-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathpred-semantic.tex
2959 lines (2550 loc) · 148 KB
/
pred-semantic.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
\chapter{Semantic Metalogic} \label{chap-sem}
%% TO DO:
%% 0. categorical => complete
%% 0. $\kappa$-categorical
%% 0. Robinson and Craig theorems -- Svenonius => Beth => Robinson =>
%% especially Robinson, with diamond diagram
%% 0. Scott definability theorem?
%% 0. cite Visser papers
%% 0. Beth's theorem = epi-surjective property for Boolean algebras.
%% Do I mention this connection already in the chapter on Boolean
%% algebras? If not, then begin discussion of Beth by pointing that
%% out. (Recall that we prove ES by means of completeness.)
%% 0. Craig interpolation theorem
%% 0. How to prove: homotopy equivalent => categorically equivalent
%% 0. perhaps prove some things about $\vDash$
%% 0. Beth theorem
%% 4. examples of models
%% 5. functors between Mod(T) and Mod(T')
%% 6. meaning of eso, faithful, etc..
%% 0. ultraproducts
%% 0. Bickle -- reduction
%% 0. van Fraassen -- embeddable
%% 7. standard versus non-standard models; intended versus non-intended
%% model
%% 8. conservative versus model-theoretically conservative ... the latter
%% same as forgetful functor being eso. TO DO: example of
%% conservative but not M-conservative
%% 10. mention completeness via Deligne embedding thm
%% Do I want to show that $(GF)^*=F^*G^*$ ??
%% Open Question
%% I doubt that the following is true -- but something close to it
%% is. See Breiner thesis
% \begin{prop} If $F$ is conservative then $F^*$ is
% superjective. \end{prop}
% \begin{proof} Take a model $M$ of $T$. The trick here is to grow $M$
% into a model $M'$ of $T'$, with an elementary embedding
% $h:M\to F^*(M')$. Intuitively speaking, since $F$ is conservative,
% $M$ is also a model of a subtheory of $T'$, namely the image of $F$.
% Then we need to show that a model of a subtheory of $T'$ can be
% extended to a model of $T'$. \end{proof}
%% TO DO: vF thought that he could explicate further notions than had
%% been in the syntactic approach. Most importantly, the notion of
%% empirical content. See vF - Muller paper, and my articles. Also,
%% that J Phil paper by Peter Turney
\section{The semantic turn}
Already in the 19th century, geometers were proving the relative
consistency of theories by interpreting them into well-understood
mathematical frameworks --- e.g.\ other geometrical theories, or the
theory of real numbers. At roughly the same time, the theory of sets
was under active development, and mathematicians were coming to
realize that the things they were talking about (numbers, functions,
etc.) could be seen to be constituted by sets. However, it was only
in the middle of the 20th century that Alfred Tarski gave a precise
definition of an {\it interpretation} of a theory in the universe of
sets.
Philosophers of science were not terribly quick to latch onto the new
discipline of logical semantics. Early adopters included the Dutch
philosopher Evert Beth, and to a lesser extent, Carnap himself. It
required a generational change for the semantic approach to take root
in philosophy of science. Here we are using ``semantic approach'' in
the broadest sense --- essentially for any approach to philosophy of
science that is reactionary against Carnap's syntax program, but that
wishes to use precise mathematical tools (set theory, model theory,
etc.) in order to explicate the structure of scientific theories.
What's most interesting for us is how the shift to the semantic
approach influenced shifts in philosophical perspective. Some of the
cases are fairly clear. For example, with the rejection of the
syntactic approach, many philosophers stopped worrying about the
``problem of theoretical terms'', i.e.\ how scientific theories (with
their abstract theoretical terms) connect up to empirical reality.
According to Putnam, among others, if you step outside the confines of
Carnap's syntax program, there is no problem of theoretical terms.
(Interestingly, debates about the conventionality of geometry all but
stopped around the 1970s, just when the move to the semantic view was
in full swing.) Other philosophers diagnosed the situation
differently. For example, van Fraassen saw the semantic approach as
providing the salvation of empiricism --- which, he thought, was
incapable of an adequate articulation from a syntactic point of view.
In reading late 20th century analytic philosophy, it can seem that
logical semantics by itself obviates many of the problems that
exercised the previous generation of philosophers. For example,
\citet[p.\ 222]{bas1989} says that ``the semantic view of theories
makes language largely irrelevant to the subject [philosophy of
science].'' Indeed, the picture typically presented to us is that
logical semantics deals with mind-independent things (viz.\
set-theoretic structures), which can stand in mind-independent
relations to concrete reality, and to which we have unmediated
epistemic access. Such a picture suggests that logical semantics
provides a bridge over which we can safely cross the notorious
mind-world gap.
But something is fishy with this picture. How could logical semantics
get any closer to ``the world'' than any other bit of mathematics?
And why think that set-theoretic structures play this privileged role
as intermediaries in our relation to empirical reality? For that
matter, why should our philosophical views on science be tied down to
some rather controversial view of the nature of mathematical objects?
Why the set-theoretic middle-man?
In what follows, I will attempt to put logical semantics back in its
place. The reconceptualization I'm suggesting begins with noting that
logical semantics is a particular version of a general mathematical
strategy called ``representation theory''. There is a representation
theory for groups, for rings, for $C^*$-algebras, etc., and the basic
idea of all these representation theories is to study one category
$\cat{C}$ of mathematical objects by studying the functors from
$\cat{C}$ to some other mathematical category, say $\cat{S}$. It
might seem strange that such an indirect approach could be helpful for
understanding $\cat{C}$, and yet, it has proven to be very frutiful.
For example, in the representation theory of groups, one studies the
representations of a group on Hilbert spaces. Similarly, in the
representation theory of rings, one studies the modules over a ring.
In all such cases, there is no suggestion that a represented
mathematical object is less linguistic than the original mathematical
object. If anything, the represented mathematical object has
superfluous structure that is not intrinsic to the original
mathematical object.
To fully understand that logical semantics is representation theory,
one needs to see theories as objects in a category, and to show that
``interpretations'' are functors from that category into some other
one. We carried out that procedure for propositional theories in
Chapter \ref{cat-prop}, where we represented each propositional theory
as a Boolean algebra. We could carry out a similar construction for
predicate logic theories, but the resulting mathematical objects would
be something more complicated than Boolean algebras. (Tarski himself
suggested representing predicate logic theories as cylindrical
algebras, but a more elegant approach involves syntactic categories in
the sense of \cite{makkai}.) Thus, we will proceed in a different
manner and directly define the arrows (in this case, translations)
between predicate logic theories. We begin, however, with a little
crash course in traditional model theory.
\begin{example} Let $T$ be the theory, in empty signature, that says:
``there are exactly two things.'' A \emph{model} of $T$ is simply a
set with two elements. However, every model of $T$ has ``redundant
information'' that is not specified by $T$ itself. To the question
``how many models does $T$ have?'' there are two correct answers:
(1) more than any cardinal number, and (2) exactly one (up to
isomorphism). \end{example}
\begin{example} Let $T_1$ be the theory of groups, as axiomatized in
Example \ref{groups}. Then a model $M$ of $T_1$ is a set $S$ with a
binary function $\cdot ^M:S\times S\to S$ and a preferred element
$e^M\in S$ which satisfy the conditions laid out in the axioms.
