-
Notifications
You must be signed in to change notification settings - Fork 5
/
Copy pathMAIN
3203 lines (2754 loc) · 275 KB
/
MAIN
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{book}
\usepackage[utf8]{inputenc}
\usepackage{amsfonts, amsthm, amsmath, amssymb, mathrsfs,mathabx}
\usepackage{enumerate}
\usepackage{hyperref}
\usepackage{units}
\usepackage{mathtools}
\usepackage{graphicx}
\graphicspath{ {images/} }
\usepackage{caption}
\usepackage{subcaption}
\usepackage{tikz}
\usepackage{tikz-3dplot}
\usepackage{standalone}
\usepackage[numbers]{natbib}
%defined functions and such
\DeclareMathOperator*{\Rcal}{\mathcal{R}}
\DeclareMathOperator*{\Q}{\mathbb{Q}}
\DeclareMathOperator*{\R}{\mathbb{R}}
\DeclareMathOperator*{\C}{\mathbb{C}}
\DeclareMathOperator*{\N}{\mathbb{N}}
\DeclareMathOperator*{\Aut}{\text{Aut}}
\DeclareMathOperator*{\Int}{\text{Int}}
\DeclareMathOperator*{\Out}{\text{Out}}
\DeclareMathOperator*{\E}{\mathbb{E}}
\DeclareMathOperator*{\Ocal}{\mathcal{O}}
\DeclareMathOperator*{\Ft}{\text{Fr}}
\DeclareMathOperator*{\Image}{\text{Im}}
\DeclareMathOperator*{\F}{\mathcal{F}}
\DeclareMathOperator*{\Diam}{\text{Diam}}
\DeclareMathOperator*{\RP}{\R\text{P}}
\DeclareMathOperator*{\Z}{\mathbb{Z}}
\DeclareMathOperator*{\Orb}{\text{Orb}}
\DeclareMathOperator*{\Stab}{\text{stab}}
\DeclareMathOperator*{\card}{\text{card}}
\DeclareMathOperator*{\Star}{\text{star}}
\DeclareMathOperator*{\car}{\text{carrier}}
\DeclareMathOperator*{\KB}{\text{KB}}
\DeclareMathOperator*{\MB}{\text{MB}}
%new commands
\newenvironment{Lemma}[1][Lemma]
{\proof[#1]\leftskip=1cm\rightskip=1cm}
{\endproof} %this make a lemma indented in proof
\newcommand{\bij}{\mathrel{\hookrightarrow\hspace{-1.8ex}\to}} %creates a bijection arrow
\newcommand*\productop{\mathbin{\Pi}} %disjoint union
\newcommand\inner[2]{\langle #1, #2 \rangle}
\newcommand\innerone[1]{\langle #1 \rangle}
%NEW ENVIRONMENTS
\newenvironment{acknowledgements}%
{\cleardoublepage\thispagestyle{empty}\null\vfill\begin{center}%
\bfseries Acknowledgements\end{center}}%
{\vfill\null}
%counters and reformating:
\setcounter{section}{-1}
\setcounter{chapter}{-1}
\title{Armstrong Topology Solutions}
\author{George Blikas \\ Gregory Grant \\ David Hong \\ Grant Baker \\ Kristy Ong }
\date{April 2016}
\begin{document}
%creates title and the empty page style allows
%for the format of the blank page after
%maketitle to actually be blank (w/o numbering).
\maketitle
\thispagestyle{empty}
\tableofcontents
\begin{acknowledgements}
George B: Much thanks to all the contributors of this book. Special thanks goes to Gregory Grant, and particularly in helping establish chapter \hyperref[chapter:intro]{\ref{chapter:intro}}, section \hyperref[sec:productspaces]{\ref{sec:productspaces}}, and problems $10$, $11$, and $12$ of section $4.2$, all of which greatly sped-up the processes of this project.
\end{acknowledgements}
\chapter{Preliminaries}
\section{Definitions}
\subsection{Functions}
\begin{enumerate}[(1)]
\item Let $f:A \rightarrow B$ be a function. Then,
\begin{itemize}
\item $f$ is onto if, and only if, there exists a function $g: B \rightarrow A$ such that $fg = 1_B$.
\item $f$ is one-to-one if, and only if, there exists a function $g: B \rightarrow A$ such that $gf = 1_A$ (provided $A$ is non-empty).
\item $f$ Is in 1-1 correspondence if there exists a function $g: B \rightarrow A$ such that $fg = 1_B$ and $gf = 1_A$. In this case, $g$ is unique and is called the inverse function of $f$, typically denoted $f^{-1}$.
\end{itemize}
\end{enumerate}
\section{Extra Proofs and Lemmas}
\subsection{Group Theory \& Abstract Algebra}
\begin{enumerate}[(1)]
\item Split Exact Sequence: \label{sec:splitexact}
Given an exact sequence of abeilean groups and homomorphisms,
\[0 \xrightarrow[]{} G \xrightarrow{\theta} H \xrightarrow{\phi} K \xrightarrow[]{} 0, \]
and a homomorphism $\psi: K \rightarrow H$, such that $\phi\psi \equiv 1_K$, we have $H \cong G \oplus K$.
\begin{proof}
Define $\alpha: G \oplus K \rightarrow H$ by $\alpha(g \oplus k) = \theta(g) + \psi(k)$: it is easy to see that $\alpha$ is a homomorphism. Also, $\alpha$ is (1-1), for if $\alpha(g \oplus k) = 0$, we have
\[0 = \phi(\theta(g) + \psi(k)) = \phi\psi(k) = k;\]
but, then $\theta(g) = 0$, so that $g = 0$ since $\theta$ is one-to-one.
\par Moreover, $\alpha$ is onto, since given $h \in H$, we have
\[\phi(h - \psi\phi(h)) = \phi(h) - \phi\psi\phi(h) = 0\]
Thus, there exists $g \in G$ such that $h - \psi\phi(k) = \theta(g)$, that is,
\[h = \theta(g) + \psi\phi(h) = \alpha(g \oplus \psi(h))\]
\end{proof}
\end{enumerate}
\subsection{Limit Points, Closure \& Density}
\begin{enumerate}[(1)]
\item Let $(X, \tau)$ be a topological space, and $A \subset X$. We show that $\overline{A} = A \cup L_A$, where $L_A$ is the set of all accumulation points of $A$:
\begin{proof} If $\overline{A} = A$, then by Theorem $2.2$, pg. $29$, $L_A \subset A$ and we are done. So, suppose that $A$ is not closed. Then, $A$ does not contain all of its limit points; consequently, $L_A \neq \emptyset$. Further, the set $B = A \cup L_A$ is closed by Theorem $2.2$, pg. $29$.
\par To conclude the proof, we show that any closed set $C$ containing $A$, contains $B$. Indeed; Let $C \subset X$ be closed, such that $A \subset C$. Now, for $a \in L_A$ and open set $O_a$ containing $a$, we have that $A \cap (O_a - \{a\}) \neq \emptyset$. But, as $A \subset C$, we have $C \cap (O_a - \{a\}) \neq \emptyset$. Therefore, $a \subset C$. So, $L_A \subset C$, implying $A \cup L_A \subset C$. By Theorem $2.3$, pg. $30$, $\overline{A} = B = A \cup L_A$.
\end{proof}
\item Let $(X, \tau)$ be a topological space, and $A \subset X$. We prove that $L_A$, defined above, contains all limit points of sequences contained in $A$:\footnote{This assumes the general definition of limit points of a set}
\begin{proof} Let $\{a_n\}_{n=1}^\infty \subset A$, such that $a_n \rightarrow a$. Then, $a \in L_{\{a_n: n \in \mathbb{N}\}}$ and every neighbourhood of $a$ contains a point of $\{a_n: n \in \mathbb{N}\} - \{a\} = B$. As $\{a_n: n \in \mathbb{N}\} \subset A$, every neighbourhood of $a$ has a point in $A - \{a\}$; $a \in L_A$.
\end{proof}
\item Let $(X, \tau)$ be a topological space, and $A \subset X$ such that $\overline{A} = X$. We prove that $A \cap O \neq \emptyset$, for all $O \neq \emptyset$, $O \in \tau$:
\begin{proof} Suppose that for some $O \in \tau$, $O \neq \emptyset$, $A \cap O = \emptyset$. By a previous lemma, we have $\overline{A} = (L_A - A) \cup O^c$, implying $O^c = A$. But then we have $X = O^c = \overline{O^c} = \overline{A}$, as $O$ is open. But, contrarily, this implies that $(O^c)^c = O = X^c = \emptyset$.
\end{proof}
\item We prove that the intersection of a closed set and a compact set is always compact:
\begin{proof} Let $(X, \tau)$ be a topological space. Let $H,K \subset X$, such that $H$ is closed and $K$ is compact. Consider $H \cap K$. Now, if $\{O_{\alpha}\}_{\alpha}$ is an open cover of $H \cap K$, then $K \subset \bigcup_{\alpha \in N} O_\alpha \cup (X - H)$. But, since $H$ is closed, $X - H$ is open. In conclusion, as
$$H \cap K \subset \bigcup_{\alpha \in N} O_\alpha \cup (X - H)$$
such a finite subcover of $H \cap K$ exists.
\end{proof}
\item We prove that if $(X, d)$ is a metric space with the induced topology, then $C \subset X$ is closed if, and only if, whenever $\{a_n\}_{n=1}^\infty$ is a sequence in $C$, with $\{a_n\} \rightarrow L$, we have $L \in C$:
\begin{proof}
\begin{enumerate}
$ $\newline
\item[] ($\implies$): Suppose, to the contrary, that $\overline{C} = C$, but $L \notin C$. Thus, there exists some $\epsilon > 0$, such that $B_\epsilon(L) \cap C = \emptyset$, as $L$ is not a limit point of $C$. But then, $a_n \notin B_\epsilon(L)$, for all $n \geq N$, $N \in \mathbb{N}$ is sufficiently large; A contradiction.
\item[] ($\impliedby$): Suppose, to the contrary, that $\overline{C} \neq C$. Then, by extra lemma\footnote{reference this}, there is some $l \in L_C$, such that $L \notin C$. Thus, for each $n \in \mathbb{N}$, we pick $a_n \in B_{\nicefrac{1}{n}}(L) \cap C \neq \emptyset$. Consequently, $\{a_n\}_{n=1}^\infty$ is a sequence in $C$ such that $a_n \rightarrow L \in C$, by hypothesis; A contradiction.
\end{enumerate}
\end{proof}
\end{enumerate}
\subsection{Separation}
\begin{enumerate}[(1)]
\item We show that a compact $T_2$ space $T_3$. Consequently, we show that it is $T_4$:
\begin{proof} Let $X$ be the compact $T_2$ space. We first show that $X$ is $T_3$:
\par Let $A \subset X$, such that $A$ is closed. Then $A$ is compact. Further, let $b \in X$ such that $b \notin A$. Now, for each $a \in A$, there exists an open set $O_a$, and some open set $O_b^a$, such that $O_a \cap O_b^a = \emptyset$. It follows that $A \subset \bigcup_{a \in N} O_a \subset \bigcup_a O_a$ and that $\bigcup_{a \in N} O_a = O_A$ is open. Further, $b \in \bigcap_{a \in N} O_b^a = O_b$, is open. And by construction, we have $O_A \cap O_b = \emptyset$. So, $X$ is $T_3$.
\par To show that $X$ is $T_4$, let $A, B \subset X$, be disjoint and closed. Then, arguing as above, we have two disjoint open subsets $O_B \in \tau$, $O_A \in \tau$ with $O_A \cap O_B = \emptyset$.
