You are on page 1of 18

Erkenn

DOI 10.1007/s10670-013-9578-5

ORIGINAL ARTICLE

Whats Right with a Syntactic Approach to Theories


and Models?

Sebastian Lutz

Received: 31 August 2012 / Accepted: 27 September 2013


Springer Science+Business Media Dordrecht 2014

Abstract Syntactic approaches in the philosophy of science, which are based on


formalizations in predicate logic, are often considered in principle inferior to
semantic approaches, which are based on formalizations with the help of structures.
To compare the two kinds of approach, I identify some ambiguities in common
semantic accounts and explicate the concept of a structure in a way that avoids
hidden references to a specific vocabulary. From there, I argue that contrary to
common opinion (i) unintended models do not pose a significant problem for
syntactic approaches to scientific theories, (ii) syntactic approaches can be at least as
language-independent as semantic ones, and (iii) in syntactic approaches, scientific
theories can be as well connected to the world as in semantic ones. Based on these
results, I argue that syntactic and semantic approaches fare equally well when it
comes to (iv) ease of application, (v) accommodating the use of models in the
sciences, and (vi) capturing the theory-observation relation.

1 Introduction

The analysis of scientific theories needs some frameworka way in which theories
and possibly the world are described and, building on that, a set of tools for analysis.
The earliest formal framework, developed within logical empiricism and especially
by Carnap and Hempel, has been dubbed the Received View. It relies on
formalizations of scientific theories in languages of predicate logic and assumes that
the non-logical vocabulary is bipartitioned into observational and theoretical terms,
where only the observational terms are directly interpreted. The interpretation of the
theoretical terms is fixed only by the interpretation of the observation terms, the

S. Lutz (&)
Munich Center for Mathematical Philosophy, Ludwig-Maximilians-Universitat Munchen,
Geschwister-Scholl-Platz 1, 80539 Munchen, Germany
e-mail: sebastian.lutz@gmx.net

123
S. Lutz

formalized theory, and additional sentences (correspondence rules) containing both


observational and theoretical terms. The Received View is often regarded as
superseded by frameworks in which scientific theories are assumed to be formalized
in model or set theory, which I will call semantic approaches.1 Calling
frameworks that rely on formalizations in predicate logic of first or higher order
syntactic, the Received View is a specific syntactic approach that additionally
assumes a bipartition of the vocabulary and allows a direct interpretation only of the
observational terms.2
Syntactic approaches have often been dismissed with reference to criticisms of
the Received View and to the alleged superiority of semantic approaches. In this
article, I will argue that this was a mistake, and defend syntactic approaches (but not
the Received View) relative to semantic approaches. Syntactic approaches are
widely thought to suffer from a number of shortcomings: Unlike semantic
approaches, it is claimed, syntactic approaches (i) cannot avoid unintended models,
(ii) are language dependent, (iii) require and have failed to provide an account of the
relation between language and the world, (iv) lead to cumbersome descriptions of
scientific theories, (v) do not allow the treatment of models, and are therefore at
odds with actual scientific practice, and (vi) presume a misleading or false relation
between theory and observation. I will argue that if the first three of these problems
can be solved for semantic approaches, they can be solved for syntactic ones as well.
That the same holds for the last three problems then follows easily. In discussing the
relative merits of syntactic approaches, I will ignore the ontological question of
whether theories can be identified with either kind of description (or, for that matter,
with platonic objects, sets of propositions, thoughts, actions, connection weights in
brains, combinations thereof,). Given that scientific theories are typically not
formalized in either kind of framework, they are probably ontologically different
from both kinds of descriptions.3

2 Translating Between Sentences and Models

In syntactic approaches, a theory is formalized by a set of sentences, in semantic


approaches by a class of structures (cf. van Fraassen 1980, 44). It seems plausible
that every syntactic description of a scientific theory can be captured by a semantic
one, because any set R of sentences determines the set S of its models through the
mapping

U : R7!S : fSjS  Rg: 1

1
I will use approach and framework interchangeably in the following.
2
Although this categorization of syntactic and semantic approaches is standard terminology, it is
somewhat incongruous: Set theory and predicate logic can both be used as foundational languages in
which to formalize other theories. Model theory, on the other hand, is one of those theories that can be so
formalized. I will discuss this point further at the end of 4.
3
The question what scientific theories really are may also simply not be well-defined, so that there is no
fact of the matter.

123
Whats Right with a Syntactic Approach?

R has a fixed vocabulary V, containing mi-place predicates Pi, nj-place functions Fj,
constants ck, and in higher order logic their respective types. V is contained in the
image of U as well, since every structure S 2 S contains a mapping from every
element of V to an extension with the corresponding arity and type (Hodges 1993,
24; Chang and Keisler 1990, 1.3). Call the arities and types of the extensions the
similarity type of V. V and its similarity type, sometimes called the signature of S,
can thus be read off uniquely from S (Hodges 1993, 4).4
However, U does lose some information because it cannot distinguish between
equivalent sets of sentences, that is, if R H, then UR UH. This can pose
problems, for example when modifying a theory: One formulation of a theory can
be vastly superior to an equivalent one when it needs to be generalized or adjusted
(van Fraassen 1980, 3.5). Relatedly, the formulation is also relevant when it comes
to the inductive support of components of the theory: If the data support one
postulate but not another, a formulation that keeps the two postulates separate is
arguably better than one that contains a single postulate equivalent to their
conjunction. The loss of distinction between equivalent sets of sentences does not
pose a problem, however, if the results of an analysis of a scientific theory are
invariant under the theorys equivalent reformulation; and outside of questions of
induction, many interesting analyses of scientific theories are so invariant. Indeed, it
is often considered a problem if an analysis of the theory is not (witness, for
example, Carnap 1956, 56).
Conversely, any set of structures S yields a set of sentences R through the
mapping

W : S7!R : fujS  u for all S 2 Sg: 2


WS is thus the set of sentences true in all elements of S. For many logics, W is not
the inverse of U. For instance, S may not be closed under elementary equivalence,
while for first order logic, UWS is. However, many non-isomorphic structures
that are elementary equivalent and hence cannot be distinguished in first order logic
can be distinguished in higher order logic. The use of logics that allow formulas
with infinitely many quantifiers, conjuncts, or disjuncts further increases the number
of structures that can be distinguished and thus the number of classes for which W is
the inverse of U. Call two structures that cannot be distinguished in some given
logic syntactically equivalent in that logic. For simplicity, I will assume in the
following that each class S of structures under discussion either can be captured by a
single set of sentences of the respective logic (so that UWS S) or is not closed
under syntactical equivalence.
As a matter of principle (that is, independently of the logic used), W always loses
information because it cannot distinguish between two non-identical, element-wise
isomorphic sets of structures, that is, if for each S 2 S, there is a T 2 T with S T
and vice versa,5 then WS WT even if S T. However, if the results of an

4
This is discussed at length in Sect. 3.
5
Note that this is a weaker condition than the point-wise isomorphism defined by Halvorson (2012, 190).

