-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathpred-metatheory2.tex
814 lines (694 loc) · 37.3 KB
/
pred-metatheory2.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
\documentclass[11pt,fleqn]{article}
\usepackage{array,multirow,amsthm,amsmath,amssymb,url}
% \usepackage[osf,sc]{mathpazo}
% Helvetica for sans serif
% (scaled to match size of Palatino)
% \usepackage[scaled=0.90]{helvet}
% Bera Mono for monospaced
% (scaled to match size of Palatino)
% \usepackage[scaled=0.85]{beramono}
% \usepackage{fullpage}
\newcommand{\RA}{\vdash}
\usepackage{mVersion}
\usepackage{natbib}
\usepackage{bussproofs}
\usepackage{tikz}
\usepackage{tikz-cd}
\usetikzlibrary{trees}
\title{First-order logic in syntactic approach}
\author{Hans Halvorson}
\date{\today}
% \setlength{\parindent}{0em}
% \setlength{\parskip}{1em}
\swapnumbers
\newtheorem{prop}{Proposition}
\newtheorem{thm}[prop]{Theorem}
\newtheorem{cor}[prop]{Corollary}
\newtheorem{lemma}[prop]{Lemma}
\newtheorem{claim}[prop]{Claim}
\newtheorem{fact}[prop]{Fact}
\newtheorem*{subthm}{Substitution Theorem}
\newtheorem*{rthm}{Replacement Theorem}
\newtheorem{conj}[prop]{Conjecture}
\theoremstyle{definition}
\newtheorem*{defn}{Definition}
\newtheorem{exercise}[prop]{Exercise}
\newtheorem*{example}{Example}
\theoremstyle{remark}
\newtheorem*{note}{Note}
\newtheorem*{disc}{Discussion}
\swapnumbers
% \usepackage{fbb}
\newcommand{\df}[1]{\textbf{#1}}
\usepackage{mathrsfs}
\newcommand{\2}{\mathscr}
\renewcommand{\emph}{\textbf}
\newcommand{\vp}{\varphi}
\usepackage[framemethod=TikZ]{mdframed}
\newcounter{axi}\setcounter{axi}{0}
\renewcommand{\theaxi}{\arabic{axi}}
\newenvironment{axi}[2][]{%
\refstepcounter{axi}
\ifstrempty{#1}%
% if condition (without title)
{\mdfsetup{%
frametitle={%
\tikz[baseline=(current bounding box.east),outer sep=0pt]
\node[anchor=east,rectangle,fill=blue!20]
{\strut Definition};}
}%
% else condition (with title)
}{\mdfsetup{%
frametitle={%
\tikz[baseline=(current bounding box.east),outer sep=0pt]
\node[anchor=east,rectangle,fill=blue!20]
{\strut Definition};}%
}%
}%
% Both conditions
\mdfsetup{%
skipabove=2em,
skipbelow=2em,
innertopmargin=10pt,linecolor=blue!20,%
linewidth=2pt,topline=true,%
frametitleaboveskip=\dimexpr-\ht\strutbox\relax%
}
\begin{mdframed}[]\relax}{%
\end{mdframed}}
\begin{document}
\maketitle
First-order logic has pride of place in our best account of the
structure of knowledge. In fact, there is reason to believe that
first-order logic is fully sufficient to encode all deductively valid
reasoning in mathematics, hence all deductively valid reasoning in the
mathematically rigorous empirical sciences. It was discovered in the
early 20th century that first-order logic is powerful enough to
axiomatize many of the theories that mathematicians use, such as
number theory, group theory, ring theory, field theory, etc.. And
although there are other mathematical theories that are overtly
second-order (e.g.\ the theory of topological spaces quantifies over
subsets, and not just individual points), nonetheless first-order
logic can be used to axiomatize set-theory, and any second-order
theory can be formalized within set theory. Thus, first-order logic
provides an expansive framework in which much, if not all, deductively
valid human reasoning can be represented.
Philosophers in the 20th century spent some time debating the question
of whether first-order logic is powerful enough to represent all such
valid reasoning. We will return to that question in Chapter ??. In
this chapter, we will study the properties of first-order logic, the
theories that can be formulated within it, and the relations that hold
between them.
Let's work from the ground up, beginning with a specific theory.
\begin{example}[The theory of partial orders] We suppose that there is
a relation, which we'll denote by $\leq$, and we then proceed to lay
down some postulates for this relation. In particular:
\begin{itemize}
\item Postulate 1: The relation $\leq$ is reflexive in the sense that
it holds between anything and itself. For example, if we were
working with numbers, we could write $2\leq 2$, or more generally,
we could write $n\leq n$ for any $n$. For this last phrase, we have
a shorthand: we abbreviate it by $\forall n(n\leq n)$, which can be
read out as, ``for all $n$, $n\leq n$.'' The symbol $\forall$ is
called the \textbf{universal quantifier}.
