Study Giude Lecture Notes 32 PDF
Study Giude Lecture Notes 32 PDF
Study Giude Lecture Notes 32 PDF
Contents
2 Sequences 15
2.1 Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Bounded and Monotonic Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4 Differentiation 37
4.1 Revision – Self Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.2 The Chain Rule and Inverse Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5 Series 43
5.1 Definitions and Basic Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.2 Convergence Tests for Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3 Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6 Riemann Integration 51
6.1 Suprema and Infima (self study) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2 Riemann Sums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.3 Refinements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.4 The Riemann Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.5 Properties of the Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Introduction
Aims
Introductory Analysis is a fundamental part of the mathematics curriculum and the results and methods
of this course will be indispensable in any further course in the School which involves Analysis. You
will have seen many of the definitions and theorems in Calculus I. So you may ask what is the difference
between Calculus I and Introductory Analysis?
In Calculus you have learnt definitions and theorems which then were used to solve standard problems.
Whenever it was not too hard, proofs were provided, but often concepts were explained rather vaguely
with some sketches. In particular the fundamental concept of limit and the definition of real numbers
have not been given formally. Indeed, it is the step from rational numbers to real numbers which is crucial
in Real Analysis (and also in Calculus).
You have seen proof techniques and axioms in Algebra I. This combined with the axiom of the real
numbers gives you the basis for all definitions, theorems and proofs in this course, spelt out rigorously
using mathematical logic.
Students therefore have to know definitions and theorems and have to know the proofs which are provided
in class. It is more important to understand definitions, theorems and proofs than learning them by heart.
Any valid reformulation in your own words will be fully accepted.
You have seen some proofs in Calculus I (or even Algebra I) of results which you will re-encounter in
Introductory Analysis. These proofs are provided in this study guide and may not be presented in class.
They will be clearly indicated by their reference to first year. Although these proofs will not be examined
directly, you nevertheless should recollect them.
In this course, you will only see naïve logic and naïve set theory, as much as needed to understand
Introductory Analysis.
If you think that the concepts in Introductory Analysis are not that obvious, you are right! Recall that
Newton and Leibniz are credited as having created calculus at the end of the 17th century, whereas the
rigorous definitions which you will encounter in Introductory Analysis were only introduced by Bolzano
and Cauchy in the early 19th century, and the axiomatic definition of the real number system only evolved
in the second half of the 19th century.
Throughout this study guide you will find additional notes. This is material which will not be examined
and, in general, not discussed in class. However, some notation may be introduced there which may be
used somewhere else in the course.
Outcomes
At the end of this course, students should be able to
• solve problems and perform computations using concepts and methods based on the theory
• apply theory
• form hypotheses
• reason mathematically
Recommended textbooks
H. Amann, J. Escher, Analysis I, Springer (electronic).
K. A. Ross, Elementary Analysis, Springer.
M. Spivak, Calculus, Cambridge University Press.
In its presentation, the book by Ross is closest to this course. However, if students would like to have
more reading on the transition from Calculus to Analysis, they may find Spivak’s book useful. As this
is only recommended reading, you may use any edition of these textbooks. The electronic version of
Amann’s textbook can be accessed via the Wits library catalogue. Volume I covers more than what will
be covered in this course, and also gives an alternative definition of the real numbers. Volumes I to III
are recommended as supplementary reading for this and further Analysis courses up to Honours level.
Chapter 1
Proof. (a) Let 0, 0′ ∈ ℝ such that 𝑎 + 0 = 𝑎 and 𝑎 + 0′ = 𝑎 for all 𝑎 ∈ ℝ. We must show that 0 = 0′ :
0 = 0 + 0′
= 0′ + 0 by (A2)
′
=0
(b) Let 𝑎 ∈ ℝ and 𝑎′ , 𝑎′′ ∈ ℝ such that 𝑎 + 𝑎′ = 0 and 𝑎 + 𝑎′′ = 0. We must show that 𝑎′ = 𝑎′′ :
𝑎′ = 𝑎′ + 0 by (A3)
′ ′′
= 𝑎 + (𝑎 + 𝑎 )
= (𝑎′ + 𝑎) + 𝑎′′ by (A1)
′ ′′
= (𝑎 + 𝑎 ) + 𝑎 by (A2)
′′
=0+𝑎
= 𝑎′′ + 0 by (A2)
′′
=𝑎 by (A3)
𝑎 + 𝑥 = 𝑎 + (𝑏 − 𝑎) = 𝑎 + (𝑏 + (−𝑎))
= 𝑎 + ((−𝑎) + 𝑏) by (A2)
= (𝑎 + (−𝑎)) + 𝑏 by (A1)
=0+𝑏 by (A4)
=𝑏+0 by (A2)
=𝑏 by (A3)
M. Axioms of multiplication
(M3) One: There is a real number 1 such that 1 ≠ 0 and 𝑎 ⋅ 1 = 𝑎 for all 𝑎 ∈ ℝ.
(M4) Multiplicative inverse: For each 𝑎 ∈ ℝ with 𝑎 ≠ 0 there is 𝑎−1 ∈ ℝ such that 𝑎𝑎−1 = 1.
Additional Notes. 1. Any set with the operations +, ⋅ satisfying the axioms (A1)–(A4), (M1)–(M4) and
(D) is called a field.
2. You will encounter detailed discussions of fields in algebra courses.
3. The multiplicative inverse of a number is also briefly called the inverse of a number.
4. The set of nonzero real numbers, ℝ ⧵ {0}, is an Abelian group with respect to multiplication.
(c) For all 𝑎, 𝑏 ∈ ℝ with 𝑎 ≠ 0, the equation 𝑎𝑥 = 𝑏 has a unique solution. This solution is 𝑥 = 𝑎−1 𝑏.
(d) ∀𝑎 ∈ ℝ ⧵ {0}, (𝑎−1 )−1 = 𝑎.
(e) ∀𝑎, 𝑏 ∈ ℝ ⧵ {0}, (𝑎𝑏)−1 = 𝑎−1 𝑏−1 .
(f) ∀𝑎 ∈ ℝ ⧵ {0}, (−𝑎)−1 = −𝑎−1 .
(g) 1−1 = 1.
Proof. See tutorials. The proofs are similar to those of Theorem 1.1.
𝑏 1
Notation. 𝑎−1 𝑏 is also written as , In particular, 𝑎−1 = .
𝑎 𝑎
Tutorial 1.1.1. 1. Prove Theorem 1.3.
Next we give the axioms for the set of positive real numbers. It is convenient to use the notation 𝑎 > 0
for positive numbers 𝑎.
O. The order axioms
(O1) Trichotomy: For each 𝑎 ∈ ℝ, exactly one of the following statements is true:
𝑎 > 0 or 𝑎 = 0 or −𝑎 > 0.
The definition of positivity of real numbers gives rise to an order relation for real numbers:
Definition. Let 𝑎, 𝑏 ∈ ℝ. Then 𝑎 is called larger than 𝑏, written 𝑎 > 𝑏, if 𝑎 − 𝑏 > 0.
Notes. 1. Since 𝑎 − 0 = 𝑎, the notation 𝑎 > 0 is consistent.
2. It is convenient to introduce the following notations:
𝑎 ≥ 𝑏 ⇔ 𝑎 > 𝑏 or 𝑎 = 𝑏,
𝑎 < 𝑏 ⇔ 𝑏 > 𝑎,
𝑎 ≤ 𝑏 ⇔ 𝑎 < 𝑏 or 𝑎 = 𝑏.
Proof. (a)
Note. There is still one axiom missing, the axiom of Dedekind completeness. However, we will postpone
the formulation of this axiom to the next section since we need some further definitions.
if 𝑥 ≥ 0,
{
𝑥
|𝑥| =
−𝑥 if 𝑥 < 0.
(b) 𝑎 ≥ 1 ⇒ 3
𝑎+1
3
𝑎
,
9 10
(c) 𝑎 > 1 ⇒ 𝑎 𝑎−1
,
1 1
(d) 𝑎 > 1 ⇒ 𝑎2 𝑎
,
(e) 𝑎 ≥ 2 ⇒ 1
𝑎2 −1
1
𝑎
,
−3 −2
(f) 𝑎 > 3 ⇒ .
5. Let 𝑥 ≥ 0 and 𝑦 ≥ 0. Show that 𝑥 < 𝑦 ⇔ 𝑥2 < 𝑦2 .
𝑎 𝑎−1
(𝑎, 𝑏) = {𝑥 ∈ ℝ ∶ 𝑎 < 𝑥 < 𝑏}, (𝑎, 𝑏] = {𝑥 ∈ ℝ ∶ 𝑎 < 𝑥 ≤ 𝑏}, [𝑎, 𝑏) = {𝑥 ∈ ℝ ∶ 𝑎 ≤ 𝑥 < 𝑏},
[𝑎, 𝑏] = {𝑥 ∈ ℝ ∶ 𝑎 ≤ 𝑥 ≤ 𝑏}, (−∞, 𝑏) = {𝑥 ∈ ℝ ∶ 𝑥 < 𝑏}, (−∞, 𝑏] = {𝑥 ∈ ℝ ∶ 𝑥 ≤ 𝑏},
(𝑎, ∞) = {𝑥 ∈ ℝ ∶ 𝑎 < 𝑥}, [𝑎, ∞) = {𝑥 ∈ ℝ ∶ 𝑎 ≤ 𝑥}.
Example 1.2.1. Let 𝑆 = (−∞, 1). Does 𝑆 have upper and lower bounds?
Solution. In class.
We have already implicitly used in our notation that maximum and minimum are unique if they exist.
The next result states this formally.
Proposition 1.6. Let 𝑆 be a nonempty subset of ℝ. If maximum or minimum of 𝑆 exist, then they are
unique.
Proof. Assume that 𝑆 has maxima 𝑀1 and 𝑀2 . We must show 𝑀1 = 𝑀2 . Since 𝑀1 ∈ 𝑆 and 𝑀2
is an upper bound of 𝑆, we have 𝑀1 ≤ 𝑀2 . Since 𝑀2 ∈ 𝑆 and 𝑀1 is an upper bound of 𝑆, we have
𝑀2 ≤ 𝑀1 . Hence 𝑀1 = 𝑀2 .
A similar proof holds for the minimum.
Example 1.2.2. Let 𝑎, 𝑏 ∈ ℝ, 𝑎 < 𝑏, and 𝑆 = [𝑎, 𝑏). Then min 𝑆 = 𝑎, but 𝑆 has no maximum.
Solution. In class.
Proof. 1. max 𝑆 is an upper bound of 𝑆 by definition. Since max 𝑆 ∈ 𝑆, max 𝑆 ≤ 𝐿 for any upper
bound 𝐿 of 𝑆. Hence max 𝑆 = sup 𝑆.
Example 1.2.3. Let 𝑎 < 𝑏 and put 𝑆 = [𝑎, 𝑏). Find inf 𝑆 and sup 𝑆 if they exist.
Solution. In class.
Note. It is crucial for Analysis that sufficiently may subsets have a supremum and/or infimum. For
have suprema. √ √
Note that the naïve approach putting 𝑆2 = (0, 2] is invalid as we do not know if 2, i. e., a real number
𝑥 > 0 with 𝑥2 = 2, exists.
Thus we have the last axiom of the real number system.
C. The Dedekind completeness axiom
Additional Note. It can be shown that the real number system is uniquely determined by the axioms
(A1)-(A4), (M1)-(M4), (D), (O1)-(O3), (C).
Theorem 1.8 (Positive square root). Let 𝑎 ≥ 0. Then there is a unique 𝑥 ≥ 0 such that 𝑥2 = 𝑎. We write
√ 1
𝑥 = 𝑎 = 𝑎2 .
Proof. If 𝑎 = 0, we have 02 = 0 and 𝑥2 > 0 if 𝑥 > 0, so that 0 = 0 is the unique number 𝑥 ≥ 0 such
√
that 𝑥2 = 0.
Now let 𝑎 > 0. Note that 𝑥 ≥ 0 and 𝑥2 = 𝑎 > 0 gives 𝑥 > 0. For the uniqueness proof let 𝑥1 > 0 and
𝑥2 > 0 such that 𝑥21 = 𝑎 and 𝑥22 = 𝑎. Then
≤ 𝑀𝑎2 − 𝑎 + 3𝑀𝑎
𝑎 − 𝑀𝑎2
4𝑀𝑎
1
= (𝑀𝑎2 − 𝑎) < 0.
4
Thus
(𝑀𝑎 + 𝜀)2 < 𝑎,
giving 𝑀𝑎 + 𝜀 ∈ 𝑆𝑎 , contradicting the fact that 𝑀𝑎 is an upper bound of 𝑆𝑎 .
Case II: 𝑀𝑎2 > 𝑎. Put
𝑀𝑎2 − 𝑎
𝜀= .
2𝑀𝑎
Then 0 < 𝜀 < 12 𝑀𝑎 and
so that any 𝑥 ≥ 𝑀𝑎 − 𝜀 does not belong to 𝑆𝑎 . Hence 𝑀𝑎 − 𝜀 is an upper bound of 𝑆𝑎 , contradicting the
fact that 𝑀𝑎 is the least upper bound of 𝑆𝑎 .
So 𝑀𝑎2 ≠ 𝑎 is impossible, and 𝑀𝑎2 = 𝑎 follows.
The completeness axiom can be thought of as ensuring that there are no ’gaps’ on the real line.
Theorem 1.9 (Characterizations of the supremum). Let 𝑆 be a nonempty subset of ℝ. Let 𝑀 ∈ ℝ. The
following are equivalent:
(a) 𝑀 = sup 𝑆;
(b) 𝑀 is an upper bound of 𝑆, and for each 𝜀 > 0, there is 𝑠 ∈ 𝑆 such that 𝑀 − 𝜀 < 𝑠;
(c) 𝑀 is an upper bound of 𝑆, and for each 𝑥 < 𝑀, there exists 𝑠 ∈ 𝑆 such that 𝑥 < 𝑠.
Proof. (a) ⇒ (b): Since (a) holds, 𝑀 is the least upper bound and thus an upper bound.
Let 𝜀 > 0. Then 𝑀 − 𝜀 is not an upper bound of 𝑆 since 𝑀 is the least upper bound of 𝑆. Hence
𝑥 ≤ 𝑀 − 𝜀 does not hold for all 𝑥 ∈ 𝑆, and therefore there must be 𝑠 ∈ 𝑆 such that 𝑀 − 𝜀 < 𝑠.
(b) ⇒ (c): Let 𝑥 < 𝑀. Then 𝜀 = 𝑀 − 𝑥 > 0, and by assumption there is 𝑠 ∈ 𝑆 with 𝑀 − 𝜀 < 𝑠. Then
𝑥 = 𝑀 − 𝜀 < 𝑠.
(c) ⇒ (a): By assumption, 𝑀 is an upper bound of 𝑆. Suppose that 𝑀 is not the least upper bound.
Then there is an upper bound 𝐿 of 𝑆 with 𝐿 < 𝑀. By assumption (c), there is 𝑠 ∈ 𝑆 such that 𝐿 < 𝑠,
contradicting that 𝐿 is an upper bound of 𝑆. So 𝑀 must be indeed the least upper bound of 𝑆.
There is an apparent asymmetry in the Dedekind completion. However, there is a version for infima,
which is obtained by reflection. To this end, if 𝑆 is a (nonempty) subset of ℝ set
−𝑆 = {−𝑥 ∶ 𝑥 ∈ 𝑆}.
−𝑆 = {𝑥 ∈ ℝ ∶ −𝑥 ∈ 𝑆}.
Theorem 1.11. Every nonempty subset of ℝ which is bounded below has an infimum.
Theorem 1.12 (Characterizations of the infimum). Let 𝑆 be a nonempty subset of ℝ. Let 𝑚 ∈ ℝ. The
following are equivalent:
(a) 𝑚 = inf 𝑆;
(b) 𝑚 is a lower bound of 𝑆, and for each 𝜀 > 0, there is 𝑠 ∈ 𝑆 such that 𝑠 < 𝑚 + 𝜀;
(c) 𝑚 is a lower bound of 𝑆, and for each 𝑥 > 𝑚, there exists 𝑠 ∈ 𝑆 such that 𝑠 < 𝑥.
Theorem 1.13 (Dedekind cut). Let 𝐴 and 𝐵 be nonempty subsets of ℝ such that
(i) 𝐴 ∩ 𝐵 = ∅,
(ii) 𝐴 ∪ 𝐵 = ℝ,
(iii) ∀ 𝑎 ∈ 𝐴 ∀ 𝑏 ∈ 𝐵, 𝑎 ≤ 𝑏.
Then there is 𝑐 ∈ ℝ such that 𝑎 ≤ 𝑐 ≤ 𝑏 for all 𝑎 ∈ 𝐴 and 𝑏 ∈ 𝐵.
Proof. 𝐴 is nonempty and bounded above (any 𝑏 ∈ 𝐵 is an upper bound of 𝐴), so 𝑐 = sup 𝐴 exists by
the Dedekind completeness axiom, and 𝑐 ≥ 𝑎 for all 𝑎 ∈ 𝐴 by definition of upper bound. (iii) says that
each 𝑏 ∈ 𝐵 is an upper bound of 𝐴, and hence 𝑐 ≤ 𝑏 since 𝑐 is the least upper bound of 𝐴.
Note. The above theorem says that Dedekind completeness implies the Dedekind cut property. Con-
versely, if the ordered field axioms and the Dedekind cut property are satisfied, then the Dedekind com-
pleteness axiom holds (see tutorials).
Tutorial 1.2.1. 1. In each of the following cases, state if the given set is bounded above or not. If a set
is bounded above, give two different upper bounds for the set, give the supremum of the set and state if
the set has a maximum or not.
(a) (−3, 2) (b) (1, ∞) (c) [10, 11]
(h) {𝑥 ∈ ℝ ∶ 𝑥2 ≤ 3}
(d) {5, 4} (e) {10, 9, 8, 7, 6, 5, 4, 3, 2, 1, 0, −1, −2, −5}
(f) (−∞, 2] (g) {𝑥 ∈ ℝ ∶ 𝑥2 < 3}
2. For each of the sets in Q. 1, state if the given set is bounded below or not. If a set is bounded below,
give two different lower bounds for the set, give the infimum of the set and state if the set has a minumum
or not.
(g) 2.5 ∈ {𝑥 ∈ ℝ ∶ 𝑥2 ≥ 4}
(d) 17 ∈ (−3, 18) (e) 17 ∈ [16, 18] (f) 2 ∈ {1, 3, 5, 7}
(h) −1 ∈ {𝑥 ∈ ℝ ∶ 2𝑥 + 7 < 5}
9∗ . Assume that the Dedekind cut property, Theorem 1.13 as well as the ordered field axioms are satisfied.
Show that the Dedekind completeness holds.
(P1) 0 ∈ ℕ,
(P2) 𝑎 ∈ ℕ ⇒ 𝑎 + 1 ∈ ℕ,
(P3) ∀ 𝑎 ∈ ℕ, 𝑎 + 1 ≠ 0,
Additional Notes. 1. Here we define natural numbers to start at 0. This is common if one uses the ax-
iomatic definition of the natural numbers. You may start the natural numbers with 1, if you are consistent
with this.
2. If one defines the natural numbers without reference to real number, there is no arithmetic on ℕ needed.
Then the number 𝑎 + 1 is called the successor of 𝑎, and axioms (P2) and (P3) can be phrased as follows:
Every natural number has a successor, and 0 is not the successor of a natural number.
3. The natural numbers are used to “count”. In everyday life, counting refers to concrete objects (‘There
are 220 students in the class’, ‘I have got no pen with me’, ‘I ate three rolls this morning’). However, they
can also be defined as cardinalities of abstract (finite) sets. Here the cardinality #𝑆 of a (finite) set 𝑆 is
the number of elements in the set. The empty set ∅ has no elements, so #∅ = 0 serves as the definition
of the number 0. To get the next number, consider the set which consists of the empty set. It has exactly
one element, namely the empty set. So #{∅} = 1. Next #{∅, {∅}} = 2, and so on.
Recall that the integers ℤ and the rational numbers ℚ are defined by
ℤ = {𝑥 ∈ ℝ ∶ 𝑥 ∈ ℕ or − 𝑥 ∈ ℕ},
{ }
ℚ = 𝑥 ∈ ℝ ∶ 𝑥 = , 𝑝, 𝑞 ∈ ℤ, 𝑞 ≠ 0 .
𝑝
𝑞
As a consequence of (P4) one has
The Principle of Mathematical Induction
Suppose we have a statement 𝐴(𝑛) associated with each 𝑛 ∈ ℕ. If
(i) the statement 𝐴(0) is true, and
(ii) whenever the statement 𝐴(𝑛) is true, also the statement 𝐴(𝑛 + 1) is true,
then the statement 𝐴(𝑛) is true for all 𝑛 ∈ ℕ.
Note. The statement 𝐴(0) is called the induction base. The induction base does not need to be taken at 0,
one can take any integer which is suitable for the statements to be proved.
𝑚 + (𝑛 + 1) = (𝑚 + 1) + 𝑛 ∈ ℕ
Finally, we state and prove two important principles which are consequences of the Dedekind complete-
ness and Peano’s axioms.
Theorem 1.15 (Well-ordering principle). Every nonempty subset of ℕ has a smallest element.
Proof. In class.
Note. One can show that if 𝑆 ⊂ ℤ, 𝑆 ≠ ∅, and 𝑆 is bounded below, then 𝑆 has a minimum.
Theorem 1.16 (The Archimedean principle). For each 𝑥 ∈ ℝ there is 𝑛 ∈ ℕ such that 𝑛 > 𝑥.
Proof. Assume the Archimedian principle is false. Then there is 𝑥 ∈ ℝ such that 𝑛 ≤ 𝑥 for all 𝑛 ∈ ℕ.
That means, ℕ is bounded above and therefore has a supremum 𝑀.
By Theorem 1.9 there is 𝑛 ∈ ℕ such that 𝑀 − 1 < 𝑛. Then
𝑛 + 1 > (𝑀 − 1) + 1 = 𝑀
and 𝑛 + 1 ∈ ℕ contradict the fact that 𝑀 is an upper bound of ℕ. Hence the Archimedian principle must
be true.
Definition 1.5. A subset 𝑆 of ℝ is said to be dense in ℝ if for all 𝑥, 𝑦 ∈ ℝ with 𝑥 < 𝑦 there is 𝑠 ∈ 𝑆
such that 𝑥 < 𝑠 < 𝑦.
Real numbers which are not rational numbers are called irrational numbers.
√
Note. 1. 2 is irrational (see tutorials).
2. ℚ is an ordered field, i. e., ℚ satisfies all axioms of ℝ except the Dedekind completeness axiom (see
tutorials).
3. If 𝑎 is rational and 𝑏 is irrational, then 𝑎 + 𝑏 is irrational (see tutorials).
Theorem 1.17. The set of rational numbers as well as the set of irrational numbers are dense in ℝ.
Proof. In class.
2} and 𝑇 = {𝑥 ∈ ℚ ∶ 𝑥 ≤
√ √
Example 1.3.1. Let 𝑆 = {𝑥 ∈ ℚ ∶ 𝑥 < 2}. Then 𝑆 = 𝑇 , and 𝑆 and 𝑇
have no supremum in ℚ.
Example 1.3.2 (Bernoulli’s inequality). For all 𝑥 ∈ ℝ with 𝑥 ≥ −1 and all 𝑛 ∈ ℕ, (1 + 𝑥)𝑛 ≥ 1 + 𝑛𝑥.
Solution. In class.
Note. We will make use of more rules of real numbers without proof. See the tutorials for the power
rules, for example.
Tutorial 1.3.1. 1. Show that if 𝑆 ⊂ ℤ, 𝑆 ≠ ∅, and 𝑆 is bounded below, then 𝑆 has a minimum.
√ √
2. Show that 2 is irrational. Hint. Assume that 2 = 𝑞𝑝 with positive integers 𝑝 and 𝑞. You may
assume that 𝑝 and 𝑞 do not have common factors, i. e., there are no positive integers 𝑘, 𝑝1 , 𝑞1 with 𝑘 ≠ 1
such that 𝑝 = 𝑝1 𝑘 and 𝑞 = 𝑞1 𝑘.