123
S. Lutz

analysis of a scientific theory are invariant under isomorphic transformations of


theories (a natural demand that is usually fulfilled),6 this does not pose a problem.
It is sometimes claimed that as a matter of principle, more information is lost in
the mapping W than the distinction between isomorphic structures. Suppe (2000,
S104), for example, argues that syntactic approaches are in principle unable to
capture some scientific theories, because for syntactic descriptions
the LowenheimSkolem theorem implie[s] that [] models must include both
intended and wildly unintended models. Unintended models provide potential
counterexamples. Blocking them more concerns eliminating syntactical-
approach artifacts than dealing with substantive analysis. [] For example,
Kitchers (1989) unification explanation account has a very simple idea. But
he develops it syntactically spending most of the paper trying to block
unintended consequences that are artifacts of his formalism. [] This is the
correct sense of [the] claim symbolic logic is an inappropriate formalism.
First and foremost, Suppes criticism is not directed at syntactic approaches (or
symbolic logic) in general, because the Lowenheim-Skolem theorem does not hold
in higher order logic. And while there are structures that even in predicate logic of
transfinite order cannot be described up to isomorphism (Enderton 2009, 2, 4), set
theory itself can be described up to isomorphism by (and only by) the addition of
axioms about the existence of specific inaccessible cardinals (Vaananen 2001, 516),
which is enough to capture the analyses in semantic approaches. Since the proof
theory of higher order logic is not complete for full semantics (Vaananen 2001,
505), entailment needs to be defined semantically in terms of structures. Luckily,
there is no good reason to disallow the use of structures in syntactic approaches,
since syntactic approaches presume only that scientific theories can be analyzed by
way of their description in predicate logic. They do not presume that the analysis
itself must proceed wholly in the object language. In fact, the use of structures for
defining entailment was even part of the Received View (Lutz 2012c, 83).
Second, for predicate logic of any order, some structures can be characterized in
that language up to isomorphism, so that there are no unintended models. The
standard examples of structures that cannot be captured in first order logic, e. g. the
natural numbers and the reals, can be described up to isomorphism in finite order
logic if a full semantics is assumed (Enderton 2009, 2). And if the theory has a
finite domain, all elementary equivalent structures are isomorphic (Hodges 1993,
2.2, ex. 5). Third, whether the existence of unintended models poses a problem
depends on the kind of analysis sought after. The answers to questions that can be
phrased in the object language, for example, depend at the most on syntactic
equivalence, since otherwise they would provide a means of syntactically
distinguishing between syntactically equivalent models. Finally, Suppes criticism
rests on an equivocation of unintended model and non-standard model, the latter
referring to a model that is syntactically equivalent but not isomorphic to a standard
model. Kitchers account is a good example of how difficult the formalization of an
idea can be, but is no example of a failure of a syntactic approach because of the

6
See, for example, van Fraassens notion of a theory (van Fraassen 2008, 238).

123
Whats Right with a Syntactic Approach?

Lowenheim-Skolem theorem. Put crudely, if the unintended consequences could be


blocked by using a predicate logic of higher order, Kitcher would probably not have
shied away from it for lack of a complete proof theory.

3 Language Independence

According to van Fraassen (1980, 44), the same class of structures could well be
described in radically different ways, each with its own limitations, and in the
words of Suppe (1989, 4), a semantic approach construes theories as what their
formulations refer to when the formulations are given a (formal) semantic
interpretation. French and Ladyman (1999, 114115) similarly assume that the
structures used in semantic approaches do not contain a vocabulary when they
discuss a criticism of semantic approaches they attribute to Mauricio Suarez: If a
semantic approach uses models as they are defined in model theory, it is still
dependent on a language, since a model is a structure and an interpretation of a
formal language in terms of that structure (that is, a map from the symbols of the
syntax to elements of the structure). If models are taken to involve such a mapping,
French and Ladyman (1999, 114) write, it is clear that the celebrated claim of the
linguistic independence of considering models (and not first-order formalizations of
theories), stressed by adherents of the semantic approach as giving it a clear
advantage over the syntactic view, is simply not true. A similar position is taken by
van Fraassen (1989, 366), and French and Ladyman (1999, 115) conclude that van
Fraassen should be interpreted as talking about structures by those who wish to
understand model in the sense of the standard logic texts.
However, it is not exactly clear what French and Ladyman mean by structure,
except that a structure contains the extensions of symbols, and not the symbols
themselves. They do claim that van Fraassens emphasis on structure is compatible
with this definition of model theory from a contemporary textbook, according to
which model theory is the study of the construction and classification of structures
within specified classes of structures. Of course, everything in this quote from
Hodges (1993, ix) depends on his definition of structure. Here is, for example, the
part of his definition that deals with relations (Hodges 1993, 2, my notation):
For each positive integer n [a structure contains] a set of n-ary relations on jAj
(i. e. subsets of jAjn), each of which is named by one or more n-ary relation
symbols. If R is a relation symbol, we write RA for the relation named by R.
It is clear that, as assumed so far and contrary to French and Ladymans claim, the
symbols play an important role in Hodges definition of a structure: They identify
the extensions by naming them.7
A definition better suited to French and Ladymans position can be found in a
less contemporary textbook (Bell and Slomson 1974, 3.2), in which

