Analysis Chap2&3
Analysis Chap2&3
Definition 2.1.1. We say that the sequence of functions (fn )n converges pointwise to
f on A if and only if for every fixed x ∈ A, the numerical sequence (fn (x))n converges
to f (x), and we write
∀x ∈ A : lim fn (x) = f (x)
n→+∞
or
fn −→ f, n → +∞.
A
In other words
fn (x) = xn , n ∈ N.
We have
43
Example 2.1.3. Let fn : R → R be defined by
nx3
fn (x) = , n ∈ N.
1 + nx3
• For x = 0, fn (0) = 0, hence limn→+∞ fn (0) = 0.
nx3
• Pour x ̸= 0, lim fn (x) = lim 1+nx 3 = 1.
n→+∞ n→+∞
In the above definition, the integer N generally depends on both ε and x. This will some-
times lead us to write N (ε, x) instead of just N . By requiring that N be independent
of x, we achieve a more restrictive mode of convergence called uniform convergence.
Definition 2.1.4. We say that the sequence of functions (fn )n converges uniformly on
A to the function f if and only if
Definition 2.1.5. We say that the sequence of functions (fn ) converges uniformly on
A to the function f if and only if
as n → +∞.
The inverse of Theorem 2.1.6 is not true, as shown by the following example.
44
Example 2.1.7. Let
fn (x) = xn , A = [0, 1], n ∈ N.
For a fixed x ∈ [0, 1), lim fn (x) = 0, and for x = 1, we have lim fn (1) = 1. The
n→+∞ n→+∞
sequence (fn )n converges pointwise to the function f defined by
(
0, x ∈ [0, 1[
f (x) =
1, x = 1
Remark. In the previous example, we observe that the pointwise limit of continuous
functions on A = [0, 1] is not a continuous function.
Definition 2.1.5 of uniform convergence assumes that the limit function f of the sequence
(fn )n is known. When f is unknown, one can utilize the highly significant criterion as
follows.
45
Theorem 2.1.10. Let fn : A → R n ∈ N be a sequence of functions. For the sequence
(fn )n to converge uniformly on A, it is necessary and sufficient that:
Proof. Suppose that (fn )n converges uniformly to a function f on A, and let ε > 0.
Then, there exists N ∈ N such that
ε
∀n ≥ N, ∀x ∈ A : |fn (x) − f (x)| < .
2
So, ∀m ≥ N, ∀n ≥ N, ∀x ∈ A
Let ε > 0, and let N ∈ N be chosen according to the condition of the theorem. For all
x∈A
∀m ≥ N, ∀n ≥ N : |fm (x) − fn (x)| < ε.
So, by letting m approach infinity, we obtain
Corollary 2.1.11. For the sequence (fn )n not to converge uniformly on A, it is nec-
essary and sufficient that
we consider xn = nπ
2 ∈ R, p = 2n (n ∈ N∗ ) and let ε = 14 . Then
xn xn π π 1 1
|f3n (xn ) − fn (xn )| = sin − sin = sin − sin = > = ε,
3n n 6 2 2 4
for all n ≥ N , which shows that the sequence (fn )n does not satisfy the Cauchy criterion
for uniform convergence on A = R.
46
2.1.2 Uniform Convergence and Continuity
Theorem 2.1.13. Let the sequence (fn )n uniformly converges to f on A. If the func-
tions fn are continuous at x0 ∈ A, then f is continuous at x0 .
|f (x) − f (x0 )| ≤ |f (x) − fn (x)| + |fn (x) − fn (x0 )| + |fn (x0 ) − f (x0 )|
ε ε ε
< + + = ε,
3 3 3
which proves the theorem.
The functions fn are continuous, and their limit f is not continuous at x = 0. Therefore,
the convergence of (fn )n is not uniform on R− .
We should point out here the existence of sequences (fn ) of continuous functions that
converge pointwise but not uniformly to a limit function f that is continuous.
47
(2) We have
Z b Z b Z b
lim fn (x)dx = lim fn (x)dx = f (x)dx
n→+∞ a a n→+∞ a
where
Mj (fn ) = sup fn ; mj (fn ) = inf fn , ∆xj = xj+1 − xj .
[xj ,xj+1 ] [xj ,xj+1 ]
ε ε
< + (b − a) = ε.
2 2(b − a)
So, f ∈ R(A).
To prove the second assertion, we have
Z b Z b Z b
fn (x)dx − f (x)dx ≤ |fn (x) − f (x)| dx
a a a
≤ (b − a) sup |fn (x) − f (x)| −→ 0, n → +∞
x∈[a,b]
This result allows, in some examples, to demonstrate the lack of uniform convergence.
Therefore, if a sequence (fn )n of integrable functions on A = [a, b] converges to an
integrable function on A = [a, b] and
Z b Z b
lim fn (x)dx ̸= f (x)dx.
n→+∞ a a
48
Example 2.1.17. Let
The sequence (fn )n converges pointwise to the function f defined by f (x) = 0, x ∈ [0, 1].
So Z 1
f (x)dx = 0.
