-
Notifications
You must be signed in to change notification settings - Fork 1
/
Copy pathnotes.tex
2623 lines (1926 loc) · 218 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[Category, Measure, and Forcing]{Category, Measure, and Forcing\\\smaller Lecture Notes}
\author{Samuel Coskey}
\author{Based on notes by Erik Holmes}
\usepackage{amsmath,amsthm}
\usepackage[vscale=.8,vmarginratio=4:3]{geometry}
\usepackage{mathpazo}
\usepackage{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{\Q}{\mathbb Q}
\newcommand{\R}{\mathbb R}
\newcommand{\BB}{\mathbb B}
\newcommand{\PP}{\mathbb P}
\newcommand{\Null}{\mathcal N}
\newcommand{\Meager}{\mathcal M}
\newcommand{\Ksigma}{\mathcal K_\sigma}
\newcommand{\forces}{\Vdash}
\DeclareMathOperator{\dom}{dom}
\DeclareMathOperator{\cod}{cod}
\DeclareMathOperator{\rng}{rng}
\DeclareMathOperator{\cl}{cl}
\DeclareMathOperator{\rk}{rk}
\DeclareMathOperator{\add}{\mathsf{add}}
\DeclareMathOperator{\non}{\mathsf{non}}
\DeclareMathOperator{\cov}{\mathsf{cov}}
\DeclareMathOperator{\cof}{\mathsf{cof}}
\DeclareMathOperator{\Diff}{Diff}
\DeclareMathOperator{\Con}{Con}
\DeclareMathOperator{\Fn}{Fn}
\DeclareMathOperator{\Nonnull}{Nonnull}
\theoremstyle{definition}
\newtheorem{exercise}{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*{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}}
\begin{document}
\frontmatter
\maketitle
\tableofcontents
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter*{Foreword}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
These notes began when the author Samuel Coskey supervised Erik Holmes in an independent studen course at Boise State University. The subject was cardinal characteristics of the continuum. In his work for the course, Erik wrote a set of notes on the subject. Afterwards, while teaching a graduate course in set theory, Samuel adapted and extended Erik's notes for use as lecture notes, resulting in the present document.
We would like to acknowledge the support of Boise State University, and the Math 522 students who have provided feedback: Rhett Barton, Kyle Beserra, Kayla Krakoff, Joshua Meier, Stephanie Potter, Max Sullivan, and Rustyn Yazdanpour.
\mainmatter
\setcounter{page}{5}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter*{Introduction}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
This course is about set theory, and its use in the study of the real line. By way of motivation, consider the question of measuring the \emph{size} of a given set of real numbers. The word ``size'' can mean many different things, depending on the context. To see what I mean, consider these three important examples.
\begin{itemize}
\item To a combinatorist or set theorist, the simplest notion of size is
\emph{cardinality}, which asks for instance whether a set is finite, countably infinite, or uncountable. Of course if the set is uncountable, a set theorist would further hope to distinguish whether it is of the first uncountable cardinality ($\aleph_1$), the second uncountable cardinality ($\aleph_2$), etc.
\item To an analyst, one standard notion of size is \emph{measure}, which asks for instance whether a set has zero length or positive length.
\item To a topologist, a key notion of size is \emph{category}, which asks whether a set is meager (topologically negligable), comeager, or else neither meager nor comeager.
\end{itemize}
In this course we will ask the natural question: How do these three kinds of size compare with one another? In answering this question, we will discover numerous relationships between measure and category, with the diagram of these relationships revealing an intricate structure.
Our investigation of measure and category will lead us to a number of statements that are \emph{independent} of the axioms of set theory. Such statements cannot be proved true or false using the usual axioms of set theory. The most central example of such a statement is the \emph{continuum hypothesis} (CH), which asserts that the set of all real numbers has cardinality equal to the first uncountable cardinal number. Cantor initially posed this problem in 1878. In 1940 G\"odel showed the axioms of set theory can't disprove CH, and in 1965 Cohen finally showed that the axioms of set theory can't prove CH.
In his proof, Cohen developed a tool for building models of set theory called \emph{forcing}. Far from being an isolated application, Cohen's contemporaries developed forcind into a general tool for proving independence results. We will develop the machinery of forcing and give a number of such applications.
With the tool of forcing in hand, we will return to the notions of measure and category. We will find that there are numerous cardinal numbers surrounding the zero length sets and the meager sets whose values, like the size of the set of all real numbers, are independent of the axioms of set theory.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Category and measure}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{The real number system}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
Since we will be working with real numbers, we should begin with a definition of the real number system. First we need some vocabulary surrounding fields.
\begin{definition}
\begin{itemize}
\item \emph{Field}: A set $F$ together with distinguished elements $0,1$, operations $+,\times$, satisfying the associative, commutative, and distributive laws, $0$ is the additive identity and $1$ is the multiplicative identity;
\item \emph{Ordered field}: A field with a linear ordering $<$ satisfying $a<b$ implies $a+c<b+c$, and if $c$ is positive then $a<b$ implies $ac<bc$;
\item \emph{Complete field}: if $A\subset F$ has an upper bound, then it has a least upper bound (supremum).
\end{itemize}
\end{definition}
It is not difficult to prove that any two complete ordered fields are isomorphic to each other, which means it makes sense for us to define the real number system as the unique such object.
\begin{definition}
The \emph{real number system} is the unique complete ordered field.
\end{definition}
Now this definition of the real number system is axiomatic---it tells you how real numbers behave, but it leaves it up to your imagination what a real number really is. In fact, we have some choice in this matter. One can construct real numbers using decimal strings, binary strings, or even Cauchy sequences. For our set-theoretic purposes, all that matters is that we give one construction of the real numbers that works.
We will give the construction of the real numbers using cuts, which was proposed by Dedekind in the 19th century. First, let's take the construction of the rational number system for granted. It is simply the set of fractions whose numerator is an integer and whose denominator is a nonzero natural number.
Now observe that each real number $r$ \emph{cuts} the rational numbers into two pieces: the set of rational numbers that are strictly less than $r$, and the set of rational numbers that are greater than or equal to $r$. Conversely, if I cut the rational numbers into two pieces, there is a real number just in the middle of the cut (the supremum of the left piece).
This means that all we have to do to define real numbers is to define cuts. To avoid circularity, we should axiomatize cuts without referring to real numbers at all. A \emph{cut} in the rational numbers is a subset $C\subset\Q$ satisfying three properties:
\begin{itemize}
\item (nontriviality) $C\neq\emptyset$ and $C\neq\Q$
\item (downwards closure) if $q\in C$ and $q'<q$ then $q'\in C$
\item (no last element) if $q\in C$ then there is $q'>q$ such that $q'\in C$
\end{itemize}
To complete the construction one must say how the $+,\times$ operations and $<$ relation can be defined purely in terms of cuts, and then prove that they satisfy all the axioms of a complete ordered field. For example, one defines $C+C'$ as follows:
\[C+C'=\set{q+q'\in\Q\mid q\in C,q'\in C'}
\]
Similarly one defines $C<C'$ if and only if $C\subset C'$ and $C\neq C'$. We omit the details verifying that these definitions work.
\begin{remark}
It is interesting to compare Dedekind's construction with the modern decimal number construction. The decimal number construction is practical for calculations, but has some oddities. For example some numbers have two decimal representations, like $.999\cdots$ and $1.000\cdots$. Also we made an arbitrary decision to use decimal expansions over other bases. Dedekind's construction is beautiful because it is uniform and doesn't involve any arbitrary decisions.
\end{remark}
When studying the foundations of numbers, one often starts with the natural numbers, uses differences to construct integers, then fractions to construct rational numbers, and finally constructs the real numbers as above. It may seem as if these steps are all similar to one another, but the final step is different. To see why we must introduce the idea of set size.
\begin{definition}
A set $A$ of real numbers is said to be \emph{countable} if its elements can be enumerated using natural number indices. In other words, $A$ is countable if there exists a sequence $(a_n)_{n\in\N}$ such that $A=\set{a_n\mid n\in\N}$.
\end{definition}
Note that in this terminology, a finite set is officially countable. If we wish to emphasize that a particular countable set is infinite, we will use the term \emph{countably infinite}. A set that is not countable is called \emph{uncountable}.
The natural numbers, the integers, and the rational numbers are all countable using well-known argument. We now give Cantor's famous proof of the fact that the set of all real numbers is uncountable. The argument goes by contradiction and uses a recursive construction to reach a contradiction one step at a time. Such arguments are often called ``diagonal'', for reasons we shall see later on.
\begin{theorem}[Cantor]
\label{thm:cantor}
The set of all real numbers is uncountable.
\end{theorem}
\begin{proof}
Suppose towards a contradiction that the set of all real numbers is countable. Then there exists a sequence $(a_n)_{n\in\N}$ such that $\R=\set{a_n\mid n\in\N}$. Our strategy will be to inductively construct a decreasing sequence of closed intervals $I_n=[l_n,r_r]$ such that for all $n$, $a_n\notin I_n$. Then if $r$ lies in the intersection of these intervals, $r$ cannot be equal to $a_n$ for any $n$, a contradiction.
The construction itself is straightforward. For the base case let $I_1=[l_1,r_1]$ be any interval which omits $a_1$. For the inductive step, if $I_{n-1}$ has been defined, let $I_n=[l_n,r_n]$ be any subinterval of $I_{n-1}$ which omits $a_n$.
