-
Notifications
You must be signed in to change notification settings - Fork 7
/
Copy pathanalysis_t2_exercises.tex
4244 lines (3741 loc) · 191 KB
/
analysis_t2_exercises.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]{article}
\title{Propositions of solutions for \textit{Analysis II} by Terence Tao}
\author{Frédéric Santos}
% General packages:
\usepackage{a4wide}
\usepackage[english]{babel}
\usepackage[utf8]{inputenc}
\usepackage[T1]{fontenc}
\usepackage{titlesec}
\titlelabel{\thetitle.\quad}
\usepackage{enumitem}
% Fonts and math packages:
\usepackage{lmodern}
\usepackage{amsmath}
\usepackage{amssymb}
\numberwithin{equation}{section}
\usepackage[matha,mathb,mathx]{mathabx}
\usepackage{mbboard}
\usepackage{stmaryrd}
\usepackage{hyperref}
% Macros:
\newcommand{\successor}[1]{#1 \! +\!\!\!+}
\newcommand{\aval}[1]{\left\lvert #1 \right\rvert}
\newcommand{\intset}[2]{\llbracket #1, #2 \rrbracket}
\newcommand{\inv}[1]{#1^{-1}}
\newcommand{\rr}{\mathbb{R}}
\newcommand{\partsof}[1]{\mathcal{P}\left( #1 \right)}
\newcommand{\minus}{\, \textrm{---}\!\textrm{--} \:}
\newcommand{\quot}{\, \textrm{/}\!\textrm{/} \:}
\newcommand{\nn}{\mathbb{N}}
\newcommand{\zz}{\mathbb{Z}}
\newcommand{\qq}{\mathbb{Q}}
\newcommand{\floor}[1]{\lfloor #1 \rfloor}
\newcommand{\formallimit}[1]{\text{LIM}_{n \to \infty} #1}
\newcommand{\seq}[2]{(#1)_{n=#2}^\infty}
\newcommand{\limit}[1]{\text{lim}_{n \to \infty} #1}
\newcommand{\extrr}{\overline{\rr}}
\newcommand{\adh}[1]{\overline{#1}}
\newcommand{\liminfp}[2]{\inf (#1^+_N)_{N=#2}^{\infty}}
\newcommand{\abs}[1]{\left\lvert #1 \right\rvert}
\renewcommand{\P}{\mathbf{P}}
\newcommand{\Q}{\mathbf{Q}}
\renewcommand{\epsilon}{\varepsilon}
\newcommand{\infint}[2]{\underline{\int}_{#2} \, #1}
\newcommand{\supint}[2]{\overline{\int}_{#2} \, #1}
\newcommand{\ddisc}{d_{\text{disc}}}
\newcommand{\inter}{\text{int}}
\newcommand{\ext}{\text{ext}}
\newcommand{\ninf}[1]{\|#1\|_\infty}
\newcommand{\vecn}[1]{(#1_i)_{1 \leq i \leq n}}
% Lemmas:
\usepackage{amsthm}
\newtheorem*{lem}{Lemma}
\newtheorem*{theorem}{Theorem}
% Environment:
\newenvironment{exo}[2]{\noindent \textsc{Exercise #1}. ---
\textit{#2} \vspace{3mm}}
%%%%%%%%%%%%%%
%%% Begin doc:
\begin{document}
\maketitle
\tableofcontents
\vskip 15mm
\noindent \textbf{Remarks.} The numbering of the Exercises follows the
fourth edition of \textit{Analysis II}. In order to make the
references to \textit{Analysis I} easier, we consider that we begin
with Chapter 12 here, as in earlier editions of the textbook. Thus, in
particular, a reference to ``Exercise 4.3.3'' (for instance) will
always mean ``Exercise 4.3.3 from \textit{Analysis I}''.
\pagebreak
\setcounter{section}{11}
\section{Metric spaces}
\label{sec:metric-spaces}
\begin{exo}{12.1.1}{Prove Lemma 12.1.1}
Consider the sequence $\seq{a_n}{m}$ defined by
$a_n := d(x_n, x) = |x_n - x|$ for all $n \geq m$. We have to prove
that $\lim_{n \to \infty} a_n = 0$ if and only if
$\lim_{n \to \infty} x_n = x$.
\begin{itemize}
\item Let be $\epsilon > 0$. If $\lim_{n \to \infty} a_n = 0$, then
there exists an $N \geq m$ such that $|a_n| < \epsilon$ whenever
$n \geq N$. Thus, there exists an $N \geq m$ such that
$|x_n - x| < \epsilon$ whenever $n \geq N$, which means that
$\lim_{n \to \infty} x_n = x$.
\item Let be $\epsilon > 0$. Conversely, if $\lim_{n \to \infty} x_n
= x$, then there exists an $N \geq m$ such that $|x_n - x| <
\epsilon$ whenever $n \geq N$. But since $|a_n| := |x_n - x|$, it
means that $\lim_{n \to \infty} a_n = 0$, as expected.
\end{itemize}
\end{exo}
\begin{exo}{12.1.2}{Show that the real line with the metric $d(x, y) :=
|x-y|$ is indeed a metric space.}
Using Proposition 4.3.3, this claim is obvious. All claims (a)--(d)
of Definition 12.1.2 are satisfied because:
\begin{enumerate}[label=(\alph*)]
\item comes from Proposition 4.3.3(e)
\item also comes from Proposition 4.3.3(e)
\item comes from Proposition 4.3.3(f)
\item comes from Proposition 4.3.3(g).
\end{enumerate}
\end{exo}
\begin{exo}{12.1.3}{Let $X$ be a set, and let
$d : X \times X \to [0, \infty)$ be a function. With respect to
Definition 12.1.2, give an example of a pair $(X,d)$ which...
\vskip -4mm}
\begin{enumerate}[label=(\alph*)]
\item obeys the axioms (bcd) but not (a).
Consider $X = \rr$, and $d$ defined by $d(x,x) = 1$ and $d(x,y) =
5$ for all $x \neq y \in \rr$.
\item obeys the axioms (acd) but not (b).
Consider $X = \rr$, and $d$ defined by $d(x,y) = 0$ for all $x, y
\in \rr$.
\item obeys the axioms (abd) but not (c).
Consider $X = \rr$, and $d$ defined by $d(x,y) = \max(x-y,0)$ for
all $x,y \in \rr$.
\item obeys the axioms (abc) but not (d).
Consider the finite set $X := \{1,2,3\}$ and the application $d$
defined by $d(1,2) = d(2,1) = d(2,3) = d(3,2) := 1$, and
$d(1,3) = d(3,1) := 5$, and $d(x,x) = 0$ for all $x \in X$.
\end{enumerate}
\end{exo}
\begin{exo}{12.1.4}{Show that the pair $(Y, d|_{Y \times Y})$ defined
in Example 12.1.5 is indeed a metric space.}
By definition, since $Y \subseteq X$, we have $x,y \in X$ whenever
$x,y \in Y$. And furthermore, since $d|_{Y \times Y}(x,y) :=
d(x,y)$, then the application $d|_{Y \times Y}$ obeys all four
statements (a)--(d) of Definition 12.1.2. Thus, $(Y, d|_{Y \times
Y})$ is indeed a metric space.