0
On the other hand, we have
Z 1 Z 1 Z 0
2 n 2
fn (x)dx = n x(1 − x) dx = n (1 − t)tn (−dt), (1 − x = t)
0 0 1
Z 1
n2
2 n n+1
2 1 1
=n t −t dt = n − = −→ 1, n → +∞,
0 n+1 n+2 (n + 1)(n + 2)
Corollary 2.1.18. Let (fn ) be a sequence of integrable functions on the compact in-
terval [a, b], converging uniformly to a function f on [a, b]. Denote
Z x Z x
Fn (x) = fn (t)dt, F (x) = f (t)dt.
a a
Theorem 2.1.19. Let (fn ) be a sequence of integrable functions on the compact interval
[a, b], converging pointwise to a function f on [a, b]. If the functions fn are all bounded
by M , and if the convergence of (fn ) to f is uniform on every compact interval [α, β] ⊂
(a, b), then f is integrable on [a, b], and we have
Z b Z b
lim fn (x)dx = f (x)dx.
n→+∞ a a
49
Proof. The limit function f is integrable on [a, b]. Indeed, it is bounded because as
we take the limit as n approaches infinity, we have |f (x)| ≤ M for all x ∈ [a, b].
Furthermore, f is integrable on any compact interval contained within (a, b) as the
uniform limit of integrable functions.
Given the number ε > 0, let’s choose
ε ε
α<a+ , β >b− such that a < α < β < b.
6M 6M
For all n ∈ N, we have
Z α
ε
(fn (x) − f (x)) dx ≤ 2M (α − a) ≤
a 3
and Z b
ε
(fn (x) − f (x)) dx ≤ 2M (b − β) ≤ .
β 3
The numbers α and β fixed, the uniform convergence of (fn ) to f on [α, β] implies the
existence of an integer N such that for all n > N , we have
Z β
ε
(fn (x) − f (x)) dx ≤ ,
α 3
Example 2.1.20. Let (fn ) be the sequence defined on [0, π/2] by fn (x) = sinn x.
This sequence converges uniformly to the zero function on any interval [0, a], for any
0 < a < π/2, because on such an interval we have |sinn x| ≤ sinn a with 0 < a < 1.
Furthermore, the functions fn are bounded by 1. Therefore, thanks to the previous
theorem, we have
Z π/2
lim sinn xdx = 0.
n→+∞ 0
1 1/2
2
fn (x) = x + 2
n
50
This sequence converges pointwise to the function f : x 7→ |x|. For each n ≥ 1, fn is
differentiable (and even of class C ∞ ) on R. Furthermore, we have:
r
1 1 1
|x| ≤ x2 + 2 ≤ |x| + , then |fn (x) − f (x)| ≤ ,
n n n
which proves that the convergence of (fn ) to f is uniform on R. However, the function
f is not differentiable at the origin!
In other words, the limit of a sequence (even uniformly convergent) of differentiable
functions is not necessarily differentiable. In this subsection, we will establish a sufficient
(but not necessary) condition for the limit of a sequence (fn ) of differentiable functions
to be differentiable.
Then the sequence (fn )n converges uniformly to a function f of class C 1 , and we have
f ′ = g. In other words, ′
′
lim fn (x) = lim f (x) .
n→+∞ n→+∞
uniformly on [a, b]. Therefore, the sequence (fn )n converges uniformly to f defined by
Z x
f (x) = lim fn (x0 ) + g(t)dt.
n→+∞ x0
ln 1 + nx2
fn (x) = , x ∈ [1, 2].
2n
Here fn (1) converges to 0 , and the sequence of derivatives
2nx x
fn′ (x) = 2
= , x ∈ [1, 2]
2n (1 + nx ) 1 + nx2
51
converges pointwise to g(x) = 0, x ∈ [0, 2].
The convergence of fn′ to g is uniform on [1, 2] because |x| ≤ 2 and 1 + nx2 ≥ n implies
x |x| 2
2
−0 = 2
≤
1 + nx 1 + nx n
Then by the above Theorem, the sequence (fn ) converges uniformly (something not
easily proved directly) and the limit function f = lim fn is differentiable with f ′ = g.
Here f (x) = 0 for x ∈ [1, 2] whose derivative is 0 .
Example 2.1.23. Let a > 0 be a fixed number, and denote by (fn )n the sequence of
functions defined by
fn (x) = e−nx , n ∈ N, x ≥ a
For x ≥ a, we have
fn′ (x) = −ne−nx ≤ ne−na .
Since the sequence (ne−na )n tends to zero, this shows the uniform convergence of (fn′ )n
to zero. We have
lim fn′ (x) = lim f ′ (x) = 0
n→+∞ n→+∞
As with numerical series, from a sequence of functions (fn )n , one can construct a
sequence (Sn )n where
Sn = f0 + . . . + fn .
P
This gives what is called a series of functions, denoted as fn .
Definition 2.2.1. Given (fn )n a sequence of functions defined on a set A and taking
real values. We call the series of functions the expression:
+∞
X
f0 (x) + f1 (x) + f2 (x) + · · · + fn (x) + · · · = fn (x), x ∈ A.
n=0
Define
n
X
Sn (x) = fk (x).
k=0
P
The function Sn is referred to as the n-th partial sum of the series fn .
