-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathfunctional_analysis.tex
1832 lines (1569 loc) · 104 KB
/
functional_analysis.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{article}
% preamble
\def\npart{III}
\def\nyear{2023}
\def\nterm{Michaelmas}
\def\nlecturer{Dr Andr\'as Zs\'ak}
\def\ncourse{Functional Analysis}
\def\draft{Incomplete}
\input{header}
\setcounter{section}{-1}
% and here we go!
\begin{document}
\maketitle
\tableofcontents
\clearpage
\section{Introduction}
\subsection*{Prerequisites}
\begin{itemize}
\item some basic functional analysis
\item a bit of measure theory
\item a bit of complex analysis
\end{itemize}
\subsection*{Books}
Books relevant to the course are:
\begin{itemize}
\item Bollob\'as, \textit{Linear Analysis}
\item Murphy, \textit{$C^*$-algebras}
\item Rudin
\item Graham-Allan
\end{itemize}
\subsection*{Notation}
We will use $\K$ to mean "either $\R$ or $\C$".
For $X$ a normed space, we define
\begin{align*}
B_X & = \{x \in X \mid \norm x \le 1\} \\
S_X & = \{x \in X \mid \norm x = 1\} \\
D_X & = \{x \in X \mid \norm x < 1\}
\end{align*}
For $X, Y$ normed spaces, we write $X \sim Y$ if $X, Y$ are isomorphic, ie there
exists a linear bijection $T : X \to Y$ such that $T$ and $T^{-1}$ are continuous. We write $X \cong Y$ if $X, Y$ are isometrically isomorphic, ie there exists a surjective linear map $T : X \to Y$ such that $\norm{T x} = \norm x$ for all $x$.
\clearpage
\section{Hahn-Banach extension theorems}
\newlec
Let $X$ be a normed space. The {\bf dual space} of $X$ is the space $X^*$ of
bounded linear functionals on $X$.
$X^*$ is always a Banach space in the operator norm: for $f \in X^*$,
$$\norm f = \sup_{x \in B_X} |f(x)|$$
\begin{egs*}~
\begin{itemize}
\item For $1 < p, q < \infty$, $p^{-1} + q^{-1} = 1$, $\ell_p^* \cong \ell_q$
\item $\ell_1^* \cong \ell_\infty$, $c_0^* \cong \ell_1$
\item If $H$ is a Hilbert space, then $H^* \cong H$ (the isomorphism is conjugate-linear in the complex case).
\end{itemize}
\end{egs*}
For $x \in X, f \in X^*$, we write $\inn x f = f(x)$. Note that
$$\inn x f = |f(x)| \le \norm f\norm x$$
\begin{dfn*}
Let $X$ be a {\it real} vector space. A functional $p : X \to \R$ is
\begin{itemize}
\item {\bf positive homogeneous} if $p(tx) = tp(x)$ for all $x \in X$, $t \ge 0$
\item {\bf subadditive} if $p(x + y) \le p(x) + p(y)$ for all $x, y \in X$
\end{itemize}
\end{dfn*}
\begin{dfn*}
Let $P$ be a preorder, $A \subseteq P, x \in P$. We say
\begin{itemize}
\item $x$ is an {\bf upper bound} for $A$ if $\for a \in A, a \le x$.
\item $A$ is a {\bf chain} if $\for a, b \in A, a \le b \lor b \le a$.
\item $x$ is a {\bf maximal element} if $\for y \in P, x \not < y$
\end{itemize}
\end{dfn*}
\begin{fact}[Zorn's lemma]
A nonempty preorder in which all nonempty chains have an upper bound has a maximal element.
\end{fact}
\begin{thm}[Hahn-Banach, positive homogeneous version]\label{thm:hb-positive}
Let $X$ be a real vector space and $p : X \to \R$ be positive homogeneous and subadditive. Let $Y$ be a subspace of $X$ and $g : Y \to \R$ be linear such that $\for y \in Y, g(y) \le p(y)$. Then there exists $f : X \to \R$ linear such that $f\restriction_Y = g$ and $\for x \in X, f(x) \le p(x)$.
\end{thm}
\begin{proof}
Let $P$ be the set of pairs $(Z, h)$ where $Z$ is a subspace of $X$ with $Y \subseteq Z$ and $h : Z \to \R$ linear, $h\restriction_Y = g$ and $\for z \in Z, h(z) \le p(z)$. $P$ is nonempty since $(Y, g) \in P$, and is partially ordered by
$$(Z_1, h_1) \le (Z_2, h_2) \iff Z_1 \subseteq Z_2 \land h_2\restriction_{Z_1} = h_1$$
If $\{(Z_i, h_i) \mid i \in I\}$ is a chain with $I$ nonempty, then we can define
$$Z := \Union_{i \in I} Z_i, h\restriction_{Z_i} = h_i$$
The definition of $h$ makes sense thanks to the chain assumption. $(Z, h) \in P$ is therefore an upper bound for the chain. \\
Hence find by Zorn a maximal element $(Z, h)$ of $P$. If $Z = X$, we won. So assume there is some $x \in X \ Z$. Let $W = \Span(Z \cup \{x\})$ and define $f : W \to \R$ by
$$f(z + \lambda x) = h(z) + \lambda\alpha$$
for some $\alpha \in \R$. Then $f$ is linear and $f\restriction_Z = h$. We now look for $\alpha$ such that $\for w \in W, f(w) \le p(w)$. We would then have $(W, f) \in P$ and $(Z, h) < (W, f)$, contradicting maximality of $(Z, h)$. \\
We need
$$h(z) + \lambda\alpha \le p(z + \lambda x) \for z \in Z, \lambda \in \R$$
Since $p$ is positive homogeneous, this becomes
\begin{align}
h(z) + \alpha \le p(z + x)
h(z) - \alpha \le p(z - x)
\end{align}
ie
$$h(z) - p(z - x) \le \alpha \le p(z + x) - h(z) \for z \in Z$$
The existence of $\alpha$ now amounts to
$$h(z_1) - p(z_1 - x) \le \alpha \le p(z_2 + x) - h(z_2) \for z_1, z_2 \in Z$$
But indeed
$$h(z_1) + h(z_2) = h(z_1 + z_2) \le p(z_1 + z_2) \le p(z_1 - x) + p(z_2 + x)$$
\end{proof}
\begin{dfn*}
Let $X$ be a $\K$-vector space. A {\bf seminorm} on $X$ is a functional $p : X \to \R$ such that
\begin{itemize}
\item $\for x \in X, p(x) \ge 0$
\item $\for x \in X, \lambda \in \K, p(\lambda x) = |\lambda| p(x)$
\item $\for x, y \in X, p(x + y) \le p(x) + p(y)$
\end{itemize}
\end{dfn*}
\begin{rmk}
$$\text{norm} \imp \text{seminorm} \imp \text{positive homogeneous}$$
\end{rmk}
\newlec
\begin{thm}[Hahn-Banach, absolute homogeneous version]\label{thm:hb-absolute}
Let $X$ be a real of complex vector space and $p$ a seminorm on $X$. Let $Y$ be a subspace of $X$, $g$ a linear functional on $Y$ such that $\for y \in Y, |g(y)| \le p(y)$. Then there exists a linear functional $f$ on $X$ such that $f\restriction_Y = g$ and $\for x \in X, |f(x)| \le p(x)$.
\end{thm}
\begin{proof}~\\
{\bf Real case}
$$\for y \in Y, g(y) \le |g(y)| \le p(y)$$
By Theorem \ref{thm:hb-positive}, there exists $f : X \to \R$ such that $f\restriction_Y = g$ and $\for x \in X, f(x) \le p(x)$. We also have
$$\for x \in X, -f(x) = f(-x) \le p(-x) = p(x)$$
Hence $|f(x)| \le p(x)$ \\
{\bf Complex case} \\
$\Re g : Y \to \R$ is real-linear.
$$\for y \in Y, |\Re g(y)| \le |g(y)| \le p(y)$$
By the real case, find $h : X \to \R$ real-linear such that $h\restriction_Y = \Re g$
\begin{claim}
There exists a unique complex-linear $f : X \to \C$ such that $h = \Re f$.
\end{claim}
\begin{proof}~\\
{\bf Uniqueness} \\
If we have such $f$, then
\begin{align*}
f(x)
& = \Re f(x) + i\Im f(x) \\
& = \Re f(x) - i\Re f(ix) \\
& = h(x) - ih(ix)
\end{align*}
{\bf Existence} \\
Define $f(x) = h(x) - ih(ix)$. Then $f$ is real-linear and $f(ix) = if(x)$, so $f$ is complex-linear with $\Re f = h$.
\end{proof}
We now have $f : X \to \C$ such that $\Re f = h$.
$$\Re f\restriction_Y = h\restriction_Y = \Re g$$
So, by uniqueness, $f\restriction_Y = g$. \\
Given $x \in X$, find $\lambda$ with $|\lambda| = 1$ such that
\begin{align*}
|f(x)|
& = \lambda f(x) \\
& = f(\lambda x) \\
& = \Re f(\lambda x) \\
& = h(\lambda x) \\
& \le p(\lambda x) \\
& = p(x)
\end{align*}
\end{proof}
\begin{rmk}
For a complex vector space $X$, if we write $X_\R$ for $X$ considered as a real vector space, the above proof shows that
$$\Re : (X^*)_\R \to X_\R^*$$
is an isometric isomorphism.
\end{rmk}
\begin{cor}\label{cor:hb-point}
Let $X$ be a $\K$-vector space, $p$ a seminorm on $X$, $x_0 \in X$. Then there exists a linear functional $f$ on $X$ such that $f(x_0) = p(x_0)$ and $\for x \in X, |f(x)| \le p(x)$.