\item Postulate 2: The relation $\leq$ is transitive in the sense
that if $x\leq y$ and $y\leq z$, then $x\leq z$. Again, we can
abbreviate this last sentence as
\[ \forall x\forall y\forall z((x\leq y\wedge y\leq z)\to x\leq z)
,\]
which can be read as, ``for all $x$, for all $y$, and for all $z$,
if \dots ''
\item Postulate 3: The relation $\leq$ is asymmetric in the sense
that if $x\leq y$ and $y\leq x$, then $x=y$. This postulate can
be formalized as:
\[ \forall x\forall y((x\leq y\wedge y\leq x)\to x=y) .\]
\end{itemize}
In these previous postulates, we see the same logical connectives
that we used in propositional logic, such as $\wedge$ and $\to$.
But now these connectives might hold between things that are not
themselves sentences. For example, $x\leq y$ is not itself a
sentence, because $x$ and $y$ aren't names of things. We say that
$x$ and $y$ are \textbf{variables}, and we say that $\leq$ is a
\textbf{relation symbol}. Finally, the familiar symbol $=$ is also
a relation symbol. \hfill \qed \end{example}
We've described just the barest of bones of the theory of a partial
order. There are a couple of further things that we would definitely
like to be able to do with this theory. First, we would like to be
able to derive consequences from the postulates --- i.e.\ we would
like to derive theorems from the axioms. In order to do so, we will
need to specify the \textbf{rules of derivation} for first-order
logic. Second, we would like to be able to identify mathematical
structures that exemplify the axioms of partial order. To do so, we
need to specify the \textbf{semantics}, or \textbf{model theory}, for
first-order logic.
%% Perhaps a couple more examples at this point? e.g. mereology?
%% e.g. group theory
\begin{example}[The theory of abelian groups] We're all familiar with
number systems such as the integers, the rational numbers, and the
real numbers. What do these number systems have in common? One
common structure between them is that they have a binary relation
$+$, a neutral element $0$, and each number has a unique inverse.
We also notice that the binary relation $+$ is associative in the
sense that $x+(y+z)=(x+y)+z$, for all $x,y,z$. We can formalize
this last statement as
\[ \forall x\forall y\forall z(x+(y+z)=(x+y)+z) .\]
In these cases, the operation $+$ is also commutative, that is
\[ \forall x\forall y(x+y=y+x ) .\] Bringing these postulates
together, we have the theory of abelian groups. Notice that in this
case, we've enlarged our vocabulary to include a symbol $+$ and a
symbol $0$. The symbol $+$ is not exactly a relation symbol, but
instead is a function symbol. Intuitively speaking, given any names
$n$ and $m$ of numbers, $n+m$ also names a number. Similarly, $0$
is taken to be the name of some specific number, and in this sense
it differs from variables. \hfill \qed
\end{example}
Abstracting from the previous examples, and many others like them
throughout mathematics, we now define the language of first-order
logic as follows.
\begin{defn} The \textbf{logical vocabulary} consists of the symbols
$\bot,\forall ,\exists ,\wedge ,\vee ,\neg ,\to ,(, )$. The symbol
$\bot$ will serve as a propositional constant. The final two
symbols here are simply punctuation symbols that will allow us to
keep track of groupings of the other symbols. \end{defn}
Please note that we intentionally excluded the equality symbol $=$
from the list of logical vocabulary. Several philosophers in the 20th
century discussed the question of whether the axioms for equality were
``analytic truths,'' or whether they should count as part of a
contingent, non-logical theory. We will not enter into that
discussion here. Nonetheless, it will be helpful on a purely
technical level to treat the axioms for equality as a particular
theory. We will also be taking a somewhat non-standard approach by
treating variables $x_1,x_2,\dots $ as part of the ``non-logical
vocabulary'' of a theory. Our reason for doing so will become clear
when we discuss the notion of translations between theories.
\begin{defn} A \textbf{signature} $\Sigma$ consists of:
\begin{enumerate} \item A countably infinite collection of
\textbf{variables}.
\item A collection of \textbf{relation symbols}, each of which is
assigned a natural number called its \textbf{arity}.
\item A collection of \textbf{function symbols}, each of which is assigned a
natural number called its arity.
\item A collection of \textbf{constant symbols}. Note: constant
symbols can also be considered as function symbols of zero
arity. \end{enumerate}
\end{defn}
Some logicians use the name \emph{similarity type} as a synonym for
\emph{signature}. There is also a tendency among philosophers to
think of a signature as an \emph{uninterpreted language}. The idea
here is that the elements of the signature are symbols that receive
meaning by means of a semantic interpretation.
