support if, and only if, there
exists a constant C > 0 and l, N ∈ N such that
X
|T (ϕ)| ≤ C sup |D α ϕ(x)|, ϕ ∈ S (Rn ).
|x|≤N |α|≤l
44
14 Distributions and Fourier transform
Proof. Let the estimate be satisfied and let ψ ∈ S (Rn ) with supp ψ ⊂ Rn \ {x ∈ Rn :
|x| ≤ N } = {x ∈ Rn : |x| > N }. Then obviously |T (ψ)| = 0 and we conclude that T has
compact support.
Let on the other hand T have compact support, i.e. supp T ⊂ {x :∈ Rn : |x| ≤ M }.
We consider an infinitely differentiable function ψ with ψ(x) = 1 for |x| ≤ M and ψ(x) = 0
for |x| ≥ M + 1. Using Theorem 14.2.3, we get
i) T ∗ ϕ is a C ∞ -function,
iii) For all multiindices α ∈ Nn0 there exist constants Cα , kα > 0, such that
τhej ψ − ψ
−h
τx+hej ϕ̃ − τx ϕ̃ S
→ (−1) · D ej (τx ϕ̃) = (−1) · τx (D ej ϕ̃)
h
and, after applying T ,
(T ∗ ϕ)(x + hej ) − (T ∗ ϕ)(x) τx+hej ϕ̃ − τx ϕ̃
=T → T (−τx (D ej ϕ̃)).
h h
We combine this with T ∗ (D ej ϕ)(x) = T (τx (D ^ ej ϕ)) = T (−τ (D ej ϕ̃)) = T (−D ej (τ ϕ̃)) =
x x
e
(D T ) ∗ ϕ(x) and finish the proof of (ii) for α = ej . Iterating this identity, we obtain that
j
45
Functional Analysis II
Finally, to show that T ∗ ϕ ∈ S (Rn ) for T with compact support, we use Lemma
14.5.5. We get for arbitrary k ∈ N
X
|(T ∗ ϕ)(x)| = |T (ϕ(x − ·))| ≤ C sup |Dyα ϕ(x − y)|
|y|≤N |α|≤l
for |x| ≥ 2N. Hence (T ∗ ϕ)(x) decays faster then any polynomial. The same argument
applied to D α (T ∗ ϕ) = T ∗ (D α ϕ) yields the same for all the derivatives of T ∗ ϕ. Hence
T ∗ ϕ ∈ S (Rn ).
Finally, we show that products and convolutions of distributions behave under Fourier
transform exactly as we expect them to do.
Theorem 14.5.7. Let T ∈ S ′ (Rn ) and ϕ ∈ S (Rn ). Then
F(ϕT ) = (2π)−n/2 F(ϕ) ∗ F(T ),
F(ϕ ∗ T ) = (2π)n/2 F(ϕ)F(T ).
Proof. We obtain for every ψ ∈ S (Rn )
F(ϕT )(ψ) = (ϕT )(Fψ) = T (ϕF(ψ))
and
g ∗ ψ) = T (F(Fϕ
F(ϕ) ∗ F(T )(ψ) = FT (Fϕ g ∗ ψ)) = (2π)n/2 T (ϕF(ψ)).
46
14 Distributions and Fourier transform
whenever the right-hand side makes sense. As e−ihx,·i is not in S (Rn ), some care is
of course necessary. If T has compact support, then T (e−ihx,·i ) might be defined as
T (e−ihx,·i ϕ(·)) for any ϕ ∈ S (Rn ) with ϕ = 1 on supp T. Obviously, this “definition” does
not depend on the choice of ϕ, i.e. T (e−ihx,·i ϕ(·)) = T (e−ihx,·i ψ(·)) if both ϕ, ψ ∈ S (Rn )
with ϕ = ψ = 1 on supp T.
First, using ψ ∈ S (Rn ) with ψ(x) = 1 for x ∈ supp T , we get immediately that
FT = F(ψT ) = (2π)−n/2 Fψ ∗ FT
is a C ∞ -function (cf. Theorem 14.5.6). Furthermore, with Φ ∈ S (Rn ) with FΦ = ψ, we
get
(2π)n/2 FT (ξ) = (FT ∗ Fψ)(ξ) = (FT ∗ Φ̃)(ξ) = FT (τξ Φ)
= T (F(τξ Φ)) = T (e−ih·,ξi ψ),
which we defined as T (e−ih·,ξi ). Therefore, also (14.10) is justified.
Let us observe that (14.10) makes sense also for z ∈ Cn instead of ξ ∈ Rn . This
mapping (which is an extension of Fourier transform to complex arguments) is usually
called Fourier-Laplace transform and is defined by
!
e−ih·,zi
(FT )(z) := T , z ∈ Cn . (14.11)
(2π)n/2
Note, that we denote it by F again.
Its properties are given in the following version of (Schwartz-)Paley-Wiener’s theorem.
Theorem 14.6.2. Let T ∈ S ′ (Rn ) have compact support. Then its Fourier-Laplace
transform is an entire function with polynomial growth on Rn .
Proof. First, we observe that
!
ej FT (z + hej ) − FT (z) 1 e−ihx,z+hej i − e−ihx,zi
D (FT )(z) = lim = lim T .
h→0 h (2π)n/2 h→0 h
The fact, that FT is an entire function, now follows from the compact support of T and
the convergence of the sequence
e−ihx,z+hej i − e−ihx,zi
x→ , x ∈ Rn ,
h
to −ixj e−ihx,zi in C ∞ (Rn ). This gives D ej (FT )(z) = T (−ixj e−ihx,zi ). This argument
might be iterated and we obtain D α (FT )(z) = T ((−ixj )α e−ihx,zi ) for all α ∈ Nn0 . Moreover,
the polynomial growth on Rn follows from (14.9).
47
Functional Analysis II
Proof. We may assume, that x0 = 0. We shall use Lemma 14.5.5. We assume that
X
|T (ϕ)| ≤ C sup |D α ϕ(x)|
|x|≤N |α|≤l
and it is rather easy to see (using (14.12)), that the last expression tends to zero as ε → 0.
Now, let ϕ ∈ S (Rn ) be arbitrary and let η ∈ S (Rn ) be a function that is equal to 1
in a neighborhood of origin. Then (using Taylor’s polynomial)
X D α ϕ(0)
ϕ(x) = η(x) xα + ψ(x) + (1 − η(x))ϕ(x),
α!
|α|≤l
Corollary 14.7.2. Let T ∈ S ′ (Rn ) with FT supported at the singleton {ξ0 }. Then T is a
linear combination of functions ξ α eihξ,ξ0 i . In particular, if FT is supported at the origin,
then T is a polynomial.
48
14 Distributions and Fourier transform
Proof. Taking the Fourier transform, we observe that |ξ|2 ·FT = 0. Hence, FT is supported
at the origin and, therefore, must be a polynomial.
Liouville’s classical theorem says that every bounded harmonic function must be con-
stant. We observe that it follows directly from Corollary 14.7.3.
A distribution T ∈ S ′ (Rn ) is called a fundamental solution of a partial differential
operator with constant coefficients L if L(T ) = δ0 .
2π n/2
∆(|x|2−n ) = −(n − 2)ωn−1 δ0 = −(n − 2) δ0
Γ(n/2)
and for n = 2,
∆(log |x|) = 2πδ0 .
Here, ωn−1 stands for the (n − 1)-dimensional measure of the unit sphere in Rn .
where Ω is a domain in Rn with smooth boundary and ∂/∂ν denotes the derivative with
respect to the outer normal vector to ∂Ω.
