-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathrealism.tex
2351 lines (2117 loc) · 131 KB
/
realism.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{Metalogic in Analytic Philosophy} \label{chap-realism}
%% TO DO: people like Psillos complain that it is no longer clear what
%% the scientific realism debate is even about. Yes, I agree with
%% this. But we can now see that it has a lot to do with how finely
%% one draws the net of "equivalence".
Much of 20th century analytic philosophy was concerned --- when not
explicitly, then implicitly --- with theories, and with the
relationships between them. For example, is every spacetime theory
equivalent to one with Euclidean background geometry? Or, is folk
psychology reducible to neuroscience? Or, can there be a good reason
to choose a theory over an empirically equivalent rival theory?
But what is a theory? And what does it mean to say that two theories
are equivalent, or that one theory is reducible to another? Carnap
had the audacious idea that philosophy can follow mathematics' method
of explication: to take an intuitive notion, and to find a nearby
neighbor in the realm of precisely defined mathematical concepts. In
this book, we've tried to follow Carnap's lead; and indeed, we've
hoped that we could do a bit better than Carnap, because mathematics
has come a long way in the past 75 years. We now have mathematical
concepts --- such as adjoint functors, or categorical equivalence ---
the likes of which Carnap could not even have dreamed about.
In this book, we've attempted to explicate the concept of a theory, as
well as some of the relations between theories that scientists and
philosophers find it useful to discuss. With these explications in
the background, we can now return to some of the big questions of
philosophy of science, such as, ``what is the proper attitude to take
toward a successful scientific theory?''
\section{Ramsey sentences} \label{ramsey}
%% van fraassen = Beth shows that there is no mysterious notion of
%% implicit definition
%% Lewis on theoretical terms
%% Functionalism (Bealer)
%% Dewar on Ramsey equivalence
No analytic philosopher's education is complete until she learns the
magic of the Ramsey sentence. The idea was proposed by Frank
\cite{ramsey}, and was reinvented by Carnap in the 1950s --- or, more
accurately, Carnap forgot that he learned about it from Herbert
Bohnert \cite[see][]{psillos2000}. Most contemporary philosophers
know of the idea because David \cite{lewis-terms} argued that it
solves the \emph{problem of theoretical terms}. In the years since
Lewis' seminal paper, Ramsey sentences have become a sort of {\it deus
ex machina} of analytic philosophy.
Let's start with a simple example. Suppose that $P$ is a theoretical
predicate, and that $O$ is an observational predicate. (Or in Lewis'
preferred terminology, $O$ is antecedently understood vocabulary, and
$P$ is new vocabulary.) Now suppose that our theory $T$ consists of a
single sentence $P(c)\to O(c)$, which might be paraphrased as saying
that $O(c)$ is an empirical sign that $P(c)$. (Here $c$ is a constant
symbol. We omit first-order quantifiers to keep things simple.) To
form the Ramsey sentence of $T$, we simply perform an instance of
second-order existential generalization:
\[ \begin{array}{c c} P(c)\to O(c) \\ \hline \exists X(X(c)\to
O(c)) \end{array} \] The sentence below the line is called the
Ramsey sentence $T^R$ of the theory $T$. Thus, while the original
theoretical statement $T$ mentions some particular property $P$, the
Ramsey sentence $T^R$ simply says that there is some or other property
that plays the appropriate role. It may feel --- and has felt to many
philosophers --- that the truth of $T^R$ somehow magically endows the
term $P$ with meaning. In particular, philosophers are wont to say
things like, ``$P$ is whatever it is that plays the role described by
$T^R$.''
Since Ramsey sentences draw upon the resources of second-order logic,
the neophyte is left to wonder: does the philosophical magic here
depend on something special that happens in second-order logic,
something that only the most technically sophisticated philosophers
can understand? We think that the answer to this question is No. In
fact, Ramsifying a theory simply weakens that theory in the same way
that existentially quantifying a first-order sentence weakens that
sentence. Consider the following pedestrian example.
\begin{example} Let $\Sigma =\{ m\}$, where the name $m$ is a
theoretical term. Let $T$ be the theory $\exists x(x=m)$ in
$\Sigma$. Then the Ramsey sentence $T^R$ of $T$ is the sentence
$\exists x(x=x)$, which is just a tautology. That is, $T^R$ is the
empty theory in the empty signature. It is easy to see that the
inclusion $I:T^R\to T$ is conservative, but is not essentially
surjective. In particular, there is no formula $\phi$ of $\Sigma$
such that $(I\phi )(x)\equiv (x=m)$. The fact that $I$ is not
essentially surjective corresponds to the fact that
$I^*:\mathrm{Mod}(T)\to \mathrm{Mod}(T^R)$ is not full. Here $I^*$
is the functor that takes a model of $T$, and forgets the extension
of $m$. In general, then, $I^*M$ has more symmetries than $M$.
We can be yet more precise about the differences between
$\mathrm{Mod}(T)$ and $\mathrm{Mod}(T^R)$. In short, a model of
$T^R$ is simply a nonempty set $X$ (and two such models are
isomorphic if they have the same cardinality). For each $p\in X$,
there is a corresponding model $X_p$ of $T$ where $X_p(m)=p$. For a
fixed $X$, and $p,q\in X$, there is an isomorphism $h:X_p\to X_q$
that maps $p$ to $q$. However, the automorphism group of $X_p$ is
smaller than the automorphism group of $X$. Indeed,
$\mathrm{Aut} (X_p)$ consists of all permutations of $X$ that fix
$p$, hence is isomorphic to $\mathrm{Aut}(X\backslash \{
p\})$.
We can see then that $T$ and $T^R$ are not intertranslatable (or
definitionally equivalent). Nonetheless, there is a sense in which
mathematicians would have no qualms about passing from $T^R$ to the
more structured theory $T$. Indeed, once we've established that the
domain $X$ is non-empty (which, of course, is a presupposition of
first-order logic), we could say, ``let $m$ be one of the elements
of $X$.'' This latter statement does not involve any further
theoretical commitment over what $T^R$ asserts. \end{example}
Our advice then to the neophyte is not to allow herself to be
intimidated by second-order quantification. In fact, we will argue
that passage from a theory $T$ to its Ramsified version $T^R$ either
forgets too much of what the original theory said, or says {\it more}
than what the original theory said --- depending on which notion of
second-order logical equivalence one adopts. Before we do this, let's
pause to recall just how much philosophical work Ramsey sentences have
been asked to do. We will look at three applications. First, Carnap
claims that Ramsey sentences solve the problem of dividing the
analytic and synthetic parts of a scientific theory. Second, Lewis
claims that Ramsey sentences solve the problem of theoretical terms,
and in particular the problem of giving meaning to ``mentalese'' in a
physical world. Third, contemporary structural realists claim that
Ramsey sentences give a way of isolating the structural claims of a
scientific theory.
