-
Notifications
You must be signed in to change notification settings - Fork 0
/
notes.tex
3379 lines (2670 loc) · 203 KB
/
notes.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
\documentclass[11pt,oneside]{amsbook}
\title{Introduction to real analysis}
\author{Course notes written by Samuel Coskey;\\Based on material from ``Understanding Analysis'',\\written by Stephen Abbott}
\usepackage[vscale=.8,vmarginratio=4:3]{geometry}
\usepackage{mathpazo,amssymb}
\usepackage{setspace}\onehalfspacing\raggedbottom
\renewcommand{\labelitemi}{$\circ$}
\renewcommand{\labelenumi}{(\alph{enumi})}
\renewcommand{\chaptername}{Part}
\renewcommand{\thechapter}{\Roman{chapter}}
\usepackage{remreset}
\makeatletter\@removefromreset{section}{chapter}\makeatother
\usepackage{etoolbox}
\makeatletter
\pretocmd{\@seccntformat}{\S}{}{}
\patchcmd{\tocsection}{#2.}{\S#2.}{}{}
\apptocmd{\tocsection}{\dotfill}{}{}
\patchcmd{\@maketitle}{\newpage}{}{}{}
\makeatother
\usepackage[linktoc=all,colorlinks,linkcolor=blue]{hyperref}
\usepackage{tikz}\usetikzlibrary{positioning,decorations.fractals,decorations.pathmorphing}
\newcommand{\set}[1]{\left\{\,#1\,\right\}}
\newcommand{\N}{\mathbb N}
\newcommand{\Z}{\mathbb Z}
\newcommand{\Q}{\mathbb Q}
\newcommand{\R}{\mathbb R}
\DeclareMathOperator{\dom}{dom}
\DeclareMathOperator{\rng}{rng}
\DeclareMathOperator{\cl}{cl}
\DeclareMathOperator{\inte}{int}
\DeclareMathOperator{\ext}{ext}
\renewcommand{\setminus}{\smallsetminus}
\theoremstyle{definition}
\newtheorem{exerc}{Exercise}[section]
\swapnumbers
\theoremstyle{plain}
\newtheorem{theorem}{Theorem}[section]
\newtheorem{corollary}[theorem]{Corollary}
\newtheorem{lemma}[theorem]{Lemma}
\newtheorem{proposition}[theorem]{Proposition}
\theoremstyle{definition}
\newtheorem{definition}[theorem]{Definition}
\theoremstyle{remark}
\newtheorem{remark}[theorem]{Remark}
\newtheorem{example}[theorem]{Example}
\newtheorem*{reading}{Reading}
\newtheorem*{readnext}{Reading for next time}
\newtheorem*{notes}{Notes and further reading}
\numberwithin{equation}{section}
\numberwithin{figure}{section}
\renewcommand{\theequation}{\arabic{section}.e\arabic{equation}}
\renewcommand{\thefigure}{\arabic{section}.f\arabic{figure}}
\newcounter{activityitem}
\newenvironment{activity}{\begin{list}{\arabic{activityitem}.}{\usecounter{activityitem}\setlength{\itemsep}{.2in}}}{\end{list}}
\begin{document}
\maketitle
\tableofcontents
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{The theory of the real number line}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{What are rational and real numbers?}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\begin{reading}
Abbott \S 1.1, 1.2, start 8.6
\end{reading}
We begin the course by briefly pondering a very deep question: what are numbers and what are they for?
The natural or counting numbers $1,2,3,\ldots$ are collectively denoted $\N$ and are central to many areas of mathematical study. \emph{Combinatorics} is the study of counting finite sets. For example: how many ways are there to write the number 10 as a sum of three natural numbers? \emph{Number theory} is the study of the arithmetic operations on the natural numbers. For example: how many prime numbers are there less than 1 billion?
But in \emph{calculus}, analysis, and geometry, numbers are used for measurement. How can we measure the distance between two points on the floor? To answer this we would first need to choose a unit, such as a straight stick of random length. With this in hand, we can try count how many copies of the unit stick fit along the striaght line segment joining the two points. But what if the number of unit sticks doesn't come out whole?
In this case we can try to use \emph{ratios} to measure the distance. For example, suppose we are trying to measure the distance $d$, and three unit sticks fit perfectly inside two copies of $d$. Then we say that $d=3/2$ units. Sometimes this telescoping process takes a long time, for example if five units fit perfectly inside four copies of $d$. The set of all ratios of natural numbers (and their negatives) is called the \emph{rational numbers} and given the symbol $\Q$ (for quotient).
This leads to the \emph{big question} of the day: are rational numbers enough to measure all distances? It turns out that the answer is NO, and this fact surprised many early mathematicians (and still blows minds of new mathematicians today). Ratios are good enough for most practical purposes because in the real world we can accept approximations. But there are ideal distances (numbers) which are not rational. The simplest classical example is $d=$ the diagonal of a square.
First notice that $d$ has the property that $d^2=2$. The explanation for this is at once simple, beautiful, and clever: If we quadruplicate the square, we see that the square with sidelength $d$ is exactly half the square with sidelength $2$. Thus $d^2=2^2/2=2$.
\begin{center}
\begin{tikzpicture}
\draw (0,0)--(1,0)--(1,1)--node[above]{$1$}(0,1)--node[left]{$1$}(0,0)--node[below right,inner sep=0]{$d$} (1,1);
\end{tikzpicture}
\quad
\begin{tikzpicture}
\draw (0,0)--(2,0)--(2,2)--node[above]{$2$}(0,2)--node[left]{$2$}(0,0);
\draw (1,0)--node[below right,inner sep=0]{$d$}(2,1)--(1,2)--(0,1)--(1,0);
\end{tikzpicture}
\end{center}
Armed with the knowledge that $d^2=2$, we can now show $d$ is irrational.
\begin{theorem}
\label{thm:root-2-irrational}
If $d^2=2$ then $d\notin\Q$.
\end{theorem}
\begin{proof}
Suppose $d$ is rational and write $d=a/b$ where $a,b$ are natural numbers and the fraction is in lowest terms. Then $a^2=2b^2$. Notice that the right-hand side is even, and therefore so is the left-hand side. It follows that $a$ is even. Hence the left-hand side is divisible by $4$, and therefore so is the right-hand side. It follows that $b$ is even. Thus both $a$ and $b$ are even, contradicting that the fraction $a/b$ is in lowest terms.
\end{proof}
We conclude that the rational number system possesses some kind of gap at $\sqrt2$. Another way to say it is that the concept of $\sqrt2$ \emph{cuts} the rationals into two pieces, those that are less than $\sqrt2$ and those that are larger than $\sqrt{2}$. (Actually, since in our development the number $\sqrt{2}$ doesn't exist yet, it would be more appropriate to say: those that are negative or have square less than $2$, and those that are nonnegative and have square greater than $2$.)
