Fourier Analysis
Fourier Analysis
Yu. Safarov
Notation
If ck , k = 1, 2, . . . , are real numbers then
lim sup ck = lim sup {ck } , lim inf ck = lim inf {ck } ,
m→∞ k>m m→∞ k>m
We shall always fix a set E (usually E = Rn ) and deal with subsets B of E and
complex functions f defined on E. Then
B c = E \ B is the complement of B in E.
If E is a metric space then
B = {x ∈ Rn : dist (x, B) = 0} is the closure (completion) of the set B.
The support of a function f on E is defined as the closure of the set {x ∈ E :
f (x) 6= 0}.
Typeset by AMS-TEX
1
2 CMMS 11 FOURIER ANALYSIS
X ∈ Nc if and only if X c = N \ X ∈ Nf .
Exercise 1.3. Let Ai , i ∈ I, be some σ-algebras over E where I is some set (not
necessarily countable). Prove that ∩i∈I Ai is a σ-algebra over E.
Let A0 be a set of subsets of E. Since 2E is a σ-algebra, it is clear that there
always exists at least one σ-algebra containing all the sets from A0 .
Definition 1.4. Denote by E(A0 ) the intersection of all σ-algebras over E which
contain all the sets from A0 . Then E(A0 ) is a σ-algebra (see Exercise 1.3) which is
called the σ-algebra generated by A0 .
In other words, E(A0 ) is the minimal σ-algebra over E which contains all the
sets from A0 .
Example 1.5. Let E = {a, b, c}, A0 = {{a}}. Then E(A0 ) = {∅, E, {a}, {b, c}}.
Indeed, since the set A1 = {∅, E, {a}, {b, c}} is a σ-algebra which contains {a}
inclusion E(A0 ) ⊂ A1 holds. Since E(A0 ) is a σ-algebra and {a} ∈ E(A0 ) we have
{∅, E} ⊂ A0 , {b, c} = ({a})c = E \ {a} and hence A1 ⊂ A0 .
CMMS 11 FOURIER ANALYSIS 3
Definition 1.6. Let E be a metric space. The Borel σ-algebra over E is the σ-
algebra generated by all open sets. We will denote this σ-algebra by B(E). The
sets B ∈ B(E) are said to be the Borel sets.
Example 1.7.
(1) Obviously, every open or closed set is a Borel set.
(2) Since any countable set {x1 , x2 , . . . } coincides with the union of closed sets
{xk }, the countable sets are also Borel sets.
1, if x ∈ B,
µ(B) =
0, if x ∈
/ B.
Definition 1.10. The measure space (E, A, µ) is said to be complete if every subset
B1 of every set B ∈ A with µ(B) = 0 also belongs to A (and then, by Lemma 1.12,
µ(B1 ) = 0).
Examples 1.11.
(1) If A = 2E (as in Example 1.9) then the measure space (E, A, µ) is always
complete.
(2) Let E = {a, b}, A = {∅, E}, and µ(∅) = µ(E) = 0. Then the measure
space (E, A, µ) is not complete.
Proof.
(1) Since B = A ∪ (B \ A), A ∩ (B \ A) = ∅ and the function µ is nonnegative,
we obtain µ(B) = µ(A) + µ(B \ A) > µ(A).
(2) Since
∪∞ m=k
k=1 Bk = (B1 ) ∪ (B2 \ B1 ) ∪ (B3 \ (B1 ∪ B2 )) ∪ . . . (Bk+1 \ ∪m=1 Bm ) . . .
are mutually disjoint, from countable additivity of the measure µ it follows that
µ(∪∞
k=1 Bk ) = µ(B1 ) + µ(B2 \ B1 ) + µ(B3 \ (B1 ∪ B2 ) + . . .
+ µ(Bk+1 \ ∪m=k
m=1 Bm ) + . . .
for all m ∈ N. The sets Bk \ Bk+1 are mutually disjoint, and therefore
∞
X
µ(Bm ) = µ (Bk \ Bk+1 ) (1.1)
k=m
2. Lebesgue’s measure
Outer measure.
Definition 2.1. Let (E, A, µ) be a measure space. The map ν : A → R̂+ is a
outer measure if
(1) ν(∅) = 0;
∞
(2) the inequality ν (∪∞
P
k=1 B k ) 6 ν(Bk ) holds for all Bk ∈ A.
k=1
where the infimum is taken over all countable families of rectangles {Pk }k∈N such
that B ⊂ ∪k∈N Pk .