\subsubsection*{Carnap's irenic realism}
One theme running throughout Carnap's work is a rejection of what he
sees as false dilemmas. In one sense, Carnap is one of the most
pragmatic philosophers ever in the western tradition, as he places
extreme emphasis on questions such as: {\it what questions are worth
asking, and what problems are worth working on?} Now, one can
imagine a philosophy graduate student asking herself: what question
should I try to answer in my dissertation? If she's a particularly
ambitious (or perhaps overconfident) student, she might decide to
determine whether materialism or dualism is true. Or she might decide
to determine whether scientific realism or instrumentalism is true.
Carnap's advice to her would be: to work on such questions is not a
good use of your time.
In the early 20th century, the debate between scientific realism and
instrumentalism centered around the question: do theoretical entities
--- i.e.\ the things named by scientific theories, but which are not
evident in our everyday experience --- exist? Or, shifting to a more
explicitly normative manner of speech: are we entitled to believe in
the existence of these entities, and perhaps even obliged to do so?
Carnap thinks that these questions are ill-posed.
Toward the end of his career, Carnap hoped that Ramsey sentences could
help show why there is no real argument between realism and
instrumentalism. In particular, if $T$ is a scientific theory
containing some theoretical terms $r_1,\dots ,r_n$, then Carnap parses
$T$ into two parts: the Ramsey sentence $T^R$, and the sentence
$T^R\to T$ that has since been dubbed the ``Carnap sentence''. Carnap
claims that the Ramsey sentence $T^R$ gives the empirical (synthetic)
content of $T$, whereas $T^R\to T$ gives the definitional (analytic)
part of $T$. The latter claim can be made plausible by realizing that
$T^R\to T$ is trivially satisfiable, simply by stipulating appropriate
extensions for $r_1,\dots ,r_n$.
\cite{psillos2000} argues that Carnap's equation of synthetic content
with the Ramsey sentence makes him a structural realist --- in which
case he is subject to Newman's objection, which impales him on the
horns of the realism-instrumentalism dilemma. \cite{friedman2011}
disagrees, arguing that Carnap's invocation of the Ramsey sentence
successfully implements his neutralist stance. Debate on this issue
continues in the literature --- see e.g.\ \cite{uebel2011,beni}.
\subsubsection*{Ramsey sentence functionalism}
In the philosophy of mind, Ramsey sentences came to play a central
role through the work of \cite{lewis-id,lewis-psy,lewis1994}, and more
generally in a point of view known as \emph{functionalism}. To be
sure, Lewis claims not to know whether or not he is a functionalist;
and most functionalists don't talk explicitly about Ramsey sentences.
However, by the 1980s, the connection between functionalism and Ramsey
had been firmly established \cite[see][]{shoemaker}.
Around 1970, materialist reductionism had gone out of style.
Philosophers concluded that folk psychology cannot, and should not, be
reduced --- neither to descriptions of behavior, nor to physiological
descriptions. However, philosophers weren't ready to give up the
physicalist project, and in particular, they didn't want to entertain
the possibility that there is an autonomous realm of mental objects or
properties. The goal then is to explain how mental properties are
anchored in physical properties, even if the former cannot be
explicitly defined in terms of the latter.
Functionalism, and functional definitions, are supposed to provide a
solution to this problem. According to functionalism, mental
properties are defined by the role that they play in our total theory
$T$, which involves both mental concepts (such as ``belief'' and
``desire'') and physical concepts (such as ``smiling'' or ``synapse
firing''). How then are we supposed to cash out this notion of being
``defined by role''? It's here that Ramsey sentences are invoked as
providing the best formal explication of functional definitions.
Contemporary analytic philosophers routinely mention Ramsey sentences
in this connection. Nonetheless, long ago, \cite{bealer} argued that
this attempt to define mental properties --- call it ``Ramsey sentence
functionalism'' --- is inconsistent. According to Bealer,
functionalism has both a negative and a positive theses. On the
negative side, functionalism is committed to the non-reductionist
thesis: mental properties (m-properties) cannot be explicitly defined
in terms of physical properties (p-properties). On the positive side,
m-properties {\it are} defined in terms of the role they play
vis-a-vis each other and the p-properties.
Let $T$ be a theory in signature $\Sigma \cup \{r_1,\dots ,r_n \}$,
where we think of $\Sigma$ as p-vocabulary, and of $r_1,\dots ,r_n$ as
m-vocabulary. We then adopt the following proposal (which defenders
of functionalism are welcome to reject or modify):
\begin{quote} $T$ provides functional definitions of $r_1,\dots ,r_n$
in terms of $\Sigma$ just in case in each model $M$ of the Ramsey
sentence $T^R$, there are unique realizing properties
$M(r_1),\dots ,M(r_n)$. \end{quote} It's easy to see then that $T$
provides functional definitions of $r_1,\dots ,r_n$ in terms of
$\Sigma$ only if $T$ implicitly defines $r_1,\dots ,r_n$ in terms of
$\Sigma$. Indeed, if $M$ and $N$ are models of $T$, then
$M|_{\Sigma}$ and $N|_{\Sigma}$ are models of $T^R$, and it follows
from the uniqueness clause that $M(r_i)=N(r_i)$. It then follows from
Beth's theorem that $T$ explicitly defines $r_1,\dots ,r_n$ in terms
of $\Sigma$.
Bealer's argument, if successful, shows that functionalism is
inconsistent: the positive thesis of functionalism entails the
negation of the negative thesis. Surprisingly, however, functionalism
lives on, apparently oblivious of this little problem of
inconsistency. In fact, functionalism hasn't just survived, it is
flourishing and spreading its tendrils --- indeed, it has become an
overarching philosophical ideology: \emph{the Canberra plan}. The
goal of the Canberra plan is to find a place in the causal nexus of
physical properties for all the stuff that makes up our daily lives
--- things like moral and aesthetic values, laws, society, love, etc..
(For further discussion, see \cite{menzies}.) One has to wonder: how
attractive would the Canberra program be if it were clearly recognized
that what it seeks is nothing more or less than the kind of full-blown
physicalistic reduction that Carnap tried to achieve in his {\it
Aufbau} program?