52
2.2.1 Pointwise convergence
P
Definition 2.2.2. We say that the series of functions n≥0 fn converges pointwise on
A if and only if the sequence (Sn )n of partial sums converges pointwise on A. In this
P
case, the limit function S of (Sn )n is called the sum function of the series fn , and it
is denoted as
+∞
X
S= fn .
n=0
• If x ̸= 0, we have
1
n
X x2 1− (1+x2 )n+1
Sn (x) = = x2 1
(1 +
k=0 x2 )k 1 − 1+x 2
2
1
= 1+x 1− .
(1 + x2 )n+1
1 1
Since 1+x2
< 1, then lim n+1 = 0, and so
n→+∞ (1+x2 )
lim Sn (x) = 1 + x2 .
n→+∞
+∞
( )
X
D= x∈A: un (x) converges
n=0
P+∞
The domain of pointwise convergence of the series n=0 un is called D. The set D can
be empty or coincide with A.
P+∞
Example 2.2.5. Study the pointwise convergence of the series of functions n=0 un ,
where
un : R → R
2 x2
x 7→ x2 e−n , n ∈ N.
Let x be fixed.
53
• If x ∈ R∗ , then
p
n 1 2
lim x2 e−n2 x2 = lim x2 n
e−nx = 0 < 1,
n→+∞ n→+∞
P+∞
which shows that the series n=0 un is absolutely
convergent (we can also remark
2 2 2 −n2 x2
that n un (x) = n x e −→ 0, n → +∞ .
P+∞
• If x = 0, the series n=0 un (0) = 0 converges.
Then +∞
P
n=0 un converges pointwise on D = R.
54
2.2.3 Uniform convergence
P
Definition 2.2.11. We say that the series fn converges uniformly on A if the se-
quence of functions (Sn )n converges uniformly on A.
xn 1
Since ∀x ∈ [−1, 1] : n+1 ≤ n+1 → 0, n → +∞, then
xn
lim Sn (x) = S(x) = lim 1− =1
n→+∞ n→+∞ n+1
Let’s study the uniform convergence of the sequence (Sn )n on [−1, 1]. Since
xn
sup |Sn (x) − S(x)| = sup 1− −1
x∈[−1,1] x∈[−1,1] n+1
|x|n 1
= sup ≤ −→ 0, n → +∞,
x∈[−1,1] n + 1 n+1
This shows that the sequence (Rn )n of remainders converges uniformly to zero on A.
55
is uniformly convergent on Ba = [−a, a] (where 0 < a < 1) because
+∞
X xn+1
dn = sup |Rn (x)| = sup xk = sup
x∈[−a,a] x∈[−a,a] k=n+1 x∈[−a,a] 1−x
|x|n+1 an+1
= sup ≤ −→ 0, n → +∞.
x∈[−a,a] 1 − x 1−a
Proof. This is nothing but the uniform Cauchy criterion applied to the sequence of
partial sums (Sn ).
k
we consider x = 1/2n. Then, for all 0 < 2n ≤ 1 < π2 , we have
k 2 k
sin ≥ .
2n π 2n
Hence,
2n 2n 2n
X sin(kx) X 1 k X 1 2 k
α
= α
sin ≥
k k 2n k α π 2n
k=n+1 k=n+1 k=n+1
2n 2n
1 X 1 1 X 1 1
= = k 1−α ≥ · n(1 + n)1−α ≥ .
nπ k α−1 nπ nπ π
k=n+1 k=n+1
P+∞ sin(nx)
This shows that the series < α ≤ 1) does not satisfy the Cauchy
n=1 nα , (0
criterion for uniform convergence on [−π, π].
56
2.2.5 Uniform and Absolute Convergence
P∞
Theorem 2.2.18. If a series n=1 |fn | converges uniformly on A, then the series
P∞
n=1 fn converges uniformly and absolutely on A.
P∞
Proof. If a series n=1 |fn | converges uniformly on A, the absolute convergence of
P∞
n=1 fn at every fixed point of A follows directly from the properties of series of
numbers. To show the uniform convergence we use the Cauchy criterion:
m
X m
X
fk (x) ≤ |fk (x)| < ε
k=n+1 k=n+1
for all x ∈ X at once and for ∀m > n > N , where N is from the Cauchy criterion for
the series ∞
P
n=1 |fn |.
is convergent.
• ∀n ∈ N, ∀x ∈ A, |fn (x)| ≤ αn ;
• The series
P
αn converges.
P
Proof. If the series fn is normally convergent, then the sequence with general term
αn = ∥fn ∥∞ satisfies the two conditions of the previous proposition.
Conversely, if the two conditions of the theorem are satisfied, then each fn is bounded,
P
and the series ∥fn ∥∞ converges as a series of positively bounded terms, being domi-
nated by the sum of a convergent series.
57
Example 2.2.22. Consider the series
+∞
X sin nx
, x ∈ R.
n!
n=0
We have
sin nx 1
∀n ∈ N, ∀x ∈ R : ≤ ,
n! n!
and since +∞
P 1 P+∞ sin nx
n=0 n! converges (according to the ratio test), then n=0 n! is normally
convergent over R.
P
Theorem 2.2.23. If the series fn converges normally on A, then it converges ab-
solutely and uniformly on A, and we have
+∞
X +∞
X
fn ≤ ∥fn ∥∞ .
n=0 ∞ n=0
P
Proof. If the series fn converges normally, there exists a positive real sequence (αn )
such that
∀n ∈ N, ∀x ∈ A,
|fn (x)| ≤ αn .