7
Indeed, Hodges (1993, 12) explicitly agrees with this assessment.

123
S. Lutz

a relational structure is an ordered pair

A hA; fRn jn 2 xgi;


where [] for n 2 x Rn is a finitary, say (n)-ary relation on A. []
The relational structure A will count as an interpretation of the language
L [ fPn : n 2 xg] if the degrees of the relations Rn correspond to the degrees
[dn] of the predicate letters Pn. That is, for n 2 x; dn kn. In this case
we say that the relational structure A is a realization of the language L and that
L is appropriate for the structure A.
This definition of what one could call indexed structures seems to be used by
many proponents of semantic approaches (e. g., da Costa and French 1990; French
and Ladyman 1999). It is free of any specific vocabulary up to the vocabularys
similarity type, that is, any appropriate vocabulary can be used with the structure.
Formally, however, an indexed structure is indistinguishable from a structure: A
structure A hA; I i contains an interpretation I mapping, say, the relations fPi gi2I
from a vocabulary V to their extensions, fPA i ji 2 Ig. An indexed structure instead
contains an indexed set fPA g
i i2I of extensions. 8
And this introduces a vocabulary
through the back door: The mapping from the index set I to the set fPA i ji 2 Ig that is
needed to define such an indexed set is formally indistinguishable from an
interpretation with the vocabulary I. Any claim that I is not a vocabulary but an
index set has to rely on commitments (or rather declarations) outside of the
formalism.
Suppes (2002) writes structures as tuples with the domain as first element, which
can avoid this hidden dependence on a specific vocabulary: Tuples may, for
example, be axiomatically introduced as primitives, or be defined through sets, as in
Kuratowskis definition. However, one can still introduce a vocabulary by
assigning, for example, natural numbers to the positions of the tuple, although
this assignment is not unique. In fact, any mapping from a well-ordered index set
with the right cardinality would do. The possibility of assigning a vocabulary to a
tuple cannot be avoided, since the ordering of the extensions is essential. If one
avoided the ordering by, for example, using a multiset of extensions instead of a
tuple, semantic approaches would be significantly less expressive than syntactic
ones. This follows from an example by Halvorson (2012, 192): Let H
f91 xx xg and K f91 xx xg [ f8xQ0 x ! Qi xji 2 Ng be theories
described in the vocabulary V fQi ji 2 N0 g. The two theories are not logically
equivalent and thus do not have the same class of models. However, if structures
were defined without any order of their extensions, then the two theories would have
the same class of models: Every model of K is trivially a model of H. Conversely,
any model of H in which the extension of every predicate contains the one element
of the domain is a model of K. If in a model of H some extension is empty, make it

8
It is clear from their use of as an index set that Bell and Slomson (1974) intend fRn jn 2 xg to be an
indexed set, in my notation fRn gn2x (rather than the set of the indexed sets elements). For if R2 = R3,
then fRi gi2f1;2;3g 6 fRi gi2f1;2g ; while fRi ji 2 f1; 2; 3gg fRi ji 2 f1; 2gg; Bell and Slomson thus need
indexed sets to allow for different names with identical extension.

123
Whats Right with a Syntactic Approach?

the extension of Q0 (which is possible because the extensions are not ordered), so
that the model is a model of K as well. Thus every model of H is a model of K and
vice versa. Therefore, if the extensions in a structure were not ordered, semantic
approaches in principle could not distinguish all those theories that can be
distinguished syntactically.
Hence for the definition of a variant of the notion of a structure that specifies no
vocabulary, something like a tuple is needed. But since structures can have infinite
vocabularies and tuples are finite, such a variant cannot be defined in terms of tuples if
it is to be able to express anything that can be expressed in terms of structures. In line
with the possibility of assigning any finite well-ordered index set to tuples, I therefore
suggest defining pure structures so that they contain mappings from arbitrary and not
fixed well-ordered index sets to extensions. That the index sets are not fixed can be
captured by identifying any two mappings that differ only in their index sets:
Definition 1 A representative of a pure structure A ^ is a triple hA; a; i; where
a : I ! I V is a mapping from the index set I to the image of an interpretation I ,
and  is a well-ordering of I. Two triples hA; a; i and hA; a0 ; 0 i represent the same
pure structure, hA; a; i hA; a0 ; 0 i, if and only if there is an order isomorphism
f : I ! I 0 and a0  f a.
The use of an interpretation I is just a convenient, non-essential way of ensuring
that a maps only to set theoretical objects that can be extensions of predicate,
function, or constant symbols. The definition determines a tuple if and only if the
index sets of the representatives are finite.
Definition 1 does not introduce a specific vocabulary, and is as language-
independent as the use of tuples. To stress the point: The infinity of representatives
with different vocabularies in definition 1 is not artificially introduced. Both indexed
sets and tuples assume some specific set or multiple sets of entities that provide a
vocabulary. Index sets can be used as vocabularies immediately, since they are
already mappings from some set to a set of extensions. Tuples allow the
introduction of such a mapping without any further assumptions except about
the elements of the index set. In definition 1, this further assumption is avoided by
the identification of representatives with different index sets.
The indexed structures given by Bell and Slomson (1974) relate to structures in a
very simple way: Their index set I plays the role of the set of relation symbols
fPi ji 2 Ig of structures.9 Definition 1 requires a somewhat more elaborate
modification of standard definitions, of which I will only give the modification
for the notion of embedding. Since pure structures are given by classes of mappings
with well-ordered index sets, embeddings between pure structures can be defined
via a bijection between the index sets of their representatives and a function from
one pure structures domain to the others. In this definition (which I give only for
pure first order structures), the position in the ordering of the index sets plays the
role of the element of the vocabulary.

9
For examples relevant in the following, see the definitions of reduct, isomorphism, and
substructure by Bell and Slomson (1974, 153, 73) and by Chang and Keisler (1990, 20-23),
respectively. The generalization to structures with function and constant symbols is straightforward.