We can now find a point in the intersection of the $I_n$ using the completeness property. First observe that since $I_n\subset I_{n-1}$ for all $n$, the set of left endpoints $\set{l_n\mid n\in\N}$ is bounded above by any and all of the right endpoints $r_n$. By the completeness property, the set of left endpoints has a least upper bound $x$. Since $x$ is an upper bound for $\set{l_n\mid n\in\N}$, we have $l_n\leq x$ for all $n$. Since $x$ is the least possible upper bound for $\set{l_n\mid n\in\N}$, we have $x\leq r_n$ for all $n$. Therefore we have $x\in[l_n,r_n]=I_n$ for all $n$.
Since $x$ lies in $I_n$ for all $n$, and since we chose $I_n$ in such a way that it omits $a_n$, we know that $x\neq a_n$ for all $n$. This contradicts the hypothesis that $\R=\set{a_n\mid n\in\N}$, and concludes the proof.
\end{proof}
This style argument appears in one form or another numerous times throughout elementary analysis and set theory. We shall next see it in the proof of the Baire category theorem.
\begin{exercise}
With the definition of $C+C'$ we gave for Dedekind cuts, show that addition is commutative and associative.
\end{exercise}
\begin{exercise}
Define multiplication $C\times C'$ for Dedekind cuts, and show that it agrees with multiplication for rational numbers. [Hint: consider four cases when $C,C'$ are negative or nonnegative.]
\end{exercise}
\begin{exercise}
Let $A$ be a set of real numbers, and assume $A$ contains an interval of nonzero length. Show that $A$ is uncountable.
\end{exercise}
\begin{exercise}
Show that any two complete ordered fields are isomorphic as ordered fields. [Hint: observe that both must contain a copy of $\Q$ which is dense.]
\end{exercise}
\begin{notes}
Although the material in this section is standard and can be located in most any analysis book, an excellent introduction is \emph{Understanding Analysis} by Stephen Abbott. For the completeness axiom see Abbott, Section~1.3. For Cantor's theorem see Abbott, Section~1.4. For Dedekind cuts see Abbott, Section~8.4.
\end{notes}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Baire category theory}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
In the previous section we introduced the real numbers, and a cardinality-based notion of size (countable and uncountable). In this section we introduce a topological notion of size. Intuitively one might try to begin with denseness, but it doesn't quite work. The reason is that some very small sets are dense (like the rationals) and some very large sets are not dense (like the half-line $[0,\infty)$). Instead we need the following.
\begin{definition}
A set of real numbers $A$ is said to be \emph{nowhere dense} if every positive-length interval $I$ contains a positive-length interval $J$ such that $J$ is disjoint from $A$.
\end{definition}
\begin{example}
Any finite subset of $\R$ is nowhere dense. The set $\set{1/n\mid n\in\N}$ is nowhere dense. In fact, any \emph{discrete} set is nowhere dense (here, $A$ is discrete if every point of $A$ is isolated). The set of rational numbers whose numerator is $1$ and whose denominator is a power of $2$ is \emph{not} nowhere dense, because such a number can be found in every positive-length subinterval of $[0,1]$.
\end{example}
Note that nowhere dense sets need not be countable. The classical Cantor middle thirds set is an archetypal example of a nowhere dense set. Recall that the \emph{Cantor set} is defined as follows. Begin with the set $C_1=[0,1]$. Let $C_2=[0,1/3]\cup[2/3,1]$ be the set $C_1$ with its open middle third removed. Let $C_3=[1,1/9]\cup[2/9,1/3]\cup[2/3,7/9]\cup[8/9,1]$ be the set $C_2$ with the open middle third removed from each component of $C_2$. Refer to Figure~\ref{fig:cantor-set} for a picture of these first few sets. Recursively, let $C_{n+1}$ be constructed from $C_n$ by removing the open middle third from each component of $C_n$. Finally the Cantor set $C$ is defined to be $C=\bigcap C_n$. We will see in a later section that the Cantor set is uncountable. % add ref
\begin{figure}[h]
\begin{center}
\begin{tikzpicture}
\draw[|-|] (0,0) -- (10,0) node[anchor=west] {\ \ $C_1$};
\draw[|-|] (0,-.5) -- (10/3,-.5);
\draw[|-|] (20/3,-.5) -- (10,-.5) node[anchor=west] {\ \ $C_2$};
\draw[|-|] (0,-1) -- (10/9,-1);
\draw[|-|] (20/9,-1) -- (30/9,-1);
\draw[|-|] (60/9,-1) -- (70/9,-1);
\draw[|-|] (80/9,-1) -- (10,-1) node[anchor=west] {\ \ $C_3$};
\node[anchor=west] at (10,-1.4) {\ \ \ $\vdots$};
\draw[decoration=Cantor set,very thick]
decorate{ decorate{ decorate{ decorate{ decorate{ decorate{
(0,-2) -- (10,-2) }}}}}} node[anchor=west] {\ \ $C$};
\end{tikzpicture}
\end{center}
\caption{The first few steps in the construction of the Cantor set, and a rough image of the Cantor set itself.\label{fig:cantor-set}}
\end{figure}
\begin{proposition}
The Cantor set $C$ is nowhere dense.
\end{proposition}
\begin{proof}
Let $I$ be a given positive-length interval. Observe that for each $n$, the set $C_n$ consists of intervals of length $1/3^n$ which are all at least $1/3^{n-1}$ units apart from one another. Since $1/3^n\to0$, we can find some value $N$ such that $C_N$ consists of intervals that are strictly shorter than the length of $I$. It follows that $I\cap (C_N)^c$ contains a positive-length interval $J$, as desired.
\end{proof}
The terminology ``nowhere dense'' may sound strange at first, but it is justified by the following fact.
\begin{proposition}
\label{prop:nwd-equiv}
Let $A$ be a set of real numbers. The following are equivalent.
\begin{enumerate}
\item $A$ is nowhere dense;
\item $\cl(A)$ (the topological closure of $A$) has no interior;
\item for every nonempty open set $O$, we have that $A$ is not dense in $O$.
\end{enumerate}
\end{proposition}
The proof is requested in Exercise~\ref{exerc:nwd-equiv}.
\begin{theorem}
The collection of nowhere dense sets is closed under the operations of subset, union, and closure.
\end{theorem}
\begin{proof}
Preservation to subsets is immediate from the definition.
For preservation to unions, suppose that $A,B$ are nowhere dense and let $I$ be a positive-length interval. Since $A$ is nowhere dense, there is a positive-length interval $J\subset I\smallsetminus A$. Since $B$ is nowhere dense there is a further positive-length interval $J'\subset J\smallsetminus B$. It follows that
\[J'\subset (I\smallsetminus A)\smallsetminus B=I\smallsetminus (A\cup B)
\]
This establishes that $A\cup B$ is again nowhere dense.
Preservation to closures is immediate from condition (b) of Proposition~\ref{prop:nwd-equiv}, but we can also give a direct argument. Let $A$ be nowhere dense and let $I$ be a given positive-length interval. Since $A$ is nowhere dense there is a positive-length interval $J$ such that $J\subset\R\smallsetminus A$. Removing endpoints if necessary, we may suppose that $J$ is open. It follows that $J\subset\R\smallsetminus\cl(A)$ (recall that since $\cl(A)$ is the intersection of all closed sets containing $A$, we have that $\R\smallsetminus\cl(A)$ is the union of all open sets disjoint from $A$). This shows that $\cl(A)$ is nowhere dense.
\end{proof}
It is easy to see that the collection of nowhere dense sets is \emph{not} preserved to infinite unions, even countably infinite ones. For example, the set of rational numbers is countable, and therefore it is a countable unions of singletons: $\Q=\bigcup_{q\in\Q}\{q\}$. Each singleton $\{q\}$ is clearly nowhere dense, so $\Q$ is a countable union of nowhere dense sets, but $\Q$ is dense.
\begin{definition}
A set of real numbers $A$ is called \emph{meager} if it is a union of countably many nowhere dense sets. A set $A$ is called \emph{comeager} if $\R\smallsetminus A$ is meager.
\end{definition}
At this point it is still conceivable that the concept of meagerness is completely trivial---maybe every set of real numbers turns out meager according to this definition. The Baire category theorem says that this is not the case. In fact, we will soon see that meager sets really are ``small'', as the name implies.
\begin{theorem}[Baire category theorem]
\label{thm:baire}
The set of all real numbers $\R$ is nonmeager. In fact, if $I$ is a positive-length interval then $I$ is nonmeager.
\end{theorem}
\begin{proof}
The proof is very similar to Cantor's proof of Theorem~\ref{thm:cantor}. Let $I$ be a positive-length interval and suppose towards a contradiction that $I$ is meager. Then there exists a sequence $(A_n)_{n\in\N}$ of nowhere dense sets such that $\bigcup A_n=I$. We will inductively construct a decreasing sequence of closed subintervals $J_n\subset I$ such that for all $n$, $J_n\cap A_n=\emptyset$.
The induction is again straightforward. First apply the definition of $A_1$ being nowhere dense to obtain $J_1$. Inductively apply the definition of $A_n$ being nowhere dense to $J_n$ to obtain $J_{n+1}$. (Each time we may shrink the newly obtained interval slightly to suppose that it is closed.)
We can now argue just as in the proof of Theorem~\ref{thm:cantor} that there exists a point $x\in\bigcap J_n$. It follows that $x\notin\bigcup A_n$, which contradicts the assumption that $I=\bigcup A_n$, and completes the proof.