Defintion 2.6. The measure µn defined in Theorem 2.5 is called the Lebesgue
measure. The sets B ⊂ B(Rn ) are said to be Lebesgue measurable.
Of course, Lebesgue’s measure is the most natural measure in Rn . It is invariant
with respect to Euclidean transformations (translations and rotations). Under the
additional assumption that the measure of the unit cube is equal to 1, it is the only
measure (defined on the σ-algebra Ln ) which has this property.
6 CMMS 11 FOURIER ANALYSIS
C1
(i.e., C1 = C0 \ (1/3, 2/3)). Next, let C2 be the set obtained from C1 after removing
the open middle third of each of the (two) intervals of which C1 is disjoint union;
C2
More generally, given Ck which is a union of 2k disjoint, closed intervals, let Ck+1
be the set which one gets by removing the open middle third of each of these closed
intervals. Finally, set C = ∩∞
k=0 Ck .
The set C is called the Cantor set. It is closed, and does not contain any
isolated points (i.e., any neighbourhood U of a point x ∈ C contains also other
points from C). There is a one-to-one correspondence between the points from C
and the sequences of 0 and 1 (the first k numbers in the sequence indicate the closed
interval in Ck which contains x). Therefore C is uncountable.
The measure of C0 \ C is equal to
2 3
1 1 2 1 2 1 2 1 1
+ · + · + · + ... = · = 1,
3 3 3 3 3 3 3 3 1 − 2/3
3. Measurable functions
In this section we fix a σ-algebra A over the set E and say that a set is measurable
if it belongs to our σ-algebra A.
Definition 3.1. We say that the real function f : E → R̂ is measurable (or, more
precisely, that f is a measurable function with respect to the σ-algebra A) if for
any real a the inverse image
f −1 ([−∞, a)) = {x ∈ Rn : f (x) < a}
is an element of A.
In the case where E = Rn and A = Ln we say that f is Lebesgue measurable,
and in the case where E is a metric space and A = B(E) we say that f is Borel
measurable (or, simply, that f is a Borel function). From inclusion B(Rn ) ⊂ Ln it
follows that any Borel function f : Rn → R̂ is Lebesgue measurable.
Remark 3.2. A measurable function can be infinite at some points, and such points
also form a measurable set. Indeed,
f −1 (−∞) = ∩∞
k=1 f
−1
([−∞, −k)) , f −1 (+∞) = ∪∞
k=1 (f
−1
([−∞, k)) )c .
Example 3.3. Let E be a metric space and let f be a continuous functon. Then f
is a Borel function. Indeed for any a ∈ R1 the set f −1 ((−∞, a)) is open and hence
is a Borel set.
Theorem 3.4. For any function f : E → R̂ the following two conditions are
equivalent:
(1) f is measurable
(2) the inverse images of all intervals (including degenerate and infinite inter-
vals) are elements of A.
Simple functions.
From the measure-theoretic standpoint, the most elementary functions are those
taking only a finite number of values.
Definition 3.5. The function ϕ : E → R1 is said to be simple if it takes a finite
number of values ak and for each ak the inverse image ϕ−1 (ak ) is a measurable set.
One can easily prove the following lemma.
Lemma 3.6.
(1) A simple function is measurable.
(2) Let ψ1 , ψ2 , . . . ψm be simple functions and let F : Rm → R1 be a continuous
function. Then the function F (ψ1 , ψ2 , . . . ψm ) is a simple function.
Lemma 3.6 immediately implies
Corollary 3.7. P
Let ψ1 , ψ2 , . . . ψm be simple functions and let a1 , a2 , . . . am be real
m
numbers. Then k=1 ak ψk is a simple function.
In other words, Corollary 3.7 says that a linear combination of simple functions
is a simple function.
The following lemma implies that every measurable function can be approxi-
mated by simple functions.
8 CMMS 11 FOURIER ANALYSIS
(ψk )+ (x) 6 (ψk+1 )+ (x) 6 f+ (x), (ψk )− (x) 6 (ψk+1 )− (x) 6 f− (x)
Obviously, 0 6 ϕk (x) 6 ϕk+1 (x) 6 f (x). Since f is measurable, the inverse images
ϕ−1
k (m/2 k
) = f −1
[m/2k
, (m + 1)/2k
) are measurable sets.
If f is bounded then every function ϕk can take only a finite number of values
(R)
0 , 1/2k , . . . , m0 /2k (where m0 /2k < sup f ). Thus, ϕk are simple functions, and
A convergence theorem.
We shall deduce all the basic properties of the measurable functions from the
following theorem.