123
S. Lutz

Definition 2 A pure rst order structure A ^ can be embedded in a pure first order
^
structure B if and only if for any two representatives A ^ hA; a; i with
^
a : I ! I V and  an ordering on I, and B hB; b;  i with b : J ! J V 0
0

and 0 an ordering on J, there is an order isomorphism g : I ! J and an injective


mapping h : A ! B (see Fig. 1) such that
1. for all c 2 I mapped to constants by a, h(a(c)) = b(g(c)),
2. for all F 2 I mapped to n-ary functions by a and all x1 ; . . .; xn
2 A; haFx1 ; . . .; xn bgFhx1 ; . . .; hxn , and
3. for all P 2 I mapped to n-ary relations by a and all x1 ; . . .; xn 2 A;
hx1 ; . . .; xn i 2 aP , hhx1 ; . . .; hxn i 2 bgP.
h is called an embedding of A ^ in B. ^ If h is surjective, A ^ are called
^ and B
isomorphic.
Embeddability of pure structures only has to be shown for one representative of
each pure structure:
Lemma 1 A pure rst order structure A ^ can be embedded in a pure rst order
^
structure B if and only if there are two representatives A ^ hA; a; i with
^
a : I ! I V and  an ordering on I, and B hB; b;  i with b : J ! J V 0
0

and 0 an ordering on J, an order isomorphism g : I ! J, and an injective


mapping h : A ! B such that conditions 13 of denition 2 are fullled.
Proof The proof from left to right is immediate. For the other direction, assume
^ with d : L ! J V 0 .
^ with c : K ! I V and hB; d; 0  i B
hA; c;  i A
By definition 1, there are order isomorphisms i : K ! I and j : J ! L (see Fig. 2).
Then there is an order isomorphism k : K ! L and a one-to-one mapping
h : A ! B such that
1. for all m 2 K mapped to constants by c; hcm h  a  im
b  g  im b  j1  j  g  im dj  g im dkm,
2. for all F 2 K mapped to n-ary functions by c and all x1 ; . . .; xn 2 A;
hcFx1 ; . . .; xn ha  iFx1 ; . . .; xn b  g  iFhx1 ; . . .; hxn
b  j1  j  g  iFhx1 ; . . .; hxn dkFhx1 ; . . .; hxn , and
3. for all P 2 I mapped to n-ary relations by c and all x1 ; . . .; xn 2 A;
hx1 ; . . .; xn i 2 cP , hx1 ; . . .; xn i 2 a  iP , hhx1 ; . . .; hxn i 2 b  g  iP
, hhx1 ; . . .; hxn i 2 b  j1  j  g  iP , hhx1 ; . . .; hxn i 2 dkP: h
The definition of embedding for pure structures respects the standard definition of
embedding for structures (Hodges 1993, 5) in that there is a connection between
structures and pure structures, and under this connection, the two definitions of

Fig. 1 Functions between


arbitrary representatives of
two structures A ^ hA; a; i;
B^ hB; b; 0 i so that A
^ can
be embedded in B ^

123
Whats Right with a Syntactic Approach?

^
^ hA; a; i hA; c;  i; B
Fig. 2 Functions between four representatives of two structures A
0
hB; b; 0 i hB; d;   i so that A ^
^ can be embedded in B

embedding are interchangeable. More precisely, any structure A gives rise to a


^ by a well-ordering of its vocabulary, and any pure
representative of a pure structure A
^
structure A contains a structure A among its representatives; embeddability of a pure
structure is equivalent to the embeddability of a structure which it contains:
Claim 2 A pure rst order structure A ^ can be embedded in a pure rst order
^
structure B if and only if there are rst order structures A hA; I i and B hB; J i
and a well-ordering  such that A ^ hB; J ; i, and A can be
^ hA; I ; i; B
embedded in B.
^
Proof For the proof from left to right, choose any representatives hA; a; i of A
0 ^
and hB; b;  i of B. Now choose for V the index set of a; I a, and J b  g,
where g is the order isomorphism from I to J. The claim follows from definition 2
because for any Q 2 V; hQA h  aQ b  gQ J Q QB in the domain
^ which is also the domain of A:
of A,
For the proof from right to left, choose g id. The claim follows with lemma 1
because for any Q 2 V; hI Q hQA QB J gQ. h
To put it slightly differently: Claim 2 and lemma 1 entail
Corollary 3 A pure rst order structure A ^ if and only if for
^ can be embedded in B
any structures A hA; I i and B hB; J i and any well-ordering  such that
^ hB; J ; i; A can be embedded in B.
^ hA; I ; i and B
A

Conversely, a structure A can be embedded in a structure B if and only if, under some
^ respectively,
^ and B,
well-ordering of their vocabulary, they represent pure structures A
^ can be embedded in B.
such that A ^ The proof assumes the axiom of choice.

123
S. Lutz

Corollary 4 First order structure A hA; I i can be embedded in B hB; J i if


and only if hA; I ; i can be embedded in hB; J ; i for some well-ordering .
Proof Note that I and J must have the same vocabulary. The proof from right to
left follows immediately from claim 2. For the proof from left to right, note that any
vocabulary V can be well-ordered (assuming the axiom of choice). The conclusion
follows from corollary 3. h
To put this again slightly differently: Corollary 4 and lemma 1 entail
Corollary 5 First order structure A hA; I i can be embedded in structure
B hB; J i if and only if hA; I ; i can be embedded in hB; J ; i for all well-
orderings .
These results show that typical model theoretic operations between language-
independent pure structures are well-defined and that nothing is lost by discussing
embeddings only for structures or only for pure structures. Language independence
can be achieved at any point by introducing an ordering for the vocabulary, turning
the structures into representatives of pure structures. Conversely, the ordering can
be eliminated at any point in the discussion by choosing, for all pure structures at
once, one of the possible vocabularies. It is clear that other model theoretic notions
besides the concept of embedding have similar analogues for pure structures, so that
model theorys restriction of each structure to a specific vocabulary is dispensable.
The independence from a specific vocabulary that French and Ladyman (1999) call
much celebrated seems indeed important, since language is inherently conventional,
or at least relative to a group of speakers. As Suppe (1974, 204205) puts it,
translating a theory from English to French changes the sentences in which the theory
is presented, but not the theory. The ideal gas law, for instance, does not change only
because one uses the names pression, volume, constante de Boltzmann, and
temperatur instead of pressure, volume, Boltzmann constant, and tempera-
ture. Semantic approaches can accommodate this kind of language independence
because the simultaneous switch of all structures to pure structures allows the choice
between an infinity of vocabularies, and thus the freedom from any specific one.
I now want to show that this kind of language independence can also be
accommodated in syntactic approaches. Assume that, as in the example of the ideal gas
law, a set H of sentences in V can be transformed into a set H0 in V 0 by renaming non-
logical constants. Assume further that two models A  H and A0  H0 represent, under
^ Then the models are reducts
some order of their vocabularies, the same pure structure A.
of a structure B in V [ V 0 that is a model of both H and H0 . But B is also a model of
identity claims between any two elements of V and V 0 that have the same position in the
pure structure. In the vocabulary V [ V 0 , each symbol of V can thus be used as the
definiens of the symbol of the same position in V 0 and vice versa. These identifications
therefore allow the definitional extension of either set of sentences to the other. Since
explicit definitions are non-creative, neither set of sentences gains content by the
addition of the definitions, and thus it is possible to go from, say, H to H0 by first
extending H through the definitions and then eliminating the V-symbols from the
resulting theory. The use of definitional extensions and reducts thus leads to an
equivalence class of sets of sentences that are taken to formalize the same theory.