Let Ωε = {x ∈ Rn : ε < |x| < 1/ε}. Furthermore, we choose (for n ≥ 3) v(x) = |x|2−n
and u = ϕ ∈ S (Rn ). This leads to
Z Z
2−n
∆(|x| )(ϕ) = v(∆ϕ) = v(x)∆ϕ(x)dx = lim |x|2−n ∆ϕ(x)dx
Rn ε→0 Ωε
and
Z Z Z
2−n 2−n ∂|x|2−n ∂ϕ(x)
|x| ∆ϕ(x)dx = ∆(|x| )ϕ(x)dx − ϕ(x) − |x|2−n ds.
Ωε Ωε ∂Ωε ∂ν ∂ν
as ε → 0. This is due to
Z
∂|x|2−n 2−n ∂ϕ(x)
ϕ(x) − |x| ds
{x∈Rn :|x|=1/ε} ∂ν ∂ν
Z
n−1 n−2 ∂ϕ(x)
= ϕ(x)(2 − n)ε −ε ds
{x∈Rn :|x|=1/ε} ∂ν
and the rapid decay of ϕ ∈ S (Rn ) and all its derivatives at infinity.
49
Functional Analysis II
Finally, the first integral tends to (2 − n)ϕ(0)ωn−1 , where ωn−1 is the (n − 1)-dimensional
measure of the unit sphere in Rn , and the second integral tends to zero.
Let us note, that the formula ωn−1 = 2π n/2 /Γ(n/2) follows easily by the following
trick. Using polar coordinates in Rn and the substitution t := r 2 , we get
Z Z ∞ Z
√ 2 2 ωn−1 ∞ −t n/2−1 ωn−1
( π)n = e−|x| dx = ωn−1 e−r r n−1 dr = e t dt = Γ(n/2).
Rn 0 2 0 2
The proof for n = 2 is very similar.
Remark 14.7.5. Due to the previous Corollary, the fundamental solution of Laplacian is
not unique. We may add to any of the fundamental solutions above a harmonic polynomial
and the Laplace operator of such a distribution would not change.
Although one could develop a full theory of fundamental solutions in the spirit of
the book of Rudin (with Theorem of Malgrange-Ehrenpreis ensuring the existence of a
fundamental solution for any partial differential operator with constant coefficients as the
main highlight), we shall give a couple of examples of this technique instead.
Let us assume that L(T ) = δ0 , i.e. that T is a fundamental solution to L. Then the
solution of the equation L(u) = f can be obtained by convolution.
L(T ∗ f ) = (LT ) ∗ f = δ0 ∗ f = f.
i.e.
where now v(ξ, t) = Fx (u(x, t))(ξ). We solve these equations for each ξ ∈ R separately. We
fix ξ ∈ R and put z(t) = v(ξ, t) and obtain z ′ (t) + kξ 2 z(t) = 0 and z(0) = (2π)−1/2 . This
50
14 Distributions and Fourier transform
2 2 2
leads to z(t) = (2π)−1/2 e−kξ t and v(ξ,√t) = (2π)−1/2 e−kξ t . If we denote h(s) = e−s /2 , i.e.
Fh = h, we get v(ξ, t) = (2π)−1/2 h( 2ktξ) and finally the fundamental solution √ to the
one-dimensional heat equation Φ(x, t) = (Fξ−1 v(ξ, t))(x) = (2π)−1/2 (2kt)−1/2 h(x/ 2kt) =
1 x2
√ exp − .
4πkt 4kt
We now claim that the solution to the heat equation with a general right hand side g
is then given as the convolution of Φ with g in the x variable, i.e.
Z
u(x, t) = Φ(x − y, t)g(y)dy, −∞ < x < ∞, 0 < t < ∞.
R
and Z Z
u(x, 0) = Φ(x − y, 0)g(y)dy = δ0 (x − y)g(y)dy = g(x).
R R
Finally, let us mention that the fundamental solution to the n-dimensional heat equa-
tion is obtained simply as a tensor product of n one-dimensional fundamental solutions,
i.e.
1 |x|2
Φ(x, t) = exp − , x ∈ Rn , 0 < t < ∞.
(4πkt)n/2 4kt
ii) A sequence (ϕj )j∈N ⊂ D(Ω) is said to converge to ϕ ∈ D(Ω) in D(Ω), if there is a
compact set K ⊂ Ω, such that
and
D α ϕj ⇉ D α ϕ for all α ∈ Nn0 .
D
We write ϕj → ϕ to denote the convergence in D(Ω).
iii) D ′ (Ω) is the collection of all complex-valued linear continuous functionals over D(Ω),
i.e. T : D(Ω) → C belongs to D ′ (Ω) if, and only if,
51
Functional Analysis II
v) Finally, we also equip the space D ′ (Ω) with the notion of convergence. Namely, we
D′
say that Tj converges to T in D ′ (Ω), Tj ⇀ T , if
The spaces D(Ω) are much more flexible due to the inclusion of the domain into their
definition. Unfortunately, a lot of the algebraic structure is lost, for example Fourier
transforms and convolutions do not have an easy counterpart on D(Ω). Also modulations
and translations are only of a limited use on Ω.
Nevertheless, one can define the derivative of a distribution T ∈ D ′ (Ω) by (D α T )(ϕ) =
(−1)|α| T (D α ϕ) and its product with a smooth function ψ ∈ C ∞ (Ω) as (ψT )(ϕ) = T (ψϕ).
If Ω = Rn , we obtain two spaces of test function, and two spaces of distributions. Obvi-
ously, D(Rn ) ⊂ S (Rn ) (including a continuous embedding) and, consequently, S ′ (Rn ) ⊂
D ′ (Rn ), again also in the sense of continuous embedding.
The space D(Ω) allows very well to explain what is meant by weak solutions to PDE’s.
Let us for example consider the equation
∂u(t, x) ∂u(t, x)
+ = 0.
∂t ∂x
In the classical setting this means that we are looking for a function u ∈ C 1 (R2 ), such
that the equation holds pointwise. In the distributional sense, we require that
∂u(t, x) ∂u(t, x)
+ (ϕ) = 0
∂t ∂x
Whenever u is smooth enough, these two notions coincide. But there are functions (i.e.
u(t, x) = |t − x|), which do satisfy the weak formulation, but can not satisfy the strong
formulation due to their lack of differentiability.
52
15 Introduction to harmonic analysis
15.1 Approximation of identity
Theorem 15.1.1. Let Ω ⊂ Rn be a domain. The set of continuous functions with compact
support contained in Ω is dense in Lp (Ω), 1 ≤ p < ∞.
ii) The space of step functions, i.e. span{χA : A ⊂ Ω, A measurable}, is dense in Lp (Ω)
for every 1 ≤ p < ∞.
S
We first consider open sets Ωj ⊂ Ω, j ∈ N, such that Ωj ⊂ Ωj ⊂ Ωj+1 ⊂ Ω and ∞ j=1 Ωj =
13
Ω. Let us take f ∈ Lp (Ω). Then f χΩj → f in Lp (Ω) and we may restrict ourselves
to f ∈ Lp (Ω) with compact support in Ω. Due to the second property of the Lebesgue
P
measure, this function may be approximated by a step function K k=1 ̺k χAk with Ak ⊂
supp f. So, it is enough to approximate characteristic functions χB with B compact in Ω.
Using the first property of the Lebesgue measure, we may restrict ourselves to bounded
open sets G ⊂ G ⊂ Ω. Then the sequence of functions x → max(0, 1 − k dist(x, G)) gives
the desired approximation.
Proof. If f is continuous with compact support, then the result follows by uniform con-
tinuity of f and the Lebesgue dominated convergence theorem. If f ∈ Lp (Rn ), we may
find for every t > 0 a continuous function g with compact support such that kf − gkp < t.