CMMS 11 FOURIER ANALYSIS 9
Now the required statement follows from Theorem 3.4 and the definition of mesura-
bility.
(2) is an easy consequence of (1) and equalities
lim sup ϕk (x) = inf sup {ϕm (x)}, lim inf ϕk (x) = sup inf {ϕm (x)}.
k m>k k m>k
for all x. Thus, F (f1 , . . . , fp ) is the pointwise limit of a sequence of simple functions
and therefore F (f1 , . . . , fp ) is measurable.
Theorem 3.10 immediately implies
Corollary 3.11. If the functions f and g are measurable then |f |, f + g and f g
are also measurable.
Complex-valued functions.
A complex-valued function is said to be measurable (simple) if its real and com-
plex parts are measurable (simple) functions. Clearly, all results of this section can
be easily extended to the complex-valued functions.
10 CMMS 11 FOURIER ANALYSIS
4. Construction of integrals
In this section we fix a measure space (E, A, µ).
Definition 4.1. Denote by Σ0 (E, A, µ) (or simply by Σ0 ) the set of all simple
functions ϕ such that µ({x ∈ E : ϕ(x) 6= 0} < ∞.
Definition 4.2. If ϕ ∈ Σ0 then we define
Z m
X
ϕ dµ = ak µ(Ωk )
E k=1
where a1 , a2 , . . . , am are the nonzero values of the function f and Ωk = ϕ−1 (ak ) =
{x ∈ E : ϕ(x) = ak }.
Example 4.3. Let (E, A, µ) = (R1 , L1 , µ1 ) and
0, if x ∈ (−∞, −1),
1, if x ∈ [−1, 1],
ϕ(x) =
3, if x ∈ (1, 2),
0, if x ∈ [2, +∞).
Lemma 4.4. Let f ∈ Σ0 and g ∈ Σ0 . Then for any (real) constants α and β
Z Z Z
(αf + βg) dµ = α f dµ + β g dµ .
for all Borel sets Ω. Therefore the first term in the right hand side of (5.2) coincides
with Z
X
ak µ(Ωk ) = f dµ .
k
By analogy, the second term in the right hand side of (5.2) is equal to
X Z
am µ(Ω
e m) = g dµ .
m
Definition of integral.
Definition 4.5. Let f be a non-negative measurable function. We define
Z Z
f dµ = sup ϕ dµ ,
E
x ∈ E0 ,
f (x),
f0 (x) =
0, x∈
/ E0 .
Further on in this
R section we shallR always assume that the integrals are taken
over E and write f dµ instead of E f dµ.
One can easily prove that Definition 4.5 does not contradict to Definition 4.2,
that is, both definitions give the same result in the case where f is a non-negative
function from Σ0 .
Definition 4.6. Let f be a measurable function. We define
R R R
(1) if R f+ dµ = +∞ andR f− dµ < ∞ then R f dµ = +∞ ;
(2) if R f+ dµ < ∞ and R f− dµ = +∞ thenR f dµ = R −∞ ; R
(3) if f+ dµ < ∞ and f− dµ < ∞ then f dµ = f+ dµ − f− dµ.
In the last case the function f is said to be integrable. The class of integrable
functions is denoted by L1 (E, A, µ) (or simply L1 ).
Remark
R 4.7. Note that a measurable function f belongs to L1 if and only if
|f | dµ < ∞.
Theorem 4.8 (elementary properties of the integral). Let f and g be inte-
grable functions. Then
R R
(1) Rf dµ 6 Rg dµ whenever f 6 g;
(2) R| f dµ| 6 R|f | dµ;
(3) R af dµ =R a f dµ Rfor every real constant a;
(4) f dµ + g dµ = f + g dµ.
12 CMMS 11 FOURIER ANALYSIS
Proof. The first three results immediately follow from the definition. The last
result is less obvious. It is deduced from Lemma 4.4 with the use of Lemma 3.8
and Theorem 5.2
The above definition of integral is due to Lebesgue. Note that Lebesgue’s integral
of a non-negative measurable functionR always exists (and
R may be infinite). But that
R f+ dµ = ∞ and f− dµ = ∞ then we cannot
is not true in the general case. If
say anything about the value of f dµ.
Example R4.9. Let f (x) = xR−1 sin x , x ∈ R1 , and µ be Lebesgue’sR measure on
∞ ∞ ∞
R1 . Then 1 f+ dµ = ∞ and 1 f− dµ = ∞, so Lebesgue’s integral 1 f dµ does
not exist. But the Riemann integral is well-defined and finite (this can be easily
proved by integrating by parts).