123
Whats Right with a Syntactic Approach?

However, semantic approaches are too liberal in their neutrality with respect to
the vocabulary, for not every renaming makes sense. In the ideal gas law, for
example, the translation should allow the renaming of temperature into
temperatur, but not into pression. For syntactic approaches, this restriction on
renaming can be expressed by introducing analytic sentences. Any renaming not
entailed by analytic sentences is not allowed. Then, temperature can be renamed
as temperatur, but not as pression.10 To avoid such incorrect renaming in
semantic approaches, the connection between pure structures and the world has to
be explicitly taken into account (see Sect. 4).
There is another problem with arbitrary renaming, for it is not only too liberal in
some respects, but also too restrictive in others. It does not, for example, avoid the
language dependence referred to by Suppe (1974, 204205), who notes that wave
and matrix mechanics, although the same theory, constitute different sets of
sentences.11 Obviously, the difference between matrix mechanics and wave
mechanics goes beyond a mere renaming of the predicate, function, and object
symbols. And this is a problem for semantic approaches, because if not only the
names of the extensions change, the pure structures change.
To accommodate the possibility of restricted renaming, I have used the simplest
kind of definition, an identification of symbols. But the comparison of two theories
can be generalized to include any kind of explicit definitions, and this generalization
can accommodate the restricted change of structure. Two theories H; H0 are
denitionally equivalent if and only if both can be extended by explicit definitions
such that they become equivalent. Then their models can be turned into each other
by a procedure analogous to the one described above: First expand the model of one
theory to include the defined symbols of the definitions. This expansion is unique.
Then reduce the resulting structure to the vocabulary of the other theory (cf. Hodges
1993, 61). This allows the identification of theories that differ not only in the
vocabulary they use, but also in the pure structures that their models represent. The
procedure therefore also allows for structural differences in the description of
theories, and thus can be used to extend not only syntactic approaches, but also
semantic ones. It leads to an equivalence class of classes of structures that are taken
to formalize the same theory. This move would also not obviously tether semantic
approaches further to syntactic ones because the notion of definitional equivalence
can be defined without reference to sets of sentences (de Bouvere 1965).

4 Sentences, Structures, and the World

A set R of sentences of predicate logic is not enough for applying a theory to the
world, for if A  R, any set of the same cardinality as jAj can be made into a model
of R as well:

10
Incidentally, if one wanted to identify theories with sets of propositions, the identification of
expressions for the same proposition would provide analytic sentences.
11
Halvorson (2012, 4.2) gives much more precise examples.

123
S. Lutz

Lemma 6 Let jAj A and g : A ! B be any bijection. Then there is a structure B


with jBj B such that g is an isomorphism from A to B.
Proof For each relation PA B A
i in A, define Pi : gPi , for each function fj and
A

each tuple hb1 ; . . .; bnj i 2 Bnj (or arguments for functions of higher order), define
fjB b1 ; . . .; bnj : gfjA g1 b1 ; . . .; g1 bnj , and for each constant cA
k , define
cBk : gc A
k . It is straightforward to show that B A: h
The bijection g may in particular be a permutation on a domain, so that any
element of A can be exchanged for any other element.
Claim 7 Let A  R and cardjAj = card(B). Then there is a B with jBj B and
B  R.
Proof (Przecki 1969, 2728) If cardjAj = card(B), then there exists a bijection
g : jAj ! B. The corollary follows from lemma 6, because sets of sentences can
determine structures at most up to isomorphism. h
In light of claim 7, it is clear that a set of sentences can be connected to the world
(beyond statements about the number of its objects) only if there is a means of
distinguishing between isomorphic structures. One way of solving this problem is
by introducing a class of possible structures M, which is determined by the
extensions of the V-terms given their meaning.12 M allows for R making more than
cardinality claims about the world by defining the possible models of R as the
possible structures that are models of R. The use of possible structures does not have
to trivialize the distinction between semantic and syntactic approaches: One can
postulate that M only distinguish between isomorphic structures, so that the
isomorphic closure of M is always the class of all V-structures. Then it is easy to
show that the isomorphic closure of the possible models of R is always the class of
all models of Rin other words, in syntactic approaches, non-isomorphic structures
have to be distinguished by sentences of the object language. This suggests that, for
the purposes of this discussion, syntactic approaches are best conceptualized as
follows:
Definition 3 Syntactic approaches describe theories and their relations to the
world using sentences in an object language and structures. Non-isomorphic
structures are only distinguished by sentences in the object language.

In the Received View, an interpretation maps the basic (observational)


vocabulary to extensions that contain objects of the world (Carnap 1939, 6). This
gives a general way of connecting syntactic descriptions to the world: The
interpretations given by their possible models map at least one symbol to an
extension that contains an object of the worlda worldly extension, as I will call it
from now on. Interpretations with a worldly extension are accordingly worldly
interpretations, and the analogous holds for models and structures.