\end{proof}
\end{enumerate}
\subsection{Compactness}
\begin{enumerate}[(1)]
\item We show that a homeomorphism between locally compact $T_2$ spaces, $X$, $Y$, extends to a homeomorphism between the Alexandroff compactifications; in other-words, locally compact homeomorphic $T_2$ spaces have homeomorphic one point compactifications.
\begin{proof} Suppose that $f: X \rightarrow Y$ is the homeomorphism. Define $g: X \cup \{\infty_1\} \rightarrow Y \cup \{\infty_2\}$ as follows:
$$ g(x) =
\begin{cases}
f(x) & x \neq \infty_1 \\
\infty_2 & x = \infty_1
\end{cases}
$$
It is clear that as $f$ is a bijection, so is $g$. We show that $g$ is a homeomorphism:
\par Suppose that $U$ is open in $Y$. Then, $g^{-1}[U] = f^{-1}[U]$, which is open in $X \cup \{\infty_1\}$. Now, if ${\Ocal}$ is open in $Y \cup \{\infty_2\}$, we have ${\Ocal} = K \cup \{\infty_2\}$, where $K \subset Y$ is compact. Then,
$$g^{-1}[{\Ocal}] = g^{-1}[K \cup \{\infty_2\}] = g^{-1}[K] \cup g^{-1}[\{\infty_2\}] =g^{-1}[K] \cup \{\infty_1\}$$
which is open in $X \cup \{\infty_1\}$.
\par Consequently, as $g$ is one-to-one and onto, since $f$ is, $g$ is a homeomorphism by Theorem $3.7$, pg. $48$.
\end{proof}
\item We show that if $(X, \tau)$ is a topological space, and $S \subset X$ is compact, then if $W$ is a neighbourhood of $S$, there exists another neighbourhood $G$, such that
$$S \subset G \subset W$$
\begin{proof} By exercise $21$, section $3$ pg. $55$, and the fact that $X \cup \{y\} \cong X$, the result follows.
\end{proof}
\end{enumerate}
\subsection{Connectedness}
\begin{enumerate}[(1)]
\item We show that if $X$ is locally connected, then every connected component of $X$ is open in $X$; hence $X$ is the disjoint union of its connected components:
\begin{proof} Let $x \in Y$, where $Y$ is a connected component of $X$. By definition, $x$ is contained in some open connected subset $U$ of $X$. Since $Y$ is a maximal connected set containing $x$, we have $x \in U \subset Y$. This shows that $Y$ is open in $X$.
\end{proof}
\end{enumerate}
\subsection{Quotient Maps \& Quotient Topology}
\begin{enumerate}[(1)]
\item We show that if $q: X \rightarrow Y$ is a quotient map\footnote{Armstrong does not do a good job describing what the topology on $Y$ is. A simple exercise shows that by letting ${\Ocal}$ be open in $Y$ whenever $q^{-1}[{\Ocal}]$ is open in $X$, we have a topology on $Y$; call this $\tau_{Y}$. Further, Armstrong does not do an adequate job describing what a quotient map is: $q: X \rightarrow Y$ is a quotient map if it is onto, continuous with respect to $\tau_Y$, and such that $g^{-1}[{\Ocal}]$ is open in $X$ implies ${\Ocal}$ open in $Y$. We can summarize by saying that $q$ is onto, and such that ${\Ocal}$ is open in $Y$ iff $g^{-1}[{\Ocal}]$ is open in $X$.}, then the topology of $Y$ is the largest which makes $q$ continuous:
\begin{proof} Suppose that $\tau$ was some other topology on $Y$, such that $q$ was continuous. We conclude by showing that $\tau \subset \tau_Y$:
\par Indeed; if ${\Ocal} \in \tau$, then $q^{-1}[{\Ocal}]$ is open in $X$. But then, $q[q^{-1}[{\Ocal}]] = {\Ocal} \in \tau_Y$. Thus, $\tau \subset \tau_Y$.
\end{proof}
\item Suppose that $(X, \tau)$ is a topological space, and $A$ a non-empty set. We show that if $f: X \rightarrow A$, then the quotient topology on $A$, $\tau_A$ is indeed a topology:
\begin{proof} By definition, as noted above, ${\Ocal}$ is open in $A$, if $f^{-1}[{\Ocal}]$ is open in $X$:
\begin{enumerate}
\item As $f^{-1}[\emptyset] = \emptyset$, and $f^{-1}[A] = X$, $\emptyset, A \in \tau_A$.
\item As $$f^{-1}[\bigcup_{j \in J} {\Ocal}_j] = \bigcup_{j \in J} f^{-1}[{\Ocal}_j],$$ we see that $\bigcup_{j \in J} {\Ocal}_j \in \tau_A$.
\item Lastly, $$f^{-1}[\bigcap_{j =1 }^\infty {\Ocal}_j] = \bigcap_{j =1}^\infty f^{-1}[{\Ocal}_j]$$ So, $\bigcap_{j =1 }^\infty {\Ocal}_j \in \tau_A$.
\end{enumerate}
\end{proof}
\item We show that projective real n-space, $\mathbb{P}^n$, is Hausdorff\footnote{Credit for the initial idea of the proof goes to Brain M. Scott. In addtion, we use the notation $B(z,z')$ for the open ball of radius $z'$, centered at $z$.}:
\begin{proof} Let $q:S^n \rightarrow \mathbb{P}^n$ be the quotient map, and suppose that $u,v \in \mathbb{P}^n$, $u \neq v$; there are such $x,y \in S^n$, such that $q^{-1}[\{u\}] = \{x, -x\}$, and $q^{-1}[\{v\}] = \{y, -y\}$. We exhibit an $\epsilon \in \mathbb{R}$ such that the open neighbourhoods of radius $\epsilon$, centered at $u,v$ are disjoint:
\par Let $$\epsilon = \frac{1}{2} \min\{ ||x - y ||, ||x +y||\},$$ and set $U = B(x,\epsilon)$, and $V = B(y, \epsilon)$. It follows form the construction of $U,V$, that $U,V,-U,-V$ are pairwise disjoint, and open neighbourhoods of $x,y,-x,-y \in S^n$. Moreover, as noted above, $q^{-1}(q[U]) = U \cup -U$, and $q^{-1}(q[V]) = V \cup -V$. To conclude, we show that $q[U] \cap q[V]$ are disjoint neighbourhoods for $u,v \in \mathbb{P}^n$.
\par Indeed;
\begin{align*}
q[U] \cap q[V] & = q[B(x, \epsilon) \cap S^n] \cap q[B(y, \epsilon) \cap S^n] \\
& = q[B(x, \epsilon)] \cap q[B(y, \epsilon)] \cap \mathbb{P}^n \\
& = B(u, q(\epsilon)) \cap B(v, q(\epsilon)) \cap \mathbb{P}^n \\
& = B_1 \cap B_2 \cap \mathbb{P}^n
\end{align*}
Now, if $B_1 \cap B_2 \neq \emptyset$, then we would have $q^{-1}[B_1 \cap B_2] = U \cap V \neq \emptyset$, contary to assumption. Thus,
$$q[U] \cap q[V] = \emptyset \cap \mathbb{P}^n = \emptyset$$
\end{proof}
\end{enumerate}
\subsection{Topological Groups} \label{chap:sec:sub:topologicalgroups}
\begin{enumerate}[(1)]
\item Suppose that $(G, \tau, *)$ is a topological group. Fixing $x \in G$, we show that $x{\Ocal}\in \tau$ if, and only if, ${\Ocal} \in \tau$:
\begin{proof}
$ $\newline
\begin{itemize}
\item [] ($\implies$): Suppose that $x{\Ocal} \in \tau$. It follows from pg. $75$, that $L_{x^{-1}}: G \rightarrow G$ is a homeomorphism and an open map: thus, $L_{x^{-1}}[{\Ocal}] \in \tau$.
\item [] ($\impliedby$): Suppose that ${\Ocal} \in \tau$. Then, as noted on pg. $75$, $L_x: G \rightarrow G$ is a homeomorphism and an open map: thus, $L_x[{\Ocal}] = x{\Ocal} \in \tau$.
\end{itemize}
\end{proof}
\item Suppose that $(G, \tau, *)$ is a topological group. We show that ${\Ocal}$ is a neighbourhood of $x \in G$, if, and only if, $x^{-1}{\Ocal}$ is a neighbourhood of $e$:
\begin{proof} Suppose that $x \in G$, and without loss of generality, ${\Ocal}$ is an open set containing $x$. Then, by a previous lemma\footnote{reference this}, $L_{x^{-1}}[{\Ocal}] = x^{-1}{\Ocal} \in \tau$, and contains $x^{-1}x = e$.
\par By considering $L_x$, similar logic shows the reverse implication.
\end{proof}
\item We show that $\R$, with the Euclidean topology and addition is a topological group:
\begin{proof} The fact that $(\R, +)$ is a group is clear. To conclude, we show that $+ \equiv m:\R \rightarrow \R$, and $-1 \equiv i: \R \rightarrow \R$ are continuous:
\par Now, $i(x) = -x$, which is polynomial, and so continuous. And, as $m(x,y) = x + y = P_1(x,y) + P_2(x,y)$, where $P_i$ is the projection mapping from $\R^{2}$, $m$ is continuous; it is the sum of two continuous functions.
\par A similar argument show that $(\R^{n}, \tau, +)$ is a topological group.
\end{proof}
\item Let $(G, \tau, m)$ be a topological group. We show that the topological automorphisms of $(G, \tau, m)$, for a subgroup of $\text(Aut)(G)$. We denote the set of all topological automorphism of $G$ by $\Aut_\tau (G)$.
\begin{proof} It is well known that $(\Aut (G), \circ)$ is a group. We show that $(\Aut_\tau (G), \circ)$ is a group:
\par As the composition of automorphism/homeomorphisms is another automorphism/homeomorphism, the fact that $(\Aut_\tau (G), \circ )$ is closed is clear. To conclude, as $f \in \Aut_\tau(G)$ is a homeomorphism, $f^{-1} \in \Aut_\tau(G)$. Furthermore, $e: (G, \tau) \rightarrow (G, \tau)$ given by $e(x) = x$ is a homeomorphism. So, $e \in \Aut_\tau(G)$.
\par By the subgroup test,
$$\Aut_\tau(G) \leq \Aut(G)$$
\end{proof}
\item Suppose that $(G, \tau, m)$ is a $T_2$, finite, topological group. We claim $\tau = \mathcal{P}(G)$:
\begin{proof} As $G$ is $T_2$, $\{x\}$ is open for each $x \in G$. Thus, as the union of open sets is open, it follows that $\tau = \mathcal{P}(G)$.
\end{proof}
\item We claim that all topological groups of order $2$ are topologically isomorphic.
\begin{proof} Let $(G, \tau_G, m)$ be the topological group of order $2$. Consider the topological group $(Z_2, \tau_{Z_2}, +)$ and the map $\phi: G \rightarrow Z_2$, given by $\phi(e_G) = 0$, and $\phi(g) = 1$. The fact that $\phi$ is a group isomorphism is well-known. Further, $\phi$ is 1-1 and onto, and $\phi^{-1}$ is continuous, as $\tau_G = \mathcal{P}(G)$ and $\tau_{Z_2} = \mathcal{P}(Z_2)$. This concludes the proof.
\end{proof}
\end{enumerate}
\subsection{Homotopy Type}
\begin{enumerate}[(1)]
\item We prove that if $A$ is a subspace of a topological space $X$ and $G: X \times I \rightarrow X$ is deformation retract relative to $A$, then $X$ and $A$ have the same homotopy type\footnote{Armstrong does not do a good job of defining a deformation retract. We use the following definition. "Strong" deformation retract: $A$ is a 'strong' deformation retract of $X$ iff there exists a map $D:X \times I \rightarrow X$ such that $D(a,t)=a$ for every $a \in A$, $D(x,0)=x$ and $D(x,1) \in A$ for all $x \in X$. See \href{http://math.stackexchange.com/questions/488527/deformation-retract-and-homotopy-equivalence}{this post on SE.}}:
\begin{proof} Indeed: Let $f:A \rightarrow X$ be the inclusion map and $g: X \rightarrow A$ be defined by $g(x) = G(x,1)$. We have previously shown that $f$ and $g$ are continuous. It is only left to show that $f \circ g \simeq 1_X$ and $g \circ f \simeq 1_A$:
\par Direct computation shows that
\begin{align*}
(g \circ f)(a) & = g(f(a)) = g(a) = G(a,1) = a,
\end{align*}
while $f \circ g \simeq_G 1_X$. This completes the proof.
