0% found this document useful (0 votes)
52 views38 pages

Defining Determinism

This document discusses three approaches to defining determinism for scientific theories: DEQN, DMAP, and DBRN. DEQN assesses theories based on their defining differential equations. DMAP assesses theories based on mappings between their possible temporal developments. DBRN assesses theories based on branching models of their possible developments. The authors argue that DBRN provides the most useful general definition, as it combines formal clarity, connection to the philosophical notion of determinism, and relevance for assessing theories in practice. They support this claim by comparing applications of the three approaches to theories like Newtonian mechanics, quantum mechanics, and general relativity, and through a formal comparison of DMAP and DBRN.

Uploaded by

paulreg
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
52 views38 pages

Defining Determinism

This document discusses three approaches to defining determinism for scientific theories: DEQN, DMAP, and DBRN. DEQN assesses theories based on their defining differential equations. DMAP assesses theories based on mappings between their possible temporal developments. DBRN assesses theories based on branching models of their possible developments. The authors argue that DBRN provides the most useful general definition, as it combines formal clarity, connection to the philosophical notion of determinism, and relevance for assessing theories in practice. They support this claim by comparing applications of the three approaches to theories like Newtonian mechanics, quantum mechanics, and general relativity, and through a formal comparison of DMAP and DBRN.

Uploaded by

paulreg
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 38

Brit. J. Phil. Sci.

69 (2018), 215–252

Defining Determinism
Thomas Müller and Tomasz Placek

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


ABSTRACT
The article puts forward a branching-style framework for the analysis of determinism and
indeterminism of scientific theories, starting from the core idea that an indeterministic
system is one whose present allows for more than one alternative possible future. We
describe how a definition of determinism stated in terms of branching models supple-
ments and improves current treatments of determinism of theories of physics. In these
treatments, we identify three main approaches: one based on the study of (differential)
equations, one based on mappings between temporal realizations, and one based on
branching models. We first give an overview of these approaches and show that current
orthodoxy advocates a combination of the mapping- and the equations-based
approaches. After giving a detailed formal explication of a branching-based definition
of determinism, we consider three concrete applications and end with a formal
comparison of the branching- and the mapping-based approach. We conclude that the
branching-based definition of determinism most usefully combines formal clarity, con-
nection with an underlying philosophical notion of determinism, and relevance for the
practical assessment of theories.

1 Introduction
2 Determinism in Philosophy of Science: Three Approaches
2.1 Determinism: The core idea and how to spell it out
2.2 The three approaches in more detail
2.3 Representing indeterminism
3 Orthodoxy: DMAP, with Invocations of DEQN
4 Branching-Style Determinism (DBRN)
4.1 Models and realizations
4.2 Faithfulness
4.3 Two types of branching topologies
5 Comparing the Approaches
5.1 Case studies
5.2 Formal comparison of the DMAP and DBRN frameworks
6 Conclusions

ß The Author 2016. Published by Oxford University Press on behalf of British Society for the Philosophy of Science.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License
(http://creativecommons.org/licenses/by-nc/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium,
doi:10.1093/bjps/axv049 provided the original work is properly cited. For commercial re-use, please contact journals.permissions@oup.com
Advance Access published on May 5, 2016
216 Thomas Müller and Tomasz Placek

1 Introduction
In this article we describe how a definition of determinism based on branch-
ing models supplements and improves current treatments of determinism of
scientific theories in physics. Our focus is on a definition of determinism that
takes a scientific theory as input, and delivers a verdict as to the theory’s
determinism as output, providing one bit of information. This may seem to
be a simple matter, but in practice a number of subtle issues are involved: In

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


which form is a theory fed into the definition? When does a (semi-)formal
definition have a claim to providing an explication of the philosophical
notion of determinism, rather than something else? And what is the use of
the definition: Does failure of determinism signal a defect in the theory (this
may be a practitioner’s sentiment), or rather a useful insight into the theory
or perhaps even the metaphysical issues (which may be a philosopher’s
view)? It turns out that in the actual assessment of a physical theory as
deterministic or indeterministic, all these matters play an important role so
that deciding about a theory’s determinism is a delicate practice rather than a
simple application of a definition. Still, a general definition provides a useful
overarching perspective on the determinism of scientific theories, functioning
both as a guideline for practical assessment and as an interface to discussions
in other areas of philosophy. Our contribution is aimed at this general level.
In current philosophy of science, there are three subtly different approaches
to defining determinism for physical theories, which we label DEQN, DMAP,
and DBRN. Figure 1 gives a schematic overview. According to DEQN, a
theory is assessed via a study of the behaviour of its defining (differential)
equations. According to DMAP, a theory is assessed in terms of mappings
between the linear temporal developments of systems allowed for by the
theory. Finally, according to DBRN, a theory is assessed in terms of partially
ordered, branching models of such systems’ behaviour. Our main claim is that
DBRN gives the most useful general definition of determinism for physical
theories, despite the fact that DMAP enjoys the status of current orthodoxy in
philosophy of science, and DEQN is typically invoked by practitioners. We
believe that this insight into the usefulness of a DBRN-type analysis of deter-
minism carries over to other areas of philosophy as well.
We argue for this main claim in the following way: In order to provide a
basis for discussion, we briefly describe the general background notion of
determinism and introduce the three approaches to determinism of scientific
theories in physics, DMAP, DEQN, and DBRN in Section 2. Section 3 offers
an overview of currently dominant definitions of determinism, thereby show-
ing in which way DMAP enjoys the status of orthodoxy, and which role
DEQN plays in the actual assessment of theories. Section 4 provides the
DBRN definition of determinism in formal detail, including questions of
Defining Determinism 217

THEORY:
Deterministic?

extract extract extract

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


Linear Partially
(Differential)
temporal ordered
equations
realizations models

Study and assess:


Study and assess:
existence of Assess: all
global
suitable models linear?
uniqueness?
mappings?

YES / YES / YES /


NO NO NO

DEQN DMAP DBRN


Figure 1. Three definitions of determinism for a theory: DEQN, based on equa-
tions; DMAP, based on mappings between linear temporal realizations; and
DBRN, based on branching models.

topology. In Section 5, which comprises the bulk of the article, we compare the
three approaches to defining determinism, with a view to making good our
claim of the usefulness of DBRN. We proceed in two steps: First (Section 5.1),
we give three examples of the application of the three approaches to physical
theories, referring to Newtonian mechanics, quantum mechanics, and general
relativity (GR). We show in which ways the DBRN definition comes natu-
rally, including the construction of an explicit mathematical model. A perhaps
surprising result from our case studies is that the DMAP definition, despite its
status as official orthodoxy, is hardly ever used, and that the DEQN definition
218 Thomas Müller and Tomasz Placek

is inappropriate when it comes to quantum mechanics. In a second step


(Section 5.2), we offer a formal comparison of the DMAP and the DBRN
definitions. We point out the subtle role that a class of isomorphisms plays for
the DMAP analysis, and we show that the DBRN and DMAP definitions can
give rise to different assessments in such a way that DMAP classifies a theory
as indeterministic too easily. Finally, Section 6 sums up the article and reca-
pitulates our conclusions.

