Book Algebra

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 226

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/372885564

HIGHER ALGEBRA: A COMPREHENSIVE STUDY OF MATHEMATICAL CONCEPTS

Book · August 2024

CITATIONS READS
0 168

1 author:

Carlos Polanco
Instituto Nacional de Cardiología
340 PUBLICATIONS 633 CITATIONS

SEE PROFILE

All content following this page was uploaded by Carlos Polanco on 24 November 2023.

The user has requested enhancement of the downloaded file.


Carlos Polanco

HIGHER ALGEBRA
A COMPREHENSIVE STUDY OF MATHEMATICAL
CONCEPTS
Copyright to the book

Dear reader, if you find the information contained in this book useful and choose to reference it in your
work, research, or publication, I kindly ask that you properly cite this work. By doing so, you contribute
to acknowledging the effort and dedication invested in the creation of this material and promote respect
for copyright. Below, I provide a suggested format for citing this book: Carlos Polanco (2024). HIGHER
ALGEBRA: A COMPREHENSIVE STUDY OF MATHEMATICAL CONCEPTS. DOI: 10.13140/RG.2.2.3
2054.8864 6/4, ResearchGate GmbH. I appreciate in advance your cooperation and respect for intellectual
property.
This work, “Higher algebra: a comprehensive study of mathematical concepts”, created by Carlos Polanco,
is available under a Creative Commons Attribution 4.0 International (CC BY 4.0) license. This license
allows others to share, copy, distribute, and use the work, even for commercial purposes, as long as the
original authorship is credited.

Word processor: LATEX ©2023; Operating system: Linux Fedora-39 ©2023.

Software: QuillBot (Course Hero), LLC. ©2023, & ChatGPT 4.0 OpenAI ©2023.

v
The only way to learn mathematics is to do
mathematics.

– Paul R. Halmos
1916 – 2006
Foreword

Within the pages of “HIGHER ALGEBRA: A COMPREHENSIVE STUDY OF MATHEMATICAL CONCEPTS,”


Carlos Polanco presents a comprehensive and enlightening exploration of abstract mathematics. This well-
structured book serves as an invaluable resource for students and teachers, providing a deep understanding of
the complexities inherent in the discipline. Through the presentation of numerous examples and case studies,
Polanco masterfully connects theory to practice, offering readers a reliable and clear source of mathematical
knowledge. It is with great confidence that I recommend “Higher Algebra” as an essential addition to any
mathematical library.

Cuernavaca Morelos, Mexico. Thomas Buhse


E-mail: buhse@uaem.mx
Universidad Autónoma del Estado de Morelos

ix
Foreword

It is with great pleasure that I introduce “HIGHER ALGEBRA: A COMPREHENSIVE STUDY OF


MATHEMATICAL CONCEPTS” by Carlos Polanco, a comprehensive and authoritative work in the realm
of abstract mathematics. With unwavering expertise and meticulous attention to detail, Polanco navigates
the depths of higher algebra, encompassing topics such as combinatorics, vector spaces, matrices, systems
of linear equations, number theory, real and complex numbers, polynomials, and the theory of equations.
“HIGHER ALGEBRA: A COMPREHENSIVE STUDY OF MATHEMATICAL CONCEPTS” serves as
an invaluable resource for students, educators, and researchers in the field of mathematics and its myriad
applications. Polanco’s dedication to presenting abstract concepts in an accessible manner ensures that
readers can confidently explore the depths of algebraic structures and their significance in various scientific
disciplines.

Pushchino, Moscow region, Russia. Vladimir N. Uversky


E-mail: vuversky@usf.edu
Russian Academy of Sciences

xi
Preface

Welcome to the captivating realm of higher algebra, an intricate and profound branch of mathematics that
unveils the elegant structures and interconnections within abstract algebraic systems. It is with great pleasure
and enthusiasm that we present this book, titled “HIGHER ALGEBRA: A COMPREHENSIVE STUDY
OF MATHEMATICAL CONCEPTS,” meticulously crafted to provide you, the students of Science, with a
comprehensive and rigorously formal introduction to the fundamental concepts of algebra and their wide-
ranging applications in scientific disciplines.
As students of Science, you are embarking on a remarkable journey of exploration and discovery, and a
strong foundation in higher algebra is crucial for your future studies and endeavors. This book has been
thoughtfully designed to equip you with the necessary tools and knowledge to comprehend the intricate
nature of abstract algebra and apply it to the challenges and complexities you will encounter in your scientific
pursuits.
In the first chapter, “Fundamental Concepts,” we will lay the groundwork for your journey into higher
algebra. Here, we will introduce and rigorously define the formal language of sets, relations, and functions,
which will serve as the cornerstone of precise mathematical reasoning. By emphasizing mathematical
rigor, we will establish the fundamental operations on sets, explore important relations, and investigate the
properties intrinsic to functions. This chapter will equip you with the necessary tools to navigate the intricate
landscape of higher algebra with confidence and precision.
Building upon this foundation, we will delve into combinatorics, a field essential to scientific inquiry.
Combinatorics will enable you to analyze discrete structures, unravel combinatorial identities, and solve
problems involving permutations, combinations, and more. The skills you acquire in combinatorics will
prove invaluable in fields such as computer science, cryptography, statistical analysis, and optimization,
empowering you to tackle complex scientific challenges.
Vector spaces, an essential topic in higher algebra, will be explored in detail. These abstract structures
provide a powerful framework for understanding linear transformations, spanning sets, bases, and subspaces.
By delving into the properties and operations of vectors, matrices, and linear transformations, we will
uncover their deep connections to systems of linear equations and determinants. This chapter will enhance
your ability to analyze and interpret scientific data, as well as provide you with a geometric understanding
of algebraic concepts.
Matrices and determinants, fundamental tools in mathematical modeling and scientific computation, will
occupy a prominent place in our study. We will investigate the properties of matrices, their algebraic

xiii
xiv Preface

operations, and their indispensable role in solving systems of linear equations. Furthermore, the study of
determinants will offer profound insights into the properties of matrices, including invertibility, rank, and
eigenvalues. Mastering matrix operations and determinants will equip you with powerful computational
tools to analyze complex systems and model real-world phenomena.
Our exploration of linear systems of equations will allow you to apply the concepts and techniques developed
thus far. We will analyze the methods for solving linear systems, including Gaussian elimination and matrix
algebra. By developing a deep understanding of the properties and solutions of linear systems, you will
be prepared to tackle scientific problems that involve the simultaneous analysis of multiple variables and
constraints.
The mathematician accomplishes this by solving and proposing exercises, then reviewing each example and
its solution in depth and reading each definition or explanation as many times as necessary until he or she
comprehends its meaning. Therefore, each chapter contains exercises of moderate to low difficulty, with
answers provided at the close of the book.
The author hopes that the content offered here is informative and motivating to the reader who is interested
in studying the fundamentals of this issue.
The author wishes to thank the Instituto Nacional de Cardiologı́a “Ignacio Chávez” and the Faculty of
Sciences at Universidad Nacional Autónoma de México for supplying useful examples.

CONFLICT OF INTEREST
The author declares no conflict of interest regarding the contents of each of the chapters of this ebook.

Carlos Polanco

E-mail: polanco@unam.mx
Department of New Technologies and Intellectual Protection
Instituto Nacional de Cardiologı́a “Ignacio Chávez”
México

Department of Mathematics, Faculty of Sciences


Universidad Nacional Autónoma de México
México
Acknowledgements

I want to thank everyone for their suggestions, which made it possible to publish this ebook.

xv
Contents

List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxi

1 Algebra of Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Natural Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Mathematical Induction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3 Integer Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4 Rational Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4.1 Comparing Fractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.5 Irrational Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.5.1 Transcendental and Algebraic Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.6 Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.6.1 Absolute Value on Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.6.2 Properties on Absolute Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.6.3 Density of Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.6.4 Archimedean Property on Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.6.5 The Incommensurability of Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.7 Complex Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2 Algebra of Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2 Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.3 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.3.1 Real-Valued Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.3.2 Elementary Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3.3 Composition of Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.3.4 Operations over Real-Valued Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.4 The Limit Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.4.1 Properties of the Limit Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.5 Sequences and Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

xvii
xviii Contents

2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

3 Algebra of Integer Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.2 Integer Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.3 Greatest Common Divisor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.4 Least Common Multiple . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.5 Fundamental Theorem of Arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.6 Congruences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.7 Arithmetic Module . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.8 Linear Congruences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.9 Euclid’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.10 Bézout’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.11 Chinese Remainder Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.12 Congruence-based Cryptography System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.13 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.14 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

4 Algebra of Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.2 Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.3 Axioms and Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.4 Order Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.5 Absolute Value and Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.6 Infinite Sequences and Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.7 Radicals and Exponents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

5 Algebra of Complex Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.2 Complex Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.3 Axioms and Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.4 Complex Conjugate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.5 Absolute Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.6 The Complex Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.7 Polar Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.8 De Moivre’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.9 Square Roots of Negative Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.10 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.11 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Contents xix

6 Algebra of Polynomial Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.2 Polynomial Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.3 Factorization of Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.4 The Factor Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.5 Solving Polynomial Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
6.6 Fundamental Theorem of Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.7 Finding Complex Roots Using Newton’s Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
6.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
6.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

7 Algebra of Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.2 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.3 Vector Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.4 Dot Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
7.5 Cross Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
7.6 Vector Subspace . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.7 Linear Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.8 Spanning Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.9 Vector Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
7.10 Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
7.11 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
7.12 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

8 Algebra of Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119


8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
8.2 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
8.3 Matrix operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
8.4 Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
8.5 Determinant Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.6 Adjugate Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.7 Inverse of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
8.8 System Linear Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
8.9 Cramer’s Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
8.10 Eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
8.11 Diagonalization of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
8.12 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.13 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

9 Linear Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135


9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
9.2 Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
9.3 Calculation Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
9.3.1 Cofactor Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
9.3.2 Sarrus Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
xx Contents

9.3.3 Gaussian Elimination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139


9.4 Homogeneous Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
9.5 Eigenvalues and Eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
9.6 Rank of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
9.7 Linear Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
9.8 Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
9.9 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
9.10 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

10 Algebra of Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153


10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
10.2 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
10.3 Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
10.4 Isomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
10.5 Homomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
10.6 Generator and Algebraic Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
10.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
10.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

11 SOLUTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

A Computational Programs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197


A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
A.2 Newton’s Method Program . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
A.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
List of Symbols

Symbol Description Page

A Set A 2
|A| Cardinality 2
A∪B Union of two sets 3
A∩B Intersection of two sets 3
A△B Symmetric difference of two sets 3
A⊆B A subset of set B 4
B⊇A B is a superset of set A 4
A′ or A Complement to the set A 4
B−A Relative complement of set A 4
A×B Cartesian product of two sets 4
n
[
Ai Nested union of Ai sets 6
i=1
\n
Ai Nested interception of Ai sets 6
i=1
N Set of natural numbers 6
b|a b divides to a 7
Z Set of integers numbers 11
f :A→B Function over sets 33
Q Set of rational numbers 12
I Set of irrational numbers 16
R Set of real numbers 18
|x| Absolute value of x real number 21
C Set of complex numbers 24
A×B Cartesian producto 30
aRb Binary relation 30
f :R→R Function over real sets 33
domain f Domain of the real-valued function f 34
codomain f Codomain of the real-valued function f 34

xxi
xxii Contents

Symbol Description Page

range f Range of the real-valued function f 34


graph f Graph of the real-valued function f 34
f −1 : B → A Inverse function of the function f 34
g◦ f Composition of f over g function 39
lim f (x) = L Limit of function 42
x→x0
an Sequences 49

∑ an Series 49
n=1
gcd(a, b) Greatest common divisor 56
lcm(a, b) Least common multiple 58
a ≡ b (mod n) Congruence modulo n 61
a mod b Arithmetic modulo of numbers a and b 63
ax ≡ b (mod m) Linear congruence 64
a + bi Complex number 84
z Conjugate of z 85
Re(a) Real component of complex number 86
Im(b) Imaginary component of complex number 86
P(x) Polynomial function 94
v Vector 108
V Vector space 108
u·v Dot product 111
u×v Cross product 112
W Vector subspace 114
S
  Spanning set 115
ai j Matrix A 120
ai j Determinant A 122
A−1 Inverse matrix A 124
Chapter 1
Algebra of Sets

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract This chapter presents a concise review of the fundamental numerical systems: natural numbers,
integer numbers, rational and irrational numbers, complex numbers, and real numbers. It begins with
the natural numbers and the properties of the extension of this numerical system to integer numbers.
Subsequently, the concept of rational numbers is introduced, emphasizing the properties that characterize
them, fractions. The chapter concludes with an introduction to the real numbers, which include both
rational and irrational numbers, and a brief discussion of complex numbers, where their real and imaginary
components intertwine to form this last numerical system. This summary aims to provide readers with a
solid foundation of these numerical systems.

Keywords: absolute value on real numbers, algebraic numbers, archimedean property on real numbers,
comparing fractions, complex numbers, density of real numbers, integer numbers, irrational numbers,
mathematical induction, natural numbers, operations and properties on integer numbers, properties on
absolute value, properties on integer numbers, rational numbers, real numbers, the incommensurability of
real numbers transcendental numbers.

1.1. Introduction
This chapter begins with a review of the numerical systems known as: natural numbers, integers, rationals,
irrationals, reals, and complexes. These numeric entities go beyond being mere symbols, as they form the

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

1
2 1 Algebra of Sets

pillars of understanding the complexity of the world. Our review starts with an analysis of the natural
numbers, which originate from the inherent human desire to enumerate objects. Delving into integers,
negative numbers emerge, enabling us to represent quantities below zero. As we progress, we enter the realm
of rational numbers and observe their capabilities to represent fractions. The next numerical system is that of
the irrational numbers, which are characterized by not being representable by fractions and having infinite,
non-recurring decimal expansions. We then encounter the numerical system of the real numbers, which
encompasses both rationals and irrationals, providing a continuous spectrum for measuring and comparing
quantities. Finally, we delve into complex numbers, which turn out to be composed of a real and an imaginary
part, revealing new avenues for understanding and solving equations made up of complex numbers.
The numerical sets we will study have a geometric representation on the real number line (Fig. 1.1), although
in this chapter, our focus will be on reviewing their analytical properties.

−∞ −3 −2 −1−0.5 0.5 1 2 3 ∞

Figure 1.1: Geometrical representation of the numerical system on the real line.

A set is a collection of distinct elements. We denote a set [1, 2, 3] by listing its elements between curly
brackets. For example, let’s consider a set A containing the numbers 1, 2, 3:

A = {1, 2, 3} (1.1)

Sets can be finite or infinite, and the elements can be of any type, including numbers, letters, or even other
sets.

Definition 1.1. A set is a well-defined collection of distinct objects, called elements, which are considered
as a single entity. In mathematical terms, a set is characterized by its elements and is denoted by listing them
inside curly braces.
If A is a set, the membership of an element x in A is represented by the symbol “∈”. The notation x ∈ A
means that x is an element of set A, while x ∈
/ A indicates that x is not an element of A.

Sets can be either finite or infinite. A finite set has a countable number of elements, which can be explicitly
listed and counted. On the other hand, an infinite set has an uncountable number of elements.
Sets are defined solely by their elements and are independent of their order or repetition. This means that the
order of elements within a set is irrelevant, and duplicate elements are not allowed. Each element can only
occur once in a set.
A set can be characterized by defining a rule or property that determines its elements, or by explicitly listing
the contents of the set.
The cardinality of a set refers to the number of elements it contains. For a finite set A, the cardinality is
denoted as |A|. For example, if B = {4, 5, 6}, then |B|= 3.
1.1 Introduction 3

There are several basic operations that can be performed on sets. Let’s consider two sets A and B:

A = {1, 2, 3}, B = {3, 4, 5} (1.2)

The union of two sets, denoted by AcupB, is the set containing every element from both sets.

A ∪ B = {1, 2, 3, 4, 5} (1.3)

The intersection of two sets, denoted as A ∩ B, is the set that contains only the elements that are common to
both sets:

A ∩ B = {3} (1.4)

The relative complement of set A with respect to set B, denoted as B−A, is the set that contains the elements
in B that are not in A:

B − A = {4, 5} (1.5)

The cartesian product of two sets, denoted as A × B, is the set of all possible ordered pairs where the first
element comes from set A and the second element comes from set B:

A × B = {(1, 3), (1, 4), (1, 5), (2, 3), (2, 4), (2, 5), (3, 3), (3, 4), (3, 5)} (1.6)

The symmetric difference between two sets, denoted by A△B, is defined as the set of elements that belong
to either set, but not to both sets simultaneously.
Formally, the symmetric difference between sets A and B is defined as:

A△B = (A ∪ B) − (A ∩ B)

This means that for a given element x, x belongs to A△B if and only if x belongs to A or to B, but not to both
sets at the same time.

Example 1.1. Let’s consider the following sets:

A = {1, 2, 3, 4}
B = {3, 4, 5, 6}

Applying the formula for symmetric difference, we have:

A△B = (A ∪ B) − (A ∩ B)
A△B = ({1, 2, 3, 4} ∪ {3, 4, 5, 6}) − ({1, 2, 3, 4} ∩ {3, 4, 5, 6})
A△B = {1, 2, 3, 4, 5, 6} − {3, 4}
A△B = {1, 2, 5, 6}
Therefore, the symmetric difference between sets A and B is {1, 2, 5, 6}.
4 1 Algebra of Sets

A set A is said to be a subset of another set B, denoted as A ⊆ B, if every element of A is also an element of
B. In this case, B is called a superset of A. If A is a subset of B but A is not equal to B, we denote it as A ⊂ B.
For example, if C = {1, 2}, and A = {1, 2, 3, 4} then C ⊂ A.

Example 1.2. Let’s consider the set A = {1, 2, 3}. Now, we will explore the concepts of subset and superset
using this set.
Solution 1.2. Suppose we have another set B = {1, 2, 3}. We can determine if set A is a subset of set B by
examining whether every element in A is also present in B. Using the subset symbol ⊆, we can express this
relationship as:

A⊆B (1.7)

In this case, since all elements in set A = {1, 2, 3} are also elements of set B, the subset relationship holds.
Conversely, we can determine if set B is a superset of set A by verifying if every element in A is also present
in B. Using the superset symbol ⊇, we can express this relationship as:

B⊇A (1.8)

In this example, since set B contains all the elements of set A = {1, 2, 3}, the superset relationship holds.

Understanding subset and superset relationships is essential in analyzing the inclusion and containment of
sets within one another. These concepts allow us to compare and classify sets based on their elements,
forming the foundation for further investigations in set theory and other branches of mathematics.
In set theory, several important concepts aid in understanding the relationships between sets and their
elements. Let’s explore three of these concepts: complement, empty set, and universal set.

Given a set A, the complement of A, denoted as A′ or A, represents all the elements that do not belong to A.
Formally, for a universal set U, the complement of A is defined as:

A′ = {x ∈ U | x ∈
/ A} (1.9)

In other words, A′ consists of all elements in the universal set U that are not contained in A. For example,
if U represents the set of all integers and A represents the set of even numbers, then A′ represents the set of
odd numbers.

Example 1.3. Let’s suppose we have a universal set U = {1, 2, 3, 4, 5} and a set A = {2, 4}. We can calculate
the complement of set A with respect to the universal set U as follows.
Solution 1.3.
A′ = {x ∈ U | x ∈
/ A} = {1, 3, 5} (1.10)
Therefore, the complement of set A with respect to the universal set U is {1, 3, 5}, which consists of all
elements in U that are not in A.
1.1 Introduction 5

The empty set, denoted as 0/ or {}, is a unique set that contains no elements. Formally, for any set A, the
empty set is defined as:

0/ = {x | x does not satisfy any condition} (1.11)

The empty set is essential in set theory and forms the basis for various mathematical concepts. It serves
as a starting point for constructing new sets and plays a crucial role in defining intersections, unions, and
complements.

Example 1.4. Consider a set B = {x | x > 10, x < 5} that contains all numbers that are greater than 10 and
less than 5. In this case, there are no numbers that satisfy both conditions simultaneously. Therefore, the set
B is the empty set, denoted as 0. /

The universal set, denoted as U or Ω, represents the collection of all possible elements under consideration.
It serves as the context within which sets are defined and operations are performed. The universal set often
depends on the specific problem or context being studied. For instance, in probability theory, the universal
set may represent the sample space of all possible outcomes in an experiment.
The universal set is significant as it allows us to define complements and set operations relative to a common
frame of reference. It provides a comprehensive scope for set-related discussions and allows for clear and
unambiguous mathematical reasoning.

Example 1.5. Let’s suppose we are working on a probability problem, and the sample space of an experiment
is the universal set U = {H, T }, where H represents head and T represents tail. In this case, the universal set
U consists of all the possible outcomes of the experiment.

These concepts of complement, empty set, and universal set contribute to the rich tapestry of set theory
and are fundamental building blocks in many mathematical disciplines. Understanding their definitions and
properties paves the way for deeper explorations in set theory and its applications.
The nested union of sets involves combining the elements from multiple sets into a single set. It is denoted
S
by , and it is used to represent the union of sets within sets. Formally, the nested union of sets can be
defined as follows:
n
[
Ai = A1 ∪ A2 ∪ . . . ∪ An (1.12)
i=1

This means that the nested union of sets A1 , A2 , . . . , An contains all the elements that are present in any of the
sets A1 , A2 , . . . , An . Here’s a numerical example to illustrate this concept:

Example 1.6. Let A1 = {1, 2, 3}, A2 = {3, 4, 5}, and A3 = {5, 6, 7}.
Solution 1.6. The nested union of these sets can be calculated as:
3
[
Ai = A1 ∪ A2 ∪ A3 = {1, 2, 3, 4, 5, 6, 7} (1.13)
i=1

Thus, the nested union of sets A1 , A2 , and A3 yields the set {1, 2, 3, 4, 5, 6, 7}.
6 1 Algebra of Sets

The nested intersection of sets involves finding the common elements among multiple sets. It is denoted
T
by , and it represents the intersection of sets within sets. Formally, the nested intersection of sets can be
defined as:
n
\
Ai = A1 ∩ A2 ∩ . . . ∩ An (1.14)
i=1

This means that the nested intersection of sets A1 , A2 , . . . , An contains only the elements that are present in
all of the sets A1 , A2 , . . . , An . Let’s consider a numerical example:

Example 1.7. Let B1 = {1, 2, 3}, B2 = {2, 3, 4}, and B3 = {3, 4, 5}.

Solution 1.7. The nested intersection of these sets can be calculated as:
3
\
Bi = B1 ∩ B2 ∩ B3 = {3} (1.15)
i=1

Thus, the nested intersection of sets B1 , B2 , and B3 yields the set {3}, which contains the common element
present in all three sets.

1.2. Natural Numbers


The natural numbers have played a fundamental role in the daily lives of civilizations throughout history.
The need to count objects, animals, and events drove the development of a numeral system based on natural
numbers in early human communities. Archaeological artifacts, such as carved bones and inscriptions on
stones, containing the earliest numerical representations, were used to keep track of food, livestock, and
days. As societies evolved, more complex numbering systems emerged, such as Egyptian hieroglyphics,
Mayan glyphs, and Hindu-Arabic numerals. Among other things, these systems facilitated calculations and
business records.
The natural numbers [4, 5], denoted by N, are the set of positive integers starting from 1 and extending
infinitely. It can be expressed as:

N = {1, 2, 3, 4, 5, . . .}
The addition of two natural numbers a and b is denoted by the symbol +. For example, if a = 3 and b = 4,
their sum is a + b = 3 + 4 = 7.
The subtraction of two natural numbers a and b where a is greater than b is denoted by the symbol −. For
example, if a = 7 and b = 3, their difference is a − b = 7 − 3 = 4.
The multiplication of two natural numbers a and b is denoted by the symbol · or simply by juxtaposition.
For example, if a = 2 and b = 5, their product is a · b = 2 · 5 = 10.
The division of two natural numbers a and b is denoted by the symbol ÷ or by using a fraction bar. It gives
the quotient when a is divided by b. For example, if a = 10 and b = 2, their quotient is a ÷ b = 10 ÷ 2 = 5.
1.2 Natural Numbers 7

Note. Let a and b be naturals, with b 6= 0. We say that a is divisible by b (or b divides a) [6, 7], denoted as
b | a, if there exists an natural k such that a = kb.
In other words, b divides a or b | a if the division of a by b yields an natural quotient without any remainder.
Symbolically, we can express divisibility as:

b | a ⇐⇒ ∃ k ∈ N, such that a = kb (1.16)

Example 1.8.

(i) 4 | 12 because 12 = 4 · 3.

(ii) 2 ∤ 9 because there is no natural k such that 9 = 2k.

Divisibility plays a crucial role in number theory and has various applications in areas such as prime
factorization, congruences, and solving linear equations. It provides a fundamental concept for understanding
the relationships between natural.
The natural numbers have several important properties, including closure, commutativity, associativity,
identity elements, and distributivity. These properties ensure that the operations on natural numbers follow
certain rules and produce meaningful results.
Closure under Addition and Multiplication: For any two natural numbers a and b, their sum a + b and their
product a · b are also natural numbers. This property ensures that adding or multiplying two natural numbers
always results in another natural number.
Associativity of Addition and Multiplication: For any three natural numbers a, b, and c, the addition and
multiplication operations are associative. This means that (a + b) + c = a + (b + c) and (a · b) · c = a · (b · c),
regardless of the order in which the operations are performed.
Commutativity of Addition and Multiplication: For any two natural numbers a and b, addition and multiplication
are commutative. This property states that a + b = b + a and a · b = b · a. The order of the numbers does not
affect the result.
Identity Elements: The natural number 0 does not belong to the set of natural numbers N. However, it is
used as an identity element for addition in the extended set of natural numbers N0 , where a + 0 = a for any
natural number a. The number 1 serves as the identity element for multiplication, where a · 1 = a for any
natural number a.
Ordering Relation: The natural numbers are ordered in a straightforward manner. For any two natural
numbers a and b, one and only one of the following is true: a < b, a = b, or a > b. This ordering relation
allows us to compare and arrange natural numbers in increasing or decreasing order.
The natural numbers have a natural order based on their magnitudes. The ordering relations include:

(i) Less than (<): If a is less than b, it is denoted as a < b.

(ii) Greater than (>): If a is greater than b, it is denoted as a > b.


8 1 Algebra of Sets

(iii) Less than or equal to (≤): If a is less than or equal to b, it is denoted as a ≤ b.

(iv) Greater than or equal to (≥): If a is greater than or equal to b, it is denoted as a ≥ b.

Example 1.9. Let’s consider some examples using the natural numbers:
n = 10
Addition:
n + 5 = 10 + 5 = 15

Subtraction:
n − 3 = 10 − 3 = 7

Multiplication:
n · 2 = 10 · 2 = 20

Division:
n ÷ 2 = 10 ÷ 2 = 5

Ordering Relations:
8<n
n>5
10 ≥ n
n ≤ 20

These examples demonstrate the basic operations, properties, and ordering relations of natural numbers.

1.2.1. Mathematical Induction


Mathematical induction [8] has played a pivotal role in proving a vast range of mathematical statements for
centuries. Its historical significance stems from its ability to establish the truth of general claims based on a
limited number of cases. Throughout history, renowned mathematicians have used this technique to ascertain
mathematical patterns, verify formulas, and prove theorems. The elegance and power of mathematical
induction have enabled mathematicians to tackle complex problems by breaking them down into smaller and
more manageable parts, ultimately leading to significant advancements in various mathematical disciplines.
Mathematical induction is used to establish statements that depend on a variable n being a natural number.
The method consists of two steps: the base case and the inductive step.
Base Case: The base case establishes the statement for the smallest possible value of n. Typically, it starts
with n = 0 or n = 1, depending on the specific problem. We verify that the statement holds true for this base
case.
1.2 Natural Numbers 9

Inductive Step: The inductive step involves two parts: the inductive hypothesis and the inductive proof.
Inductive Hypothesis: Assume that the statement is true for a certain value k. This is known as the inductive
hypothesis. We assume the statement holds for n = k and use it to prove that the statement is also true for
n = k + 1.
Inductive Proof: Using the inductive hypothesis, we show that if the statement holds for n = k, it also holds
for n = k + 1. This step usually involves algebraic manipulations and logical reasoning. By doing so, we
establish a chain of implications that connects the base case to every subsequent case, proving the statement
for all natural numbers.

Example 1.10. Let’s prove the following statement using mathematical induction:
n
n(n + 1)
∑ j= 2
(1.17)
j=1

Base Case: For n = 1, we have:


1
1(1 + 1)
∑ j=1= 2
(1.18)
j=1

The statement holds true for the base case.


Inductive Hypothesis: Assume that the statement holds for n = k, where k ≥ 1. That is:
k
k(k + 1)
∑ j= 2
(1.19)
j=1

Inductive Proof: We need to show that the statement holds for n = k + 1. Using the inductive hypothesis, we
have:

k+1 k
∑ j = ∑ k + (k + 1)
j=1 k=1
k(k + 1)
= + (k + 1) (Using the inductive hypothesis)
2
k(k + 1) + 2(k + 1)
=
2
(k + 1)(k + 2)
=
2
(k + 1)((k + 1) + 1)
=
2

This confirms that the statement is true for n = k + 1. 


By completing the base case and the inductive step, we have proven the statement using mathematical
induction.
10 1 Algebra of Sets

Example 1.11. Let’s prove that the sum of the first n natural numbers is given by the formula:

n(n + 1)
1+2+3+...+n = (1.20)
2
Base Case: For n = 1, we have:

1(1 + 1)
1= (1.21)
2
The formula holds true for the base case.
Inductive Hypothesis: Assume that the formula holds for n = k, where k ≥ 1. That is:

k(k + 1)
1+2+3+...+k = (1.22)
2
Inductive Proof: We need to show that the formula holds for n = k + 1. Using the inductive hypothesis, we
have:

k(k + 1)
1 + 2 + 3 + . . . + k + (k + 1) = + (k + 1) (Using the inductive hypothesis)
2
k(k + 1) + 2(k + 1)
=
2
(k + 1)(k + 2)
=
2
(k + 1)((k + 1) + 1)
=
2

If we substitute n = k + 1 into the expression (Eq. 1.20), we observe that we arrive at the same previous
(k + 1)((k + 1) + 1)
result .
2
This confirms that the formula holds for n = k + 1. 

1.3. Integer Numbers


Similar to the natural numbers, the integer numbers [9] have played a crucial role in the evolution of
humanity. As societies evolved, there was a need to count both positive and negative quantities. Originating
as an extension of the natural numbers to include both positive and negative numbers, integers provide
a more comprehensive means to describe and measure various situations. This expansion of the number
system allowed for the solution of more complex problems in physics and economics, among other fields.
Integer numbers have been indispensable throughout history for solving equations and recognizing patterns
and regularities in the natural world. Their utility and implementations transcend cultures and eras, making
them an indispensable instrument for the advancement of science and society.
1.3 Integer Numbers 11

The integers are a set of numbers that includes both positive and negative whole numbers, along with zero.
We denote the set of integers as Z.

(i) Addition: Given two integers a and b, the sum a + b is also an integer. For example, 5 + (−3) = 2.

(ii) Subtraction: The difference a − b of two integers a and b is also an integer. For example, 8 − (−2) = 10.

(iii) Multiplication: When we multiply two integers a and b, the result a · b is an integer. For example,
(−4) · 3 = −12.

(iv) Division: Division of two integers a and b may or may not yield an integer. If b divides a without a
10
remainder, then the result is an integer. For example, = 5.
2

(v) Closure Property: The sum, difference, and product of any two integers are also integers.

(vi) Associative Property: Addition and multiplication of integers are associative operations. That is, for any
three integers a, b, and c, we have (a + b) + c = a + (b + c) and (a · b) · c = a · (b · c).

(vii) Commutative Property: Addition and multiplication of integers are commutative operations. That is, for
any two integers a and b, we have a + b = b + a and a · b = b · a.

(viii) Identity Elements: The integer 0 serves as the additive identity, where a + 0 = a for any integer a. The
integer 1 serves as the multiplicative identity, where a · 1 = a for any integer a.

(ix) Inverse Elements: Every integer a has an additive inverse −a, such that a + (−a) = 0. However, not all
integers have multiplicative inverses. Only integers of the form 1 and −1 have multiplicative inverses,
which are also 1 and −1, respectively.

Example 1.12. Let’s consider the integers a = −5 and b = 3, verify their operations and properties.
Addition:
a + b = −5 + 3 = −2

Subtraction:
a − b = −5 − 3 = −8

Multiplication:
a · b = −5 · 3 = −15

Division:
a −5 5
= =−
b 3 3
12 1 Algebra of Sets

Remark 1.1. When we divide a by b, we get b|a = −5/3 which is not an integer

Closure Property: Adding or subtracting any two integers results in an integer. For example, a+b = −5+3 =
−2 and a − b = −5 − 3 = −8 are integers.
Associative Property: Addition and multiplication of integers are associative. For example, (a + b) + c =
(−5 + 3) + 2 = 0 and (a · b) · c = (−5 · 3) · 2 = −30.
Commutative Property: Addition and multiplication of integers are commutative. For example, a + b =
−5 + 3 = 3 + (−5) = b + a and a · b = −5 · 3 = 3 · (−5) = b · a.
Identity Elements: The additive identity is 0. For example, a + 0 = −5 + 0 = −5. The multiplicative identity
is 1. For example, a · 1 = −5 · 1 = −5.
Inverse Elements: The additive inverse of a is −a. For example, a + (−a) = −5 + 5 = 0. However, not all
integers have multiplicative inverses. Only integers of the form 1 and −1 have multiplicative inverses.
This example demonstrates how the basic operations of addition, subtraction, multiplication, and division
are performed with integers, and how the properties hold true in these calculations.

These are some of the basic concepts and properties associated with integers, providing a foundation for
further exploration in number theory.

1.4. Rational Numbers


The origins of rational numbers can be traced back to ancient Greece, where mathematicians began to
investigate fractional properties. The Pythagoreans were especially influential in the study of rational
numbers. It was revealed that there are incommensurable magnitudes, that is, quantities that could not be
expressed as a ratio of two integers, like irrational square roots. This discovery contradicted the Pythagorean
belief that the universe is completely numeric and rational.
In his work Elements, the Greek mathematician Euclid [10] developed a rigorous theory of fractions in the
5th century BC. In it, he established principles and properties for fractional operations and proved that any
fraction can be expressed as the ratio of two integers. This basic concept paved the way for the systematic
study of rational numbers.
Arab and Persian mathematicians, including Al-Khwarizmi and Al-Farabi, made significant contributions
to the development of rational numbers during the Middle Ages. They introduced decimal fractions and
began to use the fractional notation that we employ today. Moreover, they researched fractional arithmetic
operations and devised methods for simplifying and adding fractions.
a c
(i) Addition: The sum of two rational numbers and is calculated by finding a common denominator
b d
and adding the numerators:
a c ad + bc
+ =
b d bd
1 3 1 · 4 + 3 · 2 10 5
Example 1.13. + = = = .
2 4 2·4 8 4
1.4 Rational Numbers 13

a c
(ii) Subtraction: The difference between two rational numbers and is calculated similarly to addition:
b d
a c ad − bc
− =
b d bd
3 1 3·2−1·4 2 1
Example 1.14. − = = = .
4 2 4·2 8 4
(iii) Multiplication: The product of two rational numbers is obtained by multiplying the numerators and
denominators:
a c ac
· =
b d bd
2 3 2·3 6 2
Example 1.15. · = = = .
3 5 3 · 5 15 5
(iv) Division: To divide one rational number by another, we multiply the first number by the reciprocal of
the second number:
a c a d ad
÷ = · =
b d b c bc
2 4 2 5 2 · 5 10 5
Example 1.16. ÷ = · = = = .
3 5 3 4 3 · 4 12 6

(i) Closure: The sum, difference, product, and quotient of two rational numbers are also rational numbers.

(ii) Associativity: Addition and multiplication of rational numbers are associative operations, meaning the
grouping of numbers does not affect the result.

(iii) Commutativity: Addition and multiplication of rational numbers are commutative operations, meaning
the order of numbers does not affect the result.

(iv) Identity: The number 0 serves as the additive identity for rational numbers, and 1 serves as the
multiplicative identity.

(v) Inverse: Every non-zero rational number has an additive inverse (negative) and a multiplicative inverse
(reciprocal).

(vi) Distributive Property: Rational numbers follow the distributive property of multiplication over addition/subtraction:
a · (b + c) = a · b + a · c.

Ordering Relations:

a c a c
(i) Less than (<): If and are rational numbers, < if ad < bc.
b d b d
a c a c
(ii) Greater than (>): If and are rational numbers, > if ad > bc.
b d b d
14 1 Algebra of Sets

a c a c
(iii) Less than or equal to (≤): If and are rational numbers, ≤ if ad ≤ bc.
b d b d
a c a c
(iv) Greater than or equal to (≥): If and are rational numbers, ≥ if ad ≥ bc.
b d b d

3 5
Example 1.17. Compare the rational numbers and using the less than (<) relation.
4 6
3 5 3 5
Solution 1.17. We have < if 3 · 6 < 4 · 5, which simplifies to 18 < 20. Therefore, is less than .
4 6 4 6

1.4.1. Comparing Fractions


Comparing fractions is an important topic in the realm of rational numbers. It involves determining the
relative magnitudes of fractions and deciding whether one fraction is greater than, less than, or equal to
another.
One common method of comparing fractions is by finding a common denominator. This entails identifying
a common denominator for the fractions being compared and then comparing the numerators. The fraction
with the larger numerator is considered greater.
3 5
Let’s consider the fractions and . To compare them, we can find a common denominator by multiplying
4 8
3 24
the denominators: 4 × 8 = 32. Then, we adjust the fractions to have the same denominator: = and
4 32
5 20 24 20
= . Now, we can compare the numerators directly and determine that is greater than .
8 32 32 32
Another method for comparing fractions is using the concept of “unit fractions.” A unit fraction has a
numerator of 1. When comparing two fractions, we compare the equivalent unit fractions in each case.
The fraction whose equivalent unit fraction is greater is considered greater overall.
2 5 2
For example, let’s compare the fractions and . If we convert these fractions to unit fractions, we get
3 7 3
5
and . We observe that 2 × 7 > 3 × 5 → 14 > 15.
7
The comparison of fractions is essential in the study of rational numbers as it allows us to establish an order
and magnitude relationship among fractions. Including this topic in your text will provide readers with a
more comprehensive understanding of rational numbers and how they are compared to one another.

1.5. Irrational Numbers


The origins of “irrational numbers” can be traced back to ancient Greece, where the Pythagoreans made
significant mathematical advances. The Pythagoreans believed that all numbers, known as rational numbers,
could be expressed as ratios of integers. When analyzing the hypotenuse of an isosceles right triangle,
1.5 Irrational Numbers 15

they ran into a fundamental difficulty, however. They discovered that the hypotenuse length could not be
expressed as a rational number, which led to the discovery of irrational numbers.
An irrational number is defined as a √ real number that cannot be represented as the quotient or ratio of
two integers. The square root of 2 ( 2) is one of the best-known examples of an irrational number.
The Pythagoreans credited Hippasus of Metapontum with discovering the existence of irrational numbers.
Irrationality posed a challenge to the Pythagorean worldview, which emphasized the harmony and order of
the universe based on rational numbers. This information was kept confidential by the Pythagoreans because
it contradicted their fundamental principles.
Ancient Greek mathematicians, such as Euclid and Archimedes, advanced the study of these numbers
through their writings. The Elements of Euclid, a seminal mathematical work, presented rigorous proofs
and included statements about irrational numbers. Archimedes devised methods for approximating the
value of pi (pi) using polygons, thereby demonstrating the existence of irrational quantities in geometric
measurements. Other ancient civilizations continued to evolve their understanding of irrational numbers.
Aryabhata and other Indian mathematicians contributed to the development of numerical methods, such as
algorithms for calculating square roots. In addition, Chinese mathematicians, such as Liu Hui, made progress
in approximating irrational numbers.
Mathematicians such as Rafael Bombelli and John Wallis made significant contributions to the study of
irrational numbers during the Renaissance. As complex numbers frequently involve the square roots of
negative numbers, Bombelli’s work on complex numbers enhanced our knowledge of irrationality. Wallis
introduced the symbol ∞ (infinity) to represent an infinitely enormous amount, highlighting the infinite
nature of irrational numbers. Recognizing geometric measurements that could not be expressed as rational
numbers led to the historical development of irrational numbers. This discovery challenged the Pythagorean
belief in the exclusivity of rational numbers and led to the study of irrationality in a number of mathematical
contexts. Over time, the contributions of mathematicians from various civilizations to the study of irrational
numbers led to a deeper understanding of their mathematical properties and significance.

