-
Notifications
You must be signed in to change notification settings - Fork 25
/
ch-vol1-intro.tex
1452 lines (1272 loc) · 49.6 KB
/
ch-vol1-intro.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\chapter*{Introduction}
\addcontentsline{toc}{chapter}{Introduction}
\markboth{INTRODUCTION}{INTRODUCTION}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{About this book}
%mbxSTARTIGNORE
\VolOneIntroExtrahead
%mbxENDIGNORE
% pretext (mbx) is always the two volume set
%mbxlatex
%mbxlatex \subsection{About Volume I}
%mbxlatex
This first volume is a one semester course in basic analysis.
With the second volume, it is a year-long course.
The book started its life
as my lecture notes for teaching Math 444 at the
University of Illinois at Urbana-Champaign (UIUC) in the fall semester of 2009.
I added the metric space chapter for Math 521 at the University of
Wisconsin--Madison (UW)\@.
Volume II was added to teach Math 4143/4153 at Oklahoma State University
(OSU)\@.
A prerequisite for these courses is usually a basic proof course,
using, for example, \cite{Hammack}, \cite{GIAM}, or~\cite{DW}.
It should be possible to use the book for
both a basic course for students who do not necessarily wish to
go to graduate school (such as UIUC 444), and also as a more advanced one-semester
course that also covers topics such as metric spaces (such as UW 521).
Here are suggestions for a semester course.
A slower course such as UIUC 444:
\begin{center}
\S0.3, \S1.1--\S1.4, \S2.1--\S2.5, \S3.1--\S3.4, \S4.1--\S4.2,
\S5.1--\S5.3, \S6.1--\S6.3
\end{center}
A more rigorous course covering metric spaces that runs quite a bit faster
(e.g., UW 521):
\begin{center}
\S0.3, \S1.1--\S1.4, \S2.1--\S2.5, \S3.1--\S3.4, \S4.1--\S4.2,
\S5.1--\S5.3, \S6.1--\S6.2, \S7.1--\S7.6
\end{center}
It should also be possible to run a faster course without metric spaces
covering all sections of chapters 0 through 6. The approximate number of
lectures given in the section notes through Chapter 6 are a very rough
estimate and were designed for the slower course.
The first few chapters of the book can be used in an introductory proofs
course, as is done, for example, at Iowa State University Math 201, where
this book is used in conjunction with Hammack's \emph{Book of Proof}~\cite{Hammack}.
With volume II, one can run a year-long course that covers multivariable
topics. In this scenario, it may make sense to cover most of the first volume
in the first semester while leaving metric spaces for the beginning of the
second semester.
The structure of the beginning of volume I somewhat follows the
standard syllabus of UIUC Math 444 and, therefore, has some similarities with
Bartle and Sherbert,
\emph{Introduction to Real Analysis} \cite{BS}, which is the standard book
at UIUC.
A major difference is that we define the Riemann integral using
Darboux sums and not tagged partitions. The Darboux approach is far more
appropriate for a course of this level.
Our approach allows us to fit a course such as UIUC 444 within a semester
and still spend some time on the interchange of limits and end with
Picard's theorem on the existence and uniqueness of solutions of ordinary
differential equations.
This theorem is a wonderful example
that uses many results proved in the book. For more advanced students,
material may be covered faster so that we arrive at metric spaces and
prove Picard's theorem using the fixed point theorem as is usual.
Other excellent books exist. My favorite is
Rudin's excellent
\emph{Principles of Mathematical Analysis}~\cite{Rudin:baby}
or, as it is commonly and lovingly called, \emph{baby Rudin}
(to distinguish it from his other great analysis textbook,
\emph{big Rudin}). I took a
lot of inspiration and ideas from Rudin. However, Rudin is a bit more
advanced and ambitious than this present course.
For those who wish to continue
mathematics, Rudin is a fine investment.
An inexpensive and somewhat simpler alternative to Rudin is
Rosenlicht's \emph{Introduction to Analysis}~\cite{Rosenlicht}.
There is also the freely downloadable \emph{Introduction to Real
Analysis} by William Trench~\cite{Trench}.
\medskip
A note about the style of some of the proofs: Many proofs
traditionally done by contradiction, I prefer to do by
a direct proof or by contrapositive. While the book does include
proofs by contradiction, I only
do so when the contrapositive statement seemed too awkward or when
contradiction follows rather quickly.
Contradiction is more likely to get beginning students into trouble,
as we are talking about objects that do not exist.
I try to avoid unnecessary formalism where it is unhelpful.
Furthermore, the proofs and the language get slightly less formal as we
progress through the book, as more and more details are left out to avoid
clutter.
As a general rule, I use $\coloneqq$ instead of $=$ to define an
object rather than to simply show equality. I use this symbol rather more
liberally than is usual for emphasis.
I use it even when the context is \myquote{local,}
that is, I may simply define a function $f(x) \coloneqq x^2$
for a single exercise or example.