12
Note that the domains of the structures in M do not generally contain only actual objects (Lutz 2012a,
188).

123
Whats Right with a Syntactic Approach?

If there are worldly structures, syntactic descriptions can be given a formal


connection to the world by relying on worldly possible structures. These can be
determined by actually showing (pointing to) members of worldly sets and relying
on the psychological fact of intersubjective agreements about similarity between
experiences (Przeecki 1969, 3538). More commonly, the worldly sets are
described in a metalanguage (thereby leaving out the question of how the terms of
the metalanguage are connected to the world). Either solution presumes that worldly
extensions (i. e., sets containing objects of the world) are possible. Since an
interpretation maps to such extensions, it is also presumed that it is possible to have
a function from a vocabulary to worldly extensions. I will discuss these
presumptions below.
The connection to the world has sometimes been claimed to be a strong point of
semantic approaches, because they escape the problem of connecting linguistic
entities to the world (cf. Chakravartty 2001, 327). However, the connection between
pure structures and the world is not completely straightforward either, essentially
because of a point made by Newman (1928, 140, cf. Ketland 2004, n. 10), which can
be adapted to pure structures:
Claim 8 Let jAj ^ A and card(A) = card(B). Then there is a B ^ with jBj
^ B and
^
B A. ^

Proof (cf. Przecki 1969, 2627) Let hA; I ; i be a representative of A. ^ Since


card(A) = card(B), there is a bijection from A to B, and thus, by lemma 6, a structure
B hB; J i with B A hA; I i. It follows from corollary 4 that
^ hB; J ; i:
A h
Thus just as syntactic approaches must have a means of distinguishing between
isomorphic structures, semantic approaches must have a means of distinguishing
between isomorphic pure structures. This suggests that, for the purposes of this
discussion, semantic approaches are best conceptualized as follows:
Definition 4 Semantic approaches describe theories and their relations to the
world using classes of pure structures not necessarily closed under isomorphism.

How the connection between pure structures and the world is to be envisaged in
semantic approaches depends on whether the pure structures are meant to be
worldly or non-worldly. Suppes statement quoted at the beginning of Sect. 3 (that
semantic approaches construe theories as the formal referents of the theories
formulations) suggests that the pure structures used in semantic approaches contain
worldly extensions, that is, are worldly themselves. Da Costa and French (2000, fn.
2) also seem to assume that the pure structures are worldly when they state that the
set-theoretic models are constructed in set theories with Urelemente (individual[s],
systems, portions of the universe, real things,).
Representatives of a pure structure are structures, and accordingly the existence
of a worldly pure structure entails the existence of a worldly structure. This structure
then provides the connection of the linguistic entities with the world. Therefore, if a

123
S. Lutz

semantic approach successfully connects theories to the world by using pure


worldly structures, syntactic descriptions can be connected to the world as well.13
Mostly, however, the pure structures discussed in semantic approaches are not
supposed to be worldly (see, e. g., French and Ladyman 1999). They are simply
abstract set theoretic entitiesno dogs, observations, or electrons are members of
the sets. Theses non-worldly pure structures then have to be connected to the world
somehow. I will discuss four ways of doing so.
In the first way, the theory described by pure structures is supposed to be
isomorphic to some worldly pure structure.14 But because of claim 8, if the theory is
about more than just the cardinality of its domain, there has to be one distinguished
isomorphism (with an order isomorphism g between the index sets) or a set of
distinguished isomorphisms that connect the theorys non-worldly pure structures
with those worldly pure structures that the non-worldly pure structures are supposed
to represent. This results again in worldly pure structures, since for each g and
worldly pure structure hA; a; i; hA; a  g; g1 i is again a worldly pure
structure, and thus the previous discussion applies.
As another way of connecting a non-worldly pure structure to the world, Muller
(2010, 6) suggests taking a structure (which he takes to be a tuple), and allowing it
to be assigned any appropriate vocabulary. The vocabulary can hence be chosen
according to expedience. The connection to the world is then given through an
interpretation of the vocabulary. It is clear that this way of connecting semantic
descriptions to the world presumes that it is possible to connect syntactic
descriptions to the world. Note also that Muller is basically describing a non-
worldly pure structure, since the index set of the structure can be assigned any
vocabulary, that is, any other index set.15 Since this allows arbitrary renaming, it is
doubtful that Muller in fact provides a solution to the problem.
A third way of connecting non-worldly pure structures to the world is by stating
that the pure structure and the system it is meant to represent are similar. French and
Ladyman (1999) argue successfully that the similarity relation, left largely
unexplicated, is too vague to be of much use. Thus even if, within the boundaries
of the terms vagueness, theories can be connected to the world by similarity, the
success of this approach has been achieved only by significantly lowering the
precision of the analysis. Assuming that both syntactic and semantic approaches aim
for more precision, this solution therefore at best achieves a different goal.
French and Ladyman (1999, 115, cf. 113) argue instead that the connection to the
world is simply not a problem to be solved by semantic approaches:
The theoretical models [the pure structures of the theory] are held to relate to
models of the phenomena and these are just other structures. That these

13
For indexed structures, the modification of this argument is straightforward: Each structures index set
I can be used directly as a vocabulary, providing the interpretation a : I ! aI. If a specific vocabulary
V is desirable, any bijection g : V ! I leads to the interpretation I a  g : V ! aI. The discussions
below can modified analogously.
14
In semantic approaches, isomorphisms are often used as analogues to Tarskis definition of truth.
15
Muller presents his formalism as an extension of his notion of structures, in contradiction to the
conclusion reached in connection with definition 1.