Although a list of variables is technically part of a signature, we
will frequently omit mention of the variables, and defer to using the
standard list $x,y,x_1,x_2,\dots $. Only in cases where we are
comparing two theories will we need to carefully distinguish their
variables from each other.
\begin{example} For the theory of abelian groups, we used a signature
$\Sigma$ that has a binary function symbol $+$ and a constant symbol
$0$. Some other presentations of the theory of abelian groups use a
signature $\Sigma '$ that also has a unary function symbol ``$-$''
for the inverse of an element. Still other presentations of the
theory use a signature that doesn't have the constant symbol $0$.
We will discuss in Chapter ?? the sense in which these different
theories all share the right to be called \textit{the} theory of
abelian groups. \hfill \qed \end{example}
\begin{disc} Let $\Sigma$ be the signature consisting of a binary
relation symbol $r$, and let $\Sigma '$ be the signature consisting
of a binary relation symbol $R$. Are these signatures the same or
different? That depends on what implicit background conventions
that we adopt, in particular, whether our specification of a
signature is case-sensitive or not. In fact, we could adopt a
convention that was even more strict in how it individuates
signatures. For example, let $\Sigma ''$ be the signature
consisting of a binary relation symbol $r$. One could say that
$\Sigma ''$ is a different signature from $\Sigma$ because the $r$
in $\Sigma ''$ occurs at a different location on the page than the
$r$ that occurs in $\Sigma$. Of course, we would typically assume
that $\Sigma =\Sigma '$, but such a claim depends on an implicit
background assumption that there is a single letter form of which
the two occurences of $r$ are instances.
We will generally leave these implicit background assumptions
unmentioned. Indeed, to make these background assumptions explicit,
we would have to rely on further implicit background assumptions,
and we would never make progress in our study of first-order
logic. \hfill \qed
\end{disc}
In Chapter ?? we will show a precise sense in which each first-order
theory is \textbf{equivalent} to a theory in a signature that has only
relation symbols (i.e.\ no function symbols or constant symbols).
Anticipating that result, we will first study the case of first order
theories that use only relation symbols, e.g.\ the theory of a partial
order.
\section*{Formation rules}
In this section, we suppose that $\Sigma$ is a fixed signature that
has no function or constant symbols. (As we will see in Chapter ??,
there is a sense in which this assumption results in no loss of
generality.) Since there are no function symbols, the definition of
\textbf{terms} is quite simple:
\begin{defn} A symbol $t$ is called a \textbf{term} of $\Sigma$ just
in case $t$ is a variable $x$. In this case, the variable $x$ is
free in $t$. \end{defn}
We shall now give an inductive definition that simultaneously defines
the notion of a \textbf{formula}, and of the \textbf{free variables}
in that formula.
\begin{defn} Let $\Sigma$ be a signature that consists of only
relation symbols. We define the set of \textbf{$\Sigma$-formulas}
as follows:
\begin{enumerate}
\item $\bot$ is a formula that contains no free variables.
\item If $r$ is an $n$-ary relation symbol in $\Sigma$, and
$t_1,\dots ,t_n$ are terms, then $r(t_1,\dots ,t_n)$ is a formula
whose free variables are the union of all free variables in
$t_1,\dots ,t_n$.
\item If $\phi$ and $\psi$ are formulas, then $\phi\wedge \psi$ is a
formula whose free variables are the union of the free variables in
$\phi$ and $\psi$. Similarly for the other Boolean connectives
$\neg ,\vee,\to$.
\item If $\vp$ is a formula, then so are $\exists x\vp$ and
$\forall x\vp$. The free variables in $\exists x\vp$ and
$\forall x\vp$ are those in $\vp$, except possibly for $x$. In
other words, $\exists x\vp$ and $\forall x\vp$ contain no free
occurrences of $\vp$, even if $\vp$ does. \end{enumerate} We say
that a formula $\vp$ is \emph{closed}, or is a \emph{sentence}, if it
contains no free variables. \end{defn}
As usual, a more fully precise definition of $\Sigma$-formulas would
make liberal use of parentheses. For example, we would want to say
that $(\vp\wedge \psi )$ is a $\Sigma$-formula when $\vp$ and $\psi$
are $\Sigma$-formulas. Nonetheless, we will continue to allow
ourselves to omit parentheses when no confusion is likely to result.
\begin{note} Our definition of the set of formulas allows for
redundant quantification. For example, the string
$\exists x\forall x(x=x)$ is well-formed formula according to our
definition. This formula results from applying the quantifier
$\exists x$ to the sentence $\forall x(x=x)$. We will have to be
careful in our definition of derivation rules, and semantic rules,
to take the case of empty quantification into account.
\end{note}
%% Parse Tree?
