a first course in linear algebrabuzzard.ups.edu/linear/download/fcla-onea4-0.22.pdf ·...

509
A First Course in Linear Algebra by Robert A. Beezer Department of Mathematics and Computer Science University of Puget Sound Version 0.22 November 18, 2004 c 2004

Upload: others

Post on 14-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

A First Course in Linear Algebra

byRobert A. Beezer

Department of Mathematics and Computer ScienceUniversity of Puget Sound

Version 0.22November 18, 2004

c© 2004

Page 2: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Preface

This textbook is designed to teach the university mathematics student the basics ofthe subject of linear algebra. There are no prerequisites other than ordinary algebra,but it is probably best used by a student who has the “mathematical maturity” of asophomore or junior.

The text has two goals: to teach the fundamental concepts and techniques of ma-trix algebra and abstract vector spaces, and to teach the techniques of developing thedefinitions and theorems of a coherent area of mathematics. So there is an emphasis onworked examples of non-trivial size and on proving theorems carefully.

This book is free. That means you may use it at no cost, other than downloading it,and perhaps choosing to print it. But it is also free in another sense of the word, it hasfreedom. While copyrighted, this is only to allow some modicum of editorial control overthe central portions of the text. Otherwise, it has been designed to be expanded throughcontributions from others, and so that it can be easily customized for different uses. Itwill not ever go “out of print” nor will updates be designed to frustrate the used bookmarket. You may make as many copies as you want, and use them however you see fit.

Topics The first half of this text (through Chapter M [121]) is basically a course inmatrix algebra, though the foundation of some more advanced ideas is also being laidin these early sections. Vectors are presented exclusively as column vectors (since wealso have the typographic freedom to avoid the cost-cutting move of displaying columnvectors inline as the transpose of row vectors), and linear combinations are presentedvery early. Spans, null spaces and ranges are also presented very early, simply as sets,saving most of their vector space properties for later, so they are familiar objects beforebeing scrutinized carefully.

You cannot do everything early, so in particular matrix multiplication comes late.However, with a definition built on linear combinations of column vectors, it shouldseem more natural than the usual definition using dot products of rows with columns.And this delay emphasizes that linear algebra is built upon vector addition and scalarmultiplication. Of course, matrix inverses must wait for matrix multiplication, but thisdoesn’t prevent nonsingular matrices from occurring sooner. Vector space propertiesare hinted at when vectors and matrices are first defined, but the notion of a vectorspace is saved for a more axiomatic treatment later. Once bases and dimension havebeen explored in the context of vector spaces, linear transformations and their matrixrepresentations follow. The goal of the book is to go as far as canonical forms and matrixdecompositions.

Linear algebra is an ideal subject for the novice mathematics student to learn howto develop a topic precisely, with all the rigor mathematics requires. Unfortunately,much of this rigor seems to have escaped the standard calculus curriculum, so for manyuniversity students this is their first exposure to careful definitions and theorems, andthe expectation that they fully understand them, to say nothing of the expectation that

i

Page 3: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

ii

they become proficient in formulating their own proofs. We have tried to make this textas helpful as possible with this transition. Every definition is stated carefully, set apartfrom the text. Likewise, every theorem is carefully stated, and almost every one has acomplete proof. Definitions and theorems are cataloged in order of their appearance inthe front of the book, and in the index at the back. Along the way, there are discussionsof some of the more important ideas relating to formulating proofs (Proof Techniques),which is advice mostly.

Freedom This book is freely-distributable. When it is ready, I plan to release thecomplete edition under something like a Creative Commons license that will formalizethis arrangement. Once it settles down, I also plan to release the TEX source as well.This arrangement will provide many benefits unavailable with traditional texts, such as:

• No cost, or low cost, to students. With no physical vessel (i.e. paper, binding), notransportation costs (Internet bandwidth being a negligible cost) and no marketingcosts (evaluation and desk copies are free to all), it can be obtained by anyonewith a computer and an Internet connection, and a teacher can make availablepaper copies in sufficient quantities for a class. The cost to print a copy is notinsignificant, but is just a fraction of the cost of a regular textbook. Students willnot feel the need to sell back their book, and in future years can even pick up anewer edition.

• The book will not go out of print. No matter what, a teacher will be able tomaintain their own copy and use the book for as many years as they desire. Further,the naming schemes for chapters, sections, theorems, etc. is designed so that theaddition of new material will not break any course syllabi.

• With many eyes reading the book and with frequent postings of updates, the relia-bility should become very high. Please report any errors you find that persist intothe latest version.

• For those with a working installation of the popular typesetting program TEX,the book has been designed so that it can be customized. Page layout, presenceof exercises and/or solutions, presence of sections or chapters can all be easilycontrolled. Furthermore, various pieces of mathematical notation are achieved viaTEX macros. So by changing a single macro, one’s favorite notation can be reflectedthroughout the text. For example, every transpose of a matrix is coded in thesource as \transpose{A}, which when printed will yield At. However by changingthe definition of \transpose{ }, any desired alternative notation will then appearthroughout the text instead.

• The book has also been designed to make it easy for others to contribute material.Would you like to see a section on symmetric bilinear forms? Consider writing oneand contributing it. Does there need to be more exercises about the null space of amatrix? Send me some. Historical Notes? Contact me, and we’ll see about addingthose in also.

I can think of just one drawback. The author is not guaranteed any royalties.This project is the result of the confluence of several events:

Page 4: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

iii

• Having taught an introductory linear algebra course eighteen times prior, I decidedin January 2003 to type up a good set of course notes. Then I decided to makecopies for the students, necessitating just a bit more quality control. They becamemore and more complete as the semester wore on. I used them again for the Fall2003 semester, and spent a sabbatical during the Spring 2004 semester doing atotal rewrite with an eye toward book form, and incorporating many new ideas Ihad for how best to present the material.

• I’ve used TEX and the Internet for many years, so there is little to stand in the wayof typesetting, distributing and “marketing” a free book.

• In September 2003, both of the textbooks I was using for my courses came outin new editions. Trivial changes required complete rewrites of my syllabi and webpages, and substantial reorganization of the notes that preceded.

• With recreational and professional interests in software development, I’ve long beenfascinated by the open-source software movement, as exemplified by the success ofGNU and Linux, though public-domain TEX might also deserve mention. It shouldbe obvious that this is an attempt to carryover that model of creative endeavor totextbook publishing.

However, much of my motivation for writing this book is captured by H.M. Cundy andA.P. Rollet in their Preface to the First Edition of Mathematical Models (1952), especiallythe final sentence,

This book was born in the classroom, and arose from the spontaneous interestof a Mathematical Sixth in the construction of simple models. A desire toshow that even in mathematics one could have fun led to an exhibition ofthe results and attracted considerable attention throughout the school. Sincethen the Sherborne collection has grown, ideas have come from many sources,and widespread interest has been shown. It seems therefore desirable to givepermanent form to the lessons of experience so that others can benefit bythem and be encouraged to undertake similar work.

How To Use This Book Chapter, Theorems, etc. are not numbered in this book,but are instead referenced by acronyms. This means that Theorem XYZ will always beTheorem XYZ, no matter if new sections are added, or if an individual decides to removecertain other sections. Within sections, the subsections and examples are acronymsthat begin with the acronym of the section. So Example XYZ.AB will be found withinSection XYZ. At first, all the letters flying around may be confusing, but with time, youwill begin to recognize the more important ones on sight. Furthermore, there are listsof theorems, examples, etc. in the front of the book, and an index that contains everyacronym. If you are reading this in an electronic PDF version, you will see that all ofthe cross-references are hyperlinks, allowing you to click to a definition or example, andthen use the back button to return. In printed versions, you will have to rely on thepage numbers. However, note that page numbers are not permanent! Different editions,different margins, or different sized paper will affect what content is on each page. Andin time, the addition of new material will affect the page numbering.

Robert A. Beezer

Page 5: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

iv

Tacoma, WashingtonAugust, 2004

Page 6: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Contents

Preface i

Contents v

Contributors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi

Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Theorems and Corollaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

Proof Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

Computation Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

Part C Core 2

Chapter SLE Systems of Linear Equations 2

WILA What is Linear Algebra? . . . . . . . . . . . . . . . . . . . . . . . . . . 2

LA “Linear” + “Algebra” . . . . . . . . . . . . . . . . . . . . . . . . . . 2

A An application: packaging trail mix . . . . . . . . . . . . . . . . . . . . 3

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

SSSLE Solving Systems of Simultaneous Linear Equations . . . . . . . . . . . 9

PSS Possibilities for solution sets . . . . . . . . . . . . . . . . . . . . . . 11

ESEO Equivalent systems and equation operations . . . . . . . . . . . . 12

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

RREF Reduced Row-Echelon Form . . . . . . . . . . . . . . . . . . . . . . . . 22

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

TSS Types of Solution Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

HSE Homogeneous Systems of Equations . . . . . . . . . . . . . . . . . . . . . 49

SHS Solutions of Homogeneous Systems . . . . . . . . . . . . . . . . . . 49

MVNSE Matrix and Vector Notation for Systems of Equations . . . . . . 52

NSM Null Space of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . 54

NSM NonSingular Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

NSM NonSingular Matrices . . . . . . . . . . . . . . . . . . . . . . . . . 56

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

v

Page 7: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Contents vi

Chapter V Vectors 65VO Vector Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

VEASM Vector equality, addition, scalar multiplication . . . . . . . . . . 66VSP Vector Space Properties . . . . . . . . . . . . . . . . . . . . . . . . 70

LC Linear Combinations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72LC Linear Combinations . . . . . . . . . . . . . . . . . . . . . . . . . . . 72EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

SS Spanning Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80VFSS Vector Form of Solution Sets . . . . . . . . . . . . . . . . . . . . . 80SSV Span of a Set of Vectors . . . . . . . . . . . . . . . . . . . . . . . . . 86SSNS Spanning Sets of Null Spaces . . . . . . . . . . . . . . . . . . . . . 90EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

LI Linear Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96LIV Linearly Independent Vectors . . . . . . . . . . . . . . . . . . . . . . 96LDSS Linearly Dependent Sets and Spans . . . . . . . . . . . . . . . . . 100LINSM Linear Independence and NonSingular Matrices . . . . . . . . . . 102NSSLI Null Spaces, Spans, Linear Independence . . . . . . . . . . . . . . 104EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

O Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109CAV Complex arithmetic and vectors . . . . . . . . . . . . . . . . . . . . 109IP Inner products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110N Norm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112OV Orthogonal Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114GSP Gram-Schmidt Procedure . . . . . . . . . . . . . . . . . . . . . . . . 116

Chapter M Matrices 121MO Matrix Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

MEASM Matrix equality, addition, scalar multiplication . . . . . . . . . 121VSP Vector Space Properties . . . . . . . . . . . . . . . . . . . . . . . . 123TSM Transposes and Symmetric Matrices . . . . . . . . . . . . . . . . . 124MCC Matrices and Complex Conjugation . . . . . . . . . . . . . . . . . 126

RM Range of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127RSE Range and systems of equations . . . . . . . . . . . . . . . . . . . . 127RSOC Range spanned by original columns . . . . . . . . . . . . . . . . . 128RNS The range as null space . . . . . . . . . . . . . . . . . . . . . . . . . 134RNSM Range of a Nonsingular Matrix . . . . . . . . . . . . . . . . . . . 138

RSM Row Space Of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . 141RSM Row Space of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . 141EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

MM Matrix Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151MVP Matrix-Vector Product . . . . . . . . . . . . . . . . . . . . . . . . . 151MM Matrix Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . 153MMEE Matrix Multiplication, Entry-by-Entry . . . . . . . . . . . . . . . 154PMM Properties of Matrix Multiplication . . . . . . . . . . . . . . . . . 156

Page 8: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Contents vii

PSHS Particular Solutions, Homogenous Solutions . . . . . . . . . . . . . 161

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

MISLE Matrix Inverses and Systems of Linear Equations . . . . . . . . . . . . 166

IM Inverse of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

CIM Computing the Inverse of a Matrix . . . . . . . . . . . . . . . . . . 169

PMI Properties of Matrix Inverses . . . . . . . . . . . . . . . . . . . . . . 173

MINSM Matrix Inverses and NonSingular Matrices . . . . . . . . . . . . . . . 176

NSMI NonSingular Matrices are Invertible . . . . . . . . . . . . . . . . . 176

OM Orthogonal Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

Chapter VS Vector Spaces 184

VS Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

VS Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

EVS Examples of Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . 185

VSP Vector Space Properties . . . . . . . . . . . . . . . . . . . . . . . . 190

RD Recycling Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . 195

S Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

TS Testing Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197

TSS The Span of a Set . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

SC Subspace Constructions . . . . . . . . . . . . . . . . . . . . . . . . . 207

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210

B Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

LI Linear independence . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

SS Spanning Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216

B Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219

BRS Bases from Row Spaces . . . . . . . . . . . . . . . . . . . . . . . . . 223

BNSM Bases and NonSingular Matrices . . . . . . . . . . . . . . . . . . . 225

VR Vector Representation . . . . . . . . . . . . . . . . . . . . . . . . . . 226

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

D Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230

D Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230

DVS Dimension of Vector Spaces . . . . . . . . . . . . . . . . . . . . . . 235

RNM Rank and Nullity of a Matrix . . . . . . . . . . . . . . . . . . . . . 237

RNNSM Rank and Nullity of a NonSingular Matrix . . . . . . . . . . . . 238

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242

PD Properties of Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243

GT Goldilocks’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 243

RT Ranks and Transposes . . . . . . . . . . . . . . . . . . . . . . . . . . 246

OBC Orthonormal Bases and Coordinates . . . . . . . . . . . . . . . . . 248

EXC Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252

SOL Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

Page 9: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Contents viii

Chapter D Determinants 254

DM Determinants of Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . 254

CD Computing Determinants . . . . . . . . . . . . . . . . . . . . . . . . 256

PD Properties of Determinants . . . . . . . . . . . . . . . . . . . . . . . 258

Chapter E Eigenvalues 261

EE Eigenvalues and Eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . 261

EEM Eigenvalues and Eigenvectors of a Matrix . . . . . . . . . . . . . . 261

PM Polynomials and Matrices . . . . . . . . . . . . . . . . . . . . . . . . 263

EEE Existence of Eigenvalues and Eigenvectors . . . . . . . . . . . . . . 265

CEE Computing Eigenvalues and Eigenvectors . . . . . . . . . . . . . . . 268

ECEE Examples of Computing Eigenvalues and Eigenvectors . . . . . . . 272

PEE Properties of Eigenvalues and Eigenvectors . . . . . . . . . . . . . . . . . 280

ME Multiplicities of Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . 286

EHM Eigenvalues of Hermitian Matrices . . . . . . . . . . . . . . . . . . 288

SD Similarity and Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . 291

SM Similar Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291

PSM Properties of Similar Matrices . . . . . . . . . . . . . . . . . . . . . 292

D Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

OD Orthonormal Diagonalization . . . . . . . . . . . . . . . . . . . . . . 301

Chapter LT Linear Transformations 302

LT Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302

LT Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . 302

MLT Matrices and Linear Transformations . . . . . . . . . . . . . . . . . 307

LTLC Linear Transformations and Linear Combinations . . . . . . . . . 311

PI Pre-Images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314

NLTFO New Linear Transformations From Old . . . . . . . . . . . . . . 316

ILT Injective Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . 321

EILT Examples of Injective Linear Transformations . . . . . . . . . . . . 321

NSLT Null Space of a Linear Transformation . . . . . . . . . . . . . . . . 324

ILTD Injective Linear Transformations and Dimension . . . . . . . . . . 330

CILT Composition of Injective Linear Transformations . . . . . . . . . . 331

SLT Surjective Linear Transformations . . . . . . . . . . . . . . . . . . . . . . 332

ESLT Examples of Surjective Linear Transformations . . . . . . . . . . . 332

RLT Range of a Linear Transformation . . . . . . . . . . . . . . . . . . . 336

SLTD Surjective Linear Transformations and Dimension . . . . . . . . . 342

CSLT Composition of Surjective Linear Transformations . . . . . . . . . 343

IVLT Invertible Linear Transformations . . . . . . . . . . . . . . . . . . . . . 344

IVLT Invertible Linear Transformations . . . . . . . . . . . . . . . . . . . 344

IV Invertibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348

SI Structure and Isomorphism . . . . . . . . . . . . . . . . . . . . . . . . 350

RNLT Rank and Nullity of a Linear Transformation . . . . . . . . . . . . 352

SLELT Systems of Linear Equations and Linear Transformations . . . . . 355

Page 10: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Contents ix

Chapter R Representations 358VR Vector Representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358

CVS Characterization of Vector Spaces . . . . . . . . . . . . . . . . . . . 364MR Matrix Representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369

NRFO New Representations from Old . . . . . . . . . . . . . . . . . . . 370PMR Properties of Matrix Representations . . . . . . . . . . . . . . . . . 372

Chapter A Archetypes 375A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 384C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393E . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397F . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401G . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406H . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 414J . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 418K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423L . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427M . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431N . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433O . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435P . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437Q . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444U . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444V . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444

Part T Topics 447

Chapter P Preliminaries 447CNO Complex Number Operations . . . . . . . . . . . . . . . . . . . . . . . . 447

CNA Arithmetic with complex numbers . . . . . . . . . . . . . . . . . . 447CCN Conjugates of Complex Numbers . . . . . . . . . . . . . . . . . . . 448MCN Modulus of a Complex Number . . . . . . . . . . . . . . . . . . . . 448

Part A Applications 451

Page 11: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Contributors

Beezer, Robert. University of Puget Sound. http://buzzard.ups.edu/

Riegsecker, Joe. Middlebury, Indiana. joepye(at)pobox(dot)com

Phelps, Douglas. University of Puget Sound. \

x

Page 12: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Definitions

Section WILASection SSSLESSLE System of Simultaneous Linear Equations . . . . . . . . . . . . . . . . 10ES Equivalent Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12EO Equation Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

Section RREFM Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22MN Matrix Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22AM Augmented Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23RO Row Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24REM Row-Equivalent Matrices . . . . . . . . . . . . . . . . . . . . . . . . . 25RREF Reduced Row-Echelon Form . . . . . . . . . . . . . . . . . . . . . . . 27ZRM Zero Row of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . 27LO Leading Ones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

Section TSSCS Consistent System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37IDV Independent and Dependent Variables . . . . . . . . . . . . . . . . . . 40

Section HSEHS Homogeneous System . . . . . . . . . . . . . . . . . . . . . . . . . . . 49TSHSE Trivial Solution to Homogeneous Systems of Equations . . . . . . . . 50CV Column Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52ZV Zero Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52CM Coefficient Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52VOC Vector of Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53SV Solution Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53NSM Null Space of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 54

Section NSMSQM Square Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56NM Nonsingular Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56IM Identity Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

Section VOVSCM Vector Space Cm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65CVE Column Vector Equality . . . . . . . . . . . . . . . . . . . . . . . . . 66CVA Column Vector Addition . . . . . . . . . . . . . . . . . . . . . . . . . 67CVSM Column Vector Scalar Multiplication . . . . . . . . . . . . . . . . . . 68

Section LCLCCV Linear Combination of Column Vectors . . . . . . . . . . . . . . . . . 72

Section SS

xi

Page 13: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Definitions xii

SSCV Span of a Set of Column Vectors . . . . . . . . . . . . . . . . . . . . . 87

Section LIRLDCV Relation of Linear Dependence for Column Vectors . . . . . . . . . . 96LICV Linear Independence of Column Vectors . . . . . . . . . . . . . . . . . 96

Section OCCV Conjugate of a Column Vector . . . . . . . . . . . . . . . . . . . . . . 109IP Inner Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110NV Norm of a Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113OV Orthogonal Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114OSV Orthogonal Set of Vectors . . . . . . . . . . . . . . . . . . . . . . . . 115ONS OrthoNormal Set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

Section MOVSM Vector Space of m× n Matrices . . . . . . . . . . . . . . . . . . . . . 121ME Matrix Equality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121MA Matrix Addition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122SMM Scalar Matrix Multiplication . . . . . . . . . . . . . . . . . . . . . . . 122ZM Zero Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124TM Transpose of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 124SYM Symmetric Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125CCM Complex Conjugate of a Matrix . . . . . . . . . . . . . . . . . . . . . 126

Section RMRM Range of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

Section RSMRSM Row Space of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 141

Section MMMVP Matrix-Vector Product . . . . . . . . . . . . . . . . . . . . . . . . . . 151MM Matrix Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

Section MISLEMI Matrix Inverse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167SUV Standard Unit Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 169

Section MINSMOM Orthogonal Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . 178A Adjoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181HM Hermitian Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

Section VSVS Vector Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

Section SS Subspace . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

Page 14: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Definitions xiii

TS Trivial Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200LC Linear Combination . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202SS Span of a Set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203

Section BRLD Relation of Linear Dependence . . . . . . . . . . . . . . . . . . . . . . 212LI Linear Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . 212TSS To Span a Subspace . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216B Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220

Section DD Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230NOM Nullity Of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237ROM Rank Of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237

Section PDSection DMSM SubMatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254DM Determinant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254MIM Minor In a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256CIM Cofactor In a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . 256

Section EEEEM Eigenvalues and Eigenvectors of a Matrix . . . . . . . . . . . . . . . . 261CP Characteristic Polynomial . . . . . . . . . . . . . . . . . . . . . . . . 268EM Eigenspace of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 270AME Algebraic Multiplicity of an Eigenvalue . . . . . . . . . . . . . . . . . 272GME Geometric Multiplicity of an Eigenvalue . . . . . . . . . . . . . . . . . 272

Section PEESection SDSIM Similar Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291DIM Diagonal Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295DZM Diagonalizable Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 295

Section LTLT Linear Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . 302PI Pre-Image . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314LTA Linear Transformation Addition . . . . . . . . . . . . . . . . . . . . . 316LTSM Linear Transformation Scalar Multiplication . . . . . . . . . . . . . . 317LTC Linear Transformation Composition . . . . . . . . . . . . . . . . . . . 319

Section ILTILT Injective Linear Transformation . . . . . . . . . . . . . . . . . . . . . 321NSLT Null Space of a Linear Transformation . . . . . . . . . . . . . . . . . 325

Section SLTSLT Surjective Linear Transformation . . . . . . . . . . . . . . . . . . . . 332

Page 15: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Definitions xiv

RLT Range of a Linear Transformation . . . . . . . . . . . . . . . . . . . . 336

Section IVLTIDLT Identity Linear Transformation . . . . . . . . . . . . . . . . . . . . . 344IVLT Invertible Linear Transformations . . . . . . . . . . . . . . . . . . . . 344IVS Isomorphic Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . 350ROLT Rank Of a Linear Transformation . . . . . . . . . . . . . . . . . . . . 352NOLT Nullity Of a Linear Transformation . . . . . . . . . . . . . . . . . . . 352

Section VRVR Vector Representation . . . . . . . . . . . . . . . . . . . . . . . . . . 358

Section MRMR Matrix Representation . . . . . . . . . . . . . . . . . . . . . . . . . . 369

Section CNOCCN Conjugate of a Complex Number . . . . . . . . . . . . . . . . . . . . 448MCN Modulus of a Complex Number . . . . . . . . . . . . . . . . . . . . . 448

Page 16: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Theorems and Corollaries

Section WILASection SSSLE

EOPSS Equation Operations Preserve Solution Sets . . . . . . . . . . . . . . 13

Section RREF

REMES Row-Equivalent Matrices represent Equivalent Systems . . . . . . . . 25

REMEF Row-Equivalent Matrix in Echelon Form . . . . . . . . . . . . . . . . 28

Section TSS

RCLS Recognizing Consistency of a Linear System . . . . . . . . . . . . . . 41

ICRN Inconsistent Systems, r and n . . . . . . . . . . . . . . . . . . . . . . 42

CSRN Consistent Systems, r and n . . . . . . . . . . . . . . . . . . . . . . . 42

FVCS Free Variables for Consistent Systems . . . . . . . . . . . . . . . . . . 42

PSSLS Possible Solution Sets for Linear Systems . . . . . . . . . . . . . . . . 44

CMVEI Consistent, More Variables than Equations, Infinite solutions . . . . . 44

Section HSE

HSC Homogeneous Systems are Consistent . . . . . . . . . . . . . . . . . . 50

HMVEI Homogeneous, More Variables than Equations, Infinite solutions . . . 51

Section NSM

NSRRI NonSingular matrices Row Reduce to the Identity matrix . . . . . . . 57

NSTNS NonSingular matrices have Trivial Null Spaces . . . . . . . . . . . . . 59

NSMUS NonSingular Matrices and Unique Solutions . . . . . . . . . . . . . . 59

NSME1 NonSingular Matrix Equivalences, Round 1 . . . . . . . . . . . . . . . 62

Section VO

VSPCM Vector Space Properties of Cm . . . . . . . . . . . . . . . . . . . . . . 70

Section LC

SLSLC Solutions to Linear Systems are Linear Combinations . . . . . . . . . 76

Section SS

VFSLS Vector Form of Solutions to Linear Systems . . . . . . . . . . . . . . 82

SSNS Spanning Sets for Null Spaces . . . . . . . . . . . . . . . . . . . . . . 90

Section LI

LIVHS Linearly Independent Vectors and Homogeneous Systems . . . . . . . 98

MVSLD More Vectors than Size implies Linear Dependence . . . . . . . . . . . 99

LIVRN Linearly Independent Vectors, r and n . . . . . . . . . . . . . . . . . 100

DLDS Dependency in Linearly Dependent Sets . . . . . . . . . . . . . . . . 100

NSLIC NonSingular matrices have Linearly Independent Columns . . . . . . 103

NSME2 NonSingular Matrix Equivalences, Round 2 . . . . . . . . . . . . . . . 103

BNS Basis for Null Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

xv

Page 17: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Theorems and Corollaries xvi

Section OCCRVA Complex Conjugation Respects Vector Addition . . . . . . . . . . . . 109CCRSM Complex Conjugation Respects Scalar Multiplication . . . . . . . . . 110IPVA Inner Product and Vector Addition . . . . . . . . . . . . . . . . . . . 111IPSM Inner Product and Scalar Multiplication . . . . . . . . . . . . . . . . 112IPAC Inner Product is Anti-Commutative . . . . . . . . . . . . . . . . . . . 112IPN Inner Products and Norms . . . . . . . . . . . . . . . . . . . . . . . . 113PIP Positive Inner Products . . . . . . . . . . . . . . . . . . . . . . . . . . 114OSLI Orthogonal Sets are Linearly Independent . . . . . . . . . . . . . . . 116GSPCV Gram-Schmidt Procedure, Column Vectors . . . . . . . . . . . . . . . 117

Section MOVSPM Vector Space Properties of Mmn . . . . . . . . . . . . . . . . . . . . . 123SMS Symmetric Matrices are Square . . . . . . . . . . . . . . . . . . . . . 126TASM Transposes, Addition, Scalar Multiplication . . . . . . . . . . . . . . . 126

Section RMRCS Range and Consistent Systems . . . . . . . . . . . . . . . . . . . . . . 128BROC Basis of the Range with Original Columns . . . . . . . . . . . . . . . 131RNS Range as a Null Space . . . . . . . . . . . . . . . . . . . . . . . . . . 136RNSM Range of a NonSingular Matrix . . . . . . . . . . . . . . . . . . . . . 139NSME3 NonSingular Matrix Equivalences, Round 3 . . . . . . . . . . . . . . . 140

Section RSMREMRS Row-Equivalent Matrices have equal Row Spaces . . . . . . . . . . . . 142BRS Basis for the Row Space . . . . . . . . . . . . . . . . . . . . . . . . . 144RMRST Range of a Matrix is Row Space of Transpose . . . . . . . . . . . . . 145

Section MMSLEMM Systems of Linear Equations as Matrix Multiplication . . . . . . . . . 152EMP Entries of Matrix Products . . . . . . . . . . . . . . . . . . . . . . . . 154MMZM Matrix Multiplication and the Zero Matrix . . . . . . . . . . . . . . . 156MMIM Matrix Multiplication and Identity Matrix . . . . . . . . . . . . . . . 156MMDAA Matrix Multiplication Distributes Across Addition . . . . . . . . . . . 157MMSMM Matrix Multiplication and Scalar Matrix Multiplication . . . . . . . . 157MMA Matrix Multiplication is Associative . . . . . . . . . . . . . . . . . . 158MMIP Matrix Multiplication and Inner Products . . . . . . . . . . . . . . . 159MMCC Matrix Multiplication and Complex Conjugation . . . . . . . . . . . . 159MMT Matrix Multiplication and Transposes . . . . . . . . . . . . . . . . . . 160PSPHS Particular Solution Plus Homogenous Solutions . . . . . . . . . . . . 161

Section MISLETTMI Two-by-Two Matrix Inverse . . . . . . . . . . . . . . . . . . . . . . . 169CINSM Computing the Inverse of a NonSingular Matrix . . . . . . . . . . . . 172MIU Matrix Inverse is Unique . . . . . . . . . . . . . . . . . . . . . . . . . 173SS Socks and Shoes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174MIMI Matrix Inverse of a Matrix Inverse . . . . . . . . . . . . . . . . . . . . 174

Page 18: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Theorems and Corollaries xvii

MIT Matrix Inverse of a Transpose . . . . . . . . . . . . . . . . . . . . . . 174

MISM Matrix Inverse of a Scalar Multiple . . . . . . . . . . . . . . . . . . . 175

Section MINSM

PWSMS Product With a Singular Matrix is Singular . . . . . . . . . . . . . . 176

OSIS One-Sided Inverse is Sufficient . . . . . . . . . . . . . . . . . . . . . . 177

NSI NonSingularity is Invertibility . . . . . . . . . . . . . . . . . . . . . . 177

NSME4 NonSingular Matrix Equivalences, Round 4 . . . . . . . . . . . . . . . 178

SNSCM Solution with NonSingular Coefficient Matrix . . . . . . . . . . . . . . 178

OMI Orthogonal Matrices are Invertible . . . . . . . . . . . . . . . . . . . 179

COMOS Columns of Orthogonal Matrices are Orthonormal Sets . . . . . . . . 179

OMPIP Orthogonal Matrices Preserve Inner Products . . . . . . . . . . . . . 180

Section VS

ZVU Zero Vector is Unique . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

AIU Additive Inverses are Unique . . . . . . . . . . . . . . . . . . . . . . . 191

ZSSM Zero Scalar in Scalar Multiplication . . . . . . . . . . . . . . . . . . . 191

ZVSM Zero Vector in Scalar Multiplication . . . . . . . . . . . . . . . . . . . 192

AISM Additive Inverses from Scalar Multiplication . . . . . . . . . . . . . . 192

SMEZV Scalar Multiplication Equals the Zero Vector . . . . . . . . . . . . . . 193

VAC Vector Addition Cancellation . . . . . . . . . . . . . . . . . . . . . . . 194

CSSM Canceling Scalars in Scalar Multiplication . . . . . . . . . . . . . . . 194

CVSM Canceling Vectors in Scalar Multiplication . . . . . . . . . . . . . . . 194

Section S

TSS Testing Subsets for Subspaces . . . . . . . . . . . . . . . . . . . . . . 198

NSMS Null Space of a Matrix is a Subspace . . . . . . . . . . . . . . . . . . 200

SSS Span of a Set is a Subspace . . . . . . . . . . . . . . . . . . . . . . . . 203

RMS Range of a Matrix is a Subspace . . . . . . . . . . . . . . . . . . . . . 207

RSMS Row Space of a Matrix is a Subspace . . . . . . . . . . . . . . . . . . 208

Section B

SUVB Standard Unit Vectors are a Basis . . . . . . . . . . . . . . . . . . . . 220

CNSMB Columns of NonSingular Matrix are a Basis . . . . . . . . . . . . . . 225

NSME5 NonSingular Matrix Equivalences, Round 5 . . . . . . . . . . . . . . . 226

VRRB Vector Representation Relative to a Basis . . . . . . . . . . . . . . . . 227

Section D

SSLD Spanning Sets and Linear Dependence . . . . . . . . . . . . . . . . . 230

BIS Bases have Identical Sizes . . . . . . . . . . . . . . . . . . . . . . . . 234

DCM Dimension of Cm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

DP Dimension of Pn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

DM Dimension of Mmn . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

CRN Computing Rank and Nullity . . . . . . . . . . . . . . . . . . . . . . 238

RPNC Rank Plus Nullity is Columns . . . . . . . . . . . . . . . . . . . . . . 238

RNNSM Rank and Nullity of a NonSingular Matrix . . . . . . . . . . . . . . . 239

NSME6 NonSingular Matrix Equivalences, Round 6 . . . . . . . . . . . . . . . 239

Page 19: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Theorems and Corollaries xviii

Section PDELIS Extending Linearly Independent Sets . . . . . . . . . . . . . . . . . . 243G Goldilocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244RMRT Rank of a Matrix is the Rank of the Transpose . . . . . . . . . . . . . 247COB Coordinates and Orthonormal Bases . . . . . . . . . . . . . . . . . . . 249

Section DMDMST Determinant of Matrices of Size Two . . . . . . . . . . . . . . . . . . 255DERC Determinant Expansion about Rows and Columns . . . . . . . . . . . 257DT Determinant of the Transpose . . . . . . . . . . . . . . . . . . . . . . 258DRMM Determinant Respects Matrix Multiplication . . . . . . . . . . . . . . 259SMZD Singular Matrices have Zero Determinants . . . . . . . . . . . . . . . 259NSME7 NonSingular Matrix Equivalences, Round 7 . . . . . . . . . . . . . . . 259

Section EEEMHE Every Matrix Has an Eigenvalue . . . . . . . . . . . . . . . . . . . . . 265EMRCP Eigenvalues of a Matrix are Roots of Characteristic Polynomials . . . 269EMS Eigenspace for a Matrix is a Subspace . . . . . . . . . . . . . . . . . . 270EMNS Eigenspace of a Matrix is a Null Space . . . . . . . . . . . . . . . . . 271

Section PEEEDELI Eigenvectors with Distinct Eigenvalues are Linearly Independent . . . 280SMZE Singular Matrices have Zero Eigenvalues . . . . . . . . . . . . . . . . 281NSME8 NonSingular Matrix Equivalences, Round 8 . . . . . . . . . . . . . . . 281ESMM Eigenvalues of a Scalar Multiple of a Matrix . . . . . . . . . . . . . . 282EOMP Eigenvalues Of Matrix Powers . . . . . . . . . . . . . . . . . . . . . . 282EPM Eigenvalues of the Polynomial of a Matrix . . . . . . . . . . . . . . . 283EIM Eigenvalues of the Inverse of a Matrix . . . . . . . . . . . . . . . . . . 284ETM Eigenvalues of the Transpose of a Matrix . . . . . . . . . . . . . . . . 285ERMCP Eigenvalues of Real Matrices come in Conjugate Pairs . . . . . . . . . 285DCP Degree of the Characteristic Polynomial . . . . . . . . . . . . . . . . . 286NEM Number of Eigenvalues of a Matrix . . . . . . . . . . . . . . . . . . . 286ME Multiplicities of an Eigenvalue . . . . . . . . . . . . . . . . . . . . . . 286MNEM Maximum Number of Eigenvalues of a Matrix . . . . . . . . . . . . . 288HMRE Hermitian Matrices have Real Eigenvalues . . . . . . . . . . . . . . . 288HMOE Hermitian Matrices have Orthogonal Eigenvectors . . . . . . . . . . . 289

Section SDSER Similarity is an Equivalence Relation . . . . . . . . . . . . . . . . . . 293SMEE Similar Matrices have Equal Eigenvalues . . . . . . . . . . . . . . . . 294DC Diagonalization Characterization . . . . . . . . . . . . . . . . . . . . 295DMLE Diagonalizable Matrices have Large Eigenspaces . . . . . . . . . . . . 298DED Distinct Eigenvalues implies Diagonalizable . . . . . . . . . . . . . . . 300ODHM Orthonormal Diagonalization of Hermitian Matrices . . . . . . . . . . 301

Section LTLTTZZ Linear Transformations Take Zero to Zero . . . . . . . . . . . . . . . 306

Page 20: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Theorems and Corollaries xix

MBLT Matrices Build Linear Transformations . . . . . . . . . . . . . . . . . 308

MLTCV Matrix of a Linear Transformation, Column Vectors . . . . . . . . . . 309

LTLC Linear Transformations and Linear Combinations . . . . . . . . . . . 311

LTDB Linear Transformation Defined on a Basis . . . . . . . . . . . . . . . . 312

SLTLT Sum of Linear Transformations is a Linear Transformation . . . . . . 316

MLTLT Multiple of a Linear Transformation is a Linear Transformation . . . 317

VSLT Vector Space of Linear Transformations . . . . . . . . . . . . . . . . . 318

CLTLT Composition of Linear Transformations is a Linear Transformation . . 319

Section ILT

NSLTS Null Space of a Linear Transformation is a Subspace . . . . . . . . . . 326

NSPI Null Space and Pre-Image . . . . . . . . . . . . . . . . . . . . . . . . 327

NSILT Null Space of an Injective Linear Transformation . . . . . . . . . . . . 328

ILTLI Injective Linear Transformations and Linear Independence . . . . . . 329

ILTB Injective Linear Transformations and Bases . . . . . . . . . . . . . . . 330

ILTD Injective Linear Transformations and Dimension . . . . . . . . . . . . 330

CILTI Composition of Injective Linear Transformations is Injective . . . . . 331

Section SLT

RLTS Range of a Linear Transformation is a Subspace . . . . . . . . . . . . 338

RSLT Range of a Surjective Linear Transformation . . . . . . . . . . . . . . 339

SLTS Surjective Linear Transformations and Spans . . . . . . . . . . . . . . 341

RPI Range and Pre-Image . . . . . . . . . . . . . . . . . . . . . . . . . . . 341

SLTB Surjective Linear Transformations and Bases . . . . . . . . . . . . . . 342

SLTD Surjective Linear Transformations and Dimension . . . . . . . . . . . 342

CSLTS Composition of Surjective Linear Transformations is Surjective . . . . 343

Section IVLT

ILTLT Invese of a Linear Transformation is a Linear Transformation . . . . . 347

IILT Inverse of an Invertible Linear Transformation . . . . . . . . . . . . . 347

ILTIS Invertible Linear Transformations are Injective and Surjective . . . . 348

CIVLT Composition of Invertible Linear Transformations . . . . . . . . . . . 349

ICLT Inverse of a Composition of Linear Transformations . . . . . . . . . . 349

IVSED Isomorphic Vector Spaces have Equal Dimension . . . . . . . . . . . . 351

ROSLT Rank Of a Surjective Linear Transformation . . . . . . . . . . . . . . 352

NOILT Nullity Of an Injective Linear Transformation . . . . . . . . . . . . . 353

RPNDD Rank Plus Nullity is Domain Dimension . . . . . . . . . . . . . . . . 353

Section VR

VRLT Vector Representation is a Linear Transformation . . . . . . . . . . . 358

VRI Vector Representation is Injective . . . . . . . . . . . . . . . . . . . . 363

VRS Vector Representation is Surjective . . . . . . . . . . . . . . . . . . . 363

VRILT Vector Representation is an Invertible Linear Transformation . . . . . 364

CFDVS Characterization of Finite Dimensional Vector Spaces . . . . . . . . . 364

IFDVS Isomorphism of Finite Dimensional Vector Spaces . . . . . . . . . . . 365

CLI Coordinatization and Linear Independence . . . . . . . . . . . . . . . 365

CSS Coordinatization and Spanning Sets . . . . . . . . . . . . . . . . . . . 366

Page 21: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Theorems and Corollaries xx

Section MRFTMR Fundamental Theorem of Matrix Representation . . . . . . . . . . . . 369MRSLT Matrix Representation of a Sum of Linear Transformations . . . . . . 370MRMLT Matrix Representation of a Multiple of a Linear Transformation . . . 371MRCLT Matrix Representation of a Composition of Linear Transformations . 371INS Isomorphic Null Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 372IR Isomorphic Ranges . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373

Section CNOCCRA Complex Conjugation Respects Addition . . . . . . . . . . . . . . . . 448CCRM Complex Conjugation Respects Multiplication . . . . . . . . . . . . . 448

Page 22: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Notation

Section WILASection SSSLESection RREFSection TSSRREFA Reduced Row-Echelon Form Analysis . . . . . . . . . . . . . . . . . . 37

Section HSEVN Vector (u) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52ZVN Zero Vector (0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52AMN Augmented Matrix ([A | b]) . . . . . . . . . . . . . . . . . . . . . . . 54LSN Linear System (LS(A, b)) . . . . . . . . . . . . . . . . . . . . . . . . 54

Section NSMSection VOSection LCSection SSSection LISection OSection MOME Matrix Entries ([A]ij) . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

Section RMSection RSMSection MMSection MISLESection MINSMSection VSSection SSection BSection DSection PDSection DMSection EESection PEESection SDSection LTSection ILTSection SLTSection IVLTSection VRSection MRSection CNO

xxi

Page 23: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Examples

Section WILAWILA.TM Trail Mix Packaging . . . . . . . . . . . . . . . . . . . . . . . . 3

Section SSSLESSSLE.STNE Solving two (nonlinear) equations . . . . . . . . . . . . . . . . . 9SSSLE.NSE Notation for a system of equations . . . . . . . . . . . . . . . . 10SSSLE.TTS Three typical systems . . . . . . . . . . . . . . . . . . . . . . . . 11SSSLE.US Three equations, one solution . . . . . . . . . . . . . . . . . . . 16SSSLE.IS Three equations, infinitely many solutions . . . . . . . . . . . . 17

Section RREFRREF.AM A matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22RREF.AMAA Augmented matrix for Archetype A . . . . . . . . . . . . . . . . 24RREF.TREM Two row-equivalent matrices . . . . . . . . . . . . . . . . . . . . 25RREF.US Three equations, one solution . . . . . . . . . . . . . . . . . . . 26RREF.RREF A matrix in reduced row-echelon form . . . . . . . . . . . . . . 27RREF.NRREF A matrix not in reduced row-echelon form . . . . . . . . . . . . 27RREF.SAB Solutions for Archetype B . . . . . . . . . . . . . . . . . . . . . 29RREF.SAA Solutions for Archetype A . . . . . . . . . . . . . . . . . . . . . 30RREF.SAE Solutions for Archetype E . . . . . . . . . . . . . . . . . . . . . 31

Section TSSTSS.RREFN Reduced row-echelon form notation . . . . . . . . . . . . . . . . 37TSS.ISSI Describing infinite solution sets, Archetype I . . . . . . . . . . . 38TSS.CFV Counting free variables . . . . . . . . . . . . . . . . . . . . . . . 43TSS.OSGMD One solution gives many, Archetype D . . . . . . . . . . . . . . 44

Section HSEHSE.AHSAC Archetype C as a homogeneous system . . . . . . . . . . . . . . 49HSE.HUSAB Homogeneous, unique solution, Archetype B . . . . . . . . . . . 50HSE.HISAA Homogeneous, infinite solutions, Archetype A . . . . . . . . . . 50HSE.HISAD Homogeneous, infinite solutions, Archetype D . . . . . . . . . . 51HSE.NSLE Notation for systems of linear equations . . . . . . . . . . . . . 54HSE.NSEAI Null space elements of Archetype I . . . . . . . . . . . . . . . . 55

Section NSMNSM.S A singular matrix, Archetype A . . . . . . . . . . . . . . . . . . 56NSM.NS A nonsingular matrix, Archetype B . . . . . . . . . . . . . . . . 57NSM.IM An identity matrix . . . . . . . . . . . . . . . . . . . . . . . . . 57NSM.SRR Singular matrix, row-reduced . . . . . . . . . . . . . . . . . . . 58NSM.NSRR NonSingular matrix, row-reduced . . . . . . . . . . . . . . . . . 58NSM.NSS Null space of a singular matrix . . . . . . . . . . . . . . . . . . 58NSM.NSNS Null space of a nonsingular matrix . . . . . . . . . . . . . . . . 59

Section VO

xxii

Page 24: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Examples xxiii

VO.VESE Vector equality for a system of equations . . . . . . . . . . . . . 66VO.VA Addition of two vectors in C4 . . . . . . . . . . . . . . . . . . . 67VO.VSM Scalar multiplication in C5 . . . . . . . . . . . . . . . . . . . . . 68

Section LCLC.TLC Two linear combinations in C6 . . . . . . . . . . . . . . . . . . . 72LC.ABLC Archetype B as a linear combination . . . . . . . . . . . . . . . 74LC.AALC Archetype A as a linear combination . . . . . . . . . . . . . . . 75

Section SSSS.VFSAD Vector form of solutions for Archetype D . . . . . . . . . . . . . 80SS.VFSAI Vector form of solutions for Archetype I . . . . . . . . . . . . . 83SS.VFSAL Vector form of solutions for Archetype L . . . . . . . . . . . . . 85SS.SCAA Span of the columns of Archetype A . . . . . . . . . . . . . . . 87SS.SCAB Span of the columns of Archetype B . . . . . . . . . . . . . . . 89SS.SCAD Span of the columns of Archetype D . . . . . . . . . . . . . . . 91

Section LILI.LDS Linearly dependent set in C5 . . . . . . . . . . . . . . . . . . . . 96LI.LIS Linearly independent set in C5 . . . . . . . . . . . . . . . . . . . 97LI.LLDS Large linearly dependent set in C4 . . . . . . . . . . . . . . . . 99LI.RS Reducing a span in C5 . . . . . . . . . . . . . . . . . . . . . . . 101LI.LDCAA Linearly dependent columns in Archetype A . . . . . . . . . . . 103LI.LICAB Linearly independent columns in Archetype B . . . . . . . . . . 103LI.NSLIL Null space spanned by linearly independent set, Archetype L . . 105

Section OO.CSIP Computing some inner products . . . . . . . . . . . . . . . . . . 110O.CNSV Computing the norm of some vectors . . . . . . . . . . . . . . . 113O.TOV Two orthogonal vectors . . . . . . . . . . . . . . . . . . . . . . . 114O.SUVOS Standard Unit Vectors are an Orthogonal Set . . . . . . . . . . 115O.AOS An orthogonal set . . . . . . . . . . . . . . . . . . . . . . . . . . 115O.GSTV Gram-Schmidt of three vectors . . . . . . . . . . . . . . . . . . 118O.ONTV Orthonormal set, three vectors . . . . . . . . . . . . . . . . . . . 119O.ONFV Orthonormal set, four vectors . . . . . . . . . . . . . . . . . . . 120

Section MOMO.MA Addition of two matrices in M23 . . . . . . . . . . . . . . . . . . 122MO.MSM Scalar multiplication in M32 . . . . . . . . . . . . . . . . . . . . 122MO.TM Transpose of a 3× 4 matrix . . . . . . . . . . . . . . . . . . . . 124MO.SYM A symmetric 5× 5 matrix . . . . . . . . . . . . . . . . . . . . . 125

Section RMRM.RMCS Range of a matrix and consistent systems . . . . . . . . . . . . 127RM.COC Casting out columns, Archetype I . . . . . . . . . . . . . . . . . 128RM.ROCD Range with original columns, Archetype D . . . . . . . . . . . . 133RM.RNSAD Range as null space, Archetype D . . . . . . . . . . . . . . . . . 134RM.RNSAG Range as null space, Archetype G . . . . . . . . . . . . . . . . . 137

Page 25: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Examples xxiv

RM.RAA Range of Archetype A . . . . . . . . . . . . . . . . . . . . . . . 139RM.RAB Range of Archetype B . . . . . . . . . . . . . . . . . . . . . . . 139

Section RSMRSM.RSAI Row space of Archetype I . . . . . . . . . . . . . . . . . . . . . 141RSM.RSREM Row spaces of two row-equivalent matrices . . . . . . . . . . . . 144RSM.IS Improving a span . . . . . . . . . . . . . . . . . . . . . . . . . . 145RSM.RROI Range from row operations, Archetype I . . . . . . . . . . . . . 146

Section MMMM.MTV A matrix times a vector . . . . . . . . . . . . . . . . . . . . . . 151MM.NSLE Notation for systems of linear equations . . . . . . . . . . . . . 152MM.PTM Product of two matrices . . . . . . . . . . . . . . . . . . . . . . 153MM.MMNC Matrix Multiplication is not commutative . . . . . . . . . . . . 154MM.PTMEE Product of two matrices, entry-by-entry . . . . . . . . . . . . . 155MM.PSNS Particular solutions, homogenous solutions, Archetype D . . . . 161

Section MISLEMISLE.SABMI Solutions to Archetype B with a matrix inverse . . . . . . . . . 166MISLE.MWIAA A matrix without an inverse, Archetype A . . . . . . . . . . . . 167MISLE.MIAK Matrix Inverse, Archetype K . . . . . . . . . . . . . . . . . . . . 168MISLE.CMIAK Computing a Matrix Inverse, Archetype K . . . . . . . . . . . . 170MISLE.CMIAB Computing a Matrix Inverse, Archetype B . . . . . . . . . . . . 173

Section MINSMMINSM.OM3 Orthogonal matrix of size 3 . . . . . . . . . . . . . . . . . . . . 178MINSM.OPM Orthogonal permutation matrix . . . . . . . . . . . . . . . . . . 179MINSM.OSMC Orthonormal Set from Matrix Columns . . . . . . . . . . . . . . 180

Section VSVS.VSCM The vector space Cm . . . . . . . . . . . . . . . . . . . . . . . . 186VS.VSM The vector space of matrices, Mmn . . . . . . . . . . . . . . . . 186VS.VSP The vector space of polynomials, Pn . . . . . . . . . . . . . . . . 186VS.VSIS The vector space of infinite sequences . . . . . . . . . . . . . . . 187VS.VSF The vector space of functions . . . . . . . . . . . . . . . . . . . 187VS.VSS The singleton vector space . . . . . . . . . . . . . . . . . . . . 188VS.CVS The crazy vector space . . . . . . . . . . . . . . . . . . . . . . 188VS.PCVS Properties for the Crazy Vector Space . . . . . . . . . . . . . . 193

Section SS.SC3 A subspace of C3 . . . . . . . . . . . . . . . . . . . . . . . . . . 196S.SP4 A subspace of P4 . . . . . . . . . . . . . . . . . . . . . . . . . . 198S.NSC2Z A non-subspace in C2, zero vector . . . . . . . . . . . . . . . . . 199S.NSC2A A non-subspace in C2, additive closure . . . . . . . . . . . . . . 200S.NSC2S A non-subspace in C2, scalar multiplication closure . . . . . . . 200S.RSNS Recasting a subspace as a null space . . . . . . . . . . . . . . . 201S.LCM A linear combination of matrices . . . . . . . . . . . . . . . . . 202S.SSP Span of a set of polynomials . . . . . . . . . . . . . . . . . . . . 204

Page 26: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Examples xxv

S.SM32 A subspace of M32 . . . . . . . . . . . . . . . . . . . . . . . . . 205

Section BB.LIP4 Linear independence in P4 . . . . . . . . . . . . . . . . . . . . . 212B.LIM32 Linear Independence in M32 . . . . . . . . . . . . . . . . . . . . 214B.SSP4 Spanning set in P4 . . . . . . . . . . . . . . . . . . . . . . . . . 216B.SSM22 Spanning set in M22 . . . . . . . . . . . . . . . . . . . . . . . . 218B.BP Some bases for Pn . . . . . . . . . . . . . . . . . . . . . . . . . . 221B.BM A basis for the vector space of matrices . . . . . . . . . . . . . . 221B.BSP4 A basis for a subspace of P4 . . . . . . . . . . . . . . . . . . . . 221B.BSM22 A basis for a subspace of M22 . . . . . . . . . . . . . . . . . . . 222B.RSB Row space basis . . . . . . . . . . . . . . . . . . . . . . . . . . . 223B.RS Reducing a span . . . . . . . . . . . . . . . . . . . . . . . . . . 224B.CABAK Columns as Basis, Archetype K . . . . . . . . . . . . . . . . . . 225B.AVR A vector representation . . . . . . . . . . . . . . . . . . . . . . . 226

Section DD.LDP4 Linearly dependent set in P4 . . . . . . . . . . . . . . . . . . . . 234D.DSM22 Dimension of a subspace of M22 . . . . . . . . . . . . . . . . . . 235D.DSP4 Dimension of a subspace of P4 . . . . . . . . . . . . . . . . . . . 236D.VSPUD Vector space of polynomials with unbounded degree . . . . . . . 236D.RNM Rank and nullity of a matrix . . . . . . . . . . . . . . . . . . . . 237D.RNSM Rank and nullity of a square matrix . . . . . . . . . . . . . . . . 238

Section PDPD.BP Bases for Pn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245PD.BSM22 Basis for a subspace of M22 . . . . . . . . . . . . . . . . . . . . 245PD.SVP4 Sets of vectors in P4 . . . . . . . . . . . . . . . . . . . . . . . . 246PD.RRTI Rank, rank of transpose, Archetype I . . . . . . . . . . . . . . . 248PD.CROB4 Coordinatization relative to an orthonormal basis, C4 . . . . . . 250PD.CROB3 Coordinatization relative to an orthonormal basis, C3 . . . . . . 250

Section DMDM.SS Some submatrices . . . . . . . . . . . . . . . . . . . . . . . . . . 254DM.D33M Determinant of a 3× 3 matrix . . . . . . . . . . . . . . . . . . . 255DM.MC Minors and cofactors . . . . . . . . . . . . . . . . . . . . . . . . 256DM.TCSD Two computations, same determinant . . . . . . . . . . . . . . . 257DM.DUTM Determinant of an upper-triangular matrix . . . . . . . . . . . . 258DM.ZNDAB Zero and nonzero determinant, Archetypes A and B . . . . . . . 259

Section EEEE.SEE Some eigenvalues and eigenvectors . . . . . . . . . . . . . . . . . 261EE.PM Polynomial of a matrix . . . . . . . . . . . . . . . . . . . . . . . 263EE.CAEHW Computing an eigenvalue the hard way . . . . . . . . . . . . . . 266EE.CPMS3 Characteristic polynomial of a matrix, size 3 . . . . . . . . . . . 269EE.EMS3 Eigenvalues of a matrix, size 3 . . . . . . . . . . . . . . . . . . . 270EE.ESMS3 Eigenspaces of a matrix, size 3 . . . . . . . . . . . . . . . . . . . 271EE.EMMS4 Eigenvalue multiplicities, matrix of size 4 . . . . . . . . . . . . . 272

Page 27: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Examples xxvi

EE.ESMS4 Eigenvalues, symmetric matrix of size 4 . . . . . . . . . . . . . . 273EE.HMEM5 High multiplicty eigenvalues, matrix of size 5 . . . . . . . . . . . 274EE.CEMS6 Complex eigenvalues, matrix of size 6 . . . . . . . . . . . . . . . 275EE.DEMS5 Distinct eigenvalues, matrix of size 5 . . . . . . . . . . . . . . . 278

Section PEEPEE.BDE Building desired eigenvalues . . . . . . . . . . . . . . . . . . . . 283

Section SDSD.SMS5 Similar matrices of size 5 . . . . . . . . . . . . . . . . . . . . . . 291SD.SMS3 Similar matrices of size 3 . . . . . . . . . . . . . . . . . . . . . . 292SD.EENS Equal eigenvalues, not similar . . . . . . . . . . . . . . . . . . . 294SD.DAB Diagonalization of Archetype B . . . . . . . . . . . . . . . . . . 295SD.DMS3 Diagonalizing a matrix of size 3 . . . . . . . . . . . . . . . . . . 297SD.NDMS4 A non-diagonalizable matrix of size 4 . . . . . . . . . . . . . . . 300SD.DEHD Distinct eigenvalues, hence diagonalizable . . . . . . . . . . . . 300

Section LTLT.ALT A linear transformation . . . . . . . . . . . . . . . . . . . . . . 303LT.NLT Not a linear transformation . . . . . . . . . . . . . . . . . . . . 304LT.LTPM Linear transformation, polynomials to matrices . . . . . . . . . 305LT.LTPP Linear transformation, polynomials to polynomials . . . . . . . 306LT.LTM Linear transformation from a matrix . . . . . . . . . . . . . . . 307LT.MFLT Matrix from a linear transformation . . . . . . . . . . . . . . . . 308LT.MOLT Matrix of a linear transformation . . . . . . . . . . . . . . . . . 310LT.LTDB1 Linear transformation defined on a basis . . . . . . . . . . . . . 312LT.LTDB2 Linear transformation defined on a basis . . . . . . . . . . . . . 313LT.LTDB3 Linear transformation defined on a basis . . . . . . . . . . . . . 313LT.SPIAS Sample pre-images, Archetype S . . . . . . . . . . . . . . . . . . 314LT.STLT Sum of two linear transformations . . . . . . . . . . . . . . . . . 317LT.SMLT Scalar multiple of a linear transformation . . . . . . . . . . . . . 318LT.CTLT Composition of two linear transformations . . . . . . . . . . . . 319

Section ILTILT.NIAQ Not injective, Archetype Q . . . . . . . . . . . . . . . . . . . . . 321ILT.IAR Injective, Archetype R . . . . . . . . . . . . . . . . . . . . . . . 322ILT.IAV Injective, Archetype V . . . . . . . . . . . . . . . . . . . . . . . 323ILT.NSAO Null space, Archetype O . . . . . . . . . . . . . . . . . . . . . . 325ILT.TNSAP Trivial null space, Archetype P . . . . . . . . . . . . . . . . . . 326ILT.NIAQR Not injective, Archetype Q, revisited . . . . . . . . . . . . . . . 328ILT.NIAO Not injective, Archetype O . . . . . . . . . . . . . . . . . . . . . 329ILT.IAP Injective, Archetype P . . . . . . . . . . . . . . . . . . . . . . . 329ILT.NIDAU Not injective by dimension, Archetype U . . . . . . . . . . . . . 331

Section SLTSLT.NSAQ Not surjective, Archetype Q . . . . . . . . . . . . . . . . . . . . 332SLT.SAR Surjective, Archetype R . . . . . . . . . . . . . . . . . . . . . . 333SLT.SAV Surjective, Archetype V . . . . . . . . . . . . . . . . . . . . . . 334

Page 28: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Examples xxvii

SLT.RAO Range, Archetype O . . . . . . . . . . . . . . . . . . . . . . . . 336SLT.FRAN Full range, Archetype N . . . . . . . . . . . . . . . . . . . . . . 338SLT.NSAQR Not surjective, Archetype Q, revisited . . . . . . . . . . . . . . . 340SLT.NSAO Not surjective, Archetype O . . . . . . . . . . . . . . . . . . . . 340SLT.SAN Surjective, Archetype N . . . . . . . . . . . . . . . . . . . . . . 341SLT.NSDAT Not surjective by dimension, Archetype T . . . . . . . . . . . . 343

Section IVLTIVLT.AIVLT An invertible linear transformation . . . . . . . . . . . . . . . . 344IVLT.ANILT A non-invertible linear transformation . . . . . . . . . . . . . . 345IVLT.IVSAV Isomorphic vector spaces, Archetype V . . . . . . . . . . . . . . 351

Section VRVR.VRC4 Vector representation in C4 . . . . . . . . . . . . . . . . . . . . 360VR.VRP2 Vector representations in P2 . . . . . . . . . . . . . . . . . . . . 361VR.TIVS Two isomorphic vector spaces . . . . . . . . . . . . . . . . . . . 365VR.CVSR Crazy vector space revealed . . . . . . . . . . . . . . . . . . . . 365VR.ASC A subspace charaterized . . . . . . . . . . . . . . . . . . . . . . 365VR.MIVS Multiple isomorphic vector spaces . . . . . . . . . . . . . . . . . 365VR.CP2 Coordinatizing in P2 . . . . . . . . . . . . . . . . . . . . . . . . 366VR.CM32 Coordinatization in M32 . . . . . . . . . . . . . . . . . . . . . . 367

Section MRSection CNOCNO.ACN Arithmetic of complex numbers . . . . . . . . . . . . . . . . . . 447CNO.CSCN Conjugate of some complex numbers . . . . . . . . . . . . . . . 448CNO.MSCN Modulus of some complex numbers . . . . . . . . . . . . . . . . 449

Page 29: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Proof Techniques

Section WILASection SSSLED Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9T Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13GS Getting Started . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13SE Set Equality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14L Language . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

Section RREFC Constructive Proofs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

Section TSSSN Set Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39E Equivalences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40CP Contrapositives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41CV Converses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

Section HSESection NSMU Uniqueness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59ME Multiple Equivalences . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Section VOSection LCDC Decompositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

Section SSSection LISection OSection MOP Practice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

Section RMSection RSMSection MMSection MISLESection MINSMSection VSSection SSection BSection DSection PDSection DMSection EESection PEE

xxviii

Page 30: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Proof Techniques xxix

Section SDSection LTSection ILTSection SLTSection IVLTSection VRSection MRSection CNO

Page 31: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Computation Notes

Section WILASection SSSLESection RREF

ME.MMA Matrix Entry (Mathematica) . . . . . . . . . . . . . . . . . . . . . . 22

ME.TI86 Matrix Entry (TI-86) . . . . . . . . . . . . . . . . . . . . . . . . . . 23

ME.TI83 Matrix Entry (TI-83) . . . . . . . . . . . . . . . . . . . . . . . . . . 23

RR.MMA Row Reduce (Mathematica) . . . . . . . . . . . . . . . . . . . . . . 32

RR.TI86 Row Reduce (TI-86) . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

RR.TI83 Row Reduce (TI-83) . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

Section TSS

LS.MMA Linear Solve (Mathematica) . . . . . . . . . . . . . . . . . . . . . . 45

Section HSESection NSMSection VO

VLC.MMAVector Linear Combinations (Mathematica) . . . . . . . . . . . . . . 69

VLC.TI86 Vector Linear Combinations (TI-86) . . . . . . . . . . . . . . . . . . 69

VLC.TI83 Vector Linear Combinations (TI-83) . . . . . . . . . . . . . . . . . . 70

Section LCSection SSSection LISection OSection MOSection RMSection RSMSection MM

MM.MMA Matrix Multiplication (Mathematica) . . . . . . . . . . . . . . . . . 154

Section MISLE

MI.MMA Matrix Inverses (Mathematica) . . . . . . . . . . . . . . . . . . . . . 173

Section MINSMSection VSSection SSection BSection DSection PDSection DMSection EESection PEESection SDSection LTSection ILT

xxx

Page 32: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Computation Notes xxxi

Section SLTSection IVLTSection VRSection MRSection CNO

Page 33: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Part CCore

1

Page 34: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

SLE: Systems of Linear Equations

Section WILA

What is Linear Algebra?

Subsection LA“Linear” + “Algebra”

The subject of linear algebra can be partially explained by the meaning of the two termscomprising the title. “Linear” is a term you will appreciate better at the end of thiscourse, and indeed, attaining this appreciation could be taken as one of the primarygoals of this course. However for now, you can understand it to mean anything that is“straight” or “flat.” For example in the xy-plane you might be accustomed to describingstraight lines (is there any other kind?) as the set of solutions to an equation of theform y = mx + b, where the slope m and the y-intercept b are constants that togetherdescribe the line. In multivariate calculus, you may have discussed planes. Living in threedimensions, with coordinates described by triples (x, y, z), they can be described as theset of solutions to equations of the form ax + by + cz = d, where a, b, c, d are constantsthat together determine the plane. While we might describe planes as “flat,” lines inthree dimensions might be described as “straight.” From a multivariate calculus courseyou will recall that lines are sets of points described by equations such as x = 3t − 4,y = −7t + 2, z = 9t, where t is a parameter that can take on any value.

Another view of this notion of “flatness” is to recognize that the sets of points justdescribed are solutions to equations of a relatively simple form. These equations involveaddition and multiplication only. We will have a need for subtraction, and occasionallywe will divide, but mostly you can describe “linear” equations as involving only additionand multiplication. Here are some examples of typical equations we will see in the nextfew sections:

2x + 3y − 4z = 13 4x1 + 5x2 − x3 + x4 + x5 = 0 9a− 2b + 7c + 2d = −7

What we will not see are equations like:

xy + 5yz = 13 x1 + x32/x4 − x3x4x

25 = 0 tan(ab) + log(c− d) = −7

The exception will be that we will on occasion need to take a square root.

2

Page 35: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection WILA.A An application: packaging trail mix 3

You have probably heard the word “algebra” frequently in your mathematical prepa-ration for this course. Most likely, you have spent a good ten to fifteen years learning thealgebra of the real numbers, along with some introduction to the very similar algebra ofcomplex numbers. However, there are many new algebras to learn and use, and likelylinear algebra will be your second algebra. Like learning a second language, the necessaryadjustments can be challenging at times, but the rewards are many. And it will makelearning your third and fourth algebras even easier. Perhaps you have heard of “groups”and “rings” (or maybe you have studied them already), which are excellent examples ofother algebras with very interesting properties and applications. In any event, prepareyourself to learn a new algebra and realize that some of the old rules you used for thereal numbers may no longer apply to this new algebra you will be learning!

The brief discussion above about lines and planes suggests that linear algebra hasan inherently geometric nature, and this is true. Examples in two and three dimensionscan be used to provide valuable insight into important concepts of this course. However,much of the power of linear algebra will be the ability to work with “flat” or “straight”objects in higher dimensions, without concerning ourselves with visualizing the situation.While much of our intuition will come from examples in two and three dimensions, wewill maintain an algebraic approach to the subject, with the geometry being secondary.Others may wish to switch this emphasis around, and that can lead to a very fruitfuland beneficial course, but here and now we are laying our bias bare.

Subsection AAn application: packaging trail mix

We finish this section with a rather involved example that will highlight some of thepower and techniques of linear algebra. Work through all of the details with pencil andpaper, until you believe all the assertions made. However, in this introductory example,do not concern yourself with how some of the results are obtained or how you might beexpected to solve a similar problem. We will come back to this example later and exposesome of the techniques used and properties exploited. For now, use your background inmathematics to convince yourself that everything said here really is correct.

Example WILA.TMTrail Mix PackagingSuppose you are the production manager at a food-packaging plant and one of yourproduct lines is trail mix, a healthy snack popular with hikers and backpackers, containingraisins, peanuts and hard-shelled chocolate pieces. By adjusting the mix of these threeingredients, you are able to sell three varieties of this item. The fancy version is sold inhalf-kilogram packages at outdoor supply stores and has more chocolate and fewer raisins,thus commanding a higher price. The standard version is sold in one kilogram packagesin grocery stores and gas station mini-markets. Since the standard version has roughlyequal amounts of each ingredient, it is not as expensive as the fancy version. Finally,a bulk version is sold in bins at grocery stores for consumers to load into plastic bagsin amounts of their choosing. To appeal to the shoppers that like bulk items for theireconomy and healthfulness, this mix has much more raisins (at the expense of chocolate)and therefore sells for less.

Page 36: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection WILA.A An application: packaging trail mix 4

Your production facilities have limited storage space and early each morning youare able to receive and store 380 kilograms of raisins, 500 kilograms of peanuts and620 kilograms of chocolate pieces. As production manager, one of your most importantduties is to decide how much of each version of trail mix to make every day. Clearly, youcan have up to 1500 kilograms of raw ingredients available each day, so to be the mostproductive you will likely produce 1500 kilograms of trail mix each day. Also, you wouldprefer not to have any ingredients leftover each day, so that your final product is as freshas possible. But how should these ingredients be allocated to the mixing of the bulk,standard and fancy versions?

First, we need a little more information about the mixes. Workers mix the ingredientsin 15 kilogram batches, and each row of the table below gives a recipe for a 15 kilogrambatch. There is some additional information on the costs of the ingredients and the pricethe manufacturer can charge for the different versions of the trail mix.

Raisins Peanuts Chocolate Cost Sale Price(kg/batch) (kg/batch) (kg/batch) ($/kg) ($/kg)

Bulk 7 6 2 3.69 4.99Standard 6 4 5 3.86 5.50Fancy 2 5 8 4.45 6.50

Storage (kg) 380 500 620Cost ($/kg) 2.55 4.65 4.80

As production manager, it is important to realize that you only have three decisions tomake — the amount of bulk mix to make, the amount of standard mix to make and theamount of fancy mix to make. Everything else is beyond your control or is handled byanother department within the company. Principally, you are also limited by the amountof raw ingredients you can store each day. Let us denote the amount of each mix toproduce each day, measured in kilograms, by the variable quantities b, s and f . Yourproduction schedule can be described as values of b, s and f that do several things. First,we cannot make negative quantities of each mix, so

b ≥ 0 s ≥ 0 f ≥ 0.

Second, if we want to consume all of our ingredients each day, the storage capacities leadto three (linear) equations, one for each ingredient,

7

15b +

6

15s +

2

15f = 380 (raisins)

6

15b +

4

15s +

5

15f = 500 (peanuts)

2

15b +

5

15s +

8

15f = 620 (chocolate)

It happens that this system of three equations has just one solution. In other words,as production manager, your job is easy, since there is but one way to use up all of yourraw ingredients making trail mix. This single solution is

b = 300 kg s = 300 kg f = 900 kg.

We do not yet have the tools to explain why this solution is the only one, but it shouldbe simple for you to verify that this is indeed a solution. (Go ahead, we will wait.)

Page 37: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection WILA.A An application: packaging trail mix 5

Determining solutions such as this, and establishing that they are unique, will be themain motivation for our initial study of linear algebra.

So we have solved the problem of making sure that we make the best use of ourlimited storage space, and each day use up all of the raw ingredients that are shippedto us. Additionally, as production manager, you must report weekly to the CEO ofthe company, and you know he will be more interested in the profit derived from yourdecisions than in the actual production levels. So you compute,

300(4.99− 3.69) + 300(5.50− 3.86) + 900(6.50− 4.45) = 2727

for a daily profit of $2,727 from this production schedule. The computation of the dailyprofit is also beyond our control, though it is definitely of interest, and it too looks likea “linear” computation.

As often happens, things do not stay the same for long, and now the marketing de-partment has suggested that your company’s trail mix products standardize on every mixbeing one-third peanuts. Adjusting the peanut portion of each recipe by also adjustingthe chocolate portion, leads to revised recipes, and slightly different costs for the bulkand standard mixes, as given in the following table.

Raisins Peanuts Chocolate Cost Sale Price(kg/batch) (kg/batch) (kg/batch) ($/kg) ($/kg)

Bulk 7 5 3 3.70 4.99Standard 6 5 4 3.85 5.50Fancy 2 5 8 4.45 6.50

Storage (kg) 380 500 620Cost ($/kg) 2.55 4.65 4.80

In a similar fashion as before, we desire values of b, s and f so that

b ≥ 0, s ≥ 0, f ≥ 0

and

7

15b +

6

15s +

2

15f = 380 (raisins)

5

15b +

5

15s +

5

15f = 500 (peanuts)

3

15b +

4

15s +

8

15f = 620 (chocolate)

It now happens that this system of equations has infinitely many solutions, as wewill now demonstrate. Let f remain a variable quantity. Then if we make f kilogramsof the fancy mix, we will make 4f − 3300 kilograms of the bulk mix and −5f + 4800kilograms of the standard mix. let us now verify that, for any choice of f , the values ofb = 4f − 3300 and s = −5f + 4800 will yield a production schedule that exhausts all ofthe day’s supply of raw ingredients. Grab your pencil and paper and play along.

7

15(4f − 3300) +

6

15(−5f + 4800) +

2

15f = 0f +

5700

15= 380

5

15(4f − 3300) +

5

15(−5f + 4800) +

5

15f = 0f +

7500

15= 500

3

15(4f − 3300) +

4

15(−5f + 4800) +

8

15f = 0f +

9300

15= 620

Page 38: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection WILA.A An application: packaging trail mix 6

Again, right now, do not be concerned about how you might derive expressions likethose for b and s that fit so nicely into this system of equations. But do convinceyourself that they lead to an infinite number of possibilities for solutions to the threeequations that describe our storage capacities. As a practical matter, there really arenot an infinite number of solutions, since we are unlikely to want to end the day with afractional number of bags of fancy mix, so our allowable values of f should probably beintegers. More importantly, we need to remember that we cannot make negative amountsof each mix! Where does this lead us? Positive quantities of the bulk mix requires that

b ≥ 0 ⇒ 4f − 3300 ≥ 0 ⇒ f ≥ 825.

Similarly for the standard mix,

s ≥ 0 ⇒ −5f + 4800 ≥ 0 ⇒ f ≤ 960.

So, as production manager, you really have to choose a value of f from the set

{825, 826, . . . , 960}

leaving you with 136 choices, each of which will exhaust the day’s supply of raw ingredi-ents. Pause now and think about which will you would choose.

Recalling your weekly meeting with the CEO suggests that you might want to choosea production schedule that yields the biggest possible profit for the company. So youcompute an expression for the profit based on your as yet undetermined decision for thevalue of f ,

(4f − 3300)(4.99− 3.70)+ (−5f +4800)(5.50− 3.85)+ (f)(6.50− 4.45) = −1.04f +3663.

Since f has a negative coefficient it would appear that mixing fancy mix is detrimentalto your profit and should be avoided. So you will make the decision to set daily fancymix production at f = 825. This has the effect of setting b = 4(825)− 3300 = 0 and westop producing bulk mix entirely. So the remainder of your daily production is standardmix at the level of s = −5(825) + 4800 = 675 kilograms and the resulting daily profitis (−1.04)(825) + 3663 = 2805. It is a pleasant surprise that daily profit has risen to$2,805, but this is not the most important part of the story. What is important here isthat there are a large number of ways to produce trail mix that use all of the day’s worthof raw ingredients and you were able to easily choose the one that netted the largestprofit. Notice too how all of the above computations look “linear.”

In the food industry, things do not stay the same for long, and now the sales depart-ment says that increased competition has lead to the decision to stay competitive andcharge just $5.25 for a kilogram of the standard mix, rather than the previous $5.50 perkilogram. This decision has no effect on the possibilities for the production schedule,but will affect the decision based on profit considerations. So you revisit just the profitcomputation, suitably adjusted for the new selling price of standard mix,

(4f − 3300)(4.99− 3.70) + (−5f + 4800)(5.25− 3.85) + (f)(6.50− 4.45) = 0.21f + 2463.

Now it would appear that fancy mix is beneficial to the company’s profit since the valueof f has a positive coefficient. So you take the decision to make as much fancy mix aspossible, setting f = 960. This leads to s = −5(960) + 4800 = 0 and the increased

Page 39: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection WILA.A An application: packaging trail mix 7

competition has driven you out of the standard mix market all together. The remainderof production is therefore bulk mix at a daily level of b = 4(960)− 3300 = 540 kilogramsand the resulting daily profit is 0.21(960) + 2463 = 2664.6. A daily profit of $2,664.60 isless than it used to be, but as production manager, you have made the best of a difficultsituation and shown the sales department that the best course is to pull out of the highlycompetitive standard mix market completely. 4

This example is taken from a field of mathematics variously known by names such asoperations research, system science or management science. More specifically, this is anperfect example of problems that are solved by the techniques of “linear programming.”

There is a lot going on under the hood in this example. The heart of the matter is thesolution to simultaneous systems of linear equations, which is the topic of the next fewsections, and a recurrent theme throughout this course. We will return to this exampleon several occasions to reveal some of the reasons for its behavior.

Page 40: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection WILA.EXC Exercises 8

Subsection EXCExercises

TODO: Vary selling prices on mixes to get opposite conclusions.

Page 41: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section SSSLE Solving Systems of Simultaneous Linear Equations 9

Section SSSLE

Solving Systems of Simultaneous Linear Equations

We will motivate our study of linear algebra by considering the problem of solving severallinear equations simultaneously. The word “solve” tends to get abused somewhat, asin “solve this problem.” When talking about equations we understand a more precisemeaning: find all of the values of some variable quantities that make an equation, orseveral equations, true.

Example SSSLE.STNESolving two (nonlinear) equationsSuppose we desire the simultaneous solutions of the two equations,

x2 + y2 = 1

−x +√

3y = 0.

You can easily check by substitution that x =√

32

, y = 12

and x = −√

32

, y = −12

are bothsolutions. We need to also convince ourselves that these are the only solutions. To seethis, plot each equation on the xy-plane, which means to plot (x, y) pairs that make anindividual equation true. In this case we get a circle centered at the origin with radius1 and a straight line through the origin with slope 1√

3. The intersections of these two

curves are our desired simultaneous solutions, and so we believe from our plot that thetwo solutions we know already are the only ones. We like to write solutions as sets, soin this case we write the set of solutions as

S = {(√

32

, 12), (−

√3

2, −1

2)} 4

In order to discuss systems of linear equations carefully, we need a precise definition.And before we do that, we will introduce our periodic discussions about “proof tech-niques.” Linear algebra is an excellent setting for learning how to read, understand andformulate proofs. To help you in this process, we will digress, at irregular intervals, aboutsome important aspect of working with proofs.

Proof Technique DDefinitionsA definition is a made-up term, used as a kind of shortcut for some typically morecomplicated idea. For example, we say a whole number is even as a shortcut for sayingthat when we divide the number by two we get a remainder of zero. With a precisedefinition, we can answer certain questions unambiguously. For example, did you everwonder if zero was an even number? Now the answer should be clear since we have aprecise definition of what we mean by the term even.

A single term might have several possible definitions. For example, we could say thatthe whole number n is even if there is another whole number k such that n = 2k. Wesay this is an equivalent definition since it categorizes even numbers the same way ourfirst definition does.

Page 42: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section SSSLE Solving Systems of Simultaneous Linear Equations 10

Definitions are like two-way streets — we can use a definition to replace somethingrather complicated by its definition (if it fits) and we can replace a definition by its morecomplicated description. A definition is usually written as some form of an implication,such as “If something-nice-happens, then blatzo.” However, this also means that “Ifblatzo, then something-nice-happens,” even though this may not be formally stated.This is what we mean when we say a definition is a two-way street — it is really twoimplications, going in opposite “‘directions.”

Anybody (including you) can make up a definition, so long as it is unambiguous, butthe real test of a definition’s utility is whether or not it is useful for describing interestingor frequent situations.

We will talk about theorems later (and especially equivalences). For now, be sure notto confuse the notion of a definition with that of a theorem.

In this book, we will display every new definition carefully set-off from the text, andthe term being defined will be written thus: definition. Additionally, there is a fulllist of all the definitions, in order of their appearance located at the front of the book(Definitions). Finally, the acronym for each definition can be found in the index (Index).Definitions are critical to doing mathematics and proving theorems, so we’ve given youlots of ways to locate a definition should you forget its. . . uh, uh, well, . . . definition.

Can you formulate a precise definition for what it means for a number to be odd?(Don’t just say it is the opposite of even. Act as if you don’t have a definition for evenyet.) Can you formulate your definition a second, equivalent, way? Can you employ yourdefinition to test an odd and an even number for “odd-ness”? ♦

Definition SSLESystem of Simultaneous Linear EquationsA system of simultaneous linear equations is a collection of m equations in thevariable quantities x1, x2, x3, . . . , xn of the form,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm

where the values of aij, bi and xj are from the set of complex numbers, C. �

Don’t let the mention of the complex numbers, C, rattle you. We will stick with realnumbers exclusively for many more sections, and it will sometimes seem like we onlywork with integers! However, we want to leave the possibility of complex numbers open,and there will be occasions in subsequent sections where they are necessary. For now,here is an example to illustrate using the notation introduced in Definition SSLE [10].

Example SSSLE.NSENotation for a system of equationsGiven the system of simultaneous linear equations,

x1 + 2x2 + x4 = 7

x1 + x2 + x3 − x4 = 3

3x1 + x2 + 5x3 − 7x4 = 1

Page 43: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.PSS Possibilities for solution sets 11

we have n = 4 variables and m = 3 equations. Also,

a11 = 1 a12 = 2 a13 = 0 a14 = 1

a21 = 1 a22 = 1 a23 = 1 a24 = −1

a31 = 3 a32 = 1 a33 = 5 a34 = −7

b1 = 7 b2 = 3 b3 = 1.

Additionally, convince yourself that x1 = −2, x2 = 4, x3 = 2, x4 = 1 is one solution (butit is not the only one!). 4

We will often shorten the term “system of simultaneous linear equations” to “system oflinear equations” or just “system of equations” leaving the linear aspect implied.

Subsection PSSPossibilities for solution sets

The next example illustrates the possibilities for the solution set of a system of linearequations. We will not be too formal here, and the necessary theorems to back up ourclaims will come in subsequent sections. So read for feeling and come back later to revisitthis example.

Example SSSLE.TTSThree typical systemsConsider the system of two equations with two variables,

2x1 + 3x2 = 3

x1 − x2 = 4.

If we plot the solutions to each of these equations separately on the x1x2-plane, we gettwo lines, one with negative slope, the other with positive slope. They have exactly onepoint in common, (x1, x2) = (3, −1), which is the solution x1 = 3, x2 = −1. From thegeometry, we believe that this is the only solution to the system of equations, and so wesay it is unique.

Now adjust the system with a different second equation,

2x1 + 3x2 = 3

4x1 + 6x2 = 6.

A plot of the solutions to these equations individually results in two lines, one on top ofthe other! There are infinitely many pairs of points that make both equations true. Wewill learn shortly how to describe this infinite solution set precisely. Notice now how thesecond equation is just a multiple of the first.

One more minor adjustment

2x1 + 3x2 = 3

4x1 + 6x2 = 10.

A plot now reveals two lines with identical slopes, i.e. parallel lines. They have no pointsin common, and so the system has a solution set that is empty, S = ∅. 4

Page 44: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 12

This example exhibits all of the typical behaviors of a system of equations. A subsequenttheorem will tell us that every system of simultaneous linear equations has a solutionset that is empty, contains a single solution or contains infinitely many solutions. Ex-ample SSSLE.STNE [9] yielded exactly two solutions, but this does not contradict theforthcoming theorem, since the equations are not linear and do not match the form ofDefinition SSLE [10].

Subsection ESEOEquivalent systems and equation operations

With all this talk about finding solution sets for systems of linear equations, you mightbe ready to begin learning how to find these solution sets yourself. We begin with ourfirst definition that takes a common word and gives it a very precise meaning in thecontext of systems of linear equations.

Definition ESEquivalent SystemsTwo systems of simultaneous linear equations are equivalent if their solution sets areequal. �

Notice here that the two systems of equations could look very different (i.e. not be equal),but still have equal solution sets, and we would then call the systems equivalent. Twolinear equations in two variables might be plotted as two lines that intersect in a singlepoint. A different system, with three equations in two variables might have a plot thatis three lines, all intersecting at a common point, with this common point identical tothe intersection point for the first system. By our definition, we could then say thesetwo very different looking systems of equations are equivalent, since they have identicalsolution sets. It is really like a weaker form of equality, where we allow the systems to bedifferent in some respects, but use the term equivalent to highlight the situation whentheir solution sets are equal.

With this definition, we can begin to describe our strategy for solving linear systems.Given a system of linear equations that looks difficult to solve, we would like to have anequivalent system that is easy to solve. Since the systems will have equal solution sets,we can solve the “easy” system and get the solution set to the “difficult” system. Herecome the tools for making this strategy viable.

Definition EOEquation OperationsGiven a system of simultaneous linear equations, the following three operations willtransform the system into a different one, and each is known as an equation operation.

1. Swap the locations of two equations in the list.

2. Multiply each term of an equation by a nonzero quantity.

3. Multiply each term of one equation by some quantity, and add these terms to asecond equation, on both sides of the equality. Leave the first equation the sameafter this operation, but replace the second equation by the new one. �

Page 45: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 13

These descriptions might seem a bit vague, but the proof or the examples that followshould make it clear what is meant by each.

Proof Technique TTheoremsHigher mathematics is about understanding theorems. Reading them, understandingthem, applying them, proving them. We are ready to prove our first momentarily. Everytheorem is a shortcut — we prove something in general, and then whenever we find aspecific instance covered by the theorem we can immediately say that we know somethingelse about the situation by applying the theorem. In many cases, this new informationcan be gained with much less effort than if we did not know the theorem.

The first step in understanding a theorem is to realize that the statement of ev-ery theorem can be rewritten using statements of the form “If something-happens,then something-else-happens.” The “something-happens” part is the hypothesis andthe “something-else-happens” is the conclusion. To understand a theorem, it helpsto rewrite its statement using this construction. To apply a theorem, we verify that“something-happens” in a particular instance and immediately conclude that “something-else-happens.” To prove a theorem, we must argue based on the assumption that thehypothesis is true, and arrive through the process of logic that the conclusion must thenalso be true. ♦

Theorem EOPSSEquation Operations Preserve Solution SetsSuppose we apply one of the three equation operations to the system of simultaneouslinear equations

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm.

Then the original system and the transformed system are equivalent systems. �

Proof Before we begin the proof, we make two comments about proof techniques.

Proof Technique GSGetting Started“I don’t know how to get started!” is often the lament of the novice proof-builder. Hereare a few pieces of advice.

1. As mentioned in Technique T [13], rewrite the statement of the theorem in an“if-then” form. This will simplify identifying the hypothesis and conclusion, whichare referenced in the next few items.

2. Ask yourself what it is you are trying to prove. This is always part of your conclu-sion. Are you being asked to conclude that two numbers are equal, that a functionis differentiable or a set is a subset of another? You cannot bring other techniquesto bear if you do not know what type of conclusion you have.

Page 46: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 14

3. Write down reformulations of your hypotheses. Interpret and translate each defini-tion properly.

4. Write your hypothesis at the top of a sheet of paper and your conclusion at thebottom. See if you can formulate a statement that precedes the conclusion and alsoimplies it. Work down from your hypothesis, and up from your conclusion, and seeif you can meet in the middle. When you are finished, rewrite the proof nicely,from hypothesis to conclusion, with verifiable implications giving each subsequentstatement. ♦

Proof Technique SESet EqualityIn the theorem we are trying to prove, the conclusion is that two systems are equivalent.By Definition ES [12] this translates to requiring that solution sets be equal for the twosystems. So we are being asked to show that two sets are equal. How do we do this? Well,there is a very standard technique, and we will use it repeatedly through the course. Solets add it to our toolbox now.

A set is just a collection of items, which we refer to generically as elements. If A isa set, and a is one of its elements, we write that piece of information as a ∈ A. Similarly,if b is not in A, we write b 6∈ A. Given two sets, A and B, we say that A is a subset ofB if all the elements of A are also in B. More formally (and much easier to work with)we describe this situation as follows: A is a subset of B if whenever x ∈ A, then x ∈ B.Notice the use of the “if-then” construction here. The notation for this is A ⊆ B. (If wewant to disallow the possibility that A is the same as B, we use A ⊂ B.)

But what does it mean for two sets to be equal? They must be the same. Well, thatexplanation is not really too helpful, is it? How about: If A ⊆ B and B ⊆ A, then Aequals B. This gives us something to work with, if A is a subset of B, and vice versa,then they must really be the same set. We will now make the symbol “=” do double-dutyand extend its use to statements like A = B, where A and B are sets. ♦

Now we can take each equation operation in turn and show that the solution sets of thetwo systems are equal, using the technique just outlined.

1. It will not be our habit in proofs to resort to saying statements are “obvious,” butin this case, it should be. There is nothing about the order in which we write linearequations that affects their solutions, so the solution set will be equal if the systemsonly differ by a rearrangement of the order of the equations.

2. Suppose α 6= 0 is a number. Let’s choose to multiply the terms of equation i by αto build the new system of equations,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

αai1x1 + αai2x2 + αai3x3 + · · ·+ αainxn = αbi

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm.

Page 47: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 15

Let S denote the solutions to the system in the statement of the theorem, and letT denote the solutions to the transformed system.

(a) Show S ⊆ T . Suppose (x1, x2, x3, . . . , xn) = (β1, β2, β3, . . . , βn) ∈ S is asolution to the original system. Ignoring the i-th equation for a moment, weknow it makes every other equation of the transformed system true. We alsoknow that

ai1β1 + ai2β2 + ai3β3 + · · ·+ ainβn = bi

which we can multiply by α to get

αai1β1 + αai2β2 + αai3β3 + · · ·+ αainβn = αbi.

This says that the i-th equation of the transformed system is also true, so wehave established that (β1, β2, β3, . . . , βn) ∈ T , and therefore S ⊆ T .

(b) Now show T ⊆ S. Suppose (x1, x2, x3, . . . , xn) = (β1, β2, β3, . . . , βn) ∈ S isa solution to the transformed system. Ignoring the i-th equation for a moment,we know it makes every other equation of the original system true. We alsoknow that

αai1β1 + αai2β2 + αai3β3 + · · ·+ αainβn = αbi

which we can multiply by 1α, since α 6= 0, to get

ai1β1 + ai2β2 + ai3β3 + · · ·+ ainβn = bi

This says that the i-th equation of the original system is also true, so we haveestablished that (β1, β2, β3, . . . , βn) ∈ S, and therefore T ⊆ S. Locate thekey point where we required that α 6= 0, and consider what would happen ifα = 0.

3. Suppose α is a number. Let’s choose to multiply the terms of equation i by α andadd them to equation j in order to build the new system of equations,

a11x1 + a12x2 + · · ·+ a1nxn = b1

a21x1 + a22x2 + · · ·+ a2nxn = b2

a31x1 + a32x2 + · · ·+ a3nxn = b3

...

(αai1 + aj1)x1 + (αai2 + aj2)x2 + · · ·+ (αain + ajn)xn = αbi + bj

...

am1x1 + am2x2 + · · ·+ amnxn = bm.

Let S denote the solutions to the system in the statement of the theorem, and letT denote the solutions to the transformed system.

Page 48: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 16

(a) Show S ⊆ T . Suppose (x1, x2, x3, . . . , xn) = (β1, β2, β3, . . . , βn) ∈ S is asolution to the original system. Ignoring the j-th equation for a moment, weknow it makes every other equation of the transformed system true. Usingthe fact that it makes the i-th and j-th equations true, we find

(αai1 + aj1)β1 + (αai2 + aj2)β2 + · · ·+ (αain + ajn)βn =

(αai1β1 + αai2β2 + · · ·+ αainβn) + (aj1β1 + aj2β2 + · · ·+ ajnβn) =

α(ai1β1 + ai2β2 + · · ·+ ainβn) + (aj1β1 + aj2β2 + · · ·+ ajnβn) = αbi + bj.

This says that the j-th equation of the transformed system is also true, so wehave established that (β1, β2, β3, . . . , βn) ∈ T , and therefore S ⊆ T .

(b) Now show T ⊆ S. Suppose (x1, x2, x3, . . . , xn) = (β1, β2, β3, . . . , βn) ∈ Sis a solution to the transformed system. Ignoring the j-th equation for amoment, we know it makes every other equation of the original system true.We then find

aj1β1 + aj2β2 + · · ·+ ajnβn =

aj1β1 + aj2β2 + · · ·+ ajnβn + αbi − αbi =

aj1β1 + aj2β2 + · · ·+ ajnβn + (αai1β1 + αai2β2 + · · ·+ αainβn)− αbi =

aj1β1 + aj2β2 + · · ·+ ajnβn + (αai1β1 + αai2β2 + · · ·+ αainβn)− αbi =

(αai1 + aj1)β1 + (αai2 + aj2)β2 + · · ·+ (αain + ajn)βn − αbi =

αbi + bj − αbi = bj

This says that the j-th equation of the original system is also true, so we haveestablished that (β1, β2, β3, . . . , βn) ∈ S, and therefore T ⊆ S.

Why didn’t we need to require that α 6= 0 for this row operation? In other words,how does the third statement of the theorem read when α = 0? Does our proofrequire some extra care when α = 0? Compare your answers with the similarsituation for the second row operation. �

Theorem EOPSS [13] is the necessary tool to complete our strategy for solving systemsof equations. We will use equation operations to move from one system to another, allthe while keeping the solution set the same. With the right sequence of operations, wewill arrive at a simpler equation to solve. The next two examples illustrate this idea,while saving some of the details for later.

Example SSSLE.USThree equations, one solutionWe solve the following system by a sequence of equation operations.

x1 + 2x2 + 2x3 = 4

x1 + 3x2 + 3x3 = 5

2x1 + 6x2 + 5x3 = 6

α = −1 times equation 1, add to equation 2:

x1 + 2x2 + 2x3 = 4

0x1 + 1x2 + 1x3 = 1

2x1 + 6x2 + 5x3 = 6

Page 49: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 17

α = −2 times equation 1, add to equation 3:

x1 + 2x2 + 2x3 = 4

0x1 + 1x2 + 1x3 = 1

0x1 + 2x2 + 1x3 = −2

α = −2 times equation 2, add to equation 3:

x1 + 2x2 + 2x3 = 4

0x1 + 1x2 + 1x3 = 1

0x1 + 0x2 − 1x3 = −4

α = −1 times equation 3:

x1 + 2x2 + 2x3 = 4

0x1 + 1x2 + 1x3 = 1

0x1 + 0x2 + 1x3 = 4

which can be written more clearly as

x1 + 2x2 + 2x3 = 4

x2 + x3 = 1

x3 = 4

This is now a very easy system of equations to solve. The third equation requires thatx3 = 4 to be true. Making this substitution into equation 2 we arrive at x2 = −3, andfinally, substituting these values of x2 and x3 into the first equation, we find that x1 = 2.Note too that this is the only solution to this final system of equations, since we wereforced to choose these values to make the equations true. Since we performed equationoperations on each system to obtain the next one in the list, all of the systems listed hereare all equivalent to each other by Theorem EOPSS [13]. Thus (x1, x2, x3) = (2,−3, 4)is the unique solution to the original system of equations (and all of the other systemsof equations). 4

Example SSSLE.ISThree equations, infinitely many solutionsThe following system of equations made an appearance earlier in this section (Exam-ple SSSLE.NSE [10]), where we listed one of its solutions. Now, we will try to find all ofthe solutions to this system.

x1 + 2x2 + 0x3 + x4 = 7

x1 + x2 + x3 − x4 = 3

3x1 + x2 + 5x3 − 7x4 = 1

α = −1 times equation 1, add to equation 2:

x1 + 2x2 + 0x3 + x4 = 7

0x1 − x2 + x3 − 2x4 = −4

3x1 + x2 + 5x3 − 7x4 = 1

Page 50: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 18

α = −3 times equation 1, add to equation 3:

x1 + 2x2 + 0x3 + x4 = 7

0x1 − x2 + x3 − 2x4 = −4

0x1 − 5x2 + 5x3 − 10x4 = −20

α = −5 times equation 2, add to equation 3:

x1 + 2x2 + 0x3 + x4 = 7

0x1 − x2 + x3 − 2x4 = −4

0x1 + 0x2 + 0x3 + 0x4 = 0

α = −1 times equation 2:

x1 + 2x2 + 0x3 + x4 = 7

0x1 + x2 − x3 + 2x4 = 4

0x1 + 0x2 + 0x3 + 0x4 = 0

which can be written more clearly as

x1 + 2x2 + x4 = 7

x2 − x3 + 2x4 = 4

0 = 0

What does the equation 0 = 0 mean? We can choose any values for x1, x2, x3, x4 andthis equation will be true, so we just need to only consider further the first two equationssince the third is true no matter what. We can analyze the second equation withoutconsideration of the variable x1. It would appear that there is considerable latitude inhow we can choose x2, x3, x4 and make this equation true. Lets choose x3 and x4 to beanything we please, say x3 = β3 and x4 = β4. Then equation 2 becomes

x2 − β3 + 2β4 = 4 ⇒ x2 = 4 + β3 − 2β4

Now we can take these arbitrary values for x3 and x4, and this expression for x2 andemploy them in equation 1,

x1 + 2(4 + β3 − 2β4) + β4 = 7 ⇒ x1 = −1− 2β3 + 3β4

So our arbitrary choices of values for x3 and x4 (β3 and β4) translate into specific valuesof x1 and x2. The lone solution given in Example SSSLE.NSE [10] was obtained bychoosing β3 = 2 and β4 = 1. Now we can easily and quickly find many more (infinitelymore). Suppose we choose β3 = 5 and β4 = −2, then we compute

x1 = −1− 2(5) + 3(−2) = −17

x2 = 4 + 5− 2(−2) = 13

and you can verify that (x1, x2, x3, x4) = (−17, 13, 5, −2) makes all three equationstrue. The entire solution set is written as

S = {(−1− 2β3 + 3β4, 4 + β3 − 2β4, β3, β4) | β3 ∈ C, β4 ∈ C}

Page 51: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 19

It would be instructive to finish off your study of this example by taking the general formof the solutions given in this set and substituting them into each of the three equationsand verify that they are true in each case. 4

In the next section we will describe how to use equation operations to systematicallysolve any system of simultaneous linear equations. But first, one of our more importantpieces of advice about doing mathematics.

Proof Technique LLanguage

Like any science, the language of math must be understood before furtherstudy can continue.

Erin Wilson, StudentSeptember, 2004

Mathematics is a language. It is a way to express complicated ideas clearly, precisely,and unambiguously. Because of this, it can be difficult to read. Read slowly, and havepencil and paper at hand. It will usually be necessary to read something several times.While reading can be difficult, it is even hard to speak mathematics, and so that is thetopic of this technique.

I am going to suggest a simple modification to the way you use language that willmake it much, much easier to become proficient at speaking mathematics and eventuallyit will become second nature. Think of it as a training aid or practice drill you mightuse when learning to become skilled at a sport.

First, eliminate pronouns from your vocabulary when discussing linear algebra, inclass or with your colleagues. Do not use: it, that, those, their or similar sources ofconfusion. This is the single easiest step you can take to make your oral expression ofmathematics clearer to others, and in turn, it will greatly help your own understanding.

Now rid yourself of the word “thing” (or variants like “something”). When you aretempted to use this word realize that there is some object you want to discuss, and welikely have a definition for that object (see the discussion at Technique D [9]). Always“think about your objects” and many aspects of the study of mathematics will get easier.Ask yourself: “Am I working with a set, a number, a function, an operation, or what?”Knowing what an object is will allow you to narrow down the procedures you mayapply to it. If you have studied an object-oriented computer programming language,then perhaps this advice will be even clearer, since you know that a compiler will oftencomplain with an error message if you confuse your objects.

Third, eliminate the verb “works” (as in “the equation works”) from your vocabu-lary. This term is used as a substitute when we are not sure just what we are tryingto accomplish. Usually we are trying to say that some object fulfills some condition.The condition might even have a definition associated with it, making it even easier todescribe.

Last, speak slooooowly and thoughtfully as you try to get by without all these lazywords. It is hard at first, but you will get better with practice. Especially in class, whenthe pressure is on and all eyes are on you, don’t succumb to the temptation to use theseweak words. Slow down, we’d all rather wait for a slow, well-formed question or answerthan a fast, sloppy, incomprehensible one.

Page 52: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.ESEO Equivalent systems and equation operations 20

When you come to class, check your pronouns at the door, along with other weakwords. And when studying with friends, you might make a game of catching one anotherusing pronouns, “thing,” or “works.” I know I’ll be calling you on it! ♦

Page 53: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SSSLE.EXC Exercises 21

Subsection EXCExercises

TODO: intersecting circle and hyperbola 0 to 4 solution from 5 different circles

Page 54: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 22

Section RREF

Reduced Row-Echelon Form

After solving a few systems of equations, you will recognize that it doesn’t matter somuch what we call our variables, as opposed to what numbers act as their coefficients.A system in the variables x1, x2, x3 would behave the same if we changed the names ofthe variables to a, b, c and kept all the constants the same and in the same places. Inthis section, we will isolate the key bits of information about a system of equations intosomething called a matrix, and then use this matrix to systematically solve the equations.Along the way we will obtain one of our most important and useful computational tools.

Definition MMatrixAn m × n matrix is a rectangular layout of numbers from C having m rows and ncolumns. �

Notation MNMatrix NotationWe will use upper-case Latin letters from the start of the alphabet (A, B, C, . . . ) todenote matrices and squared-off brackets to delimit the layout. Many use large paren-theses instead of brackets — the distinction is not important. Rows of a matrix will bereferenced starting at the top and working down (i.e. row 1 is at the top) and columnswill be referenced starting from the left (i.e. column 1 is at the left). 5

Example RREF.AMA matrix

B =

−1 2 5 31 0 −6 1−4 2 2 −2

is a matrix with m = 3 rows and n = 4 columns. 4

A calculator or computer language can be a convenient way to perform calculations withmatrices. But first you have to enter the matrix. Here’s how it is done on variouscomputing platforms.

Computation Note ME.MMAMatrix Entry (Mathematica)Matrices are input as lists of lists, since a list is a basic data structure in Mathematica.A matrix is a list of rows, with each row entered as a list. Mathematica uses braces({ , }) to delimit lists. So the input

a = {{1, 2, 3, 4}, {5, 6, 7, 8}, {9, 10, 11, 12}}

Page 55: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 23

would create a 3× 4 matrix named a that is equal to1 2 3 45 6 7 89 10 11 12

To display a matrix named a “nicely” in Mathematica, type a//MatrixForm , and theoutput will be displayed with rows and columns. If you just type a , then you will geta list of lists, like how you input it in the first place. ⊗Computation Note ME.TI86Matrix Entry (TI-86)On the TI-86, press the MATRX key (Yellow-7) . Press the second menu key over,F2 , to bring up the EDIT screen. Give your matrix a name, one letter or many, thenpress ENTER . You can then change the size of the matrix (rows, then columns) and beginediting individual entries (which are initially zero). ENTER will move you from entry toentry, or the down arrow key will move you to the next row. A menu gives you extraoptions for editing.

Matrices may also be entered on the home screen as follows. Use brackets ([ , ]) toenclose rows with elements separated by commas. Group rows, in order, into a final setof brackets (with no commas between rows). This can then be stored in a name with theSTO key. So, for example,

[[1, 2, 3, 4] [5, 6, 7, 8] [9, 10, 11, 12]] → A

will create a matrix named A that is equal to1 2 3 45 6 7 89 10 11 12

Computation Note ME.TI83Matrix Entry (TI-83)

Contributed by Douglas PhelpsOn the TI-83, press the MATRX key. Press the right arrow key twice so that EDIT ishighlighted. Move the cursor down so that it is over the desired letter of the matrix andpress ENTER . For example, lets call our matrix B , so press the down arrow once andpress ENTER . To enter a 2× 3 matrix, press 2 ENTER 3 ENTER . To create the matrix[

1 2 34 5 6

]press 1 ENTER 2 ENTER 3 ENTER 4 ENTER 5 ENTER 6 ENTER . ⊗Definition AMAugmented MatrixSuppose we have a system of m equations in the n variables x1, x2, x3, . . . , xn written as

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm

Page 56: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 24

then the augmented matrix of the system of equations is the m× (n + 1) matrixa11 a12 a13 . . . a1n b1

a21 a22 a23 . . . a2n b2

a31 a32 a33 . . . a3n b3...

am1 am2 am3 . . . amn bm

The augmented matrix represents all the important information in the system of equa-tions, since the names of the variables have been ignored, and the only connection withthe variables is the location of their coefficients in the matrix. It is important to realizethat the augmented matrix is just that, a matrix, and not a system of equations. Inparticular, the augmented matrix does not have any “solutions,” though it will be usefulfor finding solutions to the system of equations that it is associated with. Notice toothat an augmented matrix always belongs to some system of equations, and vice versa.Here’s a quick example.

Example RREF.AMAAAugmented matrix for Archetype AArchetype A is the following system of 3 equations in 3 variables.

x1 − x2 + 2x3 = 1

2x1 + x2 + x3 = 8

x1 + x2 = 5

Here is its augmented matrix. 1 −1 2 12 1 1 81 1 0 5

4

An augmented matrix for a system of equations will save us the tedium of continuallywriting down the names of the variables as we solve the system. It will also release usfrom any dependence on the actual names of the variables. We have seen how certainoperations we can perform on equations (Definition EO [12]) will preserve their solutions(Theorem EOPSS [13]). The next two definitions and the following theorem carry overthese ideas to augmented matrices.

Definition RORow OperationsThe following three operations will transform an m× n matrix into a different matrix ofthe same size, and each is known as a row operation.

1. Swap the locations of two rows.

2. Multiply each entry of a single row by a nonzero quantity.

3. Multiply each entry of one row by some quantity, and add these values to theentry in the same column of a second row. Leave the first row the same after thisoperation, but replace the second row by the new values. �

Page 57: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 25

We will use a kind of shorthand to describe these operations:

1. Ri ↔ Rj: Swap the location of rows i and j.

2. αRi: Multiply row i by the nonzero scalar α.

3. αRi + Rj: Multiply row i by the scalar α and add to row j.

Definition REMRow-Equivalent MatricesTwo matrices, A and B, are row-equivalent if one can be obtained from the other bya sequence of row operations. �

Example RREF.TREMTwo row-equivalent matricesThe matrices

A =

2 −1 3 45 2 −2 31 1 0 6

B =

1 1 0 63 0 −2 −92 −1 3 4

are row-equivalent as can be seen from2 −1 3 4

5 2 −2 31 1 0 6

R1↔R3−−−−→

1 1 0 65 2 −2 32 −1 3 4

−2R1+R2−−−−−→

1 1 0 63 0 −2 −92 −1 3 4

We can also say that any pair of these three matrices are row-equivalent. 4

Notice that each of the three row operations is reversible, so we do not have to be carefulabout the distinction between “A is row-equivalent to B” and “B is row-equivalent to A.”The preceding definitions are designed to make the following theorem possible. It saysthat row-equivalent matrices represent systems of linear equations that have identicalsolution sets.

Theorem REMESRow-Equivalent Matrices represent Equivalent SystemsSuppose that A and B are row-equivalent augmented matrices. Then the systems oflinear equations that they represent are equivalent systems. �

Proof If we perform a single row operation on an augmented matrix, it will have thesame effect as if we did the analogous equation operation on the corresponding system ofequations. By exactly the same methods as we used in the proof of Theorem EOPSS [13]we can see that each of these row operations will preserve the set of solutions for thecorresponding system of equations. �

So at this point, our strategy is to begin with a system of equations, represent it byan augmented matrix, perform row operations (which will preserve solutions for thecorresponding systems) to get a “simpler” augmented matrix, convert back to a “simpler”system of equations and then solve that system, knowing that its solutions are those ofthe original system. Here’s a rehash of Example SSSLE.US [16] as an exercise in usingour new tools.

Page 58: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 26

Example RREF.USThree equations, one solutionWe solve the following system using augmented matrices and row operations.

x1 + 2x2 + 2x3 = 4

x1 + 3x2 + 3x3 = 5

2x1 + 6x2 + 5x3 = 6

Form the augmented matrix,

A =

1 2 2 41 3 3 52 6 5 6

and apply row operations,

−1R1+R2−−−−−→

1 2 2 40 1 1 12 6 5 6

−2R1+R3−−−−−→

1 2 2 40 1 1 10 2 1 −2

−2R2+R3−−−−−→

1 2 2 40 1 1 10 0 −1 −4

−1R3−−−→

1 2 2 40 1 1 10 0 1 4

So the matrix

B =

1 2 2 40 1 1 10 0 1 4

is row equivalent to A and by Theorem REMES [25] the system of equations below hasthe same solution set as the original system of equations.

x1 + 2x2 + 2x3 = 4

x2 + x3 = 1

x3 = 4

Solving this “simpler” system is straightforward and is identical to the process in Exam-ple SSSLE.US [16]. 4

The preceding example amply illustrates the definitions and theorems we have seen sofar. But it still leaves two questions unanswered. Exactly what is this “simpler” form fora matrix, and just how do we get it? Here’s the answer to the first question, a definitionof reduced row-echelon form.