Then
Theorem 15.1.3. The family of functions (Kε )ε>0 ⊂ L1 (Rn ) is called the approximation
of identity, if
R
(K1) Rn |Kε (x)|dx ≤ C < ∞ for all ε > 0,
R
(K2) Rn Kε (x)dx = 1 for all ε > 0,
R
(K3) limε→0+ |x|>δ |Kε (x)|dx = 0 for all δ > 0.
Then
R
i) If K ∈ L1 (Rn ) with Rn K(x)dx = 1, then Kε (x) = ε−n K(x/ε) is an approximation
of identity.
lim kKε ∗ f − f kp = 0
ε→0+
53
Functional Analysis II
R
Proof. (i) Let K ∈ L1 (Rn ) with Rn K(x)dx = 1. Then we get immediately
Z Z Z Z
−n −n
ε K(x/ε)dx = K(x)dx = 1 and ε |K(x/ε)|dx = |K(x)|dx = kKk1 < ∞.
Rn Rn Rn Rn
p/p′
=C |Kε (y)| · kf (· − y) − f (·)kpp dy
Rn
(Z Z )
p/p′
≤C |Kε (y)| · kf (· − y) − f (·)kpp dy p
+2 kf kpp |Kε (y)|dy
|y|≤δ |y|>δ
for every δ > 0. Using (K3) and previous Lemma, we obtain the conclusion of the
theorem.
Definition 15.1.4. Let Ω ⊂ Rn be a domain. Then Cc∞ (Ω) denotes the set of infinitely-
differentiable functions compactly supported in Ω.
Theorem 15.1.5. Cc∞ (Ω) is dense in Lp (Ω) for every 1 ≤ p < ∞ and every domain
Ω ⊂ Rn .
(the support property is clear, the differentiability was proven in a much more general
setting in Theorem 14.5.6). Together with the formula kωε ∗ g − gkp → 0, the conclusion
follows.
54
15 Introduction to harmonic analysis
Here stands |B(x, r)| for the Lebesgue measure of B(x, r).
Let us note, that maximal operator is not linear, but is sub-linear, i.e.
Z
1
M (f + g)(x) = sup |f (y) + g(y)|dy
r>0 |B(x, r)| B(x,r)
Z Z !
1
≤ sup |f (y)|dy + |g(y)|dy
r>0 |B(x, r)| B(x,r) B(x,r)
≤ (M f )(x) + (M g)(x).
We shall study the mapping properties of the operator M in the frame of Lebesgue
spaces Lp (Rn ), 1 ≤ p ≤ ∞. If f ∈ L∞ (Rn ), then
Z Z
1 kf k∞
|f (y)|dy ≤ 1dy = kf k∞
|B(x, r)| B(x,r) |B(x, r)| B(x,r)
holds for every x ∈ Rn and every r > 0 and kM f k∞ ≤ kf k∞ follows. To deal with other
p’s, we need some more notation first.
Let f ∈ L1 (Rn ) and let α > 0. Then
Z
n
α · |{x ∈ R : |f (x)| > α}| = αdy
{x∈Rn :|f (x)|>α}
Z Z
≤ |f (y)|dy ≤ |f (y)|dy = kf k1 .
{x∈Rn :|f (x)|>α} Rn
is denoted by L1,w (Rn ) and called weak L1 . We have just shown that L1 (Rn ) ֒→ L1,w (Rn ).
Let us mention that k · k1,w is not a norm, but it still satisfies
1
Finally, we observe, that the function x → ∈ L1,w (Rn ) \ L1 (Rn ).
kxkn
The main aim of this section is to prove the following theorem.
55
Functional Analysis II
kM f k1,w ≤ Akf k1 ,
where A is a constant which depends only on the dimension (i.e. A = 5n will do).
kM f kp ≤ Ap kf kp ,
Here C is a positive constant that depends only on the dimension n; C = 5−n will do.
Proof. We describe first the choice of B1 , B2 , . . . . We choose B1 so that it is essentially
as large as possible, i.e.
1
diam(B1 ) ≥ sup diam(B j ).
2 j
The choice of B1 is not unique, but that shall not hurt us.
If B1 , B2 , . . . , Bk were already chosen, we take again Bk+1 disjoint with B1 , . . . , Bk and
again nearly as large as possible, i.e.
1
diam(Bk+1 ) > sup{diam(B j ) : B j disjoint with B1 , . . . , Bk }.
2
In this way, we get a sequence B1 , B2 , . . . , Bk , . . . of balls. It can be also finite, if there
were no j
P balls B disjoint with B1 , B2 , . . . , Bk . P
If k |Bk | = ∞, then the conclusion of lemma is satisfied and we are done. If k |Bk | <
∞, we argue as follows.
We denote by Bk∗ the ball with the same center as Bk and diameter five times as large.
We claim that [
Bk∗ ⊃ E,
k
P P
which then immediatelySgives that |E| ≤ k |Bk∗ | = 5n k |Bk |.
We shall show that k Bk∗ ⊃ B j for every j. This is clear if B j is one P of the balls in
the preselected sequence. If it is not the case, we obtain diam(Bk ) → 0 (as k |Bk | < ∞)
and we choose the first k with diam(Bk+1 ) < 12 diam(B j ). That means, that B j must
intersect one of the balls B1 , . . . , Bk , say Bk0 . Obvious geometric arguments (based on
the inequality diam(Bk0 ) ≥ 1/2 · diam(B j )) then give that B j ⊂ Bk∗0 . This finishes the
proof.
Proof. (of Theorem 15.2.1). Let α > 0 and let us consider Eα := {x ∈ Rn : M f (x) > α}.
From the definition of M we obtain, that for every x ∈ Eα , there is a ball Bx centered at
x such that Z
|f (y)|dy > α|Bx |.
Bx
56
15 Introduction to harmonic analysis
Hence, we get |Bx | < kf k1 /α and ∪x∈Eα Bx ⊃ Eα and the balls (Bx )x∈Eα satisfy the
assumptions of Lemma 15.2.2. Using this covering lemma, we get a sequence of disjoint
balls (Bk )k , such that
X
|Bk | ≥ C|Eα |.
k
We therefore obtain
Z XZ X
kf k1 ≥ |f (y)|dy = |f | > α |Bk | ≥ αC|Eα |,
∪k Bk k Bk k
Actually, the Theorem of Marcinkiewicz says, that every operator with these two proper-
ties is then automaticaly bounded on all Lp (Rn ), 1 < p < ∞. Furthermore, this is a corner
stone of the so-called interpolation theory.
We shall restrict ourself to the necessary minimum needed to prove the theorem. This
is based on the technique of splitting the function into good and bad part, idea elaborated
in detail by Calderón and Zygmund.
Let α > 0 and put f1 (x) := f (x) if |f (x)| > α/2 and f1 (x) := 0 otherwise. Due to
|f (x)| ≤ |f1 (x)| + α/2, we have also M f (x) ≤ M f1 (x) + α/2 and also
2A
|Eα | = |{x ∈ Rn : M f (x) > α}| ≤ |{x ∈ Rn : M f1 (x) > α/2}| ≤ kf1 k1
Z α
2A
≤ |f (y)|dy.
α {x∈Rn :|f (x)|>α/2}
We use the information of the size of the level sets of M f to estimate the Lp -norm of M f .
Z Z ∞
kM f kpp = M f (x)p dx = |{x ∈ Rn : M f (x)p > α}|dα
Rn 0
Z ∞
= |{x ∈ R : M f (x) > α1/p }|dα
n
0
Z ∞
=p β p−1 |{x ∈ Rn : M f (x) > β}|dβ
0
Z Z !