In a sense, Lebesgue’s definition of integrals is more restrictive as far as the
behaviour at infinity is concerned. On the other hand, it allows us to integrate very
non-smooth functions, for which the Riemann integral does not exist. Moreover, it
works for an arbitrary measure µ whereas the Riemann integral is defined only for
Lebesgue’s measure.
Example
R 4.10. Let µ be the δ-measure at the point x (Example 1.9). Then
f dµ = f (x).
CMMS 11 FOURIER ANALYSIS 13
5. Convergence of integrals
Proof. Given δ > 0, denote Ek = {x ∈ E : fk (x) > ϕ(x) − δ}. Then, under
conditions of the lemma, Ek ⊂ Ek+1 , and ∪k Ek = E . Therefore (E \ Ek ) ⊃
(E \ Ek+1 ) and ∩k (E \ Ek ) = ∅. Now Lemma 1.12(3) implies that µ(E \ Ek ) → 0
as k → ∞.
In view of Theorem 4.8(1) and Lemma 4.4 we have
Z Z Z
fk dµ > fk dµ > (ϕ − δ) dµ
Ek Ek
Z Z Z Z Z Z
= ϕ dµ − ϕ dµ − δ dµ > ϕ dµ − C dµ − δ dµ
E E\Ek Ek E E\Ek Ek
Z
= ϕ dµ − C µ(E \ Ek ) − δ µ(Ek )
where C = max f (x). Since the last two terms in the right hand side can be
x∈E
made arbitrarily small by choosing large k and small δ respectively, the required
inequality follows.
Theorem 5.2 (the monotone convergence theorem). Let fm be non-negative
measurable functions such that
fm (x) 6 fm+1 (x) , lim fm (x) = f (x)
m→∞
R R
for all x ∈ E. Then lim fm dµ = f dµ .
m→∞
Proof. The condition fm 6 fm+1 implies that fk 6 f and hence, by Theorem
4.8(1), Z Z
fk dµ 6 f dµ , ∀k = 1, 2, . . .
for all ϕ ∈ Simp0 such that 0 6 ϕ 6 f . By Definition 4.5 this implies that
Z Z
lim fm dµ > f dµ.
m→∞
14 CMMS 11 FOURIER ANALYSIS
for all x ∈ E and m ∈ N, Now (5.2)) and the monotone convergence theorem imply
Z Z Z
lim inf fk dµ = lim hk dµ = lim hk dµ
k→∞ k→∞
Z Z
= lim inf hk dµ 6 lim inf fk dµ .
Corollary 5.4. Let fk be real measurable functions.
R Assume that there exists a
measurable function g such that fk (x) 6 g(x) and g dµ < ∞. Then
Z Z
lim sup fk dµ > lim sup fk dµ.
Proof. If we apply Fatou’s lemma to the functions (g − fk ) then (5.1) turns into
the required inequality.
Theorem 5.5 (Lebesgue’s dominated convergence theorem). Let fk be mea-
surable functions, and fk (x) → f (x) for all x. Assume thatR there exists a non-
negative measurable function g such that |fk (x)| 6 g(x) and g dµ < ∞. Then
Z
lim |f − fk | dµ = 0 , (5.5)
k→∞
Z Z
lim fk dµ = f dµ . (5.6)
k→∞
Proof. Since |fk (x)| 6 g(x), we have |f (x)| 6 g(x) and |f (x) − fk (x)| 6 2g(x). By
Corollary 5.4
Z Z Z
lim |f − fk | dµ 6 lim |f − fk | dµ = lim |f − fk | dµ = 0 ,
k→∞ k→∞ k→∞
etc.
Complex-valued functions. R R R
If f is a complex-valued function then one defines f dµ = Re f dµ+ Im f dµ.
Using the corresponding results on real-valued functions, one can easily prove sim-
ilar statements for complex-valued functions.
16 CMMS 11 FOURIER ANALYSIS
6. Comparison of measures.
The notion “almost everywhere”.
Definition 6.1. We say that an x-dependent statement about quantities on the
measure space (E, A, µ) holds almost everywhere (a.e.) if the set B of x for which
the statement fails is an element of A and µ(B) = 0. If E0 ∈ A then we say that
an x-dependent statement holds almost everywhere on E0 if the subset B ⊂ E0 of
x for which the statement fails is an element of A and µ(B) = 0.
Of course, the notion “almost everywhere” depends on the measure µ.