The situation we just described is very similary to the way a rational number $q$ cuts the rationals into three pieces: those numbers that are less than $q$, $q$ itself, and those numbers that are greater than $q$. The key difference is the presence of $q$ itself in between the two halves. The concept of a cut is made formal in the following definition.
\begin{definition}
\label{def:cut}
A \emph{cut} in the rational number system $\Q$ consists of a triple of sets $(A,P,B)$ such that $A\cup P\cup B=\Q$ and the following properties hold:
\begin{enumerate}
\item $|A|=\infty$, $|P|=0$ or $1$, and $|B|=\infty$
\item $A$ is entirely below $B$, and if $P\neq\emptyset$ then $P$ is in between $A$ and $B$
\item $A$ has no greatest element, and $B$ has no least element
\end{enumerate}
\end{definition}
\begin{example}
The following is a cut:
\begin{equation}
\label{eq:rationalcut}
\left(\set{q\in\Q\mid q<3/2},\set{3/2},\set{q\in\Q\mid q>3/2}\right)\text{.}
\end{equation}
This type of cut, with the set $P$ containing a single rational number, is called a \emph{rational cut}.
\begin{center}
\begin{tikzpicture}
\draw[<-,very thick](-3,0)--node[below]{$A$}(-.05,0);
\draw[->,very thick](.05,0)--node[below]{$B$}(3,0);
\node[circle,fill,inner sep=1pt,label=below:$P$] at (0,0) {};
\node at (-.08,0) {$)$};
\node at (.08,0) {$($};
\end{tikzpicture}
\end{center}
\end{example}
\begin{example}
The following is a cut:
\begin{equation}
\label{eq:irrationalcut}
\left(\set{q\in\Q\mid q<0\text{ or }q^2<2},\emptyset,\set{q\in\Q\mid q\geq0\text{ and }q^2>2}\right)\text{.}
\end{equation}
This is the cut at $\sqrt{2}$ discussed above. We leave it until the next section to prove that $A$ really has no greatest element and $B$ has no least element. This type of cut, with $P=\emptyset$, is called an \emph{irrational cut} or a \emph{gap}.
\begin{center}
\begin{tikzpicture}
\draw[<-,very thick](-3,0)--node[below]{$A$}(-.01,0);
\draw[->,very thick](.01,0)--node[below]{$B$}(3,0);
\node at (0,0) {$)\kern-2.5pt($};
\end{tikzpicture}
\end{center}
\end{example}
We are now prepared to define the real number system by ``filling in'' the gaps of $\Q$.
\begin{definition}
The \emph{real number system} $\R$ consists of the rational numbers $\Q$ together with, for every irrational cut $(A,\emptyset,B)$, a new point $p$ lying between the sets $A$ and $B$.
\end{definition}
Thus there is a one-to-one correspondence between real numbers and cuts, with the rational numbers corresponding to rational cuts, and the irrational numbers corresponding to irrational cuts.
To close, let us look at the number systems we know so far, from smallest to largest. Each one has progressively more power than the last:
\begin{itemize}
\item $\N$: supports addition and multiplication, but not subtraction or division
\item $\Z$: supports addition, multiplication, and subtraction, but not division
\item $\Q$: supports all four operations, but has gaps;
\item $\R$: supports all four operations, and has no gaps.
\end{itemize}
In the next section we will prove the final claim, that $\R$ has no gaps.
% We emphasize that an irrational number is \emph{defined} from its cut. That is, we don't define $\sqrt2$ to be a positive number whose square is $2$, though it's true. That would tell you what property the number has, but not where to actually find it. Instead we define $\sqrt2$ to be a point in between $\set{q\in\Q\mid q<0\text{ or }q^2<2}$ and $\set{q\in\Q\mid q\geq0\text{ and }q^2<2}$.
\newpage
\subsection*{Activity for \S \thesection}
\begin{activity}
\item (see Abbott, exercise 1.4.1) Complete and prove:
\begin{enumerate}\itemsep\fill
\item The sum of two rational numbers is\ldots
\item The product of two rational numbers is\ldots
\item The sum of a rational number and an irrational number is\ldots
\item The product of a rational number and an irrational number is\ldots unless\ldots
\item The product of two irrational numbers is\ldots? (discuss)
\vspace{\fill}
\end{enumerate}
\item For each triple, either argue that it is a cut, or show that one of the conditions in the definition of cut is not satisfied.
\begin{enumerate}\itemsep\fill
\item $(\set{q\in\Q\mid q<2},\emptyset,\set{q\in\Q\mid q>2})$
\item $(\set{q\in\Q\mid q\leq2},\emptyset,\set{q\in\Q\mid q>2})$
\item $(\set{q\in\Q\mid q\leq2},\{2\},\set{q\in\Q\mid q>2})$
\item $(\set{q\in\Q\mid q<2},\{1.5\},\set{q\in\Q\mid q>2})$
\item $(\set{q\in\Q\mid q^2<2},\emptyset,\set{q\in\Q\mid q^2>2})$
\end{enumerate}
\end{activity}
\vspace{\fill}
\begin{readnext}
Abbott \S 8.6, start \S 1.3
\end{readnext}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Cuts and least upper bounds}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\begin{reading}
Abbott \S 8.6, start \S 1.3
\end{reading}
Although we introduced the real numbers as a way to measure distances, we of course know that the real numbers should carry arithmetic operations, too. In fact, it should be an \emph{ordered field}, which means it should support the four arithmetic operations as well as the comparison $\leq$, and these operations should satisfy the usual axioms of associativity, commutativity, and so forth. The rational number system has all these properties, and the real number system does too. We have yet to \emph{define} the arithmetic operations on the real numbers, but in principle it isn't difficult.
For example, we define the ordering $\leq$ on real numbers as follows. If $\alpha_1$ and $\alpha_2$ are real numbers corresponding to cuts $(A_1,P_1,B_1)$ and $(A_2,P_2,b_2)$ are cuts, then we say $\alpha_1\leq\alpha_2$ if and only if $A_1\subseteq A_2$. It is straightforward to check this definition satisfies the usual axioms of $\leq$.
Similarly, if we define $\alpha_1+\alpha_2$ as the real number corresponding to the cut $(A,P,B)$, where
\[A=\set{a_1+a_2\mid a_1\in A_1\text{ and }a_2\in A_2}
\]
and $P$ and $B$ determined using $A$. One must verify this really is a cut, and then verify the axioms of addition hold. We will omit the tedious details of defining the last three arithmetic operations, and verifying the remaining axioms.
Thus we continue with the knowledge that $\R$ is an ordered field. But $\R$ must satisfy even more properties than just this, or else we would have been happy to stick with $\Q$. It turns out that $\R$ additionally supports the \emph{least upper bound} operation, which we now define.
% In the previous section we defined the real numbers using cuts in the rational number system. Our ``triple'' notation for cuts $(A,P,B)$ is nice and symmetric, but as it happens knowing $A$ is enough to determine $P$ and $B$ too: $P\cup B$ will always be equal to the complement $\Q\setminus A$, with $P$ containing the least element of $\Q\setminus A$ if it exists. Thus from now on we drop the $P$ and the $B$ from cuts and make the following revised definition.
%
% \begin{definition}
% A \emph{cut} in $\Q$ is a subset $A\subset\Q$ such that the following properties hold
% \begin{enumerate}
% \item $A$ is infinite, and $\Q\setminus A$ is infinite;
% \item $A$ lies entirely below $\Q\setminus A$;
% \item $A$ has no greatest element.
% \end{enumerate}
% \end{definition}
% Intuitively, a cut $A$ corresponds to a real number $\alpha$ which ``comes just after'' $A$. But what does it mean to come just after? To answer this, we must define an ordering on real numbers. This turns out to be easy using our new definition of cuts. If $\alpha,\alpha'$ are real numbers corresponding to cuts $A,A'$, we say that $\alpha\leq\alpha'$ if and only if $A\subseteq A'$.
\begin{definition}
Let $S$ be a set of real numbers. An \emph{upper bound} for $S$ is a real number $\beta$ such that for all $\alpha\in S$, $\alpha\leq\beta$. If $\beta$ is an upper bound for $S$ then it is called the \emph{least upper bound} or \emph{supremum} of $S$ if it is least among all possible upper bounds of $S$.
\end{definition}
\begin{example}
\begin{itemize}
\item The set $(-5,5]$ has a maximum element $5$. Therefore it has least upper bound $5$.
\item The set $(-5,5)$ has no maximum element. But it has upper bounds, for example $7$, $23.6$, and $100$. The least upper bound or supremum is $5$.
\item The set $\set{\frac12,\frac23,\frac34,\ldots}$ has no maximum element. But it has upper bounds, and its supremum is $1$.
% \item The set $A=\set{q\in\Q\mid q<0\text{ or }q^2<2}$ has no maximum element. But it has upper bounds, such as $1.5$, and its supremum is $\sqrt2$.
\end{itemize}
\end{example}
% Our construction of the real numbers can now be rephrased using our new definition of cut and the concept of supremum: The real number system $\R$ consists of the rational numbers $\Q$ together with, for every cut $A\subset\Q$ which does not have a rational supremum, a newly inserted supremum $\alpha$ of $A$.