\end{exo}
\begin{exo}{12.1.5}{Let $n \geq 1$, and let $a_1, a_2, \ldots, a_n$
and $b_1, b_2, \ldots, b_n$ be real numbers. Verify the identity
$\left(\sum_{i=1}^n a_i b_i\right)^2 + \frac{1}{2} \sum_{i=1}^{n}
\sum_{j=1}^{n} (a_i b_j - a_j b_i)^2 = \sum_{i=1}^{n} a_i^2
\sum_{j=1}^{n} b_j^2$, and conclude the Cauchy-Schwarz inequality.
Then use the Cauchy-Schwarz inequality to prove the triangle
inequality.}
Let's prove these three statements.
\begin{enumerate}[label=(\roman*)]
\item To prove the first identity, let's use induction on $n$.
The base case $n=1$ is obvious: on the left-hand side, we just get
$(a_1 b_1)^2$, and on the right-hand side, we get $a_1^2 b_1^2$,
hence the statement.
Now let's suppose inductively that this identity is true for a
given positive integer $n \geq 1$, and let's prove that it is
still true for $n+1$. We have to prove that
\begin{equation}
\label{eq:12.1.5a}
\underbrace{\left(\sum_{i=1}^{n+1} a_i b_i\right)^2}_{:=A} +
\underbrace{\frac{1}{2} \sum_{i=1}^{n+1} \sum_{j=1}^{n+1} (a_i
b_j - a_j b_i)^2}_{:= B}
= \underbrace{\left(\sum_{i=1}^{n+1} a_i^2\right)
\left(\sum_{j=1}^{n+1} b_j^2\right)}_{:=C}
\end{equation}
where we gave a name to each part of the identity for an easier
computation below. Indeed,
\begin{itemize}
\item for $A$, we have
\begin{align*}
A &:= \left(\sum_{i=1}^{n+1} a_i b_i\right)^2 \\
&= \left(a_{n+1} b_{n+1} + \sum_{i=1}^{n} a_i b_i\right)^2
\\
&= (a_{n+1} b_{n+1})^2 + \left(\sum_{i=1}^{n} a_i
b_i\right)^2 + 2 (a_{n+1} b_{n+1}) \sum_{i=1}^{n} a_i b_i
\end{align*}
\item for $B$, we have
\begin{align*}
B &:= \frac{1}{2} \sum_{i=1}^{n+1} \sum_{j=1}^{n+1} (a_i
b_j - a_j b_i)^2 \\
&= \frac{1}{2} \sum_{i=1}^{n} \sum_{j=1}^{n+1} (a_i b_j -
a_j b_i)^2 + \frac{1}{2} \sum_{j=1}^{n+1} (a_{n+1} b_j -
a_j b_{n+1})^2\\
&= \frac{1}{2} \sum_{i=1}^{n} \sum_{j=1}^{n} (a_i b_j - a_j
b_i)^2 +
\underbrace{\frac{1}{2} \sum_{i=1}^{n} (a_i b_{n+1} -
a_{n+1} b_i)^2}_{:= 1/2 \times S} +
\underbrace{\frac{1}{2} \sum_{j=1}^{n} (a_{n+1} b_{j} - a_{j}
b_{n+1})^2}_{:= 1/2 \times S} \\
&\quad + \underbrace{\frac{1}{2} (a_{n+1}b_{n+1} -
b_{n+1}a_{n+1})^2}_{=0} \\
&= \frac{1}{2} \sum_{i=1}^{n} \sum_{j=1}^{n} (a_i b_j - a_j
b_i)^2 + \sum_{k=1}^{n} (a_kb_{n+1} -
a_{n+1}b_k)^2
\end{align*}
\item and thus, for $A + B$, we now use the induction hypothesis (IH)
to get:
\begin{align*}
A + B
&:= (a_{n+1} b_{n+1})^2 + \left(\sum_{i=1}^{n} a_i
b_i\right)^2 +
2 (a_{n+1} b_{n+1}) \sum_{i=1}^{n} a_i b_i\\
&\quad + \frac{1}{2} \sum_{i=1}^{n} \sum_{j=1}^{n} (a_i b_j - a_j
b_i)^2 + \sum_{k=1}^{n} (a_kb_{n+1} - a_{n+1}b_k)^2\\
&= \underbrace{\left(\sum_{i=1}^{n} a_i b_i\right)^2 +
\frac{1}{2} \sum_{i=1}^{n} \sum_{j=1}^{n} (a_i b_j - a_j
b_i)^2}_{\text{apply (IH) here}} \\
&\quad + (a_{n+1} b_{n+1})^2 +
2 (a_{n+1} b_{n+1}) \sum_{i=1}^{n} a_i b_i
+ \sum_{k=1}^{n} (a_kb_{n+1} - a_{n+1}b_k)^2\\
&= \left(\sum_{i=1}^{n} a_i^2\right) \left(\sum_{j=1}^{n}
b_j^2\right)\\
&\quad + (a_{n+1} b_{n+1})^2 +
2 (a_{n+1} b_{n+1}) \sum_{i=1}^{n} a_i b_i
+ \sum_{k=1}^{n} (a_kb_{n+1} - a_{n+1}b_k)^2\\
&= \left(\sum_{i=1}^{n} a_i^2\right) \left(\sum_{j=1}^{n}
b_j^2\right) + (a_{n+1} b_{n+1})^2\\
&\quad + 2 \sum_{i=1}^{n} a_i a_{n+1} b_i b_{n+1} +
\sum_{i=1}^{n}(a_i^2 b_{n+1}^2 - 2a_ib_{n+1}a_{n+1}b_i + a_{n+1}^2b_i^2)\\
&= \left(\sum_{i=1}^{n} a_i^2\right) \left(\sum_{j=1}^{n}
b_j^2\right) + \sum_{i=1}^{n}(a_i^2 b_{n+1}^2 + a_{n+1}^2
b_i^2) \\
&= \left(\sum_{i=1}^{n+1} a_i^2\right) \left(\sum_{j=1}^{n+1}
b_j^2\right) \\
&= C
\end{align*}
so that the identity is indeed true for all natural number $n$.
\end{itemize}
\item We can use this identity to prove the Cauchy-Schwarz identity,
\begin{equation}
\label{eq:12.1.5b}
\left| \sum_{i=1}^{n} a_i b_i\right| \leq \left( \sum_{i=1}^{n}
a_i^2 \right)^{1/2} \left( \sum_{i=1}^{n}
b_i^2 \right)^{1/2}.
\end{equation}
Indeed, since $B \geq 0$ in the identity \eqref{eq:12.1.5a}, we
have
\[\left(\sum_{i=1}^{n} a_i b_i\right)^2 \leq \left(\sum_{i=1}^{n} a_i^2\right)
\left(\sum_{j=1}^{n} b_j^2\right)\]
and thus, taking the square root on both sides, we get
\eqref{eq:12.1.5b}, as expected.
\item Finally, we can use the Cauchy-Schwarz inequality to prove the
triangle inequality.