\end{proof}
Baire category theory gets its name from the above theorem, and the fact that meager sets used to be called ``first category''. Nonmeager sets used to be called ``second category''.
\begin{theorem}
\label{thm:meager-pres}
The class of meager sets is closed under the operations of subset and countable union.
\end{theorem}
We remark that the meager property is not necessarily preserved to the closure. For example the closure of $\Q$ is $\R$. In fact $\cl(A)$ is meager if and only if $A$ is nowhere dense.
As a consequence of the Baire category theorem, for any set $A$ it is possible to assign $A$ one of three distinct ``sizes'':
\begin{itemize}
\item $A$ is meager (small)
\item $A$ is comeager (large)
\item $A$ neither meager nor comeager (intermediate)
\end{itemize}
We should verify that no set of numbers can be both meager and comeager. Indeed, suppose that both $A$ and $\R\smallsetminus A$ were meager. Then by Theorem~\ref{thm:meager-pres} their union $\R$ would be meager, contradicting the Baire category theorem.
We should also verify that there exists sets of numbers that are neither meager nor comeager. Indeed, if $I$ is any bounded positive-length interval then by the Baire category theorem $I$ is nonmeager, and since $\R\smallsetminus I$ also contains a bounded interval, $I$ is also non-comeager.
\begin{exercise}
\label{exerc:cantor}
Compute the sum of the lengths of all of the intervals removed from $[0,1]$ in the construction of the Cantor set. What if some fraction other than $1/3$ is removed at each stage?
\end{exercise}
\begin{exercise}
\label{exerc:nwd-equiv}
Prove Proposition~\ref{prop:nwd-equiv}.
\end{exercise}
\begin{exercise}
We have observed that unlike the nowhere dense property, the meager property is not necessarily preserved to the closure. Prove that if $\cl(A)$ is meager, then $A$ was nowhere dense to begin with.
\end{exercise}
\begin{notes}
For more on the Cantor set see Abbott, Section~3.1. For the Baire category theorem see Abbott, Section~3.5. For Baire category theory generally see Oxtoby, Chapter~1.
\end{notes}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Lebesgue measure theory}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
Classical measure theory aims to try to extend the length function on intervals to be defined on more complicated sets. For example, if a given set is a finite or even countable union of intervals, then it is appropriate to take the sum of the lengths of the components. But what about more complex sets like the Cantor set? This was known as the \emph{measure problem}: to construct a measurement function $m$, defined on sets of real numbers and valued in $[0,\infty]$, satisfying:
\begin{enumerate}
\item (normality) if $I$ is an interval then $m(I)=\ell(I)$
\item (translation invariance) $m(x+A)=m(A)$ for all sets $A$ and real numbers $x$
\item (countable additivity) if $(A_n)_{n\in\N}$ is a sequence pairwise disjoint sets then $m(\bigcup A_n)=\sum m(A_n)$
\end{enumerate}
Perhaps surprisingly, the conditions (a)--(c) are mutually contradictory and so no such measurement function $m$ exists! Here is Vitali's clever example of a contradiction arising from these requirements. Regard $\Q$ as an additive subgroup of $\R$ and consider the cosets of $\Q$, that is, sets of the form $x+\Q$. Let $A\subset[0,1]$ be a set of coset representatives, that is, $A$ contains exactly one element from each of the cosets. (It is always possible to choose a coset representative in $[0,1]$ because $\Q$ is dense.)
Then the translations $q+A$ for $q\in\Q$ form a countable sequence of pairwise disjoint sets that cover all of $\R$. In fact, if we let $S=\Q\cap[-1,1]$ then the translations $q+A$ for $q\in S$ already cover all of $[0,1]$. We can then infer from (a) and (c) that $m[\bigcup_{q\in S}(q+A)]$ lies between $1$ and $3$. But on the other hand, by (b) and (c) we have that
\[m\left(\bigcup_{q\in S}(q+A)\right)=\sum_{q\in S}m(q+A)=\sum_{q\in S}m(A)
\]
This is a contradiction because an infinite sum of a single constant value $m(A)$ must equal either $0$ or $\infty$, and so cannot lie between $1$ and $3$.
\begin{remark}
We observe that the axiom of choice was explicitly used in Vitali's construction of the set $A$ above. In fact, it is known that the use of the axiom of choice is necessary to build a nonmeasurable set.
\end{remark}
The contradiction described above is typically resolved not by dropping one of the conditions (a)--(c), but rather by dropping the requirement that $m$ is defined on \emph{all} sets of real numbers. The justification for this decision is that sets like the $A$ constructed above should be regarded as pathological, and we don't usually need to measure them in applications.
Let us na\"ively begin to construct a measure $m$ that is at least defined on reasonable sets. First, condition (a) implies we should let $m(I)=\ell(I)$ for any interval $I$. Next, if $A=\bigcup I_n$ is a union of disjoint intervals $I_n$, then condition (c) implies we should let $m(A)=\sum\ell(I_n)$. Third, if $A=\bigcap A_n$ is an intersection of sets where $m$ is defined and finite, then it is natural to define $m(A)=\inf m(A_n)$. We now observe that all three of the above simple cases fit into the rule given by the following key formula:
\begin{equation}
\label{eq:measure}
m(A)=\inf\set{\sum\ell(I_n)\mid \text{$I_n$ are intervals and }A\subset\bigcup I_n}
\end{equation}
The next result states that this rule for defining $m$ actually works for all sets that are reasonably constructible. Recall that a \emph{Borel set} is one that can be constructed by beginning with the intervals and then executing countably many countable unions or intersections.
\begin{theorem}[Carath\'eodory's extension theorem]
\label{thm:caratheodory}
The measure $m$ defined in Equation~\eqref{eq:measure} satisfies conditions (a) and (b), and additionally satisfies condition (c) when applied to Borel sets.
\end{theorem}
\begin{proof}[Proof of conditions (a) and (b) only]
For condition (a), let $I$ be an interval. It is clear that $m(I)\leq\ell(I)$, since $I$ itself is an interval covering $I$. For the other direction, we will show that $I\subset\bigcup I_n$ implies $\sum\ell(I_n)\geq\ell(I)$. Then taking the infemum over all such coverings this allows us to conclude that $m(I)\geq\ell(I)$.
Let us first handle the case when $I=[a,b]$ is closed and bounded, and all of the $I_n$ are open intervals $(a_n,b_n)$. Then since $I$ is compact, $I$ is covered by just finitely many of the $I_n$, without loss of generality we may assume they are indexed $I_0,\ldots,I_N$. Now after further renaming or removing intervals from the list, we may suppose that $I_0$ covers the left endpoint of $I$, each $I_{n+1}$ covers the right endpoint of $I_n$, and $I_N$ covers the right endpoint of $I$. We can now compute that
\[\sum\ell(I_n)
\geq\sum_0^N(b_n-a_n)
\geq\sum_1^N(b_n-b_{n-1})-a
\geq b-a
= \ell(I)\text{.}
\]
If $I$ is not necessarily closed and the $I_n$ are not necessarily open, then we look instead at $I'=\cl(I)$ and $I_n'=(I_n)^\circ$. Then the $I_n'$ cover all but a countable subset of $I'$, so for any $\epsilon$ we can find additional open intervals $J_n$ with $\ell(J_n)\leq\epsilon/2^n$ covering these missing points. Now the previous argument shows that $\sum\ell(I_n)+\sum\ell(J_n)\geq\ell(I)$. Using the geometric series formula we obtain $\sum\ell(I_n)+\epsilon\geq\ell(I)$. Letting $\epsilon\to0$ we have the desired result.
Finally if $I$ is not bounded, we let $A_k$ be a sequence of bounded subintervals of $I$ such that $\ell(A_k)\to\infty$. The above result implies that $\sum\ell(I_n)\geq\ell(A_k)$, and letting $k\to\infty$ we once again achieve the desired result. This concludes the proof of condition (a).
Condition (b) follows directly from the fact that the length function $\ell$ is translation invariant, and the definition of $m$ depends only on $\ell$.
For the proof that $m$ satisfies condition (c), we refer the reader to any standard measure theory text.
\end{proof}
From now on we will ignore the full power of measure theory to assign a real number measure to any Borel set, and focus only on the specific value zero. Sets whose measure is zero are called null sets, and for convenience we extract the definition of null set from the above definition of arbitrary measure.
\begin{definition}
\label{defn:null}
A set of real numbers $A$ is \emph{null} if for all $\epsilon>0$ there exists a sequence of intervals $(I_n)_{n\in\N}$ such that $A\subset\bigcup I_n$ and $\sum\ell(I_n)<\epsilon$.
\end{definition}
\begin{example}
The Cantor set $C$ is null. Indeed, we have already computed that the sum of the lengths of all intervals removed from $[0,1]$ in the construction of $C$ is equal to $1$. Since $[0,1]$ has measure $1$, it follows from additivity that $C$ must have measure $0$.
\end{example}
The notion of null set bears many similarities with the notion of meager set. As was the case with meager sets, the notion of a null set allows us to assign to any given set $A$ one of three simple ``sizes'': null, conull, or nonnull and non-conull. Moreover, null sets satisfy an analog of the Baire category theorem and the preservation properties of meager sets.
\begin{corollary}
\label{cor:interval-nonnull}
The set $\R$ of all real numbers is not null. In fact, if $I$ is any positive-length interval then $I$ is not null.
\end{corollary}
\begin{corollary}
\label{cor:null-pres}
The class of null sets is closed under the operations of subset and countable union.