In the rational number system $\Q$, sets don't need to have a supremum (within $\Q$). Before we can prove this, we first show as promised that the cut at $\sqrt2$ is really a cut. The proof turns out to be more technical than one might expect. It will provide us with a not-so-soft example of the kind of arguments we will see later in the course.
\begin{theorem}
\label{thm:root-2-cut}
Let $(A,\emptyset,B)$ be the triple from Equation~\ref{eq:irrationalcut}. Then $(A,\emptyset,B)$ is a cut in $\Q$.
\end{theorem}
\begin{proof}
The key step is to verify that $A$ has no greatest element, as the proof that $B$ has no least element is symmetric. Let $q\in A$ be given; we will show $q$ is not the greatest element of $A$. It is clear that a negative number cannot be the greatest element of $A$, so we may assume that $q\geq0$ and $q^2<2$.
We wish to find a small rational number $\epsilon>0$ such that $q+\epsilon$ is in $A$ also, that is, $(q+\epsilon)^2<2$. To find what $\epsilon$ should be, we work \emph{backwards}.
\begin{align*}
(q+\epsilon)^2<2&\iff q^2+2q\epsilon+\epsilon^2<2\\
&\,\Longleftarrow\,\,\, q^2+2q\epsilon+\epsilon<2\\
&\iff \epsilon<(2-q^2)/(2q+1)
\end{align*}
There is a clever trick going on in the second implication (the $\Longleftarrow$ one) which we will comment on in a moment. But overall, this calculation is telling us that we will be safe if we choose $\epsilon<(2-q^2)/(2q+1)$. Note that we can do this and still have $\epsilon>0$, because $q^2<2$.
Now, in the second implication we replaced $\epsilon^2$ with $\epsilon$ to make it easier to solve for $\epsilon$. To do this, we actually need to make the extra assumption that $\epsilon<1$, since then $\epsilon^2<\epsilon$ and the $\Longleftarrow$ implication holds true. This is acceptible because we haven't yet chosen $\epsilon$, so we just need to additionally promise to choose $\epsilon<1$.
We now know that our proof will succeed if we choose $\epsilon$ to be any positive rational number such that $\epsilon<(2-q^2)/(2q+1)$ and $\epsilon<1$. To verify that this works, we use our preparatory work \emph{forwards} as follows.
\begin{align*}
\epsilon<(2-q^2)/(2q+1)\quad\&\quad\epsilon<1
&\implies q^2+2q\epsilon+\epsilon<2\quad\&\quad\epsilon^2<\epsilon\\
&\implies q^2+2q\epsilon+\epsilon^2<2\\
&\implies (q+\epsilon)^2<2
\end{align*}
We conclude that $q+\epsilon$ is a rational number greater than $q$ satisfying $(q+\epsilon)^2<2$. This shows $q$ is \emph{not} the greatest element of $A$, as desired.
\end{proof}
We may now conclude that if $(A,\emptyset,B)$ is the cut at $\sqrt2$, then $A$ has no rational supremum. Indeed if $\beta$ were the supremum of $A$, then since $A$ has no greatest element we have $\beta\notin A$. Similarly since $B$ has no least element we have $\beta\notin B$. Thus $\beta\notin\Q$. (In the next section, we will show that in fact $\beta=\sqrt2$.)
In the real number system $\R$, on the other hand, sets do have a supremum whenever possible.
\begin{lemma}
Let $S$ be a set of real numbers which has an upper bound. Then the supremum of $S$ exists.
\end{lemma}
\begin{proof}[Proof idea]
By our definition of the real numbers, every $\alpha\in S$ corresponds to a cut $(A_\alpha,P_\alpha,B_\alpha)$. From these cuts we define a new cut cut $(A,P,B)$ by letting $A=\bigcup A_\alpha$, $B=\bigcap B_\alpha$, and $P=$ whatever is left. We leave it to the reader to verify that $(A,P,B)$ is really a cut. Let $\beta$ be the real number corresponding to the cut $(A,P,B)$.
Now, since $A_\alpha\subset A$ for every $\alpha\in S$, we see that $\beta$ is an upper bound of $S$. Moreover $\beta$ is the least upper bound of $S$. Indeed, any other upper bound $\beta'$ corresponding to $(A',P',B')$ satisfies $A_\alpha\subset A'$ for every $\alpha\in S$, and therefore $A=\bigcup A_\alpha\subset A'$. In other words, $\alpha\leq\beta$, as desired.
\end{proof}
We are finally ready to prove the result, promised in the previous section, that $\R$ has no gaps.
\begin{theorem}
The set of real numbers has no gaps. That is, if $(A,P,B)$ is such that $A\cup P\cup B=\R$ and properties (a)(b)(c) of Definition~\ref{def:cut} satisfied, then $P\neq\emptyset$.
\end{theorem}
\begin{proof}
Let $(A,P,B)$ be such a triple. By the previous lemma, we can let $\alpha$ be the supremum of $A$. Since $A$ has no greatest element, we have $\alpha\notin A$. Since $B$ has no least element, $\alpha\notin B$. Thus we must have $\alpha\in P$, which means that $P\neq\emptyset$.
\end{proof}
\newpage
\subsection*{Activity for \S \thesection}
\begin{activity}
\item (Abbott, ex 1.3.8) Find the supremum and infemum of each of the following sets. You do not need to give a detailed proof, but please explain your reasoning. Assume for this exercise that $\N$ exludes $0$.
\begin{enumerate}\itemsep.5in
\item $\set{n\in\N\mid n^2<10}$
\item $\set{n/(m+n)\mid m,n\in\N}$
\item $\set{n/(2n+1)\mid n\in\N}$
\item $\set{n/m\mid m,n\in\N\text{ and }m+n\leq10}$
\vspace{.5in}
\end{enumerate}
\item Suppose that $A$ and $B$ are subsets of $\R$ and that $A\subset B$. Which of the following is correct, and why? ``any upper bound for $A$ is also an upper bound for $B$'', or; ``any upper bound of $B$ is also an upper bound for $A$.''
Now additionally suppose that $\sup(A)$ and $\sup(B)$ exist. Which inequality is right, and why? $\sup(A)\leq\sup(B)$, or; $\sup(B)\leq\sup(A)$.
\end{activity}
% \begin{exerc}[see also Abbott, ex 8.6.5(a)]
% Review the three-part definition of cut. Suppose that $A$ and $B$ are cuts, and let $S$ be the so-called sum-set of the two cuts, that is, let $S=\set{a+b\mid a\in A\text{ and }b\in B}$. Prove that $S$ is a cut by proving that each of the conditions in the definition of cut is satisfied.
%
% Why isn't the set $P=\set{ab\mid a\in A\text{ and }b\in B}$ a cut? What kinds of things do you have to do to fix this?
% \end{exerc}
% \question Consider the sequence defined by $x_1=0$ and $x_{n+1}=\sqrt{2+x_n}$. Prove using induction that for all $n$, $x_n<2$. Prove using induction that for all $n$, $x_n<x_{n+1}$.
% \question Give a detailed proof from the definition of supremum that the supremum of the interval $(0,1)$ is exactly $1$.
\vspace{\fill}
\begin{readnext}
Abbott \S 1.3, 1.4
\end{readnext}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{The Axiom of Completeness}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\begin{reading}
Abbott, \S 1.3, 1.4
\end{reading}
In the previous two sections we showed how one can construct the real number system using cuts. In this section we take a high-level approach and show how to work in the real number system using its properties alone. This will enable us to leave cuts behind for good.
\begin{definition}
An ordered field is said to satisfy the \emph{Axiom of Completeness} if every subset which has an upper bound in fact has a supremum.
\end{definition}
In the previous section, we saw that the ordered field $\Q$ \emph{does not} satisfy the Axiom of Completeness, since the set $A=\set{q\in\Q\mid q<0\text{ or }q^2<2}$ has no supremum in $\Q$. We also saw that the ordered field $\R$ \emph{does} satisfy the Axiom of Completeness. In fact, the following is true:
\begin{theorem}
The real number system $\R$ is an ordered field satisfying the Axiom of Completeness. Moreover, $\R$ is the unique such field up to isomorphism.