Once again, every such model $M$ carries all the structure that
$T_1$ requires of it, and then some more structure that $T_1$
doesn't care about. \end{example}
In order to precisely define the concept of a model of a theory, we
must first begin with the concept of a $\Sigma$-structure.
\begin{defn} A \emph{$\Sigma$-structure} $M$ is a mapping from
$\Sigma$ to appropriate structures in the category $\cat{Sets}$. In
particular $M$ fixes a particular set $S$, and then
\begin{itemize}
\item $M$ maps each $n$-ary relation symbol $p\in \Sigma$ to a
subset $M(p)\subseteq S^n=S\times \cdots \times S$.
\item $M$ maps each $n$-ary function symbol $f\in \Sigma$ to a
function $M(f):S^n\to S$.
\end{itemize} \label{sigstr}
\end{defn}
A $\Sigma$-structure $M$ extends naturally to all syntactic structures
built out of $\Sigma$. In particular, for each $\Sigma$-term $t$, we
define $M(t)$ to be a function, and for each $\Sigma$-formula $\phi$,
we define $M(\phi )$ to be a subset of $S^n$ (where $n$ is the number
of free variables in $\phi$). In order to do so, we need to introduce
several auxiliary constructions.
%% now contexts
\begin{defn} Let $\Gamma$ be a finite set of $\Sigma$-formulas. We
say that $\vec{x}=x_1,\dots ,x_n$ is a \emph{context} for $\Gamma$
just in case $\vec{x}$ is a duplicate-free sequence that contains
all free variables that appear in any of the formulas in $\Gamma$.
We say that $\vec{x}$ is a \emph{minimal context} for $\Gamma$ just
in case every variable $x_i$ in $\vec{x}$ occurs free in some
formula in $\Gamma$. Note: we also include, as a context for
sentences, the zero length string of variables. \end{defn}
\begin{defn} Let $\vec{x}$ and $\vec{y}$ be duplicate-free sequences
of variables. Then $\vec{x}.\vec{y}$ denotes the result of
concatenating the sequences, then deleting repeated variables in
order from left to right. Equivalently, $\vec{x}.\vec{y}$ results
from deleting from $\vec{y}$ all variables that occur in $\vec{x}$,
and then appending the resulting sequence to $\vec{x}$. \end{defn}
\begin{defn} For each term $t$, we define the \emph{canonical context}
$\vec{x}$ of $t$ as follows. First, for a variable $x$, the
canonical context is $x$. Second, suppose that for each term $t_i$,
the canonical context $\vec{x}_i$ has been defined. Then the
canonical context for $f(t_1,\dots ,t_n)$ is
$(\cdots ((\vec{x}_1.\vec{x}_2)\cdots ).\vec{x_n}$. \end{defn}
\begin{exercise} Suppose that $\vec{x}=x_1,\dots ,x_n$ is the
canonical context for $t$. Show that $FV(t)=\{ x_1,\dots
,x_n\}$. \end{exercise}
%% these contexts behave like multisets -- duplicates don't matter,
%% but order matters
\begin{defn} For each formula $\phi$, we define the \emph{canonical
context} $\vec{x}$ of $\phi$ as follows. First, if $\vec{x}_i$ is
the canonical context for $t_i$, then the canonical context for
$t_1=t_2$ is $\vec{x}_1.\vec{x}_2$, and the canonical context for
$p(t_1,\dots ,t_n)$ is
$(\cdots ((\vec{x}_1.\vec{x}_2)\cdots ).\vec{x_n}$. For the Boolean
connnectives, we also use the operation $\vec{x}_1.\vec{x}_2$ to
combine contexts. Finally, if $\vec{x}$ is the canonical context
for $\phi$, then the canonical context for $\forall x\phi$ is the
result of deleting $x$ from $\vec{x}$, if it occurs. \end{defn}
\begin{exercise} Show that the canonical context for $\phi$ does, in
fact, contain all and only those variables that are free in
$\phi$. \end{exercise}
If a $\Sigma$-structure $M$ has a domain set $S$, then it assigns
relation symbols to subsets of the Cartesian products,
\[ S,S\times S,S^3,\dots \] Of course, these sets are all connected to
each other by projection maps, such as the projection $S\times S\to S$
onto the first coordinate. We will now develop some apparatus to
handle these projection maps. To this end, let $[n]$ stand for the
finite set $\{ 1,\dots ,n\}$.
\begin{lemma} For each injective function $p:[m]\to [n]$, there is a
unique projection $\pi _p:S^n\to S^m$ defined by
\[ \pi _p\langle x_1,\dots ,x_n\rangle = \langle x_{p(1)},\dots
,x_{p(m)}\rangle . \] Furthermore, if $q:[\ell ]\to [m]$ is also
injective, then $\pi _{p\circ q}=\pi _q\circ \pi _p$. \end{lemma}
\begin{proof} The first claim is obvious. For the second claim, it's
easier if we ignore the variables $x_1,\dots ,x_n$ and note that
$\pi _p$ is defined by the coordinate projections:
\[ \pi _i\circ \pi _p \: = \: \pi _{p(i)} ,\] for $i=1,\dots ,m$.
Thus, in particular,
\[
\pi _i\circ \pi _q\circ \pi _p \: = \: \pi _{q(i)}\circ \pi _p
\: = \: \pi _{p(q(i))} \: = \: \pi _i\circ \pi _{p\circ q} ,\]
which proves the second claim.
\end{proof}
\begin{defn} Let $\vec{x}=x_1,\dots ,x_m$ and
$\vec{y}=y_1,\dots ,y_n$ be duplicate-free sequences of variables.
We say that $\vec{x}$ is a \emph{subcontext} of $\vec{y}$ just in
case each element in $\vec{x}$ occurs in $\vec{y}$. In other
words, for each $i\in [m]$, there is a unique $p(i)\in [n]$ such
that $x_i=y_{p(i)}$. Since $i\mapsto y_i$ is injective,
$p:[m]\to [n]$ is also injective. Thus, $p$ determines a unique
projection $\pi _p:S^n\to S^m$. We say that $\pi _p$ is the
\emph{linking projection} for contexts $\vec{y}$ and $\vec{x}$.
If $\vec{x}$ and $\vec{y}$ are canonical contexts of formulas or
terms, then we say that $\pi _p$ is the \emph{linking projection}
for these formulas or terms.