It is helpful to think of formulas in terms of their \textbf{parse
trees.} For example, the formula
$\forall x\exists y(r(x,x)\to r(x,y))$ has the following parse tree:
\bigskip \begin{tikzpicture}[level distance=1.5cm,
level 1/.style={sibling distance=3cm},
level 2/.style={sibling distance=1.5cm}]
\node {$\forall x\exists y(r(x,x)\to r(x,y))$}
child {node {$\exists y(r(x,x)\to r(x,y))$}
child {node {$r(x,x)\to r(x,y)$}
child {node {$r(x,x)$}}
child {node {$r(x,y)$}}}};
\end{tikzpicture}
\noindent The bottom nodes must each be elementary formulas, i.e.\
either $\bot$, or a relation symbol followed by the appropriate number
of variables. Each parent-child relationship in the tree corresponds
to one of the Boolean connectives or to one of the quantifiers.
Formulas stand in one-to-one correspondence with parse trees: each
well-formed tree ends with a specific formula, and no other tree
yields the same formula.\footnote{Of course, we are not being fully
explicit here about the identity conditions of parse trees.} Using
the identity of formulas and parse trees, we can easily define a few
further helpful notions:
\begin{defn} Let $\vp$ be a $\Sigma$-formula. The family of
\textbf{subformulas} of $\vp$ consists of all those formulas that
occur at some node in its parse tree. \end{defn}
\begin{defn} If a quantifier $\exists x$ occurs in the formula $\vp$,
then the \textbf{scope} of that occurrence is the formula which
occurs at the immediately previous node in the parse
tree. \end{defn}
For example, in the formula $\forall x\exists y(r(x,x)\to r(x,y))$,
the scope of $\exists y$ is the formula $r(x,x)\to r(x,y)$. In
contrast, in the formula $\forall x(r(x,x)\to \exists yr(x,y))$, the
scope of $\exists y$ is the formula $r(x,y)$.
We can now make the notion of free and bound variables even more
precise. In particular, each individual occurrence of a variable in
$\vp$ is either free or bound. For example, in the formula
$p(x)\wedge \exists p(x)$, $x$ occurs freely in the first subformula,
and bound in the second subformula.
\begin{defn}[Free and bound occurrences] An occurence of a variable
$x$ in $\vp$ is \textbf{bound} just in case that occurrence is
within the scope of either $\forall x$ or $\exists x$. Otherwise
that occurence of $x$ is free. \end{defn}
We could now perform a sanity check to make sure that our two notions
of bound/free variables coincide with each other.
\begin{fact} A variable $x$ is free in $\vp$ (in the sense of the
definition of $\Sigma$-formulas) if and only if there is a free
occurence of $x$ in $\vp$ (in the sense that this occurence does not
lie in the scope of any corresponding quantifier). \end{fact}
\begin{defn} We say that $y$ is free for $x$ in $\vp$ just in case no
free occurrence of $x$ in $\vp$ is in the scope of the quantifier
$\exists y$. \end{defn}
Intuitively speaking, $y$ is free for $x$ just in case subtituting $y$
for $x$ does not result in $y$ being captured by $\exists y$ or
$\forall y$. For example, in the formula $p(x)$, the variable $y$ is
free for $x$ (since $y$ is free in $p(y)$). In contrast, in the
formula $\exists yp(x)$, the variable $y$ is not free for $x$ (since
$y$ is not free in $\exists yp(y)$). We will need this notion in
order to formulate sound derivation rules. For example, the rule of
$\forall$-elim should say something like:
$\forall x\vp (x)\vdash \vp (y)$. However, if this rule were not
restricted in some way, then it would yield
\[ \forall x\exists y(x\neq y) \: \vdash \: \exists y(y\neq y) ,\]
which is clearly not valid.
It is also sometimes necessary to distinguish particular occurrences
of a subformula of a formula, and to define the \textbf{depth} at
which such an instance occurs.
\begin{defn} Let $\psi$ be a node in the parse tree of $\vp$. The
\textbf{depth} of $\psi$ is the number of steps from $\psi$ to the
root node. We say that $\psi$ is a \textbf{proper subformula} of
$\vp$ if $\psi$ occurs with depth greater than $0$. \end{defn}
The parse trees of formulas are finite, by definition. Therefore, the
depth of every occurrence of a subformula of $\vp$ is some finite
number.
There are a number of other properties of formulas that are definable
in purely syntactic terms. For example, we could define the
\textbf{length} of a formula. We could then note that the connectives
take formulas of a certain length and combine them to create formulas
of a certain greater length.
\begin{exercise} Show that no $\Sigma$-formula can occur as a proper
subformula of itself. \end{exercise}
%% Example proofs by induction on construction?