\medskip
Finally, I would like to acknowledge Jana Ma\v{r}\'ikov\'a,
Glen Pugh, Paul Vojta, Frank Beatrous, S\"{o}nmez \c{S}ahuto\u{g}lu,
Jim Brandt, Kenji Kozai, Arthur Busch, Anton Petrunin,
Mark Meilstrup, Harold P.\ Boas, Atilla Y{\i}lmaz,
Thomas Mahoney, Scott Armstrong, and Paul Sacks,
Matthias Weber, Manuele Santoprete,
Robert Niemeyer, Amanullah Nabavi,
for teaching with the book and giving me lots of useful feedback.
Frank Beatrous wrote the University of Pittsburgh version extensions, which
served as inspiration for many more recent additions.
I would also like to thank
Dan Stoneham, Jeremy Sutter, Eliya Gwetta, Daniel Pimentel-Alarc\'on,
Steve Hoerning, Yi Zhang, Nicole Caviris,
Kristopher Lee, Baoyue Bi, Hannah Lund,
Trevor Mannella, Mitchel Meyer, Gregory Beauregard,
Chase Meadors, Andreas Giannopoulos, Nick Nelsen,
Ru Wang, Trevor Fancher, Brandon Tague, Wang KP,
Wai Yan Pong, Sam Merat, Judah Nouriyelian,
Arnold Cross, Jesse Wallace,
Adnan Hashem Mohamed,
Nikita Borisov, Bob Strain,
Salven V.\ DeMartino,
an anonymous reader or two, and in general all the students in my classes
for suggestions and finding errors and typos.
%mbxlatex During the writing of this book,
%mbxlatex the author was in part supported by NSF grant DMS-0900885 and
%mbxlatex DMS-1362337.
% if in the two volume set then this will expand to the intro for vol II
%mbxSTARTIGNORE
\VolTwoIntro
%mbxENDIGNORE
% pretext (mbx) is always the two volume set
%mbxlatex
%mbxlatex \subsection{About Volume II}
%mbxlatex
%mbxlatex \input{frag-vol2-intro.tex}
%mbxlatex
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\sectionnewpage
\section{About analysis} \label{sec:aboutra}
Analysis is the branch of mathematics that deals with inequalities and
limits. The present course deals with the most basic
concepts in analysis. The goal of the course is to acquaint the reader
with rigorous proofs in analysis and also to
set a firm foundation for calculus of one variable (and several variables if
volume II is also considered).
Calculus has prepared you, the student, for using mathematics without telling
you why what you learned is true. To use, or teach, mathematics
effectively, you cannot simply know \emph{what} is true, you must know
\emph{why} it is true. This course shows you \emph{why} calculus
is true. It is here to give you a good understanding of the concept of a
limit, the derivative, and the integral.
Let us use an analogy.
An auto mechanic who has learned to change the oil, fix broken headlights,
and charge the battery, but who does not understand how the engine
works, will only be able to do those simple tasks.
He will be unable to work independently to diagnose and fix problems.
A high school teacher who does not understand the definition of the Riemann
integral or the derivative may not be able to properly answer all the
students' questions.
To this day I remember several nonsensical statements I heard
from my calculus teacher in high school, who simply did not understand
the concept of the limit, although he could \myquote{do} the problems in the
textbook.
\medskip
We start with a discussion of the real number system, most importantly
its completeness property, which is the basis for all that follows.
We then discuss the simplest form of a limit,
the limit of a sequence. Afterwards, we study
functions of one variable, continuity, and the derivative.
Next, we define the Riemann integral and prove the fundamental theorem of
calculus. We discuss sequences of functions and the
interchange of limits. Finally, we give an introduction to metric
spaces.
\medskip
Let us give the most important difference between analysis and algebra.
In algebra, we prove equalities directly;
we prove that an object, perhaps a number, is equal to another object.
In analysis, we usually prove inequalities,
and we prove those inequalities by estimating.
To illustrate the point, consider the following statement.
\medskip
\emph{Let $x$ be a real number. If $x < \epsilon$ is true for all
real numbers
$\epsilon > 0$, then $x \leq 0$}.
\medskip
This statement is the general idea of what we do in analysis.
Suppose we really wish to prove the equality
$x = 0$. In analysis, we prove two inequalities:
$x \leq 0$ and $x \geq 0$.
To
prove the inequality
$x \leq 0$, we prove
$x < \epsilon$ for all positive $\epsilon$.
To
prove the inequality
$x \geq 0$, we prove
$x > -\epsilon$ for all positive $\epsilon$.
\medskip
The term \emph{real analysis} is a little bit of a misnomer. I prefer to
use simply \emph{analysis}. The other type of analysis,
\emph{complex analysis}, really builds up on the present material, rather than
being distinct. Furthermore, a more advanced course on real
analysis would talk about complex numbers often.
I suspect the nomenclature is
historical baggage.
\medskip
Let us get on with the show\ldots
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\sectionnewpage
\section{Basic set theory} \label{sec:basicset}
\index{set theory}
%mbxINTROSUBSECTION
\sectionnotes{1--3 lectures (some material can be skipped, covered lightly,
or left as reading)}
Before we start talking about analysis, we need to fix some language.
Modern\footnote{The term \myquote{modern} refers to the late 19th century up to
the present.}
analysis uses the language of sets and, therefore, that is where we start.
We talk about sets in a rather informal way, using the so-called
\myquote{\myindex{na\"ive set theory}.} Do not worry, that is what the majority of
mathematicians use, and it is hard to get into trouble.