\end{defn}
We are now ready to complete the extension of the $\Sigma$-structure
$M$ to all $\Sigma$-terms.
\begin{defn} For each term $t$ with $n$-free variables, we define
$M(t):S^n\to S$.
\begin{enumerate}
\item Recall that a constant symbol $c\in \Sigma$ is really a
special case of a function symbol, viz.\ a $0$-ary function
symbol. Thus, $M(c)$ should be a function from $S^0$ to $S$.
Also recall that the $0$-ary Cartesian product of any set is a
one-point set $\{\ast\}$. Thus, $M(c):\{ \ast \}\to S$, which
corresponds to a unique element $c^M\in S$.
\item For each variable $x$, we let $M(x):S\to S$ be the identity
function. This might seem like a strange choice, but its utility
will soon be clear.
%% TO DO: use van oosten to get this defn straight
\item Let $t\equiv f(t_1,\dots ,t_n)$, where $M(t_i)$ has already
been defined. Let $n_i$ be the number of free variables in $t_i$.
The context for $t_i$ is a subcontext of the context for $t$.
Thus, there is a linking projection $\pi _i:S^n\to S^{n_i}$.
Whereas the $M(t_i)$ may have different domains (if
$n_i\neq n_j$), precomposition with the linking projections makes
them functions of a common domain $S^n$. Thus, we define
\[ M[f(t_1,\dots ,t_n)] \: = \: M(f)\circ \langle M(t_1)\circ \pi
_1,\dots ,M(t_n)\circ \pi _n \rangle , \] which is a function
from $S^n$ to $S$.
\end{enumerate} \end{defn}
We illustrate the definition of $M(t)$ with a couple of examples.
\begin{example} Suppose that $f$ is a binary function symbol, and
consider the two terms $f(x,y)$ and $f(y,x)$. The canonical context
for $f(x,y)$ is $x,y$, and the canonical context for $f(y,x)$ is
$y,x$. Thus, the linking projection for $f(x,y)$ and $x$ is the
projection $\pi _0:S\times S\to S$ onto the first coordinate; and
the linking projection for $f(y,x)$ and $x$ is
$\pi _1:S\times S\to S$ onto the second coordinate. Thus,
\[ M(f(x,y)) = M(f)\circ \langle \pi _0,\pi _1 \rangle = M(f) .\] A
similar calculation shows that $M(f(y,x))=M(f)$, which is as it
should be: $f(x,y)$ and $f(y,x)$ should correspond to the same
function $M(f)$.
However, it does \textit{not} follow that the formula
$f(x,y)=f(y,x)$ should be regarded as a semantic tautology.
Whenever we place both $f(x,y)$ and $f(y,x)$ into the \textit{same}
context, this context serves as a reference point by which the order
of inputs can be distinguished.
\end{example}
\begin{defn} For each formula $\phi$ of $\Sigma$ with $n$ distinct
free variables, we define $M(\phi )$ to be a subset of
$S^n=S\times \cdots\times S$.
\begin{enumerate}
\item $M(\bot )$ is the empty set $\emptyset$, considered as a subset
of the one-element set $1$.
\item Suppose that $\phi\equiv (t_1=t_2)$, where $t_1$ and $t_2$ are
terms. Let $n_i$ be the number of free variables in $t_i$. Since
the context for $t_i$ is a subcontext of that for $t_1=t_2$, there
is a linking projection $\pi _i:S^n\to S^{n_i}$. We define
$M(t_1=t_2)$ to be the equalizer of the functions
$M(t_1)\circ \pi _1$ and $M(t_2)\circ \pi _2$.
\item Suppose that $\phi\equiv p(t_1,\dots ,t_m)$, where $p$ is a
relation symbol and $t_1,\dots ,t_m$ are terms. Let $n$ be the
number of distinct free variables in $\phi$. Since the context of
$t_i$ is a subcontext of that of $\phi$, there is a linking
projection $\pi _i:S^n\to S^{n_i}$. Then
$\langle \pi _1,\dots ,\pi _m\rangle$ is a function from $S^n$ to
$S^{n_1}\times\cdots\times S^{n_m}$. We define
$M[p(t_1,\dots ,t_m)]$ to be the pullback of $M(p)\subseteq S^m$
along the function
\[ \langle M(t_1)\circ \pi _1,\dots ,M(t_m)\circ \pi _m\rangle .\]
%% TO DO: Boolean + Quantifiers
\item Suppose that $M$ has already been defined for $\phi$. Then we
define $M(\neg \phi )=S^n\backslash M(\phi )$.
\item Suppose that $\phi$ is a Boolean combination of
$\phi _1,\phi _2$, and that $M(\phi _1)$ and $M(\phi _2)$ have
already been defined. Let $\pi _i$ be the linking projection for
$\phi _i$ and $\phi$, and let $\pi _i^*$ be the corresponding
pullback (preimage) map that takes subsets to subsets. Then we define
\[ \begin{array}{l l l} M(\phi _1\wedge\phi _2 ) & = & \pi
_1^*(M(\phi
_1))\cap \pi _2^*(M(\phi _2)) , \\
M(\phi _1\vee\phi _2) & = & \pi _1^*(M(\phi _1))\cup \pi
_2^*(M(\phi
_2)) , \\
M(\phi _1\to \phi _2) & = & (S^n\backslash \pi _1^*(M(\phi
_1)))\cup \pi _2^*(M(\phi _2)) . \end{array} \]
\item Suppose that $M(\phi )$ is already defined as a subset of
$S^n$. Suppose first that $x$ is free in $\phi$, and let $\pi
:S^{n+1}\to S$ be the linking projection for $\phi$ and $\exists
x\phi$. Then we define $M(\exists x\phi )$ to be the image of
$M(\phi )$ under $\pi$, i.e.,
\[ M(\exists x\phi )\: = \: \{ \vec{a}\in S^n\mid \pi
^{-1}(\vec{a})\cap M(\phi )\neq \emptyset \} .\] If $x$ is not
free in $\phi$, then we define $M(\exists x\phi )=M(\phi )$.
Similarly, if $x$ is free in $\phi$, then we define
\[ M(\forall x\phi ) \: = \: \{ \vec{a}\in S^n \mid \pi
^{-1}(\vec{a})\subseteq M(\phi ) \} . \] If $x$ is not free in
$\phi$, then we define $M(\forall x\phi )=M(\phi )$.
\end{enumerate} \end{defn}
\begin{example} Let's unpack the definitions of $M(x=y)$ and $M(x=x)$.