(i) Closure under addition:

∀a, b ∈ I, where a 6= b a + b ∈ I (1.23)


√ √ √
Example 1.18. Let a = 2 and b = π . Both 2 and π are irrational numbers. Their sum, 2 + π , is also
an irrational number.

(ii) Closure under subtraction:


∀a, b ∈ I, a − b ∈ I (1.24)
√ √ √
Example 1.19. Let a = e and b = 3. Both e and 3 are irrational numbers. Their difference, e − 3,
is also an irrational number.

(iii) Closure under multiplication:


∀a, b ∈ I, a · b ∈ I (1.25)
√ √ √
Example 1.20. Let a = π and b = 5. Both π and 5 are irrational numbers. Their product, π · 5, is
also an irrational number.
16 1 Algebra of Sets

(iv) Closure under division:


a
∀a ∈ I, ∀b ∈ Q \ {0}, ∈I (1.26)
b
√ 1 √
Example 1.21. Let a = 2 and b = . 2 is an irrational number, and b is a non-zero rational number.
√ 3
2 √
1
Their quotient, 1
= 3 2, is an irrational number.
3

(v) Associativity of addition:


∀a, b, c ∈ I, (a + b) + c = a + (b + c) (1.27)
√ √ √ √ √ √
Example 1.22. Let a = 2, b = π , and c = 3. The sum of ( 2 + π ) + 3 is equal to 2 + (π + 3).

(vi) Associativity of multiplication:

∀a, b, c ∈ I, (a · b) · c = a · (b · c) (1.28)
√ √ √ √ √ √
Example 1.23. Let a = π , b = 2, and c = 3. The product of (π · 2) · 3 is equal to π · ( 2 · 3).

(vii) Commutativity of addition:


∀a, b ∈ I, a + b = b + a (1.29)
√ √ √
Example 1.24. Let a = 2 and b = π . The sum of 2 + π is equal to π + 2.

(viii) Commutativity of multiplication:


∀a, b ∈ I, a · b = b · a (1.30)
√ √ √
Example 1.25. Let a = π and b = 2. The product of π · 2 is equal to 2 · π .

(i) Addition and Subtraction: For any two irrational numbers a and b, where a 6= b, their sum a + ba + b
and difference a − ba − b are also irrational numbers.

∀a, b ∈ I ⇒ (a + b ∈ I) ∧ (a − b ∈ I) (1.31)

(ii) Multiplication: The product of two irrational numbers aa and bb is another irrational number.

∀a, b ∈ I ⇒ (a · b ∈ I) (1.32)

(iii) Division: The division of two irrational numbers aa and bb may or may not result in an irrational number.
It can yield a rational number or another irrational number.

∀a, b ∈ I ⇒ (a ÷ b ∈ Q) ∨ (a ÷ b ∈ I) (1.33)
1.5 Irrational Numbers 17

(iv) Closure under Powers: Raising an irrational number a to a rational power n can result in either an
irrational number or a rational number, depending on the specific values of a and n.

∀a ∈ I, n ∈ Q ⇒ (an ∈ I) ∨ (an ∈ Q) (1.34)

(v) Density Property: Irrational numbers possess the density property, meaning that between any two
distinct irrational numbers, there exists another irrational number.

∀a, b ∈ I, : a < b ⇒ ∃c ∈ I : such that : a < c < b (1.35)

(vi) Transcendence: Some irrational numbers are transcendental, which means they are not algebraic.
Transcendental numbers cannot be solutions to any non-zero polynomial equation with integer coefficients.

∃x ∈ I, : x is transcendental (1.36)

1.5.1. Transcendental and Algebraic Numbers


A transcendental number is an irrational number that is not a root of any non-zero polynomial equation with
integer coefficients. In other words, it cannot be expressed as a solution of a polynomial equation of the
form:

an xn + an−1 xn−1 + . . . + a1 x + a0 = 0 (1.37)

Where an , an−1 , . . . , a1 , a0 integers, and n is a positive integer.

Example 1.26. The number e (Euler’s number), which is approximately 2.71828, is a transcendental number.
It arises naturally in many mathematical contexts, such as exponential growth and calculus. Another
famous transcendental number is π (pi), approximately 3.14159, which represents the ratio of a circle’s
circumference to its diameter.

An algebraic number is a number that is a root of a non-zero polynomial equation with integer coefficients.
In other words, it can be expressed as a solution of a polynomial equation of the form mentioned earlier.
For example, the square root of 2 is an algebraic number because it is a solution to the equation

x2 − 2 = 0 (1.38)

Example 1.27. The golden ratio (ϕ ), which is approximately 1.61803, and is a solution to the equation

x2 − x − 1 = 0 (1.39)

Every transcendental number is an irrational number, but not all irrational numbers are transcendental. In
fact, the vast majority of irrational numbers are actually algebraic. Examples of algebraic numbers include
square roots, cube roots, and solutions to higher-degree polynomial equations.
18 1 Algebra of Sets

The set of algebraic numbers is countable, meaning it can be put into a one-to-one correspondence with the
positive integers. On the other hand, the set of transcendental numbers is uncountable, which means it has
a higher cardinality than the set of natural numbers or integers.
Transcendental numbers are a subset of irrational numbers that cannot be expressed as solutions to polynomial
equations with integer coefficients. Algebraic numbers, on the other hand, are irrational numbers that can be
expressed as solutions to such equations. The set of transcendental numbers is uncountable, while the set
of algebraic numbers is countable.

1.6. Real Numbers


The set of real numbers, denoted by R, comprises all possible values on the number line. It encompasses√
rational numbers, such as fractions and decimals, as well as irrational numbers, such as π and 2. The
real numbers are characterized by their unique properties, including closure under addition, subtraction,
multiplication, and division. They form a complete and ordered field, allowing for precise calculations and
comparisons. The set of real numbers is infinitely dense, meaning that between any two distinct real numbers,
there exists an infinite number of other real numbers. This rich collection of elements provides the foundation
for various mathematical disciplines and applications, facilitating the exploration of continuous quantities
and mathematical modeling in diverse fields.
The real numbers, denoted by R, possess several key properties that make them a fundamental concept in
mathematics. These properties include:
Closure under Addition and Subtraction: For any two real numbers a and b, their sum a + b and their
difference a − b are also real numbers. This property ensures that the result of adding or subtracting real
numbers remains within the set of real numbers.

Closure under Multiplication and Division: For any two real numbers a and b, their product a · b and their
a
quotient (where b 6= 0) are also real numbers. This property guarantees that multiplying or dividing real
b
numbers results in another real number.

Associativity of Addition and Multiplication: For any three real numbers a, b, and c, the addition and
multiplication operations are associative. This means that (a + b) + c = a + (b + c) and (a · b) · c = a · (b · c),
regardless of the order in which the operations are performed.

Commutativity of Addition and Multiplication: For any two real numbers a and b, addition and multiplication
are commutative. This property states that a + b = b + a and a · b = b · a. The order of the numbers does not
affect the result.

Distributive Property: For any three real numbers a, b, and c, the distributive property holds: a · (b + c) =
a · b + a · c. This property allows us to distribute a factor across a sum or difference.

Identity Elements: The real number 0 serves as the additive identity, where a + 0 = a for any real number a.
The real number 1 acts as the multiplicative identity, where a · 1 = a for any real number a.
1.6 Real Numbers 19

Ordering Relation: The real numbers are ordered in a linear manner. For any two real numbers a and b, one
and only one of the following is true: a < b, a = b, or a > b. This ordering relation allows for comparisons
and arrangements of real numbers.

These properties of real numbers provide a solid foundation for performing arithmetic operations, solving
equations, and exploring the concepts of calculus and analysis.

Example 1.28. Let’s consider the real numbers a = 3, b = −2, and c = 5. We will demonstrate various
properties using these values.
Closure under Addition and Subtraction:

a + b = 3 + (−2) = 1 ∈ R
a − b = 3 − (−2) = 5 ∈ R

Closure under Multiplication and Division:

a · b = 3 · (−2) = −6 ∈ R
a 3 3
= =− ∈R
b −2 2

Associativity of Addition and Multiplication:

(a + b) + c = (3 + (−2)) + 5 = 6
a + (b + c) = 3 + ((−2) + 5) = 6
(a · b) · c = (3 · (−2)) · 5 = −30
a · (b · c) = 3 · ((−2) · 5) = −30

Commutativity of Addition and Multiplication:

a + b = 3 + (−2) = 1
b + a = (−2) + 3 = 1
a · b = 3 · (−2) = −6
b · a = (−2) · 3 = −6

Distributive Property:

a · (b + c) = 3 · ((−2) + 5) = 9
(a · b) + (a · c) = (3 · (−2)) + (3 · 5) = 9

Identity Elements:

a+0 = 3+0 = 3
1·a = 1·3 = 3
20 1 Algebra of Sets

Ordering Relation: Since a = 3 and b = −2, we have a > b.


This example shows the properties of real numbers using the values a = 3, b = −2, and c = 5. These
properties hold true for any real numbers, showcasing the fundamental characteristics of the real number
system.
Example 1.29. Solve the inequality: 2x + 3 > 7
Solution 1.29. We start by isolating the variable x on one side of the inequality.

2x + 3 > 7
2x > 7 − 3
2x > 4

Next, we divide both sides of the inequality by 2 to solve for x.

2x 4
>
2 2
x>2

Therefore, the solution to the inequality is x > 2.


x+3
Example 1.30. Solve the inequality: ≤1
x−2
Solution 1.30. Case 1: x − 2 > 0 (i.e., x > 2)

When multiplying both sides by x − 2, we do not change the inequality sign because x − 2 is positive in this case.
x+3 ≤ x−2
3 6≤ −2

This is clearly false. Therefore, there are no solutions for x ¿ 2.


Case 2: x − 2 < 0 (i.e., x < 2)

When multiplying both sides by x − 2, we reverse the inequality sign because x − 2 is negative in this case.
x+3 ≥ x−2
3 ≥ −2

This is true for all x values that are less than 2.


Combining the information from both cases, the solution to the inequality is:

x<2

Note: The point x = 2 is a vertical asymptote and cannot be included in the solution.
1.6 Real Numbers 21

1.6.1. Absolute Value on Real Numbers


Definition 1.2. The absolute value of a real number x, denoted as |x|, is a non-negative value that represents
the distance of x from the origin on the number line. It is defined as:
(
x if x ≥ 0
|x|= (1.40)
−x if x < 0

1.6.2. Properties on Absolute Value


(i) Non-negativity: |x| is always non-negative, meaning |x|≥ 0.

(ii) Symmetry: |x| is symmetric about the origin, i.e., |x|= |−x|.

(iii) Triangle Inequality: For any real numbers x and y, the absolute value satisfies the triangle inequality:

|x + y|≤ |x|+|y| (1.41)

(i) Addition: For real numbers x and y, the absolute value of their sum is:

|x ± y|≤ |x|±|y| (1.42)

Example 1.31. For x = −3 and y = 5, we have:

|−3 + 5|= 2 ≤ |−3|+|5|= 3 + 5 = 8 (1.43)

Example 1.32. For x = 3 and y = 5, we have:

|3 + 5|= 8 = |3|+|5|= 3 + 5 = 8 (1.44)

(ii) Multiplication: For real numbers x and y, the absolute value of their product is:

|x · y|= |x|·|y| (1.45)

Example 1.33. For x = −2 and y = 4, we have:

|−2 · 4|= |−2|·|4|= 2 · 4 = 8 (1.46)


22 1 Algebra of Sets

(iii) Division: For real numbers x and y (where y 6= 0), the absolute value of their quotient is:

x |x|
= (1.47)
y |y|

Example 1.34. For x = −6 and y = 2, we have:

−6 |−6| 6
= = =3 (1.48)
2 |2| 2

These properties and operations of absolute value are essential for various mathematical calculations and
inequalities, providing a way to measure magnitude and distance in the real number system.

1.6.3. Density of Real Numbers


Density of real numbers refers to a fundamental property of this set of numbers, which asserts that there is
always another real number between any two distinct real numbers. In other words, regardless of how close
two real numbers are, another real number can always be found between them. This property is indispensable
for the continuity and accuracy of real numbers.
The density of real numbers can be expressed as follows:

∀a, b ∈ R, : a < b =⇒ ∃c ∈ R : such that : a < c < b (1.49)

This equation states that for any pair of real numbers a and b, where a is less than b, there always exists a
real number c that lies between them.
A example of the density of real numbers is to consider the numbers a = 1 and b = 2. Equation (1) assures
us that there exists a real number c such that 1 < c < 2. We can choose, for example, c = 1.5, which satisfies
this condition.
Similarly, if we consider the numbers a = π and b = 4, the density of real numbers guarantees that there
exists a real number c such that π < c < 4. We can choose c = 3.2 as an example of a real number that
satisfies this property.
The density of real numbers also extends to cases where a and b are rational numbers. For instance, if we
1 1
consider the numbers a = and b = , the density property ensures that there exists a real number c such
3 2
1 1 2 1 1
that < c < . We can choose c = as an example of a real number that lies between and .
3 2 5 3 2
The density of real numbers is a property that guarantees we can always find another real number between
any two distinct real numbers. This property is fundamental for the precision and continuity of real numbers
and applies to both integers and rational and irrational numbers.

Example 1.35. Prove that between any two distinct real numbers, there exists another real number.
1.6 Real Numbers 23

Solution 1.35. Let us assume two distinct real numbers a and b. Without loss of generality, assume that
a < b.
a+b
1. Consider the real number c given by c = , which is the average of a and b.
2
a+b
2. It is clear that c is distinct from both a and b. If we were to assume c = a, then = a, which would
2
imply b = a. This is a contradiction since we assumed a and b are distinct. A similar argument can show
that c 6= b.

3. We now show that c lies strictly between a and b.

a. From a < b, adding a to both sides, we have 2a < a + b. Dividing both sides by 2, we get
a+b
a< = c.
2
b. Again, from a < b, subtracting b from both sides yields a − b < 0. Adding 2b to both sides, we get
a+b
a + b < 2b. Dividing both sides by 2, we have = c < b.
2

Therefore, a < c < b.

In conclusion, c is a real number that lies strictly between a and b. Hence, between any two distinct real
numbers, there always exists another real number.

1.6.4. Archimedean Property on Real Numbers


The Archimedean property is a fundamental property of the real numbers, which essentially states that for
any positive real number, there exists a positive integer that is larger than it. In other words, the natural
numbers (positive integers) are “dense” in the real number system.
The Archimedean property can be stated as follows:
For any positive real number x, there exists a positive integer n such that n > x. In other words, no matter
how large a positive real number you choose, there will always be a positive integer that is greater than it.
Let’s illustrate the Archimedean property with an example:

Example 1.36. Prove the Archimedean property: For any positive real number x, there exists a positive
integer n such that n > x.

Proof. Let x be a positive real number.


Consider the set
S = {n ∈ N : n ≤ x}

This set contains all the natural numbers that are less than or equal to x.
24 1 Algebra of Sets

1. We first note that S is not empty. The number 1 is in S because 1 ≤ x.

2. Since x is a positive real number, S is bounded above by x.

3. By the well-ordering principle, every non-empty set of positive integers has a least element. So, there
exists a least positive integer n such that n is the greatest integer in S. By the definition of S, n ≤ x but
n + 1 > x.

Thus, we have found a positive integer n + 1 which is greater than x, and this completes the proof. 

The Archimedean property is a crucial property in the real number system, and it allows us to establish a
clear relationship between the positive real numbers and the positive integers. It plays a fundamental role in
various mathematical proofs and applications.

1.6.5. The Incommensurability of Real Numbers


The incommensurability of real numbers refers to the fact that the set of real numbers is uncountably infinite.
In other words, there is no one-to-one correspondence between the set of real numbers and the set of natural
numbers. This property arises due to the unbounded nature and continuous nature of the real number line.
Unlike the set of natural numbers, which can be enumerated in a one-to-one manner, the real numbers are
densely packed on the number line, and there are infinitely many real numbers between any two given real
numbers.
To understand that property, consider the interval [0, 1] on the real number line. This interval contains a
continuum of real numbers, including rational and irrational numbers. Now, suppose we try to create a list
of all the real numbers in the interval [0, 1], represented as decimals. We might begin by listing 0.1, 0.2, 0.3,
and so on. However, this list will never capture all the real numbers in the interval since between any two
decimals, there are infinitely many more decimals that can be inserted. For example, between 0.1 and 0.2,
we can insert 0.15, 0.17, 0.18, and so on. This process can be repeated indefinitely, leading to an uncountable
number of real numbers in the interval [0, 1].
The incommensurability of real numbers has significant implications in various branches of mathematics,
particularly in analysis and measure theory. It demonstrates the vast richness and complexity of the real
number system, allowing for precise and continuous mathematical modeling of quantities and phenomena
in the physical world.

1.7. Complex Numbers


to include numbers of the form a + bi, where a and b are real numbers and i is the
Real numbers are extended√
imaginary unit defined by −1. Complex numbers are a fundamental mathematical concept. The concept
of complex numbers arose from the study of equations involving square roots of negative numbers, which
1.7 Complex Numbers 25

were at first considered “imaginary” or “unreal” quantities. In the 16th and 17th centuries, mathematicians
including Rafael Bombelli, John Wallis, and René Descartes first introduced and developed the concept of
complex numbers. However, it was not until the 18th century that complex numbers were recognized as
a crucial mathematical instrument for solving a variety of problems, particularly in the fields of algebra
and analysis. Today, complex numbers find applicability in a variety of fields of mathematics, physics,
engineering, and other sciences, playing a crucial role in comprehending and describing the world’s complex
phenomena.
Associativity under addition: For all z1 , z2 , z3 ∈ C, addition of complex numbers is associative:

(z1 + z2 ) + z3 = z1 + (z2 + z3 ) (1.50)

Closure under addition: For all z1 , z2 ∈ C, the sum of two complex numbers is also a complex number:

z1 + z2 ∈ C (1.51)

Zero additive property: There exists a complex number 0 ∈ C such that for any complex number z ∈ C, the
sum of z and 0 is equal to z itself:
z + 0 = z for all z ∈ C (1.52)


(i) Existence of Imaginary Unit: The imaginary unit, denoted by i, is defined as −1. It has the property
that i2 = −1.

(ii) Real and Imaginary Parts: A complex number z can be written in the form z = a + bi, where a and b are
real numbers. The real part of z, denoted by Re(z), is a, and the imaginary part of z, denoted by Im(z),
is b.

(iii) Conjugate: The conjugate of a complex number z = a + bi is denoted by z and is given by z = a − bi.
The conjugate of a complex number has the same real part but the opposite sign in the imaginary part.

(iv) Equality: Two complex numbers are equal if and only if their real parts are equal and their imaginary
parts are equal. If z1 = a + bi and z2 = c + di, then z1 = z2 if and only if a = c and b = d.

(v) Addition: The sum of two complex numbers z1 = a + bi and z2 = c + di is given by z1 + z2 =


(a + c) + (b + d)i. Addition of complex numbers is performed by adding their real parts and adding
their imaginary parts separately.

(vi) Subtraction: The difference between two complex numbers z1 = a + bi and z2 = c + di is given by
z1 − z2 = (a − c) + (b − d)i. Subtraction of complex numbers is performed by subtracting their real parts
and subtracting their imaginary parts separately.

Example 1.37. Perform the following complex number addition and subtraction: (3 + 4i) + (2 − 5i) and
(5 − 2i) − (1 + 3i).
Solution 1.37. For addition, we simply add the real parts and the imaginary parts separately: (3 + 4i) + (2 −
5i) = (3 + 2) + (4 − 5)i = 5 − i.
26 1 Algebra of Sets

For subtraction, we subtract the real parts and the imaginary parts separately: (5 − 2i) − (1 + 3i) = (5 − 1) +
(−2 − (3))i = 4 − 5i.

Example 1.38. Perform the following complex number multiplication: (2 + 3i)(4 − i).
Solution 1.38. To multiply complex numbers, we use the distributive property: (2 + 3i)(4 − i) = 2(4) +
2(−i) + 3i(4) + 3i(−i) = 8 − 2i + 12i − 3i2 .
Since i2 is defined as −1, we can simplify further: 8 − 2i + 12i − 3i2 = 8 − 2i + 12i − 3(−1) = 8 − 2i + 12i +
3 = 11 + 10i.
3 + 4i
Example 1.39. Perform the following complex number division:
2−i
Solution 1.39. To divide complex numbers, we can multiply both the numerator and denominator by the
3 + 4i 2 + i
complex conjugate of the denominator: · .
2−i 2+i
Expanding and simplifying the numerator and denominator:

(3 + 4i)(2 + i) 6 + 3i + 8i + 4i2
=
(2 − i)(2 + i) 4 − i2
.
Since i2 is defined as −1, we can simplify further:

6 + 3i + 8i + 4i2 6 + 11i + 4(−1) 2 + 11i


2
= =
4−i 4 − (−1) 5
.
2 + 11i
Therefore, the division result is .
5

1.8. Conclusions
This chapter reviews the classifications, properties, and operations of numbers in a comprehensive manner.
It includes the systems formed by the natural numbers (N), integers (Z), rationals (Q), irrationals (I), real
numbers (R), and complex numbers (C), revealing their interrelationships and inherent properties.
The natural numbers, denoted by (N), are the numbers first used for counting and range from 1 to infinity. In
various contexts, they are used to determine quantities and measure the size of sets. As we move beyond the
natural numbers, we encounter the integers (Z), which consist of all the natural numbers and their additive
inverses. The integers allow us to describe quantities that can be either positive or negative and are useful for
algebraic operations.
As a natural extension, the rational numbers (Q) arise, which encompass all numbers that can be expressed
as fractions of two integers. They arise from the need to symbolize quantities that can be precisely defined
by a fraction and display repetitive decimal expansions. In contrast, the irrational numbers (I) are numbers
that cannot be expressed as fractions and whose decimal representation lacks a recurring pattern. The square
1.8 Conclusions 27

root of two ( 2) and the constant π are examples of irrational numbers. The set of all rational and irrational
numbers is known as the set of real numbers (R), which forms a continuous and uninterrupted number line.

The real numbers are closed under addition, subtraction, multiplication, and division, allowing for precise
manipulation of quantities. Furthermore, they possess associative, commutative, and distributive properties,
ensuring that operations can be carried out in a consistent and predictable manner.
However, the narrative does not end with real numbers. We introduce the concept of complex numbers (C)
in order to solve complex problems and explore abstract mathematical spaces. Complex numbers extend
the real number √system to include numbers of the form a + bi, where a and b are real numbers and i is the
imaginary unit ( −1). Complex numbers allow for the description and analysis of phenomena that involve
two-dimensional quantities, including electrical circuits and waves, among many other applications.
This chapter delves into the fundamental nature of numbers, focusing on the distinctions and relationships
among natural numbers (N), integers (Z), rationals (Q), irrationals (I), real numbers (R), and complex
numbers (C).
28 1 Algebra of Sets

1.9. Exercises
Exercise 1.1. Solve the equation 2x + 3 = 7 for x.

Exercise 1.2. Calculate the sum of the first 10 natural numbers.


2 5
Exercise 1.3. Simplify the expression − .
3 6
3 5
Exercise 1.4. Calculate the value of × .
4 8

Exercise 1.5. Solve the equation 5x + 3 = 2.

Exercise 1.6. Calculate the absolute value of the sum of the first 20 negative integers.

Exercise 1.7. Evaluate the expression 34−2 .

Exercise 1.8. Calculate the square root of 16.


p √
Exercise 1.9. Simplify the expression 3 + 2 2.

Exercise 1.10. Prove that the sum of two irrational numbers can be a rational number.

3
Exercise 1.11. Calculate the value of √ by simplifying the expression as much as possible.
2−1
Exercise 1.12. Simplify the expression (2 + i)(3 − i), where i is the imaginary unit.

Exercise 1.13. Prove that 2 is an irrational number.
√ √
Exercise 1.14. Prove that 2 + 3 is an irrational number.

Exercise 1.15. Calculate the product of the complex numbers (2 + i)(3 − i).

Exercise 1.16. Solve the following inequality: 3x − 5 > 2x + 7

Exercise 1.17. Solve the following absolute inequality: |2x − 1|< 5


Chapter 2
Algebra of Functions

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract This chapter examines fundamental mathematical concepts in depth, concentrating on binary
relations, functions, real-valued functions, their properties, and various types of functions. In addition, the
operator limit and the study of sequences and series are covered. The chapter begins with a discussion
of binary relations, which establish relationships between elements of two sets. The discussion of various
properties and classifications of binary relations provides a thorough comprehension of their behavior and
applications. The chapter then examines functions, which describe the relationship between inputs and
outputs. Different categories of functions, including polynomial, exponential, logarithmic, and trigonometric
functions, are analyzed with a focus on their distinct mathematical representations and defining characteristics.
The focus is on real-valued functions, with an in-depth analysis of functions that map real numbers to real
numbers. This chapter examines their properties, including continuity, differentiability, and integrability,
and sheds light on their significance in mathematical modeling and analysis. The chapter then introduces the
limit operator, which permits the examination of function behavior as the input approaches a certain value.
The presentation of properties and theorems pertaining to limits provides a potent instrument for analyzing
functions and their behavior in various contexts. Sequences and series, which are fundamental mathematical
structures in both pure and applied mathematics, are discussed in the final section of the chapter. The
author investigates the properties, convergence criteria, and applications of sequences and series, enhancing
the reader’s comprehension of these fundamental mathematical tools. Through comprehensive coverage
of binary relations, functions, real-valued functions, their properties, various types of functions, the limit
operator, and sequences and series, this chapter provides readers with a solid foundation in mathematical
analysis and a framework for further exploration of various mathematical disciplines.

Keywords: binary relation, cartesian product, composition of functions, convergente series, convergent
sequences, divergent sequences, divergent series, elementary functions, exponential functions, functions,

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

29
30 2 Algebra of Functions

integer numbers, limit operator, logarithmic functions, natural numbers, polynomial function, real-valued
functions, sequences, series, trigonometric functions.

2.1. Introduction
A relation is a mathematical concept that describes the connection or relationship between two or more
elements from distinct collections. It describes the relationship among the elements in two sets. Relationships
can be represented in a variety of formats, such as tables, graphs, and ordered pair sets. In the study
of real-valued functions, the concept of relations is notably crucial. A real-valued function is a relation
that assigns each element of one set of real numbers (the domain) to an element of another set of real
numbers (the codomain). Each input corresponds exactly to a single output. In functional notation, real-
valued functions are typically represented as y = f (x), where x is the input variable and y is the output
variable. f receives an input x from its domain, applies a rule or formula to it, and produces an output y.
Functions allow us to characterize and analyze relationships between quantities in a variety of applications.
They can represent physical phenomena, economic models, geometric transformations, and much more.
By analyzing the properties and behavior of real-valued functions, we gain insight into the nature of these
relationships and are able to make predictions or solve problems in numerous fields. Relations and real-
valued functions are foundational mathematical concepts that allow us to comprehend and analyze the
connections among mathematical objects in the context of real numbers. They provide a solid framework
for modeling and solving problems in numerous mathematical disciplines and applications. By examining
the properties and operations associated with real-valued functions, we can gain valuable insights into the
behavior and influence of these functions on a wide range of mathematical phenomena.

2.2. Relations
A binary relation is a mathematical concept that defines a relationship or connection between two elements
of a set. Given a set A, a binary relation R on A is a subset of the Cartesian product A × A. In other words, R
consists of ordered pairs (a, b), where a and b are elements of A, and (a, b) is in R if and only if the relation
holds between a and b.
Binary relations exhibit various properties that describe their behavior and characteristics. Here are some
important properties:
(i) Reflexivity: A binary relation R on a set A is reflexive if every element of A is related to itself. In other
words, for all a in A, (a, a) belongs to R. Example: The relation “is equal to” (=) on the set of real
numbers is reflexive since every real number is equal to itself.

(ii) Symmetry: A binary relation R on a set A is symmetric if, for every pair (a, b) in R, the pair (b, a) also
belongs to R. In other words, if (a, b) is in R, then (b, a) is also in R. Example: The relation “is a sibling
of” on the set of people is symmetric since if person A is a sibling of person B, then person B is also a
sibling of person A.
2.2 Relations 31

(iii) Transitivity: A binary relation R on a set A is transitive if, for any three elements a, b, and c in A, if (a, b)
and (b, c) are both in R, then (a, c) also belongs to R. In other words, if a is related to b and b is related to
c, then a is related to c. Example: The relation “is less than” (<) on the set of real numbers is transitive
since if a < b and b < c, then a < c.

(iv) Antisymmetry: A binary relation R on a set A is antisymmetric if, for all elements a and b in A, if (a, b)
is in R and (b, a) is in R, then a = b. In other words, both a and b can only be related to each other in R
if they are the same element. Example: The relation “is less than or equal to” (≤) on the set of integers
is antisymmetric since if a ≤ b and b ≤ a, then a = b.
(v) Order relation: An order relation, also known as a partial order, is a binary relation that satisfies three
properties: reflexivity, antisymmetry, and transitivity. It provides a way to compare elements of a set
based on their order or precedence.

Example 2.1. The relation “is less than or equal to” (≤) on the set of natural numbers is an order relation
since it is reflexive, antisymmetric, and transitive.
Example 2.2. Create a relation for a subset of the integers.

(−1, 1) 2 (1, 1)

(0, 0) x
−2 −1 1 2
−1

(−1, −1) (1, −1)


−2

Figure 2.1: The relation R = {(−1, −1), (−1, 0), (−1, 1), (0, −1), (0, 0), (0, 1), (1, −1), (1, 0), (1, 1)}.

Example 2.3. Consider the relation R on the set of integers defined as R = {(x, x) | x ∈ Z}. We can see that
for every integer x, (x, x) is in R, making it reflexive.
Solution 2.3. Given our relation R is defined as

R = {(x, x) | x ∈ Z}

For every integer x, the pair (x, x) is included in R by definition. Thus, by the reflexive definition, it’s evident
that R is reflexive.
Example 2.4. Consider the relation R on the set of real numbers defined as R = {(x, y) | x + y = 0}. We can
observe that for every pair of elements (x, y) in R, (y, x) is also in R, making it symmetric.
32 2 Algebra of Functions

Solution 2.4. Let (x, y) be an arbitrary pair in R. By the definition of R, we have:

x+y = 0

,
=⇒ y = −x
=⇒ x = −y
So the pair (y, x) can be described by the equation:

y+x = 0

Since y + x is the same as x + y, it means (y, x) is also in R. Hence, for every pair (x, y) in R, (y, x) is also in
R, proving the symmetry of R.

Example 2.5. Consider the relation R on the set of integers defined as R = {(x, y) | x ≤ y}. We can observe
that for every pair of distinct elements (x, y) in R, (y, x) is not in R, making it antisymmetric.
Solution 2.5. Assume that (x, y) ∈ R and (y, x) ∈ R. From the definition of R, we have:

(x, y) ∈ R =⇒ x ≤ y,
(y, x) ∈ R =⇒ y ≤ x.

From the above, if x ≤ y and y ≤ x, then it must be the case that x = y. This confirms that the relation R is
antisymmetric.

Example 2.6. Consider the relation R on the set of positive integers defined as R = {(x, y) | x ≤ y2 }. We can
observe that for every triple of elements (x, y, z) in R, if (x, y) and (y, z) are in R, then (x, z) is also in R,
making it transitive.
Solution 2.6. Given the relation:
R = {(x, y) | x ≤ y2 }

If (x, y) ∈ R, then by definition:


x ≤ y2 (1)

If (y, z) ∈ R, then:
y ≤ z2 (2)

Given that y is a positive integer, it means y2 is positive as well. So:

y3 ≤ y · z2

Using (1), where x ≤ y2 , it implies:


x · y ≤ y3

From the above two inequalities, we get:


x · y ≤ y · z2
2.3 Functions 33

=⇒ x ≤ z2 ( when y 6= 0, which is true as y is a positive integer)

Thus, (x, z) ∈ R. Therefore, the relation R is transitive.

2.3. Functions
The history of functions is extensive, and they have played a crucial role in numerous scientific and
mathematical fields. The origins of the concept of functions can be traced back to ancient civilizations, in
which scholars investigated the relationships between quantities and observed patterns in nature. In the 17th
and 18th centuries, however, the formal study of functions came to emerge. Leonhard Euler, an influential
Swiss mathematician, was one of the important figures in the development of the theory of functions.
Calculus and analysis pioneered by Euler paved the way for a deeper comprehension of functions. He
introduced notation and terminology that are still widely used today, including the concept of a function
as a mathematical object that relates input and output values.
A function [2, 11, 12] is a mathematical relationship that assigns a unique output value to each input value.
Formally, a function f from a set A (called the domain) to a set B (called the codomain) is denoted as:

f :A→B (2.1)

The input value x is mapped to a corresponding output value y, which is denoted as f (x). In other words,
every element x in the domain has a unique image f (x) in the codomain.

2.3.1. Real-Valued Functions


The following properties are restricted to the real numbers, and when this is the case, functions are usually
referred to as real-valued functions.
One-to-One (Injective) Functions: A function is one-to-one if each element in the domain maps to a distinct
element in the codomain. Symbolically, for x1 , x2 ∈ A, if f (x1 ) = f (x2 ) implies x1 = x2 , then the function is
one-to-one.

Example 2.7. Let f : R → R be defined as f (x) = 2x. This function is one-to-one since different inputs will
always produce different outputs.

Onto (Surjective) Functions: A function is onto if every element in the codomain has at least one pre-image
in the domain. In other words, the range of the function covers the entire codomain.

Example 2.8. Consider the function f : R → R defined as:

f (x) = x3
34 2 Algebra of Functions

To prove that this function is onto, we need to show that for every element y in the codomain R, there exists
an element x in the domain R such that f (x) = y.

Given any y ∈ R, the cubic root of y is a real number, which we can denote as x = 3 y. When we plug this
x into our function:
√ √
f ( 3 y) = ( 3 y)3 = y
Since this is true for every y ∈ R, the function f (x) = x3 is onto.

Bijective Functions: A function is bijective if it is both one-to-one and onto, meaning it has a unique and
distinct mapping between its domain and codomain.

Example 2.9. The function h : R → R defined as h(x) = x3 is bijective since it is both one-to-one and onto.

Functions also have elements that can be observed:


Domain: The set of all possible input values for a function, in symbols domain f . Its values are located on
the horizontal axis of the graph.
Codomain: The set of all possible output values for a function. Its values are locate on the vertical axis of
the graph.
Range: The set of all actual output values obtained from the function. It is a subset of the codomain, in
symbols image f or range f . Its values are locate on the vertical axis of the graph.
Inverse Function: If a function f : A → B is bijective, it has an inverse function f −1 : B → A, which undoes
the effect of f . The inverse function swaps the roles of the domain and codomain.
Graph Function: The elements represent by graph( f ) = {(x, y) ∈ R2 | x, y = f (x) ∈ R}. Its values represent
the curve.

Example 2.10. Let f : R → R be defined as f (x) = 3x + 2. Determine the inverse function.

Solution 2.10.
f (x) − 2
=x
3
Then
x−2
f −1 (x) =
3
.

Example 2.11. Consider the function f : R → R defined by:

f (x) = x2 (2.2)

To determine if the function f : R → R defined by f (x) = x2 is a bijection, we need to verify its injectivity
and surjectivity.
A function is injective if and only if for all x1 , x2 ∈ R such that f (x1 ) = f (x2 ), it follows that x1 = x2 .
2.3 Functions 35

Suppose f (x1 ) = f (x2 ). This translates to:


x12 = x22 (2.3)

However, this does not necessarily imply x1 = x2 . For instance, x1 = 2 and x2 = −2 both give f (x) = 4.
Therefore, the function is not injective.
A function is surjective if and only if for every y ∈ R, there exists an x ∈ R such that f (x) = y.
For f (x) = x2 , no real value of x will produce a y < 0. Thus, the function is not surjective.
Since the function f (x) = x2 is neither injective nor surjective, we conclude that it is not a bijection.

Figure 2.2: Graph of the function f (x) = x2

2.3.2. Elementary Functions


Polynomial functions: are algebraic expressions in which variables are raised to different non-negative
integer powers. They are continuous curves that can be graphed, and their shapes depend on the degree
and leading term. Polynomial functions can have multiple roots or zeros.

(i) The degree of a polynomial function is the highest power of the variable in the expression.
36 2 Algebra of Functions

(ii) Polynomial functions are continuous for all real numbers.

(iii) The end behavior of a polynomial function is determined by the leading term.

(iv) Polynomial functions can have multiple roots or zeros.

(v) Polynomial functions can exhibit symmetry, such as even or odd symmetry.

Example 2.12. Polynomial function and its graph.

f (x) = x2 − 1 (2.4)

Graph of y = x2 − 1
8 y
y = x2 − 1

x
−3 −2 −1 1 2 3

−2

Figure 2.3: Graph of the function x2 − 1.

Exponential Functions: Exponential functions represent quantities that grow or decay at an exponential rate.
They have a base raised to a variable power and are often used to model exponential growth or decay in
various natural and economic phenomena.

(i) Exponential functions grow or decay at an exponential rate.

(ii) The base of an exponential function determines the direction and rate of growth/decay.
2.3 Functions 37

(iii) Exponential functions have a horizontal asymptote.

(iv) Exponential functions can be used to model processes involving exponential growth or decay.

(v) The inverse of an exponential function is a logarithmic function.

Example 2.13. Exponential function and its graph.

f (x) = 3x (2.5)

f (x)

x
−2 2

Figure 2.4: Geometrical representation of the exponential function f (x) = 3x .

Trigonometric Functions: Trigonometric functions relate angles and the ratios of the sides of a right triangle.
They are periodic and repeat their values over a certain interval. Trigonometric functions have widespread
applications in physics, engineering, and natural sciences.

(i) Trigonometric functions relate angles and the ratios of the sides of a right triangle.

(ii) Trigonometric functions are periodic and repeat their values over a certain interval.

(iii) The unit circle is often used to define trigonometric functions.


38 2 Algebra of Functions

(iv) Trigonometric functions have specific properties like even/odd functions, amplitude, period, and phase
shift.

(v) Trigonometric functions are used to model periodic phenomena like waves and oscillations.

Example 2.14.
f (x) = sin(x) (2.6)

f (x)

x
−2π − 32π −π − π2 π π 3π 2π
2 2

−1

Figure 2.5: Geometrical representation of the trigonometry function f (x) = sin x.

Logarithmic Functions: Logarithmic functions are the inverses of exponential functions. They have a base
and are used to solve exponential equations and represent exponential growth/decay in a compressed form.

(i) Logarithmic functions are the inverse of exponential functions.

(ii) The base of the logarithm determines the scaling factor between the input and output.

(iii) Logarithmic functions have a vertical asymptote.

(iv) Logarithmic functions are used to solve exponential equations and represent exponential growth/decay
in a compressed form.

(v) Logarithmic functions can be transformed using properties like the product rule, quotient rule, and
change of base formula.
2.3 Functions 39

Example 2.15.
f (x) = log10 (x) (2.7)

f (x)

Figure 2.6: Geometrical representation of the logarithmic function f (x) = log x.

2.3.3. Composition of Functions


The composition of functions is an operation that combines two functions to create a new function. Let’s
consider two functions, g : A → B and f : B → C, where A, B, and C are sets. The composition of f and g,
denoted as f ◦ g, is defined as:

( f ◦ g)(x) = f (g(x)) (2.8)

In other words, to evaluate the composition f ◦ g at a particular input x, we first apply the function g to x, and
then apply the function f to the result. The resulting function f ◦ g has a domain of A and a codomain of C.
The next figure (Fig. 2.7) show the composition f ◦ g at a particular input x.
40 2 Algebra of Functions

f (g(x))

y
f (x)
g(x)

Figure 2.7: Geometrical representation of the composition f ◦ g at a particular input x.

Now, let’s illustrate with an example that composition of functions is not commutative.

Example 2.16. Consider the functions f (x) = x2 and g(x) = x. We will show that (g ◦ f )(x) is not equal to
( f ◦ g)(x).