\end{corollary}
Corollary~\ref{cor:interval-nonnull} is immediate from condition (a) of Theorem~\ref{thm:caratheodory}. Corollary~\ref{cor:null-pres} is immediate from condition (c) of Theorem~\ref{thm:caratheodory}.
\begin{exercise}
Prove that the properties (a)--(c) of a measure imply \emph{monotonicity}: if $A\subset B$ then $m(A)\leq m(B)$.
\end{exercise}
\begin{exercise}
\label{exerc:continuity-above}
Prove that the properties (a)--(c) of a measure imply \emph{continuity from below}: if $A_n$ is an increasing sequence of sets and $A=\bigcap A_n$, then $m(A)=\sup m(A_n)$. Then prove \emph{continuity from above}: if $A_n$ is a decreasing sequence of sets, $m(A_n)$ is finite, and $A=\bigcap A_n$, then $m(A)=\inf m(A_n)$.
\end{exercise}
\begin{exercise}
Give a proof directly from Definition~\ref{defn:null} that the Cantor set is null.
\end{exercise}
\begin{exercise}
Give a proof directly from Definition ~\ref{defn:null} that the class of null sets is closed under countable union.
\end{exercise}
\begin{notes}
Our presentation of Vitali's construction follows Folland, Section~1.1. Our proof of Theorem~\ref{thm:caratheodory}, part (a) follows Oxtoby, Chapter~1. For a proof of part (c), see Folland, Section~1.4.
\end{notes}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Sets, orderings, and cardinality}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
So far we have seen sets that are finite, countable, and uncountable. If a set $A$ is finite, then there is a natural number that corresponds to the number of elements of $A$. If $A$ is countable, we understand that it has exactly as many elements as there are natural numbers. But if $A$ is uncountable, is that all that needs be said or is there some kind of number that tells us just how uncountable it is?
In this section we discuss the notion of ``cardinality'' of a set $A$, which replaces the notion of ``number of elements'' in the case when $A$ is infinite. Notationally, we write $|A|$ for the cardinality of $A$. When $A$ is finite, $|A|$ will be a natural number. When $A$ is countable, $|A|$ will be the value $\aleph_0$ (pronounced aleph zero, or aleph nought). And when $A$ is uncountable $|A|$ will be one of the values $\aleph_1,\aleph_2$, etc.
We will now begin to describe several key properties of cardinality. Surprisingly, we can do this without actually defining what the $\aleph$'s are! The following definition of how $|\cdot|$ behaves is sufficient for many practical purposes. In the next section we will give a formal definition of the $\aleph$'s.
\begin{definition}
\label{defn:cardinal-rel}
Let $A$ and $B$ be sets.
\begin{itemize}
\item We write $|A|=|B|$ if there exists a bijective function $A\to B$;
\item We write $|A|\leq|B|$ if there exists an injective function $A\to B$;
\item We write $|A|<|B|$ if we have $|A|\leq|B|$ and also $|A|\neq|B|$.
\end{itemize}
\end{definition}
With this definition in hand, we can now confirm that there are many different uncountable cardinalities. Recall that if $A$ is a set, then the \emph{power set} of $A$, denoted $\mathcal P(A)$, is the set of all subsets of $A$.
\begin{theorem}[Cantor]
If $A$ is any set, then $|A|<|\mathcal P(A)|$.
\end{theorem}
\begin{proof}
It is clear that $|A|\leq|\mathcal P(A)|$, since the map $i(a)=\{a\}$ is an injection from $A$ into $\mathcal P(A)$.
To see that there is no bijection between $A$ and $\mathcal P(A)$, let $f\colon A\to\mathcal P(A)$ be any function. Then build the set $D$ of all elements $a\in A$ such that $a\notin f(a)$. We claim that $D$ is not in the range of $f$, and therefore that $f$ is not a bijection.
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}
THis classical argument is often called a ``diagonal argument'' because of the key formula in the proof: $a\notin f(a)$. To picture it, suppose you graph the $A^2$ plane, and shade the set of pairs $(x,y)\in A^2$ such that that $x\in f(y)$. Then the set $D$ is constructed by taking the unshaded elements of the diagonal of the $A^2$ plane.
\begin{definition}
Let $A$ be a set (or class).
\begin{itemize}
\item A \emph{partial ordering} of $A$ is a binary relation $\leq$ satisfying: (reflexive) $a\leq a$; (antisymmetric) $a\leq b\leq a\implies a=b$; (transitive) $a\leq b\leq c\implies a\leq c$.
\item A \emph{total ordering} of $A$ is a partial ordering such that for all $a,a'$ either $a\leq a'$ or $a'\leq a$.
\item A \emph{well-ordering} of $A$ is a total ordering $\leq$ such that every subset $B\subset A$ has a $\leq$-least element $b_0$.
\end{itemize}
\end{definition}
\begin{remark}
The property of being a well-ordering may seem obscure at first, but finding a least element is precisely what is needed in induction arguments. It is what allows us to say ``\ldots otherwise, let $x$ be the least counterexample''.
\end{remark}
In the next few results, we essentially show that the ordering $\leq$ on the cardinals is a well-ordering. Note that reflexivity holds because the identity function is an injection from $A$ into itself, and transitivity holds because the composition of two injections is an injection. The following result establishes antisymmetry.
\begin{theorem}[Cantor--Schr\"oder--Bernstein]
\label{thm:csb}
If $|A|\leq|B|$ and $|B|\leq|A|$ then $|A|=|B|$. In other words, if there exist injections $i\colon A\to B$ and $j\colon B\to A$ then there is a bijection $f\colon A\to B$.
\end{theorem}
\begin{proof}
Replacing $B$ with $j(B)$, we may suppose that that $B\subset A$. Then we have
\[A\supset B\supset i(A)\supset i(B)\supset i^2(A)\supset\cdots
\]
Now $A$ can be written as the union of the successive differences of these sets, together with the intersection of them all:
\[A=(A\mathord{\smallsetminus}B)\cup(B\mathord{\smallsetminus}i(A))
\cup(i(A)\mathord{\smallsetminus}i(B))\cup\cdots\cup \bigcap i^n(A)
\]
Meanwhile, $i$ provides bijections
\[A\mathord{\smallsetminus}B\leftrightarrow i(A)\mathord{\smallsetminus}i(B)\leftrightarrow i^2(A)\mathord{\smallsetminus}i^2(B)\leftrightarrow\cdots
\]
It follows that the map
\[f(a)=\begin{cases}i(a)&\text{if }a\in(A\mathord{\smallsetminus}B)\cup(i(A)\mathord{\smallsetminus}i(B))\cup(i^2(A)\mathord{\smallsetminus}i^2(B))\cup\cdots\\
a&\text{otherwise}
\end{cases}
\]
is a bijection from $A$ to $B$.
\end{proof}
This result also gives a simple recipe for checking whether two sets have the same cardinality. For instance to show $[0,1]$ and $(0,1)$ have the same cardinality, it is much easier to construct two injections than a single bijection!
The next result shows that the cardinalities are totally ordered. In the proof we will need \emph{Zorn's lemma}: If $P$ is a partially ordered set such that every totally ordered subset has an upper bound, then $P$ has at least one maximal element. Zorn's lemma is used perennially in analysis, and it is a consequence of the axiom of choice.
\begin{theorem}
\label{thm:card-total}
If $A,B$ are sets then either $|A|\leq|B|$ or $|B|\leq|A|$. That is, there exist injective functions $A\to B$ and $B\to A$.
\end{theorem}
\begin{proof}
Consider the family $\mathcal F$ of all injective functions whose domain is a subset of $A$ and whose range is a subset of $B$. Then $\mathcal F$ is partially ordered by extension of functions, and it is easy to check that this ordering satisfies the hypothesis of Zorn's lemma. Thus there is a maximal element $f$, and since it is maximal either the domain of $f$ is all of $A$ or the range of $f$ is all of $B$. In the first case $f$ is an injection from $A$ to $B$, and in the second case $f^{-1}$ is an injection from $B$ to $A$.
\end{proof}
Finally we show that $\leq$ is a well-ordering on cardinals. One technicality arises here that did not in the last four properties. To check that the cardinals are well-ordered, we should check that \emph{any} collection of sets has a minimal element, and that collection need not itself be a set. (Remember: the set of all sets isn't a set!)
\begin{theorem}
\label{thm:card-well}
Let $\mathcal A$ be a class of sets. Then there exists a $\leq$-minimal element $A\in\mathcal A$, that is, $A$ injects into any $B\in\mathcal A$.
\end{theorem}
\begin{proof}
We argue very similarly to Theorem~\ref{thm:card-total}, but in order to apply Zorn's lemma we must first suppose that $\mathcal A$ really is a set. Fix any element $A\in\mathcal A$ and let
\[\mathcal F=\set{(f_B)_{B\in\mathcal A}\mid\text{there is $A_0\subset A$ such that for all $B$, $f_B$ is an injection from $A_0$ to $B$}}
\]
This time $\mathcal F$ is partially ordered by \emph{coordinatewise} extension of functions. By Zorn's lemma there is a maximal element $(f_B)_{B\in\mathcal A}$, and since it is maximal either the domain $A_0$ is all of $A$ or the range of one of the functions $f_B$ is all of $B$. In the first case $f_B$ is an injection from $A$ to $B$ for all $B$. In the second case, fix $B$ such that $f_B(A_0)=B$. Then for any $C\in\mathcal A$, the composition $f_C\circ f_B^{-1}$ is an injection from $B$ to $C$, so $B$ is as desired.