∞
2A
≤p β p−1 |f (y)|dy dβ
0 β {x∈Rn :|f (x)|>β/2}
Z Z 2|f (y)| Z
p−2 2Ap
= 2Ap |f (y)| β dβdy = |f (y)| · |2f (y)|p−1 dy
R n 0 p − 1 Rn
Z
2p Ap
= |f (y)|p dy,
p − 1 Rn
57
Functional Analysis II
where we used Fubini’s theorem and the substitution β := α1/p with dα = pβ p−1 dβ. This
gives the first and the third statement of the theorem with
1/p
5n p
Ap = 2 , 1 < p < ∞.
p−1
by
∞
( Z )
[ 1 1
n
x ∈ R : lim sup f (y)dy − f (x) >
r→0 |B(x, r)| B(x,r) k
k=1
united with
∞
( Z )
[ 1 1
n
x ∈ R : lim inf f (y)dy − f (x) < − .
r→0 |B(x, r)| B(x,r) k
k=1
It is therefore enough to show, that each of these sets has measure zero. Let us fix k ∈ N
and decompose f = g + h, where g ∈ C(Rn ) and khk1 ≤ t, t > 0.
Obviously, Z
1
lim g(y)dy = g(x), x ∈ Rn ,
r→0 |B(x, r)| B(x,r)
which implies
( Z )
n 1 1
x ∈ R : lim sup f (y)dy − f (x) >
r→0 |B(x, r)|B(x,r) k
( Z )
1 1
= x ∈ Rn : lim sup h(y)dy − h(x) >
r→0 |B(x, r)| B(x,r) k
1 1
⊂ x ∈ Rn : M h(x) > ∪ x ∈ Rn : |h(x)| > .
2k 2k
The measure of the first set is smaller than 2kAkhk1 and the measure of the second is
smaller than 2kkhk1 . As khk1 might be chosen arbitrary small, the measure of the original
set is zero. The same argument works for lim inf instead of lim sup as well.
58
15 Introduction to harmonic analysis
Theorem 15.2.4. Let ϕ be a function which is non-negative, radial, decreasing (as func-
tion on (0, ∞)) and integrable. Then
Proof. We denote by Qk the collection of dyadic cubes with side length 2−k , k ∈ Z.
Furthermore, we define
X 1 Z
Ek f (x) = f χQ (x).
|Q| Q
Q∈Qk
We define also
That is, x ∈ Ωk if k is the first index with Ek f (x) > λ. As the integrability of f implies
Ek f (x) → 0 for k → −∞, such k always exists. The sets Ωk are clearly disjoint and each
can be written as the union of cubes in Qk . Together, these cubes form the system (Qj ).
14
possibly finite, or even empty
59
Functional Analysis II
This gives the third statement of the theorem. The first follows by Lebesgue differen-
tiation theorem: indeed, Ek f (x) ≤ λ for all k ∈ Z implies f (x) ≤ λ at almost every such
point. The second follows just by
[ X Z
1X 1
Qj = |Qj | ≤ f ≤ kf k1 .
λ Qj λ
j j j
We shall present two basic interpolation theorems, the Riesz-Thorin interpolation theorem
and Marcinkiewicz interpolation theorem.
The operator T mapping measurable functions to measurable functions is called sub-
linear, if
Let (X, µ) and (Y, ν) be measure spaces and let T be a sub-linear operator mapping
Lp (X, µ) into a space of measurable functions on (Y, ν). We say that T is strong type
(p, q) if it is bounded from Lp (X, µ) into Lq (Y, ν). We say, that it is of weak type (p, q),
q < ∞, if
kT f kq,w := sup λ · ν 1/q ({y ∈ Y : |T f (y)| > λ}) ≤ Ckf kp , f ∈ Lp (X, µ).
λ>0
Proof. Let λ > 0 be given and let f ∈ Lp (X, µ). Then we decompose f into f = f0 + f1
with
f0 = f χ{x:|f (x)|>λ} ,
f1 = f χ{x:|f (x)|≤λ} .
The case p1 = ∞ appeared implicitly already in the proof of the boundedness of Hardy-
Littlewood maximal operator, so we suppose that p1 < ∞. Then we have
ν({y ∈ Y : |T f (y)| > λ}) ≤ ν({y ∈ Y : |T f0 (y)| > λ/2}) + ν({y ∈ Y : |T f1 (y)| > λ/2})
and
p i
2Ai
ν({y ∈ Y : |T fi (y)| > λ/2}) ≤ kfi kpi , i = 0, 1.
λ
60
15 Introduction to harmonic analysis
and
kT f kq1 ≤ M1 kf kp1 for f ∈ Lp1 ,
then
kT f kq ≤ M01−θ M1θ kf kp for f ∈ Lp .
With this definition, Hf (x) makes sense for all smooth functions, especially for f ∈ S (R).
15
15
Observe, that the restriction to n = 1 is both natural and essential for the cancelation property.
Furthermore, from now on, we shall denote the Fourier transform of a function f also by the more usual
fb.
61
Functional Analysis II
1 1
then Hf := p.v. ∗ f. This formula suggests that we look for the Fourier transform of
π x
1
Hf. Therefore, we regularize 1/x. This can be done using complex distributions π(x±iε)
and letting ε → 0 or (and that is what we shall do) by defining
1 x
Qt (x) := · .
π t 2 + x2
1
Obviously, limt→0 Qt (x) = πx holds pointwise and, as we shall show below, also in S ′ (R).
As
Z ∞
1
F −1 (sgn(x)e−a|x| )(ξ) = √ sgn(x)e−a|x| · eixξ dx
2π −∞
Z 0 Z ∞
1 x(a+iξ) x(−a+iξ)
=√ − e dx + e dx
2π −∞ 0
r
1 −1 1 2 iξ
=√ + = · 2 ,
2π a + iξ a − iξ π a + ξ2
we obtain
b t (ξ) = √−i sgn(ξ)e−t|ξ| ,
Q
2π
√
we have also limt→0 Q bt (ξ) = −i sgn(ξ)/ 2π. As this convergence is uniform on compact
sets, it holds also in S ′ (R). Finally, due to the continuity of Fourier transform, we obtain
1 1 ∧ bt = −i√sgn(·) .
p.v. = [lim Qt ]∧ = lim Q
π x t→0 t→0 2π
Theorem 15.5.1. In S ′ (R),
1 1
lim Qt (x) = p.v. .
t→0 π x
Proof. For each ε > 0, the functions ψε (x) = x−1 χ{y:|y|>ε} (x) are bounded and define
tempered distributions with limε→0+ ψε = p.v. x1 . Therefore, it is enough to show that
lim (πQt − ψt ) = 0
t→0
for ϕ ∈ S (R). As t → 0, we apply Lebesgue dominated convergence theorem and use the
symmetry of the integrands to conclude, that the limit is zero.
62
15 Introduction to harmonic analysis
S ′ (R) 1 1
Qt −−−−→ π p.v. x
t→0
Fy Fy
S ′ (R) −i√
sgn(·)
√−i sgn(ξ)e−t|ξ| −−−−→ ,
2π t→0 2π
Summarizing, one defines the Hilbert transform Hf for f ∈ S (Rn ) by any of these
formulas:
1 1
Hf = p.v. ∗ f,
π x
Hf = lim Qt ∗ f,
t→0
(Hf ) (ξ) = −i sgn(ξ)fˆ(ξ).
∧
Using the third expression, we can extend the definition of H to L2 (R) and it holds
hHf, gi = h(Hf )∧ , ĝi = h−i sgn(·)fˆ, ĝi = hfˆ, sgn(·)ĝi = hfˆ, −(Hg)∧ i = −hf, Hgi
C
kHf k1,w ≤ Ckf k1 , i.e. |{x ∈ R : |Hf (x)| > λ}| ≤ kf k1 , λ > 0, f ∈ L1 (Rn ).
λ
(ii) (M. Riesz) H is of strong type (p, p) for every 1 < p < ∞, i.e.
kHf kp ≤ Cp kf kp .