Definition RREFReduced Row-Echelon FormA matrix is in reduced row-echelon form if it meets all of the following conditions:

Page 59: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 27

1. A row where every entry is zero is below any row containing a nonzero entry.

2. The leftmost nonzero entry of a row is equal to 1.

3. The leftmost nonzero entry of a row is the only nonzero entry in its column.

4. Consider any two different leftmost nonzero entries, in locations indexed as row i,column j and row s, column t. If s > i, then t > j. �

The principal feature of reduced row-echelon form is the pattern of leading 1’s guaranteedby conditions (2) and (4), reminiscent of a flight of geese, or steps in a staircase, or watercascading down a mountain stream. Because we will make frequent reference to reducedrow-echelon form, we make precise definitions of two terms.

Definition ZRMZero Row of a MatrixA row of a matrix where every entry is zero is called a zero row. �Definition LOLeading OnesFor a matrix in reduced row-echelon form, the leftmost nonzero entry of any row that isnot a zero row will be called a leading 1. �

Example RREF.RREFA matrix in reduced row-echelon formThe matrix C is in reduced row-echelon form.

1 −3 0 6 0 0 −5 90 0 0 0 1 0 3 −70 0 0 0 0 1 7 30 0 0 0 0 0 0 00 0 0 0 0 0 0 0

4

Example RREF.NRREFA matrix not in reduced row-echelon formThe matrix D is not in reduced row-echelon form, as it fails each of the four requirementsonce.

1 0 −3 0 6 0 7 −5 90 0 0 5 0 1 0 3 −70 0 0 0 0 0 0 0 00 0 0 0 0 0 1 7 30 1 0 0 0 0 0 −4 20 0 0 0 0 0 0 0 0

4

Proof Technique CConstructive ProofsConclusions of proofs come in a variety of types. Often a theorem will simply assertthat something exists. The best way, but not the only way, to show something existsis to actually build it. Such a proof is called constructive. The thing to realize aboutconstructive proofs is that the proof itself will contain a procedure that might be usedcomputationally to construct the desired object. If the procedure is not too cumbersome,then the proof itself is as useful as the statement of the theorem. Such is the case withour next theorem. ♦

Page 60: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 28

Theorem REMEFRow-Equivalent Matrix in Echelon FormSuppose A is a matrix. Then there is a (unique!) matrix B so that

1. A and B are row-equivalent.

2. B is in reduced row-echelon form. �

Proof Suppose that A has m rows. We will describe a process for converting A into Bvia row operations.

Set i = 1.

1. If i = m + 1, then stop converting the matrix.

2. Among all of the entries in rows i through m locate the leftmost nonzero entry(there may be several entries that tie for being leftmost). Denote the column ofthis entry by j. If this is not possible because all the entries are zero, then stopconverting the matrix.

3. If the nonzero entry found in the preceding step is not in row i, swap rows so thatrow i has a nonzero entry in column j.

4. Use the second row operation to multiply row i by the reciprocal of the value incolumn j, thereby creating a leading 1 in row i at column j.

5. Use row i and the third row operation to convert every other entry in column jinto a zero.

6. Increase i by one and return to step 1.

The result of this procedure is the matrix B. We need to establish that it has therequisite properties. First, the steps of the process only use row operations to convertthe matrix, so A and B are row-equivalent.

It is a bit more work to be certain that B is in reduced row-echelon form. At theconclusion of the i-th trip through the steps, we claim the first i rows form a matrix inreduced row-echelon form, and the entries in rows i + 1 through m in columns 1 throughj are all zero. To see this, notice that

1. The definition of j insures that the entries of rows i + 1 through m, in columns 1through j − 1 are all zero.

2. Row i has a leading nonzero entry equal to 1 by the result of step 4.

3. The employment of the leading 1 of row i in step 5 will make every element ofcolumn j zero in rows 1 through i + 1, as well as in rows i + 1 through m.

4. Rows 1 through i− 1 are only affected by step 5. The zeros in columns 1 throughj − 1 of row i mean that none of the entries in columns 1 through j − 1 for rows 1through i− 1 will change by the row operations employed in step 5.

5. Since columns 1 through j are are all zero for rows i + 1 through m, any nonzeroentry found on the next pass will be in a column to the right of column j, ensuringthat the fourth condition of reduced row-echelon form is met.

Page 61: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 29

6. If the procedure halts with i = m + 1, then every row of B has a leading 1, andhence has no zero rows. If the procedure halts because step 2 fails to find a nonzeroentry, then rows i through m are all zero rows, and they are all at the bottom ofthe matrix. �

So now we can put it all together. Begin with a system of linear equations (Defini-tion SSLE [10]), and represent it by its augmented matrix (Definition AM [23]). Use rowoperations (Definition RO [24]) to convert this matrix into reduced row-echelon form (Def-inition RREF [27]), using the procedure outlined in the proof of Theorem REMEF [28].Theorem REMEF [28] also tells us we can always accomplish this, and that the result isrow-equivalent (Definition REM [25]) to the original augmented matrix. Since the matrixin reduced-row echelon form has the same solution set, we can analyze it instead of theoriginal matrix, viewing it as the augmented matrix of a different system of equations.The beauty of augmented matrices in reduced row-echelon form is that the solution setsto their corresponding systems can be easily determined, as we will see in the next fewexamples and in the next section.

We will now run through some examples of using these definitions and theorems tosolve some systems of equations. From now on, when we have a matrix in reduced row-echelon form, we will mark the leading 1’s with a small box. In your work, you can boxthem, circle them or write them in a different color. This device will prove very usefullater and is a very good habit to develop now.

Example RREF.SABSolutions for Archetype BSolve the following system of equations.

−7x1 − 6x2 − 12x3 = −33

5x1 + 5x2 + 7x3 = 24

x1 + 4x3 = 5

Form the augmented matrix for starters,−7 −6 −12 −335 5 7 241 0 4 5

and work to reduced row-echelon form, first with i = 1,

R1↔R3−−−−→

1 0 4 55 5 7 24−7 −6 −12 −33

−5R1+R2−−−−−→

1 0 4 50 5 −13 −1−7 −6 −12 −33

7R1+R3−−−−→

1 0 4 50 5 −13 −10 −6 16 2

Now, with i = 2,

15R2−−→

1 0 4 50 1 −13

5−15

0 −6 16 2

6R2+R3−−−−→

1 0 4 5

0 1 −135

−15

0 0 25

45

Page 62: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 30

And finally, with i = 3,

52R3−−→

1 0 4 5

0 1 −135

−15

0 0 1 2

135

R3+R2−−−−−→

1 0 4 5

0 1 0 50 0 1 2

−4R3+R1−−−−−→

1 0 0 −3

0 1 0 5

0 0 1 2

This is now the augmented matrix of a very simple system of equations, namely x1 = −3,x2 = 5, x3 = 2, which has an obvious solution. Furthermore, we can see that this is theonly solution to this system, so we have determined the entire solution set. You mightcompare this example with the procedure we used in Example SSSLE.US [16]. 4

Archetypes A and B are meant to contrast each other in many respects. So lets solveArchetype A now.

Example RREF.SAASolutions for Archetype ASolve the following system of equations.

x1 − x2 + 2x3 = 1

2x1 + x2 + x3 = 8

x1 + x2 = 5

Form the augmented matrix for starters,1 −1 2 12 1 1 81 1 0 5

and work to reduced row-echelon form, first with i = 1,

−2R1+R2−−−−−→

1 −1 2 10 3 −3 61 1 0 5

−1R1+R3−−−−−→

1 −1 2 10 3 −3 60 2 −2 4

Now, with i = 2,

13R2−−→

1 −1 2 10 1 −1 20 2 −2 4

1R2+R1−−−−→

1 0 1 30 1 −1 20 2 −2 4

−2R2+R3−−−−−→

1 0 1 3

0 1 −1 20 0 0 0

The system of equations represented by this augmented matrix needs to be considered abit differently than that for Archetype B. First, the last row of the matrix is the equation0 = 0, which is always true, so we can safely ignore it as we analyze the other twoequations. These equations are,

x1 + x3 = 3

x2 − x3 = 2.

While this system is fairly easy to solve, it also appears to have a multitude of solutions.For example, choose x3 = 1 and see that then x1 = 2 and x2 = 3 will together form asolution. Or choose x3 = 0, and then discover that x1 = 3 and x2 = 2 lead to a solution.

Page 63: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 31

Try it yourself: pick any value of x3 you please, and figure out what x1 and x2 should beto make the first and second equations (respectively) true. We’ll wait while you do that.Because of this behavior, we say that x3 is a “free” or “independent” variable. But whydo we vary x3 and not some other variable? For now, notice that the third column ofthe augmented matrix does not have any leading 1’s in its column. With this idea, wecan rearrange the two equations, solving each for the variable that corresponds to theleading 1 in that row.

x1 = 3− x3

x2 = 2 + x3

To write the solutions in set notation, we have

S = {(3− x3, 2 + x3, x3) | x3 ∈ C}

We’ll learn more in the next section about systems with infinitely many solutions and howto express their solution sets. Right now, you might look back at Example SSSLE.IS [17].4

Example RREF.SAESolutions for Archetype ESolve the following system of equations.

2x1 + x2 + 7x3 − 7x4 = 2

−3x1 + 4x2 − 5x3 − 6x4 = 3

x1 + x2 + 4x3 − 5x4 = 2

Form the augmented matrix for starters,

2 1 7 −7 2−3 4 −5 −6 31 1 4 −5 2

and work to reduced row-echelon form, first with i = 1,

R1↔R3−−−−→

1 1 4 −5 2−3 4 −5 −6 32 1 7 −7 2

3R1+R2−−−−→

1 1 4 −5 20 7 7 −21 92 1 7 −7 2

−2R1+R3−−−−−→

1 1 4 −5 20 7 7 −21 90 −1 −1 3 −2

Page 64: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 32

Now, with i = 2,

R2↔R3−−−−→

1 1 4 −5 20 −1 −1 3 −20 7 7 −21 9

−1R2−−−→

1 1 4 −5 20 1 1 −3 20 7 7 −21 9

−1R2+R1−−−−−→

1 1 4 −5 20 1 1 −3 20 7 7 −21 9

−7R2+R3−−−−−→

1 1 4 −5 2

0 1 1 −3 20 0 0 0 −5

And finally, with i = 3,

− 15R3−−−→

1 1 4 −5 2

0 1 1 −3 20 0 0 0 1

−2R3+R2−−−−−→

1 1 4 −5 2

0 1 1 −3 00 0 0 0 1

−2R3+R1−−−−−→

1 1 4 −5 0

0 1 1 −3 0

0 0 0 0 1

Lets analyze the equations in the system represented by this augmented matrix. Thethird equation will read 0 = 1. This is patently false, all the time. No choice of valuesfor our variables will ever make it true. We’re done. Since we cannot even make the lastequation true, we have no hope of making all of the equations simultaneously true. Sothis system has no solutions, and its solution set is the empty set, ∅ = { }.

Notice that we could have reached this conclusion sooner. After performing therow operation −7R2 + R3, we can see that the third equation reads 0 = −5, a falsestatement. Since the system represented by this matrix has no solutions, none of thesystems represented has any solutions. However, for this example, we have chosen tobring the matrix fully to reduced row-echelon form for the practice. 4

These three examples (Example RREF.SAB [29], Example RREF.SAA [30], Exam-ple RREF.SAE [31]) illustrate the full range of possibilities for a system of linear equa-tions — no solutions, one solution, or infinitely many solutions. In the next section we’llexamine these three scenarios more closely.

After some practice by hand, you will want to use your favorite computing device todo the computations required to bring a matrix to reduced row-echelon form.

Computation Note RR.MMARow Reduce (Mathematica)If a is the name of a matrix in Mathematica, then the command RowReduce[a] will

output the reduced row-echelon form of the matrix. ⊗

Page 65: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RREF Reduced Row-Echelon Form 33

Computation Note RR.TI86Row Reduce (TI-86)If A is the name of a matrix stored in the TI-86, then the command rref A will

return the reduced row-echelon form of the matrix. This command can also be found bypressing the MATRX key, then F4 for OPS , and finally, F5 for rref . ⊗

Computation Note RR.TI83Row Reduce (TI-83)

Contributed by Douglas PhelpsSuppose B is the name of a matrix stored in the TI-83. Press the MATRX key. Press theright arrow key once so that MATH is highlighted. Press the down arrow eleven times sothat rref ( is highlighted, then press ENTER . to choose the matrix B , press MATRX ,then the down arrow once followed by ENTER . Supply a right parenthesis ( ) ) and pressENTER . ⊗

Page 66: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RREF.EXC Exercises 34

Subsection EXCExercises

For problems C10–C12, find all solutions to the system of linear equations. Write thesolutions as a set, using correct set notation. C10 Contributed by Robert Beezer

2x1 − 3x2 + x3 + 7x4 = 14

2x1 + 8x2 − 4x3 + 5x4 = −1

x1 + 3x2 − 3x3 = 4

−5x1 + 2x2 + 3x3 + 4x4 = −19

Solution [35]

C11 Contributed by Robert Beezer

3x1 + 4x2 − x3 + 2x4 = 6

x1 − 2x2 + 3x3 + x4 = 2

10x2 − 10x3 − x4 = 1

Solution [35]

C12 Contributed by Robert Beezer

2x1 + 4x2 + 5x3 + 7x4 = −26

x1 + 2x2 + x3 − x4 = −4

−2x1 − 4x2 + x3 + 11x4 = −10

Solution [35]

C30 Contributed by Robert BeezerRow-reduce the matrix below without the aid of a calculator, indicating the row opera-tions you are using at each step. 2 1 5 10

1 −3 −1 −24 −2 6 12

Solution [36]

M50 Contributed by Robert BeezerA parking lot has 66 vehicles (cars, trucks, motorcycles and bicycles) in it. There

are four times as many cars as trucks. The total number of tires (4 per car or truck,2 per motorcycle or bicycle) is 252. How many cars are there? How many bicycles?Solution [36]

TODO: Row-equivalent is an equivalence relation

Page 67: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RREF.SOL Solutions 35

Subsection SOLSolutions

C10 Exercise [34] Contributed by Robert BeezerThe augmented matrix row-reduces to

1 0 0 0 1

0 1 0 0 −3

0 0 1 0 −4

0 0 0 1 1

and we see from the locations of the leading 1’s that the system is consistent (Theo-rem RCLS [41]) and that n− r = 4−4 = 0 and so the system has no free variables (The-orem CSRN [42]) and hence has a unique solution. This solution is {(1, −3, −4, 1)}.

C11 Exercise [34] Contributed by Robert BeezerThe augmented matrix row-reduces to

1 0 1 4/5 0

0 1 −1 −1/10 0

0 0 0 0 1

and a leading 1 in the last column tells us that the system is inconsistent (Theo-rem RCLS [41]). So the solution set is ∅ = {}.

C12 Exercise [34] Contributed by Robert BeezerThe augmented matrix row-reduces to

1 2 0 −4 2

0 0 1 3 −60 0 0 0 0

(Theorem RCLS [41]) and (Theorem CSRN [42]) tells us the system is consistent andthe solution set can be described with n − r = 4 − 2 = 2 free variables, namely x2 andx4. Solving for the dependent variables (D = {x1, x3}) the first and second equationsrepresented in the row-reduced matrix yields,

x1 = 2− 2x2 + 4x4

x3 = −6 − 3x4

As a set, we write this as

{(2− 2x2 + 4x4, x2, −6− 3x4, x4) | x2, x4 ∈ C}

C30 Exercise [34] Contributed by Robert Beezer

Page 68: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RREF.SOL Solutions 36

2 1 5 101 −3 −1 −24 −2 6 12

R1↔R2−−−−→

1 −3 −1 −22 1 5 104 −2 6 12

−2R1+R2−−−−−→

1 −3 −1 −20 7 7 144 −2 6 12

−4R1+R3−−−−−→

1 −3 −1 −20 7 7 140 10 10 20

17R2−−→

1 −3 −1 −20 1 1 20 10 10 20

3R2+R1−−−−→

1 0 2 40 1 1 20 10 10 20

−10R2+R3−−−−−−→

1 0 2 4

0 1 1 20 0 0 0

M50 Exercise [34] Contributed by Robert BeezerLet c, t, m, b denote the number of cars, trucks, motorcycles, and bicycles. Then the

statements from the problem yield the equations:

c + t + m + b = 66

c− 4t = 0

4c + 4t + 2m + 2b = 252

The augmented matrix for this system is1 1 1 1 661 −4 0 0 04 4 2 2 252

which row-reduces to 1 0 0 0 48

0 1 0 0 12

0 0 1 1 6

c = 48 is the first equation represented in the row-reduced matrix so there are 48 cars.m + b = 6 is the third equation represented in the row-reduced matrix so there areanywhere from 0 to 6 bicycles. We can also say that b is a free variable, but the contextof the problem limits it to 7 integer values since cannot have a negative number ofmotorcycles.

Page 69: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 37

Section TSS

Types of Solution Sets

We will now be more careful about analyzing the reduced row-echelon form derived fromthe augmented matrix of a system of linear equations. In particular, we will see how tosystematically handle the situation when we have infinitely many solutions to a system,and we will prove that every system of linear equations has either zero, one or infinitelymany solutions. With these tools, we will be able to solve any system by a well-describedmethod.

The computer scientist Donald Knuth said, “Science is what we understand wellenough to explain to a computer. Art is everything else.” In this section we’ll removesolving systems of equations from the realm of art, and into the realm of science. Webegin with a definition.

Definition CSConsistent SystemA system of linear equations is consistent if it has at least one solution. Otherwise, thesystem is called inconsistent. �

We will want to first recognize when a system is inconsistent or consistent, and in thecase of consistent systems we will be able to further refine the types of solutions possible.We will do this by analyzing the reduced row-echelon form of a matrix, so we now initiatesome notation that will help us talk about this form of a matrix.

Notation RREFAReduced Row-Echelon Form AnalysisSuppose that B is an m× n matrix that is in reduced row-echelon form. Let r equal thenumber of rows of B that are not zero rows. Each of these r rows then contains a leading1, so let di equal the column number where row i’s leading 1 is located. For columnswithout a leading 1, let fi be the column number of the i-th column (reading from leftto right) that does not contain a leading 1. Let

D = {d1, d2, d3, . . . , dr} F = {f1, f2, f3, . . . , fn−r} 5

This notation can be a bit confusing, since we have subscripted variables that are in turnequal to subscripts used to index the matrix. However, many questions about matricesand systems of equations can be answered once we know r, D and F . An example mayhelp.

Example TSS.RREFNReduced row-echelon form notationFor the 5× 8 matrix

B =

1 5 0 0 2 8 0 5 −1

0 0 1 0 4 7 0 2 0

0 0 0 1 3 9 0 3 −6

0 0 0 0 0 0 1 4 20 0 0 0 0 0 0 0

Page 70: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 38

in reduced row-echelon form we have

r = 4

d1 = 1 d2 = 3 d3 = 4 d4 = 7

f1 = 2 f2 = 5 f3 = 6 f4 = 8 f5 = 9.

Notice that the sets D = {d1, d2, d3, d4} = {1, 3, 4, 7} and F = {f1, f2, f3, f4, f5} ={2, 5, 6, 8, 9} have nothing in common and together account for all of the columns of B(we say it is a partition of the set of column indices). 4

Before proving some theorems about the possibilities for solution sets to systems ofequations, lets analyze one particular system with an infinite solution set very carefullyas an example. We’ll use this technique frequently, and shortly we’ll refine it slightly.

Archetypes I and J are both fairly large for doing computations by hand (though notimpossibly large). Their properties are very similar, so we will frequently analyze thesituation in Archetype I, and leave you the joy of analyzing Archetype J yourself. Sowork through Archetype I with the text, by hand and/or with a computer, and thentackle Archetype J yourself (and check your results with those listed). Notice too thatthe archetypes describing systems of equations each lists the values of r, D and F . Herewe go. . .

Example TSS.ISSIDescribing infinite solution sets, Archetype IArchetype I [414] is the system of m = 4 equations in n = 7 variables

x1 + 4x2 − x4 + 7x6 − 9x7 = 3

2x1 + 8x2 − x3 + 3x4 + 9x5 − 13x6 + 7x7 = 9

2x3 − 3x4 − 4x5 + 12x6 − 8x7 = 1

−x1 − 4x2 + 2x3 + 4x4 + 8x5 − 31x6 + 37x7 = 4

has a 4×8 augmented matrix that is row-equivalent to the following matrix (check this!),and which is in reduced row-echelon form (the existence of this matrix is guaranteed byTheorem REMEF [28]),

1 4 0 0 2 1 −3 4

0 0 1 0 1 −3 5 2

0 0 0 1 2 −6 6 10 0 0 0 0 0 0 0

.

So we find that r = 3 and

D = {d1, d2, d3} = {1, 3, 4} F = {f1, f2, f3, f4, f5} = {2, 5, 6, 7, 8} .

Let i denote one of the r = 3 non-zero rows, and then we see that we can solve thecorresponding equation represented by this row for the variable xdi

and write it as alinear function of the variables xf1 , xf2 , xf3 , xf4 (notice that f5 = 8 does not reference avariable). We’ll do this now, but you can already see how the subscripts upon subscriptstakes some getting used to.

(i = 1) xd1 = x1 = 4− 4x2 − 2x5 − x6 + 3x7

(i = 2) xd2 = x3 = 2− x5 + 3x6 − 5x7

(i = 3) xd3 = x4 = 1− 2x5 + 6x6 − 6x7

Page 71: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 39

Each element of the set F = {f1, f2, f3, f4, f5} = {2, 5, 6, 7, 8} is the index of a variable,except for f5 = 8. We refer to xf1 = x2, xf2 = x5, xf3 = x6 and xf4 = x7 as “free” (or“independent”) variables since they are allowed to assume any possible combination ofvalues that we can imagine and we can continue on to build a solution to the system bysolving individual equations for the values of the other (“dependent”) variables.

Each element of the set D = {d1, d2, d3} = {1, 3, 4} is the index of a variable. Werefer to the variables xd1 = x1, xd2 = x3 and xd3 = x4 as “dependent” variables since theydepend on the independent variables. More precisely, for each possible choice of valuesfor the independent variables we get exactly one set of values for the dependent variablesthat combine to form a solution of the system.

To express the solutions as a set with elements that are 7-tuples, we write

{(4− 4x2 − 2x5 − x6 + 3x7, x2, 2− x5 + 3x6 − 5x7, 1− 2x5 + 6x6 − 6x7, x5, x6, x7) | x2, x5, x6, x7 ∈ C}

The condition that x2, x5, x6, x7 ∈ C is how we specify that the variables x2, x5, x6, x7

are “free” to assume any possible values.This systematic approach to solving a system of equations will allow us to create a

precise description of the solution set for any consistent system once we have found thereduced row-echelon form of the augmented matrix. It will work just as well when theset of free variables is empty and we get just a single solution. And we could program acomputer to do it! Now have a whack at Archetype J, mimicking the discussion in thisexample. We’ll still be here when you get back. 4Sets are an important part of algebra, and we’ve seen a few already. Being comfortablewith sets is important for understanding and writing proofs. So here’s another prooftechnique.

Proof Technique SNSet NotationSets are typically written inside of braces, as { }, and have two components. The firstis a description of the type of objects contained in a set, while the second is some sortof restriction on the properties the objects have. Every object in the set must be ofthe type described in the first part and it must satisfy the restrictions in the secondpart. Conversely, any object of the proper type for the first part, that also meets theconditions of the second part, will be in the set. These two parts are set off from eachother somehow, often with a vertical bar (|) or a colon (:). Membership of an element ina set is denoted with the symbol ∈.

I like to think of sets as clubs. The first part is some description of the type of peoplewho might belong to the club, the basic objects. For example, a bicycle club woulddescribe its members as being people who like to ride bicycles. The second part is like amembership committee, it restricts the people who are allowed in the club. Continuingwith our bicycle club, we might decide to limit ourselves to “serious” riders and onlyhave members who can document having ridden 100 kilometers or more in a single dayat least one time.

The restrictions on membership can migrate around some between the first and secondpart, and there may be several ways to describe the same set of objects. Here’s a moremathematical example, employing the set of all integers, Z, to describe the set of evenintegers.

E = {x ∈ Z | x is an even number} = {x ∈ Z | 2 divides x evenly} = {2k | k ∈ Z} .

Page 72: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 40

Notice how this set tells us that its objects are integer numbers (not, say, matrices orfunctions, for example) and just those that are even. So we can write that 10 ∈ E, while17 6∈ E once we check the membership criteria. We also recognize the question[

1 −3 52 0 3

]∈ E?

as being ridiculous. ♦

We mix our metaphors a bit when we call variables free versus dependent. Maybe weshould call dependent variables “enslaved”? Here’s the definition.

Definition IDVIndependent and Dependent VariablesSuppose A is the augmented matrix of a system of linear equations and B is a row-equivalent matrix in reduced row-echelon form. Suppose j is the number of a columnof B that contains the leading 1 for some row, and it is not the last column. Then thevariable j is dependent. A variable that is not dependent is called independent orfree. �

We can now use the values of m, n, r, and the independent and dependent variables tocategorize the solutions sets to linear systems through a sequence of theorems. First thedistinction between consistent and inconsistent systems, after two explanations of someproof techniques we will be using.

Proof Technique EEquivalencesWhen a theorem uses the phrase “if and only if” (or the abbreviation “iff”) it is ashorthand way of saying that two if-then statements are true. So if a theorem says “A ifand only if B,” then it is true that “if A, then B” while it is also true that “if B, then A.”For example, it may be a theorem that “I wear bright yellow knee-high plastic boots ifand only if it is raining.” This means that I never forget to wear my super-duper yellowboots when it is raining and I wouldn’t be seen in such silly boots unless it was raining.You never have one without the other. I’ve got my boots on and it is raining or I don’thave my boots on and it is dry.

The upshot for proving such theorems is that it is like a 2-for-1 sale, we get to do twoproofs. Assume A and conclude B, then start over and assume B and conclude A. Forthis reason, “if and only if” is sometimes abbreviated by ⇐⇒ , while proofs indicatewhich of the two implications is being proved by prefacing each with ⇒ or ⇐. A carefullywritten proof will remind the reader which statement is being used as the hypothesis, aquicker version will let the reader deduce it from the direction of the arrow. Traditiondictates we do the “easy” half first, but that’s hard for a student to know until you’vefinished doing both halves! Oh well, if you rewrite your proofs (a good habit), you canthen choose to put the easy half first.

Theorems of this type are called equivalences or characterizations, and they are someof the most pleasing results in mathematics. They say that two objects, or two situations,are really the same. You don’t have one without the other, like rain and my yellowboots. The more different A and B seem to be, the more pleasing it is to discover theyare really equivalent. And if A describes a very mysterious solution or involves a toughcomputation, while B is transparent or involves easy computations, then we’ve founda great shortcut for better understanding or faster computation. Remember that every

Page 73: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 41

theorem really is a shortcut in some form. You will also discover that if proving A ⇒ Bis very easy, then proving B ⇒ A is likely to be proportionately harder. Sometimes thetwo halves are about equally hard. And in rare cases, you can string together a wholesequence of other equivalences to form the one you’re after and you don’t even need todo two halves. In this case, the argument of one half is just the argument of the otherhalf, but in reverse.

One last thing about equivalences. If you see a statement of a theorem that says twothings are “equivalent,” translate it first into an “if and only if” statement. ♦

Proof Technique CPContrapositivesThe contrapositive of an implication A ⇒ B is the implication not(B) ⇒ not(A),where “not” means the logical negation, or opposite. An implication is true if and only ifits contrapositive is true. In symbols, (A ⇒ B) ⇐⇒ (not(B) ⇒ not(A)) is a theorem.Such statements about logic, that are always true, are known as tautologies.

For example, it is a theorem that “if a vehicle is a fire truck, then it has big tiresand has a siren.” (Yes, I’m sure you can conjure up a counterexample, but play alongwith me anyway.) The contrapositive is “if a vehicle does not have big tires or does nothave a siren, then it is not a fire truck.” Notice how the “and” became an “or” when wenegated the conclusion of the original theorem.

It will frequently happen that it is easier to construct a proof of the contrapositivethan of the original implication. If you are having difficulty formulating a proof of someimplication, see if the contrapositive is easier for you. The trick is to construct thenegation of complicated statements accurately. More on that later. ♦

Theorem RCLSRecognizing Consistency of a Linear SystemSuppose A is the augmented matrix of a system of linear equations with m equations inn variables. Suppose also that B is a row-equivalent matrix in reduced row-echelon formwith r rows that are not zero rows. Then the system of equations is inconsistent if andonly if the leading 1 of row r is located in column n + 1 of B. �

Proof (⇐) The first half of the proof begins with the assumption that the leading 1 ofrow r is located in column n + 1 of B. Then row r of B begins with n consecutive zeros,finishing with the leading 1. This is a representation of the equation 0 = 1, which is false.Since this equation is false for any collection of values we might choose for the variables,there are no solutions for the system of equations, and it is inconsistent.

(⇒) For the second half of the proof, we wish to show that if we assume the system isinconsistent, then the final leading 1 is located in the last column. But instead of provingthis directly, we’ll form the logically equivalent statement that is the contrapositive, andprove that instead (see Technique CP [41]). Turning the implication around, and negatingeach portion, we arrive at the equivalent statement: If the leading 1 of row r is not incolumn n + 1, then the system of equations is consistent.

If the leading 1 for row i is located somewhere in columns 1 through n, then everypreceding row’s leading 1 is also located in columns 1 through n. In other words, sincethe last leading 1 is not in the last column, no leading 1 for any row is in the last column,due to the echelon layout of the leading 1’s. Let bi,n+1, 1 ≤ i ≤ r denote the entriesof the last column of B for the first r rows. Employ our notation for columns of thereduced row-echelon form of a matrix (see Notation RREFA [37]) to B and set xfi

= 0,

Page 74: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 42

1 ≤ i ≤ n − r and then set xdi= bi,n+1, 1 ≤ i ≤ r. These values for the variables make

the equations represented by the first r rows all true (convince yourself of this). Rowsr + 1 through m (if any) are all zero rows, hence represent the equation 0 = 0 and arealso all true. We have now identified one solution to the system, so we can say it isconsistent. �

The beauty of this theorem being an equivalence is that we can unequivocally test to seeif a system is consistent or inconsistent by looking at just a single entry of the reducedrow-echelon form matrix. We could program a computer to do it!

Notice that for a consistent system the row-reduced augmented matrix has n+1 ∈ F ,so the largest element of F does not refer to a variable. Also, for an inconsistent system,n+1 ∈ D, and it then does not make much sense to discuss whether or not variables arefree or dependent since there is no solution.

Theorem ICRNInconsistent Systems, r and nSuppose A is the augmented matrix of a system of linear equations with m equations inn variables. Suppose also that B is a row-equivalent matrix in reduced row-echelon formwith r rows that are not completely zeros. If r = n + 1, then the system of equations isinconsistent. �

Proof If r = n + 1, then D = {1, 2, 3, . . . , n, n + 1} and every column of B containsa leading 1. In particular, the entry of column n + 1 for row r = n + 1 is a leading 1.Theorem RCLS [41] then says that the system is inconsistent. �

Theorem CSRNConsistent Systems, r and nSuppose A is the augmented matrix of a consistent system of linear equations with mequations in n variables. Suppose also that B is a row-equivalent matrix in reduced row-echelon form with r rows that are not zero rows. Then r ≤ n. If r = n, then the systemhas a unique solution, and if r < n, then the system has infinitely many solutions. �

Proof This theorem contains three implications that we must establish. Notice first thatthe echelon layout of the leading 1’s means that there are at most n + 1 leading 1’s andtherefore r ≤ n + 1 for any system. We are assuming this system is consistent, so weknow by Theorem ICRN [42] that r 6= n + 1. Together these two observations leave uswith r ≤ n.

When r = n, we find n − r = 0 free variables (i.e. F = {n + 1}) and any solutionmust equal the unique solution given by the first n entries of column n + 1 of B.

When r < n, we have n − r > 0 free variables, corresponding to columns of Bwithout a leading 1, excepting the final column, which also does not contain a leading1 by Theorem RCLS [41]. By varying the values of the free variables suitably, we candemonstrate infinitely many solutions. �

Corollary FVCSFree Variables for Consistent SystemsSuppose A is the augmented matrix of a consistent system of linear equations with mequations in n variables. Suppose also that B is a row-equivalent matrix in reducedrow-echelon form with r rows that are not completely zeros. Then the solution set canbe described with n− r free variables. �

Page 75: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 43

Proof Technique CVConversesThe converse of the implication A ⇒ B is the implication B ⇒ A. There is no guaranteethat the truth of these two statements are related. In particular, if an implication hasbeen proven to be a theorem, then do not try to use its converse too as if it were a theorem.Sometimes the converse is true (and we have an equivalence, see Technique E [40]). Butmore likely the converse is false, especially if it wasn’t included in the statement of theoriginal theorem.

For example, we have the theorem, “if a vehicle is a fire truck, then it is has big tiresand has a siren.” The converse is false. The statement that “if a vehicle has big tiresand a siren, then it is a fire truck” is false. A police vehicle for use on a sandy publicbeach would have big tires and a siren, yet is not equipped to fight fires.

We bring this up now, because Theorem CSRN [42] has a tempting converse. Doesthis theorem say that if r < n, then the system is consistent? Definitely not, asArchetype E [397] has r = 2 and n = 4 but is inconsistent. This example is then said tobe a counterexample to the converse. Whenever you think a theorem that is an impli-cation might actually be an equivalence, it is good to hunt around for a counterexamplethat shows the converse to be false. ♦

Example TSS.CFVCounting free variablesFor each archetype that is a system of equations, the values of n and r are listed. Manyalso contain a few sample solutions. We can use this information profitably, as illustratedby four examples.

1. Archetype A [379] has n = 3 and r = 2. It can be seen to be consistent by thesample solutions given. Its solution set then has n − r = 1 free variables, andtherefore will be infinite.

2. Archetype B [384] has n = 3 and r = 3. It can be seen to be consistent by thesingle sample solution given. Its solution set then has n− r = 0 free variables, andtherefore will have just the single solution.

3. Archetype H [410] has n = 2 and r = 3. In this case, r = n + 1, so Theo-rem ICRN [42] says the system is inconsistent. We should not try to apply Corol-lary FVCS [42] to count free variables, since the theorem only applies to consistentsystems. (What would happen if you did?)

4. Archetype E [397] has n = 4 and r = 3. However, by looking at the reduced row-echelon form of the augmented matrix, we find a leading 1 in row 3, column 4. ByTheorem RCLS [41] we recognize the system is then inconsistent. (Why doesn’tthis example contradict Theorem ICRN [42]?) 4

We have accomplished a lot so far, but our main goal has been the following theorem,which is now very simple to prove. The proof is so simple that we ought to call it acorollary, but the result is important enough that it deserves to be called a theorem.Notice that this theorem was presaged first by Example SSSLE.TTS [11] and furtherforeshadowed by other examples.

Page 76: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 44

Theorem PSSLSPossible Solution Sets for Linear SystemsA simultaneous system of linear equations has no solutions, a unique solution or infinitelymany solutions. �

Proof By definition, a system is either inconsistent or consistent. The first case describessystems with no solutions. For consistent systems, we have the remaining two possibilitiesas guaranteed by, and described in, Theorem CSRN [42]. �

We have one more theorem to round out our set of tools for determining solution sets tosystems of linear equations.

Theorem CMVEIConsistent, More Variables than Equations, Infinite solutionsSuppose a consistent system of linear equations has m equations in n variables. If n > m,then the system has infinitely many solutions. �

Proof Suppose that the augmented matrix of the system of equations is row-equivalentto B, a matrix in reduced row-echelon form with r nonzero rows. Because B has m rowsin total, the number that are nonzero rows is less. In other words, r ≤ m. Follow thiswith the hypothesis that n > m and we find that the system has a solution set describedby at least one free variable because

n− r ≥ n−m > 0.

A consistent system with free variables will have an infinite number of solutions, as givenby Theorem CSRN [42]. �

Notice that to use this theorem we need only know that the system is consistent, togetherwith the values of m and n. We do not necessarily have to compute a row-equivalentreduced row-echelon form matrix, even though we discussed such a matrix in the proof.This is the substance of the following example.

Example TSS.OSGMDOne solution gives many, Archetype DArchetype D is the system of m = 3 equations in n = 4 variables,

2x1 + x2 + 7x3 − 7x4 = 8

−3x1 + 4x2 − 5x3 − 6x4 = −12

x1 + x2 + 4x3 − 5x4 = 4

and the solution x1 = 0, x2 = 1, x3 = 2, x4 = 1 can be checked easily by substitution.Having been handed this solution, we know the system is consistent. This, together withn > m, allows us to apply Theorem CMVEI [44] and conclude that the system hasinfinitely many solutions. 4

These theorems give us the procedures and implications that allow us to completely solveany simultaneous system of linear equations. The main computational tool is using rowoperations to convert an augmented matrix into reduced row-echelon form. Here’s abroad outline of how we would instruct a computer to solve a system of linear equations.

Page 77: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 45

1. Represent a system of linear equations by an augmented matrix (an array is theappropriate data structure in most computer languages).

2. Convert the matrix to a row-equivalent matrix in reduced row-echelon form usingthe procedure from the proof of Theorem REMEF [28].

3. Determine r and locate the leading 1 of row r. If it is in column n + 1, output thestatement that the system is inconsistent and halt.

4. With the leading 1 of row r not in column n + 1, there are two possibilities:

(a) r = n and the solution is unique. It can be read off directly from the entriesin rows 1 through n of column n + 1.

(b) r < n and there are infinitely many solutions. If only a single solution isneeded, set all the free variables to zero and read off the dependent vari-able values from column n + 1, as in the second half of the proof of Theo-rem RCLS [41]. If the entire solution set is required, figure out some nicecompact way to describe it, since your finite computer is not big enough tohold all the solutions (we’ll have such a way soon).

The above makes it all sound a bit simpler than it really is. In practice, row operationsemploy division (usually to get a leading entry of a row to convert to a leading 1) andthat will introduce round-off errors. Entries that should be zero sometimes end up beingvery, very small nonzero entries, or small entries lead to overflow errors when used asdivisors. A variety of strategies can be employed to minimize these sorts of errors, andthis is one of the main topics in the important subject known as numerical linear algebra.

Computation Note LS.MMALinear Solve (Mathematica)Mathematica will solve a linear system of equations using the LinearSolve[ ] com-

mand. The inputs are a matrix with the coefficients of the variables (but not the columnof constants), and a list containing the constant terms of each equation. This will look abit odd, since the lists in the matrix are rows, but the column of constants is also inputas a list and so looks like a row rather than a column. The result will be a single solution(even if there are infinitely many), reported as a list, or the statement that there is nosolution. When there are infinitely many, the single solution reported is exactly thatsolution used in the proof of Theorem RCLS [41], where the free variables are all set tozero, and the dependent variables come along with values from the final column of therow-reduced matrix.

As an example, Archetype A [379] is

x1 − x2 + 2x3 = 1

2x1 + x2 + x3 = 8

x1 + x2 = 5

To ask Mathematica for a solution, enter

LinearSolve[ {{1, −1, 2}, {2, 1, 1}, {1, 1, 0}}, {1, 8, 5} ]

Page 78: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section TSS Types of Solution Sets 46

and you will get back the single solution

{3, 2, 0}

We will see later how to coax Mathematica into giving us infinitely many solutions forthis system. ⊗

In this section we’ve gained a foolproof procedure for solving any system of linear equa-tions, no matter how many equations or variables. We also have a handful of theoremsthat allow us to determine partial information about a solution set without actuallyconstructing the whole set itself.

Page 79: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection TSS.EXC Exercises 47

Subsection EXCExercises

For Exercises M50–M52 say as much as possible about each system’s solution set. Besure to make it clear which theorems you are using to reach your conclusions. M50Contributed by Robert BeezerA homogeneous system of 8 equations in 8 variables. Solution [48]

M51 Contributed by Robert BeezerA consistent system of 8 equations in 6 variables. Solution [48]

M52 Contributed by Robert BeezerA consistent system of 6 equations in 8 variables. Solution [48]

Page 80: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection TSS.SOL Solutions 48

Subsection SOLSolutions

M50 Exercise [47] Contributed by Robert BeezerSince the system is homogeneous, we know it has the trivial solution (Theorem HSC [50]).We cannot say anymore based on the information provided, except to say that thereis either a unique solution or infinitely many solutions (Theorem PSSLS [44]). SeeArchetype A [379] and Archetype B [384] to understand the possibilities.

M51 Exercise [47] Contributed by Robert BeezerConsistent means there is at least one solution (Definition CS [37]). It will have either

a unique solution or infinitely many solutions (Theorem PSSLS [44]).

M52 Exercise [47] Contributed by Robert BeezerWith 6 rows in the augmented matrix, the row-reduced version will have r ≤ 6. Since

the system is consistent, apply Theorem CSRN [42] to see that n−r ≥ 2 implies infinitelymany solutions.

Page 81: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section HSE Homogeneous Systems of Equations 49

Section HSE

Homogeneous Systems of Equations

In this section we specialize to systems of linear equations where every equation has azero as its constant term. Along the way, we will begin to express more and more ideasin the language of matrices and begin a move away from writing out whole systems ofequations. The ideas initiated in this section will carry through the remainder of thecourse.

Subsection SHSSolutions of Homogeneous Systems

As usual, we begin with a definition.

Definition HSHomogeneous SystemA system of linear equations is homogeneous if each equation has a 0 for its constantterm. Such a system then has the form,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = 0

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = 0

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = 0

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = 0 �

Example HSE.AHSACArchetype C as a homogeneous systemFor each archetype that is a system of equations, we have formulated a similar, yet

different, homogeneous system of equations by replacing each equation’s constant termwith a zero. To wit, for Archetype C, we can convert the original system of equationsinto the homogeneous system,

2x1 − 3x2 + x3 − 6x4 = 0

4x1 + x2 + 2x3 + 9x4 = 0

3x1 + x2 + x3 + 8x4 = 0

Can you quickly find a solution to this system without row-reducing the augmentedmatrix? 4

As you might have discovered by studying Example HSE.AHSAC [49], setting each vari-able to zero will always be a solution of a homogeneous system. This is the substance ofthe following theorem.

Page 82: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection HSE.SHS Solutions of Homogeneous Systems 50

Theorem HSCHomogeneous Systems are ConsistentSuppose that a system of linear equations is homogeneous. Then it is consistent. �

Proof Set each variable of the system to zero. When substituting these values into eachequation, the left-hand side evaluates to zero, no matter what the coefficients are. Sincea homogeneous system has zero on the right-hand side of each equation as the constantterm, each equation is true. With one demonstrated solution, we can call the systemconsistent. �

Since this solution is so obvious, we now define it as the trivial solution.

Definition TSHSETrivial Solution to Homogeneous Systems of EquationsSuppose a homogeneous system of linear equations has n variables. The solution x1 = 0,x2 = 0,. . . , xn = 0 is called the trivial solution. �

Here are three typical examples, which we will reference throughout this section. Workthrough the row operations as we bring each to reduced row-echelon form. Also noticewhat is similar in each example, and what differs.

Example HSE.HUSABHomogeneous, unique solution, Archetype BArchetype B can be converted to the homogeneous system,

−11x1 + 2x2 − 14x3 = 0

23x1 − 6x2 + 33x3 = 0

14x1 − 2x2 + 17x3 = 0

whose augmented matrix row-reduces to 1 0 0 0

0 1 0 0

0 0 1 0

By Theorem HSC [50], the system is consistent, and so the computation n−r = 3−3 = 0means the solution set contains just a single solution. Then, this lone solution must bethe trivial solution. 4

Example HSE.HISAAHomogeneous, infinite solutions, Archetype AArchetype A [379] can be converted to the homogeneous system,

x1 − x2 + 2x3 = 0

2x1 + x2 + x3 = 0

x1 + x2 = 0

whose augmented matrix row-reduces to 1 0 1 0

0 1 −1 00 0 0 0

Page 83: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection HSE.SHS Solutions of Homogeneous Systems 51

By Theorem HSC [50], the system is consistent, and so the computation n−r = 3−2 = 1means the solution set contains one free variable by Corollary FVCS [42], and hence hasinfinitely many solutions. We can describe this solution set using the free variable x3,

S = {(x1, x2, x3) | x1 = −x3, x2 = x3} = {(−x3, x3, x3) | x3 ∈ C}

Geometrically, these are points in three dimensions that lie on a line through the origin.4

Example HSE.HISADHomogeneous, infinite solutions, Archetype DArchetype D [393] (and identically, Archetype E [397]) can be converted to the homo-

geneous system,

2x1 + x2 + 7x3 − 7x4 = 0

−3x1 + 4x2 − 5x3 − 6x4 = 0

x1 + x2 + 4x3 − 5x4 = 0

whose augmented matrix row-reduces to 1 0 3 −2 0

0 1 1 −3 00 0 0 0 0

By Theorem HSC [50], the system is consistent, and so the computation n−r = 4−2 = 2means the solution set contains two free variables by Corollary FVCS [42], and hence hasinfinitely many solutions. We can describe this solution set using the free variables x3

and x4,

S = {(x1, x2, x3, x4) | x1 = −3x3 + 2x4, x2 = −x3 + 3x4}= {(−3x3 + 2x4, −x3 + 3x4, x3, x4) | x3, x4 ∈ C}

4

After working through these examples, you might perform the same computations forthe slightly larger example, Archetype J [418].

Example HSE.HISAD [51] suggests the following theorem.

Theorem HMVEIHomogeneous, More Variables than Equations, Infinite solutionsSuppose that a homogeneous system of linear equations has m equations and n variableswith n > m. Then the system has infinitely many solutions. �

Proof We are assuming the system is homogeneous, so Theorem HSC [50] says it isconsistent. Then the hypothesis that n > m, together with Theorem CMVEI [44], givesinfinitely many solutions. �

Example HSE.HUSAB [50] and Example HSE.HISAA [50] are concerned with homo-geneous systems where n = m and expose a fundamental distinction between the twoexamples. One has a unique solution, while the other has infinitely many. These areexactly the only two possibilities for a homogeneous system and illustrate that each ispossible (unlike the case when n > m where Theorem HMVEI [51] tells us that there isonly one possibility for a homogeneous system).

Page 84: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection HSE.MVNSE Matrix and Vector Notation for Systems of Equations 52

Subsection MVNSEMatrix and Vector Notation for Systems of Equations

Notice that when we do row operations on the augmented matrix of a homogeneoussystem of linear equations the last column of the matrix is all zeros. Any one of the threeallowable row operations will convert zeros to zeros and thus, the final column of thematrix in reduced row-echelon form will also be all zeros. This observation might sufficeas a first explanation of the reason for some of the following definitions.

Definition CVColumn VectorA column vector of size m is an m × 1 matrix. We will frequently refer to a columnvector as simply a vector. �

Notation VNVector (u)Vectors will be written in bold, usually with lower case letters u, v, w, x, y, z. Somebooks like to write vectors with arrows, such as ~u. Writing by hand, some like to putarrows on top of the symbol, or a tilde underneath the symbol, as in u

∼. 5

Definition ZVZero VectorThe zero vector is the m× 1 matrix

0 =

000...0

Notation ZVNZero Vector (0)The zero vector will be written as 0. 5

Definition CMCoefficient MatrixFor a system of linear equations,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm

Page 85: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection HSE.MVNSE Matrix and Vector Notation for Systems of Equations 53

the coefficient matrix is the m× n matrix

A =

a11 a12 a13 . . . a1n

a21 a22 a23 . . . a2n

a31 a32 a33 . . . a3n...

am1 am2 am3 . . . amn

Definition VOCVector of ConstantsFor a system of linear equations,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm

the vector of constants is the column vector of size m

b =

b1

b2

b3...

bm

Definition SVSolution VectorFor a system of linear equations,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm

the solution vector is the column vector of size m

x =

x1

x2

x3...

xm

The solution vector may do double-duty on occasion. It might refer to a list of variablequantities at one point, and subsequently refer to values of those variables that actuallyform a particular solution to that system.

Page 86: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection HSE.NSM Null Space of a Matrix 54

Notation AMNAugmented Matrix ([A | b])With these definitions, we will write the augmented matrix of system of linear equationsin the form [A | b] in order to identify and distinguish the coefficients and the constants.5

Notation LSNLinear System (LS(A, b))We will write LS(A, b) to denote the system of linear equations with A as a coefficientmatrix and b as the vector of constants. 5

Example HSE.NSLENotation for systems of linear equationsThe system of linear equations

2x1 + 4x2 − 3x3 + 5x4 + x5 = 9

3x1 + x2 + +x4 − 3x5 = 0

−2x1 + 7x2 − 5x3 + 2x4 + 2x5 = −3

has coefficient matrix

A =

2 4 −3 5 13 1 0 1 −3−2 7 −5 2 2

and vector of constants

b =

90−3

and so will be referenced as LS(A, b). 4

With these definitions and notation a homogeneous system will be notated as LS(A, 0).Its augmented matrix will be [A | 0], which when converted to reduced row-echelon formwill still have the final column of zeros. So in this case, we may be as likely to justreference only the coefficient matrix.

Subsection NSMNull Space of a Matrix

The set of solutions to a homogeneous system (which by Theorem HSC [50] is neverempty) is of enough interest to warrant its own name. However, we define it as a propertyof the coefficient matrix, not as a property of some system of equations.

Definition NSMNull Space of a MatrixThe null space of a matrix A, denoted N (A), is the set of all the vectors that aresolutions to the homogeneous system LS(A, 0). �

In the Archetypes (Chapter A [375]) each example that is a system of equations also hasa corresponding homogeneous system of equations listed, and several sample solutionsare given. These solutions will be elements of the null space of the coefficient matrix.We’ll look at one example.

Page 87: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection HSE.NSM Null Space of a Matrix 55

Example HSE.NSEAINull space elements of Archetype IThe write-up for Archetype I [414] lists several solutions of the corresponding homoge-

neous system. Here are two, written as solution vectors. We can say that they are in thenull space of the coefficient matrix for the system of equations in Archetype I [414].

x =

30−5−6001

y =

−41−3−2111

However, the vector

z =

1000002

is not in the null space, since it is not a solution to the homogeneous system. For example,it fails to even make the first equation true. 4

Page 88: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section NSM NonSingular Matrices 56

Section NSM

NonSingular Matrices

In this section we specialize to systems with equal numbers of equations and variables,which will prove to be a case of special interest.

Subsection NSMNonSingular Matrices

Our theorems will now establish connections between systems of equations (homogeneousor otherwise), augmented matrices representing those systems, coefficient matrices, con-stant vectors, the reduced row-echelon form of matrices (augmented and coefficient) andsolution sets. Be very careful in your reading, writing and speaking about systems ofequations, matrices and sets of vectors. Now would be a good time to review the discus-sion about speaking and writing mathematics in Technique L [19].

Definition SQMSquare MatrixA matrix with m rows and n columns is square if m = n. In this case, we say the

matrix has size n. To emphasize the situation when a matrix is not square, we will callit rectangular. �

We can now present one of the central definitions of linear algebra.

Definition NMNonsingular MatrixSuppose A is a square matrix. And suppose the homogeneous linear system of equationsLS(A, 0) has only the trivial solution. Then we say that A is a nonsingular matrix.Otherwise we say A is a singular matrix. �

We can investigate whether any square matrix is nonsingular or not, no matter if thematrix is derived somehow from a system of equations or if it is simply a matrix. Thedefinition says that to perform this investigation we must construct a very specific systemof equations (homogenous, with the matrix as the coefficient matrix) and look at itssolution set. We will have theorems in this section that connect nonsingular matriceswith systems of equations, creating more opportunities for confusion. Convince yourselfnow of two observations, (1) we can decide nonsingularity for any square matrix, and (2)the determination of nonsingularity involves the solution set for a certain homogenoussystem of equations.

Notice that it makes no sense to call a system of equations nonsingular (the term doesnot apply to a system of equations), nor does it make any sense to call a 5 × 7 matrixsingular (the matrix is not square).

Example NSM.SA singular matrix, Archetype AExample HSE.HISAA [50] shows that the coefficient matrix derived from Archetype A [379],

Page 89: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.NSM NonSingular Matrices 57

specifically the 3× 3 matrix,

A =

1 −1 22 1 11 1 0

is a singular matrix since there are nontrivial solutions to the homogeneous systemLS(A, 0). 4

Example NSM.NSA nonsingular matrix, Archetype BExample HSE.HUSAB [50] shows that the coefficient matrix derived from Archetype B [384],specifically the 3× 3 matrix,

B =

−7 −6 −125 5 71 0 4

is a nonsingular matrix since the homogeneous system, LS(B, 0), has only the trivialsolution. 4

Notice that we will not discuss Example HSE.HISAD [51] as being a singular or nonsin-gular coefficient matrix since the matrix is not square.

The next theorem combines with our main computational technique (row-reducing amatrix) to make it easy to recognize a nonsingular matrix. But first a definition.

Definition IMIdentity MatrixThe m × m identity matrix, Im = (aij) has aij = 1 whenever i = j, and aij = 0whenever i 6= j. �

Example NSM.IMAn identity matrixThe 4× 4 identity matrix is

I4 =

1 0 0 00 1 0 00 0 1 00 0 0 1

. 4

Notice that an identity matrix is square, and in reduced row-echelon form. So in par-ticular, if we were to arrive at the identity matrix while bringing a matrix to reducedrow-echelon form, then it would have all of the diagonal entries circled as leading 1’s.

Theorem NSRRINonSingular matrices Row Reduce to the Identity matrixSuppose that A is a square matrix and B is a row-equivalent matrix in reduced row-echelon form. Then A is nonsingular if and only if B is the identity matrix. �

Proof (⇐) Suppose B is the identity matrix. When the augmented matrix [A | 0]is row-reduced, the result is [B | 0] = [In | 0]. The number of nonzero rows is equalto the number of variables in the linear system of equations LS(A, 0), so n = r andCorollary FVCS [42] gives n − r = 0 free variables. Thus, the homogeneous systemLS(A, 0) has just one solution, which must be the trivial solution. This is exactly thedefinition of a nonsingular matrix.

Page 90: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.NSM NonSingular Matrices 58

(⇒) We will prove the contrapositive. Suppose B is not the identity matrix. Whenthe augmented matrix [A | 0] is row-reduced, the result is [B | 0]. The number ofnonzero rows is less than the number of variables for the system of equations, so Corol-lary FVCS [42] gives n− r > 0 free variables. Thus, the homogeneous system LS(A, 0)has infinitely many solutions, so the system has more solutions than just the trivialsolution. Thus the matrix is not nonsingular, the desired conclusion. �

Notice that since this theorem is an equivalence it will always allow us to determine if amatrix is either nonsingular or singular. Here are two examples of this, continuing ourstudy of Archetype A and Archetype B.

Example NSM.SRRSingular matrix, row-reducedThe coefficient matrix for Archetype A [379] is

A =

1 −1 22 1 11 1 0

which when row-reduced becomes the row-equivalent matrix

B =

1 0 1

0 1 −10 0 0

.

Since this matrix is not the 3 × 3 identity matrix, Theorem NSRRI [57] tells us that Ais a singular matrix. 4

Example NSM.NSRRNonSingular matrix, row-reducedThe coefficient matrix for Archetype B [384] is

A =

−7 −6 −125 5 71 0 4

which when row-reduced becomes the row-equivalent matrix

B =

1 0 0

0 1 0

0 0 1

.

Since this matrix is the 3 × 3 identity matrix, Theorem NSRRI [57] tells us that A is anonsingular matrix. 4

Example NSM.NSSNull space of a singular matrixGiven the coefficient matrix from Archetype A [379],

A =

1 −1 22 1 11 1 0

Page 91: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.NSM NonSingular Matrices 59

the null space is the set of solutions to the homogeneous system of equations LS(A, 0)has a solution set and null space constructed in Example HSE.HISAA [50] as

N (A) =

−x3

x3

x3

∣∣∣∣∣∣ x3 ∈ C

4

Example NSM.NSNSNull space of a nonsingular matrixGiven the coefficient matrix from Archetype B,

A =

−7 −6 −125 5 71 0 4

the homogeneous system LS(A, 0) has a solution set constructed in Example HSE.HUSAB [50]that contains only the trivial solution, so the null space has only a single element,

N (A) =

0

00

4

These two examples illustrate the next theorem, which is another equivalence.

Theorem NSTNSNonSingular matrices have Trivial Null SpacesSuppose that A is a square matrix. Then A is nonsingular if and only if the null space ofA, N (A), contains only the trivial solution to the system LS(A, 0), i.e. N (A) = {0}. �

Proof The null space of a square matrix, A, is the set of solutions to the homogeneoussystem, LS(A, 0). A matrix is nonsingular if and only if the set of solutions to the ho-mogeneous system, LS(A, 0), has only a trivial solution. These two observations may bechained together to construct the two proofs necessary for each of half of this theorem.�

Proof Technique UUniquenessA theorem will sometimes claim that some object, having some desirable property, isunique. In other words, there should be only one such object. To prove this, a standardtechnique is to assume there are two such objects and proceed to analyze the conse-quences. The end result may be a contradiction, or the conclusion that the two allegedlydifferent objects really are equal. ♦

The next theorem pulls a lot of ideas together. It tells us that we can learn a lot aboutsolutions to a system of linear equations with a square coefficient matrix by examining asimilar homogeneous system.

Theorem NSMUSNonSingular Matrices and Unique SolutionsSuppose that A is a square matrix. A is a nonsingular matrix if and only if the systemLS(A, b) has a unique solution for every choice of the constant vector b. �

Page 92: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.NSM NonSingular Matrices 60

Proof (⇐) The hypothesis for this half of the proof is that the system LS(A, b) has aunique solution for every choice of the constant vector b. We will make a very specificchoice for b: b = 0. Then we know that the system LS(A, 0) has a unique solution. Butthis is precisely the definition of what it means for A to be nonsingular. If this half ofthe proof seemed too easy, perhaps we’ll have to work a bit harder to get the implicationin the opposite direction.

(⇒) We will assume A is nonsingular, and try to solve the system LS(A, b) withoutmaking any assumptions about b. To do this we will begin by constructing a newhomogeneous linear system of equations that looks very much like the original. SupposeA has size n (why must it be square?) and write the original system as,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

... (∗)an1x1 + an2x2 + an3x3 + · · ·+ annxn = bn

form the new, homogeneous system in n equations with n+1 variables, by adding a newvariable y, whose coefficients are the negatives of the constant terms,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn − b1y = 0

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn − b2y = 0

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn − b3y = 0

... (∗∗)an1x1 + an2x2 + an3x3 + · · ·+ annxn − bny = 0

Since this is a homogeneous system with more variables than equations (m = n+1 > n),Theorem HMVEI [51] says that the system has infinitely many solutions. We will chooseone of these solutions, any one of these solutions, so long as it is not the trivial solution.Write this solution as

x1 = c1 x2 = c2 x3 = c3 . . . xn = cn y = cn+1

We know that at least one value of the ci is nonzero, but we will now show that inparticular cn+1 6= 0. We do this using a proof by contradiction. So suppose the ci form asolution as described, and in addition that cn+1 = 0. Then we can write the i-th equationof system (∗∗) as,

ai1c1 + ai2c2 + ai3c3 + · · ·+ aincn − bi(0) = 0

which becomes

ai1c1 + ai2c2 + ai3c3 + · · ·+ aincn = 0

Since this is true for each i, we have that x1 = c1, x2 = c2, x3 = c3, . . . , xn = cn isa solution to the homogeneous system LS(A, 0) formed with a nonsingular coefficient

Page 93: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.NSM NonSingular Matrices 61

matrix. This means that the only possible solution is the trivial solution, so c1 = 0, c2 =0, c3 = 0, . . . , cn = 0. So, assuming simply that cn+1 = 0, we conclude that all of the ci

are zero. But this contradicts our choice of the ci as not being the trivial solution to thesystem (∗∗). So cn+1 6= 0.

We now propose and verify a solution to the original system (∗). Set

x1 =c1

cn+1

x2 =c2

cn+1

x3 =c3

cn+1

. . . xn =cn

cn+1

Notice how it was necessary that we know that cn+1 6= 0 for this step to succeed. Now,evaluate the i-th equation of system (∗) with this proposed solution, and recognize in thethird line that c1 through cn+1 appear as if they were substituted into the left-hand sideof the i-th equation of system (∗∗),

ai1c1

cn+1

+ ai2c2

cn+1

+ ai3c3

cn+1

+ · · ·+ aincn

cn+1

=1

cn+1

(ai1c1 + ai2c2 + ai3c3 + · · ·+ aincn)

=1

cn+1

(ai1c1 + ai2c2 + ai3c3 + · · ·+ aincn − bicn+1) + bi

=1

cn+1

(0) + bi

= bi

Since this equation is true for every i, we have found a solution to system (∗). To finish,we still need to establish that this solution is unique.

With one solution in hand, we will entertain the possibility of a second solution. Soassume system (∗) has two solutions,

x1 = d1 x2 = d2 x3 = d3 . . . xn = dn

x1 = e1 x2 = e2 x3 = e3 . . . xn = en

Then,

(ai1(d1 − e1) + ai2(d2 − e2) + ai3(d3 − e3) + · · ·+ ain(dn − en))

= (ai1d1 + ai2d2 + ai3d3 + · · ·+ aindn)− (ai1e1 + ai2e2 + ai3e3 + · · ·+ ainen)

= bi − bi

= 0

This is the i-th equation of the homogeneous system LS(A, 0) evaluated with xj = dj−ej,1 ≤ j ≤ n. Since A is nonsingular, we must conclude that this solution is the trivialsolution, and so 0 = dj − ej, 1 ≤ j ≤ n. That is, dj = ej for all j and the two solutionsare identical, meaning any solution to (∗) is unique. �

This important theorem deserves several comments. First, notice that the proposedsolution (xi = ci

cn+1) appeared in the proof with no motivation whatsoever. This is just

fine in a proof. A proof should convince you that a theorem is true. It is your job toread the proof and be convinced of every assertion. Questions like “Where did that comefrom?” or “How would I think of that?” have no bearing on the validity of the proof.

Page 94: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.NSM NonSingular Matrices 62

Second, this theorem helps to explain part of our interest in nonsingular matrices. Ifa matrix is nonsingular, then no matter what vector of constants we pair it with, usingthe matrix as the coefficient matrix will always yield a linear system of equations witha solution, and the solution is unique. To determine if a matrix has this property (non-singularity) it is enough to just solve one linear system, the homogeneous system withthe matrix as coefficient matrix and the zero vector as the vector of constants.

Finally, formulating the negation of the second part of this theorem is a good exercise.A singular matrix has the property that for some value of the vector b, the systemLS(A, b) does not have a unique solution (which means that it has no solution or infinitelymany solutions). We will be able to say more about this case later.

Proof Technique MEMultiple EquivalencesA very specialized form of a theorem begins with the statement “The following areequivalent. . . ” and then follows a list of statements. Informally, this lead-in sometimesgets abbreviated by “TFAE.” This formulation means that any two of the statements onthe list can be connected with an “if and only if” to form a theorem. So if the list has nstatements then there are n(n−1)

2possible equivalences that can be constructed (and are

claimed to be true).Suppose a theorem of this form has statements denoted as A, B, C,. . .Z. To prove

the entire theorem, we can prove A ⇒ B, B ⇒ C, C ⇒ D,. . . , Y ⇒ Z and finally,Z ⇒ A. This circular chain of n equivalences would allow us, logically, if not practically,to form any one of the n(n−1)

2possible equivalences by chasing the equivalences around

the circle as far as required. ♦

Square matrices that are nonsingular have a long list of interesting properties, which wewill start to catalog in the following, recurring, theorem. Of course, singular matriceswill have all of the opposite properties.

Theorem NSME1NonSingular Matrix Equivalences, Round 1Suppose that A is a square matrix. The following are equivalent.

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the trivial solution, N (A) = {0}.

4. The linear system LS(A, b) has a unique solution for every possible choice of b.�

Proof That A is nonsingular is equivalent to each of the subsequent statements by, inturn, Theorem NSRRI [57], Theorem NSTNS [59] and Theorem NSMUS [59]. So thestatement of this theorem is just a convenient way to organize all these results. �

Page 95: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.EXC Exercises 63

Subsection EXCExercises

C30 Contributed by Robert BeezerIs the matrix below singular or nonsingular? Why?

−3 1 2 82 0 3 41 2 7 −45 −1 2 0

Solution [64]

Page 96: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection NSM.SOL Solutions 64

Subsection SOLSolutions

C30 Exercise [63] Contributed by Robert BeezerThe matrix row-reduces to

1 0 0 0

0 1 0 0

0 0 1 0

0 0 0 1

which is the 4× 4 identity matrix. By Theorem NSRRI [57] the original matrix must benonsingular.

Page 97: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

V: Vectors

Section VO

Vector Operations

We have worked extensively in the last chapter with matrices, and some with vectors.In this chapter we will develop the properties of vectors, while preparing to study vectorspaces. Initially we will depart from our study of systems of linear equations, but inSection LC [72] we will forge a connection between linear combinations and systems oflinear combinations in Theorem SLSLC [76]. This connection will allow us to understandsystems of linear equations at a higher level, while consequently discussing them lessfrequently.

In the current section we define some new operations involving vectors, and collectsome basic properties of these operations. Begin by recalling our definition of a columnvector as a matrix with just one column (Definition CV [52]). The collection of allpossible vectors of a fixed size is a commonly used set, so we start with its definition.

Definition VSCMVector Space Cm

The vector space Cm is the set of all column vectors of size m with entries from the setof complex numbers. �

When this set is defined using only entries from the real numbers, it is written as Rm

and is known as Euclidean m-space.

The term “vector” is used in a variety of different ways. We have defined it asa matrix with a single column. It could simply be an ordered list of numbers, andwritten like (2, 3, −1, 6). Or it could be interpreted as a point in m dimensions, such as(3, 4, −2) representing a point in three dimensions relative to x, y and z axes. With aninterpretation as a point, we can construct an arrow from the origin to the point whichis consistent with the notion that a vector has direction and magnitude.

All of these ideas can be shown to be related and equivalent, so keep that in mindas you connect the ideas of this course with ideas from other disciplines. For now, we’llstick with the idea that a vector is a matrix with just one column, or even more simply,just a list of numbers, in some order.

65

Page 98: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VO.VEASM Vector equality, addition, scalar multiplication 66

Subsection VEASMVector equality, addition, scalar multiplication

We start our study of this set by first defining what it means for two vectors to be thesame.

Definition CVEColumn Vector EqualityThe vectors

u =

u1

u2

u3...

um

v =

v1

v2

v3...

vm

are equal, written u = v provided that ui = vi for all 1 ≤ i ≤ m. �

Now this may seem like a silly (or even stupid) thing to say so carefully. Of course twovectors are equal if they are equal for each corresponding entry! Well, this is not as sillyas it appears. We will see a few occasions later where the obvious definition is not theright one. And besides, in doing mathematics we need to be very careful about makingall the necessary definitions and making them unambiguous. And we’ve done that here.

Notice now that the symbol ‘=’ is now doing triple-duty. We know from our earliereducation what it means for two numbers (real or complex) to be equal, and we take thisfor granted. Earlier, in Technique SE [14] we discussed at some length what it meant fortwo sets to be equal. Now we have defined what it means for two vectors to be equal, andthat definition builds on our definition for when two numbers are equal when we use thecondition ui = vi for all 1 ≤ i ≤ m. So think carefully about your objects when you seean equal sign and think about just which notion of equality you have encountered. Thiswill be especially important when you are asked to construct proofs whose conclusionstates that two objects are equal.

OK, lets do an example of vector equality that begins to hint at the utility of thisdefinition.

Example VO.VESEVector equality for a system of equationsConsider the system of simultaneous linear equations in Archetype B [384],

−7x1 − 6x2 − 12x3 = −33

5x1 + 5x2 + 7x3 = 24

x1 + 4x3 = 5

Note the use of three equals signs — each indicates an equality of numbers (the lin-ear expressions are numbers when we evaluate them with fixed values of the variablequantities). Now write the vector equality,−7x1 − 6x2 − 12x3

5x1 + 5x2 + 7x3

x1 + 4x3

=

−33245

.

Page 99: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VO.VEASM Vector equality, addition, scalar multiplication 67

By Definition CVE [66], this single equality (of two column vectors) translates into threesimultaneous equalities of numbers that form the system of equations. So with this newnotion of vector equality we can become less reliant on referring to systems of simultaneousequations. There’s more to vector equality than just this, but this is a good example forstarters and we will develop it further. 4

We will now define two operations on the set Cm. By this we mean well-defined proce-dures that somehow convert vectors into other vectors. Here are two of the most basicdefinitions of the entire course.

Definition CVAColumn Vector AdditionGiven the vectors

u =

u1

u2

u3...

um

v =

v1

v2

v3...

vm

the sum of u and v is the vector

u + v =

u1

u2

u3...

um

+

v1

v2

v3...

vm

=

u1 + v1

u2 + v2

u3 + v3...

um + vm

. �

So vector addition takes two vectors of the same size and combines them (in a naturalway!) to create a new vector of the same size. Notice that this definition is required,even if we agree that this is the obvious, right, natural or correct way to do it. Noticetoo that the symbol ‘+’ is being recycled. We all know how to add numbers, but now wehave the same symbol extended to double-duty and we use it to indicate how to add twonew objects, vectors. And this definition of our new meaning is built on our previousmeaning of addition via the expressions ui + vi. Think about your objects, especiallywhen doing proofs. Vector addition is easy, here’s an example from C4.

Example VO.VAAddition of two vectors in C4

If

u =

2−342

v =

−152−7

then

u + v =

2−342

+

−152−7

=

2 + (−1)−3 + 54 + 2

2 + (−7)

=

126−5

. 4

Page 100: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VO.VEASM Vector equality, addition, scalar multiplication 68

Our second operation takes two objects of different types, specifically a number and avector, and combines them to create another vector. In this context we call a number ascalar in order to emphasize that it is not a vector.

Definition CVSMColumn Vector Scalar MultiplicationGiven the vector

u =

u1

u2

u3...

um

and the scalar α ∈ C, the scalar multiple of u by α is

αu = α

u1

u2

u3...

um

=

αu1

αu2

αu3...

αum

. �

Notice that we are doing a kind of multiplication here, but we are defining a new type,perhaps in what appears to be a natural way. We use juxtaposition (smashing twosymbols together side-by-side) to denote this operation rather than using a symbol likewe did with vector addition. So this can be another source of confusion. When twosymbols are next to each other, are we doing regular old multiplication, the kind we’vedone for years, or are we doing scalar vector multiplication, the operation we just defined?Think about your objects — if the first object is a scalar, and the second is a vector,then it must be that we are doing our new operation, and the result of this operation willbe another vector.

Notice how consistency in notation can be an aid here. If we write scalars as lowercase Greek letters from the start of the alphabet (such as α, β, . . . ) and write vectorsin bold Latin letters from the end of the alphabet (u, v, . . . ), then we have some hintsabout what type of objects we are working with. This can be a blessing and a curse, sincewhen we go read another book about linear algebra, or read an application in anotherdiscipline (physics, economics, . . . ) the types of notation employed may be very differentand hence unfamiliar.

Again, computationally, vector scalar multiplication is very easy.

Example VO.VSMScalar multiplication in C5

If

u =

31−24−1

Page 101: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VO.VEASM Vector equality, addition, scalar multiplication 69

and α = 6, then

αu = 6

31−24−1

=

6(3)6(1)

6(−2)6(4)6(−1

=

186−1224−6

. 4

It is usually straightforward to effect these computations with a calculator or program.

Computation Note VLC.MMAVector Linear Combinations (Mathematica)

Contributed by Robert BeezerVectors in Mathematica are represented as lists, written and displayed horizontally. Forexample, the vector

v =

1234

would be entered and named via the command

v = {1, 2, 3, 4}

Vector addition and scalar multiplication are then very natural. If u and v are twolists of equal length, then

2u + (−3)v

will compute the correct vector and return it as a list. If u and v have different sizes,then Mathematica will complain about “objects of unequal length.” ⊗Computation Note VLC.TI86Vector Linear Combinations (TI-86)

Contributed by Robert BeezerVector operations on the TI-86 can be accessed via the VECTR key, which is Yellow-8

. The EDIT tool appears when the F2 key is pressed. After providing a name andgiving a “dimension” (the size) then you can enter the individual entries, one at a time.Vectors can also be entered on the home screen using brackets ( [ , ] ). To create thevector

v =

1234

use brackets and the store key ( STO ),

[1, 2, 3, 4] → v

Vector addition and scalar multiplication are then very natural. If u and v are twovectors of equal size, then

2 ∗ u + (−3) ∗ v

will compute the correct vector and display the result as a vector. ⊗

Page 102: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VO.VSP Vector Space Properties 70

Computation Note VLC.TI83Vector Linear Combinations (TI-83)

Contributed by Douglas PhelpsEntering a vector on the TI-83 is the same process as entering a matrix. You press 4

ENTER 3 ENTER for a 4× 3 matrix. Likewise, you press 4 ENTER 1 ENTER for a vectorof size 4. To multiply a vector by 8, press the number 8, then press the MATRX key, thenscroll down to the letter you named your vector (A, B, C, etc) and press ENTER .

To add vectors A and B for example, press the MATRX key, then ENTER . Thenpress the + key. Then press the MATRX key, then the down arrow once, then ENTER .[A] + [B] will appear on the screen. Press ENTER . ⊗

Subsection VSPVector Space Properties

With definitions of vector addition and scalar multiplication we can state, and prove,several properties of each operation, and some properties that involve their interplay.We now collect ten of them here for later reference.

Theorem VSPCMVector Space Properties of Cm

Suppose that u, v and w are vectors in Cm and α and β are scalars. Then

1. u + v ∈ Cm (Additive closure)

2. αu ∈ Cm (Scalar closure)

3. u + v = v + u (Commutativity)

4. u + (v + w) = (u + v) + w (Associativity of vector addition)

5. There is a vector, 0, called the zero vector, such that u + 0 = u for all u ∈ Cm.(Additive identity)

6. For each vector u ∈ Cm, there exists a vector −u ∈ Cm so that u + (−u) = 0.(Additive inverses)

7. α(βu) = (αβ)u (Associativity of scalar multiplication)

8. α(u + v) = αu + αv (Distributivity across vector addition)

9. (α + β)u = αu + βu (Distributivity across addition)

10. 1u = u (Scalar multiplication with 1) �

Proof While some of these properties seem very obvious, they all require proof. However,the proofs are not very interesting, and border on tedious. We’ll prove one version of

Page 103: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VO.VSP Vector Space Properties 71

distributivity very carefully, and you can test your proof-building skills on some of theothers.

(α + β)u = (α + β)

u1

u2

u3...

um

=

(α + β)u1

(α + β)u2

(α + β)u3...

(α + β)um

=

αu1 + βu1

αu2 + βu2

αu3 + βu3...

αum + βum

=

αu1

αu2

αu3...

αum

+

βu1

βu2

βu3...

βum

= α

u1

u2

u3...

um

+ β

u1

u2

u3...

um

= αu + βu �

Be careful with the notion of the vector −u. This is a vector that we add to u so thatthe result is the particular vector 0. This is basically a property of vector addition. Ithappens that we can compute −u using the other operation, scalar multiplication. Wecan prove this directly by writing that

−u =

−u1

−u2

−u3...

−um

= (−1)

u1

u2

u3...

um

= (−1)u

We will see later how to derive this property as a consequence of several of the tenproperties listed in Theorem VSPCM [70].

Page 104: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section LC Linear Combinations 72

Section LC

Linear Combinations

Subsection LCLinear Combinations

In Section VO [65] we defined vector addition and scalar multiplication. These twooperations combine nicely to give us a construction known as a linear combination, aconstruct that we will work with throughout this course.

Definition LCCVLinear Combination of Column VectorsGiven n vectors u1, u2, u3, . . . , un and n scalars α1, α2, α3, . . . , αn, their linear com-bination is the vector

α1u1 + α2u2 + α3u3 + · · ·+ αnun. �

So this definition takes an equal number of scalars and vectors, combines them usingour two new operations (scalar multiplication and vector addition) and creates a singlebrand-new vector, of the same size as the original vectors. When a definition or theorememploys a linear combination, think about the nature of the objects that go into itscreation (lists of scalars and vectors), and the type of object that results (a single vector).Computationally, a linear combination is pretty easy.

Example LC.TLCTwo linear combinations in C6

Suppose that

α1 = 3 α2 = −4 α3 = 2 α4 = −1

and

u1 =

24−3129

u2 =

630−214

u3 =

−5211−30

u4 =

32−5713

Page 105: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LC.LC Linear Combinations 73

then their linear combination is

α1u1 + α2u2 + α3u3 + α4u4 = (1)

24−3129

+ (−4)

630−214

+ (2)

−5211−30

+ (−1)

32−5713

=

24−3129

+

−24−1208−4−16

+

−10422−60

+

−3−25−7−1−3

=

−35−644−9−8

.

A different linear combination, of the same set of vectors, can be formed with differentscalars. Take

β1 = 3 β2 = 0 β3 = 5 β4 = −6

and form the linear combination

β1u1 + β2u2 + β3u3 + β4u4 = (3)

24−3129

+ (0)

630−214

+ (5)

−5211−30

+ (−6)

32−5713

=

612−93627

+

000000

+

−251055−150

+

−16−25−7−1−3

=

−352011−1024

.

Notice how we could keep our set of vectors fixed, and use different sets of scalars to con-struct different vectors. You might build a few new linear combinations of u1, u2, u3, u4

right now. We’ll be right here when you get back. What vectors were you able to create?Do you think you could create the vector

w =

13155−17225

with a “suitable” choice of four scalars? Do you think you could create any possible vectorfrom C6 by choosing the proper scalars? These last two questions are very fundamental,and time spent considering them now will prove beneficial later. 4

Page 106: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LC.LC Linear Combinations 74

Proof Technique DCDecompositionsMuch of your mathematical upbringing, especially once you began a study of algebra,revolved around simplifying expressions — combining like terms, obtaining common de-nominators so as to add fractions, factoring in order to solve polynomial equations. How-ever, as often as not, we will do the opposite. Many theorems and techniques will revolvearound taking some object and “decomposing” it into some combination of other objects,ostensibly in a more complicated fashion. When we say something can “be written as”something else, we mean that the one object can be decomposed into some combinationof other objects. This may seem unnatural at first, but results of this type will give usinsight into the structure of the original object by exposing its building blocks. ♦

Example LC.ABLCArchetype B as a linear combinationIn this example we will rewrite Archetype B [384] in the language of vectors, vector

equality and linear combinations. In Example VO.VESE [66] we wrote the simultaneoussystem of m = 3 equations as the vector equality−7x1 − 6x2 − 12x3

5x1 + 5x2 + 7x3

x1 + 4x3

=

−33245

.

Now we will bust up the linear expressions on the left, first using vector addition,−7x1

5x1

x1

+

−6x2

5x2

0x2

+

−12x3

7x3

4x3

=

−33245

.

Now we can rewrite each of these n = 3 vectors as a scalar multiple of a fixed vector,where the scalar is one of the unknown variables, converting the left-hand side into alinear combination

x1

−751

+ x2

−650

+ x3

−1274

=

−33245

.

We can now interpret the problem of solving the system of equations as determiningvalues for the scalar multiples that make the vector equation true. In the analysis ofArchetype B [384], we were able to determine that it had only one solution. A quickway to see this is to row-reduce the coefficient matrix to the 3 × 3 identity matrix andapply Theorem NSRRI [57] to determine that the coefficient matrix is nonsingular. ThenTheorem NSMUS [59] tells us that the system of equations has a unique solution. Thissolution is

x1 = −3 x2 = 5 x3 = 2.

So, in the context of this example, we can express the fact that these values of thevariables are a solution by writing the linear combination,

(−3)

−751

+ (5)

−650

+ (2)

−1274

=

−33245

.

Page 107: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LC.LC Linear Combinations 75

Furthermore, these are the only three scalars that will accomplish this equality, sincethey come from a unique solution.

Notice how the three vectors in this example are the columns of the coefficient matrixof the system of equations. This is our first hint of the important interplay between thevectors that form the columns of a matrix, and the matrix itself. 4

With any discussion of Archetype A [379] or Archetype B [384] we should be sure tocontrast with the other.

Example LC.AALCArchetype A as a linear combinationAs a vector equality, Archetype A [379] can be written asx1 − x2 + 2x3

2x1 + x2 + x3

x1 + x2

=

185

.

Now bust up the linear expressions on the left, first using vector addition, x1

2x1

x1

+

−x2

x2

x2

+

2x3

x3

0x3

=

185

.

Rewrite each of these n = 3 vectors as a scalar multiple of a fixed vector, where the scalaris one of the unknown variables, converting the left-hand side into a linear combination

x1

121

+ x2

−111

+ x3

210

=

185

.

Row-reducing the augmented matrix for Archetype A [379] leads to the conclusion thatthe system is consistent and has free variables, hence infinitely many solutions. So forexample, the two solutions

x1 = 2 x2 = 3 x3 = 1

x1 = 3 x2 = 2 x3 = 0

can be used together to say that,

(2)

121

+ (3)

−111

+ (1)

210

=

185

= (3)

121

+ (2)

−111

+ (0)

210

Ignore the middle of this equation, and move all the terms to the left-hand side,

(2)

121

+ (3)

−111

+ (1)

210

+ (−3)

121

+ (−2)

−111

+ (−0)

210

=

000

.

Regrouping gives

(−1)

121

+ (1)

−111

+ (1)

210

=

000

.

Page 108: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LC.LC Linear Combinations 76

Notice that these three vectors are the columns of the coefficient matrix for the system ofequations in Archetype A [379]. This equality says there is a linear combination of thosecolumns that equals the vector of all zeros. Give it some thought, but this says that

x1 = −1 x2 = 1 x3 = 1

is a nontrivial solution to the homogeneous system of equations with the coefficient matrixfor the original system in Archetype A [379]. In particular, this demonstrates that thiscoefficient matrix is singular. 4

There’s a lot going on in the last two examples. Come back to them in a while and makesome connections with the intervening material. For now, we will summarize and explainsome of this behavior with a theorem.

Theorem SLSLCSolutions to Linear Systems are Linear CombinationsDenote the columns of the m × n matrix A as the vectors A1, A2, A3, . . . , An. Then

x =

α1

α2

α3...

αn

is a solution to the linear system of equations LS(A, b) if and only if

α1A1 + α2A2 + α3A3 + · · ·+ αnAn = b �

Proof Write the system of equations LS(A, b) as

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn = b1

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn = b2

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn = b3

...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn = bm.

Now use vector equality (Definition CVE [66]) to replace the m simultaneous equalitiesby one vector equality,

a11x1 + a12x2 + a13x3 + · · ·+ a1nxn

a21x1 + a22x2 + a23x3 + · · ·+ a2nxn

a31x1 + a32x2 + a33x3 + · · ·+ a3nxn...

am1x1 + am2x2 + am3x3 + · · ·+ amnxn

=

b1

b2

b3...

bm

.

Use vector addition (Definition CVA [67]) to rewrite,a11x1

a21x1

a31x1...

am1x1

+

a12x2

a22x2

a32x2...

am2x2

+

a13x3

a23x3

a33x3...

am3x3

+ · · ·+

a1nxn

a2nxn

a3nxn...

amnxn

=

b1

b2

b3...

bm

.

Page 109: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LC.LC Linear Combinations 77

And finally, use the definition of vector scalar multiplication Definition CVSM [68],

x1

a11

a21

a31...

am1

+ x2

a12

a22

a32...

am2

+ x3

a13

a23

a33...

am3

+ · · ·+ xm

a1n

a2n

a3n...

amn

=

b1

b2

b3...

bm

.

and use notation for the various column vectors,

x1A1 + x2A2 + x3A3 + · · ·+ xnAn = b.

Each of the expressions above is just a rewrite of another one. So if we begin with asolution to the system of equations, substituting its values into the original system willmake the equations simultaneously true. But then these same values will also make thefinal expression with the linear combination true. Reversing the argument, and employingthe equations in reverse, will give the other half of the proof. �

In other words, this theorem tells us that solutions to systems of equations are linearcombinations of the column vectors of the coefficient matrix (Ai) which yield the constantvector b. Or said another way, a solution to a system of equations LS(A, b) is an answerto the question “How can I realize the vector b as a linear combination of the columnsof A?” Look through the archetypes that are systems of equations and examine a fewof the advertised solutions. In each case use the solution to form a linear combinationof the columns of the coefficient matrix and verify that the result equals the constantvector.

Page 110: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LC.EXC Exercises 78

Subsection EXCExercises

C40 Contributed by Robert BeezerFind the vector form of the solutions to the system of equations below.

2x1 − 4x2 + 3x3 + x5 = 6

x1 − 2x2 − 2x3 + 14x4 − 4x5 = 15

x1 − 2x2 + x3 + 2x4 + x5 = −1

−2x1 + 4x2 − 12x4 + x5 = −7

Solution [79]

Page 111: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LC.SOL Solutions 79

Subsection SOLSolutions

C40 Exercise [78] Contributed by Robert BeezerRow-reduce the augmented matrix representing this system, to find

1 −2 0 6 0 1

0 0 1 −4 0 3

0 0 0 0 1 −50 0 0 0 0 0

The system is consistent (no leading one in column 6, Theorem RCLS [41]). x2 and x4

are the free variables. Now apply Theorem VFSLS [82], or follow the three-step processof Example SS.VFSADI [??], Example SS.VFSAL [85] to obtain

x1

x2

x3

x4

x5

=

1030−5

+ x2

21000

+ x4

−60410

Page 112: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section SS Spanning Sets 80

Section SS

Spanning Sets

In this section we will describe a compact way to indicate the elements of an infinite setof vectors, making use of linear combinations. This will give us a nice way to describethe elements of a set of solutions to a linear system, or the elements of the null space ofa matrix.

Subsection VFSSVector Form of Solution Sets

We have recently begun writing solutions to systems of equations as column vectors. Forexample Archetype B [384] has the solution x1 = −3, x2 = 5, x3 = 2 which we now writeas

x =

x1

x2

x3

=

−352

.

Now, we will use column vectors and linear combinations to express all of the solutionsto a linear system of equations in a compact and understandable way. First, here’s anexample that will motivate our next theorem. This is a valuable technique, almost theequal of row-reducing a matrix, so be sure you get comfortable with it over the course ofthis section.

Example SS.VFSADVector form of solutions for Archetype DArchetype D [393] is a linear system of 3 equations in 4 variables. Row-reducing the

augmented matrix yields 1 0 3 −2 4

0 1 1 −3 00 0 0 0 0

and we see r = 2 nonzero rows. Also, D = {1, 2} so the dependent variables are then x1

and x2. F = {3, 4, 5} so the two free variables are x3 and x4. We will develop a linearcombination that expresses a typical solution, in three steps.

Step 1. Write the vector of variables as a fixed vector, plus a linear combination ofn− r vectors, using the free variables as the scalars.

x =

x1

x2

x3

x4

=

+ x3

+ x4

Step 2. For each free variable, use 0’s and 1’s to ensure equality for the corresponding

Page 113: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.VFSS Vector Form of Solution Sets 81

entry of the the vectors.

x =

x1

x2

x3

x4

=

00

+ x3

10

+ x4

01

Step 3. For each dependent variable, use the augmented matrix to formulate an equationexpressing the dependent variable as a constant plus multiples of the free variables.Convert this equation into entries of the vectors that ensure equality for each dependentvariable, one at a time.

x1 = 4− 3x3 + 2x4 ⇒ x =

x1

x2

x3

x4

=

4

00

+ x3

−3

10

+ x4

2

01

x2 = 0− 1x3 + 3x4 ⇒ x =

x1

x2

x3

x4

=

4000

+ x3

−3−110

+ x4

2301

This final form of a typical solution is especially pleasing and useful. For example, wecan build solutions quickly by choosing values for our free variables, and then computea linear combination. Such as

x3 = 2, x4 = −5 ⇒ x =

x1

x2

x3

x4

=

4000

+ (2)

−3−110

+ (−5)

2301

=

−12−172−5

or,

x3 = 1, x4 = 3 ⇒ x =

x1

x2

x3

x4

=

4000

+ (1)

−3−110

+ (3)

2301

=

7813

.

You’ll find the second solution listed in the write-up for Archetype D [393], and youmight check the first solution by substituting it back into the original equations.

While this form is useful for quickly creating solutions, its even better because it tellsus exactly what every solution looks like. We know the solution set is infinite, which is

pretty big, but now we can say that a solution is some multiple of

−3−110

plus a multiple

of

2301

plus the fixed vector

4000

. Period. So it only takes us three vectors to describe the

entire infinite solution set, provided we also agree on how to combine the three vectorsinto a linear combination. 4

Page 114: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.VFSS Vector Form of Solution Sets 82

We’ll now formalize the last example as a theorem.

Theorem VFSLSVector Form of Solutions to Linear SystemsSuppose that [A | b] is the augmented matrix for a consistent linear system LS(A, b) ofm equations in n variables. Denote the vector of variables as x = (xi). Let B = (bij)be a row-equivalent m×(n+1) matrix in reduced row-echelon form. Suppose that B has rnonzero rows, columns without leading 1’s having indices F = {f1, f2, f3, . . . , fn−r, n + 1},and columns with leading 1’s having indices D = {d1, d2, d3, . . . , dr}. Define vectorsc = (ci), uj = (uij), 1 ≤ j ≤ n− r of size n by

ci =

{0 if i ∈ F

bk,n+1 if i ∈ D, i = dk

uij =

1 if i ∈ F , i = fj

0 if i ∈ F , i 6= fj

−bk,fjif i ∈ D, i = dk

.

Then the system of equations represented by the vector equation

x =

x1

x2

x3...

xn

= c + xf1u1 + xf2u2 + xf3u3 + · · ·+ xfn−run−r

is equivalent to LS(A, b). �

Proof We are being asked to prove that two systems of equations are equivalent, thatis, they have identical solution sets. First, LS(A, b) is equivalent to the linear system ofequations that has the matrix B as its augmented matrix (Theorem REMES [25]). Wewill now show that the equations in the conclusion of the proof are either always true,or are simple rearrangements of the equations in the system with B as its augmentedmatrix. This will then establish that all three systems are equivalent to each other.

Suppose that i ∈ F , so i = fk for a particular choice of k, 1 ≤ k ≤ n − r. Considerthe equation given by entry i of the vector equality.

xi =ci + xf1ui1 + xf2ui2 + xf3ui3 + · · ·+ xfn−rui,n−r

xi =cfk+ xf1ufk1 + xf2ufk2 + xf3ufk3 + · · ·+ xfk

ufkfk+ · · ·+ xfn−rufk,n−r

xi =0 + xf1(0) + xf2(0) + xf3(0) + · · ·+ xfk(1) + · · ·+ xfn−r(0) = xfk

xi =xi.

This means that equality of the two vectors in entry i represents the equation xi = xi

when i ∈ F . Since this equation is always true, it does not restrict the possibilities forthe solution set.

Now consider the i-th entry, when i ∈ D, and suppose that i = dk, for some particularchoice of k, 1 ≤ k ≤ r. Consider the equation given by entry i of the vector equality.

xi =ci + xf1ui1 + xf2ui2 + xf3ui3 + · · ·+ xfn−rui,n−r

xi =cdk+ xf1udk1 + xf2udk2 + xf3udk3 + · · ·+ xfn−rudk,n−r

xi =bk,n+1 + xf1(−bk,f1) + xf2(−bk,f2) + xf3(−bk,f3) + · · ·+ xfn−r(−bk,fn−r)

xi =bk,n+1 −(bk,f1xf1 + bk,f2xf2 + bk,f3xf3 + · · ·+ bk,fn−rxfn−r

)

Page 115: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.VFSS Vector Form of Solution Sets 83

Rearranging, this becomes,

xi + bk,f1xf1 + bk,f2xf2 + bk,f3xf3 + · · ·+ bk,fn−rxfn−r = bk,n+1.

This is exactly the equation represented by row k of the matrix B. So the equationsrepresented by the vector equality in the conclusion are exactly the equations representedby the matrix B, along with additional equations of the form xi = xi that are alwaystrue. So the solution sets will be identical. �

Theorem VFSLS [82] formalizes what happened in the three steps of Example SS.VFSAD [80].The theorem will be useful in proving other theorems, and it it is useful since it tells usan exact procedure for simply describing an infinite solution set. We could program acomputer to implement it, once we have the augmented matrix row-reduced and havechecked that the system is consistent. By Knuth’s definition, this completes our con-version of linear equation solving from art into science. Notice that it even applies (butis overkill) in the case of a unique solution. However, as a practical matter, I preferthe three-step process of Example SS.VFSAD [80] when I need to describe an infinitesolution set. So lets practice some more, but with a bigger example.

Example SS.VFSAIVector form of solutions for Archetype IArchetype I [414] is a linear system of m = 4 equations in n = 7 variables. Row-reducingthe augmented matrix yields

1 4 0 0 2 1 −3 4

0 0 1 0 1 −3 5 2

0 0 0 1 2 −6 6 10 0 0 0 0 0 0 0

and we see r = 3 nonzero rows. The columns with leading 1’s are D = {1, 3, 4} sothe r dependent variables are x1, x3, x4. The columns without leading 1’s are F ={2, 5, 6, 7, 8}, so the n− r = 4 free variables are x2, x5, x6, x7.

Step 1. Write the vector of variables (x) as a fixed vector (c), plus a linear combinationof n− r = 4 vectors (u1, u2, u3, u4), using the free variables as the scalars.

x =

x1

x2

x3

x4

x5

x6

x7

=

+ x2

+ x5

+ x6

+ x7

Step 2. For each free variable, use 0’s and 1’s to ensure equality for the correspondingentry of the the vectors. Take note of the pattern of 0’s and 1’s at this stage, because thisis the best look you’ll have at it. We’ll state an important theorem in the next section

Page 116: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.VFSS Vector Form of Solution Sets 84

and the proof will essentially rely on this observation.

x =

x1

x2

x3

x4

x5

x6

x7

=

0

000

+ x2

1

000

+ x5

0

100

+ x6

0

010

+ x7

0

001

Step 3. For each dependent variable, use the augmented matrix to formulate an equationexpressing the dependent variable as a constant plus multiples of the free variables.Convert this equation into entries of the vectors that ensure equality for each dependentvariable, one at a time.

x1 = 4− 4x2 − 2x5 − 1x6 + 3x7 ⇒

x =

x1

x2

x3

x4

x5

x6

x7

=

40

000

+ x2

−41

000

+ x5

−20

100

+ x6

−10

010

+ x7

30

001

x3 = 2 + 0x2 − x5 + 3x6 − 5x7 ⇒

x =

x1

x2

x3

x4

x5

x6

x7

=

402

000

+ x2

−410

000

+ x5

−20−1

100

+ x6

−103

010

+ x7

30−5

001

x4 = 1 + 0x2 − 2x5 + 6x6 − 6x7 ⇒

x =

x1

x2

x3

x4

x5

x6

x7

=

4021000

+ x2

−4100000

+ x5

−20−1−2100

+ x6

−1036010

+ x7

30−5−6001

We can now use this final expression to quickly build solutions to the system. You mighttry to recreate each of the solutions listed in the write-up for Archetype I [414]. (Hint:look at the values of the free variables in each solution, and notice that the vector c has0’s in these locations.)

Even better, we have a description of the infinite solution set, based on just 5 vectors,which we combine in linear combinations to produce solutions.

Page 117: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.VFSS Vector Form of Solution Sets 85

Whenever we discuss Archetype I [414] you know that’s your cue to go work throughArchetype J [418] by yourself. Remember to take note of the 0/1 pattern at the conclusionof Step 2. Have fun — we won’t go anywhere while you’re away. 4This technique is so important, that we’ll do one more example. However, an importantdistinction will be that this system is homogeneous.

Example SS.VFSALVector form of solutions for Archetype LArchetype L [427] is presented simply as the 5× 5 matrix

L =

−2 −1 −2 −4 4−6 −5 −4 −4 610 7 7 10 −13−7 −5 −6 −9 10−4 −3 −4 −6 6

We’ll interpret it here as the coefficient matrix of a homogeneous system and referencethis matrix as L. So we are solving the homogeneous system LS(L, 0) having m = 5equations in n = 5 variables. If we built the augmented matrix, we would a sixth columnto L containing all zeros. As we did row operations, this sixth column would remain allzeros. So instead we will row-reduce the coefficient matrix, and mentally remember themissing sixth column of zeros. This row-reduced matrix is

1 0 0 1 −2

0 1 0 −2 2

0 0 1 2 −10 0 0 0 00 0 0 0 0

and we see r = 3 nonzero rows. The columns with leading 1’s are D = {1, 2, 3} so the rdependent variables are x1, x2, x3. The columns without leading 1’s are F = {4, 5}, sothe n− r = 2 free variables are x4, x5. Notice that if we had included the all-zero vectorof constants to form the augmented matrix for the system, then the index 6 would haveappeared in the set F , and subsequently would have been ignored when listing the freevariables.

Step 1. Write the vector of variables (x) as a fixed vector (c), plus a linear combinationof n− r = 2 vectors (u1, u2), using the free variables as the scalars.

x =

x1

x2

x3

x4

x5

=

+ x4

+ x5

Step 2. For each free variable, use 0’s and 1’s to ensure equality for the correspondingentry of the the vectors. Take note of the pattern of 0’s and 1’s at this stage, even if itis not as illuminating as in other examples.

x =

x1

x2

x3

x4

x5

=

00

+ x4

10

+ x5

01

Page 118: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SSV Span of a Set of Vectors 86

Step 3. For each dependent variable, use the augmented matrix to formulate an equationexpressing the dependent variable as a constant plus multiples of the free variables. Don’tforget about the “missing” sixth column being full of zeros. Convert this equation intoentries of the vectors that ensure equality for each dependent variable, one at a time.

x1 = 0− 1x4 + 2x5 ⇒ x =

x1

x2

x3

x4

x5

=

0

00

+ x4

−1

10

+ x5

2

01

x2 = 0 + 2x4 − 2x5 ⇒ x =

x1

x2

x3

x4

x5

=

00

00

+ x4

−12

10

+ x5

2−2

01

x3 = 0− 2x4 + 1x5 ⇒ x =

x1

x2

x3

x4

x5

=

00000

+ x4

−12−210

+ x5

2−2101

The vector c will always have 0’s in the entries corresponding to free variables. However,since we are solving a homogeneous system, the row-reduced augmented matrix has zerosin column n + 1 = 6, and hence all the entries of c are zero. So we can write

x =

x1

x2

x3

x4

x5

= 0 + x4

−12−210

+ x5

2−2101

= x4

−12−210

+ x5

2−2101

It will always happen that the solutions to a homogeneous system has c = 0 (even in thecase of a unique solution?). So our expression for the solutions is a bit more pleasing.In this example it says that the solutions are all possible linear combinations of the two

vectors u1 =

−12−210

and u2 =

2−2101

, with no mention of any fixed vector entering into

the linear combination.This observation will motivate our next subsection and definition, and after that we’ll

conclude the section by formalizing this situation. 4

Subsection SSVSpan of a Set of Vectors

In Example SS.VFSAL [85] we saw the solution set of a homogeneous system describedas all possible linear combinations of two particular vectors. This happens to be a useful

Page 119: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SSV Span of a Set of Vectors 87

way to construct or describe infinite sets of vectors, so we encapsulate this idea in adefinition.

Definition SSCVSpan of a Set of Column VectorsGiven a set of vectors S = {u1, u2, u3, . . . , ut}, their span, Sp (S), is the set of allpossible linear combinations of u1, u2, u3, . . . , ut. Symbolically,

Sp (S) = {α1u1 + α2u2 + α3u3 + · · ·+ αtut | αi ∈ C, 1 ≤ i ≤ t}

=

{t∑

i=1

αiui

∣∣∣∣∣ αi ∈ C, 1 ≤ i ≤ t

}�

The span is just a set of vectors, though in all but one situation it is an infinite set. (Justwhen is it not infinite?) So we start with a finite collection of vectors (t of them to beprecise), and use this finite set to describe an infinite set of vectors. We will see thisconstruction repeatedly, so lets work through some examples to get comfortable with it.The most obvious question about a set is if a particular item of the correct type is in theset, or not.

Example SS.SCAASpan of the columns of Archetype ABegin with the finite set of three vectors of size 3

S = {u1, u2, u3} =

1

21

,

−111

,

210

and consider the infinite set U = Sp (S). The vectors of S could have been chosen to beanything, but for reasons that will become clear later, we have chosen the three columnsof the coefficient matrix in Archetype A [379]. First, as an example, note that

v = (5)

121

+ (−3)

−111

+ (7)

210

=

22142

is in Sp (S), since it is a linear combination of u1, u2, u3. We write this succinctly asv ∈ Sp (S). There is nothing magical about the scalars α1 = 5, α2 = −3, α3 = 7, theycould have been chosen to be anything. So repeat this part of the example yourself, usingdifferent values of α1, α2, α3. What happens if you choose all three scalars to be zero?

So we know how to quickly construct sample elements of the set Sp (S). A slightlydifferent question arises when you are handed a vector of the correct size and asked if it is

an element of Sp (S). For example, is w =

185

in Sp (S)? More succinctly, w ∈ Sp (S)?

To answer this question, we will look for scalars α1, α2, α3 so that

α1u1 + α2u2 + α3u3 = w.

By Theorem SLSLC [76] solutions to this vector equality are solutions to the system ofequations

α1 − α2 + 2α3 = 1

2α1 + α2 + α3 = 8

α1 + α2 = 5.

Page 120: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SSV Span of a Set of Vectors 88

Building the augmented matrix for this linear system, and row-reducing, gives 1 0 1 3

0 1 −1 20 0 0 0

.

This system has infinitely many solutions (there’s a free variable), but all we need is one.The solution,

α1 = 2 α2 = 3 α3 = 1

tells us that(2)u1 + (3)u2 + (1)u3 = w

so we are convinced that w really is in Sp (S). Lets ask the same question again, but

this time with y =

243

, i.e. is y ∈ Sp (S)?

So we’ll look for scalars α1, α2, α3 so that

α1u1 + α2u2 + α3u3 = y.

By Theorem SLSLC [76] this linear combination becomes the system of equations

α1 − α2 + 2α3 = 2

2α1 + α2 + α3 = 4

α1 + α2 = 3.

Building the augmented matrix for this linear system, and row-reducing, gives 1 0 1 0

0 1 −1 0

0 0 0 1

This system is inconsistent (there’s a leading 1 in the last column), so there are noscalars α1, α2, α3 that will create a linear combination of u1, u2, u3 that equals y. Moreprecisely, y 6∈ Sp (S).

There are three things to observe in this example. (1) It is easy to construct vectorsin Sp (S). (2) It is possible that some vectors are in Sp (S) (e.g. w), while others are not(e.g. y). (3) Deciding if a given vector is in Sp (S) leads to solving a linear system ofequations and asking if the system is consistent.

With a computer program in hand to solve systems of linear equations, could youcreate a program to decide if a vector was, or wasn’t, in the span of a given set ofvectors? Is this art or science?

This example was built on vectors from the columns of the coefficient matrix ofArchetype A [379]. Study the determination that v ∈ Sp (S) and see if you can connectit with some of the other properties of Archetype A [379]. 4

Lets do a similar example to Example SS.SCAA [87], only now with Archetype B [384].

Page 121: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SSV Span of a Set of Vectors 89

Example SS.SCABSpan of the columns of Archetype BBegin with the finite set of three vectors of size 3

R = {v1, v2, v3} =

−7

51

,

−650

,

−1274

and consider the infinite set V = Sp (R). The vectors of R have been chosen as the threecolumns of the coefficient matrix in Archetype B [384]. First, as an example, note that

x = (2)

−751

+ (4)

−650

+ (−3)

−1274

=

−29−10

is in Sp (R), since it is a linear combination of v1, v2, v3. In other words, x ∈ Sp (R).Try some different values of α1, α2, α3 yourself, and see what vectors you can create aselements of Sp (R).

Now ask if a given vector is an element of Sp (R). For example, is z =

−33245

in

Sp (R)? Is z ∈ Sp (R)?To answer this question, we will look for scalars α1, α2, α3 so that

α1v1 + α2v2 + α3v3 = z.

By Theorem SLSLC [76] this linear combination becomes the system of equations

−7α1 − 6α2 − 12α3 = −33

5α1 + 5α2 + 7α3 = 24

α1 + 4α3 = 5.

Building the augmented matrix for this linear system, and row-reducing, gives 1 0 0 −3

0 1 0 5

0 0 1 2

.

This system has a unique solution,

α1 = −3 α2 = 5 α3 = 2

telling us that(−3)v1 + (5)v2 + (2)v3 = z

so we are convinced that z really is in Sp (R).There is no point in replacing z with another vector and doing this again. A question

about membership in Sp (R) inevitably leads to a system of three equations in the threevariables α1, α2, α3 with a coefficient matrix whose columns are the vectors v1, v2, v3.This particular coefficient matrix is nonsingular, so by Theorem NSMUS [59], it is guar-anteed to have a solution. (This solution is unique, but that’s not important here.) So no

Page 122: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SSNS Spanning Sets of Null Spaces 90

matter which vector we might have chosen for z, we would have been certain to discoverthat it was an element of Sp (R). Stated differently, every vector of size 3 is in Sp (R),or Sp (R) = C3.

Compare this example with Example SS.SCAA [87], and see if you can connect zwith some aspects of the write-up for Archetype B [384]. 4

Subsection SSNSSpanning Sets of Null Spaces

We saw in Example SS.VFSAL [85] that when a system of equations is homogeneous thesolution set can be expressed in the form described by Theorem VFSLS [82] where thevector c is the zero vector. We can essentially ignore this vector, so that the remainderof the typical expression for a solution looks like an arbitrary linear combination, wherethe scalars are the free variables and the vectors are u1, u2, u3, . . . , un−r. Which soundsa lot like a span. This is the substance of the next theorem.

Theorem SSNSSpanning Sets for Null SpacesSuppose that A is an m × n matrix, and B is a row-equivalent matrix in reduced row-echelon form with r nonzero rows. Let D = {d1, d2, d3, . . . , dr} and F = {f1, f2, f3, . . . , fn−r}be the sets of column indices where B does and does not (respectively) have leading 1’s.Construct the n− r vectors uj = (uij), 1 ≤ j ≤ n− r of size n as

uij =

1 if i ∈ F , i = fj

0 if i ∈ F , i 6= fj

−bk,fjif i ∈ D, i = dk

.

