analysis
analysis
Analysis I MATH254
Course Outline:
Fundamentals of set theory. Properties of the reals. Limits, limsup, liminf. Continuity. Functions. Differentiation.
References:
Understanding Analysis, Abbott; Introduction to Real Analysis, Bartle; Analysis I, Tao
Based on Lectures from Fall, 2023 by Prof. Vojkan Jaksic.
Contents
2 Sequences 27
2.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 Properties of Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.3 Limit Superior, Inferior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.4 Subsequences and Bolzano-Weirestrass Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.5 Cauchy Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.6 Contractive Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.7 Euler’s Number e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.8 Limit Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.9 Properly Divergent Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4 Differentiation 93
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.2 The Chain Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.3 Derivative of the Inverse Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5 Appendix 98
5.1 Interesting Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
1 Logic, Sets, and Functions
(1) 1 ∈ N 1 1
using 0 instead of 1 is also
valid, but we will use 1 here,
and throughout the rest of
(2) every natural number has a successor in N course.
(a) 1 ∈ S
(b) if n ∈ S, then n + 1 ∈ S 3
then S = N.
3
(a) is called the inductive
base; (b) the inductive step.
⊛ Example 1.1 All AI restatements are
equivalent in having both of
Prove that, for every n ∈ N, 1 + 2 + · · · + n = n(n+1)
2
(≡ (1)) these, and only differentiate
on their specific values.
Proof (via AI.i). Let S be the subset of N for which (1) holds; thus, our goal is to
show S = N, and we must prove (a) and (b) of AI.i.
§1.1 Logic, Sets, and Functions: Mathematical Induction & The Naturals p. 3
to both sides yields:
n(n + 1)
1 + 2 + · · · + n + (n + 1) = + (n + 1) (1)
2
n
= (n + 1)( + 1) (2)
2
(n + 1)(n + 2)
= (3)
2
(n + 1)((n + 1) + 1)
= (4)
2
Line (4) is equivalent to statement (1) (substituting n for n + 1), and thus if n ∈ S,
then n + 1 ∈ S and (b) holds. Thus, by AI.i, S = N and 1 + 2 + · · · + n = n(n+1)
2
holds ∀ n ∈ N. ■
⊛ Example 1.2
h i2
Prove (by induction), that for every n ∈ N, 13 + 23 + · · · + n3 = n(n+1)
2
.
Proof. Follows a similar structure to the previous example. Let S be the subset of N
for which the statement holds. 1 ∈ S by inspection ((a) holds), and we prove (b) by
assuming n ∈ S and showing n + 1 ∈ S (algebraically). Thus, by AI.i, S = N and
the statement holds ∀ n ∈ N. ■
(a) m ∈ S
(b) n ∈ S =⇒ n + 1 ∈ S
then {m, m + 1, m + 2, . . . } ⊆ S.
§1.1 Logic, Sets, and Functions: Mathematical Induction & The Naturals p. 4
⊛ Example 1.3
Using AI.ii, prove that for n ≥ 2, n2 > n + 1.
Proof. Let S ⊆ N be the set of n for which the statement holds. n = 2 =⇒ 4 > 3,
so the base case holds. Consider n2 > n + 1 for some n ≥ 2. Then, (n + 1)2 =
n2 + 2n + 1 > n + 1 + 2n + 1 = 3n + 2 > 2n + 2 > n + 2, hence S = {2, 3, 4, · · · }
(all n ≥ 2). ■
(a) 1 ∈ S
(b) if 1, 2, . . . , n − 1 ∈ S, then n ∈ S
then S = N.
(a) m ∈ S
(b) if m, m + 1, . . . , m + n ∈ S, then m + n + 1 ∈ S
then {m, m + 1, m + 2, . . . } ⊆ S.
Proof of theorem 1.1. Let S be the set of all natural numbers that can be written as a product
of one or more primes. We will use AI.iv to show S = {2, 3, . . . }.
§1.1 Logic, Sets, and Functions: Mathematical Induction & The Naturals p. 5
– if 2 + (n + 1) is prime, then 2 + (n + 1) ∈ S, as all primes are products of 1 and
themselves and are thus in S by definition.
Within Z, we can define multiplication, addition and subtraction, with the neutrals of
1 and 0, respectively. However, we cannot define division, as we are not guaranteed a
quotient in Z. This necessitates the rationals, Q. We define
p
Q = { : p ∈ Z, q ∈ Z, q ̸= 0}.
q
• x < y =⇒ x + z < y + z
Q, together with its operations and relations above, is called an ordered field.
We can consider historical reasoning for the extension of Q to R. Consider a right triangle
of legs a, b and hypotenuse c. By the Pythagorean Theorem, a2 + b2 = c2 . Consider further
the case there a = b = 1, and thus c2 = 2. Does c exist in Q?
p
c=
q
p2
c2 = 2 = 2
q
2q 2 = p2
T∞
•
T
i=1 An = n∈N An = {x : x ∈ An ∀ n ∈ N}
/ A}7
• AC = {x : x ∈ X and x ∈ 7
X is often omitted if it is clear
from context.
(a) (A ∩ B)C = AC ∪ B C
and
(b) (A ∪ B)C = AC ∩ B C .
,→ Proposition 1.2
∞
!C ∞
\ [
(a) An = AC
n
n=1 n=1
∞
!C ∞
[ \
(b) An = AC
n
n=1 n=1
Proof of proposition 1.2. Consider Proposition (b). Working from the left-hand side, we have
∞
!C
[ [
An = {x : x ∈
/ An }
n=1
= {x : x ∈
/ An ∀ n ∈ N}
\
= {x : x ∈ / An }
\
= ACn
(a) can be logically deduced from this result. Consider the RHS, n . Taking the comple-
AC
S
ment:
C
via (b)
[ \ C
AC
n = AC
n
\
= An
1.4 Functions
,→ Definition 1.1
Let A, B be sets. A function f is a rule assigned to each x ∈ A a corresponding unique
element f (x) ∈ B. We denote
f : A → B.
,→ Definition 1.2
The domain of a function f : A → B, denoted Dom(f ) = A. The range of f , denoted
⊛ Example 1.4
The function f (x) = sin x, f : R → [−1, 1]. Here, Dom(f ) = R, and Ran(f ) =
[−1, 1].
,→ Proposition 1.3
Let f : A → B, C ⊆ A, f (C) = {f (x) : x ∈ C}. We claim f (C1 ∪ C2 ) =
f (C1 ) ∪ f (C2 ).
(2) y ∈ f (C1 ) ∪ f (C2 ) =⇒ y ∈ f (C1 ) or y ∈ f (C2 ). This means that for some
x ∈ C1 , y = f (x), or for some x ∈ C2 , y = f (x). Thus, x must be in C1 ∪ C2 , and for
some x ∈ C1 ∪ C2 , y = f (x) =⇒ y ∈ f (C1 ∪ C2 ).
⊛ Example 1.6
Let An = 1, 2, . . . be a sequence of sets. Prove that f ( ∞
S S∞
n=1 An ) = n=1 f (An ).
Proof. Let y ∈ f ( ∞
S∞
n=1 An ). This implies that ∃x ∈ n=1 An s.t. f (x) = y. This
S
implies that x ∈ An for some n, and y ∈ f (An ) for that same “some” n, and thus
y must be in the union of all possible f (An ), ie y ∈ f (An ). This shows ⊆, use
S
⊛ Example 1.7
T∞ T∞
Prove that if An , n = 1, 2, . . . , f ( n=1 An ) ⊆ n=1 f (An ).
Proof (Sketch). Use the same idea as in example 1.6, but, naturally, with intersec-
tions. ■
⊛ Example 1.8
Take f (x) = sin x, A = R, B = R, and take C1 = [0, 2π], C2 = [2π, 4π]. Then,
f (C1 ) = [−1, 1], and f (C2 ) = [−1, 1]. But C1 ∩ C2 = {2π}; f ({2π}) = {sin 2π} =
{0}, and thus f (C1 ∩ C2 ) = {0}, while f (C1 ) ∩ f (C2 ) = [−1, 1], as shown in
proposition 1.4.
,→ Proposition 1.5
Given function f and sets D1 , D2 ,
(a) f −1 ( ∞
S S∞ −1
n=1 An ) = n=1 f (An )
(b) f −1 ( ∞
T T∞ −1
n=1 A n ) = n=1 f (An )
11
Proof.
(a)
∞
[ ∞
[
x ∈ f −1 ( An ) ⇐⇒ f (x) ∈ An
n=1 n=1
(b)
∞
\ ∞
\
x ∈ f −1 ( An ) ⇐⇒ f (x) ∈ An
n=1 n=1
■ 12
This is a “proof by
definitions” as I like to call it.
12
Remark 1.1. f : A → B, A1 ⊆ A. Given f (AC C Similar proof can be used to
1 ) and f (A1 ) , there is no general relation
prove proposition 1.5, less
between the two. generally.
For instance, take A = [0, 6π], B = [−1, 2], C = [0, 2π], and f (x) = sin x. Then,
f (C) = [−1, 1], and f (C C ) = f ([−1, 0)) = [−1, 1], but f (C)C = [−1, 1]C = (1, 2], and
f (C C ) ̸= f (C)C ; in fact, these sets are disjoint.
Proof.
1.5 Reals
In other words; let A ⊆ R and suppose A is bounded from above (A has at a least
upper bound). Then sup(A) exists.
Real numbers, algebraically, have the same properties as the rationals; we have addition,
multiplication, inverse of non-zero real numbers, and we have the relation <. All together,
R is an ordered field.
,→ Definition 1.4
Let A ⊆ R. A number b ∈ R is called an upper bound for A if for any x ∈ A, x ≤ B.
The least upper bound of a set A is unique, if it exists; if s and s′ are two least upper
bounds, then by (a), s and s′ are upper bound for A, and by (b), s ≤ s′ and s′ ≤ s, and
,→ Proposition 1.8
Let13 A ⊆ R and let s be an upper bound for A. Then s = sup(A) iff for any ε > 0,
there exists x ∈ A s.t. s − ε < x.
13
Note that this, and
proposition 1.9 that follows,
are not definitions: they are
Proof. We have two statements: restatements, and do
technically require proof.
I. s = sup(A);
• I =⇒ II: Let ε > 0. Then, since s = sup(A), s − ε cannot be an upper bound for A
(as s is the least upper bound, and thus s − ε < s cannot be an upper bound at all).
Thus, there exists x ∈ A such that s − ε < x, and thus if I holds, II must hold.
• II =⇒ I: suppose that this does not hold, ie II holds for an upper bound s for A, but
s ̸= sup(A). Then, there exists some upper bound b of A s.t. b < s. Take ε = s − b.
ε > 0, and since II holds, there exists x ∈ A such that s − ε < x. But since s − ε = b
and thus b < x, then b cannot be an upper bound for A, contradicting our initial
condition. So, if II =⇒ I does not hold, we have a “impossibility”, ie a value b which
is an upper bound for A which cannot be an upper bound, and thus II =⇒ I.
− sup(A) = inf(−A),
− inf(A) = sup(−A).
NB: M = max(A) need not exist, while sup(A) must exist. Consider A = [0, 1);
sup(A) = 1, but there exists no max(A).
The same logic exists for the existence of minimum vs infimum (consider (0, 1),
with no maximum nor minimum).
I1 ⊇ I2 ⊇ I3 ⊇ . . . In ⊇ In+1 ⊇ . . .
15
Proof. We have In = [an , bn ], In+1 = [an+1 , bn+1 ], . . . . And the inclusion In ⊇ In+1 . 15
Sketch: show that the left-end
points are increasing and the
an ≤ an+1 ≤ bn+1 ≤ bn , ∀ n ≥ 1. So, the sequence an (left-end) is increasing, and the right-end points are
sequence bn (right-end) is decreasing. decreasing. Show either that
all the left-end points are
bounded from above or that
We also have that for any n, k ≥ 1, an ≤ bk . We see this by considering two cases: all the right-end points are
bounded from below. As a
result, there exists a sup/inf
• Case 1: n ≤ k, then an ≤ ak (as an is increasing), and thus an ≤ ak ≤ bk . (depending on which end you
choose) of the set of all the
• Case 2: n > k, then an ≤ bn ≤ bk (again, as bn is decreasing). right/left points. For the sup
case, all upper bounds must
be ≥ sup, and thus the sup is
in all In , and thus in their
Let A = {an : n ∈ N}. Then, A is bounded from above by any bk (as in our inequality we
intersect, and thus the
showed above). Let x = sup(A), which must exist by ?? 1.5. intersect is not empty.