Example 6.2. Let µ be the δ-measure at the point x0 , and f and g be some
measurable functions. Then f > g a.e. if and only if f (x0 ) > g(x0 ).
Example 6.3. Let µ be Lebesgue’s measure on R1 , and f and g be measurable
functions on R1 such that f (x) = g(x) for all irrational x. Then f = g a.e.
Theorem 6.4. Let (E, A, µ) be a measure space, E0 ∈ A and f, g be non-negative
measurable functions on E0 . Then
R
(1) E0 f dµ = 0 if and only if f (x) = 0 a.e. on the set E0 ;
R R
(2) E0 f dµ = E0 g dµ whenever f = g a.e. on the set E0
lim g(x− +
k ) = f (x0 − 0) 6= f (x0 + 0) = lim g(xk ) ,
k→∞ k→∞
Definition 6.6. Let A be a σ-algebra over E and let µ1 and µ2 be measures defined
on A. We say that
(1) the measures µ1 and µ2 are (mutually) singular if there exists a set Ω ∈ A
such that µ1 (Ω) = µ2 (Ωc ) = 0.
(2) the measure µ1 is absolutely continuous with respect to the measure µ2 if
µ1 (Ω) = 0 whenever µ2 (Ω) = 0.
Example 6.7. Let ν1 coincides with the Lebesgue measure µn and ν2 coincides
with the restriction of the δ-measure at the point x0 to the σ-algebra Ln of Lebesgue
mesurable sets. Then the measures ν1 and ν2 are mutually singular. Indeed, we
can take Ω = {x0 }.
If two measures are singular then they live on non-overlapping subsets of E and,
in fact, have nothing to do with one another. In particular, it can happen that
f > g a.e. with respect to µ1 and f < g a.e. with respect to µ2 .
If µ1 is absolutely continuous with respect to µ2 then any statement which holds
µ2 -a.e. is also valid µ1 -a.e. Moreover, under some additional restriction the measure
µ1 can be represented as an integral with respect to the measure µ2 (see Theorem
6.10).
Definition 6.8. We say that the measure space (E, A, µ) (or the measure µ) is
σ-finite if E can be represented as the union of a countable collection of the sets
Ek , k = 1, 2, . . . such that Ek ∈ A and µ(Ek ) < ∞ for all k ∈ N.
Example 6.9. The Lebegue measure µn is σ-finite. Indeed, we can take Ek =
(−k, k)n , k ∈ N.
Theorem 6.10 (Radon–Nikodym theorem). Let A be a σ-algebra over the set
E and let µ1 and µ2 be measures defined on A. Assume that the measure µ2 is
σ-finite. Then µ1 is absolutely continuous with respect to µ2 if and only if there
exists a measurable non-negative function ρ such that
Z
µ1 (E0 ) = ρ dµ2 (6.1)
E0
where δxk are the delta-measures at the points xk and ck = µ ({xk }). The corre-
sponding distribution function is constant on the intervals which do not contain the
points xk and has the jumps ck at the points xk .
Clearly, if µ1 is continuous and µ2 is discrete then µ1 and µ2 are mutually singular
(we can take Ω = {x1 , x2 , . . . }).
Proposition 6.15. Any Borel measure µ on R1 can be uniquely represented as the
sum µd + µc , where µd is a discrete measure and µc is a continuous measure.
Proof. Let Ω be the set of points x for which µ ({x}) > ε, and
Ω = ∪ε>0 Ωε = {x : µ ({x}) > 0} .
Since the measure of a bounded set is finite, the intersection of Ωε with any bounded
interval contains only a finite number of points. This implies that the sets Ωε are
countable. Therefore Ω is also countable. Let Ω = {x1 , x2 , . . . }, and µd be the
discrete measure defined by (6.2). Then µc = µ − µd is a continuous measure, and
µ = µd + µc . The points xk and the constants ck are determined uniquely by the
measure µ, so µd and µc are uniquely defined.
Proposition 6.15 is not very helpful, since we do not have any explicit represen-
tation for an arbitrary continuous measure µc . It is easy to see that a measure
is continuous if it is absolutely continuous with respect to Lebesgue’s measure.
However, the converse is not true.
Example. Let N (λ) be the continuous function on [0, 1] such that
N (λ) = 1/2 as λ ∈ (1/3, 2/3),
N (λ) = 1/4 as λ ∈ (1/9, 2/9) and N (λ) = 3/4 as λ ∈ (7/9, 8/9),
etc. Then N (0) = 0, N (1) = 1, and N is constant on any open interval from
[0, 1] \ C, where C is the Cantor set (see Section 2). The corresponding measure µ
lives on the Cantor set, and for any fixed point x we have µ ({x}) = 0. Therefore µ
is neither discrete nor absolutely continuous with respect to Lebesgue’s measure.