\end{theorem}
From this point on, we \emph{define} the real number system $\R$ to be the unique ordered field satisfying the Axiom of Completeness. It is therefore never necessary to speak of cuts again! In the future we may describe real numbers in many different ways: with cuts, with equations, with limits, with series, and of course with decimal expansions.
In the rest of this section, we take up the task of showing that we can use our new abstract definition of $\R$ to prove all the properties and facts we will need about the real number system. We begin with the following elementary result.
\begin{theorem}[Archimedean Principle]
For any real number $x$, there exists a natural number $n$ such that $x<n$. That is, while there are real numbers between integers, there aren't any real numbers beyond the integers to the right.
\end{theorem}
\begin{proof}
Suppose to the contrary that $x$ is a real number such that for all $n\in\N$ we have $n\leq x$. This says $x$ is an upper bound for $\N$, in particular it would mean $\N$ is bounded. By the Axiom of Completeness, $\N$ should have a least upper bound, call it $\alpha$.
Now, since $\alpha$ is the \emph{least} upper bound, $\alpha-1$ is not an upper bound for $\N$. Hence there exists $n\in\N$ such that $\alpha-1<n$. It follows that $\alpha<n+1$, contradicting that $\alpha$ is an upper bound for all natural numbers.
\end{proof}
\begin{center}
\begin{tikzpicture}
\draw[<->] (-1,0)--(8,0);
\node[label=above:{\small$\alpha-1$}] at (3,0) {$|$};
\node[label=below:{\small$n$}] at (3.8,0) {$|$};
\node[label=above:{\small$\alpha$}] at (5,0) {$)$};
\node[label=below:{\small$n+1$}] at (5.8,0) {$|$};
\end{tikzpicture}
\end{center}
As a consequence of this simple fact, we can now formally establish one fact we have previously stated without proof. While the rational number system is not sufficient to measure any distance, it is sufficient to \emph{approximate} any distance. In the following statement, we let $x<y$ and imagine that they are very close to one another.
\begin{theorem}[Rational density]
For any real numbers $x,y$, if $x<y$ then there exists $q\in\Q$ such that $x<q<y$.
\end{theorem}
\begin{proof}
Using the Archimedean Principle, we can find $n\in\N$ such that $n>1/(y-x)$. Intuitively, this guarantees that increments of size $1/n$ cannot step over the interval $(x,y)$.
Now let $m\in\N$ be the smallest natural number such that $x<m/n$ (again use the Archimedean Principle). Note that by this minimality we have $m/n-1/n\leq x$. Putting these equations together, we have
\begin{align*}
x&<m/n\\
&\leq x+1/n\\
&<x+(y-x)=y
\end{align*}
Thus the rational number $q=m/n$ lies in the interval $(x,y)$.
\end{proof}
To see that this implies we can approximate real numbers by rational numbers as closely as we like, let $x$ be a given real number and apply the rational density theorem to intervals of the form $(x-\epsilon,x+\epsilon)$ where $\epsilon$ is small.
As an application of the rational density theorem, we can finally calculate the following promised supremum.
\begin{proposition}
Let $A=\set{q\in\Q\mid q<0\text{ or }q^2<2}$ and let $\beta=\sup(A)$. Then $\beta^2=2$, that is, $\beta=\sqrt2$.
\end{proposition}
\begin{proof}
If $\beta^2<2$, then using the calculation from Theorem~\ref{thm:root-2-cut}, we can find $\epsilon>0$ such that $(\beta+\epsilon)^2<2$. By rational density, we can find a rational number $q$ such that $\beta<q<\beta+\epsilon$. Then $\beta<q$ and $q^2<2$, contradicting that $\beta$ is an upper bound of $A$.
If $\beta^2>2$, we can reach a contradiction in a similar way. Thus we must have $\beta^2=2$, as desired.
\end{proof}
We close this section with a characterization of the supremum of a set that will be useful in the future.
\begin{lemma}
If $A$ is a set, then $\beta$ is the supremum of $A$ if and only if the following two conditions are satisfied:
\begin{itemize}
\item ($\beta$ is an upper bound) for all $a\in A$ we have $a\leq\beta$, and
\item ($\beta$ is the least such) for all $\epsilon>0$ there exists $a\in A$ such that $\beta-\epsilon\leq a$.
\end{itemize}
\end{lemma}
You can find the proof in Lemma~1.3.8 of the text.
\newpage
\subsection*{Activity for \S\thesection}
\begin{activity}
\item (Abbott, exercise 1.4.5) Prove the statement known as \emph{irrational density}: for every two real numbers $a$ and $b$ with $a<b$, there exists an \emph{irrational} number $r$ satisfying $a<r<b$.
\vspace{\fill}
\item Recall that the \emph{infemum} of a set is the greatest lower bound. Write a characterization of infemum in the style of the $\epsilon$-characterizaton of supremum.
\end{activity}
% \begin{exerc}
% Give a detailed proof from the definition of supremum that the supremum of the interval $(0,1)$ is exactly $1$.
% \end{exerc}
%
% \begin{exerc}
% Read the characterization of supremum described in Abbott, Lemma~1.3.8. Then, give a thorough proof that $1/2$ is the supremum of the set $\set{n/(2n+1)\mid n\in\N}$. Point out the exact moment in the proof when the Archimedean Principle is used.
% \end{exerc}
%
% \begin{exerc}
% Consider the sequence defined by $x_1=0$ and $x_{n+1}=\sqrt{2+x_n}$. Prove using induction that for all $n$, $x_n<2$. Prove using induction that for all $n$, $x_n<x_{n+1}$.
% \end{exerc}
\vspace{\fill}
\begin{readnext}
Abbott, \S 1.5
\end{readnext}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Completeness and the size of the reals}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\begin{reading}
Abbott, \S 1.5
\end{reading}
In this section we will continue with consequences of the completeness axiom. In particular we will establish one of the most surprising and historically disruptive consequences: the fact that there are more irrational numbers than there are rational numbers! In order to make this statement precise, we need to introduce the notion of countable and uncountable sets.
\begin{definition}
A set $A$ is \emph{countable} if it can be enumerated in a sequence $a_n$, whose items are indexed by natural numbers $n$. Intuitively, $A$ is countable if all its elements can be ordered up in an infinite shopping list:
\begin{align*}
1.&\quad\underline{\hspace{1in}}\\
2.&\quad\underline{\hspace{1in}}\\
3.&\quad\underline{\hspace{1in}}\\
\vdots&
\end{align*}
\end{definition}
\begin{example}
The set of natural numbers $\N$ is countable, as is any subset of $\N$.
\end{example}
\begin{example}
The set of integers $\Z$ is countable. To see this, one can simply alternate between positive and negative integers: $0,+1,-1,+2,-2,+3,-3,\ldots$.
\end{example}
\begin{example}
The set of rational numbers $\Q$ is countable. To see this, let us first show that the set of positive rational numbers $\Q_+$ is countable. Although the presences of the numerator and denominator make it seem like we would need infinitely many infinite lists, we can get around that by fixing the sum of the two and then letting it increase: begin with $\frac11$, then do $\frac12,\frac21$, then $\frac13,\frac22,\frac31$, then $\frac14,\frac23,\frac32,\frac41$, and so forth. The list does have duplicates such as $\frac11$ and $\frac22$, but this is allowed, as we could in principle run through and delete them if we wanted. Finally, we can add in the negative rationals by applying the same method as we did with $\Z$.
\end{example}
The following result can be handy for showing sets are countable.
\begin{lemma}
\begin{itemize}
\item If $A,B$ are countable, then the union $A\cup B$ is countable.
\item If $A_1,A_2,\ldots$ is a sequence of countable sets, then $\bigcup A_n$ is countable.