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


2 Determinism in Philosophy of Science: Three Approaches
The question of whether our world is deterministic or not—whether the future
is genuinely open or whether there is just one real possibility for the future—is
one of the fundamental concerns of metaphysics. And the impact of that
question is not limited to theoretical metaphysical speculation. Determinism
is a topic that cuts across many philosophical sub-disciplines, including ethics,
action theory, and philosophy of science.
In philosophy of science, the question of determinism is addressed in rela-
tion to scientific theories and provides an important means of assessing the-
ories in various respects. There are many reasons to ask whether a given
scientific theory is deterministic or not. One is metaphysical: we may be con-
vinced that the theory gives an appropriate picture of what the world is like,
and therefore use the theory in order to find out about the determinism or
indeterminism of the world as a whole. Another is epistemological: finding out
whether a theory is deterministic can tell us something about in-principle
limitations on predictions (or retrodictions) that the theory affords. Asking
about determinism is often a good way to deepen one’s understanding of the
theory itself, since many subtle technical issues have to be addressed in order
to provide a verdict on whether the theory is deterministic or not.
Assessing whether a given theory is deterministic or not is a tricky issue.
Earman ([1986]) has done a great service to the philosophy of science com-
munity by tracing out in detail the questions involved for physical theories,
and the decisions that need to be made along the way. His work closely follows
the discussions of practitioners, demonstrating the role played by, among
other things, the behaviour of differential equations, the identification of
gauge degrees of freedom, the notion of a physical state, and considerations
of the physicality of specific set-ups (we will mention some of these issues later
on). Nevertheless, what is being done in assessing the determinism of a theory,
must in some important sense be the same in all cases, for otherwise, there
would be no unity of the notion of determinism of a theory to begin with. It is
thus reasonable to ask what it is that is the same in all these cases, that is, what
is the common definition of determinism being employed?
Defining Determinism 219

2.1 Determinism: The core idea and how to spell it out


The core idea behind the philosophical (metaphysical) notion of determinism
is that given the way things are at present, there is only one possible way for
the future to turn out.1 Accordingly, indeterminism (which we equate with the
negation of determinism) amounts to there being more than one possible way
for the future to turn out, given the way things are at present. This core idea lends
itself immediately to a branching representation of indeterminism. Graphically,

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


indeterminism can be pictured as a tree-like structure of possible histories over-
lapping at present and branching into the future. This idea of branching histories
has been rigorously developed in tense logic (by Prior [1967] and Thomason
[1970]), leading also to fruitful applications in other fields such as the theory of
agency (Belnap et al. [2001]; Horty [2001]). In a branching approach, alternative
future possibilities are represented by models that are ‘modally thick’, in the sense
of containing more than one overlapping history (for a comparison with other
modal notions, see Müller [2012]). This feature is central to the DBRN definition
to be detailed below (see especially Section 4.1).
The core idea of determinism appeals to future possibilities, but this notion
does not immediately transfer to mathematical physics. After all, the concept
of future possibilities does not belong to the repertoire of physics; it thus needs
to be spelled out what this concept amounts to in the context of a given
physical theory. Laplace’s ([1951], p. 4) popular metaphor of a demon com-
puting the future of the universe suggests a characterization of determinism in
terms of laws of nature. The resulting doctrine is called nomological state-
determinism: given the laws and a state of the world, there is only one way the
world can turn out to be. Both the notion of state and the notion of laws
employed here require analysis.2
Nomological state-determinism with respect to a given theory can be spelled
out in at least two different ways. The first requires a philosophical concept of
laws of nature and the notion of a possible world satisfying (or being governed
by) such laws. Less metaphysically, the concepts of a theory and of a theory’s
models, rather than those of the laws of nature and of possible worlds, can be
taken as starting points. The DMAP analysis of determinism employs exactly
these two conceptual ingredients: a theory, T, is said to be deterministic just in
case whenever models w and v of T agree with respect to their state at one time,
then w and v agree with respect to their states at all times.
The other way to spell out nomological state-determinism begins with the
observation that a theory’s mathematical representation typically supplies all

1
It is possible to express this core idea without tensed notions, by saying that each event permits
at most one possible subsequent course of events. We will stick to the tensed version in terms of
present state and future development in what follows.
2
Thanks to Balázs Gyenis for discussion of this point.
220 Thomas Müller and Tomasz Placek

the resources needed for assessing the theory’s determinism: a theory’s defining
differential equations assume the role of laws of nature, and solutions to these
equations stand in for physically possible worlds. On the assumption that different
solutions to a theory’s defining equations represent different physically possible
worlds, determinism then boils down to the existence of a unique solution for each
appropriate initial value—this is the essence of the DEQN approach.

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


2.2 The three approaches in more detail
We will now characterize the three approaches in more detail: DEQN,
DMAP, and DBRN. All three presuppose that a theory is given to us as an
object to be diagnosed as to its determinism or indeterminism—the general
structure is depicted in Figure 1.3

2.2.1 DEQN
The equation-based definition of determinism represents the mathematical per-
spective of present physicists, focusing on a theory’s defining (differential) equa-
tions. The leading question is whether for each initial condition there exists a
unique solution for these equations. It turns out that the answer depends on
what kind of differential equations one considers. For ordinary differential
equations (ODEs), there are general methods that allow for a conditional state-
ment of the existence and uniqueness of solutions. In contrast to the tractable
landscape of ODEs, there are no useful general results concerning the existence
and uniqueness of solutions to partial differential equations.
It is important to distinguish here between global existence and local exis-
tence of solutions, where ‘global’ refers to the full range of the time parameter,
and ‘local’ indicates a neighbourhood (possibly arbitrarily small) of a given
moment of time. The question of the existence and uniqueness of solutions
thus splits into two problems. First, is there a unique local solution for each
moment of time? And if the answer to that question is positive, are such local
solutions uniquely extendible to the full, global range of the time parameter?
Now, for an ODE dx dt ¼ f ðx; tÞ, the Peano theorem establishes that for every
initial condition there is at least one local solution of the equation—provided
that the function, f, is bounded and continuous. Further, the Picard–Lindelöf
theorem states that, provided the function f satisfies the so-called Lipschitz
condition, for every initial condition an ODE has at most one local solution.
3
As Wilson ([1989]) remarks, this picture may be unjustified when it comes to assessing, for
example, the determinism or indeterminism of classical mechanics: breakdowns of the determin-
ism of the theory will normally lead to the incorporation of additional assumptions or additional
bits of theory, rather than a flat-out admission of indeterminism. The point remains, however,
that at any stage of practical assessment, one can consider ‘what’s currently on the table’, and a
definition of determinism has to apply at any such stage.
Defining Determinism 221