Note that as a result, x ≥ an for all n, and for all k, x ≤ bk , as x is the lowest upper
bound and must be ≤ all other upper bounds, and so for all n ≥ 1, an ≤ x ≤ bn , ie
x ∈ In ∀ n ≥ 1, and thus x ∈ ∞
T∞
n=1 In and so
T
n=1 ̸= ∅. ■
Remark 1.3. The proof above emphasized the left-end points; it can equivalently be proven
via the right-end points, and using y = inf({bn : n ∈ N}) = inf(B), rather than sup(A),
T
and showing that y ∈ In .
T∞
Remark 1.4 (⋆). Note too that, if x = sup(A) and y = inf(B), then x, y ∈ n=1 In ; in fact,
T∞
n=1 In = [x, y]. This can be done by
• Show x ≤ y =⇒ [x, y] ⊆
T
In
• Show
T
In ⊆ [x, y] =⇒ equality.
1
∀ x ∈ R, ∃N ∈ N s.t. N
< x. Clearly, x must be greater than 0 to exist in the intersec-
1
tion; hence, there will always exist some sufficiently large N such that N
< x =⇒ x ∈
/
(1, N1 ) =⇒ x ∈
T T
/ In =⇒ In = ∅.
1 1 1
Remark 1.7. (b) follows from (a) by taking x = y
in (a), then ∃n ∈ N s.t. n > y
=⇒ n
< y,
and thus we need only prove (a).
Remark 1.8. Recall that Q is an ordered field (operations +, · and a relation <). Q can be
extended to a larger ordered field with extended definitions of these operations/relations, such
that it contains elements that are larger than any natural numbers (ie, not bounded above).
This is impossible in R due to AC.
Proof. Suppose (a) not true in R, ie N is bounded from above in R. Let α = sup N, which
exists by AC.
Remark 1.9. If you take a ∈ R and ε > 0, then by the theorem, ∃x ∈ Q where x ∈
(a − ε, a + ε). So any real number can be approximated arbitrarily closely (via choose of ε)
by a rational number.
satisfies a < m
n
< b. ■
Proof. Let S be a non-empty subset of N. Then S has the least element; ∃m ∈ S s.t. m ≤
n, ∀ n ∈ S.
We can assume z ≥ 0; if 0 ≤ z < 1, then we are done (take m = 1), and assume that
z ≥ 1. Let now S = {n ∈ N : z < n}, ̸= ∅ by proposition 1.10, (a). Let m be the least
element of S. It exists by Well-Ordering Property; then, since m ∈ S, z < m. But, we also
have m − 1 ≤ z, otherwise, if z < m − 1 then m − 1 ∈ S and then m is not the least
element of S. Thus, we have m − 1 ≤ z < m, as required. ■
,→ Theorem 1.5
The set J of irrationals is also dense in R. That is, if a, b ∈ R, a < b, ∃ irrational y s.t.
a < y < b (noting that J = R \ Q).
,→ Theorem 1.6
∃ a unique positive real number α s.t. α2 = 2.
α2 < 2
noting that 1
n2
≤ 1
n
for n ≥ 1.
2−α2
Let y = 2α+1
, which is strictly positive. By proposition 1.10, ∃n0 ∈ N s.t.
1 2 − α2 2α + 1
< or < 2 − α2 .
n0 2α + 1 n0
Since α + 1
n0
is positive, α + 1
n0
∈ A. But, since α = sup A, α + 1
n0
≤ α, which
is impossible, so α2 < 2 cannot be true.
α2 > 2
Take n ∈ N; 2
1 2α 1 2α
α− = α2 − + 2 > α2 − .
n n n n
α2 −2
Now, let y = 2α
; y > 0, and by proposition 1.10, ∃n0 ∈ N s.t.
1 α2 − 2 2α
< , or < α2 − 2.
n0 2α n0
So, α − 1
n0
> x for all x ∈ A. So α − 1
n0
is an upper bound for A, but since
α = sup A, α − 1
n0
≥ α ie α ≥ α + 1
n0
, which is impossible. So α2 > 2 cannot
be true.
Thus, α2 = 2.
■ 16
Proof sketch: uniqueness is
clear. Existence follows from
showing that α2 cannot be
either < or > 2. This is done
by contradiction, taking some
§1.6 Logic, Sets, and Functions: Density of Rationals in Reals p. 18 number slightly
Remark 1.10. A similar argument gives that for any x ∈ R, x ≥ 0, ∃!α ∈ R, α ≥ 0 such
√
that α2 = x. This x is called the square root of x, denoted α = x.
Remark 1.11. For any natural number m ≥ 2 and x ≥ 0, ∃!α ∈ R, α ≥ 0 s.t. αm = x. The
proof is similar, and we call α the m-th root of x.
Remark 1.12. Our last proof also gives that Q cannot satisfy AC. Suppose it does, ie any set in
Q bounded from above has a supremum ∈ Q. Then, consider B = {x ∈ Q : x ≥ 0 and x2 <
2}; set α = sup B. The exact same proof can be used, but we will not be able to find an upper
bound in Q.
1.7 Cardinality
,→ Definition 1.8
Let f : A → B.
3. f bijective if both.
⊛ Example 1.10
Consider functions f, g.
1. If f, g injective, so is h = g ◦ f
2. If f, g bijective, then so is h
,→ Definition 1.10
The inverse function17 is defined only for bijective map f : A → B. y ∈ B, f −1 (y) = x
where x ∈ A s.t. f (x) = y.
17
Not the same as the inverse
image of a set by a function,
which is defined for any
§1.7 Logic, Sets, and Functions: Cardinality p. 19 function.
⊛ Example 1.11
1. A = R, B = (0, ∞), f (x) = ex . f is a bijection, and f −1 (y) = ln y, y ∈
(0, ∞).
⊛ Example 1.12
Let E = {2, 4, 6, . . . } (even natural numbers). Define f : N → E by f (n) = 2n.
Thus, f is a bijection, and N ∼ E.18
18
See these independent notes
for more.
,→ Theorem 1.7
The relation ∼ is a relation of equivalence.
1. A ∼ A
2. if A ∼ B, then B ∼ A
3. if A ∼ B and B ∼ C, then A ∼ C
Remark 1.13. According to this, finite sets are not countable; this is just a convention. Some-
times, we say a set is countable if it is finite or to above definition holds, where we say that a
set is countably infinite if it is infinite and countable.
Other times, finite sets are treated separately than countable sets.
,→ Theorem 1.8
Suppose that A ⊆ B.
,→ Proposition 1.12
N × N ∼ N; there exists a bijection f : N × N → N.
,→ Proposition 1.13
Let A be a set. The following are equivalent statements:
Proof. We proceed by proving that each statement implies the next (and thus are equiva-
lent).
• (a) =⇒ (b): Suppose A is finite and has N elements. Then there exists a bijection
h : {1, 2, . . . n} → A. We now define a map f : N → A, by setting
h(m) if m ≤ n
f (m) = .
h(n) if m > n
f is surjective, and thus (b) holds. If (a) countable, ∃ bijection f : N → A, and any
bijection is a surjection, so (b) also holds.
since h is a surjection. Then, by the well-ordering property of N, the set h−1 ({a})
has a least element.
If n is the least element of h−1 ({a}), we set f (a) =. This defines a function
f : A → N,
,→ Theorem 1.9
Let An , n = 1, 2, . . . be a sequence of sets such that each An is either finite or count-
able. Then, their union
∞
[
A= An
n=1
Proof. We will use (a) ⇐⇒ (b) from proposition 1.13 to prove this.
φn : N → An .
h(n, m) = φn (m).
h(n, m) = a,
which is a surjection as both h, f are surjections. So, there exists a surjection from N → A,
and by proposition 1.13, (b) =⇒ (a), and thus A = ∞ n=1 An is also finite or countable.
S
■
S∞
Remark 1.14. If A = n=1 An , where each An is either finite or countable, and at least one
An is countable, then A is countable.
Remark 1.15. If A1 , . . . , An are finitely many finite or countable sets then their union A1 ∪
· · · ∪ An is also finite or countable (essentially just previous proof where we use n instead of
∞ for the upper limit of the union…).
,→ Theorem 1.10
The set Q of rational numbers is countable.
Proof. We write
Q = A0 ∪ A1 ∪ A2 ,
where A0 = {0}, A1 = { m
n
: m, n ∈ N}, and A2 = {− m
n
: m, n ∈ N}.
Let us show that A1 is countable; define
m
h : N × N → A1 , f (m, n) = .
n
I1 ⊇ I2 and x1 ∈
/ I1 , x2 ∈
/ I2 .
Consider x3 ; if x3 ∈
/ I2 , then I3 = I2 . If x3 ∈ I2 , we repeat the “dividing” process as before.
/ I2′′ . If x3 ∈
/ I2′ or x3 ∈
Since x3 ∈ I2 , either x3 ∈ / I2′′ , I3 = I2′′ .
/ I2′ , I3 = I2′ . Else, if x3 ∈
We have now that
I1 ⊇ I2 ⊇ I3 and x1 ∈
/ I1 , x2 ∈
/ I2 , x3 ∈
/ I3 ,
and we can continue this construction to obtain an infinite sequence of bounded, closed
intervals In s.t.
I1 ⊇ I2 ⊇ · · · ⊇ In ⊇ In+1 ⊇ · · · ,
∞
\
In .
n=1
T∞
For every m, xm ∈
/ Im , so for every m ∈ N, xm ∈
/ n=1 In , and so R = {x1 , x2 , . . . xm , . . . }
has an empty intersection with this intersection, ie
∞
!
\
R∩ In = ∅.
n=1
T∞ T∞
Otoh, n=1 In ⊆ R, so we must have that n=1 In = ∅ contradicting the nested interval
,→ Proposition 1.15
The set (−1, 1) ⊆ R is uncountable.
S∞
Proof. We can write R = n=1 (−n, n). If each (−n, n) is countable, then R would also
be countable, as a countable union of countable sets. Thus, there must exist some n0 ∈
N s.t. (−n0 , n0 ) is not countable. The map
x
f : (−n0 , n0 ) → (−1, 1), f (x) =
n0
⊛ Example 1.13
Show that the map
x
f (x) =
1 − x2
is a bijection between (−1, 1) and R ie (−1, 1) ∼ R.
Proof. Surjection is fairly trivial (if stuck, consider the graph of the function).
Injection; given f (x) = f (y) where x, y ∈ (−1, 1),
x y
=
1 − x2 1 − y2
x − xy 2 = y − yx2
x − y = xy 2 − yx2 = xy(y − x)
x − y = −xy(x − y)
=⇒ −xy = 1 =⇒ xy = −1, or x − y = 0
,→ Proposition 1.16
Any bounded non-empty open interval (a, b) ∈ R is uncountable.
The map
2(x − a)
f (x) = −1
b−a
is a bijection between (a, b) and (−1, 1), and we have shown that (−1, 1) ∼ R, so (a, b) ∼
R, and thus any open interval has the same cardinality as R. ■
⊛ Example 1.14
Prove that ∃ bijection between [0, 1) and (0, 1), and conclude that [0, 1) ∼ (0, 1) ∼
R. Then conclude for any a < b, [a, b) ∼ R.
f : A → P(A).
Let D ⊆ A defined as
D = {a ∈ A : a ∈
/ f (a)}.
Since D ⊆ P(A), and f is surjective, there must exist some a0 ∈ A s.t. f (a0 ) = D.
We have two cases:
2. a0 ∈
/ D. But then, since D = f (a0 ), a0 ∈
/ f (a0 ), and so by definition of D, a0 ∈ D,
which is again not possible.
So, the assumption of a surjection existing has led to a0 ∈ A such that neither a0 ∈ D nor
/ D, which is impossible. Thus there can be no surjective f .
a0 ∈
Notice, though, that there exists an injection A → P(A), a 7→ {a}, and thus there is an
injection but no bijection.
Thus, we can say that P(A) is strictly bigger than A.
■
2 Sequences
2.1 Definitions
,→ Definition 2.1
Let A be a set. An A-valued sequence indexed by R is a map
x : N → A.
The value x(n) is called the n-th element of the sequence. One writes x(n) = xn , or
lists its elements
⊛ Example 2.1
Let (xn ) be a sequence defined by xn = n1 , n ∈ N, then limn→∞ xn = 0.
1 1
0< ≤ < ε.
n N
lim xn = x
n→∞
means that
( ∀ ε > 0)(∃N ∈ N)( ∀ n ≥ N )(|xn − x| < ε).