The reader has hopefully seen the very basics of set theory
and proof writing before, and this section should be a quick refresher.
\subsection{Sets}
\begin{defn}
A \emph{\myindex{set}} is a collection of objects called
\emph{elements}\index{element} or \emph{members}\index{member}. A set with
no objects is called the \emph{\myindex{empty set}} and is denoted by
$\emptyset$\glsadd{not:emptyset} (or sometimes by $\{ \}$).
\end{defn}
Think of a set as a club with a certain membership. For
example, the students who play chess are members of the chess club.
The same student can be a member of many different clubs.
However,
do not take the analogy too far. A set is only defined by the members
that form the set; two sets that have the same members are the same set.
Most of the time, we will consider sets
of numbers. For example, the set
\glsadd{not:set}
\begin{equation*}
S \coloneqq \{ 0, 1, 2 \}
\end{equation*}
is the set containing
the three elements 0, 1, and 2.
By \myquote{$\coloneqq$},\glsadd{not:def} we mean we are defining what $S$ is, rather than
just showing equality.
We write\glsadd{not:eltof}
\begin{equation*}
1 \in S
\end{equation*}
to denote that the number 1 belongs to the set $S$. That is, 1 is a member
of $S$.
At times we want to say that two elements are in a set $S$, so we
write \myquote{$1,2 \in S$} as a shorthand for
\myquote{$1 \in S$ and $2 \in S$.}
Similarly, we write\glsadd{not:noteltof}
\begin{equation*}
7 \notin S
\end{equation*}
to denote that the number 7 is not in $S$. That is, 7 is not a member of
$S$.
The elements of all sets under consideration come from some set we call the
\emph{\myindex{universe}}. For simplicity,
we often consider the universe to be the set that contains only the elements
we are interested in.
The universe is generally understood from context
and is not explicitly mentioned. In this course, our universe will
often be the set of real numbers.
Although the elements of a set are often numbers,
other objects, such as other sets, can be elements of a set.
A set may also contain some of the same elements as another set. For example,
\begin{equation*}
T \coloneqq \{ 0, 2 \}
\end{equation*}
contains the numbers 0 and 2. In this case, all elements of $T$ also
belong to $S$. We write $T \subset S$.\glsadd{not:subset}
See \figureref{fig:sets} for a diagram.
%More formally we make the
%following definition.
\begin{myfigureht}
\subimport*{figures/}{sets.pdf_t}
\caption{A diagram of the example sets $S$ and its subset $T$.\label{fig:sets}}
\end{myfigureht}
\begin{defn}
%FIXMEevillayouthack
\pagebreak[2]
\leavevmode
\begin{enumerate}[(i)]
\item
A set $A$ is a \emph{\myindex{subset}}
of a set $B$ if $x \in A$ implies $x \in B$, and we write $A \subset B$.
That is, all members of $A$ are also members of $B$. At times we
write $B \supset A$ to mean the same thing.
\item
Two sets $A$ and $B$ are \emph{\myindex{equal}}\glsadd{not:equal} if $A \subset B$ and $B
\subset A$. We write $A = B$.
That is, $A$ and $B$ contain exactly the same elements.
If it is not true that $A$ and $B$ are equal, then
we write $A \not= B$.
\item
A set $A$ is a \emph{\myindex{proper subset}} of $B$ if $A \subset B$
and $A \not= B$. We write $A \subsetneq B$.\glsadd{not:subsetneq}
\end{enumerate}
\end{defn}
For the example $S$ and $T$ defined above, $T \subset S$, but
$T \not= S$. So $T$ is a proper subset of $S$.
If $A = B$, then $A$ and $B$ are simply two names for the
same exact set.
To define new sets, one often uses
the
\emph{\myindex{set building notation}},\glsadd{not:setbuilder}
\begin{equation*}
\bigl\{ x \in A : P(x) \bigr\} .
\end{equation*}
This notation refers to a subset of the set $A$ containing all elements
of $A$ that satisfy the property $P(x)$.
Using $S = \{ 0, 1, 2 \}$ as above, $\{ x \in S : x \not= 2 \}$
is the set $\{ 0, 1 \}$.
The notation is sometimes
abbreviated as $\bigl\{ x : P(x) \bigr\}$, that is, $A$ is not mentioned when understood from context.
Furthermore, $x \in A$ is sometimes replaced with a formula to make the notation
easier to read.
\begin{example}
The following are sets including the standard notations.
\begin{enumerate}[(i)]
\item The set of \emph{\myindex{natural numbers}},
\glsadd{not:natnums}
$\N \coloneqq \{ 1, 2, 3, \ldots
\}$.
\item The set of \emph{\myindex{integers}},
\glsadd{not:ints}
$\Z \coloneqq \{ 0, -1, 1, -2, 2, \ldots
\}$.
\item The set of \emph{\myindex{rational numbers}},
\glsadd{not:rationals}
$\Q \coloneqq \bigl\{ \frac{m}{n} : m, n \in \Z \text{ and } n \not= 0 \bigr\}$.
\item The set of even natural numbers, $\{ 2m : m \in \N \}$.