For the former, the canonical context for $x=y$ is $x,y$. Thus, the
linking projection for $x=y$ and $x$ is $\pi _0:S\times S\to S$ onto
the first coordinate, and the linking projection for $x=y$ and $y$
is $\pi _1:S\times S\to S$ onto the second coordinate. By
definition, $M(x)\equiv 1_S\equiv M(y)$, and $M(x=y)$ is the
equalizer of $1_s\circ \pi _0$ and $1_S\circ \pi _1$. This
equalizer is clearly the diagonal subset of $S\times S$:
\[ M(x=y) \: \equiv \: \{ \langle a,b\rangle \in S\times S \mid a=b
\} \: \equiv \: \{ \langle a,a\rangle \mid a\in S \} .\] In
contrast, the canonical context for $x=x$ is $x$, and the linking
projection for $x=x$ and $x$ is simply the identity. Thus, $M(x=x)$
is defined to be the equalizer of $M(x)$ and $M(x)$, which is the
entire set $S$. That is, $M(x=x)\equiv S$.
\end{example}
\begin{exercise} Describe $M(f(x,y)=f(y,x))$, and explain why it won't
neccesarily be the entire set $S\times S$. \end{exercise}
We are now going to define a relation $\phi\vDash _M\psi$ of semantic
entailment in a structure $M$; and we will use that notion to define
the absolute relation $\phi\vDash\psi$ of semantic entailment. (In
short: $\phi\vDash\psi$ means that $\phi \vDash _M\psi$ in every
structure $M$.) Here $\phi$ and $\psi$ are formulas (not necessarily
sentences), so we need to take a bit of care with their free
variables. One thing we could do is to consider the sentence
$\forall \vec{x}(\phi\to\psi )$, where $\vec{x}$ is any sequence that
includes all variables free in $\phi$ or $\psi$. However, even in
that case, we would have to raise a question about whether the
definition depends on the choice of the sequence $\vec{x}$. Since we
have to deal with that question in any case, we will instead look more
directly at the relation between the formulas $\phi$ and $\psi$, which
might share some free variables in common.
As a first proposal, we might try saying that $\phi\vDash _M\psi$ just
in case $M(\phi )\subseteq M(\psi )$. But the problem with this
proposal is that $M(\phi )$ and $M(\psi )$ are typically defined to be
subsets of different sets. For example: the definition of $\vDash _M$
should imply that $p(x)\vDash _M(p(x)\vee q(y))$. However, for any
$\Sigma$-structure $M$, $M(p(x))$ is a subset of $S$ whereas
$M(p(x)\vee q(y))$ is a subset of $S\times S$. The way to fix this
problem is to realize that $M(p(x))$ can also be considered to be a
subset of $S\times S$. In particular, $p(x)$ is equivalent to
$p(x)\wedge (y=y)$, and intuitively $M(p(x)\wedge (y=y))$ should the
subset of $S\times S$ of things satisfying $p(x)$ and $y=y$. In other
words, $M(p(x)\wedge (y=y))$ should be $M(p(x))\times S$.
Here's what we will do next. First we will extend the definition of
$M$ so that it assigns a formula $\phi$ an extension
$M_{\vec{x}}(\phi )$ relative to a context $\vec{x}$. Then we will
define $\phi\vDash _M\psi$ to mean that
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$, where $\vec{x}$ is
an arbitrarily chosen context for $\phi ,\psi$. Then we will show
that this definition does not depend on which context we chose.
In order to define $M_{\vec{y}}(\phi )$ where $\vec{y}$ is an
arbitrary context for $\phi$, we will first fix the canonical context
$\vec{x}$ for $\phi$, and we will set $M_{\vec{x}}(\phi )=M(\phi )$.
Then for any other context $\vec{y}$ of which $\vec{x}$ is a
subcontext, we will use the linking projection $\pi _p$ to define
$M_{\vec{y}}(\phi )$ as a pullback of $M_{\vec{x}}(\phi )$.
\begin{defn} Let $\vec{y}=y_1,\dots ,y_n$ be a context for $\phi$, let
$\vec{x}=x_1,\dots ,x_m$ be the canonical context for $\phi$, and
let $p:[m]\to [n]$ be the corresponding injection. We define
$M_{\vec{y}}(\phi )$ to be the pullback of $M(\phi )$ along
$\pi _p$. In particular, when $\vec{y}=\vec{x}$, then
$p:[n]\to [n]$ is the identity, and $M_{\vec{x}}(\phi )=M(\phi
)$. \end{defn}
Now we are ready to define the relation $\phi\vDash _M\psi$.
\begin{defn} For each pair of formulas $\phi ,\psi$, let $\vec{x}$ be
the canonical context for $\phi\to\psi$. We say that
$\phi\vDash _M\psi$ just in case
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$.
\end{defn}
We will now show that the definition of $\phi\vDash _M\psi$ is
independent of the chosen context $\vec{x}$ for $\phi ,\psi$. In
particular, we show that for any two contexts $\vec{x}$ and $\vec{y}$
for $\phi ,\psi$, we have
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$ if and only if
$M_{\vec{y}}(\phi )\subseteq M_{\vec{y}}(\psi )$. As the details of
this argument are a bit tedious, the impatient reader may wish to skip
to Definition \ref{entailment}.
We'll first check the compatibility of the definitions
$M_{\vec{y}}(\phi )$ and $M_{\vec{z}}(\phi )$, where $\vec{y}$ and
$\vec{z}$ are contexts for $\phi$.
\begin{lemma} Suppose that $\vec{x}=x_1,\dots ,x_\ell$ is a subcontext
of $\vec{y}=y_1,\dots ,y_m$, and that $\vec{y}$ is a subcontext of
$\vec{z}=z_1,\dots ,z_n$. Suppose that $p:[\ell ]\to [m]$,
$q:[m]\to [n]$, and $r:[\ell ]\to [n]$ are the corresponding
injections. Then $r=q\circ p$. \end{lemma}
\begin{proof} By definition of $p$, $y_{p(i)}=x_i$ for $i\in [\ell ]$.
By definition of $r$, $z_{r(i)}=x_i$ for $i\in [\ell ]$. Thus,
$y_{p(i)}=z_{r(i)}$. Furthermore, by definition of $q$,
$z_{q(p(i))}=y_{p(i)}$. Therefore, $z_{q(p(i))}=z_{r(i)}$, and
$q(p(i))=r(i)$. \end{proof}
\begin{lemma} Suppose that $\vec{x}$ is a context for $\phi$, and that
$\vec{x}$ is a subcontext of $\vec{y}$. Let $\pi ^r:S^n\to S^m$ be
the projection connecting the contexts $\vec{y}$ and $\vec{x}$.