\subsubsection*{Structural realism}
In more recent times, Ramsey sentences have been invoked in support of
a trendy view in philosophy of science: \emph{structural realism}. In
the early 1990s, structural realism was the new kid on the block in
discussions of scientific realism and antirealism. As forcefully
recounted by \cite{worrall}, there are good arguments against both
scientific realism, and against scientific antirealism. Against
scientific realism, there is the {\it pessimistic metainduction},
which points to the long history of failed scientific theories as
evidence that our current favorite scientific theories will probably
also fail. Against scientific antirealism, there is the {\it no
miracles argument}, which points to the success of scientific
theories as something crying out for an explanation. In good Hegelian
fashion, Worrall seeks a synthesis of the extremes of realism and
antirealism --- a position which offers the best of both worlds. His
proposal is \emph{structural realism}, according to which the part of
a theory to take seriously is it's pronouncements on issues of {\it
structure}.
Worrall illustrates the idea of ``preserved structure'' with a
specific example. In particular, before Einstein's special theory of
relativity, it was thought that there was a substance, the ``aether,''
in which electric and magentic waves propagated. After the
Michelson-Morley experiment, and the success of special relativity,
there was no longer any use for the aether. Thus, the transition to
special relativity might be taken to be a particularly clear example
of failed reference --- showing, in particular, that pre-Einsteinian
physicists ought not to have taken their theory so seriously.
Nonetheless, says Worrall, it would have been a mistake for
pre-Einsteinian physicists to treat their theory instrumentally,
i.e.\ merely as a tool for making predictions. For the form of the
equations of motion was preserved through the transition to special
relativity --- hence, they would have done well to trust their
equations. The general lesson, says Worrall, is to trust your
theory's structure, but not the underlying stuff it purports to be
talking about.
Worrall's example is highly suggestive, and we might like to apply it
in a forward-looking direction. In particular, take one of our
current-day successful scientific theories $T$, such as quantum
mechanics. The pessimistic metainduction suggests that $T$ will be
wrong about something. But can we already make an educated guess
about which parts of $T$ will be preserved, and which part will go on
the scrap heap with other rejected theories?
\cite{zahar2001,cruse,zahar2004} provide a specific proposal for
picking out the structural commitments of a theory $T$: they are given
by its Ramsey sentence $T^R$. This idea certainly has some intuitive
appeal --- trading on an analogy to ``coordinate free'' descriptions
of space. For a naive or straightforward description of physical
space, we might use triples of real numbers, i.e.\ the mathematical
space $\7R ^3$. But now our description of space has superfluous
structure. In particular, we assigned the origin $0\in\7R ^3$ to some
particular point in space --- but we didn't mean to indicate that the
denoted point is any different than any other point in space. Thus,
our description breaks the natural symmetry of space, and it would be
natural to look for another description that respects these
symmetries. Indeed, that's precisely the idea behind the move from
using vector spaces to using affine spaces to describe space.
Now, just as a vector-space description of space breaks its symmetry,
so our theoretical descriptions in general might fail to respect the
symmetry between properties. For example, we didn't need to use the
word ``electron'' to describe those things that are found in the
energy shells around an atom's nucleus --- we could simply say that
something or other plays the relevant role. And that's exactly what
the Ramsified theory says. Thus, it might seem that $T^R$ provides a
more intrinsic description than the original theory $T$.
Nonetheless, the intuitive appeal of Ramsey sentences fades quickly in
the light of critical scrutiny. Most famously, already in 1928,
Newman argued that Bertrand Russell's structuralism trivializes, for
these structural claims are true whenever their observational
consequences are true \cite[see][]{newman1928}. The so-called Newman
objection to structural realism has been the center piece of recent
debates about Ramsey-sentence structuralism. But even before we get
to that level of scrutiny, there is something quite strange in the
idea of passing to the Ramsified theory $T^R$ to get rid of
redundancy. Let's recall that a formal theory $T$ doesn't actually
refer to thinks like electrons or protons --- it's formulated in an
uninterpreted calculus. Hence, $T$ doesn't actually have any
referring terms.
It seems that the impulse to Ramsify is no other than the original
impulse to use uninterpreted mathematical symbols to represent
physical reality. You'll recall that one of the key maneuvers in the
development of non-euclidean geometries was de-interpreting words like
``line'', thereby liberating mathematicians to focus attention solely
on the relation that ``line'' plays relative to other (uninterpreted)
terms in their formal calculus.
In any case, what's really at stake here is the question of what
attitude we should take toward the best scientific theories of our day
and age. At one extreme, radical scientific realists assert that we
should give nothing less than {\it full} assent to these theories,
interpreted literally. To draw an analogy (that scientific realists
will surely eschew), the extreme scientific realist is akin to the
radical religious fundamentalist, and in particular to those
fundamentalists who say that one must interpret scriptures literally.
The point of that injunction, we all know, is to enable religious
leaders to foist their opinions on others. At the opposite extreme,
an extreme scientific antirealist sees science as having no epistemic
authority whatsoever --- i.e.\ a successful scientific theory doesn't
call for any more epistemic attention on my part than, say,
Zoroastrianism.
In the light of this somewhat hyperbolic characterization of the
anti/realism debate, we can see various alternative positions as
granting a selective epistemic authority to successful scientific
theories. Consider an analogy: suppose that you know a highly skilled
car mechanic Jacob. You completely trust Jacob when it comes to his
opinions on automobile-related issues. For example, if he says that
you need a new alternator, then you won't doubt him, even if it costs
you a lot of money. Nonetheless, if Jacob tells you that you need a
new kidney, or that you should vote for a certain candidate, you might
well ignore his opinion --- since he's speaking on a topic that lies
outside his proper expertise.
Now, selective scientific realists consider successful scientific
theories to be epistemically authoritative, but only when they speak
on topics within their expertise. The different brands of selective
realism are distinguished by how they understand the expertise of
science. For example, a constructive empiricist (such as van
Fraassen) trusts a successful scientific theory $T$ when it makes
predictions about empirical phenomena (presupposing, as he does, that
it makes sense to speak of predictions and empirical phenomena ---
precisely the point to which Boyd and Putnam object). Similarly, a
structural realist (such as Worrall) trusts a successful scientific
theory $T$ on its structural pronouncements. But if $T$ says
something about things in themselves (or whatever is {\it not}
structure), then the structural realist treats it as no more of an
authority than your auto mechanic is on politics.
The previous considerations suggest that varieties of selective
scientific realism can be classified by means of different notions of
theoretical equivalence. For example, the strict empiricist thinks
that the important part of a theory is its empirical content; and
hence, if two theories $T_1$ and $T_2$ agree on empirical content,
then there is no epistemically relevant difference between them.