⊛ Example 2.2
Prove22 that
n2 + 1
lim = 1.
n→∞ n2
n2 + 1 n2 + 1 − n2 1 1
| 2
− 1| = | 2
| = 2 ≤ 2 < ε.
n n n N
■
22
Work backwards to start; how
can you simply the sequence
(that is, build a string of
,→ Definition 2.4: Divergent Sequences
inequalities) such that
If a sequence (xn ) does not converge to any real number x, we say that the sequence defining an N in terms of ε
becomes apparent?
is divergent. For instance, consider
xn = (−1)n , n ≥ 1.
The sequence alternates between 1 and −1 and so intuitively does not converge. How
do we prove it?
⊛ Example 2.3
Evaluate the following examples using the ε definition:
1. limn→∞ sin
√
3n
n
=0
2. limn→∞ n!
nn
=0
(1+2+···+n)2
3. limn→∞ n4
= 1
4
√ √
3 1 1
3
n ≥ N =⇒ √
n ≥ N =⇒ ≤ √3
3
n N
sin n 1 1
−1 ≤ sin n ≤ 1 =⇒ |sin n| ≤ 1 =⇒ √ ≤ √3
≤ √ 3
<ε
3
n N N
sin n
=⇒ lim √ =0
n→∞ 3 n
2. Take 1
N
≤ ε. Then, ∀ ε > 0, ∀ n ≥ N =⇒ 1
n
≤ 1
N
,
n2 (n+1)2
(1 + 2 + · · · + n)2 1 4 n4 n4 + 2n3 + n2 − n4
− = − =
n4 4 n4 n4 n4
2n3 + n2 2n + 1 2n 2 2
= 4
= 2
≤ 2 ≤ ≤ <ε
n n n n N
(1 + 2 + · · · + n)2 1
=⇒ lim =
n→∞ n4 4
,→ Theorem 2.1: ⋆
A limit of a sequence is unique. In other words, if the sequence is converging, then
its limit is unique. The sequence cannot converge to two distinct numbers x and y.24
24
Proof sketch: contradiction,
assume two distinct limits,
and take ε as their midpoint.
Proof. By contradiction; suppose ∃(xn ) s.t. limn→∞ xn = x and limn→∞ xn = y, and that Arrive at a contradiction by
using triangle inequalities to
x ̸= y. show that |x − y| < |x − y|,
|x−y| and thus the limits cannot be
Take ε = . Since x ̸= y, we have that ε > 0. Since limn→∞ xn = x, ∃N1 ∈ N s.t. for
2 distinct.
n ≥ N1 , |xn − x| < ε.
Similarly, since lim xn = y, ∃N2 ∈ N s.t for g ≥ N2 , |xn − y| < ε.
|x − y| = |x − xn + xn − y| ≤ |x − xn | + |xn − y|
< ε + ε = |x − y|
=⇒ |x − y| < |x − y|, ⊥
,→ Theorem 2.2
Any converging sequence is bounded.25
In other words, if (xn ) is a converging sequence,
lim xn = x, lim yn = y.
Then:
Proof (part 3. of proposition 2.1). Take28 lim xn = x, lim yn = y. Since (xn ) is converging, it
is bound by theorem 2.2, and there exists M > 0 s.t. ∀ n ≥ 1, |xn | ≤ M .
Now,
Thus, for n ≥ N , |xn yn − xy| < ε, and by definition of the limit, lim xn yn = xy. ■
28
Proof sketch: take an upper
bound of xn . Then, show that
|xn yn − xy| < ε, by using
,→ Theorem 2.3: Order Properties of Limits
triangle inequalities to show
Let (xn ), (yn ) be two sequences such that inequality to a combination of
M , arbitrarily small values
(by def of limits of xn , yn
lim xn = x, lim yn = y. resp,), and |y|.
1. xn ≥ 0 ∀ n =⇒ x ≥ 0.
3. c is constant since c ≤ xn ∀ n ≥ 1 =⇒ c ≤ x. xn ≤ c ∀ n ≥ 1 =⇒ xn ≤ c.
Proof of 1. Suppose 1. does not hold; suppose ∃(xn ) s.t. lim xn = x, xn ≥ 0 ∀ ≥, but
x < 0.
Let ε > 0 s.t. x < −2ε < 0. With this ε, lim xn = x gives that ∃N ∈ N s.t. ∀ n ≥
N, |xn − x| < ε, or particularly, xn − x < ε.
Then, xn < ε + x, and since x < −2ε, we have ∀ n ≥ N , xn < −ε, and thus ∀ n ≥ N ,
xn < 0, a contradiction. ■
1. Let (xn ) be an increasing sequence that is bounded from above. Then (xn ) is
converging.
2. Let (xn ) be a decreasing sequence that is bounded from below. then (xn ) is
converging.
30
Sketch: 1,2 are proven very
similarly. For 1., take the set
of all xn in the given
Proof (of 1). Let A = {xn : n ≥ 1}. Since (xn ) is bounded above by M , the set A is bounded sequence. Since the sequence
is bounded, then so is the set,
from above. Let α = sup A, which exists by AC. and so we can take its
Let ε > 0. Since α is the least upper bound for A, α − ε is not an upper bound of A supremum. Use the ε
definition of sup to show that
(α − ε < α). Hence, there must exist some N ∈ N such that α − ε < xN (if it didn’t exist, this supremum is also the
limit of the sequence
then α wouldn’t be the supremum . . . ). Then, for n ≥ N , and since (xn ) increasing, (basically, a bunch of
inequalities, and being careful
with definitions). 2. follows
α − ε < xN ≤ xn ≤ α. identically but using the
infimum.
α − ε < xn ≤ α =⇒ −ε < xn − α ≤ 0,
⊛ Example 2.5
A sequence (xn ) is called eventually increasing if there exists some N0 ∈ N s.t. ∀ n ≥
N0 , xn+1 ≥ xn . If (xn ) is eventually increasing and bounded from above, lim xn = α
exists.
⊛ Example 2.6
√ √
Let (xn ) be a sequence defined recursively by x1 = 2 and xn+1 = 2 + xn , n ≥ 1.
√ √
p q p
So x2 = 2 + 2, x3 = 2 + 2 + 2 · · · , xn = 2 cos 2n+1 π
, n ≥ 1. Show that
lim xn = 2.
Proof. We will prove this using the Monotone Convergence Thm by showing that
the xn is bounded from above and increasing, which implies that the limit exists.
We will then find the actual limit.
Recall that n ≥ 1, xn ≤ 2. We will prove this by induction. Let S ⊆ N be the set of
√
indices such that xn ≤ 2. Since x1 = 2 < 2, 1 ∈ S. Now suppose some n ∈ S, ie
√ √
xn ≤ 2. Then, we have that xn+1 = 2 + xn ≤ 2 + 2 = 2 =⇒ xn+1 ≤ 2. Thus,
by induction, n ∈ S =⇒ n + 1 ∈ S =⇒ S = N, ie xn ≤ 2 ∀ n ∈ N. Thus, our
sequence is bounded from above.
We now prove that (xn ) is increasing. Let S ⊆ N s.t. n ∈ S ⇐⇒ xn+1 ≤ xn .
p √ √
x2 = 2 + 2 ≥ 2 = x1 =⇒ x1 ≤ x2 =⇒ 1 ∈ S. Suppose n ∈ S =⇒
√ √
xn+1 ≥ xn . Then, xn+2 = 2 + xn+1 ≥ 2 + xn = xn+1 =⇒ n + 1 ∈ S. Thus,
S = N, so xn+1 ≥ xn ∀ n ∈ N.
So the sequence (xn ) is increasing and bounded from above, and thus ∃ lim xn = α.
√
To find the value of α, consider xn+1 = 2 + xn , or x2n+1 = 2 + xn . We can
also write that α = lim xn = lim xn+1 .31 We then have that lim xn+1 = α =⇒
lim x2n+1 = α2 , and thus x2n+1 = 2 + xn =⇒ lim x2n+1 = lim(2 + xn ) =⇒ α2 =
2 + α =⇒ α = 2, −1. xn ≥ 0 ∀ n, by Order Limit Theorem, and so α ≥ 0 and thus
α = 2. ■
31
Add proof
,→ Corollary 2.1
√
For a, b > 0, then 21 (a + b) ≥ ab
1 2 √
Proof. 2
(a + b) = 14 (a2 + 2ab + b2 ) ≥ ab =⇒ 1
2
(a + b) ≥ ab ■
⊛ Example 2.7
Let (xn ) be defined recursively by x1 = 2 and xn+1 = 1
2
xn + 2
xn
for n ≥ 1. Then,
√
(xn ) is converging and lim xn = 2.
Proof. We32 will show that (xn ) bounded from below and decreasing, implying the
1 2 1 1 1 1 √ 2
xn − xn+1 = xn − (xn + ) = xn − = (x2n − 2) ≥ ( 2 − 2) ≥ 0,
2 xn 2 xn 2xn 2xn
where the second-to-last inequality holds from the lower bound we found on xn .
Having shown that xn decreasing and is bounded from below, we conclude that it
converges by Monotone Convergence Theorem. To find its limit, let L := lim xn .
Then,
1 2 1 1
lim xn = lim xn−1 + = lim xn−1 + lim ,
2 xn−1 2 xn−1
and since the limit of xn is equal to the limit of xn−1 , we have that
1 1 √
L= L+ =⇒ L2 = 2 =⇒ L = ± 2,
2 L
√
but we know that xn ≥ 2 hence we can ignore the negative root, and thus xn
√
converges to 2. ■
32
This example, as well as the
more general one after it, rely
⊛ Example 2.8: ⋆ on applying 1) the monotone
convergence theorem, then 2)
Let a > 0 and let (xn ) be a sequence defined recursively by x1 is arbitrary (positive), using Algebraic Limit
Properties to turn the
and problem into an algebraic
1 a problem, using the given
xn+1 = (xn + ), n ≥ 1.
2 xn recursive relation.
√
Show that limn→∞ xn = a.
√
Proof. By corollary 2.1, xn+1 = 12 (xn + xan ) ≥ a
a, hence, xn is bounded
p
xn · xn
=
√
from below by a.
We also have that xn − xn+1 = xn − 12 xn − 2xan = x2n − 2xan = x1n (x2n − a). We
√
have that xn ≥ a =⇒ x2n ≥ a =⇒ x2n − a ≥ 0. Further, since the sequence
√
is bounded from below by a > 0( ⇐= a > 0), then x1n > 0 as well. Hence,
1
xn
(x2n − a) ≥ 0, and thus xn − xn+1 ≥ 0 =⇒ xn ≥ xn+1 and thus xn is decreasing.
Thus, by the Monotone Convergence Theorem, xn is convergent. Let X := limn→∞ xn .
We have from the recursive definition, lim xn = lim 2 (xn + xn ) . Since we know
1 a
1 a 1 a 1
lim xn = lim xn + lim = lim xn + lim
2 2xn 2 2 xn
1 a
=⇒ X = X +
2 2X
X a √
=⇒ = =⇒ X 2 = a =⇒ X = a,
2 2X
⊛ Example 2.9
Evaluate33 the limit of xn given the recursive relation xn+1 = 1
4−xn
, x1 = 3.
Proof. We aim to show that (xn ) is bounded from below and decreasing.
Bounded from below: we claim xn > 0; we proceed by induction. x1 = 3 > 0
holds; say xn > 0 for some n ≥ 1. Then, we have
1 1 1
xn > 0 =⇒ −xn < 0 =⇒ 4−xn < 4 =⇒ > > 0 =⇒ xn+1 = > 0,
4 − xn 4 4 − xn
1 1
xn−1 ≥ xn =⇒ 4−xn−1 ≤ 4−xn =⇒ ≥ = xn+1 =⇒ xn ≥ xn+1
4 − xn−1 4 − xn
and thus the sequence decreases, and by theorem 2.5 the limit exists. Let X =
1 1
limn→∞ xn = limn→∞ =⇒ X = 4−X
4−xn−1
=⇒ 4X − X 2 = 1 =⇒ 0 =
√
X 2 − 4X + 1 =⇒ X = · · · = 2 ± 3. We must take the negative root, since X is
decreasing and thus must be less than 3. ■
33
Abbott, pg 54 exercise 2.4.2
m ≤ xn ≤ M, ∀ n
and further,
y1 ≥ y2 ≥ · · · ≥ yn ≥ yn+1 ≥ · · · ;
Now, similarly, note that An is bounded below by m and thus zn = inf An exists. We
further have that zn ≤ xn ≤ M , and that zn ≥ m ∀ n, and we have
z1 ≤ z2 ≤ · · · ≤ zn ≤ zn+1 ≤ · · · ,
by a similar argument as before. So, as before, the sequence (zn ) is increasing, and
bounded from above by M . Again, by MCT, limn→∞ zn = z exists. We call z the limit
inferior of (xn ), and denote
⊛ Example 2.10
Let (xn ) = (−1)n , n ∈ N. We showed previously that this is a divergent sequence,
so the limit does not exist. However, the sequence is bounded, since −1 ≤ xn ≤
1 ∀ n. We have An = {(−1)k : k ≥ n} = {−1, 1}. So, yn = sup An = 1, and zn =
inf An = −1, ∀ n. Thus, lim sup xn = lim yn = 1, and lim inf xn = lim zn = −1,
despite lim xn not existing.