\end{cor}
\begin{proof}
Let $Y = \Span(x_0)$,
\begin{align*}
g : Y & \to \K \\
\lambda x_0 & \mapsto \lambda p(x_0)
\end{align*}
We see that $\for y \in Y, g(y) \le p(y)$. Hence find by Theorem \ref{thm:hb-absolute} a linear functional $f$ on $X$ such that $f\restriction_Y = g$ and $\for x \in X, |f(x)| \le p(x)$. We check that $f(x_0) = g(x_0) = p(x_0)$.
\end{proof}
\begin{thm}[Hahn-Banach, existence of support functionals]\label{thm:hb-support}
Let $X$ be a real or complex normed space. Then
\begin{enumerate}
\item If $Y$ is a subspace of $X$ and $g \in Y^*$, then there exists $f \in X^*$ such that $f\restriction_Y = g$ and $\norm f = \norm g$.
\item Given $x_0 \ne 0$, there exists $f \in S_{X^*}$ such that $f(x_0) = \norm{x_0}$.
\end{enumerate}
\end{thm}
\begin{proof}~
\begin{enumerate}
\item Let $p(x) = \norm g \norm x$. Then $p$ is a seminorm on $X$ and
$$\for y \in Y, |g(y)| \le \norm g \norm y = p(y)$$
Find by Theorem \ref{thm:hb-positive} a linear functional $f$ on $X$ such that $f\restriction_Y = g$ and $\for x \in X, |f(x)| \le p(x) = \norm g \norm x$. So $\norm f \le \norm g$. Since $f\restriction_Y = g$, we also have $\norm g \le \norm f$. Hence $\norm f = \norm g$.
\item Apply Corollary \ref{cor:hb-point} with $p(x) = \norm x$ to get $f \in X^*$ such that
$$\for x \in X, |f(x)| \le \norm x \text{ and } f(x_0) = \norm{x_0}$$
It follows that $\norm f = 1$.
\end{enumerate}
\end{proof}
\begin{rmks}~
\begin{itemize}
\item Part 1 is a sort of linear version of Tietze's extension theorem: Given $K$ compact Hausdorff, $L \subseteq K$ closed, $g : L \to \K$ continuous, there exists $f : K \to \K$ such that $f\restriction_L = g$ and $\norm f_\infty = \norm g_\infty$.
\item Part 2 shows that for all $x \ne y$ in $X$ there exists $f \in X^*$ such that $f(x) \ne f(y)$, namely $X^*$ {\bf separates points} of $X$. This is a sort of linear version of Urysohn: $C(K)$ separates points of $K$.
\item The $f$ in part 2 is called a {\bf norming functional}, aka {\bf support functional}, for $x_0$. The existence of support functionals shows that
$$x_0 = \max_{g \in B_{X^*}} \inn{x_0}g$$
Assuming $X$ is a real normed space and $\norm{x_0} = 1$, we have $B_X \subseteq \{x \in X| f(x) \le 1\}$. Visually, TODO: insert tangency diagram
\end{itemize}
\end{rmks}
\subsection{Bidual}
Let $X$ be a normed space. Then $X^{**}$ is called the {\bf bidual} or {\bf second dual} of $X$.
For $x \in X$, define $\hat x : X^* \to \K$, the {\bf evaluation at $x$}, by $\hat x(f) = f(x)$. $\hat x$ is linear and $|\hat x(f)| = |f(x)| \le \norm f \norm x$, so $\hat x \in X^{**}$ and $\norm{\hat x} \le \norm x$.
The map $x \mapsto \hat x : X \to X^{**}$ is called the {\bf canonical embedding} of $X$ into $X^{**}$.
\begin{thm}\label{thm:can-emb}
The canonical embedding is an isometric embedding.
\end{thm}
\begin{proof}~\\
{\bf Linearity}
\begin{align*}
\widehat{x + y}(f) & = f(x + y) = f(x) + f(y) = \hat x(f) + \hat y(f) \\
\widehat{\lambda x}(f) & = f(\lambda x) = \lambda f(x) = \lambda \hat x(f)
\end{align*}
{\bf Isometry} \\
If $x \ne 0$, there exists a support functional $f$ for $x$. Then
$$ \norm{\hat x} \ge |\hat x(f)| = |f(x)| = \norm x$$
\end{proof}
\begin{rmks}~
\begin{itemize}
\item In bracket notation, $\inn f {\hat x} = \inn x f$
\item Let $\hat X$ be the image of $X$ in $X^{**}$. Theorem \ref{thm:can-emb} says
$$X \cong \hat X \subseteq X^{**}$$
We often identify $\hat X$ with $X$ and think of $X$ as living isometrically inside $X^{**}$. Note that
$$X \text{ complete } \iff \hat X \text{ closed in } X^{**}$$
\item More generally, $\bar{\hat X}$ is a Banach space containing an isometric copy of $X$ as a dense subspace. We proved that normed spaces have completions!
\end{itemize}
\end{rmks}
\begin{dfn*}
A normed space $X$ is {\bf reflexive} if the canonical embedding $X \to X^{**}$ is surjective.
\end{dfn*}
\begin{egs*}~
\begin{itemize}
\item Some reflexive spaces are Hilbert spaces, finite-dimensional spaces, $\ell_p$ and $L_p(\mu)$ for $1 < p < \infty$.
\item Some non-reflexive spaces are $c_0, \ell_1, \ell_\infty, L_1[0, 1]$.
\end{itemize}
\end{egs*}
\begin{rmks}~
\begin{itemize}
\item If $X$ is reflexive, then $X \cong X^{**}$, so $X$ is complete.
\item There are Banach spaces $X$ such that $X \cong X^{**}$ but $X$ is not reflexive, eg {\bf James' space}. Any isomorphism to the bidual is then necessarily not the canonical embedding.
\end{itemize}
\end{rmks}
\subsection{Dual operators}
\newlec
Let $X, Y$ be normed spaces. Recall
$$\mcB(X, Y) = \{T : X \to Y \mid T \text{ linear, bounded}\}$$
This is a normed space in the operator norm:
$$\norm T = \sup_{x \in B_X} \norm{Tx}$$
If $Y$ is complete, then so is $\mcB(X, Y)$. For $T \in \mcB(X, Y)$, the {\bf dual operator} of $T$ is the map $T^* : Y^* \to X^*$ given by $T^*g = g \comp T$. In bracket notation $\inn x{T^*g} = \inn{Tx} g$ for $x \in X, g \in Y^*$.
{\bf $T^*$ is linear}
\begin{align*}
\inn x {T^*(g + h)}
& = \inn{Tx}{g + h} \\
& = \inn{Tx}g + \inn{Tx}h \\
& = \inn x{T^*g} + \and x{T^* h} \\
& = \inn x{T^*g + T^*h}
\end{align*}
\begin{align*}
\inn x {T^*(\lambda g)}
& = \inn{Tx}{\lambda g} \\
& = \lambda \inn{Tx}g \\
& = \lambda \inn x{T^*g} \\
& = \inn x{\lambda T^*g}
\end{align*}
{\bf $T^*$ is bounded}
\begin{align*}
\norm{T^*}
& = \sup_{g \in B_{Y^*}} \norm{T^*g} \\
& = \sup_{g \in B_{Y^*}} \sup_{x \in B_X} \abs{\inn x{T^*g}} \\
& = \sup_{x \in B_X} \sup_{g \in B_{Y^*}} \abs{\inn{Tx}g} \\
& = \sup_{x \in B_X} \norm{Tx} \text{ by Theorem \ref{thm:hb-support} (ii)} \\
& = \norm T
\end{align*}
\begin{rmks}~
\begin{itemize}
\item Hahn-Banach is crucial here. Without it, the dual could be $0$.
\item $\inn\cdot\cdot$ is linear in both arguments. This contrasts with the Hilbert space case where $\inn\cdot\cdot$ is conjugate-linear in one of the arguments. This comes from the conjugate-linearity of the identification $H^* \cong H$.
\item If $X, Y$ are Hilbert spaces and we identify $X, Y$ with $X^*, Y^*$, respectively, then $T^*$ is the adjoint of $T$.
\end{itemize}
\end{rmks}
\begin{eg*}
Let $1 < p, q < \infty$, $p^{-1} + q^{-1} = 1$ and define $R : \ell_p \to \ell_p$ to be the {\bf right shift operator} $(x_0, x_1, \dots) \mapsto (0, x_0, \dots)$. Then $R^* : \ell_q \to \ell_q$ is the {\bf left shift operator} $(x_0, x_1, \dots) \mapsto (x_1, x_2, \dots)$.
\end{eg*}
Some properties of the dual operator are
\begin{enumerate}
\item $\id_X^* = \id_{X^*}$
\item $(S + T)^* + S^* + T^*$, $(\lambda T)^* = \lambda T^*$
\item $(ST)^* = T^*S^*$
\item $T \mapsto T^* : \mcB(X, Y) \to \mcB(Y^*, X^*)$ is an {\it into} isomorphism.
\item The double dual of an operator commutes with the double dual embedding. \\
TODO: Insert commutative diagram
For all $x$,
$$\inn g{T^{**}\hat x} = \inn{T^*g}{\hat x} = \inn x{T^*g} = \inn{Tx}g = \inn g{\hat{Tx}}$$
So $T^{**}\hat x = \widehat{Tx}$.
\end{enumerate}
\begin{rmk}
From the above properties, if $X \sim Y$, then $X^* \sim Y^*$. Interestingly, if $X$ and $Y$ are reflexive, then we can deduce $X \sim Y$ from $X^* \sim Y^*$.
\end{rmk}
\subsection{Quotient spaces}
Let $X$ be a normed space and $Y$ be a {\it closed} subspace. Then the quotient space $X / Y$ becomes a normed space in the quotient norm:
$$\norm{x + Y} = d(x, Y) = \inf_{y \in Y} \norm{x + y}$$
The quotient map $q : X \to X/Y$ is linear and bounded: $\norm{q(x)} \le \norm x$, so $\norm q \le 1$.