CMMS 11 FOURIER ANALYSIS 19
Z 1/p
p
kf kLp (E,µ) = |f | dµ .
Definition 7.1. We say that f ∈ Lp (E, µ) if f is measurable and kf kLp (E,µ) < ∞.
When E = Rn and µ is the Lebesgue measure, we omit µ and simply write Lp (Rn ).
We shall identify two functions from Lp (E, µ) which coincide µ-a.e. (which
means, in fact, that we assume the elements of Lp (E, µ) to be the classes of func-
tions). Since kf kLp (E,µ) = 0 implies f = 0 a.e., this means that f = 0 as an element
of Lp (E, µ) whenever kf kLp (E,µ) = 0. By Minkowski’s inequality
satisfies the triangle inequality. We shall consider Lp (E, µ) as a metric space with
the metric (7.1). In particular, fk → f in Lp (E, µ) if kf − fk kLp (E,µ) → 0.
Remark 7.2. Strictly speaking, the elements of Lp are equivalence classes of func-
tions. One can say “the function f belongs to Lp ” (Definition 7.1) but bearing
in mind that there are many other functions which coincide with f as elements of
Lp . In particular, if we modify f on a set of measure zero then the new function
coincides with the initial one as an element of the Lp space. Therefore one should
be carefull making statements like “the function f ∈ Lp (Rn ) is equal to zero at the
origin” (changing the function f at the origin we do not change it as an element of
Lp (Rn )).
By Lebesgue’s dominated convergence theorem, if fk → f µ-a.e. and there is
a function g ∈ Lp (E, µ) such that |fk (x)| 6 g(x) then fk → f in Lp (E, µ). The
converse is not true; convergence in Lp (E, µ) does not imply convergence µ-a.e.
However, we have the following
CMMS 11 FOURIER ANALYSIS 21
Theorem 7.3. If kf − fk kLp (E,µ) → 0 then there is a subsequence fkj which con-
verges to f µ-a.e. Moreover, if {fk } is a Cauchy sequence in Lp (E, µ) then there
exist a subsequence fki and f ∈ Lp (E, µ) such that kf − fkj kLp (E,µ) → 0 and
fkj → f µ-a.e.
Proof. Assume that {fk } is a Cauchy sequence in Lp (E, µ), that is,
Z
p
kfm − fk kLp (E,µ) = |fm − fk |p dµ → 0 , k, m → ∞ .
kf − fkj kpLp (E,µ) 6 lim inf kfkl − fkj kpLp (E,µ) 6 sup kfkl − fkj kpLp (E,µ) → 0
l→∞ l>j
for each fixed ε. Let us choose 1 6 k1 < k2 < . . . in such a way that
If |fkj+1 (x) − fkj (x)| < 2−j−1 for all j > i then
l−1
X l−1
X
|fkl (x) − fki (x)| 6 |fkj+1 (x) − fkj (x)| < 2−j−1 < 2−i , ∀l > i.
j=i j=i
Therefore sup |fkl (x) − fki (x)| > 2−i implies sup |fkj+1 (x) − fkj (x)| > 2−j−1 ,
l>i j>i
and we obtain
−i
µ {x : sup |fkl (x) − fki (x)| > 2 }
l>i
−j−1
6 µ {x : sup |fkj+1 (x) − fkj (x)| > 2 }
j>i
−j−1
6 µ ∪ {x : |fkj+1 (x) − fkj (x)| > 2 }
j>i
∞
X ∞
X
µ {x : |fkj+1 (x) − fkj (x)| > 2−j−1 } 2−j−1 6 2−i . (7.2)
6 6
j=i j=i
Assume that the sequence fkj (x) is not convergent. Then there exists ε > 0 for
which the inequality
sup |fkl (x) − fki (x)| > ε
l>i
22 CMMS 11 FOURIER ANALYSIS
holds with an arbitrarily large i. Therefore the set of points at which {fkj } does
not converge is a subset of
lim kf − fk kLp (E,µ) 6 lim kf − fkj kLp (E,µ) + lim kfkj − fk kLp (E,µ) = 0 .
k→∞ kj →∞ k,kj →∞
Definition 7.5. We say that a function ϕ on R1 is a step function if ϕ takes only
a finite number of values a1 , . . . , am and for each ak 6= 0 the inverse image ϕ−1 (ak )
is a (possibly degenerate) interval.