2. limn→∞ xn = limn→∞ xn = x.
(1) =⇒ (2): Suppose (xn ) is converging, and limn→∞ xn = x. Let ε > 0. Then, there
exists some N ∈ N s.t. ∀ n ≥ N ,
ε
|xn − x| < ,
2
or equivalently,
ε ε
x− < xn < x + , ∀ n ≥ N.
2 2
Since An = {xk : k ≥ n}, if n ≥ N , then x + ε
2
is an upper bound for An , and x − ε
2
is a
lower bound for An . This gives that
ε ε
yn = sup An ≤ x + ; zn = inf An ≥ x − .
2 2
ε ε
x− ≤ zn ≤ xn ≤ yn ≤ x + ,
2 2
(2) =⇒ (1): Let ε > 0. Since limn→∞ yn = x, ∃N1 s.t. ∀ n ≥ N1 , |yn − x| < ε. Similarly,
since lim zn = x, ∃N2 s.t. ∀ n ≥ N2 , |zn − x| < ε.
Take N = max{N1 , N2 }. Then, for n ≥ N , we have
x − ε < zn ≤ xn ≤ yn < x + ε.
⊛ Example 2.11
Let (xn ) be a bounded sequence. Then
Proof. Recall Remark 1.2; Let An := {xk : k ≥ n} as in the definition of lim sup, lim inf.
Let yn := sup An , zn := inf An . By theorem 2.6, lim yn = lim zn . Further, sup(−An ) =
− inf(An ), where −An = {−xk : k ≥ n}; hence, lim sup(−xn ) = − lim inf xn , as
desired. ■
Remark 2.3. Given (xn ) bounded and α ≥ 0, then the following holds:
,→ Proposition 2.2
Let (xn ) and (yn ) be bounded sequences. Then,
and
(2) limn→∞ (xn + yn ) ≥ limn→∞ xn + limn→∞ yn
Proof. (1) Take An = {xk + yk : k ≥ n}, Bn = {xk : k ≥ n}, Cn = {yk : k ≥ n}. Then,
take
Bn + Cn = {xk + yj : k ≥ n, j ≥ n} ⊇ An
so tn ≤ rn + sn , that is, lim tn ≤ lim rn + lim sn , and thus limn→∞ (xn + yn ) ≤ limn→∞ xn +
limn→∞ yn , proving (1).
(2) The same argument holds, replacing each instance of limn→∞ with limn→∞ and reversing
inequalities where necessary. Alternatively, it follows directly from (1) by negating the
sequences where appropriate. ■
Remark 2.4. (2) follows from (1) by either repeating the argument used to prove (1) (changing
notation), or using the fact that limn→∞ xn = − limn→∞ (−xn ).
Remark 2.5. The set {n : xn > t} is empty or finite iff ∃ nt ∈ N s.t. ∀ n > nt , xn ≤ t. The
set is empty or finite if t is an eventual upper bound for (xn ); that is, starting with sufficiently
large nt , xn ≤ t ∀ n ≥ t.
In other words, t is an upper bound if we neglect finitely many elements. Hence, (1) states
equivalently states that limn→∞ xn is the infimum of the eventual upper bounds for (xn );
finite
t upper bounds xn for an infinite number of n’s
Proof. (Of (1)) Let A = {t : {n : xn > t} is either empty or finite }. We note that this set
is non-empty and bounded from below, hence the inf is well-defined. We can see this by
recalling that (xn ) bounded, hence ∀ n ∃m, M s.t. m ≤ xn ≤ M . Then, {n : xn > M } is
empty, hence M ∈ A. Otoh, if t < m, then the set {n : xn > t} = N since xn ≥ m > t ∀ n.
So, if t < m, then t ∈
/ A and hence m is a lower bound for A.
We have now that limn→∞ xn is a lower bound for A, and hence limn→∞ xn ≥ inf A.
Let t ∈ A. We aim to show that limn→∞ xn ≤ t.
Proof. (Of (1), More Concise) Let α = inf A and ε > 0. Then α − ε is not in A, α − ε is not
Let, now,
ym = sup{xn : n ≥ m}.
limn→∞ xn = lim yn ≥ α − ε,
m→∞
so for any ε > 0, limn→∞ xn ≥ α − ε, and thus limn→∞ xn ≥ α = inf A, and the proof is
complete. ■
Remark 2.6. k is the index of the subsequence, (xnk )k∈N , not n; xn1 is the 1st element, . . . ,
xnk is the k-th element.
⊛ Example 2.12
Let xn = n1 , ( n1 )n∈N , and let nk = 2k + 1, k ∈ N. n1 = 3, n2 = 5, n3 = 7, . . . , nk =
2k + 1. Our subsequence is then
1 1 1 1
(xn1 , xn2 , . . . , xnk , . . . ) = , ,..., ,... =
3 5 2k + 1 2k + 1 k∈N
nk ≥ k ≥ K = N,
,→ Proposition 2.4
If 0 < b < 1, then limn→∞ bn = 0.
Proof. Let xn = bn . Then xn > 0, and xn+1 = bn+1 = bxn < xn , and since 0 < b < 1,
(xn ) is decreasing and bounded from below, (xn ) converges by the Monotone Convergence
Theorem. Let x = limn→∞ xn . Again, xn+1 = bxn , so limn→∞ xn+1 = limn→∞ bxn =
b limn→∞ xn , so x = bx =⇒ (1 − b)x = 0. 0 < b < 1 =⇒ x = 0. ■
36
BW Proof (1): using Nested Interval Property. Since (xn ) bounded, ∃M > 0 s.t. |xn | ≤ M ∀ n ∈
N. Let I1 = [−M, M ] and n1 = 1. We now construct I2 , n2 as follows.
Divide I1 into two intervals of the same size, I1′ = [−M, 0], I1′′ = [0, M ]. Now, consider
the sets
Suppose now that Ik , nk are chosen, and that Ik contains infinitely many elements of
the sequence (xn ). Divide Ik into two equal sub-intervals, Ik′ , Ik′′ . We now introduce
(k) (k)
A1 = {n ∈ N : n > nk and xn ∈ Ik′ }, A2 = {n ∈ N : n > nk and xn ∈ Ik′′ },
(k) (k)
(similar to our construction of A1 , A2 ). A1 ∪ A2 must be infinite, so one of the two
(k)
must be infinite. If A1 infinite, set Ik+1 = Ik′ , nk+1 = min A1 . If A2 infinite, set Ik+1 =
(k)
Ik′′ , nk+1 = min A2 .
This gives now that Ik+1 and nk+1 , where Ik+1 ⊆ Ik , Ik+1 contains infinitely many
elements of the sequence. Further, by construction, nk+1 > nk . This gives us a sequence
of closed intervals Ik = [ak , bk ], k ∈ N such that I1 ⊇ I2 ⊇ · · · ⊇ Ik ⊇ Ik+1 ⊇ · · · , such
that xnk ∈ Ik , and that nk is a strictly increasing sequence of natural numbers, defining
subsequence (xnk ).
We claim now that our subsequence (xnk ) satisfies limk→∞ xnk = x. To see this, let
ε > 0. Since limk→∞ M
2k−2
= limk→∞ 4M
2k
= 0, by proposition 2.4, with b = 21 . There exists
K ∈ N such that ∀ k ≥ K, we have M
2k−2
= bk − ak < ε. So, since Ik is a nested sequence of
intervals, ∀ k ≥ K, xnk ∈ IK (xnk ∈ Ik ⊆ IK ). We also have that x ∈ IK , since x ∈ Ik .
T
So, x, xnk ∈ [aK , bK ] ∀ k ≥ K. So, for k ≥ K, |xnk − x| ≤ |bk − ak | < ε. So for ε > 0,
∃K ∈ N s.t. ∀ k ≥ K, |xnk − x| < ε, and so limk→∞ xnk = x, as desired. ■
36
Sketch: this proof is
somewhat diagonal in nature
(if one can say that); if you
,→ Definition 2.9: Peak
understand the proof of
Let (xn ) be a sequence of real numbers. An element xm is called a peak of this sequence Cantor’s Theorem using the
Nested Interval property, this
if xm ≥ xn ∀ n ≥ m. xm is bigger or equal then to any element of the sequence that should follow naturally. In
short, construct subsequences
follows it. such that the subsequence
If a sequence is decreasing, then any element of the sequence is a peak. has all its terms contained in
a “nest” of intervals, and
If a sequence is increasing, then there is no peak. show that the length (sts) of
these intervals converges to 0.
But these are subsets of R,
§2.4 Sequences: Subsequences and Bolzano-Weirestrass Theorem p. 44 their intersect must contain
BW Proof (2): using Peaks. Take sequence (xn ). Then,
• Case 1: (xn ) has infinitely many peaks; enumerate the indices of those peaks as
n1 < n2 < n3 < · · · , then xnk < xnk+1 ∀ k, since xnk is a peak, nk+1 > nk . This gives
a decreasing subsequence (xnk ).
• Case 2: (xn ) has finitely many peaks, with indices m1 < m2 < · · · < mr . Set
n1 = mr + 1. Then xn1 is not a peak, and so ∃ n2 > n1 s.t. xn2 > xn1 . Now, xn2 is
also not a peak, (n2 > n1 > mr ), and so there exists n3 > n2 such that xn3 > xn2 ,
and so on. In this way, we construct a subsequence (xnk ) that is strictly increasing,
that is, xnk+1 > xnk .
If in addition (xn ) is bounded, say |xn | ≤ M ∀ n, then the monotone subsequence con-
structed in Cases 1, 2 is also bounded; ie |xnk | ≤ M ∀ k. Thus, by Monotone Convergence
Theorem, (xnk ) is converging. ■
ε ε
|xn − xm | = |xn − x + x − xm | ≤ |xn − x| + |xm − x| < + =ε
2 2
=⇒ |xn − xm | < ε,
Remark 2.12. To prove ⇐= , we first introduce the following theorem(s); see section 2.5 for
the remainder.
,→ Theorem 2.10
Let (xn ) be a Cauchy sequence and suppose that (xn ) has a convergent subsequence
(xnk ). Then (xn ) is also convergent.
Proof. Let x = limn→∞ xnk . Let ε > 0. Then, ∃K ∈ N such that ∀ k ≥ K, |xnk − x| < ε.
We have too that (xn ) Cauchy, ie ∃N ∈ N s.t. ∀ n, m ≥ N , |xn − xm | < 2ε .
Let now K0 ≥ max{K, N }. Recall that nK0 ≥ K0 ≥ N . Take now n ≥ N , and estimate
Since K0 ≥ K, xnK0 − x < 2ε . Since nK0 ≥ N , we also have xn − xnK0 < 2ε . Thus, we
have
ε ε
|xn − x| < + = ε,
2 2
hence limn→∞ xn = x. ■
,→ Theorem 2.11
Any Cauchy sequence is bounded.
Proof. Let (xn ) be Cauchy. We aim to show that ∃M > 0 s.t. ∀ n ∈ N, |xn | ≤ M .
Take ε = 1 in the definition of Cauchy sequence. Let N be such that ∀ n, m ≥ N ,
|xn − xm | < 1. We can take m = N , and so for all n ≥ N , |xn − xN | < 1, which gives that
for n ≥ N ,
Let38 38
While this seems like an
arbitrary definition, this is a
common “trick” to find a
§2.5 Sequences: Cauchy Sequences p. 46
bound of a sequence based on
M = |x1 | + |x2 | + . . . |xN −1 | + |xN | + 1.