3. Show that the rational numbers satisfy the axioms (A1)–(A4), (M1)–(M4), (D), and (O1)–(O3).
4. Let 𝑎, 𝑏 ∈ ℚ with 𝑏 ≠ 0 and 𝑟 ∈ ℝ ⧵ ℚ. Show that 𝑎 + 𝑏𝑟 ∈ ℝ ⧵ ℚ.
5. For 𝑥 ∈ ℝ and 𝑛 ∈ ℕ, 𝑥𝑛 is defined inductively by
(i) 𝑥0 = 1,
(ii) 𝑥𝑛+1 = 𝑥𝑥𝑛 for 𝑛 ∈ ℕ.
Show that (a) 𝑥𝑛 𝑥𝑚 = 𝑥𝑛+𝑚 , (b) (𝑥𝑛 )𝑚 = 𝑥𝑛𝑚 , (c) 𝑥𝑛 𝑦𝑛 = (𝑥𝑦)𝑛 .
Chapter 2
Sequences
Students have seen sequences and learnt rules for limits and how to apply these rules to find limits. In
this chapter, students will learn the proper definitions of limits and the proofs of the rules for limits.
2.1 Sequences
Definition 2.1. A (real) sequence is an ordered list of infinitely many real numbers
𝑎1 , 𝑎2 , 𝑎3 , 𝑎4 , … .
We usually denote a sequence by (𝑎𝑛 ) = 𝑎1 , 𝑎2 , 𝑎3 , … .
The number 𝑎𝑛 is called the 𝑛-th term of the sequence. The subscript 𝑛 is called the index.
Note. The index of the sequence does not have to start at 1, and therefore one also writes (𝑎𝑛 )∞
𝑛=𝑛0 , where
∞ ∞ ∞
𝑛0 can be any integer, e. g., (2𝑛 − 3)𝑛=1 , (2𝑛 − 3)𝑛=0 , (2𝑛 − 3)𝑛=−5 .
Recall the following note and sketch from First Year Calculus.
Note. (i) A sequence can be identified with a function on the (positive) integers:
𝑎𝑛 = 𝑓 (𝑛).
(ii) A sequence can be plotted as points of the real axis or as the graph of the function, see (i).
( )∞
𝑛
Below is a plot for the sequence .
𝑛 + 1 𝑛=1
𝑎1 𝑎2 𝑎3
1
0 2
1
𝑎𝑛
𝑛
1 2 3 4 5 6 7 8
15
In first year calculus students have seen the following informal definition.
Informal Definition. A sequence {𝑎𝑛 } is said to converge if there is a number 𝐿 such that 𝑎𝑛 is as close
to 𝐿 as we like for all sufficiently large 𝑛.
We have to formalize “as close to” and “for all sufficiently large”: “as close to” means that the distance
does not exceed any given (small) positive number. The distance between two numbers 𝑥 and 𝑦 is 𝑦 − 𝑥
if 𝑥 < 𝑦 and 𝑥 − 𝑦 if 𝑥 > 𝑦. Hence the distance between 𝑥 and 𝑦 is |𝑦 − 𝑥|.
Therefore we have the following formal definition:
Solution. In class.
Proof. In class.
Theorem 2.2. If the sequence (𝑎𝑛 ) converges, then its limit is unique.
Proof. In class.
We are now going to prove the rules you have learnt in first year.
Theorem 2.3 (Limit Laws). Let 𝑐 ∈ ℝ and suppose that lim 𝑎𝑛 = 𝐿 and lim 𝑏𝑛 = 𝑀 both exist. Then
𝑛→∞ 𝑛→∞
(a) lim 𝑐 = 𝑐.
𝑛→∞
(b) lim (𝑎𝑛 + 𝑏𝑛 ) = lim 𝑎𝑛 + lim 𝑏𝑛 = 𝐿 + 𝑀.
𝑛→∞ 𝑛→∞ 𝑛→∞
(c) lim (𝑐𝑎𝑛 ) = 𝑐 lim 𝑎𝑛 = 𝑐𝐿.
𝑛→∞
(𝑛→∞ ) ( )
(d) lim (𝑎𝑛 𝑏𝑛 ) = lim 𝑎𝑛 lim 𝑏𝑛 = 𝐿𝑀.
𝑛→∞ 𝑛→∞ 𝑛→∞
(e) If 𝑀 ≠ 0, lim
𝑎𝑛 lim 𝑎𝑛
𝑛→∞ 𝐿
= = .
𝑛→∞ 𝑏𝑛 lim 𝑏𝑛 𝑀
𝑛→∞
Proof. In class.
Proof. In class.
1
𝑦 ∶= − 1 > 0.
𝑥
Then ( )𝑛
= (1 + 𝑦)𝑛 ≥ 1 + 𝑛𝑦
1 1
𝑛
=
𝑥 𝑥
1−𝜀
by Bernoulli’s inequality. Put 𝐾 = . Then, for 𝑛 > 𝐾,
𝑦𝜀
0 < 𝑥𝑛 ≤
1 1
< = 𝜀.
1 + 𝑛𝑦 1 + 1−𝜀
𝜀
which is impossible.
The remaining cases are left as an exercise.
≤
1 1 𝜀
0< = < 𝜀,
𝑛 𝐾 2
| |
whence | 1𝑛 | < 𝜀.
| |
Now assume the statement holds for an integer 𝑟 > 0. Then
1 1 1
lim = lim ⋅ lim 𝑟 = 0 ⋅ 0 = 0.
𝑛→∞ 𝑛𝑟+1 𝑛→∞ 𝑛 𝑛→∞ 𝑛
( )
𝑛
Tutorial 2.1.1. 1. (a) Prove, using the definition of convergence, that the sequence 𝑛+1 does not
converge to 2.
(b) Prove, using the definition of convergence, that the sequence ((−1)𝑛 ) does not converge to any 𝐿.
2. (a) Prove that if lim 𝑎𝑛 = 𝐿, then lim |𝑎𝑛 | = |𝐿|. (Hint: use (and prove) the inequality ||𝑥| − |𝑦|| ≤
𝑛→∞ 𝑛→∞
|𝑥 − 𝑦|.)
(b) Give an example to show that the converse to part (a) is not true.
3. Contractive maps. Suppose that for some 𝑐 ∈ ℝ with 0 < 𝑐 < 1, we have |𝑎𝑛+1 − 𝐿| ≤ 𝑐|𝑎𝑛 − 𝐿| for
all 𝑛 ∈ ℕ.
(a) Use induction to prove that |𝑎𝑛 − 𝐿| ≤ 𝑐 𝑛 |𝑎0 − 𝐿| for all 𝑛 ∈ ℕ.
(b) Use the Sandwich Theorem and the fact that lim 𝑐 𝑛 = 0 to prove that lim 𝑎𝑛 = 𝐿.
𝑛→∞ 𝑛→∞
√
4. Recursive algorithm for finding 𝑎. Let 𝑎 > 1 and define
( )
for 𝑛 ≥ 1.
1 𝑎
𝑎0 = 𝑎 and 𝑎𝑛 = 𝑎𝑛−1 +
2 𝑎𝑛−1
𝑎)2 for 𝑛 ≥ 1.
√ 1 √
(a) Prove that 0 < 𝑎𝑛 − 𝑎= (𝑎
2𝑎𝑛−1 𝑛−1
−
(b) Use (a) to prove that 0 ≤ 𝑎𝑛 − 𝑎 ≤ 21 (𝑎𝑛−1 − 𝑎) for 𝑛 ≥ 1.
√ √
√
(c) Deduce that lim 𝑎𝑛 = 𝑎.
𝑛→∞ √
(d) Apply four steps of the recursive algorithm with 𝑎 = 3 to approximate 3.
Proof. Here we use the fact, which is easy to prove (by induction), that a set of the form {𝑎𝑛 ∶ 𝑛 ∈
ℤ, 𝑛0 ≤ 𝑛 ≤ 𝑛1 } is bounded (and indeed has minimum and maximum). We also use that the union of
Since (𝑎𝑛 ) converges, there are 𝐿 and 𝑘 such that |𝑎𝑛 − 𝐿| < 1 for all 𝑛 ≥ 𝑘. Hence {𝑎𝑛 ∶ 𝑛 < 𝑘} and
two bounded sets is bounded.
Although each unbounded sequence diverges, there are some divergent sequences which have a limiting
behaviour which we want to exploit. Recall that we use the notation ‘𝑛 → ∞’ in the definition of the
limit. This extends to ‘𝑎𝑛 → ∞’, and we have therefore the
Definition 2.5. 1. We say that 𝑎𝑛 tends to infinity as 𝑛 tends to infinity and write 𝑎𝑛 → ∞ as 𝑛 → ∞ if
for every 𝐴 ∈ ℝ there is 𝐾 ∈ ℝ such that 𝑎𝑛 > 𝐴 for all 𝑛 ≥ 𝐾.
2. We say that 𝑎𝑛 tends to minus infinity as 𝑛 tends to infinity and write 𝑎𝑛 → −∞ as 𝑛 → ∞ if for every
𝐴 ∈ ℝ there is 𝐾 ∈ ℝ such that 𝑎𝑛 < 𝐴 for all 𝑛 ≥ 𝐾.
Notes. 1. For the definition of 𝑎𝑛 → ∞(−∞) as 𝑛 → ∞, we may restrict 𝐴 to 𝐴 > 𝐴0 (𝐴 < 𝐴0 ) for
suitable numbers 𝐴0 .
2. For 𝑎𝑛 → ∞ as 𝑛 → ∞ we also write lim 𝑎𝑛 = ∞ and say that (𝑎𝑛 ) diverges to ∞.
𝑛→∞
3. For 𝑎𝑛 → −∞ as 𝑛 → ∞ we also write lim 𝑎𝑛 = −∞ and say that (𝑎𝑛 ) diverges to −∞.
𝑛→∞
Example 2.2.1. 1. Let 𝑎𝑛 = (−1)𝑛 . Then (𝑎𝑛 ) is bounded but not convergent.
2. Let 𝑎𝑛 = 𝑛. Then (𝑎𝑛 ) is unbounded (by the Archimedean principle). Hence (𝑎𝑛 ) diverges.
Solution. In Class.
Theorem 2.8. 1. Let (𝑎𝑛 ) be a sequence with 𝑎𝑛 > 0 for all indices 𝑛. Then
1
lim 𝑎 = ∞ ⇔ lim = 0.
𝑛→∞ 𝑛 𝑛→∞ 𝑎𝑛
1
lim 𝑎 = −∞ ⇔ lim = 0.
𝑛→∞ 𝑛 𝑛→∞ 𝑎𝑛
1 1
Proof. 1. Assume that lim 𝑎𝑛 = ∞. To show that lim = 0 let 𝜀 > 0. Put 𝐴 = . Then there is 𝐾
such that 0 < 𝐴 < 𝑎𝑛 for all 𝑛 ≥ 𝐾. Then
𝑛→∞ 𝑛→∞ 𝑎𝑛 𝜀
1 1
0< < =𝜀
𝑎𝑛 𝐴
1
for these 𝑛, which shows that lim = 0.
𝑛→∞ 𝑎𝑛
1
Conversely, assume that lim = 0. Let 𝐴 ∈ ℝ and put 𝐴0 = max{𝐴, 1}. Then 𝐴0 > 0 and put
𝑛→∞ 𝑎𝑛
Apart from this rule, there are more rules for infinite limits. Some of these rules are listed below.
Definition 2.6. Let 𝐴 ⊂ ℝ be nonempty. If 𝐴 is not bounded above, we write sup 𝐴 = ∞, and if 𝐴 is
not bounded below, we write inf 𝐴 = −∞.
We know that not every sequence converges, and it may be hard to decide if a sequence converges.
However, the situation is different with monotonic sequences:
Theorem 2.10. 1. Let (𝑎𝑛 ) be an increasing sequence. If (𝑎𝑛 ) is bounded, then (𝑎𝑛 ) converges, and
lim 𝑎 = sup{𝑎𝑛 ∶ 𝑛 ∈ ℕ}. If (𝑎𝑛 ) is not bounded, then (𝑎𝑛 ) diverges to ∞.
𝑛→∞ 𝑛
2. Let (𝑎𝑛 ) be a decreasing sequence. If (𝑎𝑛 ) is bounded, then (𝑎𝑛 ) converges, and
lim 𝑎𝑛 = inf {𝑎𝑛 ∶ 𝑛 ∈ ℕ}. If (𝑎𝑛 ) is not bounded, then (𝑎𝑛 ) diverges to −∞.
𝑛→∞
Proof. In class.
( )𝑛 ( )𝑛+1
Example 2.2.3. Let 𝑎𝑛 = 1 + 1𝑛 and 𝑏𝑛 = 1 + 1𝑛 . Then 𝑎𝑛 < 𝑏𝑛 for all 𝑛 ∈ ℕ, (𝑎𝑛 ) is an
increasing sequence, (𝑏𝑛 ) is a decreasing sequence, both sequences converge to the same number, and
the limit is denoted by 𝑒, Euler’s number.
Definition 2.7. 1. With every sequence (𝑎𝑛 ) which is bounded above, we can associate the sequence of
sup{𝑎𝑘 ∶ 𝑘 ≥ 𝑛}.
numbers
Since {𝑎𝑘 ∶ 𝑘 ≥ 𝑛} = {𝑎𝑛 } ∪ {𝑎𝑘 ∶ 𝑘 ≥ 𝑛 + 1} ⊃ {𝑎𝑘 ∶ 𝑘 ≥ 𝑛 + 1}, this sequence is decreasing, and we
𝑛→∞ 𝑛→∞
inf {𝑎𝑘 ∶ 𝑘 ≥ 𝑛}
Note. 1. Since inf 𝑆 ≤ sup 𝑆 for every nonempty set 𝑆, it follows that lim inf 𝑎𝑛 ≤ lim sup 𝑎𝑛 , where
𝑛→∞ 𝑛→∞
we write −∞ < 𝑥 and 𝑥 < ∞ for each 𝑥 ∈ ℝ.
2. The sequence (𝑎𝑛 ) is bounded if and only if both lim inf 𝑎𝑛 and lim sup 𝑎𝑛 are real numbers (also called
𝑛→∞ 𝑛→∞
finite).
We have now the following characterization of the convergence of a sequence and its limit when it exists.
Theorem 2.11. 1. The sequence (𝑎𝑛 ) converges if and only if lim inf 𝑎𝑛 and lim sup 𝑎𝑛 are finite and
𝑛→∞ 𝑛→∞
equal, and then
lim inf 𝑎𝑛 = lim 𝑎𝑛 = lim sup 𝑎𝑛 .
𝑛→∞ 𝑛→∞ 𝑛→∞
Proof. In Class.
2., 3. Exercise.
Definition 2.8 (Cauchy sequence). A seqence (𝑎𝑛 ) is called a Cauchy sequence if for all 𝜀 > 0 there is
𝐾 ∈ ℝ such that for all 𝑛, 𝑚 ∈ ℕ with 𝑛, 𝑚 ≥ 𝐾, |𝑎𝑛 − 𝑎𝑚 | < 𝜀.
which gives
( ) ( )
0 ≤ lim sup 𝑎𝑛 − lim inf 𝑎𝑛 ≤ 𝑎𝐾 +
𝜀 𝜀 2𝜀
− 𝑎𝐾 − = < 𝜀.
𝑛→∞ 𝑛→∞ 3 3 3
By Lemma 2.1, lim inf 𝑎𝑛 = lim sup 𝑎𝑛 , and an application of Theorem 2.11, part 1, completes the
𝑛→∞ 𝑛→∞
proof.
𝑛3 − 3𝑛2
(a) lim (𝑛2 + 2𝑛 − 10) (b) lim (𝑛 − 1𝑛 ) (c) lim
𝑛→∞ 𝑛→∞ 𝑛→∞ 𝑛 + 1
3. Prove that if lim |𝑎𝑛 | = ∞, then (𝑎𝑛 ) diverges.
𝑛→∞
4. Prove that if 𝑝 ∈ ℕ, 𝑝 > 0, then 𝑛𝑝 → ∞ as 𝑛 → ∞.
5. Define a sequence as follows:
𝑛→∞ 𝑛→∞ ( )
𝑥 𝑛
8. (a) Show that exp(𝑥) = lim 1 + exists for all 𝑥 ∈ ℝ and that exp(1) = 𝑒.
𝑛→∞ 𝑛
Hint. Adapt the proof of Example 2.2.3.
(b) Use Bernoulli’s inequality to prove that
( )𝑛
1 + 𝑥+𝑦
𝑛
lim =1
𝑛→∞ 1 + 𝑥+𝑦 + 𝑥𝑦
𝑛 𝑛2
for all 𝑥, 𝑦 ∈ ℝ.
(c) Use (b) to show that exp(𝑥 + 𝑦) = exp(𝑥) exp(𝑦) for all 𝑥, 𝑦 ∈ ℝ.
(d) Show that exp(𝑥) ≥ 1 + 𝑥 for all 𝑥 > 0.
(e) Show that exp(𝑥) > 0 for all 𝑥 ∈ ℝ.
(f) Show that exp is strictly increasing.
(g) Show that exp(𝑛) = 𝑒𝑛 for all 𝑛 ∈ ℤ.
Chapter 3
The single most important concept in all of analysis is that of a limit. Every single notion of analysis is
encapsulated in one sense or another to that of a limit. In the previous chapter, you learnt about limits
of sequences. Here this notion is extended to limits of functions, which leads to the notion of continuity.
You have learnt an intuitive version in Calculus I. Here you will learn a precise definition and you will
learn how to prove the results you have learnt in Calculus I.
In this course all functions will have domains and ranges which are subsets of ℝ unless otherwise stated.
Such functions are called real functions.
In the definition of the limit of sequences you have seen formal definitions for ‘𝑛 tends to ∞’ and ‘𝑎𝑛
tends to 𝐿’, and the latter one has an obvious generalization to ‘𝑥 tends to 𝑎’ and ‘𝑓 (𝑥) tends to 𝐿’, which
you will encounter in the next definition.
Definition 3.2 (Limit of a function). Let 𝑓 be a real function, 𝑎, 𝐿 ∈ ℝ and assume that the domain of 𝑓
contains a deleted neighbourhood of 𝑎, that is, 𝑓 (𝑥) is defined for all 𝑥 in a deleted neighbourhood of 𝑓 .
Then ‘𝑓 (𝑥) → 𝐿 as 𝑥 → 𝑎’ is defined to mean:
∀ 𝜀 > 0 ∃ 𝛿 > 0 ∀ 𝑥 ∈ (𝑎 − 𝛿, 𝑎 + 𝛿) ⧵ {𝑎}, 𝑓 (𝑥) ∈ (𝐿 − 𝜀, 𝐿 + 𝜀),
i. e., ∀ 𝜀 > 0 ∃ 𝛿 > 0 (0 < |𝑥 − 𝑎| < 𝛿 ⇒ |𝑓 (𝑥) − 𝐿| < 𝜀).
If 𝑓 (𝑥) → 𝐿 as 𝑥 → 𝑎, then we write lim 𝑓 (𝑥) = 𝐿.
𝑥→𝑎
Note. 1. 𝐿 in the definition above is also called the limit of the function at 𝑎. For the definition of the
limit of a function at 𝑎, the number 𝑓 (𝑎) is never used, and indeed, the function need not be defined at 𝑎.
2. Observe that the number 𝛿 will depend on 𝜀, and also on 𝑎, if we vary 𝑎.
23
Exercise. Use the notation in the above definition to label the following graph:
Solution. Let 𝜀 > 0. We must show that there is 𝛿 > 0 such that 0 < |𝑥 − 2| < 𝛿 implies |𝑥2 − 4| < 𝜀.
Since we want to use |𝑥 − 2| < 𝛿 to get |𝑥2 − 4| < 𝜀, we try to factor out |𝑥 − 2| from |𝑥2 − 4|. Thus
It is often convenient to make an initial restriction on 𝛿, like 𝛿 ≤ 1. Then we can continue the above
estimate to obtain from |𝑥 − 𝑎| < 𝛿 that
If we now put { }
𝜀
𝛿 = min 1, ,
5
then we can conclude (note that we already used 𝛿 ≤ 1) for 0 < |𝑥 − 2| < 𝛿 that
Proof. In class.
Definition 3.3 (One sided limits). 1. Let 𝑓 be a real function, 𝑎, 𝐿 ∈ ℝ and assume that the domain of
𝑓 contains an interval (𝑎, 𝑑) with 𝑑 > 𝑎, that is, 𝑓 (𝑥) is defined for all 𝑥 in (𝑎, 𝑑).
Then ‘𝑓 (𝑥) → 𝐿 as 𝑥 → 𝑎+ ’ is defined to mean:
∀ 𝜀 > 0 ∃ 𝛿 > 0 (𝑎 < 𝑥 < 𝑎 + 𝛿 ⇒ |𝑓 (𝑥) − 𝐿| < 𝜀).
If 𝑓 (𝑥) → 𝐿 as 𝑥 → 𝑎+ , then we write lim+ 𝑓 (𝑥) = 𝐿.
𝑥→𝑎
2. Let 𝑓 be a real function, 𝑎, 𝐿 ∈ ℝ and assume that the domain of 𝑓 contains an interval (𝑐, 𝑎) with
𝑐 < 𝑎, that is, 𝑓 (𝑥) is defined for all 𝑥 in (𝑐, 𝑎).
Then ‘𝑓 (𝑥) → 𝐿 as 𝑥 → 𝑎− ’ is defined to mean:
∀ 𝜀 > 0 ∃ 𝛿 > 0 (𝑎 − 𝛿 < 𝑥 < 𝑎 ⇒ |𝑓 (𝑥) − 𝐿| < 𝜀).
If 𝑓 (𝑥) → 𝐿 as 𝑥 → 𝑎− , then we write lim− 𝑓 (𝑥) = 𝐿.
𝑥→𝑎
𝑥
Example 3.1.2. Let 𝑓 (𝑥) = for 𝑥 ∈ ℝ ⧵ {0}. Then 𝑓 (𝑥) = 1 if 𝑥 > 0 and 𝑓 (𝑥) = −1 if 𝑥 < 0. It is
|𝑥|
easy to prove from the definition that lim+ 𝑓 (𝑥) = 1 and lim− 𝑓 (𝑥) = −1. Now it follows from Theorem
𝑥→0 𝑥→0
3.2 that the function 𝑓 does not have a limit as 𝑥 tends to 𝑎.
1
3. For 𝑥 ∈ ℝ ⧵ {0} let 𝑓 (𝑥) = sin . Prove that 𝑓 does not tend to any limit as 𝑥 → 0.
𝑥
4. Let 𝑓 (𝑥) = 𝑥 − ⌊𝑥⌋. For each integer 𝑛, find lim− 𝑓 (𝑥) and lim+ 𝑓 (𝑥) if they exist.
𝑥→𝑛 𝑥→𝑛
Definition 3.4. 1. Let 𝑓 be a real function defined on a set containing an interval of the form (𝑐, ∞).
Then ‘𝑓 (𝑥) → 𝐿 as 𝑥 → ∞’ is defined to mean:
∀ 𝜀 > 0 ∃ 𝐾(> 0) (𝑥 > 𝐾 ⇒ |𝑓 (𝑥) − 𝐿| < 𝜀).
If 𝑓 (𝑥) → 𝐿 as 𝑥 → ∞, then we write lim 𝑓 (𝑥) = 𝐿.
𝑥→∞
2. Let 𝑓 be a real function defined on a set containing an interval of the form (−∞, 𝑐).
Then ‘𝑓 (𝑥) → 𝐿 as 𝑥 → −∞’ is defined to mean:
∀ 𝜀 > 0 ∃ 𝐾(< 0) (𝑥 < 𝐾 ⇒ |𝑓 (𝑥) − 𝐿| < 𝜀).
If 𝑓 (𝑥) → 𝐿 as 𝑥 → −∞, then we write lim 𝑓 (𝑥) = 𝐿.
𝑥→−∞
1
Example 3.2.1. Let 𝑓 (𝑥) = . Find lim 𝑓 (𝑥) and lim 𝑓 (𝑥).