\end{itemize}
\end{lemma}
\begin{proof}
For the first statement, since $A$ is countable we can enumerate its elements $a_1,a_2,a_3,\ldots$ and since $B$ is countable we can enumerate its elements $b_1,b_2,b_3,\ldots$. It follows that we can enumerate the elements of $A\cup B$ in the sequence $a_1,b_1,a_2,b_2,a_3,b_3\ldots$. (Formally we are describing $c_n$ where $c_{2n}=b_n$ and $c_{2n-1}=a_n$.) Thus $A\cup B$ is countable.
We leave the proof of the second statement as an exercise.
\end{proof}
To show how this lemma can come in handy, we can use it recast the proof that $\Z$ is countable. It is obvious that the set $\Z_+$ of nonnegative integers is countable, and ditto for the negative integers $\Z_-$. Since $\Z$ is the union of the two sets, it follows that $\Z$ is countable.
We now come to the ``surprising'' fact hinted at above.
\begin{theorem}
\label{thm:reals-uncountable}
The set of real numbers $\R$ is \textbf{not} countable.
\end{theorem}
In order to prove this theorem, we need the following consequence of the Axiom of Completeness for the real numbers called the \emph{nested interval property}. This is a fact which we'll also use a few more times throughout the course.
\begin{theorem}[Nested interval property]
Suppose that $I_n$ is a sequence of closed and bounded intervals, that is, $I_n=[a_n,b_n]$. Moreover suppose that the sequence is nested, that is, $I_{n+1}\subset I_n$. Then we have $\bigcap I_n\neq\emptyset$.
\end{theorem}
\begin{proof}
Notice that the set of left endpoints $\set{a_n\mid n\in\N}$ is bounded above, because any of the $b_n$'s is an upper bound. By the Axiom of Completeness, the set $\set{a_n\mid n\in\N}$ has a least upper bound, call it $\alpha$.
Then given any $n$, we have $a_n\leq\alpha$, since $\alpha$ is an upper bound for the set $\set{a_n\mid n\in\N}$. On the other hand, we also have $\alpha\leq b_n$, since $\alpha$ is the \emph{least} upper bound for the set $\set{a_n\mid n\in\N}$. It follows that $\alpha$ lies in the interval $I_n=[a_n,b_n]$. Since $n$ was arbitrary, we conclude that $\alpha$ lies in $\bigcap I_n$. This completes the proof.
\end{proof}
We are now ready to prove that the set of real numbers is uncountable.
\begin{proof}[Proof of Theorem~\ref{thm:reals-uncountable}]
Suppose towards a contradiction that $\R$ is countable. Then its elements can be enumerated in a sequence $x_n$. Now inductively construct a nested sequence of closed, bounded intervals as follows. Let $I_1$ be any closed, bounded interval not containing $x_1$. Let $I_2$ be any closed subinterval of $I_1$ not containing $x_2$. Inductively, let $I_n\subset I_{n-1}$ be any closed interval not containing $x_n$. This completes the construction.
Now by the NIP, there exists an element $\alpha\in\R$ such that $\alpha\in\bigcap I_n$. Now for each $n$, we have both $x_n\notin I_n$ and $\alpha\in I_n$, which implies that $\alpha\neq x_n$. This contradicts the claim that every element of $\R$ was enumerated in the sequence $x_n$!
\end{proof}
\begin{corollary}
The set of irrational numbers is uncountable.
\end{corollary}
\begin{proof}
Suppose towards a contradiction that the set of irrational numbers is countable. Then $\R=\Q\bigcup(\R\setminus\Q)$ is countable, being the union of two countable sets. This contradicts Theorem~\ref{thm:reals-uncountable}.
\end{proof}
\newpage
\subsection*{Activity for\S \thesection}
\begin{activity}
\item Let $A$ be the set of integer lattice points in the plane, that is, $A=\set{(x,y):x,y\in\Z}$. Draw a picture of the set $A$. Prove $A$ is countable by showing how to write all its elements in a list. (Write the first 10+ elements, explain how you would continue, and explain why the completed list would include every element.)
\vspace\fill
\item Find a nested sequence of nonempty bounded intervals with empty intersection.
\vspace\fill
\item Find a nested sequence of nonempty closed intervals with empty intersection.
\end{activity}
% \begin{exerc}
% Show that the set of Gaussian integers is countable. Recall that the Gaussian integers is the set of all complex numbers of the form $a+bi$ where $a,b\in\Z$.
% \end{exerc}
%
% \begin{exerc}
% Suppose that for every $n$, the set $A_n$ is countable. Show that the set $A=\bigcup A_n$ is countable. [See the hint described in Exercise~1.5.3(c).]
% \end{exerc}
%
% \begin{exerc}
% Show that the set $\N^3$ is countable. Here $\N^3$ consists of all triples of the form $(a,b,c)$, where $a,b,c\in\N$. [Hint: use the previous two exercises.]
% \end{exerc}
%
% \begin{exerc}
% \begin{enumerate}
% \item Find a nested family of nonempty bounded intervals with empty intersection.
% \item Find a nested family of nonempty closed intervals with empty intersection.
% \end{enumerate}
% \end{exerc}
\vspace{\fill}
\begin{readnext}
Abbott, \S 1.6
\end{readnext}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{More on countability and uncountability}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\begin{reading}
Abbott, \S 1.6
\end{reading}
In the previous section we saw that among the infinite sets, some of them are not too big (countable sets) while others are much larger (uncountable sets). Showing that a set is countable requires a direct argument---write down how to enumerate all its elements in a sequence. Showing that a set is uncountable often requires an indirect argument---assume the set is countable and find a contradiction.
We were able to show that $\R$ is uncountable using the Nested Interval Property. But that proof was very particular to the set of real numbers. How would we show that some other given set is uncountable? It turns out there is another and more systematic way to prove that $\R$ is uncountable, due the mathematician Cantor. The method is called \emph{diagonalization}, and uses the decimal digit system.
\begin{theorem}
The unit interval $(0,1)$ is uncountable.
\end{theorem}
\begin{proof}
Suppose towards a contradiction that $(0,1)$ is countable. Then the elements of $(0,1)$ can all be enumerated in a sequence $a_n$. We need to analyze the digits of each entry of this sequence, so let us fix some notation for them:
\begin{align*}
a_1&=0.a_{11}\,a_{12}\,a_{13}\cdots\\
a_2&=0.a_{21}\,a_{22}\,a_{33}\cdots\\
a_3&=0.a_{31}\,a_{32}\,a_{33}\cdots\\
\vdots&
\end{align*}
We now define a ``diagonal'' real number $d\in(0,1)$ with a decimal expansion $d_n$ with the property that for all $n$
\[d_n\neq a_{nn}
\]
We can even give an explicit definition of the digits $d_n$ that achieves this property. For instance we may let $d_n=2$ when $a_{nn}\neq2$ and let $d_n=3$ when $a_{nn}=2$. It then follows that for all $n$, we have $d\neq a_n$. Thus $d$ is an element of $(0,1)$ which does not appear in the sequence $a_n$, contradicting our assumption that every element of $(0,1)$ appears in the sequence $a_n$.
\end{proof}
In the exercises you will be asked to use diagonalization on other kinds of sets.
The dichotomy between countable and uncountable sets suggests that one should further investigate the possibilities for the size of an infinite set. Recall that for finite sets we write $|A|$ for the number of elements of $A$. However for infinite sets this definition of $|A|$ isn't a good one anymore. What is needed is the following.
\begin{definition}
Suppose that $A,B$ are sets. Then we introduce the following notation:
\begin{itemize}
\item We write $|A|\leq|B|$ if there is a one-to-one function $A\to B$
\item We write $|A|=|B|$ if there exists a one-to-one, onto function $A\to B$
\item We write $|A|<|B|$ if we have $|A|\leq|B|$ and $|A|\neq|B|$.
\end{itemize}
\end{definition}
This definiton is consistent with the ``number of elements'' definition of $|A|$ when $A$ is finite. However, while the terms ``number of elements'', ``more'', and ``fewer'' all lose their meaning for infinite sets, the notions of one-to-one and onto functions remain meaningful.