In the general case when $\mathcal A$ is a class, we can reduce to the set case as follows. Fix an element $D\in\mathcal A$ and let $\mathcal A'$ denote the collection of subsets of $D$ that are in bijection with some element of $\mathcal A$. Since $\mathcal A'$ is a set, we can find $A\in\mathcal A$ which injects into all elements of $\mathcal A'$. It follows that $A$ injects into all other elements of $\mathcal A$ too. Indeed if $B\in\mathcal A$ and $A$ does not inject into $B$, then by Theorem~\ref{thm:card-total}, $B$ injects into $A$. It follows that $B$ is in bijection with a subset of $D$ and hence $A$ injects into $B$ after all.
\end{proof}
Although we still can't be fully rigorous about the meaning of the symbols $\aleph_1,\aleph_2,\ldots$, the well-ordering property helps justify the use of these symbols. Essentially $\aleph_1$ is the least cardinality greater than $\aleph_0$, $\aleph_2$ is the least cardinality greater than $\aleph_1$, and so forth. In the next section, we will make this fully precise.
\begin{exercise}
Show that the following sets are all in bijection with one another: $\R$, $[0,1]$, $(0,1)$, $(0,\infty)$, $\mathcal P(\N)$, and $\{A\subset\N\mid A\text{is infinite}\}$.
\end{exercise}
\begin{exercise}
Which of the following categories satisfy the analog of the Cantor--Schr\"oder--Bernstein theorem? (That is, monomorphisms $A\to B\to A$ implies isomorphism $A\cong B$.) linear orders with order-preserving maps; groups with group homomorphisms; topological spaces with continuous maps; topological spaces with piecewise continuous maps.
\end{exercise}
% why not show the set of countable ordinals is an ordinal
\begin{exercise}
We used the well-ordering principle together with Cantor's theorem to argue that there are uncountable ordinals. Here is another way: show that the set of isomorphism equivalence classes of well-orderings of $\omega$ is itself naturally well-ordered, and that this well-ordering is uncountable.
\end{exercise}
\begin{notes}
Our proof of Theorem~\ref{thm:card-well} follows and extends the proof given in Chaim Samuel H\"onig, \emph{Proof of the well-ordering of cardinal numbers}. For a more detailed proof of Theorem~\ref{thm:csb}, see Kunen (Foundations), Section~I.10.
\end{notes}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Ordinal numbers and cardinal numbers}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
In the previous section we defined cardinal equality $|A|=|B|$ and studied the order structure $\leq$ of the cardinal equality equivalence classes. In this section we finally define $|A|$ by choosing a single representative of $A$'s equivalence class. Specifically, we will define $|A|$ to be the least ordinal number in $A$'s equivalence class.
Ordinal numbers play a central role in set theory, including both cardinality theory and forcing. While cardinal numbers are needed to measure the \emph{size} of an infinite set, ordinal numbers are needed to measure the \emph{length} of an infinite well-ordered set. The ordinals can be used to extend the notion of \emph{counting} into the infinite, and thus are an excellent choice to help us define the cardinal numbers or $\aleph$'s.
The initial goal in defining the ordinals is to provide a collection of well-ordered sets such that for any given well-ordered set $A$ there is one and only one ordinal $\alpha$ such that $A$ is isomorphic to $\alpha$. For finite well-ordered sets, we can write such a definition explicitly:
\begin{align*}
0&=\emptyset\\
1&=\{0\}\\
2&=\{0,1\}\\
\vdots\\
n+1&=\{0,\ldots,n\}
\end{align*}
This construction can be summed up in one recurrence: $n+1=n\cup\{n\}$. The first infinite ordinal, called $\omega$ or $\omega_0$, is simply the union of all the finite ordinals. The process can then continue; the successor of $\omega$ is the ordinal $\omega+1=\omega\cup\{\omega\}$.
Intuitively, infinite ordinals are all constructed in this fashion. If $\alpha$ is any ordinal then its successor is $\alpha+1=\alpha\cup\{\alpha\}$, and after infinitely many such steps we take a union. Unfortunately this prescription is not rigorous because it is circular; we cannot make a definition like $\alpha=\bigcup_{\beta<\alpha}\beta$. Instead we have the following more technical characterization.
\begin{definition}
\label{defn:ordinals}
A set $\alpha$ is an \emph{ordinal} if it satisfies the properties:
\begin{itemize}
\item $\alpha$ is well-ordered by the $\in$ relation (or more precisely by the relation $\leq$ defined as $\in$-or-equal);
\item $\alpha$ is transitive: if $\gamma\in\beta\in\alpha$ then $\gamma\in\alpha$.
\end{itemize}
\end{definition}
\begin{remark}
The first condition ensures that the ordinals are in fact well-ordered; the order relation $\in$ is simply the most convenient one available in set theory. The second condition ensures that the ordinals have no ``gaps''; for instance the set $\{0,1,3,5,9\}$ is well-ordered but not an ordinal.
\end{remark}
The following fundamental facts about ordinals together imply that Definition~\ref{defn:ordinals} achieves our initial goals in defining the ordinals.
\begin{theorem}\
\label{thm:ordinals}
\begin{itemize}
\item The class of all ordinals is itself transitive and well-ordered by $\leq$.
\item If $A$ is a well-ordered set then there exists a unique ordinal $\alpha$ such that $A$ is isomorphic to $\alpha$.\qed
\end{itemize}
\end{theorem}
An ordinal $\alpha$ is said to be \emph{successor} ordinal if it is of the form $\beta+1$ for some ordinal $\beta$. Otherwise $\alpha$ is said to be \emph{limit} ordinal. The following result confirms that our initial intuitive construction of the ordinals, though it involved circular reasoning, was nonetheless correct in hindsight.
\begin{proposition}
If $\alpha=\beta+1$ then $\alpha$ is the least ordinal greater than $\beta$. If $\alpha$ is a limit ordinal then $\alpha$ is the union of the ordinals that came before it.
\end{proposition}
\begin{proof}
It is easy to check $\alpha=\beta\cup\{\beta\}$ is indeed an ordinal (it is transitive and well-ordered by $\in$). Since the ordinals are well-ordered we may let $\alpha'$ be the least ordinal greater than $\beta$; we want to show that $\alpha=\alpha'$. If this were not the case, then by totality we would either have $\alpha'\in\alpha$ or $\alpha\in\alpha'$. In the first case either $\alpha=\beta$ or $\alpha\in\beta$, which contradicts that $\alpha'$ is greater than $\beta$. The second case contradicts that $\alpha'$ is the least such.
Next let $\alpha$ be a limit ordinal, and let $\alpha'$ be the union of all $\beta\in\alpha$. Since $\alpha$ is transitive we easily have that $\alpha'\subset\alpha$. On the other hand if $\gamma\in\alpha$ then by the previous paragraph $\gamma+1\leq\alpha$ and since $\alpha$ is limit we must have $\gamma+1\in\alpha$. Thus $\gamma\in\gamma+1\in\alpha$ and it follows that $\gamma\in\alpha'$.
\end{proof}
As we have hinted, although ordinals naturally measure the length of well-orderings, they can also be used to measure size $|A|$. Recall that the axiom of choice implies the \emph{well-ordering principle}, which states that any set $A$ admits a well-ordering $\leq$. Combining this with Theorem~\ref{thm:ordinals}, it follows that any set $A$ is in bijection with at least one ordinal. Different well-orderings of $A$ can lead to different ordinals, so we make the following definition.
\begin{definition}
If $A$ is any set, then $|A|$ is the least ordinal $\alpha$ such that $A$ is in bijection with $\alpha$.
\end{definition}
Thus a cardinal is a special type of ordinal. Most ordinals will not be cardinals, since for instance if $A$ is in bijection with $\omega+7$ then clearly it is also in bijection with $\omega$. We give $\aleph$ names to the ordinals which are cardinals: The first infinite cardinal is $\aleph_0$, the first uncountable cardinal is $\aleph_1$, the next least cardinal is $\aleph_2$, and so forth.
The pattern continues transfinitely as well, with $\aleph_\alpha$ defined for every ordinal $\alpha$. Officially, these higher cardinals are defined using \emph{transfinite recursion}. Just as natural numbers are used to index traditional recursion, ordinals are used to index transfinite recursion. While ordinary recursion requires a special ``base'' case at $n=0$, transfinite recursion requires a ``limit'' case at each limit ordinal.
\begin{definition}
The first infinite cardinal is $\aleph_0=\omega$. If $\aleph_\alpha$ is defined then $\aleph_{\alpha+1}$ is the least ordinal that is not in bijection with $\aleph_\alpha$. If $\lambda$ is a limit ordinal and $\aleph_\alpha$ has been defined for $\alpha<\lambda$, then $\aleph_\lambda=\bigcup_{\alpha<\lambda}\aleph_\alpha$.
\end{definition}
% additional topics
% discuss the value of the continuum |R|
% define cofinality
% konig's lemma limiting the value of |R|:
% the continuum has uncountable cofinality
Recall that Cantor proved $|\R|$ is uncountable. Now that we have names for the uncountable cardinals, it is interesting to ask which one $|\R|$ is equal to. Is it $\aleph_1$? Or $\aleph_{17}$? The method of forcing, discussed in Part~III, was invented to answer questions like this one.
\begin{exercise}
Show that if $\kappa$ is an infinite cardinal, then $\kappa$ is a limit ordinal.
\end{exercise}
\begin{exercise}
Use Zorn's lemma to prove the well-ordering principle.