These results extend to ODEs of arbitrary order and carry over to sets of
ODEs as well.4 As to the extendibility of local solutions to an ODE to a
global solution, in general, the answer is in the negative, though for some
classes of ODEs, under certain conditions, extendibility holds. This is highly
pertinent to the assessment of determinism, as it is global uniqueness that
naturally corresponds to determinism, whereas the mentioned theorems con-
ditionally assure the existence and uniqueness of merely local solutions. Non-

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


extendability then points to a possible failure of determinism due to lack of a
unique global solution—despite there being unique local solutions everywhere
(for some topological details, see Section 4.3). Note then that mathematics
alone indicates how subtle determinism or its failure, indeterminism, can be.
A disadvantage of DEQN is that not all theories can be described neatly in a
form that allows for the application of the DEQN recipe. A notable case in
point is quantum mechanics, for which the defining Schrödinger equation is
very well behaved, but which is often regarded as a main example of an
indeterministic theory. The Born rule, which is an integral part of quantum
mechanics, prescribes probabilities for possible measurement outcomes (see
Section 5.1.2 below for discussion).

2.2.2 DMAP
The mapping-based approach to defining determinism amounts to current
orthodoxy in philosophy of science (see Section 3). This approach takes deter-
minism to be a matter of the existence of suitable mappings in the whole space
of a theory’s temporal realizations. The approach is grounded in Montague’s
([1974]) pioneering formal investigations of deterministic theories from a logi-
cal point of view. Speaking abstractly, the diagnosis of determinism according
to DMAP is a two-stage affair. In a first step, all of the individual realizations
of the linear temporal development of systems falling under the theory are put
side by side. These are the separate possible ways a world could be that are
admitted by the theory. Depending on the theory in question, these could be
all the solutions to the theory’s defining equations, or a class of temporal
realizations that is given in some other, perhaps more complex manner
(quantum mechanics in a consistent histories formulation would be a case
in point here; see Section 5.1.2). In a second step, this class of temporal
developments is checked for instances of indeterminism, in the following
way: If there are two realizations that can be identified at one time,5 but
whose future segments after that time cannot be identified, this signals
4
For a rigorous statement of the mentioned theorems and some useful discussions, see (Arnol’d
[1992]).
5
Instead of identification at a time, for some theories it is necessary to consider identification over
an arbitrarily short interval of time, or over initial segments of temporal developments.
222 Thomas Müller and Tomasz Placek

indeterminism. If the test fails, that is, if all realizations that can be identified
at one time can also be identified at all future times, then the theory is deter-
ministic. The type of mapping that is used to identify different realizations at
different times plays a subtle but crucial role for this definition (see Section 3).
The DMAP definition corresponds to a divergence analysis of future pos-
sibilities, which is popular in current metaphysics (Lewis [1986]). Individual
realizations (ways a world could be) are ‘modally thin’, in that they harbour

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


no possibilities. Possibilities are present only extrinsically, via the existence of
suitable mappings between the realizations.

2.2.3 DBRN
An alternative understanding of future possibilities underlies the branching-
based DBRN characterization of the determinism of theories. Doing justice to
the philosophical idea of alternative future possibilities, in a branching con-
ception a single model is so construed such that it can contain multiple pos-
sibilities. The existence of possibilities is intrinsic to a model, and a model can
thus be modally thick. Rather than opting for linearly ordered temporal rea-
lizations as in DMAP, a branching model is generally only partially ordered in
a tree-like manner; the individual realizations form linear chains (histories)
within that partial ordering. Within one partial ordering, these histories are
bound together by overlapping up to a certain time, so that there is no need to
look for the identifying mappings needed for DMAP. The diagnosis of inde-
terminism is very simple: if there is a model that is not linearly ordered (such a
model contains more than one history), then the theory is indeterministic. A
deterministic theory is one all models of which are linear.

2.3 Representing indeterminism


Figure 2 shows what indeterminism—that is, failure of determinism—looks
like according to the three mentioned definitions. For DEQN, such a failure
comes down to a differential equation admitting globally different solutions
for the same initial data. For DMAP, indeterminism is witnessed by the
existence of two linear temporal realizations that can be mapped at one
time (lower arrow), but not at all future times (upper, crossed arrow). For
DBRN, a witness of indeterminism is a model that is branching rather than
linear.

3 Orthodoxy: DMAP, with Invocations of DEQN


In order to have a point of reference for our work on a formal branching-style
analysis of determinism, we first characterize the dominant approach to
Defining Determinism 223

THEORY:
Indeterministic

extract extract extract

Linear Partially

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


(Differential)
temporal ordered
equations
realizations models

DEQN DMAP DBRN

Figure 2. Failure of determinism according to the three definitions of determinism:


DEQN, DMAP, and DBRN.

defining determinism for scientific theories. This dominant approach owes


much to the work of Jeremy Butterfield and John Earman. Their views are
summarized in two influential encyclopedia entries (Butterfield [2005];
Earman [2006]) and an earlier book (Earman [1986]), on which we will focus.6
Both authors define determinism along the lines of DMAP: any two realiza-
tions (separate mathematical structures) that agree (can be suitably identified
by a mapping) at one time have to agree at all later times.
Butterfield starts by stressing the need to study determinism in terms of
isolated systems. A model of a theory is ‘a sequence of states for such a
single system, that conforms to the laws of the theory’ (Butterfield [2005]).7
The concept of state, as Butterfield notes, is philosophically loaded: states
should be maximal and intrinsic, and in many actual theories, different math-
ematical states correspond to the same physical state (see below). This pre-
cludes a direct use of DMAP in terms of identity. Butterfield ([2005])
accordingly appeals to the notion of isomorphism:
Determinism is [. . .] a matter of isomorphic instantaneous slices implying
that the corresponding final segments are isomorphic (where ‘corre-
sponding’ means ‘starting at the time of the instantaneous slice’). That is:

6
For affirmations of a similar general outlook, see, for example, (Bishop [2006]; Hoefer [2010]).
7
In this section we stick to Butterfield’s use of ‘model’ for what we generally call ‘realization’ or
‘history’, that is, for a single, linear temporal time-course of the development of a system.
224 Thomas Müller and Tomasz Placek

we say that a theory is deterministic if, and only if: for any two of its
models, if they have instantaneous slices that are isomorphic, then the
corresponding final segments are also isomorphic.