$q$ maps the open unit ball $D_X$ onto $D_{X/Y}$. Indeed, if $x \in D_X$, then $\norm{q(x)} \le \norm x < 1$. Reciprocally, if $q(x) \in D_{X/Y}$, then there exists $y \in Y$ such that $\norm{x + y} < 1$. So $x + y \in D_X$ and $q(x + y) = q(x)$. It follows that $q$ is an open map and $\norm q = 1$.
If $Z$ is another normed space, $T \in \mcB(X, Z)$ and $Y \subseteq \ker T$, then there exists a unique map $\tilde T$ is linear and $\tilde T(D_{X/Y}) = \tilde T(q(D_X)) = T(D_X)$. It follows that $\norm{\tilde T} = \norm T$.
\begin{thm}
Let $X$ be a normed space. If $X^*$ is separable, then so is $X$.
\end{thm}
\begin{rmk}
The converse is false, as $X = \ell_1, X^* = \ell_\infty$ shows.
\end{rmk}
\begin{proof}
Since $X^*$ is separable, so is $S_{X^*}$. Let $f_n$ be a dense subset of $S_{X^*}$. For every $n$, find $x_n \in B_X$ such that $f_n(x_n) > \frac 12$. Let
$$Y = \bar{\Span\{x_n \mid n \in \N\}}$$
\begin{claim}
$Y = X$
\end{claim}
Then we're done since $Y$ is separable via $Y = \bar{\Span_\Q\{x_n \mid n \in \N\}}$.
\begin{proof}
Assume not. Then we can pick $g \in (X/Y)^*$, $\norm g = 1$ (by Theorem \ref{thm:hb-support} (ii)). Let $f = g \comp q$. Then $\norm f = \norm g = 1$, ie $f \in S_{X^*}$. Thus find $n$ such that $\norm{f - f_n} < \frac 14$, so that
$$\frac 14 > \norm{f - f_n}\norm{x_n} \ge \abs{(f - f_n)(x_n)} = \abs{f_n(x_n)} > \frac 12$$
contradiction.
\end{proof}
\end{proof}
\begin{thm}
Let $X$ be a separable normed space. Then $X$ embeds isometrically into $\ell_\infty$.
\end{thm}
\begin{proof}
Let $\{x_n \mid n \in \N\}$ be dense in $X$. For every $n$, find $f_n \in S_{X^*}$, $f_n(x_n) = \norm{x_n}$ (assuming $X \ne \{0\}$). Define $T : X \to \ell_\infty$ by $(Tx)_n = f_n(x)$.
{\bf Well definition}
$$ \abs{(Tx)_n} = \abs{f_n(x)} \le \norm{f_n}\norm x = \norm x$$
Hence $\norm{Tx}_\infty \le \norm x < \infty$.
{\bf Linearity}
$$ (T(x + y))_n = f_n(x + y) = f_n(x) + f_n(y) = (Tx + Ty)_n$$
$$ (T(\lambda x))_n = f_n(\lambda x) = \lambda f_n(x)) = (\lambda Tx)_n$$
so $T(x + y) = Tx + Ty, T(\lambda x) = \lambda Tx$.
{\bf Isometry} \\
We already know $\norm{Tx}_\infty \le \norm x$. On the other hand, find $f$ a supporting functional for $x$ and $f_{n_k}$ a subsequence converging to $f$. Then
$$\norm{Tx}_\infty \ge \sup_k (Tx)_{n_k} = \sup_k \abs{f_{n_k}(x)} \ge \abs{f(x)} = \norm x$$
\end{proof}
\newlec
\begin{rmks}~
\begin{itemize}
\item The result says that $\ell_\infty$ is isometrically universal for the class $\mathcal{SB}$ of separable Banach spaces.
\item There is a dual result: Every separable Banach space is a quotient of $\ell_1$.
\end{itemize}
\end{rmks}
\begin{thm}[Vector-valued Liouville]\label{thm:vector-liouville}
Lex $X$ be a complex Banach space, $f : \C \to X$ holomorphic and bounded. Then $f$ is constant.
\end{thm}
\begin{proof}
Find $M \ge 0$ such that $\for z \in \C, \abs{f(z)} \le M$. Fix $\phi \in X^*$. $\phi \comp f : \C \to \C$ is
{\bf bounded}
$$\abs{\phi(f(z))} \le \norm\phi \norm{f(z)} \le M\norm\phi$$
{\bf holomorphic}
$$\frac{\phi(f(z)) - \phi(f(w))}{z - w} = \phi\left(\frac{f(z) - f(w)}{z - w}\right) \to \phi(f'(z))$$
By scalar Liouville, $\phi \comp f$ is constant. For every $z \in \C, \phi \in X^*$, $\phi(f(z)) = \phi(f(0))$. Since $X^*$ separates points of $X$, $f(z) = f(0)$.
\end{proof}
\begin{rmk}
This is a typical example of how to transfer a scalar result to a vector-valued one: Prove the result once composed with any functional, then go back using Hahn-Banach.
\end{rmk}
\subsection{Locally convex spaces}
\begin{dfn*}
A {\bf locally convex space} is a $\K$-vector space such that there exists a family $\mcP$ of seminorms on $X$ that separate points of $X$ in the sense that for all $x \ne 0$ there exists $p \in \mcP$ such that $p(x) \ne 0$.
\end{dfn*}
The family $\mcP$ defines a topology on $X$:
$$U \subseteq X \text{ open} \iff \for x \in U, \exists s \subseteq \mcP \text{ finite}, \eps > 0, \{y \in X \mid \for p \in s, p(x) < \eps\} \subseteq U$$
\begin{rmks}~
\begin{enumerate}
\item Addition and scalar multiplication are continuous.
\item The topology is Hausdorff as $\mcP$ separates points.
\item $x_n \to x \iff \for p \in \mcP, p(x_n - x) \to 0$
\item Let $Y$ be a subspace of $X$ and $\mcP_Y = \{p\restriction_Y \mid p \in \mcP\}$. Then $(Y, \mcP_Y)$ is a LCS and its topology is the subspace topology.
\item Let $\mcP, \mcQ$ be two families of seminorms on $X$ both separating points of $X$. We say $\mcP, \mcQ$ are {\bf equivalent}, write $\mcP \sim \mcQ$, if they induce the same topology on $X$. One interesting result is that
$$ (X, \mcP) \text{ metrisable} \iff \mcP \text{ equivalent to some countable family}$$
\item We make $\mcP$ part of the data here out of simplicity, but in grown up mathematics we instead assume that $X$ already comes with a topology and that this topology coincides with the one induced by $\mcP$.
\end{enumerate}
\end{rmks}
\begin{dfn*}
A {\bf Fréchet space} is a complete metrisable LCS.
\end{dfn*}
\begin{egs*}~
\begin{enumerate}
\item A normed space is a LCS with $\mcP = \{\norm\cdot\}$.
\item Let $U \subseteq \C$ nonempty open. Let $\mcO(U) = \{f : U \to \C \mid f \text{ holomorphic}\}$. For compact $K \subseteq U$, define $p_K(f) = \sup_{z \in K} \abs{f(z)}$. Let
$\mcP = \{p_K \mid K \subseteq U \text{ compact}\}$
Then $(\mcO(U), \mcP)$ is a LCS. If we replace $\{K \subseteq U \text{ compact}\}$ by a compact exhaustion of $U$, then we get a countable separating family equivalent to $\mcP$. So $(\mcO(U), \mcP)$ is metrisable. However it is not normable: no norm on $\mcO(U)$ induces the topology of $(\mcO(U), \mcP)$, which is the topology of uniform convergence. This is a consequence of Montel's theorem.
\item Fix $d \in \N, \Omega \subseteq \R^d$ a nonempty open set. Let
$$C^\infty(\Omega) = \{f : \Omega \to \R \mid f \text{ infinitely differentiable}\}$$
Given a multi-index $\alpha \in \Z^d$, $\alpha$ defines a differential operator
$$D^\alpha = \left(\frac\partial{\partial x_1}\right)^{\alpha_1} \dots
\left(\frac\partial{\partial x_d}\right)^{\alpha_d}$$
For a compact $K \subseteq \Omega, \alpha \in \Z^d$, define
$$p_{K, \alpha}(f) = \sup_{z \in K}\abs{D^\alpha f(z)}$$
Let
$$\mcP = \{p_{K, \alpha} \mid K \subseteq U \text{ compact}, \alpha \in \Z^d\}$$
Then $(C^\infty, \mcP)$ is a LCS. It is in fact a non-normable Fréchet space.
\end{enumerate}
\end{egs*}
\begin{lem}\label{lem:lcs-continuity}
Let $(X, \mcP), (Y, \mcQ)$ be LCS, $T : X \to Y$ linear. TFAE
\begin{enumerate}
\item $T$ is continuous
\item $T$ is continuous at $0$
\item $\for q \in \mcQ, \exists s \subseteq \mcP \text{ finite}, C \ge 0, \for x \in X, q(Tx) \le C\max_{p \in s} p(x)$
\end{enumerate}
\end{lem}
\begin{proof}~\\
{\bf (i) $\iff$ (ii)} \\
Translation is continuous.
