-
Notifications
You must be signed in to change notification settings - Fork 0
/
completeness.tex
821 lines (799 loc) · 23.6 KB
/
completeness.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
\documentclass[main.tex]{subfiles}
\begin{document}
% 4
\section{Uniform Completeness}
After Yosida's Theorem~\ref{3.14},
a reasonable question is:
When is the map~$E\ra \Cont{\Phi}$ surjective?
%
% 4.1
%
\begin{psec}{4.1}{Exercise}
Let $E$ be an Archimedean Riesz space with a unit, $e$.
\begin{enumerate}
\item \label{4.1-1}
Let $u\in E^+$
and let $E_u$ denote the principal ideal of~$E$
generated by~$u$ (\ref{2.11}).
Then~$E_u$ is Archimedean and has~$u$ as a unit.
Show that
\begin{equation*}
\| x \|_e \leq \| u \|_e \, \| x \|_u \qquad (x\in E)
\end{equation*}
so that the identity map $E_u \ra E$
is continuous relative to~$\|\cdot\|_u$ and~$\|\cdot \|_e$.
%
\item \label{4.1-2}
Now let~$u$ be a unit of~$E$.
Show that~$\|\cdot\|_e$ and~$\|\cdot\|_u$
determine the same topology on~$E$,
and, less trivially:
$E$ is complete under (the metric induced by)
the norm~$\|\cdot\|_u$
if and only if it is complete under~$\|\cdot\|_e$.
\end{enumerate}
\end{psec}
%
% 4.2
%
\begin{psec}{4.2}{Definitions}
Let $E$ be an Archimedean unitary Riesz space.
The topology induced by~$\|\cdot\|_e$,
where~$e$ is any unit,
is called the \keyword{(relatively) uniform topology}.
$E$ is said to be
\keyword{uniformly complete}
if it is complete relative to~$\|\cdot\|_e$
where~$e$ is any unit.
We use terms such as ``relatively uniformly closed''
and ``relatively uniformly convergent''
with the obvious meanings.
For instance,
if~$e$ is a unit in~$E$,
a sequence $(x_n)_n$
``converges relatively uniformly''
to an element $a$
if and only if for every $\varepsilon>0$
there is an~$N$ with
\begin{equation*}
|x_n - a| \leq \varepsilon e \qquad (n\geq N)\htam{.}
\end{equation*}
In spaces such as~$\ell^\infty$
and $\BCont{X}$
the constant function $\mathbb{1}$
serves as a unit.
As~$\|\cdot\|_\mathbb{1}$ coincides
with the sup-norm,
\emph{in these spaces,
relatively uniform convergence is just uniform convergence.}
Accordingly,
the qualification ``relatively'' is often dropped.
(However,
if~$E$ is the space of
all scalar multiples of the function
$i\colon x \mapsto x\quad (x\in [0,\infty)\,)$,
then the sequence $(n^{-1}i)_n$
converges to~$0$
relatively uniformly,
but not uniformly.
\end{psec}
%
% 4.3
%
\begin{psec}{4.3}{Exercise}
Let $E$ be an Archimedean unitary Riesz space.
Show that,
with respect to the uniform topology,
the map $x\mapsto|x|$
is continuous
and every ``interval'' $[a,b]$ is closed.
\end{psec}
%
% 4.4
%
\begin{psec}{4.4}{Exercise}
Show that in the Riesz space $\Cont{[0,1]}$
the principal Riesz ideal
generated by the function $t\mapsto t$
is not uniformly closed in $\Cont{[0,1]}$.
\end{psec}
%
% 4.5
%
\begin{psec}{4.5}{Theorem}
(Continuation of~\ref{3.14})
\statement{%
If~$E$ is uniformly complete,
then~$\hat{E}=\Cont{\Phi}$.
Thus:
Every uniformly complete (Archimedean, unitary) Riesz space
is Riesz isomorphic to $\Cont{\Phi}$
for some compact Hausdorff space~$\Phi$.
}
\end{psec}
\begin{proof}
$E$ is complete relative to~$\|\cdot\|_e$.
Hence, $\hat{E}$ is complete relative to~$\|\cdot\|_{\hat{e}}$,
which is~$\|\cdot\|_\infty$.
Then~$\hat{E}$ is closed in~$\Cont{\Phi}$
(relative to~$\|\cdot\|_\infty$).