⊛ Example 2.14
Let39 (xn ) be a sequence defined recursively by x1 = 1, x2 = 2, xn+1 = 1
2
(xn +
xn−1 ), n ≥ 2. Prove that (xn ) is a convergence sequence, and find its limit.
39
Sketch: show xn Cauchy
=⇒ xn converges, then take
a subsequence of xn (spec,
Remark 2.15. Before solving, we establish a number of properties about the sequence. odd n) and find a closed form
of it which is nicer to
evaluate. Use then
,→ Proposition 2.5: Property I theorem 2.7 to conclude that
the limit of the subsequence
1 ≤ xn ≤ 2 ∀ n ≥ 1 is equal to the limit of the
sequence.
1
xn+1 = (xn + xn+1 ),
2
1 ≤ xn+1 ≤ 2,
1
|xn+2 − xn+1 | = (xn+1 + xn ) − xn+1
2
1 1 1
= xn − xn+1 = |xn+1 − xn |
2 2 2
1 1 1
(assumption =⇒ ) = · n−1 = n ,
2 2 2
,→ Corollary 2.2
1−rk+1
For any r ̸= 1, and any k ∈ N, 1 + r + r2 + · · · + rk = 1−r
.
1−r1
Proof. We proceed by induction. k = 1 =⇒ r0 = 1−r
= 1, holds. Suppose 1 + · · · rk−1 =
1−rk
1−r
holds for some k − 1 ∈ N. Then, we have that
1 − rk 1 − rk + (1 − r)rk
1 + · · · rk−1 + rk = + rk =
1−r 1−r
k k k+1
−r + r − r
1
=
1−r
1 + 1 − rk+1
= ,
1−r
hence, the statement for k − 1 =⇒ the statement for k, hence it holds ∀ k ∈ N and the
proof is complete. ■
Proof. Let ε > 0. We need to find N ∈ N such that ∀ n, m ≥ N , |xn − xm | < ε. Let N be
such that40 2N1−2 = 4
2N
< ε. Let, now, n, m ≥ N , and suppose n > m (when n = m, we are
done; when n < m, simply switch the variables wlog). We can write
1 1 1
|xn − xm | ≤ + + · · ·
2m−1 2m 2n−2
1 1 1
= m−1 1 + + · · · + n−m−1
2 2 2
n−m !
1 − 12
1 1 1 1 1 1
1 + + · · · + n−m−1 = m−1 1 < m−2 ≤ N −2 .
2m−1 2 2 2 1− 2 2 2
1
X = lim xn = lim( (xn−1 + xn−2 ))
2
1
=⇒ X = (X + X) = X,
2
(x2k+1 )k∈N
1 1 1
x2k+1 = 1 + + 3 + · · · + k−1 , k ≥ 1. ⋆
2 2 2
Note that ∀ n ≥ 1, x2n ≥ x2n−1 and x2n ≥ x2n+1 . We can argue by induction. Let S ⊆ N
for which the relation holds. Since x1 = 1, x2 = 2, x3 = 23 , we have that x2 ≥ x1 , x2 ≥ x3 ,
and so the relation holds, ie 1 ∈ S. Suppose that n ∈ S, ie x2n ≥ x2n−1 , x2n ≥ x2n+1 for
some n ≥ 1. We can write
1 1
x2k+2 = (x2k+1 + x2k ) ≥ (x2n+1 + x2n+1 ) ≥ x2n+1
2 2
1 1
=⇒ x2n+3 = (x2n+2 + x2n+1 ) ≤ (x2n+2 + x2n+2 ) = x2n+2
2 2
Hence x2n+2 ≥ x2n+1 and x2n+2 ≥ x2n+3 , n + 1 ∈ S, and hence S = N, and our relation
holds ∀ n ∈ N.
k−1 !
1 1 1
x2k+1 =1+ 1 + 2 + ··· + .
2 2 22
1 k
!
1 1 − 4
x2k+1 =1+
2 1 − 41
k !
2 1
=1+ 1−
3 4
k
5 2 1
= −
3 3 4
Now, since limn→∞ xn = X and we showed xn convergent, then each of its subse-
quences converges to the same limit. Thus, X = 53 , ie,
5
lim xn = .
n→∞ 3
Remark 2.16. Generally, this type of approach is quite tedious. The next example(s) will try
to generalize it.
⊛ Example 2.15
Consider the recursive relation xn+1 = 12 xn + 21 xn−1 ⋆.
1 1
x2 = x + ,
2 2
1
xn = C1 an + C2 bn = C1 + C2 (− )n , ⋆⋆
2
where C1 , C2 are arbitrary constants. We claim that this sequences satisfies our
recursive relation, ⋆; note that
n+1 n−1
1 1 1 1 1 1 1
− = − · = − − +
2 2 4 2 2 2 2
n+1
1
=⇒ xn+1 = C1 + C2 −
2
n−1
C1 C1 1 1 1 1
= + + C2 − − +
2 2 2 2 2 2
n n
C1 C1 1 C2 1
= + + C2 − + −
2 2 2 2 2
n n−1
C1 C2 1 C1 C2 1
= + − + + −
2 2 2 2 2 2
xn xn−1
= +
2 2
Hence, our ⋆⋆ is our so-called general solution to ⋆. The only factor we must find,
then, are our C1 , C2 . Recall our initial x1 = 1, x2 = 2. Plugging these into ⋆⋆, then,
gives
2
1 1
x1 = C 1 + C 2 − = 1; x2 = C 1 + C 2 − = 2,
2 2
which is simply a system of two equations for two unknowns, C1 , C2 . Solving
them42 , we have
5 4
C1 = , C2 = ,
3 3
hence we have the general formula
n
5 4 1
xn = + −
3 3 2
The RHS of this sum goes to zero, and thus our limit is
5
lim xn = .
n→∞ 3
■
Remark 2.17. From this general form, we can conclude, as in example 2.14, that x2n ≥
x2n−1 , x2n ≥ x2n+1 , since x2n > 53 , x2n+1 < 35 , x2n−1 < 53 ; ie, the same property that we used
to prove the previous example holds here.
Remark 2.18. Any recursively defined sequence of the form xn+1 = Axn + Bxn−1 , n > 1
x2 = Ax + B,
√ √
A+ A2 +4B A− A2 +4B
with solutions a = 2
,b = 2
. It may be that a, b ∈ C; we shall not consider
these cases. Indeed, we have:
x1 = C1 a + C2 b, x2 = C2 a2 + C2 b2 .
a b
= ab2 − ba2 ̸= 0.
2 2
a b
In the case a = b, or a = 0 or b = 0, then the determinant is also equal to 0, and we thus have
to use a different method. As long as the determinant is nonzero, then we have a valid specific
definition.
Remark 2.19. Note that nothing in this derivation assumed xn convergent; this form can
indeed be found even if xn diverges. It will simply also diverge.
Remark 2.20. The recursive relation xn+1 = Axn +Bxn−1 is a discrete analog of a differential
equation.
,→ Theorem 2.12
Let43 (xn ) be a contractive sequence with contractive constant K. Then, (xn ) is a
Cauchy sequence, and in particular, (xn ) converges.
43
Sketch: start with |xn − xm |,
and add/subtract each term
between xn and xm , use
Proof. Let n, m ∈ N such that n > m ≥ 2. Then, we have triangle inequality,
“substitute” in contractive
constant, collect like terms,
|xn − xm | = |xn − xn−1 + xn−1 − xn−2 + xn−2 − · · · − xm+1 + xm+1 − xm | and simplify. This creates an
upper bound for |xn − xm |,
≤ |xn − xn−1 | + |xn−1 − xn−2 | + · · · + |xm+1 − xm | which converges to 0, then
use this converges to define
≤ K n−2 |x2 − x1 | + K n−3 |x2 − x1 | + · · · + K m−1 |x2 − x1 | the epsilon to use in the
Cauchy definition.
= K m−1 |x2 − x1 | 1 + K + K 2 + · · · + K n−m−1
1 − K n−m
= K m−1 |x2 − x1 | by corollary 2.2
1−K
K m−1 |x2 − x1 |
<
1−K
K m−1 |x2 − x1 |
=⇒ |xn − xm | < ∀n > m ≥ 2
1−K
K m−1 K m−1
lim |x2 − x1 | = 0 =⇒ ∀ ε > 0, ∃N s.t. ∀ m > N, |x2 − x1 | < ε
1−K 1−K
K m−1
→ n > m ≥ N =⇒ |xn − xm | ≤ |x2 − x1 | < ε
1−K
K n−1
→ m > n ≥ N =⇒ |xn − xm | ≤ |x2 − x1 | < ε
1−K
=⇒ ∀ m, n ≥ N, |xm − xn | < ε, and (xn ) Cauchy
K m−1
|xn − xm | ≤ · |x2 − x1 |,
1−K
together with the fact that limn→∞ xn = X, whose convergence also implies by Algebraic
K m−1
|X − xm | ≤ |x2 − x1 |,
1−K
||a| − |b|| ≤ |a − b| ∀ a, b ∈ R,
Remark 2.23. The result that every contractive sequence is convergent is a simple example of
the more general “Fixed Point Theorems”; this proof can be generalized to the Banach Fixed
Point Theorem on arbitrary metric spaces. This is further used to establish the existence and
uniqueness of solutions of differential, integral equations.44 44
See the Picard-Lindelöf
Theorem
Remark 2.24. In the case of the recursively defined
1
xn+1 = (xn + xn−1 ),
2
we have that
1
|xn+2 − xn+1 | > |xn+1 − xn |,
2
that is, xn is a contractive sequence with K = 21 . The argument used to prove that this in-
equality implies (xn ) Cauchy is the same as the one we used to prove a general contractive
sequence is Cauchy.
⊛ Example 2.16
Let (xn ) be a sequence defined recursively by x1 = 2, xn+1 = 2 + 1
xn
. Prove that
the sequence converges and find its limit.
Proof. First, we note that xn ≥ 2 ∀ n. Now, we aim to show that (xn ) is contractive
We can now find the limit using the recursive definition; let X = limn→∞ xn . xn ≥
2, in particular, it is ̸= 0 for any n. Then, we have:
1 1
X = lim xn = lim 2 + =2+ =X
n→∞ n→∞ xn x
1
=⇒ X = 2 + =⇒ X 2 − 2X − 1 = 0
X
√
=⇒ X = 1 ± 2
√
1− 2 < 0, which can’t hold since xn ≥ 0 ∀ n, hence it must be that X = 1 +
√
2. ■
⊛ Example 2.17
Show that the sequence xn = 1 + 12 + · · · + n1 , n ≥ 1, diverges.
1 1 1 1 1
x2n − xn = + + ··· + ≥n· ≥ , ∗
|n + 1 n +{z2 2n} 2n 2
n terms, each ≥ 2n
1
which means that the sequence cannot be Cauchy hence it cannot be convergent
(see theorem 2.9).
1
|xn − xm | < ε = .