Any time we define a set inductively, we have a corresponding method
of proof by induction. In this case, we can prove that a property
$\mathbb{P}$ holds for all $\Sigma$-formulas by showing that
$\mathbb{P}$ holds for all elementary $\Sigma$-formulas, and that
$\mathbb{P}$ is preserved by the various operations that allow us to
construct more complex $\Sigma$-formulas. What is a bit more
difficult in this case is proving that some property holds of all
$\Sigma$-sentences, when that property does not hold of all
$\Sigma$-formulas. In such cases, we'll have to look for a clever way
to modify the standard sort of proof by induction.
\section*{Derivation rules}
We suppose again that $\Sigma$ is a fixed signature that has no
function or constant symbols. We could then use any one of a myriad
of systems of derivation rules (axiomatic, natural deduction, sequent
calculus) to define the family of provable $\Sigma$-formulas. These
rules will include analogues of the rules we provided for
propositional logic, and also some specific rules for introducing and
eliminating quantifiers. For example, one might say that a
universally quantified sentence $\forall x\phi (x)$ implies any
instance $\phi (a)$; and that an instance $\phi (a)$ implies
$\forall x\phi (x)$ only if $\phi (a)$ was derived from something in
which the name ``$a$'' does not occur. There are various ways that
these rule can be formulated, but there is widespread consensus on
what should be the extension of the provability relation $\vdash$
between sentences in first-order logic. We will adopt the following
set of rules:
\begin{enumerate}
\item The rules for the Boolean connectives that we put forward in our
discussion of propositional logic. In this case, we permit
derivations between formulas with free variables. For example, we
take
\[ \vp (x)\wedge \psi (y) \: \vdash \: \vp (x) ,\] to be licensed
by the rule of $\wedge$-elim.
\item $\vp\wedge \neg p\:\vdash \: \bot$, for any formula $\vp$.
\item $\bot \: \vdash \: \vp$, for any formula $\vp$.
\item $\forall x\vp (x)\: \vdash \: \vp (t)$, where $t$ is any term
that is free for $x$.
\item $\vp \: \vdash $
\end{enumerate}
Although there are different ways we could specify the quantifier
rules, we will always have the following two results:
\begin{enumerate}
\item[(E1)] If $\vp\vdash \psi$, and $x$ does not occur free in
$\psi$, then $\exists x\vp \vdash \psi$.
\item[(E2)] If $\exists x\vp \vdash \psi$ then $\vp\vdash\psi$.
\end{enumerate}
We will make use of these facts in the proof of completeness in
Section ??.
\begin{defn} For formulas $\vp$ and $\psi$, we say that $\vp$ and
$\psi$ are \emph{logically equivalent}, written $\vp\simeq\psi$,
just in case both $\vp\vdash\psi$ and $\psi\vdash\vp$.
\end{defn}
The definition of $\simeq$ is patently symmetric. The rule of
assumptions $\varphi\vdash\varphi$ entails that $\simeq$ is reflexive,
and the transitivity of $\vdash$ entails that $\simeq$ is transitive.
Thus, $\simeq$ is an equivalence relation on the set of
$\Sigma$-formulas. Note that formulas $\vp$ and $\psi$ can be
equivalent in this sense even if they don't share all free variables
in common --- as long as the non-matching variables occur vacuously.
For example, $p(x)$ is equivalent to $p(x)\wedge (y=y)$, and it's also
equivalent to $p(x)\vee (y\neq y)$. [The issue here has nothing in
particular to do with the equality relation. The variable $y$ also
occurs vacuously in $p(y)\vee \neg p(y)$.] In contrast, the formulas
$p(x)$ and $p(y)$ are not equivalent (in the empty theory), since it's
not universally valid that
$\vdash \forall x\forall y(p(x)\leftrightarrow p(y))$.
\begin{lemma} The relation $\simeq$ is compatible with the Boolean
connectives in the following sense: if $\phi\simeq\phi '$ and
$\psi\simeq \psi '$, then
$(\phi\wedge \psi )\simeq (\phi '\wedge \psi ')$, and similarly for
the other connectives. \end{lemma}
The proof of this lemma is straightforward if you know the
introduction and elimination rules for the Boolean connectives.
%% Deduction theorem?
%% Sufficient sets of connectives
%% TO DO: Does it also work with the quantifiers? Careful, could
%% capture a free variable in one of the formulas
\section*{Translation and substitution}
Suppose that $\Sigma$ and $\Sigma '$ are predicate logic signatures.
At this point, we are working with single-sorted logic, which means
that $\Sigma$ is a list of relation symbols, function symbols, and
constant symbols. Similary for $\Sigma '$. Suppose now that we are
looking for a way to translate formulas of $\Sigma$ to formulas of
$\Sigma '$. As usual, we begin by assigning the elements of the
signature $\Sigma $ to syntactic objects built from $\Sigma '$. Now,
you might initially think that a relation symbol $r\in \Sigma$ ought
to be assigned to some corresponding relation symbol in $\Sigma '$.