Proof. Step 1.: We show the weak type (1, 1) by exploiting Theorem 15.3.1. Let f be
non-negative and let λ > 0, then Theorem 15.3.1 gives a sequence of disjoint intervals
(Ij ), such that
[
f (x) ≤ λ for a.e. x 6∈ Ω = Ij ,
j
1
|Ω| ≤ kf k1 ,
Zλ
1
λ< f ≤ 2λ.
|Ij | Ij
63
Functional Analysis II
Given this decomposition of R, we decompose f into “good” and “bad” part defined by
!
f (x),Z x 6∈ Ω, X X Z
1
g(x) = 1 b(x) = bj (x) = f (x) − f χIj (x).
f, x ∈ Ij , |Ij | Ij
|Ij | Ij j j
Then g(x) ≤ 2λ almost everywhere, and bj is supported on Ij and has zero integral. Since
Hf = Hg + Hb, we have
|{x ∈ R : |Hf (x)| > λ}| ≤ |{x ∈ R : |Hg(x)| > λ/2}| + |{x ∈ R : |Hb(x)| > λ/2}|.
XZ
|Hbj (x)|dx ≤ Ckf k1 .
j R\2Ij
Denote the center of Ij by cj and use that bj has zero integral to get
Z Z Z
b (y)
j
|Hbj (x)|dx = dy dx
R\2Ij R\2Ij Ij x − y
Z
Z 1 1
= bj (y) − dy dx
R\2Ij Ij x − y x − cj
Z Z !
|y − cj |
≤ |bj (y)| dx dy
Ij R\2Ij |x − y| · |x − cj |
Z Z !
|Ij |
≤ |bj (y)| 2
dx dy
Ij R\2Ij |x − cj |
The last inequality follows from |y −cj | < |Ij |/2 and |x−y| > |x−cj |/2. The inner integral
equals 2, so
XZ XZ
|Hbj (x)|dx ≤ 2 |bj (y)|dy ≤ 4kf k1 .
j R\2Ij j Ij
Our proof of the weak (1, 1) inequality is for non-negative f , but this is sufficient since
an arbitrary real function can be decomposed into its positive and negative parts, and a
complex function into its real and imaginary parts.
64
15 Introduction to harmonic analysis
Step 2.: Since H is weak type (1, 1) and strong type (2, 2), it is also strong type (p, p)
for 1 < p < 2. If p > 2, we apply duality, i.e.
Z
kHf kp = sup Hf · g : kgkp′ ≤ 1
ZR
= sup f · Hg : kgkp′ ≤ 1
R
≤ kf kp · sup kHgkp′ : kgkp′ ≤ 1 ≤ Cp′ kf kp .
The strong (p, p) inequality is false if p = 1 or p = ∞; this can easily be seen if we let
f = χ[0,1] . Then
1 x
Hf (x) = log ,
π x − 1
and Hf is neither integrable nor bounded.
15.6 BMO
Hilbert transform acts rather badly on L∞ (R). Not only is H unbounded on L∞ (R), it
can not be easily defined on a dense subset of L∞ (R). The definition
Z
1 f (x)
Hf (y) = dx
π R y−x
runs for f ∈ L∞ (R) into troubles for x near y and x near infinity. If we look onto
differences, the situation changes to
Z
′ 1 1 1
Hf (y) − Hf (y ) = f (x) − dx.
π R y − x y′ − x
This improves the situation for x near infinity, as 1/(x − y) − 1/(x − y ′ ) = O(1/x2 ) in this
case.
We say that f and g are equivalent modulo a constant if f (x) = g(x) + C for some
(complex) constant C and almost every x ∈ R. Given f ∈ L∞ (R) and y ∈ R. Then we
take an open interval B ⊂ R with center at zero containing y. Then f χB ∈ L2 (R) and we
define Hf (y) to be
Z
1 1 1
Hf (y) := H(f χB )(y) + f (x) + dx.
π R\B y−x x
The first term is defined by the L2 definition of H, the integral in the second term converges
absolutely. The definition depends on B, but choosing another interval B ′ ⊃ B with the
center at the origin leads to a difference
Z Z
1 1 1 1 1 1
H(f χB )(y) + f (x) + dx − H(f χB ′ )(y) − f (x) + dx
π R\B y−x x π R\B ′ y−x x
Z
1 1 1
= H(f χB − f χB ′ )(y) + f (x) + dx
π B ′ \B y−x x
Z Z
1 1 1 1 f (x)
= −H(f χB ′ \B )(y) + f (x) + dx = dx,
π B ′ \B y−x x π B ′ \B x
which does not depend on y. This defines Hf modulo constant for f ∈ L∞ (R). Of course,
such a definition does not allow to measure Hf in the usual norms, as Lp . Instead of that,
we need a space of functions defined modulo constants.
65
Functional Analysis II
where B ranges over all balls. Note that if one shifts f by a constant, the BM O norm
is unchanged, so this norm is well-defined for functions defined modulo constants. We
denote by BM O(Rn ) the space of all functions with finite BM O norm.
Example 15.6.2. Let f (x) = sgn(x). Let |y| < a/2. We take B = (−a, a) and apply the
definition of Hf as presented above. This gives (for y > 0)
Z a Z
sgn(x) 1 1
πHf (y) = p.v. dx + sgn(x) + dx
−a y − x (−∞,−a)∪(a,∞) y−x x
! Z
Z −a Z ∞
sgn(x) 1 1 1 1
= lim dx − − dx + − dx
ε→0+ (−a,a)\(y−ε,y+ε) y − x −∞ x x−y a x x−y
| {z } | {z
x=∞ }
a
− ln a+y x a
=ln x−y =− ln a−y
x=a
Z 0 Z y−ε Z a
1 1 1 a a
= lim dx − dx − − ln − ln
−a x − y x−y y+ε x − y a+y a−y
ε→0+
0
y y a−y a a
= lim ln + ln − ln − ln − ln
ε→0+ a+y ε ε a+y a−y
= 2 ln y − 2 ln a.
If y < 0, similar calculation applies as well. Hence, πHf (y) = 2 ln |y| − 2 ln a. Hence,
ignoring the constant,
2
H(sgn x) = ln |x|.
π
Let us observe, that
Z Z Z
1
f − 1
f ≈ inf
1
|f − c|
|B| |B| B c∈R |B| B
B
holds for every ball B with universal constants. Indeed, the left-hand side is obviously
larger than the right hand side. On the other hand, we get as well
Z Z Z Z Z
1 1 1 1 1
f − f ≤ |f − c| + c − f
|B| B |B| B |B| B |B| B |B| B
Z Z Z
1 1 1
= |f − c| + (c − f )
|B| B
|B| B |B| B
Z Z Z
1 1 1
≤ |f − c| + |c − f |
|B| B |B| B |B| B
Z
2
= |f − c| .
|B| B
66
15 Introduction to harmonic analysis
Proof. Due to the observation above, it is enough to show that for every ball B, there is
a constant cB such that Z
1
|Hf − cB | . kf k∞ .
|B| B
We split Hf = H(f χ2B ) + H(f χR\2B ), where 2B is a ball with the same center as B, but
twice the radius. First we get
Z Z 1/2 Z 1/2 Z 1/2
1 1 1 kf k∞ p
|H(f χ2B )| ≤ |H(f χ2B )|2 · 1 ≤p |f |2 ≤p |2B| . kf k∞ .
|B| B |B| B B |B| 2B |B|
This deals with the “local” part of Hf . For the “global” part, observe that for x ∈ B we
have (modulo constant)
Z
1 1
H(f χR\2B )(x) = f (y) − dy,
y:y6∈2B x−y γ −y
67
Functional Analysis II
The proof copies very much the proof of the same statement for the Hilbert transform,
and we leave out the details.