\item The set of real numbers,
\glsadd{not:reals}
$\R$.
\end{enumerate}
Note that $\N \subset \Z \subset \Q \subset \R$.
\end{example}
We create new sets out of old ones by applying some natural operations.
\begin{defn}
\pagebreak[2]
\leavevmode
\begin{enumerate}[(i)]
\item
A \emph{\myindex{union}} of two sets $A$ and $B$ is defined as
\glsadd{not:union}
\begin{equation*}
A \cup B \coloneqq \{ x : x \in A \text{ or } x \in B \} .
\end{equation*}
\item
An \emph{\myindex{intersection}} of two sets $A$ and $B$ is defined as
\glsadd{not:intersection}
\begin{equation*}
A \cap B \coloneqq \{ x : x \in A \text{ and } x \in B \} .
\end{equation*}
\item
A \emph{complement of $B$ relative to $A$\index{complement relative to}}
(or \emph{\myindex{set-theoretic difference}} of $A$ and $B$) is defined as
\glsadd{not:setminus}
\begin{equation*}
A \setminus B \coloneqq \{ x : x \in A \text{ and } x \notin B \} .
\end{equation*}
\item
We say
\emph{\myindex{complement}} of $B$ and write $B^c$ \glsadd{not:complement}
instead of $A \setminus B$ if
the set $A$ is either the entire universe or if it is the obvious
set containing $B$, and is understood from context.
\item
We say sets $A$ and $B$ are \emph{\myindex{disjoint}} if $A \cap B =
\emptyset$.
\end{enumerate}
\end{defn}
The notation $B^c$ may be a little vague at this point. If
the set $B$ is a subset of the real numbers $\R$, then $B^c$ means
$\R \setminus B$. If $B$ is naturally a subset of the natural numbers,
then $B^c$
is $\N \setminus B$. If ambiguity can arise, we
use the set difference notation $A \setminus B$.
\begin{myfigureht}
\subimport*{figures/}{figsetop.pdf_t}
\caption{Venn diagrams of set operations, the result of the operation is
shaded.\label{figsetop}}
\end{myfigureht}
We illustrate the operations on the
\emph{Venn diagrams}\index{Venn diagram} in \figureref{figsetop}.
Let us now establish one of the most basic theorems about sets and logic.
\begin{thm}[DeMorgan]\index{DeMorgan's theorem}
Let $A, B, C$ be sets. Then
\begin{equation*}
{(B \cup C)}^c = B^c \cap C^c , \qquad
{(B \cap C)}^c = B^c \cup C^c ,
\end{equation*}
or, more generally,
\begin{equation*}
A \setminus (B \cup C) = (A \setminus B) \cap (A \setminus C) ,
\qquad
A \setminus (B \cap C) = (A \setminus B) \cup (A \setminus C) .
\end{equation*}
\end{thm}
\begin{proof}
The first statement is proved by the second statement if we
assume the set $A$ is our \myquote{universe.}
Let us prove $A \setminus (B \cup C) = (A \setminus B) \cap (A \setminus C)$.
Remember the definition of equality of sets. First, we must show that
if $x \in A \setminus (B \cup C)$, then
$x \in (A \setminus B) \cap (A \setminus C)$. Second, we must also show that
if $x \in (A \setminus B) \cap (A \setminus C)$, then
$x \in A \setminus (B \cup C)$.
So let us assume $x \in A \setminus (B \cup C)$. Then $x$ is in
$A$, but not in $B$ nor $C$. Hence $x$ is in $A$ and not in $B$, that is,
$x \in A \setminus B$. Similarly, $x \in A \setminus C$. Thus
$x \in (A \setminus B) \cap (A \setminus C)$.
On the other hand, suppose
$x \in (A \setminus B) \cap (A \setminus C)$. In particular,
$x \in (A \setminus B)$, so
$x \in A$ and $x \notin B$. Also, as $x \in (A \setminus C)$, then $x \notin C$.
Hence $x \in A \setminus (B \cup C)$.
The proof of the other equality is left as an exercise.
\end{proof}
The result above we called a \emph{Theorem}, while most results we call
a \emph{Proposition}, and a few we call a \emph{Lemma} (a result leading to another result) or
\emph{Corollary} (a quick consequence of the preceding result).
Do not read too much into
the naming. Some of it is traditional, some of it is stylistic choice.
It is not necessarily true that a \emph{Theorem} is always
\myquote{more important} than a
\emph{Proposition} or a \emph{Lemma}.
\medskip
We will also need to intersect or union several sets at once. If there are
only finitely many, then we simply apply the union or intersection operation
several times. However, suppose we have an infinite collection
of sets (a set of sets)
$\{ A_1, A_2, A_3, \ldots \}$. We define
\glsadd{not:infunion}
\glsadd{not:infintersection}
\begin{align*}
& \bigcup_{n=1}^\infty A_n \coloneqq \{ x : x \in A_n \text{ for some } n \in \N
\} , \\
& \bigcap_{n=1}^\infty A_n \coloneqq \{ x : x \in A_n \text{ for all } n \in \N
\} .