We have
\begin{align*}
\sum_{i=1}^{n} (a_i^2 + b_i^2)
&= \sum_{i=1}^{n} a_i^2 + \sum_{i=1}^{n} b_i^2 + 2
\sum_{i=1}^{n} a_i b_i
&\\
&\leq \sum_{i=1}^{n} a_i^2 + \sum_{i=1}^{n} b_i^2 + 2
\left(\sum_{i=1}^{n} a_i^2\right)^{1/2} \left(\sum_{i=1}^{n}
b_i^2\right)^{1/2}
&\text{ (by eq. \eqref{eq:12.1.5b})}\\
&\leq \left( \left( \sum_{i=1}^{n} a_i^2 \right)^{1/2}
+ \left( \sum_{i=1}^{n} b_i^2 \right)^{1/2}\right)^2&
\end{align*}
and, since everything is positive, we get the triangle inequality
by taking square roots on both sides.
\end{enumerate}
\end{exo}
\begin{exo}{12.1.6}{Show that $(\rr^n , d_{l^2}$) in Example 12.1.6 is
indeed a metric space.}
We have to show the four axioms of Definition 12.1.2.
\begin{enumerate}[label=(\alph*)]
\item For all $x \in \rr^n$, we have
$d_{l^2}(x,x) = \sqrt{\sum_{i=1}^{n} (x_i - x_i)^2} = 0$, as expected.
\item Positivity: for all $x \neq y \in \rr^n$, there exists at
least one $1 \leq i \leq n$ such that $x_i \neq y_i$, so that
$(x_i-y_i)^2 > 0$, and
$d_{l^2}(x,y) = \sqrt{\sum_{i=1}^{n} (x_i - y_i)^2} > 0$, as
expected.
\item Symmetry: for all $x,y \in \rr^n$, we have
\[d_{l^2}(y,x) = \sqrt{\sum_{i=1}^{n} (y_i - x_i)^2} =
\sqrt{\sum_{i=1}^{n} (x_i - y_i)^2} = d_{l^2}(x,y)\] as expected.
\item Triangle inequality: for all $x,y,z \in \rr^n$, we have
\begin{align*}
d_{l^2}(x,z)
&:= \left(\sum_{i=1}^{n} (x_i - z_i)^2\right)^{1/2} &\\
&= \left(\sum_{i=1}^{n} (a_i + b_i)^2\right)^{1/2}
&\text{with $a_i := x_i-y_i$ and $b_i := y_i-z_i$}\\
&\leq \left(\sum_{i=1}^{n} a_i^2\right)^{1/2} +
\left(\sum_{i=1}^{n} b_i^2\right)^{1/2}
&\text{(Exercise 12.1.5(iii))}\\
&\leq \left(\sum_{i=1}^{n} (x_i - y_i)^2\right)^{1/2} +
\left(\sum_{i=1}^{n} (y_i-z_i)^2\right)^{1/2}& \\
&\leq d_{l^2}(x,y) + d_{l^2}(y,z)&
\end{align*}
as expected.
\end{enumerate}
Thus, $(\rr^n, d_{l^2})$ is indeed a metric space.
\end{exo}
\pagebreak
\begin{exo}{12.1.7}{Show that $(\rr^n , d_{l^1}$) in Example 12.1.7 is
indeed a metric space.}
Once again, let's show the four axioms of Definition 12.1.2.
\begin{enumerate}[label=(\alph*)]
\item For all $x \in \rr^n$, we have
$d_{l^1}(x,x) = \sum_{i=1}^{n} |x_i - x_i| = 0$, as expected.
\item Positivity: for all $x \neq y \in \rr^n$, there exists at
least one $1 \leq i \leq n$ such that $x_i \neq y_i$, so that
$|x_i-y_i| > 0$, and
$d_{l^1}(x,y) = \sum_{i=1}^{n} |x_i - y_i| > 0$, as
expected.
\item Symmetry: for all $x,y \in \rr^n$, we have
\[d_{l^1}(y,x) = \sum_{i=1}^{n} |y_i - x_i| =
\sum_{i=1}^{n} |x_i - y_i| = d_{l^1}(x,y)\]
as expected.
\item Triangle inequality: we already know from Proposition 4.3.3(g)
(generalized to real numbers) that we have the triangle inequality
$|a-c| \leq |a-b| + |b-c|$ for all $a,b,c \in \rr$. Thus, for all
$x,y,z \in \rr^n$, we have
\begin{equation*}
d_{l^1}(x,z) := \sum_{i=1}^{n} |x_i - z_i| \leq \sum_{i=1}^{n}
(|x_i - y_i| + |y_i - z_i|) =: d_{l^1}(x,y) + d_{l^1}(y,z)
\end{equation*}
as expected.
\end{enumerate}
Thus, $(\rr^n, d_{l^1})$ is indeed a metric space.
\end{exo}
\bigskip
\begin{exo}{12.1.8}{Prove the two inequalities in equation (12.1).}
We have to prove that for all $x,y \in \rr^n$, we have
\begin{equation}
\label{eq:12.1.8goal}
d_{l^2}(x,y) \leq d_{l^1}(x,y) \leq \sqrt{n} \, d_{l^2}(x,y)
\end{equation}
\begin{itemize}
\item The first inequality, since everything is non-negative, is
equivalent to $d_{l^2}(x,y)^2 \leq d_{l^1}(x,y)^2$, and we will prove
it in this form.
Indeed, using a trivial product expansion, we have
\begin{align*}
d_{l_1}(x,y)^2
&:= \left(\sum_{i=1}^{n} |x_i - y_i|\right)^2 \\
&= \left(\sum_{i=1}^{n} |x_i - y_i|\right) \times
\left(\sum_{i=1}^{n} |x_i - y_i|\right) \\
&= \sum_{i=1}^{n} |x_i - y_i|^2 + \overbrace{\sum_{1 \leq i,j \leq n ; \,
i\neq j} |x_i-y_i| \times |x_j - y_j|}^{\geq 0} \\
&\geq \sum_{i=1}^{n} |x_i - y_i|^2 =: d_{l^2}(x,y)^2
\end{align*}
as expected.
\item For the second inequality, we use the Cauchy-Schwarz
inequality, which says that
\begin{align*}
d_{l^1}(x,y) &:= \sum_{i=1}^{n} |x_i - y_i|\\
&= \left| \sum_{i=1}^{n} |x_i - y_i| \times 1 \right|\\
&\leq \left(\sum_{i=1}^{n} |x_i - y_i|^2 \right)^{1/2}
\left(\sum_{i=1}^{n} 1^2 \right)^{1/2} \\
&\leq d_{l^2}(x,y) \times \sqrt{n}
\end{align*}
as expected.
\end{itemize}
\end{exo}
\begin{exo}{12.1.9}{Show that the pair $(\rr^n, d_{l^\infty})$ in
Example 12.1.9 is a metric space.}
Once again, let's show the four axioms of Definition 12.1.2.
\begin{enumerate}[label=(\alph*)]
\item For all $x \in \rr^n$, we clearly have
$d_{l^\infty}(x,x) = \sup \{|x_i - x_i| : 1 \leq i \leq n\} = 0$,
as expected.
\item Positivity: for all $x \neq y \in \rr^n$, there exists at
least one $1 \leq j \leq n$ such that $x_j \neq y_j$. Thus
$|x_j-y_j| > 0$, and
$d_{l^\infty}(x,y) = \sup \{|x_i - y_i| : 1 \leq i \leq n\} \geq
|x_j - y_j| > 0$, as expected.