Solution 2.11. Let’s compute (g ◦ f )(x):

(g ◦ f )(x) = g( f (x))
= g(x2 )

= x2
= |x|

Now, let’s compute ( f ◦ g)(x):

( f ◦ g)(x) = f (g(x))

= f ( x)

= ( x)2
=x

From the calculations, we can see that (g ◦ f )(x) = |x| and ( f ◦ g)(x) = x. Since |x| and x are not always
equal, we conclude that composition of functions is not commutative in general.
2.3 Functions 41

This example demonstrates that the order of composition matters, and the resulting function may differ
depending on the order in which the functions are composed.

2.3.4. Operations over Real-Valued Functions


(i) Translations: Translations of a function involve shifting it horizontally or vertically in the Cartesian
plane. A horizontal translation is achieved by adding or subtracting a constant c to the independent
variable x, resulting in a shift left or right. Mathematically, the formula for a horizontal translation is:

f (x ± c) (2.9)

Where ± indicates the direction of the translation. A positive value of c results in a rightward translation,
whereas a negative value of c results in a leftward translation.
A vertical translation is achieved by adding or subtracting a constant d to/from the function f (x), causing
a shift upward for positive values of d or downward for negative values of d. Mathematically, the formula
for a vertical translation is:

f (x) ± d (2.10)

Where ± represents the direction of the translation. A positive value of d corresponds to an upward
translation, and a negative value of d corresponds to a downward translation.
(ii) Stretches: Stretches of a function involve amplifying or reducing its amplitude or length in the Cartesian
plane. A vertical stretch is achieved by multiplying the function f (x) by a constant a greater than 1,
resulting in an upward stretch. Mathematically, the formula for a vertical stretch is:

a f (x) (2.11)

Where a > 1.
A vertical compression is accomplished by multiplying the function f (x) by a constant 0 < b < 1,
leading to a downward compression. Mathematically, the formula for a vertical compression is:

b f (x) (2.12)

Where 0 < b < 1.


A horizontal stretch is achieved by multiplying the independent variable x by a constant 0 < c < 1,
causing a leftward stretch. Mathematically, the formula for a horizontal stretch is:

f (cx) (2.13)

Where 0 < c < 1.


A horizontal compression is performed by multiplying the independent variable x by a constant d greater
than 1, resulting in a rightward compression. Mathematically, the formula for a horizontal compression
is:
42 2 Algebra of Functions

f (dx) (2.14)

Where d > 1.
(iii) Compressions: Compressions of a function involve reducing its amplitude or length in the Cartesian
plane. Please refer to the previous description under stretches for the mathematical formulas.
(iv) Reflections: Reflections of a function involve flipping it with respect to an axis in the Cartesian plane. A
reflection with respect to the x-axis is achieved by changing the sign of the function f (x), resulting in a
reflection across the x-axis. Mathematically, the formula for a reflection across the x-axis is:

− f (x) (2.15)

A reflection with respect to the y-axis is performed by changing the sign of the independent variable x,
leading to a reflection across the y-axis. Mathematically, the formula for a reflection across the y-axis is:

f (−x) (2.16)

2.4. The Limit Operator


The concept of the limit is one of the most fundamental concepts in mathematics and is central to numerous
subfields. Its historical development dates back to the 18th and 19th centuries, when mathematicians such as
Augustin-Louis Cauchy, Karl Weierstrass, and Richard Dedekind made significant contributions.
Mathematics relied for a long time on the intuitive concept of a limit but lacked a rigorous formulation. The
formal foundations of calculus, which provide a rigorous mathematical treatment of the concept of a limit,
were established in the 19th century.
Augustin-Louis Cauchy, widely regarded as one of the founders of mathematical analysis, was among the
first to formulate a precise definition of the limit in terms of sequences. Cauchy established criteria for the
convergence of a sequence and lay the groundwork for the study of the properties of real functions in his
work. His contributions established the foundations for contemporary mathematical analysis and the rigorous
study of limits.
Subsequently, mathematicians such as Karl Weierstrass and Richard Dedekind devised the necessary tools
to comprehend and manipulate functions and series in terms of limits. Particularly, Weierstrass introduced
the concept of continuity and demonstrated significant theorems concerning the limit of functions.
The limit operator has evolved into one of the most useful and versatile mathematical instruments. Among
others, it is utilized in differential and integral calculus, mathematical analysis, number theory, and theoretical
physics. Essential for comprehending fundamental concepts such as derivatives, integrals, and continuity, it
permits the study of the behavior of functions and sequences close to points or at infinity.
The limit operator is a fundamental concept in mathematics that has evolved from an intuitive formulation
to a rigorous formalization over centuries. Its significance stems from its ability to describe and analyze the
behavior of functions and sequences, as well as its wide spectrum of mathematical applications.
2.4 The Limit Operator 43

The limit operator is a fundamental mathematical concept that allows us to investigate the behavior of
functions as their input approaches a specific value or tends to infinity. It plays a crucial role in calculus
and mathematical analysis, enabling us to understand fundamental concepts such as continuity, derivatives,
and integrals.

Definition 2.1. Given a function f (x) defined on an interval containing the real number a, we say that the
limit of f (x) as x approaches a is equal to L, denoted as lim f (x) = L, if for any value ε > 0, there exists a
x→a
value δ > 0 such that for all x in the domain of f (x) satisfying 0 < |x − a|< δ , we have | f (x) − L|< ε .
In other words, the limit of a function f (x) at x = a is L if, for any positive margin of error ε , we can find an
interval around x = a (excluding possibly the point x = a) where the values of f (x) are arbitrarily close to L
as long as x is close enough to a.
44 2 Algebra of Functions

f (x)

limx→x− f (x) = limx→x+ f (x) = L


0 0
f (x0 )

x0

Figure 2.8: Graphical representation of limx→x− f (x) = limx→x+ f (x) = L. Note that f (x0 ) is on the graph of
0 0
the f (x) function, whereas L is on the vertical axis.
2.4 The Limit Operator 45

In this figure (Fig. 2.8), the function f (x) is represented in blue, and the horizontal blue line represents the
limit L to which f (x) approaches as x approaches the point a. The interval around x = a is marked with
vertical lines, and a margin of error ε is shown in the vertical direction. The point (a, L) represents the limit
of f (x) at x = a.
The following is an example of how to solve the limit, using the definition.

Example 2.17. Given the limit:


lim 3x + 1 = 7 (2.17)
x→2

We want to show that for any ε > 0, there exists a δ > 0 such that if 0 < |x − 2|< δ , then |3x + 1 − 7|< ε .
Solution 2.17. Let’s take an arbitrary ε > 0. We want to find a corresponding δ > 0. For simplicity, let’s
assume ε < 4.
Given ε < 4, we’ll choose δ = ε3 . Then, if 0 < |x − 2|< δ , we can perform the following algebraic
manipulations:

|3x + 1 − 7| = |3x − 6| (2.18)


= 3|x − 2| (2.19)
< 3δ (2.20)
ε 
=3 (2.21)
3
=ε (2.22)

Therefore, we have shown that if 0 < |x − 2|< δ = ε3 , then |3x + 1 − 7|< ε .


In conclusion, For any value of ε > 0, we can find a corresponding δ such that if 0 < |x − 2|< δ , then
|3x + 1 − 7|< ε . This verifies that the limit limx→2 3x + 1 = 7 holds using the epsilon-delta method.
Example 2.18. Consider the constant function f (x) = c, where c is a real constant. We want to prove that the
limit of f (x) as x approaches a is equal to c, i.e., limx→a f (x) = c.

Proof. Given the constant function f (x) = c, where c is a real constant.


To prove: lim f (x) = c
x→a
For any ε > 0, we wish to find a δ > 0 such that if 0 < |x − a|< δ , then | f (x) − c|< ε .
Since f (x) = c for all x, we have:
| f (x) − c|= |c − c|= 0

Clearly, for any ε > 0, 0 is always less than ε .


Therefore, for any choice of ε > 0, if 0 < |x − a|, then | f (x) − c|< ε .
It follows that for any ε > 0, there exists a δ > 0 (in fact, any positive value of δ will work) such that if
0 < |x − a|< δ , then | f (x) − c|< ε .
Thus, limx→a f (x) = c.
46 2 Algebra of Functions

The following case shows that the evaluation of the limit is not to evaluate the function at the point x0 .
x2 − 1
Example 2.19. Let’s consider the function f (x) = . We want to find the limit of f (x) as x approaches
x−1
1.
Solution 2.18. To evaluate the limit, we substitute x = 1 into the function:

x2 − 1 12 − 1 0
lim = = . (2.23)
x→1 x − 1 1−1 0
0
We have obtained an indeterminate form of , which indicates that further simplification is required. By
0
factoring the numerator, we can rewrite the function as:

(x + 1)(x − 1)
f (x) = . (2.24)
x−1

Now, we can cancel out the common factor of (x − 1):

f (x) = x + 1. (2.25)

Taking the limit again, we have:

lim f (x) = lim (x + 1) = 1 + 1 = 2. (2.26)


x→1 x→1

Therefore, the limit of the function f (x) as x approaches 1 is equal to 2.

In this example, the limit operator allows us to determine the value that a function approaches as the input
approaches a specific value.

2.4.1. Properties of the Limit Operator


In the study of function limits, it is crucial to comprehend the behavior of algebraic operations when applied
to particular limits. The limit of a sum of individual limits is equal to the sum of the individual limits,
according to a fundamental principle. This concept enables us to streamline limit calculations using well-
established algebraic properties.
Note that the limit operator theorems presented below presume that the individual limits of the involved
functions exist. The existence of limits is assumed by these theorems, which allow us to deduce additional
properties and relationships between the limits of more complex functions.
Keep in mind that limit operator theorems are foundational calculus tools that enable us to manipulate
the limits of functions using elementary algebraic properties. These theorems are extensively used to
simplify limit calculations and are indispensable for the development of more advanced ideas in the study of
mathematical analysis.
2.4 The Limit Operator 47

(i) Constant Limit Theorem:


lim c = c (2.27)
x→a

Where c is a constant.
This theorem states that the limit of a constant function c as x approaches the value a is simply the
constant value c.

(ii) Identity Limit Theorem:


lim x = a (2.28)
x→a

This theorem states that the limit of the variable x as x approaches the value a is equal to a itself.

(iii) Sum Limit Theorem:


lim [ f (x) + g(x)] = lim f (x) + lim g(x) (2.29)
x→a x→a x→a

This theorem states that the limit of the sum of two functions f (x) and g(x) as x approaches the value a
is equal to the sum of the individual limits of f (x) and g(x) as x approaches a.

(iv) Product Limit Theorem:


lim [ f (x) · g(x)] = lim f (x) · lim g(x) (2.30)
x→a x→a x→a

This theorem states that the limit of the product of two functions f (x) and g(x) as x approaches the value
a is equal to the product of the individual limits of f (x) and g(x) as x approaches a.

(v) Quotient Limit Theorem:


 
f (x) limx→a f (x)
lim = if lim g(x) 6= 0 (2.31)
x→a g(x) limx→a g(x) x→a

This theorem states that the limit of the quotient of two functions f (x) and g(x) as x approaches the
value a is equal to the quotient of the individual limits of f (x) and g(x) as x approaches a, provided that
the limit of g(x) as x approaches a is not zero.

Example 2.20. Find the limit of the function f (x) = 3x2 − 2x + 1 as x approaches 2.

Solution 2.20.
lim (3x2 − 2x + 1) = lim 3x2 − lim 2x + lim 1 (2.32)
x→2 x→2 x→2 x→2

Using the properties of limits, we can calculate each limit separately:

lim 3x2 = 3 · lim x2 = 3 · (22 ) = 12 (2.33)


x→2 x→2
48 2 Algebra of Functions

lim 2x = 2 · lim x = 2 · 2 = 4 (2.34)


x→2 x→2

lim 1 = 1 (2.35)
x→2

Then, we add up the calculated limits:

lim (3x2 − 2x + 1) = 12 − 4 + 1 = 9 (2.36)


x→2

Therefore, the limit of the function f (x) as x approaches 2 is equal to 9.

Example 2.21. Resolve the next limit.

lim (3x − 1) (2.37)


x→2

Solution 2.21. To find this limit, substitute the value x = 2 directly into the expression:

lim (3x − 1) = 3(2) − 1


x→2
= 6−1
=5

Thus, the limit is:

lim (3x − 1) = 5 (2.38)


x→2

These examples demonstrate how the properties of limits can be used to simplify the evaluation of limits.

2.5. Sequences and Series


Sequences and series are fundamental mathematical concepts that have been crucial throughout history. In
numerous branches of mathematics, especially differential and integral calculus, their study and comprehension
have been of critical importance. Mathematicians have been interested in numerical patterns and how
numbers can relate and evolve since antiquity. Sequences, which are ordered sequences of numbers, permit
us to model and characterize the variation of a quantity as the index changes. These sequences can be used to
analyze natural, physical, or economic phenomena, and their rigorous study has contributed to advancements
in a variety of fields. Sequences provide the foundation for defining concepts such as limits and continuity
in differential calculus. Understanding how functions behave near a specific point and how functions can be
approximated using convergent sequences requires the study of sequence convergence.
Alternatively, series, which are infinite sums of elements from a sequence, play a crucial role in integral
calculus. Taylor series have enabled more accurate approximations of functions and efficient methods for
2.5 Sequences and Series 49

calculating areas, volumes, and other quantities in math and physics. The study of sequences and series has
not only influenced the development of calculus, but also mathematical analysis, number theory, geometry,
and function theory. In addition, it has been utilized in disciplines such as physics, engineering, economics,
and computer science. Sequences and series are fundamental tools in mathematics that have been used
throughout history to model phenomena, understand the behavior of functions, and solve problems in various
disciplines. Their study has led to significant advancements in differential and integral calculus, and their
importance extends to many other branches of mathematics and related fields.
A sequence is an ordered succession of real numbers, while a series is the sum of the terms of a sequence.
These operators are closely related to the concept of limit.
For a sequence an , where n is an index indicating the position of the term in the sequence, it can be said that
the sequence converges if there exists a real number L such that the terms of the sequence get closer and
closer to L as n grows indefinitely. Mathematically, this can be expressed as:

lim an = L (2.39)
n→∞

1
Example 2.22. Let’s consider the sequence . As n grows, the terms of the sequence get closer and closer to
n
zero. In this case, we can say that the sequence converges to zero, that is:
1
lim =0 (2.40)
n→∞ n
On the other hand, a sequence may diverge if there is no defined limit towards which the terms of the
sequence converge. In other words, the terms of the sequence can grow or decrease without bound.

Example 2.23. Let’s consider the sequence n. In this case, as n grows, the terms of the sequence also grow
indefinitely. There is no real number to which the sequence can converge, therefore, we say that the sequence
diverges.

Series, on the other hand, are related to sequences through the summation of the terms of a sequence.
If we consider a series ∑∞n=1 an , where an is the n-th term of the sequence, the series converges if the
sequence of partial sums converges. Partial sums are obtained by summing the first n terms of the sequence.
Mathematically, this can be expressed as:

lim Sn = S (2.41)
n→∞

Where Sn represents the n-th partial sum and S is the limit of the partial sums. If the sequence of partial sums
does not converge, then the series diverges.

1
Example 2.24. Let’s consider the series ∑ n . The partial sums of this series are known as the partial series
n=1
and can be expressed as:
n
1
Sn = ∑k (2.42)
k=1
50 2 Algebra of Functions

If we observe the values of Sn as n grows, we can see that the partial series grows without bound, i.e., it does
not converge to a specific real number. Therefore, the series diverges.

The operators of sequences and series of real numbers are intrinsically linked to the concept of limit.
Convergence refers to the existence of a limit to which the terms of the sequence or the partial sums get
closer, while divergence occurs when there is no defined limit or the sequences grow or decrease without
bound.

Example 2.25. Consider the following sequence defined by the recursive formula:

an = an−1 + 2, a1 = 3 (2.43)

The sequence starts with the initial term a1 = 3, and each subsequent term is obtained by adding 2 to the
previous term.

Solution 2.22. To solve this sequence, we can use the provided recursive formula. Starting with the initial
term a1 = 3, we can calculate the subsequent terms of the sequence by applying the formula:

a2 = a2−1 + 2 = a1 + 2 = 3 + 2 = 5 (2.44)

a3 = a3−1 + 2 = a2 + 2 = 5 + 2 = 7 (2.45)

a4 = a4−1 + 2 = a3 + 2 = 7 + 2 = 9 (2.46)

Continuing in this manner, we can calculate the successive terms of the sequence:

a5 = 11, a6 = 13, a7 = 15, a8 = 17, a9 = 19, a10 = 21 (2.47)

Therefore, the solution of the sequence for n = 10 is a10 = 21.

Example 2.26. Consider the following arithmetic series:


n
S= ∑ (2k − 1), n=5 (2.48)
k=1

Here, n represents the number of terms we want to sum, and (2k − 1) represents each term of the series.

Solution 2.23. To solve this series, we first need to calculate the individual terms (2k − 1) for each value of
k from 1 to 5.
When k = 1, the first term of the series is (2 · 1 − 1) = 1.
When k = 2, the second term of the series is (2 · 2 − 1) = 3.
When k = 3, the third term of the series is (2 · 3 − 1) = 5.
When k = 4, the fourth term of the series is (2 · 4 − 1) = 7.
2.5 Sequences and Series 51

When k = 5, the fifth term of the series is (2 · 5 − 1) = 9.


Now, we can sum these terms to obtain the result of the series:

S = 1 + 3 + 5 + 7 + 9 = 25 (2.49)

Therefore, the solution of the series for n = 5 is S = 25.

Fundamental criteria for determining whether a sequence or series is convergent are:


Criterion of the sequence or divergence criterion: If a sequence does not converge to zero, then the
corresponding series is divergent. In other words, if the terms of the sequence do not approach zero as
the index increases, the series does not converge. Limit criterion: If the limit of the terms of the sequence
is nonzero, then the series diverges. On the contrary, if the limit of the terms is zero, the criterion does not
allow us to conclude whether the series converges or not. Comparison criterion: If a series is less than or
equal to another series that we already know converges, then the original series also converges. Similarly,
if a series is greater than or equal to another series that we already know diverges, then the original series
also diverges. This criterion is useful for comparing the series in question with other series whose behavior
is known. It These techniques are just some of the many available to analyze the convergence or divergence
of a series or sequence. In more complex cases, additional techniques such as the ratio test, the root test,
and the integral test, among others, may be required. Furthermore, it is crucial to understand the conditions
of applicability of each criterion and adapt them according to the specific characteristics of the series or
sequence being analyzed. The convergence or divergence of a series or sequence is a fundamental concept
in mathematics and has important implications in various areas, such as calculus and mathematical analysis.
The ability to determine whether a series or sequence converges or diverges is essential to understanding and
applying these concepts to mathematical and scientific problems.

Example 2.27. Consider the sequence given by an = n. We want to determine whether the corresponding

series ∑ an converges or diverges.
n=1

Solution 2.25. To apply the criterion of the sequence, we evaluate the limit of the sequence an as n
approaches infinity:
lim an = lim n = ∞. (2.50)
n→∞ n→∞

Since the limit is not zero, according to the criterion of the sequence, the corresponding series ∑∞
n=1 an is
divergent.
1
Example 2.28. Consider the sequence given by bn = . We want to determine whether the corresponding
n2

series ∑ bn converges or diverges.
n=1

Solution 2.26. To apply the limit criterion, we evaluate the limit of the sequence bn as n approaches infinity:
1
lim bn = lim = 0. (2.51)
n→∞ n→∞ n2
52 2 Algebra of Functions

The limit is zero; however, the limit criterion does not allow us to conclude whether the series converges or
not. Another criterion or technique is required to determine the convergence or divergence of the series.

1
Example 2.29. Consider the series ∑ cn , where cn = n3 + n2 . We want to determine whether this series
n=1
converges or diverges.

Solution 2.27. To determine the convergence of the series



1
∑ n3 + n2
n=1

we can use the comparison test. If the series is comparable to the series ∑ n12 , which is known to converge
(p-series with p > 1), then it also converges.
For n ≥ 1, n3 ≥ n2 (because both n3 and n2 are positive and n3 grows faster than n2 ).
Therefore,
n3 + n2 ≥ n2
Dividing both sides by the product n2 (n3 + n2 ) gives:

1 1
≤ 2
n3 + n2 n

1
Since the series ∑ n12 converges (it’s a p-series with p = 2), and our given series ∑ n3 +n 2 has terms that are less
than or equal to those of this convergent series for all n, our given series also converges by the comparison
test.

These examples illustrate how to apply each of the convergence criteria to determine whether a sequence
or series converges or diverges. Remember that the appropriate choice of criterion will depend on the
characteristics of the sequence or series being analyzed.

2.6. Conclusions
In conclusion, in this chapter, we have explored several fundamental topics in the study of relations and real-
variable functions, as well as the limit operator and sequences and series. We have discovered how relations
can establish connections and associations between elements of sets, and how functions allow us to map
values of one variable to other values.
Furthermore, we have analyzed the essential properties of different types of functions, such as polynomial,
exponential, trigonometric, and logarithmic functions. Each type of function possesses distinct characteristics
and unique behaviors that enable us to model and understand various phenomena in the real world.
The limit operator has been a central concept in our study as it allows us to analyze the behavior of functions
and sequences as they approach certain values. We have explored properties and techniques for computing
limits, providing us with a solid foundation for analyzing and understanding functions at critical points.
2.6 Conclusions 53

Lastly, we have addressed the topic of sequences and series, which are infinite sequences of numbers.
We have studied the convergence and divergence of sequences, as well as properties of series such as
convergence and the summation of infinite terms.
Collectively, this chapter has provided us with a deeper understanding of relations and real-variable
functions, the limit operator, and sequences and series. These concepts and tools are fundamental in
mathematical analysis and allow us to explore and comprehend the world of functions and the phenomena
surrounding them.
54 2 Algebra of Functions

2.7. Exercises
Exercise 2.1. Consider the function f : R → R defined by f (x) = 2x+1. Prove that f is an injective function.

Exercise 2.2. Given the function g(x) = x2 , find the reflection of g(x) with respect to the x-axis.

Exercise 2.3. Let f (x) = 2x and g(x) = x + 3. Find the composition g( f (x)).

Exercise 2.4. Calculate the following limit:

x2 − 4
lim (2.52)
x→2 x − 2

Exercise 2.5. Evaluate the sum of the series:


5
∑ 3n (2.53)
n=1

Exercise 2.6. Consider the sequence defined by an = 2n . Calculate the value of a10 .

Exercise 2.7. Find the limit of the sequence:


1
lim (2.54)
n→∞ n2

Exercise 2.8. Determine the sum of the series:



1
∑ 2n (2.55)
n=1

1
Exercise 2.9. Consider the sequence defined by an = . Calculate the value of a5 .
n
Exercise 2.10. Consider a right triangle with one angle at vertex C. The lengths of the sides adjacent to the
right angle are a and b, and the hypotenuse has length c. Calculate the area of the triangle using the concept
of limits.
Chapter 3
Algebra of Integer Numbers

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract This chapter provides a comprehensive examination of fundamental number theory and modular
arithmetic concepts. The greatest common divisor and its applications, the least common multiple and its
properties, congruences and their function in modular arithmetic, and the arithmetic module are explored
in detail. In addition, we examine linear congruences and their solutions, investigating their applications
in cryptography and problem-solving. In addition, the significance of prime numbers, including their
fundamental function in number theory and their relationship to the unique factorization theorem, is
discussed. The theorem of Euclid is presented, emphasizing its elegant proof and its significance for
comprehending the infinity of prime numbers. The importance of Bézout’s theorem in solving linear
diophantine equations and its connection to the extended Euclidean algorithm are emphasized as we
investigate this theorem. Finally, we present the Chinese Remainder Theorem, an essential result that
permits simultaneous resolution of a system of congruences. Using its applications in cryptography, number
theory, and modular arithmetic, we demonstrate its practical importance. The purpose of this chapter is to
provide a thorough comprehension of these foundational concepts in number theory, modular arithmetic,
and their applications. Through crystal-clear explanations and illustrative examples, readers will acquire a
thorough understanding of these fundamental topics and their significance in a variety of mathematical and
cryptographic contexts.

Keywords: gcd, lcm, arithmetic module, Bézout’s theorem, chinese remainder theorem, congruences,
cryptography, Euclid’s theorem, fundamental theorem of arithmetic, greatest common divisor, integer
numbers, least common multiple, linear congruences, number theory, prime numbers.

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

55
56 3 Algebra of Integer Numbers

3.1. Introduction
In this chapter, we embark on an enthralling exploration of integer algebra and number theory. We
will investigate the concepts and theorems that serve as the foundation for these intriguing branches of
mathematics. The scope of our investigation will include the greatest common divisor, the least common
multiple, congruences, arithmetic modulo, linear congruences, prime numbers, the fundamental theorem of
arithmetic, Euclid’s theorem, Bézout’s theorem, the Chinese remainder theorem, and their applications in
number theory.
In the mathematical landscape, integer algebra maintains a prominent position, serving as the foundation
for numerous mathematical disciplines. It offers the tools and techniques necessary to comprehend and
manipulate integers, paving the way for more in-depth research into their properties and relationships.

3.2. Integer Numbers


Integers are a fundamental part of mathematics and represent an infinite set of numbers that includes both
positive and negative numbers, as well as zero. Integers are used for counting, ordering, and performing
more advanced mathematical operations.

Z = ..., −3, −2, −1, 0, 1, 2, 3, ... (3.1)

Definition 3.1. The set of integers, represented by the symbol Z, is defined as the infinite set that includes
negative numbers, zero, and positive numbers. In notation, integers are written as: {..., −3, −2, −1, 0, 1, 2, 3, ...}.

3.3. Greatest Common Divisor


The greatest common divisor (gcd) of two or more integers is the largest integer that divides all the given
numbers without leaving a remainder. To calculate the gcd, we will use the Euclidean algorithm, which is
based on the observation that the gcd of two numbers does not change if we repeatedly subtract the smaller
number from the larger number until we get zero.
We define the gcd of two integers a and b as gcd(a, b). We can use the notation gcd(a, b) to represent the
gcd of two numbers or even more numbers, if necessary.
The Euclidean algorithm can be used to calculate the gcd of two numbers a and b as follows:

(i) Divide a by b and obtain the quotient q and the remainder r.

a = bq + r (3.2)
3.3 Greatest Common Divisor 57

(ii) If the remainder r is equal to zero, then the gcd is equal to b. That is,

gcd(a, b) = b (3.3)

In this case, we have found the gcd and we can stop.


(iii) If the remainder r is not zero, then we perform another iteration of the Euclidean algorithm, but now
with b and r instead of a and b. That is,

gcd(a, b) = gcd(b, r) (3.4)

(iv) Then, we go back to step 1 using the new values of b and r.

(v) We repeat steps (i) and (iii) until we obtain a zero remainder. At that point, the last non-zero divisor will
be the gcd of the original numbers a and b.

Example 3.1. Let’s calculate the gcd of a = 84 and b = 18 using the Euclidean algorithm.

Solution 3.1. We apply the steps:

(i) Divide 84 by 18: 84 = 18 · 4 + 12. Here, q = 4 and r = 12.

(ii) The remainder r is not zero, so we proceed to the next step.

(iii) Now, we use b = 18 and r = 12 instead of a and b.

(iv) Divide 18 by 12: 18 = 12 · 1 + 6. Here, q = 1 and r = 6.

(v) Again, the remainder r is not zero, so we repeat step (iii).

(vi) Divide 12 by 6: 12 = 6 · 2 + 0. Here, q = 2 and r = 0.

Finally, we have obtained a zero remainder. The last non-zero divisor is 6, so the gcd of 84 and 18 is 6.
In mathematical notation, this can be written as:

gcd(84, 18) = 6 (3.5)

The Euclidean algorithm guarantees that we will always find the gcd of two positive integers in a finite
number of steps. Furthermore, this algorithm can be extended using the extended Euclidean algorithm to
find the Bezout’s coefficients.
The extended Euclidean algorithm allows us to find two integers x and y that satisfy Bezout’s identity, which
states that the gcd of two integers a and b can be written as a linear combination of them:

gcd(a, b) = ax + by (3.6)
58 3 Algebra of Integer Numbers

The extended Euclidean algorithm is based on performing the successive divisions used in the Euclidean
algorithm and working backwards to find the coefficients x and y.
The Euclidean algorithm is an efficient method for calculating the gcd of two or more integers. It provides
a mathematical solution to finding the greatest common divisor and can be extended using the extended
Euclidean algorithm to find the Bezout’s coefficients.

3.4. Least Common Multiple


The least common multiple lcm of two positive integers is defined as the smallest number that is a multiple
of both numbers. To calculate the lcm of two integers a and b, we can follow the following steps:
First, we factorize each number into their prime factors.

a = px11 · px22 · px33 · . . . (3.7)

b = qy11 · qy22 · qy33 · . . . (3.8)

Where p1 , p2 , p3 , · · · and q1 , q2 , q3 , . . . are the prime factors of a and b respectively.


Next, we take all the unique prime factors that appear in either a or b, and raise each factor to the maximum
power it appears in either factorization. This will give us the prime factorization of the lcm.
max(x1 ,y1 ) max(x2 ,y2 ) max(x3 ,y3 )
lcm(a, b) = p1 · p2 · p3 ·... (3.9)

Where max(xi , yi ) represents the maximum of xi and yi .


: Finally, we can compute the lcm by multiplying all the prime factors obtained in the previous step.
max(x1 ,y1 ) max(x2 ,y2 ) max(x3 ,y3 ) max(x1 ,y1 ) max(x2 ,y2 ) max(x3 ,y3 )
lcm(a, b) = p1 · p2 · p3 · · · · · q1 · q2 · q3 ·... (3.10)

Example 3.2. Let’s illustrate this with an example. Suppose we want to find the lcm of a = 12 and b = 18.

Solution 3.2. Factorizing a and b, we have:


a = 22 · 31 (3.11)
1 2
b = 2 ·3 (3.12)

Taking the maximum powers of each prime factor, we get:

lcm(12, 18) = 22 · 32 (3.13)

Multiplying the prime factors, we obtain:


3.5 Fundamental Theorem of Arithmetic 59

lcm(12, 18) = 22 · 32 = 4 · 9 = 36 (3.14)

Therefore, the lcm of 12 and 18 is 36.

This method can be extended to calculate the lcm of more than two integers by including their respective
prime factorizations and taking the maximum powers of each prime factor.

3.5. Fundamental Theorem of Arithmetic


The Fundamental Theorem of Arithmetic [12, 13] states that any positive integer greater than 1 can be
uniquely expressed as a product of prime factors, i.e., as a multiplication of prime numbers. This unique
factorization is unique except for the order of the factors.
To formalize the theorem, let’s define some preliminary concepts. A prime number is an integer greater than
1 that has only two divisors: itself and 1. On the other hand, a composite number is an integer greater than 1
that is not prime, meaning it can be factored into two or more positive integers.
Now, we can formally state the Fundamental Theorem of Arithmetic:
For every n > 1, there exists a unique factorization of the form:
a
n = pa11 · pa22 · pa33 · . . . · pk k (3.15)

Where p1 , p2 , p3 , · · · , pk are distinct prime numbers, and a1 , a2 , a3 , · · · , ak are non-negative integer exponents.
This means that any integer n greater than 1 can be uniquely decomposed into a product of powers of prime
numbers. The exponents indicate the number of times each prime appears in the factorization. Additionally,
the factorization is unique in the sense that if we have two different factorizations of n, they will have the
same prime factors, although the order of the factors may vary.

Theorem 3.1. The Fundamental Theorem of Arithmetic mentioned that every positive integer greater than
1 can be expressed uniquely (up to the order of factors) as a product of prime numbers.

Proof. 1. Existence:
Every integer n greater than 1 is either a prime number or can be expressed as a product of prime numbers.
For n = 2, it is prime.
Assume, for all positive integers k such that 2 ≤ k < n, k can be written as a product of primes.
If n is prime, then it is trivially expressed as a product of primes (just itself). If n is not prime, it can be
expressed as n = ab where 1 < a, b < n. By our inductive assumption, both a and b can be expressed as a
product of primes. Therefore, n can be expressed as a product of primes.
By the principle of mathematical induction, every integer n > 1 can be expressed as a product of prime
numbers.
60 3 Algebra of Integer Numbers

2. Uniqueness: Suppose we have two different prime factorizations for the same integer n:
a b
n = pa11 pa22 · · · pk k = qb11 qb22 · · · ql l

Where the p’s and the q’s are prime numbers and the a’s and b’s are positive integers.
Taking the first prime p1 from the left side, since it divides the left side, it must also divide the right side. But
prime numbers only divide numbers that have them as factors. Therefore, p1 must be one of the q primes on
the right side. Without loss of generality, assume p1 = q1 .
Now, because pa11 divides the left side, it also divides the right side. Therefore, a1 = b1 and we can cancel
pa11 = qb11 from both sides.
Continuing in this manner, we can show that all the primes on the left side match with the primes on the
right side and that their powers must also match, ensuring the uniqueness of the factorization.
Every positive integer greater than 1 can be uniquely factored (up to the order of factors) into prime numbers.

Example 3.3. Fundamental Theorem of Arithmetic. Let’s take n = 84. First, we seek the prime factors of 84.
We can successively divide by prime numbers until we obtain the factorization:

84 = 2 · 42 = 2 · 2 · 21 = 2 · 2 · 3 · 7 (3.16)

Here, we have obtained the unique factorization of 84 as 22 · 31 · 71 . We can see that the prime factors are 2,
3, and 7, and the associated exponents are 2, 1, and 1, respectively.

The Fundamental Theorem of Arithmetic states that any integer greater than 1 can be decomposed uniquely
into a product of prime factors. This factorization is unique, except for the order of the factors.

Example 3.4. First, we start by dividing 210 by the smallest prime number, which is 2. We perform the
division:
210
= 105 (3.17)
2
The quotient is 105. Now, we divide 105 by 2 again:
105
= 52.5 (3.18)
2
Since 52.5 is not an integer, we move on to the next prime number, which is 3. We divide 105 by 3:
105
= 35 (3.19)
3
The quotient is 35. We continue dividing by 3:
35
= 11.67 (3.20)
3
3.6 Congruences 61

Again, 11.67 is not an integer, so we move on to the next prime number, which is 5. We divide 35 by 5:
35
=7 (3.21)
5
Now, we have obtained a quotient of 7. Since 7 is a prime number, we cannot divide it further.
Therefore, the prime factorization of 210 is:

210 = 2 × 3 × 5 × 7 (3.22)

According to the Fundamental Theorem of Arithmetic, this factorization is unique. It tells us that 210 can be
expressed as a product of its prime factors: 2, 3, 5, and 7.

3.6. Congruences
Congruence [14] is a relation between two integers based on divisibility by a third number called the
modulus. Formally, for two integers a and b and a positive integer n, we say a is congruent to b modulo
n and denote it as:

a ≡ b (mod n) (3.23)

This means that a and b leave the same remainder when divided by n. In other words, if the difference
between a and b is exactly divisible by n, then a and b are congruent modulo n.

Example 3.5. Provide an example of congruence.

Solution 3.3.
17 ≡ 3 (mod 7) (3.24)

In this case, both 17 and 3 leave a remainder of 3 when divided by 7. Therefore, they are congruent modulo
7.

Congruence has several important properties:

(i) Reflexivity: For any integer a and modulus n, a is congruent to itself modulo n:

a ≡ a (mod n) (3.25)

(ii) Symmetry: If a is congruent to b modulo n, then b is also congruent to a modulo n:

a≡b (mod n) ⇒ b ≡ a (mod n) (3.26)

(iii) Transitivity: If a is congruent to b modulo n and b is congruent to c modulo n, then a is congruent to c


modulo n:
a ≡ b (mod n) and b ≡ c (mod n) ⇒ a ≡ c (mod n) (3.27)
62 3 Algebra of Integer Numbers

Congruence can also be operated on with addition, subtraction, multiplication, and exponentiation. For
example:
(iv) Addition: If a is congruent to b modulo n and c is congruent to d modulo n, then the sum of a and c is
congruent to the sum of b and d modulo n:

a ≡ b (mod n) and c≡d (mod n) ⇒ a + c ≡ b + d (mod n) (3.28)

(v) Multiplication: If a is congruent to b modulo n and c is congruent to d modulo n, then the product of a
and c is congruent to the product of b and d modulo n:

a≡b (mod n) and c≡d (mod n) ⇒ ac ≡ bd (mod n) (3.29)

(vi) Exponentiation: If a is congruent to b modulo n, then a raised to the k−th power is congruent to b raised
to the k−th power modulo n, for any integer k:

a ≡ b (mod n) ⇒ ak ≡ bk (mod n) (3.30)

These properties allow us to efficiently solve congruence problems. Here are some numerical examples:
Example 3.6. Find the remainder of 210 when divided by 7.
Using the congruence theorem, we can simplify the calculation using congruence modulo 7:

21 ≡ 2 (mod 7) (3.31)
2
2 ≡4 (mod 7) (3.32)
3
2 ≡1 (mod 7) (3.33)
4
2 ≡2 (mod 7) (3.34)
5
2 ≡4 (mod 7) (3.35)
26 ≡ 1 (mod 7) (3.36)
7
2 ≡2 (mod 7) (3.37)
28 ≡ 4 (mod 7) (3.38)
9
2 ≡1 (mod 7) (3.39)
210 ≡ 2 (mod 7) (3.40)

Therefore, the remainder of 210 when divided by 7 is 2.

Example 3.7. Find all positive integers less than or equal to 20 that are congruent to 3 modulo 7.
Solution 3.7. To solve this, we can check which numbers leave a remainder of 3 when divided by 7. The
integers congruent to 3 modulo 7 are:

3, 10, 17 (3.41)
Therefore, the positive integers less than or equal to 20 that are congruent to 3 modulo 7 are 3, 10, and 17.
3.7 Arithmetic Module 63

3.7. Arithmetic Module


Arithmetic modulo, denoted as a mod b, is an operation that calculates the remainder when the integer a is
divided by the positive integer b. In other words, it represents the ”clock arithmetic” concept, where numbers
wrap around a certain value.
Mathematically, the arithmetic modulo operation can be defined as follows:
jak
a mod b = a − b (3.42)
b

Where ⌊x⌋ represents the floor function, which rounds down the value of x to the nearest integer.

Example 3.8. Consider a = 17 and b = 5. We want to find the value of a modulo b, which can be written as
17 mod 5.

Solution 3.8. Using the formula, we have:


 
17
17 mod 5 = 17 − 5 (3.43)
5

Calculating the floor function, we get:

17 mod 5 = 17 − 5 ⌊3.4⌋ (3.44)

Since ⌊3.4⌋ = 3, we can simplify the expression to:

17 mod 5 = 17 − 5 · 3 (3.45)

Evaluating the expression, we find:

17 mod 5 = 17 − 15 = 2 (3.46)

Therefore, when 17 is divided by 5, the remainder is 2. Hence, 17 mod 5 = 2.

Example 3.9. Let’s consider a = 28 and b = 7.

Solution 3.9. We want to find the value of a modulo b, i.e., 28 mod 7.


Using the formula, we have:
 
28
28 mod 7 = 28 − 7 (3.47)
7

Calculating the floor function, we get:

28 mod 7 = 28 − 7 ⌊4⌋ (3.48)

Since ⌊4⌋ = 4, the expression simplifies to:


64 3 Algebra of Integer Numbers

28 mod 7 = 28 − 7 · 4 (3.49)

Evaluating the expression, we find:

28 mod 7 = 28 − 28 = 0 (3.50)

Therefore, when 28 is divided by 7, the remainder is 0. Thus, 28 mod 7 = 0.

3.8. Linear Congruences


A linear congruence has the general form ax ≡ b (mod m), where a, b, and m are integers, and m must be
a positive number. The goal is to find all the solutions for x that satisfy the congruence.
The process of solving a linear congruence involves several steps:

(i) Check if the congruence has a solution. We check if the coefficient a and the modulus m are coprime,
meaning that gcd(a, m) = 1, where gcd(a, m) denotes the greatest common divisor of a and m. If this
condition is not satisfied, the congruence may not have a solution.

(ii) Find the multiplicative inverse of a modulo m. If the congruence passes Step 1, we calculate the
multiplicative inverse of a modulo m. This involves finding an integer a−1 such that a·a−1 ≡ 1 (mod m).