\end{align*}
We can also have sets indexed by two natural numbers. For example, we can have
the set of sets
$\{ A_{1,1}, A_{1,2}, A_{2,1}, A_{1,3}, A_{2,2}, A_{3,1}, \ldots \}$. Then
we write
\begin{equation*}
\bigcup_{n=1}^\infty \bigcup_{m=1}^\infty A_{n,m}
=
\bigcup_{n=1}^\infty \left( \bigcup_{m=1}^\infty A_{n,m} \right) .
\end{equation*}
And similarly with intersections.
It is not hard to see that we can take the unions in any order. However,
switching the order of unions and intersections is not generally permitted without proof.
For instance,
\begin{equation*}
\bigcup_{n=1}^\infty
\bigcap_{m=1}^\infty
\{ k \in \N : mk < n \}
=
\bigcup_{n=1}^\infty \emptyset = \emptyset .
\end{equation*}
However,
\begin{equation*}
\bigcap_{m=1}^\infty
\bigcup_{n=1}^\infty
\{ k \in \N : mk < n \}
=
\bigcap_{m=1}^\infty
\N
=
\N.
\end{equation*}
Sometimes, the index set is not the set of natural numbers. In such a case, we
require a more general notation. Suppose $I$ is some set and for each $\lambda \in
I$, there is a set $A_\lambda$. Then we define
\glsadd{not:arbitunion}
\glsadd{not:arbitintersection}
\begin{equation*}
\bigcup_{\lambda \in I} A_\lambda \coloneqq \{ x : x \in A_\lambda \text{ for some }
\lambda \in I
\} ,
\qquad
\bigcap_{\lambda \in I} A_\lambda \coloneqq \{ x : x \in A_\lambda \text{ for all }
\lambda \in I
\} .
\end{equation*}
\subsection{Induction}
When a statement includes an arbitrary natural number,
a common proof method is the principle of induction.
We start with the set of natural numbers $\N = \{ 1,2,3,\ldots \}$, and we
give them their natural ordering,
that is, $1 < 2 < 3 < 4 < \cdots$.
By $S \subset \N$ having a \emph{\myindex{least element}}, we mean that
there exists an $x \in S$,
such that for every
$y \in S$, we have $x \leq y$.
The natural numbers $\N$ ordered in the natural way
possess the so-called \emph{\myindex{well ordering property}}.
We take this property
as an \myindex{axiom}; we simply assume it is true.
%mbxSTARTIGNORE
\theoremstyle{plain}
\newtheorem*{wellordprop}{Well ordering property of $\N$}
\hypertarget{wop:link}{}%
\begin{wellordprop}
Every nonempty subset of $\N$ has a least (smallest) element.
\end{wellordprop}
%mbxENDIGNORE
%mbx <axiom xml:id="wop_link" number="">
%mbx <title>Well ordering property of <m>\N</m></title>
%mbx <statement><p>
%mbx Every nonempty subset of <m>\N</m> has a least (smallest) element
%mbx </p></statement>
%mbx </axiom>
The \emph{principle of \myindex{induction}}\index{principle of induction} is
the following theorem, which is in a sense\footnote{To be completely
rigorous, this equivalence is only true if we
also assume as an axiom that $n-1$ exists for all natural numbers bigger than $1$, which we
do. In this book, we are assuming all the usual arithmetic holds.}
equivalent to the well ordering property of the natural numbers.
\begin{thm}[Principle of induction] \label{induction:thm}
Let $P(n)$ be a statement depending on a natural number~$n$. Suppose that
\begin{enumerate}[(i)]
\item \emph{(\myindex{basis statement})} $P(1)$ is true.
\item \emph{(\myindex{induction step})} If $P(n)$ is true, then $P(n+1)$ is true.
\end{enumerate}
Then $P(n)$ is true for all $n \in \N$.
\end{thm}
\begin{proof}
Let $S$ be the set of natural numbers $n$ for which $P(n)$ is
not true. Suppose for contradiction that
$S$ is nonempty. Then $S$ has a least element by the well ordering
property. Call $m \in S$ the least element of $S$. We know $1 \notin
S$ by hypothesis. So $m > 1$, and $m-1$ is a natural number as well.
Since $m$ is the least element of $S$, we know that $P(m-1)$ is true.
But the induction step says that $P(m-1+1) = P(m)$ is true,
contradicting the statement that $m \in S$. Therefore, $S$ is empty and
$P(n)$ is true for all $n \in \N$.
\end{proof}
Sometimes it is convenient to start at a different number than 1,
all that changes is the labeling. The assumption that
$P(n)$ is true in \myquote{if $P(n)$ is true,
then $P(n+1)$ is true}
is usually called the \emph{\myindex{induction hypothesis}}.
\begin{example}
Let us prove that for all $n \in \N$,
\begin{equation*}
2^{n-1} \leq n! \qquad \text{(recall } n! = 1 \cdot 2 \cdot 3 \cdots n\text{)}.
\end{equation*}
Let $P(n)$ be the statement that
$2^{n-1} \leq n!$ is true.
Plug in $n=1$ to see that $P(1)$ is true.
Suppose $P(n)$ is true. That is, suppose
$2^{n-1} \leq n!$ holds. Multiply both sides by 2 to obtain
\begin{equation*}
2^n \leq 2(n!) .