The notion of isomorphism appealed to here needs clarification. Butterfield


says that while he uses ‘model’ in the broad sense of philosophy of science, he
uses ‘isomorphism’ in the ‘usual sense used by logicians’ (Butterfield [2005]). A
theory’s model is thus not a model in the logical sense, but a realization. For

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


Newtonian, special relativistic, and general relativistic theories, a theory’s
realization takes the form hM; O1 ; . . . ; On i, where M is a differentiable mani-
fold, and Oi are geometrical object fields on M. Realizations of that form are
used quite generally in discussions of a theory’s determinism in philosophy of
science.8
An isomorphism, in the logical sense, is a structure-preserving bijection
between the domains of two models, where the relevant structure depends
on the characteristics of a language: its class of constants, its relation symbols,
and its function symbols (see Hodges [1993], pp. 5ff.). There is no intuitively
adequate notion of isomorphism that is language-independent.9 Since
Butterfield does not say which language and which symbols need to be con-
sidered, the admissible class of mappings (isomorphisms) remains underspe-
cified. Charitably, one can read the reference to isomorphisms as a promissory
note: for each given theory, a linguistic presentation shall be specified from
which the sought-for notion of isomorphism would follow.10
In view of the difficulties involved in specifying the correct notion of iso-
morphism, it is tempting to phrase the concept of determinism in terms of
identical instantaneous states, but this calls for carefully distinguishing
between states as represented within a given theory and physical states. This

8
Note that a defining feature of manifolds, local Euclidicity, together with the Hausdorff prop-
erty (typically assumed for manifolds in physics applications) implies that there is no branching
in M. As a consequence, there is no way for M to represent alternative possible events; M is
only interpretable as a totality of spatiotemporal events. One might worry that to account for
alternative possible events via branching, either local Euclidicity or the Hausdorff property for
space-times must be sacrificed. We will return to this worry and dismiss it in Section 4.3.
9
See (Halvorson [2012]) for similar issues that arise for attempts to specify a theory in language-
independent terms.
10
For GR, the promissory note is repaid (although not in a strictly logical sense, as no language is
specified) in (Butterfield [1989]). ‘Isomorphism’ is used there in the sense applicable to mani-
folds. A diffeomorphism is a smooth bijection between two manifolds M and M0 . Two models,
hM; Oi i and hM0 ; O0i i, are called isomorphic if and only if there is a diffeomorphism, d, between
the manifolds M and M0 , and for the objects Oi, we have d  ðOi Þ ¼ O0i (where d  ðOi Þ is the object
Oi dragged along by the diffeomorphism d). The definition of determinism is then as follows:

A theory with models hM; Oi i is S-deterministic, where S is a kind of region that


occurs in manifolds of the kind occurring in the models, iff: given any two models
hM; Oi i and hM0 ; O0i i containing regions S; S 0 of kind S respectively, and any
0
diffeomorphism  from S onto S 0 : If  ðOi Þ ¼ Oi on ðSÞ ¼ S 0 , then there is an
0
isomorphism  from M onto M0 that sends S to S 0 , i.e.,   ðOi Þ ¼ Oi throughout
M0 and ðSÞ ¼ S 0 . (Butterfield [1989], p. 9)
Defining Determinism 225

issue is stressed by Earman, whose point of departure is the explication of


determinism for pre-relativistic, pre-quantum theories. Such theories specify a
set of candidates for genuine physical magnitudes, call it O. It is assumed that
each such magnitude takes a definite value at every moment of time t 2 R. The
explication then is as follows:
A history H is a map from R to tuples of values of the basic magnitudes,
where for any t 2 R the state H(t) gives a snapshot of behaviour of the

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


basic magnitudes at time t. The world is Laplacean deterministic with
respect to O just in case for any pair of histories H1, H2 satisfying the
laws of physics, if H1 ðtÞ ¼ H2 ðtÞ for some t, then H1 ðtÞ ¼ H2 ðtÞ for all t.
(Earman [2006], p. 1370)11

Here, the correspondence between the set O and the set of genuine physical
states is crucial. Since elements of O may have mathematical surplus structure,
the failure of the requirement that ‘if H1 ðtÞ ¼ H2 ðtÞ for some t, then H1 ðtÞ
¼ H2 ðtÞ for all t’ need not signal indeterminism. It is a typical situation in
physics, and not some mere philosophical possibility of theoretical under-
determination, that a theory’s mathematical descriptions correspond many-
to-one to physical states, so that the identity in the above quotation needs to
be replaced by a broader notion of agreement. Consider classical electromag-
netism, where the electric field, E, and the magnetic field, B, are derived from a
scalar potential, ’, and a vector potential, A. The relation between A;  and
E; B is many-to-one: for any smooth function , the potentials A; ’, and the
potentials
qc
A0 ¼ A þ rc; ’0 ¼ ’  ;
qt
represent the same physical situation, and the same E and B fields. The trans-
formation A ° A0 ; ’ ° ’0 is called a gauge transformation. The arbitrariness
of the choice of means that the theory has surplus mathematical structure
(‘gauge freedom’). For a theory with gauge freedom, the fact that two realiza-
tions have the same mathematical state at a certain t, but different states at
some later t0 , is not of itself indicative of indeterminism; it could also be that
the states at t0 represent the same physical state by different mathematical
means. It will not do to demand that only theories without gauge degrees of
freedom are considered, since there are good scientific reasons for allowing
that kind of freedom in our physical theories.12 Thus, to decide the question of
determinism of a theory requires us to decide whether the divergence of reali-
zations results from gauge freedom or not. This is conceptually difficult, as
11
Although Earman focuses on future- and past-oriented determinism, whereas Butterfield ana-
lyses future-oriented determinism, each explication can be readily extended to accommodate
both versions of determinism.
12
As Earman ([2006], p. 1381) points out, attempts at treating gauge degrees of freedom as
physical quantities subject to dynamical laws are generally not fruitful.
226 Thomas Müller and Tomasz Placek

one of the main motivations for believing gauge freedom to be operative in a


theory is to maintain determinism. We will return to this issue in a more
formal setting in Section 5.2, where we compare the DMAP and DBRN
approaches to determinism.
Our final observation is that although both Butterfield and Earman expli-
cate determinism in terms of mappings (DMAP), they both add a gloss to the
effect that for theories given via differential equations, their definitions corre-

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


spond to the existence of unique solutions to such equations,13 which amounts
to DEQN. The nature of this correspondence is, however, left open.