𝑥 𝑥→∞ 𝑥→−∞
Solution. In class.
Definition 3.5. Let 𝑓 be a real function whose domain includes a deleted neighbourhood of the number 𝑎.
1. ‘𝑓 (𝑥) → ∞ as 𝑥 → 𝑎’ is defined to mean:
∀ 𝐾(> 0) ∃ 𝛿 > 0 (0 < |𝑥 − 𝑎| < 𝛿 ⇒ 𝑓 (𝑥) > 𝐾).
If 𝑓 (𝑥) → ∞ as 𝑥 → 𝑎, then we write lim 𝑓 (𝑥) = ∞.
𝑥→𝑎
2. ‘𝑓 (𝑥) → −∞ as 𝑥 → 𝑎’ is defined to mean:
∀ 𝐾(< 0) ∃ 𝛿 > 0 (0 < |𝑥 − 𝑎| < 𝛿 ⇒ 𝑓 (𝑥) < 𝐾).
If 𝑓 (𝑥) → −∞ as 𝑥 → 𝑎, then we write lim 𝑓 (𝑥) = −∞.
𝑥→𝑎
2
Example 3.2.2. Prove that → ∞ as 𝑥 → 0.
𝑥2
Solution. The following sketch illustrates the result.
40
35
30
25
20
15
10
5
−2 −1 0 1 2
1
We may choose 𝐾 to be sufficiently large. So choose 𝐾 > 1 and let 𝛿 = 𝐾
. Then, for 0 < |𝑥| < 𝛿, and
because of 𝛿 < 1 we have
≥
2 1 2 2 2
= > = 2𝐾 > 𝐾.
𝑥2 |𝑥| |𝑥| |𝑥| 𝛿
Definition 3.6. Let 𝑓 be a real function defined on a set containing an interval of the form (𝑐, ∞).
Then ‘𝑓 (𝑥) → ∞ as 𝑥 → ∞’ is defined to mean:
∀ 𝐴(> 0) ∃ 𝐾(> 0) (𝑥 > 𝐾 ⇒ 𝑓 (𝑥) > 𝐴).
If 𝑓 (𝑥) → ∞ as 𝑥 → ∞, then we write lim 𝑓 (𝑥) = ∞.
𝑥→∞
Similar definition hold with all other combinations of limits involving ∞ and −∞, including one-sided
limits.
𝑥 sin 𝑥
Example 3.2.3. Prove that 𝑓 (𝑥) = + 10 → ∞ as 𝑥 → ∞. A section of the function’s graph is
10 𝑥
plotted below.
20 40 60 80
Additional notes. 1. You have learnt different notations for limits depending on whether real numbers,
∞ or −∞ are considered. However, using the notion of (deleted) neighbourhood, one can give just one
definition which applies to each case. To define (deleted) neighbourhoods of ∞ and −∞, we first observe
that ∞ and −∞ are just symbols for our purposes, and thus never belong to sets of numbers which we
consider. Therefore, no distinction is made here between neighbourhoods and deleted neighbourhoods
of ∞ and −∞, respectively.
A (deleted) neighbourhood of ∞ is any set of the form (𝑐, ∞) with 𝑐 ∈ ℝ.
A (deleted) neighbourhood of −∞ is any set of the form (−∞, 𝑑) with 𝑑 ∈ ℝ.
Now let 𝑎, 𝐿 ∈ ℝ ∪ {∞, −∞} and 𝑓 be a function defined in a deleted neighbourhood of 𝑎. Then
𝑓 (𝑥) → 𝐿 as 𝑥 → 𝑎 if and only if for each neighbourhood 𝑈 of 𝐿 there is a deleted neighbourhood 𝑉 of
𝑎 such that 𝑥 ∈ 𝑉 implies 𝑓 (𝑥) ∈ 𝑈 .
If we define one-sided deleted neighbourhoods of 𝑎 ∈ ℝ to be the sets (𝑐, 𝑎) and (𝑎, 𝑑), respectively, then
the above formulation also covers one-sided limits.
2. What we call neighbourhood here should actually be called basic neighbourhood. In general topology,
a neighborhood is any subset (of a given set) which contains a basic neighbourhood. But since we do not
need general neighbourhoods, we conveniently dropped the word ’basic’.
1
Example 3.2.4. Let 𝑓 (𝑥) = . Find lim− 𝑓 (𝑥), lim+ 𝑓 (𝑥), lim 𝑓 (𝑥).
𝑥 𝑥→0 𝑥→0 𝑥→0
Solution. In class.
Theorem 3.3 (Limit Laws). Let 𝑎, 𝑐 ∈ ℝ and suppose that the real functions 𝑓 and 𝑔 are defined in a
deleted neighbourhood of 𝑎 and that lim 𝑓 (𝑥) = 𝐿 ∈ ℝ and lim 𝑔(𝑥) = 𝑀 ∈ ℝ both exist. Then
𝑥→𝑎 𝑥→𝑎
(a) lim 𝑐 = 𝑐.
𝑥→𝑎
(f) If 𝑀 ≠ 0, lim
lim 𝑓 (𝑥)
𝑓 (𝑥) 𝑥→𝑎 𝐿
= = .
𝑥→𝑎 𝑔(𝑥) lim 𝑔(𝑥) 𝑀
𝑥→𝑎
(i) lim 𝑥 = 𝑎.
𝑥→𝑎
(j) If 𝑛 ∈ ℕ, lim 𝑥𝑛 = 𝑎𝑛 .
𝑥→𝑎
√
𝑛
√
𝑛
(k) If 𝑛 ∈ ℕ, lim 𝑥= 𝑎. If 𝑛 is even, we assume that 𝑎 > 0.
𝑥→𝑎
√ √ √
𝑛
𝑛
(l) If 𝑛 ∈ ℕ, lim 𝑓 (𝑥) = 𝑛 lim 𝑓 (𝑥) = 𝐿.
𝑥→𝑎 𝑥→𝑎
Proof. The proofs are similar to those in Theorem 2.3 and we will only prove (b), (e), (f) in class.
with 𝑏𝑖 ∈ ℝ for 𝑖 = 1, 2, … , 𝑛 and 𝑛 any non-negative integer. A rational function is of the form
𝑝(𝑥)
𝑓 (𝑥) = with 𝑝(𝑥) and 𝑞(𝑥) polynomials. We then have the following as a consequence of Theorem
𝑞(𝑥)
2.3, (b),(d),(i),(j).
Corollary 3.4. If 𝑓 is a polynomial or a rational function and 𝑎 is in the domain of 𝑓 , then lim 𝑓 (𝑥) =
𝑥→𝑎
𝑓 (𝑎).
Corollary 3.5. All the limit rules in Theorem 3.3 remain true if 𝑥 → 𝑎 is replaced by any of the following:
𝑥 → 𝑎+ , 𝑥 → 𝑎− , 𝑥 → ∞, 𝑥 → −∞.
Proof. For 𝑥 → 𝑎+ and 𝑥 → 𝑎− one just has to replace 0 < |𝑥 − 𝑎| < 𝛿 with 0 < 𝑥 − 𝑎 < 𝛿 and
−𝛿 < 𝑥 − 𝑎 < 0, respectively, in the proof of each of the statements. For 𝑥 → ∞ and 𝑥 → −∞, the
proofs are very similar to those for sequences.
Similar rules hold if the functions have infinite limits. We state some of the results for 𝑥 → 𝑎, observing
that there are obvious extensions as in Corollary 3.5.
Theorem 3.6. Assume that lim 𝑓 (𝑥) = ∞, lim 𝑔(𝑥) = ∞ and lim ℎ(𝑥) = 𝑐 ∈ ℝ. Then
𝑥→𝑎 𝑥→𝑎 𝑥→𝑎
(a) 𝑓 (𝑥) + 𝑔(𝑥) → ∞ as 𝑥 → 𝑎,
(b) 𝑓 (𝑥) + ℎ(𝑥) → ∞ as 𝑥 → 𝑎,
(c) 𝑓 (𝑥)𝑔(𝑥) → ∞{as 𝑥 → 𝑎,
∞ if c>0
(d) 𝑓 (𝑥)ℎ(𝑥) → as 𝑥 → 𝑎.
−∞ if c<0
𝑓 (𝑥)𝑔(𝑥) > 1 ⋅ 𝐴 = 𝐴.
Theorem 3.7 (Sandwich Theorem). Let 𝑎 ∈ ℝ∪{∞, −∞} and assume that 𝑓 , 𝑔 and ℎ are real functions
defined in a deleted neighbourhood of 𝑎. If 𝑓 (𝑥) ≤ 𝑔(𝑥) ≤ ℎ(𝑥) for 𝑥 in a deleted neighbourhood of 𝑎
and
lim 𝑓 (𝑥) = 𝐿 = lim ℎ(𝑥), then lim 𝑔(𝑥) = 𝐿.
𝑥→𝑎 𝑥→𝑎 𝑥→𝑎
Proof. Note that 𝐿 ∈ ℝ ∪ {∞, −∞}. We will prove this theorem in the case 𝑎 ∈ ℝ and 𝐿 ∈ ℝ. The
other cases are left as an exercise.
Let 𝜀 > 0. Then there are 𝛿1 and 𝛿2 such that
(i) |𝑓 (𝑥) − 𝐿| < 𝜀 if 0 < |𝑥 − 𝑎| < 𝛿1 ,
(ii) |ℎ(𝑥) − 𝐿| < 𝜀 if 0 < |𝑥 − 𝑎| < 𝛿2 .
Put 𝛿 = min{𝛿1 , 𝛿2 }. Then, for 0 < |𝑥 − 𝑎| < 𝛿,
Theorem 3.8. Let 𝑓 be defined on an interval (𝑎, 𝑏), where 𝑎 = −∞ and 𝑏 = ∞ are allowed. With the
convenient notation 𝑎+ = −∞ if 𝑎 = −∞ and 𝑏− = ∞ if 𝑏 = ∞, we obtain
(a) if 𝑓 is increasing, then
lim− 𝑓 (𝑥) = sup{𝑓 (𝑥) ∶ 𝑥 ∈ (𝑎, 𝑏)} and lim+ 𝑓 (𝑥) = inf {𝑓 (𝑥) ∶ 𝑥 ∈ (𝑎, 𝑏)};
𝑥→𝑏 𝑥→𝑎
(b) if 𝑓 is decreasing, then
lim− 𝑓 (𝑥) = inf {𝑓 (𝑥) ∶ 𝑥 ∈ (𝑎, 𝑏)} and lim+ 𝑓 (𝑥) = sup{𝑓 (𝑥) ∶ 𝑥 ∈ (𝑎, 𝑏)}.
𝑥→𝑏 𝑥→𝑎
Proof. In class.
Note that the sketch on page 24 also illustrates the definition of continuity.
We realize that the definition of the existence of a limit of a function at a point and continuity at that
point are very similar, but that there are subtle (and important) differences:
For limits, 𝑓 does not need to be defined at 𝑎, and even if 𝑓 (𝑎) exists, this value is not used at all when
finding the limit of the function 𝑓 at 𝑎.
We conclude
𝑓 is continuous at 𝑎 ⇔ ∀ 𝜀 > 0 ∃ 𝛿 > 0 (|𝑥 − 𝑎| < 𝛿 ⇒ |𝑓 (𝑥) − 𝑓 (𝑎)| < 𝜀) (by definition)
⇔ ∀ 𝜀 > 0 ∃ 𝛿 > 0 (0 < |𝑥 − 𝑎| < 𝛿 ⇒ |𝑓 (𝑥) − 𝑓 (𝑎)| < 𝜀) (∵ 𝑥 = 𝑎 ⇒ 𝑓 (𝑥) = 𝑓 (𝑎))
⇔ lim 𝑓 (𝑥) = 𝑓 (𝑎).
𝑥→𝑎
Theorem 3.9. 𝑓 is continuous at 𝑎 if and only if the following three conditions are satisfied:
1. 𝑓 (𝑎) is defined, i. e., 𝑎 is in the domain of 𝑓 ,
2. lim 𝑓 (𝑥) exists, i.e., lim− 𝑓 (𝑥) = lim+ 𝑓 (𝑥), and
𝑥→𝑎 𝑥→𝑎 𝑥→𝑎
3. 𝑓 (𝑎) = lim 𝑓 (𝑥).
𝑥→𝑎
for 𝑥 ≠ 0 while 𝑓 is not defined at 𝑥 = 0. Then lim 𝑓 (𝑥) = 1 exists, but 𝑓 is not
sin 𝑥
2. Let 𝑓 (𝑥) =
𝑥 𝑥→0
defined at 0. Hence 𝑓 is not continuous at 0.
3. Let {
if 𝑥 ≠ 0,
sin 𝑥
𝑓 (𝑥) = 𝑥
1 if 𝑥 = 0.
Then lim 𝑓 (𝑥) = 1 exists, and 𝑓 (0) = 1. Hence 𝑓 is continuous at 0.
𝑥→0
if 𝑔(𝑎) ≠ 0 and
𝑓
4. the quotient
𝑔
5. the scalar multiple 𝑐𝑓
are functions that are also continuous at 𝑎.
Proof. The statements follow immediate from the limit laws, Theorem 3.3, and Theorem 3.9.
For example, for 3. we have
lim (𝑓 𝑔)(𝑥) = lim 𝑓 (𝑥)𝑔(𝑥) = lim 𝑓 (𝑥) lim 𝑔(𝑥) = 𝑓 (𝑎)𝑔(𝑎) = (𝑓 𝑔)(𝑎).
𝑥→𝑎 𝑥→𝑎 𝑥→𝑎 𝑥→𝑎
Recall that the composite 𝑔◦𝑓 of two functions 𝑓 and 𝑔 is defined by (𝑔◦𝑓 )(𝑥) = 𝑔(𝑓 (𝑥)).
Proof. In class.
Definition 3.8. 1. A function 𝑓 is continuous from the right at 𝑎 if lim+ 𝑓 (𝑥) = 𝑓 (𝑎).
𝑥→𝑎
2. A function 𝑓 is continuous from the left at 𝑎 if lim− 𝑓 (𝑥) = 𝑓 (𝑎).
𝑥→𝑎
if 𝑥 ≠ 0,
⎧ |𝑥| + 𝑥
⎪
𝑓 (𝑥) = ⎨ 2𝑥
⎪0 if 𝑥 = 0.
⎩
Determine the right and left continuity of 𝑓 at 𝑥 = 0.
Solution. In class.
Note. 1. It is easy to show that 𝑓 is continuous at 𝑎 if and only if 𝑓 is continuous from the right and
continuous from the left at 𝑎.
2. If 𝑎 ∈ dom(𝑓 ) and if there is 𝜀 > 0 such that dom(𝑓 ) ∩ (𝑎 − 𝜀, 𝑎 + 𝜀) = (𝑎 − 𝜀, 𝑎], then we say that 𝑓
is continuous at 𝑎 if lim− 𝑓 (𝑥) = 𝑓 (𝑎).
𝑥→𝑎
3. If 𝑎 ∈ dom(𝑓 ) and if there is 𝜀 > 0 such that dom(𝑓 ) ∩ (𝑎 − 𝜀, 𝑎 + 𝜀) = [𝑎, 𝑎 + 𝜀), then we say that 𝑓
is continuous at 𝑎 if lim+ 𝑓 (𝑥) = 𝑓 (𝑎).
𝑥→𝑎
4. The convention in 2 and 3 is consistent with what you will learn in General Topology about continuity.
Just note that the condition |𝑓 (𝑥) − 𝑓 (𝑎)| < 𝜀 has to be checked for all 𝑥 ∈ dom(𝑓 ) which satisfy
|𝑥 − 𝑎| < 𝛿.
Lemma 3.12. If 𝑓 (𝑥) → 𝑏 as 𝑥 → 𝑎 (𝑎+ , 𝑎− ) and 𝑔 is continuous at 𝑏, then 𝑔(𝑓 (𝑥)) → 𝑔(𝑏) as 𝑥 →
𝑎 (𝑎+ , 𝑎− ), which can be written, e. g., as
( )
lim 𝑔(𝑓 (𝑥)) = 𝑔 lim 𝑓 (𝑥) .
𝑥→𝑎 𝑥→𝑎
Solution. The domain of 𝑓 is {𝑥 ∈ ℝ ∶ |𝑥| ≥ 2} = (−∞, −2] ∪ [2, ∞). By Theorem 3.10, the
function 𝑥 → 𝑥2 − 4 is continuous on ℝ, and by Theorem 3.3 (k), the square root is continuous at each
positive number. So also the composite function 𝑓 is continuous on (−∞, −2) ∪ (2, ∞). Also, the proof
of Theorem 3.3 (k) can be easily adapted to show that the square root is continuous from the right at 0.
Then it easily follows that 𝑓 is continuous (from the right) at 2 and continuous (from the left) at −2.
, 𝑝 and 𝑞 ≠ 0 polynomials.
𝑝(𝑥)
2. Rational functions
𝑞(𝑥)
3. Sums, differences, products and quotients of continuous functions.
4. Root functions.
5. The trigonometric functions sin 𝑥, cos 𝑥, tan 𝑥, cosec 𝑥, sec 𝑥 and cot 𝑥.
6. The exponential function exp(𝑥).
7. The absolute value function |𝑥|.
Proof. 1, 2 and 3 easily follow from previous theorems on limits and continuity, as does 7. However, 7
can be easily proved dircetly: For each 𝜀 > 0 let 𝛿 = 𝜀. Then, for |𝑥 − 𝑎| < 𝛿 we have
The continuity of sin and cos follows from the sum of angles formulae and from the limits proved in
Calculus I (the proofs used the Sandwich Theorem, which now has been proved). The continuity of the
other trigonometric functions then follows from part 3.
Finally, the continuity of exp is a tutorial problem.
Theorem 3.14. Let 𝑎 ∈ ℝ and let 𝑓 be a real function which is defined in a neighbourhood of 𝑎. Then
𝑓 is continuous at 𝑎 if and only if for each sequence (𝑥𝑛 ) in dom(𝑓 ) with lim 𝑥𝑛 = 𝑎 the sequence 𝑓 (𝑥𝑛 )
𝑛→∞
satisfies lim 𝑓 (𝑥𝑛 ) = 𝑓 (𝑎).
𝑛→∞
Proof. In class.
𝑓 (𝑏)
𝑦=𝑘
𝑓 (𝑎)
𝑎 𝑐 𝑏
Theorem 3.15 (Intermediate Value Theorem (IVT)). Suppose that 𝑓 is continuous on the closed interval
[𝑎, 𝑏] with 𝑓 (𝑎) ≠ 𝑓 (𝑏). Then for any number 𝑘 between 𝑓 (𝑎) and 𝑓 (𝑏) there exists a number 𝑐 in the
open interval (𝑎, 𝑏) such that 𝑓 (𝑐) = 𝑘.
Proof. Let
𝑔(𝑥) = 𝑓 (𝑥) − 𝑘, (𝑥 ∈ [𝑎, 𝑏]).
Then 𝑔 is continuous, and 0 lies between 𝑔(𝑎) and 𝑔(𝑏), that is, 𝑔(𝑎) and 𝑔(𝑏) have opposite signs:
𝑔(𝑎)𝑔(𝑏) < 0.
Let [𝑎0 , 𝑏0 ] = [𝑎, 𝑏] and use bisection to define intervals [𝑎𝑛 , 𝑏𝑛 ] as follows: If [𝑎𝑛 , 𝑏𝑛 ] with 𝑔(𝑎𝑛 )𝑔(𝑏𝑛 ) < 0
has been found, let 𝑑 be the midpoint of the interval [𝑎𝑛 , 𝑏𝑛 ]. If 𝑔(𝑑) = 0, the result follows with 𝑐 = 𝑑. If
𝑔(𝑑) has the same sign as 𝑔(𝑏𝑛 ), then 𝑔(𝑎𝑛 ) and 𝑔(𝑑) have opposite signs, and putting 𝑎𝑛+1 = 𝑎𝑛 , 𝑏𝑛+1 = 𝑑,
we have 𝑔(𝑎𝑛+1 )𝑔(𝑏𝑛+1 ) < 0. Otherwise, if 𝑔(𝑑) has the opposite sign to 𝑔(𝑏𝑛 ), we put 𝑎𝑛+1 = 𝑑, 𝑏𝑛+1 = 𝑏𝑛
and get again 𝑔(𝑎𝑛+1 )𝑔(𝑏𝑛+1 ) < 0.
If this procedure does not stop, we obtain an increasing sequence (𝑎𝑛 ) and a decreasing sequence (𝑏𝑛 ),
both of which converge by Theorem 2.10. We observe that
1
𝑏𝑛 = 𝑎𝑛 + (𝑏𝑛−1 − 𝑎𝑛−1 ) = 𝑎𝑛 + 2−𝑛 (𝑏 − 𝑎).
2
Then
𝑐 ∶= lim 𝑏𝑛 = lim 𝑎𝑛 + lim 2−𝑛 (𝑏 − 𝑎) = lim 𝑎𝑛 .
Note. You have seen the definition of interval in first year and you will recall that that the definition
required several cases, depending on whether the endpoints belong to the interval or not and whether the
interval is bounded (above, below), namely (𝑎, 𝑏), [𝑎, 𝑏), (𝑎, 𝑏], [𝑎, 𝑏], (−∞, 𝑏), (−∞, 𝑏], (𝑎, ∞), [𝑎, ∞),
(−∞, ∞) where 𝑎, 𝑏 ∈ ℝ and 𝑎 < 𝑏. However, intervals can be characterized by one common property.
For this we need the following notion: A subset 𝑆 of ℝ is called a singleton if the set 𝑆 has exactly one
element.
Note. A subset 𝑆 of ℝ is an interval if and only if it contains at least two elements and if all real numbers
between any two elements in 𝑆 also belong to 𝑆.
Corollary 3.16. Let 𝐼 be an interval and let 𝑓 be a continuous real function on 𝐼. Then 𝑓 (𝐼) is either
an interval or a singleton.
Proof. In class.
Example 3.5.1. Let 𝑓 (𝑥) = 𝑥2 . Then 𝑓 ((−1, 2)) = [0, 4). Notice that 𝐼 = (−1, 2) is an open interval,
while 𝑓 (𝐼) is not.
Theorem 3.17. Let 𝑓 be a real function which is continuous on [𝑎, 𝑏], where 𝑎 < 𝑏. Then 𝑓 is bounded
on [𝑎, 𝑏], i. e., 𝑓 ([𝑎, 𝑏]) is bounded.
Proof. In class.
Theorem 3.18. A continuous function on a closed bounded interval achieves its supremum and infimum.
Proof. In Class.
Corollary 3.19. If 𝑓 is continuous on [𝑎, 𝑏], 𝑎 < 𝑏, then either 𝑓 ([𝑎, 𝑏]) is a singleton or 𝑓 ([𝑎, 𝑏]) = [𝑐, 𝑑]
with 𝑐 < 𝑑.
Proof. In class.
Theorem 3.20. Let 𝐼 be an interval and 𝑓 ∶ 𝐼 → ℝ be a stricly monotonic continuous function. Then
𝑓 (𝐼) is an interval, and the inverse function 𝑓 −1 ∶ 𝑓 (𝐼) → ℝ is continuous.
Proof. In class.
Tutorial 3.5.1. 1. Prove that a subset 𝑆 of ℝ is an interval if and only if 𝑆 has the form (𝑎, 𝑏), [𝑎, 𝑏),
(𝑎, 𝑏], [𝑎, 𝑏], (−∞, 𝑏), (−∞, 𝑏], (𝑎, ∞), [𝑎, ∞) or (−∞, ∞) where 𝑎, 𝑏 ∈ ℝ and 𝑎 < 𝑏.
Hint. Consider inf 𝑆 and sup 𝑆.
2. Let 𝑓 be a real function and let ∅ ≠ 𝐴 ⊂ 𝐵 ⊂ dom(𝑓 ) such that for each 𝑥 ∈ 𝐴 there is 𝜀 > 0 such
that (𝑥 − 𝜀, 𝑥] ⊂ 𝐴 or [𝑥, 𝑥 + 𝜀) ⊂ 𝐴 or (𝑥 − 𝜀, 𝑥 + 𝜀) ⊂ 𝐴, and the same property for 𝐵. Show that if 𝑓
is continuous on 𝐵, then 𝑓 is also continuous on 𝐴.