P P
This, together with the convergence of the series αn . The series fn (x) therefore
P
converges absolutely on A. Hence, the absolute convergence of fn on A.
Now, we have
+∞
X +∞
X +∞
X
dn = sup fk (x) ≤ sup |fk (x)| ≤ αk → 0, n → +∞.
x∈A k=n+1 x∈A k=n+1 k=n+1
P+∞
This shows that the series n=0 un is uniformly convergent on A.
set A ⊂ R. Then for any m > n ≥ 1 and for each x ∈ X the following formula is true
m
X m−1
X
ak bk = Bk (ak − ak+1 ) + Bm am − Bn an+1
k=n+1 k=n+1
P∞
Theorem 2.2.25 (Dirihlet-Abel’s test 1). Let n=1 fn be a series of functions defined
on a set A ⊂ R with fn (x) = an (x)bn (x). If
58
(1) ∃M such that
n
X
|Bn (x)| = bk (x) ≤ M,
k=1
for all n and for all x ∈ A,
(2) The sequence (an )n is monotone at every fixed x ∈ A,
(3) The sequence (an )n converges uniformly to 0 on A.
P∞
Then the series n=1 an bn is uniformly convergent on A.
m
X m−1
X
|S| = ak bk = Bk (ak − ak+1 ) + Bm am − Bn an+1
k=n+1 k=n+1
m−1
X
≤ |Bk | · |ak − ak+1 | + |Bm | · |am | + |Bn | · |an+1 | ≡ S1 .
k=n+1
By the condition (1) of the Theorem, the partial sums Bn are uniformly bounded, that
is, there exists M > 0 such that |Bn | ≤ M for ∀n ∈ N and for all x ∈ A. Then, we can
proceed with the evaluation of S1 in the following way:
m−1
!
X
S1 ≤ M |ak − ak+1 | + |am | + |an+1 | ≡ S2 .
k=n+1
Since the sequence an is monotone at each x ∈ A, all the differences ak − ak+1 have the
same sign, and consequently, the sum of the absolute values is equal to the absolute
value of the sum:
m−1
!
X
S2 = M (ak − ak+1 ) + |am | + |an+1 |
k=n+1
A
Now we use the condition of the uniform convergence of an : an ⇒ 0. By definition,
n→∞
ε
this means that for ∀ε > 0 (take for convenience ε1 = 4M ) there exists Nε such that
ε
for ∀n > Nε and simultaneously for all x ∈ A it follows that |an | < 4M . Choosing for
the evaluation of the sum S the indices ∀m > n > Nε , we have
m ε
X ε
|S| = ak bk ≤ 2M (|an+1 | + |am |) < 2M + = ε.
4M 4M
k=n+1
Hence, according to the Cauchy criterion for uniform convergence, the series
P∞
n=1 an (x)bn (x) converges uniformly on A.
59
P∞
Theorem 2.2.26 (Dirihlet-Abel’s test 2). Let n=1 fn be a series of functions defined
on a set A ⊂ R with fn (x) = an (x)bn (x). If
P∞
(1) the series n=1 bn converges uniformly on A,
(2) The sequence (an )n is monotone for every fixed x ∈ A,
(3) ∃M such that |an (x)| ≤ M, for all n and for all x ∈ A,
P∞
Then the series n=1 an bn is uniformly convergent on A.
Since the sequence an is monotone at each x ∈ A, all the differences ak − ak+1 have the
same sign, and consequently, the sum of the absolute values is equal to the absolute
value of the sum:
m−1
!
X
S2 = ε1 (ak − ak+1 ) + |am |
k=n+1
60
Finally, we employ the condition of the uniform boundedness of an : |an | ≤ M , for
ε
∀n ∈ N and for all x ∈ A, and specify ε1 = 3M to arrive at the inequality
m
X ε ε
|S| = ak bk < (|an+1 | + 2 |am |) ≤ · 3M = ε.
3M 3M
k=n+1
Let us solve some examples with application of Dirichlet’s and Abel’s tests.
First, let us apply Dirichlet’s test. For this purpose, choose an = n1 and bn (x) =
(−1)n xn . Then, the series ∞
P P∞ n n
n=1 bn (x) = n=1 (−1) x has the partial sums bounded
uniformly on [0, 1] :
n
X 1 − (−x)n x
|Bn (x)| = (−1)k xk = −x ≤2 < 2, ∀n, ∀x ∈ [0, 1]
1+x 1+x
k=1
converges uniformly over R according to the Leibniz rule. Indeed, if we define un (x) =
√ 1 for n ∈ N and x ∈ R, then we have:
2 n+sin 3x
• ∀n ∈ N, ∀x ∈ R : un (x) = √ 1
2 n+sin 3x
>0
61
• ∀x ∈ R fixed:
1 1
un+1 (x) = √ ≤ un (x) = √ ,n ∈ N
2 n + 1 + sin 3x 2 n + sin 3x
1 1
|un (x)| = √ ≤ √ → 0, n → +∞.
2 n + sin 3x 2 n−1
(1) ∀n ∈ N : fn ∈ C(A);
(2) The series +∞
P
n=0 fn (x) = S(x) converges uniformly over the set A.