But recall the case of propositional logic. In that case, we found it
convenient sometimes to translate an elementary sentence, such as $p$,
to a complex sentence in another language. The only important thing
was that sentences should be mapped to sentences.
Similarly, a \textbf{reconstrual} of $\Sigma$ into $\Sigma '$ will
assign each relation symbol $r$ of $\Sigma$ to some corresponding open
formula $r_F$ of $\Sigma '$. For example, suppose that $r$ is a
binary relation symbol in $\Sigma$, and that $\Sigma '$ has two unary
relation symbols $p$ and $q$. Then, for example, $r(x,y)$ could be
reconstrued as the open formula $p(x)\wedge \neg q(y)$. Similarly, if
$\Sigma '$ contains a ternary relation symbol $s$, then $r(x,y)$ could
be reconstrued as the open formula $\exists z\,s(x,y,z)$.
Note that, in both cases, we reconstrued $r(x,y)$ as a formula $r_F$
of $\Sigma '$ that shares the same free variables as $r$. That
assumption should strike you as a little bit suspicious. Why think
that the language built on $\Sigma $ should share exactly the same
variables as the language built on $\Sigma '$? (And your answer
shouldn't be: that's how we did it in my introduction to logic class.)
It's not like variables have some ``trans-theoretical'' meaning that
must be preserved by any reasonable translation.
But how then can variables be reconstrued in moving from one theory to
another? One natural proposal would be to include in a reconstrual a
mapping from variables of $\Sigma$ to variables of $\Sigma '$, i.e.\ a
function that assigns a variable of $\Sigma '$ to each variable of
$\Sigma$. Even so, it's a non-trivial question whether there is an
in-principle reason that a single variable in $\Sigma$ must be
reconstrued as a single variable in $\Sigma '$. Perhaps one theorist
uses several variables to do the work that the other theorist manages
to do with a single variable. Such cases are not hard to find in the
sciences --- for example, when the objects of one mathematical theory
are reconstrued as ``logical constructions'' of objects in another
mathematical theory.
\begin{enumerate}
\item In the 19th century, the German mathematician Leopold Kronecker
is reported to have said, ``God made the integers, all else is the
work of man.'' In more prosaic terms, talk about higher number
systems --- such as rational, real, and complex numbers --- can be
reduced to talk about integers. However, to effect such a
reduction, one must treat each rational number as a pair of integers
--- or, more accurately, as an equivalence class of pairs of
integers. Similarly, to reduce the complex numbers to the real
numbers, one must treat a complex number as a pair of real numbers,
viz.\ the real and imaginary parts of the complex number.
\item Now for a more controversial example, which we will take up at
greater length in Chapter ??. There are different ways that one can
write down axioms for Euclidean geometry. In one axiomatization,
the basic objects are points; and in another axiomatization, the
basic objects are lines. Is there a sense in which these two
axiomatized theories could both be Euclidean geometry, in
particular, that they could be equivalent? The answer is Yes, but
only if one allows translations that take a single variable of the
first theory to a pair of variables of the second theory. In
particular, a line needs to be treated as an equivalence class of
pairs of points; and a point needs to be treated as a pair of
intersecting lines.
\end{enumerate}
Let's proceed then under the assumption that a single variable in
one language could be reconstrued in terms of multiple variables in
another language. Thus, a \textbf{reconstrual}, in the formal
sense, should include a function that matches variables of the
signature $\Sigma$ to $n$-tuples of variables of the signature
$\Sigma'$. For simplicity, we will often indicate such a map by
using primed variables. For example, a variable $x$ is mapped to
$x'_1,\dots ,x'_n$. But recall that such a map is not given in
advance --- giving a translation includes specifying some particular
map.
So far we have: a reconstrual $F$ assigns relation symbols to open
formulas, and variables to $n$-tuples of variables. There are still
a couple more pieces of data that we wish to include in a
reconstrual.
First of all, consider again the case of reconstruing rational
numbers (i.e.\ fractions) as pairs of integers. Of course, not
\emph{every} pair of integers gives a well-defined fraction. For
example, there is no fraction of the form $\frac{1}{0}$. In that
case, the ``integer theorist'' doesn't think of the domain of
fractions as consisting of all pairs of integers; rather, she thinks
of that domain as consisting of pairs of integers where the second
entry is non-zero.
To capture this nuance --- the restriction of the domain of
quantification --- we stipulate the a reconstrual $F$ includes a
formula $u_F$ of the target language $\Sigma '$. In the running
example, the formula $u_F$ could be given by
\[ u_F(x,y) \: \equiv \: (x=x)\wedge (y\neq 0) .\] The integer
theorist can then use the formula $u_F$ to restrict her quantifiers
to the domain of well-defined fractions.