Then $M_{\vec{y}}(\phi )$ is the pullback of $M_{\vec{x}}(\phi )$
along $\pi _r$. \label{rolig} \end{lemma}
\begin{proof} Let $\pi _p$ be the projection connecting $\vec{x}$ to
the canonical context for $\phi$, and let $\pi _q$ be the projection
connecting $\vec{y}$ to the canonical context for $\phi$. Thus,
$M_{\vec{x}}(\phi )=\pi _p^*[M(\phi )]$, where $\pi _p^*$ denotes
the operation of pulling back along $\pi _p$. Similarly,
$M_{\vec{y}}(\phi )=\pi _q^*[M(\phi )]$. Furthermore,
$\pi _q=\pi _p\circ \pi _r$, and since pullbacks commute, we have
\[ M_{\vec{y}}(\phi )=\pi _q^*[M(\phi )] = \pi _r^*[\pi _p^*[M(\phi
)]] = \pi _r^*[M_{\vec{x}}(\phi )] ,\] as was to be
shown. \end{proof}
\begin{prop} Suppose that $\vec{x}$ is a context for $\phi ,\psi$, and
that $\vec{x}$ is a subcontext of $\vec{y}$. If
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$ then
$M_{\vec{y}}(\phi )\subseteq M_{\vec{y}}(\psi
)$. \label{subcon} \end{prop}
\begin{proof} Suppose that
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$. Let
$\pi _r:S^n\to S^m$ be the projection connecting the contexts
$\vec{y}$ and $\vec{x}$. By the previous lemma,
$M_{\vec{y}}(\phi )=\pi _r^*[M_{\vec{x}}(\phi )]$ and
$M_{\vec{y}}(\psi )=\pi _r^*[M_{\vec{x}}(\psi )]$. Since pullbacks
preserve set inclusion,
$M_{\vec{y}}(\phi )\subseteq M_{\vec{y}}(\psi )$. \end{proof}
Since we defined $\phi\vDash _M\psi$ using a minimal context $\vec{x}$
for $\phi ,\psi$, we now have the first half of our result: if
$\phi\vDash _M\psi$ then
$M_{\vec{y}}(\phi )\subseteq M_{\vec{y}}(\psi )$ for any context
$\vec{y}$ for $\phi ,\psi$. To complete the result, we now show that
redundant variables can be deleted from contexts.
\begin{lemma} Let $\vec{x}$ be a context for $\phi$, and suppose that
$y$ does not occur in $\vec{x}$. Then
$M_{\vec{x}.y}(\phi )=M_{\vec{x}}(\phi )\times S$. \end{lemma}
\begin{proof} Let $\vec{x}=x_1,\dots ,x_n$, and let $p:[n]\to [n+1]$
be the injection corresponding to the inclusion of $\vec{x}$ in
$\vec{x}.y$. In this case, $p(i)=i$ for $i=1,\dots ,n$, and
$\pi _p:S^{n+1}\to S^n$ projects out the last coordinate. By Lemma
\ref{rolig}, $M_{\vec{x}.y}(\phi )$ is the pullback of
$M_{\vec{x}}(\phi )$ along $\pi _p$. However, the pullback of any
set $A$ along $\pi _p$ is simply $A\times S$. \end{proof}
Now suppose that $M_{\vec{x}.y}(\phi )\subseteq M_{\vec{x}.y}(\psi )$,
where $\vec{x}$ is a context for $\phi ,\psi$, and $y$ does not occur
in $\vec{x}$. Then the previous lemma shows that
$M_{\vec{x}.y}(\phi )=M_{\vec{x}}(\phi )\times S$ and
$M_{\vec{x}.y}(\psi )=M_{\vec{x}}(\psi )\times S$. Thus,
$M_{\vec{x}.y}(\phi )\subseteq M_{\vec{x}.y}(\psi )$ if and only if
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$. A quick inductive
argument then shows that any number of appended empty variables makes
no difference.
We can now conclude the argument that
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$ if and only if
$M_{\vec{y}}(\phi )\subseteq M_{\vec{y}}(\psi )$, where $\vec{x}$ is a
subcontext of $\vec{y}$. The ``if'' direction was already shown in
Prop.\ \ref{subcon}. For the ``only if'' direction, suppose that
$M_{\vec{y}}(\phi )\subseteq M_{\vec{y}}(\psi )$. First use Prop.\
\ref{subcon} again to move any variables not in $\vec{x}$ to the end
of the sequence $\vec{y}$. (Recall that $\vec{y}$ is a subcontext of
any permutation of $\vec{y}$.) Then use the previous lemma to
eliminate these variables. The resulting sequence is a permutation of
$\vec{x}$, hence a subcontext of $\vec{x}$. Finally, use Prop.\
\ref{subcon} one more time to show that
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$. Thus, we have shown
that the definition of $\phi\vDash _M\psi$ is independent of the
context chosen for $\phi ,\psi$.
\begin{defn} We say that $\phi$ \emph{semantically entails} $\psi$,
written $\phi\vDash\psi$, just in case $\phi \vDash _M\psi$ for
every $\Sigma$-structure $M$. We write $\vDash \psi$ as
shorthand for $\top\vDash\psi$. \label{entailment} \end{defn}
\begin{note} The canonical context $\vec{x}$ for the pair
$\{\top ,\phi \}$ is simply the context for $\phi$. By definition,
$M_{\vec{x}}(\top )$ is the pullback of $1$ along the unique map
$\pi :S^n\to 1$. Thus, $M_{\vec{x}}(\top )=S^n$, and
$\top\vDash _M\phi$ if and only if $M(\phi )=S^n$.
\end{note}
We're now ready for two of the most famous definitions in
mathematical philosophy.
\begin{box-thm}[Truth in a structure]
A sentence $\phi$ has zero free variables. In this case, $M(\phi )$
is defined to be a subset of $S^0=1$, a one-element set. We say
that $\phi$ is \emph{true} in $M$ if $M(\phi )=1$, and we say that
$\phi$ is \emph{false} in $M$ if $M(\phi
)=\emptyset$. \end{box-thm}
\begin{box-thm}[Model] Let $T$ be a theory in signature $\Sigma$, and
let $M$ be a $\Sigma$-structure. We say that $M$ is a \emph{model}
of $T$ just in case: for any sentence $\phi$ of $\Sigma$, if
$T\vdash\phi$ then $M(\phi )=1$. \end{box-thm}
\section{The semantic view of theories}
In Chapter \ref{ch:syntax}, we talked about how Rudolf Carnap used
syntactic metalogic to explicate the notion of a scientific theory.
By the 1960s, people were calling Carnap's picture the ``syntactic
view of theories'', and they were saying that something was
fundamentally wrong with it. According to \cite{suppe2000}, the
syntactic view of theories died in the late 1960s (March 26, 1969, to
be precise) after having met with an overwhelming number of objections
in the previous two decades. Upon the death of the syntactic view, it
was unclear where philosophy of science would go. Several notable
philosophers --- such as Feyerabend and Hanson --- wanted to push
philosophy of science away from formal analyses of theories. However,
others such as Patrick Suppes, Bas van Fraassen, and Fred Suppe saw
formal resources for philosophy of science in other branches of
mathematics, most particularly set theory and model theory. Roughly
speaking, the ``semantic view of theories'' designates proposals to
explicate theory-hood by means of semantic metalogic.