Then S ∈ C(A)
Remark. Under the same assumptions as in Theorem 2.2.30, we can thus write:
+∞
X +∞
X +∞
X
lim fn (x) = lim fn (x) = fn (x0 ) .
x→x0 x→x0
n=0 n=0 n=0
In addition,
e−n|x|
Example 2.2.31. The function x 7→ +∞
P
n=1 n2 is continuous over R. Indeed, it is
the sum of a series of continuous functions that converges normally (hence uniformly)
over R since
e−n|x| 1
∀x ∈ R, 2
≤ 2.
n n
+∞ +∞
X 1 X 1
lim S(x) = lim 2 2
= lim 2 = 0.
x→∞ x→∞ x + (n + 1) x→∞ x + (n + 1)2
n=0 n=0
62
Example 2.2.33. The series
+∞
X x
(1 + x2 )n
n=0
x
but not uniformly over R because each un (x) = (1+x 2 )n (where n ∈ N) is continuous at
1 + x2
lim S(x) = lim = ∞ and S(0) = 0
x→0 x→0 x
tn e−t
un : [0, x] → R, t 7→ .
n!
Since
tn e−t |x|n e−|x|
sup |un (t)| = sup ≤
t∈[0,x] t∈[0,x] n! n!
|x|n e−|x|
and the series +∞ converges, then the series +∞
P P
n=0 n! n=0 un (t) converges normally,
thus uniformly over [0, x]. According to Theorem 2.2.34, we can write:
+∞ +∞ n
!
1 x n −t
Z Z x X Z x
X t −t
t e dt = e dt = et e−t dt = x.
n! 0 0 n! 0
n=0 n=0
63
Example 2.2.36. Show that
+∞
# #
X (−1)n+1
ln(1 + x) = xn , x ∈ − 1, 1 .
n
n=1
This series converges uniformly over [−α, α] (where 0 < α < 1) because
∀n ∈ N, ∀t ∈ [−α, α] : |(−1)n tn | ≤ αn
lim ln(1 + x) = ln 2
x→1
Thus
+∞
# #
X (−1)n+1 n
ln(1 + x) = x , x ∈ − 1, 1 .
n
n=1
64
(2) ∀n ∈ N : fn ∈ C 1 ([a, b])
(3) The series +∞ ′
P
n=0 fn (x) converges uniformly on [a, b].
converges uniformly on [a, b], and its sum S ∈ C 1 ([a, b]), and we have
+∞
! +∞
d X X
fn (x) = fn′ (x), x ∈ [a, b].
dx
n=0 n=0
+∞
X (−1)n
S(x) = x
n+x
n=1
P+∞ (−1)n
According to Leibniz’s rule, the series S1 (x) = n=1 n+x converges uniformly for
∗ (−1)n
every x ∈ ] 0, +∞ [ . For all n ∈ N , gn (x) = n+x is of class C 1 on ]0, +∞[, and
(−1)n+1
∀x ∈]0, +∞ : gn′ (x) =
(n + x)2
We have ∀n ∈ N∗ , ∀x ∈] 0, +∞[ :
(−1)n+1 1 1
gn′ (x) = = ≤ 2
(n + x)2 (n + x)2 n
As +∞
P 1 P+∞ ′
n=1 n2 converges, we deduce that n=1 gn (x) converges normally, hence uniformly
on ]0, +∞[. According to Theorem 2.2.11, we conclude that S1 is of class C 1 on ]0, +∞[
and consequently S(x) = xS1 (x) is also of class C 1 on ]0, +∞[. Moreover,
+∞
"
X (−1)n n
∀x ∈]0, +∞ : S ′ (x) = S1 (x) + xS1′ (x) =
(n + x)2
n=1
65
Chapter 3
Power Series
Power series are one of the most useful type of series in analysis. For example, we can
use them to define transcendental functions such as the exponential and trigonometric
functions (and many other less familiar functions).
3.1 Introduction
66
and here is a power series centered at 1
∞
X (−1)n+1 1 1 1
(x − 1)n = (x − 1) − (x − 1)2 + (x − 1)3 − (x − 1)4 + . . .
n 2 3 4
n=1
Note that any series (3.1.1) can be brought to this form through the change of variables
x − x0 → y → x.
lim an xn0 = 0.
n→+∞
∀n ∈ N : |an xn0 | ≤ M
67
x
as < 1, which completes the proof of the first assertion.
x0
P+∞ n
Theorem 3.1.4. If the power series n=0 an x converges on the interval ] − a, a[
(a > 0), then it converges normally on every interval [−r, r], where r < a.
Proof. Let’s fix R, with r < R < a. There exists M > 0, such that
∀n ∈ N : |an Rn | ≤ M.
xn x n x n r n
|an xn | = an Rn n
= |an Rn | ≤M ≤M ,
R R R R
Before we introduce the concept of the radius of convergence, we present the following
result.
be a power series. Then, there exists one and only one number R such that
Proof. Let E be the set of non-negative real numbers r ≥ 0 such that the sequence
(|an | rn )n is bounded. The set E is not empty since it contains at least 0. If E is not
upper-bounded, we set R = +∞. If E is upper-bounded, we set R = sup E, which is the
least upper bound of the set E (the existence of sup E is guaranteed by the supremum
axiom).