\end{exercise}
\begin{exercise}
If $A$ is an infinite set, show that $|A|$ is equal to $\aleph_\alpha$ for some $\alpha$.
\end{exercise}
\begin{notes}
This material on ordinals and cardinals can be found in any introductory set theory textbook. For ordinals, see for instance Section~1.7 of Devlin's \emph{The joy of sets}. For cardinals, see Section~3.6 of the same text.
\end{notes}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{The topology of Cantor and Baire space}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
If $A$ is a countable set then $A^\omega$ denotes the space of all sequences with values in $A$, that is, functions from $\omega$ to $A$. We can endow $A^\omega$ with a topology by regarding $A$ as discrete, $A^\omega$ as a product of countably many copies of $A$, and using the \emph{product topology}. The official definition of the product topology on $A^B$ states a basis consists of sets of the form
\[\set{x\in A^B\mid x(b_0)\in A_0,\ldots,x(b_{n-1})\in A_{n-1}}
\]
where $b_i\in B$ are distinct and $A_i\subset A$ are open. In the case of $A^\omega$ we can make two simplifications: since $\omega$ is countable we can replace $i_0,\ldots i_{n-1}$ with $0,\ldots,n-1$, and since $A$ is discrete we can assume the $A_i$ are singletons. Putting this all together, $A^\omega$ has a basis consisting of all
\[V_s=\set{x\in A^\omega\mid s\subset x}
\]
where $s$ is an element of $A^n$ for some $n$.
The two most important examples of sequence spaces are the \emph{Cantor space} $2^\omega$ and the \emph{Baire space} $\omega^\omega$.
\begin{proposition}
\label{prop:cantor-baire-homeo}
The Cantor space is homeomorphic to the Cantor middle thirds set. The Baire space is homeomorphic to the set of irrational real numbers.
\end{proposition}
\begin{proof}
We give the proof in the case of the Cantor space, and a brief hint in the case of the Baire space.
The Cantor middle thirds set $C$ has a natural description in terms of ternary expansions. If $a\in[0,1]$ then $a$ lies in the Cantor set if and only if it has a ternary expansion that does not contain the digit $1$. Thus there is a simple bijection $2^\omega\to C$ given by replacing the $1$'s in $x$ with $2$'s:
\[f(x)=0.(2x(0))(2x(1))(2x(2))\cdots
\]
The map is continuous: if $f(x)\in(a,b)$ then there exists $n$ such that if we round $f(x)$ down at its $n$th digit then it is still $>a$ and if we round it up at its $n$th digit then it is still $<b$. It follows that if $s=x\restriction n$, then we have $f(V_s)>a$. Finally recall that a continuous bijection between compact spaces is always a homeomorphism.
For the Baire space, it is common to use the values of $x\in(\omega\smallsetminus\{0\})^\omega$ as the entries in a continued fraction:
\[f(x)=x(0)+\cfrac{1}{x(1)+\cfrac{1}{x(2)+\cdots}}
\]
With some elementary number theory, it is possible to verify that this map is a bijection onto the set of irrational numbers, and even a homeomorphism.
\end{proof}
This result also implies that $2^\omega$ admits a complete metric, because the Cantor set is a closed and hence complete subspace of $\R$. In fact, any sequence space $A^\omega$ is completely metrizable. For an example of a complete metric, given $x,y\in A^\omega$ such that $x\neq y$, let $n$ be the least natural number such that $x(n)\neq y(n)$ and set $d(x,y)=1/n$.
\begin{proposition}
The metric on $A^\omega$ defined above is complete.
\end{proposition}
\begin{proof}
Let $x_i$ be a Cauchy sequence in $A^\omega$. We can inductively construct an increasing sequence of indices $i_0,i_1,\ldots$ such that for all $n$ and $i\geq i_n$ we have $d(x_{i_n},x_i)<1/n$. In other words for $i\geq i_0$ the $x_i(0)$ all agree, for $i\geq i_1$ the $x_i(1)$ all agree, etc. Thus we may define an element $x$ by letting $x(n)=$ this agreed upon value. Now it is easy to check that $x_i\to x$.
\end{proof}
It is not difficult to check this metric on $A^\omega$ gives the same topology as the product topology described above. While this metric was fairly natural, it is not canonical. For example, any reordering of $\omega$ would give rise to a different metric compatible with the toplogy.
We now discuss a variety of topological properties in the context of sequence space. The closed sets, nowhere dense sets, and compact sets all have special descriptions in sequence space (compact sets are discussed in the next section).
To begin, let $A^{<\omega}=\bigcup A^n$ denote the set of finite sequences of elements of $A$. This set is partially ordered by the subset relation, but we employ special terminology in this case. If $s\subset t$ we say that $s$ is an \emph{initial segment} of $t$ or alternatively that $t$ is an \emph{extension} of $s$. We use the same terminology if $s\in A^{<\omega}$, $x\in A^\omega$, and $s\subset x$: $s$ is a finite initial segment of $x$, or $x$ is an infinite extension of $s$.
A subset $T\subset A^{<\omega}$ is said to be a \emph{tree} if it is closed under initial segments. An element $x\in A^\omega$ is said to be a \emph{branch} through $T$ if all of its finite initial segments $x\restriction n$ lie in $T$. We denote by $[T]$ the subset of $A^\omega$ consisting of all branches through $T$.
\begin{proposition}
A subset $C\subset A^\omega$ is closed if and only if there exists a tree $T\subset A^{<\omega}$ such that $C=[T]$.
\end{proposition}
\begin{proof}
Conversely given any set $B\subset A^\omega$, we let $T_B$ be the tree consisting of all $s$ such that $V_s\cap B\neq\emptyset$ (that is, all initial segments of elements of $B$). We will show that the set $[T_B]$ of all branches through the tree $T_B$ is precisely the \emph{closure} of $B$, which implies the desired result. For this, note that $x$ lies in the closure of $B$ if and only if every open neighborhood of of $x$ meets $B$. This is equivalent to the statement that for every initial segment $s\subset x$, $V_s\cap B\neq\emptyset$. Finally, this is equivalent to the statement that $x$ is a branch through $T_B$.
\end{proof}
% Don't you have to check that $[T]$ is always closed?
In previous sections we have defined nowhere dense subsets of $\R$ in terms of intervals. In fact we can define nowhere dense subsets of any topological space in terms of open sets: $S\subset X$ is \emph{nowhere dense} in $X$ if every open set has an open subset disjoint from $S$. This definition can even be made with basic open sets in place of open sets. Thus we have the following characterization.
\begin{proposition}
\label{prop:cantor-space-nwd}
$S\subset A^\omega$ is nowhere dense if and only if for every $s\in A^{<\omega}$ there exists $t$ such that $s\subset t$ and $V_t$ is disjoint from $S$.
\end{proposition}
Meager sets are now defined in the same way as before: $X\subset A^{<\omega}$ is \emph{meager} if it is the union of countably many nowhere dense sets. Our proof of the Baire category theorem for the real numbers naturally extends to arbitrary complete metric spaces.
\begin{theorem}[Baire category theorem]
If $X$ is a complete metric space then $X$ is not meager. Moreover if $O$ is a nonempty open subset of $X$ then $O$ is not meager.\qed
\end{theorem}
In particular in Cantor space and Baire space we can divide the subsets into three sizes, meager, comeager, and neither. The proof of this general Baire category theorem is nearly identical to the proof of Theorem~\ref{thm:baire} with closed intervals replaced by closures of basic open sets $\overline{O_n}$. The completeness of $X$ is used to verify there is a point $x\in\bigcap\overline{O_n}$.
% compact sets next section
\begin{exercise}
Check that the metric on $A^\omega$ defined by $d(x,y)=1/{n+1}$, where $n$ is least such that $x(n)=y(n)$, gives rise to the same topology as the product topology with basis consisting of all $V_s$.
\end{exercise}
\begin{exercise}
Give an example of an open subset of $\omega^\omega$ which is not closed.
\end{exercise}
\begin{exercise}
Show that the sequence space $A^\omega$ is homeomorphic to the cartesian product with itself $A^\omega\times A^\omega$.
\end{exercise}
\begin{notes}
For introductory material on the combinatorics and topology of sequence space, see Kechris, Sections~2A and~2B. For another proof that the Baire space is homeomorphic to the irrationals, see Section~1 of Miller, \emph{Descriptive set theory and forcing}.
\end{notes}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Cardinal characteristics}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{$\sigma$-compactness in Baire space}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
As we have discussed in the previous part, the exact cardinal value of the size of the set of all real numbers is independent of the axioms of set theory. That is, if we write $\mathfrak c=|\R|$ then we know $\mathfrak c=\aleph_\alpha$ for some $\alpha\geq1$, but we do not know which one. The letter $\mathfrak c$ with its German font stands for \emph{continuum}. Since $\mathfrak c$ is also equal to the cardinality of Cantor space $2^\omega$, it is also often denoted $2^{\aleph_0}$. In this section we discuss several cardinal values other than $\mathfrak c$ that arise from the special topology of the Baire space $\omega^\omega$.
We have measured the size of sets of real numbers using cardinality, category, and measure. We now introduce a fourth notion in the Baire space. A set $A\subset\omega^\omega$ is said to be \emph{$\sigma$-compact} if it is the union of countably many compact sets.
Like the meager and null sets, the class of $\sigma$-compact sets is closed under countable unions. While the class of $\sigma$-compact sets is not closed under subsets, we instead consider the class of subsets of $\sigma$-compact sets.
\begin{definition}
Let $\mathcal K_\sigma$ denote the set of subsets of $\omega^\omega$ which are contained in a $\sigma$-compact set.