We now have the technical resources in place to state a preliminary
version of the semantic view of theories:
\begin{quote}
(SV) A scientific theory is a class of $\Sigma$-structures for some
signature~$\Sigma$.
\end{quote}
Now, proponents of the semantic view will balk at SV for a couple of
different reasons. First, semanticists stress that a scientific
theory has two components:
\begin{enumerate}
\item A theoretical definition; and
\item A theoretical hypothesis.
\end{enumerate}
The theoretical definition, roughly speaking, is intended to replace
the first component of Carnap's view of theories. That is, the
theoretical definition is intended to specify some abstract
mathematical object --- the thing that will be used to do the
representing. Then the theoretical hypothesis is some claim to the
effect that some part of the world can be represented by the
mathematical object given by the theoretical definition. So, to be
clear, SV here is only intended to give one half of a theory, viz.\
the theoretical definition. I am not speaking yet about the
theoretical hypothesis.
But proponents of the semantic view will balk for a second reason: SV
makes reference to a signature $\Sigma$. And one of the supposed
benefits of the semantic view was to free us from the language
dependence implied by the syntactic view. So, how are we to modify SV
in order to maintain the insight that a scientific theory is
independent of the language in which it is formulated?
I will give two suggestions, the first of which I think cannot
possibly succeed. The second suggestion works; but it shows that the
semantic view actually has no advantage over the syntactic view in
being ``free from language dependence.''
How then to modify SV? The first suggestion is to formulate a notion
of mathematical structure that makes no reference to language. At
first glance, it seems simple enough to do so. The paradigm case of a
mathematical structure is supposed to be an ordered $n$-tuple
$\langle X,R_1,\dots R_n\rangle$, where $X$ is a set, and
$R_1,\dots ,R_n$ are relations on $X$. (This notion of mathematical
structure follows in the footsteps of \cite{bourbaki}, which
incidentally, has been rendered obsolete by category theory.)
Consider, for example, the proposal made by Lisa Lloyd:
\begin{quote}
In our discussion, a {\it model} is not such an interpretation
[i.e.\ not an $\Sigma$-structure], matching statements to a set of
objects which bear certain relations among themselves, but the set
of objects itself. That is, models should be understood as
structures; in the cases we shall be discussing, they are
mathematical structures, i.e., a set of mathematical objects
standing in certain mathematically representable
relations. \citep[p.\ 30]{lloyd}
\end{quote}
However, it's difficult to make sense of this proposal. Consider the
following example.
\begin{example} Let $a$ be an arbitrary set, and consider the
following purported example of a mathematical structure:
\[ M \: = \: \Bigl\langle \{ a,b,\langle a,a\rangle \} ,\{ \langle
a,a\rangle \} \Bigr\rangle .\] That is, the domain $X$ consists of
three elements $a,b,\langle a,a\rangle$, and the indicated structure
is the singleton set containing $\langle a,a\rangle$. But how are
we supposed to understand this structure? Are we supposed to
consider $\{ \langle a,a\rangle \}$ to be a subset of $X$, or as a
subset of $X\times X$? The former is a structure for a signature
$\Sigma$ with a single unary predicate symbol; the latter is a
structure for a signature $\Sigma '$ with a single binary relation
symbol. In writing down $M$ as an ordered $n$-tuple, we haven't yet
fully specified an intended mathematical structure.
We conclude then that a mathematical structure cannot simply be, ``a
set of mathematical objects standing in certain mathematically
representable relations.'' To press the point further, consider
another purported example of a mathematical structure:
\[ N \: = \: \Bigl\langle \{ a,b,\langle a,b\rangle \} ,\{ \langle a,b\rangle \}
\Bigr\rangle .\] Are $M$ and $N$ isomorphic structures? Once again, the
answer is underdetermined. If $M$ and $N$ are supposed to be
structures for a signature $\Sigma$ with a single unary predicate
symbol, then the answer is Yes. If $M$ and $N$ are supposed to be
structures for a signature $\Sigma '$ with a single binary relation
symbol, then the answer is No. \end{example}
Thus, it's doubtful that there is any ``language-free'' account of
mathematical structures, and hence no plausible language-free semantic
view of theories. I propose then that we embrace the fact that we are
``suspended in language'', to borrow a phrase from Niels Bohr. To
deal with our language-dependence, we need to consider notions of
equivalence of theory-formulations --- so that the same theory to be
formulated in different languages. And note that this stratagem is
available for both semantic and syntactic views of theories. Thus,
``language independence'' is not a genuine advantage of the semantic
view of theories as against the syntactic view of theories.
\begin{box-thm}[Philosophical Moral] It is of {\it crucial} importance
that we do not think of a $\Sigma$-structure $M$ as representing the
world. To say that the world is isomorphic to, or even partially
isomorphic to, or even similar to, $M$, would be to fall into a
profound confusion.
A $\Sigma$-structure $M$ is {\it not} a ``set-theoretic structure''
in any direct sense of that phrase. Rather, $M$ is a function whose
domain is $\Sigma$ and whose range consists of some sets, subsets,
and functions between them. If one said that ``$M$ represents the
world,'' then one would be saying that the world is represented by a
mathematical object of type $\Sigma\to\cat{Sets}$. Notice, in
particular, that $M$ has ``language'' built into its very
definition. \end{box-thm}
%% It should be the case that if $T$ is defined in terms of sequences,
%% then this defn is equivalent to saying that M(\phi )\subseteq
%% M(\psi )$ for all sequents
%% TO DO: examples of models
%% TO DO: Should we prove soundness?
\section{Soundness, completeness, compactness}
We now prove versions of four central metalogical results: soundness,
completeness, compactness, and L\"ownheim-Sk{\"olem} theorems. For
these results, we will make a couple of simplifying assumptions,
merely for the sake of mathematical elegance. We will assume that
$\Sigma$ is fixed signature that is countable, and that has no
function symbols. This assumption will permit us to use the
topological techniques introduced by \cite{rasiowa}.
%% Baire category theorem
\subsection*{Soundness}
In its simplest form, the soundness theorem shows that for any
sentence $\phi$, if $\phi$ is provable ($\top\vdash\phi$) then $\phi$
is true in all $\Sigma$-structures ($\top\vDash\phi$). Inspired by
categorical logic, we derive this version of soundness as a special
case of a more general result for $\Sigma$-formulas. We show that:
for any $\Sigma$-formulas $\phi$ and $\psi$, and for any context
$\vec{x}$ for $\{ \phi ,\psi \}$, if $\phi\vdash _{\vec{x}}\psi$ then
$M_{\vec{x}}(\phi )\subseteq M_{\vec{x}}(\psi )$.