{\bf (ii) $\imp$ (iii)} \\
Given $q \in \mcQ$, let $V = \{y \in Y \mid q(y) \le 1\}$. Then $V$ is a neighborhood of $0$ in $Y$. So there exists $U$ neighborhood of $0$ in $X$ such that $T(U) \subseteq V$. WLOG
$$U = \{x \in X \mid \for p_K \in s, p_K(x) \le \eps\}$$
Let $p = \max_{p_K \in s} p_K(x)$. If $p(x) = 1$, then $p(\eps x) = \eps$, so $\eps x \in U$ and
$$q(T(\eps x)) \le 1$$
By homogeneity, $q(Tx) \le \frac 1\eps p(x)$ for all $x$ such that $p(x) > 0$. If $p(x) = 0$, then $p(\lambda x) = 0$ for all scalar $\lambda$. So $q(T(\lambda x)) \le 1$ for all $\lambda$. Hence $q(Tx) = 0 \le \frac 1\eps p(x)$.
{\bf (iii) $\imp$ (ii)} \\
Assume $t \subseteq \mcQ$ is finite, $\eps > 0$, and let $V = \{y \in Y \mid \for q \in t, q(y) \le \eps$ the corresponding neighborhood of $0$. For each $q \in t$, find $s_q \subseteq \mcP$ finite and $C_q$ so that $\for x \in X, q(Tx) \le C_q\max_{p \in s_q}p(x)$. Let
$$U = \left\{x \in X \mid \for q \in \mcQ, p \in s_q, p(x) \le \frac\eps{C_q}\right\}$$
Then $U$ is a neighborhood of $0$ and $T(U) \subseteq V$.
\end{proof}
\begin{dfn*}
Let $(X, \mcP)$ be a LCS. The {\bf dual space} of $X$ is the space of continuous linear functionals $X \to \K$.
\end{dfn*}
\newlec
\begin{lem}\label{lem:lcs-dual}
Let $f$ be a linear functional on a LCS $(X, \mcP)$. Then
$$f \in X^* \iff \ker f \text{ closed}$$
\end{lem}
\begin{proof}~\\
$\imp$ \\
$\ker f = f^{-1}(0)$ is closed since $f$ is continuous.
$\impliedby$ \\
If $\ker f = 0$, then $f = 0$ is continuous. Else fix some $x_0 \nin \ker f$. Since $(\ker f)^c$ is open, find $s \subseteq \mcP$ finite, $\eps > 0$ such that
$$\underbrace{\{x \in X \mid \for p \in s, p(x - x_0) < \eps\}}_U \subseteq (\ker f)^c$$
Then $U$ is a neighborhood of $0$ and $(x_0 + U) \cap \ker f = \empty$. Note that $U$ is convex and {\bf balanced} ($x \in U, \abs\lambda \le 1 \imp \lambda x \in U$), hence so is $f(U)$ as $f$ is linear. \\
If $f(U)$ is unbounded, then it is the whole scalar field, hence so is $f(x_0 + U) = f(x_0) + f(U)$. But $0 \in \ker f$, contradicting disjointness. \\
So find $M$ such that $\abs{f(x)} < M$ for all $x \in U$. For all $\delta > 0$, $\frac\delta M U$ is a neighborhood of $0$ and $f\left(\frac\delta M U\right) \subseteq \{\lambda \in \K \mid \abs\lambda < \delta\}$. Thus $f$ is continuous.
\end{proof}
\begin{thm}[Hahn-Banach]\label{thm:hb-lcs}
Let $(X, \mcP)$ be a LCS.
\begin{enumerate}
\item Given a subspace $Y$ of $X$ and $g \in Y^*$, there exists $f \in X^*$ such that $f\restriction_Y = g$.
\item Given a closed subspace $Y$ of $X$ and $x_0 \nin Y$, there exists $f \in X^*$ such that $f\restriction_Y = 0$, $f(x_0) \ne 0$.
\end{enumerate}
\end{thm}
\begin{rmk}
This means that $X^*$ separates points of $X$.
\end{rmk}
\begin{proof}~
\begin{enumerate}
\item By Lemma \ref{lem:lcs-continuity}, find $s \subseteq \mcP$ finite, $C \ge 0$ such that
$$\for y \in Y, \abs{g(y)} \le C\max_{p \in s}p(y)$$
Let $p(x) = C\max_{p \in s}p(x)$. Then $p$ is a seminorm on $X$ and $\for y \in Y, \abs{g(y)} \le p(y)$. By Theorem \ref{thm:hb-absolute}, find a linear functional $f$ on $X$ such that $f\restriction_Y = g, \for x \in X, \abs{f(x)} \le p(x)$. By Lemma \ref{lem:lcs-continuity}, $f \in X^*$.
\item Let $Z = \Span(Y \cup \{x_0\})$ and define a linear functional $g$ on $Z$ by
$$g(y + \lambda x_0) = \lambda, y \in Y, \lambda \in \K$$
Then $g\restriction_Y = 0, g(x_0) = 1 \ne 0$ and $\ker g = Y$ is closed, so $g \in Z^*$ by Lemma \ref{lem:lcs-dual}. By part (i), find $f \in X^*$ such that $f\restriction_Z = g$. This works.
\end{enumerate}
\end{proof}
\clearpage
\section{The dual of \texorpdfstring{$L_p(\mu)$ and $C(K)$}{Lp(mu) or C(K)}}
Let $(\Omega, \mcF, \mu)$ be measure space.
$1 \le p < \infty$
$$L_p(\mu) = \{f : \Omega \to \K \mid f \text{ measurable and } \int_\Omega \abs f^p d\mu < \infty\}$$
This is a normed space in the $L_p$-norm:
$$\norm f_p = \left(\int_\Omega \abs f^p d\mu\right)^{\frac 1p}$$
$p = \infty$ \\
A measurable function $f : \Omega \to \K$ is {\bf essentially bounded} if there exists $N \in \mcF$ such that $\mu(N) = 0$ and $f\restriction_{N^c}$ is bounded.
$$L_p(\mu) = \{f : \Omega \to \K \mid f \text{ measurable and essentially bounded}\}$$
This is a normed space in the $L_\infty$-norm:
$$\norm f_\infty = \esssup \abs f = \inf_{\abs f \le k \text{ ae}} k$$
The inf is attained: there exists some $N \in \mcF, \mu(N) = 0$ such that $\norm f_\infty = \sup_{N^c} \abs f$.
In all cases, we identify functions up to almost everywhere equality.
\begin{thm}
$L_p(\mu)$ is complete for $1 \le p \le infty$.
\end{thm}
\begin{dfn*}[Complex measures]
A {\bf complex measure} on $\mcF$ is a countably additive set function $\nu : \mcF \to \C$. \\
The {\bf total variation measure} $\abs\nu$ is defined by
$$\abs\nu(A) = \sup_{\substack{A_1, \dots, A_n \text{ measurable} \\ \text{partition of } A}} \sum_k \abs{\nu(A_k)}$$
$\abs\nu : \mcF \to [0, \infty]$ is a positive measure. Later we'll see that $\abs\nu$ is a finite measure. \\
The {\bf total variation} of $\nu$ is $\norm\nu_1 = \abs\nu(\Omega)$.
\end{dfn*}
\begin{prop}
If $\nu$ is a complex measure on $\mcF$ and $A_n \in \mcF$ for all $n$, then
\begin{itemize}
\item If $A$ is monotone, then $\nu(\Union_n A_n) = \lim_{n \to \infty} \nu(A_n)$.
\item If $A$ is antitone, then $\nu(\bigcap_n A_n) = \lim_{n \to \infty} \nu(A_n)$.
\end{itemize}
\end{prop}
\begin{dfn*}[Signed measures]
A {\bf signed measure} on $\mcF$ is a countably additive set function $\nu : \mcF \to \R$.
\end{dfn*}
\begin{thm}\label{thm:hahn-decomp}
If $\nu$ is a signed measure, then there exists a measurable partition $\Omega = P \cup N$ such that for all $A \in \mcF$
\begin{align*}
A \subseteq P & \imp \nu(A) \ge 0 \\
A \subseteq N & \imp \nu(A) \le 0
\end{align*}
\end{thm}
\begin{rmks}~
\begin{enumerate}
\item This decomposition is called the {\bf Hahn decomposition} of $\nu$.
\item Define $\nu^+(A) = \nu(A \cap P), \nu^-(A) = -\nu(A \cap N)$. Then $\nu^+, \nu^-$ are finite positive measures such that $\nu = \nu^+ - \nu^-$. This determines $\nu^+, \nu^-$ uniquely and the decomposition composition $\nu = \nu^+ - \nu^-$ is called the {\bf Jordan decomposition} of $\nu$.
\item If $\nu$ is a complex measure on $\mcF$, then $\Re\nu, \Im\nu$ are signed measures with Jordan decomposition $\nu_1 - \nu_2, \nu_3 - \nu_4$ respectively. Hence $\nu = \nu_1 - \nu_2 + i\nu_3 - i\nu_4$ is the Jordan decomposition of $\nu$.
$$\nu_1, \nu_2, \nu_3, \nu_4 \le \abs\nu \le \nu_1 + \nu_2 + \nu_3 + \nu_4$$
So $\abs\nu$ is a finite measure.
\end{enumerate}
\end{rmks}
\begin{proof}[Sketch]
Define $\nu^+(A) = \sup_{\substack{B \in \mcF \\ B \subseteq A}} \nu(B)$. $\nu^+$ is nonnegative and finitely additive.
{\bf Key step: $\nu^+(\Omega) < \infty$} \\
By contradiction, construct inductively sequences $A_n, B_n$ such that
$$\nu^+(A_n) = \infty, \nu(B_n) > n$$
by taking $A_0 = \Omega, B_{n + 1} \subseteq A_n$ such that $\nu(B_n) > n$ (exists by continuity) and $A_{n + 1} = B_{n + 1}$ or $A_n \setminus B_{n + 1}$. This contradicts countable additivity.
Now find a sequence $A_n$ such that $\nu(A_n) > \nu^+(\Omega) - 2^{-n}$ and set $P = \liminf_n A_n, N = P^c$. Check that this works.