3. A fixed point theorem. Let 𝑎 < 𝑏 and let 𝑓 be a continuous function on [𝑎, 𝑏] such that 𝑓 ([𝑎, 𝑏]) ⊂
[𝑎, 𝑏]. Show that there is 𝑥 ∈ [𝑎, 𝑏] such that 𝑓 (𝑥) = 𝑥.
4. Let 𝐼 be an interval and 𝑓 be a continuous function on 𝐼 such that 𝑓 (𝐼) is unbounded. What can you
say about 𝑓 (𝐼)? Find examples which illustrate your answer.
5. Let 𝑓 ∶ [0, 1] → ℝ be a continuous function which only assumes rational values. Show that 𝑓 is
constant.
6. Find a continuous function 𝑓 ∶ [−1, 1] → ℝ which is one-to-one when restricted to rational numbers
in [−1, 1] but which is not one-to-one on the whole interval [−1, 1].
Chapter 4
Differentiation
𝑓 (𝑥) − 𝑓 (𝑎)
𝑓 ′ (𝑥) = lim
𝑥→𝑎 𝑥−𝑎
exists. The number 𝑓 ′ (𝑥) is called the derivative of 𝑓 .
2. 𝑓 is called differentiable on 𝐴 if 𝑓 is differentiable at each 𝑎 ∈ 𝐴.
𝑑𝑓 𝑑
Note. We also use the notations or 𝑓 for 𝑓 ′ .
𝑑𝑥 𝑑𝑥
Theorem 4.1. Let 𝑓 ∶ 𝐴 → ℝ and 𝑎 ∈ 𝐴. If 𝑓 is differentiable at 𝑎, then 𝑓 is continuous at 𝑎.
Proof. From
𝑓 (𝑥) − 𝑓 (𝑎)
𝑓 (𝑥) = 𝑓 (𝑎) + (𝑥 − 𝑎)
𝑥−𝑎
it follows that
𝑓 (𝑥) − 𝑓 (𝑎)
lim 𝑓 (𝑥) = 𝑓 (𝑎) + lim lim (𝑥 − 𝑎) = 𝑓 (𝑎) + [𝑓 ′ (𝑎)](0) = 𝑓 (𝑎).
𝑥→𝑎 𝑥→𝑎 𝑥−𝑎 𝑥→𝑎
Theorem 4.2 (Rules for the derivative). Let 𝑓 , 𝑔 ∶ 𝐴 → ℝ, 𝑎 ∈ 𝐴, 𝑓 and 𝑔 differentiable at 𝑎, and
𝑐 ∈ ℝ. Then
1. Linearity of the derivative: 𝑓 + 𝑔 and 𝑐𝑓 are differentiable at 𝑎, and
37
and
(𝑐𝑓 )(𝑥) − (𝑐𝑓 )(𝑎) 𝑐𝑓 (𝑥) − 𝑐𝑓 (𝑎) 𝑓 (𝑥) − 𝑓 (𝑎)
(𝑐𝑓 )′ (𝑎) = lim = lim = 𝑐 lim = 𝑐𝑓 ′ (𝑎).
𝑥→𝑎 𝑥−𝑎 𝑥→𝑎 𝑥−𝑎 𝑥→𝑎 𝑥−𝑎
2. Observing Theorem 4.1 we have
(𝑓 𝑔)(𝑥) − (𝑓 𝑔)(𝑎)
(𝑓 𝑔)′ (𝑎) = lim
𝑥→𝑎 𝑥−𝑎
𝑓 (𝑥)𝑔(𝑥) − 𝑓 (𝑎)𝑔(𝑥) + 𝑓 (𝑎)𝑔(𝑥) − 𝑓 (𝑎)𝑔(𝑎)
= lim
𝑥→𝑎 𝑥−𝑎
[𝑓 (𝑥) − 𝑓 (𝑎)]𝑔(𝑥) 𝑓 (𝑎)[𝑔(𝑥) − 𝑔(𝑎)]
= lim + lim
𝑥→𝑎 𝑥−𝑎 𝑥→𝑎 𝑥−𝑎
𝑓 (𝑥) − 𝑓 (𝑎) 𝑔(𝑥) − 𝑔(𝑎)
= lim lim 𝑔(𝑥) + 𝑓 (𝑎) lim
𝑥→𝑎 𝑥−𝑎 𝑥→𝑎 𝑥→𝑎 𝑥−𝑎
′ ′
= 𝑓 (𝑎)𝑔(𝑎) + 𝑓 (𝑎)𝑔 (𝑎).
𝑑𝑐
Example 4.1.1. 1. If 𝑐 ∈ ℝ, then 𝑑𝑥 = 0.
𝑑
2. For 𝑛 ∈ ℕ∗ we have 𝑥𝑛 = 𝑛𝑥𝑛−1 .
𝑑𝑥
Proof. 1. follows immediately from the definition.
2. We present an alternate proof to that given in first year: proof by induction. Again, for 𝑛 = 1 the
statement is straightforward:
𝑑 𝑦−𝑥
𝑥 = lim = 1 = (1)𝑥1−1 .
𝑑𝑥 𝑦→𝑥 𝑦−𝑥
Now assume the statement is true for 𝑛 ∈ ℕ∗ . Then, making use of the Product Rule,
( )
𝑑 𝑛+1 𝑑 𝑑 𝑛 𝑑
𝑥 = ((𝑥𝑛 )(𝑥)) = 𝑥 (𝑥)+𝑥𝑛 𝑥 = 𝑛𝑥𝑛−1 (𝑥)+𝑥𝑛 (1) = (𝑛+1)𝑥𝑛 = (𝑛+1)𝑥(𝑛+1)−1 .
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑 𝑛
As in first year calculus, we now obtain the derivative 𝑥 = 𝑛𝑥𝑛−1 for negative integers 𝑛 with the aid
𝑑𝑥
of the quotient rule. Also, the derivatives of the trigonometric functions sin, cos, tan have been derived
in first year calculus and should be known.
The proof of the differentiability of 𝑒𝑥 in first year calculus was incomplete; it will be given in the next
section.
Also, the differentiability of the inverse trigonometric functions, ln 𝑥 and 𝑥𝑟 for 𝑟 ∈ ℝ ⧵ ℕ will be
postponed to the next section.
Theorem 4.3 (Fermat’s Theorem). Let 𝐼 be an interval, 𝑓 ∶ 𝐼 → ℝ, and let 𝑐 be in the interior of 𝐼. If
𝑓 has a local maximum or local minimum at 𝑐 and 𝑓 is differentiable at 𝑐, then 𝑓 ′ (𝑐) = 0.
Proof. Since 𝑐 is an interior point of 𝐼, there is 𝜀0 > 0 such that (𝑐 − 𝜀0 , 𝑐 + 𝜀0 ) ⊂ 𝐼. Assume that 𝑓 has
a local maximum at 𝑐. Then there is 𝜀 ∈ (0, 𝜀0 ) such that 𝑓 (𝑥) ≤ 𝑓 (𝑐) for all 𝑥 ∈ (𝑐 − 𝜀, 𝑐 + 𝜀). Therefore
≥ 0 for 𝑥 ∈ (𝑐 − 𝜀, 𝑐),
{
𝑓 (𝑥) − 𝑓 (𝑐)
𝑥−𝑐 ≤ 0 for 𝑥 ∈ (𝑐, 𝑐 + 𝜀).
≥ lim− 0 = 0
𝑓 (𝑥) − 𝑓 (𝑐)
𝑓 ′ (𝑐) = lim−
𝑥→𝑐 𝑥−𝑐 𝑥→𝑐
and
≤ lim+ 0 = 0.
𝑓 (𝑥) − 𝑓 (𝑐)
𝑓 ′ (𝑐) = lim+
𝑥→𝑐 𝑥−𝑐 𝑥→𝑐
Theorem 4.4 (Rolle’s Theorem). Let 𝑎 < 𝑏 be real numbers and 𝑓 ∶ [𝑎, 𝑏] → ℝ be a function having
the following properties:
1. 𝑓 is continuous on the closed interval [𝑎, 𝑏],
2. 𝑓 is differentiable on the open interval (𝑎, 𝑏),
3. 𝑓 (𝑎) = 𝑓 (𝑏).
Then there is 𝑐 ∈ (𝑎, 𝑏) such that 𝑓 ′ (𝑐) = 0.
Proof. If 𝑓 is constant, then 𝑓 ′ = 0, and the statement is true for any 𝑐 ∈ (𝑎, 𝑏). If 𝑓 is not constant, then
there is 𝑥0 ∈ [𝑎, 𝑏] such that 𝑓 (𝑥0 ) ≠ 𝑓 (𝑎). We may assume 𝑓 (𝑥0 ) > 𝑓 (𝑎). Otherwise, 𝑓 (𝑥0 ) < 𝑓 (𝑎),
and we can consider −𝑓 . Since 𝑓 is continuous on [𝑎, 𝑏] by property 1, 𝑓 has a maximum on [𝑎, 𝑏], see
Corollary 3.19. That is, there is some 𝑐 ∈ [𝑎, 𝑏] such that 𝑓 (𝑐) ≥ 𝑓 (𝑥) for all 𝑥 ∈ [𝑎, 𝑏]. In particular,
𝑓 (𝑐) ≥ 𝑓 (𝑥0 ) > 𝑓 (𝑎). Since 𝑓 (𝑎) = 𝑓 (𝑏), we have 𝑐 ≠ 𝑎 and 𝑐 ≠ 𝑏, and therefore 𝑐 ∈ (𝑎, 𝑏). Hence
𝑓 ′ (𝑐) = 0 by Fermat’s Theorem.
Theorem 4.5 (First Mean Value Theorem). Let 𝑎 < 𝑏 be real numbers and 𝑓 ∶ [𝑎, 𝑏] → ℝ be a
continuous function which is differentiable on (𝑎, 𝑏).
Then there is 𝑐 ∈ (𝑎, 𝑏) such that
𝑓 (𝑏) − 𝑓 (𝑎)
= 𝑓 ′ (𝑐).
𝑏−𝑎
Proof. The function
𝑓 (𝑏) − 𝑓 (𝑎)
𝑔(𝑥) = 𝑓 (𝑥) − (𝑥 − 𝑎)
𝑏−𝑎
is continuous on [𝑎, 𝑏], differentiable on (𝑎, 𝑏), and 𝑔(𝑎) = 𝑓 (𝑎) = 𝑔(𝑏). Hence 𝑔 satisfies the assumptions
of Rolle’s theorem. Therefore, there is 𝑐 ∈ (𝑎, 𝑏) such that 𝑔 ′ (𝑐) = 0. But
𝑓 (𝑏) − 𝑓 (𝑎)
𝑔 ′ (𝑥) = 𝑓 ′ (𝑥) − ,
𝑏−𝑎
and substituting 𝑐 completes the proof.
Tutorial 4.1.1. Let 𝑎 < 𝑏 be real numbers, 𝑓 , 𝑔 ∶ [𝑎, 𝑏] → ℝ be continuous and differentiable on (𝑎, 𝑏).
1. Show that 𝑔 is injective on [𝑎, 𝑏] if 𝑔 ′ (𝑥) ≠ 0 for all 𝑥 ∈ (𝑎, 𝑏).
2. Prove the Second Mean Value Theorem: If 𝑔 ′ (𝑥) ≠ 0 for all 𝑐 ∈ (𝑎, 𝑏), then there is 𝑐 ∈ (𝑎, 𝑏) such
that
𝑓 (𝑏) − 𝑓 (𝑎) 𝑓 ′ (𝑐)
= ′ .
𝑔(𝑏) − 𝑔(𝑎) 𝑔 (𝑐)
Hint. Consider the function
𝑓 (𝑏) − 𝑓 (𝑎)
ℎ(𝑥) = 𝑓 (𝑥) − (𝑔(𝑥) − (𝑔(𝑎)).
𝑔(𝑏) − 𝑔(𝑎)
condition that 𝑔 ′ (𝑥) ≠ 0 for all 𝑥 ∈ (𝑎, 𝑏) with the weaker condition that 𝑔(𝑎) ≠ 𝑔(𝑏) and that there is no
3. Show that the statement of the Second Mean Value Theorem remains correct if one replaces the
*Tutorial 4.1.2. Let −∞ ≤ 𝑎 < 𝑏 ≤ ∞ and let 𝑓 , 𝑔 ∶ (𝑎, 𝑏) → ℝ be differentiable on (𝑎, 𝑏) such that
𝑔 ′ (𝑥) ≠ 0 for all 𝑥 ∈ (𝑎, 𝑏). Prove the following one-sided versions of l’Hôpital’s Rule.
𝑓 ′ (𝑥)
1. If lim+ 𝑓 (𝑥) = lim+ 𝑔(𝑥) = 0 and lim+ ′ exists as a proper or improper limit, then
𝑥→𝑎 𝑥→𝑎 𝑥→𝑎 𝑔 (𝑥)
𝑓 (𝑥) 𝑓 ′ (𝑥)
lim+ = lim+ ′ .
𝑥→𝑎 𝑔(𝑥) 𝑥→𝑎 𝑔 (𝑥)
𝑓 ′ (𝑥)
2. If lim+ 𝑓 (𝑥) = lim+ 𝑔(𝑥) = ∞ and lim+ exists as a proper or improper limit, then
𝑥→𝑎 𝑥→𝑎 𝑥→𝑎 𝑔 ′ (𝑥)
𝑓 (𝑥) 𝑓 ′ (𝑥)
lim+ = lim+ ′ .
𝑥→𝑎 𝑔(𝑥) 𝑥→𝑎 𝑔 (𝑥)
is continuous at 𝑎.
(iii) There is a function 𝑓̃ ∶ 𝐼 → ℝ such that 𝑓̃ is continuous at 𝑎 and
(iv) There are 𝛾 ∈ ℝ and a function 𝜀 ∶ 𝐼 → ℝ such that 𝜀(𝑎) = 0, 𝜀 is continuous at 𝑎, and
If (ii) or (iv) hold, then the number 𝛾 is unique and 𝛾 = 𝑓 ′ (𝑎). If (iii) holds, then 𝑓̃(𝑎) = 𝑓 ′ (𝑎).
Theorem 4.7 (Chain Rule). Let 𝐼 and 𝐽 be intervals, 𝑔 ∶ 𝐽 → ℝ and 𝑓 ∶ 𝐼 → ℝ with 𝑓 (𝐼) ⊂ 𝐽 , and let
𝑎 ∈ 𝐼. Assume that 𝑓 is differentiable at 𝑎 and that 𝑔 is differentiable at 𝑓 (𝑎). Then 𝑔◦𝑓 is differentiable
at 𝑎 and
(𝑔◦𝑓 )′ (𝑎) = 𝑔 ′ (𝑓 (𝑎))𝑓 ′ (𝑎).
Proof. In class.
Theorem 4.8. Let 𝐼 be an interval, let 𝑓 ∶ 𝐼 → ℝ be continuous and strictly increasing or decreasing,
and 𝑏 ∈ 𝑓 (𝐼). Assume that 𝑓 is differentiable at 𝑎 = 𝑓 −1 (𝑏) with 𝑓 ′ (𝑎) ≠ 0. Then 𝑓 −1 is differentiable
at 𝑏 = 𝑓 (𝑎) and
1 1
(𝑓 −1 )′ (𝑏) = or, equivalently, (𝑓 −1 )′ (𝑓 (𝑎)) = .
𝑓 ′ (𝑓 −1 (𝑏)) 𝑓 ′ (𝑎)
Proof. In class.
This result now allows us to find the derivatives of arcsin and arctan.
Example 4.2.1. Show that arcsin ∶ [−1, 1] → [− 𝜋2 , 𝜋2 ] is continuous on [−1, 1] and differentiable on
𝑑
(−1, 1), and find arcsin.
𝑑𝑥
Solution. In class.
Finally, in this section we shall show that 𝑒𝑥 = exp(𝑥) is differentiable and find its derivative. This will
then allow us to find the derivative of ln 𝑥 as an inverse function.
Chapter 5
Series
∑
∞
𝑎𝑛
𝑛=1
∑
∞
Definition 5.1. Let 𝑎𝑛 be a series.
𝑛=1
∑𝑛
1. The number 𝑠𝑛 = 𝑎𝑖 is called the 𝑛-th partial sum of the series.
𝑖=1
∑
∞
2. The series 𝑎𝑛 is said to converge if the sequence (𝑠𝑛 ) converges.
𝑛=1
In this case, the number 𝑠 = lim 𝑠𝑛 is called the sum of the series and we write
𝑛→∞
∑
∞
𝑎𝑛 = 𝑠
𝑛=1
∑
∞
Note. 1. Observe that 𝑎𝑛 denotes a series (convergent or divergent) as well as its sum if it converges.
𝑛=1
2. A series does not have to start at 𝑛 = 1. The change of notation is obvious for other starting indices.
𝑎 ≠ 0, 𝑟 ∈ ℝ.
∑
∞
𝑎𝑟𝑛 ,
𝑛=0
𝑠𝑛 = 𝑎 + 𝑎𝑟 + 𝑎𝑟2 + ⋯ + 𝑎𝑟𝑛 ,
𝑟𝑠𝑛 = 𝑎𝑟 + 𝑎𝑟2 + ⋯ + 𝑎𝑟𝑛 + 𝑎𝑟𝑛+1 .
43
and diverges if |𝑟| > 1 or 𝑟 = −1. Finally, for 𝑟 = 1, 𝑠𝑛 = 𝑎(𝑛 + 1), and thus 𝑠𝑛 → ∞ as 𝑛 → ∞.
∑
∞
Example 5.1.2. Is the series 22𝑛 31−𝑛 convergent or divergent?
𝑛=1
∑
∞
Theorem 5.2. If the series 𝑎𝑛 converges, then lim 𝑎𝑛 = 0.
𝑛→∞
𝑛=1
Proof. Let
𝑠𝑛 = 𝑎1 + 𝑎2 + ⋯ + 𝑎𝑛 .
Then
𝑎𝑛 = 𝑠𝑛 − 𝑠𝑛−1 .
∑∞
Since 𝑎𝑛 converges,
𝑛=1
lim 𝑠 =𝑠
𝑛→∞ 𝑛
exists. Since also 𝑛 − 1 → ∞ as 𝑛 → ∞,
lim 𝑠 = 𝑠.
𝑛→∞ 𝑛−1
Hence
lim 𝑎 = lim (𝑠𝑛 − 𝑠𝑛−1 ) = lim 𝑠𝑛 − lim 𝑠𝑛−1 = 𝑠 − 𝑠 = 0.
𝑛→∞ 𝑛 𝑛→∞ 𝑛→∞ 𝑛→∞
Theorem 5.3 (Test for Divergence). If lim 𝑎𝑛 does not exist or if lim 𝑎𝑛 ≠ 0, then the series
∑
∞
𝑎𝑛
𝑛→∞ 𝑛→∞
𝑛=1
diverges.
∑
∞
Note. From lim 𝑎𝑛 = 0 nothing can be concluded about the convergence of the series 𝑎𝑛 .
𝑛→∞
𝑛=1
∑
∞
∑
∞
Theorem 5.4 (Sum Laws). Let 𝑐 ∈ ℝ and suppose that 𝑎𝑛 and 𝑏𝑛 both converge. Then also
𝑛=1 𝑛=1
∑
∞
∑
∞
∑
∞
(𝑐𝑎𝑛 ), (𝑎𝑛 + 𝑏𝑛 ) and (𝑎𝑛 − 𝑏𝑛 ) converge, and
𝑛=1 𝑛=1 𝑛=1
∑
∞
∑
∞
1. (𝑐𝑎𝑛 ) = 𝑐 𝑎𝑛 ,
𝑛=1 𝑛=1
∑∞
∑
∞
∑
∞
2. (𝑎𝑛 + 𝑏𝑛 ) = 𝑎𝑛 + 𝑏𝑛 ,
𝑛=1 𝑛=1 𝑛=1
∑
∞
∑
∞
∑
∞
3. (𝑎𝑛 − 𝑏𝑛 ) = 𝑎𝑛 − 𝑏𝑛 ,
𝑛=1 𝑛=1 𝑛=1
( )
∑
∞
5 1
Example 5.1.4. Find the sum of the series + .
𝑛=1
𝑛(𝑛 + 1) 5𝑛
Note. For convergence it does not matter at which index the series starts. But it matters for the sum.
∑
∞
Theorem 5.5. The series 𝑎𝑛 converges if and only if for each 𝜀 > 0 there is 𝐾 ∈ ℕ such that for all
𝑛=1
|∑ |
𝑚 ≥ 𝑘 ≥ 𝐾, | 𝑎𝑛 | < 𝜀.
𝑚
| |
| |
|𝑛=𝑘 |
Proof. In class.
Tutorial 5.1.1. 1. Test each of the following series for convergence or divergence:
∞ ( ) ( )
∑ 1 𝑛 ∑∞
1 ∑∞
𝑛2 − 1
(a) 1+ , (b) 𝑛 sin , (c) .
𝑛=1
𝑛 𝑛=1
𝑛 𝑛=1
𝑛 − 50𝑛2
2. Which of the following is valid? Justify your conclusions.
∑
∞
(a) If 𝑎𝑛 → 0 as 𝑛 → ∞, then 𝑎𝑛 is convergent.
𝑛=1
∑∞
(b) If 𝑎𝑛 ̸→ 0 as 𝑛 → ∞, then 𝑎𝑛 is divergent.
𝑛=1
∑
∞
(c) If 𝑎𝑛 is divergent, then 𝑎𝑛 ̸→ 0 as 𝑛 → ∞.
𝑛=1
Proof. In class.
∑
∞
∑
∞
Theorem 5.7 (Comparison Test). Let 𝑎𝑛 and 𝑏𝑛 be series with nonnegative terms and assume
∑
∞
∑
∞
(i) If 𝑏𝑛 converges, then also 𝑎𝑛 converges.
𝑛=1 𝑛=1
∑∞
∑
∞
(ii) If 𝑎𝑛 diverges, then also 𝑏𝑛 diverges.
𝑛=1 𝑛=1
Proof. In class.
∑
∞
sin2 𝑛 + 10
Example 5.2.1. Test the series for convergence.
𝑛=1
𝑛 + 2𝑛
∑
∞
Definition 5.2. 1. A series 𝑎𝑛 is called absolutely convergent if the series of its absolute values
𝑛=1
∑
∞
|𝑎𝑛 | converges.
𝑛=1
∑
∞
2. A series 𝑎𝑛 is called conditionally convergent if it is convergent but not absolutely convergent.
𝑛=1
Proof. In class.
Recall from Calculus I that there are convergent series which are not absolutely convergent.
with 𝑏𝑛 ≥ 0.
∑
∞
∑
∞
(−1)𝑛 𝑏𝑛 or (−1)𝑛−1 𝑏𝑛
𝑛=1 𝑛=1
∑
∞
∑
∞
𝑛
Theorem 5.9 (Alternating series test). If the alternating series (−1) 𝑏𝑛 or (−1)𝑛−1 𝑏𝑛 satisfies
(ii) lim 𝑏𝑛 = 0,
𝑛→∞
then the series converges.
∑
𝑘+2𝑚
𝑘
(−1) (−1)𝑛 𝑏𝑛 = (𝑏𝑘 − 𝑏𝑘+1 ) + (𝑏𝑘+2 − 𝑏𝑘+3 ) + ⋯ + (𝑏𝑘+2𝑚−2 − 𝑏𝑘+2𝑚−1 ) + 𝑏𝑘+2𝑚
𝑛=𝑘
= 𝑏𝑘 − (𝑏𝑘+1 − 𝑏𝑘+2 ) − ⋯ − (𝑏𝑘+2𝑚−1 − 𝑏𝑘+2𝑚 ).