68
• Suppose |x| < R. Then there exists r0 ∈ E such that |x| < r0 < R. The sequence
(|an | r0n ) is bounded. If, in Abel’s Lemma, we take x0 = r0 , we see that the series
an xn converges absolutely.
P
• Let’s suppose |x| > R. Then |x| is not in the set E. Therefore, the sequence (an xn )
an xn diverges because its
P
is unbounded. We deduce that the numerical series
general term does not tend towards 0.
Let’s prove the uniqueness of R. If there existed R and R′ such that 0 ≤ R < R′ ≤ +∞,
both satisfying properties 1) and 2), then for a chosen number r such that R < r < R′ ,
an rn would have to both converge and diverge, which is absurd.
P
the series
The following definition therefore makes sense for every power series.
converges for |x − c| < R and diverges for |x − c| > R, then 0 ≤ R ≤ ∞ is called the
radius of convergence of the power series.
Theorem 3.2.1 does not say what happens at the endpoints x = c ± R, and in general
the power series may converge or diverge there. We refer to the set of all points where
the power series converges as its interval of convergence, which is one of
In addition, Theorem 3.2.1 does not give an explicit expression for the radius of con-
vergence of a power series in terms of its coefficients. The ratio test gives a simple, but
useful, way to compute the radius of convergence, although it doesn’t apply to every
power series.
Theorem 3.2.4 (Ratio test). Suppose that an ̸= 0 for all sufficiently large n and
the limit
an
R = lim
n→∞ an+1
69
Proof. Let
an+1 (x − c)n+1 an+1
r = lim = |x − c| lim .
n→∞ an (x − c)n n→∞ an
By the ratio test, the power series converges if 0 ≤ r < 1, or |x − c| < R, and diverges
if 1 < r ≤ ∞, or |x − c| > R, which proves the result.
an nn 1 1
= = ,
an+1 (n + 1) n+1 n + 1 1 + n1 n
So,
an 1
lim = 0 · = 0.
n→+∞ an+1 e
Therefore, R = 0 and ∆ = {0}.
Remark. If the sequence with the general term an+1 an has no limit in R+ , the rule of
D’Alembert is inapplicable. This is, for example, the case for the power series
X
(2 + (−1)n ) xn .
(x − 2)2n−1
un (x) = ,
n3
we have
un+1 (x) (x − 2)2n+1 n3 2 n3
= = (x − 2) → (x − 2)2 , n → +∞
un (x) (x − 2)2n−1 (n + 1)3 (n + 1)3
70
For x = 1 , we obtain the series
+∞
X −1
n3
n=1
which is convergent.
For x = 3, we obtain the convergent series
+∞
X 1
.
n3
n=1
The root test gives an expression for the radius of convergence of a general power series.
is given by
0, ρ = +∞
R = ρ1 , 0 < ρ < +∞
+∞, ρ = 0
where
p
n
ρ = lim |an |,
n→+∞
Proof. Let
r = lim sup |an (x − c)n |1/n = |x − c| lim sup |an |1/n
n→∞ n→∞
By the root test, the series converges if 0 ≤ r < 1, or |x − c| < R, and diverges if
1 < r ≤ ∞, or |x − c| > R, which proves the result.
We have
1 √ 1 √
an = 5n
⇒ n an = 5
⇒ lim n an = 0
(ln n) (ln n) n→+∞
71
To calculate the radius of convergence, we will use the Cauchy-Hadamard rule. We have
n )n
an = (3−(−1)
ln n , and
p
n 3 − (−1)n √ 2 √ 4
|an | = √ n
; 2k √
a2k = 2k ; 2k+1 a2k+1 = 2k+1
p .
ln n ln 2k ln(2k + 1)
Hence
p
n
lim |an | = 4.
n→+∞
If we assume
( (
1 1
0, n = 2k, 22k
· ln(2k) , n = 2k,
cn = 1 ∗
, dn =
ln(2k+1) , n = 2k + 1 (k ∈ N ) 0, n = 2k + 1 (k ∈ N∗ ) ,
Then the series converges normally (hence absolutely and uniformly) in each interval
[a, b] ⊂] − R, R[.
72
Proof. Follows from the definition of the convergence radius. Indeed, let r ∈ R+ such
that the sequence (bn rn ) is bounded. Since |an rn | ≤ |bn rn |, the sequence (an rn ) is also
bounded. Thus, we have the inequality: Ra ≥ Rb .
Example 3.3.2. For every n ∈ N, we have e−1 ≤ ecos n ≤ e. Since the power series
P −1 n P n P cos n n
e x and ex both have a convergence radius equal to 1, the series e z
also has a convergence radius equal to 1.
Proof. By definition, there exists M > 0 and a rank N such that for all n ≥ N ,
M bn xn has a convergence radius of Rb . The previous
P
|an | ≤ M |bn |. The power series
theorem allows us to conclude that indeed Ra ≥ Rb .
Proof. There exist two non-zero polynomials P and Q such that F = P/Q. Denoting
λX α (resp. µX β ) as the highest-degree term of P (resp. Q), we have λµ ̸= 0, and
λ α−β
|an | ∼ n .
n→+∞ µ
Since α−β
(n + 1)α−β
1
= 1+ −→ 1
nα−β n n→+∞
an xn is also equal to 1
P
is equal to 1 according to the ratio test. Therefore, that of
according to the previous proposition.