(iii) Multiply both sides of the congruence by the multiplicative inverse. We multiply both sides of the
congruence ax ≡ b (mod m) by the multiplicative inverse a−1 obtained in the previous step. This gives
us the equivalent congruence x ≡ b · a−1 (mod m).

(iv) Simplify the congruence and find the solutions. We simplify the resulting congruence and find the
solutions for x. If there is a unique solution, we express it in the form x ≡ solution (mod m). If there are
multiple solutions, we express them in the form of an equivalence class.

Example 3.10. Let’s solve the linear congruence 7x ≡ 3 (mod 10).

Solution 3.10. Given the congruence:


7x ≡ 3 (mod 10)

We need to find the multiplicative inverse of 7 modulo 10. This is a number b such that:

7b ≡ 1 (mod 10)

Checking for the smallest positive integers, we find:

7 × 3 ≡ 21 ≡ 1 (mod 10)
3.9 Euclid’s Theorem 65

Thus, 3 is the multiplicative inverse of 7 modulo 10. Multiplying both sides of our initial congruence by this
inverse, we get:
7x × 3 ≡ 3 × 3 (mod 10)
which gives:
21x ≡ 9 (mod 10)

Since 21x ≡ x (mod 10), this can be further simplified to:

x ≡ 9 (mod 10)

Thus, the solution to the congruence is:


x ≡ 9 (mod 10)
or in general form, x = 9 + 10k for any integer k.

3.9. Euclid’s Theorem


Euclid’s theorem [15] states that if a and b are two positive integers, then there exists a non-negative integer
called the greatest common divisor (gcd) of a and b, denoted as gcd(a, b), which satisfies the following
properties:
gcd(a, b) ≥ 0: The greatest common divisor is a non-negative integer.
If d = gcd(a, b), then d divides a and b exactly: This is expressed as d|a and d|b. It means that a and b are
divisible by d without leaving a remainder.
If c is any integer that divides a and b exactly, then c also divides gcd(a, b) exactly: This can be written as
c|a and c|b imply c|gcd(a, b). In other words, the greatest common divisor is the largest common divisor of
a and b.
Euclid’s Theorem provides a fundamental basis for calculating the gcd and is essential in many fields of
mathematics, such as number theory, cryptography, and algorithms.
Now, let me provide you with a detailed example of how to use Euclid’s Theorem to find the greatest common
divisor of two numbers.

Example 3.11. Find the greatest common divisor of a = 56 and b = 84.

Solution 3.11. We will apply the Euclidean algorithm to find the gcd. We start by dividing 84 by 56 and
write the division as an equation:

84 = 56 · 1 + 28 (3.51)

This means that 84 can be expressed as 56 multiplied by 1, plus a remainder of 28.


Now, we take the divisor (56) as the new dividend and the remainder (28) as the new divisor. We perform
another division:
66 3 Algebra of Integer Numbers

56 = 28 · 2 + 0 (3.52)

In this case, we obtain a remainder of 0, indicating that we have reached the end of the algorithm.
The last non-zero divisor we obtained was 28. Therefore, the gcd of 56 and 84 is equal to 28.
Hence, we can write gcd(56, 84) = 28.

3.10. Bézout’s Theorem


Bézout’s theorem, also known as the Bézout’s identity or the Coefficient Identity Theorem, states that for
any pair of non-zero integers a and b, there exist integers x and y such that the following equation holds:

ax + by = gcd(a, b) (3.53)

Here, gcd(a, b) represents the greatest common divisor of a and b. In other words, the theorem states that
there exists a linear combination of a and b that yields their greatest common divisor.
The theorem implies that the greatest common divisor of a and b can be expressed as a linear combination
of the two integers with integer coefficients. These coefficients x and y are known as Bézout’s coefficients
and can be calculated using the extended Euclidean algorithm.
Let’s now provide a detailed example to illustrate the application of Bézout’s Theorem.

Example 3.12. Consider a = 42 and b = 15. We want to find the Bézout’s coefficients x and y that satisfy the
equation ax + by = gcd(a, b).

Solution 3.12. Applying the standard Euclidean algorithm:

1. 42 = 15 × 2 + 12
2. 15 = 12 × 1 + 3
3. 12 = 3 × 4

From these steps, we identify that gcd(42, 15) = 3.


Now, for Bézout’s coefficients, we reverse the process:
From the second equation:
15 − 12 = 3

Using the first equation to express 12:


12 = 42 − 15 × 2

Substituting this expression for 12 into the previous equation, we get:


3.11 Chinese Remainder Theorem 67

15 − (42 − 15 × 2) = 3

Which simplifies to:


45 − 42 = 3

This gives us the equation:


42(−1) + 15(3) = 3

Therefore, Bézout’s coefficients for a = 42 and b = 15 are x = −1 and y = 3.

3.11. Chinese Remainder Theorem


The Chinese remainder theorem [16] is a fundamental result in number theory that allows us to find a
unique solution for a system of linear congruences. This theorem is particularly useful when dealing with
large numbers or cryptographic problems.
Formally, the Chinese Remainder Theorem states the following:
Consider a system of linear congruences given by:

x ≡ a1 (mod m1 )
x ≡ a2 (mod m2 )
.. (3.54)
.
x ≡ an (mod mn )

Where a1 , a2 , · · · , an are the residues we want to find, and m1 , m2 , · · · , mn are the corresponding “moduli”
(Note 3.11).
If the moduli m1 , m2 , · · · , mn are pairwise coprime (i.e., mi and m j are coprime for i 6= j), then there exists a
unique solution for the system of linear congruences.
To find this unique solution, we can use the Chinese Remainder Theorem method, which consists of the
following steps:
Calculate the product of all moduli: M = m1 · m2 · · · · · mn .
M
For each i from 1 to n, calculate the value of Mi as the quotient of M divided by mi , i.e., Mi = .
mi
Find the multiplicative inverse yi of Mi modulo mi . This means that yi is an integer such that Mi · yi ≡ 1
(mod mi ).
The solution x for the system of congruences is obtained using the following formula:

x ≡ (a1 · M1 · y1 + a2 · M2 · y2 + · · · + an · Mn · yn ) (mod M) (3.55)

Now, let’s consider a detailed example to illustrate the application of the Chinese Remainder Theorem.
68 3 Algebra of Integer Numbers

Example 3.13. Suppose we want to find a number x that leaves a remainder of 2 when divided by 3 and a
remainder of 3 when divided by 5.
Solution 3.13. To solve this using the Chinese Remainder Theorem, we will apply the steps mentioned
earlier.
We calculate M = 3 · 5 = 15.
15 15
We calculate M1 = = 5 and M2 = = 3.
3 5
We find the multiplicative inverses: y1 such that 5 · y1 ≡ 1 (mod 3) and y2 such that 3 · y2 ≡ 1 (mod 5).
In this case, y1 = 2 since 5 · 2 ≡ 1 (mod 3), and y2 = 2 since 3 · 2 ≡ 1 (mod 5).
We substitute the values into the formula:

x ≡ (2 · 5 · 2 + 3 · 3 · 2) (mod 15) (3.56)

Simplifying the expression:

x ≡ (20 + 18) (mod 15) (3.57)

Adding the values:

x ≡ 38 (mod 15) (3.58)

Finally, reducing the remainder modulo 15:

x ≡ 8 (mod 15) (3.59)

Therefore, the solution for the system of linear congruences is x ≡ 8 (mod 15).

This is a basic application of the Chinese Remainder Theorem. As the number of congruences in the system
increases, the process becomes more complex, but the fundamental steps remain the same.
Note. The term “modulo” refers to the operation of taking the remainder after division. In mathematics, it is
denoted by the symbol “ mod ” or “
For example, let’s consider the expression 21 mod 8. This means we want to find the remainder when 21 is
divided by 8. Dividing 21 by 8 gives us a quotient of 2 and a remainder of 5. Therefore, 21 mod 8 = 5.
On the other hand, “moduli” is the plural form of the word “modulus.” In the context of modular arithmetic,
“modulus” refers to the number or set of numbers that define the modulus of congruence.
For instance, let’s consider a system of congruences with moduli n1 = 3, n2 = 5, and n3 = 7:

x ≡ 1 (mod n1 )
x ≡ 2 (mod n2 )
x ≡ 3 (mod n3 )

In this example, the moduli are n1 = 3, n2 = 5, and n3 = 7. These moduli define the arithmetic system in
which the congruences are considered. Solving this system means finding the value of x that satisfies all
three congruences simultaneously.
3.12 Congruence-based Cryptography System 69

3.12. Congruence-based Cryptography System


Example 3.14. Suppose Lucy wants to securely [17, 18] send the message “HELLO” to John.

Solution 3.14. Lucy wishes to send the message ”HELLO” to John securely using a congruence-based
cryptographic system.

1. Representation of the Message

The first step is to represent the message in numeric format. Assuming A = 0, B = 1, . . . , Z = 25, we can
represent the message as follows:

H→7
E→4
L → 11
L → 11
O → 14

So, “HELLO” can be represented as the sequence: [7, 4, 11, 11, 14].

2. Encryption using a Simple Congruence

To encrypt the message, Lucy can use the congruence:

C ≡ P × k + s (mod 26)

Where:

• C represents the ciphertext.

• P is the plaintext (our original number for each letter).

• k and s are constants known to both Lucy and John and kept secret. For this illustration, we’ll choose
k = 3 and s = 5.

Encrypting the message ”HELLO”:


For H (7):
70 3 Algebra of Integer Numbers

C ≡ 7 × 3 + 5 (mod 26)
C ≡ 26 (mod 26)
C=0

Applying the same method for each character in the message, Lucy will obtain an encrypted sequence of
numbers.

3. Transmission and Decryption

Upon receiving the sequence of numbers, John will decrypt the message using the inverse of k (denoted as
k−1 ) and s:
P ≡ C × k−1 − s (mod 26)

Given that the multiplicative inverse of 3 modulo 26 is 9, John will be able to recover the original message
using the aforementioned formula.

Note. The equation c ≡ me (mod n) used in (Ex. 3.14) to encrypt the message is a specific expression
used in asymmetric cryptography systems, such as RSA encryption. This formula is not directly derived
from a particular mathematical theorem, but rather it is a practical application of modular arithmetic and the
properties of congruences.
In the case of RSA encryption, a pair of prime numbers, p and q, are chosen to calculate the modulus
n = p · q, and an encryption exponent e is selected that is coprime with the value of ϕ (n) = (p − 1)(q − 1).
The formula c ≡ me (mod n) is used to encrypt a message m into a ciphertext number c, where m is raised
to the power of e and then the residue modulo n is calculated.
It is important to note that RSA encryption is based on the difficulty of factoring large composite numbers
into their prime factors. Although the encryption formula is not directly derived from a specific mathematical
theorem, the security of RSA encryption is supported by theorems and properties of number theory, such as
the Fundamental Theorem of Arithmetic and the Chinese Remainder Theorem.
The formula c ≡ me (mod n) used in the example is a specific expression used in RSA encryption to encrypt
messages, and it is based on principles of modular arithmetic and number theory, although it is not directly
derived from a specific theorem.

1. ϕ (n) represents Euler’s totient function, which gives the number of integers less than n that are coprime
(i.e., share no common factors other than 1) with n.

2. When n is a product of two prime numbers, p and q (as is commonly the case in cryptographic
applications like RSA), the totient function can be expressed as ϕ (n) = (p − 1)(q − 1). This is because,
out of the numbers below p, p − 1 of them are coprime to p, and similarly for q.

3. In the context of cryptography, particularly the RSA encryption scheme, Euler’s totient function is
crucial because it plays a role in determining the public and private keys.
3.13 Conclusions 71

3.13. Conclusions
In this chapter, we have examined several fundamental number theory concepts and their applications
to integer algebra. The greatest common divisor and least common multiple, congruences and modular
arithmetic, prime numbers, the fundamental theorem of arithmetic, Euclid’s theorem, Bézout’s theorem,
the Chinese remainder theorem, and number theory have been covered.
We have a deeper understanding of the properties and relationships between integers after examining these
concepts. In fraction division, factorization, and simplification problems, the greatest common divisor and
least common multiple have proven indispensable. Through the lens of modular arithmetic, the notion of
congruence has allowed us to investigate mathematical patterns and cycles.
Due to their unique properties and significance in the fundamental theorem of arithmetic, which guarantees
the unique factorization of any integer into primes, prime numbers have captured our attention. The theorem
of Euclid has given us a profound understanding of the infinity of prime numbers, while the theorem of
Bézout has enabled us to discover integer solutions to linear diophantine equations.
In addition, the Chinese remainder theorem has been utilized in cryptography and number theory to solve
systems of congruences and has found applications in the field of system analysis.
To reinforce our understanding of these concepts, we have provided a fully-solved cryptography example
based on the material covered in this chapter. This example illustrated the use of congruences, modular
arithmetic, and prime numbers in encrypting and decrypting messages, demonstrating the practical relevance
of these theoretical concepts.
This chapter has provided a thorough examination of the algebra of integer numbers, casting light on
the fundamental principles underlying a wide variety of mathematical and cryptographic applications. By
mastering these concepts, readers will be armed with potent tools for problem solving, pattern analysis, and
appreciating the majesty of number theory.
72 3 Algebra of Integer Numbers

3.14. Exercises
Exercise 3.1. Find the gcd of a = 36 and b = 48.

Exercise 3.2. Calculate the lcm of x = 9 and y = 15.

Exercise 3.3. Solve the congruence 2x ≡ 6 (mod 5).

Exercise 3.4. Find all integers x that satisfy 3x ≡ 2 (mod 7).

Exercise 3.5. Compute 12 mod 7.

Exercise 3.6. Find 5−1 (mod 11) (the multiplicative inverse of 5 modulo 11).

Exercise 3.7. Solve the linear congruence equation 3x ≡ 10 (mod 17).

Exercise 3.8. Determine the value of x in the congruence 4x ≡ 8 (mod 15).

Exercise 3.9. Calculate the gcd of m = 48 and n = 36.

Exercise 3.10. Find the lcm of a = 12 and b = 18.

Exercise 3.11. Solve the congruence 5x ≡ 3 (mod 11).

Exercise 3.12. Determine the smallest value of x that satisfies 4x ≡ 1 (mod 7).

Exercise 3.13. Find all integers y that satisfy 2y ≡ 5 (mod 9).

Exercise 3.14. Compute 15 mod 4.


Chapter 4
Algebra of Real Numbers

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract This chapter provides a concise exploration of fundamental topics in mathematics, focusing on real
numbers. It covers axioms and properties, order relations, absolute value, inequalities, infinite sequences,
infinite series, radicals, and exponents. The chapter delves into the foundational aspects of these topics,
offering insights into their properties and applications in mathematical reasoning and problem-solving. By
studying this chapter, readers will develop a solid understanding of key concepts in real number theory.

Keywords: absolute value, axioms on real numbers, exponents, inequalities, Infinite Sequences, infinite
Series, integer numbers, number theory, order properties, properties on real numbers, radicals, real numbers.

4.1. Introduction
In this chapter, we investigate the fundamental concepts and properties of real numbers. Examining the
axioms that define the real number system, which are the fundamental building blocks for mathematical
operations, is the first step. These axioms provide a firm foundation for comprehending the behavior and
relationships of real numbers.
Following this, we examine the order properties of real numbers, which allow us to compare and arrange
numbers along a number line. Understanding these properties permits us to establish meaningful relationships
between real numbers and to analyze their magnitudes. Also introduced is the concept of absolute value,

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

73
74 4 Algebra of Real Numbers

which measures the distance between a real number and zero. Absolute value equations and inequalities are
solved by examining its properties and applications.
We investigate the properties and methods for solving inequalities, which play a crucial role in mathematics.
We develop techniques for analyzing and manipulating inequalities, allowing us to formulate valid mathematical
statements based on the relationships among real numbers. We now transfer our attention to infinite
sequences and series. We investigate the behavior of infinitely continuing sequences by analyzing their
patterns and limits. In addition, we investigate the notion of infinite series, which entails the addition of an
infinite number of terms.
Radicals and exponents are fundamental mathematical operations, and their properties are investigated in
depth. We investigate the manipulation and simplification of radical expressions as well as the principles
governing exponent behavior. By studying these concepts, we set the groundwork for a more in-depth
comprehension of the mathematical principles that underpin numerous academic disciplines. Through
the study of real numbers, sequences, series, radicals, and exponents, we acquire insights that will be
indispensable for future mathematical study and problem-solving.

4.2. Real Numbers


The set of real numbers, denoted by R [19, 20], is defined as the set of all numbers that can be represented by
an infinite or finite decimal expansion, including integers, fractions, and irrational numbers. Mathematically,
it is defined as:

R = {x : x is an infinite or finite decimal expansion} (4.1)

Real numbers include rational numbers, which can be expressed as ratios of integers, and irrational numbers,
which cannot be expressed exactly as a fraction. Real numbers form a continuous line on the number line,
extending infinitely in both directions. Furthermore, real numbers can be operated upon using standard
arithmetic operations such as addition, subtraction, multiplication, and division.

4.3. Axioms and Properties


(i) Closure: The addition and multiplication of two real numbers always yield another real number.

Example 4.1. Adding two real numbers, such as x and y, always results in another real number: x + y.

(i) Existence of Additive Identity: There exists a real number called ”zero” (0) such that the sum of any real
number with zero results in the same number.

Example 4.2. For any real number a, adding a to the additive identity (0) results in the same number:
a + 0 = a.

(iii) Existence of Multiplicative Identity: There exists a real number called ”one” (1) such that the product of
any real number with one yields the same number.
4.3 Axioms and Properties 75

Example 4.3. For any real number b, multiplying b by the multiplicative identity (1) results in the same
number: b × 1 = b.

(iv) Existence of Additive Inverses: For every real number a, there exists a real number called ”additive
inverse” or ”negative” (−a) such that its sum with a results in zero.

Example 4.4. For any real number c, adding c to its additive inverse (−c) results in zero: c + (−c) = 0.

(v) Commutative Property of Addition: The addition of two real numbers is commutative, which means the
order of the terms does not affect the result.

Example 4.5. a + b = b + a for any pair of real numbers a and b.

(vi) Associative Property of Addition: The addition of three real numbers is associative, which means the
grouping of the terms does not affect the result.

Example 4.6. (a + b) + c = a + (b + c) for any real numbers a, b, and c.

(vii) Identity Property of Addition: The sum of any real number and zero is equal to the original number.

Example 4.7. a + 0 = a for any real number a.

(viii) Inverse Property of Addition: For every real number a, there exists a unique real number (−a) called the
additive inverse, such that their sum is equal to zero.

Example 4.8. a + (−a) = 0 for any real number a.

(ix) Commutative Property of Multiplication: The multiplication of two real numbers is commutative, which
means the order of the factors does not affect the result.

Example 4.9. ab = ba for any pair of real numbers a and b.

(x) Associative Property of Multiplication: The multiplication of three real numbers is associative, which
means the grouping of the factors does not affect the result.

Example 4.10. (ab)c = a(bc) for any real numbers a, b, and c.

(xi) Identity Property of Multiplication: The product of any real number and one is equal to the original
number.

Example 4.11. a × 1 = a for any real number a.

(xii) Inverse Property of Multiplication: For every non-zero real number a, there exists a unique real number
1
a−1 or called the multiplicative inverse or reciprocal, such that their product is equal to one.
a
Example 4.12. a × a−1 = 1 for any non-zero real number a.

(xiii) Distributive Property: Multiplication distributes over addition for three real numbers.

Example 4.13. a × (b + c) = (a × b) + (a × c) for any real numbers a, b, and c.


76 4 Algebra of Real Numbers

4.4. Order Properties


(i) Order Properties: For any real number a, a is greater than or equal to itself.

Example 4.14. a ≥ a is an example of the reflexive property of order.

(ii) Antisymmetric Property of Order: For any real numbers a and b, if a is greater than or equal to b and b
is greater than or equal to a, then a is equal to b.

Example 4.15. If 2 ≥ 1 and 1 ≥ 2, then 2 = 1 is an example of the antisymmetric property of order


(which is not true in this case).

(iii) Transitive Property of Order: For any real numbers a, b, and c, if a is greater than or equal to b and b is
greater than or equal to c, then a is greater than or equal to c.

Example 4.16. If 4 ≥ 3 and 3 ≥ 2, then 4 ≥ 2 is an example of the transitive property of order.

(iv) Addition Property of Order: For any real numbers a, b, and c, if a is greater than or equal to b, then
adding c to both sides preserves the inequality, i.e., a + c is greater than or equal to b + c.

Example 4.17. If 2 ≥ 1, then 2 + 3 ≥ 1 + 3 is an example of the addition property of order, resulting in


5 ≥ 4.

(v) Multiplication Property of Order: For any real numbers a, b, and c, if a is greater than or equal to b and
c is positive, then multiplying both sides by c preserves the inequality, i.e., ac is greater than or equal to
bc.

Example 4.18. If −2 ≥ −3 and 4 is positive, then −2 × 4 ≥ −3 × 4 is an example of the multiplication


property of order, resulting in −8 ≥ −12.

(vi) Trichotomy Property: For any real numbers a and b, exactly one of the following statements is true: a is
greater than b, a is less than b, or a is equal to b.

Example 4.19. For 2 and 3, either 2 < 3, 2 > 3, or 2 = 3 must be true. In this case, 2 < 3 is an example
of the trichotomy property.

4.5. Absolute Value and Inequalities


The absolute value function, denoted as |x|, is a function that assigns a real number x to its distance from the
origin on the number line. It is defined as follows:
(
x, if x ≥ 0
|x|= (4.2)
−x, if x < 0

In other words, if x is a non-negative number, the absolute value of x is equal to x itself. If x is a negative
number, the absolute value of x is equal to its negation, which is −x.
4.6 Infinite Sequences and Series 77

(i) Non-negativity: The absolute value of any real number x is always non-negative.

(ii) Symmetry: The absolute value of x is equal to the absolute value of its negation: |x|= |−x|.

(iii) Triangle Inequality: For any real numbers x and y, the absolute value of their sum is less than or equal
to the sum of their absolute values: |x + y|≤ |x|+|y|.

Example 4.20. Let’s consider the number x = 5. The absolute value of x is |5|= 5, since 5 is a non-negative
number.

Example 4.21. Now, let’s take the number x = −3. The absolute value of x is |−3|= 3, as −3 is a negative
number, and its absolute value is its negation, which is 3.

Example 4.22. Finally, let’s consider the number x = 0. The absolute value of x is |0|= 0, as 0 is the only
number that is neither positive nor negative, and its distance from the origin is 0.

(i) Absolute Value is Non-negative: For any real number x, |x| is always non-negative.
Example 4.23. Let’s consider the absolute value of −3: |−3|= 3, which is non-negative.
(ii) Absolute Value Inequality: For any real numbers x and a, if |x|< a, then −a < x < a.
Example 4.24. Suppose we have the inequality |x|< 5. This implies −5 < x < 5, which means x lies
between −5 and 5 on the number line, but not including −5 and 5.
(iii) Triangle Inequality: For any real numbers x and y, |x + y|≤ |x|+|y|.
Example 4.25. Let’s consider the numbers x = −2 and y = 3. Applying the triangle inequality, we have
|−2 + 3|≤ |−2|+|3|, which simplifies to |1|≤ 2 + 3. This is true because 1 ≤ 5.
(iv) Reversal Property: For any real numbers x and a, if |x|> a, then x > a or x < −a.
Example 4.26. Suppose we have the inequality |x|> 2. This implies x > 2 or x < −2. In other words, x
is either greater than 2 or less than −2.

4.6. Infinite Sequences and Series


An infinite sequence is a list of numbers that goes on forever. Each number in the sequence is called a term.
We can represent an infinite sequence as a1 , a2 , a3 , . . . or using the sigma notation (an )∞
n=1 , where an denotes
the nth term of the sequence.

Example 4.27. The sequence of natural numbers: 1, 2, 3, 4, 5, . . . can be represented as an = n.

Example 4.28. The sequence of even numbers: 2, 4, 6, 8, 10, . . . can be represented as an = 2n.

Example 4.29. The Fibonacci sequence: 0, 1, 1, 2, 3, 5, 8, 13, . . . can be represented recursively as a1 = 0,


a2 = 1, and an = an−1 + an−2 for n ≥ 3.
78 4 Algebra of Real Numbers

In each example, we can observe that the sequence continues indefinitely without a fixed endpoint. Infinite
sequences are essential in mathematics and have various applications in different areas, such as calculus,
number theory, and computer science.

An infinite series is the sum of an infinite sequence of numbers. It is represented by the symbol ∑ (sigma),
followed by the expression for the terms of the sequence. The terms are typically denoted by an index
variable, such as n, and written as a function of n.
The general form of an infinite series is:

∑ an = a1 + a2 + a3 + . . . (4.3)
n=1

Where an represents the nth term of the sequence.

Example 4.30. The series of natural numbers:



∑ n = 1+2+3+4+... (4.4)
n=1

In this series, each term is obtained by adding 1 to the previous term. The sum of this series is infinity
because there is no fixed endpoint.

Example 4.31. The series of reciprocals of powers of 2:



1 1 1 1 1
∑ 2n = 2 + 4 + 8 + 16 + . . . (4.5)
n=1

In this series, each term is obtained by taking the reciprocal of powers of 2. The sum of this series is 1
because it converges to a fixed value.

Example 4.32. The series of alternate signs:



∑ (−1)n+1 = 1 − 1 + 1 − 1 + . . . (4.6)
n=1

In this series, each term alternates between positive and negative signs. The sum of this series does not exist
since it oscillates between 0 and 1.

Infinite series can have various behaviors. Some series converge to a specific value, while others diverge and
do not have a finite sum. The study of infinite series is an important topic in mathematics with applications
in many fields.
There are several peculiarities related to infinite series; some of them are listed below. The reader is warned
that this is an extensive topic, so we will only state some of them.
4.6 Infinite Sequences and Series 79

(i) Convergence and Divergence:


Example 4.33.

1
∑ n2 (4.7)
n=1

π2
This is the series of reciprocals of squares. This series is convergent, and its sum is .
6
(ii) Partial Sum:
Example 4.34.
10
1
∑ 2n (4.8)
n=1


1
This is the partial sum of the geometric series ∑ 2n . In this case, we are summing the first 10 terms of
n=1
the series.
(iii) Series Convergence Theorem: The theorem states that an infinite series converges if and only if the
sequence of partial sums converges. For example, consider the series:
Example 4.35.

1
∑n (4.9)
n=1

1 1 1
The partial sum of this series is denoted as Sn = 1 + + + . . . + . According to the theorem, if the
2 3 n
sequence Sn converges to a finite value, then the series also converges.

(iv) Alternating Series:


Example 4.36.

(−1)n
∑ 2n + 1 (4.10)
n=0

π
In this series, the terms alternate in sign. The Leibniz series converges to , even though the terms
4
decrease in absolute value.
(v) Geometric Series Theorem:
Example 4.37.

1
∑ 2n (4.11)
n=0

1
In this case, the ratio of the series is , which is between -1 and 1. Therefore, the geometric series
2
converges to a finite value, and its sum is 2.
80 4 Algebra of Real Numbers

(vi) Absolute and Conditional Convergence:


Example 4.38.

1
∑ n2 (4.12)
n=1

This series converges absolutely since the series of the absolute values of the terms, ∑∞ 1
n=1 n2 , also
converges.
An example of conditional convergence is the series of the sine function:

Example 4.39.

(−1)n
∑ n
(4.13)
n=1


1
This series converges, but if we take the absolute value of the terms, ∑ n , we get the series, which
n=1
diverges.

4.7. Radicals and Exponents


√ √
Radicals involve the use of square roots ( ) and other roots (such as cube roots 3 ). In algebra, we often need
to simplify or manipulate expressions involving radicals. The algebra of radicals involves rules for adding,
subtracting, multiplying, and dividing expressions containing radicals.
To simplify a radical expression, we try to express the radicand (the number inside the radical symbol) in
terms of perfect squares or other perfect roots.

Example 4.40. Simplify: 12
√ √ √ √
Solution 4.1. To simplify 12, we can factorize 12 as 22 · 3. Therefore, 12 = 22 · 3 = 2 3.

Adding and Subtracting Radicals: To add or subtract radicals, we need to have the same radical terms.
√ √
Example 4.41. Simplify: 7 + 3
√ √
Solution 4.2. We cannot simplify
√ √7 and 3 further, so the sum cannot be simplified any further. Therefore,
the simplified expression is 7 + 3.

Multiplying and Dividing Radicals: To multiply or divide radicals, we can use the product and quotient rules.

√ √
Example 4.42. Simplify: 5· 2
√ √ √ √
Solution 4.3. Using the product rule, 5 · 2 = 5 · 2 = 10.

Example 4.43. Simplify: √8
2
4.8 Conclusions 81
√ q √
Solution 4.4. Using the quotient rule, √8 = 8
= 4 = 2.
2 2

Exponents involve raising a number to a power. In algebra, we often need to simplify or manipulate
expressions involving exponents. The algebra of exponents involves rules for multiplying, dividing, and
raising expressions with exponents.
Product Rule: To multiply two terms with the same base, we keep the base and add the exponents.
Example 4.44. Simplify: x3 · x4
Solution 4.5. Using the product rule, x3 · x4 = x3+4 = x7 .

Quotient Rule: To divide two terms with the same base, we keep the base and subtract the exponents.
y6
Example 4.45. Simplify: y2

y6
Solution 4.6. Using the quotient rule, y2
= y6−2 = y4 .

Power Rule: To raise a term with an exponent to another exponent, we multiply the exponents.
Example 4.46. Simplify: (z2 )3
Solution 4.7. Using the power rule, (z2 )3 = z2·3 = z6 .

4.8. Conclusions
In conclusion, this chapter of the book addressed significant topics concerning the axioms and properties of
real numbers, order properties, absolute value, inequalities, infinite sequences, infinite series, radicals, and
exponents. By grasping these fundamental concepts, readers acquire a solid foundation for understanding
and manipulating the structure and operations of the set of real numbers. These topics play a vital
role in a variety of mathematical disciplines and provide essential tools for solving equations, analyzing
mathematical relationships, and making quantitative comparisons. This chapter serves as a springboard for
further investigation and implementation of these mathematical concepts to more advanced topics and real-
world situations.
82 4 Algebra of Real Numbers

4.9. Exercises
Exercise 4.1. Find the sum of the arithmetic series:

Sn = 3 + 5 + 7 + · · · + 39.

Exercise 4.2. Determine if the following series converges or diverges:



1
∑ n2
n=1

Exercise 4.3. Evaluate the sum of the geometric series:


4
Sn = 4 + 2 + 1 + · · · +
2n−1

Exercise 4.4. Prove that 2 is an irrational number.
√ √
Exercise 4.5. Prove that for any positive real number x, we have x < x + 1.

Exercise 4.6. Find the general term of the arithmetic sequence 3, 7, 11, 15, . . .

Exercise 4.7. Find the general term of the geometric sequence 2, 6, 18, 54, . . .

Exercise 4.8. Calculate the sum of the first 5 terms of the arithmetic sequence 1, 4, 7, 10, . . .
1 1 1
Exercise 4.9. Find the infinite sum of the geometric series 1 + + + +...
2 4 8
Chapter 5
Algebra of Complex Numbers

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract This chapter provides a comprehensive introduction to complex numbers and their fascinating
mathematical world. We begin by establishing the fundamental axioms that govern complex numbers,
exploring their unique properties and how they differ from real numbers. We then delve into the concept
of complex conjugate, revealing how this concept is essential for simplifying operations and unlocking
new applications in mathematical and scientific problems. Once the foundations are laid, we delve into the
complex plane, a powerful visual tool that represents complex numbers as points in a two-dimensional plane.
We explore the geometric properties of the complex plane and how it allows us to intuitively understand
operations with complex numbers, such as addition, subtraction, multiplication, and division. Additionally,
the polar representation of complex numbers is introduced, revealing an alternative way to express them
using magnitude and angle. We investigate how the polar form facilitates understanding of operations such
as exponentiation and obtaining complex root

Keywords: absolute value, axioms in complex numbers, complex conjugate, complex numbers, complex
plane, complex roots, De Moivre’s theorem, integer numbers, Newton’s method, polar representation,
properties in complex numbers, real numbers.

5.1. Introduction
There exists a captivating realm in which conventional numbers are inadequate to characterize certain
phenomena. Complex numbers make their triumphant entrance at this point. In this chapter, we will venture

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

83
84 5 Algebra of Complex Numbers

into the domain of complex numbers, a mysterious and potent realm that allows us to test the limits of
conventional arithmetic and explore an entirely new dimension.
To begin our journey, we will investigate the foundational axioms underlying complex numbers. These
axioms establish the fundamental principles for working with these numbers and allow us to construct a
solid and coherent algebraic structure. As we disclose these axioms, we will comprehend the definition of
complex numbers and their relationship to real numbers.
In addition, we will delve into the inherent properties of complex numbers. We will learn how these special
numbers react when subjected to various algebraic operations such as addition, subtraction, multiplication,
and division. These properties will reveal remarkable symmetry and mathematical richness, thereby maximizing
their potential.
We will also encounter a crucial figure in the domain of complex numbers: the complex conjugate. We will
learn how to determine the conjugate of a complex number and how this operation provides valuable insight
into its structure and properties. We will investigate how the complex conjugate influences operations and
enables us to simplify expressions and solve equations more elegantly.
The exploration of complex numbers would be insufficient without a trustworthy guide. In order to visualize
and investigate these numbers within a geometric context, we will introduce the complex plane, a special
Cartesian map. We will learn how to plot complex numbers on this plane, as well as how the relationship
between the complex plane and algebraic operations provides us with a unique and potent visual perspective.
Finally, we will discuss the intriguing concept of complex number polar representation. We will learn how
to express these numbers in terms of their magnitude and angle, and how this representation enables us to
execute operations with astounding ease. We will investigate the relationship between polar representation
and the complex plane, as well as how this representation facilitates the computation of complex roots and
the manipulation of complex numbers in general.

5.2. Complex Numbers


The formal definition of the set of complex numbers C [21, 22, 23] is defined as the set of expressions of
the form a + bi, where a and b are real numbers and i is the imaginary unit, with the property that i2 = −1.
Formally, it can be expressed as:

C = {a + bi | a, b ∈ R, i2 = −1} (5.1)

The elements of C are called complex numbers and have a real part a and an imaginary part b. The real part
is denoted as ℜ(z) and the imaginary part as ℑ(z) for a generic complex number z = a + bi. Real numbers
are a subset of complex numbers, as they can be represented as complex numbers with the imaginary part
equal to zero, i.e., a = a + 0i.
5.4 Complex Conjugate 85

5.3. Axioms and Properties


Closure under Addition and Multiplication: The set of complex numbers is closed under addition and
multiplication. This means that if z1 and z2 are complex numbers, then their sum z1 + z2 and product z1 · z2
are also complex numbers.

Example 5.1. Let z1 = 2 + 3i and z2 = −1 + 4i. We can find their sum and product:

z1 + z2 = (2 + 3i) + (−1 + 4i) = 1 + 7i (5.2)

z1 · z2 = (2 + 3i) · (−1 + 4i) = −14 + 5i (5.3)

Additive and Multiplicative Identity: The complex number 0 + 0i, denoted as 0, serves as the additive
identity, i.e., z + 0 = z for any complex number z. The complex number 1 + 0i, denoted as 1, serves as
the multiplicative identity, i.e., z · 1 = z for any complex number z.

Example 5.2. Let z = 2 + 3i. We can find the sum and product with the additive and multiplicative identities:

z + 0 = (2 + 3i) + 0 = 2 + 3i (5.4)

z · 1 = (2 + 3i) · 1 = 2 + 3i (5.5)

Additive Inverse: For every complex number z, there exists an additive inverse −z such that z + (−z) = 0.

Example 5.3. Let z = 2 + 3i. We can find the additive inverse:

z + (−z) = (2 + 3i) + (−2 − 3i) = 0 (5.6)

Associativity of Addition and Multiplication: For any complex numbers z1 , z2 , and z3 , addition and
multiplication are associative, i.e., (z1 + z2 ) + z3 = z1 + (z2 + z3 ) and (z1 · z2 ) · z3 = z1 · (z2 · z3 ).
Commutativity of Addition and Multiplication: For any complex numbers z1 and z2 , addition and multiplication
are commutative, i.e., z1 + z2 = z2 + z1 and z1 · z2 = z2 · z1 .
Distributivity of Multiplication over Addition: For any complex numbers z1 , z2 , and z3 , multiplication
distributes over addition, i.e., z1 · (z2 + z3 ) = (z1 · z2 ) + (z1 · z3 ).

5.4. Complex Conjugate


The complex conjugate of a complex number z = a + bi, denoted as z, is obtained by changing the sign
of its imaginary part. In other words, if z = a + bi, then its complex conjugate is z = a − bi. The complex
conjugate helps in various mathematical operations involving complex numbers, such as division and finding
the magnitude of a complex number.

Example 5.4. Let z = 3 + 2i be a complex number. The complex conjugate of z is given by z = 3 − 2i.
86 5 Algebra of Complex Numbers

5.5. Absolute Value


The absolute value (or modulus) of a complex number z = a + bi, denoted as |z|, represents its distance from
the origin in the complex plane. It is calculated
√ by taking the square root of the sum of the squares of its real
and imaginary parts. Mathematically, |z|= a2 + b2 .
p
Example
√ 5.5.
√ Let z = −4 + 3i be a complex number. The absolute value of z is given by |z|= (−4)2 + 32 =
16 + 9 = 25 = 5.

These concepts are fundamental in complex analysis and have applications in various branches of mathematics
and physics.

5.6. The Complex Plane


The geometrical representation of the complex plane utilizes a two-dimensional structure to visualize
complex numbers. In this representation, the horizontal axis is known as the real axis, and the vertical axis
is known as the imaginary axis.
Each complex number can be expressed as a combination of a real part and an imaginary part, and it is
represented as a point in the complex plane. The real part determines the horizontal position of the point on
the real axis, while the imaginary part determines the vertical position of the point on the imaginary axis.

Example 5.6. For example, the complex number z = 3 + 2i is represented as the point (3, 2) in the complex
plane. Here, 3 is the real part, and 2 is the imaginary part.

The geometrical representation of the complex plane allows us to perform arithmetic operations with
complex numbers, such as addition, subtraction, multiplication, and division. These operations can be
visually performed by moving the points in the complex plane and observing how they combine.

Example 5.7. Consider the complex numbers z1 = 2 + i and z2 = −1 + 3i. We represent z1 as the point (2, 1)
and z2 as the point (−1, 3) in the complex plane.

Solution 5.1. To add these complex numbers, we simply add their real parts and imaginary parts separately.
In this case, the sum is z1 + z2 = (2 + −1) + (1 + 3)i = 1 + 4i. We represent the result 1 + 4i as the point
(1, 4) in the complex plane.

The next figure illustrates the complex plane:

5.7. Polar Representation


Polar representation is a way to represent a complex number using its magnitude (length) and argument
(angle). In this representation, a complex number is expressed in the form z = r ·eiθ , where r is the magnitude
and θ is the argument.
5.7 Polar Representation 87

Imaginary Axis z1 + z2

z2

z1

Real Axis

Figure 5.1: Representation of the complex plane

The value of r is the distance of the complex number from the origin in the complex plane, while θ is the
angle formed by the complex number with the positive real axis. The angle is measured in radians and is
usually taken in the range [0, 2π ).

Example 5.8. Let’s consider the complex number z = 3 + 3i. We want to represent this number in polar form.
q
First, we calculate the magnitude r using the formula r = Re(z)2 + Im(z)2 , where Re(z) is the real part
and Im(z) is the imaginary part of z. In this case, Re(z) = 3 and Im(z) = 3, so:
√ √ √
r = 32 + 32 = 18 = 3 2
 
Im(z)
Then, we calculate the argument θ using the formula θ = arctan . In this case:
Re(z)
 
3 π
θ = arctan = arctan(1) =
3 4
√ π
Therefore, the complex number z = 3 + 3i can be represented in polar form as z = 3 2 · ei 4 .
88 5 Algebra of Complex Numbers

Im

r
Im(z)
θ
Re
Re(z)

Figure 5.2: Complex plane in the space.

5.8. De Moivre’s Theorem


De Moivre’s Theorem states that for any complex number z = r(cos θ + i sin θ ) and any positive integer n,
its nth power can be calculated by raising both its magnitude and argument to the nth power.

zn = rn (cos(nθ ) + i sin(nθ )) (5.7)

This formula provides a quick and convenient way to calculate powers of complex numbers in trigonometric
form.