Then the null space of A is given by

N (A) = Sp ({u1, u2, u3, . . . , un−r}) . �

Proof Consider the homogeneous system with A as a coefficient matrix, LS(A, 0). Itsset of solutions is, by definition, the null space of A, N(A). Row-reducing the augmentedmatrix of this homogeneous system will create the row-equivalent matrix B′. Row-reducing the augmented matrix that has a final column of all zeros, yields B′, which isthe matrix B, along with an additional column (index n + 1) that is still totally zero.

Now apply Theorem VFSLS [82], noting that our homogeneous system is consistent(Theorem HSC [50]). The vector c has zeros for each entry that corresponds to an indexin F . For entries that correspond to an index in D, the value is −b′k,n+1, but for B′

these entries in the final column are all zero. So c = 0. This says that a solution of thehomogeneous system is of the form

x = c+xf1u1 +xf2u2 +xf3u3 + · · ·+xfn−run−r = xf1u1 +xf2u2 +xf3u3 + · · ·+xfn−run−r

where the free variables xfjcan each take on any value. Rephrased this says

N (A) ={

xf1u1 + xf2u2 + xf3u3 + · · ·+ xfn−run−r

∣∣ xf1 , xf2 , xf3 , . . . , xfn−r ∈ C}

= Sp ({u1, u2, u3, . . . , un−r}) . �

Page 123: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SSNS Spanning Sets of Null Spaces 91

Here’s an example that will exercise the span construction and Theorem SSNS [90], whilealso pointing the way to the next section.

Example SS.SCADSpan of the columns of Archetype DBegin with the set of four vectors of size 3

T = {w1, w2, w3, w4} =

2−31

,

141

,

7−54

,

−7−6−5

and consider the infinite set W = Sp (T ). The vectors of T have been chosen as the fourcolumns of the coefficient matrix in Archetype D [393]. Check that the vector

u2 =

2301

is a solution to the homogeneous system LS(D, 0) (it is the second vector of the spanningset for the null space of the coefficient matrix D, as described in Theorem SSNS [90]).Applying Theorem SLSLC [76], we can write the linear combination,

2w1 + 3w2 + 0w3 + 1w4 = 0

which we can solve for w4,w4 = (−2)w1 + (−3)w2.

This equation says that whenever we encounter the vector w4, we can replace it with aspecific linear combination of the vectors w1 and w2. So using w4 in the set T , alongwith w1 and w2, is excessive. An example of what we mean here can be illustrated bythe computation,

5w1 + (−4)w2 + 6w3 + (−3)w4 = 5w1 + (−4)w2 + 6w3 + (−3) ((−2)w1 + (−3)w2)

= 5w1 + (−4)w2 + 6w3 + (6w1 + 9w2)

= 11w1 + 5w2 + 6w3.

So what began as a linear combination of the vectors w1, w2, w3, w4 has been reducedto a linear combination of the vectors w1, w2, w3. A careful proof using our definitionof set equality (Technique SE [14]) would now allow us to conclude that this reductionis possible for any vector in W , so

W = Sp ({w1, w2, w3}) .

So the span of our set of vectors, W , has not changed, but we have described it by thespan of a set of three vectors, rather than four. Furthermore, we can achieve yet another,similar, reduction.

Check that the vector

u1 =

−3−110

Page 124: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SSNS Spanning Sets of Null Spaces 92

is a solution to the homogeneous system LS(D, 0) (it is the first vector of the spanningset for the null space of the coefficient matrix D, as described in Theorem SSNS [90]).Applying Theorem SLSLC [76], we can write the linear combination,

(−3)w1 + (−1)w2 + 1w3 = 0

which we can solve for w3,w3 = 3w1 + 1w2.

This equation says that whenever we encounter the vector w3, we can replace it with aspecific linear combination of the vectors w1 and w2. So, as before, the vector w3 is notneeded in the description of W , provided we have w1 and w2 available. In particular, acareful proof would show that

W = Sp ({w1, w2}) .

So W began life as the span of a set of four vectors, and we have now shown (utilizingsolutions to a homogeneous system) that W can also be described as the span of a set ofjust two vectors. Convince yourself that we cannot go any further. In other words, it isnot possible to dismiss either w1 or w2 in a similar fashion and winnow the set down tojust one vector.

What was it about the original set of four vectors that allowed us to declare certainvectors as surplus? And just which vectors were we able to dismiss? And why did we haveto stop once we had two vectors remaining? The answers to these questions motivate“linear independence,” our next section and next definition, and so are worth consideringcarefully now. 4

Page 125: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.EXC Exercises 93

Subsection EXCExercises

C40 Contributed by Robert Beezer

Suppose that S =

2−134

,

32−21

. Let W = Sp (S) and let x =

58−12−5

. Is x ∈ W?

If so, provide an explicit linear combination that demonstrates this. Solution [94]

C41 Contributed by Robert Beezer

Suppose that S =

2−134

,

32−21

. Let W = Sp (S) and let y =

5135

. Is y ∈ W? If

so, provide an explicit linear combination that demonstrates this. Solution [94]

C60 Contributed by Robert BeezerFor the matrix A below, find a set of vectors S so that (1) S is linearly independent,

and (2) the span of S equals the null space of A, Sp (S) = N (A).

A =

1 1 6 −81 −2 0 1−2 1 −6 7

Solution [95]

Page 126: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SOL Solutions 94

Subsection SOLSolutions

C40 Exercise [93] Contributed by Robert BeezerRephrasing the question, we want to know if there are scalars α1 and α2 such that

α1

2−134

+ α2

32−21

=

58−12−5

Theorem SLSLC [76] allows us to rephrase the question again as a quest for solutions tothe system of four equations in two unknowns with an augmented matrix given by

2 3 5−1 2 83 −2 −124 1 −5

This matrix row-reduces to

1 0 −2

0 1 30 0 00 0 0

From the form of this matrix, we can see that α1 = −2 and α2 = 3 is an affirmativeanswer to our question. More convincingly,

(−2)

2−134

+ (3)

32−21

=

58−12−5

C41 Exercise [93] Contributed by Robert BeezerRephrasing the question, we want to know if there are scalars α1 and α2 such that

α1

2−134

+ α2

32−21

=

5135

Theorem SLSLC [76] allows us to rephrase the question again as a quest for solutions tothe system of four equations in two unknowns with an augmented matrix given by

2 3 5−1 2 13 −2 34 1 5

Page 127: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SS.SOL Solutions 95

This matrix row-reduces to 1 0 0

0 1 0

0 0 10 0 0

With a leading 1 in the last column of this matrix (Theorem RCLS [41]) we can see thatthe system of equations has no solution, so there are no values for α1 and α2 that willallow us to conclude that y is in W . So y 6∈ W .

C60 Exercise [93] Contributed by Robert BeezerTheorem BNS [104] says that if we find the vector form of the solutions to the homoge-neous system LS(A, 0), then the fixed vectors (one per free variable) will have the desiredproperties. Row-reduce A, viewing it as the augmented matrix of a homogeneous systemwith an invisible columns of zeros as the last column, 1 0 4 −5

0 1 2 −30 0 0 0

Moving to the vector form of the solutions (Theorem VFSLS [82]), with free variables x3

and x4, solutions to the consistent system (it is homogeneous, Theorem HSC [50]) canbe expressed as

x1

x2

x3

x4

= x3

−4−210

+ x4

5301

Then with S given by

S =

−4−210

,

5301

Theorem SSNS [90] and Theorem BNS [104] guarantee the set has the desired properties.

Page 128: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section LI Linear Independence 96

Section LI

Linear Independence

Subsection LIVLinearly Independent Vectors

Theorem SLSLC [76] tells us that a solution to a homogeneous system of equations is alinear combination of the columns of the coefficient matrix that equals the zero vector.We used just this situation to our advantage (twice!) in Example SS.SCAD [91] wherewe reduced the set of vectors used in a span construction from four down to two, bydeclaring certain vectors as surplus. The next two definitions will allow us to formalizethis situation.

Definition RLDCVRelation of Linear Dependence for Column VectorsGiven a set of vectors S = {u1, u2, u3, . . . , un}, an equation of the form

α1u1 + α2u2 + α3u3 + · · ·+ αnun = 0

is a relation of linear dependence on S. If this equation is formed in a trivial fashion,i.e. αi = 0, 1 ≤ i ≤ n, then we say it is a trivial relation of linear dependence onS. �

Definition LICVLinear Independence of Column VectorsThe set of vectors S = {u1, u2, u3, . . . , un} is linearly dependent if there is a relationof linear dependence on S that is not trivial. In the case where the only relation of lineardependence on S is the trivial one, then S is a linearly independent set of vectors. �

Notice that a relation of linear dependence is an equation. Though most of it is a linearcombination, it is not a linear combination (that would be a vector). Linear independenceis a property of a set of vectors. It is easy to take a set of vectors, and an equal numberof scalars, all zero, and form a linear combination that equals the zero vector. When theeasy way is the only way, then we say the set is linearly independent. Here’s a couple ofexamples.

Example LI.LDSLinearly dependent set in C5

Consider the set of n = 4 vectors from C5,

S =

2−1312

,

12−152

,

21−361

,

−67−101

.

Page 129: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.LIV Linearly Independent Vectors 97

To determine linear independence we first form a relation of linear dependence,

α1

2−1312

+ α2

12−152

+ α3

21−361

+ α4

−67−101

= 0.

We know that α1 = α2 = α3 = α4 = 0 is a solution to this equation, but that is ofno interest whatsoever. That is always the case, no matter what four vectors we mighthave chosen. We are curious to know if there are other, nontrivial, solutions. Theo-rem SLSLC [76] tells us that we can find such solutions as solutions to the homogeneoussystem LS(A, 0) where the coefficient matrix has these four vectors as columns,

A =

2 1 2 −6−1 2 1 73 −1 −3 −11 5 6 02 2 1 1

.

Row-reducing this coefficient matrix yields,

1 0 0 −2

0 1 0 4

0 0 1 −30 0 0 00 0 0 0

.

We could solve this homogeneous system completely, but for this example all we needis one nontrivial solution. Setting the lone free variable to any nonzero value, such asx4 = 1, yields the nontrivial solution

x =

2−431

.

completing our application of Theorem SLSLC [76], we have

2

2−1312

+ (−4)

12−152

+ 3

21−361

+ 1

−67−101

= 0.

This is a relation of linear dependence on S that is not trivial, so we conclude that S islinearly dependent. 4

Example LI.LISLinearly independent set in C5

Page 130: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.LIV Linearly Independent Vectors 98

Consider the set of n = 4 vectors from C5,

T =

2−1312

,

12−152

,

21−361

,

−67−111

.

To determine linear independence we first form a relation of linear dependence,

α1

2−1312

+ α2

12−152

+ α3

21−361

+ α4

−67−111

= 0.

We know that α1 = α2 = α3 = α4 = 0 is a solution to this equation, but that is ofno interest whatsoever. That is always the case, no matter what four vectors we mighthave chosen. We are curious to know if there are other, nontrivial, solutions. Theo-rem SLSLC [76] tells us that we can find such solutions as solution to the homogeneoussystem LS(B, 0) where the coefficient matrix has these four vectors as columns,

B =

2 1 2 −6−1 2 1 73 −1 −3 −11 5 6 12 2 1 1

.

Row-reducing this coefficient matrix yields,1 0 0 0

0 1 0 0

0 0 1 0

0 0 0 10 0 0 0

.

From the form of this matrix, we see that there are no free variables, so the solutionis unique, and because the system is homogeneous, this unique solution is the trivialsolution. So we now know that there is but one way to combine the four vectors of Tinto a relation of linear dependence, and that one way is the easy and obvious way. Inthis situation we say that the set, T , is linearly independent. 4

Example LI.LDS [96] and Example LI.LIS [97] relied on solving a homogeneous system ofequations to determine linear independence. We can codify this process in a time-savingtheorem.

Theorem LIVHSLinearly Independent Vectors and Homogeneous SystemsSuppose that A is an m×n matrix and S = {A1, A2, A3, . . . , An} is the set of vectorsin Cm that are the columns of A. Then S is a linearly independent set if and only if thehomogeneous system LS(A, 0) has a unique solution. �

Page 131: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.LIV Linearly Independent Vectors 99

Proof (⇐) Suppose that LS(A, 0) has a unique solution. Since it is a homogeneoussystem, this solution must be the trivial solution x = 0. By Theorem SLSLC [76], thismeans that the only relation of linear dependence on S is the trivial one. So S is linearlyindependent.

(⇒) We will prove the contrapositive. Suppose that LS(A, 0) does not have a uniquesolution. Since it is a homogeneous system, it is consistent (Theorem HSC [50]), and somust have infinitely many solutions (Theorem PSSLS [44]). One of these infinitely manysolutions must be nontrivial (in fact, almost all of them are), so choose one. By Theo-rem SLSLC [76] this nontrivial solution will give a nontrivial relation of linear dependenceon S, so we can conclude that S is a linearly dependent set. �

Since Theorem LIVHS [98] is an equivalence, we can use it to determine the linearindependence or dependence of any set of column vectors, just by creating a correspondingmatrix and analyzing the row-reduced form. Let’s illustrate this with another example.

Example LI.LLDSLarge linearly dependent set in C4

Consider the set of n = 9 vectors from C4,

R =

−1312

,

71−36

,

12−1−2

,

0429

,

5−243

,

21−64

,

30−31

,

1153

,

−6−111

.

To employ Theorem LIVHS [98], we form a 4× 9 coefficient matrix, C,

C =

−1 7 1 0 5 2 3 1 −63 1 2 4 −2 1 0 1 −11 −3 −1 2 4 −6 −3 5 12 6 −2 9 3 4 1 3 1

.

To determine if the homogeneous system LS(C, 0) has a unique solution or not, we wouldnormally row-reduce this matrix. But in this particular example, we can do better.Theorem HMVEI [51] tells us that since the system is homogeneous with n = 9 variablesin m = 4 equations, and n > m, there must be infinitely many solutions. Since there isnot a unique solution, Theorem LIVHS [98] says the set is linearly dependent. 4

The situation in Example LI.LLDS [99] is slick enough to warrant formulating as atheorem.

Theorem MVSLDMore Vectors than Size implies Linear DependenceSuppose that S = {u1, u2, u3, . . . , un} is the set of vectors in Cm, and that n > m.Then S is a linearly dependent set. �

Proof Form the m×n coefficient matrix A that has the column vectors ui, 1 ≤ i ≤ n asits columns. Consider the homogeneous system LS(A, 0). By Theorem HMVEI [51] thissystem has infinitely many solutions. Since the system does not have a unique solution,Theorem LIVHS [98] says the columns of A form a linearly dependent set, which is thedesired conclusion. �

As an equivalence, Theorem LIVHS [98] gives us a straightforward way to determine ifa set of vectors is linearly independent or dependent. We can improve on it just slightly,so we will state the result as a corollary.

Page 132: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.LDSS Linearly Dependent Sets and Spans 100

Corollary LIVRNLinearly Independent Vectors, r and nSuppose that A is an m× n matrix and S = {A1, A2, A3, . . . , An} is the set of vectorsin Cm that are the columns of A. Let B be a matrix in reduced row-echelon form thatis row-equivalent to A and let r denote the number of non-zero rows in B. Then S islinearly independent if and only if n = r. �

Proof Theorem LIVHS [98] says the linear independence of S is equivalent to the homo-geneous linear system LS(A, 0) having a unique solution. Since LS(A, 0) is consistent(Theorem HSC [50]) we can apply Theorem CSRN [42] to see that the solution is uniqueexactly when n = r. �

Subsection LDSSLinearly Dependent Sets and Spans

In any linearly dependent set there is always one vector that can be written as a linearcombination of the others. This is the substance of the upcoming Theorem DLDS [100].Perhaps this will explain the use of the word “dependent.” In a linearly dependent set,at least one vector “depends” on the others (via a linear combination).

If we use a linearly dependent set to construct a span, then we can always create thesame infinite set with a starting set that is one vector smaller in size. We will illustratethis behavior in Example LI.RS [101]. However, this will not be possible if we build aspan from a linearly independent set. So in a certain sense, using a linearly independentset to formulate a span is the best possible way to go about it — there aren’t any extravectors being used to build up all the necessary linear combinations. OK, here’s thetheorem, and then the example.

Theorem DLDSDependency in Linearly Dependent SetsSuppose that S = {u1, u2, u3, . . . , un} is a set of vectors. Then S is a linearly dependentset if and only if there is an index t, 1 ≤ t ≤ n such that ut is a linear combination ofthe vectors u1, u2, u3, . . . , ut−1, ut+1, . . . , un. �

Proof (⇒) Suppose that S is linearly dependent, so there is a nontrivial relation oflinear dependence,

α1u1 + α2u2 + α3u3 + · · ·+ αnun = 0.

Since the αi cannot all be zero, choose one, say αt, that is nonzero. Then,

−αtut =α1u1 + α2u2 + α3u3 + · · ·+ αt−1ut−1 + αt+1ut+1 + · · ·+ αnun

and we can multiply by −1αt

since αt 6= 0,

ut =−α1

αt

u1 +−α2

αt

u2 +−α3

αt

u3 + · · ·+ −αt−1

αt

ut−1 +−αt+1

αt

ut+1 + · · ·+ −αn

αt

un.

Page 133: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.LDSS Linearly Dependent Sets and Spans 101

Since the values of αi

αtare again scalars, we have expressed ut as the desired linear

combination.(⇐) Suppose that the vector ut is a linear combination of the other vectors in S.

Write this linear combination as

β1u1 + β2u2 + β3u3 + · · ·+ βt−1ut−1 + βt+1ut+1 + · · ·+ βnun = ut

and move ut to the other side of the equality

β1u1 + β2u2 + β3u3 + · · ·+ βt−1ut−1 + (−1)ut + βt+1ut+1 + · · ·+ βnun = 0.

Then the scalars β1, β2, β3, . . . , βt = −1, . . . , βn provide a nontrivial linear combinationof the vectors in S, thus establishing that S is a linearly dependent set. �

This theorem can be used, sometimes repeatedly, to whittle down the size of a setof vectors used in a span construction. We have seen some of this already in Exam-ple SS.SCAD [91], but in the next example we will detail some of the subtleties.

Example LI.RSReducing a span in C5

Consider the set of n = 4 vectors from C5,

R = {v1, v2, v3, v4} =

12−132

,

21312

,

0−76−11−2

,

41216

and define V = Sp (R).To employ Theorem LIVHS [98], we form a 5× 4 coefficient matrix, D,

D =

1 2 0 42 1 −7 1−1 3 6 23 1 −11 12 2 −2 6

and row-reduce to understand solutions to the homogeneous system LS(D, 0),

1 0 0 4

0 1 0 0

0 0 1 10 0 0 00 0 0 0

.

We can find infinitely many solutions to this system, most of them nontrivial, and wechoose anyone we like to build a relation of linear dependence on R. Lets begin withx4 = 1, to find the solution

−40−11

.

Page 134: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.LINSM Linear Independence and NonSingular Matrices 102

So we can write the relation of linear dependence,

(−4)v1 + 0v2 + (−1)v3 + 1v4 = 0.

Theorem DLDS [100] guarantees that we can solve this relation of linear dependence forsome vector in R, but the choice of which one is up to us. Notice however that v2 has azero coefficient. In this case, we cannot choose to solve for v2. Maybe some other relationof linear dependence would produce a nonzero coefficient for v2 if we just had to solve forthis vector. Unfortunately, this example has been engineered to always produce a zerocoefficient here, as you can see from solving the homogeneous system. Every solution hasx2 = 0!

OK, if we are convinced that we cannot solve for v2, lets instead solve for v3,

v3 = (−4)v1 + 0v2 + 1v4 = (−4)v1 + 1v4.

We now claim that this particular equation will allow us to write

V = Sp (R) = Sp ({v1, v2, v3, v4}) = Sp ({v1, v2, v4})

in essence declaring v3 as surplus for the task of building V as a span. This claim isan equality of two sets, so we will use Technique SE [14] to establish it carefully. LetR′ = {v1, v2, v4} and V ′ = Sp (R′). We want to show that V = V ′.

First show that V ′ ⊆ V . Since every vector of R′ is in R, any vector we can constructin V ′ as a linear combination of vectors from R′ can also be constructed as a vector inV by the same linear combination of the same vectors in R. That was easy, now turn itaround.

Next show that V ⊆ V ′. Choose any v from V . Then there are scalars α1, α2, α3, α4

so that

v = α1v1 + α2v2 + α3v3 + α4v4

= α1v1 + α2v2 + α3 ((−4)v1 + 1v4) + α4v4

= α1v1 + α2v2 + ((−4α3)v1 + α3v4) + α4v4

= (α1 − 4α3)v1 + α2v2 + (α3 + α4)v4.

This equation says that v can then be written as a linear combination of the vectors inR′ and hence qualifies for membership in V ′. So V ⊆ V ′ and we have established thatV = V ′.

If R′ was also linearly dependent (its not), we could reduce the set even further.Notice that we could have chosen to eliminate any one of v1, v3 or v4, but somehow v2

is essential to the creation of V since it cannot be replaced by any linear combination ofv1, v3 or v4. 4

Subsection LINSMLinear Independence and NonSingular Matrices

We will now specialize to sets of n vectors from Cn. This will put Theorem MVSLD [99]off-limits, while Theorem LIVHS [98] will involve square matrices. Lets begin by con-trasting Archetype A [379] and Archetype B [384].

Page 135: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.LINSM Linear Independence and NonSingular Matrices 103

Example LI.LDCAALinearly dependent columns in Archetype AArchetype A [379] is a system of linear equations with coefficient matrix,

A =

1 −1 22 1 11 1 0

.

Do the columns of this matrix form a linearly independent or dependent set? By Exam-ple NSM.S [56] we know that A is singular. According to the definition of nonsingularmatrices, Definition NM [56], the homogeneous system LS(A, 0) has infinitely manysolutions. So by Theorem LIVHS [98], the columns of A form a linearly dependent set.4

Example LI.LICABLinearly independent columns in Archetype BArchetype B [384] is a system of linear equations with coefficient matrix,

B =

−7 −6 −125 5 71 0 4

.

Do the columns of this matrix form a linearly independent or dependent set? By Ex-ample NSM.NS [57] we know that B is nonsingular. According to the definition of non-singular matrices, Definition NM [56], the homogeneous system LS(A, 0) has a uniquesolution. So by Theorem LIVHS [98], the columns of B form a linearly independentset. 4

That Archetype A [379] and Archetype B [384] have opposite properties for the columnsof their coefficient matrices is no accident. Here’s the theorem, and then we will updateour equivalences for nonsingular matrices, Theorem NSME1 [62].

Theorem NSLICNonSingular matrices have Linearly Independent ColumnsSuppose that A is a square matrix. Then A is nonsingular if and only if the columns ofA form a linearly independent set. �

Proof This is a proof where we can chain together equivalences, rather than proving thetwo halves separately. By definition, A is nonsingular if and only if the homogeneoussystem LS(A, 0) has a unique solution. Theorem LIVHS [98] then says that the systemLS(A, 0) has a unique solution if and only if the columns of A are a linearly independentset. �

Here’s an update to Theorem NSME1 [62].

Theorem NSME2NonSingular Matrix Equivalences, Round 2Suppose that A is a square matrix. The following are equivalent.

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the zero vector, N (A) = {0}.

Page 136: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.NSSLI Null Spaces, Spans, Linear Independence 104

4. The linear system LS(A, b) has a unique solution for every possible choice of b.

5. The columns of A form a linearly independent set. �

Proof Theorem NSLIC [103] is yet another equivalence for a nonsingular matrix, so wecan add it to the list in Theorem NSME1 [62]. �

Subsection NSSLINull Spaces, Spans, Linear Independence

In Subsection SS.SSNS [90] we proved Theorem SSNS [90] which provided n− r vectorsthat could be used with the span construction to build the entire null space of a matrix.As we have seen in Theorem DLDS [100] and Example LI.RS [101], linearly dependentsets carry redundant vectors with them when used in building a set as a span. Ouraim now is to show that the vectors provided by Theorem SSNS [90] form a linearlyindependent set, so in one sense they are as efficient as possible a way to describe thenull space. Notice that the vectors uj, 1 ≤ j ≤ n − r first appear in the vector formof solutions to arbitrary linear systems (Theorem VFSLS [82]). The exact same vectorsappear again in the span construction in the conclusion of Theorem SSNS [90]. Sincethis second theorem specializes to homogeneous systems the main difference is that thevector c in Theorem VFSLS [82] is now the zero vector. Finally, Theorem BNS [104]now shows that these same vectors are a linearly independent set.

The proof is really quite straightforward, and relies on the “pattern” of zeros andones that arise in the vectors ui, 1 ≤ i ≤ n − r in the entries that correspond to thefree variables. So take a look at Example SS.VFSAD [80], Example SS.VFSAI [83] andExample SS.VFSAL [85], especially during the conclusion of Step 2 (temporarily ignorethe construction of the constant vector, c). It is a good exercise in showing how to provea conclusion that states a set is linearly independent.

Theorem BNSBasis for Null SpacesSuppose that A is an m × n matrix, and B is a row-equivalent matrix in reduced row-echelon form with r nonzero rows. Let D = {d1, d2, d3, . . . , dr} and F = {f1, f2, f3, . . . , fn−r}be the sets of column indices where B does and does not (respectively) have leading 1’s.Construct the n− r vectors uj = (uij), 1 ≤ j ≤ n− r of size n as

uij =

1 if i ∈ F , i = fj

0 if i ∈ F , i 6= fj

−bk,fjif i ∈ D, i = dk

.

Then the set {u1, u2, u3, . . . , un−r} is linearly independent. �

Proof To prove the linear independence of a set, we need to start with a relation oflinear dependence and somehow conclude that the scalars involved must all be zero, i.e.that the relation of linear dependence only happens in the trivial fashion. To this end,we start with

α1u1 + α2u2 + α3u3 + · · ·+ αn−run−r = 0.

Page 137: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.NSSLI Null Spaces, Spans, Linear Independence 105

For each j, 1 ≤ j ≤ n− r, consider the entry of the vectors on both sides of this equalityin position fj. On the right, it is easy since the zero vector has a zero in each entry. Onthe left we find,

α1ufj1 + α2ufj2 + α3ufj3 + · · ·+ αjufj ,j + · · ·+ αn−rufj ,n−r

= α1(0) + α2(0) + α3(0) + · · ·+ αj(1) + · · ·+ αn−r(0)

= αj

So for all j, 1 ≤ j ≤ n− r, we have αj = 0, which is the conclusion that tells us that theonly relation of linear dependence on {u1, u2, u3, . . . , un−r} is the trivial one, hence theset is linearly independent, as desired. �

Example LI.NSLILNull space spanned by linearly independent set, Archetype LIn Example SS.VFSAL [85] we previewed Theorem SSNS [90] by finding a set of two

vectors such that their span was the null space for the matrix in Archetype L [427].Writing the matrix as L, we have

N(L) = Sp

−12−210

,

2−2101

.

Solving the homogeneous system LS(L, 0) resulted in recognizing x4 and x5 as the freevariables. So look in entries 4 and 5 of the two vectors above and notice the pattern ofzeros and ones that provides the linear independence of the set. 4

Page 138: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.EXC Exercises 106

Subsection EXCExercises

Determine if the sets of vectors in Exercises C20–C22 are linearly independent or linearlydependent. C20 Contributed by Robert Beezer 1−21

,

2−13

,

150

Solution [107]

C21 Contributed by Robert Beezer−1242

,

33−13

,

73−64

Solution [107]

C22 Contributed by Robert Beezer1

51

,

6−12

,

9−38

,

28−1

,

3−20

Solution [107]

M50 Contributed by Robert BeezerConsider the set of vectors from C3, W , given below. Find a set T that contains three

vectors from W and such that W = Sp (T ).

W = Sp ({v1, v2, v3, v4, v5}) = Sp

2

11

,

−1−11

,

123

,

313

,

01−3

Solution [107]

T20 Contributed by Robert BeezerSuppose that {v1, v2, v3, v4} is a linearly independent set in C35. Prove that

{v1, v1 + v2, v1 + v2 + v3, v1 + v2 + v3 + v4}

is a linearly independent set. Solution [108]

Page 139: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.SOL Solutions 107

Subsection SOLSolutions

C20 Exercise [106] Contributed by Robert BeezerWith three vectors from C3, we can form a square matrix by making these three vectorsthe columns of a matrix. We do so, and row-reduce to obtain, 1 0 0

0 1 0

0 0 1

the 3 × 3 identity matrix. So by NSME2 t [??]he original matrix is nonsingular and itscolumns are therefore a linearly independent set.

C21 Exercise [106] Contributed by Robert BeezerCorollary LIVRN [??] says we can answer this question by putting theses vectors into amatrix as columns and row-reducing. Doing this we obtain,

1 0 0

0 1 0

0 0 10 0 0

With n = 3 (3 vectors, 3 columns) and r = 3 (3 leading 1’s) we have n = r and thecorollary says the vectors are linearly independent.

C22 Exercise [106] Contributed by Robert BeezerFive vectors from C3. Theorem MVSLD [99] says the set is linearly dependent. Boom.

M50 Exercise [106] Contributed by Robert BeezerWe want to find some relations of linear dependence on {v1, v2, v3, v4, v5} that will

allow us to “kick out” some vectors, in the spirit of Example SS.SCAD [91] and Exam-ple LI.RS [101]. To find relations of linear dependence, we formulate a matrix A whosecolumns are v1, v2, v3, v4, v5. Then we consider the homogeneous sytem of equationsLS(A, 0) by row-reducing its coefficient matrix (remember that if we formulated theaugmented matrix we would just add a column of zeros). After row-reducing, we obtain 1 0 0 2 −1

0 1 0 1 −2

0 0 1 0 0

From this we that solutions can be obtained employing the free variables x4 and x5. Withappropriate choices we will be able to conclude that vectors v4 and v5 are unnecessaryfor creating W via a span. By Theorem SLSLC [76] the choice of free variables belowlead to solutions and linear combinations, which are then rearranged.

x4 = 1, x5 = 0 ⇒ (−2)v1 + (−1)v2 + (0)v3 + (1)v4 + (0)v5 = 0 ⇒ v4 = 2v1 + v2

x4 = 0, x5 = 1 ⇒ (1)v1 + (2)v2 + (0)v3 + (0)v4 + (1)v5 = 0 ⇒ v5 = −v1 − 2v2

Page 140: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LI.SOL Solutions 108

Since v4 and v5 can be expressed as linear combinations of v1 and v2 we can say that v4

and v5 are not needed for the linear combinations used to build W . Thus

W = Sp ({v1, v2, v3}) = Sp

2

11

,

−1−11

,

123

There are other answers to this question, but notice that any nontrvial linear combina-tion of v1, v2, v3, v4, v5 will have a zero coefficient on v3, so this vector can never beeliminated from the set used to build the span.

Though it was not requested in the problem, notice that {v1, v2, v3} is a linearlyindependent set, so we cannot make the set any smaller and still span W .

T20 Exercise [106] Contributed by Robert BeezerOur hypothesis and our conclusion use the term linear independence, so it will get a work-out. To establish linear independence, we begin with the definition (Definition LICV [96])and write a relation of linear dependence (Definition RLDCV [96]),

α1 (v1) + α2 (v1 + v2) + α3 (v1 + v2 + v3) + α4 (v1 + v2 + v3 + v4) = 0

Using the distributive and commutative properties of vector addition and scalar multi-plication (Theorem VSPCM [70]) this equation can be rearranged as

(α1 + α2 + α3 + α4)v1 + (α2 + α3 + α4)v2 + (α3 + α4)v3 + (α4)v4 = 0

However, this is a relation of linear dependence (Definition RLDCV [96]) on a linearlyindependent set, {v1, v2, v3, v4} (this was our lone hypothesis). By the definition oflinear independence (Definition LICV [96]) the scalars must all be zero. This is thehomogeneous system of equations,

α1 + α2 + α3 + α4 = 0

α2 + α3 + α4 = 0

α3 + α4 = 0

α4 = 0

Row-reducing the coefficient matrix of this system (or backsolving) gives the conclusion

α1 = 0 α2 = 0 α3 = 0 α4 = 0

This means, by Definition LICV [96], that the original set

{v1, v1 + v2, v1 + v2 + v3, v1 + v2 + v3 + v4}

is linearly independent.

Page 141: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section O Orthogonality 109

Section O

Orthogonality

In this section we define a couple more operations with vectors, and prove a couple oftheorems. These definitions and results are not central to what follows, but we will makeuse of them frequently throughout the remainder of the course on various ocassions.Because we have chosen to use C as our set of scalars, this subsection is a bit more, uh,. . . complex than it would be for the real numbers. We’ll explain as we go along howthings get easier for the real numbers R. First, we extend the basics of complex numberarithmetic to our study of vectors in Cm.

Subsection CAVComplex arithmetic and vectors

We know how the addition and multiplication of complex numbers is employed in definingthe operations for vectors in Cm (Definition CVA [67] and Definition CVSM [68]). Wecan also extend the idea of the conjugate to vectors.

Definition CCVConjugate of a Column VectorSuppose that

u =

u1

u2

u3...

um

is a vector from Cm. Then the conjugate of the vector is defined as

u =

u1

u2

u3...

um

With this definition we can show that the conjugate of a column vector behaves as wewould expect with regard to vector addition and scalar multiplication.

Theorem CCRVAComplex Conjugation Respects Vector AdditionSuppose x and y are two vectors from Cm. Then

x + y = x + y �

Page 142: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.IP Inner products 110

Proof Apply the definition of vector addition (Theorem CVA [??]) and the definitionof the conjugate of a vector (Definition CCV [109]), and in each component apply thesimilar property for complex numbers (Theorem CCRA [448]). �

Theorem CCRSMComplex Conjugation Respects Scalar MultiplicationSuppose x is a vector from Cm, and α ∈ C is a scalar. Then

αx = αx �

Proof Apply the definition of scalar multiplication (Theorem CVSM [194]) and thedefinition of the conjugate of a vector (Definition CCV [109]), and in each componentapply the similar property for complex numbers (Theorem CCRM [448]). �

These two theorems together tell us how we can “push” complex conjugation throughlinear combinations.

Subsection IPInner products

Definition IPInner ProductGiven the vectors

u =

u1

u2

u3...

um

v =

v1

v2

v3...

vm

the inner product of u and v is the scalar quantity in C,

〈u, v〉 = u1v1 + u2v2 + u3v3 + · · ·+ umvm =m∑

i=1

uivi �

This operation is a bit different in that we begin with two vectors but produce a scalar.Computing one is straightforward.

Example O.CSIPComputing some inner productsThe scalar product of

u =

2 + 3i5 + 2i−3 + i

and v =

1 + 2i−4 + 5i0 + 5i

Page 143: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.IP Inner products 111

is

〈u, v〉 = (2 + 3i)(1 + 2i) + (5 + 2i)(−4 + 5i) + (3 + i)(0 + 5i)

= (2 + 3i)(1− 2i) + (5 + 2i)(−4− 5i) + (3 + i)(0− 5i)

= (8− i) + (−10− 33i) + (5 + 15i)

= 3− 19i

The scalar product of

w =

24−328

and x =

310−1−2

is

〈w, x〉 = 2(3)+4(1)+(−3)(0)+2(−1)+8(−2) = 2(3)+4(1)+(−3)0+2(−1)+8(−2) = −8.4

In the case where the entries of our vectors are all real numbers (as in the second part ofExample O.CSIP [110]), the computation of the inner product may look familiar and beknown to you as a dot product or scalar product. So you can view the inner productas a generalization of the scalar product to vectors from Cm (rather than Rn).

There are several quick theorems we can now prove, and they will each be useful later.

Theorem IPVAInner Product and Vector AdditionSuppose uv,w ∈ Cm. Then

1. 〈u + v, w〉 = 〈u, w〉+ 〈v, w〉2. 〈u, v + w〉 = 〈u, v〉+ 〈u, w〉 �

Proof The proofs of the two parts are very similar, with the second one requiring justa bit more effort due to the conjugation that occurs. We will prove part 2 and you canprove part 1.

〈u, v + w〉 =m∑

i=1

ui(vi + wi) Definition IP [110]

=m∑

i=1

ui(vi + wi) Theorem CCRA [448]

=m∑

i=1

uivi + uiwi Distributivity

=m∑

i=1

uivi +m∑

i=1

uiwi Commutativity

= 〈u, v〉+ 〈u, w〉 Definition IP [110] �

Page 144: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.N Norm 112

Theorem IPSMInner Product and Scalar MultiplicationSuppose u, v ∈ Cm and α ∈ C. Then

1. 〈αu, v〉 = α 〈u, v〉2. 〈u, αv〉 = α 〈u, v〉 �

Proof The proofs of the two parts are very similar, with the second one requiring justa bit more effort due to the conjugation that occurs. We will prove part 2 and you canprove part 1.

〈u, αv〉 =m∑

i=1

ui(αvi) Definition IP [110]

=m∑

i=1

ui(α vi) Theorem CCRM [448]

= αm∑

i=1

uivi Distributivity

= α 〈u, v〉 Definition IP [110] �

Theorem IPACInner Product is Anti-CommutativeSupose that u and v are vectors in Cm. Then 〈u, v〉 = 〈v, u〉. �

Proof

〈u, v〉 = u1v1 + u2v2 + u3v3 + · · ·+ umvm

= u1 v1 + u2 v2 + u3 v3 + · · ·+ um vm

= u1v1 + u2v2 + u3v3 + · · ·+ umvm

= u1v1 + u2v2 + u3v3 + · · ·+ umvm

= v1u1 + v2u2 + v3u3 + · · ·+ vmum

= 〈v, u〉�

Subsection NNorm

If treating linear algebra in a more geometric fashion, the length of a vector occursnaturally, and is what you would expect from its name. With complex numbers, we willdefine a similar function.

Page 145: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.N Norm 113

Definition NVNorm of a VectorThe norm of the vector

u =

u1

u2

u3...

um

is the scalar quantity in Cm

‖u‖ =

√|u1|2 + |u2|2 + |u3|2 + · · ·+ |um|2 =

√√√√ m∑i=1

|ui|2 �

Computing a norm is also easy to do.

Example O.CNSVComputing the norm of some vectorsThe norm of

u =

3 + 2i1− 6i2 + 4i2 + i

is

‖u‖ =

√|3 + 2i|2 + |1− 6i|2 + |2 + 4i|2 + |2 + i|2 =

√13 + 37 + 20 + 5 =

√75 = 5

√3.

The norm of

v =

3−124−3

is

‖v‖ =

√|3|2 + |−1|2 + |2|2 + |4|2 + |−3|2 =

√32 + 12 + 22 + 42 + 32 =

√39. 4

Notice how the norm of a vector with real number entries is just the length of the vector.Inner products and norms are related by the following theorem.

Theorem IPNInner Products and NormsSupose that u is a vector in Cm. Then ‖u‖2 = 〈u, u〉. �

Proof

‖u‖2 =

(√|u1|2 + |u2|2 + |u3|2 + · · ·+ |um|2

)2

= |u1|2 + |u2|2 + |u3|2 + · · ·+ |um|2

= u1u1 + u2u2 + u3u3 + · · ·+ umum

= 〈u, u〉 �

Page 146: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.OV Orthogonal Vectors 114

When our vectors have entries only from the real numbers Theorem IPN [113] says thatthe dot product of a vector with itself is equal to the length of the vector squared.

Theorem PIPPositive Inner ProductsSuppose that u is a vector in Cm. Then 〈u, u〉 ≥ 0 with equality if and only if u = 0.�

Proof From the proof of Theorem IPN [113] we see that

〈u, u〉 = |u1|2 + |u2|2 + |u3|2 + · · ·+ |um|2

Since each modulus squared, every term is positive, and the sum must also be positive.(Notice that in general the inner product is a complex number and cannot be comparedwith zero, but in the special case of 〈u, u〉 the result is a real number.) The phrase,“with equality if and only if” means that we want to show that the statement 〈u, u〉 = 0(i.e. with equality) is equivalent (“if and only if”) to the statement u = 0.

If u = 0, then it is a straightforward computation to see that 〈u, u〉 = 0. In the otherdirection, assume that 〈u, u〉 = 0. As before, 〈u, u〉 is a sum of moduli. So we have

0 = 〈u, u〉 = |u1|2 + |u2|2 + |u3|2 + · · ·+ |um|2

Now we have a sum of squares equaling zero, so each term must be zero. Then by similarlogic, |ui| = 0 will imply that ui = 0, since 0 + 0i is the only complex number with zeromodulus. Thus every entry of u is zero and so u = 0, as desired. �

The conditions of Theorem PIP [114] are summarized by saying “the inner product ispositive definite.”

Subsection OVOrthogonal Vectors

“Orthogonal” is a generalization of “perpendicular.” You may have used mutually per-pendicular vectors in a physics class, or you may recall from a calculus class that perpen-dicular vectors have a zero dot product. We will now extend these ideas into the realmof higher dimensions and complex scalars.

Definition OVOrthogonal VectorsA pair of vectors, u and v, from Cm are orthogonal if their inner product is zero, thatis, 〈u, v〉 = 0. �

Example O.TOVTwo orthogonal vectorsThe vectors

u =

2 + 3i4− 2i1 + i1 + i

v =

1 + i2− 3i4 + 6i

1

Page 147: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.OV Orthogonal Vectors 115

are orthogonal since

(2 + 3i)(1 + i) + (4− 2i)(2− 3i) + (1 + i)(4 + 6i) + (1 + i)(1)

= (−1 + 5i) + (2− 16i) + (−2 + 10i) + (1 + i)

= 0 + 0i. 4

We extend this definition to whole sets by requiring vectors to be pairwise orthogonal.Despite using the same word, careful thought about what objects you are using willeliminate any source of confusion.

Definition OSVOrthogonal Set of VectorsSuppose that S = {u1, u2, u3, . . . , un} is a set of vectors from Cm. Then S is orthogo-nal if every pair of different vectors from S is orthogonal, that is 〈ui, uj〉 = 0 wheneveri 6= j. �

The next example is trivial in some respects, but is still worthy of discussion since it isthe prototypical orthogonal set.

Example O.SUVOSStandard Unit Vectors are an Orthogonal SetThe standard unit vectors are the columns of the identity matrix (Definition SUV [169]).Computing the inner product of two distinct vectors, ei, ej, i 6= j, gives,

〈ei, ej〉 = 00 + 00 + · · ·+ 10 + · · ·+ 01 + · · ·+ 00 + 00

= 0(0) + 0(0) + · · ·+ 1(0) + · · ·+ 0(1) + · · ·+ 0(0) + 0(0)

= 0 4

Example O.AOSAn orthogonal setThe set

{x1, x2, x3, x4} =

1 + i1

1− ii

,

1 + 5i6 + 5i−7− i1− 6i

,

−7 + 34i−8− 23i−10 + 22i30 + 13i

,

−2− 4i6 + i4 + 3i6− i

is an orthogonal set. Since the inner product is anti-commutative (Theorem IPAC [112])we can test pairs of different vectors in any order. If the result is zero, then it will alsobe zero if the inner product is computed in the opposite order. This means there are sixpairs of different vectors to use in an inner product computation. We’ll do two and youcan practice your inner products on the other four.

〈x1, x3〉 = (1 + i)(−7− 34i) + (1)(−8 + 23i) + (1− i)(−10− 22i) + (i)(30− 13i)

= (27− 41i) + (−8 + 23i) + (−32− 12i) + (13 + 30i)

= 0 + 0i

and

〈x2, x4〉 = (1 + 5i)(−2 + 4i) + (6 + 5i)(6− i) + (−7− i)(4− 3i) + (1− 6i)(6 + i)

= (−22− 6i) + (41 + 24i) + (−31 + 17i) + (12− 35i)

= 0 + 0i 4

Page 148: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.GSP Gram-Schmidt Procedure 116

So far, this section has seen lots of definitions, and lots of theorems establishing un-surprising consequences of those definitions. But here is our first theorem that suggeststhat inner products and orthogonal vectors have some utility. It is also one of our firstillustrations of how to arrive at linear independence as the conclusion of a theorem.

Theorem OSLIOrthogonal Sets are Linearly IndependentSuppose that S = {u1, u2, u3, . . . , un} is an orthogonal set of nonzero vectors. Then Sis linearly independent. �

Proof To prove linear independence of a set of vectors, we can appeal to the defi-nition (Definition LICV [96]) and begin with a relation of linear dependence (Defini-tion RLDCV [96]),

α1u1 + α2u2 + α3u3 + · · ·+ αnun = 0.

Then, for every 1 ≤ i ≤ n, we have

0 = 0 〈ui, ui〉= 〈0ui, ui〉 Theorem IPSM [112]

= 〈0, ui〉 Theorem CVSM [194]

= 〈α1u1 + α2u2 + α3u3 + · · ·+ αnun, ui〉 Relation of linear dependence

= 〈α1u1, ui〉+ 〈α2u2, ui〉+ 〈α3u3, ui〉+ · · ·+ 〈αnun, ui〉 Theorem IPVA [111]

= α1 〈u1, ui〉+ α2 〈u2, ui〉+ α3 〈u3, ui〉+ · · ·+ αi 〈ui, ui〉+ · · ·+ αn 〈un, ui〉 Theorem IPSM [112]

= α1(0) + α2(0) + α3(0) + · · ·+ αi 〈ui, ui〉+ · · ·+ αn(0) Orthogonal set

= αi 〈ui, ui〉

So we have 0 = αi 〈ui, ui〉. However, since ui 6= 0 (the hypothesis said our vectorswere nonzero), acronymreftheoremPIP says that 〈ui, ui〉 > 0. So we must conclude thatαi = 0 for all 1 ≤ i ≤ n. But this says that S is a linearly independent set since the onlyway to form a relation of linear dependence is the trivial way, with all the scalars zero.Boom! �

Subsection GSPGram-Schmidt Procedure

TODO: Proof technique on induction.

The Gram-Schmidt Procedure is really a theorem. It says that if we begin with a linearlyindependent set of p vectors, S, then we can do a number of calculations with thesevectors and produce an orthogonal set of p vectors, T , so that Sp (S) = Sp (T ). Giventhe large number of computations involved, it is indeed a procedure to do all the necessarycomputations, and it is best employed on a computer. However, it also has value in proofswhere we may on ocassion wish to replace a linearly independent set by an orthogonalone.

Page 149: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.GSP Gram-Schmidt Procedure 117

Theorem GSPCVGram-Schmidt Procedure, Column VectorsSuppose that S = {v1, v2, v3, . . . , vp} ia a linearly independent set of vectors in Cm.Define the vectors ui, 1 ≤ i ≤ p by

ui = vi −〈vi, u1〉〈u1, u1〉

u1 −〈vi, u2〉〈u2, u2〉

u2 −〈vi, u3〉〈u3, u3〉

u3 − · · · −〈vi, ui−1〉〈ui−1, ui−1〉

ui−1

Then if T = {u1, u2, u3, . . . , up}, then T is an orthogonal set of non-zero vectors, andSp (T ) = Sp (S). �

Proof We will prove the result by using induction on p. To begin, we prove that T hasthe desired properties when p = 1. In this case u1 = v1 and T = {u1} = {v1} = S.Because S and T are equal, Sp (S) = Sp (T ). Equally trivial, T is an orthogonal set. Ifu1 = 0, then S would be a linearly dependent set, a contradiction.

Now suppose that the theorem is true for any set of p − 1 linearly independentvectors. Let S = {v1, v2, v3, . . . , vp} be a linearly independent set of p vectors. ThenS ′ = {v1, v2, v3, . . . , vp−1} is also linearly independent. So we can apply the theorem toS ′ and construct the vectors T ′ = {u1, u2, u3, . . . , up−1}. T ′ is therefore an orthogonalset of nonzero vectors and Sp (S ′) = Sp (T ′). Define

up = vp −〈vp, u1〉〈u1, u1〉

u1 −〈vp, u2〉〈u2, u2〉

u2 −〈vp, u3〉〈u3, u3〉

u3 − · · · −〈vp, up−1〉〈up−1, up−1〉

up−1

and let T = T ′ ∪ {up}. We need to now show that T has several properties by buildingon what we know about T ′. But first notice that the above equation has no problemswith the denominators (〈ui, ui〉) being zero, since the ui are from T ′, which is composedof nonzero vectors.

We show that Sp (T ) = Sp (S), by first establishing that Sp (T ) ⊆ Sp (S). Supposex ∈ Sp (T ), so

x = a1u1 + a2u2 + a3u3 + · · ·+ apup

The term apup is a linear combination of vectors from T ′ and the vector vp, while theremaining terms are a linear combination of vectors from T ′. Since Sp (T ′) = Sp (S ′),any term that is a multiple of a vector from T ′ can be rewritten as a linear combinationsof vectors from S ′. The remaining term apvp is a multiple of a vector in S. So we seethat x can be rewritten as a linear combination of vectors from S, i.e. x ∈ Sp (S).

To show that Sp (S) ⊆ Sp (T ), begin with y ∈ Sp (S), so

y = a1v1 + a2v2 + a3v3 + · · ·+ apvp

Rearrange our defining equation for up by solving for vp. Then the term apvp is a multipleof a linear combination of elements of T . The remaining terms are a linear combination ofv1, v2, v3, . . . , vp−1, hence an element of Sp (S ′) = Sp (T ′). Thus these remaining termscan be written as a linear combination of the vectors in T ′. So y is a linear combinationof vectors from T , i.e. y ∈ Sp (T ).

The elements of T ′ are nonzero, but what about up? Suppose to the contrary that

Page 150: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.GSP Gram-Schmidt Procedure 118

up = 0,

0 = up = vp −〈vp, u1〉〈u1, u1〉

u1 −〈vp, u2〉〈u2, u2〉

u2 −〈vp, u3〉〈u3, u3〉

u3 − · · · −〈vp, up−1〉〈up−1, up−1〉

up−1

vp =〈vp, u1〉〈u1, u1〉

u1 +〈vp, u2〉〈u2, u2〉

u2 +〈vp, u3〉〈u3, u3〉

u3 + · · ·+ 〈vp, up−1〉〈up−1, up−1〉

up−1

Since Sp (S ′) = Sp (T ′) we can write the vectors u1, u2, u3, . . . , up−1 on the right sideof this equation in terms of the vectors v1, v2, v3, . . . , vp−1 and we then have the vectorvp expressed as a linear combination of the other p− 1 vectors in S, implying that S is alinearly dependent set (Theorem DLDS [100]), contrary to our lone hypothesis about S.

Finally, it is a simple matter to establish that T is an orthogonal set, though it will notappear so simple looking. Think about your objects as you work through the following— what is a vector and what is a scalar. Since T ′ is an orthogonal set by induction, mostpairs of elements in T are orthogonal. We just need to test inner products between up

and ui, for 1 ≤ i ≤ p− 1. Here we go, using summation notation,

〈up, ui〉 =

⟨vp −

p−1∑k=1

〈vp, uk〉〈uk, uk〉

uk, ui

= 〈vp, ui〉 −

⟨p−1∑k=1

〈vp, uk〉〈uk, uk〉

uk, ui

⟩Theorem IPVA [111]

= 〈vp, ui〉 −p−1∑k=1

⟨〈vp, uk〉〈uk, uk〉

uk, ui

⟩Theorem IPVA [111]

= 〈vp, ui〉 −p−1∑k=1

〈vp, uk〉〈uk, uk〉

〈uk, ui〉 Theorem IPSM [112]

= 〈vp, ui〉 −〈vp, ui〉〈ui, ui〉

〈ui, ui〉 −∑k 6=i

〈vp, uk〉〈uk, uk〉

(0) T ′ orthogonal

= 〈vp, ui〉 − 〈vp, ui〉 −∑k 6=i

0

= 0 �

Example O.GSTVGram-Schmidt of three vectorsWe will illustrate the Gram-Schmidt process with three vectors. Begin with the linearlyindependent (check this!) set

S = {v1, v2, v3} =

1

1 + i1

,

−i1

1 + i

,

0ii

Page 151: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.GSP Gram-Schmidt Procedure 119

Then

u1 = v1 =

11 + i

1

u2 = v2 −

〈v2, u1〉〈u1, u1〉

u1 =1

4

−2− 3i1− i2 + 5i

u3 = v3 −

〈v3, u1〉〈u1, u1〉

u1 −〈v3, u2〉〈u2, u2〉

u2 =1

11

−3− i1 + 3i−1− i

and

T = {u1, u2, u3} =

1

1 + i1

,1

4

−2− 3i1− i2 + 5i

,1

11

−3− i1 + 3i−1− i

is an orthogonal set (which you can check) of nonzero vectors and Sp (T ) = Sp (S) (allby Theorem GSPCV [117]). Of course, as a by-product of orthogonality, the set T is alsolinearly independent (Theorem OSLI [116]). 4

One final definition related to orthogonal vectors.

Definition ONSOrthoNormal SetSuppose S = {u1, u2, u3, . . . , un} is an orthogonal set of vectors such that ‖ui‖ = 1 forall 1 ≤ i ≤ n. Then S is an orthonormal set of vectors. �

Once you have an orthogonal set, it is easy to convert it to an orthonormal set —multiply each vector by the reciprocal of its norm, and the resulting vector will havenorm 1. This scaling of each vector will not affect the orthogonality properties (applyTheorem IPSM [112]).

Example O.ONTVOrthonormal set, three vectorsThe set

T = {u1, u2, u3} =

1

1 + i1

,1

4

−2− 3i1− i2 + 5i

,1

11

−3− i1 + 3i−1− i

from Example O.GSTV [118] is an orthogonal set. We compute the norm of each vector,

‖u1‖ = 2 ‖u2‖ =1

2

√11 ‖u3‖ =

√2√11

Page 152: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection O.GSP Gram-Schmidt Procedure 120

Converting each vector to a norm of 1, yields an orthonormal set,

w1 =1

2

11 + i

1

w2 =

112

√11

1

4

−2− 3i1− i2 + 5i

=1

2√

11

−2− 3i1− i2 + 5i

w3 =

1√

2√11

1

11

−3− i1 + 3i−1− i

=1√22

−3− i1 + 3i−1− i

4

Example O.ONFVOrthonormal set, four vectorsAs an exercise convert the linearly independent set

S =

1 + i1

1− ii

,

i

1 + i−1−i

,

i−i

−1 + i1

,

−1− i

i1−1

to an orthogonal set via the Gram-Schmidt Process (Theorem GSPCV [117]) and thenscale the vectors to norm 1 to create an orthogonal set. You should get the same set youwould if you scaled the orthogonal set of Example O.AOS [115] to become an orthonormalset. 4

Over the course of the next couple of chapters we will discover that orthonormal setshave some very nice properties (in addition to being linearly independent).

Page 153: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

M: Matrices

Section MO

Matrix Operations

We have made frequent use of matrices for solving systems of equations, and we havebegun to investigate a few of their properties, such as the null space and nonsingularity.In this Chapter, we will take a more systematic approach to the study of matrices, and inthis section we will backup and start simple. We start with the definition of an importantset.

Definition VSMVector Space of m× n MatricesThe vector space Mmn is the set of all m×n matrices with entries from the set of complexnumbers. �

Subsection MEASMMatrix equality, addition, scalar multiplication

Just as we made, and used, a careful definition of equality for column vectors, so too, wehave precise definitions for matrices.

Definition MEMatrix EqualityThe m× n matrices

A = (aij) B = (bij)

are equal, written A = B provided aij = bij for all 1 ≤ i ≤ m, 1 ≤ j ≤ n. �

So equality of matrices translates to the equality of complex numbers, on an entry-by-entry basis. Notice that we now have our fourth definition that uses the symbol ‘=’for shorthand. Whenever a theorem has a conclusion saying two matrices are equal(think about your objects), we will consider appealing to this definition as a way offormulating the top-level structure of the proof. We will now define two operations onthe set Mmn. Again, we will overload a symbol (‘+’) and a convention (juxtaposition formultiplication).

121

Page 154: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MO.MEASM Matrix equality, addition, scalar multiplication 122

Definition MAMatrix AdditionGiven the m× n matrices

A = (aij) B = (bij)

define the sum of A and B to be A + B = C = (cij), where

cij = aij + bij, 1 ≤ i ≤ m, 1 ≤ j ≤ n. �

So matrix addition takes two matrices of the same size and combines them (in a naturalway!) to create a new matrix of the same size. Perhaps this is the “obvious” thing todo, but it doesn’t relieve us from the obligation to state it carefully.

Example MO.MAAddition of two matrices in M23

If

A =

[2 −3 41 0 −7

]B =

[6 2 −43 5 2

]then

A + B =

[2 −3 41 0 −7

]+

[6 2 −43 5 2

]=

[2 + 6 −3 + 2 4 + (−4)1 + 3 0 + 5 −7 + 2

]=

[8 −1 04 5 −5

]4

Our second operation takes two objects of different types, specifically a number and amatrix, and combines them to create another matrix. As with vectors, in this contextwe call a number a scalar in order to emphasize that it is not a matrix.

Definition SMMScalar Matrix MultiplicationGiven the m× n matrix A = (aij) and the scalar α ∈ C, the scalar multiple of A by αis the matrix αA = C = (cij), where

cij = αaij, 1 ≤ i ≤ m, 1 ≤ j ≤ n. �

Notice again that we have yet another kind of multiplication, and it is again writtenputting two symbols side-by-side. Computationally, scalar matrix multiplication is veryeasy.

Example MO.MSMScalar multiplication in M32

If

A =

2 8−3 50 1

and α = 7, then

αA = 7

2 8−3 50 1

=

7(2) 7(8)7(−3) 7(5)7(0) 7(1)

=

14 56−21 350 7

4

Its usually straightforrward to have a calculator do these computations.

Page 155: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MO.VSP Vector Space Properties 123

Subsection VSPVector Space Properties

To refer to matrices abstractly, we have used notation like A = (aij) to connect the nameof a matrix with names for its individual entries. As expressions for matrices becomemore complicated, we will find this notation more cumbersome. So here’s some notationthat will help us along the way.

Notation MEMatrix Entries ([A]ij)For a matrix A, the notation [A]ij will refer to the complex number in row i and columnj of A. 5

As an example, we could rewrite the defining property for matrix addition as

[A + B]ij = [A]ij + [B]ij .

Be careful with this notation, since it is easy to think that [A]ij refers to the whole matrix.It does not. It is just a number, but is a convenient way to talk about all the entries atonce. You might see some of the motivation for this notation in the definition of matrixequality, Definition ME [121].

With definitions of vector addition and scalar multiplication we can state, and prove,several properties of each operation, and some properties that involve their interplay. Wenow collect ten of them here for later reference.

Theorem VSPMVector Space Properties of Mmn

Supppose that A, B and C are m× n matrices in Mmn and α and β are scalars. Then

1. A + B ∈ Mmn (Additive closure)

2. αA ∈ Mmn (Scalar closure)

3. A + B = B + A (Commutativity)

4. A + (B + C) = (A + B) + C (Associativity of matrix addition)

5. There is a matrix, O, called the zero matrix, such that A+O = A for all A ∈ Mmn.(Additive identity)

6. For each matrix A ∈ Mmn, there exists a matrix −A ∈ Mmn so that A+(−A) = O.(Additive inverses)

7. α(βA) = (αβ)A (Associativity of scalar multiplication)

8. α(A + B) = αA + αB (Distributivity across matrix addition)

9. (α + β)A = αA + βA (Distributivity across addition)

10. 1A = A (Scalar multiplication with 1) �

Page 156: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MO.TSM Transposes and Symmetric Matrices 124

Proof While some of these properties seem very obvious, they all require proof. However,the proofs are not very interesting, and border on tedious. We’ll prove one version ofdistributivity very carefully, and you can test your proof-building skills on some of theothers. We’ll give our new notation for matrix entries a workout here. Compare the styleof the proofs here with those given for vectors in Theorem VSPCM [70] — while theobjects here are more complicated, our notation makes the proofs cleaner.

To prove that (α+β)A = αA+βA, we need to establish the equality of two matrices(see Technique GS [13]). Definition ME [121] says we need to establish the equality oftheir entries, one-by-one. How do we do this, when we do not even know how manyentries the two matrices might have? This is where Notation ME [123] comes into play.Ready? Here we go.

For any i and j, 1 ≤ i ≤ m, 1 ≤ j ≤ n,

[(α + β)A]ij = (α + β) [A]ij = α [A]ij + β [A]ij = [αA]ij + [βA]ij = [αA + βA]ij .

A one-liner! There are several things to notice here. (1) Each equals sign is an equalityof numbers. (2) The two ends of the equation, being true for any i and j, allow usto conclude the equality of the matrices. (3) There are several plus signs, and severalinstances of juxtaposition. Identify each one, and state exactly what operation is beingrepresented by each. (4) State the definition or theorem that makes each step of theproof possible. �

For now, note the similarities between Theorem VSPM [123] about matrices and Theo-rem VSPCM [70] about vectors.

The zero matrix described in this theorem, O, is what you would expect — a matrixfull of zeros.

Definition ZMZero MatrixThe m × n zero matrix is written as O = Om×n = (zij) and defined by zij = 0 for all1 ≤ i ≤ m, 1 ≤ j ≤ n. Or, equivalently, [O]ij = 0, for all 1 ≤ i ≤ m, 1 ≤ j ≤ n. �

Subsection TSMTransposes and Symmetric Matrices

We describe one more common operation we can perform on matrices. Informally, totranspose a matrix is to build a new matrix by swapping its rows and columns.

Definition TMTranspose of a MatrixGiven an m× n matrix A, its transpose is the n×m matrix At given by[

At]ij

= [A]ji , 1 ≤ i ≤ n, 1 ≤ j ≤ m. �

Example MO.TMTranspose of a 3× 4 matrix

Page 157: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MO.TSM Transposes and Symmetric Matrices 125

Suppose

D =

3 7 2 −3−1 4 2 80 3 −2 5

.

We could formulate the transpose, entry-by-entry, using the definition. But it is easierto just systematically rewrite rows as columns (or vice-versa). The form of the definitiongiven will be more useful in proofs. So we have

Dt =

3 −1 07 4 32 2 −2−3 8 5

. 4

It will sometimes happen that a matrix is equal to its transpose. In this case, we will calla matrix symmetric. These matrices occur naturally in certain situations, and also havesome nice properties, so it is worth stating the definition carefully. Informally a matrix issymmetric if we can “flip” it about the main diagonal (upper-left corner, running downto the lower-right corner) and have it look unchanged.

Definition SYMSymmetric MatrixThe matrix A is symmetric if A = At. �

Example MO.SYMA symmetric 5× 5 matrixThe matrix

E =

2 3 −9 5 73 1 6 −2 −3−9 6 0 −1 95 −2 −1 4 −87 −3 9 −8 −3

is symmetric. 4

You might have noticed that Definition SYM [125] did not specify the size of the matrixA, as has been our custom. That’s because it wasn’t necessary. An alternative wouldhave been to state the definition just for square matrices, but this is the substance of thenext proof. But first, a bit more advice about constructing proofs.

Proof Technique PPracticeHere is a technique used by many practicing mathematicians when they are teachingthemselves new mathematics. As they read a textbook, monograph or research article,they attempt to prove each new theorem themselves, before reading the proof. Often theproofs can be very difficult, so it is wise not to spend too much time on each. Maybelimit your losses and try each proof for 10 or 15 minutes. Even if the proof is not found,it is time well-spent. You become more familar with the definitions involved, and thehypothesis and conclusion of the theorem. When you do work through the proof, it mightmake more sense, and you will gain added insight about just how to construct a proof.

The next theorem is a great place to try this technique. ♦

Page 158: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MO.MCC Matrices and Complex Conjugation 126

Theorem SMSSymmetric Matrices are SquareSupppose that A is a symmetric matrix. Then A is square. �

Proof We start by specifying A’s size, without assuming it is square, since we are tyringto prove that, so we can’t also assume it. Suppose A is an m × n matrix. Because A issymmetric, we know by Definition SM [254] that A = At. So, in particular, A and At

have the same size. The size of At is n ×m, so from m × n = n ×m, we conclude thatm = n, and hence A must be square. �

We finish this section with another easy theorem, but it illustrates the interplay of ourthree new operations, our new notation, and the techniques used to prove matrix equal-ities.

Theorem TASMTransposes, Addition, Scalar MultiplicationSupppose that A and B are m× n matrices. Then

1. (A + B)t = At + Bt

2. (αA)t = αAt

3. (At)t= A �

Proof Each statement to be proved is an equality of matrices, so we work entry-by-entryand use Definition ME [121]. Think carefully about the objects involved here, the manyuses of the plus sign and juxtaposition, and the justification for each step.[

(A + B)t]ij

= [A + B]ji = [A]ji + [B]ji =[At]ij

+[Bt]ij

=[At + Bt

]ij

and [(αA)t

]ij

= [αA]ji = α [A]ji = α[At]ij

=[αAt

]ij

and [(At)t]

ij=[At]ji

= [A]ij �

Subsection MCCMatrices and Complex Conjugation

As we did with vectors (Definition CCV [109]), we can define what it means to take theconjugate of a matrix.

Definition CCMComplex Conjugate of a MatrixSuppose A is an m×n matrix. Then the conjugate of A, written A is an m×n matrixdefined by [

A]ij

= [A]ij �

TODO: An example, and two theorems (interaction with matrix addition and scalarmultiplication) belong here.

Page 159: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RM Range of a Matrix 127

Section RM

Range of a Matrix

Theorem SLSLC [76] showed us that there is a natural correspondence between solutionsto linear systems and linear combinations of the columns of the coefficient matrix. Thisidea motivates the following important definition.

Definition RMRange of a MatrixSuppose that A is an m × n matrix with columns {A1, A2, A3, . . . , An}. Then the

range of A, written R (A), is the subset of Cm containing all linear combinations of thecolumns of A,

R (A) = Sp ({A1, A2, A3, . . . , An}) �

Subsection RSERange and systems of equations

Upon encountering any new set, the first question we ask is what objects are in the set,and which objects are not? Here’s an example of one way to answer this question, andit will motivate a theorem that will then answer the question precisely.

Example RM.RMCSRange of a matrix and consistent systemsArchetype D [393] and Archetype E [397] are linear systems of equations, with an identical3× 4 coefficient matrix, which we call A here. However, Archetype D [393] is consistent,while Archetype E [397] is not. We can explain this distinction with the range of thematrix A.

The column vector of constants, b, in Archetype D [393] is

b =

8−124

.

One solution to LS(A, b), as listed, is

x =

7813

.

By Theorem SLSLC [76], we can summarize this solution as a linear combination of thecolumns of A that equals b,

7

2−31

+ 8

141

+ 1

7−54

+ 3

−7−6−5

=

8−124

= b.

Page 160: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RSOC Range spanned by original columns 128

This equation says that b is a linear combination of the columns of A, and then byDefinition RM [127], we can say that b ∈ R (A).

On the other hand, Archetype E [397] is the linear system LS(A, c), where the vectorof constants is

c =

232

and this sytem of equations is inconsistent. This means c 6∈ R (A), for if it were, then itwould equal a linear combination of the columns of A and Theorem SLSLC [76] wouldlead us to a solution of the system LS(A, c). 4

So if we fix the coefficient matrix, and vary the vector of constants, we can sometimesfind consistent systems, and sometimes inconsistent systems. The vectors of constantsthat lead to consistent systems are exactly the elements of the range. This is the contentof the next theorem, and since it is an equivalence, it provides an alternate view of therange.

Theorem RCSRange and Consistent SystemsSuppose A is an m× n matrix and b is a vector of size m. Then b ∈ R (A) if and onlyif LS(A, b) is consistent. �

Proof (⇒) Suppose b ∈ R (A). Then we can write b as some linear combination ofthe columns of A. By Theorem SLSLC [76] we can use the scalars from this linearcombination to form a solution to LS(A, b), so this system is consistent.

(⇐) If LS(A, b) is consistent, a solution may be used to write b as a linear combina-tion of the columns of A. This qualifies b for membership in R (A). �

This theorem tells us that asking if the system LS(A, b) is consistent is exactly the samequestion as asking if b is in the range of A. Or equivalently, it tells us that the rangeof the matrix A is precisely those vectors of constants, b, that can be paired with A tocreate a system of linear equations LS(A, b) that is consistent.

Given a vector b and a matrix A it is now very mechanical to test if b ∈ R (A).Form the linear system LS(A, b), row-reduce the augmented matrix, [A | b], and test forconsistency with Theorem RCLS [41].

Subsection RSOCRange spanned by original columns

So we have a foolproof, automated procedure for determining membership in R (A).While this works just fine a vector at a time, we would like to have a more usefuldescription of the set R (A) as a whole. The next example will preview the first of twofundamental results about the range of a matrix.

Example RM.COCCasting out columns, Archetype IArchetype I [414] is a system of linear equations with m = 4 equations in n = 7 variables.

Page 161: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RSOC Range spanned by original columns 129

Let I denote the 4× 7 coefficient matrix from this system, and consider the range of I,R (I). By the definition, we have

R (I) =Sp ({I1, I2, I3, I4, I5, I6, I7})

=Sp

120−1

,

480−4

,

0−122

,

−13−34

,

09−48

,

7−1312−31

,

−97−837

The set of columns of I is obviously linearly dependent, since we have n = 7 vectors fromC4 (see Theorem MVSLD [99]). So we can slim down this set some, and still create therange as the span of a set. The row-reduced form for I is the matrix

1 4 0 0 2 1 −3

0 0 1 0 1 −3 5

0 0 0 1 2 −6 60 0 0 0 0 0 0

so we can easily create solutions to the homogenous system LS(I, 0) using the free vari-ables x2, x5, x6, x7. Any such solution will correspond to a relation of linear dependenceon the columns of I. These will allow us to solve for one column vector as a linear com-bination of some others, in the spirit of Theorem DLDS [100], and remove that vectorfrom the set. We’ll set about this task methodically. Set the free variable x2 to one, andset the other free variables to zero. Then a solution to LS(I, 0) is

x =

−4100000

which can be used to create the linear combination

(−4)I1 + 1I2 + 0I3 + 0I4 + 0I5 + 0I6 + 0I7 = 0

This can then be arranged and solved for I2, resulting in I2 expressed as a linear combi-nation of {I1, I3, I4},

I2 = 4I1 + 0I3 + 0I4

This means that I2 is surplus, and we can create R (I) just as well with a smaller setwith this vector removed,

R (I) = Sp ({I1, I3, I4, I5, I6, I7})

Set the free variable x5 to one, and set the other free variables to zero. Then a solution

Page 162: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RSOC Range spanned by original columns 130

to LS(I, 0) is

x =

−20−1−2100

which can be used to create the linear combination

(−2)I1 + 0I2 + (−1)I3 + (−2)I4 + 1I5 + 0I6 + 0I7 = 0

This can then be arranged and solved for I5, resulting in I5 expressed as a linear combi-nation of {I1, I3, I4},

I5 = 2I1 + 1I3 + 2I4

This means that I5 is surplus, and we can create R (I) just as well with a smaller setwith this vector removed,

R (I) = Sp ({I1, I3, I4, I6, I7})

Do it again, set the free variable x6 to one, and set the other free variables to zero.Then a solution to LS(I, 0) is

x =

−1036010

which can be used to create the linear combination

(−1)I1 + 0I2 + 3I3 + 6I4 + 0I5 + 1I6 + 0I7 = 0

This can then be arranged and solved for I6, resulting in I6 expressed as a linear combi-nation of {I1, I3, I4},

I6 = 1I1 + (−3)I3 + (−6)I4

This means that I6 is surplus, and we can create R (I) just as well with a smaller setwith this vector removed,

R (I) = Sp ({I1, I3, I4, I7})

Set the free variable x7 to one, and set the other free variables to zero. Then a solution

Page 163: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RSOC Range spanned by original columns 131

to LS(I, 0) is

x =

30−5−6001

which can be used to create the linear combination

3I1 + 0I2 + (−5)I3 + (−6)I4 + 0I5 + 0I6 + 1I7 = 0

This can then be arranged and solved for I7, resulting in I7 expressed as a linear combi-nation of {I1, I3, I4},

I7 = (−3)I1 + 5I3 + 6I4

This means that I7 is surplus, and we can create R (I) just as well with a smaller setwith this vector removed,

R (I) = Sp ({I1, I3, I4})

You might think we could keep this up, but we have run out of free variables. Andnot coincidentally, the set {I1, I3, I4} is linearly independent (check this!). Hopefully itis clear how each free variable was used to eliminate the corresponding column from theset used to span the range, for this will be the essence of the proof of the next theorem.See if you can mimic this example using Archetype J [418]. Go ahead, we’ll go grab acup of coffee and be back before you finish up.

For extra credit, notice that the vector

b =

3914

is the vector of constants in the definition of Archetype I [414]. Since the system LS(I, b)is consistent, we know by Theorem RCS [128] that b ∈ R (I). This means that b mustbe a linear combination of just I1, I3, I4. Can you find this linear combination? Did younotice that there is just a single (unique) answer? Hmmmm. 4

We will now formalize the previous example, which will make it trivial to determinea linearly independent set of vectors that will span the range of a matrix. However,the connections made in the last example are worth working through the example (andArchetype J [418]) carefully before employing the theorem.

Theorem BROCBasis of the Range with Original ColumnsSuppose that A is an m × n matrix with columns A1, A2, A3, . . . , An, and B is arow-equivalent matrix in reduced row-echelon form with r nonzero rows. Let D ={d1, d2, d3, . . . , dr} be the set of column indices where B has leading 1’s. Let S ={Ad1 , Ad2 , Ad3 , . . . , Adr}. Then

1. R (A) = Sp (S).

Page 164: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RSOC Range spanned by original columns 132

2. S is a linearly independent set. �

Proof We have two conclusions stemming from the same hypothesis. We’ll prove thefirst conclusion first. By definition

R (A) = Sp ({A1, A2, A3, . . . , An}) .

We will find relations of linear dependence on this set of column vectors, each one involv-ing a column corresponding to a free variable along with the columns corresponding tothe dependent variables. By expressing the free variable column as a linear combinationof the dependent variable columns, we will be able to reduce the set S down to only theset of dependent variable columns while preserving the span.

Let F = {f1, f2, f3, . . . , fn−r} be the sets of column indices where B does not haveany leading 1’s. For each j, 1 ≤ j ≤ n− r construct the specific solution to the homoge-nous system LS(A, 0) given by Theorem VFSLS [82] where the free variables are chosenby the rule

xfi=

{0 if i 6= j

1 if i = j.

This leads to a solution that is exactly the vector uj as defined in Theorem SSNS [90],

uij =

1 if i ∈ F , i = fj

0 if i ∈ F , i 6= fj

−bk,fjif i ∈ D, i = dk

.

Then Theorem SLSLC [76] says this solution corresponds to the following relation oflinear dependence on the columns of A,

(−b1fj)Ad1 + (−b2fj

)Ad2 + (−b3fj)Ad3 + · · ·+ (−brfj

)Adr + (1)Afj= 0.

This can be rearranged as

Afj= b1fj

Ad1 + b2fjAd2 + b3fj

Ad3 + · · ·+ brfjAdr .

This equation can be interpreted to tell us that Afj∈ Sp (S) for all 1 ≤ j ≤ n − r,

so Sp ({A1, A2, A3, . . . , An}) ⊆ Sp (S). It should be easy to convince yourself (so goahead and do it!) that the opposite is true, i.e. Sp (S) ⊆ Sp ({A1, A2, A3, . . . , An}).These two statements about subsets combine (see Technique SE [14]) to give the desiredset equality as our conclusion.

Our second conclusion is that S is a linearly independent set. To prove this, wewill begin with a relation of linear dependence on S. So suppose there are scalarsα1, α2, α3, . . . , αr so that

α1Ad1 + α2Ad2 + α3Ad3 + · · ·+ αrAdr = 0.

To establish linear independence, we wish to deduce that αi = 0, 1 ≤ i ≤ r. This relationof linear dependence allows us to use Theorem SLSLC [76] and construct a solution, x,to the homogenous system LS(A, 0). This vector, x, has αi in entry di, and zeros inevery entry at an index in F . This is equivalent then to a solution, x, where each freevariable equals zero. What would such a solution look like?

Page 165: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RSOC Range spanned by original columns 133

By Theorem VFSLS [82], or from details contained in the proof of Theorem SSNS [90],we see that the only solution to a homogenous system with the free variables all chosento be zero is precisely the trivial solution, so x = 0. Since each αi occurs somewhere asan entry of x = 0, we conclude, as desired, that αi = 0, 1 ≤ i ≤ r, and hence the set Sis linearly independent. �

This is a very pleasing result since it gives us a handful of vectors that describe the entirerange (through the span), and we believe this set is as small as possible because we cannotcreate any more relations of linear dependence to trim it down further. Furthermore, wedefined the range (Definition RM [127]) as all linear combinations of the columns of thematrix, and the elements of the set S are still columns of the matrix (we won’t be solucky in the next two constructions of the range).

Procedurally this theorem is very easy to apply. Row-reduce the original matrix,identify r columns with leading 1’s in this reduced matrix, and grab the correspond-ing columns of the original matrix. Its still important to study the proof of Theo-rem BROC [131] and its motivation in Example RM.COC [128]. We’ll trot through anexample all the same.

Example RM.ROCDRange with original columns, Archetype DLet’s determine a compact expression for the entire range of the coefficient matrix of thesystem of equations that is Archetype D [393]. Notice that in Example RM.RMCS [127]we were only determining if individual vectors were in the range or not.

To start with the application of Theorem BROC [131], call the coefficient matrix A

A =

2 1 7 −7−3 4 −5 −61 1 4 −5

.

and row-reduce it to reduced row-echelon form,

B =

1 0 3 −2

0 1 1 −30 0 0 0

.

There are leading 1’s in columns 1 and 2, so D = {1, 2}. To construct a set that spansR (A), just grab the columns of A indicated by the set D, so

R (A) = Sp

2−31

,

141

.

That’s it.

In Example RM.RMCS [127] we determined that the vector

c =

232

was not in the range of A. Try to write c as a linear combination of the first two columnsof A. What happens?

Page 166: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RNS The range as null space 134

Also in Example RM.RMCS [127] we determined that the vector

b =

8−124

was in the range of A. Try to write b as a linear combination of the first two columns ofA. What happens? Did you find a unique solution to this question? Hmmmm. 4

Subsection RNSThe range as null space

We’ve come to know null spaces quite well, since they are the sets of solutions to homoge-nous systems of equations. In this subsection, we will see how to describe the range ofa matrix as the null space of a different matrix. Then all of our techniques for studyingnull spaces can be brought to bear on ranges. As usual, we will begin with an example,and then generalize to a theorem.

Example RM.RNSADRange as null space, Archetype DBegin again with the coefficient matrix of Archetype D [393],

A =

2 1 7 −7−3 4 −5 −61 1 4 −5

and we will describe another approach to finding the range of A.

Theorem RCS [128] says a vector is in the range only if it can be used as the vector ofconstants for a system of equations with coefficient matrix A and result in a consistentsystem. So suppose we have an arbitrary vector in the range,

b =

b1

b2

b3

∈ R (A) .

Then the linear system LS(A, b) will be consistent by Theorem RCS [128]. Let’s considersolutions to this system by first creating the augmented matrix

[A | b] =

2 1 7 −7 b1

−3 4 −5 −6 b2

1 1 4 −5 b3

.

To locate solutions we would row-reduce this matrix and bring it to reduced row-echelonform. Despite the presence of variables in the last column, there is nothing to stopus from doing this. Except our numerical routines on calculators can’t be used, andeven some of the symbolic algebra routines do some unexpected maneuvers with thiscomputation. So do it by hand. Notice along the way that the row operations are exactlythe same ones you would do if you were just row-reducing the coefficient matrix alone,say in connection with a homogenous system of equations. The column with the bi acts

Page 167: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RNS The range as null space 135

as a sort of bookkeeping device. Here’s what you should get:

[A | b] =

1 0 3 −2 −17b2 + 4

7b3

0 1 1 −3 17b2 + 3

7b3

0 0 0 0 b1 + 17b2 − 11

7b3

.

Is this a consistent system or not? There’s an expression in the last column of the thirdrow, preceded by zeros. Theorem RCLS [41] tells us to look at the leading 1 of the lastnonzero row, and see if it is in the final column. Could the expression at the end ofthe third row be a leading 1 in the last column? The answer is: maybe. It depends onb. Some vectors are in the range, some are not. For b to be in the range, the systemLS(A, b) must be consistent, and the expression in question must not be a leading 1.The only way to prevent it from being a leading 1 is if it is zero, since any nonzero valuecould be scaled to equal 1 by a row operation. So we have

b ∈ R (A) ⇐⇒ LS(A, b) is consistent ⇐⇒ b1 +1

7b2 −

11

7b3 = 0.

So we have an algebraic description of vectors that are, or are not, in the range. Andthis description looks like a single linear homogenous equation in the variables b1, b2, b3.The coefficient matrix of this (simple) homogenous system has the following coefficientmatrix

K =[1 1

7−11

7

].

So we can write that R (A) = N (K)! Example RM.RMCS [127] has a vector b ∈ R (A)and a vector c 6∈ R (A). Test each of these vectors for membership in N (K). The fourcolumns of the matrix A are definitely in the range, are they also in N (K)? (The workabove tells us these answers shouldn’t be surprising, but perhaps doing the computationsmakes it feel a bit remarkable?).

Theorem SSNS [90] tells us how to find a set that spans a null space, and Theo-rem BNS [104] tells us that the same set is linearly indepenent. If we compute this setfor K in this example, we find

R (A) = N (K) = Sp ({u1, u2}) = Sp

−1

7

10

,

117

01

.

Can you write these two new vectors (u1, u2) each as linear combinations of the columnsof A? Uniquely? Can you write each of them as linear combinations of just the first twocolumns of A? Uniquely? Hmmmm.

Doing row operations by hand with variables can be a bit error prone, so lets continuewith this example and see if we can improve on it some. Rather than having b1, b2, b3

all moving around in the same column, lets put each in its own column. So if we insteadrow-reduce 2 1 7 −7 b1 0 0

−3 4 −5 −6 0 b2 01 1 4 −5 0 0 b3

we find 1 0 3 −2 0 −1

7b2

47b3

0 1 1 −3 0 17b2

37b3

0 0 0 0 b117b2 −11

7b3

.

Page 168: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RNS The range as null space 136

If we sum the entries of the third row in columns 4, 5 and 6, and set it equal to zero, weget the equation

b1 +1

7b2 −

11

7b3 = 0

which we recognize as the previous condition for membership in the range. Perhapsyou can see the row operations reflected in the revised form of the matrix involvingthe variables. You might also notice that the variables are acting more and more likeplaceholders (and just getting in the way). Let’s try removing them. One more time.Now row-reduce, using a calculator if you like since there are no symbols, 2 1 7 −7 1 0 0

−3 4 −5 −6 0 1 01 1 4 −5 0 0 1

to obtain 1 0 3 −2 0 −1

747

0 1 1 −3 0 17

37

0 0 0 0 1 17

−117

.

The matrix K from above is now sitting down in the last row, in our “new” columns(5,6,7), and can be read off quicky. Why, we could even program a computer to do it!4

Archetype D [393] is just a tad on the small size for fully motivating the following theorem.In practice, the original matrix may row-reduce to a matrix with several nonzero rows. Ifwe row-reduced an augmented matrix having a variable vector of constants (b), we mightfind several expressions that need to be zero for the system to be consistent. Notice thatit is not enough to make just the last expression zero, as then the one above it would alsohave to be zero, etc. In this way we typically end up with several homogenous equationsprescribing elements of the range, and the coefficient matrix of this system (K) will haveseveral rows.

Here’s a theorem based on the preceding example, which will give us another proce-dure for describing the range of a matrix.

Theorem RNSRange as a Null SpaceSuppose that A is an m × n matrix. Create the m × (n + m) matrix M by placing them×m identity matrix Im to the right of the matrix A. Symbolically, M = [A | Im]. LetN be a matrix that is row-equivalent to M and in reduced row-echelon form. Supposethere are r leading 1’s of N in the first n columns. If r = m, then R (A) = Cm. Otherwise,r < m and let K be the (m − r) ×m matrix formed from the entries of N in the lastm− r rows and last m columns. Then

1. K is in reduced row-echelon form.

2. K has no zero rows, or equivalently, K has m− r leading 1’s.

3. R (A) = N (K). �

Proof Let B denote the m×n matrix that is the first n columns of N , and let J denotethe m×m matrix that is the final m columns of N . Then the sequence of row operationsthat convert M into N , will also convert A into B and Im into J .

Page 169: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RNS The range as null space 137

When r = m, there are m leading 1’s in N that occur in the first n columns, so B hasno zero rows. Thus, the linear system LS(A, b) is never inconsistent, no matter whichvector is chosen for b. So by Theorem RCS [128], every b ∈ Cm is in R (A).

Now consider the case when r < m. The final m− r rows of B are zero rows since theleading 1’s of these rows for N are located in columns n + 1 or higher. The final m− rrows of J form the matrix K.

Since N is in reduced row-echelon form, and the first n entries of each of the finalm− r rows are zero, K will have leading 1’s in an echelon pattern, any zero rows are atthe bottom (but we’ll soon see that there aren’t any), and columns with leading 1’s willbe otherwise zero. In other words, K is in reduced row-echelon form.

Theorem NSRRI [57] tells us that the matrix Im is nonsingular, since it is row-equivalent to the identity matrix (by an empty sequence of row operations!). Therefore,it cannot be row-equivalent to a matrix with a zero row. Why not? A square matrix witha zero row is the coefficient matrix of a homogenous system that has more variables thanequations, if we consider the zero row as a “nonexistent” equation. Theorem HMVEI [51]then says the system has infinitely many solutions. In turn this implies that the homoge-nous linear system LS(Im, 0) has infinitely many solutions, implying that Im is singular,a contradiction. Since K is part of J , and J is row-equivalent to Im, there can be nozero rows in K. If K has no zero rows, then it must have a leading 1 in each of its m− rrows.

For our third conclusion, begin by supposing that b is an arbitrary vector in Cm.To the vector b apply the row operations that convert M to N and call the resultingvector c. Then the linear systems LS(A, b) and LS(B, c) are equivalent. Also, the linearsystems LS(Im, b) and LS(J, c) are equivalent and the unique solution to each is simplyb. Finally, let c∗ be the vector of length m − r containing the final m − r entries of c.Then we have the following,

b ∈ R (A) ⇐⇒ LS(A, b) is consistent Theorem RCS [128]

⇐⇒ LS(B, c) is consistent Definition ES [12]

⇐⇒ c∗ = 0 Theorem RCLS [41]

⇐⇒ b is a solution to LS(K, 0) b is unique solution to LS(J, c)

⇐⇒ b ∈ N (K) . Definition NSM [54]

Running these equivalences in the two different directions will establish the subset inclu-sions needed by Technique SE [14] and so we can conclude that R (A) = N (K). �

We’ve commented that Archetype D [393] was a tad small to fully appreciate this theorem.Let’s apply it now to an Archetype where there’s a bit more action.

Example RM.RNSAGRange as null space, Archetype GArchetype G [406] and Archetype H [410] are both systems of m = 5 equations in n = 2variables. They have identical coefficient matrices, which we will denote here as thematrix G,

G =

2 3−1 43 103 −16 9

.

Page 170: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RNSM Range of a Nonsingular Matrix 138

Adjoin the 5× 5 identity matrix, I5, to form

M =

2 3 1 0 0 0 0−1 4 0 1 0 0 03 10 0 0 1 0 03 −1 0 0 0 1 06 9 0 0 0 0 1

.

This row-reduces to

N =

1 0 0 0 0 3

11133

0 1 0 0 0 − 211

111

0 0 1 0 0 0 −13

0 0 0 1 0 1 −13

0 0 0 0 1 1 −1

.

The first n = 2 columns contain r = 2 leading 1’s, so we extract K from the finalm − r = 3 rows in the final m = 5 columns. Since this matrix is guaranteed to be inreduced row-echelon form, we mark the leading 1’s.

K =

1 0 0 0 −13

0 1 0 1 −13

0 0 1 1 −1

.

Theorem RNS [136] now allows us to conclude that R (G) = N (K). But we can dobetter. Theorem SSNS [90] tells us how to find a set that spans a null space, andTheorem BNS [104] tells us that the same set is linearly indepenent. The matrix K has3 nonzero rows and 5 columns, so the homogenous system LS(K, 0) will have solutionsdescribed by two free variables x4 and x5 in this case. Applying these results in thisexample yields,

R (G) = N (K) = Sp ({u1, u2}) = Sp

0−1−110

,

1313

101

.

As mentioned earlier, Archetype G [406] is consistent, while Archetype H [410] is incon-sistent. See if you can write the two different vectors of constants as linear combinationsof u1 and u2. How about the two columns of G? They must be in the range of G also.Are your answers unique? Do you notice anything about the scalars that appear in thelinear combinations you are forming? 4

Subsection RNSMRange of a Nonsingular Matrix

Let’s specialize to square matrices and contrast the ranges of the coefficient matrices inArchetype A [379] and Archetype B [384].

Page 171: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RNSM Range of a Nonsingular Matrix 139

Example RM.RAARange of Archetype AThe coefficient matrix in Archetype A [379] is

A =

1 −1 22 1 11 1 0

which row-reduces to 1 0 1

0 1 −10 0 0

.

Columns 1 and 2 have leading 1’s, so by Theorem BROC [131] we can write

R (A) = Sp ({A1, A2}) = Sp

1

21

,

−111

.

We want to show in this example that R (A) 6= C3. So take, for example, the vector

b =

132

. Then there is no solution to the system LS(A, b), or equivalently, it is

not possible to write b as a linear combination of A1 and A2. Try one of these twocomputations yourself. (Or try both!). Since b 6∈ R (A), the range of A cannot be all ofC3. So by varying the vector of constants, it is possible to create inconsistent systems ofequations with this coefficient matrix (the vector b being one such example). 4

Example RM.RABRange of Archetype BThe coefficient matrix in Archetype B [384], call it B here, is known to be nonsingular

(see Example NSM.NS [57]). By Theorem NSMUS [59], the linear system LS(B, b) hasa (unique) solution for every choice of b. Theorem RCS [128] then says that b ∈ R (B)for all b ∈ C3. Stated differently, there is no way to build an inconsistent system withthe coefficient matrix B, but then we knew that already from Theorem NSMUS [59]. 4

Theorem RNSMRange of a NonSingular MatrixSuppose A is a square matrix of size n. Then A is nonsingular if and only if R (A) = Cn.�

Proof (⇐) Suppose A is nonsingular. By Theorem NSMUS [59], the linear systemLS(A, b) has a (unique) solution for every choice of b. Theorem RCS [128] then saysthat b ∈ R (A) for all b ∈ Cn. In other words, R (A) = Cn.

(⇒) We’ll prove the contrapositive (see Technique CP [41]). Suppose that A issingular. By Theorem NSRRI [57], A will not row-reduce to the identity matrix In. Sothe row-equivalent matrix B of Theorem RNS [136] has r < n nonzero rows and then thematrix K is a nonzero matrix (it has at least one leading 1 in it). By Theorem NSRRI [57],R (A) = N (K). If we can find one vector of Cn that is not in N (K), then we can concludethat R (A) 6= Cn, the desired conclusion for the contrapositive.

The matrix K has at least one nonzero entry, suppose it is located in column t.Let x ∈ Cn be a vector of all zeros, except a 1 in entry t. Use this vector to forma linear combination of the columns of K, and the result will be just column t of K,

Page 172: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RM.RNSM Range of a Nonsingular Matrix 140

which is nonzero. So by Theorem SLSLC [76], the vector x cannot be a solution to thehomogenous system LS(K, 0), so x 6∈ N (K). �

With this equivalence for nonsingular matrices we can update our list, Theorem NSME2 [103].

Theorem NSME3NonSingular Matrix Equivalences, Round 3Suppose that A is a square matrix of size n. The following are equivalent.

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the zero vector, N (A) = {0}.

4. The linear system LS(A, b) has a unique solution for every possible choice of b.

5. The columns of A are a linearly independent set.

6. The range of A is Cn, R (A) = Cn. �

Proof Since Theorem RNSM [139] is an equivalence, we can add it to the list in Theo-rem NSME2 [103]. �

Page 173: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section RSM Row Space Of a Matrix 141

Section RSM

Row Space Of a Matrix

The range of a matrix is sometimes called the column space since it is formed bytaking all possible linear combinations of the columns of the matrix. We can do a smilarconstruction with the rows of a matrix, and that is the topic of this section. This willprovide us with even more connections with row operations. However, we are also goingto try to parlay our knowledge of the range, so we’ll get at the rows of a matrix byworking with the columns of the transpose. A side-benefit will be a third way to describethe range of a matrix.

Subsection RSMRow Space of a Matrix

Definition RSMRow Space of a MatrixSuppose A is an m×n matrix. Then the row space of A, rs (A), is the range of At, i.e.rs (A) = R (At). �

Informally, the row space is the set of all linear combinations of the rows of A. However,we write the rows as column vectors, thus the necessity of using the transpose to makethe rows into columns. With the row space defined in terms of the range, all of the resultsof Section RM [127] can be applied to row spaces.

Notice that if A is a rectangular m× n matrix, then R (A) ⊆ Cm, while rs (A) ⊆ Cn

and the two sets are not comparable since they do not even hold objects of the sametype. However, when A is square of size n, both R (A) and rs (A) are subsets of Cn,though usually the sets will not be equal.

Example RSM.RSAIRow space of Archetype IThe coefficient matrix in Archetype I [414] is

I =

1 4 0 −1 0 7 −92 8 −1 3 9 −13 70 0 2 −3 −4 12 −8−1 −4 2 4 8 −31 37

.

To build the row space, we transpose the matrix,

I t =

1 2 0 −14 8 0 −40 −1 2 2−1 3 −3 40 9 −4 87 −13 12 −31−9 7 −8 37

Page 174: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.RSM Row Space of a Matrix 142

Then the columns of this matrix are used in a span to build the row space,

rs (I) = R(I t)

= Sp

140−107−9

,

28−139−137

,

002−3−412−8

,

−1−4248−3137

.

However, we can use Theorem BROC [131] to get a slightly better description. First,row-reduce I t,

1 0 0 −317

0 1 0 127

0 0 1 137

0 0 0 00 0 0 00 0 0 00 0 0 0

.

Since there are leading 1’s in columns with indices D = {1, 2, 3}, the range of I t can bespanned by just the first three columns of I t,

rs (I) = R(I t)

= Sp

140−107−9

,

28−139−137

,

002−3−412−8

.

The technique of Theorem RNS [136] could also have been applied to the matrix I t, byadjoining the 7×7 identity matrix, I7 and row-reducing the resulting 11×7 matrix. The4× 7 matrix K that results is

K =

1 0 0 0 −3

7−1

70

0 1 0 0 −127

−47

00 0 1 0 −3

7− 9

14−1

2

0 0 0 1 17

1328

14

.

Then rs (I) = R (I t) = N (K), and we could use Theorem SSNS [90] and Theorem BNS [104]to express the null space of K as the span of three vectors, one for each free variable inthe homogenous system LS(K, 0). 4

The row space would not be too interesting if it was simply the range of the transpose.However, when we do row operations on a matrix we have no effect on the many linearcombinations that can be formed with the rows of the matrix. This is stated morecarefully in the following theorem.

Theorem REMRSRow-Equivalent Matrices have equal Row SpacesSuppose A and B are row-equivalent matrices. Then rs (A) = rs (B). �

Page 175: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.RSM Row Space of a Matrix 143

Proof Two matrices are row-equivalent (Definition REM [25]) if one can be obtainedfrom another by a sequence of (possibly many) row operations. We will prove the theoremfor two matrices that differ by a single row operation, and then this result can be appliedrepeatedly to get the full statement of the theorem. The row spaces of A and B are spansof the columns of their transposes. For each row operation we perform on a matrix, wecan define an analagous operation on the columns. Perhaps we should call these columnoperations. Instead, we will still call them row operations, but we will apply them tothe columns of the transposes.

Refer to the columns of At and Bt as Ai and Bi, 1 ≤ i ≤ m. The row operationthat switches rows will just switch columns of the transposed matrices. This will haveno effect on the possible linear combinations formed by the columns.

Suppose that Bt is formed from At by multiplying column At by α 6= 0. In otherwords, Bt = αAt, and Bi = Ai for all i 6= t. We need to establish that two sets are equal,R (At) = R (Bt). We will take a generic element of one and show that it is contained inthe other.

β1B1 + β2B2 + β3B3 + · · ·+ βtBt + · · ·+ βmBm =

β1A1 + β2A2 + β3A3 + · · ·+ βt (αAt) + · · ·+ βmAm =

β1A1 + β2A2 + β3A3 + · · ·+ (αβt)At + · · ·+ βmAm

says that R (Bt) ⊆ R (At). Similarly,

γ1A1 + γ2A2 + γ3A3 + · · ·+ γtAt + · · ·+ γmAm =

γ1A1 + γ2A2 + γ3A3 + · · ·+(γt

αα)

At + · · ·+ γmAm =

γ1A1 + γ2A2 + γ3A3 + · · ·+ γt

α(αAt) + · · ·+ γmAm =

γ1B1 + γ2B2 + γ3B3 + · · ·+ γt

αBt + · · ·+ γmBm

says that R (At) ⊆ R (Bt). So rs (A) = R (At) = R (Bt) = rs (B) when a single rowoperation of the second type is performed.

Suppose now that Bt is formed from At by replacing At with αAs + At for someα ∈ C and s 6= t. In other words, Bt = αAs + At, and Bi = Ai for i 6= t.

β1B1 + β2B2 + β3B3 + · · ·+ βsBs + · · ·+ βtBt + · · ·+ βmBm =

β1A1 + β2A2 + β3A3 + · · ·+ βsAs + · · ·+ βt (αAs + At) + · · ·+ βmAm =

β1A1 + β2A2 + β3A3 + · · ·+ βsAs + · · ·+ (βtα)As + βtAt + · · ·+ βmAm =

β1A1 + β2A2 + β3A3 + · · ·+ βsAs + (βtα)As + · · ·+ βtAt + · · ·+ βmAm =

β1A1 + β2A2 + β3A3 + · · ·+ (βs + βtα)As + · · ·+ βtAt + · · ·+ βmAm

says that R (Bt) ⊆ R (At). Similarly,

γ1A1 + γ2A2 + γ3A3 + · · ·+ γsAs + · · ·+ γtAt + · · ·+ γmAm =

γ1A1 + γ2A2 + γ3A3 + · · ·+ γsAs + · · ·+ (−αγtAs + αγtAs) + γtAt + · · ·+ γmAm =

γ1A1 + γ2A2 + γ3A3 + · · ·+ (−αγtAs) + γsAs + · · ·+ (αγtAs + γtAt) + · · ·+ γmAm =

γ1A1 + γ2A2 + γ3A3 + · · ·+ (−αγt + γs)As + · · ·+ γt (αAs + At) + · · ·+ γmAm =

γ1B1 + γ2B2 + γ3B3 + · · ·+ (−αγt + γs)Bs + · · ·+ γtBt + · · ·+ γmBm

Page 176: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.RSM Row Space of a Matrix 144

says that R (At) ⊆ R (Bt). So rs (A) = R (At) = R (Bt) = rs (B) when a single rowoperation of the third type is performed.

So the row space is preserved by each row operation, and hence row spaces of row-equivalent matrices are equal. �

Example RSM.RSREMRow spaces of two row-equivalent matricesIn Example RREF.TREM [25] we saw that the matrices

A =

2 −1 3 45 2 −2 31 1 0 6

B =

1 1 0 63 0 −2 −92 −1 3 4

are row-equivalent by demonstrating a sequence of two row operations that converted Ainto B. Applying Theorem REMRS [142] we can say

rs (A) = Sp

2−134

,

52−23

,

1106

= Sp

1106

,

30−2−9

,

2−134

= rs (B) 4

Theorem REMRS [142] is at its best when one of the row-equivalent matrices is inreduced row-echelon form. The vectors that correspond to the zero rows can be ignored(who needs the zero vector when building a span?). The echelon pattern insures that thenonzero rows yield vectors that are linearly independent. Here’s the theorem.

Theorem BRSBasis for the Row SpaceSuppose that A is a matrix and B is a row-equivalent matrix in reduced row-echelonform. Let S be the set of nonzero columns of Bt. Then

1. rs (A) = Sp (S).

2. S is a linearly independent set. �

Proof From Theorem REMRS [142] we know that rs (A) = rs (B). If B has any zerorows, these correspond to columns of Bt that are the zero vector. We can safely tossout the zero vector in the span construction, since it can be recreated from the nonzerovectors by a linear combination where all the scalars are zero.

Suppose B has r nonzero rows and let D = {d1, d2, d3, . . . , dr} denote the columnindices of B that have a leading one in them. Denote the r column vectors of Bt, thevectors in S, as B1, B2, B3, . . . , Br. To show that S is linearly independent, start witha relation of linear dependence

α1B1 + α2B2 + α3B3 + · · ·+ αrBr = 0

Now consider this equation across entries of the vectors in location di, 1 ≤ i ≤ r. Since Bis in reduced row-echelon form, the entries of column di are all zero, except for a (leading)1 in row i. Considering the column vectors of Bt, the linear combination for entry di is

α1(0) + α2(0) + α3(0) + · · ·+ αi(1) + · · ·+ αr(0) = 0

and from this we conclude that αi = 0 for all 1 ≤ i ≤ r, establishing the linear indepen-dence of S. �

Page 177: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.RSM Row Space of a Matrix 145

Example RSM.ISImproving a spanSuppose in the course of analyzing a matrix (its range, its null space, its. . . ) we encounterthe following set of vectors, described by a span

X = Sp

12166

,

3−12−16

,

1−10−1−2

,

−32−36−10

Let A be the matrix whose rows are the vectors in X, so by design rs (A) = X,

A =

1 2 1 6 63 −1 2 −1 61 −1 0 −1 −2−3 2 −3 6 −10

Row-reduce A to form a row-equivalent matrix in reduced row-echelon form,

B =

1 0 0 2 −1

0 1 0 3 1

0 0 1 −2 50 0 0 0 0

Then Theorem BRS [144] says we can grab the nonzero columns of Bt and write

X = rs (A) = rs (B) = Sp

1002−1

,

01031

,

001−25

These three vectors provide a much-improved description of X. There are fewer vectors,and the pattern of zeros and ones in the first three entries makes it easier to determinemembership in X. And all we had to do was row-reduce the right matrix and tossout a zero row. Next to row operations themselves, this is probably the most powerfulcomputational technique at your disposal. 4

Theorem BRS [144] and the techniques of Example RSM.IS [145] will provide yet anotherdescription of the range of a matrix. First we state a triviality as a theorem, so we canreference it later

Theorem RMRSTRange of a Matrix is Row Space of TransposeSuppose A is a matrix. Then R (A) = rs (At). �

Proof Apply Theorem TASM [126] with Definition RSM [141],

rs(At)

= R((

At)t)

= R (A) . �

So to find yet another expression for the range of a matrix, build its transpose, row-reduceit, toss out the zero rows, and convert the nonzero rows to column vectors to yield animproved spanning set. We’ll do Archetype I [414], then you do Archetype J [418].

Page 178: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.RSM Row Space of a Matrix 146

Example RSM.RROIRange from row operations, Archetype ITo find the range of the coefficient matrix of Archetype I [414], we proceed as follows.

The matrix is

I =

1 4 0 −1 0 7 −92 8 −1 3 9 −13 70 0 2 −3 −4 12 −8−1 −4 2 4 8 −31 37

.

The transpose is

1 2 0 −14 8 0 −40 −1 2 2−1 3 −3 40 9 −4 87 −13 12 −31−9 7 −8 37

.

Row-reduced this becomes,

1 0 0 −317

0 1 0 127

0 0 1 137

0 0 0 00 0 0 00 0 0 00 0 0 0

.

Now, using Theorem RMRST [145] and Theorem BRS [144]

R (I) = rs(I t)

= Sp

100−31

7

,

010127

,

001137

.

This is a very nice description of the range. Fewer vectors than the 7 involved in thedefinition, and the structure of the zeros and ones in the first 3 slots can be used toadvantage. For example, Archetype I [414] is presented as consistent system of equationswith a vector of constants

b =

3914

.

Since LS(I, b) is consistent, Theorem RCS [128] tells us that b ∈ R (I). But we couldsee this quickly with the following computation, which really only involves any work inthe 4th entry of the vectors as the scalars in the linear combination are dictated by thefirst three entries of b.

b =

3914

= 3

100−31

7

+ 9

010127

+ 1

001137

Page 179: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.RSM Row Space of a Matrix 147

Can you now rapidly construct several vectors, b, so that LS(I, b) is consistent, andseveral more so that the system is inconsistent? 4

Example RM.COC [128] and Example RSM.RROI [146] each describes the range of thecoefficient matrix from Archetype I [414] as the span of a set of r = 3 linearly independentvectors. It is no accident that these two different sets both have the same size. If we(you?) were to calculate the range of this matrix using the null space of the matrix Kfrom Theorem RNS [136] then we would again find a set of 3 linearly independent vectorsthat span the range. More on this later.

So we have three different methods to obtain a description of the range of a matrix asthe span of a linearly independent set. Theorem BROC [131] is sometimes useful sincethe vectors it specifies are equal to actual columns of the matrix. Theorem RNS [136]tends to create vectors with lots of zeros, and strategically placed 1’s near the bottom ofthe vector. Finally, Theorem BRS [144] and Theorem RMRST [145] combine to createvectors with lots of zeros, and strategically placed 1’s near the top of the vector.

Page 180: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.EXC Exercises 148

Subsection EXCExercises

C30 Contributed by Robert BeezerLet A be the matrix below, and find the indicated sets by the requested methods.

A =

2 −1 5 −3−5 3 −12 71 1 4 −3

(a) A linearly independent set S so that R (A) = Sp (S) and S is composed of columnsof A.(b) A linearly independent set S so that R (A) = Sp (S) and the vectors in S have anice pattern of zeros and ones at the top of the vectors.(c) A linearly independent set S so that R (A) = Sp (S) and the vectors in S have anice pattern of zeros and ones at the bottom of the vectors.(d) A linearly independent set S so that rs (A) = Sp (S). Solution [149]

TODO: Construct an example of a matrix A where R (A) = rs (A).