4
1
|x2N − xN | < ,
4
and 1+n
1
yn = 1+ .
n
We consider the following propositions regarding the sequences.
lim xn = lim yn ,
n→∞ n→∞
(1 + x)n ≥ 1 + nx
Proof. We proceed by induction; fixing x > −1, let S ⊆ N the set for which the inequality
holds. n = 1 =⇒ (1 + x)1 ≥ 1 + x, which clearly holds, ie 1 ∈ S. Suppose n ∈ S, that is,
(1 + x)n ≥ 1 + nx
≥0
z}|{
(1 + x)n · (1 + x) = (1 + x)n+1 ≥ (1 + nx)(1 + x) = 1 + nx + x + nx2 ≥ 1 + (n + 1)x
=⇒ n + 1 ∈ S
1
n+1 n+2 n+1
xn+1 1 + n+1 n+1 n + 2 (n + 2)n nn
= n = n+1 n
= ·
1 + n1 n + 1 [(n + 1)2 ]n
xn n
n
n2 + 2n
n+2
=
n + 1 n2 + 2n + 1
n
n + 2 n2 + 2n + 1 − 1
=
n+1 n2 + 2n + 1
n
n+2 1
= 1− 2
n+1 n + 2n + 1
n
n+2 1
= 1−
n+1 (n + 1)2
n
1 n
1− ≥1− ,
(n + 1)2 (n + 1)2
n + 2 n2 + n + 1
xn+1 n+2 n
≥ 1− = ·
xn n+1 (n + 1)2 n + 1 (n + 1)2
n3 + n2 + n + 2n2 + 2n + 2
=
n3 + 3n2 + 3n + 1
n3 + 3n2 + 3n + 2
= 3
n + 3n2 + 3n + 1
3 2
n + 3n + 3n + 1 1
= 3 2
+ 3 2
n + 3n + 3n + 1 n + 3n + 3n + 1
1
=1+ 3 >1
n + 3n2 + 3n + 1
hence, xn+1
xn
> 1 =⇒ xn+1 > xn ∀ n, ie it is strictly increasing. ■
(n+1)n+1
n+1 n+1 n+1
1 + n1
yn n+1
n nn+1
= n+2 = = · (n+2)
yn+1 1 n+2 n+2
n+2 n+1
1 + n+1 n+1 (n+1)n+1
n+1 n+1
n + 1 [(n + 1)2 ] n + 1 n2 + 2n + 1
= · =
n + 2 nn+1 (n + 2)n+1 n+2 n2 + 2n
n+1
n+1 1
= · 1+ 2
n+2 n + 2n
1 yn n+1 n+1
Bernoulli’s Inequalityx = 2 =⇒ ≥ 1+ 2
n + 2n yn+1 n+2 n + 2n
2
n + 1 n + 3n + 1
= ·
n+2 n2 + 2n
n3 + 3n2 + n + n2 + 3n + 1 n3 + 4n2 + 4n + 1
= =
n3 + 2n2 + 2n2 + 4n n3 + 4n2 + 4n
1
=1+ 3 >1
n + 4n2 + 4n
Hence, ∀ n, yn
yn+1
> 1 =⇒ yn > yn+1 , ie, it is strictly decreasing. ■
- (Case 1) n = k:
n n n+1
1 1 1 1
xn = 1 + < 1+ 1+ = 1+ = yn
n n n n
- (Case 2) n > k:
xn < y1 = 4 ∀ n,
and
2 = x1 < yk ∀ k,
hence (xn ) is bounded from above (by y1 , say) and (yn ) is bounded from below (by x1 ,
say). ■
Proof. (Of Step 5) Since (xn ) increasing and bounded from above, it is converging by Mono-
tone Convergence Theorem. Similarly, (yn ) is decreasing and bounded from below, hence
it too converges. We have too that
n+1 n
1 1 1 1
yn = 1 + = 1+ 1+ = 1+ xn
n n n n
Since limn→∞ 1 + 1
= 1, we have, from proposition 2.1, that
n
1
lim yn = lim 1 + · lim xn = lim xn ,
n→∞ n→∞ n n→∞
that is, (xn ) and (yn ) converge to the same limit, which we define as
n n+1
1 1
e ≡ lim 1+ = 1+ .
n→∞ n n
that limn→∞ Sn = e.
Remark 2.28. Note that L could be an empty set; however, if xn bounded, then L ̸= 0 by
Bolzano-Weirestrass Theorem. Further, L is a bounded subset of R.
,→ Proposition 2.14
Let (xn ) be a sequence. Then, x ∈ L iff ∀ ε > 0 the set {n : |xn − x| < ε} is infinite.
Proof. Suppose first that x ∈ L and let (xnk ) be a subsequence such that limk→∞ xnk = x.
Let ε > 0. Then ∃K ∈ N s.t. ∀ k ≥ K, |xnk − x| < ε.
We now show the converse. Suppose that ∀ ε > 0, the set {n : |xn − x| < ε} is infinite.
Take ε = 1, then the set {n : |xn − x| < 1} is an infinite set. Take n1 = min{n : |xn − x| <
1}. We can now define nk , where k = 2, 3, . . . recursively. Suppose that some nk chosen.
Then, take ε = k1 , and consider
1
{n : n > nk , |xn − x| < }.
k
This set has infinitely many elements, since {n : |xn − x| < k1 } is also infinite. We then
set nk+1 = min{n : n > nk , |xn − x| < k1 }, which defines a strictly increasing sequence
(nk )k≥1 of natural numbers such that for any k, |xnk − x| < k1 . So, limk→∞ |xnk − x| = 0
which gives that limk→∞ xnk = x, so x ∈ L . ■
1. limn→∞ xn = sup L
2. limn→∞ xn = inf L
Remark 2.29. The following proof shows even more, that is, limn→∞ xn = sup L and
limn→∞ xn ∈ L (same for limn→∞ ).
1
ynk +1 − ≤ xnk+1 ≤ ynk+1 = sup{xn : n ≥ nk+1 }.
k+1
So, we have constructed a strictly increasing sequence (nk ) of natural numbers such that
∀ k ≥ 1,
1
ynk +1 − ≤ xnk+1 ≤ ynk+1 . ⋆
k+1
and so, given this and ⋆, by the The Squeeze Theorem, limk→∞ xnk+1 = limn→∞ xn , and so
limn→∞ xn ∈ L . ■
,→ Corollary 2.3
Let (xn ) be a bounded sequence and α = limn→∞ xn , β = limn→∞ xn . Then, α, β ∈ L
(that is, they are limit points of (xn )), and for any x ∈ L , α ≤ x ≤ β (that is, L is a
closed set).
lim xn = ∞.
n→∞
That is,
( ∀ α ∈ R)(∃N ∈ N)( ∀ n ≥ N )(xn ≥ α).
⊛ Example 2.19
xn = n properly diverges to ∞; xn = −n properly diverges to −∞.
⊛ Example 2.20
Let C > 1. Then, limn→∞ C n = ∞.
C n = (1 + x)n ≥ 1 + nx.
,→ Proposition 2.15
Let (xn ) be increasing. Then limn→∞ xn = ∞ iff xn not bounded from above.
,→ Proposition 2.16
Let (xn ) be decreasing. Then limn→∞ xn = −∞ ⇐⇒ (xn ) not bounded from below.
,→ Proposition 2.17
Let (xn ), (yn ) be sequences such that xn ≤ yn ∀ n. Then
1. limn→∞ xn = ∞ =⇒ limn→∞ yn = ∞
2. limn→∞ yn = −∞ =⇒ limn→∞ xn = −∞
,→ Proposition 2.18
Let (xn ) be a sequence and c > 0. Then lim xn = ∞ ⇐⇒ lim c · xn = ∞. The
converse follows for c < 0 and → −∞.
Proof. ■
,→ Proposition 2.19
Let (xn ) and (yn ) be strictly positive sequences. Suppose that for some L > 0,
xn
lim = L.
n→∞ yn
Proof. Take ε = L
2
. Then, xn
yn
→ L =⇒ ∃N s.t. ∀ n ≥ N, L − ε < xn
yn
< L + ε ⇐⇒
L− L
2
< xn
yn
< L + L2 . So, ∀ n ≥ N, L2 < xn
yn
< 3L
2
=⇒ L
y
2 n
< xn < 32 Lyn . Hence,
if xn → ∞, it must be that yn → ∞, by the previous inequality. The other side of the
implication follows similarly. ■
,→ Proposition 2.20
Let (xn ), (yn ) be two sequences such that (xn ) is properly divergent and yn bounded.
Then their sum is also diverging.
Proof. ■
⊛ Example 2.21
xn = n, yn = −n
2
. xn → ∞, yn → −∞ while xn + yn → ∞.
In the case xn not bounded above, then we define limn→∞ xn = ∞. In this way, we
define limn→∞ xn for all sequences, regardless of convergence or boundedness.
,→ Proposition 2.21
Practically all previously proven properties of limsup/liminf hold with these general-
izations:
Remark 2.32. Not all concepts defined on convergent/bounded sequences extend easily to
properly divergent sequences. For instance, limn→∞ (xn + yn ) ≤ limn→∞ xn + limn→∞ yn
holds for bounded sequences xn , yn , but does not generally hold if limn→∞ xn = ∞, etc..
,→ Proposition 3.1
Let A ⊆ R, c ∈ R. Then, TFAE:
1. c is a cluster point of A
⊛ Example 3.1
Let A = (0, 1). Then, 0 is a cluster point of A.
⊛ Example 3.2
Let A = (0, 1) ∪ {5}. Is 5 a cluster point?
Proof. No; it is impossible to find arbitrarily (ε) close points to 5 in the set; ̸ ∃x ∈
A, x ̸= 5 such that 0 < |x − 5| < ε. Then, the set of all cluster points of A is equal
[0, 1]. ■
⊛ Example 3.3
Let A = { n1 : n ∈ N}. Then, c = 0 is the only cluster point of A.
⊛ Example 3.4
Let A = Q. Then, the set of cluster points is precisely R.
Proof. Take x ∈ R, ε > 0. Consider the interval (x, x + ε); by density of the
rationals, ∃q ∈ Q s.t. q ∈ (x, x + ε). Hence, ∃q ∈ Q, q ̸= x s.t. 0 < |x − q| < ε,
hence, x a cluster point of A. ■
Remark 3.2. “As x gets closer and closer to c, f (x) gets closer and closer to L”.
sin x
Remark 3.3. The point c may or may not be in A (for instance, limx→0 x
= 1). However,
it must be that c is a cluster point of A; this is what “allows” the arbitrary closeness to L in the
definition of a limit.
Remark 3.4. This definition is often called the “ε − δ” definition of functional limits. Quan-
tified, it states
( ∀ ε > 0)(∃δ > 0)( ∀ x ∈ A)(0 < |x − c| < δ =⇒ |f (x) − L| < ε).
⊛ Example 3.5
Let A = (0, ∞), let f (x) = x1 , x ∈ A, and let c ∈ A. Prove that limx→c f (x) = 1c .
1 1 c−x |x − c| δ
− = = < ,
x c xc |xc| |xc|
,→ Theorem 3.1
Let A ⊆ R, f : A → R, and let c be a cluster point of A. Then, TFAE:
1. limx→c f (x) = L.
2. For any sequence (xn ) ∈ A, xn ̸= c, such that lim xn = c, we have that the
sequence (f (xn )) converges to L, that is, limn→∞ f (xn ) = L.
(2. =⇒ 1.) Suppose not. Then, ∀ ε > 0, we can find δ > 0 such that ∀ x ∈ A s.t. 0 <
|x − c| < δ we have |f (x) − L| < ε. But then, this means that ∃ε0 > 0 s.t. ∀ δ > 0, ∃x ∈
A, x ̸= C s.t. 0 < |x − c| < δ and |f (x) − L| ≥ ε0 . So, for this ε0 > 0, we can take δ = n1 ,
which gives xn ∈ A, xn ̸= c, such that 0 < |xn − c| < 1
n
and |f (xn ) − L| ≥ ε0 . This gives
us a sequence (xn ) ∈ A, xn ̸= c, such that limn→∞ |xn − c| = 0 =⇒ limn→∞ xn = c, and
|f (xn ) − L| ≥ ε0 ∀ n. But this means that we have a sequence xn s.t. lim xn = c, and the
sequence (f (xn )) does not converge to L. But this contradicts 2.; hence, we have come to
a contradiction, and 1. holds. ■
,→ Proposition 3.2
A functional limit is unique. That is, if f : A → R and c a cluster point of A, if
limx→c f = L an limx→c f = M, L = M .
(ε − δ) Let ε > 0. Since limx→c f (x) = L, ∃δ1 > 0 s.t. ∀ x ∈ A s.t. 0 < |x − c| < δ1 we
have |f (x) − L| < ε. Since limx→c h(x) = L, we have that ∃δ2 > 0 s.t. ∀ x ∈ A s.t. 0 <
|x − c| < δ2 we have |h(x) − L| < ε. Let δ = min{δ1 , δ2 } and let x be such that x ∈ A, 0 <
|x − c| < δ. Then,
Remark 3.5. Note the similarity between the ε − δ proofs above and the proofs of correspond-
ing properties for sequences.
This sequence does not converge, and so limx→0 sin x1 does not exist. ■
hence, limn→∞ f (xn ) = 1 ̸= limn→∞ f (yn ) = 0, so the limit does not exist.
Consider now limx→0 x cos x1 . Fix ε > 0, and take δ = ε, then, we have that
∀ x s.t. 0 < |x − 0| < δ. Then, we have by properties of cos,
1
x cos ≤ |x| < δ = ε,
x
Proof. (Seq’n) Define L := limx→c f (x). Then, we have that ∀ xn ∈ A\{c} s.t. (xn ) →
c, limn→∞ f (xn ) = L. We can write, then,
p p p 2
L = lim f (xn ) = lim f (xn ) f (xn ) = lim f (xn ) ,
n→∞ n→∞ n→∞
and taking the square root of both sides, we have the desired result. Note that this
√
used the assumption that ∃ limn→∞ xn =⇒ ∃ limn→∞ xn .