\end{definition}
In a moment we will show that the whole Baire space $\omega^\omega$ is not in $\mathcal K_\sigma$, so that $\sigma$-compactness really is a meaningful notion of a small set. First we provide a sequence of characterizations to help us better understand which subsets of Baire space are compact and which are $\sigma$-compact. First we define the natural partial ordering $\leq$ on $\omega^\omega$ by $f\leq g$ if for all $n$, we have $f(n)\leq g(n)$.
% Among the spaces we've explored so far, the notion of $\sigma$-compactness only works in the Baire space. That's because both $2^\omega$ and $\R$ both have the property that the whole space is $\sigma$-compact: $\R$ is the union of closed bounded intervals, and $2^\omega$ is itself compact. On the other hand, the Baire space $\omega^\omega$ is not $\sigma$-compact, as we now aim to show.
\begin{lemma}
\label{lem:baire-compact}
A subset $A\subset\omega^\omega$ is compact if and only if it is a closed and $\leq$-bounded.
\end{lemma}
\begin{proof}
Let $A\subset\omega^\omega$ be compact, and suppose towards a contradiction that $A$ is not $\leq$-bounded. Then there exists some coordinate $n$ such that $A$ contains a sequence of elements $f_i$ with $f_i(n)\to\infty$. Then any subsequence of $f_i$ fails to converge on coordinate $n$, so $f_i$ does not have a convergent subsequence. This contradicts that $A$ is compact.
Conversely note that if $h\in\omega^\omega$ then the set $B_h=\{f\in\omega^\omega\mid f\leq h\}$ of elements $\leq$-bounded by $h$ is compact, because it is a product of finite sets. (This follows from Tychonoff's theorem, but can easily be checked directly.) The result now follows from the fact that any closed subset of a compact set is compact.
\end{proof}
Next we defnie the \emph{eventual domination} order on Baire space: $f\leq^*g$ if $f\leq g$ almost everywhere, that is, there is some $N$ such that for all $n\geq N$ we have $f(n)\leq g(n)$. Note that eventual domination is a quasi-order not a partial order, because $f\leq^*g\leq^*f$ only implies that $f$ and $g$ agree almost everywhere (denoted $f=^*g$).
\begin{lemma}
\label{lem:sigma-compact-bounded}
A subset $A\subset\omega^\omega$ is in $\mathcal K_\sigma$ if and only if it is $\leq^*$-bounded.
\end{lemma}
\begin{proof}
Suppose $A$ is in $\mathcal K_\sigma$. Then there exist compact sets $A_i$ such that $A\subset\bigcup A_i$. By the previous lemma, each $A_i$ is $\leq$-bounded by some $h_i$. Let $h$ be a single function such that for all $n$ we have $h_i\leq^* h$ (for instance, let $h(n)=\max_{i\leq n}h_i(n)$). Then $A$ is $\leq^*$-bounded by $h$.
Conversely suppose that $A$ is $\leq^*$-bounded by $h$. Let $h_n$ enumerate all possible finite modifications of $h$. Then letting $B_n=\{f\in\omega^\omega\mid f\leq h_n\}$, the previous lemma implies $B_n$ are compact sets, and we have $A\subset\bigcup B_n$.
\end{proof}
Since the Baire space $\omega^\omega$ is not $\leq^*$-bounded, we may now conclude that $\omega^\omega$ itself is not in $\mathcal K_\sigma$.
\begin{remark}
\label{rem:ksigma-meager}
It follows from Lemma~\ref{lem:sigma-compact-bounded} that every $\Ksigma$ subset of $\omega^\omega$ is meager. Indeed, compact sets are closed, and moreover since they are bounded they cannot contain any basic open sets $V_s$. It follows that compact sets are nowhere dense, and hence $\sigma$-compact sets are meager.
% This in turn implies the Baire category property for $\sigma$-compact sets: $\omega^\omega$ is not $\sigma$-compact and any basic open neighborhood $V_s$ is not $\sigma$-compact.
\end{remark}
% Suppose towards a contradictoin that $\omega^\omega=\bigcup A_n$ where each $A_n$ is compact. Then by the Baire category theorem, at least one of the $A_n$ must have a nonempty interior, say $V_s\subset A_n$. But $V_s$ is not contained in a product of bounded intervals, so by Lemma~\ref{lem:baire-compact} this contradicts that $A_n$ is compact.
We now introduce for the first time the concept of associating \emph{cardinal characteristics} to a collection of sets, in this case $\mathcal K_\sigma$.
\begin{definition}
\begin{itemize}
\item The cardinal $\non(\Ksigma)$ is the least cardinality of any set $A\subset\omega^\omega$ which is not $K_\sigma$.
\item The cardinal $\cov(\Ksigma)$ is the least cardinality of any family $\mathcal F$ of subsets of $\omega^\omega$ whose union covers all of $\omega^\omega$.
\end{itemize}
\end{definition}
Our next result generalizes Cantor's theorem by showing that the cardinals $\non(\Ksigma)$ and $\cov(\Ksigma)$ are uncountable lower bounds for the value of the continuum.
\begin{theorem}
\label{thm:k-sigma}
We have $\aleph_1\leq\non(\Ksigma)\leq\mathfrak c$ and $\aleph_1\leq\cov(\Ksigma)\leq\mathfrak c$.
\end{theorem}
\begin{proof}
It is clear that $\aleph_1\leq\non(\Ksigma)$ because every countable set is $\Ksigma$, being the union of its points. The inequality $\non(\Ksigma)\leq\mathfrak c$ follows from the fact that $\omega^\omega$ itself is not $\sigma$-compact.
To show that $\aleph_1\leq\cov(\Ksigma)$, suppose a countable family of $\sigma$-compact sets covered all of $\omega^\omega$. Then a countable family of compact sets would also cover all of $\omega^\omega$, again contradicting that $\omega^\omega$ is not $\sigma$-compact. Finally it is clear that $\cov(\Ksigma)\leq\mathfrak c$, since the whole space $\omega^\omega$ is the union of its points and $|\omega^\omega|=\mathfrak c$.
\end{proof}
In the rest of this section we show that the cardinals $\non(\Ksigma)$ and $\cov(\Ksigma)$ may be characterized using the $\leq^*$ quasi-ordering alone, and use this to establish that $\non(\Ksigma)\leq\cov(\Ksigma)$.
\begin{definition}
\begin{itemize}
\item A subset $\mathcal F\subset\omega^\omega$ is said to be an \emph{unbounded family} if there is no single $h\in\omega^\omega$ such that $f\leq^*h$ for all $f\in\mathcal F$. The \emph{unbounding number} $\mathfrak b$ is the smallest cardinality of any unbounded family.
\item A subset $\mathcal F\subset\omega^\omega$ is said to be a \emph{dominating family} if for all $h\in\omega^\omega$ there exists $f\in\mathcal F$ such that $h\leq^*f$. The \emph{dominating number} $\mathfrak d$, is the smallest cardinality of any dominating family.
\end{itemize}
\end{definition}
The following result records the relationship between these two cardinal values.
\begin{proposition}
\label{prop:bd}
We have $\mathfrak{b}\leq\mathfrak{d}$.
\end{proposition}
\begin{proof}
It is enough to show that every dominating family is unbounded. Let us establish the contrapositive: if $\mathcal F$ is bounded, say by $h\in\omega^\omega$, then every $f\in\mathcal F$ satisfies $f\leq^*h$. Letting $h'(n)=h(n+1)$, it follows that every $f\in\mathcal F$ satisfies $h'\not\leq^*f$, and hence that $\mathcal F$ is not a dominating family.
\end{proof}
\begin{theorem}
\label{thm:bd-vs-ksigma}
We have $\mathfrak b=\non(\Ksigma)$ and $\mathfrak d=\cov(\Ksigma)$. As a consequence we have $\non(\Ksigma)\leq\cov(\Ksigma)$.
\end{theorem}
\begin{proof}
The first equality follows directly from Lemma~\ref{lem:sigma-compact-bounded}.
The second equality uses the same ideas again. If $\mathcal F$ is a dominating family then consider the collection of all $A_f=\set{g\mid g\leq f}$ for $f$ a finite modification of an element of $\mathcal F$. This latter collection is a family of compact sets that covers all of $\omega^\omega$. Conversely if $\mathcal F$ is a family of compact subsets of $\omega^\omega$ that covers all of $\omega^\omega$, then each $F\in\mathcal F$ is $\leq$-bounded by some $h_F\in\omega^\omega$. It follows that the collection of all $h_F$ is a dominating family.
\end{proof}
\begin{exercise}
Prove directly that a countable product of finite metric spaces is compact.
\end{exercise}
\begin{exercise}
Find an example of a meager subset of $\omega^\omega$ which is not $\Ksigma$.
\end{exercise}
\begin{exercise}
Consider the space $\R^\omega$ with the product topology. Is it $\sigma$-compact?
\end{exercise}
\begin{notes}
The conection between $\mathcal K_\sigma$ and the domination relation $\leq^*$ is presented in Section~2 of Blass, \emph{Combinatorial cardinal characteristics of the continuum}.
\end{notes}
\newpage
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Ideals and their cardinals}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
Let $X$ be any one of the three spaces $\R$, $2^\omega$, or $\omega^\omega$. An \emph{ideal} on $X$ is a subset $\mathcal I\subset\mathcal P(X)$ which is closed under subsets and (finite) unions. We think of an ideal as any collection of sets that captures some quality of smallness in subsets of $X$. Of course, the whole space $X$ should not be small, so we are only interested in \emph{proper} ideals, i.e.\ those do not contain $X$.