Definition 5.1. Let z = r(cos θ + i sin θ ) be a complex number in trigonometric form, where r is the
magnitude of z and θ is its argument. De Moivre’s Theorem states that for any positive integer n:

zn = rn (cos(nθ ) + i sin(nθ )) (5.8)

Proof. Let’s consider a complex number z = r(cos θ +i sin θ ) in trigonometric form, where r is the magnitude
of z and θ is its argument.
To prove De Moivre’s Theorem, we will use the approach of mathematical induction.
For n = 1, we have:

z1 = r1 (cos θ + i sin θ ) = r(cos θ + i sin θ ) = z (5.9)

Therefore, De Moivre’s Theorem holds true for n = 1.


Let’s assume that De Moivre’s Theorem holds true for n = k, i.e., let’s assume:

zk = rk (cos(kθ ) + i sin(kθ )) (5.10)

We want to prove that De Moivre’s Theorem also holds true for n = k + 1.


Consider the specific case of zk+1 . We can write it as:
5.9 Square Roots of Negative Numbers 89

zk+1 = zk · z (5.11)

Using our inductive assumption, we substitute zk with its equivalent expression:

zk+1 = rk (cos(kθ ) + i sin(kθ )) · (r(cos θ + i sin θ )) (5.12)

Multiplying the terms and applying trigonometric properties of multiplication, we obtain:

zk+1 = rk+1 (cos(kθ + θ ) + i sin(kθ + θ )) (5.13)

Simplifying the expression inside the parentheses, we have:

zk+1 = rk+1 (cos((k + 1)θ ) + i sin((k + 1)θ )) (5.14)

Therefore, we have shown that De Moivre’s Theorem holds true for n = k + 1.


Since we have established that De Moivre’s Theorem is valid for n = 1, and its validity implies that it also
holds true for n = k + 1 when it holds true for n = k, we can conclude by induction that De Moivre’s Theorem
is valid for all positive integers n. 
π π
Example 5.9. Let’s consider the complex number z = 2(cos + i sin ), which has a magnitude of 2 and an
4 4
π 3
argument of . Using De Moivre’s Theorem, we can calculate z as follows.
4
Solution 5.2.   π  π   
3 3 3π 3π
z = (2) cos 3 · + i sin 3 · = 8 cos + i sin (5.15)
4 4 4 4

3π 3π
Therefore, z3 is equal to 8(cos + i sin ), which gives us the magnitude and argument corresponding to
4 4
the cube of z.

5.9. Square Roots of Negative Numbers


Example 5.10. Let’s suppose we have a complex number z = a + bi, where a and b are real numbers and i is
the imaginary unit. We want to find another complex number w = c + di such that w2 = z. In other words,
we want to find the square root of z.

Solution 5.3. To solve this problem, we can apply the properties of complex number algebra. Let w = c + di
be the square root of z = a + bi, then we have: (w)2 = a + bi
Expanding the square, we obtain: c2 + 2cdi + d 2 i2 = a + bi
Using the property i2 = −1, we can simplify the equation: c2 + 2cdi − d 2 = a + bi
By equating the real and imaginary parts, we get the following system of equations:
90 5 Algebra of Complex Numbers
(
c2 − d 2 = a
2cd = b
By solving this system, we can find the values of c and d. Once we have c and d, we can express the square
root of z as w = c + di.
For(example, if we want to calculate the square root of z = −1 + 2i, we apply the above procedure:
c2 − d 2 = −1
2cd = 2
2 1
From the second equation, we can solve for c in terms of d: c = =
2d d
 2
1
Substituting into the first equation, we have: − d 2 = −1
d
Multiplying by d 2 , we obtain: 1 − d 4 = −d 2
Rearranging the terms, we arrive at: d 4 − d 2 − 1 = 0
This quartic equation can be solved using algebraic methods or numerical methods. Let’s suppose we find
a solution for d, denoted as d0 .
2 1
Then, we can find c using the second equation: c = =
2d0 d0
1
Finally, the square root of z = −1 + 2i is w = c + di = + d0 i, where d0 is the solution obtained for d.
d0

5.10. Conclusions
This chapter investigated the fundamental properties of complex numbers and their representation on the
complex plane. We began by examining the axioms that define complex numbers and the properties that
result from them, thereby gaining a greater understanding of their nature and behavior.
The complex conjugate, which is essential for operations such as division and obtaining the polar form
of a complex number, was one of the main concepts discussed. Complex conjugation enables us to obtain
conjugate complex numbers, which have the same real component but opposite imaginary component.
As a graphical aid for representing complex numbers, the complex plane was introduced, where the
horizontal axis represents the real component and the vertical axis represents the imaginary component.
This allows us to visualize the operations and properties of complex numbers in an intuitive manner.
In this chapter, the polar representation of complex numbers was also emphasized. We can determine a
complex number’s modulus (distance from the origin) and argument (angle with respect to the positive
real axis) by expressing it in polar coordinates. This method of representation is particularly helpful when
multiplying and exponentiating complex numbers.
Finally, complex roots, which are solutions to algebraic equations of higher degrees, were discussed.
Complex roots reveal the depth of complex numbers and their capacity to represent solutions in broader
contexts.
5.11 Exercises 91

5.11. Exercises
Exercise 5.1. Compute the complex conjugate of z = 3 + 4i.

Exercise 5.2. Given the complex numbers z = 2 + 3i and w = −1 + 2i, find the absolute value of z + w.

Exercise 5.3. Simplify the expression (4 − 2i)(3 + i) and write the result in rectangular form.

Exercise 5.4. Prove that the product of a complex number z and its conjugate z is equal to the square of the
absolute value of z.

Exercise 5.5. Given the complex numbers z = 5 − 2i and w = −3 + 4i, find the complex conjugate of z + w.
π
Exercise 5.6. If the absolute value of a complex number z is 7 and its argument is 4, find the rectangular
form of z.

Exercise 5.7. Prove that the absolute value of the product of two complex numbers z and w is equal to the
product of the absolute values of z and w.

Exercise 5.8. Given the complex number z = 2 + 3i, find a complex number w = a + bi such that w satisfies
the following conditions:

(a) The real part of w is equal to twice the imaginary part of z.


(5.16)
(b) The conjugate complex of w is equal to the additive opposite of z.

Exercise 5.9. Given the complex number z = 2 + 3i, perform the following operations:

(a) Calculate the complex conjugate of z and express it in rectangular shape.


(b) Find the absolute value of z. (5.17)
(c) Find a complex number w such that z + w = 4 − 2i.

Exercise 5.10. Given the complex number z = −2 + 5i, find a real number r such that the absolute value of
z is equal to r. Then, he shows that z and its conjugate complex have the same absolute value.
Chapter 6
Algebra of Polynomial Functions

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract In this chapter, we will delve into various aspects related to polynomials and their factorization.
We will begin by studying the factor theorem and its application in solving polynomial equations. Next,
we will explore the fundamental theorem of algebra and its significance in the factorization of complex
polynomials. Additionally, we will analyze Newton’s method, a powerful tool for finding the roots of
polynomial equations. Throughout the chapter, we will delve into the concepts of polynomial functions
and their properties, as well as the algebraic manipulation of polynomials.

Keywords: Cardano formula, complex numbers, complex roots, cubic equation, factorization of polynomials,
factor theorem, fundamental theorem of algebra, Newton’s method, polynomial equations, polynomial
functions, polynomials, quad ratic equation, synthetic division.

6.1. Introduction
In this chapter, we will investigate a variety of polynomials and their mathematical properties. We will
analyze techniques and methods for reducing polynomials to simpler factors as we investigate the fascinating
world of polynomial factorization.
The factor theorem will provide us with a fundamental instrument to identify the factors of a given
polynomial, enabling us to gain a deeper understanding of its structure and behavior. This theorem will

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

93
94 6 Algebra of Polynomial Functions

allow us to identify the roots of a polynomial and decompose it into a product of linear and/or quadratic
factors.
Moreover, we will examine the fundamental theorem of algebra, which establishes a profound connection
between polynomials and complex roots. This theorem guarantees that every polynomial with complex
coefficients has a minimum of one complex root.
Newton’s method, an iterative numerical technique for locating the roots of a polynomial equation, will
be studied in order to tackle the topic of polynomial equations. This method is based on the concept of
progressively approximating the roots through a series of refined estimates.
We will also examine the properties and characteristics of polynomial functions, which are algebraic
functions defined by polynomials. These functions are fundamental to mathematical analysis and have
diverse applications in fields such as physics, economics, and engineering.
This chapter will provide a comprehensive examination of polynomials, their factorization, polynomial
equations, complex roots, and polynomial functions. These mathematical concepts and techniques will allow
us to comprehend and solve a broad variety of problems in various fields of study.

6.2. Polynomial Functions


A polynomial [24] is a mathematical expression that is formed using variables and coefficients, and is
constructed by combining algebraic terms through addition and multiplication operations. Each term consists
of a variable raised to a non-negative integer power, multiplied by a numerical coefficient. For example, the
simplest polynomial could be something like: 3x2 + 2x − 5.
Definition 6.1. A polynomial P(x) is an algebraic expression of the form:

P(x) = an xn + an−1 xn−1 + . . . + a2 x2 + a1 x + a0 (6.1)


where an , an−1 , . . . , a2 , a1 , a0 are the coefficients of the polynomial, x is the variable, and n is the degree
of the polynomial. The coefficients can be real or complex numbers, and n is a non-negative integer. Each
term in the polynomial consists of a power of the variable x multiplied by its respective coefficient.

6.3. Factorization of Polynomials


Polynomial factorization is the process of breaking down a polynomial into a multiplication of simpler
factors. A polynomial is an algebraic expression that contains a combination of variables and coefficients,
connected through addition and multiplication operations.
The goal of polynomial factorization is to find the factors that, when multiplied together, produce the original
polynomial. This decomposition is useful in many areas of mathematics, such as algebra, calculus, and
number theory.
Polynomial factorization can be performed in different ways, depending on the type of polynomial and the
factors being sought. Some common methods of factorization include:
6.4 The Factor Theorem 95

Common Factor: If a polynomial has a common factor in all its terms, it can be factored out by dividing
each term by that common factor. For example, the polynomial 2x3 + 4x2 has a common factor of 2x2 , so it
can be factored as 2x2 (x + 2).
Factoring by Grouping: Sometimes, it is possible to group terms within the polynomial in a way that allows
factoring by groups. This method is useful when the polynomial has four terms. For example, the polynomial
x3 + x2 + 4x + 4 can be factored by grouping the terms as (x3 + x2 ) + (4x + 4) and then factoring each
group separately: x2 (x + 1) + 4(x + 1). Then, a common factor of (x + 1) can be factored out, resulting in
(x + 1)(x2 + 4).
Factoring Perfect Square Trinomials: Some polynomials are perfect square trinomials, which means they can
be factored as the square of a binomial. A perfect square trinomial has the form a2 +2ab+b2 or a2 −2ab+b2 .
For example, the polynomial x2 + 4x + 4 is a perfect square trinomial and can be factored as (x + 2)2 .
Factoring by Difference of Squares: Some polynomials are the difference of two squares, which means they
can be factored using the identity a2 − b2 = (a + b)(a − b). For example, the polynomial x2 − 9 can be
factored as (x + 3)(x − 3).

6.4. The Factor Theorem


The Factor Theorem is a fundamental theorem in algebra that allows us to find the linear factors of a
polynomial and determine if a given number is a root of the polynomial. The theorem states that if a
polynomial f (x) has a root c, then f (x) is divisible by (x − c).

Theorem 6.1. Let P(x) be a polynomial and a be a real number. Then, (x − a) is a factor of P(x) if and only
if P(a) = 0.
Now, I will proceed to prove this theorem:

Proof. Suppose that P(x) is a polynomial of degree n and a is a real number. If (x − a) is a factor of P(x), we
can write P(x) as the product of (x − a) and another polynomial Q(x):

P(x) = (x − a) · Q(x) (6.2)

Where Q(x) is a polynomial of degree n − 1.


To prove the Factor Theorem, we need to show that P(a) = 0. To do this, let’s substitute x with a in equation
(2):

P(a) = (a − a) · Q(a) = 0 · Q(a) = 0 (6.3)

We have obtained that P(a) = 0, which demonstrates that if (x − a) is a factor of P(x), then P(a) = 0.
To prove the other direction, suppose that P(a) = 0. Then, a is a root of P(x). We can write P(x) in terms of
its root factorization:

P(x) = (x − a) · R(x) (6.4)


96 6 Algebra of Polynomial Functions

Where R(x) is a polynomial of degree n − 1.


If we substitute x with a in (Eq. 6.4), we obtain:

P(a) = (a − a) · R(a) = 0 · R(a) = 0 (6.5)

Since P(a) = 0, we have shown that if P(a) = 0, then (x − a) is a factor of P(x).


In conclusion, we have demonstrated that (x − a) is a factor of P(x) if and only if P(a) = 0, which completes
the proof of the Factor Theorem. 

The Factor Theorem can be expressed as follows:

If f (c) = 0, then f (x) = (x − c) · q(x) (6.6)

Where f (x) is the original polynomial, c is the root, and q(x) is the resulting quotient after dividing f (x) by
(x − c).
This means that if c is a root of f (x), then (x − c) is a factor of f (x). Therefore, we can use the Factor
Theorem to find linear factors and simplify the factoring of polynomials.
To apply the Factor Theorem, follow these steps:
(i) Given a polynomial f (x), set f (c) = 0 and find the value of c. This is achieved by solving the equation
f (x) = 0 for x.

(ii) If you find a value of c, it means that f (x) is divisible by (x − c). You can verify this by dividing f (x)
by (x − c) using polynomial long division.

(iii) If (x − c) is a factor of f (x), divide f (x) by (x − c) to obtain the quotient q(x). The quotient q(x)
represents the other factors of the polynomial f (x).

(iv) Repeat the process with the quotient q(x) to find more linear factors if necessary.

Once you have found all the linear factors using the Factor Theorem, you can write the original polynomial
f (x) as the product of the linear factors and any resulting quadratic or higher-order factors.

6.5. Solving Polynomial Equations


(i) Quadratic Equation:
ax2 + bx + c = 0, where a 6= 0 (6.7)

(ii) Quadratic Formula: √


−b ± b2 − 4ac
x= (6.8)
2a
6.5 Solving Polynomial Equations 97

(iii) Factoring: If the quadratic equation can be factored, set each factor equal to zero and solve for x.
(iv) Cubic Equation:
ax3 + bx2 + cx + d = 0, where a 6= 0 (6.9)

(v) Cardano’s Formula:


s r  s r 
3 q q 2  p 3 3 q q 2  p 3
x= − + + + − − + (6.10)
2 2 3 2 2 3

where

3ac − b2
p=
3a2
2b − 9abc + 27a2 d
3
q=
27a3

(vi) Synthetic Division: If one root is known, synthetic division can be used to divide the cubic equation by
(x − r), where r is the known root. This will result in a quadratic equation that can be solved using the
quadratic formula.
(vii) Numerical Methods: If the exact solutions cannot be found algebraically, numerical methods such as
Newton’s method can be used to approximate the roots.

Example 6.1. Quadratic Equation: Consider the equation 2x2 − 5x + 2 = 0.


Solution 6.1. Applying the quadratic formula, we have:
p
−(−5) ± (−5)2 − 4(2)(2)
x=
2(2)

5 ± 25 − 16
=
√4
5± 9
=
4
5±3
=
4

8 2 1
Therefore, the solutions are x1 = = 2 and x2 = = .
4 4 2

We can factorize the equation as (2x − 1)(x − 2) = 0. Setting each factor equal to zero gives 2x − 1 = 0 and
1
x − 2 = 0, which leads to the same solutions x1 = 2 and x2 = .
2
98 6 Algebra of Polynomial Functions

In (Fig. 6.1) it can be seen that the geometric resolution of a polynomial means finding the zeros of the
function in the plane.

3
2x2 − 5x + 2 = 0

y 1

−1

−2
−2 −1 0 1 2 3
x

Figure 6.1: Graphical representation of the function 2x2 − 5x + 2 = 0

Example 6.2. Cubic Equation: Consider the equation x3 − 3x2 + 3x − 1 = 0.


Solution 6.2. Given the equation:
x3 − 3x2 + 3x − 1 = 0

Comparing with the general equation:


x3 + ax2 + bx + c = 0
We use the change of variable x = y − 3a to simplify the equation and eliminate the x2 term.

With a = −3, the transformation is x = y + 1.


Substituting into the original equation, we obtain:

(y + 1)3 − 3(y + 1)2 + 3(y + 1) − 1 = 0

Expanding, we get:
y3 + 3y − 1 = 0

Comparing this to the canonical form:


y3 + py + q = 0
We find:
p=3
6.5 Solving Polynomial Equations 99

q = −1

Using the provided formula:


s r  s r 
3 q q 2  p 3 3 q q 2  p 3
x= − + + + − − +
2 2 3 2 2 3

Substitute in the values for p and q:


v s v s
u 2  3 u
−1 2
  3
t1
u
3 −1 3 t1
u
3 3
x= + + + − +
2 2 3 2 2 3

Calculating inside the cube roots:


s r s r
3 1 1 3 1 1
x= + +1+ − +1
2 4 2 4

Remember, this x is in terms of y due to our change of variable. To revert to the original variable x, we need
to subtract 1 from the final result.

f (x)

x
-1 1 2

-1

Figure 6.2: Graphical representation of the function f (x) = x3 − 3x2 + 3x − 1 = 0

Example 6.3. Quadratic Equation: Given f (x) = x2 − 2, we will find the solutions using the quadratic
formula.

Solution 6.3. The equation is of the form ax2 + bx + c = 0, where a = 1, b = 0, and c = −2. Applying the
quadratic formula, we have:
100 6 Algebra of Polynomial Functions

−b ± b2 − 4ac
x=
p2a
−0 ± 02 − 4 · 1 · (−2)
=
√ 2·1
± 8
=
2

=± 2

√ √
Therefore, the solutions of f (x) = x2 − 2 are x = 2 and x = − 2.

Example 6.4. Factoring Method Given f (x) = x2 − 2, we will find the solutions through factoring.
√ √
Solution 6.4.√We observe √ that the equation can be factored as (x + 2)(x − 2) = 0. This implies that the
factors (x + 2) and (x − 2) must be equal to zero for the equation to hold.
√ √ √ √
This gives us the solutions x + 2 = 0 and x − 2 = 0, which simplify to x = − 2 and x = 2 respectively.
√ √
Therefore, the solutions of f (x) = x2 − 2 are x = 2 and x = − 2.

6.6. Fundamental Theorem of Algebra


The Fundamental theorem of Algebra is a toral result in mathematics that relates to the roots of polynomial
equations. It states that every non-constant polynomial with complex coefficients can be factored completely
into linear and quadratic factors, and hence has at least one complex root.

Theorem 6.2. Every non-constant polynomial with complex coefficients has at least one complex root.

Proof. Consider a non-constant polynomial P(z) with complex coefficients, where z is a complex variable.
Suppose, to the contrary, that P(z) has no complex roots.
If P(z) has no complex roots, then P(z) can only have real roots. In this case, we can write P(z) as a product
of irreducible linear factors over the real numbers:

P(z) = (an zn + an−1 zn−1 + . . . + a1 z + a0 ) = (bk z − r1 )(bk−1 z − r2 ) . . . (b1 z − rk ) (6.11)

Where an , an−1 , . . . , a0 , bk , bk−1 , . . . , b1 are real coefficients, and r1 , r2 , . . . , rk are the real roots of the polynomial.
However, this contradicts the fact that P(z) has no complex roots.
Therefore, our initial assumption that P(z) has no complex roots must be incorrect.
We conclude that every non-constant polynomial with complex coefficients must have at least one complex
root.
Thus, we have proved the Fundamental Theorem of Algebra. 
6.6 Fundamental Theorem of Algebra 101

To understand this theorem, let’s consider a polynomial function with complex coefficients:

P(z) = an zn + an−1 zn−1 + . . . + a1 z + a0 (6.12)

Where n is a positive integer, z is a complex variable, and an , an−1 , . . . , a1 , a0 are complex coefficients.
According to the Fundamental Theorem of Algebra, we can factorize the polynomial P(z) into linear and
quadratic factors of the form:

P(z) = an (z − z1 )(z − z2 ) . . . (z − zn ) (6.13)

Where z1 , z2 , . . . , zn are complex numbers and correspond to the roots of the polynomial equation P(z) = 0.
This theorem guarantees the existence of at least one complex root for any non-constant polynomial. It
implies that there is always a solution to the equation P(z) = 0 in the complex number system.

Example 6.5. Let’s take an example to illustrate the theorem. Consider the polynomial:

P(z) = z2 + 2z + 1 (6.14)

To find the roots of this polynomial, we set P(z) = 0 and solve for z:

z2 + 2z + 1 = 0 (6.15)

Factoring the polynomial, we get:

(z + 1)2 = 0 (6.16)

This equation has a repeated root z = −1. Thus, the polynomial P(z) has a complex root z = −1, which
satisfies P(z) = 0.

The Fundamental Theorem of Algebra has significant and far-reaching implications in mathematics and
numerous scientific disciplines. Here are some of its most important implications:
Every nonconstant polynomial with complex coefficients has at least one complex root, according to this
theorem. This implies that algebraic equations in the field of complex numbers always have solutions.
This property is essential to the study of equations and systems of equations in numerous fields, including
numerical analysis, theoretical physics, and engineering.
The theorem guarantees that any non-constant polynomial over the field of complex numbers can be
thoroughly factored. This provides an essential instrument for studying polynomials and their properties.
Polynomial factorization is crucial in algebra, calculus, number theory, and numerous other mathematical
disciplines.
The theorem has implications for algebraic geometry, the study of intersections of curves and surfaces
defined by polynomials. The fact that every polynomial has at least one complex root implies that every
algebraic curve (defined by a polynomial) has complex-plane intersection points. This permits a thorough
investigation of the geometric properties of algebraic curves and surfaces.
102 6 Algebra of Polynomial Functions

Applications in physics and engineering sciences: algebraic equations and polynomials are used to represent
numerous physical laws and phenomena. The Fundamental Theorem of Algebra guarantees that these
mathematical models have solutions in this intricate field, facilitating the analysis and comprehension of
physical systems. In addition, techniques founded on the theorem are essential for the design of algorithms
and systems in communication, signal processing, automatic control, and other engineering fields.
The Fundamental Theorem of Algebra is a significant and applicable mathematical result. Its validity
guarantees the existence of complex roots for algebraic equations, enables the factorization of polynomials,
and has a variety of applications in algebraic geometry, physics, and engineering. It is a cornerstone of the
study of mathematics and provides essential instruments for comprehending and modeling mathematical and
scientific phenomena.

6.7. Finding Complex Roots Using Newton’s Method


Consider the polynomial equation:

f (x) = x3 + 4x2 + 5x + 2 = 0 (6.17)

To find the complex roots of this equation, we can use Newton’s method. Let’s start with an initial guess of
x0 = −1. Now, let’s calculate the values of f (xn ) and f ′ (xn ) at each iteration and refine the approximation:
Iteration 1:
x0 = −1
f (x0 ) = (−1)3 + 4(−1)2 + 5(−1) + 2
= −2 (6.18)
′ 2
f (x0 ) = 3(−1) + 8(−1) + 5
=6

Using the iterative formula, we can calculate x1 as follows:

f (x0 )
x1 = x0 −
f ′ (x0 )
−2
= −1 −
6 (6.19)
1
= −1 +
3
2
=−
3
Iteration 2:
6.7 Finding Complex Roots Using Newton’s Method 103

2
x1 = −
3
2 3 2 2
     
2
f (x1 ) = − +4 − +5 − +2
3 3 3

≈ 0.1111 (6.20)

2 2
   
2
f ′ (x1 )= 3 − +8 − +5
3 3

≈ 7.3333

Using the iterative formula, we can calculate x2 as follows:

f (x1 )
x2 = x1 −
f ′ (x1 )

2 0.1111 (6.21)
=− −
3 7.3333

≈ −0.0170

By repeating this process, we can continue refining the approximation until we reach the desired level
of accuracy. In this case, we stop after three iterations. Therefore, an approximate complex root of the
polynomial equation f (x) = 0 is x ≈ −0.0170.

f (x)
f (x) = x3 + 4x2 + 5x + 2 4

x
−3 −2 −1 1

−2

−4

Figure 6.3: Real roots of the function f (x) = x3 + 4x2 + 5x + 2


104 6 Algebra of Polynomial Functions

The FORTRAN-95 program that simulates Newton’s method can be consulted at (Prog A.2).

6.8. Conclusions
In this chapter, we examined various aspects of polynomials, including their factorization and factor theorem,
as well as their applicability in solving equations and analyzing functions. The fundamental theorem of
algebra established the relationship between polynomials and complex roots. In addition, we examined
Newton’s method for approximating the roots of polynomials and learned how polynomials are used to
represent a variety of mathematical and applied science situations.
6.9 Exercises 105

6.9. Exercises
Exercise 6.1. Solve the following quadratic equation:

3x2 + 5x − 2 = 0 (6.22)

Exercise 6.2. Find the real solutions of the following cubic equation:

x3 − 4x2 + 3x + 2 = 0 (6.23)

Exercise 6.3. Determine all complex solutions of the following quadratic equation:

x2 + 2x + 5 = 0 (6.24)

Exercise 6.4. Solve the following quartic equation using the factorization method:

x4 − 5x2 + 4 = 0 (6.25)

Exercise 6.5. A farmer wants to fence off a rectangular area on his land to separate it from the rest. He
has calculated that the total area he wants to enclose is represented by the polynomial A(x) = 2x2 − 5x + 2,
where x is the length in meters of one side of the rectangle, and A(x) represents the area in square meters.
The farmer needs to determine the appropriate dimensions for his fence, so he needs to find the roots of the
polynomial A(x). Help the farmer calculate the roots of the polynomial to determine the possible lengths of
the sides of the rectangle.

Exercise 6.6. An object is thrown from an initial height of 100 meters with an initial velocity of 20 meters
per second. The vertical motion of the object is given by the equation:

h(t) = −4.9t 2 + 20t + 100 (6.26)

Where h(t) represents the height in meters and t is the time in seconds.
Find the time at which the object hits the ground, i.e., when h(t) = 0.

Exercise 6.7. Find the complex roots of f (x) = x4 + 2x3 + 3x2 + 4x + 5


Chapter 7
Algebra of Vectors

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Keywords: cross product, dot product, linear independence, linear transformations, spanning sets, vector
dimension, vectors, vector space, vector subspace.

7.1. Introduction
In this chapter, we will explore fundamental concepts related to vector spaces and linear transformations.
Understanding these concepts is essential in various fields of mathematics and its applications, such as
physics, computer science, and engineering. We will delve into the properties and operations involving
vectors, as well as the foundational principles that govern their behavior.
The chapter begins by introducing vectors as mathematical objects that possess both magnitude and
direction. We will study the dot product, a fundamental operation that allows us to quantify the relationship
between vectors, measuring their alignment and providing insights into angles and projections. Additionally,
we will investigate the cross product, which yields a new vector perpendicular to the original vectors and
plays a crucial role in three-dimensional geometry and physics.
Next, we will explore the notion of linear independence, which describes a set of vectors that cannot be
expressed as a linear combination of the others. We will examine its implications and the concept of spanning

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

107
108 7 Algebra of Vectors

sets, which are sets of vectors that can generate any vector within a given vector space. The vector dimension
will also be discussed, providing a measure of the ”size” or ”complexity” of a vector space.
Furthermore, we will delve into the concept of a vector space, which is a collection of vectors equipped with
operations of addition and scalar multiplication, satisfying specific axioms. We will examine the properties
and characteristics of vector spaces, including their subspaces, which are subsets that themselves form vector
spaces under the inherited operations.
Lastly, we will explore linear transformations, which are functions that preserve vector operations and
properties within vector spaces. We will study their properties, behavior, and the relationship between linear
transformations and matrices.
Throughout this chapter, we will provide definitions, theorems, and examples to solidify the understanding
of these concepts. Developing a firm grasp of these fundamental principles will serve as a stepping stone for
further exploration of advanced topics in linear algebra and beyond.

7.2. Vectors
Vectors [25, 26] are mathematical objects that represent both magnitude and direction. They are used to
describe quantities such as displacement, velocity, and force in physics and mathematics. Vectors can be
visualized as arrows in space, with the length of the arrow representing the magnitude and the direction of
the arrow indicating the direction.

Definition 7.1. A vector v in a vector space V is an element of V that satisfies the following properties:

(i) Addition: ∀u, w ∈ V, u + w ∈ V

(ii) Scalar Multiplication: ∀v ∈ V, ∀α ∈ R, α v ∈ V

In this formal definition, v represents a vector in a vector space V , and R represents the set of real numbers.
Property (i) states that the addition of two vectors in V results in another vector in V . Property (i) states that
multiplying a vector v by a scalar α produces another vector in V .

Vectors have magnitude (or length) and direction, and they can be added, subtracted, and multiplied by
scalars. They play a fundamental role in many branches of mathematics and physics, including linear algebra,
calculus, and mechanics.

7.3. Vector Space


Definition 7.2. A vector space V over a field F is a non-empty set of elements called vectors, along with
two operations, vector addition and scalar multiplication, that satisfy the following axioms:
7.3 Vector Space 109

(i) Closure under vector addition: For all u, v ∈ V, the sum u + v is in V.

(ii) Associativity of vector addition: For all u, v, w ∈ V, (u + v) + w = u + (v + w).

(iii) ) Existence of additive identity: There exists a vector 0 ∈ V, called the zero vector, such that v + 0 = v
for all v ∈ V.

(iv) Existence of additive inverses: For each v ∈ V, there exists a vector −v ∈ V, called the additive inverse
of v, such that v + (−v) = 0.

(v) Closure under scalar multiplication: For all α ∈ F and all v ∈ V, the scalar product α v is in V.

(vi) Distributivity of scalar multiplication over vector addition: For all α ∈ F and all u, v ∈ V, α (u + v) =
α u + α v.

(vii) Distributivity of scalar multiplication over scalar addition: For all α , β ∈ F and all v ∈ V, (α + β )v =
α v + β v.

(viii) Associativity of scalar multiplication: For all α , β ∈ F and all v ∈ V, α (β v) = (αβ )v.

3 3 are vectors
  the vector space R over the field of real numbers R. The elements of R 
Example 7.1. Consider 
x u1 + v1
of the form v = y, where x, y, z ∈ R. Vector addition in R3 is defined component-wise: u + v = u2 + v2 .
z u3 + v3
Scalarmultiplication
 is performed by multiplying each component of the vector by the corresponding scalar:
αx
α v = α y, where α ∈ R.
αz
In this example, R3 is a vector space where vector addition and scalar multiplication satisfy the axioms
defined above.
110 7 Algebra of Vectors

v u+v

Figure 7.1: Sum of two vectors in the Cartesian plane.

Example 7.2. Given the distributive equality between vectors:

a · (b + c) = a · b + a · c (7.1)

Proof. The dot product of vector a with the sum of vectors b and c is equal to the sum of the dot products of
vector aa with vector b and vector a with vector c.
Where a, b, and c are vectors in a vector space.
To prove this, let’s consider the vectors a, b, and c in coordinate form:
     
a1 b1 c1
a2  b2  c2 
a =  . , b =  . , c =  .  (7.2)
     
 ..   ..   .. 
an bn cn

Then, we can write the left-hand side expression of the distributive equality as:
  b  c 
a1 1 1
 a2   b2  c2 
· · · · 
a · (b + c) =     .  +  .  (7.3)
  
 ..   .. 
an bn cn

Using the definition of the dot product of vectors, this can be expressed as:

a · (b + c) = a1 (b1 + c1 ) + a2 (b2 + c2 ) + · · · + an (bn + cn ) (7.4)

Next, let’s expand the right-hand side expression of the distributive equality:
7.4 Dot Product 111

       
a1 b1 a1 c1
a2  b2  a2  c2 
a·b+a·c =  . · . + . · .  (7.5)
       
 ..   ..   ..   .. 
an bn an cn

Using the definition of the dot product, this can be written as:

a · b + a · c = a1 b1 + a2 b2 + · · · + an bn + a1 c1 + a2 c2 + · · · + an cn (7.6)

Now, let’s compare the expressions obtained for the left-hand side and right-hand side of the distributive
equality:

a1 (b1 + c1 ) + a2 (b2 + c2 ) + . . . + an (bn + cn )


(7.7)
= a1 b1 + a2 b2 + . . . + an bn + a1 c1 + a2 c2 + . . . + an cn

We can observe that the two expressions are identical, which demonstrates the validity of the distributive
equality between vectors. 

7.4. Dot Product


The dot product is a mathematical operation applied to two vectors in Euclidean space. It is denoted as · or
h·, ·i and is defined as the sum of the products of the corresponding components of the two vectors.

Definition 7.3. Formally, for two vectors a = (a1 , a2 , ..., an ) and b = (b1 , b2 , ..., bn ), the dot product is
defined as:
n
a · b = ∑ ai bi = a1 b1 + a2 b2 + ... + an bn (7.8)
i=1

In this expression, n is the dimension of vectors a and b, and ai and bi are the corresponding components of
the vectors.

Example 7.3. Let’s consider the vectors u = (2, 3) and v = (4, −1). To calculate their dot product, we use
the dot product formula:

u · v = (2)(4) + (3)(−1) = 8 − 3 = 5 (7.9)

Therefore, the dot product between u and v is equal to 5.

The dot product is a fundamental operation in linear algebra and has various applications in geometry,
physics, and calculus, such as calculating angles, projections, and work done by a force.
112 7 Algebra of Vectors

Theorem 7.1. If a · b = 0 ⇐⇒ a ⊥ b
Solution 7.1. Suppose a · b = 0. We want to prove that a is perpendicular to b.
We know that the dot product is defined as:

a · b = |a||b|cos(θ ) (7.10)

Where |a| and |b| are the magnitudes of vectors a and b respectively, and θ is the angle between them.
If a · b = 0, then we have:

0 = |a||b|cos(θ ) (7.11)

Since |a| and |b| are non-negative magnitudes, we can divide both sides of the equation by |a||b|, which
gives us:

cos(θ ) = 0 (7.12)

Recall that the cosine of an angle is zero if and only if the angle is a right angle (90 degrees). Therefore, we
conclude that θ = 90◦ .
If the angle between two vectors is a right angle, we say that the vectors are perpendicular or orthogonal.
Thus, if a · b = 0, then a is perpendicular to b.
For the other direction, suppose a is perpendicular to b. This implies that the angle between them is a right
angle (θ = 90◦ ).
Using the definition of the dot product again, we have:

a · b = |a||b|cos(θ ) (7.13)

If θ = 90◦ , then cos(θ ) = 0, which implies a · b = 0.


Therefore, we have shown that a · b = 0 if and only if a is perpendicular to b. 

7.5. Cross Product

Definition 7.4. The cross product, denoted as u × v, is an operation between two vectors in a three-
dimensional space that yields a new vector perpendicular to both input vectors. Formally, the cross product
is defined as:
 
u2 v3 − u3 v2
u × v = u3 v1 − u1 v3  (7.14)
u1 v2 − u2 v1
7.5 Cross Product 113
   
u1 v1
Where u = u2  and v = v2  are the input vectors.
u3 v3
   
2 4
Example 7.4. Let’s consider the vectors u = −1 and v =  2 . Applying the cross product formula, we
3 −1
have:
   
(−1)(−1) − (3)(2) −5
u × v =  (3)(4) − (2)(−1)  =  14  (7.15)
(2)(2) − (−1)(4) 0
 
−5
Therefore, the cross product of u and v is  14 .
0

Example 7.5. Suppose we have two vectors in three-dimensional space, A and B, defined by their components:

A = (A1 , A2 , A3 ) (7.16)
B = (B1 , B2 , B3 ) (7.17)

The area of the parallelogram formed by vectors A and B can be calculated using the cross product as
follows:
Compute the cross product of vectors A and B:

C = A×B (7.18)

The cross product C is calculated as:

C = (A2 · B3 − A3 · B2 , A3 · B1 − A1 · B3 , A1 · B2 − A2 · B1 ) (7.19)

Compute the magnitude of the resulting vector C:


q
|C|= C12 +C22 +C32 (7.20)

The area of the parallelogram formed by vectors A and B is equal to the magnitude of vector C:

Area = |C| (7.21)

Let’s consider a specific example:


Suppose we have vectors A = (2, 3, 1) and B = (4, −1, 2). We will calculate the area of the parallelogram
formed by these vectors:
Compute the cross product C:
114 7 Algebra of Vectors

C = (2 · 2 − 3 · (−1), 3 · 2 − 2 · 4, 2 · (−1) − 2 · 3) = (4 + 3, 6 − 8, −2 − 6) = (7, −2, −8) (7.22)

Compute the magnitude of C:


q √ √
|C|= 72 + (−2)2 + (−8)2 = 49 + 4 + 64 = 117 ≈ 10.82 (7.23)

The area of the parallelogram formed by vectors A and B is approximately 10.82 square units.

7.6. Vector Subspace


Definition 7.5. A subspace W of a vector space V is a non-empty set of vectors in V that satisfies the
following conditions:
(i) Closure under addition: for all u, v ∈ W , u + v ∈ W .

(ii) Closure under scalar multiplication: for all u ∈ W and any scalar α , α u ∈ W .

The subspace W must inherit the addition and scalar multiplication operations from the vector space V and
must itself be a vector space.
Example 7.6. Let’s consider the vector space R3 , that is, the set of all three-dimensional vectors with real
components. An example of a subspace of R3 would be the plane W defined by:

W = {(x, y, z) ∈ R3 : x + 2y − 3z = 0} (7.24)

To demonstrate that W is a subspace of R3 , we need to verify that it satisfies the two conditions mentioned
earlier. If we take two vectors (x1 , y1 , z1 ) and (x2 , y2 , z2 ) in W , we can verify that their sum also belongs to
W by satisfying the equation x1 + 2y1 − 3z1 = 0 and x2 + 2y2 − 3z2 = 0. Additionally, when we multiply a
vector (x, y, z) in W by a scalar α , it satisfies the equation α x + 2α y − 3α z = 0, which indicates that α (x, y, z)
is also in W . Therefore, W is a subspace of R3 .

7.7. Linear Independence


Linear independence is a fundamental concept in linear algebra. It refers to the relationship among a set of
vectors and how they are linearly combined.
In simple terms, a set of vectors is considered linearly independent if none of the vectors can be expressed
as a linear combination of the other vectors in the set. This means that there is no linear dependence among
them.
More formally, if we have a set of vectors {v1 , v2 , ..., vn }, where each vector has components in a vector
space, they are said to be linearly independent if the only solution to the equation:
7.8 Spanning Sets 115

a1 v1 + a2 v2 + ... + an vn = 0 (7.25)

is

a1 = a2 = ... = an = 0 (7.26)

In other words, all the coefficients of the linear combination are zero.
On the other hand, if there exists at least one linear combination of the vectors that equals the zero vector,
where at least one coefficient is nonzero, then the set of vectors is considered linearly dependent.
Linear independence means that the vectors in a set cannot be expressed as linear combinations of each
other, implying that each vector in the set contributes unique information to the vector space it belongs to.

Definition 7.6. A set of vectors {v1 , v2 , ..., vn } in a vector space V is considered linearly independent if
no nontrivial linear combination of these vectors equals the zero vector. Formally, this means that if the
constants α1 , α2 , ..., αn are not all zero, then the equation.

α1 v1 + α2 v2 + ... + αn vn = 0 (7.27)

Does not hold.


   
1 0
Example 7.7. Let’s consider the set of vectors {v1 , v2 } in R2 ,where v1 = and v2 = . To show that
0 1
these vectors are linearly independent, let’s assume that there exist constants α1 and α2 such that
 
0
α1 v1 + α2 v2 = (7.28)
0

This implies the system of equations.

α1 = 0
(7.29)
α2 = 0

The only solution to this system is α1 = α2 = 0, confirming that {v1 , v2 } is a linearly independent set.
This example illustrates that the vectors v1 and v2 cannot be expressed as a nontrivial linear combination
that results in the zero vector, demonstrating their linear independence.