Similarly, a structural realist thinks that the important part of a
theory is its pronouncements about structure; and hence, if two
theories $T_1$ and $T_2$ agree on structure, then there is no
epistemically relevant difference between them. In the particular
case of Ramsey-sentence structuralism, the structural pronouncement's
of a theory $T_i$ are captured by its Ramsey sentence $T_i^R$. Hence,
if $T_1^R\equiv T_2^R$, then there is no epistemically relevant
difference between $T_1$ and $T_2$.
Unfortunately, the statement ``$T_1^R\equiv T_2^R\,$'' doesn't have an
obvious meaning, since there is no single, obviously correct notion of
second-order logical consequence. What this means is that we get
different notions of ``same structure'' depending on which notion of
second-order consequence we adopt. Let's review, then, some salient
notions of second-order logical consequence.
Second-order logic is a complicated subject in its own right, and has
been the source of much dispute among analytic philosophers. We refer
the reader to studies such as \citep{shapiro,bueno} for more details.
For present purposes, it will suffice to make some minor modifications
of first-order logical grammar: first, we add a list of second-order
variables $X,Y,\dots $. Each second-order variable has a specific
arity $n\in\7N$, which means that it can stand in the place of an
$n$-ary relation symbol. We then permit formulas such as
$X(x_1,\dots ,x_n)$, with a second-order variable of arity $n$ applied
to $n$ first order variables. We also add an existential quantifier
$\exists X$ that can be applied to quantify over second-order
variables.
Now there are two important facts to keep in mind about second-order
logic. The first fact to keep in mind is that second-order logic has
is intrinsically incomplete --- hence there is no tractable syntactic
relation ``$\vdash$'' of second-order provability. The second fact to
keep in mind is that there are several candidates for the semantic
relation ``$\vDash$'' of entailment. Depending on which choice we
make for this relation, we will get a different notion of logical
equivalence.
\begin{defn} A second-order $\Sigma$-frame $\2F=(M,(\2E )_{n\in\7N })$
consists of a first order $\Sigma$-structure $M$, and for each
$n\in \7N$, a subset $\2E _n$ of $\2P(M^n)$. We let
$\2E ^{\2F}=\bigcup _{n\in\7N}\2E _n$. Here the sets in
$\2E ^{\2F}$ will give the domain of the second-order quantifiers in
frame $\2F$. \end{defn}
In order to define the relation $\vDash$, we will also make use of the
notion of a variable assignment. Given a $\Sigma$-frame $\2F$, a
first-order variable assignment $g$ assigns each variable $x$ to an
element $g(x)\in M$. A second-order variable assignment $G$ assigns
each variable $X$ of arity $n$ an element $G(X)\in \2E _n$. We then
define:
\begin{quote} $M[G,g] \vDash \exists X\phi$ iff for some $E\in \2E_n$,
$M[G^E_X,g]\vDash \phi$, where $G^E_X$ is the second-order variable
assignment that agrees with $G$ on everything besides $X$, which it
assigns to $E$. \end{quote} Now to define the relation $\vDash$
between sentences, we have to decide which second-order
$\Sigma$-frames to quantify over. We get three different notions,
depending on the family we choose:
\begin{enumerate}
\item For \emph{full semantics}, we permit only those $\Sigma$-frames
in which $\2E _n=\2P(M^n)$.
\item For \emph{Henkin semantics}, we permit all $\Sigma$-frames in
which $\2E _n$ is closed under first-order definability.
\item For \emph{frame semantics}, we permit all $\Sigma$-frames.
\end{enumerate}
Recall that the more structures, the more counterexamples, and hence
the fewer implications. Accordingly, full semantics has more
entailments than Henkin semantics, and Henkin semantics has more
entailments than frame semantics; hence, full semantics yields a
more liberal notion of equivalence than Henkin semantics, which
yields a more liberal notion of equivalence than frame semantics.
In the following discussion, we will take $T_i$, for $i=1,2$, as a
theory in signature $\Sigma\cup \Sigma _i$, where $\Sigma _i$ is
disjoint from $\Sigma$. We let $T_i^*$ be the result of replacing
terms in $\Sigma _i$ with (possibly second-order) variables, and we
let $T_i^R$ be the corresponding Ramsey sentence of $T_i$. We now
give a general schema for Ramsey equivalence of theories:
\begin{defn} Two theories $T_1$ and $T_2$ are \emph{Ramsey
equivalent} if $T_1^R$ is logically equivalent to
$T_2^R$. \end{defn} \noindent The three choices of frames above
give rise to three notions of Ramsey equivalence.
\begin{itemize}
\item RE$_1$ = loose Ramsey equivalence = Ramsey sentences are
equivalent relative to full semantics.
\item RE$_2$ = moderate Ramsey equivalence = Ramsey sentences are
equivalent relative to Henkin semantics.
\item RE$_3$ = strict Ramsey equivalence = Ramsey sentences are
equivalent relative to frame semantics.
\end{itemize}
Obviously then we have RE$_3$ $\Rightarrow$ RE$_2$ $\Rightarrow$
RE$_1$.
We can now give a sharpened formulation of the Newman problem --- in
the spirit of \cite{ketland} and \cite{dew-ram}. Recall that on the
old-fashioned syntactic view of theories, two theories $T_1$ and $T_2$
are considered to be empirically equivalent if they have the same
consequences in the observation language. If we now think of $\Sigma$
as the observation vocabulary, then we could formulate this criterion
as saying that $\cn{T_1}|_\Sigma=\cn{T_2}|_\Sigma$, where
$\cn{T_i}|_\Sigma$ indicates the restriction of the set of
consequences to those that contain only observation terms.
One might also wish to formulate a more semantically oriented notion
of empirical equivalence. For example, we might say that two theories
$T_1$ and $T_2$ are empirically equivalent if their models agree on
$\Sigma$-structure.
\begin{defn} We say that $T_1$ and $T_2$ are $\Sigma$-equivalent just
in case for each model $M$ of $T_1$, there is a model $N$ of $T_2$
and an isomorphism $h:M|_\Sigma\to N|_\Sigma$, and vice
versa. \end{defn}
The following result shows that this semantic notion of empirical
equivalence implies the syntactic notion.
\begin{prop} If $T_1$ and $T_2$ are $\Sigma$-equivalent then
$\cn{T_1}|_\Sigma=\cn{T_2}|_\Sigma$. \end{prop}
\begin{proof} Suppose that $T_1$ and $T_2$ are $\Sigma$-equivalent.