3.1 Four marks of orthodoxy


To summarize, the current orthodoxy in treating the question of determinism
has the following four marks:
(1) A theory is represented by the class of its realizations (‘models’)—
possible total time courses of evolution of a system to which the
theory applies. The realizations are modally thin: a single realization
does not contain different possibilities. Such possibilities (whose exis-
tence implies indeterminism) thus have to be represented via relations
between realizations.
(2) Accordingly, the question of whether a theory is deterministic or not
is understood as a question about the structure of the class of its
realizations, to be spelled out in terms of suitable mappings; deter-
minism means that agreement of two realizations at one time implies
agreement at all later times.
(3) The notion of ‘agreeing at a time’ is crucial, and gives rise to com-
plications. The agreement is meant to be with respect to physical
states of the system, but the class of realizations contains mathema-
tical objects that may have surplus structure, for instance, due to
gauge freedom.
(4) It is assumed that the practical assessment of determinism or inde-
terminism of a given theory mostly depends on the behaviour of that
theory’s defining differential equations: A strong link is claimed
between determinism in the mapping sense (DMAP) and the well-
posedness of the initial-value problem, if the laws of a theory in
question are formulated by differential equations.
13
Cf. (Butterfield [2005], p. 98): the given ‘definition [. . .] corresponds to such a set of equations
having a unique solution for future times, given the values at the initial time’. Similarly, (Earman
[2006], pp. 1371f.): ‘[. . .] the laws of physics typically take the form of differential equations, in
which case the issue of Laplacean determinism translates into the question of whether the
equations admit of an initial value formulation, i.e., whether for arbitrary initial data there
exists a unique solution agreeing with the given initial data’.
Defining Determinism 227

Points (1)–(3) show that the orthodox definition of determinism amounts to


DMAP, with point (3) constraining the suitable mappings. Point (4), however,
makes a strong link to a DEQN-style definition of determinism.

4 Branching-Style Determinism (DBRN)


We turn now to a branching characterization of determinism. As we said, that

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


characterization is an attempt to capture directly the core idea of determinism:
the present has exactly one possible future. The crucial concept of alternative
future possibilities is analysed by means of branching histories; a system is
assessed as indeterministic if and only if some initial segment of its evolution
can be continued in more than one way. Although a branching analysis of
future possibilities is rarely used in philosophy of science, it has been rigor-
ously developed in logic and its applications, for instance, in theories of
agency. The branching concept is also sometimes used in natural sciences
(see below). The general definition of determinism according to DBRN is as
follows:

Definition 1 (Determinism of a theory): A theory is indeterministic if and


only if it has at least one faithful indeterministic model. The theory is
deterministic if and only if all of its faithful models are deterministic.

So far, this is just passing the buck. To see in what way branching is
involved, we need to specify a sense of ‘model’, ‘faithful’, and ‘indeterministic’
such that a model of a theory can be both faithful and indeterministic.

4.1 Models and realizations


In the orthodox DMAP approach described in Section 3, ‘model’ is often used
as a synonym for ‘realization’ or ‘history’. Our proposed usage here is
broader: a model of a theory is whatever fulfils the requirements of that
theory. Thus, a model can be modally thick (containing structures represent-
ing alternative possibilities) or modally thin (containing no such structures).
‘Realization’, on the other hand, is tied to a linear temporal development, and
thus, realizations are always modally thin. More formally, a realization,
hT; <; S; f i, of a theory specifies a function, f, from a linearly ordered set of
times, hT; <i, to mathematical states S.14 The function, f, must be admitted by
the dynamics of the theory; typically, it must be a solution to the theory’s
dynamical equations given some initial data.
14
We will mostly take hT; <i to be hR; <R i, but we also allow for discrete time or other temporal
structures.
228 Thomas Müller and Tomasz Placek

With a view to Definition 1, we are looking for a broader formal notion of a


model of a theory that allows such a model to be intrinsically indeterministic.
We can take a lead from logic (see, for instance, Thomason [1970]), where
indeterministic models are discussed in the context of so-called ‘branching
time’,15 and from the stochastic processes literature (for example, van
Kampen [2007], Chapter 3), in which a model of a stochastic process includes
not just a single realization, but many incompatible realizations. The consis-

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


tent histories approach to quantum mechanics (see Griffiths [2002]; Section
5.1.2, below) also employs a branching time representation. In all these
approaches, a model is allowed to be modally thick by representing different,
incompatible future possibilities in one mathematical structure.
We will thus take a model, hM; <; S; f i, to specify a function, f, from a
possibly branching, tree-like partial ordering, hM; <i, to the allowed states, S,
and such that the restriction of f to each realization accords with the theory’s
dynamics. Formally, we require of hM; <i, for x, y, z 2 M:
. asymmetry: if x < y, then not y < x;
. transitivity: if x < y and y < z, then x < z;
. backwards linearity: if x < z and y < z, then either x ¼ y or x < y or y < x;
. connectedness: for any x and y, there is some z such that z  x and z  y.

A chain in hM; <i is a linear subset; by virtue of Zorn’s lemma, there are
maximal chains in M. A tree-like order hM; <i can contain more than one
maximal chain, as in the lower right of Figure 2 (see also Figure 3).16 It is
usually sensible to require that all maximal chains in M be order-isomorphic,
for example, all isomorphic to hR; <R i.
Based on this notion of a model, the following definition of determinism is
adequate:

Definition 2: A model, hM; <; S; f i, is indeterministic if and only if hM; <i


contains more than one maximal chain. The model is deterministic if and
only if hM; <i contains just one maximal chain.

The existence of more than one maximal chain in hM; <i means that there is
more than one realization in the model hM; <; S; f i, since each maximal chain
h  M in the model specifies a realization hh; <jh ; S; fjh i, simply by restricting
f to h, as hh; <jh i is a linear order. Accordingly, a deterministic model of a
theory contains just one realization, whereas an indeterministic model bundles

15
See also (Belnap et al. [2001], Chapter 7A) and, for the related framework of branching space-
times, (Belnap [1992], [2012]).
16
Provably, M contains more than one maximal chain if and only if it contains at least one upward
fork, that is, three moments, x, y, and z, for which x < y, x < z, but there is no common upper
bound for y and z, which holds if and only if neither y  z nor z < y.
Defining Determinism 229

together a number of realizations in one branching tree. These realizations


branch in the following sense: for any pair, h and h0 , of distinct maximal
chains, h \ h0 6¼ ; (by connectedness), and for every x 2 h \ h0 , we have
fjh ðxÞ ¼ f ðxÞ ¼ fjh0 ðxÞ. Viewed from such an x, the realizations hh; <jh ; S; fjh i
and hh0 ; <jh0 ; S; fjh0 i thus describe (alternative) possible future developments of
the system in question. We may thus interpret realizations as (alternative)
possibilities, and call them possible histories. As a single model can contain

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


different histories, a model in our sense can capture modality intrinsically.