\end{equation*}
As $2 \leq (n+1)$ when $n \in \N$, we have
$2(n!) \leq (n+1)(n!) = (n+1)!$. That is,
\begin{equation*}
2^n \leq 2(n!) \leq (n+1)!,
\end{equation*}
and hence $P(n+1)$ is true. By the principle of induction,
$P(n)$
is true for all $n \in \N$. In other words,
$2^{n-1} \leq n!$ is true for all $n \in \N$.
\end{example}
\begin{example} \label{example:geometricsum}
We claim that for all $c \not= 1$,
\begin{equation*}
1 + c + c^2 + \cdots + c^n = \frac{1-c^{n+1}}{1-c} .
\end{equation*}
Proof: It is easy to check that the equation holds with $n=1$. Suppose
it is true for $n$. Then
\begin{equation*}
\begin{split}
1 + c + c^2 + \cdots + c^n + c^{n+1} & =
( 1 + c + c^2 + \cdots + c^n ) + c^{n+1} \\
& = \frac{1-c^{n+1}}{1-c} + c^{n+1} \\
& = \frac{1-c^{n+1} + (1-c)c^{n+1}}{1-c} \\
& = \frac{1-c^{n+2}}{1-c} .
\end{split}
\end{equation*}
\end{example}
Sometimes, it is easier to use in the inductive step
that $P(k)$ is true for all $k = 1,2,\ldots,n$, not just for $k=n$.
This principle is called \emph{\myindex{strong induction}} and is equivalent
to the normal induction above. The proof of that
equivalence is left as an exercise.
\begin{thm}[Principle of strong induction]
\index{principle of strong induction}\index{strong induction}
Let $P(n)$ be a statement depending on a natural number $n$. Suppose that
\begin{enumerate}[(i)]
\item \emph{(\myindex{basis statement})} $P(1)$ is true.
\item \emph{(\myindex{induction step})} If $P(k)$ is true for all $k = 1,2,\ldots,n$, then $P(n+1)$ is true.
\end{enumerate}
Then $P(n)$ is true for all $n \in \N$.
\end{thm}
\subsection{Functions}
Informally,
a \emph{\myindex{set-theoretic function}}\index{function}
$f$ taking a set $A$ to a set $B$
is a mapping that to each $x \in A$ assigns a unique $y \in B$. We write
$f \colon A \to B$.\glsadd{not:func} An example function
$f \colon S \to T$ taking $S \coloneqq \{ 0, 1, 2 \}$ to $T \coloneqq \{ 0, 2 \}$
can be defined
by assigning $f(0) \coloneqq 2$, $f(1) \coloneqq 2$, and $f(2) \coloneqq 0$. That is, a function $f
\colon A \to B$ is
a black box, into which we stick an element of $A$ and the function
spits out an element of $B$.
Sometimes $f$ is called a \emph{\myindex{mapping}} or a
\emph{\myindex{map}},
and we say $f$ \emph{maps $A$ to $B$}.
Often, functions are defined by some sort of
formula; however, you should really think of a function as just a very large
table of values.
The subtle issue here is that a single function can have several
formulas, all giving the same function. Also, for many functions, there is
no formula that expresses its values.
To define a function rigorously, let us first define the Cartesian product.
\begin{defn}
Let $A$ and $B$ be sets. The \emph{\myindex{Cartesian product}}
is the set of tuples defined as
\glsadd{not:cartprod}
\begin{equation*}
A \times B \coloneqq
\bigl\{ (x,y) : x \in A, y \in B \bigr\} .
\end{equation*}
\end{defn}
For instance, $\{ a,b \} \times \{ c , d\} =
\bigl\{
(a,c), (a,d), (b,c), (b,d)
\bigr\}$.
A more complicated example is the set $[0,1] \times [0,1]$: a subset of the
plane bounded by a square with vertices $(0,0)$, $(0,1)$, $(1,0)$, and $(1,1)$.
When $A$ and $B$ are the same set, we often use a superscript 2 to denote
such a product. For example, $[0,1]^2 =
[0,1] \times [0,1]$ or $\R^2 = \R \times \R$ (the Cartesian plane).
\begin{defn}
A \emph{function} $f \colon A \to B$ is a subset $f$ of $A \times B$
such that for each $x \in A$, there exists a unique $y \in B$ for which $(x,y) \in f$.
We write $f(x) = y$. Sometimes
the set $f$ is called the \emph{\myindex{graph}} of the function rather than
the function itself.
The set $A$ is called the \emph{\myindex{domain}} of $f$ (and
sometimes confusingly denoted $D(f)$). The set
\begin{equation*}
R(f) \coloneqq \{ y \in B : \text{there exists an } x \in A \text{ such that }
f(x)=y \}
\end{equation*}
is called the \emph{\myindex{range}} of $f$.
The set $B$ is called the \emph{\myindex{codomain}} of $f$.
\end{defn}
It is possible that the range $R(f)$ is a proper subset of the codomain $B$,
while the domain of $f$ is always equal to $A$. We generally
assume that the domain of $f$ is nonempty.
\begin{example}
From calculus, you are most familiar with functions taking real numbers to real
numbers. However, you saw some other types of functions as well.
The derivative is a function that maps the set of
differentiable functions to the set of all functions.