\end{proof}
\item We show that any non-empty convex subset, $X$, of a euclidean space is homotopy equivalent to a point:
\begin{proof}
Without loss of generality, we assume that $X \subset \E^n$. We first show the existence of a deformation retract, $G: X \times I \hookrightarrow \{x\}$, where $x \in X$ is fixed:
\par Let $G$ be defined as $G(y,t) = (1-t)y + tx$, for all $y \in X$. Then, clearly $G(y,0) = y$, and $G(y,1) = x \in \{x\}$ for all $y \in X$. So, $G$ is deformation retract. By the above, $X$ and $A$ are of the same homotopy type: null-homotopic.
\end{proof}
\item We show that $(X, \tau)$ is contractible\footnote{A space $(X, \tau)$ is contractible (to a point $x_0 \in X$) provided that there exists a map $F: X \times I \rightarrow X$ such that $F(x, 0) = x$ and $F(x,1) = x_0$.} if, and only if, every map $f:X \rightarrow Y$, $(Y, \tau_Y)$ a topological space, is null-homotopic:
\begin{proof}
$ $\newline
\begin{itemize}
\item[] ($\implies$): Suppose that $(X, \tau)$ is contractible to $x_0 \in X$. Let $F$ be the contraction, and $f: X \rightarrow Y$ any map. The claim is that $f \circ F: X\times I \rightarrow Y$ is a homotopy between $f$, and $f(x_0)$.
\par Indeed; We note that the composition of two continuous functions is continuous and
$$(f \circ F)(x,0) = f(F(x,0)) = f(x)$$ As well as,
$$(f \circ F)(x,1) = f(F(x,1)) = f(x_0)$$
\item[] ($\impliedby$): Suppose that every map $f:X \rightarrow Y$ is null-homotopic. Then, in particular, $i:X \rightarrow X$ is null-homotopic. Thus, there exists a map $F:X \times I \rightarrow X$, such that $F(x,0) = x =i(x)$, and $F(x,1) = x_0 = i(x_0)$.
\end{itemize}
\end{proof}
\item Let $(X, \tau)$ be a topological space. We show that the cone on $X$, $CX$ is contractible. We consider the "cone tip" as $X \times \{0\}$: \begin{proof} Consider the function $\overline{H}: (X \times I) \times I \rightarrow X \times I$, given by $\overline{H}((x,v),s) = (x,v(1-s))$. By the product topology, it follows that $\overline{H}$ is continuous. \footnote{For every $t \in I$, we can associate $\pi(\overline{H}((x,v),t))$ with $(c,v(1-t),t)$, showing that the product topology of $CX \times I$ agrees with that of $(X \times I) \times I$.} Let $\pi:X \times I \rightarrow CX = \nicefrac{X \times I}{X \times \{0\}}$ be the canonical map. The claim is that $\pi \circ \overline{H}: (X \times I) \times I \rightarrow CX$ is the desired homotopy.
\par Direct computation shows that, for all $x \in X$, $v \in V$,
$$\pi(\overline{H}( (x,v),0)) = \pi((x,v)) = (x,v),$$
and
$$\pi(\overline{H}( (x,v),1)) = \pi((x,0)) = (x,0)$$
Thus, $H \equiv \pi \circ \overline{H}$ is the desired homotopy.
\end{proof}
\end{enumerate}
\clearpage
\chapter{Introduction} \label{chapter:intro}
\section{Euler's Theorem}
There are no exercises listed for this section.
\section{Topological Equivalence}
There are no exercises listed for this section.
\section{Surfaces}
There are no exercises listed for this section.
\section{Abstract Spaces}
There are no exercises listed for this section.
\section{A Classification Theorem}
There are no exercises listed for this section.
\section{Topological Invariants}
\begin{enumerate}
\item We prove that $v(T) - e(T) = 1$ for any tree $T$:
\begin{proof} We proceed by induction on the number of edges in $T$:
\begin{itemize}
\item Basis: If $e(T) =1$, then $v(T) = 2$, and so, $v(T) - e(T) = 1$.
\item Hypothesis: Suppose that $v(T) - e(T) = 1$ for any tree $T$, such that $e(T) = n$, for some $n \in \N$.
\item Step: Suppose that $T$ is a tree, such that $e(T) = n+1$. Now, removing any edge will disconnect the tree, since by definition, $T$ contains no loops. Say that we remove and edge $e_0 \in T$, breaking up $T$ into $T_1$, $T_2$. Now, clearly, $v(T) = v(T_1) + v(T_1)$, and $e(T) = n+1 = e(T_1) +e(T_2) +1$. Thus, by the inductive hypothesis,
\begin{align*}
v(T) - e(T) & = v(T_1) + v(T_2) - e(T_1) - e(T_2) - 1 \\
& = 1 + 1 - 1 \\
& = 1
\end{align*}
\end{itemize}
\end{proof}
\item We show that inside any graph we can always find a tree which contains all the vertexes:
\begin{proof} Let $G$ be a finite graph. If $G$ is already a tree, we are done. So, suppose that $G$ is not a tree. Then, by the comments on pg. $3$, $G$ contains a (without loss of generality, minimal) loop, $L$. Now, we remove some edge $e_L \in L$. Now, as $L$ is a loop, $e_L$ does not disconnect $L$, and so $L - e_L$ is a tree. We continue in this way, as long as the new graph formed by removing an edge is not a tree.
\par To conclude, we note that this process must stop at some loop, because $e(G) < \infty$; further, we have not removed any vertexes.
\end{proof}
\item We prove that $v(\Gamma) - e(\Gamma) \leq 1$ for any graph $\Gamma$, with equality, precisely when $\Gamma$ is a tree:
\begin{proof} We have previously shown the equality condition. Using the above proof, we select a subtree, with the same vertexes as $\Gamma$. Call this tree $\Gamma'$. Then, $v(\Gamma') - e(\Gamma') = 1$. As $\Gamma'$ was created by removing edges, and not vertexes, we have $v(\Gamma) = v(\Gamma')$, and $e(\Gamma) \leq e(\Gamma')$. So,
$$e(\Gamma) - e(\Gamma) \leq v(\Gamma') - e(\Gamma') = 1$$
\end{proof}
\item We find a tree in the polyhedron of Fig. 1.3 which contains all the vertexes and construct the dual graph $\Gamma$ and show that $\Gamma$ contains loops:
\begin{proof} Please refer to figures \hyperref[fig:sub1.1]{\ref{fig:sub1.1}} and \hyperref[fig:sub1.2]{\ref{fig:sub1.2}}
\begin{figure}
\centering
\begin{subfigure}{.5\textwidth}
\centering
\includegraphics[width=.4\linewidth]{armstrong1.jpg}
\caption{The tree that contains every vertex.}
\label{fig:sub1.1}
\end{subfigure}%
\begin{subfigure}{.5\textwidth}
\centering
\includegraphics[width=.4\linewidth]{armstrong2.jpg}
\caption{Blue: Dual graph. Green: loops.}
\label{fig:sub1.2}
\end{subfigure}
\end{figure}
\end{proof}
\item Having done Problem 4, thicken both $T$ and $\Gamma$ in the polyhedron. $T$ is a tree, so thickening it gives a disc. We investigate what happens when you thicken $\Gamma$?
\begin{proof} $\Gamma$ is basically two loops connected at a point, with some other edges connected that do not make any more loops. So thickening should produce something homeomorphic to what is shown in Problem 11 (b).
\end{proof}
\item Let $P$ be a regular polyhedron in which each face has $p$ edges and for which $q$ faces meet at each vertex. We use Euler's formula to prove that,
$$ \frac{1}{p} + \frac{1}{q} = \frac{1}{2} + \frac{1}{e} $$
\begin{proof} We count the number of faces of $P$ as follows: Since each face of $P$ has $p$ edges, $pf = 2e$. We count the number of vertexes as follows: Since each vertex has $q$ faces which meet on it, $qv = 2e$. In all, $f = \nicefrac{2e}{p}$, and $v = \nicefrac{2e}{q}$. Assuming that Euler's formula holds for $P$, we have
$$v - e + f = \frac{2e}{q} - e + \frac{2e}{p} = 2$$
And the result follows.
\end{proof}
\item We deduce that there are only $5$ regular (convex) polyhedra.
\begin{proof} If $P$ is a polyhedra, it must satisfy the above criterion.Further, we assume that $p \geq 3$, since if not, we cannot construct a polygon. We use Shl{\"u}t notation. By checking routinely, for values up to $p,q \leq 5$ we have the following set of polyhedra:
$$\{ \{3,3\}, \{3,5\}, \{5,3\}, \{3,4\} , \{4,3\}\}$$
\par Now, if $p \geq 6$, then we must have
$$ \frac{1}{6} + \frac{1}{q} = \frac{1}{2} + \frac{1}{e}$$
Further, as $0 \leq \nicefrac{1}{q} \leq \nicefrac{1}{5}$, we have
\begin{align*}
\frac{1}{6} + \frac{1}{q} & = \frac{q+6}{6q} \\
& < \frac{12}{6q} = \frac{2}{q} \\
& < \frac{1}{2} + \frac{1}{e}
\end{align*}
A contradiction.
\par Please see figures \hyperref[fig:sub1.3]{\ref{fig:sub1.3}}, \hyperref[fig:sub1.4]{\ref{fig:sub1.4}}, \hyperref[fig:sub1.5]{\ref{fig:sub1.5}}, \hyperref[fig:sub1.6]{\ref{fig:sub1.6}}, and \hyperref[fig:sub1.7]{\ref{fig:sub1.7}}.
\begin{figure}
\centering
\begin{subfigure}{.5\textwidth}
\centering
\includegraphics[width=.4\linewidth]{armstrong8.jpg}
\caption{Tetrahedron}
\label{fig:sub1.3}
\end{subfigure}%
\begin{subfigure}{.5\textwidth}
\centering
\includegraphics[width=.4\linewidth]{armstrong7.jpg}
\caption{Cube}
\label{fig:sub1.4}
\end{subfigure}
\begin{subfigure}{.5\textwidth}
\centering
\includegraphics[width=.4\linewidth]{armstrong9.jpg}
\caption{Octahedron}
\label{fig:sub1.5}
\end{subfigure}
\begin{subfigure}{.5\textwidth}
\centering
\includegraphics[width=.4\linewidth]{armstrong10.jpg}
\caption{Dodecahedron}
\label{fig:sub1.6}
\end{subfigure}
\begin{subfigure}{.5\textwidth}
\centering
\includegraphics[width=.4\linewidth]{armstrong6.jpg}
\caption{Icosahedron}
\label{fig:sub1.7}
\end{subfigure}
\label{fig:sub1.8}
\caption{All five platonic solids.}
\end{figure}
\end{proof}
\item We check that $v-e+f=0$ for the polyhedron shown in Fig. 1.3 and find a polyhedron which can be deformed into a pretzel and calculate its Euler number:
\begin{proof} For the polyhedron in Fig. 1.3, $v=20$, $e=40$, $f=20$. Therefore $v-e+f=0$. Figure \hyperref[fig:sub1.9]{\ref{fig:sub1.9}} is basically a donut with two holes, i.e. a "pretzel".
\begin{figure}
\centering
\includegraphics[width=.4\linewidth]{armstrong3.jpg}
\caption{"pretzel"}
\label{fig:sub1.9}
\end{figure}%
\par Figure \hyperref[fig:sub1.9]{\ref{fig:sub1.9}} has $38$ faces ($10$ on the top, $10$ on the bottom, $10$ inside the hole, and $8$ around the outside sides). It has $76$ edges ($29$ on the top, $29$ on the bottom, $10$ vertical ones inside the holes and $8$ vertical ones around the outside sides), and it has $36$ vertexes ($18$ on the top, $18$ on the bottom); Therefore $v-e+f=-2$.