Let $\phi$ be a $\Sigma$-sentence such that $\phi\not\in\cn{T_2}$.
By completeness, there is a model $M$ of $T_2$ such that
$M\not\vDash\phi$. Since $T_1$ and $T_2$ are $\Sigma$-equivalent,
there is a model $N$ of $T_1$ and an isomorphism
$h:M|_\Sigma\to N|_\Sigma$. But then $N\not\vDash\phi$, hence
$\phi\not\in \cn{T_1}$. It follows that
$\cn{T_1}|_\Sigma\subseteq \cn{T_2}|_{\Sigma}$. The result follows
by symmetry.
\end{proof}
However, this implication cannot be reversed, i.e.\ the syntactic
notion of empirical equivalence doesn't imply the semantic notion.
\begin{example} Let $\Sigma$ be the empty signature (with equality).
Let $\Sigma _1 = \{ c_r\mid r\in \7R \}$, and let $T _1$ be the
theory in $\Sigma\cup\Sigma _1$ with axioms $c_r\neq c_s$, for all
$r\neq s$. Let $T_2$ be the theory in $\Sigma$ that says there are
infinitely many things. Then $\cn{T_1}|_\Sigma=\cn{T_2}|_\Sigma$.
However, $T_2$ has a countable model $M$, and $T_1$ has no countable
model. Therefore, $T_1$ and $T_2$ are not $\Sigma$-equivalent.
\end{example}
The Newman problem for structural realism is usually phrased as saying
that it's too easy for a theory's Ramsey sentence to be true --- that
the Ramsey sentence is ``trivially realizable''. We can make precise
what is meant here by ``too easy'' in terms of the notion of
theoretical equivalence. In short, Ramsey equivalence --- i.e.\
having logically equivalent Ramsey sentences --- is too liberal a
notion of equivalence. In particular, empirically equivalent theories
are Ramsey equivalent.
\begin{prop}[Dewar] If $T_1$ and $T_2$ are $\Sigma$-equivalent, then
$T_1^R$ and $T_2^R$ are logically equivalent relative to full
semantics. \end{prop}
\begin{proof} Suppose that $T_1$ and $T_2$ are $\Sigma$-equivalent.
Now let $\2F$ be a full $\Sigma$-frame such that $\2F\vDash T_1^R$.
Thus, there is a second-order variable assignment $G$ such that
$\2F [G]\vDash T_1^*$. Let $M$ be the $\Sigma\cup\Sigma _1$
structure obtained by assigning $M(R)=G(X)$, where $X$ is the
variable in $T_1^*$ that replaces $R$ in $T_1$. Clearly $M$ is a
model of $T_1$. Since $T_1$ and $T_2$ are $\Sigma$-equivalent, $M$
is $\Sigma$-isomorphic to a model $N$ of $T_2$. This model $N$ of
$T_2$ defines a second order variable assignment $G'$ such that
$\2F [G']\vDash T_2^*$, and hence $\2F\vDash T_2^R$. \end{proof}
The notion of empirical equivalence imposes no constraints whatsoever
on what the theories $T_1$ and $T_2$ say in their theoretical
vocabulary --- and for this reason, nobody but the most extreme
empiricist should adopt weak Ramsey equivalence as their standard.
Moving back toward the right wing side of the spectrum of theoretical
equivalence, one might hope that moderate Ramsey equivalence would
provide a more reasonable standard. But the following result shows
that any two mutually interpretable theories satisfy RE$_2$.
\begin{prop}[Dewar] If $T_1$ and $T_2$ are $\Sigma$-equivalent and
mutually interpretable, then $T_1^R$ and $T_2^R$ are logically
equivalent relative to Henkin semantics. \end{prop}
\begin{proof} Suppose that $T_i$ is a theory in
$\Sigma\cup \Sigma _i$. We will show that if $F:T_1\to T_2$ is a
translation (which is the identity on $\Sigma$), then
$T_2^R\vDash T_1^R$, where the $\vDash$ symbol is entailment
relative to Henkin semantics, and $T_i^R$ is the result of
Ramsefying out $\Sigma _i$. Suppose then that $F:T_1\to T_2$ is a
translation, and that $\2H$ is a Henkin structure (of signature
$\Sigma$) such that $\2H\vDash T_2^R$. Thus, $\2H [G]\vDash T_2^*$
relative to some second-order variable assignment $G$. Consider
then the first-order structure $M$ for signature
$\Sigma\cup \Sigma _1$ that agrees with $\2H$ on $\Sigma$, and such
that $M(P)=G(X_P)$, for each $P\in \Sigma _2$, where $X_P$ is the
second-order variable that replaces $P$ in $T_1^*$. It is clear
then that $M\vDash T_2$. Now we will use the fact that the
translation $F:T_1\to T_2$ gives rise to a functor
$F^*:\mathrm{Mod}(T_2)\to \mathrm{Mod}(T_1)$ (\ref{mfunc}). In
particular, $(F^*M)(Q)=M(F(Q))$ for each relation symbol
$Q\in\Sigma\cup\Sigma _1$. Now define a second-order variable
assignment $G'$ by setting
\[ G'(X_Q) \: = \: (F^*M)(Q) \: = \: M(F(Q)) ,\] for each variable
$X_Q$ that occurs in $T_1^*$. (Again we use $X_Q$ to denote the
variable that replaces a relation symbol $Q$ that occurs in
$T_1$.) To see that $G'$ is a Henkin-admissible assignment, note
that $F(Q)$ is a $\Sigma _2$-formula, and so $M(F(Q))$ is a
first-order definable subset of $M$. By construction, each
first-order definable subset of $M$ is an element of $\2E ^{\2H}$.
Now, it's clear that $\2H [G']\vDash T_1^*$, and hence that
$\2H\vDash T_1^R$. Since $\2H$ was an arbitrary Henkin frame, it
follows that $T_2^R\vDash T_1^R$. By symmetry, if there is a
translation $G:T_2\to T_1$, then $T_1^R\vDash T_2^R$. Therefore,
if $T_1$ and $T_2$ are mutually interpretable, then $T_1^R$ and
$T_2^R$ are Henkin equivalent. \end{proof}
There is one last hope for the Ramsifier: that strict Ramsey
equivalence (RE$_3$) will provide the right notion of structural
equivalence. Unfortunately, RE$_3$ proves to be the worst candidate
for structuralism, since intertranslatable theories need not satisfy
RE$_3$.
\begin{example} Let $\Sigma _1 = \{ r\}$, and let
$\Sigma _2=\{ r'\}$, where both $r$ and $r'$ are unary predicates.