73
• Product series (also called Cauchy product), the power series cn xn defined by
P
n
X
∀n ∈ N, cn = ak bn−k
k=0
(1) The convergence radius R of the sum series satisfies R ≥ min (Ra , Rb ), with equal-
ity if Ra ̸= Rb . Moreover, if |x| < min (Ra , Rb ), then
+∞
X +∞
X +∞
X
n n
an x + bn x = (an + bn ) xn
n=0 n=0 n=0
(3) the convergence radius R of the product series satisfies R ≥ min (Ra , Rb ). More-
over, if |x| < min (Ra , Rb ), then
+∞
! +∞ ! +∞
X X X
n n
an x bn x = cn xn
n=0 n=0 n=0
where
n
X
∀n ∈ N, cn = ak bn−k .
k=0
Proof. (1) If |z| < Ra and |z| < Rb , both power series converge absolutely. The sum
power series also converges, as the triangle inequality yields
74
Remark. For the convergence radius R of the sum series and the product series, it is
P n P n
possible to have R > max (Ra , Rb ). For instance, the series z and −z both have
a convergence radius of 1, but the sum series is the zero series, hence with a convergence
an xn and bn xn
P P
radius equal to +∞. Similarly, consider the series
( (
2 if n = 0 −1 if n = 0
an = n
and bn =
2 if n ≥ 1 1 if n ≥ 1
For n ≥ 1, we have
1
|an |1/n = 2 and |bn |1/n = 1, hence Ra = and Rb = 1.
2
cn z n is defined by
P
The product series
n−1
X n−1
X
2k − 2n = 2 + 2 2n−1 − 1 − 2n
cn = a0 bn + ak bn−k = 2 +
k=1 k=1
so cn = 0 for all n ≥ 0. Therefore, the product series is the zero series, and its conver-
gence radius is equal to +∞.
Proof. Let R and R′ denote the convergence radii of the power series an z n and
P
(n + 1)an+1 z n respectively.
P
If |z| > R, then the sequence (an z n ) is unbounded. Consequently, the sequence
((n + 1)an+1 z n ) is also unbounded, implying |z| ≥ R′ . Therefore, R ≥ R′ .
If |z| > R′ , let’s choose a number ρ such that |z| > ρ > R′ . The sequence
ρ |z| n
n n−1
|an z | = n |an | ρ ×
n ρ
is the product of two sequences with positive terms, one of which is unbounded (since
ρ > R′ ), and the other tends to +∞ (as |z| > ρ). Consequently, the sequence (|an z n |)
is unbounded, implying |z| ≥ R. Therefore, R′ ≥ R.
Remark. By applying the previous result to a primitive power series, that is,
P an n+1 P an n+1
an z n , we observe that
P
n+1 z instead of n+1 z has the same convergence
n
P
radius as an z .
75
3.5 Properties of the sum of a power series
Proof. The series is uniformly convergent on every interval [a, b] ⊂] − R, R[. Hence, the
sum of the power series is continuous within the interval of convergence ] − R, R[.
Example 3.5.2. The ratio test easily shows that the series (z n /n!) has a convergence
P
P n
radius equal to +∞. Therefore, its sum function z 7→ (z /n!) is continuous on R.
Corollary 3.5.3. Let R > 0 be the convergence radius of the power series
+∞
X
an xn .
n=0
If this series converges for x = R, then its sum is left-continuous at this point, and
+∞
X +∞
X
n
lim an x = an Rn
x→R−
n=0 n=0
Proof. We have
+∞
X +∞
X x n
an xn = an R n · .
R
n=0 n=0
In particular,
Z x +∞
X an n+1
∀x ∈] − R, R[: S(t)dt = x .
0 n+1
n=0
Proof. Trivial!
76
Remark. On the convergence interval, one can integrate a power series infinitely many
times, and the resulting series maintains the same convergence radius.
We have
+∞ +∞ +∞ Z x
X (−1)n X (−1)n x2n+1 X
= lim = lim (−1)n t2n dt
2n + 1 x→1− 2n + 1 x→1− 0
n=0 n=0 n=0
+∞
!
Z x Z x
X
n 2n 1 π
= lim (−1) t dt = lim 2
dt = lim arctan x = .
x→1− 0 x→1− 0 1+t x→1 − 4
n=0
Corollary 3.5.7. Under the assumptions of the above theorem, the sum function S is
of class C ∞ on ] − R, R[, and
+∞
(k)
X k!
∀k ∈ N, ∀x ∈] − R, R[, S (x) = an xn−k .
(n − k)!
n=k
of 1, and
+∞
X 1
∀x ∈] − 1, 1[, xn = .
1−x
n=0
77
By immediate induction on k, we deduce that
+∞
!
1 dk 1 dk
1 1 X
= = xn
(1 − x)k+1 k! dxk 1−x k! dxk
n=0
+∞ +∞
1 X X (n + k)! n
= (n + k) . . . (n + 1)xn = x
k! k!n!
n=0 n=0
and
∞
X
g(x) = bk (x − c)k
0
agree on some interval centered at c, that is f (x) = g(x) for |x − c| < ρ and some
positive ρ. Then ak = bk for all k = 0, 1, 2, . . ..
reveals one of the key facts about the exponential function, namely that it is its own
derivative. Note simply that
∞ ∞ ∞
d x d X xk X d xk X xk−1
e = = = = ex .
dx dx k! dx k! k − 1!
k=0 k=0 k=1
In this Section we examine how functions may be expressed in terms of power series.