Another example is the Laplace transform, which also
takes functions to functions. Yet another example is the function that takes
a continuous function $g$ defined on the interval $[0,1]$ and returns the
number $\int_0^1 g(x) \,dx$.
\end{example}
\begin{defn}
Consider a function $f \colon A \to B$. Define
the \emph{\myindex{image}} (or \emph{\myindex{direct image}}) of a subset
$C \subset A$ as
\glsadd{not:dirimage}
\begin{equation*}
f(C) \coloneqq \bigl\{ f(x) \in B : x \in C \bigr\} .
\end{equation*}
Define the \emph{\myindex{inverse image}} of a subset $D
\subset B$ as
\glsadd{not:invimage}
\begin{equation*}
f^{-1}(D) \coloneqq \bigl\{ x \in A : f(x) \in D \bigr\} .
\end{equation*}
\end{defn}
In particular, $R(f) = f(A)$, the range is the direct image of
the domain $A$.
\begin{myfigureht}
\subimport*{figures/}{funcimags_full.pdf_t}
\caption{Example of direct and inverse images for the function
$f \colon \{ 1,2,3,4 \} \to \{ a,b,c,d \}$ defined by
$f(1) \coloneqq b$,
$f(2) \coloneqq d$,
$f(3) \coloneqq c$,
$f(4) \coloneqq b$.\label{figfuncimags}}
\end{myfigureht}
\begin{example}
Define the function $f \colon \R \to \R$ by
$f(x) \coloneqq \sin(\pi x)$. Then $f\bigl([0,\nicefrac{1}{2}]\bigr) = [0,1]$,
$f^{-1}\bigl(\{0\}\bigr) = \Z$, etc.
\end{example}
\begin{prop} \label{st:propinv}
Consider $f \colon A \to B$. Let $C, D$ be subsets of $B$. Then
\begin{align*}
& f^{-1}( C \cup D) = f^{-1} (C) \cup f^{-1} (D) , \\
& f^{-1}( C \cap D) = f^{-1} (C) \cap f^{-1} (D) , \\
& f^{-1}( C^c) = {\bigl( f^{-1} (C) \bigr)}^c .
\end{align*}
\end{prop}
Read the last line of the proposition as
$f^{-1}( B \setminus C) = A \setminus f^{-1} (C)$.
\begin{proof}
We start with the union. If $x \in
f^{-1}( C \cup D)$, then
$x$ is taken to $C$ or $D$, that is, $f(x) \in C$ or $f(x) \in D$. Thus
$f^{-1}( C \cup D) \subset f^{-1} (C) \cup f^{-1} (D)$. Conversely,
if $x \in f^{-1}(C)$, then $x \in f^{-1}(C \cup D)$. Similarly for
$x \in f^{-1}(D)$. Hence
$f^{-1}( C \cup D) \supset f^{-1} (C) \cup f^{-1} (D)$, and we have
equality.
The rest of the proof is left as an exercise.
\end{proof}
For direct images, the best we can do is the
following weaker result.
\begin{prop} \label{st:propfor}
Consider $f \colon A \to B$. Let $C, D$ be subsets of $A$. Then
\begin{align*}
& f( C \cup D) = f (C) \cup f (D) , \\
& f( C \cap D) \subset f (C) \cap f (D) .
\end{align*}
\end{prop}
The proof is left as an exercise.
\begin{defn}
Let $f \colon A \to B$ be a function.
The function $f$ is said to be
\emph{\myindex{injective}} or
\emph{\myindex{one-to-one}} if $f(x_1) = f(x_2)$ implies $x_1 = x_2$. In
other words,
$f$ is injective if
for all $y \in B$, the set
$f^{-1}(\{y\})$ is empty or consists of a single element.
We call such an $f$ an \emph{\myindex{injection}}.
If $f(A) = B$, then we say $f$ is
\emph{\myindex{surjective}} or
\emph{\myindex{onto}}.
In other words, $f$ is surjective if the range and the codomain of $f$ are equal.
We call such an $f$ a \emph{\myindex{surjection}}.
If $f$ is both surjective and injective, then
we say $f$ is \emph{\myindex{bijective}} or that $f$ is a
\emph{\myindex{bijection}}.
\end{defn}
When $f \colon A \to B$ is a bijection, then the inverse image of a single
element, $f^{-1}(\{y\})$, is always
a unique element of $A$. We then consider $f^{-1}$ as a function
\glsadd{not:invfunction}
$f^{-1} \colon B \to A$ and we write simply $f^{-1}(y)$.
In this case, we call $f^{-1}$ the \emph{\myindex{inverse function}} of $f$.
For instance, for the bijection $f \colon \R \to \R$ defined by
$f(x) \coloneqq x^3$, we have
$f^{-1}(x) = \sqrt[3]{x}$.
%A final piece of notation for functions that
%we need is the \emph{\myindex{composition of functions}}.
\begin{defn}
Consider $f \colon A \to B$ and $g \colon B \to C$. The
\emph{composition}\index{composition of functions}
of the functions $f$ and $g$ is the function
$g \circ f \colon A \to C$ defined as
\glsadd{not:composition}
\begin{equation*}
(g \circ f)(x) \coloneqq g\bigl(f(x)\bigr) .