We call $|A|$ the \emph{cardinality} of $A$. It is important to understand that we never defined what $|A|$ actually is, only how it compares with other set cardinalities $|B|$. Using set theory we can also define what $|A|$ actually is, but it isn't necessary for this class.
Cantor's theorem can be generalized to show that for any set there is a set of even larger cardinality. Therefore there is no bound on the number of different sizes of infinity! Recall that if $A$ is any set, the \emph{power set} $\mathcal P(A)$ is the set of all subsets of $A$. Of course if $A$ is finite, then we have $|\mathcal P(A)|=2^{|A|}$ so in particular $|A|<|\mathcal P(A)|$. It turns out that this works for infinite sets too.
\begin{theorem}
If $A$ is any set, then $|A|<|\mathcal P(A)|$.
\end{theorem}
\begin{proof}
To show that $|A|\leq|\mathcal P(A)|$ we must exhibit a one-to-one function $A\to\mathcal P(A)$. For this, we can use the function $i(a)=\{a\}$.
To show that $|A|\neq|\mathcal P(A)|$, we have to show that there is no one-to-one, onto function $f\colon A\to\mathcal P(A)$. For this, suppose that $f\colon A\to\mathcal P(A)$ is any function. Then we consider the ``diagonal'' set $D$:
\[D=\set{a\in A\mid a\notin f(a)}
\]
We claim that $D$ is not in the range of $f$, and therefore that $f$ is not onto. To see this, suppose towards a contradiction that there exists $a_0\in A$ such that $D=f(a_0)$. Then by the definition of $D$, we have that $a_0\in D$ iff $a_0\notin f(a_0)$. And since $D=f(a_0)$ we can write this as $a_0\in D$ iff $a_0\notin D$, which is a clear contradiction.
\end{proof}
It is natural to ask to what extent cardinalities behave like ordinary finite counting numbers. For instance, is it possible to have two sets which are \emph{incomparable} in size, that is $|A|\not\leq|B|$ and $|B|\not\leq|A|$? If we assume all of the standard axioms about sets (including the axiom of choice) then this situation is impossible.
\begin{theorem}
If $A,B$ are any sets, then exactly one of the following is true: $|A|<|B|$, $|A|=|B|$, or $|B|<|A|$.
\end{theorem}
Another natural question is whether the cardinality of $\R$ is the \emph{smallest} uncountable cardinality. Many very narrow or thin subsets of $\R$ still have the full cardinality of $\R$. Just consider $(0,1)$ or $(0,\epsilon)$ or the set of irrationals. So the question is whether for any set $A$ of real numbers, if $A$ is uncountable then is it necessarily true that $|A|=|\R|$?
This question is known as the \emph{continuum hypothesis} and has great historical significance. After remaining open for many decades, it was finally settled in a surprising way: the answer is \emph{undecidable} using the accepted axioms of mathematics! In other words, there is a universe of mathematics in which the continuum hypothesis is true, and also a universe of mathematics in which it is false.
Since this stunning resolution to the continuum hypothesis was found in 1965, the occurrence and nature of undecidable statements has remained of key interest to mathematicians.
\newpage
\subsection*{Activity for \S \thesection}
In the following, let $S_{\infty}$ be the set of all \emph{infinite} words made with the letters $a,b$. So $S_\infty$ contains elements like $aaaa\cdots$, $ababa\cdots$, $aabaabaab\cdots$, also non-repeating strings like $abaabaaabaaaab\cdots$, and even strings with no discernible pattern whatsoever.
\begin{activity}
\item Write any ten elements of $S_{\infty}$. Include at least ten letters in each element (with the understanding that the words continue indefinitely, either according to a pattern or arbitrarily).
\begin{align*}
s_1 & = \underline{\hspace{3in}\cdots} \\
s_2 & = \underline{\hspace{3in}\cdots} \\
s_3 & = \underline{\hspace{3in}\cdots} \\
s_4 & = \underline{\hspace{3in}\cdots} \\
s_5 & = \underline{\hspace{3in}\cdots} \\
s_6 & = \underline{\hspace{3in}\cdots} \\
s_7 & = \underline{\hspace{3in}\cdots} \\
s_8 & = \underline{\hspace{3in}\cdots} \\
s_9 & = \underline{\hspace{3in}\cdots} \\
s_{10}& = \underline{\hspace{3in}\cdots} \\
\end{align*}
\item Write the first ten letters of the \emph{diagonal} word $d$ with the property that for all $n$, $d$ disagrees with $s_n$ on the $n$th letter.
\item Write the definition of $d$ in the abstract (not just in this example).
\[d=\begin{cases}
\underline{\hspace{.5in}} & \text{if }s_{nn}=\underline{\hspace{.5in}}\\
\underline{\hspace{.5in}} & \text{if }s_{nn}=\underline{\hspace{.5in}}
\end{cases}
\]
\item Write out the proof that $S_{\infty}$ is uncountable.
\end{activity}
% Prove that R and P(N) have the same cardinality.
% (See also Abbott, ex 1.5.9). Recall that a real number $\alpha$ is called \emph{algebraic} if there exists a polynomial $p$ with integer coefficients such that $p(\alpha)=0$. Is the set $A$ of all algebraic numbers countable or uncountable? Prove your assertion.
\vspace\fill
\begin{readnext}
Abbott, \S 2.1
\end{readnext}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Sequences and series}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Infinity and arithmetic}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\label{sec:sequences}
\begin{reading}
Abbott, \S 2.1
\end{reading}
A sequence of real numbers consists of one real number $a_n$ for each natural number $n$. Officially a sequence is a \emph{function} from $\N$ to $\R$. However sequences are so special that we rarely use the function notation $a(n)$ and instead prefer the traditional notation $a_n$. We have already used sequences, for instance in the definition of a countable set. We will see later in the course that sequences play a huge role in analysis. In this section we give further motivation for studying sequences and the sibling concept of series.
To begin consider the simple addition problem $\pi+e$, which we write out in decimal notation as
\begin{align*}
&3.14159265\cdots\\
{}+&2.71828182\cdots
\end{align*}
It is unclear how to begin this problem. Our usual method of carrying would have us start at the right-most digit, but there isn't one! It is possible to handle this issue by regarding $\pi$ and $e$ as cuts. But it is even more elegant to regard each of $\pi$ and $e$ as the limit of a sequence, and perform the addition term-by-term.
Another problem with addition is the prospect of adding together infinitely many numbers at once. Sometimes this clearly isn't possible; for example the sum
\[1+1+1+1+\cdots
\]
is infinite and has no value in the real numbers. On the other hand the familiar Zeno paradox asks whether the sum
\[1+\frac12+\frac14+\frac18+\cdots
\]
has any value in the real numbers, and we usually agree that it has the value $2$. It seems that sometimes it makes sense to add together infinitely many numbers, and other times it doesn't.
But this conclusion deserves a great deal of scrutiny. To see why, consider the following simple proof that the Zeno series has a value of $2$. We first notice that
\begin{align*}
\left(1+\frac12+\frac14+\frac18+\cdots\right)-1
&=\frac12+\frac14+\frac18+\cdots\\
&=\frac12\left(1+\frac12+\frac14+\cdots\right)
\end{align*}
Thus we see that if $x=1+\frac12+\frac14+\frac18+\cdots$, then $x$ satisfies the equation $x-1=\frac12x$ and therefore $x=2$.
Now let us apply this proof to the series $1+2+4+8+\cdots$. Again write
\begin{align*}
(1+2+4+8+\cdots)-1&=2+4+8+\cdots\\
&=2(1+2+4+\cdots)
\end{align*}
This time we conclude that if $x=1+2+4+8+\cdots$ then $x$ satisfies the equation $x-1=2x$ and therefore that $x=-1$, which is nonsense! (Or is it?)