An ideal $\mathcal I$ is called a \emph{$\sigma$-ideal} if it is additionally closed under countable unions. In the past several sections, we have discussed three key notions of smallness and each one is a $\sigma$-ideal.
\begin{itemize}
\item $\Meager$ is the ideal of meager subsets of $\R$.
\item $\Null$ is the ideal of Lebesgue null subsets of $\R$.
\item $\Ksigma$ is the ideal of subsets of $\omega^\omega$ which are contained in a $\sigma$-compact subset of $\omega^\omega$.
\end{itemize}
As was the case with $\Ksigma$, we can associate cardinal characteristics to any ideal. We will consider four key characteristics.
\begin{definition} Let $\mathcal I$ be an ideal of subsets of a set $X$, such that $\mathcal I$ contains all singletons of $X$.
\begin{itemize}
\item The \emph{additivity} of $\mathcal{I}$, $\add(\mathcal{I})$, is the smallest number of sets in $\mathcal{I}$ whose union is not in $\mathcal{I}$.
\item The \emph{uniformity} of $\mathcal{I}$, $\non(\mathcal{I})$, is the smallest cardinality of any subset of $X$ that is not in $\mathcal{I}$.
\item The \emph{covering number} of $\mathcal{I}$, $\cov(\mathcal{I})$, is the smallest number of sets in $\mathcal{I}$ whose union covers all of $X$.
\item The \emph{cofinality} of $\mathcal{I}$, $\cof(\mathcal{I})$, is the smallest cardinality of any subset $\mathcal{B}\subset\mathcal{I}$ such that every element of $\mathcal{I}$ is a subset of an element of $\mathcal{B}$.
\end{itemize}
\end{definition}
\begin{remark}
The cofinality of an ideal is the least number of sets you need to generate the ideal by closing under subsets. More precisely, a subset $\mathcal B\subset\mathcal I$ is called a \emph{basis} for $\mathcal I$ if every element of $I$ is a subset of some element of $\mathcal B$. Thus the cofinality of $\mathcal I$ is the least cardinality of a basis for $\mathcal I$.
\end{remark}
As was the case with the cardinals we introduced in the previous section, assuming $\mathcal I$ is reasonable, all four of the cardinals characteristics associated with $\mathcal I$ provides an uncountable lower bound for the value of the continuum.
\begin{lemma}
\label{lem:ideal-cards}
Suppose that $\mathcal I$ is a proper $\sigma$-ideal, $\mathcal I$ contains the singletons, and that $\mathcal I$ has a basis consisting of Borel sets. Then each of the four cardinals above is uncountable and bounded above by $\mathfrak c$.
\end{lemma}
\begin{proof}
The additivity of $\mathcal I$ is uncountable simply because $\mathcal I$ is a $\sigma$-ideal. In the next lemma we will show that if $\mathcal I$ is a proper $\sigma$-ideal containing the singletons, then the additivity of $\mathcal I$ is a lower bound for all four cardinal characteristics of $\mathcal I$, and hence all four are uncountable.
Next, it is easy to verify that for any proper $\sigma$-ideal $\mathcal I$ containing the singletons, the additivity, uniformity, and covering number of $\mathcal I$ are bounded above by $\mathfrak c$. If additionally $\mathcal I$ has a basis consisting of Borel sets, then since there are only $\mathfrak c$ many Borel sets, we have that the cofinality of $\mathcal I$ is bounded above by $\mathfrak c$ too.
\end{proof}
All three of the $\sigma$-ideals we have introduced have a basis of Borel sets. For $\Ksigma$ this is clear because $\sigma$-compact sets are Borel. For $\Meager$ any meager set is contained in a countable union of \emph{closed} nowhere dense sets, which are Borel. Finally for $\Null$, note that if $A$ is null then for all $n$ there is a Borel set $A_n$ such that $A\subset A_n$ and $m(A_n)<1/n$. It follows that $A$ is contained in the Borel null set $\bigcap A_n$.
\begin{remark}
There can exist proper $\sigma$-ideals containing the singletons that don't have a basis of size $\leq\mathfrak c$. For example, it is consistent that the ideal $\mathcal{SN}$ of \emph{strongly null} sets has $\cof(\mathcal{SN})=\aleph_2$ while $\mathfrak c=\aleph_1$.
[ref] % yorioka
\end{remark}
We next establish the basic relationships between the four cardinal characteristics associated with $\mathcal I$.
\begin{lemma}
\label{lem:diamond}
Suppose that $\mathcal I$ is a proper $\sigma$-ideal and that $\mathcal I$ contains the singletons. Then we have the inequalities $\add(\mathcal I)\leq\non(\mathcal I)\leq\cof(\mathcal I)$, and also $\add(\mathcal I)\leq\cov(\mathcal I)\leq\cof(\mathcal I)$.
\end{lemma}
\begin{proof}
For the inequality $\add(\mathcal{I})\leq\non(\mathcal{I})$, we will show that for every set $A$ not in $\mathcal I$, we can find a family $\mathcal F$ of the same (or lower) cardinality such that $\bigcup\mathcal F\notin I$. This is easy: if $A\notin\mathcal I$, then the family $\mathcal F$ consisting of all $\{a\}$ such that $a\in A$ has the same cardinality as $A$ and satisfies $\bigcup\mathcal F\notin\mathcal I$.
The proofs of each of the next three inequalities follows the same form: given a set indicated by the right-hand side, find a set of the same or lower cardinality indicated by the left-hand side. So to show $\non(\mathcal I)\leq\cof(\mathcal I)$, let $\mathcal F\subset \mathcal{I}$ be a basis for $\mathcal I$. For each set $B\in\mathcal F$ choose an element $x_B\notin B$ and let $A=\set{x_B\mid B\in\mathcal F}$. Then $A$ has the same or lower cardinality as $\mathcal F$, and we claim that $A\notin\mathcal{I}$. Indeed if $A\in\mathcal{I}$ then we would have $A\subseteq B$ for some $B\in\mathcal F$, contradicting that $x_B\notin B$.
For the inequality $\add(\mathcal I)\leq\cov(\mathcal I)$, if $\mathcal F\subset\mathcal{I}$ satisfies $\bigcup\mathcal F=X$, then $\mathcal F$ itself satisfies $\bigcup\mathcal F\notin\mathcal I$.
For the inequality $\cov(\mathcal I)\leq\cof(\mathcal I)$, again suppose that $\mathcal F$ is a basis for $\mathcal I$. Since $\mathcal I$ contains the singletons, it follows that $\mathcal F$ itself covers $X$.
\end{proof}
The conclusions of the two lemmas are summarized in Figure~\ref{fig:ideal}
\begin{figure}[h]
\begin{tikzpicture}[->]
\node (a1) {$\aleph_1$};
\node (add) [right=of a1] {$\add(\mathcal I)$} edge (a1);
\node (non) [above=of add] {$\non(\mathcal I)$} edge (add);
\node (cov) [right=of add] {$\cov(\mathcal I)$} edge (add);
\node (cof) [above=of cov] {$\cof(\mathcal I)$}
edge (non) edge (cov);
\node [right=of cof] {$\mathfrak{c}$} edge (cof);
\end{tikzpicture}
\caption{The relationships between the cardinal characteristics of $\mathcal I$, assuming $\mathcal I$ satisfies the hypotheses of the two lemmas. In the figure, the arrow $\leftarrow$ represents the inequality $\leq$.\label{fig:ideal}}
\end{figure}
In the previous section we introduced $\Ksigma$, but we dealt only with the characteristics $\non(\Ksigma)$ and $\cov(\Ksigma)$. We conclude this section by giving the values of $\add(\Ksigma)$ and $\cof(\Ksigma)$ as well.
\begin{proposition}
We have $\add(\Ksigma)=\mathfrak b$ and $\cof(\Ksigma)=\mathfrak d$.
\end{proposition}
\begin{proof}
We have already shown that $\add(\Ksigma)\leq\non(\Ksigma)=\mathfrak b$. For the reverse inequality, first recall from Theorem~\ref{thm:bd-vs-ksigma} that a subset of $\omega^\omega$ is unbounded if and only if it is not $\Ksigma$. Now suppose that $\mathcal F$ is a family of $\Ksigma$ subsets of $\omega^\omega$ such that $\bigcup\mathcal F$ is not $\Ksigma$. For each $F\in\mathcal F$ let $g_F$ be a bound for $F$. Then $\set{g_F\mid F\in\mathcal F}$ is an unbounded family: Indeed, otherwise the set of things bounded by some $g_F$ would be bounded and hence so would $\bigcup\mathcal F$. This shows that $\add(\Ksigma)\geq\mathfrak b$.
The second equality is similar and left as Exercise~\ref{exerc:non-ksigma-d}.
\end{proof}
\begin{exercise}
\label{exerc:non-ksigma-d}
Show that $\cof(\Ksigma)=\mathfrak d$.
\end{exercise}
\begin{exercise}
Let $\mathcal I$ be the ideal of all countable subsets of $\R$. What are the values of the four cardinal characteristics of $\mathcal I$?
\end{exercise}
\begin{notes}
We have finally entered into the material covered in the book Bartoszy\'nski--Judah, \emph{Set theory: on the structure of the real line}. The basic properties of $\sigma$-ideals and their cardinals begins in Section~1.3.
\end{notes}