4.2 Faithfulness
There is one loose end left to tie up: So far, we have seen that it would be
possible to produce an indeterministic model from a linearly ordered, deter-
ministic model hM; <; S; f i simply by adding a disjoint copy of a final segment
of M to create a forward-branching structure, and extending f on the new
branch by copying. This would be indeterminism on the cheap. We will require
a faithful branching model that contains no difference in the ordering without
a corresponding difference in states.17 On a strict reading, faithfulness requires
a difference in states at, or immediately after, the splitting of any two histories.
In the case of what we will call case (b) branching (such that branching
histories have a first moment of disagreement, see Section 4.3), this boils
down to the requirement that the first moments of difference in two histories
have different states assigned.18 In case (a) branching, where there is a max-
imal moment in the intersection of two histories, faithfulness means that states
should be different immediately after such a maximal moment. This can be
made more precise by adding a further structure that identifies moments
occurring at the same time in different histories.19
In Section 5.2, where we construct a DBRN representation of a system out
of a DMAP representation of a system, we will use a weaker concept of
faithfulness that simply requires that two branching histories must be different
state-wise. We will also discuss how to formally represent sameness of physical
states, given that physical states might have non-unique mathematical repre-
sentations resulting from gauge freedom or from the symmetries of a theory.
17
In some cases we may want to drop the requirement of faithfulness. For example, if we know
that a process is indeterministic and thus has to be modelled by a branching model, but the
assigned states agree on two different maximal chains, this may signal that the theory is incom-
plete. In the present context, however, our aim is to diagnose (in)determinism, and so we are
methodologically required to assume completeness.
18
In the case of Figure 3(b), we thus want f ð01 Þ 6¼ f ð02 Þ.
19
See (Belnap et al. [2001], pp. 194–6) on instants as partitions of the set of moments. In the case of
Figure 3(a), if these instants have the plausible form ht1 ; t2 i for t 2 Rþ , so that we can identify
the time of moment t1 on the upper track with the time of t2 on the lower track, we thus want that
for any t > 0 there is some t0 < t; t0 > 0, such that f ðt01 Þ 6¼ f ðt02 Þ.
230 Thomas Müller and Tomasz Placek

(a) 11
-1 0

12

(b) 01 11
-1

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


02 12

Figure 3. Two types of branching. Both (a) and (b) depict two continuous histories
branching at point 0. In (a), point 0 is the shared maximum in the intersection. In
(b), the intersection of the histories has no maximum, and points 01 and 02 are
different minimal upper bounds of the intersection.

4.3 Two types of branching topologies


Consider a branching order hM; <i in which there are two maximal chains, h1
and h2, both order-isomorphic to R. What does such a model look like?
Mathematically, this is equivalent to asking how we can glue together two
copies of the real line, Ri ¼ fht; iijt 2 Rg, i ¼ 1, 2, via an equivalence relation
 on R1 [ R2 such that M ¼ R1 [ R2 = , with the obvious ordering. Let us
agree that the branching should happen at 0, so that ht; 1i  ht; 2i for t < 0 but
not for t > 0. There are two possibilities, pointing to two different sensible
options for branching topologies: Either (case (a)) we add the condition
h0; 1i  h0; 2i—in this case, the histories h1 and h2 have a maximum
0 :¼ fh0; 1i; h0; 2ig in their intersection; or (case (b)) we demand
h0; 1i¿h0; 2i—then, there is an upper bounded chain in M that has two dif-
ferent minimal upper bounds, 0i :¼ fh0; iig, i ¼ 1, 2 (see Figure 3).
Both of these cases are related to a failure of determinism in the following
sense (assuming that the histories after the branching represent physically
distinct developments of the system; the mentioned times are arbitrary):
Both in case (a) and in case (b), at t ¼ –1, there is a unique state of the
system, but at t ¼ 1, two different states are possible, corresponding to differ-
ent moments, 11 and 12. So, the system at t ¼ –1 has two alternative possible
future developments. The cases are different, however, in that at t ¼ 0, there is
a unique state in case (a), but not in case (b). In case (a), the branching of the
system’s development happens immediately after t ¼ 0, whereas in case (b), the
branching has happened already at t ¼ 0, but at no time prior to t ¼ 0.
Considered as topological spaces,20 if one imposes the plausible restriction
that open subsets of R in the copies away from the branching should be open
20
For topological notions, see, for example, (Munkres [2000]).
Defining Determinism 231

and the resulting space should be connected, one sees that case (a) fails to be
locally Euclidean (any open set containing 0 has to contain some initial
segment of both tracks), whereas case (b) is a generalized manifold
(a locally Euclidean space, in which, however, the Hausdorff condition fails
for 01 and 02).
In terms of the behaviour of differential equations—that is, taking the
function, f, to be provided by solutions to such equations—case (a) represents

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


a situation in which unique initial data can be given for some time, t, but there
is no unique local solution immediately after t. Case (b) represents a different
and somewhat trickier affair, since a system of differential equations giving
rise to this case need not exhibit any failure of local uniqueness. In Figure 3(b),
at any time t < 0, there is a unique solution for at least a little while (for
example, until t =2), and for the moments tþ þ
i , where t  0 and i ¼ 1, 2, the
development along hi is unique for all times. Indeterminism reigns, however, in
the sense that, like in case (a), the state at any t < 0 does not determine a
unique state for all t > 0. In a discussion of such cases from the point of
view of DEQN, it is not local well-posedness (this is satisfied), but the exis-
tence of solutions for some longer time, or the extendability of locally unique
solutions past some crucial time, that is decisive. A branching representation
helpfully allows one to understand indeterminism as a local affair even in case
(b): in a topologically obvious sense, the points 01 and 02, which witness the
first instance of non-uniqueness, are actually closer than any two distinct
points in case (a); they are non-Hausdorff-related, meaning that they
cannot be separated by disjoint open sets.
A worry may arise at this point that the DBRN approach permits branching
within a system’s history, or that time (or space-time, in a spatiotemporal
extension of the framework; see Belnap [1992]) is not Hausdorff. A physical
space-time without the Hausdorff property has dire consequences, described
in (Earman [2008]).21 The worry, however, is groundless, since in each possible
realization, time has the topological structure of R (or, respectively, space-
time is a Hausdorff manifold). It is just that an indeterministic model repre-
sents the different possibilities for a system’s future temporal development in
one single structure. A mapping-based representation of the system has to
represent exactly the same modal facts; it just brushes under the carpet
the topological aspect of the development of the system’s states by using a
non-overlapping representation. For a discussion of topological facts about
(in)determinism in space-time theories, as well as for a proof that each space-
time in Belnap’s ([1992]) theory of branching space-times is Hausdorff, see
(Placek and Belnap [2012]), (Müller [2013]), and (Placek et al. [2014]).

21
The same paper voices the worry that branching implies a failure of the Hausdorff property.
232 Thomas Müller and Tomasz Placek

5 Comparing the Approaches


In comparing the three approaches—DEQN, DMAP, and DBRN—we will
focus on two aspects. First, we would like to learn how the three approaches
work in actual cases, that is, how a particular theory (or a system falling under
the theory) is assessed through the lens of each approach. To this end, we
explore three examples: Norton’s dome (Section 5.1.1), quantum mechanics
(Section 5.1.2), and GR (Section 5.1.3). We draw some general conclusions

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


from these examples in Section 5.1.4.
Second, since both DMAP and DBRN offer rigorous definitions of deter-
minism, we will investigate how the respective representations of determinism
and indeterminism are formally related. Do these two approaches always
deliver the same verdicts with respect to a theory’s determinism or indetermin-
ism? Answers to these questions might cast some light on which of the
approaches provides the more adequate analysis of determinism. We turn to
a formal comparison of DMAP and DBRN in Section 5.2.