Page 181: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.SOL Solutions 149

Subsection SOLSolutions

C30 Exercise [148] Contributed by Robert Beezer(a) First find a matrix B that is row-equivalent to A and in reduced row-echelon form

B =

1 0 3 −2

0 1 1 −10 0 0 0

By Theorem BROC [131] we can choose the columns of A that correspond to dependentvariables (D = {1, 2}) as the elements of S and obtain the desired properties. So

S =

2−51

,

−131

(b) We can write the range of A as the row space of the transpose. So we row-reducethe transpose of A to obtain the row-equivalent matrix C in reduced row-echelon form

C =

1 0 80 1 30 0 00 0 0

The nonzero rows (written as columns) will be a linearly independent set that spans therow space of At, by Theorem BRS [144], ans the zeros and ones will be at the top of thevectors,

S =

1

08

,

013

(c) In preparation for Theorem RNS [136], augment A with the 3 × 3 identity matrixI3 and row-reduce to obtain, 1 0 3 −2 0 −1

838

0 1 1 −1 0 18

58

0 0 0 0 1 38

−18

Then since the first four columns of row 3 are all zeros, we extract

K =[

1 38−1

8

]Theorem RNS [136] says that R (A) = N (K). We can then use Theorem SSNS [90]and Theorem BNS [104] to construct the desired set S, based on the free variables withindices in F = {2, 3} for the homogeneous system LS(K, 0), so

S =

−3

8

10

,

18

01

Page 182: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection RSM.SOL Solutions 150

Notice that the zeros and ones are at the bottom of the vectors. (d) This is a straight-forward application of Theorem BRS [144]. Use the row-reduced matrix B from part (a),grab the nonzero rows, and write them as column vectors,

S =

103−2

,

011−1

Page 183: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section MM Matrix Multiplication 151

Section MM

Matrix Multiplication

We know how to add vectors and how to multiply them by scalars. Together, theseoperations give us the possibility of making linear combinations. Similarly, we know howto add matrices and how to multiply matrices by scalars. In this section we mix all theseideas together and produce an operation known as matrix multiplication. This will leadto some results that are both surprising and central. We begin with a definition of howto multiply a vector by a matrix.

Subsection MVPMatrix-Vector Product

We have repeatedly seen the importance of forming linear combinations of the columnsof a matrix. As one example of this, Theorem SLSLC [76] said that every solution toa system of linear equations gives rise to a linear combination of the column vectors ofthe coefficient matrix that equals the vector of constants. This theorem, and others,motivates the following central definition.

Definition MVPMatrix-Vector ProductSuppose A is an m × n matrix with columns A1, A2, A3, . . . , An and u is a vector of

size n. Then the matrix-vector product of A with u is

Au = [A1|A2|A3| . . . |An]

u1

u2

u3...

un

= u1A1 + u2A2 + u3A3 + · · ·+ unAn �

So, the matrix-vector product is yet another version of “multiplication,” at least in thesense that we have yet again overloaded juxtaposition of two symbols. Remember yourobjects, an m× n matrix times a vector of size n will create a vector of size m. So if Ais rectangular, then the size of the vector changes. With all the linear combinations wehave performed so far, this computation should now seem second nature.

Example MM.MTVA matrix times a vectorConsider

A =

1 4 2 3 4−3 2 0 1 −21 6 −3 −1 5

u =

21−23−1

Page 184: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.MVP Matrix-Vector Product 152

Then

Au = 2

1−31

+ 1

426

+ (−2)

20−3

+ 3

31−1

+ (−1)

4−25

=

716

. 4

This definition now makes it possible to represent systems of linear equations compactlyin terms of an operation.

Theorem SLEMMSystems of Linear Equations as Matrix MultiplicationSolutions to the linear system LS(A, b) are the solutions for x in the vector equationAx = b. �

Proof This theorem says (not very clearly) that two sets (of solutions) are equal. Sowe need to show that one set of solutions is a subset of the other, and vice versa (re-call Technique SE [14]). Both of these inclusions are easy with the following chain ofequivalences,

x =

x1

x2

x3...xi

is a solution to LS(A, b)

⇐⇒ x1A1 + x2A2 + x3A3 + · · ·+ xnAn = b Theorem SLSLC [76]

⇐⇒ x =

x1

x2

x3...xi

is a solution to Ax = b Definition MVP [151].

Example MM.NSLENotation for systems of linear equationsConsider the system of linear equations from Example HSE.NSLE [54].

2x1 + 4x2 − 3x3 + 5x4 + x5 = 9

3x1 + x2 + +x4 − 3x5 = 0

−2x1 + 7x2 − 5x3 + 2x4 + 2x5 = −3

has coefficient matrix

A =

2 4 −3 5 13 1 0 1 −3−2 7 −5 2 2

and vector of constants

b =

90−3

and so will be described compactly by the equation Ax = b. 4

Page 185: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.MM Matrix Multiplication 153

Subsection MMMatrix Multiplication

We now define how to multiply two matrices together. Stop for a minute and think abouthow you might define this new operation. Many books would present this definition muchearlier in the course. However, we have taken great care to delay it as long as possible andto present as many ideas as practical based mostly on the notion of linear combinations.Towards the conclusion of the course, or when you perhaps take a second course in linearalgebra, you may be in a position to appreciate the reasons for this. For now, understandthat matrix multiplication is a central definition and perhaps you will appreciate itsimportance more by having saved it for later.

Definition MMMatrix MultiplicationSuppose A is an m×n matrix and B is an n×p matrix with columns B1, B2, B3, . . . , Bp.Then the matrix product of A with B is the m×p matrix where column i is the matrix-vector product ABi. Symbolically,

AB = A [B1|B2|B3| . . . |Bp] = [AB1|AB2|AB3| . . . |ABp] . �

Example MM.PTMProduct of two matricesSet

A =

1 2 −1 4 60 −4 1 2 3−5 1 2 −3 4

B =

1 6 2 1−1 4 3 21 1 2 36 4 −1 21 −2 3 0

Then

AB =

A

1−1161

∣∣∣∣∣∣∣∣∣∣A

6414−2

∣∣∣∣∣∣∣∣∣∣A

232−13

∣∣∣∣∣∣∣∣∣∣A

12320

=

28 17 20 1020 −13 −3 −1−18 −44 12 −3

. 4

Is this the definition of matrix multiplication you expected? Perhaps our previous op-erations for matrices caused you to think that we might multiply two matrices of thesame size, entry-by-entry? Notice that our current definition uses matrices of differentsizes (though the number of columns in the first must equal the number of rows in thesecond), and the result is of a third size. Notice too in the previous example that wecannot even consider the product BA, since the sizes of the two matrices in this orderaren’t right.

But it gets weirder than that. Many of your old ideas about “multiplication” won’tapply to matrix multiplication, but some still will. So make no assumptions, and don’tdo anything until you have a theorem that says you can. Even if the sizes are right,matrix multiplication is not commutative — order matters.

Page 186: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.MMEE Matrix Multiplication, Entry-by-Entry 154

Example MM.MMNCMatrix Multiplication is not commutativeSet

A =

[1 3−1 2

]B =

[4 05 1

].

Then we have two square, 2 × 2 matrices, so Definition MM [153] allows us to multiplythem in either order. We find

AB =

[19 36 2

]BA =

[4 124 17

]and AB 6= BA. Not even close. It should not be hard for you to construct other pairsof matrices that do not commute (try a couple of 3 × 3’s). Can you find a pair ofnon-identical matrices that do commute? 4

Computation Note MM.MMAMatrix Multiplication (Mathematica)If A and B are matrices defined in Mathematica, then A.B will return the product

of the two matrices (notice the dot between the matrices). If A is a matrix and v is avector, then A.v will return the vector that is the matrix-vector product of A and v.In every case the sizes of the matrices and vectors need to be correct.

Some examples:

{{1, 2}, {3, 4}}.{{5, 6, 7}, {8, 9, 10}} = {{21, 24, 27}, {47, 54, 61}}{{1, 2}, {3, 4}}.{{5}, {6}} = {{17}, {39}}

{{1, 2}, {3, 4}}.{5, 6} = {17, 39}

Understanding the difference between the last two examples will go a long way to ex-plaining how some Mathematica constructs work. ⊗

Subsection MMEEMatrix Multiplication, Entry-by-Entry

While certain “natural” properties of multiplication don’t hold, many more do. In thenext subsection, we’ll state and prove the relevant theorems. But first, we need a theoremthat provides an alternate means of multiplying two matrices. In many texts, this wouldbe given as the definition of matrix multiplication. We prefer to turn it around and havethe following formula as a consequence of the definition. It will prove useful for proofs ofmatrix equality, where we need to examine products of matrices, entry-by-entry.

Theorem EMPEntries of Matrix ProductsSuppose A = (aij) is an m×n matrix and B = (bij) is an n× p matrix. Then the entriesof AB = C = (cij) are given by

[C]ij = cij = ai1b1j + ai2b2j + ai3b3j + · · ·+ ainbnj =n∑

k=1

aikbkj =n∑

k=1

[A]ik [B]kj �

Page 187: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.MMEE Matrix Multiplication, Entry-by-Entry 155

Proof The value of cij lies in column j of the product of A and B, and so by Defini-tion MM [153] is the value in location i of the matrix-vector product ABj. By Defini-tion MVP [151] this matrix-vector product is a linear combination

ABj = b1jA1 + b2jA2 + b3jA3 + · · ·+ bnjAn

= b1j

a11

a21

a31...

am1

+ b2j

a12

a22

a32...

am2

+ b3j

a13

a23

a33...

am3

+ · · ·+ bnj

a1n

a2n

a3n...

amn

We are after the value in location i of this linear combination. Using Definition CVA [67]and Definition CVSM [68] we course through this linear combination in location i to find

b1jai1 + b2jai2 + b3jai3 + · · ·+ bnjain.

Reversing the order of the products (regular old multiplication is commutative) yieldsthe desired expression for cij. �

Example MM.PTMEEProduct of two matrices, entry-by-entryConsider again the two matrices from Example MM.PTM [153]

A =

1 2 −1 4 60 −4 1 2 3−5 1 2 −3 4

B =

1 6 2 1−1 4 3 21 1 2 36 4 −1 21 −2 3 0

Then suppose we just wanted the entry of AB in the second row, third column:

[AB]23 =a21b13 + a22b23 + a23b33 + a24b43 + a25b53

=(0)(2) + (−4)(3) + (1)(2) + (2)(−1) + (3)(3) = −3

Notice how there are 5 terms in the sum, since 5 is the common dimension of the twomatrices (column count for A, row count for B). In the conclusion of Theorem EMP [154],it would be the index k that would run from 1 to 5 in this computation. Here’s a bitmore practice.

The entry of third row, first column:

[AB]31 =a31b11 + a32b21 + a33b31 + a34b41 + a35b51

=(−5)(1) + (1)(−1) + (2)(1) + (−3)(6) + (4)(1) = −18

To get some more practice on your own, complete the computation of the other 10 entriesof this product. Construct some other pairs of matrices (of compatible sizes) and computetheir product two ways. First use Definition MM [153]. Since linear combinations arestraightforward for you now, this should be easy to do and to do correctly. Then do itagain, using Theorem EMP [154]. Since this process may take some practice, use yourfirst computation to check your work. 4

Page 188: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PMM Properties of Matrix Multiplication 156

Theorem EMP [154] is the way most people compute matrix products by hand. It willalso be very useful for the theorems we are going to prove shortly. However, the definitionis frequently the most useful for its connections with deeper ideas like the nullspace andrange. For example, an alternative (and popular) definition of the range of an m × nmatrix A would be

R (A) = {Ax | x ∈ Cn} .

We recognize this as saying take all the matrix vector products possible with the ma-trix A. By Definition MVP [151] we see that this means take all possible linear com-binations of the columns of A — precisely our version of the definition of the range(Definition RM [127]).

Subsection PMMProperties of Matrix Multiplication

In this subsection, we collect properties of matrix multiplication and its interaction withmatrix addition (Definition MA [122]), scalar matrix multiplication (Definition SMM [122]),the identity matrix (Definition IM [57]), the zero matrix (Definition ZM [124]), conjuga-tion (Theorem CM [??]) and the transpose (Definition TM [124]). Whew! Here we go.These are great proofs to practice with, so try to concoct the proofs before reading them,they’ll get progressively harder as we go.

Theorem MMZMMatrix Multiplication and the Zero MatrixSuppose A is an m× n matrix. Then1. AOn×p = Om×p

2. Op×mA = Op×n �

Proof We’ll prove (1) and leave (2) to you. Entry-by-entry,

[AOn×p]ij =n∑

k=1

[A]ik [On×p]kj Theorem EMP [154]

=n∑

k=1

[A]ik 0 Definition ZM [124]

=n∑

k=1

0 = 0.

So every entry of the product is the scalar zero, i.e. the result is the zero matrix. �

Theorem MMIMMatrix Multiplication and Identity MatrixSuppose A is an m× n matrix. Then1. AIn = A2. ImA = A �

Page 189: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PMM Properties of Matrix Multiplication 157

Proof Again, we’ll prove (1) and leave (2) to you. Entry-by-entry,

[AIn]ij =n∑

k=1

[A]ik [In]kj Theorem EMP [154]

= [A]ij [In]jj +n∑

k=1,k 6=j

[A]ik [In]kj

= [A]ij (1) +n∑

k=1,k 6=j

[A]ik (0) Definition IM [57]

= [A]ij +n∑

k=1,k 6=j

0

= [A]ij

So the matrices A and AIn are equal, entry-by-entry, and by the definition of matrixequality (Definition ME [121]) we can say they are equal matrices. �

It is this theorem that gives the identity matrix its name. It is a matrix that behaveswith matrix multiplication like the scalar 1 does with scalar multiplication. To multiplyby the identity matrix is to have no effect on the other matrix.

Theorem MMDAAMatrix Multiplication Distributes Across AdditionSuppose A is an m× n matrix and B and C are n× p matrices and D is a p× s matrix.Then1. A(B + C) = AB + AC2. (B + C)D = BD + CD �

Proof We’ll do (1), you do (2). Entry-by-entry,

[A(B + C)]ij =n∑

k=1

[A]ik [B + C]kj Theorem EMP [154]

=n∑

k=1

[A]ik ([B]kj + [C]kj) Definition MA [122]

=n∑

k=1

[A]ik [B]kj + [A]ik [C]kj

=n∑

k=1

[A]ik [B]kj +n∑

k=1

[A]ik [C]kj

= [AB]ij + [AC]ij Theorem EMP [154]

= [AB + AC]ij Definition MA [122]

So the matrices A(B + C) and AB + AC are equal, entry-by-entry, and by the definitionof matrix equality (Definition ME [121]) we can say they are equal matrices. �

Theorem MMSMMMatrix Multiplication and Scalar Matrix MultiplicationSuppose A is an m × n matrix and B is an n × p matrix. Let α be a scalar. Thenα(AB) = (αA)B = A(αB). �

Page 190: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PMM Properties of Matrix Multiplication 158

Proof These are equalities of matrices. We’ll do the first one, the second is similar andwill be good practice for you.

[α(AB)]ij =α [AB]ij Definition SMM [122]

=αn∑

k=1

[A]ik [B]kj Theorem EMP [154]

=n∑

k=1

α [A]ik [B]kj

=n∑

k=1

[αA]ik [B]kj Definition SMM [122]

= [(αA)B]ij Theorem EMP [154]

So the matrices α(AB) and (αA)B are equal, entry-by-entry, and by the definition ofmatrix equality (Definition ME [121]) we can say they are equal matrices. �

Theorem MMAMatrix Multiplication is AssociativeSuppose A is an m × n matrix, B is an n × p matrix and D is a p × s matrix. ThenA(BD) = (AB)D. �

Proof A matrix equality, so we’ll go entry-by-entry, no surprise there.

[A(BD)]ij =n∑

k=1

[A]ik [BD]kj Theorem EMP [154]

=n∑

k=1

[A]ik

(p∑

`=1

[B]k` [D]`j

)Theorem EMP [154]

=n∑

k=1

p∑`=1

[A]ik [B]k` [D]`j

We can switch the order of the summation since these are finite sums,

=

p∑`=1

n∑k=1

[A]ik [B]k` [D]`j

As [D]`j does not depend on the index k, we can factor it out of the inner sum,

=

p∑`=1

[D]`j

(n∑

k=1

[A]ik [B]k`

)

=

p∑`=1

[D]`j [AB]i` Theorem EMP [154]

=

p∑`=1

[AB]i` [D]`j

= [(AB)D]ij Theorem EMP [154]

So the matrices (AB)D and A(BD) are equal, entry-by-entry, and by the definition ofmatrix equality (Definition ME [121]) we can say they are equal matrices. �

Page 191: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PMM Properties of Matrix Multiplication 159

Theorem MMIPMatrix Multiplication and Inner ProductsIf we consider the vectors u, v ∈ Cm as m× 1 matrices then

〈u, v〉 = utv �

Proof Write u =

u1

u2

u3...

um

and v =

v1

v2

v3...

vm

. Then

〈u, v〉 =m∑

k=1

ukvk Definition IP [110]

=m∑

k=1

[u]k1 [v]k1

=m∑

k=1

[ut]1k

[v]k1 Definition TM [124]

=m∑

k=1

[ut]1k

[v]k1 Definition CCV [109]

=[utv]11

Theorem EMP [154]

To finish we just blur the distinction between a 1× 1 matrix (utv) and its lone entry.�

Theorem MMCCMatrix Multiplication and Complex ConjugationSuppose A is an m× n matrix and B is an n× p matrix. Then AB = A B. �

Proof To obtain this matrix equality, we will work entry-by-entry,[AB]ij

= [AB]ij Definition CM [52]

=n∑

k=1

[A]ik [B]kj Theorem EMP [154]

=n∑

k=1

[A]ik [B]kj Theorem CCRA [448]

=n∑

k=1

[A]ik [B]kj Theorem CCRM [448]

=n∑

k=1

[A]ik

[B]kj

Definition CCM [126]

=[A B

]ij

Theorem EMP [154]

So the matrices AB and A B are equal, entry-by-entry, and by the definition of matrixequality (Definition ME [121]) we can say they are equal matrices. �

Page 192: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PSHS Particular Solutions, Homogenous Solutions 160

One more theorem in this style, and its a good one. If you’ve been practicing with theprevious proofs you should be able to do this one yourself.

Theorem MMTMatrix Multiplication and TransposesSuppose A is an m× n matrix and B is an n× p matrix. Then (AB)t = BtAt. �

Proof This theorem may be surprising but if we check the sizes of the matrices involved,then maybe it will not seem so far-fetched. First, AB has size m × p, so its transposehas size p × m. The product of Bt with At is a p × n matrix times an n × m matrix,also resulting in a p×m matrix. So at least our objects are compatible for equality (andwould not be, in general, if we didn’t reverse the order of the operation).

Here we go again, entry-by-entry,[(AB)t

]ij

= [AB]ji Definition TM [124]

=n∑

k=1

[A]jk [B]ki Theorem EMP [154]

=n∑

k=1

[B]ki [A]jk

=n∑

k=1

[Bt]ik

[At]kj

Definition TM [124]

=[BtAt

]ij

Theorem EMP [154]

So the matrices (AB)t and BtAt are equal, entry-by-entry, and by the definition of matrixequality (Definition ME [121]) we can say they are equal matrices. �

This theorem seems odd at first glance, since we have to switch the order of A and B.But if we simply consider the sizes of the matrices involved, we can see that the switchis necessary for this reason alone. That the individual entries of the products then comealong is a bonus.

Notice how none of these proofs above relied on writing out huge general matriceswith lots of ellipses (“. . . ”) and trying to formulate the equalities a whole matrix at atime. This messy business is a “proof technique” to be avoided at all costs.

These theorems, along with Theorem VSPM [123], give you the “rules” for howmatrices interact with the various operations we have defined. Use them and use themoften. But don’t try to do anything with a matrix that you don’t have a rule for.Together, we would infomally call all these operations, and the attendant theorems, “thealgebra of matrices.” Notice, too, that every column vector is just a n × 1 matrix, sothese theorems apply to column vectors also. Finally, these results may make us feel thatthe definition of matrix multiplication is not so unnatural.

Subsection PSHSParticular Solutions, Homogenous Solutions

Having delayed presenting matrix multiplication, we have one theorem we could havestated long ago, but its proof is much easier now that we know how to represent a system

Page 193: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PSHS Particular Solutions, Homogenous Solutions 161

of linear equations with matrix multiplication and how to mix matrix multiplication withother operations.

The next theorem tells us that in order to find all of the solutions to a linear systemof equations, it is sufficient to find just one solution, and then find all of the solutionsto the corresponding homogenous system. This explains part of our interest in the nullspace, the set of all solutions to a homogenous system.

Theorem PSPHSParticular Solution Plus Homogenous SolutionsSuppose that z is one solution to the linear system of equations LS(A, b). Then y is asolution to LS(A, b) if and only if y = z + w for some vector w ∈ N(A). �

Proof We will work with the vector equality representations of the relevant systems ofequations, as described by Theorem SLEMM [152].

(⇐) Suppose y = z + w and w ∈ N(A). Then

Ay = A(z + w) = Az + Aw = b + 0 = b

demonstrating that y is a solution.(⇒) Suppose y is a solution to LS(A, b). Then

A(y − z) = Ay − Az = b− b = 0

which says that y − z ∈ N (A). In other words, y − z = w for some vector w ∈ N(A).Rewritten, this is y = z + w, as desired. �

Example MM.PSNSParticular solutions, homogenous solutions, Archetype DArchetype D [393] is a consistent system of equations with a nontrivial null space. Thewrite-up for this system begins with three solutions,

y1 =

0121

y2 =

4000

y3 =

7813

We will choose to have y1 play the role of z in the statement of Theorem PSPHS [161],any one of the three vectors listed here (or others) could have been chosen. To illustratethe theorem, we should be able to write each of these three solutions as the vector z plusa solution to the corresponding homogenous system of equations. Since 0 is always asolution to a homogenous system we can easily write

y1 = z = z + 0.

The vectors y2 and y3 will require a bit more effort. Solutions to the homogenous systemare exactly the elements of the null space of the coefficient matrix, which is

Sp

−3−110

,

2301

Page 194: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PSHS Particular Solutions, Homogenous Solutions 162

Then

y2 =

4000

=

0121

+

4−1−2−1

=

0121

+

(−2)

−3−110

+ (−1)

2301

= z + w2

where

w2 =

4−1−2−1

= (−2)

−3−110

+ (−1)

2301

is obviously a solution of the homogenous system since it is written as a linear combinationof the vectors describing the null space of the coefficient matrix as a span (or as a check,you could just evalute the equations in the homogenous system with w2).

Again

y3 =

7813

=

0121

+

77−12

=

0121

+

(−1)

−3−110

+ 2

2301

= z + w3

where

w3 =

77−12

= (−1)

−3−110

+ 2

2301

is obviously a solution of the homogenous system since it is written as a linear combinationof the vectors describing the null space of the coefficient matrix as a span (or as a check,you could just evalute the equations in the homogenous system with w2).

Here’s another view of this theorem, in the context of this example. Grab two newsolutions of the original system of equations, say

y4 =

110−3−1

y5 =

−4242

and form their difference,

u =

110−3−1

−−4242

=

15−2−7−3

.

It is no accident that u is a solution to the homogenous system (check this!). In otherwords, the difference between any two solutions to a linear system of equations is anelement of the null space of the coefficient matrix. This is an equivalent way to stateTheorem PSPHS [161]. If we let D denote the coefficient matrix then we can use thefollowing application of Theorem PSPHS [161] as the basis of a formal proof of this

Page 195: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.PSHS Particular Solutions, Homogenous Solutions 163

assertion,

D(y4 − y5) = D ((z + w4)− (z + w5))

= D(w4 −w5)

= Dw4 −Dw5

= 0− 0 = 0.

It would be very instructive to formulate the precise statement of a theorem and fill inthe details and justifications of the proof. 4

Page 196: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.EXC Exercises 164

Subsection EXCExercises

C20 Contributed by Robert BeezerCompute the product of the two matrices below, AB. Do this using the definitions of thematrix-vector product (Definition MVP [151]) and the definition of matrix multiplication(Definition MM [153]).

A =

2 5−1 32 −2

B =

[1 5 −3 42 0 2 −3

]

Solution [165]

T40 Contributed by Robert BeezerSuppose that A is an m× n matrix and B is an n× p matrix. Prove that the null spaceof B is a subset of the null space of AB, that is N (B) ⊆ N (AB). Provide an examplewhere the opposite is false, in other words give an example where N (AB) 6⊆ N (B).Solution [165]

TODO: Converse with A nonsingular?

Page 197: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MM.SOL Solutions 165

Subsection SOLSolutions

C20 Exercise [164] Contributed by Robert BeezerBy Definition MM [153],

AB =

2 5−1 32 −2

[12

]∣∣∣∣∣∣ 2 5−1 32 −2

[50

]∣∣∣∣∣∣ 2 5−1 32 −2

[−32

]∣∣∣∣∣∣ 2 5−1 32 −2

[ 4−2

]Repeated applications of Definition MVP [151] give

=

1

2−12

+ 2

53−2

∣∣∣∣∣∣ 5

2−12

+ 0

53−2

∣∣∣∣∣∣ −3

2−12

+ 2

53−2

∣∣∣∣∣∣ 4

2−12

+ (−3)

53−2

=

12 10 4 −75 −5 9 −13−2 10 −10 14

T40 Exercise [164] Contributed by Robert BeezerTo prove that one set is a subset of another, we start with an element of the smaller setand see if we can determine that it is a member of the larger set (Technique SE [14]).Suppose x ∈ N (B). Then we know that Bx = 0 by Definition NSM [54]. Consider

(AB)x = A(Bx) Theorem MMA [158]

= A0 Hypothesis

= 0 Theorem MMZM [156]

To show that the inclusion does not hold in the opposite direction, choose B to be anynonsingular matrix of size n. Then N (B) = {0} by Theorem NSTNS [59]. Let A be thesquare zero matrix, O, of the same size. Then AB = OB = O by Theorem MMZM [156]and therefore N (AB) = Cn, and is not a subset of N (B) = {0}.

Page 198: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section MISLE Matrix Inverses and Systems of Linear Equations 166

Section MISLE

Matrix Inverses and Systems of Linear Equations

We begin with a familiar example, performed in a novel way.

Example MISLE.SABMISolutions to Archetype B with a matrix inverseArchetype B [384] is the system of m = 3 linear equations in n = 3 variables,

−7x1 − 6x2 − 12x3 = −33

5x1 + 5x2 + 7x3 = 24

x1 + 4x3 = 5

By Theorem SLEMM [152] we can represent this system of equations as

Ax = b

where

A =

−7 −6 −125 5 71 0 4

x =

x1

x2

x3

b =

−33245

.

We’ll pull a rabbit out of our hat and present the 3× 3 matrix B,

B =

−10 −12 −9132

8 112

52

3 52

and note that

BA =

−10 −12 −9132

8 112

52

3 52

−7 −6 −125 5 71 0 4

=

1 0 00 1 00 0 1

.

Now apply this computation to the problem of solving the system of equations,

Ax =b

B(Ax) =Bb

(BA)x =Bb Theorem MMA [158]

I3x =Bb

x =Bb Theorem MMIM [156]

So we have

x = Bb =

−10 −12 −9132

8 112

52

3 52

−33245

=

−352

.

Page 199: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.IM Inverse of a Matrix 167

So with the help and assistance of B we have been able to determine a solution to thesystem represented by Ax = b through judicious use of matrix multiplication. We knowby Theorem NSMUS [59] that since the coefficient matrix in this example is nonsingular,there would be a unique solution, no matter what the choice of b. The derivation aboveamplifies this result, since we were forced to conclude that x = Bb and the solutioncouldn’t be anything else. You should notice that this argument would hold for anyparticular value of b. 4

The matrix B of the previous example is called the inverse of A. When A and B arecombined via matrix multiplication, the result is the identity matrix, which in this caseleft just the vector of unknowns, x, on the left-side of the equation. This is entirelyanalagous to how we would solve a single linear equation like 3x = 12. We wouldmultiply both sides by 1

3= 3−1, the multiplicative inverse of 3. This works fine for any

scalar multiple of x, except for zero, which does not have a multiplicative inverse. Formatrices, it is more complicated. Some matrices have inverses, some do not. And when amatrix does have an inverse, just how would we compute it? In other words, just wheredid that matrix B in the last example come from? Are there other matrices that mighthave worked just as well?

Subsection IMInverse of a Matrix

Definition MIMatrix InverseSuppose A and B are square matrices of size n such that AB = In and BA = In. ThenA is invertible and B is the inverse of A, and we write B = A−1. �

Notice that if B is the inverse of A, then we can just as easily say A is the inverse of B,or A and B are inverses of each other.

Not every square matrix has an inverse. In Example MISLE.SABMI [166] the matrixB is the inverse the coefficient matrix of Archetype B [384]. To see this it only remainsto check that AB = I3. What about Archetype A [379]? It is an example of a squarematrix without an inverse.

Example MISLE.MWIAAA matrix without an inverse, Archetype AConsider the coefficient matrix from Archetype A [379],

A =

1 −1 22 1 11 1 0

Suppose that A is invertible and does have an inverse, say B. Choose the vector ofconstants

b =

132

Page 200: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.IM Inverse of a Matrix 168

and consider the system of equations LS(A, b). Just as in Example MISLE.SABMI [166],this vector equation would have the unique solution x = Bb.

However, this system is inconsistent. Form the augmented matrix [A | b] and row-reduce to 1 0 1 0

0 1 −1 0

0 0 0 1

which allows to recognize the inconsistency by Theorem RCLS [41].

So the assumption of A’s inverse leads to a logical inconsistency (the system can’t beboth consistent and inconsistent), so our assumption is false. A is not invertible.

Its possible this example is less than satisfying. Just where did that particular choiceof the vector b come from anyway? Turns out its not too mysterious. We wanted aninconsistent system, so Theorem RCS [128] suggested choosing a vector outside of therange of A (see Example RM.RAA [139] for full disclosure). 4

Lets look at one more matrix inverse before we embark on a more systematic study.

Example MISLE.MIAKMatrix Inverse, Archetype KConsider the matrix defined as Archetype K [423],

K =

10 18 24 24 −1212 −2 −6 0 −18−30 −21 −23 −30 3927 30 36 37 −3018 24 30 30 −20

.

And the matrix

L =

1 −

(94

)−(

32

)3 −6

212

434

212

9 −9−15 −

(212

)−11 −15 39

2

9 154

92

10 −1592

34

32

6 −(

192

)

.

Then

KL =

10 18 24 24 −1212 −2 −6 0 −18−30 −21 −23 −30 3927 30 36 37 −3018 24 30 30 −20

1 −(

94

)−(

32

)3 −6

212

434

212

9 −9−15 −

(212

)−11 −15 39

2

9 154

92

10 −1592

34

32

6 −(

192

)

=

1 0 0 0 00 1 0 0 00 0 1 0 00 0 0 1 00 0 0 0 1

and

LK =

1 −

(94

)−(

32

)3 −6

212

434

212

9 −9−15 −

(212

)−11 −15 39

2

9 154

92

10 −1592

34

32

6 −(

192

)

10 18 24 24 −1212 −2 −6 0 −18−30 −21 −23 −30 3927 30 36 37 −3018 24 30 30 −20

=

1 0 0 0 00 1 0 0 00 0 1 0 00 0 0 1 00 0 0 0 1

so by Definition MI [167], we can say that K is invertible and write L = K−1. 4

Page 201: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.CIM Computing the Inverse of a Matrix 169

We will now concern ourselves less with whether or not an inverse of a matrix exists, butinstead with how you can find one when it does exist. In Section MINSM [176] we willhave some theorems that allow us to more quickly and easily determine when a matrixis invertible.

Subsection CIMComputing the Inverse of a Matrix

We will have occasion in this subsection (and later) to reference the following frequentlyused vectors, so we will make a useful definition now.

Definition SUVStandard Unit VectorsLet ei ∈ Cm denote the column vector that is column i of the m×m identity matrix Im.Then the set

{e1, e2, e3, . . . , em} = {ei | 1 ≤ i ≤ m}

is the set of standard unit vectors in Cm. �

Notice that ei is a column vector full of zeros, with a lone 1 in the i-th position. We willmake reference to these vectors often.

We’ve seen that the matrices from Archetype B [384] and Archetype K [423] bothhave inverses, but these inverse matrices have just dropped from the sky. How would wecompute an inverse? And just when is a matrix invertible, and when is it not? Writing aputative inverse with n2 unknowns and solving the resultant n2 equations is one approach.Applying this approach to 2× 2 matrices can get us somewhere, so just for fun, let’s doit.

Theorem TTMITwo-by-Two Matrix InverseSuppose

A =

[a bc d

]Then A is invertible if and only if ad− bc 6= 0. When A is invertible, we have

A−1 =1

ad− bc

[d −b−c a

]. �

Proof (⇐) If ad− bc 6= 0 then the displayed formula is legitimate (we are not dividingby zero), and it is a simple matter to actually check that A−1A = AA−1 = I2.

(⇒) Assume that A is invertible, and proceed with a proof by contradiction, byassuming also that ad− bc = 0. This means that ad = bc. Let

B =

[e fg h

]

Page 202: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.CIM Computing the Inverse of a Matrix 170

be a putative inverse of A. This means that

I2 = AB =

[ae + bg af + bhce + dg cf + dh

]For the matrix on the right, multiply the top row by c and the bottom row by a. Sincewe are assuming that ad = bc, massage the bottom row by replacing ad by bc in twoplaces. The result is that the two rows of the matrix are identical. Suppose we did thesame to I2, multiply the top row by c and the bottom row by a, and then arrived arrivedat equal rows? Given the form of I2 there is only one way this could happen: a = 0 andc = 0.

With this information, the product AB simplifies to

AB =

[bg bhdg dh

]=

[1 00 1

]= I2

So bg = dh = 1 and thus b, g, d, h are all nonzero. But then bh and dg (the “othercorners”) must also be nonzero, so this is (finally) a contradiction. So our assumptionwas false and we see that ad− bc 6= 0 whenever A has an inverse. �

There are several ways one could try to prove this theorem, but there is a continualtemptation to divide by one of the eight entries involved (a through f), but we can neverbe sure if these numbers are zero or not. This could lead to an analysis by cases, whichis messy,. . . . Note how the above proof never divides, but always multiplies, and howzero/nonzero considerations are handled. Pay attention to the expression ad−bc, we willsee it again in a while.

This theorem is cute, and it is nice to have a formula for the inverse, and a conditionthat tells us when we can use it. However, this approach becomes impractical for largermatrices, even though it is possible to demonstrate that, in theory, there is a generalformula. Instead, we will work column-by-column. Let’s first work an example that willmotivate the main theorem and remove some of the previous mystery.

Example MISLE.CMIAKComputing a Matrix Inverse, Archetype KConsider the matrix defined as Archetype K [423],

A =

10 18 24 24 −1212 −2 −6 0 −18−30 −21 −23 −30 3927 30 36 37 −3018 24 30 30 −20

.

For its inverse, we desire a matrix B so that AB = I5. Emphazing the structure of thecolumns and employing the definition of matrix multiplication Definition MM [153],

AB = I5

A[B1|B2|B3|B4|B5] = [e1|e2|e3|e4|e5]

[AB1|AB2|AB3|AB4|AB5] = [e1|e2|e3|e4|e5].

Equating the matrices column-by-column we have

AB1 = e1 AB2 = e2 AB3 = e3 AB4 = e4 AB5 = e5.

Page 203: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.CIM Computing the Inverse of a Matrix 171

Since the matrix B is what we are trying to compute, we can view each column, Bi, as acolumn vector of unknowns. Then we have five systems of equations to solve, each with 5equations in 5 variables. Notice that all 5 these systems has the same coefficient matrix.We’ll now solve each system in turn,

Row-reduce the augmented matrix of the linear system LS(A, e1),10 18 24 24 −12 112 −2 −6 0 −18 0−30 −21 −23 −30 39 027 30 36 37 −30 018 24 30 30 −20 0

1 0 0 0 0 1

0 1 0 0 0 212

0 0 1 0 0 −15

0 0 0 1 0 9

0 0 0 0 1 92

→ B1 =

1212

−15992

Row-reduce the augmented matrix of the linear system LS(A, e2),

10 18 24 24 −12 012 −2 −6 0 −18 1−30 −21 −23 −30 39 027 30 36 37 −30 018 24 30 30 −20 0

1 0 0 0 0 −94

0 1 0 0 0 434

0 0 1 0 0 −212

0 0 0 1 0 154

0 0 0 0 1 34

→ B2 =

−9

4434

−212

15434

Row-reduce the augmented matrix of the linear system LS(A, e3),

10 18 24 24 −12 012 −2 −6 0 −18 0−30 −21 −23 −30 39 127 30 36 37 −30 018 24 30 30 −20 0

1 0 0 0 0 −32

0 1 0 0 0 212

0 0 1 0 0 −11

0 0 0 1 0 92

0 0 0 0 1 32

→ B3 =

−3

2212

−119232

Row-reduce the augmented matrix of the linear system LS(A, e4),

10 18 24 24 −12 012 −2 −6 0 −18 0−30 −21 −23 −30 39 027 30 36 37 −30 118 24 30 30 −20 0

1 0 0 0 0 3

0 1 0 0 0 9

0 0 1 0 0 −15

0 0 0 1 0 10

0 0 0 0 1 6

→ B4 =

39−15106

Row-reduce the augmented matrix of the linear system LS(A, e5),

10 18 24 24 −12 012 −2 −6 0 −18 0−30 −21 −23 −30 39 027 30 36 37 −30 018 24 30 30 −20 1

1 0 0 0 0 −6

0 1 0 0 0 −9

0 0 1 0 0 392

0 0 0 1 0 −15

0 0 0 0 1 −192

→ B5 =

−6−9392

−15−19

2

Page 204: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.CIM Computing the Inverse of a Matrix 172

We can now collect our 5 solution vectors into the matrix B,

B =[B1|B2|B3|B4|B5]

=

1212

−15992

∣∣∣∣∣∣∣∣∣∣

−9

4434

−212

15434

∣∣∣∣∣∣∣∣∣∣

−3

2212

−119232

∣∣∣∣∣∣∣∣∣∣

39−15106

∣∣∣∣∣∣∣∣∣∣

−6−9392

−15−19

2

=

1 −

(94

)−(

32

)3 −6

212

434

212

9 −9−15 −

(212

)−11 −15 39

2

9 154

92

10 −1592

34

32

6 −(

192

)

By this method, we know that AB = I5. Check that BA = I5, and then we will knowthat we have the inverse of A. 4

Notice how the five systems of equations in the preceding example were all solved byexactly the same sequence of row operations. Wouldn’t it be nice to avoid this obviousduplication of effort? Our main theorem for this section follows, and it mimics thisprevious example, while also avoiding all the overhead.

Theorem CINSMComputing the Inverse of a NonSingular MatrixSuppose A is a nonsingular square matrix of size n. Create the n × 2n matrix M byplacing the n×n identity matrix In to the right of the matrix A. Let N be a matrix thatis row-equivalent to M and in reduced row-echelon form. Finally, let B be the matrixformed from the final n columns of N . Then AB = In. �

Proof A is nonsingular, so by Theorem NSRRI [57] there is a sequence of row operationsthat will convert A into In. It is this same sequence of row operations that will convertM into N , since having the identity matrix in the first n columns of N is sufficient toguarantee that it is in reduced row-echelon form.

If we consider the systems of linear equations, LS(A, ei), 1 ≤ i ≤ n, we see that theaforementioned sequence of row operations will also bring the augmented matrix of eachsystem into reduced row-echelon form. Furthermore, the unique solution to each of thesesystems appears in column n + 1 of the row-reduced augmented matrix and is equal tocolumn n + i of N . Let N1, N2, N3, . . . , N2n denote the columns of N . So we find,

AB =A[Nn+1|Nn+2|Nn+3| . . . |Nn+n]

=[ANn+1|ANn+2|ANn+3| . . . |ANn+n]

=[e1|e2|e3| . . . |en]

=In

as desired. �

Does this theorem remind you of any others we’ve seen lately? (Hint: Theorem RNS [136].)We have to be just a bit careful here. This theorem only guarantees that AB = In, whilethe definition requires that BA = In also. However, we’ll soon see that this is always thecase, in Theorem OSIS [177], so the title of this theorem is not inaccurate.

We’ll finish by computing the inverse for the coefficient matrix of Archetype B [384],the one we just pulled from a hat in Example MISLE.SABMI [166]. There are moreexamples in the Archetypes (Chapter A [375]) to practice with, though notice that it issilly to ask for the inverse of a rectangular matrix (the sizes aren’t right) and not everysquare matrix has an inverse (remember Example MISLE.MWIAA [167]?).

Page 205: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.PMI Properties of Matrix Inverses 173

Example MISLE.CMIABComputing a Matrix Inverse, Archetype BArchetype B [384] has a coefficient matrix given as

B =

−7 −6 −125 5 71 0 4

Exercising Theorem CINSM [172] we set

M =

−7 −6 −12 1 0 05 5 7 0 1 01 0 4 0 0 1

.

which row reduces to

N =

1 0 0 −10 −12 −90 1 0 13

28 11

2

0 0 1 52

3 52

.

So

B−1 =

−10 −12 −9132

8 112

52

3 52

once we check that B−1B = I3 (the product in the opposite order is a consequence of thetheorem). 4

While we can use a row-reducing procedure to compute an inverse, many computationaldevices have a built-in procedure.

Computation Note MI.MMAMatrix Inverses (Mathematica)If A is a matrix defined in Mathematica, then Inverse[A] will return the inverse of

A, should it exist. In the case where A does not have an inverse Mathematica will tellyou the matrix is singular (see Theorem NSI [177]). ⊗

Subsection PMIProperties of Matrix Inverses

The inverse of a matrix enjoys some nice properties. We collect a few here. First, amatrix can have but one inverse.

Theorem MIUMatrix Inverse is UniqueSuppose the square matrix A has an inverse. Then A−1 is unique. �

Page 206: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.PMI Properties of Matrix Inverses 174

Proof As described in Technique U [59], we will assume that A has two inverses. Thehypothesis tells there is at least one. Suppose then that B and C are both inverses for A.Then, repeated use of Definition MI [167] and Theorem MMIM [156] plus one applicationof Theorem MMA [158] gives

B = BIn = B(AC) = (BA)C = InC = C

and we conclude that B and C cannot be different. So any matrix that acts like theinverse, must be the inverse. �

When most of dress in the morning, we put on our socks first, followed by our shoes. Inthe evening we must then first remove our shoes, followed by our socks. Try to connectthe conclusion of the following theorem with this everyday example.

Theorem SSSocks and ShoesSuppose A and B are invertible matrices of size n. Then (AB)−1 = B−1A−1 and AB isan invertible matrix. �

Proof At the risk of carrying our everyday analogies too far, the proof of this theorem isquite easy when we compare it to the workings of a dating service. We have a statementabout the inverse of the matrix AB, which for all we know right now might not even exist.Suppose AB was to sign up for a dating service with two requirements for a compatibledate. Upon multiplication on the left, and on the right, the result should be the identitymatrix. In other words, AB’s ideal date would be its inverse.

Now along comes the matrix B−1A−1 (which we know exists because our hypothesissays both A and B are invertible), also looking for a date. Lets see if B−1A−1 is a goodmatch for AB,

(B−1A−1)(AB) =B−1(A−1A)B = B−1InB = B−1B = In

(AB)(B−1A−1) =A(BB−1)A−1 = AInA−1 = AA−1 = In.

So the matrix B−1A−1 has met all of the requirements to be AB’s inverse (date) and wecan write (AB)−1 = B−1A−1. �

Theorem MIMIMatrix Inverse of a Matrix InverseSuppose A is an invertible matrix. Then (A−1)−1 = A and A−1 is invertible. �

Proof As with the proof of Theorem SS [174], we see if A is a suitable inverse for A−1.Apply Definition MI [167] to see that

AA−1 = In

A−1A = In

The matrix A has met all the requirements to be the inverse of A−1, so we can write(A−1)−1 = A. �

Theorem MITMatrix Inverse of a TransposeSuppose A is an invertible matrix. Then (At)−1 = (A−1)t and At is invertible. �

Page 207: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MISLE.PMI Properties of Matrix Inverses 175

Proof As with the proof of Theorem SS [174], we see if (A−1)t is a suitable inverse forAt. Apply Theorem MMT [160] to see that

(A−1)tAt =(AA−1)t = I tn = In

At(A−1)t =(A−1A)t = I tn = In

The matrix (A−1)t has met all the requirements to be the inverse of At, so we can write(At)−1 = (A−1)t. �

Theorem MISMMatrix Inverse of a Scalar MultipleSuppose A is an invertible matrix and α is a nonzero scalar. Then (αA)−1 = 1

αA−1 and

αA is invertible. �

Proof As with the proof of Theorem SS [174], we see if 1αA−1 is a suitable inverse for

αA. Apply Theorem MMSMM [157] to see that(1

αA−1

)(αA) =

(1

αα

)(AA−1

)= 1In = In

(αA)

(1

αA−1

)=

1

α

)(A−1A

)= 1In = In

The matrix 1αA−1 has met all the requirements to be the inverse of αA, so we can write

(αA)−1 = 1αA−1. �

Notice that there are some likely theorems that are missing here. For example, it wouldbe tempting to think that (A + B)−1 = A−1 + B−1, but this is false. Can you find acounterexample?

Page 208: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section MINSM Matrix Inverses and NonSingular Matrices 176

Section MINSM

Matrix Inverses and NonSingular Matrices

We saw in Theorem CINSM [172] that if a square matrix A is nonsingular, then thereis a matrix B so that AB = In. In other words, B is halfway to being an inverseof A. We will see in this section that B automatically fulfills the second condition(BA = In). Example MISLE.MWIAA [167] showed us that the coefficient matrix fromArchetype A [379] had no inverse. Not coincidentally, this coefficient matrix is singular.We’ll make all these connections precise now. Not many examples or definitions in thissection, just theorems.

Subsection NSMINonSingular Matrices are Invertible

We need a couple of technical results for starters. Some books would call these minor,but essential, results “lemmas.” We’ll just call ’em theorems.

Theorem PWSMSProduct With a Singular Matrix is SingularSuppose that A or B are matrices of size n, and one, or both, is singular. Then theirproduct, AB, is singular. �

Proof We will use the vector equation representation of the relevant systems of equationsthroughout the proof (Theorem SLEMM [152]). We’ll do the proof in two cases, and itsinteresting to notice how we break down the cases.

Case 1. Suppose B is singular. Then there is a nontrivial vector z so that Bz = 0.Then

(AB)z = A(Bz) = A0 = 0

so we can conclude that AB is singular.Case 2. Suppose B is nonsingular and A is singular. This is probably not the second

case you were expecting. Why not just state the second case as “A is singular”? Thebest answer is that the proof is easier with the more restrictive assumption that A issingular and B is nonsingular. But before we see why, convince yourself that the twocases, as stated, will cover all the possibilities allowed by our hypothesis.

Since A is singular, there is a nontrivial vector y so that Ay = 0. Now consider thelinear system LS(B, y). Since B is nonsingular, the system has a unique solution, whichwe will call w. We claim w is not the zero vector. If w = 0, then

y = Bw = B0 = 0

contrary to y being nontrivial. So w 6= 0. The pieces are in place, so here we go,

(AB)w = A(Bw) = Ay = 0

which says, since w is nontrivial, that AB is singular. �

Page 209: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MINSM.NSMI NonSingular Matrices are Invertible 177

Theorem OSISOne-Sided Inverse is SufficientSuppose A is a nonsingular matrix of size n, and B is a square matrix of size n such thatAB = In. Then BA = In. �

Proof The matrix In is nonsingular (since it row-reduces easily to In, Theorem NSRRI [57]).If B is singular, then Theorem PWSMS [176] would imply that In is singular, a contra-diction. So B must be nonsingular also. Now that we know that B is nonsingular, wecan apply Theorem CINSM [172] to assert the existence of a matrix C so that BC = In.This application of Theorem CINSM [172] could be a bit confusing, mostly because ofthe names of the matrices involved. B is nonsingular, so there must be a “right-inverse”for B, and we’re calling it C.

NowC = InC = (AB)C = A(BC) = AIn = A.

So it happens that the matrix C we just found is really A in disguise. So we can write

In = BC = BA

which is the desired conclusion. �

So Theorem OSIS [177] tells us that if A is nonsingular, then the matrix B guaranteedby Theorem CINSM [172] will be both a “right-inverse” and a “left-inverse” for A, so Ais invertible and A−1 = B.

So if you have a nonsingular matrix, A, you can use the procedure described inTheorem CINSM [172] to find an inverse for A. If A is singular, then the procedure inTheorem CINSM [172] will fail as the first n columns of M will not row-reduce to theidentity matrix.

This may feel like we are splitting hairs, but its important that we do not makeunfounded assumptions. These observations form the next theorem.

Theorem NSINonSingularity is InvertibilitySuppose that A is a square matrix. Then A is nonsingular if and only if A is invertible.�

Proof (⇐) Suppose A is invertible, and consider the homogenous system representedby Ax = 0,

Ax = 0

A−1Ax = A−10

Inx = 0

x = 0

So A has a trivial null space, which is a fancy way of saying that A is nonsingular.(⇒) Suppose A is nonsingular. By Theorem CINSM [172] we find B so that AB = In.

Then Theorem OSIS [177] tells us that BA = In. So B is A’s inverse, and by construction,A is invertible. �

So the properties of having an inverse and of having a trivial null space are one and thesame. Can’t have one without the other. Now we can update our list of equivalences fornonsingular matrices (Theorem NSME3 [140]).

Page 210: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MINSM.OM Orthogonal Matrices 178

Theorem NSME4NonSingular Matrix Equivalences, Round 4Suppose that A is a square matrix of size n. The following are equivalent.

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the zero vector, N (A) = {0}.

4. The linear system LS(A, b) has a unique solution for every possible choice of b.

5. The columns of A are a linearly independent set.

6. The range of A is Cn, R (A) = Cn.

7. A is invertible. �

In the case that A is a nonsingular coefficient matrix of a system of equations, the inverseallows us to very quickly compute the unique solution, for any vector of constants.

Theorem SNSCMSolution with NonSingular Coefficient MatrixSuppose that A is nonsingular. Then the unique solution to LS(A, b) is A−1b. �

Proof By Theorem NSMUS [59] we know already that LS(A, b) has a unique solutionfor every choice of b. We need to show that the expression given is indeed a solution.That’s easy, just “plug it in” to the corresponding vector equation representation,

A(A−1b) = (AA−1)b = Inb = b. �

Subsection OMOrthogonal Matrices

Definition OMOrthogonal Matrices

Suppose that Q is a square matrix of size n such that(Q)t

Q = In. Then we say Q isorthogonal. �

This condition may seem rather far-fetched at first glance. Would there be any matrixthat behaved this way? Well, yes, here’s one.

Example MINSM.OM3Orthogonal matrix of size 3

Q =

1+i√

53+2 i√

552+2i√

221−i√

52+2 i√

55−3+i√

22i√5

3−5 i√55

− 2√22

The computations get a bit tiresome, but if you work your way through

(Q)t

Q, you willarrive at the 3× 3 identity matrix I3. 4

Page 211: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MINSM.OM Orthogonal Matrices 179

Orthogonal matrices do not have to look quite so gruesome. Here’s a larger one that isa bit more pleasing.

Example MINSM.OPMOrthogonal permutation matrixThe matrix

P =

0 1 0 0 00 0 0 1 01 0 0 0 00 0 0 0 10 0 1 0 0

is orthogonal as can be easily checked. Notice that it is just a rearrangement of thecolumns of the 5× 5 identity matrix, I5 (Definition IM [57]).

An interesting exercise is to build another 5×5 orthogonal matrix, R, using a differentrearrangment of the columns of I5. Then form the product PR. This will be anotherorthogonal matrix (reference exercise here). If you were to build all 5 × 4 × 3 × 2 ×1! = 120 matrices of this type you would have a set that remains closed under matrixmultiplication. It is an example of another algebraic structure known as a group sinceit is closed, associative, has an identity (I5), and inverses (Theorem OMI [179]). Noticethough that the operation in this group is not commutative! 4

Orthogonal matrices have easily computed inverses. They also have columns that formorthonormal sets. Here are the theorems that show us that orthogonal matrices are notas strange as they might initially appear.

Theorem OMIOrthogonal Matrices are InvertibleSuppose that Q is an orthogonal matrix of size n. Then Q is nonsingular, and Q−1 =(Q)t. �

Proof By Definition OM [178], we know that (Q)tQ = In. If (Q)t or Q were singu-lar, then this equation, together with Theorem PWSMS [176], would have us concludethat In is singular, a contradiction, since In row-reduces to the identity matrix (Theo-rem NSRRI [57]). So Q, and (Q)t, are both nonsingular.

The equation (Q)tQ = In gets us halfway to an inverse of Q, and since we now knowthat (Q)t is nonsingular, Theorem OSIS [177] tells us that Q(Q)t = In also. So Q and(Q)t are inverses of each other. �

Theorem COMOSColumns of Orthogonal Matrices are Orthonormal SetsSuppose that A is a square matrix of size n with columns S = {A1, A2, A3, . . . , An}.Then A is an orthogonal matrix if and only if S is an orthonormal set. �

Proof The proof revolves around recognizing that a typical entry of the product (A)tA

Page 212: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MINSM.OM Orthogonal Matrices 180

is an inner product of columns of A. Here are the details to support this claim.

[(A)t

A]

ij=

n∑k=1

[(A)t

]ik

[A]kj Theorem EMP [154]

=n∑

k=1

[A]ki

[A]kj Definition TM [124]

=n∑

k=1

[A]ki [A]kj Definition CCM [126]

=n∑

k=1

[A]kj [A]ki Commutativity

= 〈Aj, Ai〉 Definition IP [110]

S = {A1, A2, A3, . . . , An} is an orthonormal set if and only if 〈Aj, Ai〉 = 0 when i 6= jand 〈Aj, Ai〉 = 1 when i = j by Definition ONS [119]. However, the above expression

for an entry of the matrix product(A)t

A as an inner product means this latter condition

is precisely the statement that(A)t

A = In. �

Example MINSM.OSMCOrthonormal Set from Matrix ColumnsThe matrix

Q =

1+i√

53+2 i√

552+2i√

221−i√

52+2 i√

55−3+i√

22i√5

3−5 i√55

− 2√22

from Example MINSM.OM3 [178] is an orthogonal matrix. By Theorem COMOS [179]its columns

1+i√

53+2 i√

552+2i√

22

,

1−i√

52+2 i√

55−3+i√

22

i√5

3−5 i√55

− 2√22

form an orthonormal set. You might find checking the six inner products of pairs of these

vectors easier than doing the matrix product(Q)t

Q. 4

When using vectors and matrices that only have real number entries, orthogonalmatrices are those matrices with inverses that equal their transpose. Similarly, the innerproduct is the familiar dot product. Keep this special case in mind as you read the nexttheorem.

Theorem OMPIPOrthogonal Matrices Preserve Inner ProductsSuppose that Q is an orthogonal matrix of size n and u and v are two vectors from Cn.Then

〈Qu, Qv〉 = 〈u, v〉 ‖Qv‖ = ‖v‖. �

Page 213: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MINSM.OM Orthogonal Matrices 181

Proof We will be a bit fast and loose with the interplay of conjugates and transposeshere, but you should be able to supply some of the missing theorems.

〈Qu, Qv〉 = (Qu)tQv [??]

= utQtQv Theorem MMT [160]

= utQtQv Theorem MMC [??]

= utQtQv Conjugation twice

= utQtQv

= utQtQv Theorem MMC [??]

= utInv Definition OM [178]

= utInv In has real entries

= utv Theorem MMIM [156]

= 〈u, v〉 Definition IP [110]

The second conclusion is just a specialization.

‖Qv‖2 = 〈Qv, Qv〉 Theorem IPN [113]

= 〈v, v〉 Previous conclusion

= ‖v‖2 Theorem IPN [113]

Now take a square root on both sides to get the result. �

Definition AAdjoint

If A is a square matrix, then its adjoint is AH =(A)t

. �

Sometimes a matrix is equal to its adjoint. A simple example would be any symmetricmatrix with real entries.

Definition HMHermitian MatrixThe square matrix A is Hermitian (or self-adjoint) if A =

(A)t �

Page 214: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MINSM.EXC Exercises 182

Subsection EXCExercises

C40 Contributed by Robert BeezerSolve the system of equations below using the inverse of a matrix.

x1 + x2 + 3x3 + x4 = 5

−2x1 − x2 − 4x3 − x4 = −7

x1 + 4x2 + 10x3 + 2x4 = 9

−2x1 − 4x3 + 5x4 = 9

Solution [183]

TODO: Product of orthogonal matrices is orthogonal (solution via product)TODO: Hermitian matrices have real entries on the diagonal

Page 215: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MINSM.SOL Solutions 183

Subsection SOLSolutions

C40 Exercise [182] Contributed by Robert BeezerThe coefficient matrix and vector of constants for the system are

1 1 3 1−2 −1 −4 −11 4 10 2−2 0 −4 5

b =

5−799

A−1 can be computed by using a calculator, or by the method of Theorem CINSM [172].Then Theorem SNSCM [178] says the unique solution is

A−1b =

38 18 −5 −296 47 −12 −5−39 −19 5 2−16 −8 2 1

5−799

=

1−213

Page 216: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

VS: Vector Spaces

Section VS

Vector Spaces

Linear algebra is the study of two fundamental objects, vector spaces and linear trans-formations (see Chapter LT [302]). Here we present an axiomatic definition of vectorspaces, which will lead to an extra increment of abstraction. The power of mathematicsis often derived from generalizing many different situations into one abstract formulation,and that is exactly what we will be doing now.

Subsection VSVector Spaces

Here is one of our two most important definitions.

Definition VSVector SpaceSuppose that V is a set upon which we have defined two operations: (1) vector addi-tion, which combines two elements of V and is denoted by “+”, and (2) scalar multi-plication, which combines a complex number with an element of V and is denoted byjuxtaposition. Then V , along with the two operations, is a vector space if the followingten requirements (better known as “axioms”) are met. Let u, v, w ∈ V and α, β ∈ C.

1. u + v ∈ V (Additive closure)

2. αu ∈ V (Scalar closure)

3. u + v = v + u (Commutativity)

4. u + (v + w) = (u + v) + w (Associativity of vector addition)

5. There is a vector, 0 ∈ V , called the zero vector, such that u+0 = u for all u ∈ V .(Additive identity)

6. For each vector u ∈ V , there exists a vector −u ∈ V so that u + (−u) = 0.(Additive inverses)

184

Page 217: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.EVS Examples of Vector Spaces 185

7. α(βu) = (αβ)u (Associativity of scalar multiplication)

8. α(u + v) = αu + αv (Distributivity across vector addition)

9. (α + β)u = αu + βu (Distributivity across addition)

10. 1u = u (Scalar multiplication with 1)

The objects in V are called vectors, no matter what else they might really be, simplyby virtue of being elements of a vector space. �

Now, there are several important observations to make. Many of these will be easierto understand on a second or third reading, and especially after carefully studying theexamples in Subsection VS.EVS [185].

An axiom is often a “self-evident” truth. Something so fundamental that we all agreeit is true and accept it without proof. Typically, it would be the logical underpinningthat we would begin to build theorems upon. Here, the use is slightly different. The tenrequirements in Definition VS [184] are the most basic properties of the objects relevantto a study of linear algebra, and all of our theorems will be built upon them, as we willbegin to see in Subsection VS.VSP [190]. So we will refer to “the vector space axioms.”After studying the remainder of this chapter, you might return here and remind yourselfhow all our forthcoming theorems and definitions rest on this foundation.

As we will see shortly, the objects in V can be anything, even though we will callthem vectors. We have been working with vectors frequently, but we should stress herethat these have so far just been column vectors — scalars arranged in a columnar list offixed length. In a similar vein, you have used the symbol “+” for many years to representthe addition of numbers (scalars). We have extended its use to the addition of columnvectors, and now we are going to recycle it even further and let it denote vector additionin any possible vector space. So when describing a new vector space, we will have todefine exactly what “+” is. Similar comments apply to scalar multiplication. Conversely,we can define our operations any way we like, so long as the ten axioms are fulfilled (seeExample VS.CVS [188]).

A vector space is composed of three objects, a set and two operations. However, weusually use the same symbol for both the set and the vector space itself. Do not let thisconvenience fool you into thinking the operations are secondary!

This discussion has either convinced you that we are really embarking on a new levelof abstraction, or they have seemed cryptic, mysterious or nonsensical. In any case, letslook at some concrete examples.

Subsection EVSExamples of Vector Spaces

Our aim in this subsection is to give you a storehouse of examples to work with, to becomecomfortable with the ten vector space axioms and to convince you that the multitude ofexamples justifies (at least initially) making such a broad definition as Definition VS [184].Some our claims will be justified by reference to previous theorems, we will prove somefacts from scratch, and we will do one non-trivial example completely. In other places,our usual thoroughness will be neglected, so grab paper and pencil and play along.

Page 218: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.EVS Examples of Vector Spaces 186

Example VS.VSCMThe vector space Cm

Set: Cm, all column vectors of size m, Definition VSCM [65].Vector Addition: The “usual” addition, given in Definition CVA [67].Scalar Multiplication: The “usual” scalar multiplication, given in Definition CVSM [68].

Does this set with these operations fulfill the ten axioms? Yes. And all we need todo is quote Theorem VSPCM [70]. That was easy. 4

Example VS.VSMThe vector space of matrices, Mmn

Set: Mmn, the set of all matrices of size m×n and entries from C, Example VS.VSM [186].Vector Addition: The “usual” addition, given in Definition MA [122].Scalar Multiplication: The “usual” scalar multiplication, given in Definition SMM [122].

Does this set with these operations fulfill the ten axioms? Yes. And all we need todo is quote Theorem VSPM [123]. Another easy one. 4

So, the set of all matrices of a fixed size forms a vector space. That entitles us to call amatrix a vector, since a matrix is an element of a vector space. This could lead to someconfusion, but it is not too great a danger. But it is worth comment.

The previous two examples may be less than satisfying. We made all the relevantdefinitions long ago. And the required verifications were all handled by quoting oldtheorems. However, it is important to consider these two examples first. We have beenstudying vectors and matrices carefully (Chapter V [65], Chapter M [121]), and bothobjects, along with their operations, have certain properties in common, as you may havenoticed in comparing Theorem VSPCM [70] with Theorem VSPM [123]. Indeed, it isthese two theorems that motivate us to formulate the abstract definition of a vector space,Definition VS [184]. Now, should we prove some general theorems about vector spaces(as we will shortly in Subsection VS.VSP [190]), we can instantly apply the conclusionsto both Cm and Mmn. Notice too how we have taken six definitions and two theoremsand reduced them down to two examples . With greater generalization and abstractionour old ideas get downgraded in stature.

Let us look at some more examples, now considering some new vector spaces.

Example VS.VSPThe vector space of polynomials, Pn

Set: Pn, the set of all polynomials of degree n or less in the variable x with coefficientsfrom C.Vector Addition:

(a0 + a1x + a2x2 + · · ·+ anx

n) + (b0 + b1x + b2x2 + · · ·+ bnx

n) =

(a0 + b0) + (a1 + b1)x + (a2 + b2)x2 + · · ·+ (an + bn)xn

Scalar Multiplication:

α(a0 + a1x + a2x2 + · · ·+ anx

n) = (αa0) + (αa1)x + (αa2)x2 + · · ·+ (αan)xn

This set, with these operations, will fulfill the ten axioms, though we will not workall the details here. However, we will make a few comments and prove one of the axioms.

Page 219: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.EVS Examples of Vector Spaces 187

First, the zero vector is what you might expect, and you can check that it has the requiredproperty.

0 = 0 + 0x + 0x2 + · · ·+ 0xn

The additive inverse is also no surprise, though consider how we have chosen to write it.

−(a0 + a1x + a2x

2 + · · ·+ anxn)

= (−a0) + (−a1)x + (−a2)x2 + · · ·+ (−an)xn

Now lets prove the associativity of vector addition. This is a bit tedious, though necessary.Throughout, the plus sign (“+”) does triple-duty. You might ask yourself what each plussign represents as you work through this proof.

u + (v + w)

= (a0 + a1x + a2x2 + · · ·+ anx

n) +((b0 + b1x + b2x

2 + · · ·+ bnxn) + (c0 + c1x + c2x

2 + · · ·+ cnxn))

= (a0 + a1x + a2x2 + · · ·+ anx

n) + ((b0 + c0) + (b1 + c1)x + (b2 + c2)x2 + · · ·+ (bn + cn)xn)

= (a0 + (b0 + c0)) + (a1 + (b1 + c1))x + (a2 + (b2 + c2))x2 + · · ·+ (an + (bn + cn))xn

= ((a0 + b0) + c0) + ((a1 + b1) + c1)x + ((a2 + b2) + c2)x2 + · · ·+ ((an + bn) + cn)xn

= ((a0 + b0) + (a1 + b1)x + (a2 + b2)x2 + · · ·+ (an + bn)xn) + (c0 + c1x + c2x

2 + · · ·+ cnxn)

=((a0 + b0) + (a1 + b1)x + (a2 + b2)x

2 + · · ·+ (an + bn)xn))

+ (c0 + c1x + c2x2 + · · ·+ cnx

n)

= (u + v) + w

Notice how it is the application of the associativity of the (old) addition of complexnumbers in the middle of this chain of equalities that makes the whole proof happen.The remainder is successive applications of our (new) definition of vector (polynomial)addition. Proving the remainder of the ten axioms is similar in style, and tedium. Youmight try proving the commutativity of vector addition, or one of the distributivityaxioms. 4

Example VS.VSISThe vector space of infinite sequencesSet: C∞ = {(c0, c1, c2, c3, . . .) | ci ∈ C}.Vector Addition: (c0, c1, c2, . . .) + (d0, d1, d2, . . .) = (c0 + d0, c1 + d1, c2 + d2, . . .)Scalar Multiplication: α(c0, c1, c2, c3, . . .) = (αc0, αc1, αc2, αc3, . . .).

This should remind you of the vector space Cm, though now our lists of scalars arewritten horizontally with commas as delimiters and they are allowed to be infinite inlength. What does the zero vector look like? Additive inverses? Can you prove theassociativity of vector addition? 4

Example VS.VSFThe vector space of functionsSet: F = {f | f : C → C}.Vector Addition: f + g is the function defined by (f + g)(x) = f(x) + g(x).Scalar Multiplication: αf is the function defined by (αf)(x) = αf(x).

So this is the set of all functions of one variable that take a complex number to acomplex number. You might have studied functions of one variable that take a realnumber to a real number, and that might be a more natural set to study. But sincewe are allowing our scalars to be complex numbers, we need to expand the domain and

Page 220: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.EVS Examples of Vector Spaces 188

range of our functions also. Study carefully how the definitions of the operation aremade, and think about the different uses of “+” and juxtaposition. As an example ofwhat is required when verifying that this is a vector space, consider that the zero vectoris the function z whose definition is z(x) = 0 for every input x.

While vector spaces of functions are very important in mathematics and physics, wewill not devote them much more attention.

Here’s a unique example.

Example VS.VSSThe singleton vector spaceSet: Z = {z}.Vector Addition: z + z = z.Scalar Multiplication: αz = z.

This should look pretty wild. First, just what is z? Column vector, matrix, polyno-mial, sequence, function? Mineral, plant, or animal? We aren’t saying! z just is. Ouronly concern is if this set, along with the definitions of two operations, fulfill the tenaxioms. Let’s check associativity of vector addition. For all u, v, w ∈ Z,

u + (v + w) = z + (z + z)

= z + z

= z

= z + z

= (z + z) + z

= (u + v) + w

What is the zero vector in this vector space? With only one element in the set, we do nothave much choice. Is z = 0? It appears that z behaves like the zero vector should, so itgets the title. Maybe now the definition of this vector space does not seem so bizarre.4

Perhaps some of the above definitions and verifications seem obvious or like splittinghairs, but the next example should convince you that they are necessary. We will studythis one carefully. Ready? Leave your preconceptions behind.

Example VS.CVSThe crazy vector spaceSet: C = {(x1, x2) | x1, x2 ∈ C}.Vector Addition: (x1, x2) + (y1, y2) = (x1 + y1 + 1, x2 + y2 + 1).Scalar Multiplication: α(x1, x2) = (αx1 + α− 1, αx2 + α− 1).

Now, the first thing I hear you say is “You can’t do that!” And my response is, “Ohyes, I can!” I am free to define my set and my operations any way I please. They maynot look natural, or even useful, but we will now verify that they provide us with anotherexample of a vector space. And that is enough. If you are adventurous, you might tryfirst checking some of the axioms yourself. What is the zero vector? Additive inverses?Can you prove associativity? Here we go.

Additive and scalar closure: The result of each operation is a pair of complex numbers,so these two first axioms are fulfilled.

Page 221: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.EVS Examples of Vector Spaces 189

Commutativity:

u + v = (x1, x2) + (y1, y2) = (x1 + y1 + 1, x2 + y2 + 1)

= (y1 + x1 + 1, y2 + x2 + 1) = (y1, y2) + (x1, x2)

= v + u

Associativity:

u + (v + w) = (x1, x2) + ((y1, y2) + (z1, z2))

= (x1, x2) + (y1 + z1 + 1, y2 + z2 + 1)

= (x1 + (y1 + z1 + 1) + 1, x2 + (y2 + z2 + 1) + 1)

= (x1 + y1 + z1 + 2, x2 + y2 + z2 + 2)

= ((x1 + y1 + 1) + z1 + 1, (x2 + y2 + 1) + z2 + 1)

= (x1 + y1 + 1, x2 + y2 + 1) + (z1, z2)

= ((x1, x2) + (y1, y2)) + (z1, z2)

= (u + v) + w

Zero Vector: The zero vector is . . .0 = (−1, −1). Now I hear you say, “No, no, thatcan’t be, it must be (0, 0)!” Indulge me for a moment and let us check my proposal.

u + 0 = (x1, x2) + (−1, −1) = (x1 + (−1) + 1, x2 + (−1) + 1) = (x1, x2) = u

Feeling better? Or worse?

Additive Inverse: For each vector, u, we must locate an inverse, −u. Here it is,−(x1, x2) = (−x1 − 2, −x2 − 2). As odd as it may look, I hope you are withholdingjudgment. Check:

u+(−u) = (x1, x2)+(−x1−2, −x2−2) = (x1+(−x1−2)+1, −x2+(x2−2)+1) = (−1, −1) = 0

Associativity of Scalar Multiplication:

α(βu) = α(β(x1, x2))

= α(βx1 + β − 1, βx2 + β − 1))

= (α(βx1 + β − 1) + α− 1, α(βx2 + β − 1) + α− 1))

= ((αβx1 + αβ − α) + α− 1, (αβx2 + αβ − α) + α− 1))

= (αβx1 + αβ − 1, αβx2 + αβ − 1))

= (αβ)(x1, x2)

= (αβ)u

Distributivity Across Vector Addition: If you have hung on so far, here’s where it gets

Page 222: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.VSP Vector Space Properties 190

even wilder. In the next two axioms we mix and mash the two operations.

α(u + v) = α ((x1, x2) + (y1, y2))

= α(x1 + y1 + 1, x2 + y2 + 1)

= (α(x1 + y1 + 1) + α− 1, α(x2 + y2 + 1) + α− 1)

= (αx1 + αy1 + α + α− 1, αx2 + αy2 + α + α− 1)

= (αx1 + α− 1 + αy1 + α− 1 + 1, αx2 + α− 1 + αy2 + α− 1 + 1)

= ((αx1 + α− 1) + (αy1 + α− 1) + 1, (αx2 + α− 1) + (αy2 + α− 1) + 1)

= (αx1 + α− 1, αx2 + α− 1) + (αy1 + α− 1, αy2 + α− 1)

= α(x1, x2) + α(y1, y2)

= αu + αv

Distributivity Across Addition:

(α + β)u = (α + β)(x1, x2)

= ((α + β)x1 + (α + β)− 1, (α + β)x2 + (α + β)− 1)

= (αx1 + βx1 + α + β − 1, αx2 + βx2 + α + β − 1)

= (αx1 + α− 1 + βx1 + β − 1 + 1, αx2 + α− 1 + βx2 + β − 1 + 1)

= ((αx1 + α− 1) + (βx1 + β − 1) + 1, (αx2 + α− 1) + (βx2 + β − 1) + 1)

= (αx1 + α− 1, αx2 + α− 1) + (βx1 + β − 1, βx2 + β − 1)

= α(x1, x2) + β(x1, x2)

= αu + βu

Multiplication by 1: After all that, this one is easy, but no less pleasing.

1u = 1(x1, x2) = (x1 + 1− 1, x2 + 1− 1) = (x1, x2) = u

That’s it, C is a vector space, as crazy as that may seem.Notice that in the case of the zero vector and additive inverses, we only had to propose

possibilities and then verify that they were the correct choices. You might try to discoverhow you would arrive at these choices, though you should understand why the processof discovering them is not a necessary component of the proof itself. 4

Subsection VSPVector Space Properties

Subsection VS.EVS [185] has provided us with an abundance of examples of vector spaces,most of them containing useful and interesting mathematical objects along with naturaloperations. In this subsection we will prove some general properties of vector spaces.Some of these results will again seem obvious, but it is important to understand why itis necessary to state and prove them. A typical hypothesis will be “Let V be a vectorspace.” From this we may assume the ten axioms, and nothing more. Its like startingover, as we learn about what can happen in this new algebra we are learning. But thepower of this careful approach is that we can apply these theorems to any vector space

Page 223: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.VSP Vector Space Properties 191

we encounter, those in the previous examples, or new ones we have not contemplated. Orperhaps new ones that nobody has ever contemplated. We will illustrate some of theseresults with examples from the crazy vector space (Example VS.CVS [188]), but mostlywe are stating theorems and doing proofs. These proofs do not get too involved, but arenot trivial either, so these are good theorems to try proving yourself before you studythe proof given here. (See Technique P [125].)

First we show that there is just one zero vector. Notice that the axioms only requirethere to be one, and say nothing about there being more. That is because we can usethe axioms to learn that there can never be more than one. To require that this extracondition be stated in the axioms would make them more complicated than they needto be.

Theorem ZVUZero Vector is UniqueSuppose that V is a vector space. The zero vector, 0, is unique. �

Proof To prove uniqueness, a standard technique is to suppose the existence of twoobjects (Technique U [59]). So let 01 and 02 be two zero vectors in V . Then

01 = 01 + 02 Property of the zero vector 02

= 02 Property of the zero vector 01

Notice that we have implicitly used the commutativity of vector addition so that we canapply the defining property of a zero vector on either side of the addition. This provesthe uniqueness since the two zero vectors are really the same. �

Theorem AIUAdditive Inverses are UniqueSuppose that V is a vector space. For each u ∈ V , the additive inverse, −u, is unique.�

Proof To prove uniqueness, a standard technique is to suppose the existence of twoobjects (Technique U [59]). So let −u1 and −u2 be two additive inverses for u. Then

−u1 = −u1 + 0 Property of the zero vector

= −u1 + (u +−u2) −u2 is an additive inverse

= (−u1 + u) +−u2 Associativity of vector addition

= 0 +−u2 −u1 is an additive inverse

= −u2 Property of the zero vector

So the two additive inverses are really the same. �

As obvious as the next theorem appears, it is not guaranteed that the zero scalar, scalarmultiplication and the zero vector all interact this way. Until we have proved it, anyway.

Theorem ZSSMZero Scalar in Scalar MultiplicationSuppose that V is a vector space and u ∈ V . Then 0u = 0. �

Page 224: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.VSP Vector Space Properties 192

Proof

0u = (0 + 0)u

0u = 0u + 0u Distributivity across addition

0 is a scalar, u is a vector, so scalar closure means 0u is again a vector. As such, it hasan additive inverse.

−(0u) + 0u = −(0u) + (0u + 0u)

−(0u) + 0u = (−(0u) + 0u) + 0u Associativity of vector addition

0 = 0 + 0u Additive inverses

0 = 0u Property of zero vector �

Theorem ZVSMZero Vector in Scalar MultiplicationSuppose that V is a vector space and α ∈ C. Then α0 = 0. �

Proof

α0 = α(0 + 0)

α0 = α0 + α0 Distributivity across vector addition

α is a scalar, 0 is a vector, so scalar closure means α0 is again a vector. As such, it hasan additive inverse.

−(α0) + α0 = −(α0) + (α0 + α0)

−(α0) + α0 = (−(α0) + α0) + α0 Associativity of vector addition

0 = 0 + α0 Additive inverses

0 = α0 Property of the zero vector �

Here’s another one that sure looks obvious. But understand that we have chosen to usecertain notation because it makes the theorem’s conclusion look so nice. The theoremis not true because the notation looks so good, it still needs a proof. If we had reallywanted to make this point, we might have defined the additive inverse of u as u]. Thenwe would have written the defining property as u+u] = 0. This theorem would becomeu] = (−1)u. Not really quite as pretty, is it?

Theorem AISMAdditive Inverses from Scalar MultiplicationSuppose that V is a vector space and u ∈ V . Then −u = (−1)u. �

Proof

(−1)u + u = (−1)u + 1u Scalar multiplication with 1

= ((−1) + 1)u Distributivity across addition

= 0u

= 0 Theorem ZSSM [191]

(−1)u is a vector in V by scalar closure, and this equation says it acts as the additiveinverse of u. Since the additive inverse of a vector is unique (Theorem AIU [191]), (−1)umust be the additive inverse of u. In other words, −u = (−1)u. �

Page 225: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.VSP Vector Space Properties 193

Because of this theorem, we will now write linear combinations like 6u1 + (−4)u2

as 6u1 − 4u2, even though we have not formally defined an operation called vectorsubtraction.

Example VS.PCVSProperties for the Crazy Vector SpaceSeveral of the above theorems have interesting demonstrations when applied to the crazyvector space, C (Example VS.CVS [188]). We are not proving anything new here, orlearning anything we did not know already about C. It is just fun to see how thesegeneral theorems apply in a specific instance. For most of our examples, the applicationsare obvious or trivial, but not with C.

Suppose u ∈ C.Then by Theorem ZSSM [191],

0u = 0(x1, x2) = (0x1 + 0− 1, 0x2 + 0− 1) = (−1,−1) = 0

By Theorem ZVSM [192],

α0 = α(−1, −1) = (α(−1)+α−1, α(−1)+α−1) = (−α+α−1, −α+α−1) = (−1, −1)

By Theorem AISM [192],

(−1)u = (−1)(x1, x2) = ((−1)x1+(−1)−1, (−1)x2+(−1)−1) = (−x1−2, −x2−2) = −u4

Our next theorem is a bit different from several of the others in the list. Rather thanmaking a declaration (“the zero vector is unique”) it is an implication (“if. . . , then. . . ”)and so can be used in proofs to move from one statement to another.

Theorem SMEZVScalar Multiplication Equals the Zero VectorSuppose that V is a vector space and α ∈ C. Then if αu = 0, then either α = 0 oru = 0. �

Proof We prove this theorem by breaking up the analysis into two cases. The first seemstoo trivial, and it is, but the logic of the argument is still legitimate.

Case 1. Suppose α = 0. In this case our conclusion is true (the first part of theeither/or is satisfied) and we are done. That was easy.

Case 2. Suppose α 6= 0.

αu = 0 Hypothesis

1

α(αu) =

1

α0 α 6= 0(

1

αα

)u =

1

α0 Associativity of scalar multiplication

1u = 0 Theorem ZVSM [192]

u = 0 Scalar multiplication with 1

So in this case, the conclusion is true (the second part of the either/or) and we are donesince the conclusion was true in each of the cases. �

Page 226: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.VSP Vector Space Properties 194

The next three theorems give us cancellation properties. The two concerned with scalarmultiplication are intimately connected with Theorem SMEZV [193]. All three are im-plications. So we will prove each once, here and now, and then we can apply them atwill in the future, saving several steps in a proof whenever we do.

Theorem VACVector Addition CancellationSuppose that V is a vector space, and u, v, w ∈ V . If w + u = w + v, then u = v. �

Proof

w + u = w + v

−w + (w + u) = −w + (w + v) Additive inverses exist

(−w + w) + u = (−w + w) + v Associativity of vector addition

0 + u = 0 + v Additive inverses

u = v Property of the zero vector

Theorem CSSMCanceling Scalars in Scalar MultiplicationSuppose V is a vector space, u, v ∈ V and α is a nonzero scalar from C. If αu = αv,then u = v. �

Proof

αu = αv Hypothesis

1

α(αu) =

1

α(αv) α 6= 0(

1

αα

)u =

(1

αα

)v Associativity of scalar multiplication

1u = 1v

u = v Scalar multiplication with 1 �

Theorem CVSMCanceling Vectors in Scalar MultiplicationSuppose V is a vector space, u 6= 0 is a vector in V and α, β ∈ C. If αu = βu, thenα = β. �

Proof

αu = βu Hypothesis

−(αu) + αu = −(αu) + βu Additive inverses exist

0 = −(αu) + βu Additive inverse

0 = (−1)(αu) + βu Theorem AISM [192]

0 = ((−1)α)u + βu Associativity of scalar multiplication

0 = (−α) + β)u Distributivity across addition

Page 227: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VS.RD Recycling Definitions 195

Applying Theorem SMEZV [193], along with the hypothesis that u 6= 0 gives

−α + β = 0

α = β �

So with these three theorems in hand, we can return to our practice of “slashing” outparts of an equation, so long as we are careful about not canceling a scalar that mightpossibly be zero, or canceling a vector in a scalar multiplication that might be the zerovector.

Subsection RDRecycling Definitions

When we say that V is a vector space, we know we have a set of objects, but we also knowwe have been provided with two operations. One combines two vectors and produces avector, the other takes a scalar and a vector, producing a vector as the result. So ifu1, u2, u3 ∈ V then an expression like

5u1 + 7u2 − 13u3

would be unambiguous in any of the vector spaces we have discussed in this section. Andthe resulting object would be another vector in the vector space. If you were temptedto call the above expression a linear combination, you would be right. Four of the defi-nitions that were central to our discussions in Chapter V [65] were stated in the contextof vectors being column vectors, but were purposely kept broad enough that they couldbe applied in the context of any vector space. They only rely on the presence of scalars,vectors, vector addition and scalar multiplication to make sense. We will restate themshortly, unchanged, except that their titles and acronyms no longer refer to column vec-tors, and the hypothesis of being in a vector space has been added. Take the timenow to look forward and review each one, and begin to form some connections to whatwe have done earlier and what we will be doing in subsequent sections and chapters.(See Definition LCCV [72] and Definition LC [202], Definition SSCV [87] and Defini-tion SS [203], Definition RLDCV [96] and Definition RLD [212], Definition LICV [96]and Definition LI [212].)

Page 228: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section S Subspaces 196

Section S

Subspaces

A subspace is a vector space that is contained within another vector space. So everysubspace is a vector space in its own right, but it is also defined relative to some other(larger) vector space. We will discover shortly that we are already familiar with a widevariety of subspaces from previous sections. Here’s the definition.

Definition SSubspaceSuppose that V and W are two vector spaces that have identical definitions of vectoraddition and scalar multiplication, and that W is a subset of V , W ⊆ V . Then W is asubspace of V . �

Lets look at an example of a vector space inside another vector space.

Example S.SC3A subspace of C3

We know that C3 is a vector space (Example VS.VSCM [186]). Consider the subset,

W =

x1

x2

x3

∣∣∣∣∣∣ 2x1 − 5x2 + 7x3 = 0

It is clear that W ⊆ V , since the objects in W are column vectors of size 3. But is W avector space? Does it satisfy the ten axioms of Definition VS [184] when we use the same

operations? That is the main question. Suppose x =

x1

x2

x3

and y =

y1

y2

y3

are vectors

from W . Then we know that these vectors cannot be totally arbitrary, they must havegained membership in W by virtue of meeting the membership test. For example, weknow that x must satisfy 2x1 − 5x2 + 7x3 = 0 while y must satisfy 2y1 − 5y2 + 7y3 = 0.Our first axiom asks the question, is x + y ∈ W? When our set of vectors was C3, thiswas an easy question to answer. Now it is not so obvious. Notice first that

x + y =

x1

x2

x3

+

y1

y2

y3

=

x1 + y1

x2 + y2

x3 + y3

and we can test this vector for membership in W as follows,

2(x1 + y1)− 5(x2 + y2) + 7(x3 + y3)

= 2x1 + 2y1 − 5x2 − 5y2 + 7x3 + 7y3

= (2x1 − 5x2 + 7x3) + (2y1 − 5y2 + 7y3)

= 0 + 0 x ∈ W, y ∈ W

= 0

and by this computation we see that x + y ∈ W . One axiom down, nine to go.

Page 229: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TS Testing Subspaces 197

If α is a scalar and x ∈ W , is it always true that αx ∈ W? Again, the answer is notas obvious as it was when our set of vectors was all of C3. Let’s see.

αx = α

x1

x2

x3

=

αx1

αx2

αx3

and we can test this vector for membership in W with

2(αx1)− 5(αx2) + 7(αx3)

= α(2x1 − 5x2 + 7x3)

= α0 x ∈ W

= 0

and we see that indeed αx ∈ W . Always.If W has a zero vector, it will be unique (Theorem ZVU [191]). The zero vector for

C3 should also perform the required duties when added to elements of W . So the likelycandidate for a zero vector in W is the same zero vector that we know C3 has. You can

check that 0 =

000

is a zero vector in W too.

With a zero vector, we can now ask about additive inverses. As you might suspect,the natural candidate for an additive inverse in W is the same as the additive inversefrom C3. However, we must insure that these additive inverses actually are elements ofW . Given x ∈ W , is −x ∈ W?

−x =

−x1

−x2

−x3

and we can test this vector for membership in W with

2(−x1)− 5(−x2) + 7(−x3)

= −(2x1 − 5x2 + 7x3)

= −0 x ∈ W

= 0

and we now believe that −x ∈ W .Is the vector addition in W commutative? Is x + y = y + x? Of course! Nothing

about restricting the scope of our set of vectors will prevent the operation from still beingcommutative. Indeed, the remaining five axioms are unaffected by the transition to asmaller set of vectors, and so remain true. That was convenient.

So W satisfies all ten axioms, is therefore a vector space, and thus earns the title ofbeing a subspace of C3. 4

Subsection TSTesting Subspaces

In Example S.SC3 [196] we proceeded through all ten of the vector space axioms beforebelieving that a subset was a subspace. But six of the axioms were easy to prove, and we

Page 230: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TS Testing Subspaces 198

can lean on some of the properties of the vector space (the superset) to make the otherfour easier. Here is a theorem that will make it easier to test if a subset is a vector space.A shortcut if there ever was one.

Theorem TSSTesting Subsets for SubspacesSuppose that V is a vector space and W is a subset of V , W ⊆ V . Endow W with thesame operations as V . Then W is a subspace if and only if three conditions are met

1. W is non-empty, W 6= ∅.

2. Whenever x ∈ W and y ∈ W , then x + y ∈ W .

3. Whenever α ∈ C and x ∈ W , then αx ∈ W . �

Proof (⇒) We have the hypothesis that W is a subspace, so by Definition VS [184] weknow that W contains a zero vector. This is enough to show that W 6= ∅. Also, since Wis a vector space it satisfies the additive and scalar multiplication closure axioms, and soexactly meets the second and third conditions. If that was easy, the the other directionmight require a bit more work.

(⇐) We have three properties for our hypothesis, and from this we should concludethat W has the ten defining properties of a vector space. The second and third conditionsof our hypothesis are exactly the first two closure axioms. The commutativity, associa-tivity (twice), distributivity (twice) and multiplication with 1 axioms that hold for Vwill continue to hold in W since passing to a subset leaves their statements unchanged.Eight down, two to go.

Since W is non-empty, we can choose some vector z ∈ W . Then by scalar closure (thethird part of our hypothesis), we know (−1)z ∈ W . By Theorem AISM [192] (−1)z = −z.Now apply the additive inverse axiom for V and then additive closure (the second partof our hypothesis) to see that 0 = z + (−z) ∈ W . So W contain the zero vector from V .Since this vector performs the required duties of a zero vector in V , it will continue inthat role as an element of W . So W has a zero vector. One axiom left.

Suppose x ∈ W . Then by scalar closure (the third part of our hypothesis), we knowthat (−1)x ∈ W . By Theorem AISM [192] (−1)x = −x, so together these statementsshow us that −x ∈ W . −x is the additive inverse of x in V , but will continue in this rolewhen viewed as element of the subset W . So every element of W has an additive inversethat is an element of W .

Three conditions, plus being a subset, gets us all ten axioms. Fabulous! �

This theorem is can be paraphrased by saying that a subspace is “a non-empty subset(of a vector space) that is closed under vector addition and scalar multiplication.”

You might want to go back and rework Example S.SC3 [196] in light of this result,perhaps seeing where we can now economize or where the work done in the examplemirrored the proof and where it did not. We will press on and apply this theorem in aslightly more abstract setting.

Example S.SP4A subspace of P4

P4 is the vector space of polynomials with degree at most 4 (Example VS.VSP [186]).Define a subset W as

W = {p(x) | p ∈ P4, p(2) = 0}

Page 231: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TS Testing Subspaces 199

so W is the collection of those polynomials (with degree 4 or less) whose graphs crossthe x-axis at x = 2. Whenever we encounter a new set it is a good idea to gain a betterunderstanding of the set by finding a few elements in the set, and a few outside it. Forexample x2 − x− 2 ∈ W , while x4 + x3 − 7 6∈ W .

Is W nonempty? Yes, x− 2 ∈ W .Additive closure? Suppose p ∈ W and q ∈ W . Is p + q ∈ W? p and q are not totally

arbitrary, we know that p(2) = 0 and q(2) = 0. Then we can check p + q for membershipin W ,

(p + q)(2) = p(2) + q(2) Addition in P4

= 0 + 0 p ∈ W, q ∈ W

= 0

so we see that p + q qualifies for membership in W .Scalar multiplication closure? Suppose that α ∈ C and p ∈ W . Then we know that

p(2) = 0. Testing αp for membership,

(αp)(2) = αp(2) Scalar multiplication in P4

= α0 p ∈ W

= 0 4

so αp ∈ W .We have shown that W meets the three conditions of Theorem TSS [198] and so

qualifies as a subspace of P4. Notice that by Definition S [196] we now know that W isalso a vector space. So all the axioms of a vector space (Definition VS [184]) and thetheorems of Section VS [184] apply in full.

Much of the power of Theorem TSS [198] is that we can easily establish new vector spacesif we can locate them as subsets of other vector spaces, such as the ones presented inSection VS [184].

It can be as instructive to consider some subsets that are not subspaces. Since The-orem TSS [198] is an equivalence (see Technique E [40]) we can be assured that a subsetis not a subspace if it violates one of the three conditions, and in any example of interestthis will not be the “non-empty” condition. However, since a subspace has to be a vectorspace in its own right, we can also search for a violation of any one of the ten definingaxioms in Definition VS [184]. Notice also that a violation need only be for a specificvector or pair of vectors.

Example S.NSC2ZA non-subspace in C2, zero vectorConsider the subset W below as a candidate for being a subspace of C2

W =

{[x1

x2

] ∣∣∣∣ 3x1 − 5x2 = 12

}The zero vector of C2, 0 =

[00

]will need to be the zero vector in W also. However,

0 6∈ W since 3(0) − 5(0) = 0 6= 12. So W has no zero vector and fails the zero vectoraxiom of Definition VS [184]. This subspace also fails to be closed under addition andscalar multiplication. Can you find examples of this? 4

Page 232: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TS Testing Subspaces 200

Example S.NSC2AA non-subspace in C2, additive closureConsider the subset X below as a candidate for being a subspace of C2

X =

{[x1

x2

] ∣∣∣∣ x1x2 = 0

}You can check that 0 ∈ X, so the approach of the last example will not get us anywhere.

However, notice that x =

[10

]∈ X and y =

[01

]∈ X. Yet

x + y =

[10

]+

[01

]=

[11

]6∈ X

So X fails the additive closure requirement of either Definition VS [184] or Theorem TSS [198],and is therefore not a subspace. 4

Example S.NSC2SA non-subspace in C2, scalar multiplication closureConsider the subset Y below as a candidate for being a subspace of C2

Y =

{[x1

x2

] ∣∣∣∣ x1 ∈ Z, x2 ∈ Z}

Z is the set of integers, so we are only allowing “whole numbers” as the constituents ofour vectors. Now, 0 ∈ Y , and additive closure also holds (can you prove these claims?).

So we will have to try something different. Note that α = 12∈ C and

[23

]∈ Y , but

αx =1

2

[23

]=

[132

]6∈ Y

So Y fails the scalar multiplication closure requirement of either Definition VS [184] orTheorem TSS [198], and is therefore not a subspace. 4

There are two examples of subspaces that are trivial. Suppose that V is any vector space.Then V is a subset of itself and is a vector space. By Definition S [196], V qualifies as asubspace of itself. The set containing just the zero vector Z = {0} is also a subspace ascan be seen by applying Theorem TSS [198] or by simple modifications of the techniqueshinted at in Example VS.VSS [188]. Since these subspaces are so obvious (and thereforenot too interesting) we will refer to them as being trivial.

Definition TSTrivial SubspacesGiven the vector space V , the subspaces V and {0} are each called a trivial subspace.�

We can also use Theorem TSS [198] to prove more general statements about subspaces,as illustrated in the next theorem.

Theorem NSMSNull Space of a Matrix is a SubspaceSuppose that A is an m × n matrix. Then the null space of A, N (A), is a subspace ofCn. �

Page 233: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TS Testing Subspaces 201

Proof We will examine the three requirements of Theorem TSS [198]. Recall thatN (A) = {x ∈ Cn | Ax = 0}.

First, 0 ∈ N (A), which can be inferred as a consequence of Theorem HSC [50]. SoN (A) 6= ∅.

Second, check additive closure by supposing that x ∈ N (A) and y ∈ N (A). So weknow a little something about x and y: Ax = 0 and Ay = 0, and that is all we know.Question: is x + y ∈ N (A)? Let’s check.

A(x + y) = Ax + Ay Theorem MMDAA [157]

= 0 + 0 x ∈ N (A) , y ∈ N (A)

= 0 Theorem VSPCM [70]

So, yes, x + y qualifies for membership in N (A).

Third, check scalar multiplication closure by supposing that α ∈ C and x ∈ N (A).So we know a little something about x: Ax = 0, and that is all we know. Question: isαx ∈ N (A)? Let’s check.

A(αx) = α(Ax) Theorem MMSMM [157]

= α0 x ∈ N (A)

= 0 Theorem ZVSM [192]

So, yes, αx qualifies for membership in N (A).

Having met the three conditions in Theorem TSS [198] we can now say that the nullspace of a matrix is a subspace (and hence a vector space in its own right!). �

Here is an example where we can exercise Theorem NSMS [200].

Example S.RSNSRecasting a subspace as a null spaceConsider the subset of C5 defined as

W =

x1

x2

x3

x4

x5

∣∣∣∣∣∣∣∣∣∣

3x1 + x2 − 5x3 + 7x4 + x5 = 0,4x1 + 6x2 + 3x3 − 6x4 − 5x5 = 0,−2x1 + 4x2 + 7x4 + x5 = 0

It is possible to show that W is a subspace of C5 by checking the three conditions ofTheorem TSS [198] directly, but it will get tedious rather quickly. Instead, give W afresh look and notice that it is a set of solutions to a homogeneous system of equations.Define the matrix

A =

3 1 −5 7 14 6 3 −6 −5−2 4 0 7 1

and then recognize that W = N (A). By Theorem NSMS [200] we can immediately seethat W is a subspace. Boom! 4

Page 234: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TSS The Span of a Set 202

Subsection TSSThe Span of a Set

The span of a set of column vectors got a heavy workout in Chapter V [65] and Chap-ter M [121]. The definition of the span depended only on being able to formulate linearcombinations. In any of our more general vector spaces we always have a definition ofvector addition and of scalar multiplication. So we can build linear combinations andmanufacture spans. This subsection contains two definitions that are just mild variantsof definitions we have seen earlier for column vectors. If you haven’t already, comparethem with Definition LCCV [72] and Definition SSCV [87].

Definition LCLinear CombinationSuppose that V is a vector space. Given n vectors u1, u2, u3, . . . , un and n scalarsα1, α2, α3, . . . , αn, their linear combination is the vector

α1u1 + α2u2 + α3u3 + · · ·+ αnun. �

Example S.LCMA linear combination of matricesIn the vector space M23 of 2× 3 matrices, we have the vectors

x =

[1 3 −22 0 7

]y =

[3 −1 25 5 1

]z =

[4 2 −41 1 1

]and we can form linear combinations such as

2x + 4y + (−1)z = 2

[1 3 −22 0 7

]+ 4

[3 −1 25 5 1

]+ (−1)

[4 2 −41 1 1

]=

[2 6 −44 0 14

]+

[12 −4 820 20 4

]+

[−4 −2 4−1 −1 −1

]=

[10 0 823 19 17

]or,

4x− 2y + 3z = 4

[1 3 −22 0 7

]− 2

[3 −1 25 5 1

]+ 3

[4 2 −41 1 1

]=

[4 12 −88 0 28

]+

[−6 2 −4−10 −10 −2

]+

[12 6 −123 3 3

]=

[10 20 −241 −7 29

]4

When we realize that we can form linear combinations in any vector space, then it isnatural to revisit our definition of the span of a set, since it is the set of all possible linearcombinations of a set of vectors.

Page 235: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TSS The Span of a Set 203

Definition SSSpan of a SetSuppose that V is a vector space. Given a set of vectors S = {u1, u2, u3, . . . , ut},their span, Sp (S), is the set of all possible linear combinations of u1, u2, u3, . . . , ut.Symbolically,

Sp (S) = {α1u1 + α2u2 + α3u3 + · · ·+ αtut | αi ∈ C, 1 ≤ i ≤ t}

=

{t∑

i=1

αiui

∣∣∣∣∣ αi ∈ C, 1 ≤ i ≤ t

}�

Theorem SSSSpan of a Set is a SubspaceSuppose V is a vector space. Given a set of vectors S = {u1, u2, u3, . . . , ut} ⊆ S, theirspan, Sp (S), is a subspace. �

Proof We will verify the three conditions of Theorem TSS [198]. First, use Theo-rem ZSSM [191] repeatedly and the defining property of the zero vector from Defini-tion VS [184],

0u1 + 0u2 + 0u3 + · · ·+ 0ut+ = 0 + 0 + 0 + . . . + 0 = 0

so 0 ∈ Sp (S) since we have written 0 as a linear combination of the vectors in S.Second, suppose x ∈ Sp (S) and y ∈ Sp (S). Can we conclude that x + y ∈ Sp (S)?

What do we know about x and y by virtue of their membership in Sp (S)? There mustbe scalars from C, α1, α2, α3, . . . , αt and β1, β2, β3, . . . , βt so that

x = α1u1 + α2u2 + α3u3 + · · ·+ αtut

y = β1u1 + β2u2 + β3u3 + · · ·+ βtut

Then

x + y = α1u1 + α2u2 + α3u3 + · · ·+ αtut

+ β1u1 + β2u2 + β3u3 + · · ·+ βtut

= α1u1 + β1u1 + α2u2 + β2u2

+ α3u3 + β3u3 + · · ·+ αtut + βtut Associativity, Commutativity

= (α1 + β1)u1 + (α2 + β2)u2

+ (α3 + β3)u3 + · · ·+ (αt + βt)ut Distributivity

Since each αi + βi is again a scalar from C we have expressed the vector sum x + y as alinear combination of the vectors from S, and therefore we can say that x + y ∈ Sp (S).

Third, suppose α ∈ C and x ∈ Sp (S). Can we conclude that αx ∈ Sp (S)? What dowe know about x by virtue of its membership in Sp (S)? There must be scalars from C,α1, α2, α3, . . . , αt so that

x = α1u1 + α2u2 + α3u3 + · · ·+ αtut

Page 236: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TSS The Span of a Set 204

Then

αx = α (α1u1 + α2u2 + α3u3 + · · ·+ αtut)

= α(α1u1) + α(α2u2) + α(α3u3) + · · ·+ α(αtut) Distributivity

= (αα1)u1 + (αα2)u2 + (αα3)u3 + · · ·+ (ααt)ut Associativity

Since each ααi is again a scalar from C we have expressed the scalar multiple αx as alinear combination of the vectors from S, and therefore we can say that αx ∈ Sp (S).

With the three conditions of Theorem TSS [198] met, we can say that Sp (S) is asubspace (and so is a vector space). �

Example S.SSPSpan of a set of polynomialsIn Example S.SP4 [198] we proved that

W = {p(x) | p ∈ P4, p(2) = 0}

is a subspace of P4, the vector space of polynomials of degree at most 4. Since W is avector space itself, lets construct a span within W . First let

S ={x4 − 16, x2 − x− 2

}and verify that S is a subset of W by checking that each of these two polynomials hasx = 2 as a root. Now, if we define U = Sp (S), then Theorem SSS [203] tells us that U isa subspace of W . So quite quickly we have built a chain of subspaces, U inside W , andW inside P4.

Rather than dwell on how quickly we can build subspaces, lets try to gain a betterunderstanding of just what the span construction creates subspaces, in the context ofthis example. We can quickly build representative elements of U ,

3(x4 − 16) + 5(x2 − x− 2) = 3x4 + 5x2 − 5x− 58

and(−2)(x4 − 16) + 16(x2 − x− 2) = −2x4 + 16x2 − 16x

and each of these polynomials must be in W since it is closed under addition and scalarmultiplication. But you might check for yourself that both of these polynomials havex = 2 as a root.

I can tell you that y = x4+2x2−5x−14 is not in U , but would you believe me? A firstcheck shows that y does have x = 2 as a root, but that only shows that y ∈ W . Whatdoes y have to do to gain membership in U = Sp (S)? It must be a linear combinationof the vectors in S, x4 − 16 and x2 − x − 2. So lets suppose that y is such a linearcombination,

x4 + 2x2 − 5x− 14 = y = α1(x4 − 16) + α2(x

2 − x− 2)

= α1x4 − α1(16) + α2x

2 − α2x− α2(2)

= α1x4 + α2x

2 − α2x− α1(16)− α2(2)

Page 237: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TSS The Span of a Set 205

Now, equate coefficients to obtain the system of linear equations

1 = α1

2 = α2

−5 = −α2

−14 = −16α1 − 2α2

We could build an augmented matrix and discover that this is an inconsistent system, butthe second and third equations make it readily apparent that the system is inconsistent.The lack of a solution means the supposed linear combination is impossible and y 6∈ U .4

Lets again examine membership in a span.

Example S.SM32A subspace of M32

The set of all 3× 2 matrices forms a vector space when we use the operations of matrixaddition (Definition MA [122]) and scalar matrix multiplication (Definition SMM [122]),as was show in Example VS.VSM [186]. Consider the subset

S =

3 1

4 25 −5

,

1 12 −114 −1

,

3 −1−1 2−19 −11

,

4 21 −214 −2

,

3 1−4 0−17 7

and define a new subset of vectors W in M32 using the span (Definition SS [203]), W =Sp (S). So by Theorem SSS [203] we know that W is a subspace of M32. While W is aninfinite set, and this is a precise description, it would still be worthwhile to investigatewhether or not W contains certain elements.

First, is

y =

9 37 310 −11

in W? To answer this, we want to determine if y can be written as a linear combinationof the five matrices in S. Can we find scalars, α1, α2, α3, α4, α5 so that 9 3

7 310 −11

= α1

3 14 25 −5

+ α2

1 12 −114 −1

+ α3

3 −1−1 2−19 −11

+ α4

4 21 −214 −2

+ α5

3 1−4 0−17 7

=

3α1 + α2 + 3α3 + 4α4 + 3α5 α1 + α2 − α3 + 2α4 + α5

4α1 + 2α2 − α3 + α4 − 4α5 2α1 − α2 + 2α3 − 2α4

5α1 + 14α2 − 19α3 + 14α4 − 17α5 −5α1 − α2 − 11α3 − 2α4 + 7α5

Using our definition of matrix equality (Definition ME [121]) we can translate this state-

Page 238: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.TSS The Span of a Set 206

ment into six equations in the five unknowns,

3α1 + α2 + 3α3 + 4α4 + 3α5 = 9

α1 + α2 − α3 + 2α4 + α5 = 3

4α1 + 2α2 − α3 + α4 − 4α5 = 7

2α1 − α2 + 2α3 − 2α4 = 3

5α1 + 14α2 − 19α3 + 14α4 − 17α5 = 10

−5α1 − α2 − 11α3 − 2α4 + 7α5 = −11

This is a linear system of equations, which we can represent with an augmented matrixand row-reduce in search of solutions. The matrix that is row-equivalent to the augmentedmatrix is

1 0 0 0 58

2

0 1 0 0 −194

−1

0 0 1 0 −78

0

0 0 0 1 178

10 0 0 0 0 00 0 0 0 0 0

So we recognize that the system is consistent since there is no leading 1 in the final column(Theorem RCLS [41]), and compute n−r = 5−4 = 1 free variables (Corollary FVCS [42]).While there are infinitely many solutions, we are only in pursuit of a single solution, solets choose the free variable α5 = 0 for simplicity’s sake. Then we easily see that α1 = 2,α2 = −1, α3 = 0, α4 = 1. So the scalars α1 = 2, α2 = −1, α3 = 0, α4 = 1, α5 = 0will provide a linear combination of the elements of S that equals y, as we can verify bychecking, 9 3

7 310 −11

= 2

3 14 25 −5

+ (−1)

1 12 −114 −1

+ (1)

4 21 −214 −2

So with one particular linear combination in hand, we are convinced that y deserves tobe a member of W = Sp (S). Second, is

x =

2 13 14 −2

in W? To answer this, we want to determine if x can be written as a linear combinationof the five matrices in S. Can we find scalars, α1, α2, α3, α4, α5 so that2 1

3 14 −2

= α1

3 14 25 −5

+ α2

1 12 −114 −1

+ α3

3 −1−1 2−19 −11

+ α4

4 21 −214 −2

+ α5

3 1−4 0−17 7

=

3α1 + α2 + 3α3 + 4α4 + 3α5 α1 + α2 − α3 + 2α4 + α5

4α1 + 2α2 − α3 + α4 − 4α5 2α1 − α2 + 2α3 − 2α4

5α1 + 14α2 − 19α3 + 14α4 − 17α5 −5α1 − α2 − 11α3 − 2α4 + 7α5

Page 239: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.SC Subspace Constructions 207

Using our definition of matrix equality (Definition ME [121]) we can translate this state-ment into six equations in the five unknowns,

3α1 + α2 + 3α3 + 4α4 + 3α5 = 2

α1 + α2 − α3 + 2α4 + α5 = 1

4α1 + 2α2 − α3 + α4 − 4α5 = 3

2α1 − α2 + 2α3 − 2α4 = 1

5α1 + 14α2 − 19α3 + 14α4 − 17α5 = 4

−5α1 − α2 − 11α3 − 2α4 + 7α5 = −2

This is a linear system of equations, which we can represent with an augmented matrixand row-reduce in search of solutions. The matrix that is row-equivalent to the augmentedmatrix is

1 0 0 0 58

0

0 1 0 0 −388

0

0 0 1 0 −78

0

0 0 0 1 −178

0

0 0 0 0 0 10 0 0 0 0 0

With a leading 1 in the last column Theorem RCLS [41] tells us that the system isinconsistent. Therefore, there are no values for the scalars that will place x in W , andso we conclude that x 6∈ W . 4

Notice how Example S.SSP [204] and Example S.SM23 [??] contained questions aboutmembership in a span, but these questions quickly became questions about solutions toa system of linear equations. This will be a common theme going forward.