Finally, and perhaps most controversially, let's consider how we
might reconstrue the equality relation $=$ of the domain theory $T$
as a relation of the target theory $T'$. Recall that the single
variables $x$ and $y$ will typically be reconstrued as $n$-tuples of
variables $\vec{x}$ and $\vec{y}$. In that case, how should we
reconstrue the formula $x=y$? One might naturally propose that
$x=y$ be reconstrued as the formula
\begin{equation} (x_1=y_1)\wedge (x_2=y_2)\wedge \cdots \wedge
(x_n=y_n) .\label{simple} \end{equation} But here we need to think
a bit harder about how and why variables of $\Sigma$ are encoded as
variables of $\Sigma '$. For this, let's consider again the example
of rational numbers being reduced to integers.
\begin{example} Consider a formula $x=y$ in the theory of rational
numbers. To the ``integer theorist,'' the variables $x$ and $y$
really represent complex entities, namely fractions. What's more,
to say that two fractions $\frac{x_1}{x_2}$ and $\frac{y_1}{y_2}$
are equal does not mean that $x_1=y_1$ and $x_2=y_2$. Rather,
$\frac{x_1}{x_2}=\frac{y_1}{y_2}$ means that
$x_1\times y_2=y_1\times x_2$. In other words, the formula $x=y$
of the language of the rational numbers is reconstrued as the
formula
\begin{equation} x_1\times y_2=y_1\times x_2
,\label{foo} \end{equation} in the language of the integers,
where $\times$ is the multiplication operation. \hfill
\qed \end{example}
This example suggests that we might not always want the formula
$x=y$ to be reconstrued as Eqn.\ \ref{simple}. Instead, we might
prefer to reconstrue $x=y$ as some other $\Sigma '$ formula
$e(x_1,\dots ,x_n;y_1,\dots ,y_n)$. Of course, not everything goes:
$e$ will need to perform the same functions in the theory $T'$ that
the formula $x=y$ performs in the theory $T$. In particular, we
will require that $e$ is an equivalence relation relative to the
theory $T'$.
We're now ready to propose a general definition of a translation
between predicate logic theories. We begin with the notion of a
\textbf{reconstrual}, which provides a dictionary matching elements
of the signature $\Sigma$ with syntactic objects built from
$\Sigma '$. \def\vp{\varphi}
\begin{axi}{} Let $\Sigma$ and $\Sigma '$ be single-sorted predicate
logic signatures. An \textbf{$n$-dimensional reconstrual} $F$ of
$\Sigma$ into $\Sigma '$ consists of:
\begin{enumerate}
\item A mapping of variables of $\Sigma$ to $n$-tuples of distinct
variables of $\Sigma '$. For simplicity, we will often suppress
any explicit mention of this mapping, and rely instead on
suggestive notation --- e.g.\ using $x_1,\dots ,x_n$ or
$\vec{x}$ for the $n$-tuple corresponding to $x$. We also
require that if $x$ and $y$ are distinct variables of $\Sigma$,
then $\vec{x}$ and $\vec{y}$ are non-overlapping $n$-tuples of
variables of $\Sigma '$.
\item An $n$-ary formula $u_F(\vec{x})$ of $\Sigma$, picking out
the target domain of the reconstrual.
\item A $2n$-ary formula $e_F(\vec{x},\vec{y})$ of $\Sigma '$,
picking out the intended interpretation of the equality relation
for $\Sigma$.
\item For each $m$-ary relation symbol $r$ of $\Sigma$, an
$mn$-ary formula $r_F$ of $\Sigma '$. (To be fully precise, we
would need to specify which variables occur in $r_F$. On an
intuitive level, a unary relation $r(x)$ will correspond to an
$n$-ary relation $r_F (x_1,\dots ,x_n)$, where $x_1,\dots ,x_n$
are the variables assigned by the reconstrual to $x$.)
\end{enumerate}
\end{axi}
As we now show, a reconstrual $F$ of $\Sigma$ into $\Sigma '$
naturally extends to a map from $\Sigma$-formulas to
$\Sigma '$-formulas. If $\vp$ is a $\Sigma$-formula, we will
typically use $\vp _F$ to designate the corresponding $\Sigma '$
formula.
\begin{itemize}
\item An elementary formula $r(x)$ of $\Sigma$ is assigned to the
formula $r_F(\vec{x})$ of $\Sigma '$. (Notice here the use of the
mapping between the variable $x$ of $\Sigma$ and the $n$-tuple
$\vec{x}$ of variables of $\Sigma '$.)
\item The elementary formula $x=y$ of $\Sigma$ is assigned to the
formula $e_F(\vec{x},\vec{y})$ of $\Sigma '$.
\item Extending recursively, define
\[ (\phi\wedge \psi )_F \: = \: \phi _F\wedge \psi _F ,\]
and similarly for the other Boolean connectives.