Let $T_1=\{ \exists xr(x) \}$, and let
$T_2=\{ \exists x \neg r'(x)\}$. The reconstrual
$F(r)=\neg r'(x)$ induces a homotopy equivalence between $T_1$ and
$T_2$, i.e.\ $T_1$ and $T_2$ are intertranslatable. However, the
Ramsey sentences of $T_1$ and $T_2$ are not frame equivalent. In
particular, consider any frame $\2F$ with first-order domain $M$,
and $\2E ^{\2F}_1=\{ M\}$, i.e.\ $M$ is the only admissible subset
of $M$. Then $\2F \vDash T_1^R$ but $\2F \not\vDash
T_2^R$. \end{example}
Since strict Ramsey equivalence (RE$_3$) is more conservative
(``right wing'') than definitional equivalence, we don't expect
structural realists to find it congenial. But what about those
hard-core realists --- like David Lewis or Ted Sider --- who pin
their theoretical hopes on natural properties and reference
magnetism? Might they actually want a criterion of equivalence that
is even more conservative than definitional equivalence? In fact,
it seems that frame semantics might be a good way to capture the
idea that to describe a possible world, you need to say not only
what things exist, but also what the natural properties are. We
should note, however, that adopting a first-order signature $\Sigma$
already goes some way to picking out natural properties. When we
specify a $\Sigma$-structure $M$, we get a natural property
$M(\phi )$ for each formula $\phi$ of $\Sigma$. It's not clear then
why a theorist who has adopted a first-order signature $\Sigma$
would need to additionally specify a notion of natural properties.
The previous results can be summarized in the following diagram:
\[ \begin{tikzcd}
\text{EE} \arrow[Rightarrow]{d} & \text{MI} \arrow[Rightarrow]{l}{} \arrow[Rightarrow]{d} & \text{IT} \arrow[Rightarrow]{l}{} \arrow[strike through,Rightarrow]{dr}{} \\
\mathrm{RE_1} & \mathrm{RE_2} \arrow[Rightarrow]{l}{} & &
\mathrm{RE_3} \arrow[Rightarrow]{ll}{} \end{tikzcd} \] Here ``EE''
is empirical equivalence (explicated semantically), ``MI'' is mutual
interpretability over $\Sigma$, and ``IT'' is intertranslatability
over $\Sigma$, which is equivalent to definitional equivalence. It
appears then that none of the notions of Ramsey equivalence gets us
near the promising area in the neighborhood of intertranslatability.
Most philosophers, we think, would agree that intertranslatability is
a reasonable --- if somewhat strict --- explication of the idea that
two theories have the same logical structure.
\section{Counting possibilities}
%% counting the number of possibilities
%% are they the same or different? A question that doesn't really
%% have a good answer
%% need to account for relations between worlds
If you page through an analytic philosophy journal, it won't be long
before you see the phrase ``possible world''. Many philosophical
discussions focus on this concept, and it is frequently used as a
basis from which to explicate other concepts --- Humean supervenience,
counterfactuals, laws of nature, determinism, physicalism, content,
knowledge, etc.. When the logically cautious philosopher encounters
this concept, she will want to know what rules govern its use. Where
things get really tricky is when philosophers start invoking facts
about the structure of the space of possible worlds --- e.g.\ how
many worlds there are, which worlds are similar, and which worlds are
identical. These sorts of assumptions play a significant role in
discussions of fundamental ontology. To take a paradigm example,
\cite{baker2010} argues that if two models are isomorphic, then they
represent the same possible world.
Analytic philosophers might be the primary users of the phrase
``possible world'', but they aren't the only ones using the concept.
Scientists talk about possible worlds all the time. However, at least
in the exact sciences, there are explicit rules governing the use of
possible worlds talk. Indeed, these rules are built into the
structure of their theories, and more particularly, in the structure
of those theories' spaces of models. Following \cite{elvis}, we think
that philosophers ought to try to understand the way that scientists'
theories guide their use of modal concepts.
Nonetheless, it's not hard to find philosophers scratching their
heads, and asking themselves questions like the following:
\begin{quote} ($\star$) Consider two general relativistic spacetimes,
$M$ and $N$, and suppose that $h:M\to N$ is an isomorphism (e.g.\ a
metric preserving diffeomorphism). Do $M$ and $N$ represent the
same possible world? \end{quote}
\begin{quote} ($\ast$) Consider two Newtonian spacetimes, $M$ and $N$,
and suppose that $h:M\to N$ is an isomorphism (e.g.\ a shift). Do
$M$ and $N$ represent the same possible world? \end{quote}
\cite{elvis} helpfully classifies philosophers into two groups
according to how they answer these questions: the {\it shiftless}
claim that isomorphisms do not generate new possibilities, and the
{\it shifty} claim that isomorphisms do generate new possibilities.
In particular, the shiftless philosopher says that if $h:M\to N$ is an
isomorphism, then $M$ and $N$ represent the same possibility. In
contrast, the shifty philosopher allows that $M$ and $N$ might
represent different possibilities, even though they are isomorphic.
While the majority of philosophers of physics and metaphysicians have
become shiftless, Belot champions the heterodox, shifty point of view.
As we will now argue, all parties to the dispute have adopted a
questionable presupposition, viz.\ that it makes sense to count
possibilia.
But first, what hangs on this dispute between the shifty and the
shiftless? In the first place, shiftless philosophers believe that
they are on the right side of history, ontologically speaking. In
particular, they believe that it would be wrong to countenance the
existence of two possibilities, represented by $M$ and $N$, when a
single one will do the job. This way of thinking trades on vague
associations with Leibniz's principle of the identity of
indiscernibles: since $M$ and $N$ are indiscernible, there is no
reason to regard them as different. Belot points out, however, that
shiftless philosophers have trouble making sense of how theories can
guide the use of modal concepts. In particular, he argues that the
shiftless view necessarily collapses the distinction between
deterministic and indeterministic theories.