Subsection SCSubspace Constructions

Several of the subsets of vectors spaces that we worked with in Chapter M [??] are alsosubspaces — they are closed under vector addition and scalar multiplication in Cm.

Theorem RMSRange of a Matrix is a SubspaceSuppose that A is an m× n matrix. Then R (A) is a subspace of Cm. �

Proof Definition RM [127] shows us that R (A) is a subset of Cm, and that it is definedas the span of a set of vectors from Cm (the columns of the matrix). Since R (A) is aspan, Theorem SSS [203] says it is a subspace. �

That was easy! Notice that we could have used this same approach to prove that the nullspace is a subspace, since Theorem SSNS [90] provided a description of the null spaceof a matrix as the span of a set of vectors. However, I much prefer the current proof ofTheorem NSMS [200]. Speaking of easy, here is a very easy theorem that exposes anotherof our constructions as creating subspaces.

Page 240: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.SC Subspace Constructions 208

Theorem RSMSRow Space of a Matrix is a SubspaceSuppose that A is an m× n matrix. Then rs (A) is a subspace of Cn. �

Proof Theorem RMRST [145] applied to At, along with Theorem TASM [126], yields

rs (A) = rs((

At)t)

= R(At)

so the row space is equal to the range of a matrix, which Theorem RMS [207] says is asubspace. �

So the span of a set of vectors, and the null space, range, and row space of a matrix are allsubspaces, and hence are all vector spaces, meaning they have all the properties detailedin Section VS [184]. We have worked with these objects as just sets in Chapter V [65]and Chapter M [121], but now we understand that they have much more structure. Inparticular, being closed under vector addition and scalar multiplication means a subspaceis also closed under linear combinations.

Page 241: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.EXC Exercises 209

Subsection EXCExercises

C20 Contributed by Robert BeezerWorking within the vector space P3 of polynomials of degree 3 or less, determine if

p(x) = x3 + 6x + 4 is in the subspace W below.

W = Sp({

x3 + x2 + x, x3 + 2x− 6, x2 − 5})

Solution [210]

M20 Contributed by Robert BeezerIn C3, the vector space of column vectors of size 3, prove that the set Z is a subspace.

Z =

x1

x2

x3

∣∣∣∣∣∣ 4x1 − x2 + 5x3 = 0

Solution [210]

Page 242: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.SOL Solutions 210

Subsection SOLSolutions

C20 Exercise [209] Contributed by Robert BeezerThe question is if p can be written as a linear combination of the vectors in W . To checkthis, we set p equal to a linear combination and massage with the definitions of vectoraddition and scalar multiplication that we get with P3 (Example VS.VSP [186])

p(x) = a1(x3 + x2 + x) + a2(x

3 + 2x− 6) + a3(x2 − 5)

x3 + 6x + 4 = (a1 + a2)x3 + (a1 + a3)x

2 + (a1 + 2a2)x + (−6a2 − 5a3)

Equating coefficients of equal powers of x, we get the system of equations,

a1 + a2 = 1

a1 + a3 = 0

a1 + 2a2 = 6

−6a2 − 5a3 = 4

The augmented matrix of this system of equations row-reduces to1 0 0 0

0 1 0 0

0 0 1 0

0 0 0 1

There is a leading 1 in the last column, so Theorem RCLS [41] implies that the systemis inconsistent. So there is no way for p to gain membership in W , so p 6∈ W .

M20 Exercise [209] Contributed by Robert BeezerThe membership criteria for Z is a single linear equation, which comprises a homogeneoussystem of equations. As such, we can recognize Z as the solutions to this system, andtherefore Z is a null space. Specifically, Z = N

([4 −1 5

]). Every null space is a

subspace by Theorem NSMS [200].A less direct solution appeals to Theorem TSS [198].

First, we want to be certain Z is non-empty. The zero vector of C3, 0 =

000

, is a

good candidate, since if it fails to be in Z, we will know that Z is not a vector space.Check that

4(0)− (0) + 5(0) = 0

so that 0 ∈ Z.

Suppose x =

x1

x2

x3

and y =

y1

y2

y3

are vectors from Z. Then we know that these

vectors cannot be totally arbitrary, they must have gained membership in Z by virtue ofmeeting the membership test. For example, we know that x must satisfy 4x1−x2+5x3 = 0

Page 243: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection S.SOL Solutions 211

while y must satisfy 4y1−y2+5y3 = 0. Our second criteria asks the question, is x+y ∈ Z?Notice first that

x + y =

x1

x2

x3

+

y1

y2

y3

=

x1 + y1

x2 + y2

x3 + y3

and we can test this vector for membership in Z as follows,

4(x1 + y1)− 1(x2 + y2) + 4(x3 + y3)

= 4x1 + 4y1 − x2 − y2 + 5x3 + 5y3

= (4x1 − x2 + 5x3) + (4y1 − y2 + 5y3)

= 0 + 0 x ∈ Z, y ∈ Z

= 0

and by this computation we see that x + y ∈ Z.If α is a scalar and x ∈ Z, is it always true that αx ∈ Z? To check our third criteria,

we examine

αx = α

x1

x2

x3

=

αx1

αx2

αx3

and we can test this vector for membership in Z with

4(αx1)− (αx2) + 5(αx3)

= α(4x1 − x2 + 5x3)

= α0 x ∈ Z

= 0

and we see that indeed αx ∈ Z. With the three conditions of Theorem TSS [198] fulfilled,we can conclude that Z is a subspace of C3.

Page 244: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section B Bases 212

Section B

Bases

A basis of a vector space is one of the most useful concepts in linear algebra. It oftenprovides a finite description of an infinite vector space. But before we can define a basiswe need to return to the idea of linear independence.

Subsection LILinear independence

Our previous definition of linear independence (Definition LI [212]) employed a relation oflinear dependence that had a linear combination on one side of the equality. As a linearcombination in a vector space (Definition LC [202]) depends only on vector additionand scalar multiplication we can extend our definition of linear independence from thesetting of Cm to the setting of a general vector space V . Compare these definitions withDefinition RLDCV [96] and Definition LICV [96].

Definition RLDRelation of Linear DependenceSuppose that V is a vector space. Given a set of vectors S = {u1, u2, u3, . . . , un}, anequation of the form

α1u1 + α2u2 + α3u3 + · · ·+ αnun = 0

is a relation of linear dependence on S. If this equation is formed in a trivial fashion,i.e. αi = 0, 1 ≤ i ≤ n, then we say it is a trivial relation of linear dependence onS. �

Definition LILinear IndependenceSuppose that V is a vector space. The set of vectors S = {u1, u2, u3, . . . , un} is linearlydependent if there is a relation of linear dependence on S that is not trivial. In the casewhere the only relation of linear dependence on S is the trivial one, then S is a linearlyindependent set of vectors. �

Notice the emphasis on the word “only.” This might remind you of the definition ofa nonsingular matrix, where if the matrix is employed as the coefficient matrix of ahomogeneous system then the only solution is the trivial one.

Example B.LIP4Linear independence in P4

In the vector space of polynomials with degree 4 or less, P4 (Example VS.VSP [186])consider the set

S ={2x4 + 3x3 + 2x2 − x + 10, −x4 − 2x3 + x2 + 5x− 8, 2x4 + x3 + 10x2 + 17x− 2

}.

Page 245: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.LI Linear independence 213

Is this set of vectors linearly independent or dependent? Consider that

3(2x4 + 3x3 + 2x2 − x + 10

)+ 4

(−x4 − 2x3 + x2 + 5x− 8

)+ (−1)

(2x4 + x3 + 10x2 + 17x− 2

)= 0x4 + 0x3 + 0x2 + 0x + 0 = 0.

This is a nontrivial relation of linear dependence (Definition RLD [212]) on the set S andso convinces us that S is linearly dependent (Definition LI [212]).

Now, I hear you say, “Where did those scalars come from?” Do not worry about thatright now, just be sure you understand why the above explanation is sufficient to provethat S is linearly dependent. The remainder of the example will demonstrate how wemight find these scalars if they had not been provided so readily. Lets look at anotherset of vectors (polynomials) from P4. Let

T ={3x4 − 2x3 + 4x2 + 6x− 1, −3x4 + 1x3 + 0x2 + 4x + 2,

4x4 + 5x3 − 2x2 + 3x + 1, 2x4 − 7x3 + 4x2 + 2x + 1}

Suppose we have a relation of linear dependence on this set,

0 = α1

(3x4 − 2x3 + 4x2 + 6x− 1

)+ α2

(−3x4 + 1x3 + 0x2 + 4x + 2

)+ α3

(4x4 + 5x3 − 2x2 + 3x + 1

)+ α4

(2x4 − 7x3 + 4x2 + 2x + 1

)Using our definitions of vector addition and scalar multiplication in P4 (Example VS.VSP [186]),we arrive at,

0x4 + 0x3 + 0x2 + 0x + 0 = (3α1 − 3α2 + 4α3 + 2α4) x4 + (−2α1 + α2 + 5α3 − 7α4) x3

+ (4α1 +−2α3 + 4α4) x2 + (6α1 + 4α2 + 3α3 + 2α4) x

+ (−α1 + 2α2 + α3 + α4) .

Equating coefficients, we arrive at the homogeneous system of equations,

0 = 3α1 − 3α2 + 4α3 + 2α4

0 = −2α1 + α2 + 5α3 − 7α4

0 = 4α1 +−2α3 + 4α4

0 = 6α1 + 4α2 + 3α3 + 2α4

0 = −α1 + 2α2 + α3 + α4

We form the augmented matrix of this system of equations and row-reduce to find1 0 0 0 0

0 1 0 0 0

0 0 1 0 0

0 0 0 1 00 0 0 0 0

We expected the system to be consistent (Theorem HSC [50]) and so can compute n−r =4 − 4 = 0 and Theorem CSRN [42] tells us that the solution is unique. Since this is ahomogeneous system, this unique solution is the trivial solution (Definition TSHSE [50]),α1 = 0, α2 = 0, α3 = 0, α4 = 0. So by Definition LI [212] the set T is linearlyindependent.

Page 246: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.LI Linear independence 214

A few observations. If we had discovered infinitely many solutions, then we couldhave used one of the non-trivial ones to provide a linear combination in the mannerwe used to show that S was linearly dependent. It is important to realize that is notinteresting that we can create a relation of linear dependence with zero scalars — wecan always do that — but that for T , this is the only way to create a relation of lineardependence. It was no accident that we arrived at a homogeneous system of equationsin this example, it is related to our use of the zero vector in defining a relation of lineardependence. It is easy to present a convincing statement that a set is linearly dependent(just exhibit a nontrivial relation of linear dependence) but a convincing statement oflinear independence requires demonstrating that there is no relation of linear dependenceother than the trivial one. Notice how we relied on theorems from Chapter SLE [2] toprovide this demonstration. Whew! There’s a lot going on in this example. Spend sometime with it, we’ll still be here when you get back. 4

Example B.LIM32Linear Independence in M32

Consider the two sets of vectors R and S from the vector space of all 3× 2 matrices, M32

(Example VS.VSM [186])

R =

3 −1

1 46 −6

,

−2 31 −3−2 −6

,

6 −6−1 07 −9

,

7 9−4 −52 5

S =

2 0

1 −11 3

,

−4 0−2 2−2 −6

,

1 1−2 12 4

,

−5 3−10 72 0

One set is linearly independent, the other is not. Which is which? Lets examine R first.Build a generic relation of linear dependence,

α1

3 −11 46 −6

+ α2

−2 31 −3−2 −6

+ α3

6 −6−1 07 −9

+ α4

7 9−4 −52 5

= 0

Massaging the left-hand side with our definitions of vector addition and scalar multipli-cation in M32 (Example VS.VSM [186]) we obtain,3α1 − 2α2 + 6α3 + 7α4 −1α1 + 3α2 − 6α3 + 9α4

1α1 + 1α2 − α3 − 4α4 4α1 − 3α2 +−5α4

6α1 − 2α2 + 7α3 + 2α4 −6α1 − 6α2 − 9α3 + 5α4

=

0 00 00 0

Using our definition of matrix equality (Definition ME [121]) and equating correspondingentries we get the homogeneous system of six equations in four variables,

3α1 − 2α2 + 6α3 + 7α4 = 0

−1α1 + 3α2 − 6α3 + 9α4 = 0

1α1 + 1α2 − α3 − 4α4 = 0

4α1 − 3α2 +−5α4 = 0

6α1 − 2α2 + 7α3 + 2α4 = 0

−6α1 − 6α2 − 9α3 + 5α4 = 0

Page 247: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.LI Linear independence 215

Form the augmented matrix of this homogeneous system and row-reduce to obtain

1 0 0 0 0

0 1 0 0 0

0 0 1 0 0

0 0 0 1 00 0 0 0 00 0 0 0 0

Analyzing this matrix we are led to conclude that α1 = 0, α2 = 0, α3 = 0, α4 = 0. Thismeans there is only a trivial relation of linear dependence on the vectors of R and so wecall R a linearly independent set (Definition LI [212]).

So it must be that S is linearly dependent. Let’s see if we can find a non-trivialrelation of linear dependence on S. We will begin as with R, by constructing a relationof linear dependence with unknown scalars,

α1

2 01 −11 3

+ α2

−4 0−2 2−2 −6

+ α3

1 1−2 12 4

+ α4

−5 3−10 72 0

= 0

Massaging the left-hand side with our definitions of vector addition and scalar multipli-cation in M32 (Example VS.VSM [186]) we obtain, 2α1 − 4α2 + α3 − 5α4 α3 + 3α4

α1 − 2α2 − 2α3 − 10α4 −α1 + 2α2 + α3 + 7α4

α1 − 2α2 + 2α3 + 2α4 3α1 − 6α2 + 4α3

=

0 00 00 0

Using our definition of matrix equality (Definition ME [121]) and equating correspondingentries we get the homogeneous system of six equations in four variables,

2α1 − 4α2 + α3 − 5α4 = 0

+α3 + 3α4 = 0

α1 − 2α2 − 2α3 − 10α4 = 0

−α1 + 2α2 + α3 + 7α4 = 0

α1 − 2α2 + 2α3 + 2α4 = 0

3α1 − 6α2 + 4α3 = 0

Form the augmented matrix of this homogeneous system and row-reduce to obtain1 −2 0 −4 0

0 0 1 3 00 0 0 0 00 0 0 0 00 0 0 0 00 0 0 0 0

Analyzing this we see that the system is consistent (we expected that since the systemis homogeneous, Theorem HSC [50]) and has n− r = 4− 2 = 2 free variables, namely α2

and α4. This means there are infinitely many solutions, and in particular, we can find anon-trivial solution, so long as we do not pick all of our free variables to be zero. Themere presence of a nontrivial solution for these scalars is enough to conclude that S a

Page 248: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.SS Spanning Sets 216

linearly dependent set (Definition LI [212]). But let’s go ahead and explicitly constructa non-trivial relation of linear dependence.

Choose α2 = 1 and α4 = −1. There is nothing special about this choice, thereare infinitely many possibilities, some “easier” than this one, just avoid picking bothvariables to be zero. Then we find the corresponding dependent variables to be α1 = −2and α3 = 3. So the relation of linear dependence,

(−2)

2 01 −11 3

+ (1)

−4 0−2 2−2 −6

+ (3)

1 1−2 12 4

+ (−1)

−5 3−10 72 0

=

0 00 00 0

is an iron-clad demonstration that S is linearly dependent. Can you construct anothersuch demonstration? 4

Subsection SSSpanning Sets

In a vector space V , suppose we are given a set of vectors S ⊆ V . Then we can im-mediately construct a subspace, Sp (S), using Definition SS [203] and then be assuredby Theorem SSS [203] that the construction does provide a subspace. We now turn thesituation upside-down. Suppose we are first given a subspace W ⊆ V . Can we find a setS so that Sp (S) = W? Typically W is infinite and we are searching for a finite set ofvectors S that we can combine in linear combinations and “build” all of W .

I like to think of S as the raw materials that are sufficient for the construction of W .If you have nails, lumber, wire, copper pipe, drywall, plywood, carpet, shingles, paint(and a few other things), then you can combine them in many different ways to create ahouse (or infinitely many different houses for that matter). A fast-food restaurant mayhave beef, chicken, beans, cheese, tortillas, taco shells and hot sauce and from this smalllist of ingredients build a wide variety of items for sale. Or maybe a better analogy comesfrom Ben Cordes — the additive primary colors (red, green and blue) can be combinedto create many different colors by varying the intensity of each. The intensity is like asclar multiple, and the combination of the three intensities is like vector addition. Thethree individual colors, red, green and blue, are the elements of the spanning set.

Because we will use terms like “spanned by” and “spanning set,” there is the potentialfor confusion with “the span.” Come back and reread the first paragraph of this subsec-tion whenever you are uncertain about the difference. Here’s the working definition.

Definition TSSTo Span a SubspaceSuppose V is a vector space and W is a subspace. A subset S of W is a spanning setfor W if Sp (S) = W . In this case, we also say S spans W . �

The definition of a spanning set requires that two sets (subspaces actually) be equal. IfS is a subset of W , then Sp (S) ⊆ W , always. Thus it is usually only necessary to provethat W ⊆ Sp (S). Now would be a good time to review Technique SE [14].

Example B.SSP4Spanning set in P4

Page 249: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.SS Spanning Sets 217

In Example S.SP4 [198] we showed that

W = {p(x) | p ∈ P4, p(2) = 0}

is a subspace of P4, the vector space of polynomials with degree at most 4 (Exam-ple VS.VSP [186]). In this example, we will show that the set

S ={x− 2, x2 − 4x + 4, x3 − 6x2 + 12x− 8, x4 − 8x3 + 24x2 − 32x + 16

}is a spanning set for W . To do this, we require that W = Sp (S). This is an equalityof sets. We can check that every polynomial in S has x = 2 as a root and thereforeS ⊆ W . Since W is closed under addition and scalar multiplication, Sp (S) ⊆ W also.So it remains to show that W ⊆ Sp (S) (Technique SE [14]). So choose an arbitrarypolynomial in W , say r(x) = ax4 + bx3 + cx2 + dx + e ∈ W . This polynomial is notas arbitrary as it would appear, since we also know it must have x = 2 as a root. Thistranslates to

0 = a(2)4 + b(2)3 + c(2)(2) + d(2) + e = 16a + 8b + 4c + 2d + e

as a condition on r.

We wish to show that r is a polynomial in Sp (S), that is, we want to show that rcan be written as a linear combination of the vectors (polynomials) in S. So let’s try.

r(x) = ax4 + bx3 + cx2 + dx + e

= α1 (x− 2) + α2

(x2 − 4x + 4

)+ α3

(x3 − 6x2 + 12x− 8

)+ α4

(x4 − 8x3 + 24x2 − 32x + 16

)= α4x

4 + (α3 − 8α4) x3 + (α2 − 6α3 + 24α2) x2

+ (α1 − 4α2 + 12α3 − 32α4) x + (−2α1 + 4α2 − 8α3 + 16α4)

Equating coefficients (vector equality in P4) gives the system of five equations in fourvariables,

α4 = a

α3 − 8α4 = b

α2 − 6α3 + 24α2 = c

α1 − 4α2 + 12α3 − 32α4 = d

−2α1 + 4α2 − 8α3 + 16α4 = e

Any solution to this system of equations will provide the linear combination we need todetermine if r ∈ Sp (S), but we need to be convinced there is a solution for any valuesof a, b, c, d, e that qualify r to be a member of W . So the question is: is this systemof equations consistent? We will form the augmented matrix, and row-reduce. (Weprobably need to do this by hand, since the matrix is symbolic — reversing the order ofthe first four rows is the best way to start). We obtain a matrix in reduced row-echelon

Page 250: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.SS Spanning Sets 218

form1 0 0 0 32a− 12b + 4c + d

0 1 0 0 −24a + 6b + c

0 0 1 0 8a + b

0 0 0 1 a0 0 0 0 16a + 8b + 4c + 2d + e

=

1 0 0 0 32a− 12b + 4c + d

0 1 0 0 −24a + 6b + c

0 0 1 0 8a + b

0 0 0 1 a0 0 0 0 0

where the last entry of the last column has been simplified to zero according to the onecondition we were able to impose on an arbitrary polynomial from W . So with no leading1’s in the last column, Theorem RCLS [41] tells us this system is consistent. Therefore,W ⊆ Sp (S) and so W = Sp (S) and S is a spanning set for W .

Notice that an alternative to row-reducing the augmented matrix by hand would beto appeal to Theorem RNS [136] by expressing the range of the coefficient matrix as anull space, and then verifying that the condition on r guarantees that r is in the range,thus implying that the system is always consistent. Give it a try, we’ll wait. This hasbeen a complicated example, but worth studying carefully. 4

Given a subspace and a set of vectors, as in Example B.SSP4 [216] it can take some workto determine that the set actually is a spanning set. An even harder problem is to beconfronted with a subspace and required to construct a spanning set with no guidance.We will now work an example of this flavor, but some of the steps will be unmotivated.Fortunately, we will have some better tools for this type of problem later on.

Example B.SSM22Spanning set in M22

In the space of all 2× 2 matrices, M22 consider the subspace

Z =

{[a bc d

] ∣∣∣∣ a + 2b− 7d = 0, 3a− b + 7c− 7d = 0

}We will construct a limited number of matrices in Z and hope that they form a spanningset. We want to avoid any linear combinations of matrices we have already chosen asthey will not provide us with anything new to build Z with. And the zero matrix shouldbe avoided at all costs, it is of no help. Notice that if we choose a and b, the the firstcondition defining the set will determine the value d and then the second condition willdetermine c. This reminds me of how free variables behave, and so I am going to choosea = 1 and b = 0, and then a = 0 with b = 1. This creates two matrices[

1 0−2

717

] [0 137

27

]Because it will be easier to work with integers, I will multiply each matrix by the scalar 7(Definition SMM [122]), which should not change any of the properties relative to beinga spanning set. The new matrices are then

Q =

{[7 0−2 1

],

[0 73 2

]}

Page 251: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.B Bases 219

The question is: if we take an arbitrary element of Z, can we write it as a linear combi-nation of these two matrices in Q? Let’s try.[

a bc d

]= α1

[7 0−2 1

]+ α2

[0 73 2

]=

[7α1 7α2

−2α1 + 3α2 α1 + 2α2

]Using our definition of matrix equality (Definition ME [121]) we equate correspondingentries and get a system of four equations in two variables,

7α1 = a

7α2 = b

−2α1 + 3α2 = c

α1 + 2α2 = d

Form the augmented matrix and row-reduce (by hand),1 0 a

7

0 1 b7

0 0 17(2a− 3b + 7c)

0 0 17(a + 2b− 7d)

In order to obtain a spanning set, we want this system to have a solution for any matrix[a bc d

]∈ Z. However, this matrix is not as arbitrary as it appears, we may also assume

(or require) that a + 2b − 7d = 0 and 3a − b + 7c − 7d = 0. To be a consistent system,Theorem RCLS [41] tells us that we need to have no leading 1’s in the final column ofthe row-reduced matrix. We see that the expression in the fourth row is

1

7(a + 2b− 7d) =

1

7(0) = 0.

With a bit more finesse, the expression in the third row is

1

7(2a− 3b + 7c) =

1

7(2a− 3b + 7c + 0)

=1

7(2a− 3b + 7c + (a + 2b− 7d))

=1

7(3a− b + 7c− 7d)

=1

7(0)

= 0.

So the system is consistent, the desired scalars can be found as a solution in every event,so Z ⊆ Sp (Q) and Q is a spanning set for Z. 4

Subsection BBases

We now have all the tools in place to define a basis of a vector space.

Page 252: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.B Bases 220

Definition BBasisSuppose V is a vector space. Then a subset S ⊆ V is a basis of V if it is linearlyindependent and spans V . �

So, a basis is a linearly independent spanning set for a vector space. The requirementthat the set spans insures that S has enough raw material to build V , while the linearindependence requirement insures that we do not have any more raw material than weneed. As we shall see soon in Section D [230], a basis is a minimal spanning set.

You may have noticed that we used the term basis for some of the titles of previoustheorems (e.g. Theorem BNS [104], Theorem BROC [131], Theorem BRS [144]) and ifyou review each of these theorems you will see that their conclusions provide linearly inde-pendent spanning sets for sets that we now recognize as subspaces of Cm. Examples asso-ciated with these theorems include Example LI.NSLIL [105], Example RM.ROCD [133]and Example RSM.IS [145]. As we will see, these three theorems will continue to bepowerful tools, even in the setting of more general vector spaces.

Furthermore, the archetypes contain an abundance of bases. For each coefficientmatrix of a system of equations, and for each archetype defined simply as a matrix, thereis a basis for the null space, three bases for the range, and a basis for the row space. Forthis reason, our subsequent examples will concentrate on bases for vector spaces otherthan Cm. Notice that Definition B [220] does not preclude a vector space from havingmany bases, and this is the case, as hinted above by the statement that the archetypescontain three bases for the range of a matrix. More generally, we can grab any basis fora vector space, multiply any one basis vector by a non-zero scalar and create a slightlydifferent set that is still a basis. For “important” vector spaces, it will be convenientto have a collection of “nice” bases. When a vector space has a single particularly nicebasis, it is sometimes called the standard basis though there is nothing precise enoughabout this term to allow us to define it formally — it is a question of style. Here aresome nice bases for important vector spaces.

Theorem SUVBStandard Unit Vectors are a BasisThe set of standard unit vectors for Cm, B = {e1, e2, e3, . . . , em} = {ei | 1 ≤ i ≤ m} isa basis for the vector space Cm. �

Proof We must show that the set B is both linearly independent and a spanning setfor Cm. First, the vectors in B are, by Definition SUV [169], the columns of the identitymatrix, which we know is nonsingular (since it row-reduces to the identity matrix!). Andthe columns of a nonsingular matrix are linearly independent by Theorem NSLIC [103].

Suppose we grab an arbitrary vector from Cm, say

v =

v1

v2

v3...

vm

.

Page 253: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.B Bases 221

Can we write v as a linear combination of the vectors in B? Yes, and quite simply.v1

v2

v3...

vm

= v1

100...0

+ v2

010...0

+ v3

001...0

+ · · ·+ vm

000...1

v = v1e1 + v2e2 + v3e3 + · · ·+ vmem

this shows that Cm ⊆ Sp (B), which is sufficient to show that B is a spanning set forCm. �

Example B.BPSome bases for Pn

The vector space of polynomials with degree at most n, Pn, has the basis

B ={1, x, x2, x3, . . . , xn

}.

Another nice basis for Pn is

C ={1, 1 + x, 1 + x + x2, 1 + x + x2 + x3, . . . , 1 + x + x2 + x3 + · · ·+ xn

}.

Checking that each of B and C is a linearly independent spanning set are good exercises.4

Example B.BMA basis for the vector space of matricesIn the vector space Mmn of matrices (Example VS.VSM [186]) define the matrices Bk`,1 ≤ k ≤ m, 1 ≤ ` ≤ n by

[Bk`]ij =

{1 if k = i, ` = j

0 otherwise.

So these matrices have entries that are all zeros, with the exception of a lone entry thatis one. The set of all mn of them,

B = {Bk` | 1 ≤ k ≤ m, 1 ≤ ` ≤ n}

forms a basis for Mmn. 4

The bases described above will often be convenient ones to work with. However a basisdoesn’t have to obviously look like a basis.

Example B.BSP4A basis for a subspace of P4

In Example B.SSP4 [216] we showed that

S ={x− 2, x2 − 4x + 4, x3 − 6x2 + 12x− 8, x4 − 8x3 + 24x2 − 32x + 16

}

Page 254: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.B Bases 222

is a spanning set for W = {p(x) | p ∈ P4, p(2) = 0}. We will now show that S is alsolinearly independent in W . Begin with a relation of linear dependence,

0 + 0x + 0x2 + 0x3 + 0x4 = α1 (x− 2) + α2

(x2 − 4x + 4

)+ α3

(x3 − 6x2 + 12x− 8

)+ α4

(x4 − 8x3 + 24x2 − 32x + 16

)= α4x

4 + (α3 − 8α4) x3 + (α2 − 6α3 + 24α4) x2

+ (α1 − 4α2 + 12α3 − 32α4) x + (−2α1 + 4α2 − 8α3 + 16α4)

Equating coefficients (vector equality in P4) gives the homogeneous system of five equa-tions in four variables,

α4 = 0

α3 − 8α4 = 0

α2 − 6α3 + 24α4 = 0

α1 − 4α2 + 12α3 − 32α4 = 0

−2α1 + 4α2 − 8α3 + 16α4 = 0

We form the augmented matrix, and row-reduce to obtain a matrix in reduced row-echelon form

1 0 0 0 0

0 1 0 0 0

0 0 1 0 0

0 0 0 1 00 0 0 0 0

With only the trivial solution to this homogeneous system, we conclude that only scalarsthat will form a relation of linear dependence are the trivial ones, and therefore the setS is linearly independent. Finally, S has earned the right to be called a basis for W . 4

Example B.BSM22A basis for a subspace of M22

In Example B.SSM22 [218] we discovered that

Q =

{[7 0−2 1

],

[0 73 2

]}is a spanning set for the subspace

Z =

{[a bc d

] ∣∣∣∣ a + 2b− 7d = 0, 3a− b + 7c− 7d = 0

}of the vector space of all 2× 2 matrices, M22. If we can also determine that Q is linearlyindependent in Z (or in M22), then it will qualify as a basis for Z. Let’s begin with arelation of linear dependence.[

0 00 0

]= α1

[7 0−2 1

]+ α2

[0 73 2

]=

[7α1 7α2

−2α1 + 3α2 α1 + 2α2

]

Page 255: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.BRS Bases from Row Spaces 223

Using our definition of matrix equality (Definition ME [121]) we equate correspondingentries and get a homogeneous system of four equations in two variables,

7α1 = 0

7α2 = 0

−2α1 + 3α2 = 0

α1 + 2α2 = 0

We could row-reduce the augmented matrix of this system, but it is not necessary. Thefirst two equations tell us that α1 = 0, α2 = 0 is the only solution to this homogeneoussystem. This qualifies the set Q as being linearly independent, since the only relation oflinear dependence is trivial. Therefore Q is a basis for Z. 4

Subsection BRSBases from Row Spaces

We have seen several examples of bases in different vector spaces. In this subsection,and the next (Subsection B.BNSM [225]), we will consider building bases for Cm and itssubspaces.

Suppose we have a subspace of Cm that is expressed as the span of a set of vectors,S, and S is not necessarily linearly independent, or perhaps not very attractive. The-orem REMRS [142] says that row-equivalent matrices have identical row spaces, whileTheorem BRS [144] says the nonzero rows of a matrix in reduced row-echelon form are abasis for the row space. These theorems together give us a great computational tool forquickly finding a basis for a subspace that is expressed originally as a span.

Example B.RSBRow space basisWhen we first defined the span of a set of vectors, in Example SS.SCAD [91] we lookedat the set

W = Sp

2−31

,

141

,

7−54

,

−7−6−5

with an eye towards realizing W as the span of a smaller set. By building relationsof linear dependence (though we did not know them by that name then) we were ableto remove two vectors and write W as the span of the other two vectors. These tworemaining vectors formed a linearly independent set, even though we did not know thatat the time.

Now we know that W is a subspace and must have a basis. Consider the matrix, C,whose rows are the vectors in the spanning set for W ,

C =

2 −3 11 4 17 −5 4−7 −6 −5

Then, by Definition RSM [141], the row space of C will be W , rs (C) = W . Theo-rem BRS [144] tells us that if we row-reduce C, the nonzero rows of the row-equivalent

Page 256: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.BRS Bases from Row Spaces 224

matrix in reduced row-echelon form will be a basis for rs (C), and hence a basis for W .Let’s do it — C row-reduces to

1 0 711

0 1 111

0 0 00 0 0

If we convert the two nonzero rows to column vectors then we have a basis,

B =

1

0711

,

01111

and

W = Sp

1

0711

,

01111

4

Example RSM.IS [145] provides another example of this flavor, though now we can noticethat X is a subspace, and that the resulting set of three vectors is a basis. This is sucha powerful technique that we should do one more example.

Example B.RSReducing a spanIn Example LI.RS [101] we began with a set of n = 4 vectors from C5,

R = {v1, v2, v3, v4} =

12−132

,

21312

,

0−76−11−2

,

41216

and defined V = Sp (R). Our goal in that problem was to find a relation of lineardependence on the vectors in R, solve the resulting equation for one of the vectors, andre-express V as the span of a set of three vectors.

Here is another way to accomplish something similar. The row space of the matrix

A =

1 2 −1 3 22 1 3 1 20 −7 6 −11 −24 1 2 1 6

is equal to Sp (R). By Theorem BRS [144] we can row-reduce this matrix, ignore anyzero rows, and use the non-zero rows as column vectors that are a basis for the row spaceof A. Row-reducing A creates the matrix

1 0 0 − 117

3017

0 1 0 2517

− 217

0 0 1 − 217

− 817

0 0 0 0 0

Page 257: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.BNSM Bases and NonSingular Matrices 225

So

100− 1

173017

,

0102517

− 217

,

001− 2

17

− 817

is a basis for V . Our theorem tells us this is a basis, there is no need to verify that thesubspace spanned by three vectors (rather than four) is the identical subspace, and thereis no need to verify that we have reached the limit in reducing the set, since the set ofthree vectors is guaranteed to be linearly independent. 4

Subsection BNSMBases and NonSingular Matrices

A quick source of diverse bases for Cm is the set of columns of a nonsingular matrix.

Theorem CNSMBColumns of NonSingular Matrix are a BasisSuppose that A is a square matrix. Then the columns of A are a basis of Cm if and onlyif A is nonsingular. �

Proof (⇒) Suppose that the columns of A are a basis for Cm. Then Definition B [220]says the set of columns is linearly independent. Theorem NSLIC [103] then says that Ais nonsingular.

(⇐) Suppose that A is nonsingular. Then by Theorem NSLIC [103] this set of columnsis linearly independent. Theorem RNSM [139] says that for a nonsingular matrix, R (A) =Cm. This is equivalent to saying that the columns of A are a spanning set for the vectorspace Cm. As a linearly independent spanning set, the columns of A qualify as a basisfor Cm (Definition B [220]). �

Example B.CABAKColumns as Basis, Archetype KArchetype K [423] is the 5× 5 matrix

K =

10 18 24 24 −1212 −2 −6 0 −18−30 −21 −23 −30 3927 30 36 37 −3018 24 30 30 −20

which is row-equivalent to the 5× 5 identity matrix I5. So by Theorem NSRRI [57], Kis nonsingular. Then Theorem CNSMB [225] says the set

1012−302718

,

18−2−213024

,

24−6−233630

,

240−303730

,

−12−1839−30−20

is a (novel) basis of C5. 4

Page 258: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.VR Vector Representation 226

Perhaps we should view the fact that the standard unit vectors are a basis (Theo-rem SUVB [220]) as just a simple corollary of Theorem CNSMB [225]?

With a new equivalence for a nonsingular matrix, we can update our list of equiva-lences (Theorem NSME4 [178]).

Theorem NSME5NonSingular Matrix Equivalences, Round 5Suppose that A is a square matrix of size n. The following are equivalent.

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the zero vector, N (A) = {0}.

4. The linear system LS(A, b) has a unique solution for every possible choice of b.

5. The columns of A are a linearly independent set.

6. The range of A is Cn, R (A) = Cn.

7. A is invertible.

8. The columns of A are a basis for Cn. �

Subsection VRVector Representation

In Chapter R [358] we will take up the matter of representations fully. Now we willprove a critical theorem that tells us how to represent a vector. This theorem could wait,but working with it now will provide some extra insight into the nature of a basis as aminimal spanning set. First an example, then the theorem.

Example B.AVRA vector representationThe set

−751

,

−650

,

−1274

is a basis for C3. (You have the tools to check this right now, but this fact can beassumed for the purpose of this example.) This set comes from the columns of the

coefficient matrix of Archetype B [384]. Because x =

−352

is a solution to this system,

we can use Theorem SLSLC [76]−33245

= (−3)

−751

+ (5)

−650

+ (2)

−1274

.

Page 259: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.VR Vector Representation 227

Further, we know this is the only way we can express

−33245

as a linear combination

of these basis vectors, since the nonsingularity of the coefficient matrix tells that thissolution is unique. This is all an illustration of the following theorem. 4

Theorem VRRBVector Representation Relative to a BasisSuppose that V is a vector space with basis B = {v1, v2, v3, . . . , vm} and that w is avector in V . Then there exist unique scalars a1, a2, a3, . . . , am such that

w = a1v1 + a2v2 + a3v3 + · · ·+ amvm. �

Proof That w can be written as a linear combination of the basis vectors follows fromthe spanning property of the basis (Definition B [220], Definition TSS [216]). This isgood, but not the meat of this theorem. We now know that for any choice of the vectorw there exist some scalars that will create w as a linear combination of the basis vectors.The real question is: Is there more than one way to write w as a linear combination of{v1, v2, v3, . . . , vm}? Are the scalars a1, a2, a3, . . . , am unique? (Technique U [59])

Assume there are two ways to express w as a linear combination of {v1, v2, v3, . . . , vm}.In other words there exist scalars a1, a2, a3, . . . , am and b1, b2, b3, . . . , bm so that

w = a1v1 + a2v2 + a3v3 + · · ·+ amvm

w = b1v1 + b2v2 + b3v3 + · · ·+ bmvm.

Then notice that (using the vector space axioms of associativity and distributivity)

0 = w −w

= (a1v1 + a2v2 + a3v3 + · · ·+ amvm)− (b1v1 + b2v2 + b3v3 + · · ·+ bmvm)

= (a1 − b1)v1 + (a2 − b2)v2 + (a3 − b3)v3 + · · ·+ (am − bm)vm

But this is a relation of linear dependence on a linearly independent set of vectors! Nowwe are the other half of the assumption that {v1, v2, v3, . . . , vm} is a basis (Defini-tion B [220]). So by Definition LI [212] it must happen that the scalars are all zero. Thatis,

(a1 − b1) = 0 (a2 − b2) = 0 (a3 − b3) = 0 . . . (am − bm) = 0

a1 = b1 a2 = b2 a3 = b3 . . . am = bm.

And so we find that the scalars are unique. �

This is a very typical use of the hypothesis that a set is linear independent — obtaina relation of linear dependence and then conclude that the scalars must all be zero.The result of this theorem tells us that we can write any vector in a vector space asa linear combination of the basis vectors, but only just. There is only enough rawmaterial in the spanning set to write each vector one way as a linear combination. Thistheorem will be the basis (pun intended) for our future definition of coordinate vectorsin Definition XXX [??].

Page 260: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.EXC Exercises 228

Subsection EXCExercises

C20 Contributed by Robert BeezerIn the vector space of 2 × 2 matrices, M22, determine if the set S below is linearly

independent.

S =

{[2 −11 3

],

[0 4−1 2

],

[4 21 3

]}Solution [229]

Page 261: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection B.SOL Solutions 229

Subsection SOLSolutions

C20 Exercise [228] Contributed by Robert BeezerBegin with a relation of linear dependence on the vectors in S and massage it accordingto the definitions of vector addition and scalar multiplication in M22,

O = a1

[2 −11 3

]+ a2

[0 4−1 2

]+ a3

[4 21 3

][0 00 0

]=

[2a1 + 4a3 −a1 + 4a2 + 2a3

a1 − a2 + a3 3a1 + 2a2 + 3a3

]By our definition of matrix equality (Definition ME [121]) we arrive at a homogeneoussystem of linear equations,

2a1 + 4a3 = 0

−a1 + 4a2 + 2a3 = 0

a1 − a2 + a3 = 0

3a1 + 2a2 + 3a3 = 0

The coefficient matrix of this system row-reduces to the matrix,1 0 0

0 1 0

0 0 10 0 0

and from this we conclude that the only solution is a1 = a2 = a3 = 0. Since the relationof linear dependence (Definition RLD [212]) is trivial, the set S is linearly independent(Definition LI [212]).

Page 262: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section D Dimension 230

Section D

Dimension

Almost every vector space we have encountered has been infinite in size (an exceptionis Example VS.VSS [188]). But some are bigger and richer than others. Dimension,once suitably defined, will be a measure of the size of a vector space, and a usefultool for studying its properties. You probably already have a rough notion of what amathematical definition of dimension might be — try to forget these imprecise ideas andgo with the new ones given here.

Subsection DDimension

Definition DDimensionSuppose that V is a vector space and {v1, v2, v3, . . . , vt} is a basis of V . Then thedimension of V is defined by dim (V ) = t. If V has no finite bases, we say V has infinitedimension. �

This is a very simple definition, which belies its power. Grab a basis, any basis, and countup the number of vectors it contains. That’s the dimension. However, this simplicitycauses a problem. Given a vector space, you and I could each construct different bases— remember that a vector space might have many bases. And what if your basis andmy basis had different sizes? Applying Definition D [230] we would arrive at differentnumbers! With our current knowledge about vector spaces, we would have to say thatdimension is not “well-defined.” Fortunately, there is a theorem that will correct thisproblem.

In a strictly logical progression, the next two theorems would precede the definition ofdimension. Here is a fundamental result that many subsequent theorems will trace theirlineage back to.

Theorem SSLDSpanning Sets and Linear DependenceSuppose that S = {v1, v2, v3, . . . , vt} is a finite set of vectors from the vector spacewhich spans W . Then any set of t + 1 or more vectors from V is linearly dependent. �

Proof We want to prove that any set of t+1 or more vectors from V is linearly dependent.So we will begin with a totally arbitrary set of vectors from V , R = {u1, u2, u3, . . . , um},where m > t. We will now construct a nontrivial relation of linear dependence on R.

Each vector u1, u2, u3, . . . , um can be written as a linear combination of v1, v2, v3, . . . , vt

since S is a spanning set of V . This means there exist scalars aij, 1 ≤ i ≤ t, 1 ≤ j ≤ m,

Page 263: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.D Dimension 231

so that

u1 = a11v1 + a21v2 + a31v3 + · · ·+ at1vt

u2 = a12v1 + a22v2 + a32v3 + · · ·+ at2vt

u3 = a13v1 + a23v2 + a33v3 + · · ·+ at3vt

...

um = a1mv1 + a2mv2 + a3mv3 + · · ·+ atmvt

Now we form, unmotivated, the homogeneous system of t equations in the m variables,x1, x2, x3, . . . , xm, where the coefficients are the just-discovered scalars aij,

a11x1 + a12x2 + a13x3 + · · ·+ a1mxm = 0

a21x1 + a22x2 + a23x3 + · · ·+ a2mxm = 0

a31x1 + a32x2 + a33x3 + · · ·+ a3mxm = 0

...

at1x1 + at2x2 + at3x3 + · · ·+ atmxm = 0

This is a homogeneous system with more variables than equations (our hypothesis isexpressed as m > t), so by Theorem HMVEI [51] there are infinitely many solutions.Choose a nontrivial solution and denote it by x1 = c1, x2 = c2, x3 = c3, . . . , xm = cm.As a solution to the homogeneous system, we then have

a11c1 + a12c2 + a13c3 + · · ·+ a1mcm = 0

a21c1 + a22c2 + a23c3 + · · ·+ a2mcm = 0

a31c1 + a32c2 + a33c3 + · · ·+ a3mcm = 0

...

at1c1 + at2c2 + at3c3 + · · ·+ atmcm = 0

As a collection of nontrivial scalars, c1, c2, c3, . . . , cm will provide the nontrivial relation

Page 264: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.D Dimension 232

of linear dependence we desire,

c1u1 + c2u2 + c3u3 + · · ·+ cmum

= c1 (a11v1 + a21v2 + a31v3 + · · ·+ at1vt)

+ c2 (a12v1 + a22v2 + a32v3 + · · ·+ at2vt)

+ c3 (a13v1 + a23v2 + a33v3 + · · ·+ at3vt)

...

+ cm (a1mv1 + a2mv2 + a3mv3 + · · ·+ atmvt)

= c1a11v1 + c1a21v2 + c1a31v3 + · · ·+ c1at1vt

+ c2a12v1 + c2a22v2 + c2a32v3 + · · ·+ c2at2vt

+ c3a13v1 + c3a23v2 + c3a33v3 + · · ·+ c3at3vt

...

+ cma1mv1 + cma2mv2 + cma3mv3 + · · ·+ cmatmvt

= (c1a11 + c2a12 + c3a13 + · · ·+ cma1m)v1

+ (c1a21 + c2a22 + c3a23 + · · ·+ cma2m)v2

+ (c1a31 + c2a32 + c3a33 + · · ·+ cma3m)v3

...

+ (c1at1 + c2at2 + c3at3 + · · ·+ cmatm)vt

= (a11c1 + a12c2 + a13c3 + · · ·+ a1mcm)v1

+ (a21c1 + a22c2 + a23c3 + · · ·+ a2mcm)v2

+ (a31c1 + a32c2 + a33c3 + · · ·+ a3mcm)v3

...

+ (at1c1 + at2c2 + at3c3 + · · ·+ atmcm)vt

= 0v1 + 0v2 + 0v3 + · · ·+ 0vt

= 0 + 0 + 0 + · · ·+ 0

= 0

That does it. R has been undeniably shown to be a linearly dependent set. �

The proof just given has some rather monstrous expressions in it, mostly owing to thedouble subscripts present. Now is a great opportunity to show the value of a morecompact notation. We will rewrite the key steps of the previous proof using summationnotation, resulting in a more economical presentation, and hopefully even greater insightinto the key aspects of the proof. So here is an alternate proof — study it carefully.

Proof (Alternate Proof of Theorem SSLD) We want to prove that any set of t + 1or more vectors from V is linearly dependent. So we will begin with a totally arbitraryset of vectors from V , R = {uj | 1 ≤ j ≤ m}, where m > t. We will now construct anontrivial relation of linear dependence on R.

Each vector uj, 1 ≤ j ≤ m can be written as a linear combination of vi, 1 ≤ i ≤ tsince S is a spanning set of V . This means there are scalars aij, 1 ≤ i ≤ t, 1 ≤ j ≤ m,

Page 265: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.D Dimension 233

so that

uj =t∑

i=1

aijvi 1 ≤ j ≤ m

Now we form, unmotivated, the homogeneous system of t equations in the m variables,xj, 1 ≤ j ≤ m, where the coefficients are the just-discovered scalars aij,

m∑j=1

aijxj = 0 1 ≤ i ≤ t

This is a homogeneous system with more variables than equations (our hypothesis isexpressed as m > t), so by Theorem HMVEI [51] there are infinitely many solutions.Choose one of these solutions that is not trivial and denote it by xj = cj, 1 ≤ j ≤ m.As a solution to the homogeneous system, we then have

∑mj=1 aijcj = 0 for 1 ≤ i ≤ t.

As a collection of nontrivial scalars, cj, 1 ≤ j ≤ m, will provide the nontrivial relation oflinear dependence we desire,

m∑j=1

cjuj =m∑

j=1

cj

(t∑

i=1

aijvi

)Substitution for uj

=m∑

j=1

t∑i=1

cjaijvi Distributivity (scalar)

=t∑

i=1

m∑j=1

cjaijvi Commutativity (vector addition)

=t∑

i=1

m∑j=1

aijcjvi Commutativity (scalar)

=t∑

i=1

(m∑

j=1

aijcj

)vi Distributivity (scalar multiplication)

=t∑

i=1

0vi Solution to homogeneous system

=t∑

i=1

0 Theorem ZSSM [191]

= 0 Property of the zero vector

That does it. R has been undeniably shown to be a linearly dependent set. �

Notice how the swap of the two summations is so much easier in the third step above, asopposed to all the rearranging and regrouping that takes place in the previous proof. Inabout half the space. And there are no ellipses (. . .).

Theorem SSLD [230] can be viewed as a generalization of Theorem MVSLD [99].We know that Cm has a basis with m vectors in it (Theorem SUVB [220]), so it is aset of m vectors that spans Cm. By Theorem SSLD [230], any set of more than mvectors from Cm will be linearly dependent. But this is exactly the conclusion we have in

Page 266: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.D Dimension 234

Theorem MVSLD [99]. Maybe this is not a total shock, as the proofs of both theoremsrely heavily on Theorem HMVEI [51]. The beauty of Theorem SSLD [230] is that itapplies in any vector space. We illustrate the generality of this theorem, and hint at itspower, in the next example.

Example D.LDP4Linearly dependent set in P4

In Example B.SSP4 [216] we showed that

S ={x− 2, x2 − 4x + 4, x3 − 6x2 + 12x− 8, x4 − 8x3 + 24x2 − 32x + 16

}is a spanning set for W = {p(x) | p ∈ P4, p(2) = 0}. So we can apply Theorem SSLD [230]to W with t = 4. Here is a set of five vectors from W , as you may check by verifyingthat each is a polynomial of degree 4 or less and has x = 2 as a root,

T = {p1, p2, p3, p4, p5} ⊆ W

p1 = x4 − 2x3 + 2x2 − 8x + 8

p2 = −x3 + 6x2 − 5x− 6

p3 = 2x4 − 5x3 + 5x2 − 7x + 2

p4 = −x4 + 4x3 − 7x2 + 6x

p5 = 4x3 − 9x2 + 5x− 6

By Theorem SSLD [230] we conclude that T is linearly dependent with no further com-putations. 4

Theorem SSLD [230] is indeed powerful, but our main purpose in proving it right nowwas to make sure that our definition of dimension (Definition D [230]) is well-defined.Here’s the theorem.

Theorem BISBases have Identical SizesSuppose that V is a vector space with a finite basis B and a second basis C. Then Band C have the same size. �

Proof Suppose that C has more vectors than B. (Allowing for the possibility that C isinfinite, we can replace C by a subset of that has more vectors than B.) As a basis, B isa spanning set for V (Definition B [220]), so Theorem SSLD [230] says that C is linearlydependent. However, this contradicts the fact that as a basis C is linearly independent(Definition B [220]). So C must also be a finite set, with size less than, or equal to, thatof B.

Suppose that B has more vectors than C. As a basis, C is a spanning set for V(Definition B [220]), so Theorem SSLD [230] says that B is linearly dependent. However,this contradicts the fact that as a basis B is linearly independent (Definition B [220]).So C cannot be strictly smaller than B.

The only possibility left for the sizes of B and C is for them to be equal. �

Theorem BIS [234] tells us that if we find one finite basis in a vector space, then they allhave the same size. This (finally) makes Definition D [230] unambiguous.

Page 267: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.DVS Dimension of Vector Spaces 235

Subsection DVSDimension of Vector Spaces

We can now collect the dimension of some common, and not so common, vector spaces.

Theorem DCMDimension of Cm

The dimension of Cm (Example VS.VSCM [186]) is m. �

Proof Theorem SUVB [220] provides a basis with m vectors. �

Theorem DPDimension of Pn

The dimension of Pn (Example VS.VSP [186]) is n + 1. �

Proof Example B.BP [221] provides two bases with n + 1 vectors. Take your pick. �

Theorem DMDimension of Mmn

The dimension of Mmn (Example VS.VSM [186]) is mn. �

Proof Example B.BM [221] provides a basis with mn vectors. �

Example D.DSM22Dimension of a subspace of M22

It should now be plausible that

Z =

{[a bc d

] ∣∣∣∣ 2a + b + 3c + 4d = 0, −a + 3b− 5c− d = 0

}is a subspace of the vector space M22 (Example VS.VSM [186]). (It is.) To find thedimension of Z we must first find a basis, though any old basis will do.

First concentrate on the conditions relating a, b, c and d. They form a homogeneoussytem of two equations in four variables with coefficient matrix[

2 1 3 4−1 3 −5 −1

]We can row-reduce this matrix to obtain[

1 0 2 2

0 1 −1 0

]Rewrite the two equations represented by each row of this matrix, expressing the depen-dent variables (a and b) in terms of the free variables (c and d), and we obtain,

a = −2c− 2d

b = c

Page 268: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.DVS Dimension of Vector Spaces 236

We can now write a typical entry of Z strictly in terms of c and d, and we can decomposethe result,[

a bc d

]=

[−2c− 2d c

c d

]=

[−2c cc 0

]+

[−2d 00 d

]= c

[−2 11 0

]+ d

[−2 00 1

]this equation says that an arbitrary matrix in Z can be written as a linear combinationof the two vectors in

S =

{[−2 11 0

],

[−2 00 1

]}so we know that

Z = Sp (S) = Sp

({[−2 11 0

],

[−2 00 1

]})Are these two matrices (vectors) also linearly independent? Begin with a relation oflinear dependence on S,

a1

[−2 11 0

]+ a2

[−2 00 1

]= O[

−2a1 − 2a2 a1

a1 a2

]=

[0 00 0

]From the equality of the two entries in the last row, we conclude that a1 = 0, a2 = 0.Thus the only possible relation of linear dependence is the trivial one, and therefore Sis linearly independent (Definition LI [212]). So S is a basis for V (Definition B [220]).Finally, we can conclude that dim (Z) = 2 (Definition D [230]) since S has two elements.4

Example D.DSP4Dimension of a subspace of P4

In Example B.BSP4 [221] we showed that

S ={x− 2, x2 − 4x + 4, x3 − 6x2 + 12x− 8, x4 − 8x3 + 24x2 − 32x + 16

}is a basis for W = {p(x) | p ∈ P4, p(2) = 0}. Thus, the dimension of W is four, dim (W ) =4. 4

It is possible for a vector space to have no finite bases, in which case we say it has infinitedimension. Many of the best examples of this are vector spaces of functions, which leadto constructions like Hilbert spaces. We will focus excluively on finite-dimensional vectorspaces. OK, one example, and then we will focus exclusively on finite-dimensional vectorspaces.

Example D.VSPUDVector space of polynomials with unbounded degreeDefine the set P by

P = {p | p(x) is a polynomial in x}

Our operations will be the same as those defined for Pn (Example VS.VSP [186]).With no restrictions on the possible degrees of our polynomials, any finite set that is

a candidate for spanning P will come up short. Suppose S is a potential finite spanningset and let m be the maximum degree of all the polynomials in S. If q is a polynomial

Page 269: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.RNM Rank and Nullity of a Matrix 237

of degree m + 1, then it will be impossible to form a linear combination that equals qby using elements of S. So no finite set can span P and no finite bases exist. Thus,dim (P ) = ∞. 4

Subsection RNMRank and Nullity of a Matrix

For any matrix, we have seen that we can associate several subspaces — the null space(Theorem NSMS [200]), the range (Theorem RMS [207]) and the row space (Theo-rem RSMS [208]). As vector spaces, each of these has a dimension, and for the nullspace and range, they are important enough to warrant names.

Definition NOMNullity Of a MatrixSuppose that A is an m× n matrix. Then the nullity of A is the dimension of the nullspace of A, n (A) = dim (N (A)). �

Definition ROMRank Of a MatrixSuppose that A is an m× n matrix. Then the rank of A is the dimension of the rangeof A, r (A) = dim (R (A)). �

Example D.RNMRank and nullity of a matrixLet’s compute the rank and nullity of

A =

2 −4 −1 3 2 1 −41 −2 0 0 4 0 1−2 4 1 0 −5 −4 −81 −2 1 1 6 1 −32 −4 −1 1 4 −2 −1−1 2 3 −1 6 3 −1

To do this, we will first row-reduce the matrix since that will help us determine bases forthe null space and range.

1 −2 0 0 4 0 1

0 0 1 0 3 0 −2

0 0 0 1 −1 0 −3

0 0 0 0 0 1 10 0 0 0 0 0 00 0 0 0 0 0 0

From this row-equivalent matrix in reduced row-echelon form we record D = {1, 3, 4, 6}and F = {2, 5, 7}.

For each index in D, Theorem BROC [131] creates a single basis vector. In total thebasis will have 4 vectors, so the range of A will have dimension 4 and we write r (A) = 4.

For each index in F , Theorem SSNS [90] and Theorem BNS [104] creates a single basisvector. In total the basis will have 3 vectors, so the null space of A will have dimension3 and we write n (A) = 3. 4

Page 270: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.RNNSM Rank and Nullity of a NonSingular Matrix 238

There were no accidents or coincidences in the previous example — with the row-reducedversion of a matrix in hand, the rank and nullity are easy to compute.

Theorem CRNComputing Rank and NullitySuppose that A is an m × n matrix and B is a row-equivalent matrix in reduced row-

echelon form with r nonzero rows. Then r (A) = r and n (A) = n− r. �

Proof Theorem BROC [131] provides a basis for the range by choosing columns of Athat correspond to the dependent variables in a description of the solutions to LS(A, 0).In the analysis of B, there is one dependent variable for each leading 1, one per nonzerorow. So there are r column vectors in a basis for R (A).

Theorem SSNS [90] and Theorem BNS [104] provide a basis for the null space bycreating basis vectors of the null space of A from entries of B, one for each independentvariable, one per column with out a leading 1. So there are n − r column vectors in abasis for n (A).

Every archetype (Chapter A [375]) that involves a matrix lists its rank and nullity. Youmay have noticed as you studied the archetypes that the larger the range is the smallerthe null space is. A simple corollary states this trade-off succinctly.

Theorem RPNCRank Plus Nullity is ColumnsSuppose that A is an m× n matrix. Then r (A) + n (A) = n. �

Proof Let r be the number of nonzero rows in a row-equivalent matrix in reduced row-echelon form. By Theorem CRN [238],

r (A) + n (A) = r + (n− r) = n �

When we first introduced r as our standard notation for the number of nonzero rows ina matrix in reduced row-echelon form you might have thought r stood for “rows.” Notreally — it stands for “rank”!

Subsection RNNSMRank and Nullity of a NonSingular Matrix

Let’s take a look at the rank and nullity of a square matrix.

Example D.RNSMRank and nullity of a square matrixThe matrix

E =

0 4 −1 2 2 3 12 −2 1 −1 0 −4 −3−2 −3 9 −3 9 −1 9−3 −4 9 4 −1 6 −2−3 −4 6 −2 5 9 −49 −3 8 −2 −4 2 48 2 2 9 3 0 9

Page 271: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.RNNSM Rank and Nullity of a NonSingular Matrix 239

is row-equivalent to the matrix in reduced row-echelon form,

1 0 0 0 0 0 0

0 1 0 0 0 0 0

0 0 1 0 0 0 0

0 0 0 1 0 0 0

0 0 0 0 1 0 0

0 0 0 0 0 1 0

0 0 0 0 0 0 1

With n = 7 columns and r = 7 nonzero rows Theorem CRN [238] tells us the rank isr (E) = 7 and the nullity is n (E) = 7− 7 = 0. 4

The value of either the nullity or the rank are enough to characterize a nonsingularmatrix.

Theorem RNNSMRank and Nullity of a NonSingular MatrixSuppose that A is a square matrix of size n. The following are equivalent.

1. A is nonsingular.

2. The rank of A is n, r (A) = n.

3. The nullity of A is zero, n (A) = 0. �

Proof (1 ⇒ 2) Theorem RNSM [139] says that if A is nonsingular then R (A) = Cn. IfR (A) = Cn, then the range has dimension n by Theorem DCM [235], so the rank of Ais n.(2 ⇒ 3) Suppose r (A) = n. Then Theorem RPNC [238] gives

n = r (A) + n (A) = n + n (A) ⇒ n (A) = 0

(3 ⇒ 1) Suppose n (A) = 0, so a basis for the null space of A is the empty set. Thisimplies that N (A) = {0} and Theorem NSTNS [59] says A is nonsingular. �

With a new equivalence for a nonsingular matrix, we can update our list of equivalences(Theorem NSME4 [178]) which now becomes a list requiring into double digits to number.

Theorem NSME6NonSingular Matrix Equivalences, Round 6Suppose that A is a square matrix of size n. The following are equivalent.

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the zero vector, N (A) = {0}.

4. The linear system LS(A, b) has a unique solution for every possible choice of b.

5. The columns of A are a linearly independent set.

Page 272: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.RNNSM Rank and Nullity of a NonSingular Matrix 240

6. The range of A is Cn, R (A) = Cn.

7. A is invertible.

8. The columns of A are a basis for Cn.

9. The rank of A is n, r (A) = n.

10. The nullity of A is zero, n (A) = 0. �

Page 273: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.EXC Exercises 241

Subsection EXCExercises

M20 Contributed by Robert BeezerM22 is the vector space of 2× 2 matrices. Let S22 denote the set of all 2× 2 symmetricmatrices. That is

S22 ={

A ∈ M22 | At = A}

(a) Show that S22 is a subspace of M22.(b) Exhibit a basis for S22 and prove that it has the required properties.(c) What is the dimension of S22? Solution [242]

Page 274: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection D.SOL Solutions 242

Subsection SOLSolutions

M20 Exercise [241] Contributed by Robert Beezer(a) We will use the three criteria of Theorem TSS [198]. The zero vector of M22 is thezero matrix, O (Definition ZM [124]), which is a symmetric matrix. So S22 is not empty,since O ∈ S22.

Suppose that A and B are two matrices in S22. Then we know that At = A andBt = B. We want to know if A + B ∈ S22, so test A + B for membership,

(A + B)t = At + Bt Theorem TASM [126]

= A + B A, B ∈ S22

So A + B is symmetric and qualifies for membership in S22.Suppose that A ∈ S22 and α ∈ C. Is αA ∈ S22? We know that At = A. Now check

that,

αAt = αAt Theorem TASM [126]

= αA A ∈ S22

So αA is also symmetric and qualifies for membership in S22.With the three criteria of Theorem TSS [198] fulfilled, we see that S22 is a subspace

of M22.

(b) An arbitrary matrix from S22 can be written as

[a bb d

]. We can express this

matrix as [a bb d

]=

[a 00 0

]+

[0 bb 0

]+

[0 00 d

]= a

[1 00 0

]+ b

[0 11 0

]+ d

[0 00 1

]this equation says that the set

T =

{[1 00 0

],

[0 11 0

],

[0 00 1

]}spans S22. Is it also linearly independent?

Write a relation of linear dependence on S,

O = a1

[1 00 0

]+ a2

[0 11 0

]+ a3

[0 00 1

][0 00 0

]=

[a1 a2

a2 a3

]The equality of these two matrices (Definition ME [121]) tells us that a1 = a2 = a3 = 0,and the only relation of linear dependence on T is trivial. So T is linearly independent,and hence is a basis of S22.

(c) The basis T found in part (b) has size 3. So by Definition D [230], dim (S22) = 3.

Page 275: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section PD Properties of Dimension 243

Section PD

Properties of Dimension

Once the dimension of a vector space is known, then the determination of whether ornot a set of vectors is linearly independent, or if it spans the vector space, can often bemuch easier. In this section we will state a workhorse theorem and then apply it to therange and row space of a matrix. It will also help us describe a super-basis for Cm.

Subsection GTGoldilocks’ Theorem

We begin with a useful theorem that we will need later, and in the proof of the maintheorem in this subsection. This theorem says that we can extend linearly independentsets, one vector at a time, by simply by adding vectors from outside the span of thelinearly independent set and the linear independence of the set is preserved.

Theorem ELISExtending Linearly Independent SetsSuppose V is vector space and S is a linearly independent set of vectors from V . Supposew is a vector such that w 6∈ Sp (S). Then the set S ′ = S∪{w} is linearly independent.�

Proof Suppose S = {v1, v2, v3, . . . , vm} and begin with a relation of linear dependenceon S ′,

a1v1 + a2v2 + a3v3 + · · ·+ amvm + am+1w = 0.

There are two cases to consider. First suppose that am+1 = 0. Then the relation of lineardependence on S ′ becomes

a1v1 + a2v2 + a3v3 + · · ·+ amvm = 0.

and by the linear independence of the set S, we conclude that a1 = a2 = a3 = · · · =am = 0. So all of the scalars in the relation of linear dependence on S ′ are zero.

In the second case, suppose that am+1 6= 0. Then the relation of linear dependenceon S ′ becomes

am+1w = −a1v1 − a2v2 − a3v3 − · · · − amvm

w = − a1

am+1

v1 −a2

am+1

v2 −a3

am+1

v3 − · · · −am

am+1

vm

This equation expresses w as a linear combination of the vectors in S, contrary to theassumption that w 6∈ Sp (S), so this case leads to a contradiction.

The first case yielded only a trivial relation of linear dependence on S ′ and the secondcase led to a contradiction. So S ′ is a linearly independent set since any relation of lineardependence is trivial. �

Page 276: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.GT Goldilocks’ Theorem 244

In the story Goldilocks and the Three Bears, the young girl Goldilocks visits the emptyhouse of the three bears while out walking in the woods. One bowl of porridge is toohot, the other too cold, the third is just right. One chair is too hard, one too soft, thethird is just right. So it is with sets of vectors — some are too big (linearly dependent),some are too small (they don’t span), and some are just right (bases). Here’s Goldilocks’Theorem.

Theorem GGoldilocksSuppose that V is a vector space of dimension t. Let S = {v1, v2, v3, . . . , vm} be a setof vectors from V . Then

1. If m > t, then S is linearly dependent.

2. If m < t, then S does not span V .

3. If m = t and S is linearly independent, then S spans V .

4. If m = t and S spans V , then S is linearly independent. �

Proof Let B be a basis of V . Since dim (V ) = t, Definition B [220] and Theorem BIS [234]imply that B is a linearly independent set of t vectors that spans V .

1. Suppose to the contrary that S is linearly independent. Then B is a smaller set ofvectors that spans V . This contradicts Theorem SSLD [230].

2. Suppose to the contrary that S does span V . Then B is a larger set of vectors thatis linearly independent. This contradicts Theorem SSLD [230].

3. Suppose to the contrary that S does not span V . Then we can choose a vector wsuch that w ∈ V and w 6∈ Sp (S). By Theorem ELIS [243], the set S ′ = S ∪ {w}is again linearly independent. Then S ′ is a set of m + 1 = t + 1 vectors that arelinearly independent, while B is a set of t vectors that span V . This contradictsTheorem SSLD [230].

4. Suppose to the contrary that S is linearly dependent. Then by Theorem DLDS [100](which can be upgraded, with no changes in the proof, to the setting of a generalvector space), there is a vector in S, say vk that is equal to a linear combination ofthe other vectors in S. Let S ′ = S \{vk}, the set of “other” vectors in S. Then it iseasy to show that V = Sp (S) = Sp (S ′). So S ′ is a set of m−1 = t−1 vectors thatspans V , while B is a set of t linearly independent vectors in V . This contradictsTheorem SSLD [230]. �

There is a tension in the construction of basis. Make the set too big and you will endup with relations of linear dependence among the vectors. Make the set too small andyou will not have enough raw material to span the entire vector space. Make it just theright size (the dimension) and you only need to have linear independence or spanning,and you get the other property for free. These roughly-stated ideas are made precise byTheorem G [244].

The structure and proof of this theorem also deserve comment. The hypotheses seeminnocuous. We presume we know the dimension of the vector space in hand, then we

Page 277: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.GT Goldilocks’ Theorem 245

mostly just look at the size of the set S. From this we get big conclusions about span-ning and linear independence. Each of the four proofs relies on ultimately contradictingTheorem DLDS [100], so in a way we could think of this entire theorem as a corollaryof Theorem DLDS [100]. The proofs of the third and fourth parts parallel each other instyle (add w, toss vk) and then turn on Theorem ELIS [243] and Theorem DLDS [100]before contradicting Theorem DLDS [100].

Theorem G [244] is useful in both concrete examples and as a tool in other proofs.We will use it often to bypass verifying linear independence or spanning.

Example PD.BPBases for Pn

In Example B.BP [221] we claimed that

B ={1, x, x2, x3, . . . , xn

}C =

{1, 1 + x, 1 + x + x2, 1 + x + x2 + x3, . . . , 1 + x + x2 + x3 + · · ·+ xn

}.

were both bases for Pn (Example VS.VSP [186]). Suppose we had first verified that Bwas a basis, so we would then know that dim (Pn) = n + 1. The size of C is n + 1, theright size to be a basis. We could then verify that C is linearly independent. We wouldnot have to make any special efforts to prove that C spans Pn, since Theorem G [244]would allow us to conclude this property of C directly. Then we would be able to saythat C is a basis of Pn also. 4Example PD.BSM22Basis for a subspace of M22

In Example D.DSM22 [235] we showed that

B =

{[−2 11 0

],

[−2 00 1

]}is a basis for the subspace Z of M22 (Example VS.VSM [186]) given by

Z =

{[a bc d

] ∣∣∣∣ 2a + b + 3c + 4d = 0, −a + 3b− 5c− d = 0

}This tells us that dim (Z) = 2. In this example we will find another basis. We canconstruct two new matrices in Z by forming linear combinations of the matrices in B.

2

[−2 11 0

]+ (−3)

[−2 00 1

]=

[2 22 −3

]3

[−2 11 0

]+

[−2 00 1

]=

[−8 33 1

]Then the set

C =

{[2 22 −3

],

[−8 33 1

]}has the right size to be a basis of Z. Let’s see if it is a linearly independent set. Therelation of linear dependence

a1

[2 22 −3

]+ a2

[−8 33 1

]= O[

2a1 − 8a2 2a1 + 3a2

2a1 + 3a2 −3a1 + a2

]=

[0 00 0

]

Page 278: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.RT Ranks and Transposes 246

leads to the homogeneous system of equations whose coefficient matrix2 −82 32 3−3 1

row-reduces to

1 0

0 10 00 0

So with a1 = a2 = 0 as the only solution, the set is linearly independent. Now we canapply Theorem G [244] to see that C also spans Z and therefore is a second basis forZ. 4

Example PD.SVP4Sets of vectors in P4

In Example B.BSP4 [221] we showed that

B ={x− 2, x2 − 4x + 4, x3 − 6x2 + 12x− 8, x4 − 8x3 + 24x2 − 32x + 16

}is a basis for W = {p(x) | p ∈ P4, p(2) = 0}. So dim (W ) = 4.

The set {3x2 − 5x− 2, 2x2 − 7x + 6, x3 − 2x2 + x− 2

}is a subset of W (check this) and it happens to be linearly independent (check this, too).However, by Theorem G [244] it cannot span W .

The set{3x2 − 5x− 2, 2x2 − 7x + 6, x3 − 2x2 + x− 2, −x4 + 2x3 + 5x2 − 10x, x4 − 16

}is another subset of W (check this) and Theorem G [244] tells us that it must be linearlydependent.

The set {x− 2, x2 − 2x, x3 − 2x2, x4 − 2x3

}4

is a third subset of W (check this) and is linearly independent (check this). Since it hasthe right size to be a basis, and is linearly independent, Theorem G [244] tells us that italso spans W , and therefore is a basis of W .

Subsection RTRanks and Transposes

With Theorem G [244] in our arsenal, we prove one of the most surprising theoremsabout matrices.

Page 279: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.RT Ranks and Transposes 247

Theorem RMRTRank of a Matrix is the Rank of the TransposeSuppose A is an m× n matrix. Then r (A) = r (At). �

Proof We will need a bit of notation before we really get rolling. Write the columns ofA as individual vectors, Aj, 1 ≤ j ≤ n, and write the rows of A as individual columnvectors Ri, 1 ≤ i ≤ m. Pictorially then,

A = [A1|A2|A3| . . . |An] At = [R1|R2|R3| . . . |Rm]

Let d denote the rank of A, i.e. d = r (A) = dim (R (A)). Let C = {v1, v2, v3, . . . , vd} ⊆Cm be a basis for R (A).

Every column of A is an element of R (A) and C is a spanning set for R (A), so theremust be scalars, bij, 1 ≤ i ≤ d, 1 ≤ j ≤ n such that

Aj = b1jv1 + b2jv2 + b3jv3 + · · ·+ bdjvd

Define V to be the m× d matrix whose columns are the vectors vi, 1 ≤ i ≤ d. Let B bethe d× n matrix whose entries are the scalars, bij. More precisely, [B]ij = bij, 1 ≤ i ≤ d,1 ≤ j ≤ n. Then the previous equation expresses column j of A as a linear combinationof the columns of V , where the scalars come from column j of B. Let Bj denote columnj of B and then we are in a position to use Definition MVP [151] to write

Aj = V Bj 1 ≤ j ≤ n

Since column j of A is the product of the matrix V with column j of B, Defini-tion MM [153] tells us that A = V B. We are now in position to do all of the hardwork in this proof — take the transpose of both sides of this equation.

A = V B

At = (V B)t

[R1|R2|R3| . . . |Rm] = BtV t Theorem MMT [160]

Column i of At is row i of A, and by Definition MM [153] this row (expressed as a columnvector) is equal to the product of the n × d matrix Bt with column i of V t. ApplyingDefinition MVP [151] we see that then each row of A is a linear combination of the dcolumns of Bt. This is the key observation in this proof.

Since each row of A is a linear combination of the d columns of Bt, and the rows of Aare a spanning set for the row space of A, rs (A), we can conclude that the d columns ofBt are a spanning set for rs (A). Now we need Theorem G [244] (or Theorem SSLD [230]).Any basis for the row space of A cannot have more vectors than the size of a spanningset (or it would be linearly dependent), so the dimension of the row space cannot exceedd. Using Definition RSM [141],

r(At)

= dim(R(At))

= dim (rs (A)) ≤ d = r (A)

so r (At) ≤ r (A).This relationship only assumed that A is a matrix. It would apply equally well to At,

in which case we use Theorem TASM [126] and write

r (A) = r((

At)t) ≤ r

(At).

These two inequalities together imply that r (A) = r (At). �

Page 280: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.OBC Orthonormal Bases and Coordinates 248

This says that the row space and the column space of a matrix have the same dimension,which may be surprising. It does not say that column space and the row space areidentical. Indeed, if the matrix is not square, then the sizes (number of slots) of thevectors in each space are different, so the sets are not even comparable.

It is not hard to construct by yourself examples of matrices that illustrate Theo-rem RMRT [247], since it applies equally well to any matrix. Grab a matrix, row-reduceit, count the nonzero rows or the leading 1’s. That’s the rank. Transpose the matrix, row-reduce that, count the nonzero rows or the leading 1’s. That’s the rank of the transpose.The theorem says the two will be equal. Here’s an example anyway.

Example PD.RRTIRank, rank of transpose, Archetype IArchetype I [414] has a 4× 7 coefficient matrix which row-reduces to

1 4 0 0 2 1 −3

0 0 1 0 1 −3 5

0 0 0 1 2 −6 60 0 0 0 0 0 0

so the rank is 3. Row-reducing the transpose yields

1 0 0 −317

0 1 0 127

0 0 1 137

0 0 0 00 0 0 00 0 0 00 0 0 0

.

demonstrating that the rank of the transpose is also 3. 4

Subsection OBCOrthonormal Bases and Coordinates

We learned about orthogonal sets of vectors in Cm back in Section O [??], and we alsolearned that they were automatically linearly independent (Theorem OSLI [116]). Whenthey also span a subspace of Cm, then they are bases. And when the set is orthonormal,then they are incredibly nice bases. We will back up this claim with a theorem, but firstconsider how you might manufacture such a set.

Suppose that W is a subspace of Cm with basis B. Then B spans W and is alinearly independent set of nonzero vectors. We can apply the Gram-Schmidt Pro-cedure (Theorem GSPCV [117]) and obtain a linearly independent set T such thatSp (T ) = Sp (B) = W and T is orthogonal. In other words, T is a basis for W , andis an orthogonal set. By scaling each vector of T to norm 1, we can convert T into anorthonormal set, without destroying the properties that make it a basis of W . In short,we can convert any basis into an orthonormal basis. Example O.GSTV [118], followedby Example O.ONTV [119], illustrates this process.

Orthogonal matrices (Definition OM [178]) are another good source of orthonormalbases (and vice versa). Suppose that Q is an orthogonal matrix of size n. Then the n

Page 281: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.OBC Orthonormal Bases and Coordinates 249

columns of Q form an orthonormal set (Theorem COMOS [179]) that is therefore lin-early independent (Theorem OSLI [116]). Since Q is invertible (Theorem OMI [179]), weknow Q is nonsingular (Theorem NSI [177]), and then the columns of Q span Cn (Theo-rem RNSM [139]). So the columns of an orthogonal matrix of size n are an orthonormalbasis for Cn.

Why all the fuss about orthonormal bases? Theorem VRRB [227] told us that anyvector in a vector space could be written, uniquely, as a linear combination of basisvectors. For an orthonormal basis, finding the scalars for this linear combination isextremely easy, and this is the content of the next theorem. Furthermore, with vectorswritten this way (as linear combinations of the elements of an orthonormal set) certaincomputations and analysis become much easier. Here’s the promised theorem.

Theorem COBCoordinates and Orthonormal BasesSuppose that B = {v1, v2, v3, . . . , vp} is an orthonormal basis of the subspace W ofCm. For any w ∈ W ,

w = 〈w, v1〉v1 + 〈w, v2〉v2 + 〈w, v3〉v3 + · · ·+ 〈w, vp〉vp �

Proof Because B is a basis of W , Theorem VRRB [227] tells us that we can write wuniquely as a linear combination of the vectors in B. So it is not this aspect of theconclusion that makes this theorem interesting. What is interesting is that the particularscalars are so easy to compute. No need to solve big systems of equations — just do aninner product of w with vi to arrive at the coefficient of vi in the linear combination.

So begin the proof by writing w as a linear combination of the vectors in B, usingunknown scalars,

w = a1v1 + a2v2 + a3v3 + · · ·+ apvp

and compute,

〈w, vi〉 =

⟨p∑

k=1

akvk, vi

=

p∑k=1

〈akvk, vi〉 Theorem IPVA [111]

=

p∑k=1

ak 〈vk, vi〉 Theorem IPSM [112]

= ai 〈vi, vi〉+∑k 6=i

ak 〈vk, vi〉 Isolate term with k = i

= ai(1) +∑k 6=i

ak(0) T orthonormal

= ai

So the (unique) scalars for the linear combination are indeed the inner products advertisedin the conclusion of the theorem’s statement. �

Example PD.CROB4Coordinatization relative to an orthonormal basis, C4

Page 282: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.OBC Orthonormal Bases and Coordinates 250

The set

{x1, x2, x3, x4} =

1 + i1

1− ii

,

1 + 5i6 + 5i−7− i1− 6i

,

−7 + 34i−8− 23i−10 + 22i30 + 13i

,

−2− 4i6 + i4 + 3i6− i

was proposed, and partially verified, as an orthogonal set in Example O.AOS [115]. Let’sscale each vector to norm 1, so as to form an orthonormal basis of C4. (Notice that byTheorem OSLI [116] the set is linearly independent. Since we know the dimension of C4

is 4, Theorem G [244] tells us the set is just the right size to be a basis of C4.) The normsof these vectors are,

‖x1‖ =√

6 ‖x2‖ =√

174 ‖x3‖ =√

3451 ‖x4‖ =√

119

So an orthonormal basis is

B = {v1, v2, v3, v4}

=

1√6

1 + i

11− i

i

,1√174

1 + 5i6 + 5i−7− i1− 6i

,1√3451

−7 + 34i−8− 23i−10 + 22i30 + 13i

,1√119

−2− 4i6 + i4 + 3i6− i

Now, choose any vector from C4, say w =

2−314

, and compute

〈w, v1〉 =−5i√

6, 〈w, v2〉 =

−19 + 30i√174

, 〈w, v3〉 =120− 211i√

3451, 〈w, v4〉 =

6 + 12i√119

then Theorem COB [249] guarantees that2−314

=−5i√

6

1√6

1 + i

11− i

i

+

−19 + 30i√174

1√174

1 + 5i6 + 5i−7− i1− 6i

+120− 211i√

3451

1√3451

−7 + 34i−8− 23i−10 + 22i30 + 13i

+

6 + 12i√119

1√119

−2− 4i6 + i4 + 3i6− i

as you might want to check (if you have unlimited patience). 4

A slightly less intimidating example follows, in three dimensions and with just real num-bers.

Example PD.CROB3Coordinatization relative to an orthonormal basis, C3

The set

{x1, x2, x3} =

1

21

,

−101

,

211

Page 283: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.OBC Orthonormal Bases and Coordinates 251

is a linearly independent set, which the Gram-Schmidt Process (Theorem GSPCV [117])converts to an orthogonal set, and which can then be converted to the orthonormal set,

{v1, v2, v3} =

1√6

121

,1√2

−101

,1√3

1−11

which is therefore an orthonormal basis of C3. With three vectors in C3, all with realnumber entries, the inner product (Definition IP [110]) reduces to the usual “dot product”(or scalar product) and the orthogonal pairs of vectors can be interpreted as perpendicularpairs of directions. So the vectors in B serve as replacements for our usual 3-D axes, orthe usual 3-D unit vectors ~i,~j and ~k. We would like to decompose arbitrary vectors into“components” in the directions of each of these basis vectors. Its Theorem COB [249]that tells us how to do this.

Suppose that we choose w =

2−15

. Compute

〈w, v1〉 =5√6

〈w, v2〉 =3√2

〈w, v3〉 =8√3

then Theorem COB [249] guarantees that 2−15

=5√6

1√6

121

+3√2

1√2

−101

+8√3

1√3

1−11

which you should be able to check easily, even if you do not have much patience. 4

Page 284: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.EXC Exercises 252

Subsection EXCExercises

T60 Contributed by Joe RiegseckerSuppose that W is a vector space with dimension 5, and U and V are subspaces of W ,

each of dimension 3. Prove that U ∩ V contains a non-zero vector. State a more generalresult. Solution [253]

TODO: Every subspace has a finite basisTODO: Theorem ELIS ”forever” in a finite-dim vsp?TODO: rank of mxn matrix less than min of m and nTODO: LS(A,b) consistent iff rank A = rank A—b

Page 285: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PD.SOL Solutions 253

Subsection SOLSolutions

T60 Exercise [252] Contributed by Robert BeezerLet {u1, u2, u3} and {v1, v2, v3} be bases for U and V (respectively). Then, the set{u1, u2, u3, v1, v2, v3} is linearly dependent, since Theorem G [244] says we cannot have6 linearly independent vectors in a vector space of dimension 5. So we can assert thatthere is a non-trivial relation of linear dependence,

a1u1 + a2u2 + a3u3 + b1v1 + b2v2 + b3v3 = 0

where a1, a2, a3 and b1, b2, b3 are not all zero.We can rearrange this equation as

a1u1 + a2u2 + a3u3 = −b1v1 − b2v2 − b3v3

This is an equality of two vectors, so we can give this common vector a name, say w,

w = a1u1 + a2u2 + a3u3 = −b1v1 − b2v2 − b3v3

This is the desired non-zero vector, as we will now show.First, since w = a1u1 + a2u2 + a3u3, we can see that w ∈ U . Similarly, w =

−b1v1 − b2v2 − b3v3, so w ∈ V . This establishes that w ∈ U ∩ V .Is w 6= 0? Suppose not, in other words, suppose w = 0. Then

0 = w = a1u1 + a2u2 + a3u3

Because {u1, u2, u3} is a basis for U , it is a linearly independent set and the relation oflinear dependence above means we must conclude that a1 = a2 = a3 = 0. By a similarprocess, we would conclude that b1 = b2 = b3 = 0. But this is a contradiction sincea1, a2, a3, b1, b2, b3 were chosen so that some were nonzero. So w 6= 0.

How does this generalize? All we really needed was the original relation of lineardependence that resulted because we had “too many” vectors in W . A more generalstatement would be: Suppose that W is a vector space with dimension n, U is a subspaceof dimension p and V is a subspace of dimension q. If p + q > n, then U ∩ V contains anon-zero vector.

Page 286: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

D: Determinants

Section DM

Determinants of Matrices

The determinant is a function that takes a square matrix as an input and produces ascalar as an output. So unlike a vector space, it is not an algebraic structure. However,it has many beneficial properties for studying vector spaces, matrices and systems ofequations, so it is hard to ignore (though some have tried). While the properties of adeterminant can be very useful, they are also complicated to prove. We’ll begin witha definition, do some computations and then establish some properties. The definitionof the determinant function is recursive, that is, the determinant of a large matrix isdefined in terms of the determinant of smaller matrices. To this end, we will make a fewdefinitions.

Definition SMSubMatrixSuppose that A is an m× n matrix. Then the submatrix Aij is the (m− 1)× (n− 1)matrix obtained from A by removing row i and column j. �

Example DM.SSSome submatricesFor the matrix

A =

1 −2 3 94 −2 0 13 5 2 1

we have the submatrices

A23 =

[1 −2 93 5 1

]A31

[−2 3 9−2 0 1

]4

Definition DMDeterminantSuppose A is a square matrix. Then its determinant, det (A) = |A|, is an element of Cdefined recursively by:

If A =[a]

is a 1× 1 matrix, then det (A) = a.

254

Page 287: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section DM Determinants of Matrices 255

If A = (aij) is a matrix of size n with n ≥ 2, then

det (A) = a11 det (A11)− a12 det (A12) + a13 det (A13)− · · ·+ (−1)n+1a1n det (A1n) �

So to compute the determinant of a 5×5 matrix we must build 5 submatrices, each of size4. To compute the determinants of each the 4×4 matrices we need to create 4 submatriceseach, these now of size 3 and so on. To compute the determinant of a 10×10 matrix wouldrequire computing the determinant of 10! = 10×9×8×7×6×5×4×3×2 = 3, 628, 8001×1 matrices. Fortunately there are better ways. However this does suggest an excellentcomputer programming exercise to write a recursive procedure to compute a determinant.

Lets compute the determinant of a reasonable sized matrix by hand.

Example DM.D33MDeterminant of a 3× 3 matrixSuppose that we have the 3× 3 matrix

A =

3 2 −14 1 6−3 −1 2

Then

det (A) = |A| =

∣∣∣∣∣∣3 2 −14 1 6−3 −1 2

∣∣∣∣∣∣= 3

∣∣∣∣ 1 6−1 2

∣∣∣∣− 2

∣∣∣∣ 4 6−3 2

∣∣∣∣+ (−1)

∣∣∣∣ 4 1−3 −1

∣∣∣∣= 3

(1∣∣2∣∣− 6

∣∣−1∣∣)− 2

(4∣∣2∣∣− 6

∣∣−3∣∣)− (4 ∣∣−1

∣∣− 1∣∣−3

∣∣)= 3 (1(2)− 6(−1))− 2 (4(2)− 6(−3))− (4(−1)− 1(−3))

= 24− 52 + 1

= −27 4

In practice it is a bit silly to decompose a 2 × 2 matrix down into a couple of 1 × 1matrices and then compute the exceedingly easy determinant of these puny matrices. Sohere is a simple theorem.

Theorem DMSTDeterminant of Matrices of Size Two

Suppose that A =

[a bc d

]. Then det (A) = ad− bc �

Proof Applying Definition DM [254],∣∣∣∣a bc d

∣∣∣∣ = a∣∣d∣∣− b

∣∣c∣∣ = ad− bc �

Do you recall seeing the expression ad− bc before? (Hint: Theorem TTMI [169])

Page 288: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection DM.CD Computing Determinants 256

Subsection CDComputing Determinants

For any given matrix, there are a variety of ways to compute the determinant, by “ex-panding” about any row or column. The determinants of the submatrices used in thesecomputations are used so often that they have their own names. The first is the deter-minant of a submatrix, the second differs only by a sign.

Definition MIMMinor In a MatrixSuppose A is an n × n matrix and Aij is the (n − 1) × (n − 1) submatrix formed byremoving row i and column j. Then the minor for A at location i j is the determinantof the submatrix, MA,ij = det (Aij). �

Definition CIMCofactor In a MatrixSuppose A is an n × n matrix and Aij is the (n − 1) × (n − 1) submatrix formed byremoving row i and column j. Then the cofactor for A at location i j is the signeddeterminant of the submatrix, CA,ij = (−1)i+j det (Aij). �

Example DM.MCMinors and cofactorsFor the matrix,

A =

2 4 2 1−1 2 3 −13 1 0 53 6 3 2

we have minors

MA,4,2 =

∣∣∣∣∣∣2 2 1−1 3 −13 0 5

∣∣∣∣∣∣ = 2(15)− 2(−2) + 1(−9) = 25

MA,3,4 =

∣∣∣∣∣∣2 4 2−1 2 33 6 3

∣∣∣∣∣∣ = 2(−12)− 4(−12) + 2(−12) = 0

and so two cofactors are

CA,4,2 = (−1)4+2MA,4,2 = (1)(25) = 25

CA,3,4 = (−1)3+4MA,3,4 = (−1)(0) = 0

A third cofactor is

CA,1,2 = (−1)1+2MA,1,2 = (−1)

∣∣∣∣∣∣−1 3 −13 0 53 3 2

∣∣∣∣∣∣= (−1)((−1)(−15)− (3)(−9) + (−1)(9)) = −33 4

Page 289: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection DM.CD Computing Determinants 257

With this notation in hand, we can state

Theorem DERCDeterminant Expansion about Rows and ColumnsSuppose that A = (aij) is a square matrix of size n. Then

det (A) = ai1CA,i1 + ai2CA,i2 + ai3CA,i3 + · · ·+ ainCA,in 1 ≤ i ≤ n

which is known as expansion about row i, and

det (A) = a1jCA,1j + a2jCA,2j + a3jCA,3j + · · ·+ anjCA,nj 1 ≤ j ≤ n

which is known as expansion about column j. �

Proof TODO �

That the determinant of an n×n matrix can be computed in 2n different (albeit similar)ways is nothing short of remarkable. For the doubters among us, we will do an example,computing a 4× 4 matrix in two different ways.

Example DM.TCSDTwo computations, same determinantLet

A =

−2 3 0 19 −2 0 11 3 −2 −14 1 2 6

Then expanding about the fourth row (Theorem DERC [257] with i = 4) yields,

|A| = (4)(−1)4+1

∣∣∣∣∣∣3 0 1−2 0 13 −2 −1

∣∣∣∣∣∣+ (1)(−1)4+2

∣∣∣∣∣∣−2 0 19 0 11 −2 −1

∣∣∣∣∣∣+ (2)(−1)4+3

∣∣∣∣∣∣−2 3 19 −2 11 3 −1

∣∣∣∣∣∣+ (6)(−1)4+4

∣∣∣∣∣∣−2 3 09 −2 01 3 −2

∣∣∣∣∣∣= (−4)(10) + (1)(−22) + (−2)(61) + 6(46) = 92

while expanding about column 3 (Theorem DERC [257] with j = 3) gives

|A| = (0)(−1)1+3

∣∣∣∣∣∣9 −2 11 3 −14 1 6

∣∣∣∣∣∣+ (0)(−1)2+3

∣∣∣∣∣∣−2 3 11 3 −14 1 6

∣∣∣∣∣∣+(−2)(−1)3+3

∣∣∣∣∣∣−2 3 19 −2 11 3 −1

∣∣∣∣∣∣+ (2)(−1)4+3

∣∣∣∣∣∣−2 3 19 −2 14 1 6

∣∣∣∣∣∣= 0 + 0 + (−2)(61) + (−2)(−107) = 92

Notice how much easier the second computation was. By choosing to expand about thethird column, we have two entries that are zero, so two 3× 3 determinants need not becomputed at all! 4

Page 290: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection DM.PD Properties of Determinants 258

When a matrix has all zeros above (or below) the diagonal, exploiting the zeros byexpanding about the proper row or column makes computing a determinant insanelyeasy.

Example DM.DUTMDeterminant of an upper-triangular matrixSuppose that T=

2 3 −1 3 30 −1 5 2 −10 0 3 9 20 0 0 −1 30 0 0 0 5

We will compute the determinant of this 5 × 5 matrix by consistently expanding aboutthe first column for each submatrix that aries and does not have a zero entry multiplyingit.

det (T ) =

∣∣∣∣∣∣∣∣∣∣2 3 −1 3 30 −1 5 2 −10 0 3 9 20 0 0 −1 30 0 0 0 5

∣∣∣∣∣∣∣∣∣∣= 2(−1)1+1

∣∣∣∣∣∣∣∣−1 5 2 −10 3 9 20 0 −1 30 0 0 5

∣∣∣∣∣∣∣∣= 2(−1)(−1)1+1

∣∣∣∣∣∣3 9 20 −1 30 0 5

∣∣∣∣∣∣= 2(−1)(3)(−1)1+1

∣∣∣∣−1 30 5

∣∣∣∣= 2(−1)(3)(−1)(−1)1+1

∣∣5∣∣= 2(−1)(3)(−1)(5) = 30 4

Subsection PDProperties of Determinants

The determinant is of some interest by itself, but it is of the greatest use when employed todetermine properties of matrices. To that end, we list some theorems here. Unfortunately,mostly without proof at the moment.

Theorem DTDeterminant of the TransposeSuppose that A is a square matrix. Then det (At) = det (A). �

Page 291: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection DM.PD Properties of Determinants 259

Proof We will prove this result by induction on the size of the matrix. For a matrix ofsize 1, the transpose and the matrix itself are equal, so matter what the definition of adeterminant might be, their determinants are equal.

Now suppose the theorem is true for matrices of size n− 1. By Theorem DERC [257]we can write the determinant as a product of entries from the first row with their cofactorsand then sum these products. These cofactors are signed determinants of matrices of sizen−1, which by the induction hypothesis, are equal to the determinant of their transposes,and commutativity in the sum in the exponent of −1 means the cofactor is equal to acofactor of the transpose.

det (A) = [A]11 CA,11 + [A]12 CA,12

+ [A]13 CA,13 + · · ·+ [A]1n CA,1n Theorem DERC [257], row 1

=[At]11

CA,11 +[At]21

CA,12

+[At]31

CA,13 + · · ·+[At]n1

CA,1n Definition TM [124]

=[At]11

CAt,11 +[At]21

CAt,21

+[At]31

CAt,31 + · · ·+[At]n1

CAt,n1 Induction hypothesis

= det(At)

Theorem DERC [257], column 1 �

Theorem DRMMDeterminant Respects Matrix MultiplicationSuppose that A and B are square matrices of size n. Then det (AB) = det (A) det (B).�

Proof TODO: �

Its an amazing thing that matrix multiplication and the determinant interact this way.Might it also be true that det (A + B) = det (A) + det (B)?

Theorem SMZDSingular Matrices have Zero DeterminantsLet A be a square matrix. Then A is singular if and only if det (A) = 0. �

Proof TODO: �

For the case of 2×2 matrices you might compare the application of Theorem SMZD [259]with the combination of the results stated in Theorem DMST [255] and Theorem TTMI [169].

Example DM.ZNDABZero and nonzero determinant, Archetypes A and BThe coefficient matrix in Archetype A [379] has a zero determinant (check this!) whilethe coefficient matrix Archetype B [384] has a nonzero determinant. These matrices aresingular and nonsingular, respectively. This is exactly what Theorem SMZD [259] says,and continues our list of contrasts between these two archetypes. 4

Since Theorem SMZD [259] is an equivalence (Technique E [40]) we can expand on ourgrowing list of equivalence about nonsingular matrices.

Theorem NSME7NonSingular Matrix Equivalences, Round 7Suppose that A is a square matrix of size n. The following are equivalent.

Page 292: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection DM.PD Properties of Determinants 260

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the zero vector, N (A) = {0}.

4. The linear system LS(A, b) has a unique solution for every possible choice of b.

5. The columns of A are a linearly independent set.

6. The range of A is Cn, R (A) = Cn.

7. A is invertible.

8. The columns of A are a basis for Cn.

9. The rank of A is n, r (A) = n.

10. The nullity of A is zero, n (A) = 0.

11. The determinant of A is nonzero, det (A) 6= 0. �

Proof Theorem SMZD [259] says A is singular if and only if det (A) = 0. If we negateeach of these statements, we arrive at two contapositives that we can combine as theequivalence, A is nonsingular if and only if det (A) 6= 0. This allows us to add a newstatement to the list. �

Computationally, row-reducing a matrix is the most efficient way to determine if a matrixis nonsingular, though the effect of using division in a computer can lead to round-off errors that confuse small quantities with critical zero quantities. Conceptually, thedeterminant may seem the most efficient way to determine if a matrix is nonsingular.The definition of a determinant uses just addition, subtraction and multiplication, sodivision is never a problem. And the final test is easy: is the determinant zero or not?However, the number of operations involved in computing a determinant very quicklybecomes so excessive as to be impractical.

Page 293: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

E: Eigenvalues

Section EE

Eigenvalues and Eigenvectors

When we have a square matrix of size n, A, and we multiply it by a vector from Cn, x,to form the matrix-vector product (Definition MVP [151]), the result is another vector inCn. So we can adopt a functional view of this computation — the act of mutiplying bya square matrix is a function that converts one vector (x) into another one (Ax) of thesame size. For some vectors, this seemingly complicated computation is really no morecomplicated than scalar multiplication. The vectors vary according to the choice of A, sothe question is to determine, for an individual choice of A, if there are any such vectors,and if so, which ones. It happens in a variety of situations that these vectors (and thescalars that go along with them) are of special interest.

Subsection EEMEigenvalues and Eigenvectors of a Matrix

Definition EEMEigenvalues and Eigenvectors of a MatrixSuppose that A is a square matrix of size n, x 6= 0 is a vector from Cn, and λ is a scalarfrom C such that

Ax = λx

Then we say x is an eigenvector of A with eigenvalue λ. �

Before going any further, perhaps we should convince you that such things ever happenat all. Believe the next example, but do not concern yourself with where the piecescome from. We will have methods soon enough to be able to discover these eigenvectorsourselves.

Example EE.SEESome eigenvalues and eigenvectors

261

Page 294: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.EEM Eigenvalues and Eigenvectors of a Matrix 262

Consider the matrix

A =

204 98 −26 −10−280 −134 36 14716 348 −90 −36−472 −232 60 28

and the vectors

x =

1−125

y =

−34−104

z =

−3708

w =

1−140

Then

Ax =

204 98 −26 −10−280 −134 36 14716 348 −90 −36−472 −232 60 28

1−125

=

4−4820

= 4

1−125

= 4x

so x is an eigenvector of A with eigenvalue λ = 4. Also,

Ay =

204 98 −26 −10−280 −134 36 14716 348 −90 −36−472 −232 60 28

−34−104

=

0000

= 0

−34−104

= 0y

so y is an eigenvector of A with eigenvalue λ = 0. Also,

Az =

204 98 −26 −10−280 −134 36 14716 348 −90 −36−472 −232 60 28

−3708

=

−614016

= 2

−3708

= 2z

so z is an eigenvector of A with eigenvalue λ = 2. Also,

Aw =

204 98 −26 −10−280 −134 36 14716 348 −90 −36−472 −232 60 28

1−140

=

2−280

= 2

1−140

= 2w

so w is an eigenvector of A with eigenvalue λ = 2.So we have demonstrated four eigenvectors of A. Are there more? Yes, any nonzero

scalar multiple of an eigenvector is again an eigenvector. In this example, set u = 30x.Then

Au = A(30x) Substitution

= 30Ax Distributivity

= 30(4x) x is an eigenvector of A

= 4(30x) Associativity, Commutativity

= 4u Substitution

so that u is also an eigenvector of A for the same eigenvalue, λ = 4.

Page 295: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.PM Polynomials and Matrices 263

The vectors z and w are both eigenvectors of A for the same eigenvalue λ = 2, yet thisis not as simple as the two vectors just being scalar multiples of each other (they aren’t).Look what happens when we add them together, to form v = z + w, and multiply by A,

Av = A(z + w) Substitution

= Az + Aw Distributivity

= 2z + 2w z, w eigenvectors of A

= 2(z + w) Distributivity

= 2v Substitution

so that v is also an eigenvector of A for the eigenvalue λ = 2. So it would appear that theset of eigenvectors that are associated with a fixed eigenvalue is closed under the vectorspace operations of Cn. Hmmm.

The vector y is an eigenvector of A for the eigenvalue λ = 0, which essentially is aresult of the fact that Ay = 0. But this also means that y ∈ N (A). There would appearto be a connection here. 4

Example EE.SEE [261] hints at a number of intriguing properties, and there are manymore. We will explore the general properties of eigenvlaues and eigenvectors in Sec-tion PEE [280], but for now we will concern ourselves with the question of actuallycomputing eigenvalues and eigenvectors. First we need a bit of background material onpolynomials and matrices.

Subsection PMPolynomials and Matrices

A polynomial is a combination of powers, multiplication by scalar coefficients, and ad-dition (with subtraction just being the inverse of addition). We never have ocassion todivide in a polynomial. So it is with matrices. We can add and subtract, we can multiplyby scalars, and we can form powers by repeated uses of matrix multiplication. We do notnormally divide matrices (though sometimes we can multiply by an inverse). If a matrixis square, all the operations of a polynomial will preserve the size of the matrix. We’lldemonstrate with an example,

Example EE.PMPolynomial of a matrixLet

p(x) = 14 + 19x− 3x2 − 7x3 + x4 D =

−1 3 21 0 −2−3 1 1

Page 296: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.PM Polynomials and Matrices 264

and we will compute p(D). First, the necessary powers of D. Notice that D0 is definedto be the multiplicative identity, I3, as will be the case in general.

D0 = I3 =

1 0 00 1 00 0 1

D1 = D =

−1 3 21 0 −2−3 1 1

D2 = DD1 =

−1 3 21 0 −2−3 1 1

−1 3 21 0 −2−3 1 1

=

−2 −1 −65 1 01 −8 −7

D3 = DD2 =

−1 3 21 0 −2−3 1 1

−2 −1 −65 1 01 −8 −7

=

19 −12 −8−4 15 812 −4 11

D4 = DD3 =

−1 3 21 0 −2−3 1 1

19 −12 −8−4 15 812 −4 11

=

−7 49 54−5 −4 −30−49 47 43

Then

p(D) = 14 + 19D − 3D2 − 7D3 + D4

= 14

1 0 00 1 00 0 1

+ 19

−1 3 21 0 −2−3 1 1

− 3

−2 −1 −65 1 01 −8 −7

− 7

19 −12 −8−4 15 812 −4 11

+

−7 49 54−5 −4 −30−49 47 43

=

−139 193 16627 −98 −124−193 118 20

Notice that p(x) factors as

p(x) = 14 + 19x− 3x2 − 7x3 + x4 = (x− 2)(x− 7)(x + 1)2

Because A commutes with itself (AA = AA), we can use distributivity of matrix multi-plication across matrix addition without being careful with any of the matrix products,and just as easily evaluate p(D) using the factored form of p(x),

p(x) = 14 + 19D − 3D2 − 7D3 + D4 = (D − 2I3)(D − 7I3)(D + I3)2

=

−3 3 21 −2 −2−3 1 −1

−8 3 21 −7 −2−3 1 −6

0 3 21 1 −2−3 1 2

2

=

−139 193 16627 −98 −124−193 118 20

Page 297: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.EEE Existence of Eigenvalues and Eigenvectors 265

This example is not meant to be too profound. It is meant to show you that it is naturalto evaluate a polynomial with a matrix, and that the factored form of the polynomial isas good as (or maybe better than) the expanded form. And do not forget that constantterms in polynomials are really multiples of the identity matrix when we are evaluatingthe polynomial with a matrix. 4

Subsection EEEExistence of Eigenvalues and Eigenvectors

Before we embark on computing eigenvalues and eigenvectors, we will prove that everymatrix has at least one eigenvalue (and an eigenector to go with it). Later, in Theo-rem MNEM [288], we will determine the maximum number of eigenvalues a matrix mayhave.

The determinant (Definition D [230]) will be a powerful tool in Subsection EE.CEE [268]when it comes time to compute eigenvalues. However, it is possible, with some more ad-vanced machinery, to compute eigenvalues without ever making use of the determinant.Sheldon Axler does just that in his book, Linear Algebra Done Right. Here and now, wegive Axler’s “determinant-free” proof that every matrix has an eigenvalue. The result isnot too startling, but the proof is most enjoyable.

Theorem EMHEEvery Matrix Has an EigenvalueSuppose A is a square matrix. Then A has at least one eigenvalue. �

Proof Suppose that A has size n, and choose x as any nonzero vector from Cn. (Noticehow much latitude we have in our choice of x. Only the zero vector is off-limits.) Considerthe set

S ={x, Ax, A2x, A3x, . . . , Anx

}This is a set of n+1 vectors from Cn, so by Theorem MVSLD [99], S is linearly dependent.Let a0, a1, a2, . . . , an be a collection of n + 1 scalars from C, not all zero, that providea relation of linear dependence on S. In other words,

a0x + a1Ax + a2A2x + a3A

3x + · · ·+ anAnx = 0

Some of the ai are nonzero. Suppose that just a0 6= 0, and a1 = a2 = a3 = · · · = an = 0.Then a0x = 0 and by Theorem SMEZV [193], either a0 = 0 or v = 0, which are bothcontradictions. So ai 6= 0 for some i ≥ 1. Let m be the largest integer such that am 6= 0.From this discussion we know that m ≥ 1. We can also assume that am = 1, for if not,replace each ai by ai/am to obtain scalars that serve equally well in providing a relationof linear dependence on S.

Define the polynomial

p(x) = a0 + a1x + a2x2 + a3x

3 + · · ·+ amxm

Because we have consistently used C as our set of scalars (rather than R), we know thatwe can factor p(x) into linear factors of the form (x − bi), where bi ∈ C. So there are

Page 298: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.EEE Existence of Eigenvalues and Eigenvectors 266

scalars, b1, b2, b3, . . . , bm, from C so that,

p(x) = (x− bm)(x− bm−1) · · · (x− b3)(x− b2)(x− b1)

Put it all together and

0 = a0x + a1Ax + a2A2x + a3A

3x + · · ·+ anAnx

= a0x + a1Ax + a2A2x + a3A

3x + · · ·+ amAmx ai = 0 for i > m

=(a0In + a1A + a2A

2 + a3A3 + · · ·+ amAm

)x Theorem MMDAA [157]

= p(A)x Definition of p(x)

= (A− bmIn)(A− bm−1In) · · · (A− b3In)(A− b2In)(A− b1In)x

Let k be the smallest integer such that

(A− bkIn)(A− bk−1In) · · · (A− b3In)(A− b2In)(A− b1In)x = 0.

From the preceding equation, we know that k ≤ m. Define the vector z by

z = (A− bk−1In) · · · (A− b3In)(A− b2In)(A− b1In)x

Notice that by the definition of k, the vector z must be nonzero. In the event that k = 1,we take z = x, and z is still nonzero. Now

(A− bkIn)z = (A− bkIn)(A− bk−1In) · · · (A− b3In)(A− b2In)(A− b1In)x = 0

which we can rearrange as

(A− bkIn)z = 0

Az− bkInz = 0 Theorem MMDAA [157]

Az− bkz = 0 Theorem MMIM [156]

Az = bkz

Since z 6= 0, this final line says that z is an eigenvector of A for the eigenvalue λ = bk,so A does have at least one eigenvalue. �

The proof of Theorem EMHE [265] is constructive (it contains an unambiguous procedurethat leads to an eigenvalue), but it is not meant to be practical. We will illustrate thetheorem with an example, the purpose being to provide a companion for studying theproof and not to suggest this is the best procedure for computing an eigenvalue.

Example EE.CAEHWComputing an eigenvalue the hard wayThis example illustrates the proof of Theorem EMHE [265], so will employ the samenotation as the proof — look there for full explanations. It is not meant to be anexample of a reasonable computational approach to finding eigenvalues and eigenvectors.OK, warnings in place, here we go.