But it is also dense,
so~$\hat{E}$ equals~$\Cont{\Phi}$. \xqed
\end{proof}
%
% 4.6
%
\begin{psec}{4.6}{Examples}
\begin{enumerate}
\item \label{4.6-1}
For every set~$X$
the space $\ell^\infty(X)$ (\ref{1.5}\ref{1.5-2})
is uniformly complete.
(A brief proof:
Let $f_1,f_2,\dotsc\in \ell^\infty(X)$
and let $(\varepsilon_n)_n$ be a sequence in~$(0,\infty)$,
tending to~$0$ and with
\begin{equation*}
\|f_n-f_N\|_\infty \leq \varepsilon_N \qquad (n\geq N)\htam{.}
\end{equation*}
For every $x\in X$ we have $|f_n(x)-f_N(x)|\leq\varepsilon_N\quad (n\geq N)$,
so that $(f_n(x))_n$ is a Cauchy sequence in~$\R$;
let~$f(x)$ be its limit.
Thus we obtain a function~$f$.
For all~$x$,
$|f(x)-f_1(x)|=\lim_{n\ra \infty} |f_n(x)-f_1(x)|\leq \varepsilon_1$,
so~$f$ is bounded: $f\in \ell^\infty(X)$.
Take $N\in\N$;
then for all~$x$,
$|f(x)-f_N(x)|
=\lim_{n\ra\infty}|f_n(x)-f_N(x)|
\leq\varepsilon_N$.
Thus, $\| f-f_N \|_\infty \leq \varepsilon_N \ra 0$.)
%
\item \label{4.6-2}
If $X$ is a topological space,
$\BCont{X}$ is a $\|\cdot\|_\infty$-closed subset of~$\ell^\infty(X)$,
hence is uniformly complete.
(Proof of the closedness:
Let $f\in \ell^\infty(X)$ lie in the closure of~$\BCont{X}$.
Let $x_i\ra a$ in~$X$.
Question:
must $f(x_i)\ra f(a)$?
Take $\varepsilon>0$.
Question:
Is there an~$i_0$
such that $|f(x_i)-f(a)|<\varepsilon\quad(i\geq i_0)$?
Choose $g\in\BCont{X}$
with $\|f-g\|_\infty\leq \sfrac{\varepsilon}{3}$.
There is an~$i_0$ such that $|g(x_i)-g(a)|<\sfrac{\varepsilon}{3}$
for all $i\geq i_0$.
For such~$i$ we have
$|f(x_i)-f(a)|
\leq|f(x_i)-g(x_i)| + |g(x_i)-g(a)| + |g(a)-f(a)|
\leq \varepsilon$.)
%
\item \label{4.6-3}
In particular,
of course,
the~$\Cont{\Phi}$ of Yosida's Theorem is uniformly complete.
%
\item \label{4.6-4}
More examples follow later on in this chapter.
\end{enumerate}
\end{psec}
%
% 4.7
%
\begin{psec}{4.7}{Lemma}
\statement{%
Let $E$ be a unitary Archimedean Riesz space.
If~$E$ is uniformly complete,
then so is every principal ideal of~$E$.
}
\end{psec}
\begin{proof}
By Theorem~\ref{4.5},
$E$ is Riesz isomorphic to~$\Cont{X}$
for some compact space~$X$,
so we may as well assume that~$E$ \emph{is}~$\Cont{X}$.
Let~$E_u$ be the principal ideal of~$E$
generated by an element~$u$ of~$\Cont{X}^+$.
Let~$(f_n)_n$ and~$(\varepsilon_n)_n$
be sequences in~$E_u$ and~$(0,\infty)$,
respectively,
with $\varepsilon_n\downarrow 0$, and
\begin{equation*}
\| f_n - f_N \|_u \leq \varepsilon_N \qquad (n\geq N)\htam{,}
\end{equation*}
i.e.,
\begin{equation*}
\label{eq4.7} \tag{$*$}
|f_n(x)-f_N(x)|\leq \varepsilon_N \,u(x) \qquad (x\in X;\ n\geq N)\htam{.}
\end{equation*}
For $n,N\in\N$ with $n\geq N$
we have $\|f_n-f_N\|_\infty \leq \varepsilon_N \|u\|_\infty$,
by~\ref{4.1}\ref{4.1-1};
so that~$(f_n)_n$ is a Cauchy sequence relative to~$\|\cdot\|_\infty$.