7.8. Spanning Sets


A spanning set, or a set that generates a vector space, is a set of vectors that, through linear combinations,
can form any vector within the vector space. In other words, a set S is a spanning set for a vector space V if
every vector in V can be expressed as a linear combination of the vectors in S.
116 7 Algebra of Vectors

Definition 7.7. Formally, given a vector space V , a set S = v1 , v2 , ..., vn is a spanning set for V if any vector
v in V can be expressed as a linear combination of the vectors in S:

v = α1 v1 + α2 v2 + ... + αn vn (7.30)

Where α1 , α2 , ..., αn are scalars.


Example 7.8. Consider the vector space R2 , which represents the two-dimensional Euclidean plane. A
spanning set for this vector space can be the following set:
   
1 0
S={ , } (7.31)
0 1
2
 vector in R can be formed through linear combinations of the vectors in S. For example, the vector
Every

2
can be expressed as:
3
     
2 1 0
=2 +3 (7.32)
3 0 1

This example demonstrates that the set S is a spanning set for the vector space R2 .

7.9. Vector Dimension


Definition 7.8. The dimension of a vector refers to the number of components required to fully describe the
vector in a given space. It is denoted as dim(V ), where V is the vector space.
Definition 7.9. The basis of a vector is a set of linearly independent vectors that span the entire vector space.
Each vector in the space can be uniquely expressed as a linear combination of the basis vectors.
Example 7.9. Let’s consider the vector space R3 , which represents three-dimensional space. The dimension
of this space is 3, as it requires 3 components (coordinates x, y, and z) to fully describe any vector in this
space.
A possible basis for R3 is e1 , e2 , e3 , where:
     
1 0 0
e1 = 0 , e2 = 1 , e3 = 0 (7.33)
0 0 1

These basis vectors are linearly independent, and any vector inR3 
can be uniquely expressed as a linear
2
combination of these basis vectors. For example, the vector v = −1 can be written as:
3
v = 2e1 − e2 + 3e3 (7.34)

Here, the coefficients 2, -1, and 3 represent the x, y, and z coordinates, respectively. The vectors e1 , e2 , and
e3 form a canonical basis for the three-dimensional vector space R3 .
7.11 Conclusions 117

7.10. Linear Transformations


Definition 7.10. A linear transformation, also known as a linear map, is a function between two vector
spaces that preserves the vector addition and scalar multiplication operations. Formally, a linear transformation
T : V → W between vector spaces V and W satisfies the following properties:

T (u + v) = T (u) + T (v) ∀u, v ∈ V (7.35)

T (α u) = α T (u) ∀u ∈ V, ∀α ∈ R (7.36)

These properties indicate that a linear transformation preserves the vector structure, meaning that the sum of
vectors remains the same after applying the transformation, and scaled vectors are transformed by the same
factor.

Example 7.10. Let’s consider a linear transformation T : R2 → R2 defined by the matrix:


 
1 2
A= (7.37)
−1 3

This linear transformation can be represented as:

T (x) = Ax
 
x
Where x = 1 is a column vector in R2 .
x2
 
2
For example, if we take x = , the linear transformation is calculated as:
1
      
2 2 −1 2 3
T = =
1 3 0 1 6

   
2 3
Therefore, the linear transformation T maps the vector x = to the vector in the vector space R2 .
1 6

7.11. Conclusions
This chapter of the book explored fundamental mathematical concepts related to vectors and vector spaces.
Topics covered included properties and operations such as the cross product and dot product, as well as
linear independence and linear transformations. Additionally, key concepts such as spanning sets, vector
dimension, and vector subspaces were addressed. These concepts are essential for understanding and
applying various branches of mathematics and physics, providing powerful tools for analysis and problem-
solving in these fields.
118 7 Algebra of Vectors

7.12. Exercises
   
1 −2
Exercise 7.1. Find the cross product of the vectors u =  2  and v =  3 .
−1 4
   
2 −1
Exercise 7.2. Calculate the dot product of the vectors u = −3 and v =  4 .
1 2
     
1 2 −1
Exercise 7.3. Determine if the vectors v1 = 2, v2 = −1, and v3 =  3  are linearly independent.
3 4 5
 
  2x
x
Exercise 7.4. Consider the linear transformation T : R2 → R3 defined by T =  3y . Calculate
y
  x+y
1
T .
−2
 
x  
3 2 2x + y
Exercise 7.5. Consider the linear transformation T : R → R defined by T   y   = . Find the
3y − z
z
 
1
image of −2 under the transformation T .
4
 
  3x
x
Exercise 7.6. Consider the linear transformation T : R2 → R3 defined by T =  −2y . Determine
y
x+y
if T preserves vector addition, i.e., if T (u + v) = T (u) + T (v) for any u, v ∈ R2 .
 
x  
x+y+z
Exercise 7.7. Consider the linear transformation T : R3 → R2 defined by T y = . Find the
2x − y
z
matrix of the linear transformation T .

Exercise 7.8. Suppose you have two forces applied to an object in a two-dimensional plane. The first force
F1 has a magnitude of 10 N and an angle of 30 degrees with respect to the positive x-axis. The second force
F2 has a magnitude of 15 N and an angle of 60 degrees with respect to the positive x-axis.
For this exercise, you need to find the resultant force Fr acting on the object, using only vector algebra.
Chapter 8
Algebra of Matrices

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract This chapter focuses on the fundamental concepts of linear algebra, including Cramer’s rule,
determinants, diagonalization of a matrix, eigenvectors, the inverse of a matrix, matrices, matrix operations,
and solving linear equation systems. This chapter provides a thorough overview of these fundamental ideas
and their practical applications. It examines Cramer’s rule as an effective method for solving systems of
linear equations, introduces determinants as essential determinants of matrix properties, and illustrates
the significance of diagonalization and eigenvectors in matrix transformations. The inverse of a matrix is
investigated as a crucial operator for solving linear systems. Throughout the chapter, matrix operations such
as addition, multiplication, and manipulation of matrices through fundamental operations are discussed in
depth. This chapter lays the essential groundwork for further study of advanced topics in linear algebra and
their applications in various disciplines, including computer science, engineering, and physics.

Keywords: Cramer’s rule, determinant operations, determinant operator, diagonalization of a matrix, eigenvalues,
eigenvectors, inverse of a matrix, matrices, matrix operations, matriz operator, square matrix, system linear
equations.

8.1. Introduction
In this chapter, we examine the fundamental concepts of linear algebra, which serve as the foundation
for a vast array of mathematical applications. Cramer’s rule is an effective method for solving systems

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

119
120 8 Algebra of Matrices

of linear equations by employing determinants. Through the lens of determinants, we gain insight into the
properties and behavior of matrices, which are rectangular arrays of numbers on which a variety of algebraic
operations are defined. We investigate matrix operations, such as addition and multiplication, which allow us
to manipulate matrices and reveal their concealed relationships. As we progress, we uncover the significance
of the diagonalization of a matrix and its relationship to eigenvectors, which play an essential role in
numerous mathematical and scientific fields. In addition, we explain the concept of a matrix’s inverse and its
significance when solving linear systems. Join us as we explore the intricate world of linear algebra, where
the study of matrices, determinants, diagonalization, eigenvectors, and matrix operations brings systems of
linear equations to life.

8.2. Matrices
A matrix is an ordered collection of elements arranged in rows and columns.

Definition 8.1. From a mathematical standpoint, a matrix is a rectangular arrangement of elements organized
into rows and columns. It is generally represented using a capital letter with subscripts to indicate the position
of each element. A matrix of size m × n has m rows and n columns. The individual elements of a matrix can
be represented as ai j , where i indicates the row position and j indicates the column position. For example, a
matrix A of size 3 × 2 can be represented as:
 
a11 a12
A = a21 a22  (8.1)
a31 a32

Here, a11 , a12 , a21 , a22 , a31 , and a32 are the individual elements of the matrix. Matrices are used in various
fields of mathematics and science to represent data, perform calculations, and solve systems of equations,
among other applications.

Example 8.1. Let’s suppose we have a matrix A of size 2 × 3 and a matrix B of size 3 × 2:
 
  12
246
A= , B = 3 4 (8.2)
135
56

We can perform mathematical operations with these matrices, such as matrix multiplication. If we multiply
A by B, we obtain a new matrix C of size 2 × 2:
 
  12  
246   44 56
C = A·B = · 34 = (8.3)
135 29 38
56

This matrix C is the result of the multiplication of A and B. Matrix calculations and manipulations are
fundamental in linear algebra and have numerous applications in science, engineering, and other fields.
8.3 Matrix operations 121

8.3. Matrix operations


Definition 8.2. The addition of two matrices of the same size is performed by adding the corresponding
elements. Formally, if A and B are matrices of the same size m × n, then the sum C = A + B is defined as:

Ci j = Ai j + Bi j (8.4)

Where Ci j represents the element in the position (i, j) of the resulting matrix C.
   
24 12
Example 8.2. Let’s consider A = and B = . The sum C = A + B is calculated as follows:
13 35
   
2+1 4+2 36
C= = (8.5)
1+3 3+5 48

Definition 8.3. The subtraction of two matrices of the same size is performed by subtracting the corresponding
elements. Formally, if A and B are matrices of the same size m × n, then the difference C = A − B is defined
as:

Ci j = Ai j − Bi j (8.6)

Where Ci j represents the element in the position (i, j) of the resulting matrix C.

Example 8.3. Using the matrices A and B from the previous example, the difference D = A − B is calculated
as follows:
   
2−1 4−2 1 2
D= = (8.7)
1−3 3−5 −2 −2

Definition 8.4. The multiplication of two matrices A and B is performed by multiplying the rows of A by the
columns of B and summing the corresponding products. Formally, if A is a matrix of size m × p and B is a
matrix of size p × n, then the product C = A · B is defined as:
p
Ci j = ∑ Aik · Bk j (8.8)
k=1

Where Ci j represents the element in the position (i, j) of the resulting matrix C.
   
12 56
Example 8.4. Let’s consider the matrices A = and B = . The product C = A · B is calculated as
34 78
follows:
   
(1 · 5) + (2 · 7) (1 · 6) + (2 · 8) 19 22
C= = (8.9)
(3 · 5) + (4 · 7) (3 · 6) + (4 · 8) 43 50
122 8 Algebra of Matrices

8.4. Determinants
The determinant is a numerical measure associated with a square matrix. It provides information about how
the matrix transforms space and whether linear equations have a unique solution. The determinant helps us
determine if a matrix is singular or invertible.
The determinant is a mathematical function applied to a square matrix, resulting in a number. Formally,
the determinant is denoted as det(A), where A is a square matrix. The value of the determinant depends on
the size and elements of the matrix and is calculated using specific formulas based on the matrix size. The
determinant can be positive, negative, or zero, and it provides information about the properties and structure
of the matrix.

Example 8.5. Let’s consider the square matrix A:


 
32
A= (8.10)
14

To calculate the determinant of A, we use the specific formula for 2 × 2 matrices:

det(A) = (3 × 4) − (2 × 1) = 12 − 2 = 10 (8.11)

Therefore, the determinant of matrix A is 10. This value indicates that matrix A has a linear transformation
that does not distort space and has unique solutions for the associated linear equations.

Example 8.6. Calculate the value of the determinant of the following 3x3 matrix:
 
2 4 1
A = −1 0 3  (8.12)
5 2 −2

Solution 8.1. Using the method of cofactors for computing the determinant:

0 3 −1 3 −1 0
det(A) = 2 × −4× +1×
2 −2 5 −2 5 2

For a 2 × 2 matrix:
ab
= ad − bc
cd

Applying this:
det(A) = 2(−6) − 4(2 − 15) + 1(−2 − 0)
det(A) = −12 + 52 − 2
det(A) = 38

Thus, the determinant of A is 38.


8.6 Adjugate Matrix 123

8.5. Determinant Operations


Linearity property with respect to a row (or column):
     
a b c abc abc
det  d + e f + g h + i  = det  d f h  + det  e g i 
j k l j k l jkl
     
1 2 3 1 2 3 1 2 3
Example 8.7. det  4 + 5 6 + 7 8 + 9  = det  4 6 8  + det  5 7 9 
10 11 12 10 11 12 10 11 12

Interchange property of rows (or columns):


   
abc gh i
det  d e f  = − det  d e f 
gh i abc
   
123 78 9
Example 8.8. det  4 5 6  = − det  4 5 6
789 12 3

Multiplication property of a row (or column) by a constant:


   
a b c abc
det  k · d k · e k · f  = k · det  d e f 
g h i gh i
   
1 2 3 12 3
Example 8.9. det  2 · 4 2 · 5 2 · 6  = 2 · det  4 5 6
7 8 9 78 9

8.6. Adjugate Matrix


The adjugate matrix is obtained by interchanging the rows and columns of the matrix of cofactors.
For each element of the original matrix, its cofactor is calculated by multiplying the determinant of the matrix
obtained by removing the row and column of that element by the corresponding sign. Then, the cofactors
are placed in the opposite position to that of the original elements.
The adjugate matrix has applications in calculating the inverse matrix and solving systems of linear equations
using Cramer’s rule.
 
21
Example 8.10. Given the matrix: A =
13
124 8 Algebra of Matrices

Solution 8.2. To calculate the adjugate matrix of A, we first need to find the cofactor matrix and then
transpose it.
The cofactor matrix is obtained by taking the cofactor of each element in the original matrix. The cofactor of
an element ai j is calculated as (−1)i+ j multiplied by the determinant of the submatrix obtained by removing
the i-th row and j-th column.
 
3 −1
In this case, the cofactor matrix is: C =
−1 2
Finally, to obtain the adjugate matrix of A, we simply transpose the cofactor matrix:
 T  
3 −1 3 −1
adj(A) = = (8.13)
−1 2 −1 2

 
3 −1
Therefore, the adjugate matrix of the given matrix is: adj(A) =
−1 2

8.7. Inverse of a Matrix


The inverse of a matrix is a concept used in linear algebra for square matrices. If a matrix has an inverse, it
can be considered as a kind of “opposite” of the original matrix. Multiplying a matrix by its inverse results
in the identity matrix. In other words, if you have a matrix A and its inverse A−1 , then A · A−1 equals the
identity matrix.
The inverse of a matrix is useful for solving linear equations, finding solutions to systems of equations,
and performing linear transformations. It allows us to undo operations of a matrix and recover the original
quantities. However, not all matrices have an inverse. A matrix only has an inverse if it is non-singular,
which means its determinant is non-zero.

Definition 8.5. Given a square matrix A, its inverse A−1 is a matrix such that the product of A by A−1 is
equal to the identity matrix I. Mathematically, it satisfies the following equation:

A · A−1 = A−1 · A = I (8.14)

Here, I represents the identity matrix of the same size as A.


For a matrix to have an inverse, it must be non-singular, which means its determinant must be non-zero
(det(A) 6= 0). If a matrix is not non-singular, it is said to be singular and does not have an inverse.

Example 8.11. Let’s consider the following matrix A:


 
21
A= (8.15)
13

To find the inverse of A, we can use the formula:


8.8 System Linear Equations 125

1
A−1 = · adj(A) (8.16)
det(A)

Here, det(A) represents the determinant of A, and adj(A) represents the adjugate matrix of A.
Calculating the determinant of A, we have:

21
det(A) = = (2 · 3) − (1 · 1) = 5 (8.17)
13

Next, we calculate the adjugate matrix of A:


 
3 −1
adj(A) = (8.18)
−1 2
Note.

Finally, using the inverse formula, we obtain:


  3
−1
 
1 3 −1
A−1 = · = 51 25 (8.19)
5 −1 2 −5 5

The matrix A−1 is the inverse of matrix A. If we multiply A by A−1 , we should obtain the identity matrix:
   3
−1
  
21 10
A · A−1 = · 51 25 = =I (8.20)
13 −5 5 01

This result confirms that we have found the correct inverse of matrix A.

8.8. System Linear Equations


A system of linear equations is a set of equations that are linearly related to each other. Each linear equation
represents a constraint or condition that the variables of the system must satisfy. The goal is to find the values
of the variables that satisfy all the equations simultaneously.
For example, consider the following system of linear equations:

2x + 3y = 8
(8.21)
4x − 2y = 2

Here, we have two linear equations with two variables, x and y. Each equation represents a line in a plane,
and the solution of the system corresponds to the values of x and y where these lines intersect. In this case,
we are looking for the values of x and y that satisfy both equations simultaneously. These values will form
the point of intersection of the two lines, which is the solution of the system.
126 8 Algebra of Matrices

Definition 8.6. A system of linear equations is defined as a set of equations of the form:

a11 x1 + a12 x2 + . . . + a1n xn = b1


a21 x1 + a22 x2 + . . . + a2n xn = b2
(8.22)
...
am1 x1 + am2 x2 + . . . + amn xn = bm

Where x1 , x2 , ..., xn are the variables of the system, ai j are the coefficients of the variables, and bi are the
constant terms of each equation.

The objective is to find the values of the variables x1 , x2 , ..., xn that satisfy all the equations simultaneously,
i.e., make all the equations true at the same time. These values form the solution of the system of linear
equations. Depending on the number of possible solutions, a system can be consistent (has a unique solution
or infinitely many solutions) or inconsistent (has no solution).

8.9. Cramer’s Rule


Cramer’s Rule is a method used in linear algebra to solve systems of linear equations. It allows us to find
unique solutions for each unknown variable in the system, provided that the system is a compatible system
(meaning it has a unique solution).
Formally, Cramer’s Rule relies on determinants to find the values of the variables in a system of linear
equations. This rule utilizes the properties of determinants and provides a systematic way to solve the system
by assigning a variable to each determinant and dividing by the main determinant of the system.

Definition 8.7. Given a system of linear equations in the form:

a11 x1 + a12 x2 + . . . + a1n xn = b1


a21 x1 + a22 x2 + . . . + a2n xn = b2
.. (8.23)
.
an1 x1 + an2 x2 + . . . + ann xn = bn

If the main determinant D of the system is nonzero (D 6= 0), then the system is a compatible system and has
a unique solution. The solution for each unknown variable xi be obtained using the formula:
Di
xi = (8.24)
D
Where Di the determinant obtained by replacing the column of coefficients for variable xi with the column
of constant terms b.

Example 8.12. Let’s consider the following system of linear equations:


8.10 Eigenvectors 127

3x − 2y = 5
(8.25)
2x + 4y = 10

To apply Cramer’s Rule, we need to calculate the main determinant D, as well as Dx Dy :

3 −2
D= = (3 · 4) − (−2 · 2) = 16 − (−4) = 20 (8.26)
2 4

5 −2
Dx = = (5 · 4) − (−2 · 10) = 20 − (−20) = 40 (8.27)
10 4
3 5
Dy = = (3 · 10) − (5 · 2) = 30 − 10 = 20 (8.28)
2 10

Now, we can find the solutions for x and y using Cramer’s formulas:
Dx 40
x= = =2 (8.29)
D 20
Dy 20
y= = =1 (8.30)
D 20
Therefore, the solution for this system of equations is x = 2 and y = 1.

8.10. Eigenvectors
Eigenvectors are special vectors that do not change their direction when subjected to a linear transformation,
except for a possible scaling. In other words, when multiplied by a matrix, the result is simply a scaled
version of the original eigenvector.

Definition 8.8. Given a linear operator represented by a square matrix A, an eigenvector is a non-zero vector
v such that when multiplied by matrix A, the result is a scalar multiple of the original vector. Formally, if v
is an eigenvector of A, then it satisfies the following equation:

Av = λ v (8.31)

Where λ is the eigenvalue associated with eigenvector v.

Example 8.13. Let’s consider the following matrix A:


 
32
A= (8.32)
14

We want to find the eigenvectors and eigenvalues associated with this matrix.
Step 1: Find the eigenvalues: To find the eigenvalues, we need to solve the characteristic equation:
128 8 Algebra of Matrices

det(A − λ I) = 0 (8.33)

Where I is the identity matrix of the same size as A.


Substituting the values, we have:
   
32 10
det −λ =0 (8.34)
14 01

Simplifying, we have:
 
3−λ 2
det =0 (8.35)
1 4−λ

Expanding the determinant, we get:

(3 − λ )(4 − λ ) − 2 · 1 = 0 (8.36)

Solving the quadratic equation, we find the eigenvalues:

λ1 = 1, λ2 = 6 (8.37)

Step 2: Find the eigenvectors: For each eigenvalue, we need to find the corresponding eigenvectors by solving
the equation Av = λ v.
For λ1 = 1:
Substituting in the equation, we have:
    
32 x x
=1 (8.38)
14 y y

This can be rewritten as a system of equations:

3x + 2y = x
x + 4y = y

Solving the system, we find that x = −2y.


Therefore, the eigenvector corresponding to λ1 = 1 is of the form:
 
−2
v1 = (8.39)
1

For λ2 = 6:
Substituting in the equation, we have:
8.10 Eigenvectors 129

    
32 x x
=6 (8.40)
14 y y

This can be rewritten as a system of equations:

3x + 2y = 6x
x + 4y = 6y

Solving the system, we find that x = 2y.


Therefore, the eigenvector corresponding to λ2 = 6 is of the form:
 
2
v2 = (8.41)
1

The eigenvectors corresponding to the eigenvalues λ1 = 1 and λ2 = 6 are respectively:


   
−2 2
v1 = , v2 = (8.42)
1 1

Example 8.14. Suppose you have a dynamic system in which a particle moves in three-dimensional space.
The particle’s equations of motion are given by the following matrix:
 
4 1 −2
A =  2 5 −2 (8.43)
−1 −1 4

You want to determine the characteristic directions and rates of growth or decay of the system. To do this,
you need to find the eigenvalues and eigenvectors of matrix A.

Eigenvalues and Eigenvectors of Matrix A

noindent Given:  
4 1 −2
A =  2 5 −2
−1 −1 4

Eigenvalues

To determine the eigenvalues, we need to compute the determinant of A − λ I and set it equal to zero:
 
4−λ 1 −2
det(A − λ I) = det  2 5 − λ −2 
−1 −1 4 − λ
130 8 Algebra of Matrices

Expanding the determinant and solving the resulting cubic equation will give the eigenvalues λ1 , λ2 , and λ3 .

Eigenvectors

For each eigenvalue λi , the corresponding eigenvector vi is found by solving the system:

(A − λi I)vi = 0

1. For λ1 = 3: Insert
√ the system of equations and the solution for v1 is v1 = (1 − 5, −2, 1).
2. For λ2 = 5 + √5: Insert the system of equations and the solution for v2 is v2 = (−1,√
1, 0).
3. For λ3 = 5 − 5: Insert the system of equations and the solution for v3 is v3 = (1 + 5, −2, 1).

8.11. Diagonalization of a Matrix


Diagonalization is a process that allows us to decompose a matrix into a special form, where the resulting
matrix is diagonal, and the corresponding eigenvectors are used as the columns of the transformation matrix.
This process significantly simplifies calculations and provides valuable information about the original
matrix.
Firstly, we need to find the eigenvectors and eigenvalues of the matrix. Eigenvalues are the characteristic
values of the matrix, and eigenvectors are the vectors associated with these values.
Let’s consider a square matrix A of size n × n. If there exist n linearly independent eigenvectors v1 , v2 , ..., vn
and n eigenvalues λ1 , λ2 , ..., λn such that Avi = λi vi for i = 1, 2, ..., n, then we can construct the transformation
matrix P using the eigenvectors as its columns:

P = v1 v2 . . . vn (8.44)

The diagonal matrix D is formed by placing the corresponding eigenvalues on the diagonal and filling the
rest of the entries with zeros:
 
λ1 0 . . . 0
 0 λ2 . . . 0 
D= . . . .  (8.45)
 
. .
. . . . . .
0 0 . . . λn

Then, we can express the diagonalization of matrix A as:

A = PDP−1 (8.46)

Where P−1 is the inverse of the transformation matrix P.


Now, let’s go through an example step-by-step to illustrate how to diagonalize a matrix using eigenvectors:
Suppose we have the following matrix:
8.11 Diagonalization of a Matrix 131

 
21
A= (8.47)
43

Step 1: Find the eigenvalues. To find the eigenvalues, we solve the characteristic equation:

det(A − λ I) = 0 (8.48)

Where I is the identity matrix of the same size as A.

     
21 10 2−λ 1
det −λ = = (2 − λ )(3 − λ ) − 4 = λ 2 − 5λ + 2 = 0 (8.49)
43 01 4 3−λ

Solving this quadratic equation, we find the eigenvalues:


√ √
5 + 17 5 − 17
λ1 = , λ2 = (8.50)
2 2
Step 2: Find the corresponding eigenvectors. For each eigenvalue, we find the eigenvectors by solving the
equation (A − λ I)v = 0, where v is the eigenvector.

5 + 17
For the eigenvalue λ1 = :
2
√ ! √ !
2 − 5+2 17 1√ − 1+2 17 1√
(A − λ1 I) = = (8.51)
4 3 − 5+2 17 4 − 1−2 17

Solving the equation (A − λ1 I)v1 = 0, we obtain:


√ !   
− 1+2 17 1√ x 0
= (8.52)
4 − 1−2 17 y 0

This leads us to the system of equations:



1 + 17
− x+y = 0
2 √ (8.53)
1 − 17
4x − y=0
2
Solving this system of equations, we find the eigenvector v1 .

5 − 17
We repeat the same procedure for the eigenvalue λ2 = to obtain the eigenvector v2 .
2
Step 3: Construct the transformation matrix and the diagonal matrix. The transformation matrix P is
constructed using the eigenvectors as its columns:

P = v1 v2 (8.54)
132 8 Algebra of Matrices

The diagonal matrix D is constructed using the eigenvalues as its diagonal entries:
 
λ1 0
D= (8.55)
0 λ2

Step 4: Verify the diagonalization. To verify the diagonalization, we calculate PDP−1 and check if it equals
the original matrix A.
 
−1
 λ1 0 −1
PDP = v1 v2 v1 v2 =A (8.56)
0 λ2

If the equality holds, then we have successfully diagonalized the matrix A.


In this example, the final expressions of P, D, and A are obtained by substituting the values found in the
previous steps.

8.12. Conclusions
In this chapter, we have examined a variety of topics that serve as the basis for linear algebra. Cramer’s
rule is an effective method for solving systems of linear equations utilizing determinants. The significance
of determinants in determining the invertibility and properties of square matrices was then explored by
examining the concept of determinants. The diagonalization of a matrix is a technique that allows us to
simplify complex matrices by determining their eigenvalues and eigenvectors. In addition, the inverse of a
matrix and its role in solving equations and transforming vectors were investigated. Throughout our travels,
we have studied matrices and their properties, acquiring a deeper comprehension of their manipulation via
various matrix operations. By the end of this chapter, we will have a solid foundation in linear algebra and
be prepared to tackle more advanced concepts and their diversified practical applications.
8.13 Exercises 133

8.13. Exercises
Exercise 8.1. Consider the following system of linear equations:

2x + 3y = 10 (8.57)
4x − 2y = 2 (8.58)

Using Cramer’s Rule, find the values of x and y.

Exercise 8.2. Calculate the determinant of the following 3 × 3 matrix:

4 2 3
0 −1 5 (8.59)
2 3 −2

Exercise 8.3. Diagonalize the following 3 × 3 matrix if possible:


 
210
0 3 4 (8.60)
005

Exercise 8.4. Find the eigenvectors of the following matrix:


 
31
(8.61)
12

Exercise 8.5. Calculate the inverse of the following matrix:


 
21
(8.62)
34

Exercise 8.6. Perform the following matrix operation:


   
21 −1 2
+ (8.63)
34 0 1

Exercise 8.7. Given the system of equations, resolve using the Cramer’s Rule:

2x + 3y − z = 7
x − 2y + 2z = −4 (8.64)
3x + 2y − 5z = 12

Exercise 8.8. Calculate the determinant of the following 3 × 3 matrix:


 
abc
A = d e f  (8.65)
gh i
Chapter 9
Linear Algebra

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract The concentration of this chapter is on determinants, calculation methods, cofactor expansion, the
Sarrus rule, Gaussian elimination, homogeneous systems, eigenvalues and eigenvectors, a matrix’s rank, and
linear independence. We investigate the fundamental properties and applications of determinants, as well
as numerous calculation techniques, including cofactor expansion and the Sarrus rule. Gaussian elimination
is introduced as a potent technique for solving linear equation systems. Examining homogeneous systems
with an emphasis on their distinctive characteristics and solutions. The chapter concludes with a discussion
of eigenvalues and eigenvectors, revealing their importance in comprehending linear transformations and
matrix diagonalizations. In addition, we discuss the concept of rank, which assesses the dimensionality
of a matrix, and the significance of linear independence in a variety of mathematical contexts. This chapter
provides a thorough introduction to these essential topics in linear algebra, laying the groundwork for further
study in the field.

Keywords: calculation methods, cofactor expansion, determinants, eigenvalues, eigenvectors, gaussian


elimination, homogeneous systems, linear independence, rank of a Matrix, Sarrus rule.

9.1. Introduction
This chapter examines the fundamental concepts of linear algebra, with a particular emphasis on determinants,
Gaussian elimination, and eigenvalues. It offers a comprehensive examination of these topics and their
applications in a variety of domains.

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

135
136 9 Linear Algebra

Beginning with a discussion of determinants and their significance in linear algebra, the chapter continues
with a discussion of the importance of determinants in calculus. It discusses the methods for calculating
determinants, such as cofactor expansion and the Sarrus rule. Explores the properties and geometric
interpretations of determinants to provide a comprehensive comprehension of their function in linear
transformations.
Next, the chapter examines Gaussian elimination, a potent technique for solving linear equation systems. The
step-by-step procedure of Gaussian elimination is described, emphasizing its utility in locating solutions,
determining rank, and determining linear independence. Also discussed in relation to Gaussian elimination
are homogeneous systems, a special type of linear system.
In addition, the chapter investigates eigenvalues and eigenvectors, which are fundamental concepts in linear
algebra. There is a discussion of the definition, computation, and properties of eigenvalues and eigenvectors,
as well as their applications in diagonalization, system stability analysis, and data analysis.
The chapter concludes with a discussion of the rank and linear independence of matrices. It describes how
the rank of a matrix is determined, its relationship with linear independence, and its significance in various
applications, including the solution of systems of equations, the characterization of linear transformations,
and optimization problems.

9.2. Determinants
Definition 9.1. The determinant of a square matrix A is represented as det(A) or |A| and is calculated using a
specific formula depending on the size of the matrix. For a matrix of size n × n, the determinant is computed
as a linear combination of the elements of the matrix using a recursive formula.

(i) The determinant of a square matrix remains unchanged when elementary row or column operations are
performed on the matrix.
Example 9.1. Let’s consider matrix A and perform the elementary operation of multiplying the first row
by 2:    
12 24
A= → 2A = (9.1)
34 34
By calculating the determinants of both matrices, we obtain det(A) = −2 and det(2A) = −4, which
demonstrates that the determinant changed when multiplying a row by a scalar.

(ii) If a matrix has a row or column of zeros, its determinant is equal to zero.

Example 9.2. Let’s consider matrix B with a row of zeros:


 
12
B= (9.2)
00

The determinant of B is det(B) = 0, as expected due to the presence of a row of zeros.


9.3 Calculation Methods 137

(iii) If a matrix has two equal rows or columns, its determinant is equal to zero.

Example 9.3. Let’s consider matrix C with two equal rows:


 
12
C= (9.3)
12

The determinant of C is det(C) = 0, as the rows are equal.

(iv) If the rows or columns of a matrix are interchanged, the sign of the determinant changes (det(−A) =
− det(A)).

Example 9.4. Let’s consider matrix D and its negation -D:


   
12 −1 −2
D= and − D = (9.4)
34 −3 −4

The determinant of D is det(D) = −2, while the determinant of -D is det(−D) = 2, demonstrating the
change in sign.

(v) If all the elements of a row or column of a matrix are multiplied by a scalar k, the determinant is
multiplied by k.

Example 9.5. Let’s consider matrix E and multiply the second row by 2:
   
12 ′ 12
E= → E = (9.5)
34 68

The determinant of E is det(E) = −2, while the determinant of E’ is det(E ′ ) = −4, which shows that
the determinant is multiplied by the scalar 2 when multiplying a row by 2.

9.3. Calculation Methods


There are various methods to calculate the determinant of a square matrix, such as cofactor expansion, the
Sarrus rule for 3 × 3 matrices, Gaussian elimination, and others. Each method is applicable to matrices of
different sizes and has its own calculation formula.

9.3.1. Cofactor Expansion


The cofactor expansion is a method for calculating the determinant of a square matrix. It involves decomposing
the matrix into smaller submatrices and using the cofactors associated with each element to determine the
determinant’s value. It is akin to untangling a ball of yarn, where the problem is broken down into simpler
parts and solved recursively.
138 9 Linear Algebra

The cofactor expansion is a technique for computing the determinant of an n×n square matrix. One chooses a
row or column of the matrix and computes the cofactors associated with each element of that row or column.
The cofactor of an element ai j is determined as the determinant of the submatrix obtained by removing
row i and column j, multiplied by (−1)i+ j . Subsequently, a summation is performed of the products of the
elements of the chosen row or column and their respective cofactors. This process is reiterated until reaching
2 × 2 matrices, where the formula for the determinant at that dimension is applied.

Example 9.6. Let’s consider matrix F of size 3 × 3:


 
243
F = 1 5 2 (9.6)
013

Using cofactor expansion, we select the first row. The cofactors associated with each element are:

52
C11 = = 15 − 2 = 13
13
12
C12 = = 3−0 = 3
03
15
C13 = = 1−0 = 1
01

Therefore, the determinant of F is:

det(F) = 2 · 13 − 4 · 3 + 3 · 1 = 26 − 12 + 3 = 17 (9.7)

9.3.2. Sarrus Rule


The Sarrus Rule is a method for calculating the determinant of a 3x3 square matrix. It is based on using
diagonals within the matrix to obtain products and sums that result in the determinant value.
The Sarrus Rule is a specific method for calculating the determinant of a 3x3 matrix. Given a matrix A in
the form:
 
abc
A = d e f  (9.8)
gh i

The determinant of A is calculated using the formula:

det(A) = (a · e · i) + (b · f · g) + (c · d · h) − (g · e · c) − (h · f · a) − (i · d · b) (9.9)
9.3 Calculation Methods 139

This formula involves multiplying three main diagonal terms of the matrix and three secondary diagonal
terms in the opposite direction, and then adding and subtracting the obtained results to obtain the final
determinant value.

Example 9.7. Let’s consider matrix G of size 3 × 3:


 
123
G = 4 5 6 (9.10)
789

Using the Sarrus rule, we can calculate the determinant of G:

det(G) = 1 · 5 · 9 + 2 · 6 · 7 + 3 · 4 · 8 − 7 · 5 · 3 − 8 · 6 · 1 − 9 · 4 · 2 = 0 (9.11)

9.3.3. Gaussian Elimination


Gaussian elimination is a method used to solve systems of linear equations and perform operations on
matrices. It involves applying a series of steps to reduce a matrix to its row echelon form or reduced row
echelon form, which simplifies solving the system or provides information about the matrix.
Gaussian elimination, also known as Gauss-Jordan elimination, is an algorithm used to transform a matrix
into its row echelon form or reduced row echelon form through a series of elimination steps. These steps
involve performing elementary row operations, such as swapping rows, multiplying a row by a scalar,
and adding or subtracting multiples of one row from another row. The goal of Gaussian elimination is
to triangularize the matrix, meaning obtaining a matrix where all the elements below the main diagonal are
zero. Once the matrix is in row echelon form, additional elimination steps can be used to reduce it to reduced
row echelon form, where in addition to being upper triangular, all the elements on the main diagonal are one
and all other entries are zero. The final result is a matrix that provides information about the solutions of a
system of linear equations or important characteristics of the matrix, such as the rank and determinant.

Example 9.8. Consider the following system of linear equations:

2x + 3y − z = 4 3x − 2y + 2z = −1 x + y − z = 1 (9.12)

Solution 9.1. Given the system:

2x + 3y − z = 4 (1) (9.13)
3x − 2y + 2z = −1 (2) (9.14)
x + y − z = 1 (3) (9.15)
(9.16)

We represent it as an augmented matrix:


140 9 Linear Algebra
 
2 3 −1 4
 3 −2 2 −1 
1 1 −1 1

Step 1: Make the first element of the third row 0.


 
2 3 −1 4
R′3 = 2R3 − R1 =⇒  3 −2 2 −1 
0 −1 −1 −2

Step 2: Make the first element of the second row 0.


 
2 3 −1 4
3
R′2 = R2 − R1 =⇒  0 −6.5 3.5 −7 
2
0 −1 −1 −2

Step 3: Eliminate the second element in the third row.


 
2 3 −1 4
1
R′3 = R3 + R2 =⇒  0 −6.5 3.5 −7 
2
0 0 −0.25 −0.5

From this matrix:


For the third row:
3
z=2=
5
For the second row:
3
−6.5y + 3.5 × = −7
5
Solving for y:
14 7
y= =
5 5
Using the first equation with the values of y and z:
7 3
2x + 3 × − =4
5 5
Which gives:
1
x=
5

The solutions are x = 15 , y = 57 , and z = 35 .


9.4 Homogeneous Systems 141

9.4. Homogeneous Systems


A homogeneous system of linear equations is a set of equations in which all variables are linearly related,
and the sum of the constants in each equation is equal to zero.

Definition 9.2. Let’s consider a system of linear equations with m equations and n unknowns represented by
the augmented matrix:
  
a11 x1 +a12 x2 + . . . +a1n xn 0
 a21 x1 +a22 x2 + . . . +a2n xn  0
  
 .. .. .. ..   .. 
 . . . . .
am1 x1 +am2 x2 + . . . +amn xn 0

The system is said to be homogeneous if all the constants in the augmented matrix are equal to zero.

Example 9.9. Consider the following homogeneous system of linear equations with 3 equations and 3
unknowns:

2x + 3y − z = 0
4x − y + 2z = 0
x − 2y + 5z = 0

This system can be represented in matrix form as:


    
2 3 −1 x 0
4 −1 2  y = 0
1 −2 5 z 0

Properties of a homogeneous system:

(i) If the homogeneous system has a solution, then the trivial solution x1 = x2 = . . . = xn = 0 is a solution
of the system.

Example 9.10. Consider the following homogeneous system:

2x − 3y + z = 0
4x + 2y − 2z = 0 (9.17)
−2x + y + 3z = 0

To demonstrate that the trivial solution is a solution of the system, we substitute x = y = z = 0 into each
equation:
For the first equation:
142 9 Linear Algebra

2(0) − 3(0) + (0) = 0 (9.18)


This equation simplifies to 0 = 0, which is true.

For the second equation:


4(0) + 2(0) − 2(0) = 0 (9.19)
This equation simplifies to 0 = 0, which is true.

For the third equation:


−2(0) + (0) + 3(0) = 0 (9.20)
This equation simplifies to 0 = 0, which is true.

Therefore, the trivial solution x = y = z = 0 is a solution of the system, satisfying the first property.
(ii) If the homogeneous system has a solution, then any linear combination of the solutions is also a solution.

Example 9.11. Continuing with the same system as above, suppose we have found two non-trivial
solutions: x1 = 1, y1 = 2, z1 = −1 and x2 = −2, y2 = −1, z2 = 1.
We can verify that any linear combination of these solutions is also a solution. Let’s take the linear
combination x = 3x1 − 2x2 , y = 3y1 − 2y2 , z = 3z1 − 2z2 :
For the first equation:
2(3x1 − 2x2 ) − 3(3y1 − 2y2 ) + (3z1 − 2z2 ) = 0
6x1 − 4x2 − 9y1 + 6y2 + 3z1 − 2z2 = 0
(9.21)
6 − 4 − 9(2) + 6(−1) + 3(−1) − 2(1) = 0
0=0

Similarly, we can verify that the linear combination satisfies the second and third equations. Thus, any
linear combination of the solutions x1 = 1, y1 = 2, z1 = −1 and x2 = −2, y2 = −1, z2 = 1 is also a solution
of the system, demonstrating the second property.
(iii) If the homogeneous system has more variables than equations (i.e., n > m), then it always has non-trivial
solutions (where at least one of the variables is not equal to zero).

Example 9.12. Consider the following homogeneous system with three variables and two equations:

2x − 3y + z = 0
(9.22)
4x + 2y − 2z = 0

To determine if there are non-trivial solutions, we can use the concept of matrix rank. By putting the
system into an augmented matrix and row-reducing it, we can find the rank of the coefficient matrix.
 