Let

A =

−7 −1 11 0 −44 1 0 2 0−10 −1 14 0 −48 2 −15 −1 5−10 −1 16 0 −6

Page 299: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.EEE Existence of Eigenvalues and Eigenvectors 267

and choose

x =

303−54

It is important to notice that the choice of x could be anything, so long as it is not the zerovector. We have not chosen x totally at random, but so as to make our illustration of thetheorem as general as possible. You could replicate this example with your own choice andthe computations are guaranteed to be reasonable, provided you have a computationaltool that will factor a fifth degree polynomial for you.

The set

S ={x, Ax, A2x, A3x, A4x, A5x

}

=

303−54

,

−42−44−6

,

6−66−210

,

−1014−10−2−18

,

18−30181034

,

−3462−34−26−66

is guaranteed to be linearly dependent, as it has six vectors from C5 (Theorem MVSLD [99]).We will search for a non-trivial relation of linear dependence by solving a homogeneoussystem of equations whose coefficient matrix has the vectors of S as columns throughrow operations,

3 −4 6 −10 18 −340 2 −6 14 −30 623 −4 6 −10 18 −34−5 4 −2 −2 10 −264 −6 10 −18 34 −66

RREF−−−→

1 0 −2 6 −14 30

0 1 −3 7 −15 310 0 0 0 0 00 0 0 0 0 00 0 0 0 0 0

There are four free variables for describing solutions to this homogenous system, so wehave our pick of solutions. The most expedient choice would be to set x3 = 1 andx4 = x5 = x6 = 0. However, we will again opt to maximize the generality of ourillustration of Theorem EMHE [265] and choose x3 = −8, x4 = −3, x5 = 1 and x6 = 0.The leads to a solution with x1 = 16 and x2 = 12.

This relation of linear dependence then says that

0 = 16x + 12Ax− 8A2x− 3A3x + A4x + 0A5x

0 =(16 + 12A− 8A2 − 3A3 + A4

)x

So we define p(x) = 16 + 12x − 8x2 − 3x3 + x4, and as advertised in the proof ofTheorem EMHE [265], we have a polynomial of degree m = 4 > 1 such that p(A)x = 0.Now we need to factor p(x) over C. If you made your own choice of x at the start, thisis where you might have a fifth degree polynomial, and where you might need to use acomputational tool to find roots and factors. We have

p(x) = 16 + 12x− 8x2 − 3x3 + x4 = (x− 4)(x + 2)(x− 2)(x + 1)

Page 300: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.CEE Computing Eigenvalues and Eigenvectors 268

So we know that

0 = p(A)x = (A− 4I5)(A + 2I5)(A− 2I5)(A + 1I5)x

We apply one factor at a time, until we get the zero vector, so as to determine the valueof k described in the proof of Theorem EMHE [265],

(A + 1I5)x =

−6 −1 11 0 −44 2 0 2 0−10 −1 15 0 −48 2 −15 0 5−10 −1 16 0 −5

303−54

=

−12−1−1−2

(A− 2I5)(A + 1I5)x =

−9 −1 11 0 −44 −1 0 2 0−10 −1 12 0 −48 2 −15 −3 5−10 −1 16 0 −8

−12−1−1−2

=

4−8448

(A + 2I5)(A− 2I5)(A + 1I5)x =

−5 −1 11 0 −44 3 0 2 0−10 −1 16 0 −48 2 −15 1 5−10 −1 16 0 −4

4−8448

=

00000

So k = 3 and

z = (A− 2I5)(A + 1I5)x =

4−8448

is an eigenvector of A for the eigenvalue λ = −2, as you can check by doing the com-putation Az. If you work through this example with your own choice of the vector x(strongly recommended) then the eigenvalue you will find may be different, but will bein the set {3, 0, 1, −1, −2}. 4

Subsection CEEComputing Eigenvalues and Eigenvectors

Fortunately, we need not rely on the procedure of Theorem EMHE [265] each time weneed an eigenvalue. It is the determinant, and specifically Theorem SMZD [259], thatprovide the main tool. First a key definition.

Definition CPCharacteristic PolynomialSuppose that A is a square matrix of size n. Then the characteristic polynomial ofA is the polynomial pA (x) defined by

pA (x) = det (A− xIn) �

Page 301: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.CEE Computing Eigenvalues and Eigenvectors 269

Example EE.CPMS3Characteristic polynomial of a matrix, size 3Consider

F =

−13 −8 −412 7 424 16 7

Then

pF (x) = det (F − xI3)

=

∣∣∣∣∣∣−13− x −8 −4

12 7− x 424 16 7− x

∣∣∣∣∣∣ Definition CP [268]

= (−13− x)

∣∣∣∣7− x 416 7− x

∣∣∣∣+ (−8)(−1)

∣∣∣∣12 424 7− x

∣∣∣∣+ (−4)

∣∣∣∣12 7− x24 16

∣∣∣∣ Definition DM [254]

= (−13− x)((7− x)(7− x)− 4(16))

+ (−8)(−1)(12(7− x)− 4(24))

+ (−4)(12(16)− (7− x)(24)) Theorem DMST [255]

= 3 + 5x + x2 − x3

= −(x− 3)(x + 1)2 4

The characteristic polynomial is our main computational tool for finding eigenvalues, andwill sometimes be used to aid us in determining the properties of eigenvalues.

Theorem EMRCPEigenvalues of a Matrix are Roots of Characteristic PolynomialsSuppose A is a square matrix. Then λ is an eigenvalue of A if and only if pA (λ) = 0. �

Proof Suppose A has size n.

pA (λ) = 0 ⇐⇒ det (A− λIn) = 0 Definition CP [268]

⇐⇒ A− λIn is singular Theorem SMZD [259]

⇐⇒ there exists x 6= 0 so that (A− λIn)x = 0 Definition NM [56]

⇐⇒ Ax− λInx = 0 Theorem MMDAA [157]

⇐⇒ Ax− λx = 0 Theorem MMIM [156]

⇐⇒ Ax = λx, x 6= 0

⇐⇒ λ is an eigenvalue of A Definition EEM [261] �

Example EE.EMS3Eigenvalues of a matrix, size 3In Example EE.CPMS3 [269] we found the characteristic polynomial of

F =

−13 −8 −412 7 424 16 7

Page 302: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.CEE Computing Eigenvalues and Eigenvectors 270

to be pF (x) = −(x − 3)(x + 1)2. Factored, we can find all of its roots easily, they arex = 3 and x = −1. By Theorem EMRCP [269], λ = 3 and λ = −1 are both eigenvaluesof F , and these are the only eigenvalues of F . We’ve found them all. 4

Let us now turn our attention to the computation of eigenvectors.

Definition EMEigenspace of a MatrixSuppose that A is a square matrix and λ is an eigenvalue of A. Then the eigenspace ofA for λ, EA (λ), is the set of all the eigenvectors of A for λ, with the addition of the zerovector. �

Example EE.SEE [261] hinted that the set of eigenvectors for a single eigenvalue mighthave some closure properties, and with the addition of the non-eigenvector, 0, we indeedget a whole subspace.

Theorem EMSEigenspace for a Matrix is a SubspaceSuppose A is a square matrix of size n and λ is an eigenvalue of A. Then the eigenspaceEA (λ) is a subspace of the vector space Cn. �

Proof We will check the three conditions of Theorem TSS [198]. First, Definition EM [270]explicitly includes the zero vector in EA (λ), so the set is non-empty.

Suppose that x, y ∈ EA (λ), that is, x and y are two eigenvectors of A for λ. Then

A (x + y) = Ax + Ay Theorem MMDAA [157]

= λx + λy x, y eigenvectors of A

= λ (x + y) Distributivity

which says that x + y is an eigenvector of A for λ, so x + y ∈ EA (λ), and we haveadditive closure.

Suppose that α ∈ C, and that x ∈ EA (λ), that is, x is an eigenvector of A for λ.Then

A (αx) = α (Ax) Theorem MMSMM [157]

= αλx x an eigenvector of A

= λ (αx) Distributivity

which says that αx is an eigenvector of A for λ, so αx ∈ EA (λ), and we have scalarclosure.

With the three conditions of Theorem TSS [198] met, we know EA (λ) is a subspace.�

Theorem EMS [270] tells us that an eigenspace is a subspace (and hence a vector spacein its own right). Our next theorem tells us how to quickly construct this subspace.

Theorem EMNSEigenspace of a Matrix is a Null SpaceSuppose A is a square matrix of size n and λ is an eigenvalue of A. Then

EA (λ) = N (A− λIn) �

Page 303: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.CEE Computing Eigenvalues and Eigenvectors 271

Proof The conclusion of this theorem is an equality of sets, so normally we would followthe advice of Technique SE [14]. However, in this case we can construct a sequence ofequivalences which will together provide the two subset inclusions we need. First, noticethat 0 ∈ EA (λ) by Definition EM [270] and 0 ∈ N (A− λIn) by Theorem HSC [50]. Nowconsider a nonzero vector x ∈ Cn,

x ∈ EA (λ) ⇐⇒ Ax = λx Definition EM [270]

⇐⇒ Ax− λx = 0

⇐⇒ Ax− λInx = 0 Theorem MMIM [156]

⇐⇒ (A− λIn)x = 0 Theorem MMDAA [157]

⇐⇒ x ∈ N (A− λIn) Definition NSM [54] �

You might notice the close parallels (and differences) between the proofs of Theorem EM-RCP [269] and Theorem EMNS [271]. Since Theorem EMNS [271] describes the set of allthe eigenvectors of A as a null space we can use techniques such as Theorem SSNS [90]and Theorem BNS [104] to provide concise descriptions of eigenspaces.

Example EE.ESMS3Eigenspaces of a matrix, size 3Example EE.CPMS3 [269] and Example EE.EMS3 [270] describe the characteristic poly-nomial and eigenvalues of the 3× 3 matrix

F =

−13 −8 −412 7 424 16 7

We will now take the each eigenvalue in turn and compute its eigenspace. To do this,we row-reduce the matrix F − λI3 in order to determine solutions to the homogeneoussystem LS(F − λI3, 0) and then express the eigenspace as the null space of F − λI3

(Theorem EMNS [271]). Theorem SSNS [90] and Theorem BNS [104] then tell us howto write the null space as the span of a basis.

λ = 3 F − 3I3 =

−16 −8 −412 4 424 16 4

RREF−−−→

1 0 12

0 1 −12

0 0 0

EF (3) = N (F − 3I3) = Sp

−1

212

1

λ = −1 F + 1I3 =

−12 −8 −412 8 424 16 8

RREF−−−→

1 23

13

0 0 00 0 0

EF (−1) = N (F + 1I3) = Sp

−2

3

10

,

−13

01

Eigenspaces in hand, we can easily compute eigenvectors by forming nontrivial linearcombinations of the basis vectors describing each eigenspace. In particular, notice thatwe can “pretty up” our basis vectors by using scalar multiples to clear out fractions. 4

Page 304: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 272

Subsection ECEEExamples of Computing Eigenvalues and Eigenvectors

No theorems in this section, just a selection of examples meant to illustrate the rangeof possibilities for the eigenvalues and eigenvectors of a matrix. These examples canall be done by hand, though the computation of the characteristic polynomial wouldbe very time-consuming and error-prone. It can also be difficult to factor an arbitrarypolynomial, though if we were to suggest that most of our eigenvalues are going to beintegers, then it can be easier to hunt for roots. These examples are meant to looksimilar to a concatenation of Example EE.CPMS3 [269], Example EE.EMS3 [270] andExample EE.ESMS3 [271]. First, we will sneak in two similar definitions and illustratethem throughout this sequence of examples.

Definition AMEAlgebraic Multiplicity of an EigenvalueSuppose that A is a square matrix and λ is an eigenvalue of A. Then the algebraicmultiplicity of λ, αA (λ), is the highest power of (x− λ) that divides the characteristicpolynomial, pA (x). �

Since an eigenvalue λ is a root of the characteristic polynomial, there is always a factorof (x−λ), and the algebraic multiplicity is just the power of this factor in a factorizationof pA (x). So in particular, αA (λ) ≥ 1. Compare the definition of algebraic multiplicitywith the next definition.

Definition GMEGeometric Multiplicity of an EigenvalueSuppose that A is a square matrix and λ is an eigenvalue of A. Then the geometricmultiplicity of λ, γA (λ), is the dimension of the eigenspace EA (λ). �

Since every eigenvalue must have at least one eigenvector, the associated eigenspacecannot be trivial, and so γA (λ) ≥ 1.

Example EE.EMMS4Eigenvalue multiplicities, matrix of size 4Consider the matrix

B =

−2 1 −2 −412 1 4 96 5 −2 −43 −4 5 10

then

pB (x) = 8− 20x + 18x2 − 7x3 + x4 = (x− 1)(x− 2)3

So the eigenvalues are λ = 1, 2 with algebraic multiplicities αB (1) = 1 and αB (2) = 3.

Page 305: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 273

Computing eigenvectors,

λ = 1 B − 1I4 =

−3 1 −2 −412 0 4 96 5 −3 −43 −4 5 9

RREF−−−→

1 0 1

30

0 1 −1 0

0 0 0 10 0 0 0

EB (1) = N (B − 1I4) = Sp

−1

3

110

λ = 2 B − 2I4 =

−4 1 −2 −412 −1 4 96 5 −4 −43 −4 5 8

RREF−−−→

1 0 0 1/2

0 1 0 −1

0 0 1 1/20 0 0 0

EB (2) = N (B − 2I4) = Sp

−1

2

−1−1

2

1

So each eigenspace has dimension 1 and so γB (1) = 1 and γB (2) = 1. This example isof interest because of the discrepancy between the two multiplicities for λ = 2. In manyof our examples the algebraic and geometric multiplicities will be equal for all of theeigenvalues (as it was for λ = 1 in this example), so keep this example in mind. We willhave some explanations for this phenomenon later (see Example SD.NDMS4 [300]). 4

Example EE.ESMS4Eigenvalues, symmetric matrix of size 4Consider the matrix

C =

1 0 1 10 1 1 11 1 1 01 1 0 1

then

pC (x) = −3 + 4x + 2x2 − 4x3 + x4 = (x− 3)(x− 1)2(x + 1)

So the eigenvalues are λ = 3, 1, −1 with algebraic multiplicities αC (3) = 1, αC (1) = 2and αC (−1) = 1.

Page 306: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 274

Computing eigenvectors,

λ = 3 C − 3I4 =

−2 0 1 10 −2 1 11 1 −2 01 1 0 −2

RREF−−−→

1 0 0 −1

0 1 0 −1

0 0 1 −10 0 0 0

EC (3) = N (C − 3I4) = Sp

1111

λ = 1 C − 1I4 =

0 0 1 10 0 1 11 1 0 01 1 0 0

RREF−−−→

1 1 0 0

0 0 1 10 0 0 00 0 0 0

EC (1) = N (C − 1I4) = Sp

−1100

,

00−11

λ = −1 C + 1I4 =

2 0 1 10 2 1 11 1 2 01 1 0 2

RREF−−−→

1 0 0 1

0 1 0 1

0 0 1 −10 0 0 0

EC (−1) = N (C + 1I4) = Sp

−1−111

So the eigenspace dimensions yield geometric multiplicities γC (3) = 1, γC (1) = 2 andγC (−1) = 1, the same as for the algebraic multiplicities. This example is of interestbecause A is a symmetric matrix, and will be the subject of Theorem HMRE [288]. 4

Example EE.HMEM5High multiplicty eigenvalues, matrix of size 5Consider the matrix

E =

29 14 2 6 −9−47 −22 −1 −11 1319 10 5 4 −8−19 −10 −3 −2 87 4 3 1 −3

then

pE (x) = −16 + 16x + 8x2 − 16x3 + 7x4 − x5 = −(x− 2)4(x + 1)

So the eigenvalues are λ = 2, −1 with algebraic multiplicities αE (2) = 4 and αE (−1) = 1.

Page 307: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 275

Computing eigenvectors,

λ = 2 E − 2I5 =

27 14 2 6 −9−47 −24 −1 −11 1319 10 3 4 −8−19 −10 −3 −4 87 4 3 1 −5

RREF−−−→

1 0 0 1 0

0 1 0 −32−1

2

0 0 1 0 −10 0 0 0 00 0 0 0 0

EE (2) = N (E − 2I5) = Sp

−132

010

,

012

101

λ = −1 E + 1I5 =

30 14 2 6 −9−47 −21 −1 −11 1319 10 6 4 −8−19 −10 −3 −1 87 4 3 1 −2

RREF−−−→

1 0 0 2 0

0 1 0 −4 0

0 0 1 1 00 0 0 0 10 0 0 0 0

EE (−1) = N (E + 1I5) = Sp

−24−110

So the eigenspace dimensions yield geometric multiplicities γE (2) = 2 and γE (−1) = 1.4

Example EE.CEMS6Complex eigenvalues, matrix of size 6Consider the matrix

F =

−59 −34 41 12 25 301 7 −46 −36 −11 −29

−233 −119 58 −35 75 54157 81 −43 21 −51 −39−91 −48 32 −5 32 26209 107 −55 28 −69 −50

then

pF (x) = −50 + 55x + 13x2 − 50x3 + 32x4 − 9x5 + x6

= (x− 2)(x + 1)(x2 − 4x + 5)2

= (x− 2)(x + 1)((x− (2 + i))(x− (2− i)))2

= (x− 2)(x + 1)(x− (2 + i))2(x− (2− i))2

So the eigenvalues are λ = 2, −1, 2 + i, 2 − i with algebraic multiplicities αF (2) = 1,αF (−1) = 1, αF (2 + i) = 2 and αF (2− i) = 2.

Page 308: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 276

Computing eigenvectors,

λ = 2

F − 2I6 =

−61 −34 41 12 25 301 5 −46 −36 −11 −29

−233 −119 56 −35 75 54157 81 −43 19 −51 −39−91 −48 32 −5 30 26209 107 −55 28 −69 −52

RREF−−−→

1 0 0 0 0 15

0 1 0 0 0 0

0 0 1 0 0 35

0 0 0 1 0 −15

0 0 0 0 1 45

0 0 0 0 0 0

EF (2) = N (F − 2I6) = Sp

−1

5

0−3

515

−45

1

λ = −1

F + 1I6 =

−58 −34 41 12 25 301 8 −46 −36 −11 −29

−233 −119 59 −35 75 54157 81 −43 22 −51 −39−91 −48 32 −5 33 26209 107 −55 28 −69 −49

RREF−−−→

1 0 0 0 0 12

0 1 0 0 0 −32

0 0 1 0 0 12

0 0 0 1 0 0

0 0 0 0 1 −12

0 0 0 0 0 0

EF (−1) = N (F + I6) = Sp

−1

232

−12

012

1

Page 309: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 277

λ = 2 + i

F − (2 + i)I6 =

−61− i −34 41 12 25 30

1 5− i −46 −36 −11 −29−233 −119 56− i −35 75 54157 81 −43 19− i −51 −39−91 −48 32 −5 30− i 26209 107 −55 28 −69 −52− i

RREF−−−→

1 0 0 0 0 15(7 + i)

0 1 0 0 0 15(−9− 2i)

0 0 1 0 0 1

0 0 0 1 0 −1

0 0 0 0 1 10 0 0 0 0 0

EF (2 + i) = N (F − (2 + i)I6) = Sp

−1

5(7 + i)

15(9 + 2i)−11−11

λ = 2− i

F − (2− i)I6 =

−61 + i −34 41 12 25 30

1 5 + i −46 −36 −11 −29−233 −119 56 + i −35 75 54157 81 −43 19 + i −51 −39−91 −48 32 −5 30 + i 26209 107 −55 28 −69 −52 + i

RREF−−−→

1 0 0 0 0 15(7− i)

0 1 0 0 0 15(−9 + 2i)

0 0 1 0 0 1

0 0 0 1 0 −1

0 0 0 0 1 10 0 0 0 0 0

EF (2− i) = N (F − (2− i)I6) = Sp

15(−7 + i)

15(9− 2i)−11−11

So the eigenspace dimensions yield geometric multiplicities γF (2) = 1, γF (−1) = 1,γF (2 + i) = 1 and γF (2− i) = 1. This example demonstrates some of the possibili-ties for complex eigenvalues appearing, even when all the entries of the matrix are real.

Page 310: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 278

Notice how all the numbers in the analysis of λ = 2 − i are conjugates of the corre-sponding number in the analysis of λ = 2 + i. This is the content of the upcomingTheorem ERMCP [285]. 4

Example EE.DEMS5Distinct eigenvalues, matrix of size 5Consider the matrix

H =

15 18 −8 6 −55 3 1 −1 −30 −4 5 −4 −2−43 −46 17 −14 1526 30 −12 8 −10

then

pH (x) = −6x + x2 + 7x3 − x4 − x5 = x(x− 2)(x− 1)(x + 1)(x + 3)

So the eigenvalues are λ = 2, 1, 0, −1, −3 with algebraic multiplicities αH (2) = 1,αH (1) = 1, αH (0) = 1, αH (−1) = 1 and αH (−3) = 1.

Computing eigenvectors,

λ = 2 H − 2I5 =

13 18 −8 6 −55 1 1 −1 −30 −4 3 −4 −2−43 −46 17 −16 1526 30 −12 8 −12

RREF−−−→

1 0 0 0 −1

0 1 0 0 1

0 0 1 0 2

0 0 0 1 10 0 0 0 0

EH (2) = N (H − 2I5) = Sp

1−1−2−11

λ = 1 H − 1I5 =

14 18 −8 6 −55 2 1 −1 −30 −4 4 −4 −2−43 −46 17 −15 1526 30 −12 8 −11

RREF−−−→

1 0 0 0 −1

2

0 1 0 0 0

0 0 1 0 12

0 0 0 1 10 0 0 0 0

EH (1) = N (H − 1I5) = Sp

12

0−1

2

−11

Page 311: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection EE.ECEE Examples of Computing Eigenvalues and Eigenvectors 279

λ0 = H − 0I5 =

15 18 −8 6 −55 3 1 −1 −30 −4 5 −4 −2−43 −46 17 −14 1526 30 −12 8 −10

RREF−−−→

1 0 0 0 1

0 1 0 0 −2

0 0 1 0 −2

0 0 0 1 00 0 0 0 0

EH (0) = N (H − 0I5) = Sp

−12201

λ = −1 H + 1I5 =

16 18 −8 6 −55 4 1 −1 −30 −4 6 −4 −2−43 −46 17 −13 1526 30 −12 8 −9

RREF−−−→

1 0 0 0 −1/2

0 1 0 0 0

0 0 1 0 0

0 0 0 1 1/20 0 0 0 0

EH (−1) = N (H + 1I5) = Sp

12

00−1

2

1

λ = −3 H + 3I5 =

18 18 −8 6 −55 6 1 −1 −30 −4 8 −4 −2−43 −46 17 −11 1526 30 −12 8 −7

RREF−−−→

1 0 0 0 −1

0 1 0 0 12

0 0 1 0 1

0 0 0 1 20 0 0 0 0

EH (−3) = N (H + 3I5) = Sp

1−1

2

−1−21

So the eigenspace dimensions yield geometric multiplicities γH (2) = 1, γH (1) = 1,γH (0) = 1, γH (−1) = 1 and γH (−3) = 1, identical to the algebraic multiplicities. Thisexample is of interest for two reasons. First, λ = 0 is an eigenvalue, illustrating the up-coming Theorem SMZE [281]. Second, all the eigenvalues are distinct, yielding algebraicand geometric multiplicities of 1 for each eigenvalue, illustrating Theorem DED [300].4

Page 312: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section PEE Properties of Eigenvalues and Eigenvectors 280

Section PEE

Properties of Eigenvalues and Eigenvectors

The previous section introduced eigenvalues and eigenvectors, and concentrated on theirexistence and determination. This section will be more about theorems, and the variousproperties eigenvalues and eigenvectors enjoy. Like a good 4 × 100 meter relay, we willlead-off with one of our better theorems and save the very best for the anchor leg.

Theorem EDELIEigenvectors with Distinct Eigenvalues are Linearly IndependentSuppose that A is a square matrix and S = {x1, x2, x3, . . . , xp} is a set of eigenvectorswith eigenvalues λ1, λ2, λ3, . . . , λp such that λi 6= λj whenever i 6= j. Then S is alinearly independent set. �

Proof If p = 1, then the set S = {x1} is linearly independent since eigenvectors arenonzero (Definition EEM [261]), so assume for the remainder that p ≥ 2.

Suppose to the contrary that S is a linearly dependent set. Define k to be the smallestinteger, such that {x1, x2, x3, . . . , xk−1} is linearly independent and {x1, x2, x3, . . . , xk}is linearly dependent. Since eigenvectors are nonzero, the set {x1} is linearly independent,so k ≥ 2. Since we are assuming that S is linearly dependent, there is such a k and k ≤ p.So 2 ≤ k ≤ p.

Since {x1, x2, x3, . . . , xk} is linearly dependent there are scalars, a1, a2, a3, . . . , ak,some non-zero, so that

0 = a1x1 + a2x2 + a3x3 + · · ·+ akxk (∗)

In particular, we know that ak 6= 0, for if ak = 0, the scalars a1, a2, a3, . . . , ak−1 wouldinclude some nonzero values and would give a nontrivial relation of linear dependence on{x1, x2, x3, . . . , xk−1}, contradicting the linear dependence of {x1, x2, x3, . . . , xk−1}.Now

0 = A0 Theorem MMZM [156]

= A (a1x1 + a2x2 + a3x3 + · · ·+ akxk) Substitute (∗)= A(a1x1) + A(a2x2) + A(a3x3) + · · ·+ A(akxk) Theorem MMDAA [157]

= a1Ax1 + a2Ax2 + a3Ax3 + · · ·+ akAxk Theorem MMSMM [157]

= a1λ1x1 + a2λ2x2 + a3λ3x3 + · · ·+ akλkxk xi eigenvector of A for λi (∗∗)

Also,

0 = λk0 Theorem ZVSM [192]

= λk (a1x1 + a2x2 + a3x3 + · · ·+ akxk) Substitute (∗)= λka1x1 + λka2x2 + λka3x3 + · · ·+ λkakxk Distributivity (∗ ∗ ∗)

Page 313: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section PEE Properties of Eigenvalues and Eigenvectors 281

Put it all together,

0 = 0− 0 Property of zero vector

= (a1λ1x1 + a2λ2x2 + a3λ3x3 + · · ·+ akλkxp)

− (λka1x1 + λka2x2 + λka3x3 + · · ·+ λkakxk) Substitute (∗∗), (∗ ∗ ∗)= (a1λ1x1 − λka1x1) + (a2λ2x2 − λka2x2) + (a3λ3x3 − λka3x3)

+ · · ·+ (ak−1λk−1xk−1 − λkak−1xk−1) + (akλkxk − λkakxk) Commutativity

= a1 (λ1 − λk)x1 + a2 (λ2 − λk)x2 + a3 (λ3 − λk)x3

+ · · ·+ ak−1 (λk−1 − λk)xk−1 Distributivity

This is a relation of linear dependence on the linearly independent set {x1, x2, x3, . . . , xk−1},so the scalars must all be zero. That is, ai (λi − λk) = 0 for 1 ≤ i ≤ k − 1. However, theeigenvalues were assumed to be distinct, so λi 6= λk for 1 ≤ i ≤ k − 1. Thus ai = 0 for1 ≤ i ≤ k− 1. Earlier, we deduced that ak = 0 also. Now all these scalars are zero, whilethey were introduced as having some nonzero values. This is our desired contradiction,and therefore S is linearly independent. �

There is a simple connection between the eigenvalues of a matrix and whether or not itis nonsingular.

Theorem SMZESingular Matrices have Zero EigenvaluesSuppose A is a square matrix. Then A is singular if and only if λ = 0 is an eigenvalue ofA. �

Proof We have the following equivalences:

A is singular ⇐⇒ there exists x 6= 0, Ax = 0 Definition NSM [54]

⇐⇒ there exists x 6= 0, Ax = 0x Theorem MMZM [156]

⇐⇒ λ = 0 is an eigenvalue of A Definition EEM [261] �

With an equivalence about singular matrices we can update our list of equivalences aboutnonsingular matrices.

Theorem NSME8NonSingular Matrix Equivalences, Round 8Suppose that A is a square matrix of size n. The following are equivalent.

1. A is nonsingular.

2. A row-reduces to the identity matrix.

3. The null space of A contains only the zero vector, N (A) = {0}.

4. The linear system LS(A, b) has a unique solution for every possible choice of b.

5. The columns of A are a linearly independent set.

6. The range of A is Cn, R (A) = Cn.

7. A is invertible.

Page 314: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section PEE Properties of Eigenvalues and Eigenvectors 282

8. The columns of A are a basis for Cn.

9. The rank of A is n, r (A) = n.

10. The nullity of A is zero, n (A) = 0.

11. The determinant of A is nonzero, det (A) 6= 0.

12. λ = 0 is not an eigenvalue of A. �

Proof The equivalence of the first and last statements is the contrapositive of Theo-rem SMZE [281]. �

Certain changes to a matrix change its eigenvalues in a predictable way.

Theorem ESMMEigenvalues of a Scalar Multiple of a MatrixSuppose A is a square matrix and λ is an eigenvalue of A. Then αλ is an eigenvalue ofαA. �

Proof Let x 6= 0 be one eigenvector of A for λ. Then

(αA)x = α (Ax) Theorem MMSMM [157]

= α (λx) x eigenvector of A

= (αλ)x Associativity

So x 6= 0 is an eigenvector of αA for the eigenvalue αλ. �

Unfortunately, there are not parallel theorems about the sum or product of arbitrarymatrices. But we can prove a similar result for powers of a matrix.

Theorem EOMPEigenvalues Of Matrix PowersSuppose A is a square matrix, λ is an eigenvalue of A, and s ≥ 0 is an integer. Then λs

is an eigenvalue of As. �

Proof Let x 6= 0 be one eigenvector of A for λ. Suppose A has size n. Then we proceedby induction on s. First, for s = 0,

Asx = A0x

= Inx

= x Theorem MMIM [156]

= 1x Property of 1

= λ0x

= λsx

Page 315: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section PEE Properties of Eigenvalues and Eigenvectors 283

so λs is an eigenvalue of As in this special case. If we assume the theorem is true for s,then we find

As+1x = AsAx

= As (λx) x eigenvector of A for λ

= λ (Asx) Theorem MMSMM [157]

= λ (λsx) Induction hypothesis

= (λλs)x Associativity

= λs+1x

So x 6= 0 is an eigenvector of As+1 for λs+1, and induction (Technique XX [??]) tells usthe theorem is true for all s ≥ 0. �

While we cannot prove that the sum of two arbitrary matrices behaves in any reason-able way with regard to eigenvalues, we can work with the sum of dissimilar powersof the same matrix. We have already seen two connections between eigenvalues andpolynomials, in the proof of Theorem EMHE [265] and the characteristic polynomial(Definition CP [268]). Our next theorem strengthens this connection.

Theorem EPMEigenvalues of the Polynomial of a MatrixSuppose A is a square matrix and λ is an eigenvalue of A. Let q(x) be a polynomial inthe variable x. Then q(λ) is an eigenvalue of the matrix q(A). �

Proof Let x 6= 0 be one eigenvector of A for λ, and write q(x) = a0 + a1x + a2x2 + · · ·+

amxm. Then

q(A)x =(a0A

0 + a1A1 + a2A

2 + · · ·+ amAm)x

= (a0A0)x + (a1A

1)x + (a2A2)x + · · ·+ (amAm)x Theorem MMDAA [157]

= a0(A0x) + a1(A

1x) + a2(A2x) + · · ·+ am(Amx) Theorem MMSMM [157]

= a0(λ0x) + a1(λ

1x) + a2(λ2x) + · · ·+ am(λmx) Theorem EOMP [282]

= (a0λ0)x + (a1λ

1)x + (a2λ2)x + · · ·+ (amλm)x Associativity

=(a0λ

0 + a1λ1 + a2λ

2 + · · ·+ amλm)x Distributivity

= q(λ)x

So x 6= 0 is an eigenvector of q(A) for the eigenvalue q(λ). �

Example PEE.BDEBuilding desired eigenvaluesIn Example EE.ESMS4 [273] the 4× 4 symmetric matrix

C =

1 0 1 10 1 1 11 1 1 01 1 0 1

is shown to have the three eigenvalues λ = 3, 1, −1. Suppose we wanted a 4× 4 matrixthat has the three eigenvalues λ = 4, 0, −2. We can employ Theorem EPM [283] by

Page 316: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section PEE Properties of Eigenvalues and Eigenvectors 284

finding a polynomial that converts 3 to 4, 1 to 0, and −1 to −2. Such a polynomial iscalled an interpolating polynomial, and in this example we can use

r(x) =1

4x2 + x− 5

4

We will not discus how to concoct this polynomial, but a text on numerical analysisshould provide the details. In our case, simply verify that r(3) = 4, r(1) = 0 andr(−1) = −2.

Now compute

r(C) =1

4C2 + C − 5

4I4

=1

4

3 2 2 22 3 2 22 2 3 22 2 2 3

+

1 0 1 10 1 1 11 1 1 01 1 0 1

− 5

4

1 0 0 00 1 0 00 0 1 00 0 0 1

=1

2

1 1 3 31 1 3 33 3 1 13 3 1 1

Theorem EPM [283] tells us that if r(x) transforms the eigenvalues in the desired man-ner, then r(C) will have the desired eigenvalues. You can check this by computing theeigenvalues of r(C) directly. Furthermore, notice that the multiplicities are the same,and the eigenspaces of C and r(C) are identical. 4

Inverses and transposes also behave predictably with regard to their eigenvalues.

Theorem EIMEigenvalues of the Inverse of a MatrixSuppose A is a square nonsingular matrix and λ is an eigenvalue of A. Then 1

λis an

eigenvalue of the matrix A−1. �

Proof Notice that since A is assumed nonsingular, A−1 exists by Theorem NSI [177],but more importantly, 1

λdoes not invlove division by zero since Theorem SMZE [281]

prohibits this possibility.Let x 6= 0 be one eigenvector of A for λ. Suppose A has size n. Then

A−1x = A−1(1x) Property of 1

= A−1(1

λλx)

=1

λA−1(λx) Theorem MMSMM [157]

=1

λA−1(Ax) x eigenvector of A

=1

λ(A−1A)x Associativity

=1

λInx Definition MI [167]

=1

λx Theorem MMIM [156]

So x 6= 0 is an eigenvector of A−1 for the eigenvalue 1λ. �

Page 317: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section PEE Properties of Eigenvalues and Eigenvectors 285

The theorems above have a similar style to them, a style you should consider using whenconfronted with a need to prove a theorem about eigenvalues and eigenvectors. So farwe have been able to reserve the characteristic polynomial for strictly computationalpurposes. However, the next theorem, whose statement resembles the preceding theo-rems, has an easier proof if we employ the characteristic polynomial and results aboutdeterminants.

Theorem ETMEigenvalues of the Transpose of a MatrixSuppose A is a square matrix and λ is an eigenvalue of A. Then λ is an eigenvalue of thematrix At. �

Proof Let x 6= 0 be one eigenvector of A for λ. Suppose A has size n. Then

pA (x) = det (A− xIn) Definition CP [268]

= det((A− xIn)t) Theorem DT [258]

= det(At − (xIn)t) Theorem TASM [126]

= det(At − xI t

n

)Theorem TASM [126]

= det(At − xIn

)Definition IM [57]

= pAt (x) Definition CP [268]

So A and At have the same characteristic polynomial, and by Theorem EMRCP [269],their eigenvalues are identical and have equal algebraic multiplicities. Notice that whatwe have proved here is a bit stronger than the conclusion of the theorem. �

If a matrix has only real entries, then the computation of the characteristic polynomial(Definition CP [268]) will result in a polynomial with coefficients that are real numbers.Complex numbers could result as roots of this polynomial, but they are roots of quadraticfactors with real coefficients, and as such, come in conjugate pairs. The next theoremproves this, and a bit more, without mentioning the characteristic polynomial.

Theorem ERMCPEigenvalues of Real Matrices come in Conjugate PairsSuppose A is a square matrix with real entries and x is an eigenvector of A for theeigenvalue λ. Then x is an eigenvector of A for the eigenvalue λ. �

Proof

Ax = Ax A has real entries

= Ax Theorem MMCC [159]

= λx x eigenvector of A

= λx Theorem CCRSM [110]

So x is an eigenvector of A for the eigenvalue λ. �

This phenomenon is amply illustrated in Example EE.CEMS6 [275], where the fourcomplex eigenvalues come in two pairs, and the two basis vectors of the eigenspacesare complex conjugates of each other. Theorem ERMCP [285] can be a time-saver forcomputing eigenvalues and eigenvectors of real matrices with complex eigenvalues, sincethe conjugate eigenvalue and eigenspace can be inferred from the theorem rather thancomputed.

Page 318: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PEE.ME Multiplicities of Eigenvalues 286

Subsection MEMultiplicities of Eigenvalues

A polynomial of degree n will have exactly n roots. From this fact about polynomialequations we can say more about the algebraic multiplicities of eigenvalues.

Theorem DCPDegree of the Characteristic PolynomialSuppose that A is a square matrix of size n. Then the characteristic polynomial of A,pA (x), has degree n. �

Proof TODO: �

Theorem NEMNumber of Eigenvalues of a MatrixSuppose that A is a square matrix of size n with distinct eigenvalues λ1, λ2, λ3, . . . , λk.Then

k∑i=1

αA (λi) = n �

Proof By the definition of the algebraic multiplicity (Definition AME [272]), we canfactor the characteristic polynomial as

pA (x) = (x− λ1)αA(λ1)(x− λ2)

αA(λ2)(x− λ3)αA(λ3) · · · (x− λk)

αA(λk)

The left-hand side is a polynomial of degree n by Theorem DCP [286] and the rifgt-handside is a polynomial of degree

∑ki=1 αA (λi) = n, so the equality of the polynomials’

degrees gives the result. �

Theorem MEMultiplicities of an EigenvalueSuppose that A is a square matrix of size n and λ is an eigenvalue. Then

1 ≤ γA (λ) ≤ αA (λ) ≤ n �

Proof Since λ is an eigenvalue of A, there is an eigenvector of A for λ, x. Then x ∈EA (λ), so γA (λ) ≥ 1, since we can extend {x} into a basis of EA (λ).

To show that γA (λ) ≤ αA (λ) is the most involved portion of this proof. To thisend, let g = γA (λ) and let x1, x2, x3, . . . , xg be a basis for the eigenspace of λ, EA (λ).Construct another n− g vectors, y1, y2, y3, . . . , yn−g, so that

{x1, x2, x3, . . . , xg, y1, y2, y3, . . . , yn−g}

is a basis of Cn. This can be done by repeated applications of Theorem ELIS [243].Finally, define a matrix S by

S = [x1|x2|x3| . . . |xg|y1|y2|y3| . . . |yn−g] = [x1|x2|x3| . . . |xg|R]

Page 319: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PEE.ME Multiplicities of Eigenvalues 287

where R is an n× (n− g) matrix whose columns are y1, y2, y3, . . . , yn−g. The columnsof S are linearly independent by design, so S is nonsingular (Theorem NSLIC [103]) andtherefore invertible (Theorem NSI [177]). Then,

[e1|e2|e3| . . . |en] = In

= S−1S

= S−1[x1|x2|x3| . . . |xg|R]

= [S−1x1|S−1x2|S−1x3| . . . |S−1xg|S−1R]

So

S−1xi = ei 1 ≤ i ≤ g (∗)

Preparations in place, we compute the characteristic polynomial of A,

pA (x) = det (A− xIn) Definition CP [268]

= 1 det (A− xIn)

= det (In) det (A− xIn) Definition DM [254]

= det(S−1S

)det (A− xIn) Definition MI [167]

= det(S−1

)det (S) det (A− xIn) Theorem DRMM [259]

= det(S−1

)det (A− xIn) det (S) Commutativity in C

= det(S−1 (A− xIn) S

)Theorem DRMM [259]

= det(S−1AS − S−1xInS

)Theorem MMDAA [157]

= det(S−1AS − xS−1InS

)Theorem MMSMM [157]

= det(S−1AS − xS−1S

)Theorem MMIM [156]

= det(S−1AS − xIn

)Definition MI [167]

= pS−1AS (x) Definition CP [268]

What can we learn then about the matrix S−1AS?

S−1AS = S−1A[x1|x2|x3| . . . |xg|R]

= S−1[Ax1|Ax2|Ax3| . . . |Axg|AR] Definition MM [153]

= S−1[λx1|λx2|λx3| . . . |λxg|AR] xi eigenvectors of A

= [S−1λx1|S−1λx2|S−1λx3| . . . |S−1λxg|S−1AR] Definition MM [153]

= [λS−1x1|λS−1x2|λS−1x3| . . . |λS−1xg|S−1AR] Theorem MMSMM [157]

= [λe1|λe2|λe3| . . . |λeg|S−1AR] S−1S = In, ((∗ above)

Now imagine computing the characteristic polynomial of A by computing the character-istic polynomial of S−1AS using the form just obtained. The first g columns of S−1ASare all zero, save for a λ on the diagonal. So if we compute the determinant by expand-ing about the first column, successively, we will get succesive factors of (x − λ). Moreprecisely, let T be the square matrix of size n− g that is formed from the last n− g rowsof S−1AR. Then

pA (x) = pS−1AS (x) = (x− λ)gpT (x) .

Page 320: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PEE.EHM Eigenvalues of Hermitian Matrices 288

This says that (x− λ) is a factor of the characteristic polynomial at least g times, so thealgebraic multiplicity of λ as an eigenvalue of A is greater than or equal to g (Defini-tion AME [272]). In other words,

γA (λ) = g ≤ αA (λ)

as desired.

Theorem NEM [286] says that the sum of the algebraic multiplicities for all theeigenvalues of A is equal to n. Since the algebraic multiplicity is a positive quantity,no single algebraic multiplicity can exceed n without the sum of all of the algebraicmultiplicities doing the same. �

Theorem MNEMMaximum Number of Eigenvalues of a MatrixSuppose that A is a square matrix of size n. Then A cannot have more than n distincteigenvalues. �

Proof Suppose that A has k distinct eigenvalues, λ1, λ2, λ3, . . . , λk. Then

k =k∑

i=1

1

≤k∑

i=1

αA (λ) Theorem ME [286]

= n Theorem NEM [286] �

Subsection EHMEigenvalues of Hermitian Matrices

Recall that a matrix is Hermitian (or self-adjoint) if A =(A)t

(Definition HM [181]).In the case where A is a matrix whose entries are all real numbers, being Hermitian isidentical to being symmetric (Definition SYM [125]). Keep this in mind as you read thenext two theorems. Their hypotheses could be changed to “suppose A is a real symmetricmatrix.”

Theorem HMREHermitian Matrices have Real EigenvaluesSuppose that A is a Hermitian matrix and λ is an eigenvalue of A. Then λ ∈ R. �

Page 321: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PEE.EHM Eigenvalues of Hermitian Matrices 289

Proof Let x 6= 0 be one eigenvector of A for λ. Then

λ 〈x, x〉 = 〈λx, x〉 Theorem IPSM [112]

= 〈Ax, x〉 x eigenvector of A

= (Ax)t x Theorem MMIP [159]

= xtAtx Theorem MMT [160]

= xt((

A)t)t

x Definition HM [181]

= xtAx Theorem TASM [126]

= xtAx Theorem MMCC [159]

= 〈x, Ax〉 Theorem MMIP [159]

= 〈x, λx〉 x eigenvector of A

= λ 〈x, x〉 Theorem IPSM [112]

Since x 6= 0, Theorem PIP [114] says that 〈x, x〉 6= 0, so we can “cancel” 〈x, x〉 fromboth sides of this equality. This leaves λ = λ, so λ has a complex part equal to zero, andtherefore is a real number. �

Look back and compare Example EE.ESMS4 [273] and Example EE.CEMS6 [275]. InExample EE.CEMS6 [275] the matrix has only real entries, yet the characteristic poly-nomial has roots that are complex numbers, and so the matrix has complex eigenvalues.However, in Example EE.ESMS4 [273], the matrix has only real entries, but is alsosymmetric. So by Theorem HMRE [288], we were guaranteed eigenvalues that are realnumbers.

In many physical problems, a matrix of interest will be real and symmetric, or Her-mitian. Then if the eigenvalues are to represent physical quantities of interest, Theo-rem HMRE [288] guarantees that these values will not be complex numbers.

The eigenvectors of a Hermitian matrix also enjoy a pleasing property that we willexploit later.

Theorem HMOEHermitian Matrices have Orthogonal EigenvectorsSuppose that A is a Hermitian matrix and x and y are two eigenvectors of A for differenteigenvalues. Then x and y are orthogonal vectors. �

Proof Let x 6= 0 be an eigenvector of A for λ and let y 6= 0 be an eigenvector of A for

Page 322: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection PEE.EHM Eigenvalues of Hermitian Matrices 290

ρ. By Theorem HMRE [288], we know that ρ must be a real number. Then

λ 〈x, y〉 = 〈λx, y〉 Theorem IPSM [112]

= 〈Ax, y〉 x eigenvector of A

= (Ax)t y Theorem MMIP [159]

= xtAty Theorem MMT [160]

= xt((

A)t)t

y Definition HM [181]

= xtAy Theorem TASM [126]

= xtAy Theorem XX [??], conj. matrix mult

= 〈x, Ay〉 Theorem MMIP [159]

= 〈x, ρy〉 y eigenvector of A

= ρ 〈x, y〉 Theorem IPSM [112]

= ρ 〈x, y〉 Theorem HMRE [288]

Since λ 6= ρ, we conclude that 〈x, y〉 = 0 and so x and y are orthogonal vectors (Defini-tion OV [114]). �

Page 323: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section SD Similarity and Diagonalization 291

Section SD

Similarity and Diagonalization

This section’s topic will perhaps seem out of place at first, but we will make the connectionsoon with eigenvalues and eigenvectors. This is also our first look at one of the centralideas of Chapter R [358].

Subsection SMSimilar Matrices

The notion of matrices being “similar” is a lot like saying two matrices are row-equivalent.Two similar matrices are not equal, but they share many important properties. Thissection, and later sections in Chapter R [358] will be devoted in part to discovering justwhat these common properties are.

First, the main definition for this section.

Definition SIMSimilar MatricesSuppose A and B are two square matrices of size n. Then A and B are similar if thereexists a nonsingular matrix of size n, S, such that A = S−1BS. �

We will say “A is similar to B via S” when we want to emphasize the role of S in therelationship between A and B. Also, it doesn’t matter if we say A is similar to B, orB is similar to A. If one statement is true then so is the other, as can be seen by usingS−1 in place of S (see Theorem SER [293] for the careful proof). Finally, we will refer toS−1BS as a similarity transformation when we want to emphasize the way S changesB. OK, enough about language, lets build a few examples.

Example SD.SMS5Similar matrices of size 5If you wondered if there are examples of similar matrices, then it won’t be hard toconvince you they exist. Define

B =

−4 1 −3 −2 21 2 −1 3 −2−4 1 3 2 2−3 4 −2 −1 −33 1 −1 1 −4

S =

1 2 −1 1 10 1 −1 −2 −11 3 −1 1 1−2 −3 3 1 −21 3 −1 2 1

Page 324: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.PSM Properties of Similar Matrices 292

Check that S is nonsingular and then compute

A = S−1BS

=

10 1 0 2 −5−1 0 1 0 03 0 2 1 −30 0 −1 0 1−4 −1 1 −1 1

−4 1 −3 −2 21 2 −1 3 −2−4 1 3 2 2−3 4 −2 −1 −33 1 −1 1 −4

1 2 −1 1 10 1 −1 −2 −11 3 −1 1 1−2 −3 3 1 −21 3 −1 2 1

=

−10 −27 −29 −80 −25−2 6 6 10 −2−3 11 −9 −14 −9−1 −13 0 −10 −111 35 6 49 19

So by this construction, we know that A and B are similar. 4

Let’s do that again.

Example SD.SMS3Similar matrices of size 3Define

B =

−13 −8 −412 7 424 16 7

S =

1 1 2−2 −1 −31 −2 0

Check that S is nonsingular and then compute

A = S−1BS

=

−6 −4 −1−3 −2 −15 3 1

−13 −8 −412 7 424 16 7

1 1 2−2 −1 −31 −2 0

=

−1 0 00 3 00 0 −1

So by this construction, we know that A and B are similar. But before we move on, look athow pleasing the form of A is. Not convinced? Then consider that several computationsrelated to A are especially easy. For example, in the spirit of Example DM.DUTM [258],det (A) = (−1)(3)(−1) = 3. Similarly, the characteristic polynomial is straightforwardto compute by hand, pA (x) = (−1 − x)(3 − x)(−1 − x) = −(x − 3)(x + 1)2 and sincethe result is already factored, the eigenvalues are transparently λ = 3, −1. Finally, theeigenvectors of A are just the standard unit vectors (Definition SUV [169]). 4

Subsection PSMProperties of Similar Matrices

Similar matrices share many properties and it is these theorems that justify the choiceof the word “similar.” First we will show that similarity is an equivalence relation.

Page 325: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.PSM Properties of Similar Matrices 293

Equivalence relations are important in the study of various algebras and can always beregarded as a kind of weak version of equality. Sort of alike, but not quite equal. Thenotion of two matrices being row-equivalent is an example of an equivalence relationwe have been working with since the beginning of the course. Row-equivalent matricesare not equal, but they are a lot alike. For example, row-equivalent matrices have thesame rank. Formally, an equivalence relation requires three conditions hold: reflexive,symmetric and transitive. We will illustrate these as we prove that similarity is anequivalence relation.

Theorem SERSimilarity is an Equivalence RelationSuppose A, B and C are square matrices of size n. Then

1. A is similar to A. (Reflexive)

2. If A is similar to B, then B is similar to A. (Symmetric)

3. If A is similar to B and B is similar to C, then A is similar to C. (Transitive) �

Proof To see that A is similar to A, we need only demonstrate a nonsingular matrixthat effects a similarity transformation of A to A. In is nonsingular (since it row-reducesto the identity matrix, Theorem NSRRI [57]), and

I−1n AIn = InAIn = A

If we assume that A is similar to B, then we know there is a nonsingular matrix S sothat A = S−1BS by Definition SIM [291]. By Theorem MIMI [174], S−1 is invertible,and by Theorem NSI [177] is therefore nonsingular. So

(S−1)−1A(S−1) = SAS−1 Theorem MIMI [174]

= SS−1BSS−1 Substitution for A

=(SS−1

)B(SS−1

)Theorem MMA [158]

= InBIn Definition MI [167]

= B Theorem MMIM [156]

and we see that B is similar to A.Assume that A is similar to B, and B is similar to C. This gives us the existence

of two nonsingular matrices, S and R, such that A = S−1BS and B = R−1CR, byDefinition SIM [291]. (Notice how we have to assume S 6= R, as will usually be thecase.) Since S and R are invertible, so too RS is invertible by Theorem SS [174] andthen nonsingular by Theorem NSI [177]. Now

(RS)−1C(RS) = S−1R−1CRS Theorem SS [174]

= S−1(R−1CR

)S Theorem MMA [158]

= S−1BS Substitution of B

= A

so A is similar to C via the nonsingular matrix RS. �

Page 326: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.PSM Properties of Similar Matrices 294

Here’s another theorem that tells us exactly what sorts of properties similar matricesshare.

Theorem SMEESimilar Matrices have Equal EigenvaluesSuppose A and B are similar matrices. Then the characteristic polynomials of A and Bare equal, that is pA (x) = pB (x). �

Proof Suppose A and B have size n and are similar via the nonsingular matrix S, soA = S−1BS by Definition SIM [291].

pA (x) = det (A− xIn) Definition CP [268]

= det(S−1BS − xIn

)Substitution for A

= det(S−1BS − xS−1InS

)Theorem MMIM [156]

= det(S−1BS − S−1xInS

)Theorem MMSMM [157]

= det(S−1 (B − xIn) S

)Theorem MMDAA [157]

= det(S−1

)det (B − xIn) det (S) Theorem DRMM [259]

= det(S−1

)det (S) det (B − xIn) Commutativity in C

= det(S−1S

)det (B − xIn) Theorem DRMM [259]

= det (In) det (B − xIn) Definition MI [167]

= 1 det (B − xIn) Definition DM [254]

= pB (x) Definition CP [268] �

So similar matrices not only have the same set of eigenvalues, the algebraic multiplicitiesof these eigenvalues will also be the same. However, be careful with this theorem. Itis tempting to think the converse is true, and argue that if two matrices have the sameeigenvalues, then they are similar. Not so, as the following example illustrates.

Example SD.EENSEqual eigenvalues, not similarDefine

A =

[1 10 1

]B =

[1 00 1

]and check that

pA (x) = pB (x) = 1− 2x + x2 = (x− 1)2

and so A and B have equal characteristic polynomials. If the converse of Theorem SMEE [294]was true, then A and B would be similar. Suppose this was the case. In other words,there is a nonsingular matrix S so that A = S−1BS. Then

A = S−1BS = S−1I2S = S−1S = I2 6= A

this contradiction tells us that the converse of Theorem SMEE [294] is false. 4

Page 327: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.D Diagonalization 295

Subsection DDiagonalization

Good things happen when a matrix is similar to a diagonal matrix. For example, theeigenvalues of the matrix are the entries on the diagonal of the diagonal matrix. Andit can be a much simpler matter to compute high powers of the matrix. Diagonalizablematrices are also of interest in more abstract settings. Here are the relevant definitions,then our main theorem for this section.

Definition DIMDiagonal MatrixSuppose that A = (aij) is a square matrix. Then A is a diagonal matrix if aij = 0whenever i 6= j. �Definition DZMDiagonalizable MatrixSuppose A is a square matrix. Then A is diagonalizable if A is similar to a diagonalmatrix. �

Example SD.DABDiagonalization of Archetype BArchetype B [384] has a 3× 3 coefficient matrix

B =

−7 −6 −125 5 71 0 4

and is similar to a diagonal matrix, as can be seen by the following computation withthe nonsingular matrix S,

S−1BS =

−5 −3 −23 2 11 1 1

−1 −7 −6 −125 5 71 0 4

−5 −3 −23 2 11 1 1

=

−1 −1 −12 3 1−1 −2 1

−7 −6 −125 5 71 0 4

−5 −3 −23 2 11 1 1

=

−1 0 00 1 00 0 2

4

Example SD.SMS4 [??] provides another example of a matrix that is subjected to asimilarity transformation and the result is a diagonal matrix. Alright, just how wouldwe find the magic matrix S that can be used in a similarity transformation to produce adiagonal matrix? Before you read the statement of the next theorem, you might studythe eigenvalues and eigenvectors of Archetype B [384] and compute the eigenvalues andeigenvectors of the matrix in Example SD.SMS4 [??].

Theorem DCDiagonalization CharacterizationSuppose A is a square matrix of size n. Then A is diagonalizable if and only if thereexists a linearly independent set S that contains n eigenvectors of A. �

Page 328: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.D Diagonalization 296

Proof (⇒) Let S = {x1, x2, x3, . . . , xn} be a linearly independent set of eigenvectorsof A for the eigenvalues λ1, λ2, λ3, . . . , λn. Recall Definition SUV [169] and define

R = [x1|x2|x3| . . . |xn]

D =

λ1 0 0 · · · 00 λ2 0 · · · 00 0 λ3 · · · 0...

......

...0 0 0 · · · λn

= [λ1e1|λ2e2|λ3e3| . . . |λnen]

The columns of R are the vectors of the linearly independent set S and so by Theo-rem NSLIC [103] the matrix R is nonsingular. By Theorem NSI [177] we know R−1

exists.

R−1AR = R−1A [x1|x2|x3| . . . |xn]

= R−1[Ax1|Ax2|Ax3| . . . |Axn] Definition MM [153]

= R−1[λ1x1|λ2x2|λ3x3| . . . |λnxn] xi eigenvector of A for λi

= R−1[λ1Re1|λ2Re2|λ3Re3| . . . |λnRen] Definition MVP [151]

= R−1[R(λ1e1)|R(λ2e2)|R(λ3e3)| . . . |R(λnen)] Theorem MMSMM [157]

= R−1R[λ1e1|λ2e2|λ3e3| . . . |λnen] Definition MM [153]

= InD Definition MI [167]

= D Theorem MMIM [156]

This says that A is similar to the diagonal matrix D via the nonsingular matrix R. ThusA is diagonalizable (Definition DZM [295]).

(⇐) Suppose that A is diagonalizable, so there is a nonsingular matrix of size n

T = [y1|y2|y3| . . . |yn]

and a diagonal matrix (recall Definition SUV [169])

E =

d1 0 0 · · · 00 d2 0 · · · 00 0 d3 · · · 0...

......

...0 0 0 · · · dn

= [d1e1|d2e2|d3e3| . . . |dnen]

such that T−1AT = E.

[Ay1|Ay2|Ay3| . . . |Ayn] = A [y1|y2|y3| . . . |yn] Definition MM [153]

= AT

= InAT Theorem MMIM [156]

= TT−1AT Definition MI [167]

= TE Substitution

= T [d1e1|d2e2|d3e3| . . . |dnen]

= [T (d1e1)|T (d2e2)|T (d3e3)| . . . |T (dnen)] Definition MM [153]

= [d1Te1|d2Te2|d3Te3| . . . |dnTen] Definition MM [153]

= [d1y1|d2y2|d3y3| . . . |dnyn] Definition MVP [151]

Page 329: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.D Diagonalization 297

This equality of matrices allows us to conclude that the columns are equal vectors. Thatis, Ayi = diyi for 1 ≤ i ≤ n. In other words, yi is an eigenvector of A for the eigenvaluedi. (Why can’t yi = 0?). Because T is nonsingular, the set containing T ’s columns,S = {y1, y2, y3, . . . , yn}, is a linearly independent set (Theorem NSLIC [103]). So theset S has all the required properties. �

Notice that the proof of Theorem DC [295] is constructive. To diagonalize a matrix, weneed only locate n linearly independent eigenvectors. Then we can construct a nonsin-gular matrix using the eigenvectors as columns (R) so that R−1AR is a diagonal matrix(D). The entries on the diagonal of D will be the eigenvalues of the eigenvectors usedto create R, in the same order as the eigenvectors appear in R. We illustrate this bydiagonalizing some matrices.

Example SD.DMS3Diagonalizing a matrix of size 3Consider the matrix

F =

−13 −8 −412 7 424 16 7

of Example EE.CPMS3 [269], Example EE.EMS3 [270] and Example EE.ESMS3 [271].F ’s eigenvalues and eigenspaces are

λ = 3 EF (3) = Sp

−1

212

1

λ = −1 EF (−1) = Sp

−2

3

10

,

−13

01

Define the matrix S to be the 3× 3 matrix whose columns are the three basis vectors inthe eigenspaces for F ,

S =

−12−2

3−1

312

1 01 0 1

Check that S is nonsingular (row-reduces to the identity matrix, Theorem NSRRI [57]or has a nonzero determinant, Theorem SMZD [259]). Then the three columns of S area linearly independent set (Theorem NSLIC [103]). By Theorem DC [295] we now knowthat F is diagonalizable. Furthermore, the construction in the proof of Theorem DC [295]tells us that if we apply the matrix S to F in a similarity transformation, the result willbe a diagonal matrix with the eigenvalues of F on the diagonal. The eigenvalues appear

Page 330: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.D Diagonalization 298

on the diagonal of the matrix in the same order as the eigenvectors appear in S. So,

S−1FS =

−12−2

3−1

312

1 01 0 1

−1 −13 −8 −412 7 424 16 7

−12−2

3−1

312

1 01 0 1

=

6 4 2−3 −1 −1−6 −4 −1

−13 −8 −412 7 424 16 7

−12−2

3−1

312

1 01 0 1

=

3 0 00 −1 00 0 −1

Note that the above computations can be viewed two ways. The proof of Theorem DC [295]tells us that the four matrices (F , S, F−1 and the diagonal matrix) will interact the waywe have written the equation. Or as an example, we can actually perform the computa-tions to verify what the theorem predicts. 4

The dimension of an eigenspace can be no larger than the algebraic multiplicity of theeigenvalue by Theorem ME [286]. When every eigenvalue’s eigenspace is this big, thenwe can diagonalize the matrix, and only then. Three examples we have seen so far in thissection, Example SD.SMS5 [291], Example SD.DAB [295] and Example SD.DMS3 [297],illustrate the diagonalization of a matrix, with varying degrees of detail about just howthe diagonalization is achieved. However, in each case, you can verify that the geometricand algebraic multiplicities are equal for every eigenvalue. This is the substance of thenext theorem.

Theorem DMLEDiagonalizable Matrices have Large EigenspacesSuppose A is a square matrix. Then A is diagonalizable if and only if γA (λ) = αA (λ)for every eigenvalue λ of A. �

Proof Suppose A has size n and k distinct eigenvalues, λ1, λ2, λ3, . . . , λk.(⇐) Let Si =

{xi1, xi2, xi3, . . . , xiγA(λi)

}, be a basis for the eigenspace of λi, EA (λi),

1 ≤ i ≤ k. ThenS = S1 ∪ S2 ∪ S3 ∪ · · · ∪ Sk

is a set of eigenvectors for A. A vector cannot be an eigenvector for two different eigen-values (why not?) so the sets Si have no vectors in common. Thus the size of S is

k∑i=1

γA (λi) =k∑

i=1

αA (λi) Hypothesis

= n Theorem NEM [286]

We now want to show that S is a linearly independent set. So we will begin with arelation of linear dependence on S, using doubly-subscripted eigenvectors,

0 =(a11x11 + a12x12 + · · ·+ a1γA(λ1)x1γA(λ1)

)+(a21x21 + a22x22 + · · ·+ a2γA(λ2)x2γA(λ2)

)+ · · ·+

(ak1xk1 + ak2xk2 + · · ·+ akγA(λk)xkγA(λk)

)

Page 331: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.D Diagonalization 299

Define the vectors yi, 1 ≤ i ≤ k by

y1 =(a11x11 + a12x12 + a13x13 + · · ·+ aγA(1λ1)x1γA(λ1)

)y2 =

(a21x21 + a22x22 + a23x23 + · · ·+ aγA(2λ2)x2γA(λ2)

)y3 =

(a31x31 + a32x32 + a33x33 + · · ·+ aγA(3λ3)x3γA(λ3)

)...

yk =(ak1xk1 + ak2xk2 + ak3xk3 + · · ·+ aγA(kλk)xkγA(λk)

)Then the relation of linear dependence becomes

0 = y1 + y2 + y3 + · · ·+ yk

Since the eigenspace EA (λi) is closed under vector addition and scalar multiplication,yi ∈ EA (λi), 1 ≤ i ≤ k. Thus, for each i, the vector yi is an eigenvector of A forλi, or is the zero vector. Recall that sets of eigenvectors whose eigenvalues are distinctform a linearly independent set by Theorem EDELI [280]. Should any (or some) yi benonzero, the previous equation would provide a nontrivial relation of linear dependenceon a set of eigenvectors with distinct eigenvalues, contradicting Theorem EDELI [280].Thus yi = 0, 1 ≤ i ≤ k.

Each of the k equations, yi = 0 is a relation of linear dependence on the correspond-ing set Si, a set of basis vectors for the eigenspace EA (λi), which is therefore linearlyindependent. From these relations of linear dependence on linearly independent sets weconclude that aij = 0, 1 ≤ j ≤ γA (λi) for 1 ≤ i ≤ k. This establishes that our originalrelation of linear dependence on S has only the trivial solution, and hence S is a linearlyindependent set.

We have determined that S is a set of n linearly independent eigenvectors for A, andso by Theorem DC [295] is diagonalizable.

(⇒) Now we assume that A is diagonalizable. Aiming for a contradiction, supposethat there is at least one eigenvalue, say λt, such that γA (λt) 6= αA (λt). By Theo-rem ME [286] we must have γA (λt) < αA (λt), and γA (λi) ≤ αA (λi) for 1 ≤ i ≤ k,i 6= t.

Since A is diagonalizable, Theorem DC [295] guarantees a set of n linearly independentvectors, all of which are eigenvectors of A. Let ni denote the number of eigenvectors inS that are eigenvectors for λi, and recall that a vector cannot be an eigenvector for twodifferent eigenvalues. S is a linearly independent set, so the the subset Si containing theni eigenvectors for λi must also be linearly independent. Because the eigenspace EA (λi)has dimension γA (λi) and Si is a linearly independent subset in EA (λi), ni ≤ γA (λi),1 ≤ i ≤ k. Now,

n = n1 + n2 + n3 + · · ·+ nt + · · ·+ nk Size of S

≤ γA (λ1) + γA (λ2) + γA (λ3) + · · ·+ γA (λt) + · · ·+ γA (λk) Si linearly independent

< αA (λ1) + αA (λ2) + αA (λ3) + · · ·+ αA (λt) + · · ·+ αA (λk) Assumption about λt

= n Theorem NEM [286]

This is a contradiction (we can’t have n < n!) and so our assumption that someeigenspace had less than full dimension was false. �

Example EE.SEE [261], Example EE.CAEHW [266], Example EE.ESMS3 [271], Exam-ple EE.ESMS4 [273], Example EE.DEMS5 [278], Archetype B [384], Archetype F [401],

Page 332: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.D Diagonalization 300

Archetype K [423] and Archetype L [427] are all examples of matrices that are diagonal-izable and that illustrate Theorem DMLE [298]. While we have provide many examplesof matrices that are diagonalizable, especially among the archetypes, there are manymatrices that are not diagonalizable. Here’s one now.

Example SD.NDMS4A non-diagonalizable matrix of size 4In Example EE.EMMS4 [272] the matrix

B =

−2 1 −2 −412 1 4 96 5 −2 −43 −4 5 10

was determined to have characteristic polynomial

pB (x) = (x− 1)(x− 2)3

and an eigenspace for λ = 2 of

EB (2) = Sp

−1

2

−1−1

2

1

So the geometric multiplicity of λ = 2 is γB (2) = 1, while the algebraic multiplicity isαB (2) = 3. By Theorem DMLE [298], the matrix B is not diagonalizable. 4Archetype A [379] is the lone archetype with a square matrix that is not diagonaliz-able, as the algebraic and geometric multiplicities of the eigenvalue λ = 0 differ. Ex-ample EE.HMEM5 [274] is another example of a matrix that cannot be diagonalizeddue to the difference between the geometric and algebraic multiplicities of λ = 2, as isExample EE.CEMS6 [275] which has two complex eigenvalues, each with differing multi-plicities. Likewise, Example EE.EMMS4 [272] has an eigenvalue with different algebraicand geometric multiplicities and so cannot be diagonalized.

Theorem DEDDistinct Eigenvalues implies DiagonalizableSuppose A is a square matrix of size n with n distinct eigenvalues. Then A is diagonal-izable. �

Proof Let λ1, λ2, λ3, . . . , λn denote the n distinct eigenvalues of A. Then by Theo-rem NEM [286] we have n =

∑ni=1 αA (λi), which implies that αA (λi) = 1, 1 ≤ i ≤ n.

From Theorem ME [286] it follows that γA (λi) = 1, 1 ≤ i ≤ n. So γA (λi) = αA (λi),1 ≤ i ≤ n and Theorem DMLE [298] says A is diagonalizable. �

Example SD.DEHDDistinct eigenvalues, hence diagonalizableIn Example EE.DEMS5 [278] the matrix

H =

15 18 −8 6 −55 3 1 −1 −30 −4 5 −4 −2−43 −46 17 −14 1526 30 −12 8 −10

Page 333: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SD.OD Orthonormal Diagonalization 301

has characteristic polynomial

pH (x) = x(x− 2)(x− 1)(x + 1)(x + 3)

and so is a 5× 5 matrix with 5 distinct eigenvalues. By Theorem DED [300] we know Hmust be diagonalizable. But just for practice, we exhibit the diagonalization itself. Thematrix S contains eigenvectors of H as columns, one from each eigenspace, guaranteeinglinear independent columns and thus the nonsingularity of S. The diagonal matrix hasthe eigenvalues of H in the same order that their respective eigenvectors appear as thecolumns of S. Notice that some of the eigenvectors given in Example EE.DEMS5 [278]as basis vectors for the eigenspaces have been multiplied by suitable nonzero scalars toclear out fractions. These are of course still eigenvectors for the respective eigenvaluesby Theorem EMS [270].

S−1HS

=

2 1 −1 1 1−1 0 2 0 −1−2 0 2 −1 −2−4 −1 0 −2 −12 2 1 2 1

−1

15 18 −8 6 −55 3 1 −1 −30 −4 5 −4 −2−43 −46 17 −14 1526 30 −12 8 −10

2 1 −1 1 1−1 0 2 0 −1−2 0 2 −1 −2−4 −1 0 −2 −12 2 1 2 1

=

−3 −3 1 −1 1−1 −2 1 0 1−5 −4 1 −1 210 10 −3 2 −4−7 −6 1 −1 3

15 18 −8 6 −55 3 1 −1 −30 −4 5 −4 −2−43 −46 17 −14 1526 30 −12 8 −10

2 1 −1 1 1−1 0 2 0 −1−2 0 2 −1 −2−4 −1 0 −2 −12 2 1 2 1

=

−3 0 0 0 00 −1 0 0 00 0 0 0 00 0 0 1 00 0 0 0 2

4

Archetype B [384] is another example of a matrix that has as many distinct eigenvaluesas its size, and is hence diagonalizable by Theorem DED [300].

Powers of a diagonal matrix are easy to compute, and when a matrix is diagonalizable,it is almost as easy. We could state a theorem here perhaps, but we will settle insteadfor an example that makes the point just as well.

TODO: a computational example of this

Subsection ODOrthonormal Diagonalization

Theorem ODHMOrthonormal Diagonalization of Hermitian MatricesSuppose that A is a Hermitian matrix of size n. Then A can be diagonalized by asimilarity transformation using an orthonormal transformation. �

Proof TODO: �

Page 334: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

LT: Linear Transformations

Section LT

Linear Transformations

In the next linear algebra course you take, the first lecture might be a reminder aboutwhat a vector space is (Definition VS [184]), the ten axioms, basic properties and thensome examples. The second lecture would likely be all about linear transformations.While it may seem we have waited a long time to present what must be a central topic,in truth we have already been working with linear transformations for some time.

Functions are important objects in the study of calculus, but have been absent fromthis course until now (well, not really, it just seems that way). In your study of moreadvanced mathematics it is nearly impossible to escape the use of functions — they areas fundamental as sets are.

Subsection LTLinear Transformations

Here’s a key definition.

Definition LTLinear TransformationA linear transformation, T : U 7→ V , is a function that carries elements of the vectorspace U (called the domain) to the vector space V (called the codomain), and whichhas two additional properties

1. T (u1 + u2) = T (u1) + T (u2) for all u1, u2 ∈ V

2. T (αu) = αT (u) for all u ∈ V and all α ∈ C �

The two defining conditions of the definition of a linear transformations should “feellinear,” whatever that means. Conversely, these two conditions could be taken as aexactly what it means to be linear. As every vector space property derives from vectoraddition and scalar multiplication, so too, every property of a linear transformationderives from these two defining properties. While these conditions may be reminiscentof how we test subspaces, they really are quite different, so do not confuse the two.

302

Page 335: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.LT Linear Transformations 303

Here are two diagrams that convey the essence of the two defining properties of alinear transformation. In each case, begin in the upper left-hand corner, and follow thearrows around the rectangle to the lower-right hand corner, taking two different routesand doing the indicated operations labeled on the arrows. There are two results there.For a linear transformation these two expressions are always equal.

u1, u2T−−−→ T (u1) , T (u2)

+

y y+

u1 + u2T−−−→ T (u1) + T (u2),

T (u1 + u2)

uT−−−→ T (u)

α

y yα

αuT−−−→ αT (u),

T (αu)

A couple of words about notation. T is the name of the linear transformation, andshould be used when we want to discuss the function as a whole. T (u) is how we talkabout the output of the function, it is a vector in the vector space V . When we writeT (x + y) = T (x) + T (y), the plus sign on the left is the operation of vector addition inthe vector space U , since x and y are elements of U . The plus sign on the right is theoperation of vector addition in the vector space V , since T (x) and T (y) are elements ofthe vector space V . These two instances of vector addition might be wildly different.

Let’s examine several examples and begin to form a catalog of known linear transfor-mations to work with.

Example LT.ALTA linear transformationDefine T : C3 7→ C2 by describing the output of the function for a generic input withthe formula

T

x1

x2

x3

=

[2x1 + x3

−4x2

]

Page 336: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.LT Linear Transformations 304

and check the two defining properties.

T (x + y) = T

x1

x2

x3

+

y1

y2

y3

= T

x1 + y1

x2 + y2

x3 + y3

=

[2(x1 + y1) + (x3 + y3)

−4(x2 + y2)

]=

[(2x1 + x3) + (2y1 + y3)

−4x2 + (−4)y2)

]=

[2x1 + x3

−4x2

]+

[2y1 + y3

−4y2

]

= T

x1

x2

x3

+ T

y1

y2

y3

= T (x) + T (y)

and

T (αx) = T

α

x1

x2

x3

= T

αx1

αx2

αx3

=

[2(αx1) + (αx3)−4(αx2)

]=

[α(2x1 + x3)

α(−4x2)

]= α

[2x1 + x3

−4x2

]

= αT

x1

x2

x3

= αT (x)

So by Definition LT [302], T is a linear transformation. 4

It can be just as instructive to look at functions that are not linear transformations.Since the defining conditions must be true for all vectors and scalars, it is enough to findjust one situation where the properties fail.

Example LT.NLTNot a linear transformation

Page 337: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.LT Linear Transformations 305

Define S : C3 7→ C3 by

S

x1

x2

x3

=

4x1 + 2x2

0x1 + 3x3 − 2

This function “looks” linear, but consider

3 S

123

= 3

808

=

24024

while

S

3

123

= S

3

369

=

24028

So the second required property fails for the choice of α = 3 and x =

123

and by

Definition LT [302], S is not a linear transformation. It is just about as easy to find anexample where the first defining property fails (try it!). Notice that it is the “-2” in thethird component of the definition of S that prevents the function from being a lineartransformation. 4

Example LT.LTPMLinear transformation, polynomials to matricesDefine a linear transformation T : P3 7→ M22 by

T(a + bx + cx2 + dx3

)=

[a + b a− 2c

d b− d

]

T (x + y) = T((a1 + b1x + c1x

2 + d1x3) + (a2 + b2x + c2x

2 + d2x3))

= T((a1 + a2) + (b1 + b2)x + (c1 + c2)x

2 + (d1 + d2)x3))

=

[(a1 + a2) + (b1 + b2) (a1 + a2)− 2(c1 + c2)

d1 + d2 (b1 + b2)− (d1 + d2)

]=

[(a1 + b1) + (a2 + b2) (a1 − 2c1) + (a2 − 2c2)

d1 + d2 (b1 − d1) + (b2 − d2)

]=

[a1 + b1 a1 − 2c1

d1 b1 − d1

]+

[a2 + b2 a2 − 2c2

d2 b2 − d2

]= T

(a1 + b1x + c1x

2 + d1x3)

+ T(a2 + b2x + c2x

2 + d2x3)

= T (x) + T (y)

Page 338: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.LT Linear Transformations 306

and

T (αx) = T(α(a + bx + cx2 + dx3)

)= T

((αa) + (αb)x + (αc)x2 + (αd)x3

)=

[(αa) + (αb) (αa)− 2(αc)

αd (αb)− (αd)

]=

[α(a + b) α(a− 2c)

αd α(b− d)

]= α

[a + b a− 2c

d b− d

]= αT

(a + bx + cx2 + dx3

)= αT (x)

So by Definition LT [302], T is a linear transformation. 4

Example LT.LTPPLinear transformation, polynomials to polynomialsDefine a function S : P4 7→ P5 by

S(p(x)) = (x− 2)p(x)

Then

S (p(x) + q(x)) = (x− 2)(p(x) + q(x)) = (x− 2)p(x) + (x− 2)q(x) = S (p(x)) + S (q(x))

S (αp(x)) = (x− 2)(αp(x)) = (x− 2)αp(x) = α(x− 2)p(x) = αS (p(x))

So by Definition LT [302], S is a linear transformation. 4

Linear transformations have many amazing properties, which we will investigate throughthe next few sections. However, as a taste of things to come, here is a theorem we canprove now and put to use immediately.

Theorem LTTZZLinear Transformations Take Zero to ZeroSuppose T : U 7→ V is a linear transformation. Then T (0) = 0. �

Proof The two zero vectors in the conclusion of the theorem are different. The first isfrom U while the second is from V . We will subscript the zero vectors throughout thisproof to highlight the distinction. Think about your objects.

T (0U) = T (0U + 0U) Zero vector in U

T (0U) + 0V = T (0U + 0U) Zero vector in V

T (0U) + 0V = T (0U) + T (0U) Definition LT [302]

0V = T (0U) Theorem VAC [194] in V �

Return to Example LT.NLT [304] and compute S

000

=

00−2

to quickly see again

that S is not a linear transformation, while in Example LT.LTPM [305] and compute

S (0 + 0x + 0x2 + 0x3) =

[0 00 0

]as an example of Theorem LTTZZ [306] at work.

Page 339: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.MLT Matrices and Linear Transformations 307

Subsection MLTMatrices and Linear Transformations

If you give me a matrix, then I can quickly build you a linear transformation. Always.First a motivating example and then the theorem.

Example LT.LTMLinear transformation from a matrixLet

A =

3 −1 8 12 0 5 −21 1 3 −7

and define a function P : C4 7→ C3 by

P (x) = Ax

So we are using an old friend, the matrix-vector product (Definition MVP [151]) as away to convert a vector with 4 components into a vector with 3 components. ApplyingDefinition MVP [151] allows us to write the defining formula for P in a slightly differentform,

P (x) = Ax =

3 −1 8 12 0 5 −21 1 3 −7

x1

x2

x3

x4

= x1

321

+ x2

−101

+ x3

853

+ x4

1−2−7

So we recognize the action of the function P as using the components of the vector(x1, x2, x3, x4) as scalars to form the output of P as a linear combination of the fourcolumns of the matrix A, which are all members of C3, so the result is a vector inC3. We can rearrange this expression further, using our definitions of operations in C3,Definition CVSM [68] and Definition CVA [67].

P (x) = Ax

= x1

321

+ x2

−101

+ x3

853

+ x4

1−2−7

=

3x1

2x1

x1

+

−x2

0x2

+

8x3

5x3

3x3

+

x4

−2x4

−7x4

=

3x1 − x2 + 8x3 + x4

2x1 + 5x3 − 2x4

x1 + x2 + 3x3 − 7x4

You might recognize this final expression as being similar in style to some previous exam-ples (Example LT.ALT [303]) and some linear transformations defined in the archetypes.But the expression expressing the output as a linear combination of the columns of A isprobably the most powerful way of thinking about examples of this type.

Page 340: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.MLT Matrices and Linear Transformations 308

Almost forgot — we should verify that P is indeed a linear transformation. Thisis easy with our matrix operations from Section MO [121], specifically distributivityproperties contained in Theorem VSPM [123].

P (x + y) = A (x + y) = Ax + Ay = P (x) + P (y)

P (αx) = A (αx) = α (Ax) = αP (x)

So by Definition LT [302], P is a linear transformation. 4

So the multiplication of a vector by a matrix “transforms” the input vector into anoutput vector, possibly of a different size, by performing a linear combination. Andthis transformation happens in a “linear” fashion. This “functional” view of the matrix-vector product is the most important shift you can make right now in how you thinkabout linear algebra. Here’s the theorem.

Theorem MBLTMatrices Build Linear TransformationsSuppose that A is an m × n matrix. Define a function T : Cn 7→ Cm by T (x) = Ax.Then T is a linear transformation. �

Proof

P (x + y) = A (x + y) = Ax + Ay = P (x) + P (y)

P (αx) = A (αx) = α (Ax) = αP (x)

So by Definition LT [302], T is a linear transformation. �

So Theorem MBLT [308] gives us a rapid way to construct linear transformations. Graban m× n matrix A, define T (x) = Ax and Theorem MBLT [308] tells that T is a lineartransformation from Cn to Cm, without any further checking.

We can turn Theorem MBLT [308] around. You give me a linear transformation andI will give you a matrix.

Example LT.MFLTMatrix from a linear transformationDefine the function R : C3 7→ C4 by

R

x1

x2

x3

=

2x1 − 3x3 + 4x3

x1 + x2 + x3

−x1 + 5x2 − 3x− 3x2 − 4x− 3

You could verify that R is a linear transformation by applying the definition, but we willinstead massage the expression defining a typical output until we recognize the form of

Page 341: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.MLT Matrices and Linear Transformations 309

a known class of linear transformations.

R

x1

x2

x3

=

2x1 − 3x2 + 4x3

x1 + x2 + x3

−x1 + 5x2 − 3x3

x2 − 4x3

=

2x1

x1

−x1

0

+

−3x2

x2

5x2

x2

+

4x3

x3

3x3

−4x3

Definition CVA [67]

= x1

21−10

+ x2

−3151

+ x3

41−3−4

Definition CVSM [68]

=

2 −3 41 1 1−1 5 30 1 −4

x1

x2

x3

Definition MVP [151]

So if we define the matrix

B =

2 −3 41 1 1−1 5 30 1 −4

then R (x) = Bx. By Theorem MBLT [308], we can easily recognize R as a lineartransformation since it has the form described in the hypothesis of the theorem. 4

Example LT.MLT [??] was not accident. Consider any one of the archetypes where boththe domain and codomain are sets of column vectors (Cm) and you should be able tomimic the previous example. Here’s the theorem, which is notable since it is our firstoccasion to use the full power of the defining properties of a linear transformation whenour hypothesis includes a linear transformation.

Theorem MLTCVMatrix of a Linear Transformation, Column VectorsSuppose that T : Cn 7→ Cm is a linear transformation. Then there is an m× n matrix

A such that T (x) = Ax. �

Proof The conclusion says a certain matrix exists. What better way to prove somethingexists than to actually build it? So our proof will be constructive, and the procedurethat we will use abstractly in the proof can be used concretely in specific examples.

Let e1, e2, e3, . . . , en be the columns of the identity matrix of size n, In (Defini-tion SUV [169]). Evaluate the linear transformation T with each of these standard unitvectors as an input, and record the result. In other words, define n vectors in Cm, Ai,1 ≤ i ≤ n by

Ai = A (ei)

Then package up these vectors as the columns of a matrix

A = [A1|A2|A3| . . . |An]

Page 342: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.MLT Matrices and Linear Transformations 310

Does A have the desired properties? First, it is an m× n matrix. Then

T (x) = T (Inx) Theorem MMIM [156]

= T

[e1|e2|e3| . . . |en]

x1

x2

x3...

xn

Definition SUV [169]

= T (x1e1 + x2e2 + x3e3 + · · ·+ xnen) Definition MVP [151]

= T (x1e1) + T (x2e2) + T (x3e3) + · · ·+ T (xnen) Definition LT [302]

= x1T (e1) + x2T (e2) + x3T (e3) + · · ·+ xnT (en) Definition LT [302]

= x1A1 + x2A2 + x3A3 + · · ·+ xnAn Definition of Ai

= [A1|A2|A3| . . . |An]

x1

x2

x3...

xn

Definition MVP [151]

= Ax

as desired. �

So in the case of vector spaces of column vectors, every matrix leads to a linear trans-formation (Theorem MBLT [308]), while every linear transformation leads to a matrix(Theorem MLTCV [309]). So matrices and linear transformations are fundamentally thesame. We call the matrix A of Theorem MLTCV [309] the matrix representation ofT .

We have defined linear transformations for more general vector spaces than justCm, can we extend this correspondence to more general linear transformations? Yes,that is what Chapter R [??] is all about. Stay tuned. For now, lets illustrate Theo-rem MLTCV [309] with an example.

Example LT.MOLTMatrix of a linear transformationSuppose S : C3 7→ C4 is defined by

S

x1

x2

x3

=

3x1 − 2x2 + 5x3

x1 + x2 + x3

9x1 − 2x2 + 5x3

4x2

Page 343: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.LTLC Linear Transformations and Linear Combinations 311

Then

C1 = S (e1) = S

100

=

3190

C2 = S (e2) = S

010

=

−21−24

C3 = S (e3) = S

001

=

5150

so define

C = [C1|C2|C3] =

3 −2 51 1 19 −2 50 4 0

and Theorem MLTCV [309] guarantees that S (x) = Cx.

As an illuminating exercise, let z =

2−33

and compute S (z) two different ways.

First, return to the definition of S and evaluate S (z) directly. Then do the matrix-

vector product Cz. In both cases you should obtain the vector S (z) =

27239−12

. 4

Subsection LTLCLinear Transformations and Linear Combinations

It is the interaction between linear transformations and linear combinations that lies atthe heart of many of the important theorems of linear algebra. The next theorem distillsthe essence of this. The proof is not deep, the result is hardly startling, but it will bereferenced frequently. We have already passed by one occasion to employ it, in the proofof Theorem MLTCV [309]. Paraphrasing, this theorem says that we can “push” lineartransformations “down into” linear combinations, or “pull” linear transformations “upout” of linear combinations. We’ll have opportunities to both push and pull.

Theorem LTLCLinear Transformations and Linear CombinationsSuppose that T : U 7→ V is a linear transformation, u1, u2, u3, . . . , ut are vectors fromU and a1, a2, a3, . . . , at are scalars from C. Then

T (a1u1 + a2u2 + a3u3 + · · ·+ atut) = a1T (u1)+a2T (u2)+a3T (u3)+ · · ·+atT (ut) �

Page 344: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.LTLC Linear Transformations and Linear Combinations 312

Proof

T (a1u1 + a2u2 + a3u3 + · · ·+ atut)

= T (a1u1) + T (a2u2) + T (a3u3) + · · ·+ T (atut) Definition LT [302]

= a1T (u1) + a2T (u2) + a3T (u3) + · · ·+ atT (ut) Definition LT [302] �

Our next theorem says, informally, that it is enough to know how a linear transformationbehaves for inputs from a basis of the domain, and every other output is described by alinear combination of these values. Again, the theorem and its proof are not remarkable,but the insight that goes along with it is fundamental.

Theorem LTDBLinear Transformation Defined on a BasisSuppose that T : U 7→ V is a linear transformation, B = {u1, u2, u3, . . . , un} is a basisfor U and w is a vector from U . Let a1, a2, a3, . . . , an be scalars from C such that

w = a1u1 + a2u2 + a3u3 + · · ·+ anun

ThenT (w) = a1T (u1) + a2T (u2) + a3T (u3) + · · ·+ anT (un) �

Proof For any w ∈ U , Theorem VRRB [227] says there are (unique) scalars such thatw is a linear combination of the basis vectors in B. The result then follows from astraightforward application of Theorem LTLC [311] to the linear combination. �

Example LT.LTDB1Linear transformation defined on a basisSuppose you are told that T (C3) C2 is a linear transformation and given the three values,

T

100

=

[21

]T

010

=

[−14

]T

001

=

[60

]Because

B =

1

00

,

010

,

001

is a basis for C3 (Theorem SUVB [220]), Theorem LTDB [312] says we can compute anyoutput of T with just this information. For example, consider,

w =

2−31

= (2)

100

+ (−3)

010

+ (1)

001

so

T (w) = (2)

[21

]+ (−3)

[−14

]+ (1)

[60

]=

[13−11

]Any other value of T could be computed in a similar manner. So rather than being givena formula for the outputs of T , the requirement that T behave as a linear transformation,along with its values on a handful of vectors (the basis), are just as sufficient as a formulafor computing any value of the function. You might notice some parallels between thisexample and Example MOLT [??] or Theorem MLTCV [309]. 4

Page 345: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.LTLC Linear Transformations and Linear Combinations 313

Example LT.LTDB2Linear transformation defined on a basisSuppose you are told that R : C3 7→ C2 is a linear transformation and given the threevalues,

R

121

=

[5−1

]R

−151

=

[04

]R

314

=

[23

]You can check that

D =

1

21

,

−151

,

314

is a basis for C3 (make the vectors the columns of a square matrix and check that thematrix is nonsingular, Theorem CNSMB [225]). By Theorem LTDB [312] we can computeany output of R with just this information. However, we have to work just a bit harderto take an input vector and express it as a linear combination of the vectors in D. Forexample, consider,

y =

8−35

Then we must first write y as a linear combination of the vectors in D and solve for theunknown scalars, to arrive at

y =

8−35

= (3)

121

+ (−2)

−151

+ (1)

314

ThenTheorem LTDB [312] gives us

R (y) = (3)

[5−1

]+ (−2)

[04

]+ (1)

[23

]=

[17−8

]Any other value of R could be computed in a similar manner. 4

Here is a third example of a linear transformation defined by its action on a basis, onlywith more abstract vector spaces involved.

Example LT.LTDB3Linear transformation defined on a basisThe set W = {p(x) ∈ P3 | p(1) = 0, p(3) = 0} ⊆ P3 is a subspace of the vector space ofpolynomials P3. This subspace has C = {3− 4x + x2, 12− 13x + x3} as a basis (checkthis!). Suppose we define a linear transformation S : P3 7→ M22 by the values

S(3− 4x + x2

)=

[1 −32 0

]S(12− 13x + x3

)=

[0 11 0

]To illustrate a sample computation of S, consider q(x) = 9− 6x− 5x2 + 2x3. Verify thatq(x) is an element of W (does it have roots at x = 1 and x = 3?), then find the scalarsneeded to write it as a linear combination of the basis vectors in C. Because

q(x) = 9− 6x− 5x2 + 2x3 = (−5)(3− 4x + x2) + (2)(12− 13x + x3)

Page 346: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.PI Pre-Images 314

Theorem LTDB [312] gives us

S (q) = (−5)

[1 −32 0

]+ (2)

[0 11 0

]=

[−5 17−8 0

]And every other output of S could be computed in the same manner. Every output ofS will have a zero in the second row, second column. Can you see why this is so? 4

Subsection PIPre-Images

The definition of a function requires that for each input in the domain there is exactlyone output in the codomain. However, the correspondence does not have to behave theother way around. A member of the codomain might have many inputs from the domainthat create it, or it may have none at all. To formalize our discussion of this aspect oflinear transformations, we define the pre-image.

Definition PIPre-ImageSuppose that T : U 7→ V is a linear transformation. For each v, define the pre-imageof v to be the subset of U given by

T−1 (v) = {u ∈ U | T (u) = v} �

In other words, T−1 (v) is the set of all those vectors in the domain U that get “sent”to the vector v. TODO: All preimages form a partition of U , an equivalence relation isabout.

Example LT.SPIASSample pre-images, Archetype SArchetype S [444] is the linear transformation defined by

T : C3 7→ M22, T

abc

=

[a− b 2a + 2b + c

3a + b + c −2a− 6b− 2c

]We could compute a pre-image for every element of the codomain M22. However, evenin a free textbook, we do not have the room to do that, so we will compute just two.

Choose

v =

[2 13 2

]∈ M22

for no particular reason. What is T−1 (v)? Suppose u =

u1

u2

u3

∈ T−1 (v). That T (u) = v

becomes[2 13 2

]= v = T (u) = T

u1

u2

u3

=

[u1 − u2 2u1 + 2u2 + u3

3u1 + u2 + u3 −2u1 − 6u2 − 2u3

]

Page 347: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.PI Pre-Images 315

Using matrix equality (Definition ME [121]), we arrive at a system of four equations inthe three unknowns u1, u2, u3 with an augmented matrix that we can row-reduce in thehunt for solutions,

1 −1 0 22 2 1 13 1 1 3−2 −6 −2 2

RREF−−−→

1 0 1

454

0 1 14−3

4

0 0 0 00 0 0 0

We regognize this system as having infinitely many solutions described by the single freevariable u3. Eventually obtaining the vector form of the solutions (Theorem VFSLS [82]),we can describe the preimage precisely as,

T−1 (v) ={u ∈ C3

∣∣ T (u) = v}

=

u1

u2

u3

∣∣∣∣∣∣ u1 =5

4− 1

4u3, u2 = −3

4− 1

4u3

=

5

4− 1

4u3

−34− 1

4u3

u3

∣∣∣∣∣∣ u3 ∈ C3

=

5

4

−34

0

+ u3

−14

−14

1

∣∣∣∣∣∣ u3 ∈ C3

=

54

−34

0

+ Sp

−1

4

−14

1

This last line is merely a suggestive way of describing the set on the previous line. Youmight create three or four vectors in the preimage, and evaluate T with each. Was theresult what you expected? For a hint of things to come, you might try evaluating T withjust the lone vector in the spanning set above. What was the result? Now take a lookback at Theorem PSPHS [161]. Hmmmm.

OK, let’s compute another preimage, but with a different outcome this time. Choose

v =

[1 12 4

]∈ M22

What is T−1 (v)? Suppose u =

u1

u2

u3

∈ T−1 (v). That T (u) = v becomes

[1 12 4

]= v = T (u) = T

u1

u2

u3

=

[u1 − u2 2u1 + 2u2 + u3

3u1 + u2 + u3 −2u1 − 6u2 − 2u3

]Using matrix equality (Definition ME [121]), we arrive at a system of four equations inthe three unknowns u1, u2, u3 with an augmented matrix that we can row-reduce in the

Page 348: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.NLTFO New Linear Transformations From Old 316

hunt for solutions, 1 −1 0 12 2 1 13 1 1 2−2 −6 −2 4

RREF−−−→

1 0 1

40

0 1 14

0

0 0 0 10 0 0 0

By Theorem RCLS [41] we recognize this system as inconsistent. So no vector u is amember of T−1 (v) and so

T−1 (v) = ∅ 4

The preimage is just a set, it is rarely a subspace of U (you might think about just whenit is a subspace). We will describe its properties going forward, but in some ways it willbe a notational convenience as much as anything else.

Subsection NLTFONew Linear Transformations From Old

We can combine linear transformations in natural ways to create new linear transforma-tions. So we will define these combinations and then prove that the results really are stilllinear transformations. First the sum of two linear transformations.

Definition LTALinear Transformation AdditionSuppose that T : U 7→ V and S : U 7→ V are two linear transformations with the samedomain and codomain. Then their sum is the function T + S : U 7→ V whose outputsare defined by

(T + S) (u) = T (u) + S (u) �

Notice that the first plus sign in the definition is the operation being defined, while thesecond one is the vector addition in V . (Vector addition in U will appear just now in theproof that T+S is a linear transformation.) Definition LTA [316] only provides a function.It would be nice to know that if the constituents (T , S) are linear transformations, thenso is T + S.

Theorem SLTLTSum of Linear Transformations is a Linear TransformationSuppose that T : U 7→ V and S : U 7→ V are two linear transformations with the samedomain and codomain. Then T + S : U 7→ V is a linear transformation. �

Proof We simply check the defining properties of a linear transformation (Definition LT [302]).This is a good place to consistently ask yourself which objects are being combined withwhich operations.

(T + S) (x + y) = T (x + y) + S (x + y) Definition LTA [316]

= T (x) + T (y) + S (x) + S (y) Definition LT [302]

= T (x) + S (x) + T (y) + S (y) Commutativity in V

= (T + S) (x) + (T + S) (y) Definition LTA [316]

Page 349: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.NLTFO New Linear Transformations From Old 317

and

(T + S) (αx) = T (αx) + S (αx) Definition LTA [316]

= αT (x) + αS (x) Definition LT [302]

= α (T (x) + S (x)) Distributivity in V

= α(T + S) (x) Definition LTA [316]

Example LT.STLTSum of two linear transformationsSuppose that T : C2 7→ C3 and S : C2 7→ C3 are defined by

T

([x1

x2

])=

x1 + 2x2

3x1 − 4x2

5x1 + 2x2

S

([x1

x2

])=

4x1 − x2

x1 + 3x2

−7x1 + 5x2

Then by Definition LTA [316], we have

(T+S)

([x1

x2

])= T

([x1

x2

])+S

([x1

x2

])=

x1 + 2x2

3x1 − 4x2

5x1 + 2x2

+

4x1 − x2

x1 + 3x2

−7x1 + 5x2

=

5x1 + x2

4x1 − x2

−2x1 + 7x2

and by Theorem SLTLT [316] we know T + S is also a linear transformation from C2 toC3. 4

Definition LTSMLinear Transformation Scalar MultiplicationSuppose that T : U 7→ V is a linear transformation and α ∈ C. Then the scalarmultiple is the function αT : U 7→ V whose outputs are defined by

(αT ) (u) = αT (u) �

Given that T is a linear transformation, it would be nice to know that αT is also a lineartransformation.

Theorem MLTLTMultiple of a Linear Transformation is a Linear TransformationSuppose that T : U 7→ V is a linear transformation and α ∈ C. Then (αT ) : U 7→ V isa linear transformation. �

Proof We simply check the defining properties of a linear transformation (Definition LT [302]).This is another good place to consistently ask yourself which objects are being combinedwith which operations.

(αT ) (x + y) = α (T (x + y)) Definition LTSM [317]

= α (T (x) + T (y)) Definition LT [302]

= αT (x) + αT (y) Distributivity in V

= (αT ) (x) + (αT ) (y) Definition LTSM [317]

Page 350: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.NLTFO New Linear Transformations From Old 318

and

(αT ) (βx) = αT (βx) Definition LTSM [317]

= α (βT (x)) Definition LT [302]

= (αβ) T (x) Associativity in V

= (βα) T (x) Commutativity

= β (αT (x)) Associativity in V

= β(αT ) (x) Definition LTSM [317]

Example LT.SMLTScalar multiple of a linear transformationSuppose that T : C4 7→ C3 is defined by

T

x1

x2

x3

x4

=

x1 + 2x2 − x3 + 2x4

x1 + 5x2 − 3x3 + x4

−2x1 + 3x2 − 4x3 + 2x4

For the sake of an example, choose α = 2, so by Definition LTSM [317], we have

αT

x1

x2

x3

x4

= 2T

x1

x2

x3

x4

= 2

x1 + 2x2 − x3 + 2x4

x1 + 5x2 − 3x3 + x4

−2x1 + 3x2 − 4x3 + 2x4

=

2x1 + 4x2 − 2x3 + 4x4

2x1 + 10x2 − 6x3 + 2x4

−4x1 + 6x2 − 8x3 + 4x4

and by Theorem MLTLT [317] we know T + S is also a linear transformation from C4 toC3. 4

Now, lets imagine we have two vector spaces, U and V , and we collect every possible lineartransformation from U to V into one big set, and call it LT (U, V ). Definition LTA [316]and Definition LTSM [317] tell us how we can “add” and “scalar multiply” two elementsof LT (U, V ). Theorem SLTLT [316] and Theorem MLTLT [317] tell us that if we dothese operations, then the resulting functions are linear transformations that are also inLT (U, V ). Hmmmm, sounds like a vector space to me! A set of objects, an addition anda scalar multiplication. Why not?

Theorem VSLTVector Space of Linear TransformationsSuppose that U and V are vector spaces. Then the set of all linear transformations

from U to V , LT (U, V ) is a vector space when the operations are those given in Defini-tion LTA [316] and Definition LTSM [317]. �

Proof Theorem SLTLT [316] and Theorem MLTLT [317] provide two of the ten axiomsin Definition VS [184]. However, we still need to verify the remaining eight axioms. Byand large, the proofs are straightforward and rely on concocting the obvious object, orby reducing the question to the same vector space axiom in the vector space V .

The zero vector is of some interest, though. What linear transformation would weadd to any other linear transformation, so as to keep the second one unchanged? Theanswer is Z : U 7→ V defined by Z (u) = 0V for every u ∈ U . Notice how we do notneed to know any specifics about U and V to make this definition. �

Page 351: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.NLTFO New Linear Transformations From Old 319

Definition LTCLinear Transformation CompositionSuppose that T : U 7→ V and S : V 7→ W are linear transformations. Then thecomposition of S and T is the function (S ◦T ) : U 7→ W whose outputs are defined by

(S ◦ T ) (u) = S (T (u)) �

Given that T and S are linear transformations, it would be nice to know that S ◦ T isalso a linear transformation.

Theorem CLTLTComposition of Linear Transformations is a Linear TransformationSuppose that T : U 7→ V and S : V 7→ W are linear transformations. Then (S ◦ T ) :U 7→ W is a linear transformation. �

Proof We simply check the defining properties of a linear transformation (Definition LT [302]).

(S ◦ T ) (x + y) = S (T (x + y)) Definition LTC [319]

= S (T (x) + T (y)) Definition LT [302] for T

= S (T (x)) + S (T (y)) Definition LT [302] for S

= (S ◦ T ) (x) + (S ◦ T ) (y) Definition LTC [319]

and

(S ◦ T ) (αx) = S (T (αx)) Definition LTC [319]

= S (αT (x)) Definition LT [302] for T

= αS (T (x)) Definition LT [302] for S

= α(S ◦ T ) (x) Definition LTC [319] �

Example LT.CTLTComposition of two linear transformationsSuppose that T : C2 7→ C4 and S : C4 7→ C3 are defined by

T

([x1

x2

])=

x1 + 2x2

3x1 − 4x2

5x1 + 2x2

6x1 − 3x2

S

x1

x2

x3

x4

=

2x1 − x2 + x3 − x4

5x1 − 3x2 + 8x3 − 2x4

−4x1 + 3x2 − 4x3 + 5x4

Then by Definition LTC [319]

(S ◦ T )

([x1

x2

])= S

(T

([x1

x2

]))

= S

x1 + 2x2

3x1 − 4x2

5x1 + 2x2

6x1 − 3x2

=

2(x1 + 2x2)− (3x1 − 4x2) + (5x1 + 2x2)− (6x1 − 3x2)5(x1 + 2x2)− 3(3x1 − 4x2) + 8(5x1 + 2x2)− 2(6x1 − 3x2)−4(x1 + 2x2) + 3(3x1 − 4x2)− 4(5x1 + 2x2) + 5(6x1 − 3x2)

=

−2x1 + 13x2

24x1 + 44x2

15x1 − 43x2

Page 352: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection LT.NLTFO New Linear Transformations From Old 320

and by Theorem CLTLT [319] S ◦ T is a linear transformation from C2 to C3. 4

Here is an interesting exercise that will presage an important result later. In Exam-ple LT.STLT [317] compute (via Theorem MLTCV [309]) the matrix of T , S and T + S.Do you see a relationship between these three matrices?

In Example LT.SMLT [318] compute (via Theorem MLTCV [309]) the matrix of Tand 2T . Do you see a relationship between these two matrices?

Here’s the tough one. In Example LT.CTLT [319] compute (via Theorem MLTCV [309])the matrix of T , S and S ◦T . Do you see a relationship between these three matrices???

Page 353: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section ILT Injective Linear Transformations 321

Section ILT

Injective Linear Transformations

Linear transformations have two key properties, which go by the names injective andsurjective. We will see that they are closely related to ideas like linear independence andspanning, and sets like the null space and the range. In this section we will define aninjective linear transformation and analyze the resulting consequences. The next sectionwill do the same for the surjective property. In the final section of this chapter we willsee what happens when we have the two properties simultaneously.

As usual, we lead with a definition.

Definition ILTInjective Linear TransformationSuppose T : U 7→ V is a linear transformation. Then T is injective if wheneverT (x) = T (y), then x = y. �

Given an arbitrary function, it is possible for two different inputs to yield the same output(think about the function f(x) = x2 and the inputs x = 3 and x = −3). For an injectivefunction, this never happens. If we have equal outputs (T (x) = T (y)) then we musthave achieved those equal outputs by employing equal inputs (x = y). Some authorsprefer the term one-to-one where we use injective, and we will sometimes refer to aninjective linear transformation as an injection.

Subsection EILTExamples of Injective Linear Transformations

It is perhaps most instructive to examine a linear transformation that is not injectivefirst.

Example ILT.NIAQNot injective, Archetype QArchetype Q [439] is the linear transformation

T : C5 7→ C5, T

x1

x2

x3

x4

x5

=

−2x1 + 3x2 + 3x3 − 6x4 + 3x5

−16x1 + 9x2 + 12x3 − 28x4 + 28x5

−19x1 + 7x2 + 14x3 − 32x4 + 37x5

−21x1 + 9x2 + 15x3 − 35x4 + 39x5

−9x1 + 5x2 + 7x3 − 16x4 + 16x5

Notice that for

x =

13−124

y =

47057

Page 354: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.EILT Examples of Injective Linear Transformations 322

we have

T

13−124

=

455727731

T

47057

=

455727731

So we have two vectors from the domain, x 6= y, yet T (x) = T (y), in violation ofDefinition ILT [321]. This is another example where you should not concern yourself withhow x and y were selected, as this will be explained shortly. However, do understandwhy these two vectors provide enough evidence to conclude that T is not injective. 4

To show that a linear transformation is not injective, it is enough to find a single pair ofinputs that get sent to the identical output, as in Example ILT.NIAQ [321]. However, toshow that a linear transformation is injective we must establish that this coincidence ofoutputs never occurs. Here is an example that shows how to establish this.

Example ILT.IARInjective, Archetype RArchetype R [442] is the linear transformation

T : C5 7→ C5, T

x1

x2

x3

x4

x5

=

−65x1 + 128x2 + 10x3 − 262x4 + 40x5

36x1 − 73x2 − x3 + 151x4 − 16x5

−44x1 + 88x2 + 5x3 − 180x4 + 24x5

34x1 − 68x2 − 3x3 + 140x4 − 18x5

12x1 − 24x2 − x3 + 49x4 − 5x5

Page 355: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.EILT Examples of Injective Linear Transformations 323

To establish that R is injective we must begin with the assumption that T (x) = T (y)and somehow arrive from this at the conclusion that x = y. Here we go,

T (x) = T (y)

T

x1

x2

x3

x4

x5

= T

y1

y2

y3

y4

y5

−65x1 + 128x2 + 10x3 − 262x4 + 40x5

36x1 − 73x2 − x3 + 151x4 − 16x5

−44x1 + 88x2 + 5x3 − 180x4 + 24x5

34x1 − 68x2 − 3x3 + 140x4 − 18x5

12x1 − 24x2 − x3 + 49x4 − 5x5

=

−65y1 + 128y2 + 10y3 − 262y4 + 40y5

36y1 − 73y2 − y3 + 151y4 − 16y5

−44y1 + 88y2 + 5y3 − 180y4 + 24y5

34y1 − 68y2 − 3y3 + 140y4 − 18y5

12y1 − 24y2 − y3 + 49y4 − 5y5

−65x1 + 128x2 + 10x3 − 262x4 + 40x5

36x1 − 73x2 − x3 + 151x4 − 16x5

−44x1 + 88x2 + 5x3 − 180x4 + 24x5

34x1 − 68x2 − 3x3 + 140x4 − 18x5

12x1 − 24x2 − x3 + 49x4 − 5x5

−−65y1 + 128y2 + 10y3 − 262y4 + 40y5

36y1 − 73y2 − y3 + 151y4 − 16y5

−44y1 + 88y2 + 5y3 − 180y4 + 24y5

34y1 − 68y2 − 3y3 + 140y4 − 18y5

12y1 − 24y2 − y3 + 49y4 − 5y5

=

00000

−65(x1 − y1) + 128(x2 − y2) + 10(x3 − y3)− 262(x4 − y4) + 40(x5 − y5)

36(x1 − y1)− 73(x2 − y2)− (x3 − y3) + 151(x4 − y4)− 16(x5 − y5)−44(x1 − y1) + 88(x2 − y2) + 5(x3 − y3)− 180(x4 − y4) + 24(x5 − y5)34(x1 − y1)− 68(x2 − y2)− 3(x3 − y3) + 140(x4 − y4)− 18(x5 − y5)

12(x1 − y1)− 24(x2 − y2)− (x3 − y3) + 49(x4 − y4)− 5(x5 − y5)

=

00000

−65 128 10 −262 4036 −73 −1 151 −16−44 88 5 −180 2434 −68 −3 140 −1812 −24 −1 49 −5

x1 − y1

x2 − y2

x3 − y3

x4 − y4

x5 − y5

=

00000

Now we recognize that we have a homogenous system of 5 equations in 5 variables (theterms xi − yi are the variables), so we row-reduce the coefficient matrix to

1 0 0 0 0

0 1 0 0 0

0 0 1 0 0

0 0 0 1 0

0 0 0 0 1

So the only solution is the trivial solution

x1 − y1 = 0 x2 − y2 = 0 x3 − y3 = 0 x4 − y4 = 0 x5 − y5 = 0

and we conlude that indeed x = y. By Definition ILT [321], T is injective. 4

Lets now examine an injective linear transformation between abstract vector spaces.