The mind boggles at the existence of shiftless philosophers ---
because the view so patently conflicts with the standard reading of
physical theories, not to speak of plain common sense. Take, for
example, a Galilean spacetime $M$, and let $\gamma :\7R\to M$ be an
inertial worldline in $M$. Now, a boost $x\mapsto x+vt$ for some
fixed $v>0$ is represented by an isomorphism $h:M\to M$. Does this
boost generate a new possibility? The question might seem confusing
because the model on the right side of $h:M\to M$ is the same as the
model on the left side. It might seem to be trivially true, then,
that $h:M\to M$ does {\it not} generate a new possibility. But let's
see what happens if we adopt the shiftless view. If $h$ does {\it
not} generate a new possibility, then we ought to say, of a particle
in inertial motion that it could {\it not} be in some other state of
inertial motion (because there is no other such state of inertial
motion). But that claim is completely contrary to the way that
physicists use this theory to guide their modal reasoning. When a
physicist adopts Galilean relativity, she commits to the claim that
there are many distinct possible states of inertial motion, and that a
thing which is in one state of inertial motion {\it could be} in some
other state of inertial motion. In other words, it matters to
physicists that the isomorphism $h:M\to M$ is not the identity
isomorphism, and in particular, that the worldline $h\circ \gamma$ is
not the same as the worldline $\gamma$. Nonetheless, shiftless
philosophers can't make sense of these modal claims, because they
insist that isomorphisms don't generate new possibilities.
Despite the implausibility of the shiftless view, there are some very
serious and smart philosophers who defend it. What is it, then, that
really drives their insistence on saying that isomorphism (at the
level of representations) implies identity (at the level of the
represented)? We suspect that the shiftless are fumbling their way
toward an insight --- but an insight that is difficult to articulate
when one is operating with mistaken views about mathematical objects,
and in particular about the relation between abstract and concrete
objects. Of course, we blame a lot of this confusion on Quine and his
followers, who decided that we have no need of the abstract-concrete
distinction --- in particular, that belief in the existence of
abstracta is no different in principle from belief in the existence of
concreta. But without that distinction, we are bound to say some
silly things about the representation relation.
At risk of oversimplifying, we will first give a simple formulation of
the basic insight toward which we think the shiftless philosophers are
fumbling:
\begin{quote} ($\dagger$) A theory $T$ is indifferent to the question
of the identity of its models. In other words, if $M$ and $N$ are
models of $T$, then $T$ neither says that $M=N$, nor that $M\neq N$.
The only question $T$ understands is: are these models isomorphic,
or not?
\end{quote}
Now please don't get us wrong: ($\dagger$) does not say that
isomorphic models are identical, nor does it say that the theory $T$
treats isomorphic models as if they were identical. No, from the
point of view of $T$, the question, ``are they identical?'' simply
does not make sense. And claims of identity, or non-identity of
models, play {\it no} explanatory role in the theory.
We realize that this claim is controversial, and that it might take
some time for philosophers to become comfortable with it. The problem
is that we learned a little bit of set theory in our young years, and
we seem to assume that everything lives in a world of sets --- where
questions of the form ``is $M$ equal to $N$'' always have a definite
answer. Indeed, the rigid grip of set theory makes philosophers
profoundly uncomfortable with contemporary mathematics, which likes to
play a fast and loose game with identity conditions. Consider a
simple example (due to John Burgess): suppose that we ask two
different mathematicians two different questions:
\begin{quote} (Q1) How many groups are there with two
elements? \end{quote}
\begin{quote} (Q2) Inside the group $\7Z _2\oplus \7Z _2$, how many
subgroups are there with two elements? \end{quote} What we are
likely to find is that mathematicians will give apparently conflicting
answers. On the one hand, they will tell us that there is only one
group with two elements. On the other hand, they will tell us that
$\7Z _2\oplus \7Z _2$ has two distinct subgroups with two elements.
Obviously, if taken literally, these two answers contradict each
other. But of course, there is no genuine conflict, and
mathematicians are not in crisis about the number of groups with two
elements. No, the fact is, mathematicians use words and symbols in a
different way than we use them in everyday life --- e.g.\ when we
count the number of apples in a basket.
To reinforce this point, recall that categorical equivalence doesn't
respect the number of objects in a category. Consider for example the
following two categories: let $\cat{C}$ be the category with one
object, and one identity morphism. Let $\cat{D}$ be the category with
two objects $a,b$, one identity morphism from each object to itself,
and a pair of morphisms $f:a\to b$ and $g:b\to a$ that are inverse to
each other. Then $\cat{C}$ and $\cat{D}$ are equivalent categories
--- which entails that ``this category'' doesn't really have a
definite number of objects. It is not correct to say that it has one
object, and it's not correct to say that it has two. Or perhaps
better: it is just as correct to say that it has one object as it is
to say that it has two.
Here then is our positive proposal:
\begin{quote} For the purposes of interpreting a theory $T$, the
collection $\mathrm{Mod}(T)$ of its set-theoretic models should be
treated as nothing more nor less than a {\it category}. In
particular, the philosopher of science shouldn't say things about
$\mathrm{Mod}(T)$ that are not invariant under categorical
equivalence, nor should they argue over questions --- such as ``how
many models does $T$ have?'' --- whose answer is not invariant under
categorical equivalence. \end{quote} If this proposal is adopted
then there is no debate to be had between the shifty and the
shiftless. The question they are asking --- do isomorphisms generate
new possibilities? --- depends on a notion (the number of isomorphic
possibilities) that is not invariant under categorical equivalence.
The rationale for this proposal is our belief that models of a theory
$T$ in $\cat{Sets}$ are {\it representations} of that theory; the
set-theoretic description of these models is not {\it not} itself a
further theory that attempts to describe the world at an even finer
grained level of detail than was done by $T$. We can further clarify
these points by means of a simple example.
\begin{example} Suppose that Berit is a scientist with a very simple
theory. Her language $\Sigma$ has a single predicate symbol $P$,
and her theory $T$ says that there are exactly two things, one of
which is a $P$:
\[ \exists x\exists y(P(x)\wedge \neg P(y)\wedge \forall z((z=x)\vee
(z=y))) .\] Now we metatheorists know that a set-theoretic model
$M$ of $T$ consists of a two-element set, say $X=\{ a,b\}$, with a
singleton set $M(P)$. Let $M$ be the model such that $M(P)=\{ a\}$,
and let $N$ be the model such that $N(P)=\{ b\}$. Then the
permutation $h(a)=b,h(b)=a$ gives a $\Sigma$-isomorphism $h:M\to N$.
(But the permutation $h$ is not an automorphism of $M$.)
Let's consider the shifty-shiftless dilemma with regard to the
models $M$ and $N$, with the isomorphism $h:M\to N$. The shifty
philosopher (e.g.\ Belot) says that $M$ and $N$ represent distinct
possibilities. The shiftless philosopher (e.g.\ Baker) says that
$M$ and $N$ represent the same possibility. Who is on the side of
truth?