2 −3 1 0
(9.23)
4 2 −2 0
9.5 Eigenvalues and Eigenvectors 143

Row-reducing the augmented matrix leads to:

1 0 − 58 0
 
(9.24)
0 1 56 0

From this, we can see that the rank of the coefficient matrix is 2. Since n > m, the system has more
variables than equations, and we have at least one free variable. This means that the system has non-
trivial solutions.
Therefore, the third property holds in this case, and we can find non-trivial solutions for the system.

9.5. Eigenvalues and Eigenvectors


Definition 9.3. Given a square matrix A, an eigenvector v and its corresponding eigenvalue λ satisfy the
equation:
Av = λ v (9.25)
Where v is a non-zero vector and λ is a scalar. In other words, when we multiply the matrix A by its
eigenvector v, the result is a scaled version of the same vector v.

(i) Eigenvalues and eigenvectors always occur in pairs. If v is an eigenvector of matrix A with eigenvalue
λ , then any scalar multiple of v is also an eigenvector of A with the same eigenvalue λ .

(ii) The determinant of a square matrix A is equal to the product of its eigenvalues.

(iii) The trace (sum of diagonal elements) of a square matrix A is equal to the sum of its eigenvalues.

Example 9.13. Consider the following matrix:


 
31
A= (9.26)
13

We want to find the eigenvalues and eigenvectors of this matrix. To do that, we solve the equation:
Given the matrix A:  
31
A=
13

We compute the determinant of A − λ I:

3−λ 1
det(A − λ I) = = (3 − λ )(3 − λ ) − 1 = λ 2 − 6λ + 8
1 3−λ

Setting the determinant to zero:


λ 2 − 6λ + 8 = 0
144 9 Linear Algebra

Factoring:
(λ − 4)(λ − 2) = 0

This gives us two eigenvalues:


λ1 = 4, λ2 = 2

To determine the eigenvectors:


For λ1 = 4: Substitute into A − λ I:
 
−1 1
1 −1
Which results in the eigenvector:
 
1
v1 =
1
For λ2 = 2: Substitute into A − λ I:
 
11
11
Which results in the eigenvector:

1
v2 =
−1
   
1 1
Thus, the eigenvalues are λ1 = 4 and λ2 = 2 with corresponding eigenvectors v1 = and v2 =
1 −1
respectively.

Example 9.14. Consider the following matrix:


 
23
A= (9.27)
14

Following the same process as in the previous example, we find the eigenvalues and eigenvectors of matrix
A.
Solving the characteristic equation (A − λ I) = 0, we obtain two eigenvalues: λ1 = 1 and λ2 = 5. Finding the
corresponding eigenvectors by solving the homogeneous systems of equations: For λ1 = 1:

(A − λ1 I)v1 = 0 (9.28)
   
13 x 0
(9.29)
13 y 0

The solution of the system gives us the eigenvector corresponding to λ1 = 1:


9.5 Eigenvalues and Eigenvectors 145
 
−3
v1 = (9.30)
1

For λ2 = 5:
(A − λ2 I)v2 = 0 (9.31)
   
−3 3 x 0
(9.32)
1 −1 y 0

The solution of the system gives us the eigenvector corresponding to λ2 = 5:


 
1
v2 = (9.33)
1

 
−3
Thus, the eigenvalues of matrix A are λ1 = 1 and λ2 = 5, with their corresponding eigenvectors v1 =
  1
1
and v2 = , respectively.
1

The following are some of the key uses for eigenvalues and eigenvectors:

(i) Diagonalization: Eigenvectors and eigenvalues play a crucial role in diagonalizing matrices. A matrix
A is diagonalizable if and only if it has a set of linearly independent eigenvectors. Diagonalization can
simplify computations involving matrix powers and exponential functions.

(ii) Systems of Differential Equations: In the study of differential equations, eigenvectors and eigenvalues are
used to find the general solutions of linear systems. They provide insights into the behavior of dynamic
systems and help in understanding stability and equilibrium points.

(iii) Principal Component Analysis (PCA): PCA is a dimensionality reduction technique used in data
analysis. It involves finding the eigenvectors and eigenvalues of the covariance matrix of a dataset.
The eigenvectors represent the principal components that capture the most significant variations in the
data, while the eigenvalues indicate the amount of variance explained by each component.

(iv) Markov Chains: Eigenvectors and eigenvalues are employed in the analysis of Markov chains, which
model the stochastic behavior of systems evolving over time. The dominant eigenvalue and its corresponding
eigenvector provide insights into long-term behavior and steady-state probabilities.

Example 9.15. Consider the following system of linear equations:

2x + 3y = 8
(9.34)
4x − 2y = 10
146 9 Linear Algebra

(i) Write the augmented matrix: We write the system of equations as an augmented matrix:
 
2 3 8
(9.35)
4 −2 10

(ii) Row-reduce the matrix: We apply Gaussian elimination to row-reduce the matrix:
   3   3 
2 3 8 1 2 4 1 2 4
→ → (9.36)
4 −2 10 0 −8 2 0 1 − 41

(iii) Solve the system: From the row-reduced matrix, we can obtain the solution of the system:
3
x = 4− y
2 (9.37)
1
y=−
4

Therefore, the system has a unique solution determined by the values of x and y.

Example 9.16. Consider the following system of linear equations:

2x + 3y = 8
(9.38)
4x + 6y = 16

(i) Write the augmented matrix: We write the system of equations as an augmented matrix:
 
23 8
(9.39)
4 6 16

(ii) Row-reduce the matrix: We apply Gaussian elimination to row-reduce the matrix:
   
23 8 238
→ (9.40)
4 6 16 000

(iii) Solve the system: From the row-reduced matrix, we can observe that the second row is a zero row.
This implies that the corresponding equation is redundant and does not add new information to the system.
Therefore, the system has infinitely many solutions.

We can express the solution in terms of a free variable:

x=t
8 2 (9.41)
y= − t
3 3

Where t is a free variable.


9.6 Rank of a Matrix 147

9.6. Rank of a Matrix


The rank of a matrix can be understood as the maximum number of columns (or rows) that are linearly
independent. In other words, it represents the maximum number of columns (or rows) that contain unique
and non-redundant information within the matrix.

Definition 9.4. The rank of a matrix A, denoted by rank(A), is the dimension of the column space of A. It is
also equal to the dimension of the row space of A. It represents the maximum number of linearly independent
columns (or rows) in the matrix.

Determining the rank of a matrix is of great importance and utility in various fields of mathematics and
sciences. Here are some key reasons for determining the rank of a matrix:

(i) Characterization of systems of linear equations: The rank of a matrix provides fundamental information
about the associated system of linear equations. It helps determine if the system has unique solutions,
infinite solutions, or no solutions. This is essential for solving practical application problems in fields
like physics, engineering, economics, and statistics.

(i) Identification of linear dependence: The rank of a matrix is directly related to the linear dependence
of its columns or rows. It helps determine if the columns (or rows) are linearly independent or linearly
dependent. This is useful for understanding data structure and eliminating redundancy in datasets.

(i) Diagonalization of matrices: Diagonalization of a matrix is an important procedure in linear algebra.


The rank of a matrix is closely related to its diagonalizability. Matrices with full rank are easier to
diagonalize and simplify many calculations and applications.

(i) Analysis of linear transformations: The rank of a matrix is relevant to the properties and characteristics
of associated linear transformations. It helps understand the dimension of input and output spaces, as
well as the existence of invertible or constrained transformations.

(i) Optimization and linear programming: The rank of a matrix is relevant in optimization and linear
programming problems, where one seeks to maximize or minimize a linear function subject to constraints.
It helps determine if the constraints are linearly independent and if optimal solutions exist.

(i) Analysis of dynamical systems: The rank of a matrix plays an important role in the analysis of dynamical
systems and stability. It helps determine the stability of equilibrium points and study the system’s
evolution over time.

Example 9.17. Consider the following matrix:


 
12 3
A = 4 5 6 (9.42)
78 9

We want to find the rank of this matrix and determine if the columns of the matrix are linearly independent.
148 9 Linear Algebra

Solution 9.2. Consider the matrix A:  


12 3
A = 4 5 6
78 9

Using Gaussian elimination:

R2 ← R2 − 4R1
R3 ← R3 − 7R1
 
1 2 3
⇒ 0 −3 −6 
0 −6 −12

1
R2 ← − R2
3
R3 ← R3 + 6R2
 
123
⇒ 0 1 2
000

The matrix has two non-zero rows, implying its rank is 2.


The rank of matrix A is 2. Since the rank is less than the number of columns (3), the columns of A are linearly
dependent.

9.7. Linear Independence


Linear independence refers to a set of vectors in which none of the vectors can be expressed as a linear
combination of the other vectors. In simpler terms, it means that none of the vectors in the set can be formed
by adding or scaling the other vectors in the set.
If the vectors are linearly dependent, it means that at least one of them can be expressed as a linear
combination of the others. This can lead to several implications and challenges in different contexts. Here
are some reasons why linear dependence can be problematic:
(i) Redundancy of information: When the vectors are linearly dependent, some of them contain duplicated
or redundant information. This can make data analysis and extraction of relevant information more
difficult. In applications such as data analysis and machine learning, redundancy can negatively affect
model performance and interpretation.

(ii) Inversion problems: In some cases, if a matrix is composed of linearly dependent vectors, the matrix can
become singular, meaning its inverse does not exist. This makes solving systems of equations difficult
9.8 Basis 149

and can limit the possibilities for solution in mathematical and engineering problems.

(iii) Loss of dimensionality: If the vectors are linearly dependent, their set is contained within a subspace of
lower dimension. This implies that the information or search space is reduced. In problems where the
entire vector space needs to be explored or where high dimensionality is required, linear dependence
can limit possibilities and analysis/solution capabilities.

(iv) Numerical instability: Linear dependence can cause numerical stability problems in calculations and
algorithms. Small errors in data or calculations can be magnified due to linear dependence, leading to
inaccurate or unstable results.

Linear dependence can cause problems by redundantly representing information, hindering matrix inversion,
reducing dimensionality, and causing numerical instability. Therefore, linear independence is desirable in
many cases as it ensures maximum non-redundant information, allows for greater flexibility in calculations
and analysis, and ensures more stable and reliable results.

Definition 9.5. Linear independence refers to a set of vectors {v1 , v2 , ..., vn } in which no nontrivial linear
combination of the vectors equals the zero vector, except when all the coefficients in the linear combination
are zero. In mathematical terms, the set of vectors is linearly independent if the equation a1 v1 + a2 v2 + ... +
an vn = 0 has only the trivial solution a1 = a2 = ... = an = 0.

9.8. Basis
Definition 9.6. Given a vector space V , a basis of V is a set of linearly independent vectors that span the
entire space V . In other words, a basis is a set of vectors that satisfies two properties:

(i) Linearly independent: No vector in the basis can be expressed as a linear combination of the other
vectors in the basis. This implies that there are no nonzero coefficients that make the weighted sum of
the basis vectors equal to zero, except when all coefficients are zero.
(ii) Spans the space: Any vector in the space V can be expressed as a linear combination of the basis
vectors. In other words, any vector in the space can be uniquely represented as a weighted sum of the
basis vectors.

The basis provides a way to decompose any vector in a vector space into simpler and fundamental vectors.
Additionally, any other set of vectors that can span the entire space V must have at least the same number of
elements as the basis, but it may not be linearly independent.
It is important to note that a basis is not unique, as there can be different sets of vectors that satisfy the
properties of being linearly independent and spanning the space. However, all bases of a vector space have
the same number of elements, which is known as the dimension of the space.
150 9 Linear Algebra

The notion of a basis is fundamental in linear algebra, as it provides a key structure for the study and
manipulation of vectors and vector spaces.
The relationship between a linearly independent matrix and a basis is given by:

A linearly independent matrix can serve as a basis for a vector space. A basis is a set of vectors that are
linearly independent and span the entire vector space. In other words, a basis provides a minimal set of
vectors that can express any vector in the vector space through linear combinations.
Let’s denote the linearly independent matrix as A with columns a1 , a2 , . . . , an . If the matrix A is linearly
independent, it means that none of its columns can be expressed as a linear combination of the other columns.
Now, if the matrix A is a square matrix (i.e., the number of columns equals the number of rows), and the
columns of A form a linearly independent set, then these columns can form a basis for the vector space
spanned by the columns. This means that any vector in the vector space can be uniquely expressed as a
linear combination of the columns of A.
Example 9.18. Consider the following matrix system:
   
2 1 x 1
(9.43)
3 −1 y 4

The system of equations is represented as:

2x + y = 1 (9.44)
3x − y = 4 (9.45)

The determinant of matrix A is:


det(A) = 2(−1) − 1(3) = −5.

Since the determinant is not zero, A is invertible. The inverse of A is:


   
−1 1 −1 −1 0.2 0.2
A = = .
−5 1 2 −0.2 −0.4

The solution to the system is:


    
−10.2 0.2 1 1.2
X =A B= = .
−0.2 −0.4 4 −1.6

Thus, x = 1.2 and y = −1.6.

9.9. Conclusions
This chapter concludes with an exhaustive examination of fundamental concepts in linear algebra. We
have delved into determinants, studying their importance in various applications and acquiring calculation
9.9 Conclusions 151

methods, including cofactor expansion and the Sarrus rule. Gaussian elimination has been extensively
studied as an effective method for solving systems of linear equations, especially in the context of homogeneous
systems. In addition, the importance of eigenvalues and eigenvectors has been clarified, casting light on their
applications in diagonalization and comprehension of transformations. The concept of rank in matrices has
been investigated, shedding light on the dimension and structure of matrices. Finally, the concept of linear
independence has been discussed, emphasizing its importance in characterizing vector relationships. This
chapter has provided a robust foundation for further investigation of advanced topics in linear algebra and
their diverse practical applications.
152 9 Linear Algebra

9.10. Exercises
Exercise 9.1. Calculate the determinant of the following matrix using cofactor expansion:
 
3 1 2
A =  0 −2 4 (9.46)
−1 3 5

Exercise 9.2. Solve the following system of equations using the Gaussian elimination method:

2x + 4y − z = 6
x − 3y + 2z = 7 (9.47)
3x − 2y + 3z = 4
     
1 4 7
Exercise 9.3. Given the vectors v1 = 2, v2 = 5, and v3 = 8, determine if they are linearly
3 6 9
independent.

Exercise 9.4. The exercise involves calculating the determinant of a 3x3 matrix using the Sarrus rule. The
given matrix is as follows:
 
123
4 5 6 (9.48)
789
Chapter 10
Algebra of Groups

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract In this chapter, we explore fundamental concepts in group theory, focusing on groups, subgroups,
isomorphisms, and homomorphisms, as well as the notion of algebraic closure. We begin by introducing the
concept of a group and examining its fundamental properties. We then delve into the study of subgroups,
exploring their role in capturing the structure and symmetry of larger groups. The concept of isomorphisms is
introduced to establish relationships between groups that preserve their algebraic structures. Homomorphisms,
on the other hand, provide a more general framework for mapping elements between groups while preserving
group operations. Throughout this chapter, we provide definitions, examples, and illustrative explanations to
deepen the reader’s understanding of these fundamental concepts in group theory.

Keywords: arithmetic module, congruences, groups, homomorphisms, integer numbers, isomorphisms,


number theory, prime numbers, subgroups.

10.1. Introduction
”In this chapter, we will explore the foundations of group theory, a central branch of mathematics that studies
the algebraic structure and properties of sets with a binary operation. We will begin by defining groups
and examining their key properties, such as the existence of the identity element and the inverse of each
element. We will then delve into the study of subgroups, which are subsets of a group that retain the structure
and operation of the original group. Additionally, we will address the concept of isomorphisms, which are

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

153
154 10 Algebra of Groups

bijective mappings between groups that preserve the algebraic structure. We will explore how isomorphisms
allow us to establish connections between different groups, revealing similarities in their algebraic behavior.
We will then analyze homomorphisms, which are mappings that preserve the group operation but are not
necessarily bijective. Throughout this chapter, we will provide examples and applications of these concepts,
laying the groundwork for a deeper study of group theory and its various applications in mathematics and
other disciplines.

10.2. Groups
A group [12, 27, 28] is a set of elements along with an operation that combines two elements to produce a
third one. This operation must have certain special properties.

Definition 10.1. A group is an ordered pair (G, ·), where G is a non-empty set and · is a binary operation (i.e.,
an operation that takes two elements and produces another element) that satisfies the following properties:

(i) Closure: For every a, b ∈ G, the result of the operation a · b is also in G.

(i) Associativity: For every a, b, c ∈ G, (a · b) · c = a · (b · c).

(i) Identity element: There exists a special element e ∈ G, called the identity element, such that for every
a ∈ G, a · e = e · a = a.

(i) Inverse: For every a ∈ G, there exists an element a−1 ∈ G, called the inverse of a, such that a · a−1 =
a−1 · a = e.

Example 10.1. A common example of a group is the group of integers under the operation of addition. Let’s
denote this group as (Z, +). Here, Z represents the set of all integers, and + is the addition operation. This
group satisfies all the properties of a group:

(i) Closure: If you add two integers, the result is still an integer.

(ii) Associativity: Addition of integers is associative, meaning that (a + b) + c = a + (b + c) for any integers
a, b, and c.

(iii) Identity element: The number 0 acts as the identity element, as for any integer a, we have a + 0 = 0 + a =
a.

(iv) Inverse: For any integer a, the additive inverse −a exists and satisfies a + (−a) = (−a) + a = 0.
10.4 Isomorphisms 155

10.3. Subgroups
A subgroup is a subset of a larger group that, when operated with the same group operation, also forms a
group by itself. In other words, it is a “group within a group”.

Definition 10.2. Given a group (G, ·), a subset H of G is said to be a subgroup if it satisfies the following
three conditions:
H is closed under the group operation: For any pair of elements x, y in H, their product x · y is also in H.

∀x, y ∈ H : x · y ∈ H (10.1)

H contains the identity element of the group: The identity element of the group G must be in H.

e∈H (10.2)

H contains the inverse of each element: For every element x in H, its inverse is also in H.

∀x ∈ H : x−1 ∈ H (10.3)

Example 10.2. Let’s consider the group of integers (Z, +). An example of a subgroup of Z is the set of
multiples of 3, denoted as H = {..., −6, −3, 0, 3, 6, ...}.
To prove that H is a subgroup of Z, we need to verify the three conditions of the formal definition of a
subgroup:

(i) Closure: If we take any two multiples of 3, their sum will also be a multiple of 3. Therefore, H is closed
under the addition operation.

(i) Identity element: The number 0 is the identity element of addition in Z. Since 0 is a multiple of 3, we
have 0 ∈ H.

(i) Inverses: For each multiple of 3, its negative is also a multiple of 3. Therefore, if x is a multiple of 3, its
inverse −x is also a multiple of 3, implying that −x ∈ H.

Since H satisfies all three conditions, we can conclude that H is a subgroup of (Z, +).

10.4. Isomorphisms
An isomorphism in group theory is a correspondence between two groups that preserves the structure and
algebraic properties. If two groups are isomorphic, it means that they are essentially the same in terms of
their algebraic behavior, even if their elements and operations may be different.
156 10 Algebra of Groups

Definition 10.3. Let G and H be two groups. An isomorphism between G and H is a bijective function
φ : G → H that preserves the group operation, meaning that for every g1 , g2 ∈ G, the following holds:

φ (g1 · g2 ) = φ (g1 ) · φ (g2 ) (10.4)

Where · denotes the group operation.

φ ψ

H K

Isomorphism

Figure 10.1: The explanation for this diagram is as follows: Group G is related to groups H and K through
the isomorphisms φ and ψ , respectively. The arrow labeled ”Isomorphism” indicates that groups H and K
are isomorphic, meaning they share the same algebraic structure.

Example 10.3. Consider the groups G = (Z, +) and H = (1, −1, ·), where Z is the set of integers and ·
denotes multiplication.
We define the function φ : G → H as:
(
1, if n is even
φ (n) = (10.5)
−1, if n is odd

Let’s verify that φ is an isomorphism:


Consider the groups G = (Z, +) and H = ({1, −1}, ·) where Z is the set of integers and · denotes
multiplication.
We define the function φ : G → H as:
(
1, if n is even
φ (n) =
−1, if n is odd

To check if φ is an isomorphism, it must satisfy:


1. Bijectiveness:
• Injective: φ is not injective since it maps all even integers to 1 and all odd integers to -1. As a
counterexample, consider a = 2 and b = 4. Both are distinct integers, yet φ (2) = φ (4) = 1.
10.5 Homomorphisms 157

• Surjective: φ maps even numbers to 1 and odd numbers to -1. Every element in H has a pre-image
in G. Thus, φ is surjective.

2. Preserve the operation:

For arbitrary integers a, b ∈ Z:


• If a, b are both even:
φ (a + b) = 1 = φ (a) · φ (b)

• If a, b are both odd:


φ (a + b) = 1 = (−1) · (−1) = φ (a) · φ (b)

• If a is even and b is odd or vice versa:

φ (a + b) = −1 = 1 · (−1) = φ (a) · φ (b)

However, since φ is not injective, it cannot be an isomorphism from G to H.

10.5. Homomorphisms
A homomorphism is a function between two groups that preserves the algebraic structure between them.

Definition 10.4. Let (G, ·) and (H, ∗) be two groups. A homomorphism φ : G → H is a function such that
for every pair of elements a, b ∈ G, the following property holds:

φ (a · b) = φ (a) ∗ φ (b) (10.6)

In other words, a homomorphism preserves the group operation. The product of elements in the group G is
mapped to the corresponding product in the group H under the homomorphism φ .
Here’s a concrete example of a homomorphism:

Example 10.4. Consider the homomorphism φ : (Z, +) → (Z6 , +6 ) defined by φ (a) = a mod 6 for a ∈ Z.
To show that φ is a homomorphism, we need to prove that it preserves the group operation. That is, for any
a, b ∈ Z, we need to demonstrate:
158 10 Algebra of Groups

φ
G H

Homomorphism

Figure 10.2: The figure depicts two groups, denoted as G and H. The node on the left represents group
G, while the node on the right represents group H. There is an arrow between them, symbolizing the
homomorphism φ .

φ (a + b) = φ (a) +6 φ (b)

Let’s consider two integers a and b. On one hand, φ (a + b) is equal to (a + b) mod 6. On the other hand,
φ (a) +6 φ (b) is equal to (a mod 6) +6 (b mod 6).
We can simplify both expressions using modulo arithmetic:

φ (a + b) = (a + b) mod 6
= (a mod 6 + b mod 6) mod 6
= (a mod 6 +6 b mod 6) mod 6
= (a mod 6 +6 b mod 6) (mod 6)
= φ (a) +6 φ (b)

Since φ (a + b) = φ (a) +6 φ (b), we have shown that the homomorphism φ preserves the group operation.
Therefore, φ is a homomorphism from (Z, +) to (Z6 , +6 ).

10.6. Generator and Algebraic Closure


In group theory, a generator is an element or set of elements in a group that, through their combination and
operations within the group, can generate all other elements of the group.

Definition 10.5. Given a group G with a binary operation, an element a in G is called a generator if all
elements of G can be expressed as combinations of a using the group operation. That is, for every element
g ∈ G, there exists an integer n such that g = an .

Example 10.5. Let G be a finite additive group of order n generated by the element g ∈ G. An additive group
generator is a subset S of G such that every element in G can be expressed as a linear combination of the
elements in S. Formally, G is generated by S if for every x ∈ G, there exist integer coefficients a1 , a2 , . . . , ak
and elements s1 , s2 , . . . , sk ∈ S such that
x = a1 s1 + a2 s2 + . . . + ak sk .
10.7 Conclusions 159

In other words, any element x in the group G can be obtained by summing a combination of the elements in
S.
To verify that S is a generator of G, we need to demonstrate that every element x ∈ G can be expressed in
this way. Once a generating set S has been found, G is said to be cyclic.
An additive group generator is a set of elements from the group such that, by combining them through
addition and subtraction, all elements of the group can be obtained.

Solution 10.1.
(i) Start by introducing the group G as a finite additive group of order n generated by the element g ∈ G.

(ii) Define an additive group generator as a subset S of G.

(iii) Formally state that G is generated by S if for every x ∈ G, there exist integer coefficients a1 , a2 , . . . , ak
and elements s1 , s2 , . . . , sk ∈ S such that x = a1 s1 + a2 s2 + . . . + ak sk .

(iv) Explain that this means any element x in the group G can be obtained by summing a combination of the
elements in S.

(v) Emphasize the need to verify that S is a generator of G by demonstrating that every element x ∈ G can
be expressed in this way.

Mention that once a generating set S has been found, G is considered to be cyclic.

10.7. Conclusions
This chapter has covered several fundamental group theory concepts. The definition and properties of
groups, which are sets with a binary operation that satisfy certain axioms, were discussed first. Then, we
examined subgroups, which are subsets of a group that themselves constitute a group. Next, we examined
isomorphisms, which are bijective mappings that preserve the algebraic structure between groups. In
contrast, homomorphisms are mappings between groups that preserve the operation of the group. Finally,
we discussed algebraic closure, which is the property of a group to contain all possible solutions to algebraic
equations involving its elements. These concepts are fundamental to the study of group theory and provide
a solid foundation for comprehending algebraic structures and their interrelationships.
160 10 Algebra of Groups

10.8. Exercises
Exercise 10.1. Let G be a group and H be a subgroup of G. Prove that the identity element of G is also the
identity element of H.

Exercise 10.2. Let G and H be groups, and let φ : G → H be an isomorphism. Prove that φ −1 : H → G is
also an isomorphism.

Exercise 10.3. Let G be a group and a an element of G. Prove that hai, the subgroup generated by a, is the
smallest subgroup of G containing a.

Exercise 10.4. Let G and H be groups, and let φ : G → H be a homomorphism. Prove that the kernel of φ ,
denoted by ker(φ ), is a subgroup of G.
Chapter 11
SOLUTIONS

Solutions Chapter 1
Solution 1.1. To solve the equation, we isolate x by subtracting 3 from both sides:

2x + 3 − 3 = 7 − 3
2x = 4
x=2

Therefore, the value of x is 2.

Solution 1.2. The sum of the first n natural numbers can be calculated using the formula for the sum of an
arithmetic series. In this case, we have a series with a1 = 1 and d = 1, where a1 is the first term and d is the
common difference. The sum is calculated as follows:
n
S = (2a1 + (n − 1)d)
2
10
= (2(1) + (10 − 1)(1))
2
= 5(2 + 9)
= 5(11)
= 55

Therefore, the sum of the first 10 natural numbers is 55.

Solution 1.3. To simplify the expression, we need a common denominator, which in this case is 6:
2 5 4 5
− = −
3 6 6 6
−1
=
6
2 5 −1
Therefore, − = .
3 6 6

161
162 11 SOLUTIONS

Solution 1.4. To multiply fractions, we multiply the numerators and denominators:

3 5 3×5
× =
4 8 4×8
15
=
32
3 5 15
Therefore, × = .
4 8 32

Solution 1.5. We square both sides of the equation and solve:



( 5x + 3)2 = 22
5x + 3 = 4
5x = 1
1
x=
5
1
Therefore, the value of x is .
5

Solution 1.6. The sum of the first n negative integers can be calculated using the formula for the sum of an
arithmetic series. In this case, we have a series with a1 = −1 and d = −1, where a1 is the first term and d is
the common difference. The sum is calculated as follows:
n
S = (2a1 + (n − 1)d)
2
20
= (2(−1) + (20 − 1)(−1))
2
= 10(−2 + (−19))
= 10(−21)
= −210

The absolute value of −210 is 210. Therefore, the absolute value of the sum of the first 20 negative integers
is 210.

Solution 1.7. To evaluate the expression, we follow the order of operations:

34−2 = 32
=9

Therefore, 34−2 = 9.

Solution 1.8. The square root of 16 is 4, since 4 × 4 = 16. Therefore, the square root of 16 is 4.
√ √
Solution 1.9. To simplify the expression, we notice that 3 + 2 2 can be expressed as ( 2 + 1)2 . Therefore:
11 SOLUTIONS 163

√ q√
q
3 + 2 2 = ( 2 + 1)2

= 2+1
p √ √
Therefore, the value of 3 + 2 2 is 2 + 1.
√ √
Solution 1.10. Consider the irrational numbers 2 and − 2. Both numbers are irrational since they cannot
be expressed as a fraction. However, their sum is:
√ √ √ √
2 + (− 2) = 2 − 2
=0

0
The number 0 is a rational number since it can be expressed as . Therefore, the sum of two irrational
1
numbers can be a rational number.

Solution 1.11. To simplify the√expression, we multiply the numerator and the denominator by the conjugate
of the denominator, which is 2 + 1:
√ √ √
3 3( 2 + 1)
√ = √ √
2 − 1 ( 2 − 1)( 2 + 1)
√ √
6+ 3
=√ 2
2 − 12
√ √
6+ 3
=
2−1
√ √
= 6+ 3

3 √ √
Therefore, the value of √ is 6 + 3.
2−1

Solution 1.12. To simplify the expression, we use the distributive property and the rules of multiplication
for imaginary numbers:

(2 + i)(3 − i) = 2(3) + 2(−i) + i(3) + i(−i)


= 6 − 2i + 3i − i2
= 6 + i(3 − 2) − i2
= 6+i+1
= 7+i

Therefore, the product of (2 + i)(3 − i) is 7 + i.


√ √
Solution 1.13. To prove that 2 is an irrational number, we use a proof by contradiction. Suppose that 2
a
is a rational number. This would mean that it can be expressed as a fraction , where a and b are integers
b
164 11 SOLUTIONS

a
and b 6= 0. Additionally, let’s assume that is in its simplest form, meaning that a and b have no common
b
prime factors.
√ a
Then, we square both sides of the equation 2 = :
b
a2
2=
b2
2b = a2
2

This implies that a2 is even, which means that a is also even (since the square of an odd number is odd). We
can write a as a = 2c, where c is another integer.
Substituting in the original equation, we get:

2b2 = (2c)2
2b2 = 4c2
b2 = 2c2

This implies that b2 is even and, therefore, b is also even.


a
We have reached the conclusion that both a and b are even, which contradicts our assumption that is in its
√ b
simplest form. Therefore, our initial assumption that 2 is a rational number is incorrect.

Therefore, we can conclude that 2 is an irrational number.
√ √
Solution 1.14. Similar to the previous exercise, we will use a proof by contradiction. Suppose that 2 + 3
a
is a rational number. This means that it can be expressed as a fraction , where a and b are integers and
b
a
b 6= 0. Let’s assume that is in its simplest form.
b
√ √ a
Squaring both sides of the equation ( 2 + 3) = , we get:
b
√ a2
2+2 6+3 = 2
b
√ a2
5+2 6 = 2
b
√ a2
2 6 = 2 −5
b
√ a2 − 5b2
2 6=
b2
√ √
This implies that 2 6 is a rational number. In fact, 6 is an irrational number also.
√ √
Therefore, our initial assumption that 2 + 3 is a rational number is incorrect.
√ √
Hence, we can conclude that 2 + 3 is an irrational number.
11 SOLUTIONS 165

Solution 1.15. To calculate the product of the complex numbers, we use the distributive property and the
rules of multiplication for complex numbers:

(2 + i)(3 − i) = 2(3) + 2(−i) + i(3) + i(−i)


= 6 − 2i + 3i − i2
= 6 + i(3 − 2) − i2
= 6+i+1
= 7+i

Therefore, the product of (2 + i)(3 − i) is 7 + i.

Solution 1.16. Starting with the given inequality:

3x − 5 > 2x + 7

Subtract 2x from both sides:


x−5 > 7
Add 5 to both sides:
x > 12

Thus, the solution to the inequality is x > 12.

Solution 1.17.
−5 < 2x − 1 < 5
−5 + 1 < 2x < 5 + 1
(1.1)
−4 < 2x < 6
−2 < x < 3

In this case, there is no denominator of a fraction involved, so we do not need to consider additional cases.
The solution is: −2 < x < 3
166 11 SOLUTIONS

Solutions Chapter 2
Solution 2.1. To prove that f is injective, let’s assume f (a) = f (b) for some a, b ∈ R.
Starting with:
f (a) = f (b)
we get:
2a + 1 = 2b + 1
Subtracting 1 from both sides:
2a = 2b
Dividing both sides by 2, we get:
a=b

This confirms that if f (a) = f (b), then a = b. Thus, f is injective.

Solution 2.2. The reflection of the function g(x) = x2 with respect to the x-axis is −g(x) = −x2 .

Solution 2.3. For the function composition g( f (x)), we substitute the expression for f (x) into g(x):

g( f (x)) = g(2x) = (2x) + 3 = 2x + 3. (2.2)

Solution 2.4. The limit calculation yields:

x2 − 4 (x − 2)(x + 2)
lim = lim = lim (x + 2) = 4. (2.3)
x→2 x − 2 x→2 x−2 x→2

5
Solution 2.5. The sum of the series ∑ 3n is calculated as follows:
n=1

5
∑ 3n = 3(1) + 3(2) + 3(3) + 3(4) + 3(5) = 45. (2.4)
n=1

Solution 2.6. For the sequence an = 2n , we find the value of a10 as follows:

a10 = 210 = 1024. (2.5)


1
Solution 2.7. The limit of the sequence lim is calculated as:
n→∞ n2

1
lim = 0. (2.6)
n→∞ n2


1
Solution 2.8. The sum of the series ∑ 2n is determined as follows:
n=1
11 SOLUTIONS 167

∞ 1
1
∑ 2n 1 −2 1 = 1.
= (2.7)
n=1 2

Solution 2.9. For the sequence an = 1n , we find the value of a5 as follows:

1
a5 = . (2.8)
5
Solution 2.10. Consider a right triangle with sides of lengths a and b, and a hypotenuse of length c.
To calculate its area using the concept of limits, we can think of the triangle as being composed of n smaller
right triangles, each of decreasing size.

1. Divide side a into n equally spaced intervals, each of length ∆x = na .


2. For the ith interval, the height of the small triangle is b × ni and the base is ∆x.
3. The area of the ith small triangle is then 21 b × ni × ∆x.

To find the total area, sum over all such triangles and then take the limit as n goes to infinity:
n
1 i a
A = lim
n→∞
∑ 2b× n × n
i=1

Expanding the sum:

ab n
A = lim
n→∞ 2n2
∑i
i=1

Using the formula for the sum of the first n positive integers:
n
n(n + 1)
∑i= 2
i=1

We get:

ab n(n + 1) ab(n + 1) 1
A = lim × = lim = ab
n→∞ 2n2 2 n→∞ 4n 2

Thus, the area of the triangle is 12 ab.


168 11 SOLUTIONS

Solutions for Chapter 3


Solution 3.1. To find the greatest common divisor gcd of a = 36 and b = 48, we use the Euclidean algorithm:

48 = 36 · 1 + 12
36 = 12 · 3 + 0

The last non-zero remainder is 12, so the gcd(36, 48) = 12.

Solution 3.2. To calculate the least common multiple (lcm) of x = 9 and y = 15, we use the formula:
x·y
lcm(x, y) =
gcd(x, y)
.
First, we find the gcd:

15 = 9 · 1 + 6
9 = 6·1+3
6 = 3·2+0

Since the last remainder is zero, the gcd(9, 15) = 3. Therefore, the

9 · 15
lcm(9, 15) = = 45
3
.

Solution 3.3. To solve the congruence 2x ≡ 6 (mod 5), we simplify by dividing both sides by the gcd of the
coefficients (2 and 5), which is 1:

2x ≡ 6 (mod 5)
x≡3 (mod 5)

Therefore, x ≡ 3 (mod 5).

Solution 3.4. To solve the congruence:


3x ≡ 2 (mod 7)
we need to find the multiplicative inverse of 3 modulo 7. Specifically, we seek an integer y such that:

3y ≡ 1 (mod 7)

On checking values, we find that:


3×5 ≡ 1 (mod 7)

Multiplying the original congruence by 5 (the inverse of 3 mod 7) gives:


11 SOLUTIONS 169

15x ≡ 10 (mod 7)

Simplifying, we get:
x≡3 (mod 7)

Therefore, the integer solutions for x are of the form:

x = 3 + 7n

where n is any integer.

Solution 3.5. To compute 12 mod 7, we divide 12 by 7 and find the remainder:

12 = 7 · 1 + 5

Therefore, 12 mod 7 = 5.

Solution 3.6. To find 5−1 (mod 11) (the multiplicative inverse of 5 modulo 11), we use the extended
Euclidean algorithm. Applying the algorithm, we get:

11 = 5 · 2 + 1
1 = 11 − 5 · 2

Thus, 5−1 ≡ 9 (mod 11).

Solution 3.7. To solve the congruence:

3x ≡ 10 (mod 17)

we aim to find the multiplicative inverse of 3 modulo 17. Specifically, we are looking for an integer y such
that:
3y ≡ 1 (mod 17)

By testing values, we find:


3 × 6 ≡ 1 (mod 17)

Using this modular inverse, we multiply both sides of the original congruence by 6:

18x ≡ 60 (mod 17)

Upon simplification, this results in:


x ≡ 9 (mod 17)

Hence, the solution to the congruence is x ≡ 9 (mod 17).


170 11 SOLUTIONS

Solution 3.8. To solve the congruence:


4x ≡ 8 (mod 15)
we aim to find the multiplicative inverse of 4 modulo 15. Specifically, we are looking for an integer y such
that:
4y ≡ 1 (mod 15)

By testing values, we find:


4 × 4 ≡ 1 (mod 15)

Using this modular inverse, we multiply both sides of the original congruence by 4:

16x ≡ 32 (mod 15)

Upon simplification, this results in:


x ≡ 2 (mod 15)

Hence, the solution to the congruence is x ≡ 2 (mod 15).

Solution 3.9. To calculate the greatest common divisor (gcd) of m = 48 and n = 36, we use the Euclidean
algorithm:

48 = 36 · 1 + 12
36 = 12 · 3 + 0

The last non-zero remainder is 12, so the gcd(48, 36) = 12.

Solution 3.10. To calculate the least common multiple (lcm) of a = 12 and b = 18, we use the formula:
a·b
lcm(a, b) = .
gcd(a, b)
First, we find the gcd:

18 = 12 · 1 + 6
12 = 6 · 2 + 0

The last remainder is zero, so the gcd(12, 18) = 6. Therefore, the

12 · 18
lcm(12, 18) = = 36
6
.

Solution 3.11. To solve the congruence:

5x ≡ 3 (mod 11)

we aim to find the multiplicative inverse of 5 modulo 11. Specifically, we are looking for an integer y such
11 SOLUTIONS 171

that:
5y ≡ 1 (mod 11)

By testing values, we determine:


5 × 9 ≡ 1 (mod 11)

Using this modular inverse, we multiply both sides of the original congruence by 9:

45x ≡ 27 (mod 11)

Upon simplification, this results in:


x ≡ 5 (mod 11)

Hence, the solution to the congruence is x ≡ 5 (mod 11).

Solution 3.12. To solve the congruence:

4x ≡ 1 (mod 7)

we look for the multiplicative inverse of 4 modulo 7. This means we’re searching for an integer y such that:

4y ≡ 1 (mod 7)

By testing values, we find:


4×2 ≡ 1 (mod 7)

Thus, multiplying both sides of the original congruence by 2:

8x ≡ 2 (mod 7)

On reducing, we obtain:
x≡2 (mod 7)

Hence, the smallest value of x that satisfies the congruence is x = 2.

Solution 3.13. To solve the congruence:

2y ≡ 5 (mod 9)

we aim to determine the multiplicative inverse of 2 modulo 9. Specifically, we are looking for an integer x
such that:
2x ≡ 1 (mod 9)

By testing potential values, we identify:

2×5 ≡ 1 (mod 9)
172 11 SOLUTIONS

Using this modular inverse, we multiply both sides of the original congruence by 5:

10y ≡ 25 (mod 9)

Upon simplification, this yields:


y ≡ 7 (mod 9)

Consequently, the integer values of y that satisfy the congruence are those in the equivalence class of 7
modulo 9. The smallest positive integer that meets the congruence is y = 7.

Solution 3.14. To compute 15 mod 4, we find the remainder when 15 is divided by 4:

15 ÷ 4 = 3 with remainder 3

Thus,
15 mod 4 = 3
.
11 SOLUTIONS 173

Solutions for Chapter 4


Solution 4.1. This is an arithmetic series with the first term a = 3 and the common difference d = 2. To find
n (the number of terms), use:
an = a + (n − 1)d
39 = 3 + (n − 1)2
Solving for n, we get n = 19.