As~$\Cont{X}$ is complete,
this sequence converges to some~$f$ in~$\Cont{X}$.
By letting~$n$ tend to~$\infty$
in~\eqref{eq4.7} we obtain,
for all~$N$,
\begin{equation*}
|f(x)-f_N(x)|\leq\varepsilon_N\,u(x)\qquad(x\in X)\htam{.}
\end{equation*}
Then $|f|\leq |f_1|+\varepsilon_1 u$,
so that $f\in E_u$;
and $\| f - f_N\|_u\leq \varepsilon_N$,
so that~$(f_n)_n$
converges to~$f$ in the sense of~$\|\cdot\|_u$. \xqed
\end{proof}
%
% 4.8
%
\begin{psec}{4.8}{Comment}
There is room for confusion here:
It may happen that~$E_u$ is uniformly complete
but not uniformly closed;
see, e.g., \ref{4.4}.
The tricky part is that the word ``uniformly''
refers first to the unit of~$E_u$,
later to the unit of~$E$.
\end{psec}
%
% 4.9
%
\noindent
Let us turn to uniform completeness of quotients.
There we hit a snag:
a quotient of a uniformly complete Riesz space
need not even be Archimedean.
We have the following result.
\begin{psec}{4.9}{Theorem}
\statement{Let $D$ be a Riesz ideal
in a unitary Riesz space~$E$.
Then the quotient Riesz space~$E/D$
is Archimedean if and only if~$D$
is a uniformly closed subset of~$E$.}
\end{psec}
\begin{proof}
Let $Q$ be the quotient map $E\ra E/D$.
Choose a unit~$e$ in~$E$.
Observe that~$Qe$ is a unit in~$E/D$.
\begin{enumerate}[label=(\Roman*)]
\item \label{4.9-I}
Suppose $E/D$ is Archimedean.
Then~$Qe$ defines a norm~$\|\cdot\|_{Qe}$ on~$E/D$,
and from the definitions it is apparent that
\begin{equation*}
\|Qx\|_{Qe} \leq \| x \|_e \qquad (x\in E)\htam{.}
\end{equation*}
(Actually,
$\|\cdot\|_{Qe}$ is the quotient norm:
$\|Qx\|_{Qe}=\inf\{ \|y\|_e\colon Qy=Qx\}$.)
If~$a\in E$ lies in the $\|\cdot\|_e$-closure of~$D$,
then~$Qa$ lies in the $\|\cdot\|_{Qe}$-closure of~$Q(D)=\{0\}$,
so that~$Qa=0$ and~$a\in D$.
%
\item \label{4.9-II}
Suppose $D$ is $\|\cdot\|_e$-closed in~$E$.
Let $a,b\in E^+$,
$Qa\leq n^{-1} Qb$ for all $n$;
we prove $Qa=0$.
As~$b$ is smaller than some multiple of~$e$,
we may assume~$b\leq e$.
For each~$n$,
put $x_n := (a-n^{-1}b)^+$;
then $x_n \in D$
because $Qx_n=0$.
We have $a=a^+ \geq x_n \geq a-n^{-1} b\quad (n\in \N)$
and $b\leq e$;
it follows that $\| a - x_n \|_e\leq n^{-1}$ for each $n$.
Hence, $a$ lies in the $\|\cdot\|_e$-closure of~$D$,
which is~$D$. \xqed
\end{enumerate}
\end{proof}
%
% 4.10
%
\noindent For uniformly closed ideals everything is fine:
\begin{psec}{4.10}{Theorem}
\statement{Let $D$ be a uniformly closed Riesz ideal
in a uniformly complete (unitary Archimedean) Riesz space~$E$.
Then~$E/D$ is (unitary, Archimedean, and) uniformly complete.}
\end{psec}
%
% 4.11
%
\noindent Before turning to the proof
we consider a special situation:
\begin{psec}{4.11}{Theorem}
\statement{%
Let $X$ be a compact Hausdorff space,
$D$ a uniformly closed Riesz ideal of~$\Cont{X}$.