\end{proof}
\newlec
\begin{dfn*}[Absolute continuity]
Let $(\Omega, \mcF, \mu)$ be a measure space and $\nu : \mcF \to \C$ a complex measure. $\nu$ is {\bf absolutely continuous} with respect to $\mu$, written $\nu \ll \mu$, if $\for A \in \mcF, \mu(A) = 0 \imp \nu(A) = 0$.
\end{dfn*}
\begin{rmks}~
\begin{itemize}
\item $\nu \ll \mu \imp \abs\nu \ll \mu$, so if $\nu$ has Jordan decomposition $\nu = \nu_1 - \nu_2 + i\nu_3 - i\nu_4$ then $\nu_1, \nu_2, \nu_3, \nu_4 \ll \mu$.
\item If $\nu \ll \mu$, then $\for \eps > 0, \exists \delta > 0, \for A \in \mcF, \mu(A) < \delta \imp \abs{\nu(A)} < \eps$.
\end{itemize}
\end{rmks}
\begin{eg*}
Let $f \in L_1(\mu)$. Define $\nu(A) = \int_A f d\mu$ for $A \in \mcF$. By Dominated Convergence, $\nu$ is a complex measure and $\mu(A) = 0 \imp \nu(A) = 0$. So $\nu \ll \mu$.
\end{eg*}
\begin{dfn*}
$A \in \mcF$ is {\bf $\sigma$-finite} if there exists $A_n$ with $\mu(A_n) < \infty$ such that $A = \Union_n A_n$. Say $\mu$ is {\bf $\sigma$-finite} if $\Omega$ is $\sigma$-finite.
\end{dfn*}
\begin{thm}[Radon-Nikodym]\label{thm:rn}
Let $\mu$ be a $\sigma$-finite measure and $\nu$ a complex measure such that $\nu \ll \mu$. Then there exists a unique $f \in L_1(\mu)$ such that, for all $A \in \mcF$, $\nu(A) = \int_A f d\mu$. Moreover, $f$ takes values in $\C$/$\R$/$\R^+$ depending on where $\nu$ is valued.
\end{thm}
\begin{proof}~\\
{\bf Uniqueness} \\
standard
{\bf Existence} \\
$\nu$ is a finite measure (by the Jordan decomposition). WLOG $\mu$ is a finite measure (by $\sigma$-finiteness). Let
$$\mcH = \curlybrack{h : \Omega \to \R^+ \,\middle|\, h \text{ integrable}, \for A \in \mcF, \int_A h d\mu \ll \nu(A)}$$
$\mcH \ne \emptyset$ (eg $0 \in \mcH$). Let $\alpha = \sup_{h \in \mcH} \int_\Omega h d\mu$. We see $0 \le \alpha \le \nu(\Omega)$.
{\bf Claim} \\
There exists $f \in \mcH$ such that $\alpha = \int_\Omega f d\mu$.
{\bf Idea} \\
If $\int_A fd\mu < \nu(A)$, then $f + \frac 1 n 1_A \in \mcH$ (morally, not literally), contradicting the definition of $\alpha$.
Pick that $f$. Define $\nu_n(A) = \nu(A) - \int_A fd\mu - \frac 1n \mu(A)$. $\nu_n$ has Hahn decomposition $\Omega = P_n \cup N_n$. Then $f + \frac 1n P_n \in \mcH$. By definition of $\alpha$, $\mu(P_n) = 0$. Since $\nu \ll \mu$, $\nu(P_n) = 0$. Let $P = \Union_n P_n, N = \Inter_n N_n$. Then $\Omega = P \cup N, \mu(P) = \nu(P) = 0$. For $A \in \mcF$,
\begin{align*}
A \subseteq P & \imp \int_A fd\mu = \nu(A) = 0 \\
A \subseteq N & \imp \for n, \nu_n \le 0 \imp \int_A fd\mu \ge \nu(A)
\end{align*}
\end{proof}
\begin{rmks}~
\begin{itemize}
\item Without assuming $\nu \ll \mu$, the proof shows there is a decomposition $\nu = \nu_1 + \nu_2$ where $\nu_1(A) = \int_A fd\mu$ and $\nu_2 \perp \mu$ ({\bf orthogonal}, ie there exists a measurable decomposition $\Omega = P \cup N$ such that $\mu(P) = 0, \abs{\nu_2}(N) = 0$). $\nu = \nu_1 + \nu_2$ is the {\bf Lebesgue decomposition} of $\nu$.
\item The unique $f$ in Theorem \ref{thm:rn} is the {\bf Radon-Nikodym derivative} of $\nu$ with respect to $\mu$, denoted $\frac{d\nu}{d\mu}$. The result says
$$\nu(A) = \int_\Omega 1_A d\nu = \int_A \frac{d\nu}{d\mu} d\mu = \int_\Omega 1_A \frac{d\nu}{d\mu} d\mu$$
Hence a measurable function $g$ is $\nu$-integrable iff $g\frac{d\nu}{d\mu}$ is $\mu$-integrable and then
$$\int_\Omega f d\nu = \int_\Omega g \frac{d\nu}{d\mu} d\mu$$
\end{itemize}
\end{rmks}
\subsection{Dual space of \texorpdfstring{$L_p(\mu)$}{Lp(mu)}}
Let $(\Omega, \mcF, \mu)$ be a measure space, $1 \le p < \infty, 1 < q \le \infty$ such that $p^{-1} + q^{-1} = 1$. For $g \in L_q$, define $\phi_g : L_p \to \K$ by $\phi_g(f) = \int_\Omega fg d\mu$. By Hölder, $fg \in L_1$, and $\abs{\phi_g(f)} \le \norm f_p \norm g_q$. So $\phi_g$ is well-defined, linear and bounded with $\norm{\phi_g} \le \norm g_q$. Hence $\phi_g \in L_p^*$ and $\phi : L_q \to L_p^*$ is linear and bounded with $\norm\phi \le 1$.
\begin{thm}\label{thm:lp-dual}~
\begin{enumerate}
\item If $1 < p < \infty$, then $\phi$ is an isometric isomorphism. So $L_p^* \cong L_q$.
\item If $p = 1$ and $\mu$ is $\sigma$-finite, then $\phi$ is an isometric isomorphism. So $L_1^* \cong L_\infty$.
\end{enumerate}
\end{thm}
\begin{proof}~
\begin{enumerate}
\item {\bf $\phi$ is isometric} \\
Let $g \in L_1$. We know $\norm{\phi_g} \le \norm g_q$. Let $\lambda$ be a measurable function with $\abs\lambda = 1, \lambda g = \abs g$. let $f = \lambda\abs g^{q - 1}$. Then
$$\norm f_p^p = \int_\Omega \abs g^{pq - p} d\mu = \int_\Omega \abs g^q d\mu = \norm g_q^q$$
So $f \in L_p$ and $\norm f_p = \norm g_q^{\frac qp}$. Then
$$\norm q_q^{\frac qp}\norm{\phi_g} \ge \abs{\phi_g(f)} = \int_\Omega \abs g^q d\mu = \norm g_q^q$$
So $\norm{\phi_g} \ge \norm g_q^{q - \frac qp} = \norm g_q$.
{\bf $\phi$ is onto} \\
Fix $\psi \in L_p^*$. We seek $g \in L_q$ such that $\psi = \phi_g$. Idea: We want $\psi(1_A) = \int_A g d\mu$.
{\bf Case 1: $\mu$ is finite} \\
For $A \in \mcF$, $1_A \in L_p$, so define $\nu(A) = \psi(1_A)$. $\nu(\empty) = 0$ and, if $A = \Union_p A_n \in \mcF$, then $\sum_k 1_{A_k} = 1_A$ in $L_p$, so
$$\sum_k \nu(A_k) = \sum_k \psi(1_{A_k}) = \psi(1_A)$$
Hence $\nu$ is a complex measure. \\
If $A \in \mcF, \mu(A) = 0$, then $1_A = 0$ ae in $L_p$, so $\nu(A) = \psi(1_A) = 0$. Hence $\nu \ll \mu$. \\
By Theorem \ref{thm:rn}, find $g \in L_1$ such that $\for A \in \mcF, \nu(A) = \int_A g d\mu$. Hence
\begin{align*}
\psi(1_A) & = \int_\Omega 1_A g d\mu \text{ for all } A \in \mcF \\
\psi(f) & = \int_\Omega fg d\mu \text{ for all simple function } f
\end{align*}
Given $f \in L_\infty$, find simple functions $f_n$ tending to $f$ in $L_\infty$. So $\psi(f_n) \to \psi(f)$ and $f_ng \to fg$ (by Hölder for $\infty, 1$), meaning that
$$\psi(f) = \int_\Omega fg d\mu \text{ for all } f \in L_\infty$$
For $n \in \N$, let $A = \{\abs g \le n\}$ and $f_n = \lambda 1_{A_n} \abs g^{q - 1}$ where $\abs\lambda = 1, \lambda g = \abs g$. As $f_n \in L_\infty$,
$$\int_\Omega f_ng d\mu = \int_{A_n} \abs g^q d\mu = \psi(f_n)$$
So $\left(\int_A \abs g^q d\mu\right)^{q^{-1}} \le \norm\psi$. By Monotone Convergence, $g \in L_q$. \\
Given $f \in L_p$, find simple functions $f_n$ tending to $f$ in $L_p$. So $\psi(f_n) \to \psi(f)$ and $f_ng \to fg$ in $L_1$ (by Hölder for $p, q$). Hence $\psi(f) = \int_\Omega fg d\mu$, as wanted.
Before going onto Case 2, for $A \in \mcF$, let $\mcF_A = \{B \in \mcF \mid B \subseteq A\}$ and $\mu_A = \mu\restriction_{\mcF_A}$ so that $(A, \mcF_A, \mu_A)$ is a measure space. Then $L_p(\mu_A) \subseteq L_p(\mu)$ (by extending $f \in L_p(\mu_A)$ by $0$ outside $A$). Let $\psi_A = \psi\restriction_{L_p(\mu_A)}$.