123
Whats Right with a Syntactic Approach?

represent real events and processes cannot be determined by the content of the
theory, but is a pragmatic fact about our language [] and it is unreasonable
to demand that the semantic view explains the nature of representation in
general.
This clearly falls short of the idea that semantic approaches are easier to connect to
the world than syntactic approaches.
In French and Ladymans approach, structures are simply set theoretic constructs,
and thus in this respect their view is identical to that of the structuralists following
Sneed (1971) and Stegmuller (1979), whose formal core commitment is to the
expression of scientific theories in terms of set theory. In such a structuralist
approach, there is simply one set of set theoretic sentences (those describing the
phenomena) that are distinguished as representing real events. But if neither
predicate logic nor set theory are given a formal connection to the world, they are
simply languages that are assumed to describe the world in some not further
specified way. Then, for example, the question of non-standard models does not
even arise (neither for first order nor higher order logic, nor set theory) unless an
additional metalanguage is artificially introduced (Vaananen 2001). And if there is a
metalanguage, set theory and higher order logic are equally expressive (Vaananen
2001, 506507). The discussion in Sect. 2 about the possibilities to capture
structures up to isomorphism therefore stacks the deck against syntactic approaches:
It rests on the heroic assumption that set theory is not a language that describes the
world, but is the world (more precisely: the world is a worldly structure), and the
task of syntactic approaches is to describe the set theoretic world. If one instead
treats, with French and Ladyman, set theory as just another language (which it is), it
cannot describe the world more precisely than higher order logic.16
If the relation to the world is excluded from the formalisms of syntactic and semantic
approaches, then the difference between the two reduces to the question which syntax
(and which axioms) to use in formalizations. Semantic approaches assume the axioms of
(possibly nave) set theory and often assume that all further (mathematical) concepts
have been introduced by explicit definitions. Hilbert (1900, 10921093) calls this the
genetic method of formalization. Syntactic approaches instead rely on predicate
logic. In principle, syntactic approaches could then follow the genetic method and first
formalize set theory, then introduce further concepts by explicit definitions. But
syntactic approaches can also follow Hilberts axiomatic method and only provide the
axioms for those terms that actually appear in the theory that is to be formalized, without
going through set theory first. Therefore, semantic approaches are a subclass of syntactic
ones if the relation to the world is excluded from the formalisms.
In conclusion, the connection to the world is equally problematic for syntactic
and semantic approaches. And a solution for one kind of approach also provides a
solution for the other.17

16
Vaananen (2001) further argues that neither set theory nor higher order logic can describe the world
more precisely than first order logic.
17
I have argued that solutions for semantic approaches are also solutions for syntactic approaches. The
argument for opposite direction can rely on the fact that every structure represents a pure structure and, if
needed, corollary 5.

123
S. Lutz

5 Corollaries

The remaining criticisms of syntactic approaches, that they are too cumbersome,
cannot capture scientific models, and cannot capture the relation between theory and
observations, had been developed for the Received View and were then leveled
against all syntactic approaches without further justification. The following
discussion can hence be kept short.
That formalizations in the Received View are too cumbersome has been inferred
from the (incorrect) assumptions that these formalizations have to be given in first
order logic and have to explicitly contain all axioms for all terms that occur in a
theory (Lutz 2012c, 2, 3). Syntactic approaches in general, however, can clearly
avail themselves to higher order logic, and second order logic is already enough to
capture directly most all mathematical practice (Leivant 1994, 260; cf. Vaananen
2001, 515).18 Since capturing mathematical practice only becomes easier for higher
orders, reliance on predicate logic does not lead to cumbersome formalizations. The
possible counterargument that higher order logics do not have a complete proof
theory has been addressed in Sect. 2. Syntactic approaches can clearly also leave
some axioms implicit, to be spelled out in detail when needed. Conversely, the
axioms of set theory and model theory can also be explicitly spelled out, and so can
the additional definitions and axioms for further mathematical concepts (cf. Lutz
2012c, 2).
The claim that the Received View is less hospitable to scientific models than
semantic approaches stems from a misunderstanding of some statements by
proponents of the Received View (Lutz 2012c, 4) and, it seems, the identification
of models in the model theoretic sense with scientific models. The former point is
irrelevant for syntactic approaches in general; the latter point is wrong because the
different meanings and uses of the term model in the natural sciences clearly do
not fit the definition of model in model theory.19 Without this identification, the
question becomes whether semantic approaches can describe the world (specifically,
the scientific models in the world) better than syntactic approaches. I have answered
this question in the negative in Sects. 2 and 4.
After an analysis of the Received Views reliance on correspondence rules and a
bipartitioned vocabulary, Suppe (1974, 114) concludes that if formalization is
desirable in a philosophical analysis of theories, it must be of a semantic sort. But
syntactic approaches do not need to assume a bipartition of the vocabulary, and
hence also do not need to assume a connection of theory and observations by
correspondence rules. And not only is Suppes argument invalid, Suppe himself
shows that his conclusion is also wrong: His criticism of correspondence rules
(Suppe 1974, II.E) is based on a syntactic description of a relation between theory
and observations that he takes to be more adequate (cf. Suppe 1974, 108, n. 225).
This trivially establishes the possibility of capturing the relation syntactically.

18
Andrews (2002, xii-xiii) even argues that higher order logic provides a more natural formalization of
mathematical reasoning than set theory.
19
The influential argument to the contrary by Suppes (1960) establishes at best that the formal part of
scientific models can, with reformulations, be expressed in set theory (cf. Lutz 2012c, 9799).

123
Whats Right with a Syntactic Approach?

Conversely, semantic approaches can avail themselves to a bipartition of the


vocabulary, as a criterion of meaning by Suppes demonstrates (cf. Lutz 2012b,
6.5).

6 On Peaceful Coexistence

The most important result of the preceding discussion is the close connection between
syntactic and semantic approaches. It is easy to go from pure structures to indexed
structures, to structures, to sentences, and back again. Accordingly, it is easy to move
from a semantic approach to a syntactic one and back. On the one hand, that means that
the problems of syntactic approaches reappear in semantic approaches, as the
problems of connecting the respective formalisms to the world show. On the other
hand, it means that the solutions found in syntactic approaches can be used in semantic
approaches as well, for example the use of explicit definitions for the identification of
sets of sentences and classes of structures.
I hope that in the future more solutions from one kind of approach will find their
ways to the other. For on a meta-level, I do hold the position of the critics of
syntactic approaches: The language in which an analysis is phrased, whether it uses
pure or indexed structures, structures, or possible structures and an object language,
often matters very little.