\end{proof}
\item This is left to the reader as an exercise.
\item We find a homeomorphism from the real line to the open interval $(0,1)$, and show that any two open intervals are homeomorphic:
\begin{proof} We show that the real line, $\R$ is homeomorphic to $(\nicefrac{-\pi}{2}, \nicefrac{\pi}{2})$.This amount to showing that $arctan: \R \rightarrow (\nicefrac{-\pi}{2}, \nicefrac{\pi}{2})$ is continuous, 1-1, onto and that $tan \equiv arctan^{-1}$ is continuous. Thus, $\R \simeq (\nicefrac{-\pi}{2}, \nicefrac{\pi}{2})$.
\par To conclude the proof, we show that any open subset, $(a,b) \subset \R$ is homeomorphic to $(\nicefrac{-\pi}{2}, \nicefrac{\pi}{2})$, $a \neq b$:
\par Indeed, under the mapping $f(x) = (b-a)(\frac{x}{\pi} +\frac{1}{2}) +a$, we have
$$(\nicefrac{-\pi}{2}, \nicefrac{\pi}{2}) \rightarrow (a,b)$$
As $f$ is a degree $1$ polynomial in $\R$, it follows that it is 1-1, onto and continuous with a continuous inverse.
\end{proof}
\item This is left to the reader as an exercise.
\item We find a homeomorphism from $S^2 - N = S^2 - \{(0,0,1)\}$ to $\E^2$:
\begin{proof} As pointed out in Armstrong, we find a formula for stereographic projection. From the fact that $S^2 = \{x \in \R^3: ||x||_2 =1\}$, and rules of calculus, any line passing through $v = (x,y,z) \in S^2 - N$ and $N$ is of the following form:
\begin{equation}
p(t) = (0,0,1) + t(x,y,z) \equiv <tx,ty,1-t(1-x)>
\end{equation}
For $t \in [0, \infty)$. As the $z$ component of $p$ is zero exactly when $t_0 = \nicefrac{1}{1-z}$, we substitute $t_0$ back in $(1)$ to obtain our formula for stereographic projection:
\[\Pi (x,y,z) = \big(\frac{x}{1-z},\frac{y}{1-z} \big) \]
It is left to show that $\Pi$ is 1-1, onto, continuous, with a continuous inverse.
\begin{itemize}
\item[] 1-1: By component wise comparison, the result follows.
\item[] Onto: Given $(x_0,y_0) \in \R^2$, we have, after some simple algebra,
\[ \Pi \Big(\frac{x_0}{r}, \frac{y_0}{r}, \frac{x_0^2 + y_0^2}{r}\Big) = (x_0,y_0),\]
where $r = 1+x_0^2 +y_0^2$.
\item[] Continuity: As $\Pi$ is a linear transformation in $\R^2$, we know that from rules of calculus/real analysis that it is continuous.
\item[] Cont. Inverse: $\Pi^{-1}$ is as outline above in the derivation of onto. For the same reasons the function is continuous.
\end{itemize}
\end{proof}
\item Let $x,y \in S^2$. We find a homeomorphism from $S^2$ to $S^2$ which takes $x$ to $y$. We do the same for the plane, and torus:
\begin{proof} We assume that results about matrix groups are applicable.
\begin{itemize}
\item $S^2$: Consider $S^2$ and $SO(3)$. Let $x \in S^2$. Then, by the Gram-Schmidt Orthogonalization process, there exists $X \in SO(3)$ for which $x$ is the first column of $X$. Likewise for $y$, its corresponding matrix $Y$. Then, $YX^{-1} \in SO(3)$ and maps $x$ to $y$. As $SO(3)$ is a group, $(YX^{-1})^{-1} \in SO(3)$. Thus, $YX^{-1}$ is a homeomorphism which takes $x$ to $y$.
\item $\E^2$: Fix $(x,y),(z,w) \in \R^2$. Then, the function, $f$ that sends $(a,b) \in \R^2$ to \[((x+z)-a, (y+w)-b)\]
sends $(x,y)$ to $(z,w)$. As $f$ is linear, it is continuous. Likewise, its inverse is continuous.
\item Torus: Consider the torus as $S^1 \times S^1$. Then component wise examination shows that the example above proves the result.
\end{itemize}
\end{proof}
\item This is left to the reader as an exercise.
\item This is left to the reader as an exercise.
\item This is left to the reader as an exercise.
\item Define $f:[0,1) \rightarrow C$, by $f(x) = e^{2\pi i x}$. We prove that $f$ is a continuous bijection. In addition, we find a point $x \in [0,1)$ and a neighbourhood $N$, of $x$ in $[0,1)$, such that $f[N]$ is not a neighbourhood of $f(x)$ in $C$; consequently, this means $f$ is not a homeomorphism.
\begin{proof} We first show that $f$ is a continuous bijection:
\par Via Euler's formula,
\[ e^{2\pi i x} = \cos(2\pi x) + i \sin(2\pi x), \quad \forall x \in \R \]
Using this fact, we show that $f$ is 1-1 and onto.
\begin{itemize}
\item 1-1: Suppose that $\cos(2\pi x) + i \sin(2\pi x) = \cos(2\pi y) + i \sin(2\pi y)$ for some $x, y \in [0,1)$. Then, by component-wise comparison of complex numbers we have $\cos(2\pi y) = \cos(2\pi x)$, iff, $2 \pi y = 2\pi x$, iff $x = y$, as $x,y \in [0,1)$; likewise for $\sin$. Thus, $f$ is 1-1.
\item Onto: Let $y \in C$. Then, $y = \cos(2\pi y_0) + i \sin(2\pi y_1)$ for some $(y_0, y_1) \in \R^2$. But as $f: [0,1) \rightarrow C$, we have $0 \leq 2 \pi y_0, 2\pi y_1 \leq 2\pi$. Consequently, $f$ is onto.
\item Continuity: By the definition of continuity in $\mathbb{C}$, and the fact that $\cos$, $\sin$ are continuous, we have that $f$ is continuous.
\end{itemize}
Next, consider $[0, 1/2) \subset [0,1)$. This is a open neighbourhood in $[0,1)$, by definition. But, $f[[0,1/2)]$ is $C$ intersected with the upper half of the plane minus $z = -1$. Which is not a neighbourhood of $z =1 \in C$, because any open ball centered around $z =1$ must contain the lower plane. Therefore, $f$ is not a homeomorphism.
\end{proof}
\item This is left to the reader as an exercise.
\item This is left to the reader as an exercise.
\item We prove that the radial projection of the tetrahedron to the sphere is a homeomorphism:
\begin{proof} A solid proof of this relies on the fact that a circumscribed polygon with center of mass $0$ is homeomorphic to $S^n$ via radial projection. The result of which are shown in chapter $5$.
\end{proof}
\item Let $C$ denote the unit circle in the complex plane and $D$ the disc which it bounds. Given two points $x,y\in D-C$, we find a homeomorphism from $D$ to $D$ which interchanges $x$ and $y$ and leaves all points of $C$ fixed:
\begin{proof} This is more or less intuitively obvious. But writing down an explicit function is not so easy. First note that for any $a \in \mathbb C$, the function $f(z)=\frac{z-a}{1-\overline{A}z}$ takes $S^1$ to itself. To see this suppose $|z|=1$. then
\begin{align*}
\left|\frac{z-a}{1-\overline{a}z}\right|&=\left|\frac{z-a}{\overline{z}z-\overline{a}z}\right|\\
&=\left|\frac{z-a}{(\overline{z}-\overline{a})z}\right|\\
&=\left|\frac{z-a}{\overline{z-a}}\right|\frac{1}{|z|}\\
&=1
\end{align*}
If $|a|<1$ then the denominator never vanishes so this is a continuous function on $D$. Also,
$f(0)=-a$, so if $|a|<1$ then since $f$ takes $C$ to itself and $0$ maps to $-a\in(D-C)$, $f$ must take all of $D$ to itself. The inverse $f^{-1}$ is therefore also a continuous function from $D$ to $D$ that takes $C$ to $C$.
Now suppose $x,y\in\mathbb C$ with $|x|<1$ and $|y|<1$, let
$$f_1(z)=\frac{z-x}{1-\bar xz}$$
and
$$f_2(z)=\frac{z-ty_1}{1-t\bar{y_1}z}$$
where $y_1=f_1(y)$ and $t=\frac{1-\sqrt{1-|y_1|^2}}{|y_1|^2}$.
As shown above, both $f_1$ and $f_2$ take $C$ to itself.
Finally let
$$g(z)=ze^{i(1-|z|)\pi/(1-|x_2|)}$$
where $x_2=f_2(0)$.
Then $g(x_2)=-x_2$ and $g(-x_2)=x_2$ and $g(z)=z$ $\forall$ $z\in C$. Since $z\mapsto|z|$ is continuous, $g$ is built up from sums, products and compositions of continuous functions and therefore $g$ is continuous.
\par The function $f_1^{-1}f_2^{-1}gf_2f_1$ is therefore a continuous function from $D$ to $D$ that switches $x$ and $y$ and fixes $C$.
\end{proof}
\item With $C,D$ as above (in Problem 21), define $h:D-C\rightarrow D-C$ by
\begin{align*}
h(0) & = 0 \\
h(re^{i\theta}) & =r\exp\left[i\left(\theta+\frac{2\pi r}{1-r}\right)\right]
\end{align*}
We show that $h$ is a homeomorphism, but that $h$ cannot be extended to a homeomorphism from $D$ to $D$ and draw a picture which shows the effect of $h$ on a diameter of $D$:
\begin{proof} The function $h$ restricted to a circle of radius $r$ acts by rotation of $2\pi r/(1-r)$ radians. Now $2\pi r/(1-r)\rightarrow\infty$ as $r\rightarrow 1$, so as the circle radius grows towards one, it gets rotated to greater and greater angles approaching infinity. Thus intuitively it's pretty obvious this could not be extended to the boundary. See figure \hyperref[fig:1.9]{\ref{fig:1.9}}.
\begin{figure}
\centering
\includegraphics[width=.4\linewidth]{armstrong4.jpg}
\label{fig:1.9}
\end{figure}
\par Now, we can think of $(r,\theta)$ as polar coordinates in $\mathbb R^2$. And the topology on $\mathbb C$ is the same as that on $\mathbb R^2$. Thus as a function of two variables $r$ and $\theta$ this is just a combination of continuous functions by sums, products, quotients and composition. Since the only denominator involved does not vanish for $|r|<1$, this is a continuous function of $r$ and $\theta$ on $D$ which is clearly onto. Since it is a simple rotation on each circle of radius $r$, it is also clearly one-to-one. The inverse is evidently
\begin{align*}
h^{-1}(0) & =0 \\
h^{-1}(re^{i\theta}) & = r\exp\left[i\left(\theta-\frac{2\pi r}{1-r}\right)\right] \\
\end{align*}
\par Now, let $r_n=\frac{n}{n+2}$ for $n$ odd and $r_n=\frac{n-1}{n}$ for $n$ even. Then for $n$ odd, $\frac{r}{1-r}=\frac n2$, and for $n$ even $\frac{r}{1-r}=n-1$. Therefore $\exp\left[i\left(\frac{2\pi r_n}{1-r_n}\right)\right]$ equals 1 if $r$ is even and $-1$ if $r$ is odd. Now, $r_n\rightarrow 1$. So if $h$ could be extended to all of $D$ we must have $h(1)=\lim h(r_n)$. But $h(r_n)$ does not converge, it alternates between 1 and $-1$. Therefore, there is no way to extend $h$ to $C$ to be continuous.
\end{proof}
\item Using the intuitive notion of connectedness, we argue that a circle and a circle with a spike attached cannot be homeomorphic (Fig. 1.26):
\begin{proof} In the circle, if we remove any one point what remains is still connected. However in the circle with a spike attached there is one point we can remove that renders the space not-connected. Since this property of being able to remove a point and retain connectedness must be a topological property preserved by homeomorphism, the two spaces cannot be homeomorphic.