\end{equation*}
\end{defn}
For example, if $f \colon \R \to \R$ is $f(x)\coloneqq x^3$ and $g \colon \R \to \R$
is $g(y) = \sin(y)$, then $(g \circ f)(x) = \sin(x^3)$.
It is left to the reader as an easy exercise
to show that composition of one-to-one maps is
one-to-one and composition of onto maps is onto.
Therefore, the composition of bijections is a bijection.
\subsection{Relations and equivalence classes}
We often compare two objects in some way. For instance, we say $1 < 2$
for natural numbers, $\nicefrac{1}{2} = \nicefrac{2}{4}$ for rational
numbers, or $\{ a,c \} \subset \{ a,b,c \}$ for sets. The `$<$', `$=$', and
`$\subset$' are examples of
relations.
\begin{defn}
Given a set $A$, a \emph{\myindex{binary relation}}\index{relation} on $A$
is a subset $\sR \subset A \times A$,
which consists of those pairs where the relation is said to hold.
Instead of $(a,b) \in \sR$, we write
$a \, \sR \, b$.
\end{defn}
\begin{example}
Take $A \coloneqq \{ 1,2,3 \}$.
Consider the relation `$<$'. The corresponding set
of pairs is $\bigl\{ (1,2), (1,3), (2,3) \bigr\}$. So $1 < 2$ holds as $(1,2)$
is in the corresponding set of pairs, but $3 < 1$ does not hold as $(3,1)$ is
not in the set.
Similarly, the relation `$=$'
is defined by the set of pairs $\bigl\{ (1,1), (2,2), (3,3) \big\}$.
Any subset of $A \times A$ is a relation. If we define the relation
$\dagger$ via $\bigl\{ (1,2), (2,1), (2,3), (3,1) \bigr\}$, then $1 \dagger 2$ and
$3 \dagger 1$ are
true, but $1 \dagger 3$ is not.
\end{example}
\begin{defn}
%FIXMEevillayouthack
\pagebreak[2]
A relation
$\sR$ on a set $A$ is said to be
\begin{enumerate}[(i)]
\item
\emph{Reflexive}\index{reflexive relation} if $a\, \sR \, a$ for
all $a \in A$.
\item
\emph{Symmetric}\index{symmetric relation} if $a\, \sR \, b$ implies
$b \, \sR \, a$.
\item
\emph{Transitive}\index{transitive relation} if $a\, \sR \, b$ and
$b \, \sR \, c$ implies $a\, \sR \, c$.
\end{enumerate}
If $\sR$ is reflexive, symmetric, and transitive, then it is said to be
an \emph{\myindex{equivalence relation}}.
\end{defn}
\begin{example}
Let $A \coloneqq \{ 1,2,3 \}$.
The relation `$<$' is transitive but neither reflexive nor symmetric. The
relation `$\leq$' defined by
$\bigl\{ (1,1), (1,2), (1,3), (2,2), (2,3), (3,3) \bigr\}$
is reflexive and transitive, but not symmetric.
Finally, a relation `$\star$' defined by
$\bigl\{ (1,1), (1,2), \allowbreak (2,1), \allowbreak (2,2), \allowbreak (3,3) \bigr\}$ is
an equivalence relation.
\end{example}
Equivalence relations are useful as they divide a set into sets of
\myquote{equivalent} elements.
\begin{defn}
Let $A$ be a set and $\sR$ an equivalence relation.
An \emph{\myindex{equivalence class}} of $a \in A$, often denoted
by $[a]$,\glsadd{not:equivclass} is the set $\{ x \in A : a\, \sR \, x \}$.
\end{defn}
For example, given the relation `$\star$' above, there are two equivalence classes,
$[1] = [2] = \{ 1,2 \}$ and $[3] = \{ 3 \}$.
Reflexivity guarantees that $a \in [a]$. Symmetry guarantees
that if $b \in [a]$, then $a \in [b]$. Finally, transitivity guarantees that
if $b \in [a]$ and $c \in [b]$, then $c \in [a]$.
In particular, we have the following proposition, whose proof is an
exercise.
\begin{prop} \label{prop:equivclasses}
If $\sR$ is an equivalence relation on a set $A$,
then every $a \in A$ is in exactly one
equivalence class. Moreover, $a\, \sR \, b$ if and only if $[a] = [b]$.
\end{prop}
\begin{example} \label{example:ratnums}
The set of rational numbers can be defined as equivalence classes of
a pair of an integer and a natural number,
that is, elements of $\Z \times \N$. The relation
is defined by $(a,b) \sim (c,d)$ whenever $ad = bc$.
It is left as an exercise to prove that `$\sim$' is an equivalence relation.
Usually, the equivalence class $\bigl[(a,b)\bigr]$ is written as $\nicefrac{a}{b}$.
\end{example}
\subsection{Cardinality}
A subtle but fundamental issue in set theory
and one that generates a considerable amount of
confusion among beginning students is that of cardinality,
or \myquote{size} of sets.
%The concept of cardinality is fundamental in modern mathematics in general and
%in analysis in particular.
Indeed, in this section, we will see the first really
unexpected theorem.
\begin{defn}