\newlec
\begin{claim}
If $A, B \in \mcF$ are disjoint, then
$$\norm{\psi_{A \cup B}} = \left(\norm{\psi_A}^q + \norm{\psi_B}^q\right)^{\frac 1q}$$
\end{claim}
\begin{proof}
\begin{align*}
\left(\norm{\psi_A}^q + \norm{\psi_B}^q\right)^{\frac 1q}
& = \sup_{\substack{a, b \ge 0 \\ a^p + b^p \le 1}} a\norm{\psi_A} + b\norm{\psi_B} \\
& = \sup_{\substack{a, b \ge 0 \\ a^p + b^p \le 1 \\ f \in B_{L_p(\mu_A)} \\ g \in B_{L_p(\mu_B)}}} a\abs{\psi_A(f)} + b\abs{\psi_B(g)} \\
& = \sup_{\substack{\abs a^p + \abs b^p \le 1 \\ f \in B_{L_p(\mu_A)} \\ g \in B_{L_p(\mu_B)}}} \underbrace{\abs{a\psi_A(f) + b\psi_B(g)}}_{\psi_{A \cup B}(af + bg)} \\
& = \sup_{h \in L_p(\mu_{A \cup B})} \abs{\psi_{A \cup B}(h)} \\
& = \norm{\psi_{A \cup B}}
\end{align*}
\end{proof}
{\bf Case 2: $\mu$ is $\sigma$-finite} \\
Find a measurable partition $\Omega = \Union_n A_n$ such that $\mu(A_n) < \infty$. By Case 1, find, for each $n$, $g_n \in L_q(A_n)$ such that $\psi_{A_n} = \phi_{g_n}$, ie
$$\psi(f) = \int_{A_n} fg_n d\mu \text{ for all } f \in L_q(\mu_{A_n})$$
If we define $g$ on $\Omega$ by $g = g_n$ on $A_n$, then $g \in L_q$ and
$$\psi(f) = \phi_g(f) \text{ for all } f \in L_p(\mu_{A_n})$$
Hence $\psi = \phi_g$ on $\bar{\Span}\Union_n L_p(\mu_{A_n}) = L_p(\mu)$.
{\bf Case 3: General $n$} \\
First observe that, for $f \in L_p(\mu)$, $\{f \ne 0\}$ is $\sigma$-finite. Indeed,
$$\{f \ne 0\} = \Union_n \curlybrack{\frac 1n < \abs f}$$
and
$$\mu\curlybrack{\frac 1n < \abs f} \le \abs{n ^ p} \norm f_p^p < \infty \text{ by Markov}$$
Choose $f_n \in B_{L_p}$ such that $\psi(f_n) \to \norm\psi$. Then $A = \Union_n\{f_n \ne 0\}$ is $\sigma$-finite and $\norm{\psi_A} = \norm\psi$. By the claim,
$$\norm\psi = \left(\norm{\psi_A}^q + \norm{\psi_{A^c}}^q\right)^{\frac 1q}$$
So $\Psi_{A^c} = 0$. By Case 2, find $g \in L_q(\mu_A) \subseteq L_q(\mu)$ such that $\psi_A = \phi_g$, so that
$$\psi(f) = \psi_A{f\restriction_A} + \psi{A^c}(f\restriction{A^c}) = \int_A fg d\mu + 0 = \int_\Omega fg d\mu$$
\item {\bf $p = 1, \mu$ is $\sigma$-finite} \\
{\bf $\phi$ is isometric} \\
Let $g \in L_\infty$. We know $\norm{\phi_g} \le \norm g_\infty$ (by Hölder) Fix $s < \norm g_\infty$. Then $\mu\{s < \abs g\} > 0$. Since $\mu$ is $\sigma$-finite, find $A \subseteq \{s < \abs g\}$ such that $0 < \mu(A) < \infty$. Choose a measurable function $\lambda$ such that $\abs\lambda = 1, \lambda g = \abs g$. Then $\lambda 1_A \in L_1, \norm{\lambda 1_A}_1 = \mu(A)$. Now,
$$\mu(A)\norm{\phi_g} \ge \abs{\phi_g(\lambda 1_A)} = \int_A \abs g d\mu \ge s\mu(A)$$
So $\norm{\phi_g} \ge s$. Taking the sup, $\norm{\phi_g} \ge \norm g_\infty$.
{\bf $\phi$ is onto} \\
Fix $\psi \in L_q^*$. We seek $g \in L_\infty$ such that $\psi = \phi_g$.
{\bf Case 1: $\mu$ is finite} \\
Define $\nu(A) = \psi(1_A)$ for all $A \in \mcF$. Follow the same steps as for $1 < p < \infty$.
{\bf Case 2: $\mu$ is $\sigma$-finite} \\
This time, prove that
$$\norm{\psi_{A \cup B}} = \max(\norm{\psi_A}, \norm{\psi_B})$$
for all $A, B \in \mcF$ disjoint and proceed as before.
\end{enumerate}
\end{proof}
\begin{cor}
For $1 < p < \infty$, $L_p(\mu)$ is reflexive.
\end{cor}
\begin{proof}
Let $\psi \in L_p^{**}$. Then $g \mapsto \inn{\phi_g}\psi : L_q \to \K$ is in $L_q^*$. By Theorem \ref{thm:lp-dual}.i, find $f \in L_p$ such that
$$\inn{\phi_g}\psi = \int_\Omega fg d\mu = \inn f{\phi_g} = \inn{\phi_g}{\hat f}$$
Since $L_p^* = \{\phi_g \mid g \in L_q\}$, this proves $\psi = \hat f$.
\end{proof}
\clearpage
\subsection{Dual space of \texorpdfstring{$C(K)$}{C(K)}}
Throughout, $K$ will be a compact Hausdorff topological space. Define
\begin{align*}
C(K) & = \{f : K \to \C \mid f \text{ continuous}\} \\
C^\R(K) & = \{f : K \to \R \mid f \text{ continuous}\} \\
C^+(K) & = \{f : K \to \R^+ \mid f \text{ continuous}\} \\
M(K) & = C(K)^* \\
M^\R(K) & = \{\phi \in M(K) \mid \for f \in C^\R(K), \phi(f) \in \R\} \\
M^+(K) & = \{\phi : C(K) \to \C \mid \phi \text{ is $\R$-linear}, \for f \in C^+(K), 0 \le \phi(f) \in \R\}
\end{align*}
$C(K), C^\R(K)$ are complex/real Banach spaces in the sup norm: $\norm f_\infty = \sup_K \abs f$. $M(K)$ is a complex Banach space in the operator norm. $M^\R(K)$ is a closed real-linear subspace of $M(k)$. Elements of $M^+(K)$ are called {\bf positive linear functionals}.
\begin{aim}
Identify $M(K), M^\R(K)$.
\end{aim}
\newlec
The next lemma tells us that it's enough to understand $M^+(K)$.
\begin{lem}\label{lem:m-real-pos}~
\begin{enumerate}
\item For all $\phi \in M(K)$, there are unique $\phi_1, \phi_2 \in M^\R(K)$ such that $\phi = \phi_1 + i\phi_2$.
\item $\phi \mapsto \phi\restriction_{C^\R(K)} : M^\R(K) \to C^\R(K)^*$ is an isometric isomorphism.
\item $M^+(K) \subseteq M(K)$ and $M^+(K) = \{\phi \in M(K) \mid \norm\phi = \phi(1)\}$
\item For all $\phi \in M^\R(K)$, there are unique $\phi^+, \phi^- \in M^+(K)$ such that $\phi = \phi^+ - \phi^-$ and $\norm\phi = \norm{\phi^+} + \norm{\phi^-}$.
\end{enumerate}
\end{lem}
\begin{proof}~
\begin{enumerate}
\item Let $\phi \in M(K)$. Then $\bar\phi$ sending $f \mapsto \phi(\bar f)$ is in $M(K)$ as well and $\phi \in M^\R(K) \iff \bar\phi = \phi$.
{\bf Uniqueness} \\
Assume $\phi = \phi_1 + i\phi_2$ where $\phi_1, \phi_2 \in M^\R(K)$. Then $\bar\phi = \phi_1 - i\phi_2$, so
$$\phi_1 = \frac{\phi + \bar\phi}2, \phi_2 = \frac{\phi - \bar\phi}{2i}$$
{\bf Existence} \\
Check that the above works
\item Let $\phi \in M^\R(K)$. We show $\norm{\phi\restriction_{C^\R(K)}} = \norm\phi$. Clearly, $\norm{\phi\restriction_{C^\R(K)}} \le \norm\phi$. Let $f \in B_{C(K)}$. Choose $\lambda \in \C, \abs\lambda = 1, \lambda\phi(f) = \abs{\phi(f)}$, so that
\begin{align*}
\abs{\phi(f)}
& = \lambda\phi(f) \\
& = \phi(\lambda f) \\
& = \phi(\Re(\lambda f)) + \cancelto 0{\phi(\Im(\lambda f))} \\
& \le \norm{\phi\restriction_{C^\R(K)}} \norm{\Re(\lambda f)}_\infty \\
& \le \norm{\phi\restriction_{C^\R(K)}}
\end{align*}
Hence $\norm\phi \le \norm{\phi\restriction_{C^\R(K)}}$. \\
Finally, given $\psi \in C^\R(K)$, define $\phi(f) = \psi(\Re f) + i\psi(\Im f)$. Then $\phi \in M(K)$ and $\phi\restriction_{C^\R(K)} = \psi$.