\item Symmetry: for all $x,y \in \rr^n$, we have
\[d_{l^\infty}(x, y) = \sup \{|x_i - y_i| : 1 \leq i \leq n\} =
\sup \{|y_i - x_i| : 1 \leq i \leq n\} = d_{l^\infty}(y, x)\]
as expected.
\item Triangle inequality. Let be $x,y,z \in \rr^n$. We have
$|x_i - z_i| \leq |x_i - y_i| + |y_i - z_i|$ for all
$1 \leq i \leq n$, by Proposition 4.3.3(g). But, by definition of
the supremum, we have $|x_i - y_i| \leq d_{l^\infty}(x,y)$
and $|y_i - z_i| \leq d_{l^\infty}(y,z)$ for all $1 \leq i \leq
n$. Thus, we have $|x_i - z_i| \leq d_{l^\infty}(x,y) +
d_{l^\infty}(y,z)$ for all $1 \leq i \leq n$; i.e., $d_{l^\infty}(x,y) +
d_{l^\infty}(y,z)$ is an upper bound of the set $\{|x_i -
z_i| : 1 \leq i \leq n\}$. By definition of the supremum, it
implies that
\[d_{l^\infty}(x,z) := \sup \{|x_i - z_i| : 1 \leq i \leq n\} \leq d_{l^\infty}(x,y) +
d_{l^\infty}(y,z)\]
as expected.
\end{enumerate}
Thus, $(\rr^n, d_{l^1})$ is indeed a metric space.
\end{exo}
\bigskip
\begin{exo}{12.1.10}{Prove the two inequalities in equation (12.2).}
We have to prove that for all $x,y \in \rr^n$,
\[ \frac{1}{\sqrt{n}} d_{l^2}(x,y) \leq d_{l^\infty} (x,y) \leq
d_{l^2}(x,y).\]
First, a preliminary remark. By definition, we have
$d_{l^\infty}(x,y) := \sup \{|x_i - y_i| : 1 \leq i \leq n\}$ for
all $x,y \in \rr^n$. Since this distance is defined as the supremum
of a finite set, we know (see Chapter 8 of \textit{Analysis I}) that
there exists a $1 \leq m \leq n$ such that
$d_{l^\infty}(x,y) = |x_m - y_m|$ (the supremum belongs to the set).
The index ``$m$'' will have this meaning below.
\begin{itemize}
\item Let's prove the first inequality.
\begin{align*}
\frac{1}{\sqrt{n}} d_{l^2}(x,y)
&:= \sqrt{\frac{1}{n} \sum_{i=1}^{n} (x_i - y_i)^2} \\
&\leq \sqrt{\frac{1}{n} \sum_{i=1}^{n} (x_m - y_m)^2} \\
&\leq \sqrt{\frac{n}{n} (x_m - y_m)^2} \\
&= |x_m - y_m| =: d_{l^\infty} (x,y)
\end{align*}
as expected.
\item Now we prove the second one. We have
\begin{align*}
d_{l^2}(x,y) &:= \sqrt{\sum_{i=1}^{n} (x_i - y_i)^2} \\
&= \sqrt{(x_m - y_m)^2 + \sum_{1 \leq i \leq n ; \, i
\neq m} (x_i - y_i)^2}\\
&\geq \sqrt{(x_m - y_m)^2} = |x_m - y_m| =: d_{l^\infty}(x,y)
\end{align*}
as expected.
\end{itemize}
\end{exo}
\begin{exo}{12.1.11}{Show that the discrete metric $(X, \ddisc)$ in
Example 12.1.11 is indeed a metric space.}
Once again, let's show the four axioms of Definition 12.1.2.
\begin{enumerate}[label=(\alph*)]
\item For all $x \in X$, we have
$\ddisc(x,x) := 0$ by definition, so that there is nothing to
prove here.
\item Positivity: for all $x \neq y \in X$, we have
$\ddisc(x,y) := 1 > 0$ by definition, so that there's still
nothing to prove.
\item Symmetry: for all $x,y \in X$, we have $\ddisc(x,y) =
\ddisc(y,x) = 1$, so that $\ddisc$ obeys the symmetry property.
\item Triangle inequality. Let be $x,y,z \in X$, and let's consider
$\ddisc(x,z)$.
\begin{itemize}
\item If $x=z$, then $\ddisc(x,z) = 0$. And since $\ddisc$ is a
non-negative application, we clearly have $0 =: \ddisc(x,z) \leq
\ddisc(x,y) + \ddisc(y,z)$ for all $y \in X$.
\item If $x \neq z$, then we cannot have both $x=y$ and $y=z$ (it
would be a clear contradiction with $x \neq z$). Thus, at least
one of the propositions ``$x \neq y$'', ``$y \neq z$'' is true.
Another way to say that is $\ddisc(x,y) + \ddisc(y,z) \geq 1$.
But since $\ddisc(x,z) := 1$, we have actually $\ddisc(x,y) +
\ddisc(y,z) \geq \ddisc(x,z)$, as expected.
\end{itemize}
\end{enumerate}
\end{exo}
\begin{exo}{12.1.12}{Prove Proposition 12.1.18.}
First, recall that for all $x,y \in \rr^n$, we have, from Examples
12.1.7 and 12.1.9,
\begin{equation*}
\frac{1}{\sqrt{n}} \, d_{l^2} (x,y) \leq
d_{l^\infty}(x,y) \leq d_{l^2}(x,y) \leq
d_{l^1}(x,y) \leq \sqrt{n} \, d_{l^2}(x,y).
\end{equation*}
Note that $n$ is a real constant here.
\begin{itemize}
\item Let's prove that $(a) \implies (b)$. If
$\lim_{k \to \infty} d_{l^2}(x^{(k)}, x) = 0$, then by the limit
laws, the sequence $t_k := \sqrt{n} \, d_{l^2}(x^{(k)}, x)$ also
converges to $0$ as $k \to \infty$, since $\sqrt{n}$ is a constant
real number. Thus, we have
\[d_{l^2}(x^{(k)}, x) \leq d_{l^1}(x^{(k)}, x) \leq \sqrt{n} \,
d_{l^2}(x^{(k)}, x)\] and, by the squeeze test, this implies
that $\lim_{k \to \infty} d_{l^1}(x^{(k)}, x)$ as expected.
\item Let's prove that $(b) \implies (c)$. If
$\lim_{k \to \infty} d_{l^1}(x^{(k)}, x) = 0$, then we have
\[0 \leq d_{l^\infty}(x^{(k)}, x) \leq d_{l^1}(x^{(k)}, x)\]
and, by the squeeze test, this implies
that $\lim_{k \to \infty} d_{l^\infty}(x^{(k)}, x)$ as expected.
\item Let's prove that $(c) \implies (d)$. Suppose that
$\lim_{k \to \infty} d_{l^\infty}(x^{(k)}, x) = 0$. Then, for all
$1 \leq j \leq n$, we have
$0 \leq |x_j^{k} - x_j| \leq d_{l^\infty}(x^{(k)}, x)$. Still by
the squeeze test, this implies that
$\lim_{k \to \infty} |x_j^{k} - x_j| = 0$, i.e. that
$(x_j^{k})_{k=m}^\infty$ converges to $x_j$ as $k \to \infty$ (by
Lemma 12.1.1), as expected.