The condition (CZ2) is sometimes called Hörmander condition. By the help of mean
value theorem, it is satisfied for example if
C
k∇K(x)k2 ≤ , x 6= 0.
kxkn+1
2
An important and non-trivial generalisation of the theory of singular integrals is given
by considering the vector-valued analogues. By this, we mean the following.
• H, H̃ are (complex) Hilbert spaces.
• For 0R< p < ∞, Lp (Rn → H) is the set of measurable functions f : Rn → H, such
that Rn kf (x)kpH dx < ∞.
Z
n
• Let K : R → L (H, H̃). Then T f (x) = K(y)f (x − y)dy takes values in H̃.
Rn
68
15 Introduction to harmonic analysis
where we have used the multinomial theorem (a generalisation of the binomial theorem to
a bigger number of summands) and the fact that
Z 1
r1α1 (t) . . . rm
αm
(t)dt
0
is equal to zero if some of the αi ’s is odd and equal to 1 if all of them are even.
Let α = (α1 , . . . , αm ) ∈ Nm
0 be integers with |α| = k, then
and 1/(2k)
1/(2k) (2k)!
E ≤ kak2 .
2k k!
Hence, the statement holds with 16
1/(2k)
(2k)!
B2k := .
2k k!
Finally, we have to show the existence of A1 > 0. We proceed by a nice duality trick
using the (alreadyPmproven) first part of this theorem.
Let f (t) := n=1 an rn (t). By Hölder’s inequality for p = 3/2 and p′ = 3, we have
Z 1 Z 1 Z 1 2/3 Z 1 1/3
2 2/3 4/3 4
|f (t)| dt = |f (t)| · |f (t)| dt ≤ |f (t)|dt · |f (t)|
0 0 0 0
Z 1 2/3 Z 1 2/3
4/3 4/3 4/3 4/3
≤ |f (t)|dt B4 · kak2 = |f (t)|dt B4 · kf k2 .
0 0
Therefore,
Z 1 2/3 Z 1 1/3
−4/3 2
|f (t)|dt ≥ B4 |f (t)| dt ,
0 0
that is Z 1/2
Z 1 1
|f (t)|dt ≥ B4−2 2
|f (t)| dt = B4−2 kak2 .
0 0
Hence, A1 ≥ B4−2 .
69
Functional Analysis II
Choosing p large enough, this estimate gives very quickly the so-called tail bound estimates
on sum of independent Rademacher variables, i.e. the assymptotic estimates of
!
Xm
P ai εi > t
i=1
for t → ∞.
We use this reformulation of Khintchine’s inequalities to give another proof of Theorem
15.8.1.
Proof. (of the upper estimate in Theorem 15.8.1).
We normalize to kak2 = 1. Then
m
! m m m
X Y Y Y
E exp ai εi = E exp(ai εi ) = E exp(ai εi ) = cosh(ai ).
i=1 i=1 i=1 i=1
70
15 Introduction to harmonic analysis
Furthermore, we denote Sj∗ = Sj−1 + Sj + Sj+1 . Let us observe that this implies Sj∗ Sj =
Sj Sj∗ = Sj .
Finally, we addopt this concept also to smooth dyadic decompositions. Let ψ ∈ S (R)
be non-negative, have support in 1/2 ≤ kξk2 ≤ 4 and be equal to 1 on 1 ≤ kξk2 ≤ 2. Then
we define
ψj (ξ) = ψ(2−j ξ) and (S̃j f )∧ (ξ) = ψj (ξ)fˆ(ξ), ξ ∈ R.
Theorem 15.9.1. (Littlewood-Paley Theory) Let 1 < p < ∞.
i) Then there exist two constants Cp > cp > 0, such that
!1/2
X
cp kf kp ≤
|Sj f |2
≤ Cp kf kp .
j p
X
iii) Finally, if |ψ(2−j ξ)|2 = 1, then there is also a constant cp > 0, such that
j
!1/2
X
2
cp kf kp ≤
|S̃j f |
.
j p
Proof. Step 1.
We know, that (Sj f )∧ = χIj fˆ, where Ij was defined by (15.2). We define
71
Functional Analysis II
leading to
(sgn(ξ − 2j )fˆ(ξ))∨ = M2j (sgn ·τ−2j fˆ)∨ = (2π)−1/2 · M2j (sgn(·)∨ ∗ (τ−2j fˆ)∨ )
√ !
2π 1 1
= (2π)−1/2 M2j p.v. ∗ M−2j f
−i π x
1 1
= iM2j p.v. ∗ M−2j f
π x
= iM2j HM−2j f.
Using the boundedness of H on Lp (R) for 1 < p < ∞, we immediately obtain that
kSj f kp ≤ ckf kp , and the same is true also for Sj+ and Sj− .
Step 2.
We combine Step 1. with (15.1) to obtain
X 1/2
1
X 1/2
X 1/2
|Sj+ fj |2
≤
|(sgn(ξ − 2j )fˆj (ξ))∨ |2
+
|(sgn(ξ − 2j+1 )fˆj (ξ))∨ |2
2
j∈Z p j∈Z p j∈Z p
1
X 2
1/2
≤
|M2j HM−2j fj |
+ ...
2
j∈Z p
1
X 1/2
=
|HM−2j fj |2
+ ...
2
j∈Z p
X 1/2
X 1/2
≤ cp
|M−2j fj |2
+ . . . = c′p
|fj |2
.
j∈Z p j∈Z p
The same holds of course for Sj− and, therefore, also for Sj .
Step 3.
This, together with the identity Sj = Sj S̃j implies
X 1/2
X 1/2
X 1/2
|Sj f |2
=
|Sj S̃j f |2
≤
|S̃j f |2
.
j∈Z p j∈Z p j∈Z p
Step 4.
This shows, that the second inequality in part (i) of the theorem follows from (ii). There-
fore, we prove (ii) now.
Let Ψ̂ = ψ and Ψj (x) = 2j Ψ(2j x). Then Ψ̂j = ψj and S̃j f = Ψj ∗ f. It is enough to
prove that the vector-valued mapping
f → (S̃j f )j
The proof for p 6= 2 is based on the Hörmander’s condition for vector-valued singular
integrals, i.e. that
kΨ′j (x)kℓ2 ≤ Ckxk−2
2 .
72
15 Introduction to harmonic analysis
Step 5.
Finally, we prove the first inequality in part (i) and part (iii) of the theorem. Surprisingly
enough, they follow quite quickly from previous steps and duality.
The identity
X 1/2
|Sj f |2
= kf k2 (15.3)
j∈Z 2
follows. Using this, and the first part of the theorem for p′ with 1/p + 1/p′ = 1 allows the
following estimate.
Z
kf kp = sup f g : kgkp′ ≤ 1
R
Z X
= sup Sj f · Sj g : kgkp′ ≤ 1
R
j
X 1/2
X 1/2
≤ sup
|Sj f |2
·
|Sj g|2
: kgkp′ ≤ 1
j∈Z p j∈Z p
X 1/2
≤ cp
|Sj f |2
.
j∈Z p
Part (iii) of the theorem follows in exactly the same way - the assumption of the theorem
gives exactly the identity (15.3).
73
16 Selected topics
16.1 Riesz representation theorem
We prove the Riesz Representation Theorem, which was used to construct the Borel mea-
surable functional calculus. We follow essentially [14], where one can find a lot more to
measure and integration theory.
74
16 Selected topics
Since M(γ) is closed under complements and under finite intersections, it is also closed
under finite unions.
Step 3.: To show that γ is σ-additive on M(γ), choose Mn ∈ M(γ) pairwise disjoint.
Setting T = M1 ∪ M2 and using γ-measurability of M1 , we obtain
γ(M1 ∪ M2 ) = γ(M1 ) + γ(M2 ).
Thus γ is finitely additive. Further
∞ k k
! ∞
!