\end{proof}
\item Let $X,Y$ be the subspace of the plane shown in Fig. 1.27. Under the assumption that any homeomorphism from the annulus to itself must send the points of the two boundary circles among themselves, we argue that $X$ and $Y$ cannot be homeomorphic:
\begin{proof} The two points that connect the two spikes to the two boundary circles in $X$ must go to the two points that connect the two spikes to the boundary circles in $Y$, because those are the only two points on the boundary circles that can be removed to result in a disconnected space, and because by assumption the circles go to the circles. Since the two points lie on the same circle in $Y$ but on different circles in $X$, some part of the outer circle in $Y$ must go to the outer circle in $X$ and the rest must go to the inner circle in $X$. But then some part of the outer circle in $Y$ must go to the interior of $X$. I'm not sure exactly how Armstrong expects us to prove this but it basically follows from the intermediate value theorem, applied to the two coordinates thinking of these shapes as embedded in $\mathbb R^2$.
\end{proof}
\item With $X$ and $Y$ as above, consider the following two subspaces of $\mathbb E^3$:
$$X\times[0,1] = \{(x,y,z)\mid(x,y)\in X, 0\leq z\leq 1\},$$
$$Y\times[0,1] = \{(x,y,z)\mid(x,y)\in Y, 0\leq z\leq 1\}.$$
Convince yourself that if these spaces are made of rubber then they can be deformed into one another, and hence that they are homeomorphic:
\begin{proof}With the extra dimension, the squareness can be continuously deformed so that it is a solid torus, with two flat rectangular shapes sticking off. One has both rectangles pointing out and one has one pointing out and the other pointing in. Since the torus is round, the first space made from $X$ can be rotated at the location where the inner rectangle is a full half turn to point the rectangle out, and as parallel slices (discs) of the torus move away from where the rectangle is attached, the rotation gradually gets less and less until it becomes zero before reaching the other rectangle. In this way the inner rectangle can be rotated to point out without affecting the other rectangle and with a gradual change in rotation angle between them guaranteeing the operation is continuous.
\end{proof}
\item This is left to the reader as an exercise.
\item This is left to the reader as an exercise.
\end{enumerate}
\clearpage
\chapter{Continuity}
\section{Open and Closed Sets}
\begin{enumerate}[(1)]
\item We verify each of the following for arbitrary subsets $A,B$ of a space $X$:
\begin{itemize}
\item $\overline{A\cup B}=\overline{A}\cup \overline{B}$:
\begin{proof} $\overline{A}$ and $\overline{B}$ are closed by theorem ($2.3$). Thus $\overline{A} \cup \overline{B}$ is closed. Now $A\subseteq \overline{A}$ and $B\subseteq \overline{B}$. Therefore $\overline{A} \cup \overline{B}$ is a closed set containing $A\cup B$. By Theorem 2.3 $\overline{A\cup B}$ is the smallest closed set containing $A\cup B$, thus it must be that $\overline{A\cup B}\subseteq \overline{A}\cup \overline{B}$. Conversely, $\overline{A\cup B}$ is a closed set that contains $A$, so $\overline{A\cup B}\supseteq\overline{A}$. Similarly $\overline{A\cup B}\supseteq\overline{B}$. Thus $\overline{A\cup B}\supseteq\overline{A}\cup\overline{B}$. Thus $\overline{A\cup B}=\overline{A}\cup\overline{B}$.
\end{proof}
\item $\overline{A\cap B} \subseteq \overline{A}\cap \overline{B}$:
\begin{proof} $\overline{A}$ is a closed set that contains $A\cap B$, so $\overline{A\cap B}\subseteq\overline{A}$. Likewise $\overline{A\cap B}\subseteq\overline{B}$. Thus $\overline{A\cap B}\subseteq\overline{A}\cap\overline{B}$.
\par To see that equality does not hold, let $A=\mathbb Q$ and let $B=\mathbb R - \mathbb Q$. Then $A\cap B=\emptyset$, so $\overline{A\cap B}=\emptyset$. But $\overline{A}=\mathbb R$ and $\overline{B}=\mathbb R$, so $\overline{A}\cap \overline{B}=\mathbb R$.
\end{proof}
\item $\bar{\overline{A}}=\overline{A}$:
\begin{proof} $\bar{\overline{A}}$ is the smallest closed set containing $\overline{A}$ by corollary ($2.4$), and $\overline{A}$ is closed by theorem ($2.3$), that contains $\overline{A}$. Thus $\overline{A}=\bar{\overline{A}}$.
\end{proof}
\item $(A\cup B)^{\circ}\supseteq\overset{\circ}{A}\cup\overset{\circ}{B}$:
\begin{proof} Let $x\in\overset{\circ}{A}\cup\overset{\circ}{B}$. Assume, without loss of generality, that $x\in\overset{\circ}{A}$. Then there is an open set $U\subseteq A$ such that $x\in A$. But then $x\in U\subseteq A\cup B$. So $x\in\overset{\circ}{A}\cup\overset{\circ}{B}$.
\par Thus
\[\overset{\circ}{A}\cup\overset{\circ}{B}\subseteq(A\cup B)^{\circ}\]
\par To see that equality does not hold, let $A=\mathbb Q$ and let $B=\mathbb R - \mathbb Q$. Then $\overset{\circ}{A}=\emptyset$ and $\overset{\circ}{B}=\emptyset$. And $A\cup B=\mathbb R$, so $(A\cup B)^{\circ}=\mathbb R$. Therefore $(A\cup B)^\circ=\mathbb R$ but $\overset{\circ}{A}\cap \overset{\circ}{B}=\emptyset$.
\end{proof}
\item $(A\cap B)^{\circ}=\overset{\circ}{A}\cap\overset{\circ}{B}$:
\begin{proof} Let $U$ be an open set in $A\cap B$. Then $U\subseteq A$ and $U\subseteq B$. Thus $(A\cap B)^\circ \subseteq \overset{\circ}{A}\cap\overset{\circ}{B}$. Conversely suppose $x\in\overset{\circ}{A}\cap\overset{\circ}{B}$. Then $\exists$ open sets $U$ and $V$ s.t.\ $x\in U\subseteq A$ and $x\in V\subseteq B$. Then $U\cap V$ is open and $x\in U\cap V\subseteq A\cap B$. Thus $(A\cap B)^\circ \supseteq \overset{\circ}{A}\cap\overset{\circ}{B}$.
\par Thus
\[(A\cap B)^\circ = \overset{\circ}{A}\cap\overset{\circ}{B}\]
\end{proof}
\item $(\overset{\circ}{A})^\circ=\overset{\circ}{A}$:
\begin{proof} Clearly
\[(\overset{\circ}{A})^\circ\subseteq\overset{\circ}{A}\]
Let $x\in\overset{\circ}{A}$. Then there exists an open set $U$, such that $x\in U\subseteq A$. Now $\overset{\circ}{A}$ is a union of open sets so is open. Let $V=\overset{\circ}{A}\cap U$. Then $x\in V\subseteq \overset{\circ}{A}$. Therefore $x\in(\overset{\circ}{A})^\circ$ and so, \[\overset{\circ}{A}\subseteq(\overset{\circ}{A})^\circ\]
\par Thus,
\[\overset{\circ}{A}=(\overset{\circ}{A})^\circ\]
\end{proof}
\end{itemize}
\item This is left to the reader as an exercise.
\item We specify the interior, closure and frontier of the following subsets.
\setcounter{equation}{0}
\begin{align} \
& \{(x,y): 1 < x^2+y^2 \leq 2\} \\
& {\E}^2 - \{(0,t), (t,0): t\in \R \} \\
& {\E}^2 - \{(x, sin(1/x): x>0\}
\end{align}
\begin{itemize}
\item $\{(x,y): 1 < x^2+y^2 \leq 2\} = A$:
\begin{align*}
\Ft A & = \overline{A} \cap \overline{(\E - A)} \\
& = \overline{A} \cap \big\{ \{(x,y): 0 \leq x^2+y^2 \leq 1\} \cup \{(x,y): 2 \leq x^2+y^2\} \big\}\\
& = \{(x,y): x^2+y^2 = 1 \lor x^2+y^2 = 2 \}
\end{align*}
The rest are left to the reader.
\item ${\E}^2 - \{(0,t), (t,0): t\in \R \} = B$:
\par From the fact that $\overline{{\E}^2 - B} = \E^2$, we have
\[\Ft B = B\]
The rest are left to the reader.
\item ${\E}^2 - \{(x, sin(1/x): x>0\} = C$:
\par From the fact that $\overline{{\E}^2 - C} = \E^2$, we have
\[\Ft C = C\]
The rest are left to the reader as an exercise.
\end{itemize}
\item This is left to the reader as an exercise.
\item We show that if $A$ is a dense subset of a space $(X, \tau)$, and if ${\Ocal} \in \tau$, that ${\Ocal} \subset \overline{A \cap {\Ocal}}$:
\begin{proof} Suppose, to the contrary, that ${\Ocal} \not\subset \overline{A \cap {\Ocal}}$. Then, there exist some $x \in {\Ocal}$, such that $x \notin \overline{A \cap {\Ocal}}$.
\par As $\overline{A \cap {\Ocal}}$ is closed, $x \in (\overline{A \cap {\Ocal}})^c$ and so, there exists some ${\Ocal}_x \in \tau$ such that $x \in {\Ocal}_x$, and
\[\overline{A \cap {\Ocal}} \cap (\underset{x}{\Ocal} - \{x\}) = \emptyset \]
But, as $x \notin \overline{A \cap {\Ocal}}$, we have
\[ \overline{A \cap {\Ocal}} \cap {\Ocal}_x = \emptyset \]
and consequently, $A \cap {\Ocal} \cap {\Ocal}_x = \emptyset$. But then, setting $B = {\Ocal} \cap {\Ocal}_x$, we have $x \in B$, $B \in \tau$, but $A \cap B = \emptyset$, contrary to extra lemma; $\overline{A} \neq X$, a contradiction.
\end{proof}
\item We prove that if $Y$ is subspace of $X$, and $Z$ is a subspace of $Y$, that $Z$ is a subspace of $X$:
\begin{proof} The open sets in $Y$ are exactly the sets $\Ocal \cap Y$ where $O$ is open in $X$. The open sets in $A$ as a subspace of $Y$ are therefore sets of the form $A\cap(Y\cap \Ocal)$ where $O$ is open in $X$. But $A\subseteq Y$, so $A\cap(Y\cap \Ocal)=A\cap \Ocal$. Therefore the open sets in $A$ as a subspace of $X$ are exactly the same as the open sets of $A$ as a subspace of $Y$.
\end{proof}
\item Suppose that $Y$ is a subspace of $(X, \tau)$. We show that a subset $A$ of $Y$ is closed in $Y$ if it is the intersection of $Y$ with a closed set in $X$. Further, we show that we get the same result if we take the closer in $Y$ or $X$:
\begin{proof} If $A \subset Y$ is closed in $Y$, then $Y - A$ is open in $Y$. But then, by the definition of subspace topology, $Y - A = Y \cap {\Ocal}_y$, for some open ${\Ocal}_y \in X$. Consequently, we have
\begin{align*}
A & = Y - (Y \cap {\Ocal}_y) \\
& = Y \cap (Y \cap {\Ocal}_y)^c \\
& = Y \cap (X - {\Ocal}_y)
\end{align*}
And, as $X - {\Ocal}_y$ is closed in $X$, this proves the result. We note that a similar case holds in the case where $A$ is open.
\par Letting $\overline{A_y}$ and $\overline{A_x}$ denote the closure of $A \subset Y$ in $Y$ and $X$ respectively, we show that $\overline{A_y} = \overline{A_x}$:
\newline \par From the previous part of this proof, we have that $\overline{A_y} = Y \cap C$, where $C$ is closed in $X$. Now, since $C$ is a closed set in $X$ containing $A$, we have $\overline{A_x} \subset C$, and so $Y \cap \overline{A_x} \subset \overline{A_y} = Y \cap C$.