Hence
𝑛=𝑘
Similarly,
𝑛=𝑘
Now let 𝜀 > 0. Since 𝑏𝑘 → 0 as 𝑘 → ∞, there is 𝐾 ∈ ℕ such that 𝑏𝐾 < 𝜀. Hence for all 𝑙 ≥ 𝑘 ≥ 𝐾:
|∑ |
| (−1)𝑛 𝑏 | ≤ 𝑏 ≤ 𝑏 < 𝜀.
| 𝑙 |
| 𝑛| 𝑘 𝐾
|𝑛=𝑘 |
| |
Hence the alternating series converges.
Note. lim 𝑏𝑛 = 0 is necessary by Theorem 5.3 since lim (−1)𝑛 𝑏𝑛 = 0 ⇔ lim 𝑏𝑛 = 0 ⇔ lim (−1)𝑛−1 𝑏𝑛 =
𝑛→∞ 𝑛→∞ 𝑛→∞ 𝑛→∞
0.
∑∞
3𝑛
Example 5.2.2. The alternating series (−1)𝑛−1 does not converge since
𝑛=1
4𝑛 − 1
= ≠0
3𝑛 3
lim
𝑛→∞ 4𝑛 − 1 4
and thus (ii) is not satisfied, which is necessary for convergence. See the note following the statement of
the Alternating Series Test.
∑
∞
𝑛2
Worked Example 5.2.3. Find whether the series (−1)𝑛−1 is convergent.
𝑛=1
𝑛3 + 1
| 𝑎𝑛+1 | ∑∞
Theorem 5.10 (Ratio Test). (i) If lim sup || | = 𝐿 < 1, then the series
| 𝑎𝑛 converges absolutely.
𝑛→∞ | 𝑎𝑛 | 𝑛=1
| 𝑎𝑛+1 | ∑∞
(ii) If lim inf || | = 𝑙 > 1, then the series 𝑎𝑛 diverges.
𝑛→∞ | 𝑎𝑛 ||
𝑛=1
Proof. In class.
√ ∑
∞
Theorem 5.11 (Root Test). (i) If lim sup 𝑛
|𝑎𝑛 | = 𝐿 < 1, then the series 𝑎𝑛 converges absolutely.
𝑛→∞ 𝑛=1
√ ∑
∞
(ii) If lim sup 𝑛
|𝑎𝑛 | = 𝐿 > 1, then the series 𝑎𝑛 diverges.
𝑛→∞ 𝑛=1
Proof. In class.
{( )𝑛 ( )𝑛 } ( )
∑
∞
1 3 1
Tutorial 5.2.1. 1. Test +5 𝑛 sin for convergence.
𝑛=1
2 4 𝑛
2. Prove that the sequence (𝑎𝑛 )∞
𝑛=1
converges if and only if
(i) (𝑎2𝑛 )∞
𝑛=1
converges,
(ii) (𝑎2𝑛−1 )∞
𝑛=1
converges,
(iii) (𝑎𝑛 − 𝑎𝑛−1 ) → 0 as 𝑛 → ∞.
3. Use Tutorial 2 to prove the Alternating Series Test.
Hint. Show that (𝑠2𝑛 )∞
𝑛=1
and (𝑠2𝑛−1 )∞
𝑛=1
are monotonic sequences.
4. Use the alternating series test, ratio test or root test to test for convergence:
∞ ( )2𝑛
∑ (−1)𝑛 𝑛 ∑∞
𝑛!2𝑛
(a) , (b) ,
𝑛=1
2𝑛 + 1 𝑛=1
(2𝑛)!
( ( ) )
∑∞
1 𝑛 ∑∞
𝑛𝑛
𝑛
(c) (−1) 𝑒 − 1 + , (d) ,
𝑛=1
𝑛 𝑛=1
(𝑛!)2
∑
∞ 𝑛
2
(e) .
𝑛=1
𝑛
∑
∞
𝑐𝑛 (𝑥 − 𝑎)𝑛
𝑛=0
Note that (𝑥 − 𝑎)0 = 1. For 𝑥 = 𝑎, all terms from the second onwards are 0, so the series converges to 𝑐0
for 𝑥 = 𝑎.
Each power series defines a function whose domain is those 𝑥 ∈ ℝ for which the series converges.
∑
∞
Notation. With each power series 𝑐𝑛 (𝑥 − 𝑎)𝑛 we associate a number 𝑅 or ∞, called the radius of
𝑛=0
convergence of the series, which is defined as follows:
√
(i) 𝑅 = 0 if lim sup 𝑛 |𝑐𝑛 | = ∞,
𝑛→∞
1 √
(ii) 𝑅 = √ if 0 < lim sup 𝑛
|𝑐𝑛 | ∈ ℝ,
lim sup 𝑛 |𝑐𝑛 | 𝑛→∞
𝑛→∞
√
(iii) 𝑅 = ∞ if lim sup 𝑛
|𝑐𝑛 | = 0.
𝑛→∞
∑
∞
Theorem 5.12. There are three alternatives for the domain of a power series 𝑐𝑛 (𝑥 − 𝑎)𝑛 :
𝑛=0
(i) If 𝑅 = 0, then the series converges only for 𝑥 = 𝑎.
(ii) If 𝑅 = ∞, then the series converges absolutely for all 𝑥 ∈ ℝ.
(iii) If 0 < 𝑅 ∈ ℝ, then the series converges absolutely if |𝑥 − 𝑎| < 𝑅 and diverges if |𝑥 − 𝑎| > 𝑅.
Note that (iii) says that the series converges for 𝑥 in the interval (𝑎 − 𝑅, 𝑎 + 𝑅) and diverges outside
[𝑎 − 𝑅, 𝑎 + 𝑅]. For 𝑥 = 𝑎 − 𝑅 and 𝑥 = 𝑎 + 𝑅 anything can happen, see Calculus I. In any case, the
domain of the series is an interval, called the interval of convergence.
√ √
lim sup 𝑛
|𝑐𝑛 (𝑥 − 𝑎)𝑛 | = lim sup 𝑛
|𝑐𝑛 | |𝑥 − 𝑎| = ∞.
𝑛→∞ 𝑛→∞
𝑛→∞
√ √
lim sup 𝑛
|𝑐𝑛 (𝑥 − 𝑎)𝑛 | = |𝑥 − 𝑎| lim sup 𝑛 |𝑐𝑛 |.
𝑛→∞ 𝑛→∞
√
Hence the power series converges for all 𝑥 ∈ ℝ if lim sup 𝑛 |𝑐𝑛 | = 0, proving (ii).
√ 𝑛→∞
If finally 0 < lim sup 𝑛 |𝑐𝑛 | ∈ ℝ, then, by the Root Test, the series converges if
𝑛→∞
√ 1
|𝑥 − 𝑎| lim sup 𝑛 |𝑐𝑛 | < 1, i. e., |𝑥 − 𝑎| < √ ,
𝑛→∞ lim sup 𝑛 |𝑐𝑛 |
𝑛→∞
and the series diverges if
1
|𝑥 − 𝑎| > √ .
lim sup 𝑛 |𝑐𝑛 |
𝑛→∞
√
Tutorial 5.3.1. 1. (a) Let 𝑐 > 1 and put 𝑐𝑛 = 𝑛 𝑐 − 1.
(i) Show that 𝑐𝑛 ≥ 0.
(ii) Show that lim sup 𝑐𝑛 ≤ 0. Hint. Use Bernoulli’s inequality.
𝑛→∞ √
(iii) Conclude that lim 𝑛 𝑐 = 1.
𝑛→∞ √
(b) Use (a) to show that lim 𝑛 𝑐 = 1 for all 𝑐 > 0.
𝑛→∞
| 𝑎𝑛+1 |
What can you say if lim || | exists or is ∞?
𝑛→∞ | 𝑎𝑛 ||
5. Find the radius and interval of convergence for each of the following power series:
∑∞
(2𝑥)𝑛 ∑∞
𝑥𝑛 ∑∞
(a) , (b) 𝑛
, (c) 𝑛𝑛 𝑥𝑛 ,
𝑛=1
𝑛 𝑛=0
𝑛 𝑛=0
∑∞
(𝑥 − 1)𝑛 ∑
∞
(−2𝑥)𝑛 ∑
∞
(d) √ , (e) , (f) (−1)𝑛 𝑥𝑛 ,
𝑛=1 3
𝑛 𝑛
𝑛=1
𝑛3 𝑛=0
∞ ( )𝑛2
∑ 𝑛 ∑ (𝑥 +
∞
3)𝑛 ∑ (𝑛𝑥)𝑛
∞
(g) 𝑥𝑛 , (h) , (i) .
𝑛=0
𝑛 + 1 𝑛=1
𝑛3 𝑛=0
(2𝑛)!
Chapter 6
Riemann Integration
Lemma 6.1. Let 𝐴 and 𝐵 be non-empty bounded sets of real numbers, let 𝑐 ∈ ℝ, and define
Proof. (a) Let 𝑀 = sup 𝐴. Then 𝑥 ≤ 𝑀 for all 𝑥 ∈ 𝐴 and therefore 𝑐𝑥 ≤ 𝑐𝑀 for all 𝑐𝑥 ∈ 𝑐𝐴. Hence
𝑐𝐴 is bounded above with bound 𝑐𝑀. This shows that 𝐿 = sup(𝑐𝐴) satisfies 𝐿 ≤ 𝑐𝑀.
If 𝑐 = 0 then 𝐿 = sup{0} = 0 = 0 ⋅ 𝑀 = 𝑐𝑀, and if 𝑐 > 0, the the above inequality gives
Proof. Assume 𝐵 is bounded. Let 𝑎 ∈ 𝐴. Then 𝑎 ∈ 𝐵, and hence inf 𝐵 ≤ 𝑎 ≤ sup 𝐵. It follows that 𝐴
is bounded, that inf 𝐵 is a lower bound of 𝐴 and that sup 𝐴 is an upper bound of 𝐴. Hence inf 𝐴 ≥ inf 𝐵
and sup 𝐴 ≤ sup 𝐵. The cases that inf 𝐵 = −∞ or sup 𝐵 = ∞ are similar.
51
Lemma 6.3. Let 𝐴 and 𝐵 be nonempty subsets of ℝ such that 𝑥 ≤ 𝑦 for all 𝑥 ∈ 𝐴 and 𝑦 ∈ 𝐵. Then 𝐴
is bounded above, 𝐵 is bounded below, and
sup 𝐴 ≤ inf 𝐵.
Proof. Let 𝑏 ∈ 𝐵. Then 𝑎 ≤ 𝑏 for all 𝑎 ∈ 𝐴. So 𝑏 is an upper bound of 𝐴. Hence 𝐴 is bounded above
and sup 𝐴 ≤ 𝑏 since sup 𝐴 is the least upper bound. Then sup 𝐴 ≤ 𝑏 for all 𝑏 ∈ 𝐵, and therefore sup 𝐴
is a lower bound for 𝐵. Thus sup 𝐴 ≤ inf 𝐵 follows.
The set of all partitions of the interval [𝑎, 𝑏] will be denoted by (𝑎, 𝑏).
Next we consider the upper and lower Riemann sums with respect to a partition.
Definition 6.2. Let 𝑓 be a bounded function on [𝑎, 𝑏], (i.e. 𝑓 ∶ [𝑎, 𝑏] → ℝ and there exists a constant 𝐾
such that |𝑓 (𝑥)| ≤ 𝐾 for all 𝑥 ∈ [𝑎, 𝑏]) and let 𝑃 = {𝑎 = 𝑥0 < ⋯ < 𝑥𝑁 = 𝑏} be a partition of the closed
finite interval [𝑎, 𝑏].
(a) We define the upper sum of 𝑓 over [𝑎, 𝑏] with respect to the partition 𝑃 by
∑
𝑁
𝑈 (𝑓 , 𝑃 ) ∶= (𝑥𝑗 − 𝑥𝑗−1 ) sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}.
𝑗=1
(b) We define the lower sum of 𝑓 over [𝑎, 𝑏] with respect to the partition 𝑃 by
∑
𝑁
𝐿(𝑓 , 𝑃 ) ∶= (𝑥𝑗 − 𝑥𝑗−1 ) inf {𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}.
𝑗=1
𝑓
𝑦
𝑥
𝑥0 𝑥1 𝑥2 𝑥3 𝑥𝑛
=
𝑎 𝑏
Observe that if 𝑓 (𝑥) ≥ 0 on [𝑎, 𝑏], then both 𝐿(𝑓 , 𝑃 ) and 𝑈 (𝑓 , 𝑃 ) are approximations of the area under
the curve 𝑦 = 𝑓 (𝑥) for 𝑥 ∈ [𝑎, 𝑏]. Here 𝐿(𝑓 , 𝑃 ) under-approximates the area, while 𝑈 (𝑓 , 𝑃 ) over-
approximates the area.
Worked Example 6.2.1. Find 𝑈 (𝑓 , 𝑃 ) and 𝐿(𝑓 , 𝑃 ) for 𝑓 (𝑥) = 𝑥2 − 1 and 𝑃 = {0 < 1∕4 < 1∕2 < 1}.
Answer: 𝑈 (𝑓 , 𝑃 ) = −27∕64 and 𝐿(𝑓 , 𝑃 ) = −55∕64.
∫
2
Remark. From 1st year: (𝑥2 − 1) 𝑑𝑥 = − so we observe that
3
0
𝐿(𝑥2 − 1, 𝑃 ) ≤ (𝑥2 − 1) 𝑑𝑥 ≤ 𝑈 (𝑓 , 𝑃 ).
∫
0
Lemma 6.4. Let 𝑓 be a bounded function on the closed interval [𝑎, 𝑏] and let 𝑃 be a partition of [𝑎, 𝑏].
Then
Also, since [𝑥𝑗−1 , 𝑥𝑗 ] ⊂ [𝑎, 𝑏], it follows from Lemma 6.2 that
(𝑥𝑗 − 𝑥𝑗−1 )inf 𝑓 (𝑥) ≤ (𝑥𝑗 − 𝑥𝑗−1 ) inf 𝑓 (𝑥) ≤ (𝑥𝑗 − 𝑥𝑗−1 )sup 𝑓 (𝑥) ≤
∑
𝑛
∑
𝑛
∑
𝑛
∑
𝑛
(𝑥𝑗 − 𝑥𝑗−1 )sup 𝑓 (𝑥)
𝑗=1 𝑥∈[𝑎,𝑏] 𝑗=1 𝑥∈[𝑥𝑗−1 ,𝑥𝑗 ] 𝑗=1 𝑥∈[𝑥𝑗−1 ,𝑥𝑗 ] 𝑗=1 𝑥∈[𝑎,𝑏]
=
Worked Example 6.2.2. Let 𝑓 (𝑥) = 𝑥2 and 𝑃 = {1 < 4∕3 < 5∕3 < 2}. Then 𝑈 (𝑓 , 𝑃 ) = 77∕27,
𝐿(𝑓 , 𝑃 ) = 50∕27 while (2 − 1) sup{𝑓 (𝑥) ∣ 𝑥 ∈ [1, 2]} = 4 and (2 − 1) inf {𝑓 (𝑥) ∣ 𝑥 ∈ [1, 2]} = 1,
verifying the above lemma.
Tutorial 6.2.1. 1. Let 𝑓 (𝑥) = 𝑥 − 21 and 𝑃 = {0 < 1∕2 < 1 < 3∕2 < 2}. Find 𝑈 (𝑓 , 𝑃 ) and 𝐿(𝑓 , 𝑃 ) and
compare them with ∫0 𝑓 (𝑥) 𝑑𝑥.
2
1 2 3𝑁
2. Let 𝑓 (𝑥) = 𝑥 and 𝑃𝑁 = {−1 < −1 + 𝑁
< −1 + 𝑁
< ⋯ < −1 + 𝑁
= 2}, 𝑁 ∈ ℕ∗ .
{ [ ]} { [ ]}
𝑗−1 𝑗 𝑗−1 𝑗
(i) Find sup 𝑓 (𝑥) ∣ 𝑥 ∈ −1 + 𝑁
, −1 + 𝑁
and inf 𝑓 (𝑥) ∣ 𝑥 ∈ −1 + 𝑁
, −1 + 𝑁
.
⎧1 if 𝑥 > 0
⎪
3. Verify Lemma 6.4 for the function sgn(𝑥) ∶= ⎨ 0 if 𝑥 = 0 on the interval [−2, 3] with partition
⎪ −1 if 𝑥 < 0
⎩
𝑃 = {−2 < −1 < 0 < 1 < 2 < 3}.
6.3 Refinements
Definition 6.3. Let 𝑃 = {𝑎 = 𝑥0 < 𝑥1 < ⋯ < 𝑥𝑚 = 𝑏} and 𝑄 = {𝑎 = 𝑦0 < 𝑦1 < ⋯ < 𝑦𝑛 =
𝑏} be partitions of [𝑎, 𝑏]. Then 𝑄 is said to be a refinement of 𝑃 if 𝑃 ⊂ 𝑄 i.e. {𝑥0 , 𝑥1 , … , 𝑥𝑚 } ⊂
{𝑦0 , 𝑦1 , … , 𝑦𝑛 }.
Worked Example 6.3.1. 1. 𝑃 = {0 < 1∕2 < 1} and 𝑄 = {0 < 1∕3 < 2∕3 < 1} are both partitions of
[0, 1] but 𝑄 is not a refinement of 𝑃 .
2. 𝑃 = {0 < 1∕2 < 1} and 𝑄 = {0 < 1∕4 < 2∕4 < 3∕4 < 1} are both partitions of [0, 1] and 𝑄 is a
refinement of 𝑃 .
𝑛
3. Let 𝑃𝑛 = {0 < 21𝑛 < 22𝑛 < ⋯ < 22𝑛 }. Then 𝑃𝑛 is a partition of [0, 1] for each 𝑛 ∈ ℕ and 𝑃𝑛 ⊂ 𝑃𝑛+1
𝑗 2𝑗
since 2𝑛
= 2𝑛+1
.
Lemma 6.5. Let 𝑓 be a bounded function on [𝑎, 𝑏] and let 𝑃 and 𝑅 be partitions of [𝑎, 𝑏] with 𝑅 a
refinement of 𝑃 . Then
𝐿(𝑓 , 𝑃 ) ≤ 𝐿(𝑓 , 𝑅) ≤ 𝑈 (𝑓 , 𝑅) ≤ 𝑈 (𝑓 , 𝑃 ).
Proof. In class.
Definition 6.4. If 𝑃 and 𝑄 are partitions of [𝑎, 𝑏], then 𝑃 ∪ 𝑄 (with elements listed in increasing order
and without repetitions) is called the common refinement of 𝑃 and 𝑄.
Worked Example 6.3.2. Let 𝑃 = {−1 < 0 < 1} and 𝑄 = {−1 < −1∕3 < 1∕3 < 1}. Then the common
refinement of 𝑃 and 𝑄 is 𝑃 ∪ 𝑄 = {−1 < −1∕3 < 0 < 1∕3 < 1}.
Lemma 6.6. If 𝑃 and 𝑄 are partitions of [𝑎, 𝑏] and 𝑓 is a bounded function on [𝑎, 𝑏], then
𝐿(𝑓 , 𝑄) ≤ 𝑈 (𝑓 , 𝑃 ).
Proof. Since
𝑃 ⊂𝑃 ∪𝑄 and 𝑄 ⊂ 𝑃 ∪ 𝑄,
Lemma 6.5 gives
𝐿(𝑓 , 𝑄) ≤ 𝐿(𝑓 , 𝑃 ∪ 𝑄) ≤ 𝑈 (𝑓 , 𝑃 ∪ 𝑄) ≤ 𝑈 (𝑓 , 𝑃 ).
Hence
Tutorial 6.3.1. 1. Let 𝑃𝑛 = {0 < 1𝑛 < 2𝑛 < ⋯ < 𝑛𝑛 }. Then 𝑃𝑛 is a partition of [0, 1] for each 𝑛 ∈ ℕ.
Show that 𝑃𝑛+1 is not a refinement of 𝑃𝑛 for 𝑛 ≥ 2.
2. Let 𝑓 (𝑥) = 𝑥3 + 𝑥2 , 𝑃 = {−1 < 0 < 1}, 𝑄 = {−1 < −1∕2 < 0 < 1}. Compute 𝐿(𝑓 , 𝑃 ), 𝐿(𝑓 , 𝑄),
𝑈 (𝑓 , 𝑃 ), 𝑈 (𝑓 , 𝑄) and compare their values.
∫
𝑓 (𝑡) 𝑑𝑡 ∶= sup{𝐿(𝑓 , 𝑃 ) ∣ 𝑃 a partition of [𝑎, 𝑏]}.
𝑎
∫
𝑓 (𝑡) 𝑑𝑡 ∶= inf {𝑈 (𝑓 , 𝑃 ) ∣ 𝑃 a partition of [𝑎, 𝑏]}.
𝑎
𝑏 𝑏
∫ ∫
𝑓 (𝑡) 𝑑𝑡 = 𝑓 (𝑡) 𝑑𝑡,
𝑎 𝑎
∫
in which case the common value is denoted by 𝑓 (𝑡) 𝑑𝑡 and is called the Riemann integral of 𝑓 over
𝑎
[𝑎, 𝑏].
Note. By the properties of suprema and infima, see Lemmas 6.6 and 6.3,
𝑏 𝑏
𝑓 (𝑡) 𝑑𝑡 ≤
∫ ∫
𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
𝑏 𝑏
𝑓 (𝑡) 𝑑𝑡 ≤
∫ ∫
𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
Theorem 6.7 (Integrability Condition). A bounded function 𝑓 on the interval [𝑎, 𝑏] is Riemann integrable
if and only if for each 𝜀 > 0 there exists a partition 𝑃 of [𝑎, 𝑏] such that
𝑈 (𝑓 , 𝑃 ) − 𝐿(𝑓 , 𝑃 ) < 𝜀.
Proof. ⇒: Let 𝑓 be Riemann integrable on [𝑎, 𝑏] and let 𝜀 > 0. Then there is a partition 𝑄 of [𝑎, 𝑏] such
that
𝑏 𝑏
𝑓 (𝑡) 𝑑𝑡 ≤ 𝑈 (𝑓 , 𝑄) <
∫ ∫
𝜀
𝑓 (𝑡) 𝑑𝑡 +
2
𝑎 𝑎
and there is a partition 𝑅 of [𝑎, 𝑏] such that
𝑏 𝑏
𝑏 𝑏
≤ 𝑈 (𝑓 , 𝑃 ) ≤ 𝑈 (𝑓 , 𝑄) < 𝑓 (𝑡) 𝑑𝑡 + =
∫ 2 ∫
𝜀 𝜀
𝑓 (𝑡) 𝑑𝑡 + .
2
𝑎 𝑎
In particular,
𝑏
∫
𝜀
𝑈 (𝑓 , 𝑃 ) < 𝑓 (𝑡) 𝑑𝑡 + ,
2
𝑎
𝑏
∫
𝜀
−𝐿(𝑓 , 𝑃 ) < − 𝑓 (𝑡) 𝑑𝑡 + .
2
𝑎
Summing up gives
𝑈 (𝑓 , 𝑃 ) − 𝐿(𝑓 , 𝑃 ) < 𝜀.
⇐: Suppose that there are 𝜀 > 0 and a partition 𝑃𝜀 of [𝑎, 𝑏] such that
𝑈 (𝑓 , 𝑃𝜀 ) − 𝐿(𝑓 , 𝑃𝜀 ) < 𝜀.
Then from
𝑏
𝑓 (𝑡) 𝑑𝑡 ≥ 𝐿(𝑓 , 𝑃𝜀 ),
∫
𝑎
𝑓 (𝑡) 𝑑𝑡 ≤ 𝑈 (𝑓 , 𝑃𝜀 )
∫
𝑎
we get
𝑏 𝑏
Hence
𝑏 𝑏
0≤ 𝑓 (𝑡) 𝑑𝑡 ≤ 𝜀,
∫ ∫
∀𝜀 > 0 𝑓 (𝑡) 𝑑𝑡 −
𝑎 𝑎
giving
𝑏 𝑏
∫ ∫
𝑓 (𝑡) 𝑑𝑡 = 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
Theorem 6.8. If 𝑓 is an increasing function on the interval [𝑎, 𝑏], then 𝑓 is Riemann integrable on [𝑎, 𝑏].