\item Finally, let's prove that $(d) \implies (a)$. Using the
definition of convergence is more appropriate here. Let be
$\epsilon > 0$ a positive real number, and let be
$1 \leq j \leq n$. By definition, there exists a natural number
$N \geq m$ such that $|x_j^{(k)} - x_j| \leq \epsilon / \sqrt{n}$
whenever $k \geq N$. Thus, if $k \geq N$, we have
\[d_{l^2}(x^{(k)}, x) := \sqrt{\sum_{j=1}^{n} (x^{(k)}_j - x_j)^2}
\leq \sqrt{\sum_{j=1}^{n} \frac{\epsilon^2}{n}} \leq \epsilon\]
so that $\lim_{k \to \infty} d_{l^2}(x^{(k)}, x) = 0$, i.e.,
$(x^{k})_{k=m}^\infty$ converges to $x$ as $k \to \infty$ in the
$l^2$ metric (by Lemma 12.1.1), as expected.
\end{itemize}
\end{exo}
\begin{exo}{12.1.13}{Prove Proposition 12.1.19.}
Let be $\seq{x^{(n)}}{m}$ a sequence of elements of a set $X$.
\begin{itemize}
\item First suppose that $\seq{x^{(n)}}{m}$ is eventually constant.
Thus, by definition, there exists an $N \geq m$ and an element
$x \in X$ such that $\seq{x^{(n)}}{m} = x$ for all $n \geq N$.
This implies that we have $\ddisc(x^{(n)}, x) = 0$ for all $n \geq
N$. In particular, for all $\epsilon > 0$, we have
$\ddisc(x^{(n)}, x) \leq \epsilon$ whenever $n \geq N$, so that
$\seq{x^{(n)}}{m}$ indeed converges to $x$ with respect to
$\ddisc$.
\item Conversely, suppose that $\seq{x^{(n)}}{m}$ converges to $x$
with respect to $\ddisc$. Let be $\epsilon = 1/2$. By definition,
there exists an $N \geq m$ such that $\ddisc(x^{(n)}, x) \leq 1/2$
whenever $n \geq N$. Since $\ddisc(x^{(n)}, x)$ cannot be $1$, it
is necessarily equal to $0$, so that $x^{(n)} = x$ whenever
$n \geq N$. Thus, the sequence $x^{(n)}$ is indeed eventually
constant.
\end{itemize}
\end{exo}
\begin{exo}{12.1.14}{Prove Proposition 12.1.20.}
Suppose that we have $\lim_{n \to \infty} d(x^{(n)}, x) = 0$ and
$\lim_{n \to \infty} d(x^{(n)}, x') = 0$. Suppose, for the sake
of contradiction, that we have $x \neq x'$. Thus, the real number
$\epsilon := \frac{d(x,x')}{3}$ is positive.
Since $x^{(n)}$ converges to $x$, there exists a $N_1 \geq m$ such
that $d(x^{(n)}, x) \leq \epsilon$ whenever $n \geq N_1$.
Similarly, since $x^{(n)}$ converges to $x'$, there exists a
$N_2 \geq m$ such that $d(x^{(n)}, x') \leq \epsilon$ whenever
$n \geq N_2$.
By the triangle inequality, we thus have, for all $n \geq \max(N_1,
N_2)$,
\[d(x, x') \leq d(x, x^{(n)}) + d(x^{(n)}, x') \leq \epsilon +
\epsilon = \frac{2}{3}d(x,x')\]
which is a contradiction (since $d(x,x') > 0$ by hypothesis).
Thus, the limit is unique, and we must have $x=x'$.
\end{exo}
\bigskip
\begin{exo}{12.1.15}{Let be
$X := \{\seq{a_n}{0} : \sum_{n=0}^{\infty} |a_n| < \infty\}$. We
define on this space the metrics
$d_{l^1}(\seq{a_n}{0}, \seq{b_n}{0}) := \sum_{n=0}^{\infty} |a_n -
b_n|$, and $d_{l^\infty}(\seq{a_n}{0}, \seq{b_n}{0}) := \sup_{n \in
\nn} |a_n - b_n|$. Then...}
We have to prove the following statements.
\begin{enumerate}
\item $d_{l^1}$ is a metric on $X$.
We have to prove the four axioms of Definition 12.1.2.
\begin{enumerate}[label=(\alph*)]
\item Let be $\seq{a_n}{0} \in X$. We have $d_{l^1}(\seq{a_n}{0},
\seq{a_n}{0}) = \sum_{n=0}^{\infty} |a_n - a_n| = 0$, as
expected.
\item Let be $\seq{a_n}{0}, \seq{b_n}{0}$ two distinct elements of
$X$. Since they are distinct, there exists at least one $m \in
\nn$ such as $|a_m - b_m| > 0$. Thus, $d_{l^1}(\seq{a_n}{0},
\seq{b_n}{0}) = \sum_{n=0}^{\infty} |a_n - b_n| \geq |a_m - b_m|
> 0$, as expected.
\item Symmetry: we clearly have
\[d_{l^1}(\seq{b_n}{0}, \seq{a_n}{0}) = \sum_{n=0}^{\infty}
|b_n - a_n| = \sum_{n=0}^{\infty} |a_n - b_n| =
d_{l^1}(\seq{a_n}{0}, \seq{b_n}{0}).\]
\item Finally, let's prove the triangle inequality. Let be
$\seq{a_n}{0}, \seq{b_n}{0}, \seq{c_n}{0} \in X$. Since we have
the triangle inequality for the usual distance $d$ on $\rr$
(i.e., we have $|a_n - c_n| \leq |a_n - b_n| + |b_n - c_n|$ for
all $n \in \nn$), we have immediately
\begin{align*}
d_{l^1}(\seq{a_n}{0}, \seq{c_n}{0})
&:= \sum_{n=0}^{\infty} |a_n - c_n|\\
&\leq \sum_{n=0}^{\infty} (|a_n - b_n| + |b_n - c_n|)
\; \text{ (consequence of Prop. 7.1.11(h))}\\
&\leq \sum_{n=0}^{\infty} |a_n - b_n| + \sum_{n=0}^{\infty}
|b_n - c_n|
\; \text{ (by Proposition 7.2.14(a))}\\
&\leq d_{l^1}(\seq{a_n}{0}, \seq{b_n}{0}) +
d_{l^1}(\seq{b_n}{0}, \seq{c_n}{0}).
\end{align*}
\end{enumerate}
Thus, $d_{l^1}$ is indeed a metric on $X$.
\item $d_{l^\infty}$ is a metric on $X$.
Once again, we have to prove the four axioms of Definition 12.1.2.
\begin{enumerate}[label=(\alph*)]
\item Let be $\seq{a_n}{0} \in X$. We have
$d_{l^\infty}(\seq{a_n}{0}, \seq{a_n}{0}) = \sup_{n \in \nn}
|a_n - a_n| = 0$, as expected.