X X [ [
γ(Mn ) = lim γ(Mn ) = lim γ Mn ≤γ Mn ,
k→∞ k→∞
n=1 n=1 n=1 n=1
and since the reverse inequality always holds, we reach the conclusion. S
Step 4.: Let now Mn ∈ M(γ) be pairwise disjoint. Our aim is to show that ∞ n=1 Mn ∈
M(γ). Choosing a test set T ⊂ X, we have
k
! k
! ∞
! k
[ [ [ X
γ(T ) = γ T \ Mn + γ T ∩ Mn ≥ γ T \ Mn + γ(T ∩ Mn )
n=1 n=1 n=1 n=1
and b = 0 follows.
75
Functional Analysis II
As µ∗A is monotone on open sets, the second line of this definition extends the first line
and the notation is correct. Such a set function is called outer Radon measure (associated
to A). We shall show, that this mapping is really an outer measure (in the sense described
in Definition 16.1.2), and we shall prove its regularity properties. The construction of
Carathéodory will then give a σ-algebra (and we shall show, that it contains all open sets)
on which µ∗A is a measure. Finally, we shall prove, that this measure has the property
from the statement of the theorem. Let us work out this program.
Step 1.: We shall show the following
ii) µ∗A (G) = sup{µ∗A (C) : C compact, C ⊂ G} for every open set G ⊂ K.
To prove the reverse inequality, select an open set G ⊃ C. Urysohn’s lemma provides a
function g ∈ C(K), 0 ≤ g ≤ 1, g = 0 on K \ G and g = 1 on C. The definition of µ∗A (G)
for G open gives Ag ≤ µ∗A (G). Finally, we observe that
Step 1.(ii): Suppose we are given an open set G ⊂ K and ε > 0. Let f ∈ C(K) be
a function such that 0 ≤ f ≤ 1, f = 0 on K \ G. Since fn := min(f, 1/n) ց 0, an
appeal to Daniell’s property gives the existence of an n ∈ N, such that Afn < ε. If
C := {x ∈ K : f (x) ≥ 1/n}, then C is a compact subset of K. Due to (i), there exists
g ∈ C(K) such that 0 ≤ g ≤ 1, g = 1 on C and Ag ≤ µ∗A (C) + ε. Since f − fn ≤ g, we get
76
16 Selected topics
Hence,
and we finish the proof of one part of (ii). The second one follows directly by monotonicity
of µ∗A .
Step 1.(iii): Finally, we prove that µ∗A is an outer measure. Clearly, we have µ∗A (∅) = 0
and µ∗A (S) ≤ µ∗A (T ) whenever S ⊂ T. It remains to show that µ∗A is σ-subadditive.
To this end, let ε > 0 and C1 , C2 ⊂ K be two given compact sets. By (i), we find
fj ∈ C(K) such that fj = 1 on Cj and Afj ≤ µ∗A (Cj ) + ε; j = 1, 2. Then (by (i) again)
µ∗A (C1 ∪ C2 ) ≤ A(min(f1 + f2 , 1)) ≤ A(f1 + f2 ) = Af1 + Af2 ≤ µ∗A (C1 ) + µ∗A (C2 ) + 2ε
We can also find a compact set C ⊂ V1 ∩ G such that µ∗A (C) + ε > µ∗A (V1 ∩ G) and
an open set W with a compact closure W such that C ⊂ W ⊂ W ⊂ V1 ∩ G. Set
77
Functional Analysis II
µ∗A (T ∩ G) + µ∗A (T \ G) ≤ µ∗A (V1 ∩ G) + µ∗A (V1 \ G) < µ∗A (W ) + ε + µ∗A (W0 )
= µ∗A (W ∪ W0 ) + ε ≤ µ∗A (V1 ) + ε < µ∗A (T ) + 2ε.
Step 3.: We denote by MA := M(µ∗A ) the σ-algebra of measurable sets of µ∗A . The
restriction of µ∗A to MA will be denoted by µA . We shall show, that this measure satisfies
the statement of the theorem.
First, µA is a complete Radon measure. As we restricted ourselves to K compact, it
is easy to see that C(K) ⊂ L1 (µA ). Hence, we only have to show that
Z
Af = f dµA , f ∈ C(K).
K
Using the definition of µ∗A (Gk−1 ) and n(fk − fk−1 ) = 0 on K \ Gk−1 , we get
1
≤ A(fk − fk−1 ) ≤ µA (Gk−1 )
n
for each k = 1, . . . , n. On the other hand, as n(fk − fk−1 ) = 1 on Gk , we first obtain
µA (C) ≤ nA(fk − fk−1 ) for every compact C ⊂ Gk , which implies also µA (Gk ) ≤ nA(fk −
fk−1 ) for all k = 1, . . . , n.
As χGk ≤ n(fk − fk−1 ) ≤ χGk−1 , we get also
Z
1 1
µA (Gk ) ≤ (fk − fk−1 )dµA ≤ µA (Gk−1 )
n K n
for each k = 1, . . . , n.
Thus (using f0 = 0 and fn = f )
Z X Z
n
Af − f dµA = A(fk − fk−1 ) − (fk − fk−1 )dµA
K k=1 K
n
X 1 1 1
≤ (µA (Gk−1 ) − µA (Gk )) = µA (G0 ) = µA ({x ∈ K : f (x) > 0}).
n n n
k=1
Since RµA ({x ∈ K : f (x) > 0}) < ∞ and n may be chosen arbitrarily large, we get
Af = K f dµA .
Step 4.: Finally, we prove the uniqueness.
If ν is a complete Radon measure on K satisfying the statement of the theorem, we have
Z
Af = f dν ≤ χG dν = ν(G)
K
78
16 Selected topics
Wir benutzen nur normalverteilte Zufallsvariablen N (0, 1), d.h. Zufallsvariablen mit
Wahrscheinlichkeitsdichte
1 2
p(x) = √ · e−x /2 .
2π
2 √
Lemma 16.2.1. (i) Sei ω ∼ N (0, 1). Dann gilt E (eλω ) = 1/ 1 − 2λ für −∞ < λ <
1/2.
(ii) (2-Stabilität der Normalverteilung) Sei k eine natürliche Zahl, sei λ = (λ1 , . . . , λk ) ∈
Rk und seien ω1 , . . . , ωk unabhängige identisch
P verteilte (u.i.v.) N (0, 1) Zufallsvari-
ablen. Dann gilt λ1 ω1 + · · · + λk ωk ∼ ( ki=1 λ2i )1/2 · N (0, 1).
√
Proof. Der Beweis von (i) folgt aus der Rechnung (y := 1 − 2λ · x)
Z ∞ Z ∞ Z ∞
1 λx2 −x2 /2 1 (λ−1/2)x2 1 2 dy 1
√ e ·e dx = √ e dx = √ e−y /2 · √ =√ .
2π −∞ 2π −∞ 2π −∞ 1 − 2λ 1 − 2λ
Um (ii) zu beweisen, reicht es nur den Fall k = 2 zu betrachten. Der Rest folgt dann
durch Induktion. Sei also λ = (λ1 , λ2 ) ∈ R2 , λ 6= 0 fest und seien ω1 und ω2 u.i.v. N (0, 1)
Zufallsvariablen. Wir setzen S := λ1 ω1 + λ2 ω2 . Sei t ∈ R beliebig. Wir berechnen
Z
1 2 2
P(S ≤ t) = e−(x +y )/2 dxdy
2π (x,y):λ1 x+λ2 y≤t
Z
1 2 2
= e−(x +y )/2 dxdy
2π x≤c;y∈R
Z
1 2
=√ e−x /2 dx.
2π x≤c
2 2
In dem letzten Schritt haben wir die Rotationsinvarianz der Funktion (x, y) → e−(x +y )/2
ausgenutzt. Den Wert c berechnen wir aus der Skizze (!Tafel!) als die Länge der Normalen
von Null auf die Gerade {(x, y) : λ1 x + λ2 y = t}. Es ergibt sich
λ1 t λ2 t
t
c=
λ2 + λ2 λ2 + λ2
= pλ2 + λ2 .