Proof. By assumption, 𝑓 (𝑎) ≤ 𝑓 (𝑥) ≤ 𝑓 (𝑏) for all 𝑥 ∈ [𝑎, 𝑏]. Hence 𝑓 is bounded.
Let 𝜀 > 0 and choose 𝑁 ∈ ℕ such that
𝑓 (𝑏) − 𝑓 (𝑎)
(𝑏 − 𝑎) < 𝜀.
𝑁
Consider the partition
{ }
𝑏−𝑎 2(𝑏 − 𝑎) (𝑁 − 1)(𝑏 − 𝑎) 𝑁(𝑏 − 𝑎)
𝑃 = 𝑎<𝑎+ <𝑎+ <⋯<𝑎+ <𝑎+ =𝑏 .
𝑁 𝑁 𝑁 𝑁
Observing that over each subinterval of [𝑎, 𝑏] the increasing function 𝑓 takes its minimum at the left
endpoint and its maximum at the right endpoint of the interval, it follows that
{ | [ ]} ( )
∑
𝑁
𝑏−𝑎 | (𝑗 − 1)(𝑏 − 𝑎) 𝑗(𝑏 − 𝑎) ∑ 𝑁
𝑏−𝑎 𝑗(𝑏 − 𝑎)
𝑈 (𝑓 , 𝑃 ) = sup 𝑓 (𝑥) | 𝑥 ∈ 𝑎 + ,𝑎 + = 𝑓 𝑎+ ,
𝑁 | 𝑁 𝑁 𝑁 𝑁
𝑗=1 | 𝑗=1
{ | [ ]} ( )
∑𝑁
𝑏−𝑎 | (𝑗 − 1)(𝑏 − 𝑎) 𝑗(𝑏 − 𝑎) ∑𝑁
𝑏−𝑎 (𝑗 − 1)(𝑏 − 𝑎)
𝐿(𝑓 , 𝑃 ) = inf 𝑓 (𝑥) | 𝑥 ∈ 𝑎 + ,𝑎 + = 𝑓 𝑎+ .
𝑁 | 𝑁 𝑁 𝑁 𝑁
𝑗=1 | 𝑗=1
Hence
( ) ( )
∑
𝑁
𝑏−𝑎 𝑗(𝑏 − 𝑎) ∑ 𝑏−𝑎
𝑁−1
𝑗(𝑏 − 𝑎)
𝑈 (𝑓 , 𝑃 ) − 𝐿(𝑓 , 𝑝) = 𝑓 𝑎+ − 𝑓 𝑎+
𝑗=1
𝑁 𝑁 𝑗=0
𝑁 𝑁
𝑏−𝑎
= (𝑓 (𝑏) − 𝑓 (𝑎)) < 𝜀.
𝑁
Hence 𝑓 is integrable by Theorem 6.7.
Worked Example 6.4.1. 1. Let 𝑓 (𝑥) = 𝑥. Then 𝑓 is Riemann integrable on [0, 1] and ∫0 𝑓 (𝑡) 𝑑𝑡 =
1 1
.
2
To see this let 𝑃𝑛 = {0 < 1𝑛 < 2𝑛 < ⋯ < 𝑛𝑛 }. Then 𝐿(𝑓 , 𝑃𝑛 ) = 21 − 2𝑛1
and 𝑈 (𝑓 , 𝑃𝑛 ) = 12 1
+ 2𝑛 . Thus
1
𝑈 (𝑓 , 𝑃𝑛 ) − 𝐿(𝑓 , 𝑃𝑛 ) = , which can be made arbitrarily small by taking 𝑛 sufficiently large. Thus 𝑓
𝑛
is Riemann integrable. Also the integral is sandwiched between 𝐿(𝑓 , 𝑃𝑛 ) and 𝑈 (𝑓 , 𝑃𝑛 ), both of which
converge to , giving ∫0 𝑓 (𝑡) 𝑑𝑡 = .
1 1 1
2 2
{
0 if 𝑥 ∉ ℚ
2. Let 𝑓 (𝑥) = . Then 𝑓 is not Riemann integrable on [0, 1], since for each partition 𝑃
1 if 𝑥 ∈ ℚ
of [0, 1], 𝑈 (𝑓 , 𝑃 ) = 1, while 𝐿(𝑓 , 𝑃 ) = 0.
2
3. Let 𝑓 (𝑥) = sgn(𝑥) on [−1, 1]. Let 𝑃𝑛 ∶= {−1 < − 1𝑛 < 1𝑛 < 1} then 𝑈 (sgn, 𝑃𝑛 ) = and 𝐿(sgn, 𝑃𝑛 ) =
𝑛
2 4
− . Thus 𝑈 (sgn, 𝑃𝑛 ) − 𝐿(sgn, 𝑃𝑛 ) = → 0 as 𝑛 → ∞. For each 𝜀 > 0 there exists 𝑛 ∈ ℕ such that
𝑛 𝑛
4
< 𝜀. Hence sgn is Riemann integrable on [−1, 1]. Finally, lim 𝑈 (sgn, 𝑃𝑛 ) = lim 𝐿(sgn, 𝑃𝑛 ) = 0
gives ∫−1 sgn(𝑥) 𝑑𝑥 = 0.
𝑛 𝑛→∞ 𝑛→∞
1
Tutorial 6.4.1. Let 𝑓 ∶ [𝑎, 𝑏] → ℝ be bounded. For 𝑛 ∈ ℕ∗ let 𝑃𝑛 ∈ (𝑎, 𝑏) such that lim 𝐿(𝑓 , 𝑃𝑛 ) =
lim 𝑈 (𝑓 , 𝑃𝑛 ). Show that 𝑓 is Riemann integrable on [𝑎, 𝑏] and that ∫𝑎 𝑓 (𝑥) 𝑑𝑥 = lim 𝐿(𝑓 , 𝑃𝑛 ) =
𝑛→∞
𝑏
𝑛→∞ 𝑛→∞
lim 𝑈 (𝑓 , 𝑃𝑛 ).
𝑛→∞
Tutorial 6.4.2. Show that 𝑓 (𝑥) = 𝑥2 is Riemann integrable on [0, 1] and that ∫0 𝑓 (𝑡) 𝑑𝑡 = .
1 1
3
∫ ∫
𝑐𝑓 (𝑡) 𝑑𝑡 = 𝑐 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
∑
𝑁
𝑈 (𝑐𝑓 , 𝑃 ) = (𝑥𝑗 − 𝑥𝑗−1 ) sup{𝑐𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}
𝑗=1
∑
𝑁
= (𝑥𝑗 − 𝑥𝑗−1 )𝑐 sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}
𝑗=1
∑
𝑁
=𝑐 (𝑥𝑗 − 𝑥𝑗−1 ) sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}
𝑗=1
= 𝑐𝑈 (𝑓 , 𝑃 ).
Thus
∫
=𝑐 𝑓 (𝑡) 𝑑𝑡.
𝑎
∫ ∫
𝑐𝑓 (𝑡) 𝑑𝑡 = 𝑐 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
𝑏 𝑏
∫ ∫
But 𝑓 (𝑡) 𝑑𝑡 = 𝑓 (𝑡) 𝑑𝑡 now gives
𝑎 𝑎
𝑏 𝑏 𝑏 𝑏 𝑏
∫ ∫ ∫ ∫ ∫
𝑐𝑓 (𝑡) 𝑑𝑡 = 𝑐 𝑓 (𝑡) 𝑑𝑡 = 𝑐 𝑓 (𝑡) 𝑑𝑡 = 𝑐 𝑓 (𝑡) 𝑑𝑡 = 𝑐𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎 𝑎 𝑎 𝑎
𝑏 𝑏
∫ ∫
Hence 𝑐𝑓 is Riemann integrable and 𝑐𝑓 (𝑡) 𝑑𝑡 = 𝑐 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
Case II: 𝑐 = −1. Let 𝑃 = {𝑎 = 𝑥0 < 𝑥1 < ⋯ < 𝑥𝑁 = 𝑏} be a partition of [𝑎, 𝑏]. Then, using the results
of Section 6.1,
∑
𝑁
𝑈 (−𝑓 , 𝑃 ) = (𝑥𝑗 − 𝑥𝑗−1 ) sup{−𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}
𝑗=1
∑
𝑁
=− (𝑥𝑗 − 𝑥𝑗−1 ) inf {𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}
𝑗=1
= −𝐿(𝑓 , 𝑃 ).
Thus
𝑏
∫
= − 𝑓 (𝑡) 𝑑𝑡.
𝑎
Since so far we have only used that 𝑓 and −𝑓 are bounded, we can replace 𝑓 by −𝑓 and obtain
𝑏 𝑏 𝑏
∫ ∫ ∫
𝑓 (𝑡) 𝑑𝑡 = (−(−𝑓 (𝑡))) 𝑑𝑡 = − (−𝑓 (𝑡)) 𝑑𝑡,
𝑎 𝑎 𝑎
and therefore
𝑏 𝑏
∫ ∫
(−𝑓 (𝑡)) 𝑑𝑡 = − 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
Thus
𝑏 𝑏 𝑏 𝑏 𝑏
∫ ∫ ∫ ∫ ∫
(−𝑓 (𝑡)) 𝑑𝑡 = − 𝑓 (𝑡) 𝑑𝑡 = − 𝑓 (𝑡) 𝑑𝑡 = − 𝑓 (𝑡) 𝑑𝑡 = (−𝑓 (𝑡)) 𝑑𝑡.
𝑎 𝑎 𝑎 𝑎 𝑎
∫ ∫
(−𝑓 (𝑡)) 𝑑𝑡 = − 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
∫ ∫
(−𝑐)𝑓 (𝑡) 𝑑𝑡 = (−𝑐) 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
∫ ∫
(−(−𝑐)𝑓 (𝑡)) 𝑑𝑡 = − (−𝑐)𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎
∫ ∫ ∫ ∫
𝑐𝑓 (𝑡) 𝑑𝑡 = − (−𝑐)𝑓 (𝑡) 𝑑𝑡 = −(−𝑐) 𝑓 (𝑡) 𝑑𝑡 = 𝑐 𝑓 (𝑡) 𝑑𝑡.
𝑎 𝑎 𝑎 𝑎
Lemma 6.10. If 𝑓 and 𝑔 are bounded functions on [𝑎, 𝑏] and 𝑃 is a partition of [𝑎, 𝑏], then
𝑈 (𝑓 + 𝑔, 𝑃 ) ≤ 𝑈 (𝑓 , 𝑃 ) + 𝑈 (𝑔, 𝑃 ),
𝐿(𝑓 + 𝑔, 𝑃 ) ≥ 𝐿(𝑓 , 𝑃 ) + 𝐿(𝑔, 𝑃 ).
Proof. Let 𝑃 = {𝑥0 < 𝑥1 < ⋯ < 𝑥𝑁 } be a partion of [𝑎, 𝑏]. For each 𝑗 ∈ {1, … , 𝑁} and 𝑡 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]
whence
sup{𝑓 (𝑥) + 𝑔(𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]} ≤ sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]} + sup{𝑔(𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}.
𝑈 (𝑓 + 𝑔, 𝑃 ) ≤ 𝑈 (𝑓 , 𝑃 ) + 𝑈 (𝑔, 𝑃 ).
𝐿(𝑓 + 𝑔, 𝑃 ) ≥ 𝐿(𝑓 , 𝑃 ) + 𝐿(𝑔, 𝑃 ) can be shown similarly, or we may use the estimate for upper sums
and 𝐿(𝑓 , 𝑃 ) = −𝑈 (−𝑓 , 𝑃 ).
Theorem 6.11 (Additivity). If 𝑓 and 𝑔 are bounded Riemann integrable functions on the interval [𝑎, 𝑏],
then 𝑓 + 𝑔 is Riemann integrable and
𝑏 𝑏 𝑏
∫ ∫ ∫
(𝑓 (𝑡) + 𝑔(𝑡)) 𝑑𝑡 = 𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡.
𝑎 𝑎 𝑎
𝑓 (𝑡) 𝑑𝑡 ≥ 𝑈 (𝑓 , 𝑃1 ) − ,
∫
𝜀
2
𝑎
𝑏
𝑔(𝑡) 𝑑𝑡 ≥ 𝑈 (𝑔, 𝑃2 ) − .
∫
𝜀
2
𝑎
𝑈 (𝑓 + 𝑔, 𝑃 ) ≤ 𝑈 (𝑓 , 𝑃 ) + 𝑈 (𝑔, 𝑃 ) ≤ 𝑈 (𝑓 , 𝑃1 ) + 𝑈 (𝑔, 𝑃2 ).
Thus
≤
∫ ∫
𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡 + 𝜀.
𝑎 𝑎
𝑏 𝑏 𝑏
(𝑓 (𝑡) + 𝑔(𝑡)) 𝑑𝑡 ≤
∫ ∫ ∫
𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡.
𝑎 𝑎 𝑎
Similarly,
𝑏 𝑏 𝑏
(𝑓 (𝑡) + 𝑔(𝑡)) 𝑑𝑡 ≥
∫ ∫ ∫
𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡.
𝑎 𝑎 𝑎
𝑏 𝑏 𝑏 𝑏 𝑏 𝑏
∫ ∫ ∫ ∫ ∫ ∫
𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡 = 𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡 = 𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡,
𝑎 𝑎 𝑎 𝑎 𝑎 𝑎
and therefore
𝑏 𝑏 𝑏 𝑏
∫ ∫ ∫
(𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡)) 𝑑𝑡 = 𝑓 (𝑡) 𝑑𝑡 + 𝑔(𝑡) 𝑑𝑡.
𝑎 𝑎 𝑎
Theorem 6.12. If 𝑓 is a bounded Riemann integrable function on the interval [𝑎, 𝑏] and on the interval
[𝑏, 𝑐], then 𝑓 is Riemann integrable on [𝑎, 𝑐] and
𝑐 𝑏 𝑐
∫ ∫ ∫
𝑓 (𝑥) 𝑑𝑥 = 𝑓 (𝑥) 𝑑𝑥 + 𝑓 (𝑥) 𝑑𝑥.
𝑎 𝑎 𝑏
∑
𝑛
𝑈 (𝑓 , 𝑅) = (𝑥𝑗 − 𝑥𝑗−1 ) sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}
𝑗=1
∑𝑚
∑
𝑛
= (𝑥𝑗 − 𝑥𝑗−1 ) sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]} + (𝑥𝑗 − 𝑥𝑗−1 ) sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]}
𝑗=1 𝑗=𝑚+1
= 𝑈 (𝑓 , 𝑃 ) + 𝑈 (𝑓 , 𝑄).
∫ ∫
= 𝑓 (𝑥) 𝑑𝑥 + 𝑓 (𝑥) 𝑑𝑥.
𝑎 𝑏
Similarly
𝑐
∫ ∫
= 𝑓 (𝑥) 𝑑𝑥 + 𝑓 (𝑥) 𝑑𝑥.
𝑎 𝑏
Since 𝑓 is integrable over [𝑎, 𝑏] and over [𝑏, 𝑐], it follows from above that
𝑐 𝑏 𝑐 𝑐
𝑓 (𝑥) 𝑑𝑥 ≤ 𝑓 (𝑥) 𝑑𝑥 ≤
∫ ∫ ∫ ∫
𝑓 (𝑥) 𝑑𝑥 + 𝑓 (𝑥) 𝑑𝑥.
𝑎 𝑎 𝑏 𝑎
𝑐 𝑏 𝑐
∫ ∫ ∫
𝑓 (𝑥) 𝑑𝑥 = 𝑓 (𝑥) 𝑑𝑥 + 𝑓 (𝑥) 𝑑𝑥.
𝑎 𝑎 𝑏
Lemma 6.13. Let 𝑓 be a continuous function on [𝑎, 𝑏] and let 𝜀 > 0. Then there is a partition = {𝑥0 <
𝑥1 < ⋯ < 𝑥𝑁 } of [𝑎, 𝑏] such that |𝑓 (𝑥) − 𝑓 (𝑥𝑗−1 )| ≤ 𝜀 for all 𝑗 ∈ {1, … , 𝑁} and all 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ].
Proof. Let 𝜀 > 0 and put 𝑥0 = 𝑎. For 𝑗 ∈ ℕ∗ we define numbers 𝑥𝑗 recursively as follows: Let
Claim 2. If 𝑥𝑗−1 < 𝑏, then |𝑓 (𝑥) − 𝑓 (𝑥𝑗−1 )| < 𝜀 for all 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ), |𝑓 (𝑥𝑗 ) − 𝑓 (𝑥𝑗−1 )| ≤ 𝜀, and
|𝑓 (𝑥𝑗 ) − 𝑓 (𝑥𝑗−1 )| = 𝜀 if 𝐸𝑗 ≠ ∅.
Indeed, if 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ), then 𝑥 ∈ [𝑥𝑗−1 , 𝑏] and 𝑥 < inf 𝐸𝑗 , and so 𝑥 ∉ 𝐸𝑗 . By definition of 𝐸𝑗 it follows
that |𝑓 (𝑥) − 𝑓 (𝑥𝑗−1 )| < 𝜀, and, since 𝑥𝑗−1 < 𝑥𝑗 by Claim 1, the continuity of 𝑓 and the absolute value
function give
|𝑓 (𝑥𝑗 ) − 𝑓 (𝑥𝑗−1 )| = lim− |𝑓 (𝑥) − 𝑓 (𝑥𝑗−1 )| ≤ 𝜀.
𝑥→𝑥𝑗
lim 𝑎𝑘 = 𝑥𝑗 .
𝑘→∞
Hence there is 𝑗 such that |𝑓 (𝑥𝑗 ) − 𝑓 (𝑥𝑗−1 )| < 𝜀, contradicting Claim 2 since 𝑥𝑗 ≠ 𝑏 implies 𝐸𝑗 ≠ ∅.
Putting 𝑁 = min{𝑗 ∈ ℕ ∣ 𝑥𝑗 = 𝑏} completes the proof.
Theorem 6.14. If 𝑓 is a continuous function on [𝑎, 𝑏], then 𝑓 is Riemann integrable on [𝑎, 𝑏].
Proof. From Theorem 3.17 we know that a continuous function on a closed interval [𝑎, 𝑏] is bounded.
Let 𝜀 > 0 and let 𝑃 = {𝑥0 < ⋯ < 𝑥𝑁 } be a partition of [𝑎, 𝑏] according to Lemma 6.13. Then, for
𝑗 ∈ {1, … , 𝑁}, with Lemma 6.1,
sup 𝑓 (𝑥) − inf 𝑓 (𝑥) = sup [𝑓 (𝑥) − 𝑓 (𝑥𝑗−1 )] − inf [𝑓 (𝑥) − 𝑓 (𝑥𝑗−1 )]
𝑥∈[𝑥𝑗−1 ,𝑥𝑗 ] 𝑥∈[𝑥𝑗−1 ,𝑥𝑗 ] 𝑥∈[𝑥𝑗−1 ,𝑥𝑗 ] 𝑥∈[𝑥𝑗−1 ,𝑥𝑗 ]
≤ 2𝜀
≤
∑
𝑁
(𝑥𝑗 − 𝑥𝑗−1 )2𝜀 = 2(𝑏 − 𝑎)𝜀.
𝑗=1
Theorem 6.15 (Fundamental Theorem of Calculus). Let 𝑓 be a differentiable function on [𝑎, 𝑏] such that
𝑓 ′ is continuous on [𝑎, 𝑏]. Then 𝑓 ′ is Riemann integrable on [𝑎, 𝑏] and
∫
𝑓 ′ (𝑡) 𝑑𝑡 = 𝑓 (𝑏) − 𝑓 (𝑎).
𝑎
Proof. Since 𝑓 ′ is continuous, 𝑓 ′ is Riemann integrable by the previous theorem. Let 𝑃 = {𝑥0 < ⋯ <
𝑥𝑁 } be a partition of [𝑎, 𝑏]. By the First Mean Value Theorem, Theorem 4.5, for each 𝑗 ∈ {1, … , 𝑁}
there is 𝑐𝑗 ∈ (𝑥𝑗−1 , 𝑥𝑗 ) such that
Then
∑
𝑁
𝐿(𝑓 ′ , 𝑃 ) = (𝑥𝑗 − 𝑥𝑗−1 ) inf 𝑓 ′ (𝑥)
𝑥∈[𝑥𝑗−1 ,𝑥𝑗 ]
𝑗=1
≤
∑
𝑁
(𝑥𝑗 − 𝑥𝑗−1 )𝑓 ′ (𝑐𝑗 )
𝑗=1
∑
𝑁
= 𝑓 (𝑥𝑗 ) − 𝑓 (𝑥𝑗−1 ) = 𝑓 (𝑏) − 𝑓 (𝑎).
𝑗=1
Hence
𝑏
Similarly,
𝑏
Since 𝑓′ is Riemann integrable, the upper and lower Riemann integrals are equal, and
𝑏
∫
𝑓 ′ (𝑥) 𝑑𝑥 = 𝑓 (𝑏) − 𝑓 (𝑎)
𝑎
follows.
Tutorial 6.5.1. Let 𝑓 be a bounded Riemann integrable function on [𝑎, 𝑏] and let 𝑃 be a partition of
[𝑎, 𝑏]. Then prove that
𝑏
(𝑏 − 𝑎) inf {𝑓 (𝑥) ∣ 𝑥 ∈ [𝑎, 𝑏]} ≤ 𝐿(𝑓 , 𝑃 ) ≤ 𝑓 (𝑡) 𝑑𝑡 ≤ 𝑈 (𝑓 , 𝑃 ) ≤ (𝑏 − 𝑎) sup{𝑓 (𝑥) ∣ 𝑥 ∈ [𝑎, 𝑏]}.
∫
𝑎
*Tutorial 6.5.2. Let 𝑎 < 𝑏, let 𝑓 ∶ [𝑎, 𝑏] → ℝ and let 𝑃 = {𝑎 = 𝑥0 < 𝑥1 < ⋯ < 𝑥𝑛 = 𝑏} ∈ (𝑎, 𝑏).
Define
∑
𝑛
𝑉 (𝑓 , 𝑃 ) ∶= |𝑓 (𝑥𝑗 ) − 𝑓 (𝑥𝑗−1 )|
𝑗=1
One sets 𝑉𝑎𝑎 𝑓 = 0. The function 𝑓 is said to be of bounded variation on [𝑎, 𝑏] if 𝑉𝑎𝑏 𝑓 < ∞.
1. Let 𝑎 ≤ 𝑥 < 𝑦 ≤ 𝑏. Show that 𝑉𝑎𝑥 𝑓 + |𝑓 (𝑦) − 𝑓 (𝑥)| ≤ 𝑉𝑎𝑦 𝑓 .
2. Assume that 𝑓 is of bounded variation on [𝑎, 𝑏] and for 𝑥 ∈ [𝑎, 𝑏] define
𝑓1 (𝑥) = 𝑉𝑎𝑥 𝑓 ,
𝑓2 (𝑥) = 𝑉𝑎𝑥 𝑓 − 𝑓 (𝑥).
Chapter 7
(i) 𝑑(𝑥, 𝑦) = 0 ⇔ 𝑥 = 𝑦,
Proof. From
0 = 𝑑(𝑥, 𝑥) ≤ 𝑑(𝑥, 𝑦) + 𝑑(𝑦, 𝑥) = 𝑑(𝑥, 𝑦) + 𝑑(𝑥, 𝑦) = 2𝑑(𝑥, 𝑦)
(i) (iii) (ii)
Worked Example 7.1.1. 1. For 𝑥, 𝑦 ∈ ℝ let 𝑑(𝑥, 𝑦) = |𝑥 − 𝑦|. Then (ℝ, 𝑑) is a metric space.
2. For 𝑥, 𝑦 ∈ ℂ let 𝑑(𝑥, 𝑦) = |𝑥 − 𝑦|. Then (ℂ, 𝑑) is a metric space.
( 𝑛 )1
∑ 2
3. For 𝑥 = (𝑥𝑗 )𝑛𝑗=1 , 𝑦 = (𝑦𝑗 )𝑛𝑗=1 ∈ ℂ𝑛 let 𝑑(𝑥, 𝑦) = |𝑥𝑗 − 𝑦𝑗 |2 . Then (ℂ𝑛 , 𝑑) is a metric space.
𝑗=1
Solution. Parts 1 and 2 are clear if one observes the triangle inequality for real and complex numbers.