Then there exists a closed subset~$A$ of~$X$ such that
\begin{equation*}
D = \{ f\in\Cont{X}\colon f|A=0\}\htam{.}
\end{equation*}}%
\end{psec}
\begin{proof}
Define
\begin{equation*}
A := \bigcap_{f\in D} \bigl\{ x\colon f(x)=0 \bigr\}\htam{.}
\end{equation*}
$A$ is a closed set (possibly $\varnothing$).
Trivially, $D\subseteq\{ f\colon f|A=0\}$.
Conversely, suppose
\begin{equation*}
f\in \Cont{X}\htam{, }\qquad f|A=0\htam{; }
\end{equation*}
we prove $f\in D$.
For convenience,
assume~$f\geq 0$
(or replace~$f$ by~$|f|$).
Let~$\varepsilon\in(0,\infty)$;
we are done if we can find an~$f_\varepsilon$ in~$D$
with~$|f-f_\varepsilon|\leq \varepsilon\mathbb{1}$.
Claim:
$(f-\varepsilon\mathbb{1})^+$ is such an~$f_\varepsilon$.
First,
note that $f=f^+\geq (f-\varepsilon\mathbb{1})^+
\geq f-\varepsilon\mathbb{1}$,
so that $0\leq f - (f-\varepsilon \mathbb{1})^+
\leq \varepsilon \mathbb{1}$.
We are done if $(f-\varepsilon \mathbb{1})^+\in D$.
The set $B:=\{ x\colon f(x)\geq \varepsilon \}$ is closed.
For every point~$b$ of~$B$
we have~$b\notin A$,
so there is a~$g$ in~$D^+$ with~$g(b)>0$;
then there is also a~$g$ in~$D^+$
with $g(b)>f(b)$.
Hence:
\begin{align*}
B \subseteq &\bigcup_{\lsub{g\in D^+}{g\in D}}
\bigl\{ x\colon g(x) > f(x) \bigr\}\htam{.}\\
\intertext{%
By the compactness of~$B$,
there exist $g_1,\dotsc,g_N$ in~$D^+$ such that}
B \subseteq &\bigcup_{\lsub{\phantom{g}\,n}{g\in D}}
\bigl\{ x\colon g_n(x)>f(x) \bigr\}\htam{.}%
\end{align*}
Setting $g:=g_1\vee \dotsb \vee g_N$,
we get~$g\in D^+$
and $g>f$ on~$B$;
then $g\geq(f-\varepsilon\mathbb{1})^+$,
and $(f-\varepsilon \mathbb{1})^+ \in D$. \xqed
\end{proof}
\begin{proof} of Theorem~\ref{4.10}:
As in our proof of Lemma~\ref{4.7},
we assume $E=\Cont{X}$ for some compact~$X$.
Then $D=\{ f\colon f|A = 0 \}$ for some closed $A\subseteq X$,
and $E/D$ is Riesz isomorphic to~$\Cont{A}$~(\ref{2.9}).
But $\Cont{A}$ is uniformly complete (\ref{4.6}\ref{4.6-1}). \xqed
\end{proof}
%
% 4.12
%
\noindent The following is an application of Theorem~\ref{4.5}.
\begin{psec}{4.12}{The Stone--\v{C}ech compactification}
Let~$X$ be a topological space.
\begin{enumerate}[label=(\Roman*)]
\item \label{4.12-I}
The Stone--\v{C}ech compactification of~$X$,
denoted
\begin{equation*}
\beta X
\end{equation*}
is the compact Hausdorff space
obtained from~$\BCont{X}$
by applying the Yosida Theorem.
Explicitly:
$\beta X$ is
the set of all Riesz homomorphisms
$\BCont{X}\ra \R$ that send~$\mathbb{1}_X$ to~$1$,
topologized as a subset of~$\R^{\BCont{X}}$.
It is,
indeed,
a compact Hausdorff space,
and we have a map
$\beta\colon X \ra \beta X$ given by
\begin{equation*}
\tag{$*$} \label{eq4.12*}
\beta(x)\ \htam{ is the evaluation }\
f\mapsto f(x)
\quad(f\in \BCont{X}\,)\htam{.}
\end{equation*}
(The notation is standard but deplorable:
the range of~$\beta$ is~$\beta(X)$, not $\beta X$.)