\item Let be $\seq{a_n}{0}, \seq{b_n}{0}$ two distinct elements of
$X$. Since they are distinct, there exists at least one $m \in
\nn$ such as $|a_m - b_m| > 0$. Thus, $d_{l^\infty}(\seq{a_n}{0},
\seq{b_n}{0}) = \sup_{n \in \nn} |a_n - b_n| \geq |a_m - b_m|
> 0$, as expected.
\item Symmetry: we clearly have
\[d_{l^\infty}(\seq{b_n}{0}, \seq{a_n}{0}) = \sup_{n \in \nn}
|b_n - a_n| = \sup_{n \in \nn} |a_n - b_n| =
d_{l^\infty}(\seq{a_n}{0}, \seq{b_n}{0}).\]
\item Finally, let's prove the triangle inequality. Let be
$\seq{a_n}{0}, \seq{b_n}{0}, \seq{c_n}{0} \in X$. Since we have
the triangle inequality for the usual distance $d$ on $\rr$
(i.e., we have $|a_n - c_n| \leq |a_n - b_n| + |b_n - c_n|$ for
all $n \in \nn$), we have immediately
$|a_m - c_m| \leq \sup_{n \in \nn} |a_n - b_n| + \sup_{n \in
\nn} |b_n - c_n|$ for all $m \in \nn$, by definition of the
supremum. In other words,
$(\sup_{n \in \nn} |a_n - b_n| + \sup_{n \in \nn} |b_n - c_n|)$
is an upper bound for the set $\{|a_m - c_m| : m \in \nn\}$.
Thus we have, still by definition of the supremum,
$\sup_{n \in \nn} |a_n - c_n| \leq \sup_{n \in \nn} |a_n - b_n| +
\sup_{n \in \nn} |b_n - c_n|$, as expected.
\end{enumerate}
Thus, $d_{l^\infty}$ is indeed a metric on $X$.
\item There exist sequences $x^{(1)}$, $x^{(2)}$, ..., of elements
of $X$ (i.e., sequences of sequences) which are convergent with
respect to $d_{l^\infty}$, but are not convergent with respect to
$d_{l^1}$.
Here we are dealing with sequences of sequences: we have a
sequence $(x^{(k)})_{k=1}^\infty$ where each $x^{(k)}$ is
itself a sequence of real numbers. Thus, let's define
$(x^{(k)})_{k=1}^\infty$ as follows:
\[x^{(k)}_n := \left\{
\begin{array}{ll}
1/(k+1) & \text{ if } 0 \leq n \leq k \\
0 & \text{ if } n > k.
\end{array}
\right.\]
Just to make things clearer, we have for instance
\begin{align*}
x^{(1)} &:= \frac{1}{2}, \; \frac{1}{2}, \; 0, \; 0, \; 0, \; \ldots \\
x^{(2)} &:= \frac{1}{3}, \; \frac{1}{3}, \; \frac{1}{3}, \; 0, \; 0, \; \ldots \\
x^{(3)} &:= \frac{1}{4}, \; \frac{1}{4}, \; \frac{1}{4}, \; \frac{1}{4}, \; 0, \; \ldots
\end{align*}
Also, let be the null sequence $\seq{a_n}{0}$ defined by $a_n := 0$ for
all $n \in \nn$. Thus:
\begin{itemize}
\item $(x^{(k)})_{k=1}^\infty$ converges to $\seq{a_n}{0}$ w.r.t.
the metric $d_{l^\infty}$. Indeed, if we consider a given
positive integer $k$ (fixed), we have
\begin{equation*}
|x^{(k)} - a_n| = |x^{(k)}| = \left\{
\begin{array}{ll}
1/(k+1) & \text{ if } 0 \leq n \leq k \\
0 & \text{ if } n > k.
\end{array}
\right.
\end{equation*}
so that $d_{l^\infty}\left(\seq{x^{(k)}_n}{0},
\seq{a_n}{0}\right) := \sup_{n \in \nn} |x^{(k)} - a_n| =
\frac{1}{k+1}$.
Thus,
$\lim_{k \to \infty} d_{l^\infty}\left(\seq{x^{(k)}_n}{0},
\seq{a_n}{0}\right) = 0$, or in other words,
$(x^{(k)})_{k=1}^\infty$ converges to $\seq{a_n}{0}$ w.r.t. the
metric $d_{l^\infty}$ in $X$.
\item But $(x^{(k)})_{k=1}^\infty$ does not converges to
$\seq{a_n}{0}$ w.r.t. the metric $d_{l^1}$. Indeed, we have, for
each given (fixed) $k$,
\begin{align*}
d_{l^1} \left(\seq{x^{(k)}_n}{0}, \seq{a_n}{0}\right)
= \sum_{n=0}^{k} \frac{1}{k+1} = 1
\end{align*}
Thus, we clearly do not have
$\lim_{k \to \infty} d_{l^1}\left(\seq{x^{(k)}_n}{0},
\seq{a_n}{0}\right) = 0$, i.e., $(x^{(k)})_{k=1}^\infty$ does
not converge to $\seq{a_n}{0}$ w.r.t. the metric $d_{l^1}$.
\end{itemize}
\item Conversely, any sequence which converges with respect to
$d_{l^1}$ also converges with respect to $d_{l^\infty}$.
Suppose, for the sake of contradiction, that
$(x^{(k)})_{k=1}^\infty$ does not converge to $\seq{a_n}{0}$
w.r.t. the metric $d_{l^\infty}$, but does converge to $\seq{a_n}{0}$
w.r.t. the metric $d_{l^1}$.
In this case, there exists a $\epsilon > 0$ such that, for all
$k \geq 1$, we have
$(\sup_{n \geq 0} |x^{(k)}_n - a_n|) > \epsilon$. In particulier,
for all $k \geq 1$ and all $n \geq 0$, we have
$|x^{(k)}_n - a_n| > \epsilon$. Thus,
$\sum_{n=0}^{\infty} |x^{(k)}_n - a_n|$ is not even a convergent
series, and we cannot have
$\lim_{k \to \infty} \left(\sum_{n=0}^{\infty} |x^{(k)}_n - a_n|
\right) = 0$.
\end{enumerate}
Note that this exercise actually shows that in this space $X$, the
metrics are not equivalent; instead, the convergence in the taxi cab
metric is stronger than the convergence in the sup norm metric.
Thus, Proposition 12.1.18 is not true for \emph{any} metric space.
\end{exo}
\bigskip
\begin{exo}{12.1.16}{Let $\seq{x_n}{1}$ and $\seq{y_n}{1}$ be two
sequences in a metric space $(X, d)$. Suppose that $\seq{x_n}{1}$
converges to a point $x \in X$, and $\seq{y_n}{1}$ converges to a
point $y \in X$. Show that
$\lim_{n \to \infty} d(x_n, y_n) = d(x, y)$.}
On the one hand, the triangle inequality applied two times to $d$
gives us
\[d(x_n, y_n) \leq d(x_n, x) + d(x,y) + d(y, y_n)\]
but this is only half of what we need to prove the result.
Similarly, we have
\[d(x, y) \leq d(x, x_n) + d(x_n,y_n) + d(y_n, y)\]
so that we can combine the previous two inequalities to get
\[-d(x_n, x) - d(y_n, y) \leq d(x_n, y_n) - d(x, y) \leq d(x_n, x)
+ d(y_n, y)\]
i.e.,
\[|d(x_n, y_n) - d(x, y)| \leq d(x_n, x) + d(y_n, y).\]
Let be $\epsilon > 0$. By hypothesis, there exists a $N_1 \geq 1$
such that $d(x_n, x) \leq \epsilon/2$ whenever $n \geq N_1$.