\par Again by the first problem, we have $Y \cap \overline{A_x}$ is closed in $Y$ and contains $A$. So,
\[ \overline{A_y} \subset Y \cap \overline{A_x} \subset \overline{A_x}\]
Thus, $\overline{A_y} = \overline{A_x}$.
\end{proof}
\item Let $Y$ be a subspace of $(X, \tau)$. Given $A \subset Y$, we show that $A_X^\circ \subset A_Y^\circ$, and give an example when the two may not be equal:
\begin{proof} Let $x \in A_x^\circ$. Then, there exists and open ${\Ocal}_x \subset X$, with $x \in {\Ocal}_x$ and ${\Ocal}_x \subset A$. Now, since ${\Ocal}_x \subset A \subset Y$, $x \in {\Ocal}_x \cap Y \subset A$ and, ${\Ocal}_x \cap Y$ is open in $Y$, by definition. Thus, $x \in A_y^\circ$.
\par An example when they might not be equal is in the following case: Let $X = \R$, $Y = \mathbb{Z}$, and $A = \{0\}$. Then, $A_x^\circ = \emptyset$. But, every point of $\mathbb{Z}$ is open in the subspace topology, and so $A_y^\circ = \{ 0\}$.
\end{proof}
\item Let $Y$ be a subspace of $X$. We show that if $A$ is open (closed) in $Y$, and if $Y$ is open (closed) in $X$, that $A$ is open (closed) in $X$:
\begin{proof} Suppose $A\subset Y\subset X$ and $A$ open in $Y$. Then $A=Y\cap U$ where $U$ is open in $X$. The intersection of two open sets is open, so if $Y$ is open in $X$ then $A$ is open in $X$. Similarly, if $A\subset Y\subset X$ and $A$ closed in $Y$, then $A=Y\cap C$ where $C$ is closed in $X$ (by exercise $7$). The intersection of two closed sets is closed, so if $Y$ is closed in $X$ then $A$ is closed in $X$
\end{proof}
\item We show that the frontier of a set always contains the frontier of its interior, and describe the relationship between $\Ft (A \cup B)$ and $\Ft A$, $\Ft B$:
\begin{proof} Let $(X, \tau)$ be a topological space, and let $A \subset X$. We want to show that $\Ft A^\circ \subset \Ft A$.
\par Let $x \in \Ft A^\circ$. Then,
\[ x \in \overline{A^\circ} \cap \overline{(X - A^\circ)} = \overline{A^\circ} \cup \overline{(X-A) \cup (A - A^\circ)}\]
Now, if $x \in \overline{A^\circ}$ and $x \in \overline{X -A}$, we are done. So suppose that $x \in \overline{A^\circ}$ and $x \in \overline{(A - A^\circ)}$. But then, $x \in \overline{A^\circ} \cup \overline{(A - A^\circ)} = \overline{A}$. Thus, the result follows.
\newline \par
The inclusion $\Ft (A \cup B) \subset \Ft A \cup \Ft B$ always holds, and is left to the reader. To show that the reverse inclusion does not always hold, consider $X = \R$, and $A = \mathbb{Q}$. Then,
\begin{align*}
\Ft (A \cup A^c) & = \Ft \R = \emptyset \\
& \neq \Ft A \cup \Ft A^c = \R
\end{align*}
\end{proof}
\item This main part of this exercise is left to the reader. However, we do show that this topology does not have a countable base:
\begin{proof} Let $B = \tau$ be the topology specified. Suppose, to the contrary, that $\{B_n\}_{n = 1}^\infty$ is a countable base for the topology $\tau$. Define the function $f: \R \rightarrow \mathbb{N}$ as follows: for each $x \in \R$, let $f(x) = n$, such that $B_n \subset [x,1+x)$. Now, we show that $f$ is 1-1 to arrive at a contradiction:
\par Indeed; Suppose to the contrary, without loss of generality, that $x < y$. Then, if $f(x) = f(y)$, $f(x) = [x, x+1) \subset B_{f(y)} = [y, y+1)$, which is impossible. Thus, $x = y$.
\end{proof}
\item We show that if a topological space $(X, \tau)$ has a countable base for its topology, then $X$ contains a countable dense subset. I.e. A second countable space is separable:
\begin{proof} Let $\{B_n\}_n$ be a countable base for $\tau$. By the Axiom of Choice, let $A$ be the collection of elements $\{a_i\}_i$ such that $a_i \in B_i$. The claim is that $\overline{A} = X$.
\par Indeed; let ${\Ocal} \in \tau$. Then, ${\Ocal} = \bigcup_j B_j$, where $B_j \in B$. Now, as $A = \bigcup_i x_i$, where $x_i \in B_i$, we have $A \cap {\Ocal} \neq \emptyset$. Thus, by extra lemma, $\overline{A} = X$.
\end{proof}
\end{enumerate}
\section{Continuous Functions}
\begin{enumerate}[(1)]
\item We show that if $f: \R \rightarrow \R$ is a map, then the set of points left fixed by $f$ is closed. Further, the kernel of $f$ is closed.
\begin{proof} Define $f_0(x) = f(x) - x$. The rest is left to the reader.
\end{proof}
\item We show that the function $h(x) = \frac{e^x}{1+e^x}$ is a homeomorphism from the real line to the open interval $(0,1)$:
\begin{proof} Let
\[f(x)=\frac{e^x}{1+e^x}, \quad y\in(0,1), \quad x=\ln\left(\frac{y}{1-y}\right)\]
\par We show that $f$ is onto and 1-1:
\begin{itemize}
\item As $f(x)=y$, $f$ is onto.
\item Suppose $f(x)=f(y)$. Then,
\[\frac{e^x}{1+e^x}=\frac{e^y}{1+e^y} \text{ implies } e^x+e^{x+y}=e^y+e^{x+y} \]
Therefore $e^x=e^y$ and $f$ is 1-1.
\end{itemize}
For $a,b\in\mathbb R$,
\[f^{-1}((a,b))=\left(\ln\frac{a}{1-a},\ln\frac{b}{1-b}\right), \]
which is open. Since the intervals $(a,b)$ are a basis, it follows from Theorem ($2.9 (b)$) that $f$ is continuous. And $f((a,b)) = (f(a),f(b))$, so $f$ takes open sets in the base to open sets. Therefore $f^{-1}$ is continuous, and so, $f$ is a homeomorphism.
\end{proof}
\item Let $f: \E \rightarrow \R$ be a map and define $\Gamma_f: \E \rightarrow \E^2$ by $\Gamma_f (x) = (x, f(x))$. Whe show that $\Gamma _f$ is continuous and that its image, with the induced topology, is homeomorphic to $\E$:
\begin{proof} We show that $\Gamma_f$ is continuous. To do this, we use the sequential criterion for continuity in $\E^n$. Let $\{x_n\}_n$ be a sequence in $\E$, such that $x_n \rightarrow x \in \E$. Then $\Gamma_f$ is clearly continuous by the fact that $f$ is, and component wise comparison. That is,
\[ \Gamma_f (x_n) = (x_n, f(x_n)) \rightarrow (x, f(x)) = \Gamma_f (x) \]
Next, we show that $\Image \Gamma_f$ is homeomorphic to $\E$:
\par We claim that the function $p_1: \Image \Gamma_f \rightarrow \E$, defined by $p_1((x, f(x)) = x$ is the desired homeomorphism.
\begin{itemize}
\item[] Continuity: We have previously shown that the projection map is continuous.
\item[] 1-1: This is true, by construction.
\item[] Onto: This is clear, as $f$ a map from $\E$ to $\E$.
\item[] Cont. Inverse: This is clear, as $p_1^{-1} \equiv \Gamma_f$, and we have shown $\Gamma_f$ is continuous.
\end{itemize}
This proves the result.\footnote{A similar argument applies to $\E^n$.}
\end{proof}
\item We determine what topology on $X$ implies that every real-valued function definied on $X$ is continuous:
\begin{proof} $X$ must have the discrete topology, where every subset is open. To show this, it suffices to show points in $X$ are open.
\par Indeed; Fix $x\in X$ and define $f: X\rightarrow \mathbb R$ by
\[f(x) =
\begin{cases}
f(x) = 0 & \\
f(y) = 1 & y \neq x
\end{cases}
\]
Then,
\[f^{-1}((-1/2,1/2))=\{x\}\] and thus $f$ is continuous, since $(-1/2,1/2)$ is open and $\{x\}$ is open.
\end{proof}
\item Consider $X = \R$ with the co-finite topology, $(\R, CO)$. We show that $f: \E \rightarrow X$ defined by $f(x) = x$ is a map, but not a homeomorphism:
\begin{proof} To see that $f$ is continuous, let ${\Ocal} \in CO$. Then, $X - {\Ocal} = \R - {\Ocal}$ is finite. But then, $f^{-1}[{\Ocal}] = {\Ocal}$, and ${\Ocal}^c$ is finite in $\E$. This implies that ${\Ocal} = \overline{{\Ocal}}$. Thus, ${\Ocal}^c = {\Ocal}$ is open in $\E$. The fact that $f$ is 1-1 is clear.
\par Consider the inverse function of $f$. Note that $(a,b)$, $a < b$ is open in $\E$. But, $f^{-1}[(a,b)] = (a,b) = f[(a,b)]$, which is not open in $X$. Thus, $f^{-1}$ is not continuous.
\end{proof}
\item Suppose $X=A_1\cup A_2 \cup\dots$, where $A_n\subseteq\overset{\circ}{A}_{n+1}$ for each $n$. We show that if $f:X\rightarrow Y$ is a function such that, for each $n$, $f|A_n:A_n\rightarrow Y$ is continuous with respect to the induced topology on $A_n$, then $f$ is itself continuous:
\begin{proof} Let $x\in X$. Then $x\in A_n$ for some $n$ and $A_n\subseteq \overset{\circ}{A}_{n+1}$. Thus $x\in\overset{\circ}{A}_{n+1}$, and $X=\cup \overset{\circ}{A}_n$.
\par Now, let $U$ be open in $Y$. Since $f$ is continuous on $A_n$, by theorem ($2.8$), $f$ is continuous on $\overset{\circ}{A}_n$. Consequently,
\[f^{-1}(U)\cap\overset{\circ}{A}_n=f|_{\overset{\circ}{A}_n}^{-1}(U)\]
is open in $\overset{\circ}{A}_n$. So, there exists open sets $V_n$ such that \[V_n\cap\overset{\circ}{A}_n=f^{-1}(U)\cap\overset{\circ}{A}_n\]
And so,
\[f^{-1}(U) =f^{-1}(U)\cap \bigcup_n(\overset{\circ}{A}_n)=\bigcup_n(f^{-1}(U)\cap \overset{\circ}{A}_n)=\bigcup_n(V_n\cap \overset{\circ}{A}_n),\]
which is a union of open sets and so is open.
\end{proof}
\item Let $(X, \tau)$ be a topological space, $A \subset X$, and $\chi_A$ its characteristic function. We describe the frontier of $A$ in terms of $\chi$:
\begin{proof} We note that if $\overline{(X -A)} \cap \overline{A} \neq \emptyset$, then there exists $a \in \overline{(X -A)}$, $a \in \overline{A}$. With this in mind, we claim that $\chi_A$ is continuous at $a \in X$ if, and only if, $a \notin \Ft A$.
\par Indeed;
\begin{itemize}
\item[] ($\implies$): Suppose, to the contrary, that $\chi_A$ is continuous at $a \in X$, but that $a \in \Ft A$. Then, $a \in \overline{X-A}$, and $a \in \overline{A}$. Now, let ${\Ocal} = (.99, 1.99) \subset \R$, without loss of generality. Then, $\chi_A(a) = 1 \in {\Ocal}$. Now, since $\chi_A$ is continuous, $\chi_A^{-1} [{\Ocal}]$ is open, and such that $a \in \chi_A^{-1} [{\Ocal}]$. But then, $\overline{X-A} \cap \chi_A^{-1} [{\Ocal}] \neq \emptyset$. Further, this implies that $X-A \cap \chi_A^{-1} [{\Ocal}] \neq \emptyset$. But then, $\chi_A(a) = 0 \in {\Ocal}$, a contradiction.