The map $f\mapsto \hat{f}$ of~$\BCont{X}$ into~$\Cont{\beta X}$
satisfies $\hat{f}(\beta(x))=(\beta(x))(f)$, i.e.,
\begin{alignat*}{2}
\tag{$**$} \label{eq4.12**}
&\hat{f}(\beta(x)) = f(x) \qquad (x\in X\htam{,}\ f\in\BCont{X}\,)\htam{.}\\
&\xymatrix{
X\ar[r]^f \ar[d]_\beta & \R \\
\beta X \ar[ru]_{\hat f}}
\end{alignat*}
As $\BCont{X}$ is uniformly complete (\ref{4.6}\ref{4.6-2}),
the correspondence $f\mapsto \hat{f}$
is a Riesz isomorphism of~$\BCont{X}$ onto~$\Cont{\beta X}$.
%
\item \label{4.12-II}
$\beta(X)$ is a subset of~$\beta X$.
If $x,y\in X$ and $\beta(x)=\beta(y)$,
then $f(x)=f(y)$ for all $f$~in~$\BCont{X}$.
Hence,
for every~$f$ in~$\BCont{X}$
the formula $\beta(x)\mapsto f(x)$
determines a function on~$\beta(X)$.
By the above,
this function (is continuous and)
extends to the continuous
function $\hat{f}\colon \beta X \ra \R$.
Often (see below) $\beta$ is injective
and a homeomorphism of~$X$ onto~$\beta(X)$;
then~$X$ and~$\beta(X)$ may be identified,
and one may view~$X$ as a subset of~$\beta X$,
and~$\hat f$ as an extension of~$f$.
%
\item \label{4.12-III}
If $X$ itself is a compact Hausdorff space,
then~$\beta$ is a homeomorphism of~$X$ onto~$\beta X$;
that is~\ref{3.15}.
%
\item \label{4.12-IV}
$\beta(X)$ is always a dense subset of~$\beta X$.
Indeed, suppose it is not.
By Urysohn's Lemma
there exists a nonzero~$g$ in $\Cont{\beta X}$
such that $g=0$ on (the closure of) $\beta(X)$;
then we have a nonzero~$f$ in $\BCont{X}$
with $\hat f = 0$ on~$\beta (X)$;
but then (with \eqref{eq4.12**})
$f = \hat f \circ \beta = 0$:
contradiction.
\item \label{4.12-V}
Let~$(x_i)_i$ be a net in~$X$
and let $a\in X$.
As $\{\hat f\colon f\in\BCont{X}\}=\Cont{\beta X}$,
we have,
thanks to~\ref{2.18},
\begin{alignat*}{2}
\beta(x_i) \longrightarrow \beta(a)
& \quad\iff\quad \hat f(\beta(x_i))\longrightarrow \hat f(\beta(a))\qquad
(f\in \BCont{X}\,)\\
& \quad\iff\quad f(x_i)\longrightarrow f(a) \qquad
(f\in \BCont{X}\,)
\end{alignat*}%
In particular,
this means that~$\beta$ is continuous.
%
\item \label{4.12-VI}
Since $\beta X$ is completely regular,
$\beta(X)$ is completely regular too.
So if $\beta$ is a homeomorphism $X\ra \beta(X)$,
$X$ is completely regular.
Conversely, $\beta$ is a homeomorphism $X\rightarrow \beta(X)$,
if $X$ is completely regular (e.g., metrizable).
Indeed, suppose $X$ is completely regular.
If $a,b\in X$ and $a\neq b$,
there is a function~$f$ in $\BCont{X}$
with $f(a)\neq f(b)$,
then $(\beta(a))(f)\neq (\beta(b))(f)$,
so $\beta(a)\neq\beta(b)$.
Thus,
$\beta$ is injective,
hence a bijection $X\rightarrow \beta(X)$.
If~$(x_i)_i$ is a net in~$X$
and $a\in X$,
and if $\beta(x_i)\rightarrow \beta(a)$,
then $f(x_i)\rightarrow f(a)$ for all~$f$ in $\BCont{X}$;
Thus by lemma~\ref{2.18},
we obtain $x_i\rightarrow a$.
Thus,
$\beta^{-1}$ is continuous.
\end{enumerate}
\end{psec}
%
% 4.13
%
\begin{psec}{4.13}{Exercise}
One might expect the Stone--\v{C}ech compactification of
$(0,1]$ to be (homeomorphic to) $[0,1]$.