Theorem 7.6. The set of step functions is dense in Lp (R1 ) for all 1 6 p < ∞.
The proof of this theorem is based on the following
Lemma 7.7. Any open set U ∈ R1 is the union of a countable collection of mutu-
ally disjoint open intervals.
Proof. If U ⊂ R1 is open and x ∈ U , let Ix be the maximal connected component
of U containing x. Since U is open, Ix is an open interval. Obviously, for any
x, y ∈ U , either Ix = Iy or Ix ∩ Iy = ∅.
We have U = ∪x Ix where the union is taken over all the rational x. Since the
set of rational numbers is countable (and the intervals Ix either coincide or do not
intersect), we have obtained the required representation of U .
Proof of Theorem 7.6. From the definition of Lebesgue’s integral it follows that any
function f ∈ Lp (R1 ) is approximated in Lp (R1 ) by compactly supported simple
functions. Any compactly supported simple function is a linear combination of the
characteristic functions of some measurable sets. Therefore it is sufficient to prove
that the characteristic function χ of an arbitrary measurable set Ω ∈ R1 can be
approximated by step functions.
By Theorem 2.5 for all ε > 0 there exist open sets Ωε ⊃ Ω such that µ1 (Ωε \Ω) 6
ε. Since the characteristic functions χε of the sets Ωε converge in Lp (R1 ) to χ as
ε → 0, we can assume without loss of generality that Ω is an open set.
CMMS 11 FOURIER ANALYSIS 23
By Lemma 7.7 an open set Ω coincides with the union of a countable collection
of open intervals Ik , k = 1, 2, . . . If χk is the characteristic function of Ik then the
Pj
ej = k=1 χk converge to the characteristic function of Ω in Lp (R1 ) as
functions χ
j → ∞.
Let C0∞ (R1 ) be the space of smooth functions with compact supports. Obviously,
the characteristic function of a bounded interval I can be approximated in Lp (R1 )
by C0∞ -functions fk (for example, we can choose fk in such a way that fk (x) = 1
for all k and x ∈ I, 0 6 fk (x) 6 1 for all x, and fk (x) = 0 if dist(x, I) > k −1 ).
Therefore Theorem 7.6 implies
Corollary 7.8. The space C0∞ (R1 ) is dense in Lp (R1 ) for all 1 6 p < ∞.
24 CMMS 11 FOURIER ANALYSIS
8. Distributions
Convergence in S(R1 ).
We say that the sequence of functions ϕk ∈ S(R1 ) converges to ϕ ∈ S(R1 ) in
S(R1 ) if
dm (ϕ − ϕk )
sup | xl | → 0, k → ∞,
x dxm
for all l, m = 0, 1, . . . . If ϕk → ϕ in S(R1 ) then for all l and m we have
dm (ϕ − ϕk )
| m
| 6 ck,l,m (1 + x2 )−l ,
dx
where ck,l,m → 0 as k → ∞. This implies, in particular, that ϕk → ϕ in Lp (R1 ) for
all p.
if and only if µ1 = µ2 . Therefore the Borel measures can be also identified with the
corresponding distributions.
CMMS 11 FOURIER ANALYSIS 25
we δx in the right hand side stands for the δ-measure at x (Example 1.9).
hf 0 , ϕi = − hf, ϕ0 i , ∀ ϕ ∈ S(R1 ) .
In the same manner we can define many other operators in S 0 (R1 ), in particular,
the operator f (x) → g(x) = f (y(x)).
26 CMMS 11 FOURIER ANALYSIS
ck = f (ak + 0) − f (ak − 0)
Pm
Therefore f 0 = g + k=1 ck δak where δak are the δ-functions at the points ak .
Example 8.10. Let f = δx . Then hf 0 , ϕi = −ϕ0 (x). The distribution f 0 cannot
be described in any simpler way. This distribution is called the derivative of the
δ-function at x.
Supports of distributions.
Generally speaking, one cannot say what is the value of a distribution at one
fixed point. However, we can single out a class of distributions which coincide on
an open set.
Definition 8.11. We say that the distribution f vanishes on an open set Ω and
write f |Ω = 0 if hf, ϕi = 0 for all ϕ ∈ S(R1 ) with supp ϕ ⊂ Ω. We say that f
coincides with a distribution g on Ω if (f − g)|Ω = 0 .
In particular, we say that the distribution f coincides with a function g on Ω if
(f − g)|Ω = 0 in the sense of distributions.