(ε − δ)
lim f (x) = L,
x→c+
lim f (x) = L,
x→c−
Remark 3.6. Sometimes, but not always, the right/left endpoints are equivalent to the “usual”
limit.
Let ε > 0. Take δ > 0. Then, ∀ x s.t. 0 < x < δ, |f (x) − 1| = |1 − 1| = 0 < ε,
hence limx→0+ f (x) = 0.
,→ Proposition 3.3
Let A ⊆ R, f : A → R, and let c be a cluster point of the sets A ∩ (c, ∞) and
A ∩ (−∞, c). TFAE:
1. limx→c f (x) = L
Proof. (1. =⇒ 2.) Let ε > 0 and δ s.t. ∀ x ∈ A s.t. 0 < |x − c| < δ =⇒ |f (x) − L| < ε.
Then, we have that |f (x) − L| < ε ⇐= 0 < x − c < δ, and moreover, |f (x) − L| <
ε ⇐= −δ < x − c < 0, that is, limx→c+ f (x) = limx→c− f (x) = L, hence 2. holds.
,→ Theorem 3.4
Let56 A ⊆ R, f : A → R, and let c be a cluster point of the set A ∩ (c, ∞).Then, TFAE:
1. limx→c+ f (x) = L
2. For any sequence (xn ) ∈ A s.t. xn > c ∀ n, and limn→∞ xn = c, we have that
limn→∞ f (xn ) = L.
Proof. ( =⇒ )
( ⇐= ) ■
56
Abbott, 4.3E1 (Theorem 4.3.2)
⊛ Example 3.10
Let A = (−∞, 0) ∪ (0, ∞) and let f (x) = 1
x2
. Show that limx→0 f (x) = ∞.
⊛ Example 3.11
1. Give a sequential characterization of limx→c f (x) = ∞ and − ∞.
2. Give the definition of right/left hand limits going to infinity, limx→c+ f (x) =
∞ and − ∞, limx→c− f (x) = ∞ and − ∞.
Proof. (1.) Let M ∈ R. Since limx→c f (x) = ∞, ∃δ > 0 s.t. ∀ x ∈ A s.t. 0 < |x − c| <
δ, f (x) ≥ M . But then g(x) ≥ f (x) ∀ x, hence g(x) ≥ M =⇒ limx→c g(x) = ∞. ■
sin x
⊛ Example 3.12: x
at infinity
Let A = (0, ∞) and f (x) = sin x
x
. Prove that limx→∞ f (x) = 0.
⊛ Example 3.13
“Sequentialize” limits at infinity.
lim f (x) = 0.
x→∞
Proof.
3.4 Continuity
( ∀ ε > 0)(∃δ > 0)( ∀ x ∈ A)(|x − c| < δ =⇒ |f (x) − f (c)| < ε).
1. f continuous at c
Remark 3.7.
Remark 3.8. This theorem can be directly deduced from sequential characterization of func-
tional limits.
1. ∀ k ∈ R, kf continuous at c
2. h = f + g continuous at c
3. h = f · g continuous at c
4. If g(x) ̸= 0 ∀ x ∈ A, h = f
g
continuous at c.
Proof. (Of 3.) Let (xn ) ∈ A s.t. limn→∞ xn = c. Since f, g continuous at c, we have that
limn→∞ f (xn ) = f (c) and limn→∞ g(xn ) = g(c). By algebraic properties of limits, then,
limn→∞ f (xn )g(xn ) = f (c)g(c) and so ∀ (xn ) ∈ A s.t. limn→∞ xn = c, we have that
limn→∞ h(xn ) = h(c) where h = f · g and thus h continuous at c. ■
f (A) = {f (x) : x ∈ A} ⊆ B
so that the composite function h(x) = g ◦ f (x) = g(f (x)) is well defined on A.
Suppose c ∈ A such that f continuous at c and g continuous at f (c). Then, h also
continuous at c.
(ε − δ) Fix ε > 0. Since g is continuous at f (c), ∃δ ′ > 0 s.t. ∀ y ∈ B s.t. |y − f (c)| < δ,
|g(y) − g(f (c))| < ε.
Since f continuous at c, ∃δ ′ > 0 s.t. ∀ x ∈ A s.t. |x − c| < δ ′ , |f (x) − f (c)| < δ ′ . Then,
for such x, |g(f (x)) − g(f (c))| < ε and the proof is complete, taking h = g ◦ f . ■
⊛ Example 3.15
f : R → R, f (x) = x.
Proof. ■
Proof. Let a > 0 be rational. Then, f (a) > 0, by construction of the function. Let
(xn ) ∈ J s.t. (xn ) → a. Then, limn→∞ f (xn ) = 0, despite f (a) > 0, hence f is not
continuous at a. ■
Remark 3.9. Since c a cluster point of A∪{c}, we have that F continuous at c iff limx→c F (x) =
L ⇐⇒ limx→c f (x) = L. Hence, if f : A → R, c a cluster point of A, c ∈
/ A, and
limx→c = L, F is continuous at c. If limx→c f (x) DNE, f cannot be extended.
⊛ Example 3.19
f (x) = x sin x1 , defined on A = (−∞, 0) ∪ (0, ∞). 0 a cluster point of A. Note that
limx→0 f (x) = 0, since |f (x)| ≤ |x|, so if we extend f to 0 by setting f (0) = 0,
then the extended function is continuous at 0.
§3.5 Functional Limits and Continuity: Continuity on Bounded & Closed Interval p. 78
,→ Definition 3.10: Bounded Function
Let A be a set. A function f : A → R is called bounded if ∃M > 0 s.t. |f (x)| ≤
M ∀ x ∈ A.
So, for any n, ∃xn ∈ [a, b] s.t. |f (xn )| > n. Then, (xn ) is a sequence in [a, b] and by
the Bolzano-Weirestrass Theorem, this sequence has a subsequence (xnk ) that converges to
some x ∈ [a, b], that is,
lim xnk = x.
k→∞
so (f (xnk )) is a converging sequence of real numbers that is not bounded, which contradicts
the fact that any converging sequence is bounded. ■
§3.5 Functional Limits and Continuity: Continuity on Bounded & Closed Interval p. 79
,→ Theorem 3.8
Let f : [a, b] → R be a continuous function. Then, f has an absolute maximum and
absolute minimum on [a, b].
By theorem 3.7, f is bounded on [a, b], so there exists some M > 0 such that
So, the set f ([a, b]) is bounded, and by Axiom Of Completeness, s = sup(f ([a, b])) exists.
Hence, s ≥ f (x) ∀ x ∈ [a, b]. We aim to show then that ∃x ∈ [a, b] s.t. s = f (x).
Let n ∈ N. Since s − 1
n
is not an upper bound of f ([a, b]),
1
∃xn ∈ [a, b] s.t. s − < f (xn ) ≤ s. ⊛
n
By Bolzano-Weirestrass Theorem, (xn ) has a converging subsequence (xnk ). Let x = limk→∞ xnk .
By the sequential characterization of continuity, then, we have that f (x) = limk→∞ f (xnk ).
By ⊛, we have
1
s− < f (xnk ) ≤ s.
nk
Moreover, we have that nk ≥ k =⇒ 1
nk
≤ 1
k
=⇒ − n1k ≥ − k1 . Hence,
1
s− < f (xnk ) ≤ s,
k
and so by the squeeze theorem, limk→∞ f (xnk ) = s = f (x), and the proof is complete. ■
Proof. Let S = {x ∈ [a, b] : f (x) ≤ 0}. S ̸= ∅ since a ∈ S. S also bounded (it is a subset
of a bounded interval). Let c = sup S (exists by AC). We claim this c is the point as defined
§3.5 Functional Limits and Continuity: Continuity on Bounded & Closed Interval p. 80
in the theorem; we aim to show that f (c) = 0.
So, ∀ x ∈ (b − δ, b], we have that f (x) > 0. So, if we take [b − 2δ , b], then for every x ∈ this
interval f (x) > 0 and so S ⊆ [a, b − 2δ ), and thus c = sup S ≤ b − 2δ . Hence, ∃δ such that
a + δ ≤ c ≤ b − 2δ . So, c satisfies a < c < b.
(f (c) ≥ 0) Since c < b, we can find (xn ) ∈ [a, b] s.t. xn > c, limn→∞ xn = c (xn = c + n1 ,
for instance). We must have that f (xn ) > 0; otherwise, f (xn ) ≤ 0 =⇒ xn ∈ S, and since
we have xn > c (by construction), this would contradict the fact that c an upper bound for
S. So, we have that limn→∞ xn = c =⇒ limn→∞ f (xn ) = f (c) ≥ 0, that is, f (c) ≥ 0.
Thus, having show both f (c) ≤ 0 and f (c) ≥ 0, we conclude that ∃c ∈ [a, b] s.t. a <
c < b, where f (c) = 0, and the proof is complete. ■
3.6 Intervals in R
• [a, b] = {x : a ≤ x ≤ b} ⊆ R
• [a, b) = {x : a ≤ x < b} ⊆ R
• (a, b] = {x : a < x ≤ b} ⊆ R
(Unbounded Intervals)
• [a, ∞) = {x : x ≥ a}
• (a, ∞) = {x : x > a}
• (−∞, a] = {x : x ≤ a}
• (−∞, a) = {x : x < a}
• R = (−∞, ∞)
Remark 3.10. If you take any interval and any two points x < y in the interval, then [x, y]
is completely contained within the given interval.
,→ Theorem 3.10
Let S ⊆ R that contains more than two points. Suppose S has the property that
∀ x, y ∈ S s.t. x < y, [x, y] ⊆ S. Then, S is an interval.
Proof. Suppose S bounded. Then, a = inf S, b = sup S exist. Then, for any x ∈ S, a ≤
x ≤ b, so S ⊆ [a, b]. Let now a < z < b. z < b =⇒ z not an upper bound of S so
∃y ∈ S s.t. z < y. z > a =⇒ z not a lower bound of S so ∃x ∈ S s.t. x < z. Then,
x < z < y, x, y ∈ S, so [x, y] ⊆ S =⇒ z ∈ S. So, (a, b) ⊆ S ⊆ [a, b] and thus S must be
a bounded interval (one of those types defined above). ■
Proof. • (Case 1 : a < b) Consider h(x) = f (x) − k on the closed and bounded interval
[a, b]. Note that h(a) = f (a) − k < 0, and h(b) = f (b) − k > 0. By Location of the
Roots, there exists a a < c < b s.t. h(c) = 0 = f (c) − k =⇒ f (c) = k, as desired.
• (Case 2: a > b) Consider h(x) = k − f (x) on the closed and bounded interval [b, a].
The remainder of the proof follows identically to (Case 1).
,→ Theorem 3.12
Let I = [a, b] and f : I → R a continuous function. Let k be s.t. inf f (I) ≤ k ≤
sup f (I). Then, ∃c ∈ I s.t. f (c) = k.
Proof. Recall that m = inf f (I) is the absolute minimum of f on I and M = sup f (I) is
the absolute maximum of f on I. Moreover, ∃x, x ∈ I s.t. f (x) = M, f (x) = m. Hence,
we have that our k satisfies
f (x) ≤ k ≤ f (x).
,→ Theorem 3.13
Let I = [a, b] and f : I → R a continuous function. Then, f (I) is also a bounded and
closed interval.
Proof. Let m = inf f (I), M = sup f (I). Then, for any x ∈ [a, b], m ≤ f (x) ≤ M , hence,
f (I) ⊆ [m, M ]. OTOH, by theorem 3.12, for any m ≤ k ≤ M , ∃c ∈ [a, b] s.t. f (c) = k,
hence, [m, M ] ⊆ f (I), and thus f (I) = [m, M ] and the proof is complete. Moreover, f (I)
is precisely [inf f (I), sup f (I)]. ■
,→ Theorem 3.14
Let I be an interval in R. Let f : I → R be a continuous function. Then, f (I) is also
an interval.
Proof. We assume f is not just a constant function. We aim to show that if α, β ∈ f (I), α <
β, then [α, β] ⊆ f (I), that is, f (I) an interval.
Let a, b ∈ I be such that f (a) = α, f (b) = β. We have that f (a) < f (b), so for any
k s.t. α ≤ k ≤ β, by Bolzano’s Intermediate Value Theorem, ∃c ∈ I s.t. f (c) = k. Hence,
[α, β] ⊆ f (I), and the proof is complete. ■
Remark 3.11. This argument does not specify the actual “shape” of the intervals f (I) look
like.