In our opinion, both the shifty and the shiftless say misleading
things about this example. On the one hand, the shifty claim is
misleading, because the user of $T$ doesn't have the language to say
what would be different between $M$ and $N$. She cannot say, ``in
$M$, $a$ is $P$, and in $N$, $a$ is not $P$,'' because she herself
doesn't have the name ``$a$''. The shiftless wants us to start
counting how many models there are, but the theory $T$ doesn't
answer that question.
On the other hand, the shiftless would insist that there is only one
possibility, represented redundantly by $M$ and $N$. But that claim
is misleading for the following reason. Berit's theory $T$ is an
extension of the theory $T_0$, in empty signature, that says there
are exactly two things. Let $I:T_0\to T$ be the translation of
$T_0$ into $T$, and let $I^*:\mathrm{Mod}(T)\to\mathrm{Mod}(T_0)$ be
the functor that forgets the assignment of $P$. Here $I^*M$ and
$I^*M$ are both the bare two-point set $X$, and the isomorphism
$I^*h=h:X\to X$ is the non-trivial permutation. Recall, though,
that functors map identity morphisms to identity morphisms. Hence,
if the isomorphism $h:M\to N$ is considered to be an identity (as
the shiftless seem to do), then it would follow that $I^*h$ is the
identity morphism. Thus, contra the shiftless, we cannot identify
$M$ and $N$, and forget that there was a non-identity isomorphism
$h:M\to N$. If we do that, then we won't be able to see how the
theory $T$ is related to the theory $T_0$.
The confusion here is somewhat similar to Sk{\o}lem's paradox (about
the existence of uncountable sets in models of ZF set theory), where
we run into trouble if we don't distinguish between claims made in
the object language and claims made in the metalanguage. In the
present case, one might be tempted to think of the theory $T$ as
saying things such as:
\begin{quote} In model $M$, $a$ is a $P$. \end{quote} Of course, $T$
says no such thing, since it doesn't have names for models or for
elements in models.
The other problem here is in the way that we've set up the problem
--- by speaking as if the representation relation holds between $M$
(or $N$) and the world. To the contrary, the representation
relation holds between Berit's language and the world, and we (the
metatheorists) are representing Berit's theorizing using our own
little toy theory (which presumably includes some fragment of set
theory, because that's a convenient way to talk about collections of
formulas etc.). Berit herself doesn't claim that $M$ (or $N$)
represents the world --- rather, the metatheorist claims that $M$
and $N$ represent ways that Berit's language could represent the
world. Accordingly, Berit doesn't claim that $M=N$, or that
$M\neq N$; those are metatheoretical assertions --- and do not add
to the stock of knowledge about the world. \end{example}
Before proceeding, we should deal with an obvious objection to the
view we've put forward. Some philosophers will point out that it is
simply false to say that physicists don't count the number of
possibilities. Indeed, it's precisely by counting the number of
possibilities that physicists derive notions such as entropy.
We do not disagree with this point, but it doesn't conflict at all
with our positive proposal (to talk about models of a theory as a
category). Just remember that category theory is a framework that is
almost infinitely flexible: what we can talk about in a
categorically-invariant way depends on how we --- or physicists ---
define the relevant category. For the case at hand, if $X$ is a
classical phase space, then it is assumed that $X$ is a discrete
category --- i.e.\ that there are no non-trivial isomorphisms between
elements of $X$. Thus, in this case, there is no question about
whether to count two isomorphic possibilities as the same; because we
(or better, the physicists) have chosen not to admit isomorphic
possibilities.
To be clear, we explicitly reject the idea that there is a single
relation ``being isomorphic'' that either holds or does not hold
between concrete objects. On the contrary, the notion of isomorphism
applies to abstractions, and different notions of isomorphism are
valid for different levels of abstraction. It's up to us to decide
which level of abstraction serves our purposes in reasoning about
concrete, physical reality. (In particular, models of a theory are
not concrete realities, and that's why they cannot either be identical
or non-identical.)
For all of its other virtues, one of the defects of the semantic view
of theories is that it obscures the object/meta-language distinction,
a distinction that is absolutely necessary to make sense of the notion
of symmetry of representations. To be more accurate, the target of
this criticism are advocates of the ``language-free'' or
``semantic-L'' view \cite[see][]{halvorson2013}. The picture we get
from the language-free semantic view is that mathematical structures
are out there in the world, and that they are either isomorphic to
each other, or they are not. Of course, that picture completely
ignores the fact that isomorphisms are defined in terms of language;
or, to put it more accurately, that isomorphisms relate mappings
$M:\Sigma\to\cat{Sets}$ and $N:\Sigma\to\cat{Sets}$, which have a
common domain $\Sigma$. Thus, in particular, arbitrary mathematical
structures are neither isomorphic nor non-isomorphic.
The object language $\Sigma$ serves as the reference point in defining
a notion of symmetry. The object language tells us what must be held
fixed, and the metalanguage tells us what can be varied. In
particular, a model $M$ of a theory $T$ can have a non-trivial
automorphism group because of two features of the formal setup:
\begin{enumerate}
\item The metalanguage describes the world in finer-grained language
than the object language.
\item Distinctions that are not made by the object language are not
significant for the kinds of explanations that the theory $T$ gives.
\end{enumerate}
If we drop either one of those components, then we will most likely
make a hash of the notion of symmetry. Without the metalanguage,
there is no way to see any difference between $a$ and $b$, and so no
way to express the change the occurs in the permutation $a\mapsto b$.
But if we think of the metalanguage as a better object language, then
we shouldn't count $a\mapsto b$ as a symmetry, since these two things
are distinguishable in the metalanguage. Thus, it's precisely the
mismatch between object language and metalanguage that provides us
with a rich notion of symmetry; and conversely, the importance of the
notion of symmetry gives us reason to maintain a distinction between
object language and metalanguage.
The distinction between object- and meta-language is one of the most
interesting ideas in 20th century logic and philosophy --- and it
remains one of the least well understood. Obviously, Carnap made a
lot of this distinction, and in fact, he seems to use it as his
primary analogy in formulating the distinction between internal and
external questions, and more generally in understanding the
relationship between theories in the exact sciences and our other,
non-scientific beliefs and attitudes. In contrast, Quine seems to
reject the idea that there is an important difference of status
between object and metalanguage. He seems to propose, instead, that
the ascent to metalanguage should be seen as an extension of one's
object language --- and so assertions in the metalanguage have exactly
the same force as assertions in the object language.\footnote{I fear
that I may be missing the nuance in Quine's position. I would
encourage the interested reader to consult more careful historical
studies such as \cite[][pp 182ff]{becker} and \cite{hintikka}.}
\section{Putnam's paradox} \label{putnam}
Perhaps the most notorious argument from logical metatheory to