CMMS 11 FOURIER ANALYSIS 27
dm ψ̂k
sup | ξ l | → 0, k→∞ (9.1)
ξ dξ m
dm fˆ
l\
d (xm f )
l
ξ m
= (−i)l+m , ∀f ∈ S(R1 ) ,
dξ dxl
l
and f → xm ddxfl is a continuous operator in S(R1 ), it is sufficient to obtain (9.1)
only for l = m = 0. This follows from the estimates
Z
−1/2
sup |ψ̂k (ξ)| = (2π) |e−ixξ ψk (x) dx |
ξ
Z Z
−1/2 −1/2
6 (2π) |ψk (x)| dx 6 ck (2π) (1 + x2 )−1 dx ,
If f ∈ S(R1 ) then
Let now f be a distribution. Then, in view of Lemma 9.1, (9.2) defines a linear
continuous functional fˆ on S(R1 ).
Definition 9.2. The Fourier transform of f ∈ S 0 (R1 ) is the distribution fˆ defined
by (9.2).
Obviously, (9.2) holds if f belongs to Lp (R1 ), p = 1, 2, and if the Fourier trans-
form is understood in the Lp sense. Therefore the Fourier transform in S 0 (R1 )
coincides with the Fourier transform in Lp (R1 ) if f ∈ Lp (R1 ).
By analogy, the inverse Fourier transform F −1 f of f ∈ S 0 (R1 ) is defined by
hF −1 f, ϕi = hf, F −1 ϕi , ∀ ϕ ∈ S(R1 ) .
We have
hF −1 fˆ, ϕi = hfˆ, F −1 ϕi = hf, F
\ −1 ϕi = hf, ϕi ,
for all ϕ ∈ S(R1 ), which means that F −1 fˆ = f . Since F −1 f (ξ) = fˆ(−ξ), all the
results concerning the Fourier transform in S 0 (R1 ) can be easily reformulated for
the inverse Fourier transform.
CMMS 11 FOURIER ANALYSIS 29
Therefore δ̂0 is the function identically equal to (2π)−1/2 . In the same way we obtain
that F −1 δ0 = δ̂0 . This implies that the direct and inverse Fourier transforms of the
function which is identically equal to a, coincide with a (2π)1/2 δ0 .
Proposition 9.4. For all f ∈ S 0 (R1 )
Proof. The equalities (9.3) are valid for all the functions from S(R1 ). Therefore for
all ϕ ∈ S(R1 ) we have
c = i hξ fˆ, ϕi ,
hfb0 , ϕi = hf 0 , ϕ̂i = − hf, (ϕ̂)0 i = i hf, xϕi
Clearly, if P is not identically equal to zero and fˆ is a function then (9.5) implies
fˆ = 0 almost everywhere. Thus, we cannot obtain from (9.5) any non-trivial
solution of (9.4) if we assume fˆ to be a function. But the equation (9.5) can have
non-trivial solutions in the class of distributions. For example, if P (ξ) has a zero
of the second order at the point η then (9.5) is fulfilled for fˆ = αδη + βδη0 with all
α, β ∈ C1 . Taking the inverse Fourier transform we obtain non-trivial solutions of
(9.4), which appear to be infinitely smooth functions.
Divergent integrals.
We have defined the Fourier transform for any distribution f ∈ S 0 (R1 ). In
particular, the distribution f can be a bounded function. This means, in fact, that
we have defined the integral Z
e−ixξ f (x) dx
for an arbitrary bounded measurable function f . Of course, the integral does not
converge in the classical sense, but it can be understood as a distribution in ξ.
This idea can be generalized in different directions. It often happens that a
divergent integral depending on a parameter can be interpreted as a distribution in
this parameter. Moreover, the corresponding distribution may well be a function
even if the integral does not converge.
Example 9.6. Let f be a bounded measurable function. Let us consider the
integral Z
2
eitx f (x) dx . (9.6)
We formally write
!
Z Z 1 Z
2 2 d 2
e−itx f (x) dx = e−itx f (x) dx + ix−2 e−itx f (x) dx (9.7)
−1 dt |x|>1
and accept the right hand side of (9.7) as the definition of the integral (9.6). The
first term in the right hand side of (9.7) is a function of t, the second term is a
derivative of a function which we understand as a distribution. One can prove
that if f is infinitely differentiable with uniformly bounded derivatives, then the
distribution on the right hand side of (9.7) coincides with an infinitely smooth
function for t 6= 0.