5.1 Case studies


5.1.1 Norton’s dome
Norton’s ([2008]) dome is a system of Newtonian physics whose indetermin-
ism results from a failure of the Lipschitz condition mentioned above in
Section 2. At time t ¼ 0, a point particle of unit mass is at the apex r ¼ 0 of
a dome, whose surface satisfies the constraint h ¼ ð2=3gÞr3=2 , where g is the
strength of the homogeneous vertical gravitational field, r is the radial distance
from the apex, and h is the vertical distance from the apex. Since the dome is
discussed in the context of point particle mechanics and the mass point is
restricted to one dimension, the set of system states is the set of possible
radial distances, S ¼ R.22
Newton’s second law yields
d 2r
¼ r1=2 : ð1Þ
dt2

Given the initial data rð0Þ ¼ 0 and dr=dtð0Þ ¼ 0, Equation (1) has a station-
ary solution r1 ðtÞ ¼ 0, as well as a family of solutions rb, parametrized by the
real-valued parameter b  0:

22
We stick to this simple choice of S here in order to keep the exposition simple. Nothing of
substance is changed if we take the system’s state at a moment to specify not just the particle’s
radial distance from the apex (S ¼ R), but that radial distance together with the particle’s
instantaneous momentum (S ¼ R2 ).
Defining Determinism 233
8
<0 if t  b
rb ðtÞ ¼ ðt  bÞ4 ð2Þ
: if t > b:
144

Let us look at the three approaches to defining determinism in turn. DEQN


offers the most straightforward analysis, delivering the verdict of indetermin-

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


ism: there is no unique solution to Equation (1) for the initial data. On its own,
DEQN does not say how massive this indeterminism is, that is, how many
different possibilities there are. There are continuum many functions
rb ðtÞ; b 2 Rþ 0 , which suggests continuum many possibilities. But on the
assumption that the point particle has been located at the apex of the dome
forever before t ¼ 0, all rb-type solutions are related by a time-translation
symmetry, which suggests that all these solutions represent a single possi-
bility.23 In that case, there would be exactly two possibilities, r1 and rb.
To apply the DMAP analysis, one needs first to have a grip on the set of
realizations. Recall that the canonical form of a realization is based on mani-
folds (see Section 3), and observe that the real line is a differentiable manifold.
Accordingly, each hR; <R ; rb i as well as hR; <R ; r1 i is a realization. To decide
on determinism, we then ask if realizations with isomorphic initial segments
are globally isomorphic. As above, any two rb-type realizations are plausibly
related by time-translation trt : rb !rbþt , a clear case of isomorphism. Thus,
these realizations are not witnesses for indeterminism. Nevertheless, DMAP
delivers the verdict of indeterminism, which is secured by the existence of the
r1 solution; this solution cannot be derived from any of the rb by a time-
translation.
DBRN calls for a construction of a branching-style model for Norton’s dome.
We begin with an auxiliary set M ~ :¼ fht; bijt; b 2 R; b  0g, define the relation
& on M ~ by putting ht; bi&ht ; b0 i if and only if t ¼ t0 and ðb ¼ b0 or ðt  b and
0

t  b0 ÞÞ (which is provably reflexive, symmetric, and transitive) and define our


base set, M, as the quotient structure (i) M :¼ M ~ = &. We write elements of M
0 0
as ½t; b :¼ fht ; b i 2 M ~ jht ; b i&ht; big and define the relation l on M as (ii)
0 0

½t; b l ½t ; b  if and only if t < t0 and ½t; b ¼ ½t; b0 . It can be proved that hM; li
0 0

satisfies the postulates for a tree-like partial ordering. hM; li has a family
þ
fhb gb2Rþ of maximal chains, where Rþ 0;1 ¼ R [ f0g [ f1g. We will associ-
0;1
ate h1 with the stationary solution, and hb, for 0  b < 1, with a solution in
which the mass point begins to move immediately after time t ¼ b.

23
It is important that the particle has been at the apex for all t < 0. If there is some first time t0 < 0
at which the particle is placed on the apex, the solutions have to satisfy rb ðtÞ ¼ 0 only for
t0  t  b, and will not be time-translation symmetric. This would speak against counting
them as just one possibility.
234 Thomas Müller and Tomasz Placek

We now construct the actual branching model. The function, f, for the
model is
8
<0 if t < b
(iii) f ð½t; bÞ ¼ ðt  bÞ4 ð3Þ
: if t  b:
144

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


Now we can define:

Definition 3: The branching model for Norton’s dome is the quadruple


hM; l; R; f i, where M; l, and f are defined by conditions (i), (ii), and
(iii), respectively.

Note that the model has a ‘stationary’ history, hh1 ; l; R; fjh1 i, representing
the mass point remaining stationary on the dome’s apex, as well as a family of
‘dynamic’ histories of the form hhb ; l; R; fjhb i (0  b < 1), representing the
mass point remaining on the apex until time b, and then moving in accordance
with Equation 2.24 The model exhibits (a)-type branching, since every inter-
section hb \ hb0 (with b 6¼ b0 ) has a maximum ½c; c (where c ¼ minfb; b0 g).
Finally, the model constructed above is faithful in our sense: if two maximal
chains hb and hb0 branch at a moment, then there is a difference in states
assigned to elements of hb and of hb0 immediately after that moment. The
verdict thus is that Norton’s dome is indeterministic.
We draw some morals from the application of DEQN, DMAP, and DBRN
to Norton’s dome in Section 5.1.4—after discussing two more cases.

5.1.2 Quantum mechanics


With respect to the determinism question, standard quantum mechanics is the
odd one out. That theory is based on a very well-behaved differential equation
(suggesting determinism), but its essential ingredient is a probabilistic algo-
rithm that answers what, and how probable, are the possible results of a
measurement (which suggests indeterminism). In order to pass a final verdict
about the determinism of quantum mechanics one would thus need to resolve
the conflict between these two aspects of the theory (known as the measure-
ment problem).
It is our contention, however, that controversies surrounding determinism
of quantum mechanics partially derive from a failure to distinguish between
various senses of determinism, as captured in the three approaches, DEQN,
DMAP, and DBRN. Without proposing a solution to the measurement
24
Note that the model involves continuous branching, since if the mass point is at rest at some
t  0, then it can start moving at any later time.
Defining Determinism 235

problem or any other grand thing, we will sketch how determinism of quan-
tum mechanics is to be analysed through the lens of each of the three
approaches, focusing in particular on a branching-style representation of
that theory.
For the DEQN approach, the main fact is the form of the Schrödinger
equation,
qc