For the triangle inequality in 3, we use the notation
Then ‖𝑥 + 𝑦‖ ≤ ‖𝑥‖ + ‖𝑦‖ can be shown as was done for ℝ𝑛 in Multivariable Calculus. But rather than
̂ = (|𝑥𝑗 |)𝑛𝑗=1 ∈ ℝ𝑛 . Obviously,
repeating that proof we use it to prove the result for ℂ𝑛 . To this end, put 𝑥
𝑥‖ = ‖𝑥‖ and ‖𝑥 + 𝑦‖ ≤ ‖̂
‖̂ 𝑥 + 𝑦̂‖ by part 2. Then, from Multivarible Calculus results applicable to 𝑥 ̂
and 𝑦̂, it follows that
‖𝑥 + 𝑦‖ ≤ ‖̂ 𝑥 + 𝑦̂‖ ≤ ‖̂ 𝑥‖ + ‖̂𝑦‖ = ‖𝑥‖ + ‖𝑦‖.
67
Then, for 𝑥, 𝑦, 𝑧 ∈ ℂ𝑛 ,
𝑑(𝑥, 𝑧) = ‖𝑥 − 𝑧‖ = ‖(𝑥 − 𝑦) + (𝑦 − 𝑧)‖ ≤ ‖𝑥 − 𝑦‖ + ‖𝑦 − 𝑧‖
= 𝑑(𝑥, 𝑦) + 𝑑(𝑦, 𝑧).
Worked Example 7.1.2. Let 𝑌 be a nonempty set and let 𝐵(𝑌 ) = 𝐵(𝑌 , 𝔽 𝑛 ) be the set of all bounded
functions from 𝑌 to 𝔽 𝑛 . Here a function 𝑓 from 𝑌 to 𝔽 𝑛 is called bounded if there is a real number 𝑀
( 𝑛 )1
≤ 𝑀 for all 𝑥 ∈ 𝑌 , where 𝑓𝑗 (𝑥) is the 𝑗-th component of 𝑓 (𝑥).
∑ 2
Solution. In class.
Proposition 7.1. Let 𝑋 = (𝑋, 𝑑) be a metric space and let 𝑌 ⊂ 𝑋 be nonempty. For 𝑥, 𝑦 ∈ 𝑌 let
𝑑𝑌 (𝑥, 𝑦) = 𝑑(𝑥, 𝑦). Then 𝑌 = (𝑌 , 𝑑𝑌 ) = (𝑌 , 𝑑) is a metric space.
Note. When it is clear what the metric is, it is convenient to write briefly 𝑋 for a metric space, and
𝑌 ⊂ 𝑋 is called a (metric) subspace of 𝑋.
Theorem 7.2. Let (𝑋, 𝑑) be a metric space and let (𝑥𝑘 ) be a convergent sequence in 𝑋. Then
1. lim 𝑥𝑘 is unique.
𝑘→∞
2. (𝑥𝑘 ) is a Cauchy sequence.
Proof. 1. Let 𝑥, 𝑦 ∈ 𝑋 such that lim 𝑑(𝑥𝑘 , 𝑥) = 0 and lim 𝑑(𝑥𝑘 , 𝑦) = 0. Then
𝑘→∞ 𝑘→∞
⎧ arctan 𝑥 if 𝑥 ∈ ℝ,
⎪
⎪ 𝜋
𝑓 (𝑥) = ⎨ − if 𝑥 = −∞ ,
2
⎪𝜋
⎪ if 𝑥 = ∞,
⎩2
and let 𝑑 be as in part 1. Show that a sequence (𝑥𝑘 ) in ℝ converges in (ℝ, 𝑑) if and only if either (𝑥𝑘 )
converges in ℝ in the usual sense or 𝑥𝑘 → −∞ or 𝑥𝑘 → ∞.
3. Let 𝑋 be a nonempty set. For 𝑥, 𝑦 ∈ 𝑋 define
{
0 if 𝑥 = 𝑦,
1 if 𝑥 ≠ 𝑦.
𝑑(𝑥, 𝑦) =
a) Show that (𝑋, 𝑑) is a metric space. (𝑋, 𝑑) is called a discrete metric space.
b) Show that a sequence (𝑥𝑘 ) converges in (𝑋, 𝑑) if and only if there is 𝑁 ∈ ℕ such that 𝑥𝑘 = 𝑥𝑁 for all
𝑘 > 𝑁.
4. Let 𝑥, 𝑦 ∈ ℂ𝑛 . Prove directly that |(𝑥, 𝑦)| ≤ ‖𝑥‖‖𝑦‖ and ‖𝑥 + 𝑦‖ ≤ ‖𝑥‖ + ‖𝑦‖.
Proof. Let (𝑧𝑘 ) be a Cauchy sequence in ℂ𝑛 . Write 𝑧𝑘 = (𝑥𝑘,𝑗 + 𝑖𝑦𝑘,𝑗 )𝑛𝑗=1 with 𝑥𝑘,𝑗 , 𝑦𝑘,𝑗 ∈ ℝ. Then
Thus (𝑥𝑘,𝑗 )𝑛𝑗=1 is a Cauchy sequence in ℝ for each 𝑗 = 1, … , 𝑛. By Theorem 2.12, the sequence converges,
so that there is 𝑎𝑗 ∈ ℝ such that
lim 𝑥𝑘,𝑗 = 𝑎𝑗 .
𝑘→∞
Put
𝑧 = (𝑎𝑗 + 𝑖𝑏𝑗 )𝑛𝑗=1 .
Let 𝜀 > 0. Then there is 𝐾 ∈ ℝ such that for all 𝑘 ≥ 𝐾 and all 𝑗 = 1, … , 𝑛,
𝜀 𝜀
|𝑥𝑘,𝑗 − 𝑎𝑗 | < √ , |𝑦𝑘,𝑗 − 𝑏𝑗 | < √ .
2𝑛 2𝑛
Hence, for 𝑘 ≥ 𝐾,
( 𝑛 )1
∑( ) 2
𝑥∈𝑌 𝑥∈𝑌
Therefore 𝑓 is bounded.
Finally, for 𝑥 ∈ 𝑌 choose 𝑚 ≥ 𝑛 𝜀 such that
3
𝜀
‖𝑓𝑚 (𝑥) − 𝑓 (𝑥)‖ < .
4
Then, for 𝑘 ≥ 𝑛 𝜀 ,
4
≤ 𝜀
1
2
< 𝜀.
This proves that (𝑓𝑘 ) converges in 𝐵(𝑌 ). Since (𝑓𝑘 ) was an arbitrary Cauchy sequence in 𝐵(𝑌 ), it follows
that 𝐵(𝑌 ) is complete.
Definition 7.4. 1. A subset 𝐴 of a metric space (𝑋, 𝑑) is called open if for each 𝑥 ∈ 𝐴 there is 𝜀 > 0
such that 𝐵(𝑥, 𝜀) ⊂ 𝐴, where 𝐵(𝑥, 𝜀) = {𝑦 ∈ 𝑋 ∶ 𝑑(𝑦, 𝑥) < 𝜀} is the 𝜀-ball with centre 𝑥.
2. A subset 𝐴 of a metric space (𝑋, 𝑑) is called closed if 𝑋 ⧵ 𝐴 is open.
Theorem 7.5. A subset 𝐴 of a metric space (𝑋, 𝑑) is closed if and only if for each sequence (𝑥𝑘 ) in 𝐴
which converges to some 𝑥 ∈ 𝑋 it follows that 𝑥 ∈ 𝐴.
Proof. Let 𝐴 be closed and let (𝑥𝑘 ) be a sequence in 𝐴 which converges to some 𝑥 ∈ 𝑋. We must show
𝑥 ∈ 𝐴. Assume that 𝑥 ∉ 𝐴. Since 𝑥 ∈ 𝑋 ⧵ 𝐴 and 𝑋 ⧵ 𝐴 is open, there is 𝜀 > 0 such that 𝐵(𝑥, 𝜀) ⊂ 𝑋 ⧵ 𝐴.
Since 𝑥𝑘 ∈ 𝐴 for all 𝑘 ∈ ℕ, it follows that 𝑑(𝑥𝑘 , 𝑥) ≥ 𝜀, which contradicts 𝑥𝑘 → 𝑥.
Conversely, if 𝐴 ⊂ 𝑋 is not closed, then there is 𝑥 ∈ 𝑋 ⧵ 𝐴 such that 𝐵(𝑥, 𝜀) ⊄ 𝑋 ⧵ 𝐴 for all 𝜀 > 0.
Therefore, for positive integers 𝑘, there is 𝑥𝑘 ∈ 𝐴 ∩ 𝐵(𝑥, 𝑘1 ). Thus, for each 𝜀 > 0 and 𝑘 ≥ 1𝜀 we have
𝑑(𝑥𝑘 , 𝑥) < 𝑘1 ≤ 𝜀. Thus we have found a sequence (𝑥𝑘 ) in 𝐴 which converges to some 𝑥 ∈ 𝑋 which does
not belong to 𝐴.
Theorem 7.6. Let (𝑋, 𝑑) be a complete metric space and let 𝐴 ⊂ 𝑋. Then 𝐴 is complete if and only if
𝐴 is closed.
Proof. Assume that 𝐴 is complete and let (𝑥𝑘 ) be a sequence in 𝐴 which converges to some 𝑥 ∈ 𝑋.
We must show that 𝑥 ∈ 𝐴. But since (𝑥𝑘 ) is a convergent sequence in 𝑋, it is a Cauchy sequence in 𝑋
by Theorem 7.2, 2, and therefore also in 𝐴. Since 𝐴 is complete, (𝑥𝑘 ) converges to some 𝑎 ∈ 𝐴, and
therefore also in 𝑋. Since limits are unique by Theorem 7.2, 1, it follows that 𝑥 = 𝑎 ∈ 𝐴. Hence, by
Theorem 7.5, 𝐴 is closed.
Conversely, assume that 𝐴 is closed. Let (𝑥𝑘 ) be a Cauchy sequence in 𝐴. Then (𝑥𝑘 ) is a Cauchy sequence
in 𝑋, and since 𝑋 is complete, there is 𝑥 ∈ 𝑋 such that 𝑥𝑘 → 𝑥. Since 𝐴 is closed, it follows that 𝑥 ∈ 𝐴.
Hence each Cauchy sequence in 𝐴 converges in 𝐴. It follows that 𝐴 is complete.
Definition 7.5. Let 𝐴 ⊂ ℝ and 𝑓 ∶ 𝐴 → 𝔽 𝑛 . Then 𝑓 is uniformly continuous if for each 𝜀 > 0 there is
𝛿 > 0 such that for all 𝑥, 𝑦 ∈ 𝐴 (|𝑥 − 𝑦| < 𝛿 ⇒ ‖𝑓 (𝑥) − 𝑓 (𝑦)‖ < 𝜀).
Theorem 7.7. Let 𝑎, 𝑏 ∈ ℝ, 𝑎 < 𝑏, and let 𝑓 ∶ [𝑎, 𝑏] → 𝔽 𝑛 be continuous. Then 𝑓 is uniformly
continuous.
Proof. Because of Remark 7.2.1 we only have to consider 𝑓 ∶ [𝑎, 𝑏] → ℝ. Let 𝜀 > 0 and choose
𝑎 = 𝑥0 < 𝑥1 < ⋯ < 𝑥𝑁 = 𝑏 according to Lemma 6.13 with respect to 4𝜀 . Put
Now let 𝑥, 𝑦 ∈ [𝑎, 𝑏] such that |𝑥 − 𝑦| < 𝛿. We want to show |𝑓 (𝑥) − 𝑓 (𝑦)| < 𝜀. This is trivial if 𝑥 = 𝑦.
Thus let 𝑥 ≠ 𝑦, and we may assume 𝑥 < 𝑦.
There is 𝑗 ∈ {1, … , 𝑁} such that 𝑥 ∈ [𝑥𝑗−1 , 𝑥𝑗 ]. Hence
Case II: 𝑦 > 𝑥𝑗 . Then 𝑗 < 𝑁, and 𝑥𝑗 < 𝑦 < 𝑥𝑗 + 𝛿 ≤ 𝑥𝑗+1 . Again by Lemma 6.13,
Theorem 7.8. Let 𝑎, 𝑏 ∈ ℝ and 𝑛 ∈ ℕ ⧵ {0}. Then 𝐶([𝑎, 𝑏], 𝔽 𝑛 ) is a closed subspaces of 𝐵([𝑎, 𝑏], 𝔽 𝑛 ).
In particular, 𝐶([𝑎, 𝑏], 𝔽 𝑛 ) is complete.
Proof. By Remark 7.2.1, it suffices to consider 𝐶([𝑎, 𝑏], ℝ). By Theorem 3.17, each continuous function
on [𝑎, 𝑏] is bounded, so that 𝐶([𝑎, 𝑏], ℝ) ⊂ 𝐵([𝑎, 𝑏], ℝ).
To show that 𝐶([𝑎, 𝑏], ℝ) is closed in 𝐵([𝑎, 𝑏], ℝ) let (𝑓𝑘 ) be a sequence in 𝐶([𝑎, 𝑏], ℝ) which converges
to some 𝑓 in 𝐵([𝑎, 𝑏], ℝ). We must show that 𝑓 is continuous.
𝜀
Let 𝜀 > 0. Then there is 𝑘 ∈ ℕ such that 𝑑(𝑓𝑘 , 𝑓 ) < . Since 𝑓𝑘 is uniformly continuous, there is 𝛿 > 0
3
such that
𝜀
|𝑓𝑘 (𝑥) − 𝑓𝑘 (𝑦)| < for 𝑥, 𝑦 ∈ [𝑎, 𝑏], |𝑥 − 𝑦| < 𝛿.
3
Then, for 𝑥, 𝑦 ∈ [𝑎, 𝑏] and |𝑥 − 𝑦| < 𝛿,
|𝑓 (𝑥) − 𝑓 (𝑦)| ≤ |𝑓 (𝑥) − 𝑓𝑘 (𝑥)| + |𝑓𝑘 (𝑥) − 𝑓𝑘 (𝑦)| + |𝑓𝑘 (𝑦) − 𝑓 (𝑦)|
≤ 𝑑(𝑓 , 𝑓𝑘 ) + |𝑓𝑘 (𝑥) − 𝑓𝑘 (𝑦)| + 𝑑(𝑓𝑘 , 𝑓 )
𝜀 𝜀 𝜀
< + +
3 3 3
= 𝜀.
Tutorial 7.2.1. 1. Let (𝑋, 𝑑) be the metric space from Tutorial 7.1.1, 1. Show that (𝑋, 𝑑) is complete if
and only if range 𝑓 is closed in ℝ.
2. Show that the metric space (ℝ, 𝑑) from Tutorial 7.1.1, 2, is complete.
3. Show that ℝ is not complete with respect to the metric given in Tutorial 7.1.1, 2.
Since it is clear that 𝑑𝑌 (𝑇 𝑥, 𝑇 𝑦) ≤ (‖𝑇 ‖ + 𝜀)𝑑𝑋 (𝑥, 𝑦) for all 𝜀 > 0, it follows that
𝑑𝑌 (𝑇 𝑥, 𝑇 𝑦)
‖𝑇 ‖ = sup .
𝑥,𝑦∈𝑋 𝑑𝑋 (𝑥, 𝑦)
𝑥≠𝑦
Proof. Let
{ }
∶ 𝑥, 𝑦 ∈ 𝑋, 𝑥 ≠ 𝑦
𝑑𝑌 (𝑇 𝑥, 𝑇 𝑦)
𝐴= , 𝛼 = sup 𝐴.
𝑑𝑋 (𝑥, 𝑦)
all 𝑥, 𝑦 ∈ 𝑋. Hence 𝛼 ≥ ‖𝑇 ‖.
𝑑𝑋 (𝑥, 𝑦)
𝑑𝑍 ((𝑆𝑇 )𝑥, (𝑆𝑇 )𝑦) = 𝑑𝑍 (𝑆(𝑇 𝑥), 𝑆(𝑇 𝑦)) ≤ ‖𝑆‖𝑑𝑌 (𝑇 𝑥, 𝑇 𝑦) ≤ ‖𝑆‖ ‖𝑇 ‖𝑑𝑋 (𝑥, 𝑦).
∑
∞
‖𝑇 𝑗 ‖ < ∞.
𝑗=0
Proof. If the sum converges, then ‖𝑇 𝑗 ‖ → 0 as 𝑗 → ∞, and therefore there is 𝑚 ∈ ℕ such that ‖𝑇 𝑚 ‖ < 1.
Conversely, if ‖𝑇 𝑚 ‖ < 1, then we group the sum into blocks of 𝑚 terms, writing every natural number
as a multiple of 𝑚 plus its remainder, use
‖𝑇 𝑘𝑚+𝑙 ‖ = ‖𝑇 𝑘𝑚 𝑇 𝑙 ‖ ≤ ‖(𝑇 𝑚 )𝑘 ‖ ‖𝑇 𝑙 ‖ ≤ ‖𝑇 𝑚 ‖𝑘 ‖𝑇 𝑙 ‖
and get
∑
∞
∑ ∑
∞ 𝑚−1
‖𝑇 𝑗 ‖ = ‖𝑇 𝑘𝑚+𝑙 ‖
𝑗=0 𝑘=0 𝑙=0
≤
∑ ∑
∞ 𝑚−1
‖𝑇 𝑚 ‖𝑘 ‖𝑇 𝑙 ‖
𝑘=0 𝑙=0
( ) (𝑚−1 )
∞
∑ ∑
= ‖𝑇 𝑚 ‖𝑘 ‖𝑇 𝑙 ‖
𝑘=0 𝑙=0
1 ∑ 𝑚−1
= ‖𝑇 𝑙 ‖ < ∞.
1 − ‖𝑇 𝑚 ‖ 𝑙=0
Theorem 7.11 (Banach’s Fixed Point Theorem). Let (𝑋, 𝑑) be a complete metric space and let 𝑇 ∶ 𝑋 →
𝑋 be a generalized contraction. Then there is a unique 𝑎 ∈ 𝑋 such that 𝑇 𝑎 = 𝑎.
𝑇 𝑘 𝑎 = 𝑇 𝑇 𝑘−1 𝑎 = 𝑇 𝑎 = 𝑎 and 𝑇 𝑘 𝑏 = 𝑏,
Since
∑
∞
‖𝑇 𝑙 ‖ < ∞,
𝑙=0
it follows that (𝑇 𝑘 𝑥0 ) is a Cauchy sequence. Since 𝑋 is complete, this Cauchy sequence has a limit,
say 𝑎. Finally, to show that 𝑇 𝑎 = 𝑎, let 𝜀 > 0. Then there is 𝐾 such that for 𝑗 ≥ 𝐾,
𝜀
𝑑(𝑇 𝑗 𝑥0 , 𝑎) < .
1 + ‖𝑇 ‖
For 𝑘 ≥ 𝐾 + 1 we get
Tutorial 7.3.1. Let (𝑋, 𝑑) be a complete metric space and let 𝑇 be a generalized contraction on 𝑋. Let
𝑚 ∈ ℕ∗ such that ‖𝑇 𝑚 ‖ < 1 and let 𝑥 be the fixed point of 𝑇 . Prove the following a priori estimate: For
any 𝑥0 ∈ 𝑋,
𝑑(𝑥, 𝑥0 ) ≤
𝑑(𝑇 𝑥0 , 𝑥0 ) ∑ 𝑗
𝑚−1
‖𝑇 ‖.
1 − ‖𝑇 𝑚 ‖ 𝑗=0
Hint. Use that 𝑇 𝑚 𝑥 = 𝑥 and use a "telescoping" sum of distances involving the points 𝑇 𝑚 𝑥, 𝑇 𝑚 𝑥0 ,
𝑇 𝑚−1 𝑥0 , … , 𝑇 𝑥0 .
𝑓 (𝜉)𝑓 ′′ (𝜉)
𝑇 𝑥1 − 𝑇 𝑥2 = (𝑥1 − 𝑥2 ) .
(𝑓 ′ (𝜉))2
2. Show that ‖𝑇 ‖ < 1 if 𝑐 and 𝑑 are sufficiently close to the zero 𝑦 of 𝑓 .
3. Show that one can choose 𝑐 and 𝑑 in such a way that 𝑇 ([𝑐, 𝑑]) ⊂ [𝑐, 𝑑].
Hint. You may show that 𝑇 ([𝑐, 𝑑]) ⊂ [𝑐, 𝑑] if 𝑑 − 𝑦 = 𝑦 − 𝑐, i. e., 𝑦 is the midpoint of the interval [𝑐, 𝑑].
However, this is not very practical since you do not know 𝑦 (otherwise, the whole procedure would be
pointless). The easiest way out of this is to choose suitable values (by trial and error), 𝑐1 < 𝑑1 in [𝑐, 𝑑]
such that 𝑓 (𝑐1 ) and 𝑓 (𝑐2 ) have opposite sign, 𝑇 𝑐1 , 𝑇 𝑐2 ∈ [𝑐1 , 𝑐2 ], and 𝑓 𝑓 ′′ has no zeros on [𝑐1 , 𝑑1 ] ⧵ {𝑦}.
∗ 4. Assume that statement 2 holds and define
𝑐 + ‖𝑇 ‖𝑑 𝑑 + ‖𝑇 ‖𝑐
𝑐1 = , 𝑑1 = .
1 + ‖𝑇 ‖ 1 + ‖𝑇 ‖
Show that 𝑐 < 𝑐1 < 𝑦 < 𝑑1 < 𝑑 if 𝑓 (𝑐1 ) and 𝑓 (𝑑1 ) have opposite signs. Then show that 𝑇 ([𝑐, 𝑑]) ⊂ [𝑐, 𝑑].
Hence show that for any 𝑥 ∈ [𝑐, 𝑑], 𝑇 𝑘 𝑥 → 𝑦 as 𝑘 → ∞.
√ √ √
Tutorial 7.3.3. Let 𝑐 > 0. Show that 𝑐 + 𝑐 + 𝑐 + … converges by using the following proof.
√ [ ) [ )
(a) Let 𝑇𝑐 𝑥 = 𝑐 + 𝑥, 𝑥 ∈ [0, ∞). Show that 𝑇𝑐 𝑥 > 𝑥 for 𝑥 ∈ [0, 1] and that 𝑇𝑐 ∶ 41 , ∞ → 14 , ∞ is
Show that if 𝑥𝑘 ≤ 1 for all 𝑘 ∈ ℕ, then (𝑥𝑘 ) is (strictly) increasing. (Actually, this never happens.)
a contraction.
𝑦′ = 𝑓 (𝑥, 𝑦) (7.4.1)
is called a first-order system of differential equations, and any continuously differentiable function
𝑦 ∶ 𝐼 → 𝑈 satisfying (7.4.1) is called a solution of (7.4.1).
Let 𝑎 ∈ 𝐼 and 𝑏 ∈ 𝔽 𝑛 . Then a solution of (7.4.1) satisfying
𝑦(𝑎) = 𝑏 (7.4.2)
Lemma 7.12. 𝑦 is a a solution of the IVP (7.4.1), (7.4.2) if and only if 𝑦 ∈ 𝐶(𝐼, 𝔽 𝑛 ), 𝑦(𝐼) ⊂ 𝑈 , and
∫
𝑦(𝑥) = 𝑏 + 𝑓 (𝑡, 𝑦(𝑡)) 𝑑𝑡 (𝑥 ∈ 𝐼). (7.4.3)
𝑎
Proof. We only have to note that we can apply the Fundamental Theorem of Calculus componentwise and
to real and imaginary parts separately. Also, if 𝑦 ∈ 𝐶(𝐼, 𝔽 𝑛 ) satisfies (7.4.3), then 𝑓 is differentiable.
We are now going to show that under certain conditions, (7.4.3) and thus the IVP (7.4.1), (7.4.2) has a
unique solution. We write 𝐶 1 (𝐼, 𝔽 𝑛 ) for the set of all continuously differentiable functions from 𝐼 to 𝔽 𝑛 .