Prove that this is false
by showing that the sequence $(\beta(n^{-1}))_n$ in $\beta(0,1]$
has no converging subsequence.
\end{psec}
%
% 4.14
%
\begin{psec}{4.14}{Exercise}
\begin{enumerate}
\item \label{4.14-1}
Let $X,Y$ be topological spaces,
$\tau$ a continuous map $X\rightarrow Y$.
Prove that the following formula
\begin{equation*}
\beta\tau\colon\varphi\mapsto\varphi\circ\tau\qquad(\varphi\in\BCont{Y}\,)
\end{equation*}
defines a continuous map $\beta\tau\colon\beta X\ra\beta Y$,
satisfying $(\beta\tau)\circ\beta=\beta\circ\tau$:
\begin{equation*}
\xymatrix{
X\ar[r]^\tau \ar[d]_\beta& Y\ar[d]_\beta \\
\beta X\ar[r]^{\beta\tau} & \beta Y
}
\end{equation*}
%
\item \label{4.14-2}
Let~$X$ be a topological space,
$Y$ a compact Hausdorff space.
Show that every continuous $\tau\colon X\ra Y$
induces a continuous $\hat\tau\colon\beta X\ra Y$
with $\hat\tau\circ\beta=\tau$.
\end{enumerate}
\end{psec}
%
% 4.15
%
\begin{psec}{4.15}{Exercise}
$\ell^\infty$ is $\BCont{\N}$,
so the space~$\Phi$ considered in~\ref{3.19} is~$\beta\N$
(on which a book has been written).
More about~$\beta\N$:
\begin{enumerate}
\item \label{4.15-1}
Show that for every $n\in\N$
the singleton set $\{\beta(n)\}$ is open in~$\beta \N$.
Thus~$\beta(\N)$ is a cloud of isolated points,
dense in~$\beta\N$.
It turns out that~$\beta\N$ is very large:
its cardinality is
that of the power set of the power set of~$\N$.
As a modest step in this direction
\emph{we make an injection} $\R\ra\beta\N$:
\item \label{4.15-2}
Show: For every $A\subseteq \N$,
$\hat{\mathbb 1}_A$ is~${\mathbb 1}_{\hat A}$
for some compact open subset~$\hat A$ of~$\beta\N$.
(This was also in~\ref{3.19}\ref{3.19-1}.) We have
\begin{equation*}
\hat A \subseteq \beta(\N)
\quad\htam{ if and only if }\quad
A\htam{ is finite.}
\end{equation*}
%
\item \label{4.14-3}
For each $\sr\in \R$,
choose a sequence $(\sr_n)_n$ in $\mathbb{Q}$
that converges to~$\sr$,
such that the set $S_\sr:=\{\sr_1,\sr_2,\dotsc\}$
is infinite.
Then $S_\sr \cap S_s$ is finite whenever $\sr,s\in \R$, $\sr\neq s$.
Via a bijection $\mathbb{Q}\leftrightarrow \N$ we see:
there is a family $(A_\sr)_{\sr\in\R}$
of infinite subsets of~$\N$
such that $A_\sr\cap A_s$ is finite if $\sr\neq s$.
Now prove:
The sets ${\hat A}_\sr \backslash \beta(\N)$,
where~$\sr$ runs through $\R$,
are nonempty and pairwise disjoint.
\end{enumerate}
\end{psec}
%
% 4.16
%
\noindent We return to the general theory
for a discussion of a property that,
for unitary spaces,
implies completeness.
\begin{psec}{4.16}{Definitions}
A Riesz space is called \keyword{Dedekind complete}
if every nonempty subset that is bounded from above
has a supremum;
it is \keyword{$\sigma$-Dedekind complete}
if every countable nonempty subset
that is bounded from above
has a supremum.
\end{psec}
%
% 4.17
%
\begin{psec}{4.17}{Comments}
\begin{enumerate}
\item \label{4.17-1}
Of course,
a Riesz space is [$\sigma$-]Dedekind complete
if and only if every [countable] nonempty subset
that is bounded from below has an infimum.
%
\item \label{4.17-2}
There are various alternative ways to describe
[$\sigma$-]Dedekind completeness.
One of them:
Call a subset~$W$ of~$E$ \keyword{upwards directed}
if for all $w_1,w_2\in W$
there exists a $w\in W$ with $w_1\leq w$, $w_2\leq w$.