This is an extremely useful way of expressing a function since (as we shall see) we
can then replace ’complicated’ functions in terms of ’simple’ polynomials. The only
requirement (of any significance) is that the ’complicated’ function should be smooth;
this means that at a point of interest, it must be possible to differentiate the function
as often as we please.
78
Definition 3.6.1. Let f be a function of class C ∞ in a neighborhood of x0 . The power
series
+∞ (n)
X f (x0 )
(x − x0 )n
n!
n=0
f (n) (x0 )
∀n ∈ N : an = .
n!
We can thus state
P+∞
Theorem 3.6.2. Every power series n=0 an (x − x0 )n is the Taylor series of its sum
in the neighborhood of x0 .
The question is: under what conditions can we express the function f as a Taylor series?
In other words, what are the conditions under which we can write
∞
X f (k) (x0 )
f (x) = (x − x0 )k
k!
k=0
However, the converse is not generally true; a function f possessing derivatives of every
order at x0 is not necessarily equal to the corresponding power series
∞
X f (k) (x0 )
(x − x0 )k
k!
k=0
The following example shows that the Taylor series for f (x) converges, but its sum
differs from f (x) at every point, except the central one. For example, one can consider
the function ( 2
e−1/x , x ̸= 0,
f (x) =
0, x = 0.
This function is of class C ∞ in R. Indeed, for x ̸= 0, we have
2 2 6 4 2
f ′ (x) = 3 e−1/x , f ′′ (x) = − 4 + 6 e−1/x , . . .
x x x
79
By induction, we have
1 2
f (n)
(x) = Pn e−1/x (n = 1, 2, . . .),
x
where Pn x1 is a polynomial of degree 3n. Let’s show that all derivatives of this function
Remark. One can also use the integral form of the remainder in the Taylor expansion,
that is,
1 x
Z
Rn (x) = (x − t)n f (n+1) (t)dt
n! x0
80
Theorem 3.6.4. If f (x) is infinitely differentiable on an interval (−R, R), R > 0 and
its derivatives satisfy the evaluations
n!
f (n) (x) ≤ M , ∀x ∈ (−R, R), ∀n ∈ N, (3.6.1)
Rn
where M is a constant, then f (x) is analytic on (−R, R).
Proof. It suffices to show that lim Rn (x) = 0. Using the remainder (at a = 0 ) in the
n→∞
f (n+1) (c) n+1
Lagrange form Rn (x) = (n+1)! x , we obtain for every fixed x in (−R, R) :
n+1
f (n+1) (c) n+1
|x|
|Rn (x)| = x ≤M .
(n + 1)! R
n+1
Since |x|
R < 1, ∀x ∈ (−R, R), it follows that
|x|
R → 0, which implies that
n→∞
Rn (x) → 0, ∀x ∈ (−R, R).
n→∞
Remark. As usual, inequalities (3.6.1) can be imposed starting from some index n-
the violation of these inequalities for a finite number of derivatives does not affect the
resulting convergence of Rn .
3.6.1 Discussion
In practice, weaker versions of Theorem3.6.4 are frequently used, which impose stronger
restrictions on derivative. One of these formulations uses the condition
It is easy to show that the last restriction guarantees inequality (3.6.2), which in turn
guarantees condition (3.6.1) of Theorem 3.6.4. Indeed, without a loss of generality, we
can consider the constant q > 1 in (3.6.2) (if (3.6.2) holds for q ≤ 1, it is also valid
for q > 1 ). Then M < M q n , that is, (3.6.2) is more restrictive than (3.6.3). At the
next step, we compare q n and Rn!n . Notice that for any two constants q and R the ratio
qn (qR)n
n! = n! approaches 0 . To show this, we choose a fixed index N = [qR] and use the
Rn
representation n = N + k (n > N ) to obtain
k
(qR)n (qR)N +k (qR)N (qR)k (qR)N
qR
= = · < · → 0
n! (N + k)! N! (N + 1) · . . . · (N + k) N! N +1 k→∞
81
The hypothesis of the previous theorem is satisfied; thus, we have
x3 x5 x2n+1
sin x = x − + − · · · + (−1)n + ··· , x ∈ R.
3! 5! (2n + 1)!
x x2 xn
f (x) = 1 + + + ··· + + Rn (x)
1! 2! n!
where
eθx
Rn (x) = xn+1 , 0 < θ < 1
(n + 1)!
for any fixed x ∈ R
e|x|
|Rn (x)| ≤ |x|n+1 → 0, n → +∞
(n + 1)!
Hence,
x x2 xn
∀x ∈ R : ex = 1 + + + ··· + + ···
1! 2! n!
Similarly,
x x2 xn
∀x ∈ R : e−x = 1 − + + · · · + (−1)n + ···
1! 2! n!
and as a conclusion
ex − e−x x3 x5 x2n+1
∀x ∈ R : sinh x = =x+ + − ··· + + ··· ,
2 3! 5! (2n + 1)!
ex + e−x x2 x4 x2n
∀x ∈ R : cosh x = =1+ + + ··· + + ···
2 2! 4! (2n)!
82