\item {\bf $M^+(K) \subseteq M(K)$} \\
Let $\phi \in M^+(K)$. For $f \in B_{C^\R(K)}$, we have $1 \pm f \ge 0$, so $\phi(1 \pm f \ge 0$. Hence $\phi(f) \in \R$ and $\abs{\phi(f)} \le \phi(1)$. So $\phi\restriction_{C^\R(K)} \in C^\R(K)^*$ and $\norm{\phi\restriction_{C^\R(K)}} = \phi(1)$. By (ii), $\phi \in M(K), \norm\phi = \phi(1)$.
{\bf $M^+(K) = \{\phi \in M(K) \mid \norm\phi = \phi(1)$} \\
We have already checked one inclusion. Let $\phi \in M(K)$ with $\norm\phi = \phi(1)$. WLOG $\norm\phi = \phi(1) = 1$. Let $f \in B_{C^\R(K)}$ and write $\phi(f) = a + ib$ where $a, b \in \R$. We want $b = 0$. For $t \in \R$,
\begin{align*}
\abs{\phi(f + it)}^2
& = a^2 + (b + t)^2 = a^2 + b^2 + t^2 + 2bt \\
& \le \norm{f + it}_\infty \le 1 + t^2
\end{align*}
So $b = 0$. \\
Given $f \in C^+(K)$ with $0 \le f \le 1$, we have $-1 \le 2f - 1 \le 1$, so $\abs{\phi(2f - 1)} \le \norm{2f - 1}_\infty \le 1$, ie $-1 \le 2\phi(f) - 1 \le 1$. So $\phi(f) \ge 0$.
\item Let $\phi \in M^\R(K)$. Assume for a moment that $\phi = \psi_1 - \psi_2$ where $\psi_1, \psi_2 \in M^+(K)$. For $f, g \in C^+(K)$ with $0 \le g \le f$, we have $\psi_1(f) \ge \psi_1(g) = \phi(g) + \psi_2(g) \ge \phi(g)$. So
$$\psi_1(f) \ge \sup_{0 \le g \le f} \phi(g)$$
For $f \in C^+(K)$, define
$$\phi^+(f) = \sup_{0 \le g \le f} \phi(g)$$
Observe that $\phi^+ \ge 0, \phi^+(f) \le \norm\phi \norm f_\infty, \phi^+(f) \ge \phi(f), \phi^+$ is linear. \\
Next, for $f \in C^\R(K)$, write $f = f_1 - f_2$ where $f_1, f_2 \in C^+(K)$ and define $\phi^+(f) = \phi^+(f_1) - \phi^+(f_2)$. This is well-defined and $\R$-linear. Then $\phi$ is $\C$-linear since $\phi^+(f) \ge 0$. For all $f \in C^+(K)$ and $\phi^+ \in M^+(K)$. \\
Define $\phi^- = \phi^+ - \phi$. For $f \in C^+(K)$, $\phi^+(f) \ge \phi(f)$, so $\phi^-(f) \ge 0$, namely $\phi^- \in M^+(K)$. \\
We now see that $\norm\phi \le \norm{\phi^+} + \norm{\phi^-}$.
Given $f \in C^+(K), 0 \le f \le 1$, we have $-1 \le 2f - 1 \le 1$, so
$$2\phi(f) - \phi(1) = \phi(2f - 1) \le \norm\phi$$
Taking the sup over $f$, we thus check that
$$\norm{\phi^+} + \norm{\phi^-} = \phi^+(1) + \phi^-(1) = 2\phi^+(1) - \phi(1) \le \norm\phi$$
{\bf Uniqueness} \\
Assume $\phi = \psi_1 - \psi_2, \psi_1, \psi_2 \in M^+(K), \norm\phi = \norm{\psi_1} + \norm{\psi_2}$. From the initial observation, $\psi_1 \ge \phi^+$, hence $\psi_2 = \psi_1 - \phi \ge \phi^+ - \phi = \phi^-$. Therefore $\psi_1 - \phi^+, \psi_2 - \phi^- \in M^+(K)$. By (iii),
$$\norm{\psi_1 - \phi^+} + \norm{\psi_2 - \phi^-} = \psi_1(1) - \phi^+(1) + \psi_2(1) - \phi^-(1) = \norm\phi - \norm\phi = 0$$
Hence $\psi_1 = \phi^+, \psi_2 = \phi^-$.
\end{enumerate}
\end{proof}
{\bf Topological preliminaries}
\begin{enumerate}
\item $K$ being compact Hausdorff, it is {\bf normal}: given disjoint closed sets $E, F$ in $K$, there are disjoint open sets $U, V$ such that $E \subseteq U, F \subseteq V$. Equivalently, given $E \subseteq U \subseteq K$, $E$, closed, $U$ open, there exists $V$ open such that $E \subseteq V \subseteq \bar V \subseteq U$.
\item Urysohn says: given disjoint closed sets $E, F$, there is a continuous function $f : K \to [0, 1]$ such that $f = 0$ on $E$, $f = 1$ on $F$.
\item Write $f \prec U$ to mean that $U$ is an open set, $f$ is continuous and $\supp f \subseteq U$. Write $E \prec f$ to mean that $E$ is closed, $f$ is continuous and $f = 1$ on $E$.
\item Urysohn then becomes: Given $E \subseteq U$, there exists $f$ such that $E \prec f \prec U$.
\end{enumerate}
\begin{lem}\label{lem:partition-of-unity}
Let $E$ closed, $U_1, \dots, U_n$ open such that $E \subseteq \Union_n U_n$. Then
\begin{enumerate}
\item There exist open sets $V_j$ such that $\bar{V_j} \subseteq U_j$ and $E \subseteq \Union_j V_j$.
\item There exist $f_j \prec U_j$ such that $0 \le \sum_j f_j \le 1$ and $\sum_j f_j = 1$ on $E$.
\end{enumerate}
\end{lem}
\begin{proof}~
\begin{enumerate}
\item Induction on $n$:
{\bf $n = 0$} \\
Obvious.
{\bf $n > 0$} \\
$E \setminus U_n \subseteq \Union_{j < n} U_j$ so, by induction, find open sets $V_j$ such that $\bar{V_j} \subseteq U_j$ for all $j < n$ and $E \setminus U_n \subseteq \Union_{j < n} U_j$. So $\underbrace{E \setminus \Union_{j < n} V_j}_{\text{closed}} \subseteq \underbrace{U_n}_{\text{open}}$. By Urysohn, find an open $V_n$ such that
$$E \setminus \Union_{j < n} V_j \subseteq V_n \subseteq \bar{V_n} \subseteq U_n$$
\item Find the $V_j$ as in $(i)$ for $1 \le j \le n$ and by Urysohn find $h_j$ such that $\bar{V_j} \prec h_j \prec U_j$. By Urysohn again, find $h_0$ such that $\left(\Union_j U_j\right)^c \prec h_0 \prec E^c$. Let $h = \sum_{j = 0}^n h_j \ge 1$ and $f_j = \frac{h_j}h$ for $1 \le j \le n$. Then $0 \le \sum_{j = 1}^n \le 1$, $f_j \prec U_j$ and $\sum_{j = 1}^n f_j = 1$ on $E$.
\end{enumerate}
\end{proof}
\begin{dfn*}[Borel measures]
Let $X$ be a Hausdorff space and $\mcG$ its family of open sets. The {\bf Borel $\sigma$-algebra} is $\mcB := \sigma(\mcG)$, the $\sigma$-algebra generated by open sets. Elements of $\mcB$ are called {\bf Borel sets}. A {\bf Borel measure} on $X$ is a measure $\mu$ on $\mcB$. We say $\mu$ is {\bf regular} if
\begin{enumerate}
\item $\mu(E) < \infty$ for all compact $E \subseteq X$
\item $\mu(A) = \inf_{\substack{U \text{ open} \\ A \subseteq U}} \mu(A)$ for all Borel set $A$
\item $\mu(U) = \sup_{\substack{E \text{ compact} \\ E \subseteq U}} \mu(E)$ for all open $U$
\end{enumerate}
A complex Borel measure $\nu$ is {\bf regular} if $\abs\nu$ is regular.
\end{dfn*}
If $X$ is compact and $\mu$ is a Borel measure on $X$, then
\begin{align*}
\mu \text{ regular}
& \iff \mu(X) < \infty \text{ and } \for A \in \mcB, \mu(A) = \inf_{\substack{U \text{ open} \\ A \subseteq U}} \mu(U) \\
& \iff \mu(X) < \infty \text{ and } \for A \in \mcB, \mu(A) = \sup_{\substack{E \text{ closed} \\ E \subseteq A}} \mu(E)
\end{align*}
\begin{dfn*}[Integration with respect to a complex measure]
Let $\Omega$ be a set, $\mcF$ a $\sigma$-algebra on $\Omega$, $\nu$ a complex measure on $\mcF$. Write $\nu = \nu_1 - \nu_2 + i\nu_3 - i\nu_4$ the Jordan decomposition. Say a measurable function is {\bf $\nu$-integrable} if $f$ is $\abs\nu$-integrable, or equivalently if $f$ is $\nu_1, \nu_2, \nu_3, \nu_4$-integrable. Define
$$\int_\Omega f d\nu = \int_\Omega f d\nu_1 - \int_\Omega f d\nu_2 + i\int_\Omega f d\nu_3 - i\int_\Omega f d\nu_4$$
\end{dfn*}
\newlec
\begin{prop}~
\begin{enumerate}
\item $\int_\Omega d\nu = \nu(A)$ for all $A \in \mcF$.