Similarly, there exists a $N_2 \geq 1$ such that $d(y_n, y) \leq
\epsilon/2$ whenever $n \geq N_2$. Thus, if we set $N :=
\max(N_1,N_2)$, then for all $n \geq N$ we have
\[|d(x_n, y_n) - d(x, y)| \leq d(x_n, x) + d(y_n, y) \leq 2
\epsilon/2 \ leq \epsilon\]
which shows that $\lim_{n \to \infty} d(x_n, y_n) = d(x, y)$, as
expected.
\end{exo}
\pagebreak
\begin{exo}{12.2.1}{Verify the claims in Example 12.2.8}
Let be $(X, \ddisc)$ a metric space, and $E$ a subset of $X$.
\begin{itemize}
\item Let be $x \in E$. Then $x$ is an interior point of $E$.
Indeed, we have $B(x, 1/2) = \{x\} \subseteq E$.
\item Let be $y \notin E$. Then $y$ is an exterior point of $E$.
Indeed, we have $B(y, 1/2) \cap E = \{y\} \cap E = \emptyset$.
\item Finally, there are no boundary points of $E$ in $(X, \ddisc)$.
Indeed, let be $r > 0$ and any $x \in X$. We will always have
$B(x, r) = \{x\}$ by definition of the discrete metric $\ddisc$.
Thus, we have either $x \in E$ and then $x \in \inter (E)$, or $x
\notin E$ and then $x \in \ext(E)$. Thus, $E$ has no boundary
points.
\end{itemize}
\end{exo}
\begin{exo}{12.2.2}{Prove Proposition 12.2.10.}
We have to prove the following implications:
\begin{itemize}
\item Let's show that $(a) \implies (b)$. We will use the
contrapositive, assuming that $x_0$ is neither an interior point
of $E$, nor a boundary point of $E$. By definition, it means that
$x_0$ is an exterior point of $E$, i.e. that there exists $r > 0$
such that $B(x_0, r) \cap E = \emptyset$. This is precisely the
negation of $x_0$ being an adherent point of $E$. Thus, we have
showed that if $x_0$ is an adherent of of $E$, it is either an
interior point of a boundary point.
\item Let's show that $(b) \implies (c)$. Let be a positive integer
$n > 0$, and suppose that $x_0$ is either an interior point of
$E$, or a boundary point of $E$. In either case, the set
$A_n := B(x_0, 1/n) \cap E$ is non empty, i.e., there exists
$a_n \in X$ such that $d(a_n, x_0) < 1/n$. By the (countable)
axiom of choice, we can define a sequence $\seq{a_n}{1}$ such that
$a_n \in A_n$ for all $n \geq 1$.
Let be $\epsilon > 0$. There exists $N > 0$ such that $1/N <
\epsilon$ (Exercise 5.4.4). Thus, for all $n \geq N$, we have
\[d(a_n, x_0) < \frac{1}{n} \leq \frac{1}{N} < \epsilon\]
i.e., the sequence $\seq{a_n}{1}$ converges to $x_0$ with respect
to the metric $d$, as expected.
\item Finally, let's show that $(c) \implies (a)$. Let be $r > 0$.
If $\seq{a_n}{1}$ in $E$ converges to $x_0$ with respect to $d$,
then there exists a $n$ such that $d(x_0, a_n) < r$. But since
$a_n \in E$, it means that $B(x_0, r) \cap E$ is non empty, i.e.
that $x_0$ is an adherent point of $E$.
\end{itemize}
\end{exo}
\begin{exo}{12.2.3}{Prove Proposition 12.2.5.}
Let be $(X,d)$ a metric space.
\begin{enumerate}[label=(\alph*)]
\item Let be $E \subseteq X$. First suppose that $E$ is open; this
means that $E \cap \partial E = \emptyset$. Let be $x \in E$, then we
have $x \notin \partial E$. But since $x \in E$, we have $x \in
\adh{E}$, and thus $x \in \inter(E)$ by Proposition 12.2.10(b). We
have shown that $x \in E \implies x \in \inter(E)$, and since the
converse implication is trivial (Remark 12.2.6), we have $E =
\inter(E)$ as expected.
Now suppose that $E = \inter(E)$. Let be $x \in E$. We thus have
$x \in \inter(E)$. By definition, $x$ is thus not a boundary point
of $E$, i.e. x $\notin \partial E$. This means that $E \cap
\partial E = \emptyset$, i.e. that $E$ is open, as expected.
\item Let be $E \subseteq X$. First suppose that $E$ is closed; i.e.
that $\partial E \subseteq E$. Let be $x \in \adh{E}$. By
Proposition 12.2.10, we have $\adh{E} = \inter(E) \cup \partial
E$; such that $\adh{E}$ is the union of two subsets of $E$, and
thus is itself a subset of $E$, as expected.
Conversely, suppose that $\adh{E} \subseteq E$. It means that
$\inter(E) \cup \partial E \subseteq E$, and in particular that
$\partial E \subseteq E$, i.e. that $E$ is closed, as expected.
\item Let be $x_0 \in X$, $r > 0$ and $E := B(x_0, r)$. To show that
$E$ is open, we must show that $E = \inter(E)$ (by Proposition
1.2.15(a)), and in particular that $E \subseteq \inter(E)$ (the
converse inclusion being trivial). Let be $x \in E$, and let's
show that $x \in \inter(E)$. By definition, we have
$d(x, x_0) < r$, so that $\epsilon := r - d(x, x_0)$ is a positive
real number. Thus, let be $y \in B(x, \epsilon)$. By the triangle
inequality, we have
\begin{align*}
d(x_0, y) &< d(x, x_0) + d(x,y) \\
&< d(x, x_0) + \epsilon \\
&< d(x, x_0) + r - d(x, x_0) = r
\end{align*}
so that $y \in E$. Thus, there exists $\epsilon > 0$ such that
$B(x, \epsilon) \subseteq E$, i.e., $x$ is an interior point of
$E$. This shows that $E \subseteq \inter(E)$, as expected.
Now let be $F := \{x \in X : d(x, x_0) \leq r\}$, and let be
$\seq{a_n}{1}$ a convergent sequence in $F$. To show that $F$ is
closed, we have to show that $\ell := \lim_{n \to \infty} a_n$
lies in $F$ (Proposition 12.2.15(b)). Suppose, for the sake of
contradiction, that $\ell \notin F$. We thus have $d(\ell, x_0) >
r$, so that $\epsilon := d(\ell, x_0) - r$ is a positive real
number. Since $\seq{a_n}{1}$ converges to $\ell$, there exists a
$N > 0$ such that $d(a_n, \ell) < \epsilon$ whenever $n \geq N$.
By the triangle inequality, for $n \geq N$, we have
\begin{align*}
d(x_0, \ell) &\leq d(x_0, a_n) + d(a_n, \ell) \\
d(x_0, a_n) &\geq d(x_0, \ell) - d(a_n, \ell) \\
&\geq d(x_0, \ell) - \epsilon \\
&\geq d(x_0, \ell) + r - d(\ell, x_0) \\
&\geq r
\end{align*}
and thus, $a_n \notin B(x_0, r)$, a contradiction. Thus, we must
have $\ell \in F$, so that $F$ is indeed a closed set.