Finally, even if we do accept some infinite sums as valid, we still should decide whether the familiar laws of associativity and commutativity of addition are applicable to the terms of infinite sums. To see the potential difficulty, consider two ways of grouping the terms of the same infinite summation:
\begin{align*}
S_1&=(1-1)+(1-1)+(1-1)+\cdots\\
S_2&=1+(-1+1)+(-1+1)+\cdots
\end{align*}
Using the first grouping, we obtain a value of $0$. Using the second grouping, we obtain a value of $1$.
Commutativity has similar issues. For example, consider two ways of summing the same series:
\begin{align*}
S_1&=1-\frac12+\frac13-\frac14+\frac15-\frac16+-\cdots\\
S_2&=1+\frac13-\frac12+\frac15+\frac17-\frac14++-\cdots
\end{align*}
In the second summation we first add two of the positive terms before then adding one of the negative terms. A quick approximation with a calculator reveals that $S_1\approx0.69$ while $S_2\approx1.03$. This time we conclude that even if infinite summations are possible, one must be very careful when applying the usual laws of arithmetic to them!
\newpage
\subsection*{Activity for \S \thesection}
\begin{activity}
\item Consider the following infinite summation:
\[1-\frac12+\frac13-\frac14+\frac15-+\cdots
\]
\begin{itemize}
\item Try to reorder the terms to make the sum as close as possible to $0.5$. Use at least 15 terms to be sure you are actually representing your pattern.
\item Try to reorder the terms to make the sum as close as possible to $1$.
\item Try to reorder the terms to make the sum as close as possible to $1.5$.
\end{itemize}
Were you (approximately) successful in each case? If so, write down your pattern and result. If not, what barrier did you encounter?
\vspace{2in}
\item This time consider the following infinite summation:
\[1-\frac12+\frac14-\frac18+\frac{1}{16}-+\cdots
\]
\begin{itemize}
\item Try to reorder the terms to make the sum as close as possible to $0.5$. Use at least 15 terms to be sure you are actually representing your pattern.
\item Try to reorder the terms to make it come as close as possible to $1$.
\item Try to reorder the terms to make it come as close as possible to $1.5$.
\end{itemize}
Were you (approximately) successful in each case? If so, write down your pattern and result. If not, what barrier did you encounter?
\end{activity}
\vspace\fill
\begin{readnext}
Abbott, \S 2.2
\end{readnext}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Convergence}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\begin{reading}
Abbott, \S 2.2
\end{reading}
Recall that a \emph{sequence} is a list of one real number $a_n$ for every natural number $n$. The most important characteristic of a sequence is whether it converges or diverges. Intuitively a sequence $a_n$ converges to a limit point $L$ if the sequence eventually comes as close to $L$ as one could ever wish.
\begin{definition}
The sequence $a_n$ \emph{converges} to $L$ if the following holds:
\[(\forall\epsilon>0)\;(\exists N\in\N)\;(\forall n\geq N)\;|a_n-L|<\epsilon
\]
In this case we write both $\lim a_n=L$ and also $a_n\to L$, and we say that $L$ is the \emph{limit} of the sequence.
\end{definition}
Since the definition is both of central importance and also somewhat subtle, we will analyze it in great detail.
\begin{itemize}
\item The $(\forall\epsilon>0)$ clause. We think of $\epsilon$ as a very small \emph{tolerance}. Thus the definition of convergence begins by saying that for all possible tolerances, something will happen.
\item The $(\exists N\in\N)\;(\forall n\geq N)$ clause. This means that \emph{eventually} in the timeline of the sequence, something will become true, and then continue to be true forever.
\item The $|a_n-L|<\epsilon$ clause. This means that $a_n$ lies within a small \emph{neighborhood} of $L$. Formally, if $x$ is any real number and $\epsilon>0$, then the \emph{$\epsilon$-neighborhood} of $x$, denoted $N_\epsilon(x)$, is the open interval centered at $x$ of radius $\epsilon$. In symbols:
\[N_\epsilon(x)=(x-\epsilon,x+\epsilon)=\set{y\in\R\mid|y-x|<\epsilon}
\]
Thus the conclusion of the definition of convergence says precisely that $a_n$ lies within the $\epsilon$-neighborhood of $L$.
\end{itemize}
To summarize, the statement that $a_n$ converges to $L$ says:
\begin{quotation}
``For all tolerances $\epsilon>0$, eventually, the terms of the sequence lie in the $\epsilon$-neighborhood of $L$.''
\end{quotation}
So how does one show that a given sequence converges to a specified limit? One way to think of verifying the three-quantifier definition holds true is to play a three-step game. Your enemy takes the first turn and gives you some (small) value of $\epsilon$. You respond by calculating a (large enough) natural number $N$. Your enemy then responds by choosing a number $n\geq N$. Finally, you win if you can show that $|a_n-L|<\epsilon$, or in other words, $L-\epsilon<a_n<L+\epsilon$.
\begin{example}
Let us confirm from the definition that $\lim1/\sqrt{n}=0$. Suppose that $\epsilon$ is given to us. We must find a value $N$ (depending on this $\epsilon$) with the property that whenever $n\geq N$ we have $-\epsilon<1/\sqrt{n}<\epsilon$. Since $1/\sqrt{n}$ is always positive, we only have to worry about the second inequality.
To get an idea of how this is going to work, if $\epsilon=.1$ then we would need to take $N=101$. Similarly if $\epsilon=.01$ then we would need to take $N=10,001$. In general we can figure out the needed relationship by working backwards. We want
\begin{align*}
1/\sqrt{n}<\epsilon &\iff \sqrt{n}>1/\epsilon\\
&\iff n>1/\epsilon^2
\end{align*}
This tells us we should choose $N>1/\epsilon^2$. Technically $N$ has to be a natural number, but we can just take $N$ to be the first natural number $N>1/\epsilon^2$.
Finally we can write the proof forwards as follows. Given any $\epsilon>0$, we let $N$ be any natural number such that $N>1/\epsilon^2$. Then for all $n\geq N$ we have
\begin{align*}
n>1/\epsilon^2&\implies\sqrt{n}>1/\epsilon\\
&\implies1/\sqrt{n}<\epsilon
\end{align*}
Since we always have $-\epsilon<1/\sqrt{n}$, the proof is complete.\qed
\end{example}
\begin{example}
Let us confirm from the definition that $n/(n+1)\to1$. So let $\epsilon$ be given; we wish to conclude that eventually $1-\epsilon<n/(n+1)<1+\epsilon$. Since $n/(n+1)$ is always less than $1$, we only have to worry about the first inequality. To see the needed value of $N$ we work backwards:
\begin{align*}
1-\epsilon<n/(n+1)&\iff 1-n/(n+1)<\epsilon \\
&\iff 1/(n+1)<\epsilon \\
&\iff n+1>1/\epsilon \\
&\iff n>1/\epsilon-1
\end{align*}
Thus the needed value of $N$ is $1/\epsilon-1$. To simplify the calculation, it will be good enough to take $N>1/\epsilon$.
Now to write the proof forwards, let $\epsilon$ be given and let $N$ be a natural number such that $N>1/\epsilon$. Then for all $n\geq N$ we have
\begin{align*}
1-n/(n+1) &= 1/(n+1)\\
&< 1/n\\
&\leq 1/N
&<\epsilon
\end{align*}
Thus $1-\epsilon < n/(n+1)$, as desired.\qed
\end{example}
So far we have discussed only convergence. What happens when a sequence $a_n$ does \emph{not} converge to $L$? Formally negating the definition of convergence, this says
\[(\exists\epsilon>0)\;(\forall N\in\N)\;(\exists n\geq N)\;|a_n-L|\geq\epsilon
\]
Intuitively this says that there is some tolerance $\epsilon$ such that the sequence frequently lies outside the $\epsilon$-neighborhood of $L$.
\begin{example}
Consider the following sequence with the terms, and discuss whether it converges to $0$ or not:
\[1,-\frac12,+\frac13,-\frac14,+\frac15,-\frac15,+\frac15,-\frac15,\ldots
\]
If $\epsilon$ has a given value of $.5$ it is possible to find a suitable response $N$ which satisfies the definition of convergence. The same is true if $\epsilon=.25$. But if $\epsilon=.2$ then it is \emph{not} possible, so we conclude that the sequence does not converge to $0$
\end{example}
In general, we say that a sequence \emph{diverges} if for every $L$, the sequence does not converge to $L$.