Acknowledgments A very early version of this section has been presented at Herman Philipses Dutch
Research Seminar in Analytic Philosophy at Utrecht University. Parts have been presented at the EPSA 09
at the Vrije Universiteit, Amsterdam, on October 23, 2009 and at the workshop Perspectives on
Structuralism at the Center for Advanced Studies/ Munich Center for Mathematical Philosophy, Ludwig-
Maximilians-Universitat Munchen, on February 17, 2012. I thank the participants for helpful discussions.
This article has also profited a lot from a reading group at Tilburg University with Reinhard Muskens and
Stefan Wintein. It is based on 4.1 of my dissertation prepared at Utrecht University (Lutz 2012b), which
contains elaborations of some of the claims of this article and additional references. I thank my advisers
Janneke van Lith, Albert Visser, and especially Thomas Muller for extensive and helpful comments and
discussions, and F. A. Muller for his trenchant comments on my dissertation in general and the themes of
this article in particular. I thank Arno Bastenhof for helpful discussions, an anonymous reviewer for
helpful comments, and Alana Yu for helpful suggestions in matters of style. Parts of the research for this
article have been supported by the Alexander von Humboldt Foundation.

References

Andrews, P.B. (2002). An introduction to mathematical logic and type theory: To truth through proof (2nd
ed.). Berlin: Springer.
Bell, J. L., & Slomson, A. B. (1974). Models and ultraproducts: An introduction (3rd ed.). Amsterdam:
North-Holland.
Carnap, R. (1956). The methodological character of theoretical concepts. In H. Feigl & M. Scriven (Eds.),
The foundations of science and the concepts of psychology and psychoanalysis. Volume 1 of
Minnesota studies in the philosophy of science. Minneapolis, MN: University of Minnesota Press.
Carnap, R. (1939). Foundations of logic and mathematics. Volume I, Issue 3 of Foundations of the Unity
of Science. Toward an International Encyclopedia of Unified Science. Chicago, London: University
of Chicago Press. References are to the two-volume edition.
Chakravartty, A. (2001). The semantic or model-theoretic view of theories and scientific realism.
Synthese, 127, 325345.

123
S. Lutz

Chang, C. C., & Keisler, H. J. (1990). Model theory. Volume 73 of Studies in logic and the foundations of
mathematics (3rd ed.). Amsterdam: North Holland. 3rd impression 1992.
da Costa, N., & French, S. (1990). The model-theoretic approach in the philosophy of science. Philosophy
of Science, 57, 248265.
da Costa, N., & French, S. (2000). Models, theories, and structures: Thirty years on. Philosophy of
Science, 67(Proceedings), S116S127.
de Bouvere, K. (1965). Synonymous theories. In: J. W. Addison, L. Henkin, & A. Tarski (eds.), The
theory of models. Proceedings of the 1963 international symposium at Berkeley, (pp. 402406)
Amsterdam: North-Holland Publishing.
Enderton, H. B. (2009). Second-order and higher-order logic. In: E. N. Zalta (Ed.), The Stanford
encyclopedia of philosophy. Center for the study of language and information. Stanford University,
Spring 2009 edition.
van Fraassen, B.C. (1980). The scientic image. Oxford: Clarendon Press.
van Fraassen, B. C. (1989). Laws and symmetry. Oxford: Clarendon Press.
van Fraassen, B.C. (2008). Scientic representation: Paradoxes of perspective. Oxford: Clarendon Press.
French, S., & Ladyman, J. (1999). Reinflating the semantic approach. International Studies in the
Philosophy of Science, 13(2), 103121.
Halvorson, H. (2012). What scientific theories could not be. Philosophy of Science, 79(2), 183206.
Hilbert, D. (1900). Uber den Zahlenbegriff. In G. Hauck, & A. Gutzmer (Eds.), Jahresbericht der
deutschen Mathematiker-Vereinigung (Vol. 8, pp. 180183). Leipzig. B. G. Teubner. References are
to the translation (Hilbert, 1996).
Hilbert, D. (1996). On the concept of number. In: W. Ewald (Ed.), From Kant to Hilbert: A source book in
the foundations of mathematics, volume II. (pp. 10921095). Clarendon Press, Oxford. Digitally
reprinted in 2005.
Hodges, W. (1993). Model theory. Cambridge: Cambridge University Press. Digitally printed in 2008.
Ketland, J. (2004). Empirical adequacy and ramsification. The British Journal for the Philosophy of
Science,, 55, 287300.
Kitcher, P. & Salmon, W., (eds) (1989). Scientic explanation. Volume 13 of Minnesota studies in the
philosophy of science. Minneapolis, MN: University of Minnesota Press.
Leivant, D. (1994). Higher order logic. In D. M. Gaby, C. Hogger & J. Robinson (Eds.), Deduction
methodologies. Volume 2 of Handbook of logic in artificial intelligence and logic programming (pp.
229321). Oxford: Oxford University Press.
Lutz, S. (2012a). Artificial language philosophy of science. European Journal for Philosophy of Science, 2(2),
181203.
Lutz, S. (2012b). Criteria of empirical significance: Foundations, relations, applications. PhD thesis,
Utrecht University. http://philsci-archive.pitt.edu/id/eprint/9117.
Lutz, S. (2012c). On a straw man in the philosophy of science: A defense of the Received View. HOPOS:
The Journal of the International Society for the History of Philosophy of Science, 2(1), 77120.
Muller, F. A. (2010). Reflections on the revolution at Stanford. Synthese, 183(1), 87114.
Newman, M. H. A. (1928). Mr. Russells causal theory of perception. Mind, 37(146), 137148.
Przeecki, M. (1969). The logic of empirical theories. London, New York: Routledge & Kegan Paul/
Humanities Press.
Sneed, J. D. (1971). The logical structure of mathematical physics. Dordrecht, The Netherlands: D. Reidel
Publishing Co..
Stegmuller, W. (1979). The structuralist view of theories. New York: Springer.
Suppe, F. (1974). The search for philosophic understanding of scientific theories. In F. Suppe (Ed.), The
structure of scientic theories (pp. 3241). Urbana, IL: University of Illinois Press.
Suppe, F. (1989). The semantic conception of theories and scientic realism. Urbana, IL: University of
Illinois Press.
Suppe, F. (2000). Understanding scientific theories: An assessment of developments, 19691998.
Philosophy of Science, 67, S102S115.
Suppes, P. (1960). A comparison of the meaning and uses of models in mathematics and the empirical
sciences. Synthese, 12, 287301.
Suppes, P. (2002). Representation and invariance of scientic structures. Stanford, CA: CSLI
Publications.
Vaananen, J. (2001). Second-order logic and foundations of mathematics. The Bulletin of Symbolic Logic,
7(4), 504520.

123

You might also like