Example ILT.IAVInjective, Archetype V

Page 356: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.NSLT Null Space of a Linear Transformation 324

Archetype V [444] is defined by

T : P3 7→ M22, T(a + bx + cx2 + dx3

)=

[a + b a− 2c

d b− d

]To establish that the linear transformation is injective, begin by supposing that twopolynomial inputs yield the same output matrix,

T(a1 + b1x + c1x

2 + d1x3)

= T(a2 + b2x + c2x

2 + d2x3)

Then

O =

[0 00 0

]= T

(a1 + b1x + c1x

2 + d1x3)− T

(a2 + b2x + c2x

2 + d2x3)

Hypothesis

= T((a1 + b1x + c1x

2 + d1x3)− (a2 + b2x + c2x

2 + d2x3))

T linear transformation

= T((a1 − a2) + (b1 − b2)x + (c1 − c2)x

2 + (d1 − d2)x3)

Operations in P3

=

[(a1 − a2) + (b1 − b2) (a1 − a2)− 2(c1 − c2)

(d1 − d2) (b1 − b2)− (d1 − d2)

]Definition of T

This single matrix equality translates to the homogenous system of equations in thevariables ai − bi,

(a1 − a2) + (b1 − b2) = 0

(a1 − a2)− 2(c1 − c2) = 0

(d1 − d2) = 0

(b1 − b2)− (d1 − d2) = 0

This system of equations can be rewritten as the matrix equation1 1 0 01 0 −2 00 0 0 10 1 0 −1

(a1 − a2)(b1 − b2)(c1 − c2)(d1 − d2)

=

0000

Since the coefficient matrix is nonsingular (check this) the only solution is trivial, i.e.

a1 − a2 = 0 b1 − b2 = 0 c1 − c2 = 0 d1 − d2 = 0

so that

a1 = a2 b1 = b2 c1 = c2 d1 = d2

so the two inputs must be equal polynomials. By Definition ILT [321], T is injective. 4

Subsection NSLTNull Space of a Linear Transformation

For a linear transformation T : U 7→ V , the null space is a subset of the domain U .Informally, it is the set of all inputs that the transformation sends to the zero vector of

Page 357: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.NSLT Null Space of a Linear Transformation 325

the codomain. It will have some natural connections with the null space of a matrix, sowe will keep the same notation, and if you think about your objects, then there shouldbe little confusion. Here’s the careful definition.

Definition NSLTNull Space of a Linear TransformationSuppose T : U 7→ V is a linear transformation. Then the null space of T is the setN (T ) = {u ∈ U | T (u) = 0} �

Example ILT.NSAONull space, Archetype OArchetype O [435] is the linear transformation

T : C3 7→ C5, T

x1

x2

x3

=

−x1 + x2 − 3x3

−x1 + 2x2 − 4x3

x1 + x2 + x3

2x1 + 3x2 + x3

x1 + 2x3

To determine the elements of C3 in N (T ), find those vectors u such that T (u) = 0, thatis,

T (u) = 0−u1 + u2 − 3u3

−u1 + 2u2 − 4u3

u1 + u2 + u3

2u1 + 3u2 + u3

u1 + 2u3

=

00000

Vector equality (Definition CVE [66]) leads us to a homogenous system of 5 equations inthe variables ui,

−u1 + u2 − 3u3 = 0

−u1 + 2u2 − 4u3 = 0

u1 + u2 + u3 = 0

2u1 + 3u2 + u3 = 0

u1 + 2u3 = 0

Row-reducing the coefficient matrix gives1 0 2

0 1 −10 0 00 0 00 0 0

The null space of T is the set of solutions to this homogenous system of equations, whichby Theorem SSNS [90] and Theorem BNS [104] can be expressed as

N (T ) = Sp

−2

11

4

Page 358: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.NSLT Null Space of a Linear Transformation 326

We know that the span of a set of vectors is always a subspace (Theorem SSS [203]),so the null space computed in Example ILT.NSAO [325] is also a subspace. This is noaccident, the null space of a linear transformation is always a subspace.

Theorem NSLTSNull Space of a Linear Transformation is a SubspaceSuppose that T : U 7→ V is a linear transformation. Then the null space of T , N (T ), isa subspace of U . �

Proof We can apply the three-part test of Theorem TSS [198]. First T (0U) = 0V byTheorem LTTZZ [306], so 0U ∈ N (T ) and we know that the null space is non-empty.

Suppose we assume that x, y ∈ N (T ). Is x + y ∈ N (T )?

T (x + y) = T (x) + T (y) Definition LT [302]

= 0 + 0 x, y ∈ N (T )

= 0 Property of zero vector

This qualifies x + y for membership in N (T ). So we have additive closure.Suppose we assume that α ∈ C and x ∈ N (T ). Is αx ∈ N (T )?

T (αx) = αT (x) Definition LT [302]

= α0 x ∈ N (T )

= 0 Theorem ZVSM [192] �

This qualifies αx for membership in N (T ). So we have scalar closure and Theorem TSS [198]tells us that N (T ) is a subspace of U .

Let’s compute another null space, now that we know in advance that it will be a subspace.

Example ILT.TNSAPTrivial null space, Archetype PArchetype P [437] is the linear transformation

T : C3 7→ C5, T

x1

x2

x3

=

−x1 + x2 + x3

−x1 + 2x2 + 2x3

x1 + x2 + 3x3

2x1 + 3x2 + x3

−2x1 + x2 + 3x3

To determine the elements of C3 in N (T ), find those vectors u such that T (u) = 0, thatis,

T (u) = 0−u1 + u2 + u3

−u1 + 2u2 + 2u3

u1 + u2 + 3u3

2u1 + 3u2 + u3

−2u1 + u2 + 3u3

=

00000

Page 359: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.NSLT Null Space of a Linear Transformation 327

Vector equality (Definition CVE [66]) leads us to a homogeneous system of 5 equationsin the variables ui,

−u1 + u2 + u3 = 0

−u1 + 2u2 + 2u3 = 0

u1 + u2 + 3u3 = 0

2u1 + 3u2 + u3 = 0

−2u1 + u2 + 3u3 = 0

Row-reducing the coefficient matrix gives1 0 0

0 1 0

0 0 10 0 00 0 0

The null space of T is the set of solutions to this homogeneous system of equations, whichis simply the trivial solution u = 0, so

N (T ) = {0} = Sp ({ }) 4

Our next theorem says that if a preimage is a non-empty set then we can constructit by picking any one element and adding on elements of the null space.

Theorem NSPINull Space and Pre-ImageSuppose T : U 7→ V is a linear transformation and v ∈ V . If the preimage T−1 (v) isnon-empty, and u ∈ T−1 (v) then

T−1 (v) = {u + z | z ∈ N (T )} = u + N (T ) �

Proof Let M = {u + z | z ∈ N (T )}. First, we show that M ⊆ T−1 (v). Suppose thatw ∈ M , so w has the form w = u + z, where z ∈ N (T ). Then

T (w) = T (u + z)

= T (u) + T (z) Definition LT [302]

= v + 0 u ∈ T−1 (v) , z ∈ N (T )

= v Property of 0

which qualifies w for membership in the preimage of v, w ∈ T−1 (v).For the opposite inclusion, suppose x ∈ T−1 (v). Then,

T (x− u) = T (x)− T (u) Definition LT [302]

= v − v x, u ∈ T−1 (v)

= 0

This qualifies x − u for membership in the null space of T , N (T ). So there is a vectorz ∈ N (T ) such that x−u = z. Rearranging this equation gives x = u+z and so x ∈ M .So T−1 (v) ⊆ M and we see that M = T−1 (v), as desired. �

Page 360: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.NSLT Null Space of a Linear Transformation 328

This theorem, and its proof, should remind you very much of Theorem PSPHS [161].Additionally, you might go back and review Example LT.SPIAS [314]. Can you tell nowwhich is the only preimage to be a subspace?

Theorem NSILTNull Space of an Injective Linear TransformationSuppose that T : U 7→ V is a linear transformation. Then T is injective if and only ifthe null space of T is trivial, N (T ) = {0}. �

Proof (⇒) Suppose x ∈ N (T ). Then by Definition NSLT [325], T (x) = 0. By Theo-rem LTTZZ [306], T (0) = 0. Now, since T (x) = T (0), Definition ILT [321] we concludethat x = 0. Therefore N (T ) = {0}.

(⇐) To establish that T is injective, appeal to Definition ILT [321] and begin withthe assumption that T (x) = T (y). Then

0 = T (x)− T (y) Hypothesis

= T (x− y) Definition LT [302]

so by Definition NSLT [325] and the hypothesis that the null space is trivial,

x− y ∈ N (T ) = {0}

which means that

0 = x− y

x = y

thus establishing that T is injective. �

Example ILT.NIAQRNot injective, Archetype Q, revisitedWe are now in a position to revisit our first example in this section, Example ILT.NIAQ [321].In that example, we showed that Archetype Q [439] is not injective by constructing twovectors, which when used to evaluate the linear transformation provided the same output,thus violating Definition ILT [321]. Just where did those two vectors come from?

The key is the vector

z =

34133

which you can check is an element of N (T ) for Archetype Q [439]. Choose a vectorx at random, and then compute y = x + z (verify this computation back in Exam-ple ILT.NIAQ [321]). Then

T (y) = T (x + z)

= T (x) + T (z) Definition LT [302]

= T (x) + 0 z ∈ N (T )

= T (x) Property of 0

Page 361: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.NSLT Null Space of a Linear Transformation 329

Whenever the null space of a linear transformation is non-trivial, we can employ thisdevice and conclude that the linear transformation is not injective. This is another wayof viewing Theorem NSILT [328]. For an injective linear transformation, the null spaceis trivial and our only choice for z is the zero vector, which will not help us create twodifferent inputs for T that yield identical outputs. For every one of the archetypes thatis not injective, there is an example presented of exactly this form. 4

Example ILT.NIAONot injective, Archetype OIn Example ILT.NSAO [325] the null space of Archetype O [435] was determined to be

Sp

−2

11

a subspace of C3 with dimension 1. Since the null space is not trivial, Theorem NSLIT [??]tells us that T is not injective. 4

Example ILT.IAPInjective, Archetype PIn Example ILT.TNSAP [326] it was shown that the linear transformation in Archetype P [437]has a trivial null space. So by Theorem NSLIT [??], T is injective. 4

There is a connection between injective linear transformations and linear independentsets that we will make precise in the next two theorems. However, more informally, wecan get a feel for this connection when we think about how each property is defined.A set of vectors is linearly independent if the only relation of linear dependence is thetrivial one. A linear transformation is injective if the only way two input vectors canproduce the same output is if the trivial way, both input vectors are equal.

Theorem ILTLIInjective Linear Transformations and Linear IndependenceSuppose that T : U 7→ V is an injective linear transformation and S = {u1, u2, u3, . . . , ut}is a linearly independent subset of U . Then R = {T (u1) , T (u2) , T (u3) , . . . , T (ut)} isa linearly independent subset of V . �

Proof Begin with a relation of linear dependence on S (Definition RLD [212], Defini-tion LI [212]),

a1T (u1) + a2T (u2) + a3T (u3) + . . . + atT (ut) = 0

T (a1u1 + a2u2 + a3u3 + · · ·+ atut) = 0 Theorem LTLC [311]

a1u1 + a2u2 + a3u3 + · · ·+ atut ∈ N (T ) Definition NSLT [325]

a1u1 + a2u2 + a3u3 + · · ·+ atut ∈ {0} Theorem NSILT [328]

a1u1 + a2u2 + a3u3 + · · ·+ atut = 0

Since this is a relation of linear dependence on the linearly independent set S, we canconclude that

a1 = 0 a2 = 0 a2 = 0 . . . at = 0

and this establishes that R is a linearly independent set. �

Page 362: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.ILTD Injective Linear Transformations and Dimension 330

Theorem ILTBInjective Linear Transformations and BasesSuppose that T : U 7→ V is a linear transformation and B = {u1, u2, u3, . . . , um} is abasis of U . Then T is injective if and only if C = {T (u1) , T (u2) , T (u3) , . . . , T (um)}is a linearly independent subset of V . �

Proof (⇒) Assume T is injective. Since B is a basis, we know B is linearly independent(Definition B [220]). Then Theorem ILTLI [329] says that C is a linearly independentsubset of V .

(⇐) Assume that C is linearly independent. To establish that T is injective, we willshow that the null space of T is trivial (Theorem NSILT [328]). Suppose that u ∈ N (T ).As an element of U , we can write u as a linear combination of the basis vectors in B(uniquely). So there are are scalars, a1, a2, a3, . . . , am, such that

u = a1u1 + a2u2 + a3u3 + · · ·+ amum

Then,

0 = T (u) u ∈ N (T )

= T (a1u1 + a2u2 + a3u3 + · · ·+ amum) B spans U

= a1T (u1) + a2T (u2) + a3T (u3) + · · ·+ amT (um) Theorem LTLC [311]

This is a relation of linear dependence (Definition RLD [212]) on the linearly independentset C, so the scalars are all zero: a1 = a2 = a3 = · · · = am = 0. Then

u = a1u1 + a2u2 + a3u3 + · · ·+ amum

= 0u1 + 0u2 + 0u3 + · · ·+ 0um

= 0 + 0 + 0 + · · ·+ 0 Theorem ZSSM [191]

= 0 Property of 0

Since u was chosen as an arbitrary vector from N (T ), we have N (T ) = {0} and Theo-rem NSILT [328] tells us that T is injective. �

Subsection ILTDInjective Linear Transformations and Dimension

Theorem ILTDInjective Linear Transformations and DimensionSuppose that T : U 7→ V is an injective linear transformation. Then dim (U) ≤dim (V ). �

Proof Suppose to the contrary that m = dim (U) > dim (V ) = t. Let B be a basis of U ,which will then contain m vectors. Apply T to each element of B to form a set C thatis a subset of V . By Theorem ILTB [330], C is linearly independent and therefore mustcontain m distinct vectors. So we have a set of m linearly independent vectors in V , avector space of dimension t, with m > t. However, this contradicts Theorem G [244], soour assumption is false and dim (U) ≤ dim (V ). �

Page 363: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection ILT.CILT Composition of Injective Linear Transformations 331

Example ILT.NIDAUNot injective by dimension, Archetype UThe linear transformation in Archetype U [444] is

T : M23 7→ C4, T

([a b cd e f

])=

a + 2b + 12c− 3d + e + 6f

2a− b− c + d− 11fa + b + 7c + 2d + e− 3fa + 2b + 12c + 5e− 5f

Since dim (M23) = 6 > 4 = dim (C4), T cannot be injective for then it would violateTheorem ILTD [330]. 4

Notice that the previous example made no use of the actual formula defining the func-tion. Merely a comparison of the dimensions of the domain and codomain are enoughto conclude that the linear transformation is not injective. Archetype M [431] andArchetype N [433] are two more examples of linear transformations that have “big”domains and “small” codomains, resulting in “collisions” of outputs and thus are non-injective linear transformations.

Subsection CILTComposition of Injective Linear Transformations

In Subsection LT.NLTFO [316] we saw how to combine linear transformations to buildnew linear transformations, specifically, how to build the composition of two linear trans-formations (Definition LTC [319]). It will be useful later to know that the compositionof injective linear transformations is again injective, so we prove that here.

Theorem CILTIComposition of Injective Linear Transformations is InjectiveSuppose that T : U 7→ V and S : V 7→ W are injective linear transformations. Then(S ◦ T ) : U 7→ W is an injective linear transformation. �

Proof That the composition is a linear transformation was established in Theorem CLTLT [319],so we need only establish that the composition is injective. Applying Definition ILT [321],choose x, y from U . Then

(S ◦ T ) (x) = (S ◦ T ) (y)

S (T (x)) = S (T (y)) Definition LTC [319]

T (x) = T (y) S is injective

x = y T is injective �

Page 364: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section SLT Surjective Linear Transformations 332

Section SLT

Surjective Linear Transformations

The companion to an injection is a surjection. Surjective linear transformations areclosely related to spanning sets and ranges. So as you read this section reflect backon Section ILT [321] and note the parallels and the contrasts. In the next section,Section IVLT [344], we will combine the two properties.

As usual, we lead with a definition.

Definition SLTSurjective Linear TransformationSuppose T : U 7→ V is a linear transformation. Then T is surjective if for every v ∈ Vthere exists a u ∈ U so that T (u) = v. �

Given an arbitrary function, it is possible for there to be an element of the codomain thatis not an output of the function (think about the function f(x) = x2 and the codomainelement x = −3). For a surjective function, this never happens. If we choose any elementof the codomain (v ∈ V ) then there must be an input from the domain u ∈ U which willcreate the output when used to evaluate the linear transformation (T (u) = v). Someauthors prefer the term onto where we use surjective, and we will sometimes refer to asurjective linear transformation as a surjection.

Subsection ESLTExamples of Surjective Linear Transformations

It is perhaps most instructive to examine a linear transformation that is not injectivefirst.

Example SLT.NSAQNot surjective, Archetype QArchetype Q [439] is the linear transformation

T : C5 7→ C5, T

x1

x2

x3

x4

x5

=

−2x1 + 3x2 + 3x3 − 6x4 + 3x5

−16x1 + 9x2 + 12x3 − 28x4 + 28x5

−19x1 + 7x2 + 14x3 − 32x4 + 37x5

−21x1 + 9x2 + 15x3 − 35x4 + 39x5

−9x1 + 5x2 + 7x3 − 16x4 + 16x5

We will demonstrate that

v =

−123−14

Page 365: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.ESLT Examples of Surjective Linear Transformations 333

is an unobtainable element of the codomain. Suppose to the contrary that u is an elementof the domain such that T (u) = v. Then

−123−14

= v = T (u) = T

u1

u2

u3

u4

u5

=

−2u1 + 3u2 + 3u3 − 6u4 + 3u5

−16u1 + 9u2 + 12u3 − 28u4 + 28u5

−19u1 + 7u2 + 14u3 − 32u4 + 37u5

−21u1 + 9u2 + 15u3 − 35u4 + 39u5

−9u1 + 5u2 + 7u3 − 16u4 + 16u5

=

−2 3 3 −6 3−16 9 12 −28 28−19 7 14 −32 37−21 9 15 −35 39−9 5 7 −16 16

u1

u2

u3

u4

u5

Now we recognize the appropriate input vector u as a solution to a linear system ofequations. Form the augmented matrix, and row-reduce to

1 0 0 0 −1 0

0 1 0 0 −43

0

0 0 1 0 −13

0

0 0 0 1 −1 00 0 0 0 0 1

With a leading 1 in the last column, Theorem RCLS [41] tells us the system is inconsistent.From this we conclude that no such vector u exists, and by Definition SLT [332], T isnot surjective.

Again, do not concern yourself with how v was selected, as this will be explainedshortly. However, do understand why this vector provides enough evidence to concludethat T is not surjective. 4

To show that a linear transformation is not surjective, it is enough to find a single ele-ment of the codomain that is never created by any input, as in Example SLT.NSAQ [332].However, to show that a linear transformation is surjective we must establish that ev-ery element of the codomain occurs as an output of the linear transformation for someappropriate input.

Example SLT.SARSurjective, Archetype RArchetype R [442] is the linear transformation

T : C5 7→ C5, T

x1

x2

x3

x4

x5

=

−65x1 + 128x2 + 10x3 − 262x4 + 40x5

36x1 − 73x2 − x3 + 151x4 − 16x5

−44x1 + 88x2 + 5x3 − 180x4 + 24x5

34x1 − 68x2 − 3x3 + 140x4 − 18x5

12x1 − 24x2 − x3 + 49x4 − 5x5

Page 366: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.ESLT Examples of Surjective Linear Transformations 334

To establish that R is surjective we must begin with a totally arbitrary element of thecodomain, v and somehow find an input vector u such that T (u) = v. We desire,

T (u) = v−65u1 + 128u2 + 10u3 − 262u4 + 40u5

36u1 − 73u2 − u3 + 151u4 − 16u5

−44u1 + 88u2 + 5u3 − 180u4 + 24u5

34u1 − 68u2 − 3u3 + 140u4 − 18u5

12u1 − 24u2 − u3 + 49u4 − 5u5

=

v1

v2

v3

v4

v5

−65 128 10 −262 4036 −73 −1 151 −16−44 88 5 −180 2434 −68 −3 140 −1812 −24 −1 49 −5

u1

u2

u3

u4

u5

=

v1

v2

v3

v4

v5

We recognize this equation as a system of equations in the variables ui, but our vectorof constants contains symbols. In general, we would have to row-reduce the augmentedmatrix by hand, due to the symbolic final column. However, in this particular example,the 5 × 5 coefficient matrix is nonsingular and so has an inverse (Theorem NSI [177],Definition MI [167]).

−65 128 10 −262 4036 −73 −1 151 −16−44 88 5 −180 2434 −68 −3 140 −1812 −24 −1 49 −5

−1

=

−47 92 1 −181 −1427 −55 7

22214

11−32 64 −1 −126 −1225 −50 3

21992

99 −18 1

2712

4

so we find that

u1

u2

u3

u4

u5

=

−47 92 1 −181 −14

27x− 55 72

2214

11−32 64 −1 −126 −1225 −50 3

21992

99 −18 1

2712

4

v1

v2

v3

v4

v5

=

−47v1 + 92v2 + v3 − 181v4 − 14v5

27v1 − 55v2 + 72v3 + 221

4v4 + 11v5

−32v1 + 64v2 − v3 − 126v4 − 12v5

25v1 − 50v2 + 32v3 + 199

2v4 + 9v5

9v1 − 18v2 + 12v3 + 71

2v4 + 4v5

This establishes that if we are given any output vector v, we can use its components in thisfinal expression to formulate a vector u such that T (u) = v. So by Definition SLT [332]we now know that T is surjective. You might try to verify this condition in its fullgenerality (i.e. evaluate T with this final expression and see if you get v as the result),or tet it more specifially for some numerical vector v. 4

Lets now examine a surjective linear transformation between abstract vector spaces.

Example SLT.SAVSurjective, Archetype V

Page 367: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.ESLT Examples of Surjective Linear Transformations 335

Archetype V [444] is defined by

T : P3 7→ M22, T(a + bx + cx2 + dx3

)=

[a + b a− 2c

d b− d

]To establish that the linear transformation is injective, begin by choosing an arbitraryoutput. In this example, we need to choose an arbitrary 2× 2 matrix, say

v =

[x yz w

]and we would like to find an input polynomial

u = a + bx + cx2 + dx3

so that T (u) = v. So we have,[x yz w

]= v

= T (u)

= T(a + bx + cx2 + dx3

)=

[a + b a− 2c

d b− d

]Matrix equality leads us to the system of four equations in the four unknowns, x, y, z, w,

a + b = x

a− 2c = y

d = z

b− d = w

which can be rewritten as a matrix equation,1 1 0 01 0 −2 00 0 0 10 −1 0 1

abcd

=

xyzw

The coefficient matrix is nonsingular, hence it has an inverse,

1 1 0 01 0 −2 00 0 0 10 −1 01

−1

=

1 0 −1 10 0 1 −112−1

2−1

212

0 0 1 0

so we have

abcd

=

1 0 −1 10 0 1 −112−1

2−1

212

0 0 1 0

xyzw

=

x− z + w

z − w12(x− y − z + w)

z

Page 368: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.RLT Range of a Linear Transformation 336

So the input polynomial u = (x− z +w)+ (z−w)x+ 12(x− y− z +w)x3 + zx4 will yield

the output matrix v, no matter what form v takes. This means by Definition SLT [332]that T is surjective. All the same, lets do a concrete demonstration and evaluate T withu,

T (u) = T

((x− z + w) + (z − w)x +

1

2(x− y − z + w)x3 + x4

)=

[(x− z + w) + (z − w) (x− z + w)− 2(1

2(x− y − z + w))

z (z − w)− z

]=

[x yz w

]= v 4

Subsection RLTRange of a Linear Transformation

For a linear transformation T : U 7→ V , the range is a subset of the codomain V .Informally, it is the set of all outputs that the transformation creates when fed everypossible input from the domain. It will have some natural connections with the range ofa matrix, so we will keep the same notation, and if you think about your objects, thenthere should be little confusion. Here’s the careful definition.

Definition RLTRange of a Linear TransformationSuppose T : U 7→ V is a linear transformation. Then the range of T is the set R (T ) ={T (u) | u ∈ U} �

Example SLT.RAORange, Archetype OArchetype O [435] is the linear transformation

T : C3 7→ C5, T

x1

x2

x3

=

−x1 + x2 − 3x3

−x1 + 2x2 − 4x3

x1 + x2 + x3

2x1 + 3x2 + x3

x1 + 2x3

Page 369: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.RLT Range of a Linear Transformation 337

To determine the elements of C5 in R (T ), find those vectors v such that T (u) = v forsome u ∈ C3,

v = T (u)

=

−u1 + u2 − 3u3

−u1 + 2u2 − 4u3

u1 + u2 + u3

2u1 + 3u2 + u3

u1 + 2u3

=

−u1

−u1

u1

2u1

u1

+

u2

2u2

u2

3u2

0

+

−3u3

−4u3

u3

u3

2u3

= u1

−1−1121

+ u2

12130

+ u3

−3−4112

This says that every output of T (v) can be written as a linear combination of the threevectors

−1−1121

12130

−3−4112

using the scalars u1, u2, u3. Furthermore, since u can be any element of C3, every suchlinear combination is an output. This means that

R (T ) = Sp

−1−1121

,

12130

,

−3−4112

The three vectors in this spanning set for R (T ) form a linearly dependent set (checkthis!). So we can find a more economical presentation by any of the various methodsfrom Section RM [127] and Section RSM [141]. We will place the vectors into a matrixas rows, row-reduce, toss out zero rows and appeal to Theorem BRS [144], so we candescribe the range of T with a basis,

R (T ) = Sp

10−3−7−2

,

01251

4

Page 370: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.RLT Range of a Linear Transformation 338

We know that the span of a set of vectors is always a subspace (Theorem SSS [203]), sothe range computed in Example SLT.RAO [336] is also a subspace. This is no accident,the range of a linear transformation is always a subspace.

Theorem RLTSRange of a Linear Transformation is a SubspaceSuppose that T : U 7→ V is a linear transformation. Then the range of T , R (T ), is asubspace of V . �

Proof We can apply the three-part test of Theorem TSS [198]. First T (0U) = 0V byTheorem LTTZZ [306], so 0V ∈ R (T ) and we know that the range is non-empty.

Suppose we assume that x, y ∈ R (T ). Is x + y ∈ R (T )? If x, y ∈ R (T ) then weknow there are vectors w, z ∈ U such that T (w) = x and T (z) = y. Because U is avector space, additive closure implies that w + z ∈ U . Then

T (w + z) = T (w) + T (z) Definition LT [302]

= x + y Definition of w and z

So we have found an input (w + z) which when fed into T creates x + y as an output.This qualifies x + y for membership in R (T ). So we have additive closure.

Suppose we assume that α ∈ C and x ∈ R (T ). Is αx ∈ R (T )? If x ∈ R (T ), thenthere is a vector w ∈ U such that T (w) = x. Because U is a vector space, scalar closureimplies that αw ∈ U . Then

T (αw) = αT (w) Definition LT [302]

= αx Definition of w �

So we have found an input (αw) which when fed into T creates αx as an output. Thisqualifies αx for membership in R (T ). So we have scalar closure and Theorem TSS [198]tells us that R (T ) is a subspace of V .

Let’s compute another range, now that we know in advance that it will be a subspace.

Example SLT.FRANFull range, Archetype NArchetype N [433] is the linear transformation

T : C5 7→ C3, T

x1

x2

x3

x4

x5

=

2x1 + x2 + 3x3 − 4x4 + 5x5

x1 − 2x2 + 3x3 − 9x4 + 3x5

3x1 + 4x3 − 6x4 + 5x5

Page 371: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.RLT Range of a Linear Transformation 339

To determine the elements of C3 in R (T ), find those vectors v such that T (u) = v forsome u ∈ C5,

v = T (u)

=

2u1 + u2 + 3u3 − 4u4 + 5u5

u1 − 2u2 + 3u3 − 9u4 + 3u5

3u1 + 4u3 − 6u4 + 5u5

=

2u1

u1

3u1

+

u2

−2u2

4u2

+

3u3

3u3

4u3

+

−4u4

−9u4

−6u4

+

5u5

3u5

5u5

= u1

213

+ u2

1−24

+ u3

334

+ u4

−4−9−6

+ u5

535

This says that every output of T (v) can be written as a linear combination of the fivevectors 2

13

1−24

334

−4−9−6

535

using the scalars u1, u2, u3, u4, u5. Furthermore, since u can be any element of C5, everysuch linear combination is an output. This means that

R (T ) = Sp

2

13

,

1−24

,

334

,

−4−9−6

,

535

The five vectors in this spanning set for R (T ) form a linearly dependent set (Theo-rem MVSLD [99]). So we can find a more economical presentation by any of the variousmethods from Section RM [127] and Section RSM [141]. We will place the vectors intoa matrix as rows, row-reduce, toss out zero rows and appeal to Theorem BRS [144], sowe can describe the range of T with a (nice) basis,

R (T ) = Sp

1

00

,

010

,

001

4

In contrast to injective linear transformations having small (trivial) null spaces (Theo-rem NSILT [328]), surjective linear transformations have large ranges, as indicated in thenext theorem.

Theorem RSLTRange of a Surjective Linear TransformationSuppose that T : U 7→ V is a linear transformation. Then T is surjective if and only ifthe range of T equals the codomain, R (T ) = V . �

Proof (⇒) By Definition RLT [336], we know that R (T ) ⊆ V . To establish the reverseinclusion, assume v ∈ V . Then since T is surjective (Definition SLT [332]), there exists

Page 372: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.RLT Range of a Linear Transformation 340

a vector u ∈ U so that T (u) = v. However, the existence of u gains v membership inR (T ), so V ⊆ R (T ). Thus, R (T ) = V .

(⇐) To establish that T is surjective, choose v ∈ V . Since we are assuming thatR (T ) = V , v ∈ R (T ). This says there is a vector u ∈ U so that T (u) = v, i.e. T issurjective. �

Example SLT.NSAQRNot surjective, Archetype Q, revisitedWe are now in a position to revisit our first example in this section, Example SLT.NSAQ [332].In that example, we showed that Archetype Q [439] is not surjective by constructing avector in the codomain where no element of the domain could be used to evaluate thelinear transformation to create the output, thus violating Definition SLT [332]. Justwhere did this vector come from?

The short answer is that the vector

v =

−123−14

was constructed to lie outside of the range of T . How was this accomplished? First, therange of T is given by

R (T ) = Sp

10001

,

0100−1

,

0010−1

,

00012

Suppose an element of the range v∗ has its first 4 components equal to −1, 2, 3,−1, inthat order. Then to be an element of R (T ), we would have

v∗ = (−1)

10001

+ (2)

0100−1

+ (3)

0010−1

+ (−1)

00012

=

−123−1−8

So the only vector in the range with these first four components specified, must have −8in the fifth component. To set the fifth component to any other value (say, 4) will resultin a vector (v in Example SLT.NSAQ [332]) outside of the range. Any attempt to findan input for T that will produce v as an output will be doomed to failure.

Whenever the range of a linear transformation is not the whole codomain, we canemploy this device and conclude that the linear transformation is not injective. This isanother way of viewing Theorem RSLT [339]. For a surjective linear transformation, therange is all of the codomain and there is no choice for a vector v that lies in V , yet notin the range. For every one of the archetypes that is not surjective, there is an examplepresented of exactly this form. 4

Example SLT.NSAONot surjective, Archetype O

Page 373: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.RLT Range of a Linear Transformation 341

In Example SLT.RAO [336] the range of Archetype O [435] was determined to be

R (T ) = Sp

10−3−7−2

,

01251

a subspace of dimension 2 in C5. Since R (T ) 6= C5, Theorem RSLT [339] says T is notsurjective. 4

Example SLT.SANSurjective, Archetype NThe range of Archetype N [433] was computed in Example SLT.FRAN [338] to be

R (T ) =

1

00

,

010

,

001

Since the basis for this subspace is the set of standard unit vectors for C3 (Theo-rem SUV [??]), we have R (T ) = C3 and by Theorem RSLT [339], T is surjective. 4

Just as injective linear transformations are allied with linear independence (Theorem ILTLI [329],Theorem ILTB [330]), surjective linear transformations are allied with spanning sets.

Theorem SLTSSurjective Linear Transformations and SpansSuppose that T : U 7→ V is a surjective linear transformation and S = {u1, u2, u3, . . . , ut}spans U . Then R = {T (u1) , T (u2) , T (u3) , . . . , T (ut)} spans R (T ). �

Proof We need to establish that every element of R (T ) can be written as a linearcombination of the vectors in R. To this end, choose v ∈ R (T ). Then there exists avector u ∈ U , such that T (u) = v (Definition SLT [332]).

Because S spans V there are scalars, a1, a2, a3, . . . , at, such that

u = a1u1 + a2u2 + a3u3 + · · ·+ atut

Then

v = T (u) T surjective

= T (a1u1 + a2u2 + a3u3 + · · ·+ atut) S spans V

= a1T (u1) + a2T (u2) + a3T (u3) + . . . + atT (ut) Theorem LTLC [311]

which establishes that R spans the range of T , R (T ). �

TODO: Great way to span range.Elements of the range are precisely those elements of the codomain with non-empty

preimages.

Theorem RPIRange and Pre-ImageSuppose that T : U 7→ V is a linear transformation. Then v ∈ R (T ) if and only ifT−1 (v) 6= ∅. �

Page 374: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.SLTD Surjective Linear Transformations and Dimension 342

Proof (⇒) If v ∈ R (T ), then there is a vector u ∈ U such that T (u) = v. This qualifiesu for membership in T−1 (v), and thus the preimage of v is not empty.

(⇐) Suppose the preimage of v is not empty, so we can choose a vector u ∈ U suchthat T (u) = v. Then v ∈ R (T ). �

Theorem SLTBSurjective Linear Transformations and BasesSuppose that T : U 7→ V is a linear transformation and B = {u1, u2, u3, . . . , um} is abasis of U . Then T is surjective if and only if C = {T (u1) , T (u2) , T (u3) , . . . , T (um)}is a spanning set for V . �

Proof (⇒) Assume T is surjective. Since B is a basis, we know B is a spanning set ofU (Definition B [220]). Then Theorem ILTLI [329] says that C spans R (T ). But thehypothesis that T is surjective means V = R (T ) (Theorem RSLT [339]), so C spans V .

(⇐) Assume that C spans V . To establish that T is surjective, we will show thatevery element of V is an output of T for some input (Definition SLT [332]). Supposethat v ∈ V . As an element of V , we can write v as a linear combination of the spanningset C. So there are are scalars, b1, b2, b3, . . . , bm, such that

v = b1T (u1) + b2T (u2) + b3T (u3) + · · ·+ bmT (um)

Now define the vector u ∈ U by

u = b1u1 + b2u2 + b3u3 + · · ·+ bmum

Then

T (u) = T (b1u1 + b2u2 + b3u3 + · · ·+ bmum)

= b1T (u1) + b2T (u2) + b3T (u3) + · · ·+ bmT (um) Theorem LTLC [311]

= v

So, given any choice of a vector v ∈ V , we can design an input u ∈ U to produce v asan output of T . Thus, by Definition SLT [332], T is surjective. �

Subsection SLTDSurjective Linear Transformations and Dimension

Theorem SLTDSurjective Linear Transformations and DimensionSuppose that T : U 7→ V is a surjective linear transformation. Then dim (U) ≥dim (V ). �

Proof Suppose to the contrary that m = dim (U) < dim (V ) = t. Let B be a basis ofU , which will then contain m vectors. Apply T to each element of B to form a set Cthat is a subset of V . By Theorem SLTB [342], C is spanning set of V with m or fewervectors. So we have a set of m or fewer vectors that span V , a vector space of dimensiont, with m < t. However, this contradicts Theorem G [244], so our assumption is falseand dim (U) ≥ dim (V ). �

Page 375: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection SLT.CSLT Composition of Surjective Linear Transformations 343

Example SLT.NSDATNot surjective by dimension, Archetype TThe linear transformation in Archetype T [444] is

T : P4 7→ P5, T (p(x)) = (x− 2)p(x)

Since dim (P4) = 5 < 6 = dim (P5), T cannot be surjective for then it would violateTheorem SLTD [342]. 4

Notice that the previous example made no use of the actual formula defining the func-tion. Merely a comparison of the dimensions of the domain and codomain are enoughto conclude that the linear transformation is not surjective. Archetype O [435] andArchetype P [437] are two more examples of linear transformations that have “small”domains and “big” codomains, resulting in an inability to create all possible outputs andthus are non-surjective linear transformations.

Subsection CSLTComposition of Surjective Linear Transformations

In Subsection LT.NLTFO [316] we saw how to combine linear transformations to buildnew linear transformations, specifically, how to build the composition of two linear trans-formations (Definition LTC [319]). It will be useful later to know that the compositionof surjective linear transformations is again surjective, so we prove that here.

Theorem CSLTSComposition of Surjective Linear Transformations is SurjectiveSuppose that T : U 7→ V and S : V 7→ W are surjective linear transformations. Then(S ◦ T ) : U 7→ W is a surjective linear transformation. �

Proof That the composition is a linear transformation was established in Theorem CLTLT [319],so we need only establish that the composition is surjective. Applying Definition SLT [332],choose w ∈ W .

Because S is surjective, there must be a vector v ∈ V , such that S (v) = w. Withthe existence of v established, that T is injective guarantees a vector u ∈ U such thatT (u) = v. Now,

(S ◦ T ) (u) = S (T (u)) Definition LTC [319]

= S (v) T surjective

= w S surjective

This establishes that any element of the codomain (w) can be created by evaluating S ◦Twith the right input (u). Thus, by Definition SLT [332], T is surjective. �

Page 376: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section IVLT Invertible Linear Transformations 344

Section IVLT

Invertible Linear Transformations

In this section we will conclude our introduction to linear transformations by bringingtogether the twin properties of injectivity and surjectivity and consider linear transfor-mations with both of these properties.

Subsection IVLTInvertible Linear Transformations

One preliminary definition, and then we will have our main definition for this section.

Definition IDLTIdentity Linear TransformationThe identity linear transformation on the vector space W is defined as

IW : W 7→ W, IW (w) = w �

Informally, IW is the “do-nothing” function. You should check that IW is really a lineartransformation, as claimed, and then compute its null space and range to see that it isboth injective and surjective. All of these facts should be straightforward to verify. Withthis in hand we can make our main definition.

Definition IVLTInvertible Linear TransformationsSuppose that T : U 7→ V is a linear transformation. If there is a function S : V 7→ Usuch that

S ◦ T = IU T ◦ S = IV

then T is invertible. In this case, we call S the inverse of T and write S = T−1. �

Informally, a linear transformation T is invertible if there is a companion linear trans-formation, S, which ”undoes” the action of T . When the two linear transformations areapplied consecutively (composition), in either order, the result is to have no real effect.It is entirely analgous to squaring a number and then taking its (positive) square root.

Here is an example of a linear transformation that is invertible. As usual at thebeginning of a section, do not be concerned with where S came from, just understandhow it illustrates Definition IVLT [344].

Example IVLT.AIVLTAn invertible linear transformationArchetype V [444] is the linear transformation

T : P3 7→ M22, T(a + bx + cx2 + dx3

)=

[a + b a− 2c

d b− d

]

Page 377: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.IVLT Invertible Linear Transformations 345

Define the function S : M22 7→ P3 defined by

S

([a bc d

])= (a− c− d) + (c + d)x +

1

2(a− b− c− d)x2 + cx3

Then

(T ◦ S)

([a bc d

])= T

(S

([a bc d

]))= T

((a− c− d) + (c + d)x +

1

2(a− b− c− d)x2 + cx3

)=

[(a− c− d) + (c + d) (a− c− d)− 2(1

2(a− b− c− d))

c (c + d)− c

]=

[a bc d

]= IM22

([a bc d

])

And

(S ◦ T )(a + bx + cx2 + dx3

)= S

(T(a + bx + cx2 + dx3

))= S

([a + b a− 2c

d b− d

])= ((a + b)− d− (b− d)) + (d + (b− d))x

+

(1

2((a + b)− (a− 2c)− d− (b− d))

)x2 + (d)x3

= a + bx + cx2 + dx3

= IP3

(a + bx + cx2 + dx3

)For now, understand why these computations show that T is invertible, and that S = T−1.Maybe even be amazed by how S works so perfectly in concert with T ! We will see laterjust how to arrive at the correct form of S (when it is possible). 4

It can be as instructive to study a linear transformation that is not invertible.

Example IVLT.ANILTA non-invertible linear transformationConsider the linear transformation T : C3 7→ M22 defined by

T

abc

=

[a− b 2a + 2b + c

3a + b + c −2a− 6b− 2c

]Suppose we were to search for an inverse function S : M22 7→ C3.

First verify that the 2 × 2 matrix A =

[5 38 2

]is not in the range of T . This will

Page 378: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.IVLT Invertible Linear Transformations 346

amount to finding an input to T ,

abc

, such that

a− b = 5

2a + 2b + c = 3

3a + b + c = 8

−2a− 6b− 2c = 2

As this system of equations is inconsistent, there is no input column vector, and A 6∈R (T ). How should we define S (A)? Note that

T (S (A)) = (T ◦ S) (A) = IM22 (A) = A

So any definition we would provide for S (A) must then be a column vector that T sendsto A and we would have A ∈ R (T ), contrary to the definition of T . This is enough tosee that there is no function S that will allow us to conclude that T is invertible, sincewe cannot provide a consistent definition for S (A) if we assume T is invertible.

Even though we now know that T is not invertible, let’s not leave this example justyet. Check that

T

1−24

=

[3 25 2

]= B T

0−38

=

[3 25 2

]= B

How would we define S (B)?

S (B) = S

T

1−24

= (S ◦ T )

1−24

= IC3

1−24

=

1−24

or

S (B) = S

T

0−38

= (S ◦ T )

0−38

= IC3

0−38

=

0−38

Which definition should we provide for S (B)? Both are necessary. But then S is not afunction. So we have a second reason to know that there is no function S that will allowus to conclude that T is invertible. It happens that there are infinitely many columnvectors that S would have to take to B. Construct the null space of T ,

N (T ) = Sp

−1−14

Now choose either of the two inputs used above for T and add to it a scalar multiple ofthe basis vector for the null space of T . For example,

x =

1−24

+ (−2)

−1−14

=

30−4

Page 379: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.IVLT Invertible Linear Transformations 347

then verify that T (x) = B. Practice creating a few more inputs for T that would besent to B, and see why it is hopeless to think that we could ever provide a reasonabledefinition for S (B)! There is a “whole subspace’s worth” of values that S (B) wouldhave to take on. 4

In Example IVLT.ANILT [345] you may have noticed that T is not surjective, since thematrix A was not in the range of T . And T is not injective since there are two differentinput column vectors that T sends to the matrix B. Linear transformations T that arenot surjective lead to inverse putative functions S that are undefined on inputs outside ofthe range of T . Linear transformations T that are not injective lead to putative inversefunctions S that are multiply-defined on each of their inputs. We will formalize theseideas in Theorem ILTIS [348].

But first notice in Definition IVLT [344] that we only require the inverse (when itexists) to be a function. When it does exist, it too is a linear transformation.

Theorem ILTLTInvese of a Linear Transformation is a Linear TransformationSuppose that T : U 7→ V is an invertible linear transformation. Then the functionT−1 : V 7→ U is a linear transformation. �

Proof We work through verifying Definition LT [302] for T−1, employing as we go prop-erties of T given by Definition LT [302]. To this end, suppose x, y ∈ V and α ∈ C.

T−1 (x + y) = T−1(T(T−1 (x)

)+ T

(T−1 (y)

))T , T−1 inverse functions

= T−1(T(T−1 (x) + T−1 (y)

))T a linear transformation

= T−1 (x) + T−1 (y) T−1, T inverse functions

Now check the second defining property of a linear transformation for T−1,

T−1 (αx) = T−1(αT(T−1 (x)

))T , T−1 inverse functions

= T−1(T(αT−1 (x)

))T a linear transformation

= αT−1 (x) T−1, T inverse functions �

So T−1 fullfills the requirements of Definition LT [302] and is therefore a linear trans-formation. So when T has an inverse, T−1 is also a linear transformation. Additionally,T−1 is invertible and its inverse is what you might expect.

Theorem IILTInverse of an Invertible Linear TransformationSuppose that T : U 7→ V is an invertible linear transformation. Then T−1 is an invertiblelinear transformation and (T−1)

−1= T . �

Proof Because T is invertible, Definition IVLT [344] tells us there is a function T−1 :V 7→ U such that

T−1 ◦ T = IU T ◦ T−1 = IV

Additionally, Theorem ILTLT [347] tells us that T−1 is more than just a function, itis a linear transformation. Now view these two statements as properties of the lineartransformation T−1. In light of Definition IVLT [344], they together say that T−1 isinvertible (let T play the role of S in the statement of the definition). Furthermore, theinverse of T−1 is then T , i.e. (T−1)

−1= T . �

Page 380: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.IV Invertibility 348

Subsection IVInvertibility

We now know what an inverse linear transformation is, but just which linear transfor-mations have inverses? Here is a theorem we have been preparing for all chapter.

Theorem ILTISInvertible Linear Transformations are Injective and SurjectiveSuppose T : U 7→ V is a linear transformation. Then T is invertible if and only if T isinjective and surjective. �

Proof (⇒) Since T is presumed invertible, we can employ its inverse, T−1 (Defini-tion IVLT [344]). To see that T is injective, suppose x, y ∈ U

T (x) = T (y) Definition ILT [321]

T−1 (T (x)) = T−1 (T (y)) Apply T−1(T−1 ◦ T

)(x) =

(T−1 ◦ T

)(y) Definition LTC [319]

IU (x) = IU (y) T−1, T inverse functions

x = y Definition IDLT [344]

So by Definition ILT [321] T is injective. To check that T is surjective, suppose v ∈ V .Employ T−1 by defining u = T−1 (v). Then

T (u) = T(T−1 (v)

)Substitution for u

=(T ◦ T−1

)(v) Definition LTC [319]

= IV (v) T , T−1 inverse functions

= v Definition IDLT [344]

So there is an input to T , u, that produces the chosen output, v, and hence T is surjectiveby Definition SLT [332].

(⇐) Now assume that T is both injective and surjective. We will build a functionS : V 7→ U that will establish that T is invertible. To this end, choose any v ∈ V .Since T is surjective, Theorem RSLT [339] says R (T ) = V , so we have v ∈ R (T ).Theorem RPI [341] says that the pre-image of v, T−1 (v), is nonempty. So we can choosea vector from the pre-image of v, say u. In other words, there exists u ∈ T−1 (v).

Since T−1 (v) is non-empty, Theorem NSPI [327] then says that

T−1 (v) = {u + z | z ∈ N (T )}

However, because T is injective, by Theorem NSILT [328] the null space is trivial, N (T ) ={0}. So the pre-image is a set with just one element, T−1 (v) = {u}. Now we can defineS by S (v) = u. This is the key to this half of this proof. Normally the preimage ofa vector from the codomain might be an empty set, or an infinite set. But surjectivityrequires that the preimage not be empty, and then injectivity limits the preimage to asingleton. Since our choice of v was arbitrary, we know that every pre-image for T is aset with a single element. This allows us to construct S as a function. Now that it isdefined, verifying that it is the inverse of T will be easy. Here we go.

Page 381: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.IV Invertibility 349

Choose u ∈ U . Define v = T (u). Then T−1 (v) = {u}, so that S (v) = u and,

(S ◦ T ) (u) = S (T (u)) = S (v) = u = IU (u)

and since our choice of u was arbitary we have function equality, S ◦ T = IU .Now choose v ∈ V . Define u to be the single vector in the set T−1 (v), in other words,

u = S (v). Then T (u) = v, so

(T ◦ S) (v) = T (S (v)) = T (u) = v = IV (v) �

and since our choice of v was arbitary we have function equality, T ◦ S = IV .

We will make frequent use of this characterization of invertible linear transformations.The next theorem is a good example of this, and we will use it often, too.

Theorem CIVLTComposition of Invertible Linear TransformationsSuppose that T : U 7→ V and S : V 7→ W are invertible linear transformations. Thenthe composition, (S ◦ T ) : U 7→ W is an invertible linear transformation. �

Proof Since S and T are both linear transformations, S◦T is also a linear transformationby Theorem CLTLT [319]. Since S and T are both invertible, Theorem ILTIS [348] saysthat S and T are both injective and surjective. Then Theorem CILTI [331] says S ◦ T isinjective, and Theorem CSLTS [343] says S ◦T is surjective. Now apply the “other half”of Theorem ILTIS [348] and conclude that S ◦ T is invertible. �

When a composition is invertible, the inverse is easy to construct.

Theorem ICLTInverse of a Composition of Linear TransformationsSuppose that T : U 7→ V and S : V 7→ W are invertible linear transformations. ThenS ◦ T is invertible and (S ◦ T )−1 = T−1 ◦ S−1. �

Proof Compute, for all w ∈ W((S ◦ T ) ◦

(T−1 ◦ S−1

))(w) = S

(T(T−1

(S−1 (w)

)))= S

(IV

(S−1 (w)

))Definition IVLT [344]

= S(S−1 (w)

)Definition IDLT [344]

= w Definition IVLT [344]

= IW (w) Definition IDLT [344]

so (S ◦ T ) ◦ (T−1 ◦ S−1) = IW and also((T−1 ◦ S−1

)◦ (S ◦ T )

)(u) = T−1

(S−1 (S (T (u)))

)= T−1 (IV (T (u))) Definition IVLT [344]

= T−1 (T (u)) Definition IDLT [344]

= u Definition IVLT [344]

= IU (u) Definition IDLT [344]

so (T−1 ◦ S−1)◦(S ◦ T ) = IU . By Definition IVLT [344], S◦T is invertible and (S ◦ T )−1 =T−1 ◦ S−1. �

Page 382: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.SI Structure and Isomorphism 350

Notice that this theorem not only establishes what the inverse of S◦T is, it also duplicatesthe conclusion of Theorem CIVLT [349] and also establishes the invertibility of S ◦ T .But somehow, the proof of Theorem CIVLT [349] is nicer way to get this property.

Does this Theorem ICLT [349] remind you of the flavor of any theorem we have seenabout matrices? (Hint: Think about getting dressed.) Hmmmm.

Subsection SIStructure and Isomorphism

A vector space is defined (Definition VS [184]) as a set of objects (“vectors”) endowedwith a definition of vector addition (+) and a definition of scalar multiplication (jux-taposition). Many of our definitions about vector spaces involve linear combinations(Definition LC [202]), such as the span of a set (Definition SS [203]) and linear inde-pendence (Definition LI [212]). Other definitions are built up from these ideas, such asbases (Definition B [220]) and dimension (Definition D [230]). The defining propertiesof a linear transformation require that a function “respect” the operations of the twovector spaces that are the domain and the codomain (Definition LT [302]). Finally, aninvertible linear transformation is one that can be “undone” — it has a companion thatreverses its effect. In this subsection we are going to begin to roll all these ideas into one.

A vector space has “structure” derived from definitions of the two operations and therequirement that these operations interact in ways that satisfy the ten axioms of Defini-tion VS [184]. When two different vector spaces have an invertible linear transformationdefined between them, then we can translate questions about linear combinations (spans,linear independence, bases, dimension) from the first vector space to the second. Theanswers obtained in the second vector space can then be translated back, via the inverselinear transformation, and interpreted in the setting of the first vector space. We say thatthese invertible linear transformations “preserve structure.” And we say that the twovector spaces are “structurally the same.” The precise term is “isomorphic,” from Greekmeaning “of the same form.” Let’s begin to try to understand this important concept.

Definition IVSIsomorphic Vector SpacesTwo vector spaces U and V are isomorphic if there exists an invertible linear transfor-mation T with domain U and codomain V , T : U 7→ V . In this case, we write U ∼= V ,and the linear transformation T is known as an isomorphism between U and V . �

A few comments on this definition. First, be careful with your language (Technique L [19]).Two vector spaces are isomrphic, or not. It is a yes/no situation and the term only ap-plies to a pair of vector spaces. Any invertible linear transformation can be called anisomorphism, it is a term that applies to functions. Second, a given pair of vector spacesthere might be several different isomorphisms between the two vector spaces. But it onlytakes the existence of one to call the pair isomorphic. Third, U isomorphic to V , or Visomorphic to U? Doesn’t matter, since the inverse linear transformation will providethe needed isomorphism in the “opposite” direction. Being “isomorphic to” is an equiv-alence relation on the set of all vector spaces (see Theorem SER [293] for a reminderabout equivalence relations).

Page 383: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.SI Structure and Isomorphism 351

Example IVLT.IVSAVIsomorphic vector spaces, Archetype VArchetype V [444] is a linear transformation from P3 to M22,

T : P3 7→ M22, T(a + bx + cx2 + dx3

)=

[a + b a− 2c

d b− d

]Since it is injective and surjective, Theorem ILTIS [348] tells us that it is an invertiblelinear transformation. By Definition IVS [350] we say P3 and M22 are isomorphic.

At a basic level, the term “isomorphic” is nothing more than a codeword for thepresence of an invertible linear transformation. However, it is also a description of apowerful idea, and this power only becomes apparent in the course of studying examplesand related theorems. In this example, we are led to believe that there is nothing “struc-turally” different about P3 and M22. In a certain sense they are the same. Not equal,but the same. One is as good as the other. One is just as interesting as the other.

Here is an extremely basic application of this idea. Suppose we want to compute thefollowing linear combination of polynomials in P3,

5(2 + 3x− 4x2 + 5x3) + (−3)(3− 5x + 3x2 + x3)

Rather than doing it straight-away (which is very easy), we will apply the transformationT to convert into a linear combination of matrices, and then compute in M22 accordingto the definitions of the vector space axioms there (Example VS.VSM [186]),

T(5(2 + 3x− 4x2 + 5x3) + (−3)(3− 5x + 3x2 + x3)

)= 5T

(2 + 3x− 4x2 + 5x3

)+ (−3)T

(3− 5x + 3x2 + x3

)Theorem LTLC [311]

= 5

[5 105 −2

]+ (−3)

[−2 −31 −6

]Definition of T

=

[31 5922 8

]Operations in M22

Now we will translate our answer back to P3 by appling T−1, which we found in Exam-ple IVLT.AIVLT [344],

T−1 : M22 7→ P3, T−1

([a bc d

])= (a− c− d) + (c + d)x +

1

2(a− b− c− d)x2 + cx3

We compute,

T−1

([31 5922 8

])= 1 + 30x− 29x2 + 22x3

which is, as expected, exactly what we would have computed for the original linear combi-nation had we just used the definitions of the operations in P3 (Example VS.VSP [186]).4

Checking the dimensions of two vector spaces can be a quick way to establish that theyare not isomorphic. Here’s the theorem.

Theorem IVSEDIsomorphic Vector Spaces have Equal DimensionSuppose U and V are isomorphic vector spaces. Then dim (U) = dim (V ). �

Page 384: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.RNLT Rank and Nullity of a Linear Transformation 352

Proof If U and V are isomorphic, there is an invertible linear transformation T : U 7→V (Definition IVS [350]). T is injective by Theorem ILTIS [348] and so by Theo-rem ILTD [330], dim (U) ≤ dim (V ). Similarly, T is surjective by Theorem ILTIS [348]and so by Theorem SLTD [342], dim (U) ≥ dim (V ). The net effect of these two inequal-ities is that dim (U) = dim (V ). �

The contrapositive of Theorem IVSED [351] says that if U and V have different dimen-sions, then they are not isomorphic. Dimension is the simplest “structural” characteristicthat will allow you to distinguish non-isomorphic vector spaces. For example P6 is notisomorphic to M34 since their dimensions (7 and 12, respectively) are not equal. Withtools developed in Section VR [??] we will be able to establish that the converse ofTheorem IVSED [351] is true. Think about that one for a moment.

Subsection RNLTRank and Nullity of a Linear Transformation

Just as a matrix has a rank and a nullity, so too do linear transformations. And justlike the rank and nullity of a matrix are related (they sum to the number of columns,Theorem RPNC [238]) the rank and nullity of a linear transformation are related. Hereare the definitions and theorems, see the Archetypes (Chapter A [375]) for loads ofexamples.

Definition ROLTRank Of a Linear TransformationSuppose that T : U 7→ V is a linear transformation. Then the rank of T , r (T ), is thedimension of the range of T ,

r (T ) = dim (R (T )) �

Definition NOLTNullity Of a Linear TransformationSuppose that T : U 7→ V is a linear transformation. Then the nullity of T , n (T ), is thedimension of the null space of T ,

n (T ) = dim (N (T )) �

Here are two quick theorems.

Theorem ROSLTRank Of a Surjective Linear TransformationSuppose that T : U 7→ V is a linear transformation. Then the rank of T is the dimensionof V , r (T ) = dim (V ), if and only if T is surjective. �

Proof By Theorem RSLT [339], T is surjective if and only if R (T ) = V . ApplyingDefinition ROLT [352], R (T ) = V if and only if r (T ) = dim (R (T )) = dim (V ). �

Theorem NOILTNullity Of an Injective Linear TransformationSuppose that T : U 7→ V is an injective linear transformation. Then the nullity of T iszero, n (T ) = 0, if and only if T is injective. �

Page 385: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.RNLT Rank and Nullity of a Linear Transformation 353

Proof By Theorem NSILT [328], T is injective if and only if N (T ) = {0}. ApplyingDefinition NOLT [352], N (T ) = {0} if and only if n (T ) = 0. �

Just as injectivity and surjectivity come together in invertible linear transformations,there is a clear relationship between rank and nullity of a linear transformation. If oneis big, the other is small.

Theorem RPNDDRank Plus Nullity is Domain DimensionSuppose that T : U 7→ V is a linear transformation. Then

r (T ) + n (T ) = dim (U) �

Proof Let r = r (T ) and s = n (T ). Suppose that R = {v1, v2, v3, . . . , vr} ⊆ V is abasis of the range of T , R (T ), and S = {u1, u2, u3, . . . , us} ⊆ U is a basis of the nullspace of T , N (T ). Note that R and S are possibly empty, which means that some of thesums in this proof are “empty” and are equal to the zero vector.

Because the elements of R are all in the range of T , each must have a non-empty pre-image by Theorem RPI [341]. Choose vectors wi ∈ U , 1 ≤ i ≤ r such that wi ∈ T−1 (vi).So T (wi) = vi, 1 ≤ i ≤ r. Consider the set

B = {u1, u2, u3, . . . , us, w1, w2, w3, . . . , wr}

We claim that B is a basis for U .To establish linear independence for B, begin with a relation of linear dependence on

B. So suppose there are scalars a1, a2, a3, . . . , as and b1, b2, b3, . . . , br

0 = a1u1 + a2u2 + a3u3 + · · ·+ asus + b1w1 + b2w2 + b3w3 + · · ·+ brwr

Then

0 = T (0) Theorem LTTZZ [306]

= T (a1u1 + a2u2 + a3u3 + · · ·+ asus

+b1w1 + b2w2 + b3w3 + · · ·+ brwr) Substitution

= a1T (u1) + a2T (u2) + a3T (u3) + · · ·+ asT (us)

+ b1T (w1) + b2T (w2) + b3T (w3) + · · ·+ brT (wr) Theorem LTLC [311]

= a10 + a20 + a30 + · · ·+ as0

+ b1T (w1) + b2T (w2) + b3T (w3) + · · ·+ brT (wr) ui ∈ N (T )

= 0 + 0 + 0 + · · ·+ 0

+ b1T (w1) + b2T (w2) + b3T (w3) + · · ·+ brT (wr) Theorem ZVSM [192]

= b1T (w1) + b2T (w2) + b3T (w3) + · · ·+ brT (wr) Property zero vector

= b1v1 + b2v2 + b3v3 + · · ·+ brvr wi ∈ T−1 (vi)

This is a relation of linear dependence on R (Definition RLD [212]), and since R is alinearly independent set (Definition LI [212]), we see that b1 = b2 = b3 = . . . = br = 0.Then the original relation of linear dependence on B becomes

0 = a1u1 + a2u2 + a3u3 + · · ·+ asus + 0w1 + 0w2 + . . . + 0wr

= a1u1 + a2u2 + a3u3 + · · ·+ asus + 0 + 0 + . . . + 0 Theorem ZSSM [191]

= a1u1 + a2u2 + a3u3 + · · ·+ asus Property zero vector

Page 386: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.RNLT Rank and Nullity of a Linear Transformation 354

But this is again a relation of linear independence (Definition RLD [212]), now on theset S. Since S is linearly independent (Definition LI [212]), we have a1 = a2 = a3 =. . . = ar = 0. Since we now know that all the scalars in the relation of linear depen-dence on B must be zero, we have established the linear independence of S throughDefinition LI [212].

To now establish that B spans U , choose an arbitrary vector u ∈ U . Then T (u) ∈R(T ), so there are scalars c1, c2, c3, . . . , cr such that

T (u) = c1v1 + c2v2 + c3v3 + · · ·+ crvr

Use the scalars c1, c2, c3, . . . , cr to define a vector y ∈ U ,

y = c1w1 + c2w2 + c3w3 + · · ·+ crwr

Then

T (u− y) = T (u)− T (y) Theorem LTLC [311]

= T (u)− T (c1w1 + c2w2 + c3w3 + · · ·+ crwr) Substitution

= T (u)− (c1T (w1) + c2T (w2) + · · ·+ crT (wr)) Theorem LTLC [311]

= T (u)− (c1v1 + c2v2 + c3v3 + · · ·+ crvr) wi ∈ T−1 (vi)

= T (u)− T (u) Substitution

= 0 Additive inverses

So the vector u − y is sent to the zero vector by T and hence is an element of the nullspace of T . As such it can be written as a linear combination of the basis vectors forN (T ), the elements of the set S. So there are scalars d1, d2, d3, . . . , ds such that

u− y = d1u1 + d2u2 + d3u3 + · · ·+ dsus

Then

u = (u− y) + y

= d1u1 + d2u2 + d3u3 + · · ·+ dsus + c1w1 + c2w2 + c3w3 + · · ·+ crwr

This says that for any vector, u, from U , there exist scalars (d1, d2, d3, . . . , ds, c1, c2, c3, . . . , cr)that form u as a linear combination of the vectors in the set B. In other words, B spansU (Definition SS [203]).

So B is a basis (Definition B [220]) of U with s + r vectors, and thus

dim (U) = s + r = n (T ) + r (T )

as desired. �

Theorem RPNC [238] said that the rank and nullity of a matrix sum to the number ofcolumns of the matrix. This result is now an easy consequence of Theorem RPNDD [353]when we consider the linear transformation T : Cn 7→ Cm defined with the m×n matrixA by T (x) = Ax. The range and null space of T are identical to the range and nullspace of the matrix A (can you prove this?), so the rank and nullity of the matrix A areidentical to the rank and nullity of the linear transformation T . The dimension of thedomain of T is the dimension of Cn, exactly the number of columns for the matrix A.

Page 387: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.SLELT Systems of Linear Equations and Linear Transformations355

This theorem can be especially useful in determing basic properties of linear transfor-mations. For example, suppose that T : C6 7→ C6 is a linear transformation and you areable to quickly establish that the null space is trivial. Then n (T ) = 0. First this meansthat T is injective by Theorem NOILT [353]. Also, Theorem RPNDD [353] becomes

6 = dim(C6)

= r (T ) + n (T ) = r (T ) + 0 = r (T )

So the rank of T is equal to the rank of the codomain, and by Theorem ROSLT [352]we know T is surjective. Finally, we know T is invertible by Theorem ILTIS [348].So from the determination that the null space is trivial, and consideration of variousdimensions, the theorems of this section allow us to conclude the existence of an inverselinear transformation for T .

Similarly, Theorem RPNDD [353] can be used to provide alternative proofs for The-orem ILTD [330], Theorem SLTD [342] and Theorem IVSED [351]. It would be aninteresting exercise to construct these proofs.

It would be instructive to study the archetypes that are linear transformations and seehow many of their properties can be deduced just from considering the dimensions of thedomain and codomain, and possibly with just the nullity or rank. The table precedingall of the archetypes could be a good place to start this analysis.

Subsection SLELTSystems of Linear Equations and Linear Transformations

This subsection does not really belong in this section, or any other section, for thatmatter. Its just the right time to have a discussion about the connections between thecentral topic of linear algebra, linear transformations, and our motivating topic fromChapter SLE [2], systems of linear equations. We will discuss several theorems we haveseen already, but we will also make some forward-looking statements that will be justifiedin Chapter R [358].

Archetype D [393] and Archetype E [397] are ideal examples to illustrate connectionswith linear transformations. Both have the same coefficient matrix,

D =

2 1 7 −7−3 4 −5 −61 1 4 −5

To apply the theory of linear transformations to these two archetypes, employ matrixmultiplication (Definition MM [153]) and define the linear transformation,

T : C4 7→ C2, T (x) = Dx = x1

2−31

+ x2

141

+ x3

7−54

+ x4

−7−6−5

Theorem MBLT [308] tells us that T is indeed a linear transformation. Archetype D [393]

asks for solutions to LS(D, b), where b =

8−12−4

. In the language of linear transforma-

tions this is equivalent to asking for T−1 (b). In the language of vectors and matrices it

Page 388: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.SLELT Systems of Linear Equations and Linear Transformations356

asks for a linear combination of the four columns of D that will equal b. One solution

listed is w =

7813

. With a non-empty preimage, Theorem NSPI [327] tells us that the

complete solution set of the linear system is the preimage of b,

w + N (T ) = {w + z | z ∈ N (T )}

The null space of the linear transformation T is exactly the null space of the matrixD (Theorem XX [??]), so this approach to the solution set should be reminiscent ofTheorem PSPHS [161]. The null space of the linear transformation is the preimage ofthe zero vector, exactly equal to the solution set of the homogeneous system LS(D, 0).Since D has a null space of dimension two, every preimage (and in particular the preimageof b) is as “big” as a subspace of dimension two (but is not a subspace).

Archetype E [397] is identical to Archetype D [393] but with a different vector of

constants, d =

232

. We can use the same linear transformation T to discuss this system

of equations since the coefficient matrix is identical. Now the set of solutions to LS(D, d)is the pre-image of d, T−1 (d). However, the vector d is not in the range of the lineartransformation (nor is it in the range of the matrix, since these two ranges are equal setsby Theorem XX [??]). So the empty pre-image is equivalent to the inconsistency of thelinear system.

These two archetypes each have three equations in four variables, so either the re-sulting linear systems are inconsistent, or they are consistent and application of Theo-rem CMVEI [44] tells us that the system has infinitely many solutions. Considering thesesame parameters for the linear transformation, the dimension of the domain, C4, is four,while the codomain, C3, has dimension three. Then

n (T ) = dim(C4)− r (T ) Theorem RPNDD [353]

= 4− dim (R (T )) Definition ROLT [352]

≥ 4− 3 R (T ) subspace of C3

= 1

So the null space of T is nontrivial simply by considering the dimensions of the do-main (number of variables) and the codomain (number of equations). Pre-images ofelements of the codomain that are not in the range of T are empty (inconsistent sys-tems). For elements of the codomain that are in the range of T (consistent systems),Theorem NSPI [327] tells us that the pre-images are built from the null space, and witha non-trivial null space, these pre-images are infinite (infinitely many solutions).

When do systems of equations have unique solutions? Consider the system of linearequations LS(C, f) and the linear transformation S (x) = Cx. If S has a trivial nullspace, then pre-images will either be empty or be finite sets with single elements. Corre-spondingly, the coefficient matrix C will have a trivial null space and solution sets witheither be empty (inconsistent) or contain a single solution (unique solution). Should thematrix be square and have a non-trivial null space then we recognize the matrix as beingnonsingular. A square matrix means that the corresponding linear transformation, T ,has equal-sized domain and codomain. With a nullity of zero, T is injective, and also

Page 389: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection IVLT.SLELT Systems of Linear Equations and Linear Transformations357

Theorem RPNDD [353] tells us that rank of T is equal to the dimension of the domain,which in turn is equal to the dimension of the codomain. In other words, T is surjective.Injective and surjective, and Theorem ILTIS [348] tells us that T is invertible. Just aswe can use the inverse of the coefficient matrix to find the unique solution of any linearsystem with a nonsingular coefficient matrix (Theorem SNSCM [178]), we can use theinverse of the linear transformation to construct the unique element of any pre-image(proof of Theorem ILTIS [348]).

The executive summary of this discussion is that to every coefficient matrix of asystem of linear equations we can associate a natural linear transformation. Solutionsets for systems with this coefficient matrix are preimages of elements of the codomain ofthe linear transformation. For every theorem about systems of linear equations there isan analogue about linear transformations. The theory of linear transformations providesall the tools to recreate the theory of solutions to linear systems of equations.

We will continue this adventure in Chapter R [358].

Page 390: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

R: Representations

Section VR

Vector Representations

Definition VRVector RepresentationSuppose that V is a vector space with a basis B = {v1, v2, v3, . . . , vn}. Define a functionρB : V 7→ Cn as follows. For w ∈ V , find some scalars a1, a2, a3, . . . , an so that

w = a1v1 + a2v2 + a3v3 + · · ·+ anvn

then

ρB (w) =

a1

a2

a3...

an

We need to show that ρB is really a function (since “find some scalars” sounds like it couldbe accomplished in many ways, or perhaps not at all) and right now we want to establishthat ρB is a linear transformation. We will wrap up both objectives in one theorem, eventhough the first part is working backwards to make sure that ρB is well-defined.

Theorem VRLTVector Representation is a Linear TransformationThe function ρB (Definition VR [358]) is a linear transformation. �

Proof The definition of ρB (Definition VR [358]) appears to allow considerable latitudein selecting the scalars a1, a2, a3, . . . , an. However, since B is a basis for V , Theo-rem VRRB [227] says this can be done, and done uniquely. So despite appearances, ρB

is indeed a function.

Suppose that x and y are two vectors in V and α ∈ C. Then the vector space axioms(Definition VS [184]) assure us that the vectors x + y and αx are also vectors in V .Theorem VRRB [227] then provides the follow sets of scalars for the four vectors x, y,x + y and αx, and tells us that each set of scalars is the only way to express the given

358

Page 391: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section VR Vector Representations 359

vector as a linear combination of the basis vectors in B.

x = a1v1 + a2v2 + a3v3 + · · ·+ anvn

y = b1v1 + b2v2 + b3v3 + · · ·+ bnvn

x + y = c1v1 + c2v2 + c3v3 + · · ·+ cnvn

αx = d1v1 + d2v2 + d3v3 + · · ·+ dnvn

Then

c1v1 + c2v2 + c3v3 + · · ·+ cnvn

= x + y

= (a1v1 + a2v2 + a3v3 + · · ·+ anvn)

+ (b1v1 + b2v2 + b3v3 + · · ·+ bnvn)

= a1v1 + b1v1 + a2v2 + b2v2

+ a3v3 + b3v3 + · · ·+ anvn + bnvn Commutativity in V

= (a1 + b1)v1 + (a2 + b2)v2

+ (a3 + b3)v3 + · · ·+ (an + bn)vn Distributivity in V

By the uniqueness of the expression of x+y as a linear combination of the vectors in B,we conclude that ci = ai + bi, 1 ≤ i ≤ n. So employing our definition of vector additionin Cm (Definition CVA [67]), we have

ρB (x + y) =

c1

c2

c3...cn

=

a1 + b1

a2 + b2

a3 + b3...

an + bn

=

a1

a2

a3...

an

+

b1

b2

b3...bn

= ρB (x) + ρB (y)

so we have the first necessary property for ρB to be a linear transformation (Defini-tion LT [302]) . Similarly,

d1v1 + d2v2 + d3v3 + · · ·+ dnvn

= αx

= α (a1v1 + a2v2 + a3v3 + · · ·+ anvn)

= α(a1v1) + α(a2v2) + α(a3v3) + · · ·+ α(anvn) Distributivity in V

= (αa1)v1 + (αa2)v2 + (αa3)v3 + · · ·+ (αan)vn Associativity in V

By the uniqueness of the expression of αx as a linear combination of the vectors in B, weconclude that di = αai, 1 ≤ i ≤ n. So employing our definition of scalar multiplicationin Cm (Definition CVSM [68]), we have

ρB (αx) =

d1

d2

d3...

dn

=

αa1

αa2

αa3...

αan

= α

a1

a2

a3...

an

= αρB (x)

Page 392: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section VR Vector Representations 360

and so, with the second property of a linear transformation (Definition LT [302]) con-firmed we can conclude that ρB is a linear transformation. �

Example VR.VRC4Vector representation in C4

Consider the vector y ∈ C4

y =

61467

We will find several coordinate representations of y in this example. Notice that y neverchanges, but the representations of y do change.

One basis for C4 is

B = {u1, u2, u3, u4} =

−212−3

,

3−62−4

,

1205

,

4316

as can be seen by making these vectors the column of a matrix, checking that the matrixis nonsingular and applying Theorem CNSMB [225]. To find ρB (y), we need to findscalars, a1, a2, a3, a4 such that

y = a1u1 + a2u2 + a3u3 + a4u4

By Theorem SLSLC [76] the desired scalars are a solution to the linear system of equa-tions with a coefficient matrix whose columns are the vectors in B and with a vector ofconstants y. With a nonsingular coefficient matrix, the solution is unique, but this is nosurprise as this is the content of Theorem VRRB [227]. This unique solution is

a1 = 2 a2 = −1 a3 = −3 a4 = 4

Then by Definition VR [358], we have

ρB (y) =

2−1−34

Suppose now that we construct a representation of y relative to another basis of C4,

C =

−159−4−2

,

16−1452

,

−2614−6−3

,

14−1346

As with B, it is easy to check that C is a basis. Writing y as a linear combination ofthe vectors in C leads to solving a system of four equations in the four unknown scalarswith a nonsingular coefficient matrix. The unique solution can be expressed as

y =

−6−14−6−7

= (−28)

−159−4−2

+ (−8)

16−1452

+ 11

−2614−6−3

+ 0

14−1346

Page 393: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section VR Vector Representations 361

so that Definition VR [358] gives

ρC (y) =

−28−8110

We often perform representations relative to standard bases, but for vectors in Cm itsa little silly. Let’s find the vector representation of y relative to the standard basis(Theorem SUVB [220]),

D = {e1, e2, e3, e4}

Then, without any computation, we can check that

y =

61467

= 6e1 + 14e2 + 6e3 + 7e4

so by Definition VR [358],

ρC (y) =

61467

which is not very exciting. Notice however that the order in which we place the vectorsin the basis is critical to the representation. Let’s keep the standard unit vectors as ourbasis, but rearrange the order we place them in the basis. So a fourth basis is

E = {e3, e4, e2, e1}

Then,

y =

61467

= 6e3 + 7e4 + 14e2 + 6e1

so by Definition VR [358],

ρE (y) =

67146

So for every basis we could find for C4 we could construct a representation of y. 4

Vector representations are most interesting for vector spaces that are not Cm.

Example VR.VRP2Vector representations in P2

Consider the vector u = 15 + 10x− 6x2 ∈ P2 from the vector space of polynomials withdegree at most 2 (Example VS.VSP [186]). A nice basis for P2 is

B ={1, x, x2

}

Page 394: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section VR Vector Representations 362

so thatu = 15 + 10x− 6x2 = 15(1) + 10(x) + (−6)(x2)

so by Definition VR [358]

ρB (u) =

1510−6

Another nice basis for P2 is

B ={1, 1 + x, 1 + x + x2

}so that now it takes a bit of computation to determine the scalars for the representation.We want a1, a2, a3 so that

15 + 10x− 6x2 = a1(1) + a2(1 + x) + a3(1 + x + x2)

Performing the operations in P2 on the right-hand side, and equating coefficients, givesthe three equations in the three unknown scalars,

15 = a1 + a2 + a3

10 = a2 + a3

−6 = a3

The coefficient matrix of this sytem is nonsingular, leading to a unique solution (nosurprise there, see Theorem VRRB [227]),

a1 = 5 a2 = 16 a3 = −6

so by Definition VR [358]

ρC (u) =

516−6

While we often form vector representations relative to “nice” bases, nothing prevents usfrom forming representations relative to “nasty” bases. For example, the set

D ={−2− x + 3x2, 1− 2x2, 5 + 4x + x2

}can be verified as a basis of P2 by checking linear independence with Definition LI [212]and then arguing that 3 vectors from P2, a vector space of dimension 3 (Theorem DP [235]),must also be a spanning set (Theorem G [244]). Now we desire scalars a1, a2, a3 so that

15 + 10x− 6x2 = a1(−2− x + 3x2) + a2(1− 2x2) + a3(5 + 4x + x2)

Performing the operations in P2 on the right-hand side, and equating coefficients, givesthe three equations in the three unknown scalars,

15 = −2a1 + a2 + 5a3

10 = −a1 + 4a3

−6 = 3a1 − 2a2 + a3

Page 395: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section VR Vector Representations 363

The coefficient matrix of this sytem is nonsingular, leading to a unique solution (nosurprise there, see Theorem VRRB [227]),

a1 = −2 a2 = 1 a3 = 2

so by Definition VR [358]

ρD (u) =

−212

4

Theorem VRIVector Representation is InjectiveThe function ρB (Definition VR [358]) is an injective linear transformation. �

Proof We will appeal to Theorem NSILT [328]. Suppose U is a vector space of dimensionn, so vector representation is of the form ρB : U 7→ Cn. Let B = {u1, u2, u3, . . . , un}be the basis of U used in the definition of ρB. Suppose u ∈ N (ρB). Finally, since B is abasis for U , by Theorem VRRB [227] there are (unique) scalars, a1, a2, a3, . . . , an suchthat

u = a1u1 + a2u2 + a3u3 + · · ·+ anun

Then

0 = ρB (u) u ∈ N (ρB)000...0

=

a1

a2

a3...

an

From this vector equality (Definition CVE [66]) we find that ai = 0, 1 ≤ i ≤ n. So

u = a1u1 + a2u2 + a3u3 + · · ·+ anun

= 0u1 + 0u2 + 0u3 + · · ·+ 0un

= 0 + 0 + 0 + · · ·+ 0 Theorem ZSSM [191]

= 0 Property of 0

So N (ρB) = {0} and by Theorem NSILT [328], ρB is injective. �

Theorem VRSVector Representation is SurjectiveThe function ρB (Definition VR [358]) is a surjective linear transformation. �

Proof We will appeal to Theorem RSLT [339]. Suppose U is a vector space of dimensionn, so vector representation is of the form ρB : U 7→ Cn. Let B = {u1, u2, u3, . . . , un}be the basis of U used in the definition of ρB. Suppose v ∈ Cn. Write

v =

v1

v2

v3...

vn

Page 396: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VR.CVS Characterization of Vector Spaces 364

and defineu = v1u1 + v2u2 + v3u3 + · · ·+ vnun

Then

ρB (u) = ρB (v1u1 + v2u2 + v3u3 + · · ·+ vnun)

=

v1

v2

v3...

vn

Definition VR [358]

= v

this demonstrates that v ∈ R (T ), so Cn ⊆ R (T ). Since R (T ) ⊆ Cn by Defini-tion RLT [336], we have R (T ) = Cn and Theorem RSLT [339] says T is surjective. �

We will have many occasions later to employ the inverse of vector representation, so wewill record the fact that vector representation is an invertible linear transformation.

Theorem VRILTVector Representation is an Invertible Linear TransformationThe function ρB (Definition VR [358]) is an invertible linear transformation. �

Proof The function ρB (Definition VR [358]) is a linear transformation (Theorem VRLT [358])that is injective (Theorem VRI [363]) and surjective (Theorem VRS [363]) with domainV and codomain Cn. By Theorem ILTIS [348] we then know that ρB is an invertiblelinear transformation. �

Informally, we will refer to the application of ρB as coordinatizing a vector, while theapplication of ρ−1

B will be referred to as un-coordinatizing a vector.

Subsection CVSCharacterization of Vector Spaces

Limiting our attention to vector spaces with finite dimension, we now describe everypossible vector space. All of them. Really.

Theorem CFDVSCharacterization of Finite Dimensional Vector SpacesSuppose that V is a vector space with dimension n. Then V is isomorphic to Cn. �

Proof Since V has dimension n we can find a basis of V of size n (Definition D [230])which we will call B. The linear transformation ρB is an invertible linear transformationfrom V to Cn, so by Definition IVS [350], we have that V and Cn are isomorphic. �

Theorem CFDVS [364] is the first of several surprises in this chapter, though it mightbe a bit demoralizing too. It says that there really are not all that many different (finitedimensional) vector spaces, and none are really any more complicated than Cn. Hmmm.The following examples should make this point.

Page 397: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VR.CVS Characterization of Vector Spaces 365

Example VR.TIVSTwo isomorphic vector spacesThe vector space of polynomials with degree 8 or less, P8, has dimension 9 (Theo-rem DP [235]). By Theorem CFDVS [364], P8 is isomorphic to C9. 4

Example VR.CVSRCrazy vector space revealedThe crazy vector space, C of Example VS.CVS [188], has dimension 2 (can you provethis?). By Theorem CFDVS [364], C is isomorphic to C2. 4

Example VR.ASCA subspace charaterizedIn Example D.DSP4 [236] we determined that a subspace W of P4 has dimension 4. ByTheorem CFDVS [364], W is isomorphic to C4. 4

Theorem IFDVSIsomorphism of Finite Dimensional Vector SpacesSuppose U and V are both finite-dimensional vector spaces. Then U and V are isomorphicif and only if dim (U) = dim (V ). �

Proof (⇒) This is just the statement proved in Theorem IVSED [351].

(⇐) This is the advertised converse of Theorem IVSED [351]. We will assume Uand V have equal dimension and discover that they are isomorphic vector spaces. Letn be the common dimension of U and V . Then by Theorem CFDVS [364] there areisomorphisms T : U 7→ Cn and S : V 7→ Cn.

T is therefore an invertible linear transformation by Definition IVS [350]. Also, S isan invertible linear transformation, so S−1 is an invertible linear transformation (The-orem IILT [347]). Invertible linear transformations are injective and surjective (The-orem ILTIS [348]), so the composition of two invertible linear transformations will beinjective and surjective (Theorem CILTI [331], Theorem CSLTS [343]), and hence byTheorem ILTIS [348] the composition of S−1 with T is invertible.

So (S−1 ◦ T ) : U 7→ V is an invertible linear transformation from U to V and Defini-tion IVS [350] says U and V are isomorphic. �

Example VR.MIVSMultiple isomorphic vector spacesC10, P9, M2,5 and M5,2 are all vector spaces and each has dimension 10. By Theo-rem IFDVS [365] each is isomorphic to any other.

The subspace of M4,4 that contains all the symmetric matrices (Definition SYM [125])has dimension 10, so this subspace is also isomorphic to each of the four vector spacesabove. 4

Theorem CLICoordinatization and Linear IndependenceSuppose that U is a vector space with a basis B of size n. Then S = {u1, u2, u3, . . . , uk}is a linearly independent subset of U if and only if R = {ρB (u1) , ρB (u2) , ρB (u3) , . . . , ρB (uk)}is a linearly independent subset of Cn. �

Page 398: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VR.CVS Characterization of Vector Spaces 366

Proof The linear transformation ρB is an isomorphism between U and Cn (Theorem VRILT [364]).As an invertible linear transformation, ρB is an injective linear transformation (Theo-rem ILTIS [348]), and ρ−1

B is also an injective linear transformation (Theorem IILT [347],Theorem ILTIS [348]).

(⇒) Since ρB is an injective linear transformation and S is linearly independent,Theorem ILTLI [329] says that R is linearly independent.

(⇐) If we apply ρ−1B to each element of R, we will create the set S. Since we are

assuming R is linearly independent and ρ−1B is injective, Theorem ILTLI [329] says that

S is linearly independent. �

Theorem CSSCoordinatization and Spanning SetsSuppose that U is a vector space with a basis B of size n. Then u ∈ Sp ({u1, u2, u3, . . . , uk})if and only if ρB (u) ∈ Sp ({ρB (u1) , ρB (u2) , ρB (u3) , . . . , ρB (uk)}). �

Proof (⇒) Suppose u ∈ Sp ({u1, u2, u3, . . . , uk}). Then there are scalars, a1, a2, a3, . . . , ak,such that

u = a1u1 + a2u2 + a3u3 + · · ·+ akuk

Then,

ρB (u) = ρB (a1u1 + a2u2 + a3u3 + · · ·+ akuk)

= a1ρB (u1) + a2ρB (u2) + a3ρB (u3) + · · ·+ akρB (uk) Theorem LTLC [311]

which says that ρB (u) ∈ Sp ({ρB (u1) , ρB (u2) , ρB (u3) , . . . , ρB (uk)}).(⇐) Suppose that ρB (u) ∈ Sp ({ρB (u1) , ρB (u2) , ρB (u3) , . . . , ρB (uk)}). Then

there are scalars b1, b2, b3, . . . , bk such that

ρB (u) = b1ρB (u1) + b2ρB (u2) + b3ρB (u3) + · · ·+ bkρB (uk)

Recall that ρB is invertible (Theorem VRILT [364]), so

u = IU (u) Definition IDLT [344]

=(ρ−1

B ◦ ρB

)(u) Definition IVLT [344]

= ρ−1B (ρB (u)) Definition LTC [319]

= ρ−1B (b1ρB (u1) + b2ρB (u2) + b3ρB (u3) + · · ·+ bkρB (uk)) Substitution

= b1ρ−1B (ρB (u1)) + b2ρ

−1B (ρB (u2)) + b3ρ

−1B (ρB (u3))

+ · · ·+ bkρ−1B (ρB (uk)) Theorem LTLC [311]

= b1IU (u1) + b2IU (u2) + b3IU (u3) + · · ·+ bkIU (uk) Definition IVLT [344]

= b1u1 + b2u2 + b3u3 + · · ·+ bkuk Definition IDLT [344]

which says that u ∈ Sp ({u1, u2, u3, . . . , uk}). �

Here’s a fairly simple example that illustrates a very, very important idea.

Example VR.CP2Coordinatizing in P2

In Example VR.VRP2 [361] we needed to know that

D ={−2− x + 3x2, 1− 2x2, 5 + 4x + x2

}

Page 399: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VR.CVS Characterization of Vector Spaces 367

is a basis for P2. With Theorem CLI [365] and Theorem CSS [366] this task is much easier.First, choose a known basis for P2, a basis that is easy to form vector representationswith. We will choose

B ={1, x, x2

}Now, form the subset of C3 that is the result of applying ρB to each element of D,

F ={ρB

(−2− x + 3x2

), ρB

(1− 2x2

), ρB

(5 + 4x + x2

)}=

−2−13

,

10−2

,

541

and ask if F is a linearly independent spanning set for C3. This is easily seen to be thecase by forming a matrix A whose columns are the vectors of F , row-reducing A to theidentity matrix I3, and then using the nonsingularity of A to assert that F is a basis forC3 (Theorem CNSMB [225]). Now, since F is a basis for C3, Theorem CLI [365] andTheorem CSS [366] tell us that D is also a basis for P2. 4

Example VR.CP2 [366] illustrates the broad notion that computations in abstract vectorspaces can be reduced to computations in Cm. You may have noticed this phenomenon asyou worked through examples in Chapter VS [184] or Chapter LT [302]employing vectorspaces of matrices or polynomials. These computations seemed to invariable result in sys-tems of equations or the like from Chapter SLE [2], Chapter V [65] and Chapter M [121].It is vector representation, ρB, that allows us to make this connection formal and precise.

Knowing that vector representation allows us to translate questions about linear com-binations, linear indepencence and spans from general vector spaces to Cm allows us toprove a great many theorems about how to translate other properties. Rather thanprove these theorems, each of the same style as the other, we will offer some generalguidance about how to best employ Theorem VRLT [358], Theorem CLI [365] and Theo-rem CSS [366]. This comes in the form of a “principle”: a basic truth, but most definitelynot a theorem (hence, no proof).

The Coordinatization Principle Suppose that U is a vector space with a basis B ofsize n. Then any question about U , or its elements, which depends on the vector additionor scalar multiplication in U , or depends on linear independence or spanning, may betranslated into the same question in Cn by application of the linear transformation ρB tothe relevant vectors. Once the question is answered in Cn, the answer may be translatedback to U (if necessary) through application of the inverse linear transformation ρ−1

B .

Example VR.CM32Coordinatization in M32

This is a simple example of the Coordinatization Principle [367], depending only on thefact that coordinatizing is an invertible linear transformation (Theorem VRILT [364]).Suppose we have a linear combination to perform in M32, the vector space of 3 × 2matrices, but we are adverse to doing the operations of M32 (Definition MA [122], Defi-nition SMM [122]). More specifically, suppose we are faced with the computation

6

3 7−2 40 −3

+ 2

−1 34 8−2 5

Page 400: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection VR.CVS Characterization of Vector Spaces 368

We choose a nice basis for M32 (or a nasty basis if we are so inclined),

B =

1 0

0 00 0

,

0 01 00 0

,

0 00 01 0

,

0 10 00 0

,

0 00 10 0

,

0 00 00 1

and apply ρB to each vector in the linear combination. This gives us a new computation,now in the vector space C6,

6

3−2074−3

+ 2

−14−2385

which we can compute with the operations of C6 (Definition CVA [67], Definition CVSM [68]),to arrive at

16−4−44840−8

We are after the result of a computation in M32, so we now can apply ρ−1

B to obtain a3× 2 matrix,

16

1 00 00 0

+(−4)

0 01 00 0

+(−4)

0 00 01 0

+48

0 10 00 0

+40

0 00 10 0

+(−8)

0 00 00 1

=

16 48−4 40−4 −8

which is exactly the matrix we would have computed had we just performed the matrixoperations in the first place. 4

Page 401: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Section MR Matrix Representations 369

Section MR

Matrix Representations

Definition MRMatrix RepresentationSuppose that T : U 7→ V is a linear transformation, B = {u1, u2, u3, . . . , un} is a basisfor U of size n, and C is a basis for V of size m. The the matrix representation of Trelative to B and C is the m× n matrix,

RTB,C = [ρC (T (u1))| ρC (T (u2))| ρC (T (u3))| . . . |ρC (T (un)) ] �

We may choose to use whatever terms we want when we make a definition. Someare arbitrary, others make sense, but only in light of subsequent theorems. Matrix rep-resentation is in the latter category. We begin with a linear transformation and producea matrix. So what? Here’s the theorem that justifies the term ”matrix representation.”

Theorem FTMRFundamental Theorem of Matrix RepresentationSuppose that T : U 7→ V is a linear transformation, B is a basis for U , C is a basis forV and RT

B,C is the matrix representation of T relative to B and C. Then, for any u ∈ U ,

ρC (T (u)) = RTB,C (ρB (u)) �

Proof Let B = {u1, u2, u3, . . . , un} be the basis of U . Since u ∈ U , there are scalarsa1, a2, a3, . . . , an such that

u = a1u1 + a2u2 + a3u3 + · · ·+ anun

Then,

RTB,C (ρB (u))

= [ρC (T (u1))| ρC (T (u2))| ρC (T (u3))| . . . |ρC (T (un)) ]u Definition MR [369]

= [ρC (T (u1))| ρC (T (u2))| ρC (T (u3))| . . . |ρC (T (un)) ]

a1

a2

a3...

an

Definition VR [358]

= a1ρC (T (u1)) + a2ρC (T (u2)) + a3ρC (T (u3))

+ · · ·+ anρC (T (un)) Definition MVP [151]

= ρC (a1T (u1) + a2T (u2) + a3T (u3) + · · ·+ anT (un)) Theorem LTLC [311]

= ρC (T (a1u1 + a2u2 + a3u3 + · · ·+ anun)) Theorem LTLC [311]

= ρC (T (u)) Substitution �

Page 402: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MR.NRFO New Representations from Old 370

This theorem says that we can apply T to u and coordinatize the result relative toC in V , or we can first coordinatize u relative to B in U , then multiply by the matrixrepresentation. Either way, the result is the same. So the effect of a linear transformationcan always be accomplished by a matrix-vector product (Definition MVP [151]). That’simportant enough to say again. The effect of a linear transformation is a matrix-vectorproduct. A minor rearrangement of this result might be even more striking. Apply ρ−1

C

to both sides of the conclusion of Theorem FTMR [369] and get

T (u) = ρ−1C

(RT

B,C (ρB (u)))

To effect a linear transformation (T ) of a vector (u), coordinatize (ρB), do a matrix-vector product (RT

B,C), and un-coordinatize (ρ−1C ). So, absent some bookkeeping about

vector representations, a linear transformation is a matrix. Another way to view this iswith a diagram similar to those we saw just after defining a linear transformation,

We will use Theorem FTMR [369] frequently in the next few sections. A typicalapplication will feel like the linear transformation T “commutes” with a vector repre-sentation, ρC , and as it does the transformation morphs into a matrix, RT

B,C , while thevector representation changes to a new basis, ρB. Or vice-versa.

Subsection NRFONew Representations from Old

In Subsection LT.NLTFO [316] we built new linear transformations from other lineartransformations. Sums, scalar multiples and compositions. These new linear transfor-mations will have matrix represntations as well. How do the new matrix representationsrelate to the old matrix representations? Here are the three theorems.

Theorem MRSLTMatrix Representation of a Sum of Linear TransformationsSuppose that T : U 7→ V and S : U 7→ V are linear transformations, B is a basis of Uand C is a basis of V . Then

RT+SB,C = RT

B,C + RSB,C �

Proof Let x be any vector in Cn. Define u ∈ U by u = ρ−1B (x), so x = ρB (u). Then,

RT+SB,C x = RT+S

B,C ρB (u) Substitution

= ρC ((T + S) (u)) Theorem FTMR [369]

= ρC (T (u) + S (u)) Definition LTA [316]

= ρC (T (u)) + ρC (S (u)) Definition LT [302]

= RTB,C (ρB (u)) + RS

B,C (ρB (u)) Theorem FTMR [369]

=(RT

B,C + RSB,C

)ρB (u) Theorem MMDAA [157]

=(RT

B,C + RSB,C

)x Substitution

Since the matrices RT+SB,C and RT

B,C + RSB,C have equal matrix-vector products for every

vector in Cn, they have equal products on some basis of Cn, and by Theorem XX [??],they are equal matrices. �

Page 403: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MR.NRFO New Representations from Old 371

Theorem MRMLTMatrix Representation of a Multiple of a Linear TransformationSuppose that T : U 7→ V is a linear transformation, α ∈ C, B is a basis of U and C is abasis of V . Then

RαTB,C = αRT

B,C �

Proof Let x be any vector in Cn. Define u ∈ U by u = ρ−1B (x), so x = ρB (u). Then,

RαTB,Cx = RαT

B,CρB (u) Substitution

= ρC ((αT ) (u)) Theorem FTMR [369]

= ρC (αT (u)) Definition LTSM [317]

= αρC (T (u)) Definition LT [302]

= α(RT

B,CρB (u))

Theorem FTMR [369]

=(αRT

B,C

)ρB (u) Theorem MMSMM [157]

=(αRT

B,C

)x Substitution

Since the matrices RαTB,C and αRT

B,C have equal matrix-vector products for every vectorin Cn, they have equal products on some basis of Cn, and by Theorem XX [??], they areequal matrices. �

The vector space of all linear transformations from U to V is now isomorphic to thevector space of all m× n matrices.

Theorem MRCLTMatrix Representation of a Composition of Linear TransformationsSuppose that T : U 7→ V and S : V 7→ W are linear transformations, B is a basis of U ,C is a basis of V , and D is a basis of W . Then

RS◦TB,D = RS

C,DRTB,C �

Proof Let x be any vector in Cn. Define u ∈ U by u = ρ−1B (x), so x = ρB (u). Then,

RS◦TB,Dx = RS◦T

B,DρB (u) Substitution

= ρD ((S ◦ T ) (u)) Theorem FTMR [369]

= ρD (S (T (u))) Definition LTC [319]

= RSC,DρC (T (u)) Theorem FTMR [369]

= RSC,D

(RT

B,CρB (u))

Theorem FTMR [369]

=(RS

C,DRTB,C

)ρB (u) Theorem MMA [158]

=(RS

C,DRTB,C

)x Substitution

Since the matrices RS◦TB,D and RS

C,DRTB,C have equal matrix-vector products for every vector

in Cn, they have equal products on some basis of Cn, and by Theorem XX [??], they areequal matrices. �

Page 404: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MR.PMR Properties of Matrix Representations 372

This is the second great surprise of introductory linear algebra. Matrices are linear trans-formations (functions, really), and matrix multiplication is function composition! We canform the composition of two linear transformations, then form the matrix representationof the result. Or we can form the matrix representation of each linear transformationseparately, then multiply the two representations together via Definition MM [153]. Ineither case, we arrive at the same result.

One of our goals in the first part of this book is to make the definition of matrixmultiplication (Definition MVP [151], Definition MM [153]) seem as natural as possi-ble. Many are brought up with an entry-by-entry description of matrix multiplication(Theorem ME [286]) as the definition of matrix multiplication, and then theorems aboutcolumns of matrices and linear combinations follow from that definition. With this un-motivated definition, the realization that matrix multiplication is function compositionis quite remarkable. It is an interesting exercise to begin with the question, “What is thematrix representation of the composition of two linear transformations?” and then, with-out using any theorems about matrix multiplication, finally arrive at the entry-by-entrydescription of matrix multiplication. Try it yourself.

Subsection PMRProperties of Matrix Representations

It will not be a surprise to discover that the null space and range of a linear transfor-mation are closely related to the null space and range of the transformation’s matrixrepresentation. Perhaps this idea has been bouncing around in your head already, evenbefore seeing the definition of a matrix representation. However, with a formal definitionof a matrix representation (Definition MR [369]), and a fundamental theorem to go withit (Theorem FTMR [369]) we can be formal about the relationship, using the idea ofisomorphic vector spaces (Definition IVS [350]). Here are the twin theorems.

Theorem INSIsomorphic Null SpacesSuppose that T : U 7→ V is a linear transformation, B is a basis for U of size n, and Cis a basis for V . Then the null space of T is isomorphic to the null space of RT

B,C ,

N (T ) ∼= N(RT

B,C

)�

Proof To establish that two vector spaces are isomorphic, we must find an isomorphismbetween them, an invertible linear transformation (Definition IVS [350]). The null spaceof the linear transformation T , N (T ), is a subspace of U , while the null space of thematrix representation, N

(RT

B,C

)is a subspace of Cn. The function ρB is defined as a

function from U to Cn, but we can just as well employ the definition of ρB as a functionfrom N (T ) to N

(RT

B,C

).

The restriction in the size of the domain and codomain will not affect the fact thatρB is a linear transformation (Theorem VRLT [358]), nor will it affect the fact that ρB

is injective (Theorem VRI [363]). Something must done though to verify that ρB issurjective. To this end, appeal to the definition of surjective (Definition SLT [332]), andsuppose that we have an element of the codomain, x ∈ N

(RT

B,C

)⊆ Cn and we wish to

Page 405: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MR.PMR Properties of Matrix Representations 373

find an element of the domain with x as its image. We now show that the desired elementof the domain is u = ρ−1

B (x). First, verify that u ∈ N (T ),

T (u) = T(ρ−1

B (x))

Substitution

= IV

(T(ρ−1

B (x)))

Definition IDLT [344]

= ρ−1C

(ρC

(T(ρ−1

B (x))))

Definition IVLT [344]

= ρ−1C

(RT

B,C

(ρB

(ρ−1

B (x))))

Theorem FTMR [369]

= ρ−1C

(RT

B,C (ICn (x)))

Definition IVLT [344]

= ρ−1C

(RT

B,Cx)

Definition IDLT [344]

= ρ−1C (0Cn) x ∈ N

(RT

B,C

)= 0V Theorem LTTZZ [306]

Second, verify that the proposed isomorphism, ρB, takes u to x,

ρB (u) = ρB

(ρ−1

B (x))

Substitution

= ICn (x) Definition IVLT [344]

= x Definition IDLT [344]

With ρB demonstrated to be an injective and surjective linear transformation from N (T )to N

(RT

B,C

), Theorem ILTIS [348] tells us ρB is invertible, and so by Definition IVS [350],

we say N (T ) and N(RT

B,C

)are isomorphic. �

Theorem IRIsomorphic RangesSuppose that T : U 7→ V is a linear transformation, B is a basis for U of size n, and Cis a basis for V of size m. Then the range of T is isomorphic to the range of RT

B,C ,

R (T ) ∼= R(RT

B,C

)�

Proof To establish that two vector spaces are isomorphic, we must find an isomorphismbetween them, an invertible linear transformation (Definition IVS [350]). The range ofthe linear transformation T , R (T ), is a subspace of V , while the range of the matrixrepresentation, N

(RT

B,C

)is a subspace of Cm. The function ρC is defined as a function

from V to Cm, but we can just as well employ the definition of ρC as a function fromR (T ) to R

(RT

B,C

).

The restriction in the size of the domain and codomain will not affect the fact that ρC

is a linear transformation (Theorem VRLT [358]), nor will it affect the fact that ρC is in-jective (Theorem VRI [363]). Something must done though to verify that ρC is surjective.this all gets a bit confusing, since the domain of our isomorphism is the range of the lineartransformation, so think about your objects as you go. To establish that ρC is surjective,appeal to the definition of a surjective linear transformation (Definition SLT [332]), andsuppose that we have an element of the codomain, y ∈ R

(RT

B,C

)⊆ Cm and we wish to

find an element of the domain with y as its image. Since y ∈ R(RT

B,C

), there exists a

vector, x ∈ Cn with RTB,Cx = y. We now show that the desired element of the domain is

Page 406: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection MR.PMR Properties of Matrix Representations 374

v = ρ−1C (y). First, verify that v ∈ R (T ) by applying T to u = ρ−1

B (x),

T (u) = T(ρ−1

B (x))

Substitution

= IV

(T(ρ−1

B (x)))

Definition IDLT [344]

= ρ−1C

(ρC

(T(ρ−1

B (x))))

Definition IVLT [344]

= ρ−1C

(RT

B,C

(ρB

(ρ−1

B (x))))

Theorem FTMR [369]

= ρ−1C

(RT

B,C (ICn (x)))

Definition IVLT [344]

= ρ−1C

(RT

B,Cx)

Definition IDLT [344]

= ρ−1C (y) y ∈ R

(RT

B,C

)= v Substitution

Second, verify that the proposed isomorphism, ρC , takes v to y,

ρC (v) = ρC

(ρ−1

C (y))

Substitution

= ICm (y) Definition IVLT [344]

= y Definition IDLT [344]

With ρC demonstrated to be an injective and surjective linear transformation from R (T )to R

(RT

B,C

), Theorem ILTIS [348] tells us ρC is invertible, and so by Definition IVS [350],

we say R (T ) and R(RT

B,C

)are isomorphic. �

These two theorems can be viewed as further formal evidence for the CoordinatizationPrinciple [367], though they are not direct consequences.

THIS SECTION NOT COMPLETE

Page 407: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

A: Archetypes

The American Heritage Dictionary of the English Language (Third Edition) gives twodefinitions of the word “archetype”: 1. An original model or type after which other similarthings are patterned; a prototype; and 2. An ideal example of a type; quintessence.

Either use might apply here. Our archetypes are typical examples of systems ofequations, matrices and linear transformations. They have been designed to demonstratethe range of possibilities, allowing you to compare and contrast them. Several are of asize and complexity that is usually not presented in a textbook, but should do a betterjob of being “typical.”

We have made frequent reference to many of these throughout the text, such as thefrequent comparisons between Archetype A [379] and Archetype B [384]. Some we haveleft for you to investigate, such as Archetype J [418], which parallels Archetype I [414].

How should you use the archetypes? First, consult the description of each one as it ismentioned in the text. See how other facts about the example might illuminate whateverproperty or construction is being described in the example. Second, Each property hasa short description that usually includes references to the relevant theorems. Performthe computations and understand the connections to the listed theorems. Third, eachproperty has a small checkbox in front of it. Use the archetypes like a workbook andchart your progress by “checking-off” those properties that you understand.

The next page has a chart that summarizes some (but not all) of the propertiesdescribed for each archetype. Notice that while there are several types of objects, thereare fundamental connections between them. That some lines of the table do double-dutyis meant to convey some of these connections. Consult this table when you wish toquickly find an example of a certain phenomenon.

375

Page 408: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Chapter A Archetypes 376

Page 409: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Chapter A Archetypes 377

AB

CD

EF

GH

IJ

KL

MN

OP

QR

Type

SS

SS

SS

SS

SS

MM

LL

LL

LL

Var

s,C

ols,

Dom

ain

33

44

44

22

79

55

55

33

55

Eqns,

Row

s,C

oDom

33

33

34

55

46

55

33

55

55

Con

sist

ent

IU

II

NU

UN

II

Ran

k2

33

22

42

23

45

32

32

34

5N

ullity

10

12

20

33

45

02

32

10

10

Inje

ctiv

eX

XN

YN

YSurj

ecti

veN

YX

XN

YFull

Ran

kN

YY

NN

YY

YN

NY

NN

onsi

ngu

lar

NY

YY

NIn

vert

ible

NY

YY

NN

YD

eter

min

ant

0-2

-18

160

Dia

gonal

izab

leN

YY

YY

Arc

het

ype

Fac

tsS=

Syst

emof

Equat

ions,

M=

Mat

rix,L=

Lin

ear

Tra

nsf

orm

atio

nU

=U

niq

ue

solu

tion

,I=

Infinitel

ym

any

solu

tion

s,N

=N

oso

luti

ons

Y=

Yes

,N

=N

o,X

=Im

pos

sible

,bla

nk=

Not

Applica

ble

Page 410: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Chapter A Archetypes 378

Page 411: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype A 379

Archetype A

Summary Linear system of three equations, three unknowns. Singular coefficent ma-trix with dimension 1 null space. Integer eigenvalues and a degenerate eigenspace forcoefficient matrix.