Remark 3.12. Recall that in the definition of continuity, the “choice” of δ depended both on
c (the point in the domain) and ε. Uniform continuity defines a manner in which δ can be
chosen without relying on c; if this is the case for a function f : A → R, we say that f is
uniformly continuous on A.
( ∀ ε > 0)(∃δ > 0)( ∀ c ∈ A)( ∀ x ∈ A)(|x − c| < δ =⇒ |f (x) − f (c)| < ε).
Remark 3.13. The difference, quantifiers-wise, is the position of the ( ∀ c ∈ A); since here the
“choice” of c comes after the choice of δ, δ is independent, in contrast with “local” continuity.
⊛ Example 3.21
Let f (x) = x. Then, f is uniformly continuous on R.
⊛ Example 3.22
Let f (x) = x2 . Then, f is not uniformly continuous on R.
δ
|x − c| = < δ,
2
but
1 1 δ
x2 − c2 = ( + δ)2 − ( + )2
δ δ 2
3δ 2
= ··· = 1 + > 1,
4
⊛ Example 3.23
√
Let f (x) = x. Then, f is uniformly continuous on [0, ∞).
ε2
Proof. Let ε > 0. Take δ = 2
. Let x, c ≥ 0, and suppose |x − c| < δ. We consider
two cases:
2
• (Case 1) x, c ∈ [0, ε4 ). Then,
√ √ √ √
x− c ≤ x+ c
r r
ε2 ε2 ε
< + = 2 · = ε.
4 4 2
ε2
• (Case 2) Either x or c ≥ 4
. then
√ √
√ √ √ √ x+ c |x − c|
x − c = ( x − c) √ √ =√ √
x+ c x+ c
ε2
δ 2
< ε = ε = ε.
2 2
⊛ Example 3.24
Let f (x) = sin x1 is not uniformly continuous on (0, 1].
π
1 1 2
|x − c| = − π = <δ
nπ nπ + 2
nπ(nπ + π2 )
Then, we have
1 1
f (x) − f (c) = sin 1 − sin 1
nπ (2n+1) π2
1
= |(−1)n | = 1 > ,
2
a contradiction. ■
,→ Theorem 3.15
Let f : A → R be a continuous function. TFAE:
2. ∃ε0 > 0 and two sequences (xn ), (yn ) ∈ A s.t. limn→∞ (xn − yn ) = 0 and
|f (xn ) − f (yn )| ≥ ε0 ∀ n.
Proof. (1. =⇒ 2.) For f to be not uniformly continuous, then it is not true that ∀ ε >
0∃δ > 0 s.t. ∀ x, y ∈ A, if |x − y| < δ, |f (x) − f (y)| < ε. That is, ∃ε0 > 0 s.t. ∀ δ > 0, one
can find x, y ∈ A s.t. |x − y| < δ and |f (x) − f (y)| ≥ ε0 .
1 1
f (xn ) − f (yn ) = n2 + 2 + 2
− n2 = 2 + 2 ≥ 2,
n n
Proof. Let xn = 1
nπ+ π2
, yn = 1
nπ
. Both of these converge to 0, hence their differences
do as well. OTOH, |f (xn ) − f (yn )| = |−1 − 0| = 1 ≥ 1, hence, f is not uniformly
continuous with ε0 = 1.s ■
,→ Theorem 3.16
Let f : [a, b] → R be a continuous function. Then, f is uniformly continuous on [a, b].
Since [a, b] bounded, by Bolzano-Weirestrass Theorem, the sequence (xn ) has a conver-
gent subsequence (xnk ) that converges to z ∈ [a, b] (since [a, b] closed). We can write
Hence, by the The Squeeze Theorem, |ynk − z| converges to 0 so (ynk ) also converges to z,
that is
lim xnk = lim ynk = z.
k→∞ k→∞
Since f continuous on [a, b], we have that limk→∞ f (xnk ) = f (z) and limk→∞ f (ynk ) =
f (z), and so
lim (f (xnk ) − f (ynk )) = z − z = 0,
k→∞
Proof. Let ε > 0. Since f uniformly continuous on A, there is δ > 0 s.t. ∀ x, y ∈ A, |x − y| <
2. f can be at the end points a, b such that it is continuous on the closed interval
[a, b].
Proof. (2. =⇒ 1.) If f can be extended to a, b so that it is continuous on [a, b], then it is
also uniformly continuous by theorem 3.16. Then, f is also uniformly continuous on any
subset [a, b], in particular, on (a, b) ⊆ [a, b].
(1. =⇒ 2.) Let (xn ) ∈ (a, b) that converges to a. (xn ) Cauchy, and by theorem 3.17, (f (xn ))
also Cauchy, hence L = limn→∞ f (xn ) exists; define (“extend”) f (a) = L. It remains to
show that f continuous with this extension.
Let (un ) be an arbitrary sequence in (a, b) such that limn→∞ un = a. Let ε > 0. Since
f is uniformly continuous on (a, b), then ∃δ > 0 s.t. ∀ x, y ∈ (a, b), |x − y| < δ =⇒
|f (x) − f (y)| < 2ε . Now, we have that limn→∞ un = a, limn→∞ xn = a, so limn→∞ (un −
xn ) = 0. Hence, ∃N1 s.t. ∀ n ≥ N1 , |un − xn | < δ. This implies, then, that ∀ n ≥ N1 , we
have |f (un ) − f (xn )| < 2ε .
We have, by our extension, that limn→∞ f (xn ) = L, hence, ∃N2 s.t. ∀ n ≥ N2 , |f (xn ) − L| <
ε
2
. Let, now, N = max{N1 , N2 }. Then, ∀ n ≥ N ,
that is, ∀ n ≥ N, |f (un ) − L| < ε. Hence, for any arbitrary (un ) ∈ (a, b) such that (un ) →
a, lim n → ∞un = L, hence, as we have set f (a) = L, by sequential characterization of
continuity, f is continuous at a.
,→ Proposition 3.6
f : A → R increasing on A ⇐⇒ g = −f decreasing on A.
,→ Theorem 3.19
Let I ⊆ R, f : I → R be increasing. Let c ∈ I, where c not an endpoint of I. Then:
Proof. We prove for 2.; 1. follows identically. Let A := {f (x) : x ∈ I, x > c}. Note that
A ̸= ∅, since c not an endpoint of I by construction and hence ∃x ∈ I s.t. x > c.
Since f increasing, we have that x > c =⇒ f (x) ≥ f (c) hence A bounded below by
f (c), and thus L := inf A exists. Let ε > 0; since L + ε not a lower bound for A, there
exists some xε ∈ I s.t. L + ε > f (xε ) ≥ L. Take δ = xε − c. Since f increasing, we have
that
c < x < c + δ = xε =⇒ |f (x) − L| = f (x) − L ≤ f (xε) − L < ε.
lim f (x) = L.
x→c+
,→ Corollary 3.1
Let I ⊆ R and f : I → R be increasing on I. Take c ∈ I such that c not an endpoint
of I. TFAE:
1. f continuous at c
Proof. Note first that 1. ⇐⇒ 2. does not relate to f increasing; rather, it follows from the
left-hand limit equals right-hand limit iff limit holds; this holds if f continuous at c.
,→ Theorem 3.20
Let I ⊆ R be an interval and f : I → R be increasing. Then the set D ⊆ I of points
at which f is discontinuous is either finite or countable.
a ≤ x1 < x2 < · · · xn ≤ b.
jf (x1 ) + · · · + jf (xn ) = lim+ f (x) − lim− f (x) + · · · lim+ f (x) − lim− f (x)
x→x1 x→x1 x→xn x→xn
n−1
!
X
= lim+ f (x) − lim− f (x) + lim f (x) − lim
−
f (x)
x→xn x→x1 x→x+
k x→xk+1
k=1
| {z }
≤0
≤ f (b) − f (a) ⊛
From this, we have that for any k ∈ N, there are at most k points in I such that jf (x) ≥
f (b)−f (a)
k
; suppose there were k + 1 points; then,
k+1
f (b) − f (a) ≥ jf (x1 ) + · · · + jf (xk+1 ) ≥ (f (b − f (a))) > f (b) − f (a)⊥.
k
S∞
Let D := {x ∈ I : f discontinuous at x} = {x ∈ I : jf (x) > 0} = k=1 {x ∈ I : jf (x) ≥
f (b)−f (a)
k
}. This is a countable union of finite sets, hence D itself is finite or countable given
I = [a, b].
∞
[
I= [ak , bk ],
k=1
for some sequences an , bn , that is, as a countable union of bounded and closed intervals ⊖.
Hence, we can write our set DI defined above as
[
DI = DS[an ,bn ] = D[an ,bn ] ,
S∞
• R= k=1 [−k, k]
S∞
• (a, b) = k=1 [a + b−a
3k
,b − b−a
3k
]
S∞
• (−∞, b] = k=1 [−k − |b|, b]
S∞
• (−∞, b) = k=1 [−k − |b|, b − 1
2k
]
• ...
,→ Theorem 3.21
Let I ⊆ R be an interval and let f : I → R be a continuous function. Let S :=
f (I). Suppose f strictly increasing. Then, for any y ∈ S, there is precisely one
x ∈ I s.t. f (x) = y.
Proof. Suppose x1 , x2 s.t. f (x1 ) = f (x2 ) = y. f strictly increasing, hence both x1 > x2
and x1 < x2 are impossible, hence x1 = x2 . ■
(f ◦ g)(y) = y ∀ y ∈ S; (g ◦ f )(x) = x ∀ x ∈ I.
,→ Proposition 3.7
If f strictly increasing, so is f −1 .
Then, there is no y ∈ S s.t. g(y) = x, by our construction. But this contradicts the fact that
g(S) = I, and hence g must be continuous on S, ■
4 Differentiation
4.1 Introduction
,→ Theorem 4.1
If f : I → R has a derivative at c ∈ I, then f is continuous at c.
f (x) − f (c)
f (x) − f (c) = (x − c).
x−c
f (x) − f (c)
lim = f ′ (c),
x→c x−c
f (x) − f (c)
lim(f (x) − f (c)) = (lim lim(x − c)) = f ′ (c) · 0 = 0,
x→c x→c x−c x→c
2. f + g is differentiable at c;
Proof. 1.
2.
f (x)
h(x) − h(c) g(x)
− fg(c)
(c)
=
x−c x−c
f (x)g(c) − f (c)g(x)
=
(x − c)(g(x)g(c))
z }| { z }| {
f (x)g(c) − f (c)g(c) + f (c)g(c) − f (c)g(x)
=
(x − c)g(x)g(c)
(f (x) − f (c))g(c) (g(x) − g(c))f (c)
= − ⊛
(x − c)g(x)g(c) (x − c)g(x)g(c)
f ′ (c)g(c) g ′ (c)f (c)
lim ⊛ = lim −
x→c x→c g(x)g(c) g(x)g(c)
f (c)g(c) − g ′ (c)f (c)
′
=
[g(c)]2
,→ Definition 4.2
If f ′ exists on every point c ∈ I, then we say that f is differentiable on I. This gives
a function
f ′ : I → R.
f (x) = xn+1 = xn x
power rule
=⇒ f ′ (x) = nx n−1
| {z } ·x + x
n
assumption
= (n + 1)xn
⊛ Example 4.2
Prove that d
dx
sin x = cos x and d
dx
cos x = − sin x.
1. f is differentiable at c;
We have, then,
f (x) − f (c)
lim φ(x) = lim = f ′ (c) = φ(c),
x→c x→c x−c
hence, φ is continuous at c. For x ̸= c, the desired relation f (x) = f (c) + φ(x)(x − c) holds
by definition.
h = g ◦ f, h : J → R,
Proof. Given f ′ (c) exists, the Caratheodory theorem gives that there exists a function φ :
,→ Theorem 4.5
Let I be an interval and f : I → R be a strictly increasing continuous function. Let
J = f (I) and g : J → R be the inverse of f . Suppose f differentiable at c, f ′ (c) ̸= 0.
Then, g is differentiable at f (c), and g ′ (f (c)) = 1
f ′ (c)
.
n
x1 + x2 + · · · + xn 1X
yn = = xk
n n k=1
lim xn = lim yn .
n→∞ n→∞
xn+1 − xn
lim =L
n→∞ yn+1 − yn
xn+m ≤ xn + xm
holds. For any subadditive sequence (xn ), its limit exists, and moreover,
xn
lim xn = inf{ : n ∈ N}.
n→∞ n
xn+1
≥ q > 1.
xn