Proof. Since 𝑈 is a closed subset of 𝔽 𝑛 , 𝐶(𝐼, 𝑈 ) is a complete metric space. For 𝑦 ∈ 𝐶(𝐼, 𝑈 ) define
∫
(𝑇 𝑦)(𝑥) = 𝑏 + 𝑓 (𝑡, 𝑦(𝑡)) 𝑑𝑡 (𝑥 ∈ 𝐼). (7.4.4)
𝑎
By Lemma 7.12, the theorem is proved if we show that 𝑇 maps 𝐶(𝐼, 𝑈 ) into 𝐶(𝐼, 𝑈 ) and that 𝑇 has
exactly one fixed point.
For the induction step assume that (7.4.5) is true for 𝑘. Then
‖ 𝑥 ‖
‖ ] ‖
‖ 𝑘+1 ‖ ‖ [ ‖
‖ ‖∫
𝑘+1 𝑘 𝑘
‖(𝑇 𝑦1 )(𝑥) − (𝑇 𝑦2 )(𝑥)‖ = ‖ 𝑓 (𝑡, (𝑇 𝑦1 )(𝑡)) − 𝑓 (𝑡, (𝑇 𝑦2 )(𝑡)) 𝑑𝑡‖
‖ ‖
‖𝑎 ‖
‖ ‖
| 𝑥 |
| |
≤ | ‖𝑓 (𝑡, (𝑇 𝑦1 )(𝑡)) − 𝑓 (𝑡, (𝑇 𝑦2 )(𝑡))‖ 𝑑𝑡|
| ‖ ‖ |
|∫ ‖
𝑘 𝑘
‖ |
|𝑎 |
| |
| 𝑥 |
| |
≤ | 𝐿 ‖(𝑇 𝑘 𝑦1 )(𝑡) − (𝑇 𝑘 𝑦2 )(𝑡)‖ 𝑑𝑡| (by assumption (iii))
| ‖ ‖ |
|∫ ‖ ‖ |
|𝑎 |
| |
| 𝑥 |
| 𝐿𝑘 |𝑡 − 𝑎|𝑘 ||
≤| 𝐿
|
|∫
𝑑𝑡| 𝑑(𝑦1 , 𝑦2 ) (by induction hypothesis)
𝑘! |
|𝑎 |
| |
𝐿𝑘+1 |𝑥 − 𝑎|𝑘+1
= 𝑑(𝑦1 , 𝑦2 ).
(𝑘 + 1)!
Taking the maximum over all 𝑥 ∈ 𝐼 in (7.4.5) gives
𝑑(𝑇 𝑘 𝑦1 , 𝑇 𝑘 𝑦2 ) ≤
𝐿 𝑘 𝐴𝑘
𝑑(𝑦1 , 𝑦2 ),
𝑘!
and therefore
‖𝑇 𝑘 ‖ ≤
𝐿𝑘 𝐴 𝑘
,
𝑘!
which leads to
‖𝑇 ‖ ≤
∑
∞
∑
∞
(𝐿𝐴)𝑘
𝑘
= 𝑒𝐿𝐴 < ∞.
𝑘=0 𝑘=0
𝑘!
Hence 𝑇 is a generalized contraction on 𝐶(𝐼, 𝑈 ). An application of Banach’s Fixed Point Theorem
completes the proof.
so that
|𝑓 (𝑥, 𝑦1 ) − 𝑓 (𝑥, 𝑦2 )| = |𝑦1 + 𝑦2 | |𝑦1 − 𝑦2 |.
Hence the Lipschitz condition is satisfied if and only if
𝑐(1 + 𝐵 2 ) ≤ 𝐵 𝑐≤
𝐵
or .
1 + 𝐵2
But
2𝐵 (1 + 𝐵 2 ) − (1 − 𝐵)2 (1 − 𝐵)2
= = 1 −
1 + 𝐵2 1 + 𝐵2 1 + 𝐵2
1
has a maximum of 1 for 𝐵 = 1, so that the optimal 𝑐 is 𝑐 = .
2 ( )
𝜋 𝜋
b) The solution of the IVP is 𝑦(𝑥) = tan 𝑥. So the maximal interval where a solution exists is − , .
2 2
Let 𝑀𝑛 (𝔽 ) be the set of 𝑛 × 𝑛 matrices with values in 𝔽 . Then the norm ‖𝑀‖ of 𝑀 ∈ 𝑀𝑛 (𝔽 ) is the norm
of 𝑀 ∶ 𝔽 𝑛 → 𝔽 𝑛 . It is easy to see that
Corollary 7.14. Let 𝐼 be an interval, 𝑎 ∈ 𝐼, 𝑏 ∈ 𝔽 𝑛 , and consider the linear system of differential
equations
𝑦′ = 𝐹 (𝑥)𝑦 + 𝑔(𝑥),
where 𝐹 ∶ 𝐼 → 𝑀𝑛 (𝔽 ) and 𝑔 ∶ 𝐼 → 𝔽 𝑛 are continuous.
Then the initial value problem
𝑦′ = 𝐹 (𝑥)𝑦 + 𝑔(𝑥),
has a unique solution 𝑦 ∈ 𝐶 1 (𝐼, 𝔽 𝑛 ).
Hence the Lipschitz condition is satisfied with 𝐵 = ∞. Now apply Theorem 7.13.
If 𝐼 is arbitrary, we can choose a decreasing sequence (𝑐𝑘 ) and an increasing sequence (𝑑𝑘 ) such that
𝑐1 ≤ 𝑎 ≤ 𝑑1 and
⋃∞
𝐼= [𝑐𝑘 , 𝑑𝑘 ].
𝑘=1
[If, e. g., the right endpoint 𝑑 of 𝐼 belongs to 𝐼, choose 𝑑𝑘 = 𝑑.]
By the first part of the proof, the IVP has a unique solution 𝑦𝑘 on [𝑐𝑘 , 𝑑𝑘 ]. Therefore, 𝑦𝑘+𝑚 (𝑥) = 𝑦𝑘 (𝑥)
for all 𝑚 > 0 and 𝑥 ∈ [𝑐𝑘 , 𝑑𝑘 ], so that 𝑦(𝑥) = 𝑦𝑘 (𝑥) for 𝑥 ∈ [𝑐𝑘 , 𝑑𝑘 ] is well-defined and solves the IVP.
Because this solution is unique on [𝑐𝑘 , 𝑑𝑘 ] for each 𝑘, 𝑦 is unique.
⎛ 𝜂 ⎞ ⎛ 𝑦2 ⎞
⎜ 𝜂′ ⎟ ⎜ .. ⎟
𝑦 = ⎜ . ⎟, 𝐹 (𝑥, 𝑦) = ⎜ . ⎟.
⎜ .. ⎟ ⎜ 𝑦 ⎟
⎜ (𝑛−1) ⎟ ⎜ 𝑛 ⎟
⎝𝜂 ⎠ ⎝ 𝑓 (𝑥, 𝑦) ⎠
Then
𝜂 (𝑛) = 𝑓 (𝑥, 𝜂, 𝜂 ′ , … , 𝜂 (𝑛−1) ) ⇔ 𝑦′ = 𝐹 (𝑥, 𝑦).
Let 𝑏0 , … , 𝑏𝑛−1 ∈ 𝔽 . Then
where
⎛ 𝑏0 ⎞
⎜ ⎟
𝑏 = ⎜ ... ⎟.
⎜ ⎟
⎝ 𝑏𝑛−1 ⎠
Clearly,
and
Corollary 7.16. Let 𝐼 be an interval, 𝑎 ∈ 𝐼, 𝑏0 , … , 𝑏𝑛−1 ∈ 𝔽 , and consider the linear differential
equation
𝜂 (𝑛) = 𝑓𝑛−1 (𝑥)𝜂 (𝑛−1) + 𝑓𝑛−2 (𝑥)𝑦(𝑛−2) + … 𝑓1 (𝑥)𝑦′ + 𝑓0 (𝑥)𝑦 + 𝑔(𝑥),
with initial conditions
𝑦(𝑗) (𝑎) = 𝑏𝑗 , (𝑗 = 0, … , 𝑛 − 1),
where 𝑓𝑗 ∶ 𝐼 → 𝔽 (𝑗 = 0, … , 𝑛 − 1) and 𝑔 ∶ 𝐼 → 𝔽 are continuous.
Then the initial value problem has a unique solution 𝜂 ∈ 𝐶 𝑛 (𝐼, 𝔽 ).
2. Let 𝑀1 , 𝑀2 ∈ 𝑀𝑛 (𝔽 ).
a) Prove that ‖𝑀1 + 𝑀2 ‖ ≤ ‖𝑀1 ‖ + ‖𝑀2 ‖.
b) Let 𝑑(𝑀1 , 𝑀2 ) = ‖𝑀1 − 𝑀2 ‖. Prove that 𝑑 is a metric on 𝑀𝑛 (𝔽 ).
3. Prove that 𝑀𝑛 (𝔽 ) is complete.
∑
∞
Definition 7.8. Let (𝑎𝑗 ) be a sequence in 𝐵. Then 𝑎𝑗 is called a series in 𝐵.
𝑗=0
∑
𝑘
The series is said to converge in 𝐵 if the sequence of its partial sums 𝑎𝑗 converges in 𝐵 as 𝑘 → ∞.
𝑗=0
∑
∞
The series is said to converge absolutely if ‖𝑎𝑗 ‖ < ∞.
𝑗=0
∑
∞
Theorem 7.17. Let 𝑎𝑗 be a series in 𝐵.
𝑗=0
𝑎𝑗 converges in 𝐵 if and only if for all 𝜀 > 0 there is 𝐾 ∈ ℝ such that for all 𝑘 ≥ 𝑚 ≥ 𝐾,
∑
∞
1. The series
𝑗=0
‖∑ ‖
‖ 𝑘 ‖
‖ 𝑎𝑗 ‖
‖ ‖ < ∞.
‖𝑗=𝑚 ‖
‖ ‖
∑
∞
2. If 𝑎𝑗 converges absolutely, then it converges in 𝐵.
𝑗=0
∑
∞
√
Theorem 7.18. Let 𝑎𝑗 be a series in 𝐵, assume that 𝜌 = lim sup 𝑗
‖𝑎𝑗 ‖ < ∞ and put 𝑅 = 𝜌1 . Let
𝑗=0 𝑗→∞
𝑎, 𝑥 ∈ ℝ. Then
1. The series
∑
∞
𝑎𝑗 (𝑥 − 𝑎)𝑗
𝑗=0
∑
∞
𝑗𝑎𝑗 (𝑥 − 𝑎)𝑗−1
𝑗=1
∑
∞
𝑓 (𝑥) = 𝑎𝑗 (𝑥 − 𝑎)𝑗 ,
𝑗=0
∑
∞
𝑔(𝑥) = 𝑗𝑎𝑗 (𝑥 − 𝑎)𝑗−1 .
𝑗=1
∑
𝑘
[ ]
𝑓 (𝑥 + ℎ) − 𝑓 (𝑥) − 𝑔(𝑥)ℎ = lim 𝑎𝑗 (𝑥 + ℎ − 𝑎)𝑗 − (𝑥 − 𝑎)𝑗 − 𝑗(𝑥 − 𝑎)𝑗−1 ℎ
𝑘→∞
𝑗=1
∑
𝑘
[ ]
= lim 𝑎𝑗 (𝑥 + ℎ − 𝑎)𝑗 − (𝑥 − 𝑎)𝑗 − 𝑗(𝑥 − 𝑎)𝑗−1 ℎ .
𝑘→∞
𝑗=2
By the Mean Value Theorem, for each 𝑗 ≥ 2 there is ℎ𝑗 ∈ ℝ with 0 < |ℎ𝑗 | < |ℎ| such that
so that
‖∑ ‖
‖ 𝑎 (𝑥 + ℎ − 𝑎)𝑗 − (𝑥 − 𝑎)𝑗 − 𝑗(𝑥 − 𝑎)𝑗−1 ℎ ‖ ≤
‖ 𝑘 [ ]‖ ∑ 𝑘
≤ ℎ2
∑
𝑘
𝑗(𝑗 − 1)𝑟𝑗−2 ‖𝑎𝑗 ‖.
𝑗=2
‖𝑓 (𝑥 + ℎ) − 𝑓 (𝑥) − 𝑔(𝑥)ℎ‖ ≤ ℎ2 𝑐,
and therefore
𝑓 ′ (𝑥) = 𝑔(𝑥).
𝑑 ∑ 𝑗 ∞
∑ 1
exp(𝐴𝑥) = 𝐴𝑗 𝑥𝑗−1 = 𝐴𝑗+1 𝑥𝑗
𝑑𝑥 𝑗=1
𝑗! 𝑗=0
𝑗!
∑∞
1 𝑗 𝑗
=𝐴 𝐴 𝑥 = 𝐴 exp(𝐴).
𝑗=0
𝑗!
Similarly, (∞ )
𝑑 ∑ 1
exp(𝐴𝑥) = 𝐴𝑗 𝑥𝑗 𝐴 = exp(𝐴)𝐴.
𝑑𝑥 𝑗=0
𝑗!
2 and 3. A direct proof is quite cumbersome. Instead, we consider the matrix function
and observe that exp(𝐴 + 𝐶), exp(𝐴) and exp(𝐶) commute with 𝐴 and 𝐶 since 𝐴 + 𝐶, 𝐴 and 𝐶 commute
with 𝐴 and 𝐶. The product rule for derivatives and commutativity lead to
𝑑
Since 𝐹 (0) = (exp(0))3 = 𝐼𝑛3 = 𝐼𝑛 , where 𝐼𝑛 is the 𝑛 × 𝑛 identity matrix, and since 𝐼 = 0, the
𝑑𝑥 𝑛
uniqueness part of Corollary 7.16 gives 𝐹 (𝑥) = 𝐼𝑛 for all 𝑥 ∈ ℝ.
Putting 𝐶 = 0 and 𝑥 = 1, we get
How can one find exp(𝐴𝑥)? Recall from Linear Algebra that 𝜒(𝜆) = det(𝐴 − 𝜆𝐼𝑛 ) is the characteristic
function of 𝐴. Note that 𝜒 is a polynomial of degree 𝑛 with leading coefficient (−𝜆)𝑛 . In Complex
Analysis you will learn about the Fundamental Theorem of Algebra which says for this case that there
are 𝑛 complex numbers 𝜆1 , … , 𝜆𝑛 such that
∏
𝑛
𝜒(𝜆) = (𝜆𝑗 − 𝜆).
𝑗=1
Jordan’s canonical form, see Linear Algebra, says that the 𝜆𝑗 can be sorted into 𝑘 groups of equal number
𝜇1 , … , 𝜇𝑘 , i. e.,
∏ 𝑘
𝜒(𝜆) = (𝜇𝑗 − 𝜆)𝑝𝑗 ,
𝑗=1
where
𝐷𝑗 = 𝜇𝑗 𝐼𝑝𝑗 + 𝐽𝑝𝑗
and 𝐽𝜈 is the 𝜈 × 𝜈 matrix which has entries 1 just above the diagonal and 0 elsewhere. If we write, for
𝑙 ∈ ℕ, 𝐽𝜈𝑙 = (𝑚𝑟𝑠 )𝜈𝑟,𝑠=1 , then 𝑚𝑟𝑠 = 1 if 𝑠 − 𝑟 = 𝑙 and 𝑚𝑟𝑠 = 0 otherwise. In particular, 𝐽𝜈𝑙 = 0 if 𝑙 ≥ 𝜈.
Then ( 𝑘 )𝑚 ( 𝑘 )
⨁ ⨁
𝐴𝑚 = 𝑇 𝐷𝑗 𝑇 −1 = 𝑇 𝐷𝑗𝑚 𝑇 −1 ,
𝑗=1 𝑗=1
which gives
( 𝑘 )
⨁
exp(𝐴𝑥) = 𝑇 exp(𝐷𝑗 𝑥) 𝑇 −1 .
𝑗=1
It remains to find exp(𝐷𝑗 𝑥). For convenience, we drop the index and observe
𝑚 ( )
∑ 𝑚 𝑚−𝑙 𝑙
𝑚
(𝜇𝐼𝜈 + 𝐽𝜈 ) = 𝜇 𝐽𝜈 .
𝑙=0
𝑙
Using ( )
1 𝑚 𝑚! 1
= = ,
𝑚! 𝑙 𝑚!𝑙!(𝑚 − 𝑙)! 𝑙!(𝑚 − 𝑙)!
it follows that
∑∞
1
exp(𝐷𝑥) = (𝜇𝐼𝜈 + 𝐽𝜈 )𝑚 𝑥𝑚
𝑚=0
𝑚!
𝑚 ( )
∑∞
1 ∑ 𝑚 𝑚−𝑙 𝑙 𝑚
= 𝜇 𝐽𝜈 𝑥
𝑚=0
𝑚! 𝑙=0 𝑙
( )
∑ ∑ 1 𝑚
𝜈−1 ∞
= 𝜇 𝑚−𝑙 𝐽𝜈𝑙 𝑥𝑚
𝑙=0 𝑚=𝑙
𝑚! 𝑙
∑
𝜈−1
1∑
∞
1
= 𝜇 𝑚−𝑙 𝐽𝜈𝑙 𝑥𝑚
𝑙=0
𝑙! 𝑚=𝑙
(𝑚 − 𝑙)!
∑
𝜈−1
1∑ 1
∞
= (𝜇𝑥)𝑘 𝐽𝜈𝑙 𝑥𝑙
𝑙=0
𝑙! 𝑘=0
𝑘!
∑
𝜈−1 𝑙
𝑥
= exp(𝜇𝑥)𝐽𝜈𝑙 .
𝑙=0
𝑙!
Theorem 7.21. Assume that 𝐹 (𝑥) = 𝐹 (𝑦) for all 𝑥, 𝑦 ∈ 𝐼. Let 𝑐 ∈ 𝐼 and define
⎛ 𝑥 ⎞
⎜∫
𝑌 (𝑥) = exp ⎜ 𝐹 (𝑡) 𝑑𝑡⎟ .
⎟
⎝𝑐 ⎠
Then 𝑌 is a fundamental matrix of 𝑦′ = 𝐹 𝑦.
Proof. 𝑌 (𝑐) = exp(0) = 𝐼𝑛 is invertible. Since 𝐹 (𝑥) commutes with 𝐹 (𝑡) for all 𝑥, 𝑡 ∈ 𝐼, it follows that
𝑗 𝑗−1
⎡⎛ 𝑥 ⎞ ⎤ ⎛ 𝑥 ⎞
𝑑𝑥 ⎢⎜∫ ⎜∫
𝑑 ⎢⎜
𝐹 (𝑡) 𝑑𝑡⎟ ⎥ = 𝑗𝐹 (𝑥) ⎜ 𝐹 (𝑡) 𝑑𝑡⎟ ,
⎟ ⎥ ⎟
⎣⎝ 𝑐 ⎠ ⎦ ⎝𝑐 ⎠
so that
𝑗
⎛ 𝑥 ⎞
𝑑 ∑ 1 ⎜
∞
𝑑𝑥 𝑗=0 𝑗! ⎜∫
′
𝑌 (𝑥) = 𝐹 (𝑡) 𝑑𝑡⎟
⎟
⎝𝑐 ⎠
𝑗
∑∞ ⎡⎛ 𝑥 ⎞ ⎤
𝑗! 𝑑𝑥 ⎢⎜∫
1 𝑑 ⎢⎜
= 𝐹 (𝑡) 𝑑𝑡⎟ ⎥
⎟ ⎥
𝑗=0 ⎣⎝ 𝑐 ⎠ ⎦
𝑗−1
∑∞ ⎛ 𝑥 ⎞
𝑗
⎜∫
= 𝐹 (𝑥) ⎜ 𝐹 (𝑡) 𝑑𝑡⎟
𝑗! ⎟
𝑗=1 ⎝𝑐 ⎠
⎛ 𝑥 ⎞
⎜∫
= 𝐹 (𝑥) exp ⎜ 𝐹 (𝑡) 𝑑𝑡⎟
⎟
⎝𝑐 ⎠
= 𝐹 (𝑥)𝑌 (𝑥).
Theorem 7.22. Let 𝑌 be a fundamental matrix of 𝑦′ = 𝐹 𝑦, 𝑓 ∈ 𝐶(𝐼, 𝔽 𝑛 ), 𝑐 ∈ 𝐼 and 𝑑 ∈ 𝔽 𝑛 . Then the
unique solution of the initial value problem 𝑦′ = 𝐹 𝑦 + 𝑓 , 𝑦(𝑐) = 𝑑 is
𝑥
∫
−1
𝑦(𝑥) = 𝑌 (𝑥)𝑌 (𝑐)𝑑 + 𝑌 (𝑥) 𝑌 −1 (𝑡)𝑓 (𝑡) 𝑑𝑡.
𝑐
Proof. We could simply substitute 𝑦 into the differential equation. However, we give a constructive proof,
the so-called method of variation of the constants. It is easy to see that any solution of the homogeneous
equation is 𝑦(𝑥) = 𝑌 (𝑥)𝑧 with 𝑧 ∈ 𝔽 𝑛 . (we are not going to prove this separately because it is a special
case of this theorem and therefore follows anyway). Now replace the constant vector 𝑧 by a vector
function on 𝐼, i. e., 𝑦(𝑥) = 𝑌 (𝑥)𝑧(𝑥), where 𝑧 is the solution of the initial value problem.
Hence the proof goes as follows. From Corollary 7.14 we know that the initial value problem has a
unique solution 𝑦 ∈ 𝐶 1 (𝐼, 𝔽 𝑛 ). Define 𝑧 = 𝑌 −1 𝑦. Then
𝑦′ (𝑥) = 𝑌 ′ (𝑥)𝑧(𝑥) + 𝑌 (𝑥)𝑧′ (𝑥)
= 𝐹 (𝑥)𝑌 (𝑥)𝑧(𝑥) + 𝑌 (𝑥)𝑧′ (𝑥),
so that
𝑧′ (𝑥) = 𝑌 −1 (𝑥)𝑌 (𝑥)𝑧′ (𝑥)
= 𝑌 −1 (𝑥)[𝑦′ (𝑥) − 𝐹 (𝑥)𝑦(𝑥)]
= 𝑌 −1 (𝑥)𝑓 (𝑥).
Since
𝑧(𝑐) = 𝑌 −1 (𝑐)𝑦(𝑐) = 𝑌 −1 (𝑐)𝑑,
it follows that
𝑥
∫
𝑧(𝑥) = 𝑌 −1 (𝑡)𝑓 (𝑡) 𝑑𝑡 + 𝑌 −1 (𝑐)𝑑.
𝑐
Multiplication by 𝑌 (𝑥) completes the proof.
Note. 1. If 𝑛 > 1 and 𝐹 is not constant, then, in general, there is no closed form formula for the
fundamental matrix.
2. With the procedure outlined before Corollary 7.15, the results of this section also apply to linear 𝑛-
th order differential equations. In particular, the elements of the first row of a fundamental matrix are
called a fundamental system, say 𝑦1 , … , 𝑦𝑛 of the 𝑛-th order differential equation, and any solution of
the homogeneous 𝑛-th order equation is a linear combination of the fundamental system.
Conversely, if 𝑦1 , … , 𝑦𝑛 is a fundamental system of the 𝑛-th order differential equation, then
( )𝑛
𝑌 = 𝑦(𝑖−1)
𝑗 𝑖,𝑗=1
one can use Jordan’s canonical form to find a fundamental system. However, this is very cumbersome.
One can proceed as follows: Consider the characteristic polynomial
∑
𝑛−1
𝜇𝑛 − 𝑎𝑗 𝜇𝑗
𝑗=0
and find its distinct (complex) zeros, say 𝜇1 , … , 𝜇𝑘 , with multiplicities 𝑝1 , … , 𝑝𝑘 . Then the 𝑛 functions
𝑥𝑙 𝑒𝜇𝑗 𝑥 , 𝑗 = 1 … , 𝑘, 𝑙 = 0, … , 𝑝𝑗 − 1
form a fundamental system of the 𝑛-th order linear differential equations.