The proof proceeds by induction on the construction of proofs, i.e.\
over the definition of the relation $\vdash$. Most cases are trivial
verifications, and we leave them to the reader. We will just consider
the case of the existential elimination rule, which we consider in the
simple form:
\[ \begin{array}{c c c} \phi\vdash _{x,y}\psi \\ \hline \exists
y\phi\vdash _x \psi \end{array} \] assuming that $y$ is not free
in $\psi$. We assume then that the result holds for the top line,
i.e.\ $M_{x,y}(\phi )\subseteq M_{x,y}(\psi )$. By definition,
$M_x(\exists y\phi )$ is the image of $M_{x,y}(\phi )$ under the
projection $X\times Y\to X$. And since $y$ is not free in $\psi$,
$M_{x,y}(\psi )=M_x(\psi )\times Y$.
To complete the argument, it will suffice to make the following
general observation about sets: If $A\subseteq X\times Y$ and
$B\subseteq X$, then the following inference is valid:
\[ \begin{array}{l l l}
A\subseteq \pi ^{-1}(B) \\ \hline
\pi (A)\subseteq B
\end{array} \]
Indeed, suppose that $z\in \pi (A)$, which means that there is a $y\in
Y$ such that $\langle z,y\rangle \in A$. By the top line, $\langle
z,y\rangle\in\pi ^{-1}(B)$, which means that $z=\pi \langle
z,y\rangle \in B$. Now set $A=M_{x,y}(\phi )$ and $B=M_x(\psi )$,
and it follows that existential elimination is sound.
We leave the remaining steps of this proof to the reader, and briefly
comment on the philosophical significance (or lack thereof) of the
soundness theorem.\footnote{The discussion here borrows from the
ideas of Michaela McSweeney. See \cite{mmm}.} Philosophers often
gloss this theorem as showing that the derivation rules are ``safe'',
i.e.\ that they don't permit derivations which are not valid, or even
more strongly, that the rules won't permit us to derive a false
conclusion from true premises. But now we have a bit of a
philosophical conundrum. What is this standard of validity against
which we are supposed to measure $\vdash$? Moreover, why think that
this other standard of validity is epistemologically prior to the
standard of validity we have specified with the relation $\vdash$?
Philosophers often gloss the relation $\vDash$ in terms of ``truth
preservation.'' They say that $\vp\vDash\psi$ \textit{means that}
whenever $\vp$ is true, then $\psi$ is true. Such statements can be
highly misleading, if they cause the reader to think that $\vDash$ is
the intuitive notion of truth preservation. No, the relation
$\vDash$ is yet another attempt to capture, in a mathemtically
precise fashion, our intuitive notion of logical consequence. We
have two distinct ways of representing this intuitive notion: the
relation $\vdash$ and the relation $\vDash$. The soundness and
completeness theorems happily show that we've captured the same
notion with two different definitions.
The important point here is that: {\it logical syntax and logical
semantics are enterprises of the same kind}. The soundness and
completeness theorems are not theorems about how mathematics relates
to the world, nor are they theorems about how a mathematical notion
relates to an intuitive notion. No, these theorems demonstrate a
relationship between mathematical things.
The soundness theorem has sometimes been presented as an ``absolute
consistency'' result, i.e.\ that the predicate calculus is consistent
\textit{tout court}. But such presentations are misleading: The
soundness theorem shows only that the predicate calculus is consistent
relative to the relation $\vDash$, i.e.\ that the relation $\vdash$
doesn't exceed the relation $\vDash$. It doesn't prove that there is
no sentence $\vp$ such that $\vDash \vp$ and $\vDash \neg \vp$. We
agree, then, with David Hilbert: the only kind of formal consistency
is relative consistency.
\subsection*{Completeness}
In Chapter \ref{cat-prop}, we saw that the completeness theorem for
propositional logic is equivalent to the Boolean ultrafilter axiom
(i.e.\ every nonzero element in a Boolean algebra is contained in an
ultrafilter). In many textbooks of logical metatheory, the
completeness theorem for predicate logic uses Zorn's lemma, which is a
variant of the axiom of choice (AC). It is known, however, that the
completeness theorem does not require the full strength of AC. The
proof we give here uses the Baire category theorem, which is derivable
in ZF with the addition of the axiom of dependent choices, a slightly
weaker choice principle. (Exercise: can you see where in the proof we
make use of a choice principle?)
%% TO DO: check that we've defined all the vocabulary used here
\begin{thm}[Baire Category Theorem] Let $X$ be a compact Hausdorff
space, and let $U_1,U_2,\dots $ be a countable family of sets, all
of which are open and dense in $X$. Then
$\bigcap _{i=1}^{\infty}U_i$ is dense in $X$. \end{thm}
\begin{proof} Let $U=\bigcap _{i=1}^{\infty}U_i$, and let $O$ be a
nonempty open subset of $X$. We need only show that $O\cap U$ is
nonempty. To this end, we inductively define a family $O_i$ of open
subsets of $X$ as follows:
\begin{itemize}
\item $O_1 =O\cap U_1$, which is open, and nonempty since $U_1$ is
dense;
\item Assuming that $O_n$ is open and nonempty, it has nonempty
intersection with $U_{n+1}$, since the latter is dense. But any
point $x\in O_n\cap U_{n+1}$ is contained in a neighborhood
$O_{n+1}$ such that $O_{n+1}\subseteq U_{n+1}$, and
$\overline{O}_{n+1}\subseteq O_n$, using the regularity of $X$.
\end{itemize}
It follows then that the collection $\{ \overline{O}_i:i\in \7N \}$
satisfies the finite intersection property. Since $X$ is compact,
there is a $p$ in $\bigcap _{i=1}^{\infty}\overline{O}_i$. Since
$\overline{O}_{i+1}\subseteq O_i$, it also follows that $p\in
O_i\subseteq U_i$, for all $i$. Therefore, $O\cap U$ is nonempty.
\end{proof}
Our proof of the completess theorem for predicate logic is similar in
conception to the proof for propositional logic. First we construct a
Boolean algebra $B$ of provably-equivalent formulas. Using the
definition of $\vdash$, it is not difficult to see that the
equivalence relation is compatible with the Boolean operations. Thus,
we may define Boolean operations as follows:
\[ {[}\phi ]\cap {[}\psi ] = {[}\phi\wedge\psi ] ,\qquad {[}\phi
]\cup {[}\psi ] = {[}\phi\vee\psi ] , \qquad - {[}\phi ] = {[}\neg
\phi ] . \] If we let $0=[\bot ]$ and $1=[\top ]$, then it's easy to
see that $\langle B,0,1,\cap ,\cup ,-\rangle $ is a Boolean algebra.