,
1 2 1 2 1 2
79
Functional Analysis II
Lemma 16.2.2. Sei k eine natürliche Zahl und seien ω1 , . . . , ωk unabhängige normalverteilte
Zufallsvariablen. Sei 0 < ε < 1. Dann gilt
k 2 /2−ε3 /3]
P(ω12 + · · · + ωk2 > (1 + ε)k) ≤ e− 2 [ε
und
k 2 /2−ε3 /3]
P(ω12 + · · · + ωk2 < (1 − ε)k) ≤ e− 2 [ε .
Proof. Wir beweisen nur die erste Ungleichung. Der Beweis der zweiten Ungleichung folgt
sehr ähnlich. Wir setzen β := 1 + ε > 1 und rechnen
P(ω12 + · · · + ωk2 > βk) = P(ω12 + · · · + ωk2 − βk > 0) = P(λ(ω12 + · · · + ωk2 − βk) > 0)
= P(exp(λ(ω12 + · · · + ωk2 − βk)) > 1)
≤ E exp(λ(ω12 + · · · + ωk2 − βk)),
wobei λ > 0 eine beliebige positive reelle Zahl ist und erst später bestimmt wird. Wir
haben im letzten Schritt die Markov-Ungleichung benutzt. Weiter benutzen wir die Eigen-
schaften der Exponentialfunktion und die Unabhängigkeit von ω1 , . . . , ωk . Es ergibt sich
2 2 2
E exp(λ(ω12 + · · · + ωk2 − βk)) = e−λβk · E eλω1 · · · eλωk = e−λβk · (E eλω1 )k
und mit Hilfe von Lemma 16.2.1 erhalten wir schließlich (für 0 < λ < 1/2)
Wir suchen jetzt einen Wert von 0 < λ < 1/2, für den der letzte Ausdruck möglichst klein
wird. Wir berechnen also die Ableitung von e−λβk · (1 − 2λ)−k/2 und setzen sie gleich Null.
So erhalten wir
t2 t3
ln(1 + t) ≤ t − + , −1 < t < 1.
2 3
Remark 16.2.3. Überlegen Sie sich, dass die Unabhängigkeit von ω1 , . . . , ωk in Lemma
16.2.2 nicht weggelassen werden darf.
80
16 Selected topics
Lemma 16.2.4 (Johnson und Lindenstrauss, 1984). Sei 0 < ε < 1 und seien k, n und d
natürliche Zahlen mit
k ≥ 4(ε2 /2 − ε3 /3)−1 ln n.
Dann existiert für jede Menge P = {x1 , . . . , xn } eine Abbildung f : Rd → Rk , so dass
(1 − ε)kxi − xj k22 ≤ kf (xi ) − f (xj )k22 ≤ (1 + ε)kxi − xj k22 , i, j ∈ {1, . . . , n}. (16.1)
wobei (ωu,v ) u.i.v. N (0, 1) Zufallsvariablen sind. Wir zeigen, dass diese Wahl von f
mit positiver Wahrscheinlichkeit (16.1) erfüllt. Damit wäre die Existenz einer solchen
Abbildung bewiesen.
x −x
Sei i, j ∈ {1, . . . , n} beliebig mit xi 6= xj . Dann setzen wir y = kxii−xjjk2 und berech-
nen die Wahrscheinlichkeit, dass die rechte Ungleichung in (16.1) nicht erfüllt ist. Wir
benutzen die 2-Stabilität der Normalverteilung und erhalten
P(kf (xi ) − f (xj )k22 > (1 + ε)kxi − xj k22 ) = P(kf (y)k2 > (1 + ε)) = P(kM yk22 > (1 + ε)k)
= P((ω1,1 y1 + · · · + ω1,d yd )2 + · · · + (ωk,1 y1 + · · · + ωk,d yd )2 > (1 + ε)k)
= P(ω12 + · · · + ωk2 > (1 + ε)k)
für unabhängige normalverteilte ω1 , . . . , ωk . Mit Hilfe von Lemma 16.2.2 läßt sich der
letzte Ausdruck von oben durch
k 2 3
e− 2 [ε /2−ε /3]
abschätzen. Dieselbe Abschätzung gilt auch für die linke Ungleichung in (16.1) und für
n
alle 2 Paare {i, j} ⊂ {1, . . . , n} mit i 6= j. Die Wahrscheinlichkeit, dass eine der Ungle-
ichungen in (16.1) nicht gilt, ist also höchstens
n k 2 3 k 2 3 k
2· · e− 2 [ε /2−ε /3] < n2 · e− 2 [ε /2−ε /3] = exp(2 ln n − [ε2 /2 − ε3 /3]) ≤ e0 = 1
2 2
Remark 16.2.5. Wie ändert sich die Wahrscheinlichkeit, dass f = √1 · M die Formel
k
(16.1) erfüllt, wenn k größer wird, d.h. wenn
k ≥ c(ε2 /2 − ε3 /3)−1 ln n
mit c > 4?
Remark 16.1. Der Beweis des Lemmas 16.2.4 benutzt die sogenannte Probabilis-
tische Methode: Die Existenz von einem Objekt (in unserem Fall der Abbildung
f : Rd → Rk mit (16.1)) wird dadurch bewiesen, dass man zeigt, dass zufällig kon-
struierte Objekte mit positiver Wahrscheinlichkeit die gewünschten Eigenschaften
erfüllen. Als Urvater dieser Methode wird meistens Paul Erdös bezeichnet, es wurde
aber auch schon vorher benutzt. Bis jetzt gibt es keine konstruktive Beweise von
Lemma 16.2.4.
81
Functional Analysis II
Der ursprüngliche Beweis von Johnson und Lindenstrauss ist viel mehr geometrisch.
Man beweist, dass eine (richtig normierte) Projektion auf einen zufälligen k-dim.
Unterraum von Rd mit positiver Wahrscheinlichkeit (16.1) erfüllt.
Seit ca. 10-15 Jahren gibt es viele Varianten von diesem Beweis (meistens durch
spezifische Anwendungen motiviert).
i) Die Auswertung von f (x) soll möglichst schnell sein (in unserem Fall nur k × d,
sonst d ln(k) durch FFT, Fast Fourier Transform)
ii) Die Anzahl der Zufallsvariablen soll minimiert werden (in unserem Fall k × d)
iii) Die Implementierung soll möglichst einfach sein
iv) u.s.w.
82
References
[1] H. W. Alt, Lineare Funktionalanalysis. Eine anwendungsorientierte Einführung, 6. Aufl., Springer, 2012
[8] E. M. Stein and G. Weiss, Introduction to Fourier analysis on Euclidean spaces, Prince-
ton Mathematical Series, No. 32. Princeton University Press, Princeton, N.J., 1971.
[9] H. Triebel, Higher analysis, University Books for Mathematics, Johann Ambrosius
Barth Verlag GmbH, Leipzig, 1992.
[10] D. D. Haroske and H. Triebel, Distributions, Sobolev spaces, elliptic equations, EMS
Textbooks in Mathematics. European Mathematical Society (EMS), Zürich, 2008.
[12] L. Grafakos, Classical Fourier analysis, Second edition, Graduate Texts in Mathe-
matics, 249. Springer, New York, 2008.
[13] L. Grafakos, Modern Fourier analysis, Second edition, Graduate Texts in Mathemat-
ics, 250. Springer, New York, 2009.
[14] J. Lukeš and J. Malý, Measure and Integral, Matfyzpress, 2005; available at
”http://www.mff.cuni.cz/fakulta/mfp/download/books/lukes-maly - measure and integral.pdf”.
83