\item[] ($\impliedby$):
\end{itemize}
\end{proof}
\item We determine which of the following maps are open or closed:
\begin{itemize}
\item $x \mapsto e^x$, on $\R$:
\begin{proof} Open: Let $C=f(\R)$. Since $f(A\cup B)=f(A)\cup f(B)$, it suffices to check that $f$ is open on a base of open sets. Let $B$ be an open interval in $\R$. If $B$ has length greater than $2\pi$ then $f(B)$ is all of $C$. So $f(B)$ is open in $C$. Otherwise $f(B)$ is an open arc of the circle. Also open. Thus in all cases $f$ maps $B$ to an open set. Since open balls are a base $f$ must be an open map.
\par Not Closed: To show $f$ is not closed, for each $n\in \mathbb N$ let $E_n=[2n\pi+1/n,(2n+1)\pi-1/n]$. Let $E=\cup_nE_n$. Suppose $x$ is a limit point of $E$. Then $(x-1/2,x+1/2)$ intersects at most one $E_n$. Thus $x$ is a limit point of $E_n$. Thus $x\in E_n$ since $E_n$ is closed. Thus $E$ is closed. But $f(E)$ is equal to $C$ intersected with the upper half plane $\text{im}(z)>0$. This is an open set in $C$, and not closed since $z=1$ is a limit point of $f(E)$ not in $f(E)$. Thus $f(E)$ is not closed.
\end{proof}
\item $f: \E^2 \rightarrow \E^2$, $f(x,y) = (x, |y|)$:
\begin{proof} Not Open: To see $f$ is not open, let $D$ be the open unit disc. Let $H^+=\{(x,y)\mid y\geq 0\}$ and $H^-=\{(x,y)\mid y\leq 0\}$. Then $f(D)=D\cap H^+$. Therefore $z=0\in f(D)$ and every open ball containing $z$ intersects ${H^+}^c\subset f(D)^c$. Thus $f(D)$ is not open.
\par Closed: Now, suppose $E$ is closed in $\mathbb E^2$. Let $E'=E\cap H^-$ and let $E''$ be $E'$ reflected about the $x$-axis. Then $f^{-1}(E)= (E\cap H^+)\cup(E'')$. Now $E$ and $H^+$ are closed so $E\cap H^+$ is closed. And $E''$ is closed because $H^-$ is closed, so $E'$ is closed, and reflection is a homeomorphism. Thus $f^{-1}(E)$ is closed.
\end{proof}
\item $z \mapsto z^3$ on $\C$:
\begin{proof} Open: We first show it is open. Let $A_{\theta_1,\theta_2,r_1,r_2}=\{re^{i\theta}\in \mathbb C\mid \theta_1,\theta_2,r_1,r_2\in[0,\infty)\text{ and } \theta_1<\theta<\theta_2, r_1<r<r_2\}$. Then the set $\beta=\{A_{\theta_1,\theta_2,r_1,r_2}\mid\theta_2-\theta_1<2\pi$ and $0\leq r_1<r_2\}$ form a base for the usual topology on $\mathbb C$. The sets are clearly open and the intersection of any two of them is another one. Also for any $z\in A_{\theta_1,\theta_2,r_1,r_2}$ there is an open disc $D$ s.t.\ $z\in D\in A_{\theta_1,\theta_2,r_1,r_2}$ and for any open disc $D$ with $z\in D$ we can find ${\theta_1,\theta_2,r_1,r_2}$ s.t.\ $z\in A_{\theta_1,\theta_2,r_1,r_2}\in D$. Now consider what happens to $A_{\theta_1,\theta_2,r_1,r_2}$ under the function $f$. If $\theta_2-\theta_1>2\pi/3$ then $f(A_{\theta_1,\theta_2,r_1,r_2})$ is a full open annulus. Otherwise $f(A_{\theta_1,\theta_2,r_1,r_2})=A_{3\theta_1,3\theta_2,r_1,r_2}$ which is open. Thus $f$ takes basic open sets in $\beta$ to open sets. Since $f(A\cup B)=f(A)\cup f(B)$ for any sets $A$ and $B$, it suffices to check open-ness on the base $\beta$. Thus $f$ is an open map.
\par Closed: We now show $f$ is a closed map. Let $E$ be any closed set in $\mathbb C$. Let $Q_1$ be the first quadrant $\{x+iy\mid x,y\geq 0\}$. Let $Q_2,Q_3,Q_4$ be the other (closed) quadrants. Then $Q_1$ is closed, so $E\cap Q_1$ is closed. The map $f$ restricted to $Q_1$ is a homeomorphism from $Q_1$ to its image $f|_{Q_1}:Q_1\rightarrow f(Q_1)=Q_1\cup Q_2\cup Q_3$ because its inverse $re^{i\theta}\mapsto re^{i\theta/3}$ is continuous. Thus $f(E\cap Q_1)$ is closed in $Q_1$. Likewise $f(E\cap Q_i)$ is closed for all $i=1,2,3,4$. Functions respect unions, thus since $E=\cup_i(E\cap Q_i)$ it follows that $f(E)=\cup_i f(E\cap Q_i)$. But each $f(E\cap Q_i)$ is closed. Thus $f(E)$ is closed.
\end{proof}
\end{itemize}
\item We show that the {\it unit ball}\/ in $\mathbb E^n$ (the set of points whose coordinates satisfy $x_1^2+\cdots+x_n^2\leq1$) and the {\it unit cube} (points whose coordinates satisfy $|x_i|\leq1$, $1\leq i\leq n$) are homeomorphic if they are both given the subspace topology from $\mathbb E^n$:
\begin{proof} Note that nowhere in the proof of lemma ($2.10$) is it used that we are in two dimensions. The proof goes through basically without change to any finite dimension where we replace "disc" with "ball".
Now, let
\[f:\mathbb E^n-{0}\rightarrow\mathbb E^n-{0}\]
be given by
\[f({\bf v})=\frac{1}{||{\bf v}||}{\bf v}\]
Then $f$ is continuous and the image of $f$ is the unit sphere. Let $g$ be $f$ restricted to {\it the surface} of the unit cube. Then $g$ is one-to-one continuous. The intersection of an open ball with the surface of the cube maps to the intersection of an open ball with the sphere. Thus $g$ is an open map. Therefore the inverse of $g$ must be continuous. Therefore $g$ is a homeomorphism. By the generalization of lemma ($2.10$), $g$ may be extended from the boundaries to a homeomorphism from the whole cube to the whole ball.
\end{proof}
\end{enumerate}
\section{Space-Filling Curves}
\begin{enumerate}[(1)]
\item We find a Peano curve which fills out the unit square in $\E^2$:
\begin{proof} We apply lemma $2.10$. By a previous exercise, we have that $\partial [0,1]^2$ is homeomorphic to $\partial S^1$. Further, the boundary of the unit triangle is homeomorphic to $\partial S^1$. Therefore, by lemma $2.10$, there is a homeomorphism, $f$, from the unit triangle to the unit disc. Let $h: [0,1] \rightarrow T$ be the space filling curve of the triangle as mentioned in Armstrong. Then, $f \circ h$ is a continuous mapping from $[0,1] \rightarrow [0,1]^2$. Further, it is space filling.
\end{proof}
\item We find an onto, continuous function from $[0,1]$ to $S^2$:
\begin{proof} From a previous exercise, we have that $\E^2 \cong_f (0,1) \times (0,1)$. Further, we have shown that $\E^2 \cong_\pi S^2 - (0,0,1)$. Thus, we extend $f$ to $g: [0,1]^2 \rightarrow S^2$ by
\[g(x) =
\begin{cases}
f(x) & x \notin \partial [0,1]^2 \\
(0,0,1) & x \in \partial [0,1]^2
\end{cases}
\]
Then, if ${\Ocal} \subset S^2$ is open, and contains $(0,0,1)$, then $g^{-1}[{\Ocal}]$ is the entire square via homeomorphism. Now, if ${\Ocal} \subset S^2$ does not contain $(0,0,1)$ and is open, then $g^{-1}[{\Ocal}] = f^{-1}[{\Ocal}]$ which is clearly open.
\par
Thus, $g \circ h$, where $h$ is the triangle space-filling curve, is the desired function.
\end{proof}
\item We determine whether or no a space-filling curve can fill out the plane:
\begin{proof} Note, $[0,1]$ is compact, while $\E^2$ is not.
\end{proof}
\item We determine whether or not a space filling curve ca fill out all of the unit cube in $\E^3$:
\begin{proof} Let $I=[0,1]$. Let $f:I\rightarrow I\times I$ be a space filling curve. Let $g:I\times I\rightarrow I\times I\times I\times I$ be the map $(x,y)\mapsto (f(x),f(y))$ with the natural identifications $(I\times I)\times(I\times I)$ with $I\times I\times I\times I$. Let $p$ be projection onto the first three coordinates of $I\times I\times I\times I$. Then $p\circ g\circ f$ is a continuous function from $I$ onto $I^3$.
\end{proof}
\item To proved a rigorous proof of this, at this point, is out of the question. However, via theorems ($3.3$), and ($3.7$), it is not true.
\end{enumerate}
\section{The Tietze Extension Theorem}
Throughout these exercises, we assume that $(X, d)$ is a metric space, and that if $A \subset X$, it has the subspace metric. Further, the topology on $X$ is the induced topology, if not stated otherwise.
\begin{enumerate}[(1)]
\item We show that $d(x,A) = 0$ if, and only if, $x \in \overline{A}$:
\begin{proof} $ $\newline
\begin{itemize}
\item[] Suppose that $d(x,A) = 0 = \inf_{a \in A} d(x,a)$. Then, for every $\epsilon > 0$, there exists $a \in A$, such that
\[|d(x,a) - 0| = d(x,a) < \epsilon \]
Now, let ${\Ocal}$ be an open subset of $X$, $x \in {\Ocal}$. Choose $a \in A$ such that, $d(x,a) < \epsilon$. Then, $a \in {\Ocal}$, and ${\Ocal} \cap A \neq \emptyset$. Thus, $x \in \overline{A}$.
\item[] Suppose that $x \in \overline{A}$. Then, we do the canonical ball construction and pick a sequence $\{a_n\}_n$ of elements of $A$, such that $d(x, a_n) \rightarrow 0$, $n \rightarrow \infty$.
\end{itemize}
\end{proof}
\item We show that if $A,B \subset X$ are disjoint and closed, there exists disjoint open sets $U,V$ such that $A \subset U$ and $B \subset V$:
\begin{proof} By lemma $2.13$, there exists a continuous function $f: X \rightarrow [-1,1]$, such that $f[A] \equiv 1$, and $f[B] \equiv -1$. Let $O_1 = [-1,0]$, and $O_2 = [0,1]$. Then, $O_1, O_2$ are open in $[-1,1]$. Thus, $f^{-1}[O_i]$ is open, as $f$ is continuous. Further, $f^{-1}[O_1] \cap f^{-1}[O_2] = \emptyset$, and $A \subset f^{-1}[O_1]$, $B \subset f^{-1}[O_2]$.
\end{proof}
\item We consider what topology the discrete metric gives a space:
\begin{proof} We first show $d$ is a metric. It is real-valued. And clearly
\[ d(x,y)\geq 0 \text{ iff } x=y\]
Also clearly $d(x,y)=d(y,x)$. Finally, the only way we could have
\[d(x,y)+d(y,z)<d(x,z)\]
is if the left-hand-side is $0$. But then $x=y=z$ and so the right-hand-side is also zero. Thus $d$ is a metric. \par Since
\[\{x\}=\{y\mid d(x,y)<1/2\},\]
the sets $\{x\}$ are open. Since every set is a union of its points, every set is open. Thus this metric gives the discrete topology.