Now we want to show that if $\phi$ is not provably equivalent to a
contradition, then there is a $\Sigma$-structure $M$ such that
$M(\phi )$ is not empty. In the case of propositional logic, it was
enough to show that there is a homomorphism $f:B\to 2$ such that
$f(\phi )=1$. But that won't suffice for predicate logic, because
once we have this homomorphism $f:B\to 2$, we need to use it to build
a $\Sigma$-structure $M$, and to show that $M(\phi )$ is not empty.
As we will now see, to ensure that $M(\phi )$ is not empty, we must
choose a homomorphism $f:B\to 2$ that is ``continuous on
existentials''.
\begin{defn} Let $f:B\to 2$ be a homomorphism. We say that $f$ is
\emph{smooth on existentials} just in case for each formula $\psi$,
if $f(\exists x \psi )=1$ then $f(\psi [x_i/x])=1$ for some
$i\in \7N$. \end{defn}
We will see now that these ``smooth on existentials'' homomorphisms
are dense in the Stone space $X$ of $B$. In fact, the argument here
is quite general. We first show that for any particular convergent
family $a_i\to a$ in a Boolean algebra, the set of non-smooth
homomorphisms is closed and has empty interior. By saying that
$a_i\to a$ is convergent, we mean that $a_i\leq a$ for all $i$, and
for any $b\in B$, if $a_i\leq b$ for all $i$, then $a\leq b$. That
is, $a$ is the least upper bound of the $a_i$.
Let's say that a homomorphism $f:B\to 2$ is \emph{smooth} relative to
the convergent family $a_i\to a$ just in case $f(a_i)\to f(a)$ in the
Boolean algebra $2$. Now let $D$ be the set of homomorphisms
$f:B\to 2$ such that $f$ is \textit{not} smooth on $a_i\to a$. We
will show that $D$ is a closed subset of $X$ with empty interior. Any
homomorphism $f:B\to 2$ preserves order, and hence $f(a_i)\leq f(a)$
for all $i$. Thus, if $f(a_i)=1$ for any $i$, then $f$ is smooth on
$a_i\to a$. It follows that
\[ D \: = \: E_a \cap \left[ \bigcap _{i\in I}E_{-a_i} \right] .\] As
an intersecton of closed sets, $D$ is closed. To see that $D$ has
empty interior, suppose that $f\in E_b\subseteq D$, where $E_b$ is a
basic open subset of $X$. Then we have $E_b\subseteq E_{-a_i}$, which
implies that $a_i\leq \neg b$; and since $a_i\leq a$, we have
$a_i\leq a\wedge \neg b$. Thus, $a\wedge \neg b$ is an upper bound
for the family $\{ a_i\}$. Moreover, if $a=a\wedge \neg b$ then
$a\wedge b=0$ in contradiction with the fact that $f(a\wedge b)=1$.
Therefore $a$ is not the upper bound of $\{ a_i\}$, a contradiction.
We conclude that $D$ contains no basic open subsets, and hence it has
empty interior.
Now, this general result about smooth homomorphisms is of special
importance for the Boolean algebra of equivalence classes of formulas.
For in this case, existential formulas are the least upper bound of
their instances.
\begin{lemma} Let $\phi$ be a $\Sigma$-formula, and let $I$ be the set
of indices such that $x_i$ does not occur free in $\phi$. Then in
the Lindenbaum algebra, $E_{(\exists x\phi )}$ is the least upper
bound of $\{ E_{(\phi [x_i/x])} \mid i\in I \}$. \end{lemma}
\begin{proof} For simplicity, set $E=E_{(\exists x\phi )}$ and
$E_i=E_{(\phi [x_i/x])}$. The $\exists$-intro rule shows that
$E_i\leq E$. Now suppose that $E_\psi \in B$ such that
$E_i\leq E_\psi $ for all $i\in \7N$. That is,
$\phi [x_i/x]\vdash \psi$ for all $i\in I$. Since $\phi$ and $\psi$
have a finite number of free variables, there is some $i\in I$ such
that $x_i$ does not occur free in $\psi$. By the $\exists$-elim
rule, $\exists x_i\phi [x_i/x]\vdash \psi$. Since $x_i$ does not
occur free in $\phi$, $\exists x_i\phi [x_i/x]$ is equivalent to
$\exists x\phi$. Thus, $\exists x\phi\vdash \psi$, and
$E\leq E_\psi$. Therefore, $E$ is the least upper bound of
$\{ E_i \mid i\in I \}$. \end{proof}
Thus, for each existential $\Sigma$-formula $\phi$, the clopen set
$E_\phi$ is the union of the clopen subsets corresponding to the
instances of $\phi$, plus the meager set $D_\phi$ of homomorphisms
that are not smooth relative to $\phi$. Since the signature $\Sigma$
is countable, there are countably many such existential formulas, and
countably many of these sets $D_\phi$ of non-smooth homomorphisms.
Since each $D_\phi$ is meager, the Baire category theorem entails
that their union also is meager. Thus, the set $U$ of homomorphisms
that are smooth on {\it all} existentials is open and dense in the
Stone space $X$.
We are now ready to continue with the completeness theorem. Let
$\phi$ be our arbitrary formula that is not provably equivalent to a
contradiction. We know that the set $E_\phi$ of homomorphisms
$f:B\to 2$ such that $f([\phi ])=1$ is open and non-empty. Hence,
$E_\phi$ has non-empty intersection with $U$. Let
$f\in E_\phi\cap U$. That is, $f([\phi ])=1$, and $f$ is smooth on
all existentials. We now use $f$ to define a $\Sigma$-structure $M$.
\begin{itemize}
\item Let the domain $S$ of $M$ be the set of natural numbers.
\item For an $n$-ary relation symbol $R\in \Sigma$, let
$\vec{a}\in M(R)$ if and only if $f(R(x_{a_1},\dots
,x_{a_n}))=1$. \end{itemize}
% \begin{tomt} For the following proof, we will need to use a
% slightly
% stronger version of induction on the construction of formulas.
% In
% our original recipe, the step for quantifiers goes like this:
% \begin{quote} Assume R for $\phi$; show R for $\exists x
% \phi$. \end{quote} The stronger version goes like this:
% \begin{quote} Assume R for $\phi [y/x]$, for all variables $y$
% that do not occur free in $\phi$; show R for $\exists x
% \phi$. \end{quote} We claim that this stronger rule is also
% valid. Indeed, the weaker rule is implicitly schematic: if the
% result shown doesn't depend on the specific choice of variable
% $x$, then what we really have shown is that
% \[ R(\phi [y/x])\Longrightarrow R(\exists y\phi ) ,\]
% for any variable $y$ that doesn't occur free in $\phi$.
% \end{tomt}
\begin{lemma} For any $\Sigma$-formula $\phi$ with canonical context
$x_{c_1},\dots ,x_{c_n}$, if $f(\phi )=1$ then
$\vec{c}\in M(\phi )$. \end{lemma}