Claim:
\statement{$E$ is already [$\sigma$-]Dedekind complete
if every [countable] nonempty subset of~$E^+$ that is upwards
directed and bounded from above has a supremum.}
Indeed,
suppose the latter condition is satisfied.
Let $\varnothing\neq W\subseteq E$
and let~$W$ be bounded from above;
for the $\sigma$-Dedekind completeness,
assume~$W$ is countable.
Choose~$w_0$ in~$W$.
The set
\begin{equation*}
W_0 :=\bigl\{w_0\vee w_1\vee \dotsb \vee w_N\colon
N\in \N,\ w_1,\dotsc,w_N\in W\bigr\}
\end{equation*}
is upwards directed and has the same upper bounds~$W$ has.
By the given property of~$E$,
the set $W_0 - w_0$ has a supremum.
Then $\sup(W_0-w_0)+w_0$ is the supremum of~$W_0$ and of $W$.
\end{enumerate}
\end{psec}
%
% 4.18
%
\begin{psec}{4.18}{Theorem} \statement{
\begin{enumerate}
\item \label{4.18-1}
Dedekind complete spaces are $\sigma$-Dedekind complete.
%
\item \label{4.18-2}
$\sigma$-Dedekind complete spaces are Archimedean.
%
\item \label{4.18-3}
Unitary $\sigma$-Dedekind complete spaces are uniformly complete.
\end{enumerate}
} \end{psec}
\begin{proof}
\begin{enumerate}
\item is trivial.
\item Let~$E$ be a $\sigma$-Dedekind complete Riesz space.
Let $a,b\in E^+$ be such that $na\leq b$ for all~$n$ in~$\N$;
we prove $a=0$. (See~\ref{3.2}.)
The set $\{na\colon n\in\N\}$ has a supremum, $c$, say.
For all~$n$ in~$\N$ we have $(n+1)a\leq c$,
so that $na\leq c-a$.
Then $c-a$ is an upper bound for $\{na\colon n\in \N\}$,
$c-a\geq c$, $a\leq 0$, and $a=0$.
%
\item We turn the argument of~\ref{4.5} inside out
and use Yosida's Theorem to prove uniform completeness.
Let~$E$ be a unitary $\sigma$-Dedekind complete Riesz space;
we show that it is uniformly complete.
By Yosida's Theorem we may assume
that~$E$ is a dense Riesz subspace of some $\Cont{X}$,
with $\mathbb{1}_X\in E$,
and by~\ref{4.6}\ref{4.6-2}
it suffices to prove $E=\Cont{X}$.
Take $f\in\Cont{X}$.
For each $n\in\N$ there is an~$f_n$ in~$E$
with $\|f-f_n\|_\infty \leq n^{-1}$, i.e.,
\begin{equation*}
f_n-n^{-1}\mathbb{1}\ \leq\ f\ \leq\ f_n+n^{-1}\mathbb{1}\htam{.}
\end{equation*}
The set $\{f_n - n^{-1}\mathbb{1}\colon n\in\N\}$ has a supremum,
$u$, in~$E$.
The set $\{f_n+n^{-1} \mathbb{1}\colon n\in \N\}$ has an infimum,
$v$, in~$E$.
For all~$n$ we have
\begin{equation*}
f_n-n^{-1}\mathbb{1}\ \leq\ u\ \leq\ v\ \leq\ f_n+n^{-1}\mathbb{1}\htam{.}
\end{equation*}
Then $f=u=v\in E$. \xqed
\end{enumerate}
\end{proof}
%
% 4.19
%
\begin{psec}{4.19}{Examples}
\begin{enumerate}
\item \label{4.19-1}
It is perfectly clear that~$\R^X$ and $\ell^\infty(X)$
are Dedekind complete.
\item \label{4.19-2}
Riesz ideals of a [$\sigma$-]Dedekind complete Riesz spaces
are [$\sigma$-]Dedekind complete.
\item \label{4.19-3}
Thus, the spaces~$\ell^1$ and~$\cseq_0$ (\ref{1.5}\ref{1.5-5})
are Dedekind complete.
\end{enumerate}
\end{psec}
%
% 4.20
%
\begin{psec}{4.20}{Exercise}
\begin{enumerate}
\item \label{4.20-1}
Let $\cseq$ be the Riesz space of all converging number sequences.