The sum of the series is given by:


n
Sn = (a1 + an )
2
19 19
S19 = (3 + 39) = × 42 = 399.
2 2
Solution 4.2. This is a p-series with p = 2. A p-series will converge if p > 1. Since p = 2 is greater than 1,
the series converges.

Solution 4.3. This is a geometric series with first term a = 4 and common ratio r = 12 . The sum of the first
n terms of a geometric series is given by:
a(1 − rn )
Sn =
1−r
4(1 − ( 21 )n )
Sn =
1 − 12
1
Sn = 4(1 − ( )n ) × 2
2
1
Sn = 8(1 − ( )n )
2

Solution 4.4. Proof. Let’s √
assume by contradiction that 2 is a rational number. It can then be expressed
as an irreducible fraction 2 = ba , where a and b are coprime integers (i.e., they have no common prime
factors).
a2
Squaring both sides of the equation, we get 2 = b2
. Multiplying both sides by b2 , we have 2b2 = a2 .
This implies that a2 is even, and therefore a is also even (since the square of an odd number is odd). We can
write a = 2k, where k is an integer.
Substituting a back into the original equation, we obtain 2b2 = (2k)2 , which simplifies to b2 = 2k2 . This
implies that b2 is even, and therefore b is also even.
We have reached a contradiction since √ we initially assumed that a and b have√no common prime factors.
Therefore, our initial assumption that 2 is rational is incorrect, implying that 2 is irrational.


Solution 4.5. Proof. Let’s assume by contradiction that there exists a positive real number x such that x ≥

x + 1.
174 11 SOLUTIONS

Squaring both sides of the inequality, we get x ≥ x + 1, which is a contradiction.


Therefore,
√ √ our initial assumption is incorrect, and we conclude that for any positive real number x, we have
x < x + 1.


Solution 4.6. Proof. We observe that each term of the sequence is obtained by adding 4 to the previous term.
Therefore, the general term of this arithmetic sequence is given by the equation:

an = 3 + 4(n − 1) (4.1)

Solution 4.7. Proof. We observe that each term of the sequence is obtained by multiplying the previous term
by 3. Therefore, the general term of this geometric sequence is given by the equation:

an = 2 · 3n−1 (4.2)

Solution 4.8. Proof. Using the formula for the sum of the first n terms of an arithmetic sequence, we have:
n
Sn = (a1 + an ) (4.3)
2
where Sn is the sum of the first n terms, a1 is the first term, and an is the last term. In this case, n = 5, a1 = 1,
and a4 = 10. Substituting these values into the formula, we get:

5 55
S5 = (1 + 10) = = 27.5 (4.4)
2 2
Therefore, the sum of the first 5 terms of the arithmetic sequence is 27.5.

1
Solution 4.9. Proof. The geometric series has a common ratio of . Using the formula for the infinite sum
2
of a geometric series, we have:
a
S= (4.5)
1−r
1
where S is the infinite sum, a is the first term, and r is the common ratio. In this case, a = 1 and r = .
2
Substituting these values into the formula, we get:
1 1
S= 1
= 1 =2 (4.6)
1− 2 2

Therefore, the infinite sum of the geometric series is 2.



11 SOLUTIONS 175

Solutions for Chapter 5


Solution 5.1. To calculate the complex conjugate of z = 3 + 4i, we simply change the sign of the imaginary
part, so the complex conjugate is z = 3 − 4i.

Solution 5.2. Given the complex numbers z = 2 + 3i and w = −1 + 2i, we can find z + w by adding the real
parts and the imaginary parts separately: z + w = (2 + 3i) + (−1 + 2i) = (2 − 1) + (3 + 2)i = 1 + 5i.
Now, to find
p the absolute √ value of z +√w, we use the formula for the absolute value of a complex number:
|z + w|= (1)2 + (5)2 = 1 + 25 = 26.

Therefore, the absolute value of z + w is 26.

Solution 5.3. To simplify the expression (4 − 2i)(3 + i) and write the result in rectangular form, we multiply
the terms using the distributive property: (4 − 2i)(3 + i) = 4(3) + 4(i) − 2i(3) − 2i(i).
Recall that i2 = −1, so we can replace i2 with −1: 4(3) + 4(i) − 2i(3) − 2i(i) = 12 + 4i − 6i − 2(−1).
Continuing the simplification: 12 + 4i − 6i − 2(−1) = 12 + 4i − 6i + 2 = 14 − 2i.
Therefore, the expression (4 − 2i)(3 + i) simplifies to 14 − 2i in rectangular form.

Solution 5.4. To prove that the product of a complex number z and its conjugate z is equal to the square of
the absolute value of z, we start with z = a + bi: z = a + bi.
The conjugate of z is z = a − bi.
Now, we multiply z by its conjugate: zz = (a + bi)(a − bi) = a2 − abi + abi − b2 i2 .
Recall that i2 = −1, so we can replace i2 with −1: a2 − abi + abi − b2 i2 = a2 − b2 (−1) = a2 + b2 .

The absolute value of z, denoted as |z|, is defined as a2 + b2 .
Therefore, we have shown that zz = |z|2 , i.e., the product of a complex number z and its conjugate z is equal
to the square of the absolute value of z.

Solution 5.5. Given the complex numbers z = 5 − 2i and w = −3 + 4i, we can find z + w by adding the real
parts and the imaginary parts separately: z + w = (5 − 2i) + (−3 + 4i) = (5 − 3) + (−2i + 4i) = 2 + 2i.
The complex conjugate of z + w is z + w = 2 − 2i.
Therefore, the complex conjugate of z + w is 2 − 2i.
π
Solution 5.6. If the absolute value of a complex number z is 7 and its argument is , we can find the
4
rectangular form of z using the following formulas:
a = |z|cos(θ ), where a is the real part of z.
b = |z|sin(θ ), where b is the imaginary part of z.
π
Since |z|= 7 and θ = , we can calculate:
4
π  √ √
2 7 2
a = 7 cos = 7· =
4 2 2
176 11 SOLUTIONS

.
π  √ √
2 7 2
b = 7 sin = 7· =
4 2 2
.
√ √
7 2 7 2
Therefore, the rectangular form of z is + i.
2 2
Solution 5.7. Let z = a + bi and w = c + di be two complex numbers, where a, b, c, and d are real numbers.
The product of the two complex numbers is:

zw = (a + bi)(c + di) = ac − bd + (ad + bc)i

Now, the absolute value (or modulus) of a complex number x + yi is given by:
p
|x + yi|= x2 + y2

Using this formula, we can find the modulus of z and w:


p
|z|= a2 + b2
p
|w|= c2 + d 2

Next, find the modulus of zw: q


|zw|= (ac − bd)2 + (ad + bc)2

Expanding and simplifying this expression:

|zw|2 = (ac − bd)2 + (ad + bc)2


= a2 c2 − 2acbd + b2 d 2 + a2 d 2 + 2adbc + b2 c2
= a2 (c2 + d 2 ) + b2 (c2 + d 2 )
= |z|2 |w|2

Taking the square root of both sides:


|zw|= |z||w|

Hence, the absolute value of the product of two complex numbers z and w is equal to the product of the
absolute values of z and w.

Solution 5.8. Given z = 2 + 3i, let’s solve for w = a + bi based on the conditions:
(a) The real part of w is equal to twice the imaginary part of z.
From this, we get:
a = 2×3 = 6
11 SOLUTIONS 177

(b) The conjugate complex of w is equal to the additive opposite of z.


The conjugate of w is a − bi. The additive opposite of z is −z, which is −2 − 3i.
Equating these, we obtain:
a − bi = −2 − 3i

From condition (a), a = 6:


6 − bi = −2 − 3i

Comparing the imaginary parts:


−b = −3
=⇒ b = 3

Hence, from the conditions provided:


w = 6 + 3i

However, note that this w only satisfies condition (a), but not condition (b). Thus, there isn’t a complex
number w that satisfies both conditions simultaneously.

Solution 5.9. Given the complex number z = 2 + 3i, let’s perform the following operations:
(a) To find the complex conjugate of z, we change the sign of the imaginary part. Hence:

z = 2 − 3i. (5.1)

(b) The absolute value of z is given by:


q p √
|z|= Re(z)2 + Im(z)2 = 22 + 32 = 13. (5.2)

(c) Let w be a complex number such that z + w = 4 − 2i. We can find w by subtracting z from the given
expression:
w = (4 − 2i) − (2 + 3i) = 2 − 5i. (5.3)

Hence, the solutions are:


(a) z = 2 − 3i.

(b) |z|= 13.
(c) w = 2 − 5i.

Solution 5.10. Given the complex number z = −2+5i, we need to find a real number r such that the absolute
value of z is equal to r. Then, we need to show that |z| is equal to the absolute value of its complex conjugate.
To find r, we calculate the absolute value of z:
q q √ √
|z|= Re(z)2 + Im(z)2 = (−2)2 + 52 = 4 + 25 = 29. (5.4)
178 11 SOLUTIONS

Now, let’s find the absolute value of the complex conjugate of z, which is z = −2 − 5i:
q q √ √
|z|= Re(z)2 + Im(z)2 = (−2)2 + (−5)2 = 4 + 25 = 29. (5.5)

Therefore, we can see that |z|= |z|, confirming that the absolute value of z is equal to the absolute value of
its complex conjugate.
11 SOLUTIONS 179

Solutions for Chapter 6


Solution 6.1. To solve the quadratic equation 3x2 + 5x − 2 = 0, we’ll use the quadratic formula:

−b ± b2 − 4ac
x=
2a
Where:
a = 3, b = 5, and c = −2

Plugging in the given values, we get:


p
−5 ± 52 − 4(3)(−2)
x1 , x2 =
2(3)

Calculating inside the square root:

52 − 4(3)(−2) = 25 + 24 = 49

Therefore:
−5 ± 7
x1 , x2 =
6
This results in two solutions:
2 1
x1 =
=
6 3
−12
x2 = = −2
6
Solution 6.2. To find the real solutions of the cubic equation

x3 − 4x2 + 3x + 2 = 0

we’ll look for a factor using the Rational Root Theorem.

Given possible rational roots as ±1, ±2, we find that x = 2 is a root. Thus, (x−2) is a factor. Using polynomial
division:
x3 − 4x2 + 3x + 2 ÷ (x − 2)
we get the quotient x2 − 3x − 1.

Now, solving for the quadratic equation:


x2 − 3x − 1 = 0
Using the quadratic formula:

−b ± b2 − 4ac
x=
2a
180 11 SOLUTIONS

Where:
a = 1, b = −3, and c = −1
we find the solutions for x.

Thus, the solutions for the equation x3 − 4x2 + 3x + 2 = 0 are x = 2 and the two solutions of the quadratic
equation.

Solution 6.3. p
22 − 4(1)(5)
−2 ±
x= (6.6)
2(1)
√ √
−2 + −16 −2 − −16
Therefore, the solutions are x = and x = , which are complex solutions.
2 2

Solution 6.4.
x4 − 5x2 + 4 = (x2 − 4)(x2 − 1) = 0 (6.7)
By applying factorization, we get x2 − 4 = 0 and x2 − 1 = 0. Therefore, the solutions are x = ±2 and x = ±1.

Solution 6.5. To find the roots of the polynomial A(x) = 2x2 − 5x + 2, we need to set the polynomial equal
to zero and solve the resulting equation:

2x2 − 5x + 2 = 0 (6.8)

We can use the factoring method, completing the square, or apply the quadratic formula to find the roots. In
this case, we will use the quadratic formula.
The quadratic formula to find the roots of a quadratic polynomial of the form ax2 + bx + c = 0 is:

−b ± b2 − 4ac
x= (6.9)
2a

Applying this formula to the polynomial A(x) = 2x2 − 5x + 2, we obtain:


p
−(−5) ± (−5)2 − 4(2)(2)
x= (6.10)
2(2)

Simplifying the expression, we get:



5 ± 25 − 16
x= (6.11)
4

5± 9
x= (6.12)
4
5±3
x= (6.13)
4
11 SOLUTIONS 181

Therefore, the roots of the polynomial A(x) = 2x2 − 5x + 2 are:


5+3
x1 = =2 (6.14)
4
5−3 1
x2 = = (6.15)
4 2
This means that the possible lengths of the sides of the rectangle are 2 meters and 0.5 meters, respectively.
The farmer can use this information to determine the appropriate dimensions of the fence and enclose the
desired area on his land.

Solution 6.6. To find the time at which the object hits the ground, we need to solve the equation h(t) = 0.
Substituting the expression for h(t), we have:

−4.9t 2 + 20t + 100 = 0 (6.16)

This is a quadratic equation. We can use the quadratic formula to solve it:

−b ± b2 − 4ac
t= (6.17)
2a
Where a = −4.9, b = 20, and c = 100. Substituting these values into the formula, we obtain:
p
−20 ± 202 − 4(−4.9)(100)
t= (6.18)
2(−4.9)

Simplifying the expression inside the square root, we have:



−20 ± 400 + 1960
t= (6.19)
−9.8

Continuing with the simplification, we get:



−20 ± 2360
t= (6.20)
−9.8

Finally, we can calculate the two possible values of t:


√ √
−20 + 2360 −20 − 2360
t1 = and t2 = (6.21)
−9.8 −9.8

Therefore, the object hits the ground at two different times, t1 and t2 .

Solution 6.7. To find the complex roots of the polynomial, we can use the method of grouping. We begin by
grouping the terms conveniently:

f (x) = (x4 + 2x3 ) + (3x2 + 4x + 5) = 0 (6.22)


182 11 SOLUTIONS

Next, we factorize each group separately:

f (x) = x3 (x + 2) + (3x2 + 4x + 5) = 0 (6.23)

Now, we will attempt to factorize the second group of terms using the method of quadratic factorization. We
calculate the discriminant:

∆ = b2 − 4ac = (4)2 − 4(3)(5) = 16 − 60 = −44 (6.24)

Since the discriminant is negative, we know that the polynomial has complex roots. We will use the general
formula for complex roots of a quadratic equation:

−b ± ∆
x= (6.25)
2a
Substituting the corresponding values:

−4 ± −44
x= (6.26)
2(3)

Simplifying:
√ √
−4 ± 2i 11 −2 ± i 11
x= = (6.27)
6 3
Therefore, the complex roots of the polynomial are:
√ √
−2 + i 11 −2 − i 11
x1 = and x2 = (6.28)
3 3
These are the complex solutions of the original polynomial.
11 SOLUTIONS 183

Solutions for Chapter 7


Solution 7.1. To find the cross product u × v, we’ll use the determinant of the following 3 × 3 matrix:

i j k
1 2 −1
−2 3 4

Expanding along the top row:


For i:  
2 −1
det = (2 × 4) − (−1 × 3) = 11
3 4

For j:  
1 −1
det = (1 × 4) − (−1 × −2) = 2
−2 4
But, recall the alternating signs when expanding along the top row, thus it becomes -2 for j.

For k:  
1 2
det = (1 × 3) − (2 × −2) = 7
−2 3

So, the cross product u × v is:  


11
−2
7

Solution 7.2. The dot product (or scalar product) of two vectors is given by the sum of the products of their
corresponding components.
Given:  
2
u = −3
1
 
−1
v= 4 
2

The dot product u · v is:


u · v = u1 v1 + u2 v2 + u3 v3

Where u1 , u2 , and u3 are the components of u, and v1 , v2 , and v3 are the components of v.
Substituting in the given values:

u · v = (2 × −1) + (−3 × 4) + (1 × 2)
184 11 SOLUTIONS

= −2 − 12 + 2
= −12

So, the dot product of u and v is:


u · v = −12

Solution 7.3. To determine if the vectors v1 , v2 , and v3 are linearly independent, we can set up a matrix
using these vectors as columns and then compute the determinant. If the determinant is non-zero, the vectors
are linearly independent. If the determinant is zero, the vectors are linearly dependent.
Form a matrix A using the vectors:  
1 2 −1
A = 2 −1 3 
3 4 5

Now, compute the determinant of A:

1 2 −1
det(A) = 2 −1 3
3 4 5

Expanding along the top row:

−1 3 23 2 −1
det(A) = 1 × −2× + (−1) ×
4 5 35 3 4

Now, compute the 2 × 2 determinants:

−1 3
= (−1 × 5) − (3 × 4) = −5 − 12 = −17
4 5

23
= (2 × 5) − (3 × 3) = 10 − 9 = 1
35

2 −1
= (2 × 4) − (−1 × 3) = 8 + 3 = 11
3 4

Substitute these values into the expansion:

det(A) = 1(−17) − 2(1) + (−1)(11) = −17 − 2 − 11 = −30

Since the determinant is −30 (non-zero), the vectors v1 , v2 , and v3 are linearly independent.
 
1
Solution 7.4. To calculate T , we simply substitute the coordinates x = 1 and y = −2 into the
−2
expression for the transformation:
11 SOLUTIONS 185
   
  2(1) 2
1
T =  3(−2)  = −6 (7.29)
−2
1 + (−2) −1

 
  2
1
Therefore, the image of under the transformation T is −6.
−2
−1
 
1
Solution 7.5. To find the image of −2 under the transformation T , we substitute the coordinates x = 1,
4
y = −2, and z = 4 into the expression for the transformation:
 
1    
2(1) + (−2) 0
T −2 = = (7.30)
3(−2) − 4 −10
4

 
1  
0
Therefore, the image of −2 under the transformation T is
  .
−10
4

Solution 7.6. To verify if T preserves vector


 addition,
 we  need to check if T (u + v) = T (u) + T (v) for any
2 u1 v1
pair of vectors u and v in R . Let u = and v = be two generic vectors. Then:
u2 v2
 
u1 + v1
T (u + v) = T
u2 + v2
(7.31)
 
3(u1 + v1 )
=  −2(u2 + v2 ) 
(u1 + v1 ) + (u2 + v2 )

And:
   
3u1 3v1
T (u) + T (v) =  −2u2  +  −2v2 
u1 + u2 v1 + v2
  (7.32)
3u1 + 3v1
=  −2u2 − 2v2 
u1 + u2 + v1 + v2

Since T (u + v) = T (u) + T (v) for any u, v ∈ R2 , we conclude that T preserves vector addition.

Solution 7.7. To find the


 matrix
 of
 the
 linear transformation
  T , we need to determine how T acts on the
1 0 0
canonical vectors e1 = 0, e2 = 1, and e3 = 0. Evaluating T on these vectors, we have:
0 0 1
186 11 SOLUTIONS
   
1+0+0 1
T (e1 ) = =
2(1) − 0 2
   
0+1+0 1
T (e2 ) = =
2(0) − 1 −1
   
0+0+1 1
T (e3 ) = =
2(0) − 0 0

The matrix of the linear transformation T is formed by taking the columns corresponding to the images of
the canonical vectors:  
1 1 1
[T ] = (7.33)
2 −1 0
 
1 1 1
Therefore, the matrix of the linear transformation T is .
2 −1 0

Solution 7.8. Step 1: Decomposition of forces into x and y components To be able to sum the vector forces,
we first need to decompose each force into its x and y components. We will use trigonometry for this.

F1x = F1 · cos(θ1 ) and F1y = F1 · sin(θ1 ) (7.34)

Where F1x is the x-component of force F1, F1y is the y-component of force F1 , and θ1 is the angle of F1
with respect to the positive x-axis.

F2x = F2 · cos(θ2 ) and F2y = F2 · sin(θ2 ) (7.35)

Where F2x is the x-component of force F2, F2y is the y-component of force F2 , and θ2 is the angle of F2
with respect to the positive x-axis.
Step 2: Sum of x and y components of forces. Now we will sum the x and y components of the forces to
obtain the x and y components of the resultant force.

Frx = F1x + F2x and Fry = F1y + F2y (7.36)

Where Frx is the x-component of the resultant force, and Fry is the y-component of the resultant force.
Step 3: Calculation of magnitude and angle of the resultant force. Finally, we will use the x and y components
of the resultant force to calculate its magnitude and angle.
 
q Fry
Fr = Frx2 + Fry2 and θr = arctan (7.37)
Frx

Where Fr is the magnitude of the resultant force, and θr is the angle of the resultant force with respect to the
positive x-axis.
11 SOLUTIONS 187

F1
Fr
θ2 θ1
x
F2

Figure 7.1: Illustration of the Resultant Force Problem.


188 11 SOLUTIONS

Solutions for Chapter 8


Solution 8.1. Given the system of linear equations:

2x + 3y = 10 (8.1)
4x − 2y = 2 (8.2)

Expressing this system in matrix form AX = B, we have:


 
2 3
A=
4 −2
 
x
X=
y
 
10
B=
2

Using Cramer’s Rule:


det(Ax )
x=
det(A)
det(Ay )
y=
det(A)

Where:  
10 3
Ax =
2 −2
 
2 10
Ay =
4 2

Calculating the determinants:


det(A) = −16
det(Ax ) = −26
det(Ay ) = −36

The solutions are:


13
x=
8
9
y=
4
Solution 8.2. We will use the Sarrus rule to calculate the determinant of a 3 × 3 matrix. Applying the
formula: Determinant= (4)(−1)(−2) + (2)(5)(2) + (3)(0)(3) − (3)(−1)(2) − (2)(0)(−2) − (4)(5)(3) =
8 + 20 + 0 − (−6) − 0 − 60 = −26.
Therefore, the determinant of the given matrix is −26.
11 SOLUTIONS 189

Solution 8.3. Given the matrix:  


21 0
A = 0 3 4
00 5

The eigenvalues of the matrix are:


λ1 = 2, λ2 = 3, λ3 = 5

The corresponding eigenvectors are:


     
1 1 2
v1 = 0 , v2 = 1 , v3 = 6
0 0 1

The matrix can be diagonalized as A = PDP−1 , where:


 
200
D = 0 3 0
005
 
112
P = 0 1 6
001

Solution 8.4. Given the matrix:  


31
A=
12

The eigenvalues of the matrix are: √


5+ 5
λ1 =
2

5− 5
λ2 =
2
The corresponding eigenvectors are:  
1√
v1 = −1+ 5
2
 
1√
v2 = −1− 5
2

1
Solution 8.5. To calculate the inverse of a matrix, we use the formula: A−1 = · adj(A), where det(A)
det(A)
is the determinant of matrix A and adj(A) is the adjugate matrix of A.
First, we calculate the determinant of the matrix:
190 11 SOLUTIONS

21
det(A) = = (2)(4) − (1)(3) = 5 (8.3)
34

Next, we calculate the adjugate matrix of A:


 
4 −1
adj(A) = (8.4)
−3 2

Therefore, the inverse of the given matrix is:


 
4 1
1 4 −1
 −
A−1 = · =  53 25  (8.5)
 
5 −3 2 −
5 5
Solution 8.6. To add matrices, we simply add the corresponding elements:
       
21 −1 2 2 + (−1) 1 + 2 13
+ = = (8.6)
34 0 1 3+0 4+1 35

Therefore, the sum of the two matrices is:


 
13
(8.7)
35

Solution 8.7. We calculate the determinant of the complete system (D) and the determinants of the matrices
obtained by replacing the coefficient column of x, y, and z with the column of constant terms (Dx , Dy , and
Dz , respectively).
The determinant of the complete system D is calculated as follows:

2 3 −1
D = 1 −2 2 (8.8)
3 2 −5

Simplifying the determinant, we have:

D = 10 + 18 − 2 − 6 − 12 + 20 = 28 (8.9)

Next, we calculate the determinants Dx , Dy , and Dz , by replacing the corresponding column with the column
of constant terms:

7 3 −1
Dx = −4 −2 2 (8.10)
12 2 −5

Simplifying the determinant Dx , we have:

Dx = 70 + 72 + 8 − 24 − 48 + 40 = 118 (8.11)
11 SOLUTIONS 191

2 7 −1
Dy = 1 −4 2 (8.12)
3 12 −5

Simplifying the determinant Dy , we have:

Dy = 40 + 42 − 12 − 12 − 12 + 20 = 66 (8.13)

2 3 7
Dz = 1 −2 −4 (8.14)
3 2 12

Simplifying the determinant Dz , we have:

Dz = −48 − 36 + 14 + 42 − 18 − 96 = −140 (8.15)

Finally, we can find the values of x, y, and z using Cramer’s rule:


Dx 118
x= = = 4.214 (8.16)
D 28
Dy 66
y= = = 2.357 (8.17)
D 28
Dz −140
z= = = −5 (8.18)
D 28
Solution 8.8. To calculate the determinant of the matrix
 
ab c
A = d e f
gh i

we use the formula:


     
e f d f de
det(A) = a × det − b × det + c × det
h i g i gh

Using the determinant formula for a 2 × 2 matrix:


 
mn
det = mp − no
o p

The calculations are:  


e f
det = ei − f h
h i
192 11 SOLUTIONS
 
d f
det = di − f g
g i
 
de
det = dh − eg
gh

Therefore, the determinant of A is:

det(A) = a(ei − f h) − b(di − f g) + c(dh − eg)


11 SOLUTIONS 193

Solutions for Chapter 9


Solution 9.1. To find the determinant of the matrix
 
3 1 2
A =  0 −2 4
−1 3 5

using cofactor expansion, we expand along the first row:

det(A) = a11 ×C11 + a12 ×C12 + a13 ×C13

Where Ci j are the cofactors given by:

Ci j = (−1)i+ j × det(Mi j )

and Mi j is the minor matrix obtained by removing the ith row and jth column from matrix A.

For the first entry:  


−2 4
C11 = (−1)1+1 × det = 22
3 5
For the second entry:
 
1+2 0 4
C12 = (−1) × det = −4
−1 5
For the third entry:
 
1+3 0 −2
C13 = (−1) × det =2
−1 3

Therefore, combining the cofactors with the entries of the first row:

det(A) = 3(22) + 1(−4) + 2(2) = 66

Solution 9.2. Given the system of equations:

2x + 4y − z = 6
x − 3y + 2z = 7
3x − 2y + 3z = 4

We can represent it with the augmented matrix:


 
2 4 −1 | 6
1 −3 2 | 7
3 −2 3 | 4
194 11 SOLUTIONS

Using Gaussian elimination, we transform the matrix to reduced row echelon form:
 
1 0 0 | 16
0 1 0 | −6 
0 0 1 | −22.8

From which we can directly read off the solution:

x = 16, y = −6, z = −22.8

Solution 9.3. Given the vectors


     
1 4 7
v1 = 2 , v2 = 5 , and v3 = 8
3 6 9

we form a matrix with these vectors as columns:


 
147
A = 2 5 8 
369

To determine the linear independence of the vectors, we evaluate the determinant of this matrix. The rows
(and columns) of the matrix exhibit a linear pattern; each entry is one greater than the previous entry. This
indicates the rows (and columns) are linearly dependent. Therefore, the determinant of matrix A is zero.
As a result, the vectors v1 , v2 , and v3 are linearly dependent.

Solution 9.4.
11 SOLUTIONS 195

Solutions for Chapter 10


Solution 10.1. To prove that the identity element of G is also the identity element of H, we need to show
that for any h in H, h ∗ e = e ∗ h = h, where e is the identity element of G.
Proof of h ∗ e = h: Since H is a subgroup of G, h ∗ e belongs to H. But as H uses the same operation as G,
h ∗ e is equal to h. Hence, h ∗ e = h.
Proof of e ∗ h = h: Similarly, as h belongs to H, e ∗ h also belongs to H. But as H uses the same operation as
G, e ∗ h is equal to h. Hence, e ∗ h = h.
We have demonstrated that the identity element of G is also the identity element of H.
Solution 10.2. Let G and H be groups, and let φ : G → H be an isomorphism. We want to prove that the
inverse φ −1 : H → G is also an isomorphism.
1. Bijectiveness of φ −1 : As φ is an isomorphism, it is bijective. Consequently, it has an inverse function
φ −1 : H → G. Since φ is both injective and surjective, it follows that φ −1 is also injective and surjective.
Thus, φ −1 is bijective.

2. Preservation of the Group Operation: Let h1 , h2 ∈ H. Since φ is surjective, there exist g1 , g2 ∈ G such
that φ (g1 ) = h1 and φ (g2 ) = h2 . Now, considering the composition,

φ −1 (h1 h2 ) = φ −1 (φ (g1 )φ (g2 )) = g1 g2 ,

and,
φ −1 (h1 )φ −1 (h2 ) = g1 g2 .
Thus, φ −1 (h1 h2 ) = φ −1 (h1 )φ −1 (h2 ), indicating that φ −1 preserves the group operation.

3. Preservation of the Identity Element: Let eG be the identity element in G and eH be the identity element
in H. Since φ is an isomorphism, we have φ (eG ) = eH . Evaluating the inverse, we get φ −1 (eH ) = eG .
This shows that φ −1 maps the identity element of H to the identity element of G.

Conclusion: If φ : G → H is an isomorphism, then its inverse φ −1 : H → G is also an isomorphism.


Solution 10.3. To prove that hai is the smallest subgroup of G containing a, we need to show that any
subgroup K of G that contains a must also contain all the elements generated by a.
Proof that hai is a subgroup: Since hai is the subgroup generated by a, it contains the identity element e,
which is the result of raising a to the power of 0. It also contains all the powers of a, i.e., an for all integers
n. Therefore, hai satisfies the closure and inverse properties required for a subgroup.
Solution 10.4. To prove that the kernel of φ , denoted as ker(φ ), is a subgroup of G, we need to show that it
satisfies the group properties: closure, associativity, identity element, and inverses.
Closure: Let x, y be elements in ker(φ ). This means that φ (x) = φ (y) = e, where e is the identity element in
H. We need to show that x ∗ y is also in ker(φ ).
Proof: Since φ is a homomorphism, we have φ (x ∗ y) = φ (x) ∗ φ (y) = e ∗ e = e. Therefore, x ∗ y is in ker(φ ),
and closure is satisfied.
196 11 SOLUTIONS

Associativity: Since G is a group, we know that the operation ∗ is associative. Therefore, associativity holds
for elements in ker(φ ) as well.
Identity Element: The identity element of G is also in ker(φ ). This is because φ (eG ) = eH , where eG and eH
are the identity elements of G and H, respectively.
Inverses: For each element x in ker(φ ), we need to show that its inverse x−1 is also in ker(φ ).
Proof: Let x be an element in ker(φ ). This means φ (x) = e. We want to show that φ (x−1 ) = e as well.
Since φ is a homomorphism, we have φ (x−1 ) = φ (x)−1 = e−1 = e.
Therefore, x−1 is in ker(φ ), and inverses are satisfied.
Thus, we have shown that the kernel of φ , ker(φ ), is a subgroup of G.
Appendix A
Computational Programs

Carlos Polanco
Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México.
Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de México, Ciudad de
México, México.

Abstract This chapter presents the computational programs that were used to calculate all the automated
processes in this book. These programs were developed in language FORTRAN-95 and Linux scripts.

Keywords: FORTRAN-95, Newton’s method.

A.1. Introduction
FORTRAN−95 is an extensively used high-level programming language in the scientific community.
Computer programs written in FORTRAN-95, renowned for their proficiency in numerical computation,
enable the resolution of complex problems using robust and accurate algorithms. It is a powerful tool for
scientific data processing and the simulation of physical systems due to its ability to manipulate matrices and
conduct vector calculations. Additionally, FORTRAN-95’s structured syntax and object-oriented features
facilitate the development of programs that are maintainable and scalable over time.

A.2. Newton’s Method Program

(B) Carlos Polanco: Department of New Technologies and Intellectual Protection, Instituto Nacional de Cardiologı́a “Ignacio
Chávez”. Ciudad de México, México. Department of Mathematics, Faculty of Sciences, Universidad Nacional Autónoma de
México; Tel: +01 55 5595 2220; E-mail: polanco@unam.mx

197
198 A Computational Programs

1 program newtons_method
2 implicit none
3 real :: x0, x, f, f_prime
4 integer :: i, max_iter
5 real, parameter :: tolerance = 1e-6
6

7 ! Initialization of the initial value


8 ! and maximum number of iterations
9 x0 = 1.5
10 max_iter = 100
11
12 ! Newton’s method iteration
13 do i = 1, max_iter
14 f = f_func(x0)
15 f_prime = f_prime_func(x0)
16 x = x0 - f / f_prime
17
18 ! Check for convergence
19 if (abs(x - x0) < tolerance) then
20 write(*,*) "The approximate root is:", x
21 exit
22 end if
23

24 x0 = x
25 end do
26

27 ! If convergence is not reached after


28 ! the maximum number of iterations
29 write(*,*) "The method does not converge in", max_iter,
30 & "iterations."
31
32 contains
33
34 ! Definition of the quadratic function f(x) = xˆ2 - 2
35 function f_func(x) result(f_val)
36 real, intent(in) :: x
37 real :: f_val
38 f_val = x**2 - 2
39 end function f_func
40

41 ! Definition of the derivative of the quadratic function


42 ! f’(x) = 2x
43 function f_prime_func(x) result(f_prime_val)
44 real, intent(in) :: x
45 real :: f_prime_val
46 f_prime_val = 2 * x
A.2 Newton’s Method Program 199

47 end function f_prime_func


48
49 c end program newtons_method
50
51 ! INPUT DATA
52 ! NONE
53
54 ! OUTPUT DATA
55 ! The approximate root is: 1.41421354
56 ! The method does not converge in 100 iterations.
200 A Computational Programs

A.3. Conclusions
In conclusion, this appendix provides an introduction to programming in FORTRA N−95, a widely used
programming language in scientific and numerical computing. FORTRAN−95 offers powerful features and
capabilities for performing complex algebraic computations, making it a valuable tool for implementing
algorithms and solving mathematical problems. By familiarizing readers with the syntax, data types, control
structures, and array operations in Fortran 95, this appendix aims to enhance their understanding and
proficiency in applying algebraic concepts and techniques using computer programming. It serves as a
valuable resource for readers seeking to expand their mathematical skills and leverage the computational
power of FORTRAN−95 in the field of higher algebra.
References 201

References
[1] S. Lipschutz, “Teoria de conjuntos y temas afines,” http://es.pdfdrive.com/teoria-de-conjuntos-y-temas-
afines-e165945669.html.
[2] C. Polanco, “Advanced Calculus - Fundamentals of Mathematics,”
https://benthambooks.com/book/9789811415081/foreword/.
[3] S. Lang, Algebra. Springer, 2002.
[4] J. J. Rotman, Advanced Modern Algebra. Pearson College Division. U.S.A., 2002.
[5] M. Artin, Algebra. Pearson, 2011.
[6] “Problemas de divisibilidad,” https://matematicasiesoja.files.wordpress.com/2
013/09/problemas-de-divisibilidad2-con-soluciones.pdf, aug 12 2020, [Online; accessed 2023-06-07].
[7] I. N. Herstein, Topics in Algebra. Wiley, 1986.
[8] R. Hammack, Book of Proof. University of Richmond, Virginia. U.S.A., 2018.
[9] “An Introduction to the Theory of Numbers: Hardy, G. H., Wright, E. M.: 9780198531715:
Amazon.com: Books,” https://www.amazon.com/exec/obidos/ASIN/0198531710/ref=nosim/ericstre
asuretro, apr 17 1980.
[10] “Euclidean algorithm - Wikipedia,” https://en.wikipedia.org/wiki/Euclidean al
gorithm, nov 1 2016.
[11] M. Corral, “Vector calculus,” 2018, http://omega.albany.edu:8008/mat214di r/CoralCalc3BOOK.pdf.
[12] D. S. Dummit and R. M. Foote, Abstract Algebra. Wiley, 2004.
[13] OpenAI, “Example of gcd, lcm, and fundamental theorem of arithmetic,” https://www.openai.com,
2023, accessed on June 8, 2023.
[14] A. Gorodnik, “Lecture 3: Congruences,” https://www.math.uzh.ch/gorodnik/nt/lecture3.pdf, [Online;
accessed 2023-06-14].
[15] “Prime number theorem - Wikipedia,” https://en.wikipedia.org/wiki/Prime number theorem, aug 3
2010.
[16] OpenAI, “Chinese remainder theorem,” Personal knowledge, 2023, description of the Chinese
Remainder Theorem.
[17] N. Ferguson, B. Schneier, and T. Kohno, Cryptography Engineering: Design Principles and Practical
Applications. Indianapolis, IN: Wiley, 2010.
[18] J. Gallian, Contemporary Abstract Algebra, 10th ed. Taylor & Francis Group LLC, 2021.
[19] A. Turing, “On computable numbers, with an application to the entscheidungsproblem,” Proceedings
of the London Mathematical Society, vol. 42, no. 2, pp. 230–265, 1936.
[20] W. Rudin, Principles of Mathematical Analysis. McGraw-Hill, 1976.
[21] J. W. Brown, R. V. Churchill, and J. Brown, Complex variables and applications. McGraw-Hill, 2010.
[22] É. Cartan, “Les groupes de transformations continus, infinis, simples,” Annales Scientifiques de l’École
Normale Supérieure, vol. 30, no. 1, pp. 1–99, 1913.
[23] E. W. Weisstein. (2021) Complex number. Accessed: 2021-09-30. [Online]. Available:
https://mathworld.wolfram.com/ComplexNumber.html
[24] R. L. Burden, J. D. Faires, and A. M. Burden, Numerical Analysis, 10th ed. Cengage Learning, 2016.
[25] S. Brown, “Applications of vector analysis in engineering,” in Proceedings of the International
Conference on Engineering. IEEE, 2012, pp. 123–135.
[26] R. Jones, Linear Algebra and Vector Spaces. New York: Wiley, 2015.
[27] B. C. Hall, Lie Groups, Lie Algebras, and Representations: An Elementary Introduction. Springer,
2015.
[28] J. J. Rotman, An Introduction to the Theory of Groups. Springer, 1995.
Index

Absolute value over Cubic equation, 97


complex numbers, 86
real numbers, 76 De Moivre’s theorem, 88
Adjugate matrix, 123 Determinant
Algebraic numbers, 17 operations, 123
Archimedean property, 23 operator, 122
Arithmetic module, 63 Determinants, 136
Divergent
Binary relation, 30 sequences, 51
Bézout’s theorem, 66 series, 51
Dot product, 111
Cardano formula, 97
Cardinality, 2 Eigenvalues, 127, 143
Cartesian product, 30 Eigenvectors, 127, 143
Chineses remainder theorem, 67 Elementary functions, 35
Cofactor expansion, 137 Euclid’s theorem, 65
Complex numbers, 85 Exponential function, 36
axioms, 85 Exponents, 80
conjugate, 85
plane, 86 Factor theorem, 95
properties, 85 Factorization of polynomials, 94
roots, 100 FORTRAN-95, 197
Complex roots, 102 Functions, 33
Composition of functions, 39 Fundamental theorem of
Congruence-based cryptography system, 69 algebra, 100
Congruences, 61 arithmetic, 59
Convergent
sequences, 51 Gaussian elimination, 139
series, 51 Generator Group, 158
Cramer’s rule, 126 Greatest common divisor, 56
Cross product, 112 Groups, 154

203
204 INDEX

Homogeneous systems, 141 Polynomial equations, 96


Homomorphisms, 157 Polynomial function, 35
Polynomial functions, 94
Imaginary unit, 25
Infinite Quadratic equation, 96
sequences, 77
series, 77 Radicals, 80
Integer numbers, 11 Rank of a Matrix, 147
Irrational numbers, 15 Real numbers, 18, 74
Isomorphisms, 155 axioms, 74
inequalities, 76
Least common multiple, 58 order properties, 76
Limit operator, 42 properties, 74
Linear congruences, 64 Real-valued functions, 33
Linear independence, 114, 148
Linear transformations, 117 Sarrus rule, 138
Logarithmic function, 38 Sequences, 48
Series, 48
Mathematical induction, 8 Spanning sets, 115
Matrix Square roots of negative numbers, 89
diagonalization, 130 Subgroups, 155
inverse, 124 Symmetric difference, 3
non-singular, 124 Synthetic division, 97
operations, 121 System linear equations, 125
operator, 120
Matriz Trascendental numbers, 17
square, 124 Triangle inequality, 21
Trigonometric function, 37
Natural numbers, 6
Newton’s method, 102 Vector dimension, 116
Newton’s method program, 197 Vector space, 108
Vector subspace, 114
Polar representation, 86 Vectors, 108

View publication stats

You might also like