A system of linear equations (Definition SSLE [10]):

x1 − x2 + 2x3 = 1

2x1 + x2 + x3 = 8

x1 + x2 = 5

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = 2, x2 = 3, x3 = 1

x1 = 3, x2 = 2, x3 = 0

Augmented matrix of the linear system of equations (Definition AM [23]):1 −1 2 12 1 1 81 1 0 5

Matrix in reduced row-echelon form, row-equivalent to augmented matrix: 1 0 1 3

0 1 −1 20 0 0 0

Analysis of the augmented matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = {3, 4}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.x1

x2

x3

=

320

+ x3

−111

Given a system of equations we can always build a new, related, homogeneous system

Page 412: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype A 380

(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

x1 − x2 + 2x3 = 0

2x1 + x2 + x3 = 0

x1 + x2 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0

x1 = −1, x2 = 1, x3 = 1

x1 = −5, x2 = 5, x3 = 5

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros: 1 0 1 0

0 1 −1 00 0 0 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 2 D = {1, 2} F = {3, 4}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.1 −1 2

2 1 11 1 0

Matrix brought to reduced row-echelon form: 1 0 1

0 1 −10 0 0

Page 413: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype A 381

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = {3}

Matrix (coefficient matrix) is nonsingular or singular? (Theorem NSRRI [57]) at thesame time, examine the size of the set F above.Notice that this property does not applyto matrices that are not square.

Singular.

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp

−1

11

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

1

21

,

−111

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[1 −2 3

]Sp

−3

01

,

210

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as column

Page 414: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype A 382

vectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

1

0−1

3

,

0123

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

1

01

,

01−1

Inverse matrix, if it exists. The inverse is not defined for matrices that are not square,and if the matrix is square, then the matrix must be nonsingular. (Definition MI [167],Theorem NSI [177])

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 3 Rank: 2 Nullity: 1

Determinant of the matrix, which is only defined for square matrices. The matrix isnonsingular if and only if the determinant is nonzero (Theorem SMZD [259]). (Productof all eigenvalues?)

Determinant = 0

Eigenvalues, and bases for eigenspaces. (Definition EEM [261],Definition EM [270])

λ = 0 EA (0) = Sp

−1

11

λ = 2 EA (2) = Sp

1

53

Geometric and algebraic multiplicities. (Definition GME [272]Definition AME [272])

Page 415: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype A 383

γA (0) = 1 αA (0) = 2

γA (2) = 1 αA (2) = 1

Diagonalizable? (Definition DZM [295])

No, γA (0) 6= αB (0), Theorem DMLE [298].

Page 416: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype B 384

Archetype B

Summary System with three equations, three unknowns. Nonsingular coefficent ma-trix. Distinct integer eigenvalues for coefficient matrix.

A system of linear equations (Definition SSLE [10]):

−7x1 − 6x2 − 12x3 = −33

5x1 + 5x2 + 7x3 = 24

x1 + 4x3 = 5

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = −3, x2 = 5, x3 = 2

Augmented matrix of the linear system of equations (Definition AM [23]):−7 −6 −12 −335 5 7 241 0 4 5

Matrix in reduced row-echelon form, row-equivalent to augmented matrix: 1 0 0 −3

0 1 0 5

0 0 1 2

Analysis of the augmented matrix (Notation RREFA [37]):

r = 3 D = {1, 2, 3} F = {4}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.x1

x2

x3

=

−352

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships with

Page 417: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype B 385

various properties of the original system.

−11x1 + 2x2 − 14x3 = 0

23x1 − 6x2 + 33x3 = 0

14x1 − 2x2 + 17x3 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros: 1 0 0 0

0 1 0 0

0 0 1 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 3 D = {1, 2, 3} F = {4}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.−7 −6 −12

5 5 71 0 4

Matrix brought to reduced row-echelon form: 1 0 0

0 1 0

0 0 1

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 3 D = {1, 2, 3} F = { }

Page 418: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype B 386

Matrix (coefficient matrix) is nonsingular or singular? (Theorem NSRRI [57]) at thesame time, examine the size of the set F above.Notice that this property does not applyto matrices that are not square.

Nonsingular.

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp ({ })

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

−7

51

,

−650

,

−1274

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[]

Sp

1

00

,

010

,

001

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Page 419: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype B 387

Sp

1

00

,

010

,

001

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

1

00

,

010

,

001

Inverse matrix, if it exists. The inverse is not defined for matrices that are not square,and if the matrix is square, then the matrix must be nonsingular. (Definition MI [167],Theorem NSI [177])−10 −12 −9

132

8 112

52

3 52

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 3 Rank: 3 Nullity: 0

Determinant of the matrix, which is only defined for square matrices. The matrix isnonsingular if and only if the determinant is nonzero (Theorem SMZD [259]). (Productof all eigenvalues?)

Determinant = −2

Eigenvalues, and bases for eigenspaces. (Definition EEM [261],Definition EM [270])

λ = −1 EB (−1) = Sp

−5

31

λ = 1 EB (1) = Sp

−3

21

λ = 2 EB (2) = Sp

−2

11

Geometric and algebraic multiplicities. (Definition GME [272]Definition AME [272])

Page 420: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype B 388

γB (−1) = 1 αB (−1) = 1

γB (1) = 1 αB (1) = 1

γB (2) = 1 αB (2) = 1

Diagonalizable? (Definition DZM [295])

Yes, distinct eigenvalues, Theorem DED [300].

The diagonalization. (Theorem DC [295])−1 −1 −12 3 1−1 −2 1

−7 −6 −125 5 71 0 4

−5 −3 −23 2 11 1 1

=

−1 0 00 1 00 0 2

Page 421: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype C 389

Archetype C

Summary System with three equations, four variables. Consistent. Null space ofcoefficient matrix has dimension 1.

A system of linear equations (Definition SSLE [10]):

2x1 − 3x2 + x3 − 6x4 = −7

4x1 + x2 + 2x3 + 9x4 = −7

3x1 + x2 + x3 + 8x4 = −8

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = −7, x2 = −2, x3 = 7, x4 = 1

x1 = −1, x2 = −7, x3 = 4, x4 = −2

Augmented matrix of the linear system of equations (Definition AM [23]):2 −3 1 −6 −74 1 2 9 −73 1 1 8 −8

Matrix in reduced row-echelon form, row-equivalent to augmented matrix: 1 0 0 2 −5

0 1 0 3 1

0 0 1 −1 6

Analysis of the augmented matrix (Notation RREFA [37]):

r = 3 D = {1, 2, 3} F = {4, 5}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.

x1

x2

x3

x4

=

−5160

+ x4

−2−311

Given a system of equations we can always build a new, related, homogeneous system

Page 422: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype C 390

(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

2x1 − 3x2 + x3 − 6x4 = 0

4x1 + x2 + 2x3 + 9x4 = 0

3x1 + x2 + x3 + 8x4 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0, x4 = 0

x1 = −2, x2 = −3, x3 = 1, x4 = 1

x1 = −4, x2 = −6, x3 = 2, x4 = 2

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros: 1 0 0 2 0

0 1 0 3 0

0 0 1 −1 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 3 D = {1, 2, 3} F = {4, 5}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.2 −3 1 −6

4 1 2 93 1 1 8

Matrix brought to reduced row-echelon form: 1 0 0 2

0 1 0 3

0 0 1 −1

Page 423: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype C 391

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 3 D = {1, 2, 3} F = {4}

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp

−2−311

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

2

43

,

−311

,

121

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[ ]

Sp

1

00

,

010

,

001

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Page 424: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype C 392

Sp

1

00

,

010

,

001

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

1002

,

0103

,

001−1

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 4 Rank: 3 Nullity: 1

Page 425: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype D 393

Archetype D

Summary System with three equations, four variables. Consistent. Null space ofcoefficient matrix has dimension 2. Coefficient matrix identical to that of Archetype E,vector of constants is different.

A system of linear equations (Definition SSLE [10]):

2x1 + x2 + 7x3 − 7x4 = 8

−3x1 + 4x2 − 5x3 − 6x4 = −12

x1 + x2 + 4x3 − 5x4 = 4

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = 0, x2 = 1, x3 = 2, x4 = 1

x1 = 4, x2 = 0, x3 = 0, x4 = 0

x1 = 7, x2 = 8, x3 = 1, x4 = 3

Augmented matrix of the linear system of equations (Definition AM [23]): 2 1 7 −7 8−3 4 −5 −6 −121 1 4 −5 4

Matrix in reduced row-echelon form, row-equivalent to augmented matrix: 1 0 3 −2 4

0 1 1 −3 00 0 0 0 0

Analysis of the augmented matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = {3, 4, 5}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.

Page 426: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype D 394

x1

x2

x3

x4

=

4000

+ x3

−3−110

+ x4

2301

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

2x1 + x2 + 7x3 − 7x4 = 0

−3x1 + 4x2 − 5x3 − 6x4 = 0

x1 + x2 + 4x3 − 5x4 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0, x4 = 0

x1 = −3, x2 = −1, x3 = 1, x4 = 0

x1 = 2, x2 = 3, x3 = 0, x4 = 1

x1 = −1, x2 = 2, x3 = 1, x4 = 1

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros: 1 0 3 −2 0

0 1 1 −3 00 0 0 0 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 2 D = {1, 2} F = {3, 4, 5}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations. 2 1 7 −7−3 4 −5 −61 1 4 −5

Page 427: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype D 395

Matrix brought to reduced row-echelon form: 1 0 3 −2

0 1 1 −30 0 0 0

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = {3, 4}

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp

−3−110

,

2301

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

2−31

,

141

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[1 1

7−11

7

]Sp

11

7

01

,

−17

10

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reduced

Page 428: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype D 396

row-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

1

0711

,

01111

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

103−2

,

011−3

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 4 Rank: 2 Nullity: 2

Page 429: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype E 397

Archetype E

Summary System with three equations, four variables. Inconsistent. Null space ofcoefficient matrix has dimension 2. Coefficient matrix identical to that of Archetype D,constant vector is different.

A system of linear equations (Definition SSLE [10]):

2x1 + x2 + 7x3 − 7x4 = 2

−3x1 + 4x2 − 5x3 − 6x4 = 3

x1 + x2 + 4x3 − 5x4 = 2

Some solutions to the system of linear equations (not necessarily exhaustive):

None. (Why?)

Augmented matrix of the linear system of equations (Definition AM [23]): 2 1 7 −7 2−3 4 −5 −6 31 1 4 −5 2

Matrix in reduced row-echelon form, row-equivalent to augmented matrix: 1 1 4 −5 0

0 1 1 −3 0

0 0 0 0 1

Analysis of the augmented matrix (Notation RREFA [37]):

r = 3 D = {1, 2, 5} F = {3, 4}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.

Inconsistent system, no solutions exist.

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships with

Page 430: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype E 398

various properties of the original system.

2x1 + x2 + 7x3 − 7x4 = 0

−3x1 + 4x2 − 5x3 − 6x4 = 0

x1 + x2 + 4x3 − 5x4 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0, x4 = 0

x1 = 4, x2 = 13, x3 = 2, x4 = 5

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros: 1 0 3 −2 0

0 1 1 −3 00 0 0 0 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 2 D = {1, 2} F = {3, 4, 5}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations. 2 1 7 −7−3 4 −5 −61 1 4 −5

Matrix brought to reduced row-echelon form: 1 0 3 −2

0 1 1 −30 0 0 0

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = {3, 4}

Page 431: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype E 399

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp

−3−110

,

2301

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

2−31

,

141

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[1 1

7−11

7

]Sp

11

7

01

,

−17

10

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

1

0711

,

01111

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.

Page 432: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype E 400

(Theorem BRS [144])

Sp

103−2

,

011−3

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 4 Rank: 2 Nullity: 2

Page 433: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype F 401

Archetype F

Summary System with four equations, four variables. Nonsingular coefficient matrix.Integer eigenvalues, one has “high” multiplicity.

A system of linear equations (Definition SSLE [10]):

33x1 − 16x2 + 10x3 − 2x4 = −27

99x1 − 47x2 + 27x3 − 7x4 = −77

78x1 − 36x2 + 17x3 − 6x4 = −52

−9x1 + 2x2 + 3x3 + 4x4 = 5

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = 1, x2 = 2, x3 = −2, x4 = 4

Augmented matrix of the linear system of equations (Definition AM [23]):33 −16 10 −2 −2799 −47 27 −7 −7778 −36 17 −6 −52−9 2 3 4 5

Matrix in reduced row-echelon form, row-equivalent to augmented matrix:1 0 0 0 1

0 1 0 0 2

0 0 1 0 −2

0 0 0 1 4

Analysis of the augmented matrix (Notation RREFA [37]):

r = 4 D = {1, 2, 3, 4} F = {5}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.

Page 434: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype F 402

x1

x2

x3

x4

=

12−24

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

33x1 − 16x2 + 10x3 − 2x4 = 0

99x1 − 47x2 + 27x3 − 7x4 = 0

78x1 − 36x2 + 17x3 − 6x4 = 0

−9x1 + 2x2 + 3x3 + 4x4 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0, x4 = 0

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros:

1 0 0 0 0

0 1 0 0 0

0 0 1 0 0

0 0 0 1 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 4 D = {1, 2, 3, 4} F = {5}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.

33 −16 10 −299 −47 27 −778 −36 17 −6−9 2 3 4

Page 435: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype F 403

Matrix brought to reduced row-echelon form:1 0 0 0

0 1 0 0

0 0 1 0

0 0 0 1

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 4 D = {1, 2, 3, 4} F = { }

Matrix (coefficient matrix) is nonsingular or singular? (Theorem NSRRI [57]) at thesame time, examine the size of the set F above.Notice that this property does not applyto matrices that are not square.

Nonsingular.

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp ({ })

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

339978−9

,

−16−47−362

,

1027173

,

−2−7−64

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[]

Page 436: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype F 404

Sp

1000

,

0100

,

0010

,

0001

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

1000

,

0100

,

0010

,

0001

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

1000

,

0100

,

0010

,

0001

Inverse matrix, if it exists. The inverse is not defined for matrices that are not square,and if the matrix is square, then the matrix must be nonsingular. (Definition MI [167],Theorem NSI [177])−(

863

)383

−(

113

)73

−(

1292

)863

−(

172

)316

−13 6 −2 1−(

452

)293

−(

52

)136

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 4 Rank: 4 Nullity: 0

Determinant of the matrix, which is only defined for square matrices. The matrix isnonsingular if and only if the determinant is nonzero (Theorem SMZD [259]). (Productof all eigenvalues?)

Determinant = −18

Eigenvalues, and bases for eigenspaces. (Definition EEM [261],Definition EM [270])

Page 437: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype F 405

λ = −1 EF (−1) = Sp

1201

λ = 2 EF (2) = Sp

2521

λ = 3 EF (3) = Sp

1107

,

1745210

Geometric and algebraic multiplicities. (Definition GME [272]Definition AME [272])

γF (−1) = 1 αF (−1) = 1

γF (2) = 1 αF (2) = 1

γF (3) = 2 αF (3) = 2

Diagonalizable? (Definition DZM [295])

Yes, large eigenspaces, Theorem DMLE [298].

The diagonalization. (Theorem DC [295])12 −5 1 −1−39 18 −7 3

277

−137

67

−17

267

−127

57

−27

33 −16 10 −299 −47 27 −778 −36 17 −6−9 2 3 4

1 2 1 172 5 1 450 2 0 211 1 7 0

=

−1 0 0 00 2 0 00 0 3 00 0 0 3

Page 438: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype G 406

Archetype G

Summary System with five equations, two variables. Consistent. Null space of co-efficient matrix has dimension 0. Coefficient matrix identical to that of Archetype H,constant vector is different.

A system of linear equations (Definition SSLE [10]):

2x1 + 3x2 = 6

−x1 + 4x2 = −14

3x1 + 10x2 = −2

3x1 − x2 = 20

6x1 + 9x2 = 18

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = 6, x2 = −2

Augmented matrix of the linear system of equations (Definition AM [23]):2 3 6−1 4 −143 10 −23 −1 206 9 18

Matrix in reduced row-echelon form, row-equivalent to augmented matrix:1 0 6

0 1 −20 0 00 0 00 0 0

Analysis of the augmented matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = {3}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the set

Page 439: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype G 407

F for the larger examples.[x1

x2

]=

[6−2

]

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

2x1 + 3x2 = 0

−x1 + 4x2 = 0

3x1 + 10x2 = 0

3x1 − x2 = 0

6x1 + 9x2 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros:

1 0 0

0 1 00 0 00 0 00 0 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 2 D = {1, 2} F = {3}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.

2 3−1 43 103 −16 9

Page 440: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype G 408

Matrix brought to reduced row-echelon form:

1 0

0 10 00 00 0

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = { }

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp ({ })

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

2−1336

,

3410−19

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =

1 0 0 0 −13

0 1 0 1− 13

0 0 1 1 −1

Page 441: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype G 409

Sp

1313

101

,

0−1−110

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

10213

,

011−10

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

({[10

],

[01

]})

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 2 Rank: 2 Nullity: 0

Page 442: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype H 410

Archetype H

Summary System with five equations, two variables. Inconsistent, overdetermined.Null space of coefficient matrix has dimension 0. Coefficient matrix identical to that ofArchetype G, constant vector is different.

A system of linear equations (Definition SSLE [10]):

2x1 + 3x2 = 5

−x1 + 4x2 = 6

3x1 + 10x2 = 2

3x1 − x2 = −1

6x1 + 9x2 = 3

Some solutions to the system of linear equations (not necessarily exhaustive):

None. (Why?)

Augmented matrix of the linear system of equations (Definition AM [23]):2 3 5−1 4 63 10 23 −1 −16 9 3

Matrix in reduced row-echelon form, row-equivalent to augmented matrix:1 0 0

0 1 0

0 0 10 0 00 0 0

Analysis of the augmented matrix (Notation RREFA [37]):

r = 3 D = {1, 2, 3} F = { }

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the set

Page 443: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype H 411

F for the larger examples.

Inconsistent system, no solutions exist.

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

2x1 + 3x2 = 0

−x1 + 4x2 = 0

3x1 + 10x2 = 0

3x1 − x2 = 0

6x1 + 9x2 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros:

1 0 0

0 1 00 0 00 0 00 0 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 2 D = {1, 2} F = {3}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.

2 3−1 43 103 −16 9

Page 444: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype H 412

Matrix brought to reduced row-echelon form:

1 0

0 10 00 00 0

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 2 D = {1, 2} F = { }

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp ({ })

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

2−1336

,

3410−19

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[]

Page 445: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype H 413

Sp

1313

101

,

0−1−110

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

10213

,

011−10

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =

1 0 0 0 −13

0 1 0 1− 13

0 0 1 1 −1

Sp

1313

101

,

0−1−110

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

({[10

],

[01

]})

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 2 Rank: 2 Nullity: 0

Page 446: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype I 414

Archetype I

Summary System with four equations, seven variables. Consistent. Null space ofcoefficient matrix has dimension 4.

A system of linear equations (Definition SSLE [10]):

x1 + 4x2 − x4 + 7x6 − 9x7 = 3

2x1 + 8x2 − x3 + 3x4 + 9x5 − 13x6 + 7x7 = 9

2x3 − 3x4 − 4x5 + 12x6 − 8x7 = 1

−x1 − 4x2 + 2x3 + 4x4 + 8x5 − 31x6 + 37x7 = 4

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = −25, x2 = 4, x3 = 22, x4 = 29, x5 = 1, x6 = 2, x7 = −3

x1 = −7, x2 = 5, x3 = 7, x4 = 15, x5 = −4, x6 = 2, x7 = 1

x1 = 4, x2 = 0, x3 = 2, x4 = 1, x5 = 0, x6 = 0, x7 = 0

Augmented matrix of the linear system of equations (Definition AM [23]):1 4 0 −1 0 7 −9 32 8 −1 3 9 −13 7 90 0 2 −3 −4 12 −8 1−1 −4 2 4 8 −31 37 4

Matrix in reduced row-echelon form, row-equivalent to augmented matrix:1 4 0 0 2 1 −3 4

0 0 1 0 1 −3 5 2

0 0 0 1 2 −6 6 10 0 0 0 0 0 0 0

Analysis of the augmented matrix (Notation RREFA [37]):

r = 3 D = {1, 3, 4} F = {2, 5, 6, 7, 8}

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.

Page 447: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype I 415

x1

x2

x3

x4

x5

x6

x7

=

4021000

+ x2

−4100000

+ x5

−20−1−2100

+ x6

−1036010

+ x7

30−5−6001

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

x1 + 4x2 − x4 + 7x6 − 9x7 = 0

2x1 + 8x2 − x3 + 3x4 + 9x5 − 13x6 + 7x7 = 0

2x3 − 3x4 − 4x5 + 12x6 − 8x7 = 0

−x1 − 4x2 + 2x3 + 4x4 + 8x5 − 31x6 + 37x7 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0, x4 = 0, x5 = 0, x6 = 0, x7 = 0

x1 = 3, x2 = 0, x3 = −5, x4 = −6, x5 = 0, x6 = 0, x7 = 1

x1 = −1, x2 = 0, x3 = 3, x4 = 6, x5 = 0, x6 = 1, x7 = 0

x1 = −2, x2 = 0, x3 = −1, x4 = −2, x5 = 1, x6 = 0, x7 = 0

x1 = −4, x2 = 1, x3 = 0, x4 = 0, x5 = 0, x6 = 0, x7 = 0

x1 = −4, x2 = 1, x3 = −3, x4 = −2, x5 = 1, x6 = 1, x7 = 1

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remainzeros:

1 4 0 0 2 1 −3 0

0 0 1 0 1 −3 5 0

0 0 0 1 2 −6 6 00 0 0 0 0 0 0 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 3 D = {1, 3, 4} F = {2, 5, 6, 7, 8}

Page 448: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype I 416

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.

1 4 0 −1 0 7 −92 8 −1 3 9 −13 70 0 2 −3 −4 12 −8−1 −4 2 4 8 −31 37

Matrix brought to reduced row-echelon form:

1 4 0 0 2 1 −3

0 0 1 0 1 −3 5

0 0 0 1 2 −6 60 0 0 0 0 0 0

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 3 D = {1, 3, 4} F = {2, 5, 6, 7}

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp

−4100000

,

−20−1−2100

,

−1036010

,

30−5−6001

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Page 449: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype I 417

Sp

120−1

,

0−122

,

−13−34

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[1 −12

31−13

31731

]

Sp

− 7

31

001

,

1331

010

,

1231

100

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

100−31

7

,

010127

,

001137

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

140021−3

,

00101−35

,

00012−66

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 7 Rank: 3 Nullity: 4

Page 450: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype J 418

Archetype J

Summary System with six equations, nine variables. Consistent. Null space of coeffi-cient matrix has dimension 5.

A system of linear equations (Definition SSLE [10]):

x1 + 2x2 − 2x3 + 9x4 + 3x5 − 5x6 − 2x7 + x8 + 27x9 = −5

2x1 + 4x2 + 3x3 + 4x4 − x5 + 4x6 + 10x7 + 2x8 − 23x9 = 18

x1 + 2x2 + x3 + 3x4 + x5 + x6 + 5x7 + 2x8 − 7x9 = 6

2x1 + 4x2 + 3x3 + 4x4 − 7x5 + 2x6 + 4x7 − 11x9 = 20

x1 + 2x2 + +5x4 + 2x5 − 4x6 + 3x7 + 8x8 + 13x9 = −4

−3x1 − 6x2 − x3 − 13x4 + 2x5 − 5x6 − 4x7 + 13x8 + 10x9 = −29

Some solutions to the system of linear equations (not necessarily exhaustive):

x1 = 6, x2 = 0, x3 = −1, x4 = 0, x5 = −1, x6 = 2, x7 = 0, x8 = 0, x9 = 0

x1 = 4, x2 = 1, x3 = −1, x4 = 0, x5 = −1, x6 = 2, x7 = 0, x8 = 0, x9 = 0

x1 = −17, x2 = 7, x3 = 3, x4 = 2, x5 = −1, x6 = 14, x7 = −1, x8 = 3, x9 = 2

x1 = −11, x2 = −6, x3 = 1, x4 = 5, x5 = −4, x6 = 7, x7 = 3, x8 = 1, x9 = 1

Augmented matrix of the linear system of equations (Definition AM [23]):1 2 −2 9 3 −5 −2 1 27 −52 4 3 4 −1 4 10 2 −23 181 2 1 3 1 1 5 2 −7 62 4 3 4 −7 2 4 0 −11 201 2 0 5 2 −4 3 8 13 −4−3 −6 −1 −13 2 −5 −4 13 10 −29

Matrix in reduced row-echelon form, row-equivalent to augmented matrix:

1 2 0 5 0 0 1 −2 3 6

0 0 1 −2 0 0 3 5 −6 −1

0 0 0 0 1 0 1 1 −1 −1

0 0 0 0 0 1 0 −2 −3 20 0 0 0 0 0 0 0 0 00 0 0 0 0 0 0 0 0 0

Analysis of the augmented matrix (Notation RREFA [37]):

r = 4 D = {1, 3, 5, 6} F = {2, 4, 7, 8, 9, 10}

Page 451: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype J 419

Vector form of the solution set to the system of equations (Theorem VFSLS [82]).Notice the relationship between the free variables and the set F above. Also, notice thepattern of 0’s and 1’s in the entries of the vectors corresponding to elements of the setF for the larger examples.

x1

x2

x3

x4

x5

x6

x7

x8

x9

=

60−10−12000

+ x2

−210000000

+ x4

−502100000

+ x7

−10−30−10100

+ x8

20−50−12010

+ x9

−306013001

Given a system of equations we can always build a new, related, homogeneous system(Definition HS [49]) by converting the constant terms to zeros and retaining the coeffi-cients of the variables. Properties of this new system will have precise relationships withvarious properties of the original system.

x1 + 2x2 − 2x3 + 9x4 + 3x5 − 5x6 − 2x7 + x8 + 27x9 = 0

2x1 + 4x2 + 3x3 + 4x4 − x5 + 4x6 + 10x7 + 2x8 − 23x9 = 0

x1 + 2x2 + x3 + 3x4 + x5 + x6 + 5x7 + 2x8 − 7x9 = 0

2x1 + 4x2 + 3x3 + 4x4 − 7x5 + 2x6 + 4x7 − 11x9 = 0

x1 + 2x2 + +5x4 + 2x5 − 4x6 + 3x7 + 8x8 + 13x9 = 0

−3x1 − 6x2 − x3 − 13x4 + 2x5 − 5x6 − 4x7 + 13x8 + 10x9 = 0

Some solutions to the associated homogenous system of linear equations (not neces-sarily exhaustive):x1 = 0, x2 = 0, x3 = 0, x4 = 0, x5 = 0, x6 = 0, x7 = 0, x8 = 0, x9 = 0

x1 = −2, x2 = 1, x3 = 0, x4 = 0, x5 = 0, x6 = 0, x7 = 0, x8 = 0, x9 = 0

x1 = −23, x2 = 7, x3 = 4, x4 = 2, x5 = 0, x6 = 12, x7 = −1, x8 = 3, x9 = 2

x1 = −17, x2 = −6, x3 = 2, x4 = 5, x5 = −3, x6 = 5, x7 = 3, x8 = 1, x9 = 1

Form the augmented matrix of the homogenous linear system, and use row operationsto convert to reduced row-echelon form. Notice how the entries of the final column remain

Page 452: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype J 420

zeros:

1 2 0 5 0 0 1 −2 3 0

0 0 1 −2 0 0 3 5 −6 0

0 0 0 0 1 0 1 1 −1 0

0 0 0 0 0 1 0 −2 −3 00 0 0 0 0 0 0 0 0 00 0 0 0 0 0 0 0 0 0

Analysis of the augmented matrix for the homogenous system (Notation RREFA [37]).Notice the slight variation for the same analysis of the original system only when the orig-inal system was consistent:

r = 4 D = {1, 3, 5, 6} F = {2, 4, 7, 8, 9, 10}

Coefficient matrix of original system of equations, and of associated homogenoussystem. This matrix will be the subject of further analysis, rather than the systems ofequations.

1 2 −2 9 3 −5 −2 1 272 4 3 4 −1 4 10 2 −231 2 1 3 1 1 5 2 −72 4 3 4 −7 2 4 0 −111 2 0 5 2 −4 3 8 13−3 −6 −1 −13 2 −5 −4 13 10

Matrix brought to reduced row-echelon form:

1 2 0 5 0 0 1 −2 3

0 0 1 −2 0 0 3 5 −6

0 0 0 0 1 0 1 1 −1

0 0 0 0 0 1 0 −2 −30 0 0 0 0 0 0 0 00 0 0 0 0 0 0 0 0

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 4 D = {1, 3, 5, 6} F = {2, 4, 7, 8, 9}

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix

Page 453: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype J 421

(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp

−210000000

,

−502100000

,

−10−30−10100

,

20−50−12010

,

−306013001

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Sp

12121−3

,

−23130−1

,

3−11−722

,

−5412−4−5

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =

[1 0 186

13151131

−188131

77131

0 1 −272131

− 45131

58131

− 14131

]

Sp

− 77

13114131

0001

,

188131

− 58131

0010

,

− 51

13145131

0100

,

−186

131272131

1000

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Page 454: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype J 422

Sp

1000−1−29

7

,

0100−11

2

−947

,

00101022

,

000132

3

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

1205001−23

,

001−20035−6

,

00001011−1

,

0000010−2−3

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 9 Rank: 4 Nullity: 5

Page 455: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype K 423

Archetype K

Summary Square matrix of size 5. Nonsingular. 3 distinct eigenvalues, 2 of multiplic-ity 2.

A matrix:10 18 24 24 −1212 −2 −6 0 −18−30 −21 −23 −30 3927 30 36 37 −3018 24 30 30 −20

Matrix brought to reduced row-echelon form:1 0 0 0 0

0 1 0 0 0

0 0 1 0 0

0 0 0 1 0

0 0 0 0 1

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 5 D = {1, 2, 3, 4, 5} F = { }

Matrix (coefficient matrix) is nonsingular or singular? (Theorem NSRRI [57]) at thesame time, examine the size of the set F above.Notice that this property does not applyto matrices that are not square.

Nonsingular.

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp ({ })

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set Dabove. (Theorem BROC [131])

Page 456: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype K 424

Sp

1012−302718

,

18−2−213024

,

24−6−233630

,

240−303730

,

−12−1839−30−20

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =[]

Sp

10000

,

01000

,

00100

,

00010

,

00001

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

10000

,

01000

,

00100

,

00010

,

00001

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

10000

,

01000

,

00100

,

00010

,

00001

Inverse matrix, if it exists. The inverse is not defined for matrices that are not square,and if the matrix is square, then the matrix must be nonsingular. (Definition MI [167],Theorem NSI [177])

Page 457: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype K 425

1 −

(94

)−(

32

)3 −6

212

434

212

9 −9−15 −

(212

)−11 −15 39

2

9 154

92

10 −1592

34

32

6 −(

192

)

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 5 Rank: 5 Nullity: 0

Determinant of the matrix, which is only defined for square matrices. The matrix isnonsingular if and only if the determinant is nonzero (Theorem SMZD [259]). (Productof all eigenvalues?)

Determinant = 16

Eigenvalues, and bases for eigenspaces. (Definition EEM [261],Definition EM [270])

λ = −2 EK (−2) = Sp

2−2101

,

−12−210

λ = 1 EK (1) = Sp

4−10702

,

−418−1750

λ = 4 EK (4) = Sp

1−1011

Geometric and algebraic multiplicities. (Definition GME [272]Definition AME [272])

γK (−2) = 2 αK (−2) = 2

γK (1) = 2 αK (1) = 2

γK (4) = 1 αK (4) = 1

Diagonalizable? (Definition DZM [295])

Page 458: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype K 426

Yes, large eigenspaces, Theorem DMLE [298].

The diagonalization. (Theorem DC [295])−4 −3 −4 −6 7−7 −5 −6 −8 101 −1 −1 1 −31 0 0 1 −22 5 6 4 0

10 18 24 24 −1212 −2 −6 0 −18−30 −21 −23 −30 3927 30 36 37 −3018 24 30 30 −20

2 −1 4 −4 1−2 2 −10 18 −11 −2 7 −17 00 1 0 5 11 0 2 0 1

=

−2 0 0 0 00 −2 0 0 00 0 1 0 00 0 0 1 00 0 0 0 4

Page 459: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype L 427

Archetype L

Summary Square matrix of size 5. Singular, nullity 2. 2 distinct eigenvalues, each of“high” multiplicity.

A matrix:−2 −1 −2 −4 4−6 −5 −4 −4 610 7 7 10 −13−7 −5 −6 −9 10−4 −3 −4 −6 6

Matrix brought to reduced row-echelon form:1 0 0 1 −2

0 1 0 −2 2

0 0 1 2 −10 0 0 0 00 0 0 0 0

Analysis of the row-reduced matrix (Notation RREFA [37]):

r = 5 D = {1, 2, 3} F = {4, 5}

Matrix (coefficient matrix) is nonsingular or singular? (Theorem NSRRI [57]) at thesame time, examine the size of the set F above.Notice that this property does not applyto matrices that are not square.

Singular.

This is the null space of the matrix. The set of vectors used in the span constructionis a linearly independent set of column vectors that spans the null space of the matrix(Theorem SSNS [90], Theorem BNS [104]). Solve the homogenous system with this ma-trix as the coefficient matrix and write the solutions in vector form (Theorem VFSLS [82])to see these vectors arise.

Sp

−12−210

,

2−2101

Range of the matrix, expressed as the span of a set of linearly independent vectorsthat are also columns of the matrix. These columns have indices that form the set D

Page 460: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype L 428

above. (Theorem BROC [131])

Sp

−− 2−610−7−4

,

−1−57−5−3

,

−2−47−6−4

The range of the matrix, as it arises from row operations on an augmented matrix.The matrix K is computed as described in Theorem RNS [136]. This is followed by therange described by a set of linearly independent vectors that span the null space of K,computed as according to Theorem SSNS [90]. When r = m, the matrix K has no rowsand the range is all of Cm.

K =

[1 0 −2 −6 50 1 4 10 −9

]

Sp

−59001

,

6−10010

,

2−4100

Range of the matrix, expressed as the span of a set of linearly independent vec-tors. These vectors are computed by row-reducing the transpose the matrix into reducedrow-echelon form, tossing out the zero rows, and writing the remaining rows as columnvectors. By Theorem RMRST [145] and Theorem BRS [144], and in the style of Exam-ple RSM.IS [145], this yields a linearly independent set of vectors that span the range.

Sp

1009452

,

0105432

,

00112

1

Row space of the matrix, expressed as a span of a set of linearly independent vectors,obtained from the nonzero rows of the equivalent matrix in reduced row-echelon form.(Theorem BRS [144])

Sp

1001−2

,

010−22

,

0012−1

Inverse matrix, if it exists. The inverse is not defined for matrices that are not square,

Page 461: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype L 429

and if the matrix is square, then the matrix must be nonsingular. (Definition MI [167],Theorem NSI [177])

Subspace dimensions associated with the matrix. (Definition NOM [237], Defini-tion ROM [237]) Verify Theorem RPNC [238]

Matrix columns: 5 Rank: 3 Nullity: 2

Determinant of the matrix, which is only defined for square matrices. The matrix isnonsingular if and only if the determinant is nonzero (Theorem SMZD [259]). (Productof all eigenvalues?)

Determinant = 0

Eigenvalues, and bases for eigenspaces. (Definition EEM [261],Definition EM [270])

λ = −1 EL (−1) = Sp

−59001

,

6−10010

,

2−4100

λ = 0 EL (0) = Sp

2−2101

,

−12−210

Geometric and algebraic multiplicities. (Definition GME [272]Definition AME [272])

γL (−1) = 3 αL (−1) = 3

γL (0) = 2 αL (0) = 2

Diagonalizable? (Definition DZM [295])

Page 462: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype L 430

Yes, large eigenspaces, Theorem DMLE [298].

The diagonalization. (Theorem DC [295])4 3 4 6 −67 5 6 9 −10−10 −7 −7 −10 13−4 −3 −4 −6 7−7 −5 −6 −8 10

−2 −1 −2 −4 4−6 −5 −4 −4 610 7 7 10 −13−7 −5 −6 −9 10−4 −3 −4 −6 6

−5 6 2 2 −19 −10 −4 −2 20 0 1 1 −20 1 0 0 11 0 0 1 0

=

−1 0 0 0 00 −1 0 0 00 0 −1 0 00 0 0 0 00 0 0 0 0

Page 463: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype M 431

Archetype M

Summary Linear transformation with bigger domain than codomain, so it is guaran-teed to not be injective. Happens to not be surjective.

A linear transformation: (Definition LT [302])

T : C5 7→ C3, T

x1

x2

x3

x4

x5

=

x1 + 2x2 + 3x3 + 4x4 + 4x5

3x1 + x2 + 4x3 − 3x4 + 7x5

x1 − x2 − 5x4 + x5

A basis for the null space of the linear transformation: (Definition NSLT [325])

−2−1001

,

2−3010

,

−1−1100

Injective: No. (Definition ILT [321])

Since the null space is nontrivial Theorem XXoneonetrivial [??] tells us that the lineartransformation is not injective. Also, since the rank can not exceed 3, we are guaranteedto have a nullity of at least 2, just from checking dimensions of the domain and thecodomain. In particular, verify that

T

12−145

=

3824−16

T

0−3056

=

3824−16

This demonstration that T is not injective is constructed with the observation that

0−3056

=

12−145

+

−1−5111

and

z =

−1−5111

∈ N(T )

Page 464: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype M 432

so the vector z effectively “does nothing” in the evaluation of T .

A basis for the range of the linear transformation: (Definition RLT [336])

Evaluate the linear transformation on a standard basis to get a spanning set for the range:

1

31

,

21−1

,

340

,

4−3−5

,

471

If the linear transformation is injective, then the set above is guaranteed to be linearlyindependent (XXXX). If this spanning set is not linearly dependent at first, use relationsof linear dependence on the set to remove redundant elements while preserving the span-ning properties. Continue this until the set is linear independent. A basis for the rangeis:

10−4

5

,

0135

Surjective: No. (Definition SLT [332])

Notice that the range is not all of C3 since its dimension 2, not 3. In particular, verify

that

345

6∈ R (T ), by setting the output equal to this vector and seeing that the resulting

system of linear equations has no solution, i.e. is inconsistent. This alone is sufficient tosee that the linear transformation is not onto.

Subspace dimensions associated with the linear transformation. Examine parallelswith earlier results for matrices.

Domain dimension: 5 Rank: 2 Nullity: 3

Invertible: No.

Not injective or surjective.

Page 465: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype N 433

Archetype N

Summary Linear transformation with domain larger than its codomain, so it is guar-anteed to not be injective. Happens to be onto.

A linear transformation: (Definition LT [302])

T : C5 7→ C3, T

x1

x2

x3

x4

x5

=

2x1 + x2 + 3x3 − 4x4 + 5x5

x1 − 2x2 + 3x3 − 9x4 + 3x5

3x1 + 4x3 − 6x4 + 5x5

A basis for the null space of the linear transformation: (Definition NSLT [325])

1−1−201

,

−2−1310

Since the null space is nontrivial Theorem XXoneonetrivial [??] tells us that the lineartransformation is not injective. Also, since the rank can not exceed 3, we are guaranteedto have a nullity of at least 2, just from checking dimensions of the domain and thecodomain. In particular, verify that

T

−31−2−31

=

6196

T

−4−4−2−14

=

6196

This demonstration that T is not injective is constructed with the observation that

−4−4−2−14

=

−31−2−31

+

−1−5023

and

z =

−1−5023

∈ N(T )

Page 466: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype N 434

so the vector z effectively “does nothing” in the evaluation of T . A basis for therange of the linear transformation: (Definition RLT [336])

Evaluate the linear transformation on a standard basis to get a spanning set for the range:

2

13

,

1−20

,

334

,

−4−9−6

,

535

If the linear transformation is injective, then the set above is guaranteed to be linearlyindependent (XXXX). If this spanning set is not linearly dependent at first, use relationsof linear dependence on the set to remove redundant elements while preserving the span-ning properties. Continue this until the set is linear independent. A basis for the rangeis:

100

,

010

,

001

Surjective: Yes. (Definition SLT [332])

Notice that the basis for the range above is the standard basis for C3. So the range is allof C3 and thus the linear transformation is surjective.

Subspace dimensions associated with the linear transformation. Examine parallelswith earlier results for matrices.

Domain dimension: 5 Rank: 3 Nullity: 2

Invertible: No.

Not surjective, and the relative sizes of the domain and codomain mean the linear trans-formation cannot be injective.

Page 467: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype O 435

Archetype O

Summary Linear transformation with a domain smaller than the codomain, so it isguaranteed to not be onto. Happens to not be one-to-one.

A linear transformation: (Definition LT [302])

T : C3 7→ C5, T

x1

x2

x3

=

−x1 + x2 − 3x3

−x1 + 2x2 − 4x3

x1 + x2 + x3

2x1 + 3x2 + x3

x1 + 2x3

A basis for the null space of the linear transformation: (Definition NSLT [325])

−2

11

Injective: No. (Definition ILT [321])

Since the null space is nontrivial Theorem XXoneonetrivial [??] tells us that the lineartransformation is not injective. Also, since the rank can not exceed 3, we are guaranteedto have a nullity of at least 2, just from checking dimensions of the domain and thecodomain. In particular, verify that

T

5−13

=

−15−1971011

T

115

=

−15−1971011

This demonstration that T is not injective is constructed with the observation that1

15

=

5−13

+

−422

and

z =

−422

∈ N(T )

so the vector z effectively “does nothing” in the evaluation of T .

A basis for the range of the linear transformation: (Definition RLT [336])

Page 468: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype O 436

Evaluate the linear transformation on a standard basis to get a spanning set for the range:

−1−1121

,

12130

,

−3−4112

If the linear transformation is injective, then the set above is guaranteed to be linearlyindependent (XXXX). If this spanning set is not linearly dependent at first, use relationsof linear dependence on the set to remove redundant elements while preserving the span-ning properties. Continue this until the set is linear independent. A basis for the rangeis:

10−3−7−2

,

01251

Subspace dimensions associated with the linear transformation. Examine parallelswith earlier results for matrices.

Domain dimension: 3 Rank: 2 Nullity: 1

Surjective: No. (Definition SLT [332])

The dimension of the range is 2, and the codomain (C5) has dimension 5. So the transfor-mation is not onto. Notice too that since the domain C3 has dimension 3, it is impossiblefor the range to have a dimension greater than 3, and no matter what the actual definitionof the function, it cannot possibly be onto.

To be more precise, verify that

23111

6∈ R(T ), by setting the output equal to this

vector and seeing that the resulting system of linear equations has no solution, i.e. isinconsistent. This alone is sufficient to see that the linear transformation is not onto.

Invertible: No.

Not injective, and the relative dimensions of the domain and codomain prohibit anypossibility of being surjective.

Page 469: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype P 437

Archetype P

Summary Linear transformation with a domain smaller that its codomain, so it isguaranteed to not be surjective. Happens to be injective.

A linear transformation: (Definition LT [302])

T : C3 7→ C5, T

x1

x2

x3

=

−x1 + x2 + x3

−x1 + 2x2 + 2x3

x1 + x2 + 3x3

2x1 + 3x2 + x3

−2x1 + x2 + 3x3

A basis for the null space of the linear transformation: (Definition NSLT [325])

{ }

Injective: Yes. (Definition ILT [321])

Since N (T ) = {0}, Theorem XXoneonetrivial [??] tells us that T is injective.

A basis for the range of the linear transformation: (Definition RLT [336])

Evaluate the linear transformation on a standard basis to get a spanning set for the range:

−1−112−2

,

12131

,

12313

If the linear transformation is injective, then the set above is guaranteed to be linearlyindependent (XXXX). If this spanning set is not linearly dependent at first, use relationsof linear dependence on the set to remove redundant elements while preserving the span-ning properties. Continue this until the set is linear independent. A basis for the rangeis:

100−106

,

0107−3

,

001−11

Surjective: No. (Definition SLT [332])

The dimension of the range is 3, and the codomain (C5) has dimension 5. So the trans-formation is not surjective. Notice too that since the domain C3 has dimension 3, it is

Page 470: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype P 438

impossible for the range to have a dimension greater than 3, and no matter what theactual definition of the function, it cannot possibly be surjective in this situation.

To be more precise, verify that

21−326

6∈ R (() T ), by setting the output equal to this

vector and seeing that the resulting system of linear equations has no solution, i.e. isinconsistent. This alone is sufficient to see that the linear transformation is not onto.

Subspace dimensions associated with the linear transformation. Examine parallelswith earlier results for matrices.

Domain dimension: 3 Rank: 3 Nullity: 0

Invertible: No.

Not surjective.

Page 471: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype Q 439

Archetype Q

Summary Linear transformation with equal-sized domain and codomain, so it hasthe potential to be invertible, but in this case is not. Neither injective nor surjective.Diagonalizable, though.

A linear transformation: (Definition LT [302])

T : C5 7→ C5, T

x1

x2

x3

x4

x5

=

−2x1 + 3x2 + 3x3 − 6x4 + 3x5

−16x1 + 9x2 + 12x3 − 28x4 + 28x5

−19x1 + 7x2 + 14x3 − 32x4 + 37x5

−21x1 + 9x2 + 15x3 − 35x4 + 39x5

−9x1 + 5x2 + 7x3 − 16x4 + 16x5

A basis for the null space of the linear transformation: (Definition NSLT [325])

34133

Injective: No. (Definition ILT [321])

Since the null space is nontrivial Theorem XXoneonetrivial [??] tells us that the lineartransformation is not injective. Also, since the rank can not exceed 3, we are guaranteedto have a nullity of at least 2, just from checking dimensions of the domain and thecodomain. In particular, verify that

T

13−124

=

455727731

T

47057

=

455727731

This demonstration that T is not injective is constructed with the observation that

47057

=

13−124

+

34133

Page 472: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype Q 440

and

z =

34133

∈ N(T )

so the vector z effectively “does nothing” in the evaluation of T .

A basis for the range of the linear transformation: (Definition RLT [336])

Evaluate the linear transformation on a standard basis to get a spanning set for the range:

−2−16−19−21−9

,

39795

,

31214157

,

−6−28−32−35−16

,

328373916

If the linear transformation is injective, then the set above is guaranteed to be linearlyindependent (XXXX). If this spanning set is not linearly dependent at first, use relationsof linear dependence on the set to remove redundant elements while preserving the span-ning properties. Continue this until the set is linear independent. A basis for the rangeis:

10001

,

0100−1

,

0010−1

,

00012

Surjective: No. (Definition SLT [332])

The dimension of the range is 4, and the codomain (C5) has dimension 5. So R (T ) 6= C5

and by Theorem XXrangecodomain [??] the transformation is not surjective.

To be more precise, verify that

−123−14

6∈ R(T ), by setting the output equal to this

vector and seeing that the resulting system of linear equations has no solution, i.e. isinconsistent. This alone is sufficient to see that the linear transformation is not onto.

Subspace dimensions associated with the linear transformation. Examine parallelswith earlier results for matrices.

Domain dimension: 5 Rank: 4 Nullity: 1

Invertible: No.

Page 473: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype Q 441

Neither injective nor surjective. Notice that since the domain and codomain have thesame dimesion, either the transformation is both onto and one-to-one (making it invert-ible) or else it is both not onto and not one-to-one (as in this case) by Theorem XXsumofrankandnullity [??].

Page 474: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype R 442

Archetype R

Summary Linear transformation with equal-sized domain and codomain. Injective,surjective, invertible, diagonalizable, the works.

A linear transformation: (Definition LT [302])

T : C5 7→ C5, T

x1

x2

x3

x4

x5

=

−65x1 + 128x2 + 10x3 − 262x4 + 40x5

36x1 − 73x2 − x3 + 151x4 − 16x5

−44x1 + 88x2 + 5x3 − 180x4 + 24x5

34x1 − 68x2 − 3x3 + 140x4 − 18x5

12x1 − 24x2 − x3 + 49x4 − 5x5

A basis for the null space of the linear transformation: (Definition NSLT [325])

{ }

Injective: Yes. (Definition ILT [321])

Since the null space is trivial Theorem XXoneonetrivial [??] tells us that the lineartransformation is injective.

A basis for the range of the linear transformation: (Definition RLT [336])

Evaluate the linear transformation on a standard basis to get a spanning set for the range:

−6536−443412

,

128−7388−68−24

,

10−15−3−1

,

−262151−18014049

,

40−1624−18−5

If the linear transformation is injective, then the set above is guaranteed to be linearlyindependent (XXXX). If this spanning set is not linearly dependent at first, use relationsof linear dependence on the set to remove redundant elements while preserving the span-ning properties. Continue this until the set is linear independent. A basis for the rangeis:

10000

,

01000

,

00100

,

00010

,

00001

Surjective: Yes/No. (Definition SLT [332])

A basis for the range is the standard basis of C5, so R (T ) = C5 and Theorem XXrange-

Page 475: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype R 443

codomain [??] tells us T is surjective. Or, the dimension of the range is 5, and thecodomain (C5) has dimension 5. So the transformation is surjective.

Subspace dimensions associated with the linear transformation. Examine parallelswith earlier results for matrices.

Domain dimension: 5 Rank: 5 Nullity: 0

Invertible: Yes.

Both injective and surjective. Notice that since the domain and codomain have the samedimesion, either the transformation is both injective and surjective (making it invertible,as in this case) or else it is both not injective and not surjective.

Page 476: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype S 444

Archetype S

Summary Domain is column vectors, codomain is matrices. Domain is dimension 3and codomain is dimension 4. Not injective, not surjective.

A linear transformation: (Definition LT [302])

T : C3 7→ M22, T

abc

=

[a− b 2a + 2b + c

3a + b + c −2a− 6b− 2c

]

Archetype T

Summary Domain and codomain are polynomials. Domain has dimension 5, whilecodomain has dimension 6. Is injective, can’t be surjective.

A linear transformation: (Definition LT [302])

T : P4 7→ P5, T (p(x)) = (x− 2)p(x)

Archetype U

Summary Domain is matrices, codomain is column vectors. Domain has dimension 6,while codomain has dimension 4. Can’t be injective, is surjective.

A linear transformation: (Definition LT [302])

T : M23 7→ C4, T

([a b cd e f

])=

a + 2b + 12c− 3d + e + 6f

2a− b− c + d− 11fa + b + 7c + 2d + e− 3fa + 2b + 12c + 5e− 5f

Archetype V

Summary Domain is polynomials, codomain is matrices. Domain and codomain bothhave dimension 4. Injective, surjective, invertible, (eigenvalues, diagonalizable???).

A linear transformation: (Definition LT [302])

T : P3 7→ M22, T(a + bx + cx2 + dx3

)=

[a + b a− 2c

d b− d

]

Page 477: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Archetype V 445

When invertible, the inverse linear transformation. (Definition IVLT [344])

T−1 : M22 7→ P3, T−1

([a bc d

])= (a− c− d) + (c + d)x +

1

2(a− b− c− d)x2 + cx3

Page 478: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Part TTopics

446

Page 479: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

P: Preliminaries

Section CNO

Complex Number Operations

In this section we review of the basics of working with complex numbers.

Subsection CNAArithmetic with complex numbers

A complex number is a linear combination of 1 and i =√−1, typically written in the

form a + bi. Complex numbers can be added, subtracted, multiplied and divided, justlike we are used to doing with real numbers, including the restriction on division by zero.We will not define these operations carefully, but instead illustrate with examples.

Example CNO.ACNArithmetic of complex numbers

(2 + 5i) + (6− 4i) = (2 + 6) + (5 + (−4))i = 8 + i

(2 + 5i)− (6− 4i) = (2− 6) + (5− (−4))i = −4 + 9i

(2 + 5i)(6− 4i) = (2)(6) + (5i)(6) + (2)(−4i) + (5i)(−4i) = 12 + 30i− 8i− 20i2

= 12 + 22i− 20(−1) = 32 + 22i

Division takes just a bit more care. We multiply the denominator by a complex numberchosen to produce a real number and then we can produce a complex number as a result.

2 + 5i

6− 4i=

2 + 5i

6− 4i

6 + 4i

6 + 4i=−8 + 38i

52= − 8

52+

38

52i = − 2

13+

19

26i 4

In this example, we used 6 + 4i to convert the denominator in the fraction to a realnumber. This number is known as the conjugate, which we now define.

447

Page 480: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection CNO.CCN Conjugates of Complex Numbers 448

Subsection CCNConjugates of Complex Numbers

Definition CCNConjugate of a Complex NumberThe conjugate of the complex number c = a+bi ∈ C is the complex number c = a−bi.�

Example CNO.CSCNConjugate of some complex numbers

2 + 3i = 2− 3i 5− 4i = 5 + 4i −3 + 0i = −3 + 0i 0 + 0i = 0 + 0i 4

Notice how the conjugate of a real number leaves the number unchanged. The conju-gate enjoys some basic properties that are useful when we work with linear expressionsinvolving addition and multiplication.

Theorem CCRAComplex Conjugation Respects AdditionSuppose that c and d are complex numbers. Then c + d = c + d. �

Proof Let c = a + bi and d = r + si. Then

c + d = (a + r) + (b + s)i = (a + r)− (b + s)i = (a− bi) + (r − si) = c + d �

Theorem CCRMComplex Conjugation Respects MultiplicationSuppose that c and d are complex numbers. Then cd = cd. �

Proof Let c = a + bi and d = r + si. Then

cd = (ar − bs) + (as + br)i = (ar − bs)− (as + br)i

= (ar − (−b)(−s)) + (a(−s) + (−b)r)i = (a− bi)(r − si) = cd �

Subsection MCNModulus of a Complex Number

We define one more operation with complex numbers that may be new to you.

Definition MCNModulus of a Complex NumberThe modulus of the complex number c = a + bi ∈ C, is the real number

|c| =√

cc =√

a2 + b2. �

Page 481: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Subsection CNO.MCN Modulus of a Complex Number 449

Example CNO.MSCNModulus of some complex numbers

|2 + 3i| =√

13 |5− 4i| =√

41 |−3 + 0i| = 3 |0 + 0i| = 0 4

The modulus can be interpreted as a version of the absolute value for complex numbers,as is suggested by the notation employed. You can see this in how |−3| = |−3 + 0i| = 3.Notice too how the modulus of the complex zero, 0 + 0i, has value 0.

Page 482: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Part AApplications

450

Page 483: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Part A Applications 451

Page 484: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

Index

A(archetype), 379

A (chapter), 375A (definition), 181A (part), 451A (subsection of WILA), 3additive inverse

from scalar multiplicationtheorem AISM, 192

additive inversesunique

theorem AIU, 191adjoint

definition A, 181AISM (theorem), 192AIU (theorem), 191AM (definition), 23AME (definition), 272AMN (notation), 54Archetype A

definition, 379linearly dependent columns, 103range, 139singular matrix, 56solving homogeneous system, 50system as linear combination, 75

archetype Aaugmented matrix

example RREF.AMAA, 24archetype A:solutions

example RREF.SAA, 30Archetype B

definition, 384inverse

example MISLE.CMIAB, 173linearly independent columns, 103nonsingular matrix, 57not invertible

example MISLE.MWIAA, 167range, 139solutions via inverse

example MISLE.SABMI, 166solving homogeneous system, 50system as linear combination, 74vector equality, 66

archetype Bsolutions

example RREF.SAB, 29Archetype C

definition, 389homogeneous system, 49

Archetype Ddefinition, 393range as null space, 134range, original columns, 133solving homogeneous system, 51vector form of solutions, 80

Archetype Edefinition, 397

archetype E:solutionsexample RREF.SAE, 31

Archetype Fdefinition, 401

Archetype Gdefinition, 406

Archetype Hdefinition, 410

Archetype Idefinition, 414null space, 55range from row operations, 146row space, 141vector form of solutions, 83

Archetype I:casting out columns, range,128

Archetype Jdefinition, 418

Archetype Kdefinition, 423inverse

example MISLE.CMIAK, 170example MISLE.MIAK, 168

452

Page 485: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 453

Archetype Ldefinition, 427null space span, linearly independent,

105vector form of solutions, 85

Archetype Mdefinition, 431

Archetype Ndefinition, 433

Archetype Odefinition, 435

Archetype Pdefinition, 437

Archetype Qdefinition, 439

Archetype Rdefinition, 442

Archetype Sdefinition, 444

Archetype Tdefinition, 444

Archetype Udefinition, 444

Archetype Vdefinition, 444

augmented matrixnotation AMN, 54

B(archetype), 384

B (definition), 220B (section), 212B (subsection of B), 219B.AVR (example), 226B.BM (example), 221B.BP (example), 221B.BSM22 (example), 222B.BSP4 (example), 221B.CABAK (example), 225B.LIM32 (example), 214B.LIP4 (example), 212B.RS (example), 224B.RSB (example), 223B.SSM22 (example), 218B.SSP4 (example), 216basis

columns nonsingular matrixexample B.CABAK, 225

common sizetheorem BIS, 234

definition B, 220matrices

example B.BM, 221example B.BSM22, 222

polynomialsexample B.BP, 221example B.BSP4, 221example PD.BP, 245example PD.SVP4, 246

subspace of matricesexample PD.BSM22, 245

BIS (theorem), 234BNS (theorem), 104BNSM (subsection of B), 225BROC (theorem), 131BRS (subsection of B), 223BRS (theorem), 144

C(archetype), 389

C (part), 2C (technique), 28cancellation

vector additiontheorem VAC, 194

CAV (subsection of O), 109CCM (definition), 126CCN (definition), 448CCN (subsection of CNO), 448CCRA (theorem), 448CCRM (theorem), 448CCRSM (theorem), 110CCRVA (theorem), 109CCV (definition), 109CD (subsection of DM), 256CEE (subsection of EE), 268CFDVS (theorem), 364characteristic polynomial

definition CP, 268degree

theorem DCP, 286size 3 matrix

example EE.CPMS3, 269CILT (subsection of ILT), 331CILTI (theorem), 331CIM (definition), 256

Page 486: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 454

CIM (subsection of MISLE), 169CINSM (theorem), 172CIVLT (theorem), 349CLI (theorem), 365CLTLT (theorem), 319CM (definition), 52CMVEI (theorem), 44CNA (subsection of CNO), 447CNO (section), 447CNO.ACN (example), 447CNO.CSCN (example), 448CNO.MSCN (example), 449CNSMB (theorem), 225COB (theorem), 249coefficient matrix

definition CM, 52nonsingular

theorem SNSCM, 178COMOS (theorem), 179complex m-space

example VS.VSCM, 186complex arithmetic

example CNO.ACN, 447complex number

conjugateexample CNO.CSCN, 448

modulusexample CNO.MSCN, 449

complex numberconjugate

definition CCN, 448modulus

definition MCN, 448complex vector space

dimensiontheorem DCM, 235

compositioninjective linear transformations

theorem CILTI, 331surjective linear transformations

theorem CSLTS, 343computation

LS.MMA, 45ME.MMA, 22ME.TI83, 23ME.TI86, 23MI.MMA, 173MM.MMA, 154

RR.MMA, 32RR.TI83, 33RR.TI86, 33VLC.MMA, 69VLC.TI83, 70VLC.TI86, 69

conjugateaddition

theorem CCRA, 448column vector

definition CCV, 109matrix

definition CCM, 126multiplication

theorem CCRM, 448scalar multiplication

theorem CCRSM, 110vector addition

theorem CCRVA, 109consistent linear system, 41consistent linear systems

theorem CSRN, 42consistent system

definition CS, 37constructive proofs

technique C, 28contrapositive

technique CP, 41converse

technique CV, 43coordinates

orthonormal basistheorem COB, 249

coordinatizationlinear combination of matrices

example VR.CM32, 367linear independence

theorem CLI, 365orthonormal basis

example PD.CROB3, 250example PD.CROB4, 250

spanning setstheorem CSS, 366

coordinatization principle, 367coordinatizing

polynomialsexample VR.CP2, 366

corollary

Page 487: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 455

FVCS, 42LIVRN, 100

CP (definition), 268CP (technique), 41crazy vector space

example VR.CVSR, 365properties

example VS.PCVS, 193CRN (theorem), 238CS (definition), 37CSLT (subsection of SLT), 343CSLTS (theorem), 343CSRN (theorem), 42CSS (theorem), 366CSSM (theorem), 194CV (definition), 52CV (technique), 43CVA (definition), 67CVE (definition), 66CVS (subsection of VR), 364CVSM (definition), 68CVSM (theorem), 194

D(archetype), 393

D (chapter), 254D (definition), 230D (section), 230D (subsection of D), 230D (subsection of SD), 295D (technique), 9D.DSM22 (example), 235D.DSP4 (example), 236D.LDP4 (example), 234D.RNM (example), 237D.RNSM (example), 238D.VSPUD (example), 236DC (technique), 74DC (theorem), 295DCM (theorem), 235DCP (theorem), 286decomposition

technique DC, 74DED (theorem), 300definition

A, 181AM, 23AME, 272

B, 220CCM, 126CCN, 448CCV, 109CIM, 256CM, 52CP, 268CS, 37CV, 52CVA, 67CVE, 66CVSM, 68D, 230DIM, 295DM, 254DZM, 295EEM, 261EM, 270EO, 12ES, 12GME, 272HM, 181HS, 49IDLT, 344IDV, 40ILT, 321IM, 57IP, 110IVLT, 344IVS, 350LC, 202LCCV, 72LI, 212LICV, 96LO, 27LT, 302LTA, 316LTC, 319LTSM, 317M, 22MA, 122MCN, 448ME, 121MI, 167MIM, 256MM, 153MN, 22MR, 369

Page 488: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 456

MVP, 151NM, 56NOLT, 352NOM, 237NSLT, 325NSM, 54NV, 113OM, 178ONS, 119OSV, 115OV, 114PI, 314REM, 25RLD, 212RLDCV, 96RLT, 336RM, 127RO, 24ROLT, 352ROM, 237RREF, 27RSM, 141S, 196SIM, 291SLT, 332SM, 254SMM, 122SQM, 56SS, 203SSCV, 87SSLE, 10SUV, 169SV, 53SYM, 125TM, 124TS, 200TSHSE, 50TSS, 216VOC, 53VR, 358VS, 184VSCM, 65VSM, 121ZM, 124ZRM, 27ZV, 52

definitionstechnique D, 9

DERC (theorem), 257determinant

computed two waysexample DM.TCSD, 257

definition DM, 254expansion

theorem DERC, 257matrix multiplication

theorem DRMM, 259nonsingular matrix, 259size 2 matrix

theorem DMST, 255size 3 matrix

example DM.D33M, 255transpose

theorem DT, 258zero

theorem SMZD, 259zero versus nonzero

example DM.ZNDAB, 259determinant, upper triangular matrix

example DM.DUTM, 258diagonal matrix

definition DIM, 295diagonalizable

definition DZM, 295distinct eigenvalues

example SD.DEHD, 300theorem DED, 300

large eigenspacestheorem DMLE, 298

notexample SD.NDMS4, 300

diagonalizationArchetype B

example SD.DAB, 295criteria

theorem DC, 295example SD.DMS3, 297

DIM (definition), 295dimension

definition D, 230polynomial subspace

example D.DSP4, 236DLDS (theorem), 100DM (definition), 254DM (section), 254DM (theorem), 235

Page 489: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 457

DM.D33M (example), 255DM.DUTM (example), 258DM.MC (example), 256DM.SS (example), 254DM.TCSD (example), 257DM.ZNDAB (example), 259DMLE (theorem), 298DMST (theorem), 255DP (theorem), 235DRMM (theorem), 259DT (theorem), 258DVS (subsection of D), 235DZM (definition), 295

E(archetype), 397

E (chapter), 261E (technique), 40ECEE (subsection of EE), 272EDELI (theorem), 280EE (section), 261EE.CAEHW (example), 266EE.CEMS6 (example), 275EE.CPMS3 (example), 269EE.DEMS5 (example), 278EE.EMMS4 (example), 272EE.EMS3 (example), 270EE.ESMS3 (example), 271EE.ESMS4 (example), 273EE.HMEM5 (example), 274EE.PM (example), 263EE.SEE (example), 261EEE (subsection of EE), 265EEM (definition), 261EEM (subsection of EE), 261EHM (subsection of PEE), 288eigenspace

as null spacetheorem EMNS, 271

definition EM, 270subspace

theorem EMS, 270eigenvalue

algebraic multiplicitydefinition AME, 272

complexexample EE.CEMS6, 275

definition EEM, 261

existenceexample EE.CAEHW, 266theorem EMHE, 265

geometric multiplicitydefinition GME, 272

multiplicitiesexample EE.EMMS4, 272

powertheorem EOMP, 282

root of characteristic polynomialtheorem EMRCP, 269

scalar multipletheorem ESMM, 282

symmetric matrixexample EE.ESMS4, 273

zerotheorem SMZE, 281

eigenvaluesbuilding desired

example PEE.BDE, 283distinct

example EE.DEMS5, 278example EE.SEE, 261Hermitian matrices

theorem HMRE, 288inverse

theorem EIM, 284maximum number

theorem MNEM, 288multiplicities

example EE.HMEM5, 274theorem ME, 286

numbertheorem NEM, 286

of a polynomialtheorem EPM, 283

size 3 matrixexample EE.EMS3, 270example EE.ESMS3, 271

transposetheorem ETM, 285

eigenvector, 261eigenvectors, 262

Hermitian matricestheorem HMOE, 289

linearly independenttheorem EDELI, 280

EILT (subsection of ILT), 321

Page 490: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 458

EIM (theorem), 284ELIS (theorem), 243EM (definition), 270EMHE (theorem), 265EMNS (theorem), 271EMP (theorem), 154EMRCP (theorem), 269EMS (theorem), 270EO (definition), 12EOMP (theorem), 282EOPSS (theorem), 13EPM (theorem), 283equation operations

definition EO, 12theorem EOPSS, 13

equivalencetechnique E, 40

equivalent systemsdefinition ES, 12

ERMCP (theorem), 285ES (definition), 12ESEO (subsection of SSSLE), 12ESLT (subsection of SLT), 332ESMM (theorem), 282ETM (theorem), 285EVS (subsection of VS), 185example

B.AVR, 226B.BM, 221B.BP, 221B.BSM22, 222B.BSP4, 221B.CABAK, 225B.LIM32, 214B.LIP4, 212B.RS, 224B.RSB, 223B.SSM22, 218B.SSP4, 216CNO.ACN, 447CNO.CSCN, 448CNO.MSCN, 449D.DSM22, 235D.DSP4, 236D.LDP4, 234D.RNM, 237D.RNSM, 238D.VSPUD, 236

DM.D33M, 255DM.DUTM, 258DM.MC, 256DM.SS, 254DM.TCSD, 257DM.ZNDAB, 259EE.CAEHW, 266EE.CEMS6, 275EE.CPMS3, 269EE.DEMS5, 278EE.EMMS4, 272EE.EMS3, 270EE.ESMS3, 271EE.ESMS4, 273EE.HMEM5, 274EE.PM, 263EE.SEE, 261HSE.AHSAC, 49HSE.HISAA, 50HSE.HISAD, 51HSE.HUSAB, 50HSE.NSEAI, 55HSE.NSLE, 54ILT.IAP, 329ILT.IAR, 322ILT.IAV, 323ILT.NIAO, 329ILT.NIAQ, 321ILT.NIAQR, 328ILT.NIDAU, 331ILT.NSAO, 325ILT.TNSAP, 326IVLT.AIVLT, 344IVLT.ANILT, 345IVLT.IVSAV, 351LC.AALC, 75LC.ABLC, 74LC.TLC, 72LI.LDCAA, 103LI.LDS, 96LI.LICAB, 103LI.LIS, 97LI.LLDS, 99LI.NSLIL, 105LI.RS, 101LT.ALT, 303LT.CTLT, 319LT.LTDB1, 312

Page 491: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 459

LT.LTDB2, 313LT.LTDB3, 313LT.LTM, 307LT.LTPM, 305LT.LTPP, 306LT.MFLT, 308LT.MOLT, 310LT.NLT, 304LT.SMLT, 318LT.SPIAS, 314LT.STLT, 317MINSM.OM3, 178MINSM.OPM, 179MINSM.OSMC, 180MISLE.CMIAB, 173MISLE.CMIAK, 170MISLE.MIAK, 168MISLE.MWIAA, 167MISLE.SABMI, 166MM.MMNC, 154MM.MTV, 151MM.NSLE, 152MM.PSNS, 161MM.PTM, 153MM.PTMEE, 155MO.MA, 122MO.MSM, 122MO.SYM, 125MO.TM, 124NSM.IM, 57NSM.NS, 57NSM.NSNS, 59NSM.NSRR, 58NSM.NSS, 58NSM.S, 56NSM.SRR, 58O.AOS, 115O.CNSV, 113O.CSIP, 110O.GSTV, 118O.ONFV, 120O.ONTV, 119O.SUVOS, 115O.TOV, 114PD.BP, 245PD.BSM22, 245PD.CROB3, 250PD.CROB4, 250

PD.RRTI, 248PD.SVP4, 246PEE.BDE, 283RM.COC, 128RM.RAA, 139RM.RAB, 139RM.RMCS, 127RM.RNSAD, 134RM.RNSAG, 137RM.ROCD, 133RREF.AM, 22RREF.AMAA, 24RREF.NRREF, 27RREF.RREF, 27RREF.SAA, 30RREF.SAB, 29RREF.SAE, 31RREF.TREM, 25RREF.US, 26RSM.IS, 145RSM.RROI, 146RSM.RSAI, 141RSM.RSREM, 144S.LCM, 202S.NSC2A, 200S.NSC2S, 200S.NSC2Z, 199S.RSNS, 201S.SC3, 196S.SM32, 205S.SP4, 198S.SSP, 204SD.DAB, 295SD.DEHD, 300SD.DMS3, 297SD.EENS, 294SD.NDMS4, 300SD.SMS3, 292SD.SMS5, 291SLT.FRAN, 338SLT.NSAO, 340SLT.NSAQ, 332SLT.NSAQR, 340SLT.NSDAT, 343SLT.RAO, 336SLT.SAN, 341SLT.SAR, 333SLT.SAV, 334

Page 492: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 460

SS.SCAA, 87SS.SCAB, 89SS.SCAD, 91SS.VFSAD, 80SS.VFSAI, 83SS.VFSAL, 85SSSLE.IS, 17SSSLE.NSE, 10SSSLE.STNE, 9SSSLE.TTS, 11SSSLE.US, 16TSS.CFV, 43TSS.ISSI, 38TSS.OSGMD, 44TSS.RREFN, 37VO.VA, 67VO.VESE, 66VO.VSM, 68VR.ASC, 365VR.CM32, 367VR.CP2, 366VR.CVSR, 365VR.MIVS, 365VR.TIVS, 365VR.VRC4, 360VR.VRP2, 361VS.CVS, 188VS.PCVS, 193VS.VSCM, 186VS.VSF, 187VS.VSIS, 187VS.VSM, 186VS.VSP, 186VS.VSS, 188WILA.TM, 3

EXC (subsection of B), 228EXC (subsection of D), 241EXC (subsection of LC), 78EXC (subsection of LI), 106EXC (subsection of MINSM), 182EXC (subsection of MM), 164EXC (subsection of NSM), 63EXC (subsection of PD), 252EXC (subsection of RREF), 34EXC (subsection of RSM), 148EXC (subsection of S), 209EXC (subsection of SS), 93EXC (subsection of SSSLE), 21

EXC (subsection of TSS), 47EXC (subsection of WILA), 8

F(archetype), 401

free variablesexample TSS.CFV, 43

free variables, numbercorollary FVCS, 42

FTMR (theorem), 369FVCS (corollary), 42

G(archetype), 406

G (theorem), 244getting started

technique GS, 13GME (definition), 272goldilocks

theorem G, 244Gram-Schmidt

column vectorstheorem GSPCV, 117

three vectorsexample O.GSTV, 118

GS (technique), 13GSP (subsection of O), 116GSPCV (theorem), 117GT (subsection of PD), 243

H(archetype), 410

hermitiandefinition HM, 181

HM (definition), 181HMOE (theorem), 289HMRE (theorem), 288HMVEI (theorem), 51homogeneous system

consistenttheorem HSC, 50

definition HS, 49infinitely many solutions

theorem HMVEI, 51homogeneous systems

linear independence, 98homogenous system

Archetype Cexample HSE.AHSAC, 49

Page 493: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 461

HS (definition), 49HSC (theorem), 50HSE (section), 49HSE.AHSAC (example), 49HSE.HISAA (example), 50HSE.HISAD (example), 51HSE.HUSAB (example), 50HSE.NSEAI (example), 55HSE.NSLE (example), 54

I(archetype), 414

ICLT (theorem), 349ICRN (theorem), 42identity matrix

example NSM.IM, 57IDLT (definition), 344IDV (definition), 40IFDVS (theorem), 365IILT (theorem), 347ILT (definition), 321ILT (section), 321ILT.IAP (example), 329ILT.IAR (example), 322ILT.IAV (example), 323ILT.NIAO (example), 329ILT.NIAQ (example), 321ILT.NIAQR (example), 328ILT.NIDAU (example), 331ILT.NSAO (example), 325ILT.TNSAP (example), 326ILTB (theorem), 330ILTD (subsection of ILT), 330ILTD (theorem), 330ILTIS (theorem), 348ILTLI (theorem), 329ILTLT (theorem), 347IM (definition), 57IM (subsection of MISLE), 167inconsistent linear systems

theorem ICRN, 42independent, dependent variables

definition IDV, 40indexstring

example B.LIM32, 214example D.DSM22, 235example LT.STLT, 317theorem ERMCP, 285

theorem MMCC, 159theorem SMEE, 294

infinite solution setexample TSS.ISSI, 38

infinite solutions, 3× 4example SSSLE.IS, 17

injectiveexample ILT.IAP, 329example ILT.IAR, 322not

example ILT.NIAO, 329example ILT.NIAQ, 321example ILT.NIAQR, 328

not, by dimensionexample ILT.NIDAU, 331

polynomials to matricesexample ILT.IAV, 323

injective linear transformationbases

theorem ILTB, 330injective linear transformations

dimensiontheorem ILTD, 330

inner productanti-commutative

theorem IPAC, 112example O.CSIP, 110norm

theorem IPN, 113positive

theorem PIP, 114scalar multiplication

theorem IPSM, 112vector addition

theorem IPVA, 111INS (theorem), 372inverse

composition of linear transformationstheorem ICLT, 349

invertible linear transformationscomposition

theorem CIVLT, 349IP (definition), 110IP (subsection of O), 110IPAC (theorem), 112IPN (theorem), 113IPSM (theorem), 112IPVA (theorem), 111

Page 494: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 462

IR (theorem), 373isomorphic

multiple vector spacesexample VR.MIVS, 365

vector spacesexample IVLT.IVSAV, 351

isomorphic vector spacesdimension

theorem IVSED, 351example VR.TIVS, 365

IV (subsection of IVLT), 348IVLT (definition), 344IVLT (section), 344IVLT (subsection of IVLT), 344IVLT.AIVLT (example), 344IVLT.ANILT (example), 345IVLT.IVSAV (example), 351IVS (definition), 350IVSED (theorem), 351

J(archetype), 418

K(archetype), 423

L(archetype), 427

L (technique), 19LA (subsection of WILA), 2language

technique L, 19LC (definition), 202LC (section), 72LC (subsection of LC), 72LC.AALC (example), 75LC.ABLC (example), 74LC.TLC (example), 72LCCV (definition), 72LDSS (subsection of LI), 100leading ones

definition LO, 27LI (definition), 212LI (section), 96LI (subsection of B), 212LI.LDCAA (example), 103LI.LDS (example), 96LI.LICAB (example), 103LI.LIS (example), 97

LI.LLDS (example), 99LI.NSLIL (example), 105LI.RS (example), 101LICV (definition), 96linear combination

system of equationsexample LC.ABLC, 74

definition LC, 202definition LCCV, 72example LC.TLC, 72linear transformation, 311matrices

example S.LCM, 202system of equations

example LC.AALC, 75linear combinations

solutions to linear systemstheorem SLSLC, 76

linear dependencemore vectors than size

theorem MVSLD, 99linear independence

definition LI, 212definition LICV, 96homogeneous systems

theorem LIVHS, 98injective linear transformation

theorem ILTLI, 329orthogonal, 116r and n

corollary LIVRN, 100linear solve

mathematica, 45linear system

consistenttheorem RCLS, 41

notation LSN, 54linear systems

notationexample HSE.NSLE, 54example MM.NSLE, 152

linear transfgormationnot invertible

example IVLT.ANILT, 345linear transformation

polynomials to polynomialsexample LT.LTPP, 306

addition

Page 495: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 463

definition LTA, 316theorem MLTLT, 317theorem SLTLT, 316

checkingexample LT.ALT, 303

compositiondefinition LTC, 319theorem CLTLT, 319

defined by a matrixexample LT.LTM, 307

defined on a basisexample LT.LTDB1, 312example LT.LTDB2, 313example LT.LTDB3, 313theorem LTDB, 312

definition LT, 302identity

definition IDLT, 344injection

definition ILT, 321inverse

theorem ILTLT, 347inverse of inverse

theorem IILT, 347invertible

definition IVLT, 344example IVLT.AIVLT, 344

invertible, injective and surjectivetheorem ILTIS, 348

linear combinationtheorem LTLC, 311

matrix of, 309example LT.MFLT, 308example LT.MOLT, 310

notexample LT.NLT, 304

polynomials to matricesexample LT.LTPM, 305

rank plus nullitytheorem RPNDD, 353

scalar multipleexample LT.SMLT, 318

scalar multiplicationdefinition LTSM, 317

surjectiondefinition SLT, 332

vector space of, 318zero vector

theorem LTTZZ, 306linear transformations

compositionsexample LT.CTLT, 319

from matricestheorem MBLT, 308

linearly dependent columnsArchetype A

example LI.LDCAA, 103linearly dependent set

example LI.LDS, 96linear combinations within

theorem DLDS, 100polynomials

example D.LDP4, 234linearly independent

extending setstheorem ELIS, 243

polynomialsexample B.LIP4, 212

linearly independent columnsArchetype B

example LI.LICAB, 103linearly independent set

example LI.LIS, 97example LI.LLDS, 99

LINSM (subsection of LI), 102LIV (subsection of LI), 96LIVHS (theorem), 98LIVRN (corollary), 100LO (definition), 27LS.MMA (computation), 45LSN (notation), 54LT (chapter), 302LT (definition), 302LT (section), 302LT (subsection of LT), 302LT.ALT (example), 303LT.CTLT (example), 319LT.LTDB1 (example), 312LT.LTDB2 (example), 313LT.LTDB3 (example), 313LT.LTM (example), 307LT.LTPM (example), 305LT.LTPP (example), 306LT.MFLT (example), 308LT.MOLT (example), 310LT.NLT (example), 304

Page 496: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 464

LT.SMLT (example), 318LT.SPIAS (example), 314LT.STLT (example), 317LTA (definition), 316LTC (definition), 319LTDB (theorem), 312LTLC (subsection of LT), 311LTLC (theorem), 311LTSM (definition), 317LTTZZ (theorem), 306

M(archetype), 431

M (chapter), 121M (definition), 22MA (definition), 122mathematica

linear solvecomputation LS.MMA, 45

matrix entrycomputation ME.MMA, 22

matrix inversescomputation MI.MMA, 173

matrix multiplicationcomputation MM.MMA, 154

row reducecomputation RR.MMA, 32

vector linear combinationscomputation VLC.MMA, 69

matrixaddition

definition MA, 122augmented

definition AM, 23cofactor

definition CIM, 256definition M, 22equality

definition ME, 121example RREF.AM, 22identity

definition IM, 57minor

definition MIM, 256minors, cofactors

example DM.MC, 256nonsingular

definition NM, 56

of a linear transformationtheorem MLTCV, 309

orthogonaldefinition OM, 178

orthogonal is invertibletheorem OMI, 179

productexample MM.PTM, 153example MM.PTMEE, 155

rectangular, 56scalar multiplication

definition SMM, 122singular, 56square

definition SQM, 56submatrices

example DM.SS, 254submatrix

definition SM, 254symmetric

definition SYM, 125transpose

definition TM, 124zero

definition ZM, 124zero row

definition ZRM, 27matrix addition

example MO.MA, 122matrix entries

notation ME, 123matrix entry

mathematica, 22ti83, 23ti86, 23

matrix inverseArchetype B, 173Archetype K, 168, 170computation

theorem CINSM, 172nonsingular matrix

theorem NSI, 177of a matrix inverse

theorem MIMI, 174one-sided

theorem OSIS, 177product

theorem SS, 174

Page 497: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 465

size 2 matricestheorem TTMI, 169

transposetheorem MISM, 175theorem MIT, 174

uniquenesstheorem MIU, 173

matrix inversesmathematica, 173

matrix multiplicationassociativity

theorem MMA, 158distributivity

theorem MMDAA, 157entry-by-entry

theorem EMP, 154identity matrix

theorem MMIM, 156inner product

theorem MMIP, 159mathematica, 154scalar matrix multiplication

theorem MMSMM, 157systems of linear equations

theorem SLEMM, 152transposes

theorem MMT, 160zero matrix

theorem MMZM, 156matrix mutiplication

noncommutativeexample MM.MMNC, 154

matrix notationdefinition MN, 22

matrix representationcomposition of linear transformations

theorem MRCLT, 371definition MR, 369multiple of a linear transformation

theorem MRMLT, 371sum of linear transformations

theorem MRSLT, 370theorem FTMR, 369

matrix scalar multiplicationexample MO.MSM, 122

matrix vector spacedimension

theorem DM, 235

matrix-vector productexample MM.MTV, 151

matrix:inversedefinition MI, 167

matrix:multiplicationdefinition MM, 153

matrix:product with vectordefinition MVP, 151

matrix:rangedefinition RM, 127

matrix:row spacedefinition RSM, 141

MBLT (theorem), 308MCC (subsection of MO), 126MCN (definition), 448MCN (subsection of CNO), 448ME (definition), 121ME (notation), 123ME (subsection of PEE), 286ME (technique), 62ME (theorem), 286ME.MMA (computation), 22ME.TI83 (computation), 23ME.TI86 (computation), 23MEASM (subsection of MO), 121MI (definition), 167MI.MMA (computation), 173MIM (definition), 256MIMI (theorem), 174MINSM (section), 176MINSM.OM3 (example), 178MINSM.OPM (example), 179MINSM.OSMC (example), 180MISLE (section), 166MISLE.CMIAB (example), 173MISLE.CMIAK (example), 170MISLE.MIAK (example), 168MISLE.MWIAA (example), 167MISLE.SABMI (example), 166MISM (theorem), 175MIT (theorem), 174MIU (theorem), 173MLT (subsection of LT), 307MLTCV (theorem), 309MLTLT (theorem), 317MM (definition), 153MM (section), 151MM (subsection of MM), 153

Page 498: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 466

MM.MMA (computation), 154MM.MMNC (example), 154MM.MTV (example), 151MM.NSLE (example), 152MM.PSNS (example), 161MM.PTM (example), 153MM.PTMEE (example), 155MMA (theorem), 158MMCC (theorem), 159MMDAA (theorem), 157MMEE (subsection of MM), 154MMIM (theorem), 156MMIP (theorem), 159MMSMM (theorem), 157MMT (theorem), 160MMZM (theorem), 156MN (definition), 22MNEM (theorem), 288MO (section), 121MO.MA (example), 122MO.MSM (example), 122MO.SYM (example), 125MO.TM (example), 124more variables than equations

example TSS.OSGMD, 44theorem CMVEI, 44

MR (definition), 369MR (section), 369MRCLT (theorem), 371MRMLT (theorem), 371MRSLT (theorem), 370multiple equivalences

technique ME, 62MVNSE (subsection of HSE), 52MVP (definition), 151MVP (subsection of MM), 151MVSLD (theorem), 99

N(archetype), 433

N (subsection of O), 112NEM (theorem), 286NLTFO (subsection of LT), 316NM (definition), 56NOILT (theorem), 353NOLT (definition), 352NOM (definition), 237nonsingular

columns as basistheorem CNSMB, 225

nonsingular matriceslinearly independent columns

theorem NSLIC, 103nonsingular matrix

Archetype Bexample NSM.NS, 57

equivalencestheorem NSME1, 62theorem NSME2, 103theorem NSME3, 140theorem NSME4, 178theorem NSME5, 226theorem NSME6, 239theorem NSME7, 259theorem NSME8, 281

matrix inverse, 177null space

example NSM.NSNS, 59nullity, 239range, 139rank

theorem RNNSM, 239row-reduced

theorem NSRRI, 57trivial null space

theorem NSTNS, 59unique solutions

theorem NSMUS, 59nonsingular matrix, row-reduced

example NSM.NSRR, 58norm

example O.CNSV, 113inner product, 113

notationAMN, 54LSN, 54ME, 123RREFA, 37VN, 52ZVN, 52

notation for a linear systemexample SSSLE.NSE, 10

NRFO (subsection of MR), 370NSI (theorem), 177NSILT (theorem), 328NSLIC (theorem), 103

Page 499: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 467

NSLT (definition), 325NSLT (subsection of ILT), 324NSLTS (theorem), 326NSM (definition), 54NSM (section), 56NSM (subsection of HSE), 54NSM (subsection of NSM), 56NSM.IM (example), 57NSM.NS (example), 57NSM.NSNS (example), 59NSM.NSRR (example), 58NSM.NSS (example), 58NSM.S (example), 56NSM.SRR (example), 58NSME1 (theorem), 62NSME2 (theorem), 103NSME3 (theorem), 140NSME4 (theorem), 178NSME5 (theorem), 226NSME6 (theorem), 239NSME7 (theorem), 259NSME8 (theorem), 281NSMI (subsection of MINSM), 176NSMS (theorem), 200NSMUS (theorem), 59NSPI (theorem), 327NSRRI (theorem), 57NSSLI (subsection of LI), 104NSTNS (theorem), 59null space

Archetype Iexample HSE.NSEAI, 55

basistheorem BNS, 104

injective linear transformationtheorem NSILT, 328

linear transformationexample ILT.NSAO, 325

matrixdefinition NSM, 54

nonsingular matrix, 59of a linear transformation

definition NSLT, 325singular matrix, 58spanning set

theorem SSNS, 90subspace

theorem NSLTS, 326

theorem NSMS, 200trivial

example ILT.TNSAP, 326null space span, linearly independent

Archetype Lexample LI.NSLIL, 105

null spacesisomorphic, transformation, represen-

tationtheorem INS, 372

nullitycomputing, 238injective linear transformation

theorem NOILT, 353linear transformation

definition NOLT, 352matrix, 237

definition NOM, 237square matrix, 238

NV (definition), 113

O(archetype), 435

O (section), 109O.AOS (example), 115O.CNSV (example), 113O.CSIP (example), 110O.GSTV (example), 118O.ONFV (example), 120O.ONTV (example), 119O.SUVOS (example), 115O.TOV (example), 114OBC (subsection of PD), 248OD (subsection of SD), 301ODHM (theorem), 301OM (definition), 178OM (subsection of MINSM), 178OMI (theorem), 179OMPIP (theorem), 180ONS (definition), 119orthogonal

linear independencetheorem OSLI, 116

permutation matrixexample MINSM.OPM, 179

setexample O.AOS, 115

set of vectors

Page 500: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 468

definition OSV, 115size 3

example MINSM.OM3, 178vector pairs

definition OV, 114orthogonal matrices

columnstheorem COMOS, 179

orthogonal matrixinner product

theorem OMPIP, 180orthogonal vectors

example O.TOV, 114orthonormal

definition ONS, 119matrix columns

example MINSM.OSMC, 180orthonormal diagonalization

theorem ODHM, 301orthonormal set

four vectorsexample O.ONFV, 120

three vectorsexample O.ONTV, 119

OSIS (theorem), 177OSLI (theorem), 116OSV (definition), 115OV (definition), 114OV (subsection of O), 114

P(archetype), 437

P (chapter), 447P (technique), 125particular solutions

example MM.PSNS, 161PD (section), 243PD (subsection of DM), 258PD.BP (example), 245PD.BSM22 (example), 245PD.CROB3 (example), 250PD.CROB4 (example), 250PD.RRTI (example), 248PD.SVP4 (example), 246PEE (section), 280PEE.BDE (example), 283PI (definition), 314PI (subsection of LT), 314

PIP (theorem), 114PM (subsection of EE), 263PMI (subsection of MISLE), 173PMM (subsection of MM), 156PMR (subsection of MR), 372polynomial

of a matrixexample EE.PM, 263

polynomial vector spacedimension

theorem DP, 235practice

technique P, 125pre-image

definition PI, 314pre-images

example LT.SPIAS, 314null space

theorem NSPI, 327PSHS (subsection of MM), 161PSM (subsection of SD), 292PSPHS (theorem), 161PSS (subsection of SSSLE), 11PSSLS (theorem), 44PWSMS (theorem), 176

Q(archetype), 439

R(archetype), 442

R (chapter), 358range

Archetype Aexample RM.RAA, 139

Archetype Bexample RM.RAB, 139

as null spacetheorem RNS, 136

as null space, Archetype Dexample RM.RNSAD, 134

as null space, Archetype Gexample RM.RNSAG, 137

as row spacetheorem RMRST, 145

basis of original columnstheorem BROC, 131

consistent systemsexample RM.RMCS, 127

Page 501: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 469

theorem RCS, 128full

example SLT.FRAN, 338linear transformation

example SLT.RAO, 336matrix, 127nonsingular matrix

theorem RNSM, 139of a linear transformation

definition RLT, 336original columns, Archetype D

example RM.ROCD, 133pre-image

theorem RPI, 341removing columns

example RM.COC, 128row operations, Archetype I

example RSM.RROI, 146subspace

theorem RLTS, 338theorem RMS, 207

surjective linear transformationtheorem RSLT, 339

rangesisomorphic, transformation, represen-

tationtheorem IR, 373

rankcomputing

theorem CRN, 238linear transformation

definition ROLT, 352matrix

definition ROM, 237example D.RNM, 237

of transposeexample PD.RRTI, 248

square matrixexample D.RNSM, 238

surjective linear transformationtheorem ROSLT, 352

transposetheorem RMRT, 247

rank+nullitytheorem RPNC, 238

RCLS (theorem), 41RCS (theorem), 128RD (subsection of VS), 195

reduced row-echelon formanalysis

notation RREFA, 37definition RREF, 27example RREF.NRREF, 27example RREF.RREF, 27notation

example TSS.RREFN, 37reducing a span

example LI.RS, 101relation of linear dependence

definition RLD, 212definition RLDCV, 96

REM (definition), 25REMEF (theorem), 28REMES (theorem), 25REMRS (theorem), 142RLD (definition), 212RLDCV (definition), 96RLT (definition), 336RLT (subsection of SLT), 336RLTS (theorem), 338RM (definition), 127RM (section), 127RM.COC (example), 128RM.RAA (example), 139RM.RAB (example), 139RM.RMCS (example), 127RM.RNSAD (example), 134RM.RNSAG (example), 137RM.ROCD (example), 133RMRST (theorem), 145RMRT (theorem), 247RMS (theorem), 207RNLT (subsection of IVLT), 352RNM (subsection of D), 237RNNSM (subsection of D), 238RNNSM (theorem), 239RNS (subsection of RM), 134RNS (theorem), 136RNSM (subsection of RM), 138RNSM (theorem), 139RO (definition), 24ROLT (definition), 352ROM (definition), 237ROSLT (theorem), 352row operations

definition RO, 24

Page 502: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 470

row reducemathematica, 32ti83, 33ti86, 33

row spaceArchetype I

example RSM.RSAI, 141basis

example B.RSB, 223theorem BRS, 144

matrix, 141row-equivalent matrices

theorem REMRS, 142subspace

theorem RSMS, 208row-equivalent matrices

definition REM, 25example RREF.TREM, 25row space, 142row spaces

example RSM.RSREM, 144theorem REMES, 25

row-reduced matricestheorem REMEF, 28

RPI (theorem), 341RPNC (theorem), 238RPNDD (theorem), 353RR.MMA (computation), 32RR.TI83 (computation), 33RR.TI86 (computation), 33RREF (definition), 27RREF (section), 22RREF.AM (example), 22RREF.AMAA (example), 24RREF.NRREF (example), 27RREF.RREF (example), 27RREF.SAA (example), 30RREF.SAB (example), 29RREF.SAE (example), 31RREF.TREM (example), 25RREF.US (example), 26RREFA (notation), 37RSE (subsection of RM), 127RSLT (theorem), 339RSM (definition), 141RSM (section), 141RSM (subsection of RSM), 141RSM.IS (example), 145

RSM.RROI (example), 146RSM.RSAI (example), 141RSM.RSREM (example), 144RSMS (theorem), 208RSOC (subsection of RM), 128RT (subsection of PD), 246

S(archetype), 444

S (definition), 196S (section), 196S.LCM (example), 202S.NSC2A (example), 200S.NSC2S (example), 200S.NSC2Z (example), 199S.RSNS (example), 201S.SC3 (example), 196S.SM32 (example), 205S.SP4 (example), 198S.SSP (example), 204SC (subsection of S), 207scalar multiplication

canceling scalarstheorem CSSM, 194

canceling vectorstheorem CVSM, 194

zero scalartheorem ZSSM, 191

zero vectortheorem ZVSM, 192

zero vector resulttheorem SMEZV, 193

SD (section), 291SD.DAB (example), 295SD.DEHD (example), 300SD.DMS3 (example), 297SD.EENS (example), 294SD.NDMS4 (example), 300SD.SMS3 (example), 292SD.SMS5 (example), 291SE (technique), 14SER (theorem), 293set equality

technique SE, 14set notation

technique SN, 39shoes, 174SHS (subsection of HSE), 49

Page 503: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 471

SI (subsection of IVLT), 350SIM (definition), 291similar matrices

equal eigenvaluesexample SD.EENS, 294

example SD.SMS3, 292example SD.SMS5, 291

similaritydefinition SIM, 291equivalence relation

theorem SER, 293singular matrix

Archetype Aexample NSM.S, 56

null spaceexample NSM.NSS, 58

product withtheorem PWSMS, 176

singular matrix, row-reducedexample NSM.SRR, 58

SLE (chapter), 2SLELT (subsection of IVLT), 355SLEMM (theorem), 152SLSLC (theorem), 76SLT (definition), 332SLT (section), 332SLT.FRAN (example), 338SLT.NSAO (example), 340SLT.NSAQ (example), 332SLT.NSAQR (example), 340SLT.NSDAT (example), 343SLT.RAO (example), 336SLT.SAN (example), 341SLT.SAR (example), 333SLT.SAV (example), 334SLTB (theorem), 342SLTD (subsection of SLT), 342SLTD (theorem), 342SLTLT (theorem), 316SLTS (theorem), 341SM (definition), 254SM (subsection of SD), 291SMEE (theorem), 294SMEZV (theorem), 193SMM (definition), 122SMS (theorem), 126SMZD (theorem), 259SMZE (theorem), 281

SN (technique), 39SNSCM (theorem), 178socks, 174SOL (subsection of B), 229SOL (subsection of D), 242SOL (subsection of LC), 79SOL (subsection of LI), 107SOL (subsection of MINSM), 183SOL (subsection of MM), 165SOL (subsection of NSM), 64SOL (subsection of PD), 253SOL (subsection of RREF), 35SOL (subsection of RSM), 149SOL (subsection of S), 210SOL (subsection of SS), 94SOL (subsection of TSS), 48solution set

theorem PSPHS, 161solution sets

possibilitiestheorem PSSLS, 44

solution vectordefinition SV, 53

solving homogeneous systemArchetype A

example HSE.HISAA, 50Archetype B

example HSE.HUSAB, 50Archetype D

example HSE.HISAD, 51solving nonlinear equations

example SSSLE.STNE, 9span

definition SS, 203improved

example RSM.IS, 145reduction

example B.RS, 224set of polynomials

example S.SSP, 204subspace

theorem SSS, 203span of columns

Archetype Aexample SS.SCAA, 87

Archetype Bexample SS.SCAB, 89

Archetype D

Page 504: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 472

example SS.SCAD, 91span:definition

definition SSCV, 87spanning set

definition TSS, 216matrices

example B.SSM22, 218more vectors

theorem SSLD, 230polynomials

example B.SSP4, 216SQM (definition), 56SS (definition), 203SS (section), 80SS (subsection of B), 216SS (theorem), 174SS.SCAA (example), 87SS.SCAB (example), 89SS.SCAD (example), 91SS.VFSAD (example), 80SS.VFSAI (example), 83SS.VFSAL (example), 85SSCV (definition), 87SSLD (theorem), 230SSLE (definition), 10SSNS (subsection of SS), 90SSNS (theorem), 90SSS (theorem), 203SSSLE (section), 9SSSLE.IS (example), 17SSSLE.NSE (example), 10SSSLE.STNE (example), 9SSSLE.TTS (example), 11SSSLE.US (example), 16SSV (subsection of SS), 86subspace

as null spaceexample S.RSNS, 201

characterizedexample VR.ASC, 365

definition S, 196in P4

example S.SP4, 198not, additive closure

example S.NSC2A, 200not, scalar closure

example S.NSC2S, 200not, zero vector

example S.NSC2Z, 199testing

theorem TSS, 198trivial

definition TS, 200verification

example S.SC3, 196example S.SM32, 205

surjectiveArchetype N

example SLT.SAN, 341example SLT.SAR, 333not

example SLT.NSAQ, 332example SLT.NSAQR, 340

not, Archetype Oexample SLT.NSAO, 340

not, by dimensionexample SLT.NSDAT, 343

polynomials to matricesexample SLT.SAV, 334

surjective linear transformationbases

theorem SLTB, 342span

theorem SLTS, 341surjective linear transformations

dimensiontheorem SLTD, 342

SUV (definition), 169SUVB (theorem), 220SV (definition), 53SYM (definition), 125symmetric matrices

theorem SMS, 126symmetric matrix

example MO.SYM, 125system of equations

vector equalityexample VO.VESE, 66

system of linear equationsdefinition SSLE, 10

T(archetype), 444

T (part), 447T (technique), 13TASM (theorem), 126

Page 505: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 473

techniqueC, 28CP, 41CV, 43D, 9DC, 74E, 40GS, 13L, 19ME, 62P, 125SE, 14SN, 39T, 13U, 59

theoremAISM, 192AIU, 191BIS, 234BNS, 104BROC, 131BRS, 144CCRA, 448CCRM, 448CCRSM, 110CCRVA, 109CFDVS, 364CILTI, 331CINSM, 172CIVLT, 349CLI, 365CLTLT, 319CMVEI, 44CNSMB, 225COB, 249COMOS, 179CRN, 238CSLTS, 343CSRN, 42CSS, 366CSSM, 194CVSM, 194DC, 295DCM, 235DCP, 286DED, 300DERC, 257DLDS, 100

DM, 235DMLE, 298DMST, 255DP, 235DRMM, 259DT, 258EDELI, 280EIM, 284ELIS, 243EMHE, 265EMNS, 271EMP, 154EMRCP, 269EMS, 270EOMP, 282EOPSS, 13EPM, 283ERMCP, 285ESMM, 282ETM, 285FTMR, 369G, 244GSPCV, 117HMOE, 289HMRE, 288HMVEI, 51HSC, 50ICLT, 349ICRN, 42IFDVS, 365IILT, 347ILTB, 330ILTD, 330ILTIS, 348ILTLI, 329ILTLT, 347INS, 372IPAC, 112IPN, 113IPSM, 112IPVA, 111IR, 373IVSED, 351LIVHS, 98LTDB, 312LTLC, 311LTTZZ, 306MBLT, 308

Page 506: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 474

ME, 286MIMI, 174MISM, 175MIT, 174MIU, 173MLTCV, 309MLTLT, 317MMA, 158MMCC, 159MMDAA, 157MMIM, 156MMIP, 159MMSMM, 157MMT, 160MMZM, 156MNEM, 288MRCLT, 371MRMLT, 371MRSLT, 370MVSLD, 99NEM, 286NOILT, 353NSI, 177NSILT, 328NSLIC, 103NSLTS, 326NSME1, 62NSME2, 103NSME3, 140NSME4, 178NSME5, 226NSME6, 239NSME7, 259NSME8, 281NSMS, 200NSMUS, 59NSPI, 327NSRRI, 57NSTNS, 59ODHM, 301OMI, 179OMPIP, 180OSIS, 177OSLI, 116PIP, 114PSPHS, 161PSSLS, 44PWSMS, 176

RCLS, 41RCS, 128REMEF, 28REMES, 25REMRS, 142RLTS, 338RMRST, 145RMRT, 247RMS, 207RNNSM, 239RNS, 136RNSM, 139ROSLT, 352RPI, 341RPNC, 238RPNDD, 353RSLT, 339RSMS, 208SER, 293SLEMM, 152SLSLC, 76SLTB, 342SLTD, 342SLTLT, 316SLTS, 341SMEE, 294SMEZV, 193SMS, 126SMZD, 259SMZE, 281SNSCM, 178SS, 174SSLD, 230SSNS, 90SSS, 203SUVB, 220TASM, 126TSS, 198TTMI, 169VAC, 194VFSLS, 82VRI, 363VRILT, 364VRLT, 358VRRB, 227VRS, 363VSLT, 318VSPCM, 70

Page 507: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 475

VSPM, 123ZSSM, 191ZVSM, 192ZVU, 191

theoremstechnique T, 13

ti83matrix entry

computation ME.TI83, 23row reduce

computation RR.TI83, 33vector linear combinations

computation VLC.TI83, 70ti86

matrix entrycomputation ME.TI86, 23

row reducecomputation RR.TI86, 33

vector linear combinationscomputation VLC.TI86, 69

TM (definition), 124trail mix

example WILA.TM, 3transpose

additiontheorem TASM, 126

example MO.TM, 124matrix inverse, 174, 175scalar multiplication, 126

trivial solutionsystem of equations

definition TSHSE, 50TS (definition), 200TS (subsection of S), 197TSHSE (definition), 50TSM (subsection of MO), 124TSS (definition), 216TSS (section), 37TSS (subsection of S), 202TSS (theorem), 198TSS.CFV (example), 43TSS.ISSI (example), 38TSS.OSGMD (example), 44TSS.RREFN (example), 37TTMI (theorem), 169typical systems, 2× 2

example SSSLE.TTS, 11

U

(archetype), 444U (technique), 59unique solution, 3× 3

example RREF.US, 26example SSSLE.US, 16

uniquenesstechnique U, 59

unit vectorsbasis

theorem SUVB, 220definition SUV, 169orthogonal

example O.SUVOS, 115

V(archetype), 444

V (chapter), 65VAC (theorem), 194VEASM (subsection of VO), 66vector

additiondefinition CVA, 67

columndefinition CV, 52

equalitydefinition CVE, 66

inner productdefinition IP, 110

normdefinition NV, 113

notation VN, 52of constants

definition VOC, 53product with matrix, 151, 153scalar multiplication

definition CVSM, 68vector addition

example VO.VA, 67vector form of solutions

Archetype Dexample SS.VFSAD, 80

Archetype Iexample SS.VFSAI, 83

Archetype Lexample SS.VFSAL, 85

theorem VFSLS, 82vector linear combinations

mathematica, 69

Page 508: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 476

ti83, 70ti86, 69

vector representationexample B.AVR, 226example VR.VRC4, 360injective

theorem VRI, 363invertible

theorem VRILT, 364linear transformation

definition VR, 358theorem VRLT, 358

surjectivetheorem VRS, 363

theorem VRRB, 227vector representations

polynomialsexample VR.VRP2, 361

vector scalar multiplicationexample VO.VSM, 68

vector spacecharacterization

theorem CFDVS, 364definition VS, 184infinite dimension

example D.VSPUD, 236linear transformations

theorem VSLT, 318vector space of functions

example VS.VSF, 187vector space of infinite sequences

example VS.VSIS, 187vector space of matrices

example VS.VSM, 186vector space of matrices:definition

definition VSM, 121vector space of polynomials

example VS.VSP, 186vector space properties

matricestheorem VSPM, 123

vectorstheorem VSPCM, 70

vector space, crazyexample VS.CVS, 188

vector space, singletonexample VS.VSS, 188

vector space:definition

definition VSCM, 65vector spaces

isomorphicdefinition IVS, 350theorem IFDVS, 365

VFSLS (theorem), 82VFSS (subsection of SS), 80VLC.MMA (computation), 69VLC.TI83 (computation), 70VLC.TI86 (computation), 69VN (notation), 52VO (section), 65VO.VA (example), 67VO.VESE (example), 66VO.VSM (example), 68VOC (definition), 53VR (definition), 358VR (section), 358VR (subsection of B), 226VR.ASC (example), 365VR.CM32 (example), 367VR.CP2 (example), 366VR.CVSR (example), 365VR.MIVS (example), 365VR.TIVS (example), 365VR.VRC4 (example), 360VR.VRP2 (example), 361VRI (theorem), 363VRILT (theorem), 364VRLT (theorem), 358VRRB (theorem), 227VRS (theorem), 363VS (chapter), 184VS (definition), 184VS (section), 184VS (subsection of VS), 184VS.CVS (example), 188VS.PCVS (example), 193VS.VSCM (example), 186VS.VSF (example), 187VS.VSIS (example), 187VS.VSM (example), 186VS.VSP (example), 186VS.VSS (example), 188VSCM (definition), 65VSLT (theorem), 318VSM (definition), 121VSP (subsection of MO), 123

Page 509: A First Course in Linear Algebrabuzzard.ups.edu/linear/download/fcla-oneA4-0.22.pdf · 2004-11-19 · iii • Having taught an introductory linear algebra course eighteen times prior,

INDEX 477

VSP (subsection of VO), 70VSP (subsection of VS), 190VSPCM (theorem), 70VSPM (theorem), 123

WILA (section), 2WILA.TM (example), 3

zero vectordefinition ZV, 52notation ZVN, 52unique

theorem ZVU, 191ZM (definition), 124ZRM (definition), 27ZSSM (theorem), 191ZV (definition), 52ZVN (notation), 52ZVSM (theorem), 192ZVU (theorem), 191