\item Linearity: If $f, g : \Omega \to \C$ are $\nu$-integrable and $\lambda \in \C$, then
$$\int_\Omega f + g d\nu = \int_\Omega fd\nu + \int_\Omega gd\nu, \int_\Omega \lambda fd\nu = \lambda \int_\Omega fd\nu$$
\item Dominated Convergence: Let $f_n, f, g$ be measurable functions $\Omega \to \C$ such that $f_n \to f$ ae (with respect to $\abs\nu$), $g \in L_1$ and $\for n, f_n \le g$ ae. Then $f$ is $\nu$-integrable and $\int_\Omega f_n d\nu \to \int_\Omega f d\nu$
\item $\abs{\int_\Omega f d\nu} \le \int_\Omega \abs f d\abs\nu$ for all $\nu$-integrable $f$. This is true for simple functions by properties 1 and 2. For general $f$, use Dominated Convergence.
\end{enumerate}
\end{prop}
Let $\nu$ be a complex Borel measure on $K$. Then for $f \in C(K)$ we have
$$\int_K \abs f d\abs\nu \le \norm f_\infty \abs\nu(K) = \norm f_\infty \norm\nu_1$$
So $f$ is $\nu$-integrable. Define $\phi : C(K) \to \C$ by $\phi(f) = \int_\Omega f d\nu$. Then $\phi \in M(K)$ and $\norm\phi \le \norm\nu_1$. If $\nu$ is a signed measure, then $\phi \in M^\R(K)$. If $\nu$ is a positive measure, then $\phi \in M^+(K)$.
\begin{thm}[Riesz Representation Theorem]\label{thm:riesz-rep}
For every $\phi \in M^+(K)$, there exists a unique regular Borel measure $\mu$ on $K$ that represents $\phi$: $\phi(f) = \int_K f d\mu$ for all $f \in C(K)$. Moreover, $\norm\phi = \mu(K) = \norm\mu_1$.
\end{thm}
\begin{proof}~\\
{\bf Uniqueness} \\
Assume $\mu_1, \mu_2$ both represent $\phi$. Let $E \subseteq U \subseteq K$ where $E$ closed, $U$ open. By Urysohn, find $f$ such that $E \prec f \prec U$. Now,
$$\mu_1(E) \le \int_K f d\mu_1 = \phi(f) = \int_K f d\mu_2 \le \mu_2(U)$$
Taking the inf over $U$, we get $\mu_1(E) \le \mu_2(E)$. By symmetry, $\mu_1(E) = \mu_2(E)$. By regularity, $\mu_1 = \mu_2$.
{\bf Existence} \\
For $U$ open, define $\mu^*(U) = \sup_{f \prec U} \phi(f)$. Note that
$$\mu^*(U) \ge 0, \mu \text{ monotone}, \mu^*(K) = \phi(1)$$
It follows that, for $V$ open, $\mu^*(V) = \inf_{U \supseteq V} \mu^*(U)$. Hence extend the definition of $\mu^*$ to
$$\mu^*(A) = \inf_{U \supseteq A} \mu^*(U)$$
We will show that $\mu^*$ is an outer measure.
\begin{itemize}
\item $\mu(\emptyset) = 0$
\item If $A \subseteq B$, then $\mu^*(A) \le \mu^*(B)$.
\item Do we have $\mu^*\left(\Union_n A_n\right) = \sum_n \mu^*(A_n)$? \\
First assume that the $A_n = U_n$ are open. Let $U = \Union_n U_n$. Assume $f \prec U$ and let $E = \supp f$. $E \subseteq \Union_n U_n$, so by compactness find $N$ such that $E \subseteq \Union_{n = 1}^N U_n$. By Lemma \ref{lem:partition-of-unity}, find $h_n \prec U_n$ with $\sum_{n = 1}^N h_n \le 1$ and $\sum_{n = 1}^N h_n = 1$ on $E$. So $f = \sum_{n = 1}^N fh_n$ and
\begin{align*}
\phi(f)
& = \sum_{n = 1}^N \phi(fh_n) \\
& \le \sum_{n = 1}^N \mu^*(U_j) \text{ as } fh_n \prec U_n \\
& \le \sum_n \mu^*(U_n)
\end{align*}
Taking the sup over $f$, we get $\mu^*(U) \le \sum_n \mu^*(U_n)$. It follows that
$$\mu^*(\Union_n A_n) \le \sum_n \mu^*(A_n)$$
We now let $\mcM$ be the set of $\mu^*$-measurable sets. Then $\mcM$ is a $\sigma$-algebra and $\mu^*\restriction_{\mcM}$ is a measure on $\mcM$. \\
To restrict it further to a Borel, we now show that $\mcB \subseteq \mcM$. It's enough to show that $\mcG \subseteq \mcM$. \\
Let $U$ open. We need
$$\mu^*(A) \ge \mu^*(A \inter U) + \mu^*(A \setminus U) \text{ for all } A$$
First, let $A = V \in \mcG$. Fix $f \prec V \inter U$ and $g \prec V \setminus \supp f$. Then $f + g \prec V$, thus
$$\mu^*(V) \ge \phi(f + g) = \phi(f) + \phi(g)$$
Taking the sup over $g$,
$$\mu^*(V) \ge \phi(f + g) = \phi(f) + \mu^*(V \setminus \supp f) \ge \phi(f) + \mu^*(V \setminus U)$$
Taking the sup over $f$,
$$\mu^*(V) \ge \mu^*(V \inter U) + \mu^*(V \setminus U)$$
Now let $A$ be arbitrary. Fix $V$ open such that $A \subseteq V$. then
$$\mu^*(V) \ge \mu^*(V \inter U) + \mu^*(V \setminus U) \ge \mu^*(A \inter U) + \mu^*(A \setminus U)$$
Taking the inf over $V$,
$$\mu^*(A) \ge \mu^*(A \inter U) + \mu^*(A \setminus U)$$
Now, $\mu := \mu^*\restriction_{\mcB}$ is a Borel measure on $K$. We have
$$\mu(K) = \phi(1) = \norm\phi < \infty$$
and by definition $\mu$ is regular. It remains to show that $\phi(f) = \int_K f d\mu$ for all $f \in C(K)$. It is enough to check that for $f \in C^\R(K)$ and enough to check that $\phi(f) \le \int_K f d\mu$ (apply this to $-f$). \\
Fix $0 < a < b$ in $\R$ such that $\phi(1) \in [a, b]$. Let $\eps > 0$. Choose $0 \le y_0 < a \le y_1 < \dots < y_n = b$ such that
$$y_j < y_{j - 1} + \eps$$
Let $A_j = f^{-1}\mathopen]y_{j - 1}, y_j\mathclose]$. Those sets form a measurable partition of $K$. Choose closed sets $E_j$ and open sets $U_j$ such that $E_j \subseteq A_j \subseteq U_j$ and $\mu(U_j \setminus E_j) < \frac\eps n$ (by regularity) and $f(U_j) \subseteq \mathopen]y_{j - 1}, y_j\mathclose]$. By Lemma \ref{lem:partition-of-unity}, find $h_j \prec U_j$ for each $j$ such that $\sum_j h_j = 1$. Now,
\begin{align*}
\phi(f)
& = \sum_j \phi(f_j) \\
& \le \sum_j (y_j + \eps)\phi(h_j) \\
& \le \sum_j (y_{j - 1} + 2\eps)\left(\mu(E_j) + \frac\eps n\right) \\
& = \sum_j y_{j - 1}\mu(E_j) + \underbrace{\sum_j (b + \eps) + 2\eps\mu(K) + 2\eps^2}_{o(1)} \\
& = \int_K \sum_j y_{j - 1}1_{E_j} d\mu + o(1)
& \le \int_K f d\mu + o(1)
\end{align*}
since $f \le y_j + \eps$ on $U_j$, $h_j \prec U_j$ and $\phi \in M^+(K)$. So $\phi(f) \le \int_K f d\mu$.
\end{itemize}
\end{proof}
\newlec
\begin{cor}
For every $\phi \in M(K)$, there exists a unique regular complex Borel measure $\nu$ on $K$ that represents $\phi$: $\phi(f) = \int_K f d\nu$ for all $f \in C(K)$. Moreover, $\norm\phi = \norm\nu_1$ and if $\phi \in M^\R(K)$ then $\nu$ is a signed measure.
\end{cor}
\begin{proof}~\\
{\bf Existence} \\
Apply Lemma \ref{lem:m-real-pos} and Theorem \ref{thm:riesz-rep} to obtain a regular complex Borel measure representing $\phi$. We now want $\norm\phi = \norm\nu_1$. \\
We already know $\norm\phi \le \norm\nu_1$. Take a measurable partition $K = \Union_{j = 1}^n A_j$. Fix $\eps > 0$ and closed sets $E_j$, open sets $U_j$ such that $E_j \subseteq A_j \subseteq U_j$, $\abs\nu(U_j \setminus E_j) < \frac\eps n$ ($\nu$ is regular). We can also assume $U_i \subseteq \Inter_{j \ne i} E_j^c$. Fix $\lambda_j \in \C$ such that $\abs{\lambda_j} = 1, \lambda_j\nu(E_j) = \abs{\nu(E_j)}$. By Lemma \ref{lem:partition-of-unity}, find $h_j \prec U_j$ such that $\sum_{j = 1}^n h_j = 1$. Then $E_j \prec h_j$, hence
\begin{align*}
\abs{\int_K \left(\sum_{j = 1}^n \lambda_j 1_{E_j} - \sum_{j = 1}^n \lambda_jh_j\right) d\nu}
& \le \sum_{j = 1}^n \int_K \abs{1_{E_j} - h_j} d\abs\nu \\
& \le \sum_{j = 1}^n \abs\nu(U_j \setminus E_j) < \eps
\end{align*}
Now,
\begin{align*}
\sum_{j = 1}^n \abs{\nu(A_j)}
& \le \sum_{j = 1}^n \abs{\nu(E_j)} + \eps \\
& = \sum_{j = 1}^n \lambda_j\nu(E_j) + \eps \\
& = \int_K \sum_{j = 1}^n \lambda_j1_{E_j} d\nu + \eps \\