\item Extending recursively, define
\[ (\exists x \phi (x))_F \: \equiv \: \exists
\vec{x}(u_F(\vec{x})\wedge \phi _F ) .\]
and
\[ (\forall x\phi (x))_F\:\equiv \: \forall \vec{x}(u_F(\vec{x})\to \phi
_F ) .\]
\end{itemize}
The quantifier clauses of this definition could use some
clarification. Roughly speaking, the quantifier $\exists x$ is
translated as $\exists x_1\cdots \exists x_n$, relativized to the
domain $u_F$. Similarly, $\forall x$ is translated as
$\forall x_1\cdots \forall x_n$, again relativized to the domain
$u_F$. Keep in mind, however, that the meaning of the resulting
quantifiers $\exists x_1\cdots \exists x_n$ and
$\forall x_1\cdots \forall x_n$ also depends on the corresponding
equality relation, which in this case is $e_F$.
\begin{thm} If $F$ is a reconstrual and $\vp\vdash\psi$ then
$\vp _F\vdash \psi _F$. \end{thm}
\begin{proof} We will prove this result by induction on the definition
of the relation $\vdash$. For the base case, the rule of
assumptions justifies not only $\vp\vdash\vp$, but also
$\vp _F\vdash\vp _F$. The inductive cases for the Boolean
connectives involve no special complications, and so we omit them.
Consider then the case of $\exists$-elim. Suppose that
$\exists x\vp (x)\vdash \psi$ results from $\vp (x)\vdash \psi$,
where $x$ is not free in $\psi$. The inductive hypothesis here says
that $(\vp (x))_F\vdash \psi _F$, which can be rewritten as
$\vp _F(\vec{x})\vdash \psi _F$. We need to show that
$(\exists x\vp (x))_F\vdash \psi _F$, which can be rewritten as
\[ \exists \vec{x}(u_F(\vec{x})\wedge \vp _F(\vec{x}))\vdash \psi _F \]
\end{proof}
%% substitution -- see Belnap, maybe Kleene
\begin{defn} Let $\phi$ be a $\Sigma$-formula, and let $x$ and $y$ be
variables. We will use the notation $\phi (y/x)$ to denote the
result of a ``safe substitution'' of $y$ for all free occurrences of
$x$ in $\phi$. By ``safe substitution,'' we mean to prohibit cases
where a free occurrence of $x$ would become a bound occurrence of
$y$. For example, in the formula $\exists y\, r(x,y)$, replacing
$x$ with $y$ would result in the new variable $y$ becoming captured.
Such can be prevented by choosing a variable $z$ that does not occur
in $\phi$, replacing occurrences of $\exists y$ with $\exists z$,
and corresponding bound occurrences of $y$ with $z$, then replacing
all free occurences of $x$ with $y$.
\end{defn}
%% seems that I need to know that \phi (y/x) is equivalent to \phi
%% Step 1: alpha equivalence -- this is easy
%% Step 2: replacement by equivalent subformulas -- this is hard? yes,
%% a little bit
\subsection*{Replacement}
Intuitively speaking, it seems obvious that if a subformula $A$ of
$\vp _A$ is replaced with an equivalent subformula $B$, then the
resulting formula $\vp _B$ will be equivalent to the original. Here
we will state the result more precisely, and sketch its proof.
When we say that $A$ and $B$ are \textbf{logically equivalent}, we
simply mean that $\vdash A\leftrightarrow B$. Or, if using a proof
system that applies only to closed sentences, then we say that $A$ and
$B$ are logically equivalent just in case
\[ \vdash \forall x_1\cdots \forall x_n(A\leftrightarrow B) , \] where
$x_1,\dots ,x_n$ is the list of variables with free occurrences in
either $A$ or in $B$.
The key lemmas for proving replacement are Lemma ?? and the following:
\begin{lemma} If $A(x)\simeq B(x)$ then
$\exists xA(x)\simeq \exists xB(x)$. \end{lemma}
\begin{proof}[Sketch of proof] Given $A(x)\simeq B(x)$, we have
$\vdash A(x)\to B(x)$, and hence $\vdash A(x)\to \exists xB(x)$ by
$\exists$-intro. Since $x$ does not occur free in $\exists x B(x)$,
it follows that $\vdash \exists xA(x)\to \exists xB(x)$ by
$\exists$-elim. A similar argument shows that
$\vdash \exists x B(x) \to \exists x A(x)$. Therefore
$\exists x A(x)\simeq \exists x B(x)$. \end{proof}
\begin{rthm} Suppose that $\vp _B$ is the formula that results from
replacing $A$ in $\vp _A$ with $B$. If $A\simeq B$ then $\vp
_A\simeq \vp _B$.
\end{rthm}
\section*{{\S} The syntactic view of theories}
%% Observation and theoretical vocabulary
%% Putnam, Achinstein, van Fraassen critique
\end{document}