\item Let be $\{x_0\}$ a singleton with $x_0 \in X$. To show that
$E$ is closed, we may use Proposition 12.2.15(b), and show that
$\{x_0\}$ contains all its adherent points. Let be $\seq{a_n}{1}$
a convergent sequence in $\{x_0\}$; it can only be the constant
sequence $x_0, x_0, \ldots$. Since it is a constant sequence, its
limit can only be $x_0$ itself, and this limit belongs to
$\{x_0\}$. Thus, $\{x_0\}$ is a closed set.
\item First we can form a lemma: for any subset $E \subseteq X$, we
have $\inter(E) = \ext(X \backslash E)$. This is a direct
consequence of Definition 12.2.5. Indeed, $x \in \inter(E)$ iff
there exists a $r > 0$ such that $B(x,r) \subseteq E$, which is
equivalent to
``$\exists r>0 : B(x,r) \cap (X \backslash E) = \emptyset$'',
which is equivalent to $x \in \ext(X \backslash E)$.
This implies that the interior points of $E$ are the exterior points
of $X \backslash E$, and conversely, that the exterior points of
$E$ are the interior points of $X \backslash E$. Thus, in
particular, we have this useful fact:
\begin{equation}
\label{eq:12.2.3}
\partial E = \partial(X \backslash E).
\end{equation}
Now we go back to the main proof. First suppose that $E$ is open.
Thus, by Definition 12.2.12, we have
$E \cap \partial E = \emptyset$, so that
$\partial E \subseteq X \backslash E$, which means that
$X \backslash E$ is a closed set. The converse also applies: if we
suppose that $X \backslash E$ is closed, then
$\partial (X \backslash E) \subseteq X \backslash E$. By equation
\eqref{eq:12.2.3} above, this is equivalent to
$\partial E \subseteq X \backslash E$, and thus we have
$\partial E \cap E = \emptyset$. This means that $E$ is open, as
expected.\footnote{This important result will be used in future
proofs to turn any statement on closed sets into a statement on
open sets.}
\item Let $E_1, \ldots, E_n$ be open sets. Thus, for all
$1 \leq i \leq n$, if $x \in E_i$, there exists a $r_i > 0$ such
that $B(x, r_i) \subseteq E_i$. Let's define $r := \min(r_1,
\ldots, r_n)$. We have $B(x, r) \subseteq B(x, r_i) \subseteq E_i$
for all $1 \leq i \leq n$, i.e. $B(x,r) \subseteq E_1 \cap \ldots
\cap E_n$. Thus, $E_1 \cap \ldots \cap E_n$ is an open set.
Also, let $F_1, \ldots, F_n$ be closed sets. By the previous
result (e), the complementary sets
$X\backslash F_1, \ldots X\backslash F_n$ are open sets. Thus, we
have just proved that
$(X \backslash F_1) \cap \ldots \cap (X \backslash F_n)$ is an
open set. But we have
$(X \backslash F_1) \cap \ldots \cap (X \backslash F_n) = X
\backslash (F_1 \cup \ldots \cup F_n)$, and this set is open.
Thus, by (e), its complementary set, $F_1 \cup \ldots \cup F_n$,
is closed, as expected.
\item Let $(E_\alpha)_{\alpha \in I}$ be open sets. Suppose that we
have $x \in \bigcup_{\alpha \in I} E_\alpha$. By definition, there
exists a $i \in I$ such that $x \in E_i$. Since $E_i$ is an open
set, there exists $r_i > 0$ such that
$B(x, r_i) \subseteq E_i \subseteq \bigcup_{\alpha \in I}
E_\alpha$. Thus, by (a), $\bigcup_{\alpha \in I} E_\alpha$ is an
open set, as expected.
Now let be $(F_\alpha)_{\alpha \in I}$ be closed sets. Suppose
that we have a convergent sequence $\seq{x_n}{1}$ such that
$x_n \in \bigcap_{\alpha \in I} F_\alpha$ for all $n \geq 1$.
Thus, for all $\alpha \in I$, the sequence $\seq{x_n}{1}$ entirely
belongs to the closed set $F_\alpha$, so that its limit $\ell$
also lies in $F_\alpha$ according to (b). Thus,
$\ell \in \bigcup_{\alpha \in I} F_\alpha$, so that
$\bigcap_{\alpha \in I} F_\alpha$ is a closed set, as expected.
\item Let be $E \subseteq X$.
\begin{itemize}
\item Let's show that $\inter(E)$ is the largest open set included
in $E$. It has not clearly be proved in the main text that
$\inter(E)$ is an open set, so we begin by proving it. Let be
$x \in \inter(E)$. By definition, there exists $r > 0$ so that
$B(x, r) \subseteq E$. But by (c), we know that $B(x, r)$ is an
open set, so that any point $y$ of $B(x, r)$ is an interior
point of this open ball, and thus an interior point of $E$.
Thus, $\inter(E)$ is open.
Now consider another open set $V \subseteq E$, and let's show
that $V \subseteq \inter(E)$. If $x \in \inter(V)$, then there
exists $r > 0$ such that $B(x,r) \subseteq V \subseteq E$, so
that $x \in \inter(E)$. This shows that $V \subseteq \inter(E)$,
as expected.
\item Similarly, let's show that $\adh{E}$ is the smallest closed
set that contains $E$. First we show that $\adh{E}$ is closed,
i.e. that $\adh{\adh{E}} \subseteq \adh{E}$. (Hint: see Exercise
9.1.6 for an intuition.) Let be $x \in \adh{\adh{E}}$. By
definition, for all $r > 0$,
$B(x,r) \cap \adh{E} \neq \emptyset$. Thus, there exists
$y \in B(x,r)$ such that $y \in \adh{E}$. Thus, because $B(x,r)$
is an open set and $y$ is adherent to $E$, there exists
$\epsilon > 0$ such that $B(y, \epsilon) \subseteq B(x,r)$ and
$B(y, \epsilon) \cap E \neq \emptyset$; i.e., there exists
$z \in B(y, \epsilon) \subseteq B(x,r)$ such that $z \in E$. We
have showed that whenever $x \in \adh{\adh{E}}$, we have $B(x,r)
\cap E \neq \emptyset$ for all $r > 0$, i.e. that $x$ is an
adherent point of $E$, as expected. Thus, $\adh{E}$ is closed.
Now we consider a closed set $K$ such that $E \subseteq K$, and
we have to show that $\adh{E} \subseteq K$. Let be
$x \in \adh{E}$. By definition, for all $r > 0$, we have
$B(x,r) \cap E \neq \emptyset$. But since $E \subseteq K$, we
also have $B(x,r) \cap K \neq \emptyset$ for all $r > 0$. Thus,
$x$ is an adherent point of $K$, i.e., $x \in \adh{K}$. But
since $K$ is closed, we have $K = \adh{K}$, and thus $x \in K$.
This shows that $\adh{E} \subseteq K$, as expected.
\end{itemize}
\end{enumerate}
\end{exo}