In this section we made a formal definition of limit, and used it to verify several routine instances of convergence and non-convergence. This may make it seem that the purpose of the definition is to make routine calculations. Instead the opposite is true: we make routine calculations to help affirm that the definition is a good one.
\newpage
\subsection*{Activity for \S \thesection}
\begin{activity}
\item In each case, prove the given sequence converges to the proposed limit directly from the definition of convergence.
\begin{enumerate}\itemsep\fill
\item $\displaystyle\lim\frac{1}{6n^2+1}=0$
\item $\displaystyle\lim\frac{3n-1}{2n+5}=\frac32$
\item $\displaystyle\lim\frac{\sin(n)}{n}=0$
\item $\displaystyle\lim2^{1/n}=1$
\end{enumerate}
\vspace\fill
\item Prove the given sequence \emph{does not} converge to the proposed limit directly from the \emph{negation} of the definition of convergence.
\begin{center}
Sequence: $0.1,0.01,0.1,0.01,0.1,0.01,\ldots$. Proposed limit: $0$.
\end{center}
\end{activity}
\vspace\fill
\begin{readnext}
Abbott, \S 2.3
\end{readnext}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Limit calculus}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\begin{reading}
Abbott, \S 2.3
\end{reading}
In the previous section we gave a rigorous definition of the limit of a sequence, and checked that it corresponds with our intuition for several special examples. In this section we begin our study of \emph{calculus}, that is, the laws or rules for calculating limits, derivatives, and integrals. Of course calculus usually focuses on the derivatives and integrals, but before we can get there we must first understand limits.
Before we begin recall the \emph{triangle inequality} states that for any $x,y$ we have $||x|-|y||\leq|x-y|\leq|x|+|y|$. The triangle inequality is actually best visualized in two (or more) dimensions, where it says that if $x$ is $|x|$ units from the origin, and $y$ is $|y|$ units from the origin, then the distance between $x$ and $y$ cannot be less than $||x|-|y||$, and cannot be more than $|x|+|y|$. Since the sign of $y$ doesn't matter, we can write:
\[||x|-|y||\leq|x\pm y|\leq|x|+|y|
\]
\begin{figure}[h]
\centering
\begin{tikzpicture}
\draw (0,0) circle (.8);
\draw (0,0) circle (1.3);
\node[circle,draw,inner sep=1] at (0,0) {};
\node[circle,fill,draw,inner sep=1,label=left:$x$] (x) at (30:.8) {};
\node[circle,fill,draw,inner sep=1,label=right:$y$] (y) at (30:1.3) {};
\draw (x)--(y);
\end{tikzpicture}
\quad
\begin{tikzpicture}
\draw (0,0) circle (.8);
\draw (0,0) circle (1.3);
\node[circle,draw,inner sep=1] at (0,0) {};
\node[circle,fill,draw,inner sep=1,label=right:$x$] (x) at (30:.8) {};
\node[circle,fill,draw,inner sep=1,label=left:$y$] (y) at (30:-1.3) {};
\draw (x)--(y);
\end{tikzpicture}
\caption{Intuitive justification for the triangle inequality. On the left, $x$ and $y$ are as close together as possible at $||x|-|y||$ units. On the right, $x$ and $y$ are as far apart as possible at $|x|+|y|$ units.}
\end{figure}
We can now state the limit calculus laws. The proof is long, but the method of proof is very important and worth exploring at least until it becomes familiar (and somewhat tedious).
\begin{theorem}[Limit calculus]
Suppose that $\lim a_n=a$ and $\lim b_n=b$. Then we have
\begin{enumerate}
\item $\lim a_n+b_n=a+b$
\item $\lim ka_n=ka$ for any $k\in\R$
\item $\lim a_nb_n=ab$
\item $\lim a_n/b_n=a/b$ provided $b_n,b\neq 0$
\item $\lim \sqrt{a_n}=\sqrt{a}$ provided $a_n\geq0$
\end{enumerate}
\end{theorem}
\begin{proof}
\begin{enumerate}
\item This is our first example of a so-called ``$\epsilon/2$-argument.'' If $\epsilon>0$ is given, we cleverly apply the convergence of $a_n$ and $b_n$ to the tolerance value $\epsilon/2$. Thus there exists an $N_1$ such that for all $n\geq N_1$ we have $|a_n-a|<\epsilon/2$, and there exsits an $N_2$ such that for all $n\geq N_2$ we have $|b_n-b|<\epsilon/2$. Thus if $n\geq\max(N_2,N_2)$, we have
\begin{align*}
|(a_n+b_n)-(a+b)|&=|(a_n-a)+(b_n-b)|\\
&\leq|a_n-a|+|b_n-b|\\
&<\epsilon/2+\epsilon/2\\
&=\epsilon
\end{align*}
In other words, we use the triangle inequality to \emph{split} the distance between $a_n+b_n$ and $a+b$ into two portions, each of which we can bound.
\item This is similar but easier; we just use an $\epsilon/|k|$ argument. Given $\epsilon$, we can find $N$ such that for all $n\geq N$ we have $|a_n-a|<\epsilon/|k|$. It follows that $|ka_n-ka|<\epsilon$, as desired.
\item Given $\epsilon$, we wish to show that eventually we have $|a_nb_n-ab|<\epsilon$. To see what we will need, we calculate
\begin{align*}
|a_nb_n-ab|&\leq|a_nb_n-ab_n+ab_n-ab|\\
&\leq|a_nb_n-ab_n|+|ab_n-ab|\\
&=|a_n-a||b_n|+|a||b_n-b|
\end{align*}
Since $b_n$ converges, we can find a bound $M$ such that $|b_n|\leq M$ for all $n$ (see below). Now since $a_n$ and $b_n$ converge, we can find an $N$ large enough so that for $n\geq N$ we have both $|a_n-a|<\epsilon/2M$ and $|b_n-b|<\epsilon/2|a|$. Together with the calculation above, we thus conclude that $|a_nb_n-ab|<\epsilon$, as desired.
\item[(d),(e)] The proof of (d) is similar to (c), and the proof of (e) is left as an exercise.
\end{enumerate}
\end{proof}
\begin{example}
We can now verify that $\frac{n}{n+1}\to1$ as follows:
\begin{align*}
\lim\frac{n}{n+1}&=\frac1{\lim\frac{n+1}{n}}\\
&=\frac1{\lim 1+1/n}\\
&=\frac1{1+\lim 1/n}\\
&=1
\end{align*}
\end{example}
Before we continue, we owe
\begin{definition}
The sequence $a_n$ is \emph{bounded} if there exists some $M\in\R$ such that for all $n$ we have $|a_n|\leq M$.
\end{definition}
It is equivalent to say that $a_n$ is bounded if there exists an interval $[a,b]$ such that $\{a_n\}\subset[a,b]$.
\begin{proposition}
If $a_n$ is convergent, then $a_n$ is bounded.
\end{proposition}
\begin{proof}
Suppose $a_n\to L$, so that for all $\epsilon$ we can find $N$ such that $|a_n-L|<\epsilon$ whenever $n>N$. In particular whenever $n\geq N$ we have that $|a_n|\leq|a_n-L|+|L|<\epsilon+|L|$. Thus we may take $M_0=\epsilon+|L|$ as a bound for these terms.
Meanwhile all that remains of the sequence is the finite initial segment $a_1,\ldots,a_N$. Letting $M_1=\max(|a_1|,\ldots,|a_N|)$, we of course have for all $n\leq N$ that $|a_n|\leq M_1$. It follows that the full sequence is bounded by the constant $M=\max(M_0,M_1)$.
\end{proof}
The converse of this Proposition is false, with a simple example being the sequence with terms $+1,-1,+1,-1,\ldots$.