(See~\ref{1.5}\ref{1.5-5}.)
Show that~$\cseq$ is unitary and uniformly complete
but not $\sigma$-Dedekind complete.
(It may be convenient to view~$\cseq$ as the space of all Cauchy sequences.)
%
\item \label{4.20-2}
Let $E$ be the Riesz space of all bounded functions~$f$ on~$\R$
with the property:
\begin{alignat*}{2}
&\htam{There exists a countable set }A\subset\R \\
&\htam{such that }f\htam{ is constant on }\R\backslash A\htam{.}
\end{alignat*}
Show that~$E$ is $\sigma$-Dedekind complete but not Dedekind complete.
\end{enumerate}
\end{psec}
%
% 4.21
%
We return to Dedekind completeness in Chapter~6.
% chapter 6 does not exist yet
For the moment,
by way of nontrivial examples,
we prove the following:
\begin{psec}{4.21}{Theorem}\statement{
$\Mod(\Lambda)$ (\ref{1.10}),
$\BA(\mathcal{A})$ (\ref{1.12})
and $\sigma(\mathcal{A})$ (\ref{1.14})
are Dedekind complete.
}\end{psec}
\begin{proof}
$\BA(\mathcal A)$ is a special case of $\Mod(\Lambda)$,
and $\sigma(\mathcal A)$ is an ideal in $\BA(\mathcal A)$,
so we only consider $\Mod(\Lambda)$:
Let~$\Lambda$ be a distributive lattice with a smallest element, $0$.
If $\varphi,\psi\in\Mod(\Lambda)$ and $\varphi\leq\psi$,
then $\psi-\varphi$ is increasing,
so that for all $s\in\Lambda$,
$(\psi-\varphi)(s)\geq(\psi-\varphi)(0)=0$.
Thus:
\begin{equation}
\label{eq4.21} \tag{$*$}
\begin{alignedat}{2}
&\htam{For }s\in\Lambda\htam{ the function }\varphi \mapsto \varphi(s) \\
&\htam{on }\Mod(\Lambda)\htam{ is increasing.}
\end{alignedat}
\end{equation}
Let $\Omega$ be a nonempty subset of $\Mod(\Lambda)^+$,
upwards directed and bounded from above.
Invoking~\ref{4.17}\ref{4.17-2},
we see that we are done if we prove that~$\Omega$ has a supremum.
Every $s\in\Lambda$ induces
a net $(\omega(s))_{\omega\in\Omega}$ in~$\R$ that
by~\eqref{eq4.21} is increasing and bounded from above,
hence converging.
Define $\omega_0\colon\Lambda\ra\R$ by
\begin{equation*}
\omega_0(s):=\lim_{\omega\in\Omega}\omega(s)
=\sup_{\omega\in\Omega}(\omega(s))\qquad(s\in \Lambda)\htam{.}
\end{equation*}
From this, it is clear that $\omega_0(0)=0$
and that~$\omega_0$ is modular (\ref{1.8}\ref{1.8-2}) and increasing;
hence $\omega_0 \in\Mod(\Lambda)^+$.
Take~$\varphi$ in~$\Omega$.
Then $\Omega_\varphi:=\{\omega\in \Omega\colon \omega\geq \varphi\}$
is cofinal in~$\Omega$, so
\begin{equation*}
\omega_0(s)=\lim_{\omega\in\Omega_\varphi} \omega(s)
\qquad (s\in\Lambda)\htam{.}
\end{equation*}
By the definition of the ordering in $\Mod(\Lambda)$,
for every~$\omega$ in~$\Omega_\varphi$
the function $\omega-\varphi$ is increasing:
then so is $\omega_0-\varphi$.
Thus, $\omega_0$ is an upper bound for~$\Omega$ in $\Mod(\Lambda)$.
Finally,
let $\omega'$ be any upper bound for~$\Omega$.
Then for all~$\omega$ in~$\Omega$ we have
that the function $\omega'-\omega$ is increasing.
From the definition of~$\omega_0$ it follows that $\omega'-\omega_0$
is increasing, i.e., $\omega'\geq\omega_0$ in $\Mod(\Lambda)$.
Thus, $\omega_0=\sup\Omega$. \xqed
\end{proof}
\clearpage
\emptypage
\end{document}