\end{proof}
\item We show that every closed subset of a metric space is the intersection of a countable number of open sets:
\begin{proof} Let $A$ be a closed subset of $X$. Define
\[A_n = \{x \in X : d(x,A) < \frac{1}{n} \}\]
Then, $A_n$ is open, for each $n \in \mathbb{N}$. The claim is that $\bigcap_n A_n = A$:
\newline
\par Clearly, $A \subset A_n$, for each $n$ so, $A \subset \bigcap_n A_n$. Next, suppose that $x \in \bigcap_n A_n$, but $x \notin A$. Then,
\[1 > \inf_{a \in \bigcap_n A_n} d(x,a) = \epsilon > 0\]
By the Archimedian Principal, there exists $n_0$ large enough so that $\nicefrac{1}{n_0} < \epsilon $. But then,
\[x \notin \bigcap_n A_n\]
Thus, $\bigcap_n A_n \subset A$.
\par This concludes the proof.
\end{proof}
\item If $A,B$ are subsets of a metric space, their {\it distance apart} $d(A,B)$ is the infinum of the numbers $d(x,y)$ where $x\in A$ and $y\in B$. We find two disjoint closed subsets of the plane which are zero distance apart and check that both of the closed sets which you have just found have infinite diameter:
\begin{proof} Let $A$ be the $x$-axis and $B$ be the set
\[\{(x,1/x)\mid x>0\}\]
The functions $x\mapsto 0$ and $x\mapsto\frac{1}{x}$ are continuous on $(0,\infty)$ so $A$ and $B$ are closed by chapter $2$, problem $15$.
\par Now, let
\[\{a_n\}_{n=1}^\infty := (n,0), \quad \{b_n\}_{n=1}^\infty : = (n,1/n)\]
Then $a_n\in A$ and $b_n\in B$ and $d(a_n,b_n)=\frac{1}{n}$ and thusly,
\[ d(A,B)<\frac{1}{n}, \quad \forall n\]
Consequently, $d(A,B)=0$. Both sets clearly have infinite diameter.
\end{proof}
\item We show that if $A$ is a closed subset of $X$, then any map $f: A \rightarrow \E^n$ can be extended over $X$:
\begin{proof} The is the Tietze extension theorem applied component-wise.
\end{proof}
\item We find a map from $\mathbb E^1 - \{0\}$ to $\mathbb E^1$ which cannot be extended over $\mathbb E^1$:
\begin{proof} Let $f(x)=1/x$. Then $f$ is continuous on $\mathbb E^1-\{0\}$ by Theorem 2.9 (b) because $f^{-1}$ of an open interval is an open interval, or the union of two open intervals. Now suppose $g$ extends $f$ to all of $\mathbb E$. Let $a_n=\frac1n$. Then $a_n\rightarrow 0$. Thus $g(a_n)\rightarrow g(0)$. But $g(a_n)=f(a_n)=n\rightarrow\infty$. Thus no such $g$ can exist.
\end{proof}
\item This is left to the reader as an exercise.
\item Given a map $f:X\rightarrow \mathbb E^{n+1}-\{0\}$ we find a map $g:X\rightarrow S^n$ which agrees with $f$ on the set $f^{-1}(S^n)$:
\begin{proof} Let $h:\mathbb E^{n+1}-\{0\}\rightarrow S^n$ be given by
\[\mathbf v\mapsto\frac{1}{||\mathbf v||}\mathbf v\]
Then $h$ is continuous and $h$ is the identity on $S^n$. Let $g=h\circ f$. Then $g$ is continuous and agrees with $f$ on $f^{-1}(S^n)$.
\end{proof}
\item If $X$ is a metric space and $A$ closed in $X$, we show that a map $f:A\rightarrow S^n$ can always be extended over a {\it neighborhood of} $A$, in other words over a subset of $X$ which is a neighborhood of each point of $A$. (Think of $S^n$ as a subspace of $\mathbb E^{n+1}$ and extend $f$ to a map of $X$ into $\mathbb E^{n+1}$. now use Problem 35.):
\begin{proof} Following the hint we think of $S^n$ as a subspace of $\mathbb E^{n+1}$. Then $f=(f_1,\dots,f_n)$ and \[f_i \equiv p_i\circ f,\]
where $p_i$ is the $i$-th projection. The solution to problem $32$ shows $p_i$ is continuous, so that each $f_i$ is continuous.
\par By theorem ($2.15$), each component $f_i$ can be extended to a function $g_i$ on all of $X$ such that $g_i$ agrees with $f_i$ on $A$. Then $g=(g_1,\dots,g_n)$ extends $f$ on $A$ to a map from $X$ to $\mathbb E^{n+1}$.
\par The same argument as in problem $32$ shows $g$ is continuous. Note that
\[g^{-1}(0)\cap A=\emptyset\]
because $f$ maps $A$ into $S^n$. And $g^{-1}(0)$ is a closed set in $X$ (theorem ($2.9 (e)$)).
\par Thus by problem $28$ we can find {\it disjoint} open sets $U$ and $V$ in $X$ such that
\[A\subseteq U, \quad g^{-1}(0)\subseteq V\]
Let $h$ be the map from problem $35$. Then $h\circ g$ is well-defined as long as $g(x)\neq0$. Thus $h\circ g$ is well-defined on $U$. And
\[h\circ g|_U\] agrees with $f$ on $A$, since $h$ is the identity on $S^n$.
\end{proof}
\end{enumerate}
\newpage
\chapter{Compactness and Connectedness}
\section{Closed and Bounded Subsets of $\E^n$}
There are not exercises for this section.
\section{The Heine-Borel Theorem}
\begin{enumerate}[(1)]
\item This is left to the reader as an exercise.
\item Let $S\supseteq S_1\supseteq S_2\supseteq \cdots$ be a nested sequence of squares in the plane whose diameters tend to zero as we proceed along the sequence. We prove that the intersection of all these squares consists of exactly one point:
\begin{proof} Each $S_n$ is a square so is of the form $I_n\times J_n$ for closed one-dimensional intervals. And $S_n\supset S_{n+1}$ means $I_n\supset I_{n+1}$ and $J_n\supset J_{n+1}$. Thus we can apply the one-dimensional argument given above to the $I_n$'s and $J_n$'s. We get a sequence of points $x_n$ converging to $p$ and $y_n$ converging to $q$. So $(x_n,y_n)\in S_n$ therefore converges to $(p,q)$ which must be in $S_n$ for all $n$ since $S_n$ is closed, so $(p,q)\in\bigcap_{n=1}^\infty S_n$. Similarly the one-dimensional argument shows there can be only a unique $x$ and $y$ coordinate of anything in $\bigcap_{n=1}^\infty S_n$. Thus,
\[\bigcap_{n=1}^\infty S_n=\{(p,q)\}\]
\end{proof}
\item We use the Heine-Borel theorem to show that an infinite subset of a closed interval must have a limit point:
\begin{proof} By 'infinite', we assume that the author means that the cardinality of the set is non-finite and countable. We proceed by contradiction:
\par Let $C$ be the closed interval. Suppose that such an infinite subset, $A$, of $C$ does not have any limit points. Then, for each $x \in C$, there is an open set ${\Ocal}_x$ in $C$, such that $x \in {\Ocal}_x$ and ${\Ocal}_x \cap (A - \{x\}) = \emptyset$. In addition,
\[C \subset \bigcup_{x \in C} {\Ocal}_x \]
So, $\bigcup_{x \in C} {\Ocal}_x$ is an open cover of $C$, and consequently, has a finite subcover:
\[\mathcal{F} = \bigcup_{\substack{x_i \\ i \in \{1,2, \dots, N\}}} {\Ocal}_{x_i}\]
Then, $A \subset \mathcal{F}$ and ${\Ocal}_{x_i} \cap A = x_i$ for each $i$. Thus,
\[A = \mathcal{F} \cap A = \{x_1, x_2, \dots, x_N\}\]
A contradiction.
\end{proof}
\item We rephrase the definition of compactness in terms of closed sets:\footnote{Definition; The finite intersection propert (FIP): Let $\mathcal{F}$ be a collection of sets. Then, $\mathcal{F}$ has the finite intersection property if whenever $F_1, F_2, \dots, F_n \in \mathcal{F}$, $F_1 \cap F_2 \cap \dots \cap F_n \neq \emptyset$.}
\begin{proof} We claim that $X$ is compact if, and only if, for every collection $\{C_i\}_{i \in I}$ of closed sets in $X$ with the FIP, $\bigcap_i C_i \neq \emptyset$.
Indeed:
\begin{itemize}
\item[] ($\implies$) We proceed by contradiction: Let $X$ be compact, and let $\{C_i\}_{i \in I}$ be a collection of closed sets with the FIP, such that, $\bigcap_i C_i = \emptyset$.
\par Then, by De-Morgan's Laws, as $C_i^c$ is open,
\[\bigcup_i C_i^c = \Big( \bigcap_i C_i \Big)^c = \emptyset^c = X\]
Thus, there exists a finite subcover,
\[\{C_{i_1}^c, C_{i_2}^c, \dots, C_{i_n}^c\},\]
with $X = \bigcup_k C_{i_k}^c$. And, so,
\[X^c = \emptyset = \Big(\bigcup_k C_{i_k}^c\Big)^c = \bigcap_k C_{i_k} \neq \emptyset\]
However, this is clearly a contradiction. Thus, $\bigcap_i C_i \neq \emptyset$.
\item[] ($\impliedby$) Let $\mathcal{F} = \{ {\Ocal}_i\}_{i \in I}$ be an open cover of $X$. Then, $\{{\Ocal}_i^c\}_{i \in I}$ is a collection of closed sets, such that
\[\bigcap_i {{\Ocal}}_i^c = \Big( \bigcup_i {\Ocal}_i \Big)^c = X^c = \emptyset\]
Thus, there exists a finite set $\{{\Ocal}_1^c, {\Ocal}_2^c, \dots, {\Ocal}_n^c\}$, such that
$$\bigcap_{k = 1,2, \dots , n} {\Ocal}_k^c = \emptyset $$
\par But then,
\[\Big(\bigcap_{k = 1,2, \dots , n} {\Ocal}_k^c\Big)^c = \bigcup_{k = 1,2, \dots , n} {\Ocal}_k = (X^c)^c = X\]
So, $\mathcal{F}$ contains a finite subcover.
\end{itemize}
\end{proof}
\end{enumerate}
\section{Properties of Compact Spaces}
\begin{enumerate}[(1)]
\item We determine which of the following are compact:
\begin{itemize}
\item $\Q \subset \E$:
\begin{proof} By theorem ($3.9$), a compact subset of $\E$ is closed and bounded, $\Q$ is neither.
\end{proof}
\item $S^n$ with a finite number of points removed:
\begin{proof} $S^n-\{p_1,\dots p_n\}$ is not compact; by theorem ($3.9$), a compact subset of $\mathbb R^{n+1}$ is closed and bounded. Since $S^n$ lives in $\mathbb R^{n+1}$, the theorem applies. $S^n$ is bounded, but $S^n-\{p_1,\dots, p_n\}$ is not closed, because one can find a sequence in $S^n$ that converges to any of the removed points.
\end{proof}
\item the torus with an open disc removed:
\begin{proof} Yes, the torus with an open disc removed is compact. The torus can be embedded in $\mathbb R^3$ as a bounded subset. And since we are removing an open disc, what remains is a closed subset of the torus and therefore (by chapter $2$, problem $7$, page $31$) is a closed subset of $\mathbb R^3$. Therefore, by theorem ($3.9$), it is compact.
\end{proof}
\item the Klein Bottle:
\begin{proof} The Klein bottle is compact. It is the continuous image of a closed finite rectangle. By theorem ($3.9$) a closed finite rectangle is compact. So by theorem ($3.4$) the Klein bottle is compact. Alternatively, the Klein bottle can be embedded into $\mathbb R^4$ as a closed and bounded set. Therefore, it is compact.
\end{proof}