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


i
h ^ c;
¼H
qt
which governs the temporal evolution of isolated quantum-mechanical sys-
tems. The question then is whether for different Hamiltonians H ^ (which char-
acterize different quantum-mechanical systems), the Schrödinger equation
yields a set of differential equations with a unique solution for each appro-
priate initial value. There is a wealth of information on this subject (and
related subjects) in the mathematical literature.25 One class of results points
to Newtonian systems for which uniqueness of solutions does not obtain, but
the quantum counterparts of which are deterministic: in contrast to the former
systems, for the latter systems there exist unique solutions to the ensuing set of
equations. Such results suggest that quantum mechanics is more deterministic
than classical mechanics. On the other hand, certain Hamiltonians allow for
multiple temporal evolutions.26 It is then a matter of controversy whether
those Hamiltonians are physically meaningful. Although this question is per-
tinent to the issue of determinism of quantum mechanics, it is clear that
considerations of the evolution equation alone fail to provide an adequate
picture of quantum mechanics—DEQN simply ignores the quantum prob-
abilistic algorithm.
One might hope that DMAP may be more helpful as it is not restricted to
comparing solutions to the Schrödinger equation. Its data for comparison are
the theory’s realizations, and there seems to be no obstacle to including in the
latter other entities referred to by quantum mechanics, in particular, measure-
ment results. There is, however, virtually no DMAP-style analysis of quantum
mechanics, although a Montague-style language-based notion of realization
seems to be readily available.
In contrast to the above mentioned modally-flat notion of realization,
DBRN calls for producing branching models for quantum mechanics.
There is a formalism for standard quantum mechanics—the so-called con-
sistent histories approach (Gell-Mann and Hartle [1993]; Griffiths [2002])—
that explicitly employs a branching time representation. A ‘family of

25
A good introduction to this field is (Earman [1986]). For a more mathematically advanced
treatment, see (Earman [2006]).
26
Those that admit multiple self-adjoint extensions; cf. (Earman [2006], p. 1401).
236 Thomas Müller and Tomasz Placek

histories’ is a set fY h jh 2 Hg of mutually exclusive histories (chains of pro-


jectors) of the form

Y h ¼ Ph1 (Ph2 ( (Phm ;27

where at each chosen time ti, i ¼ 1; . . . ; m, the family contains projectors Phi
from an exhaustive set of ni mutually exclusive projectors P1i ; . . . ; Pni i ,
X
ni

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


Pji Pki ¼ djk Pji ; Pji ¼ I:
j¼1

A consistency condition constrains the set of admissible families of his-


tories. Such a family of histories directly specifies a model in the sense of
Definition 2, and when such a family has more than one member, that
model is indeterministic. Topologically, the indeterminism is of the case (b)
variety, as at the times chosen, different projectors are assigned. This assign-
ment of different projectors also ensures that the branching model is faithful in
our sense.
The formalism can be, and usually is, extended to allow for a branch-
dependent selection of the set of projectors at any given moment. This also
leads to a faithful branching model in our sense, and the same criterion for
(in)determinism applies: once a family contains more than one history, it is
indeterministic. Note that the formalism allows one to assign probabilities for
the members of a branching family in agreement with experimental results.

5.1.3 General relativity


In philosophical discussions of the determinism of GR, the DMAP approach
is the most prominent one. As the theory has realizations of the required form
(differentiable manifolds), and a notion of isomorphism for segments of such
models can be rigorously defined (via diffeomorphisms), GR is amenable to
the DMAP analysis. Since the late 1980s, the DMAP analysis has been applied
to GR in a particular way in the philosophy of physics, having been driven by
Einstein’s hole argument.28 This argument is an appeal to a special transfor-
mation on manifolds—the so-called hole diffeomorphism—which is used to
produce a different manifold out of a given one, with the resulting pair of
manifolds (seemingly) witnessing indeterminism of GR. Importantly, all par-
ties in the debate were unanimous that this phenomenon does not show inde-
terminism in GR. The consensus is rather that the argument raises the
27
A more general set-up in terms of positive operator valued measures is possible; see, for exam-
ple, (Peres [2000]). See also (Müller [2007]) for more details about the relation of consistent
histories and branching time representations.
28
See (Earman and Norton [1987]; Butterfield [1989]).
Defining Determinism 237

question of whether two differential manifolds related by a diffeomorphism


represent one physical space-time or two.
In contrast to philosophers’ uses of the DMAP analysis in the hole argu-
ment, physicists concerned with the (in)determinism of GR appeal to the
DEQN analysis. They investigate whether GR admits a globally well-posed
initial-value problem, that is, whether the data on an appropriate space-like
slice of a space-time can be uniquely extended to that space-time. This ques-

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


tion is typical of the DEQN approach.29 In this section, we investigate whether
the practitioners’ approach to the initial-value problem exhibits similarities to
the DBRN characterization of determinism, and whether these can be used to
construct branching models for GR.
The basic context of the practitioners’ debate on the initial-value problem in
GR is provided by so-called globally hyperbolic space-times.30 Such space-
times admit a global time function; surfaces of constant value of such a func-
tion are Cauchy surfaces. Thus for a globally hyperbolic space-time we are in
the familiar setting of a global time with momentary states (identified with
Cauchy surfaces with appropriate data on them).31 In this context, the starting
point of both the determinism issue and the initial-value problem is the ques-
tion of whether a three-dimensional Riemannian manifold with appropriate
data on it fixes a unique four-dimensional Lorentzian manifold of a specified
kind. ‘A specified kind’ refers here to a particular form of Einstein’s field
equations, which depends on whether a space-time includes matter and, if
so, what the model for this matter is, and whether the equations include a
non-zero cosmic constant . These decisions are also highly relevant to what
the appropriate initial data (three-dimensional Riemannian manifolds with
some objects on them) are.
We consider here only globally hyperbolic space-times that are vacuum
solutions of the Einstein equations, that is, for which the Ricci curvature
tensor R ¼ 0. In this case the initial data are triples hM; g; ki, where M is
a three-dimensional manifold, g is a Riemannian metric, and k a symmetric
covariant tensor (coding incremental changes of g in the direction normal to
the manifold). Further, for hM; g; ki to be embeddable in a globally hyperbolic
space-time satisfying Einstein’s equations, it should satisfy certain equations,
called initial-value constraints. Triples hM; g; ki satisfying these constraints
are called ‘vacuum data sets’.
29
Since physicists see the well-posedness of the initial-value problem as a criterion any good theory
should satisfy, their interest is biased towards determinism. It is thus striking to see them
concede that in some cases GR is indeterministic.
30
A space-time is globally hyperbolic if it admits a Cauchy surface—for further pertinent defini-
tions see (Wald [1984], pp. 200ff.).
31
Another determinism-friendly feature of a globally hyperbolic space-time is that the domain of
dependence of a Cauchy surface is, by definition, identical to the whole space-time, which,
roughly speaking, excludes influences coming from nowhere.
238 Thomas Müller and Tomasz Placek

The following theorem (Choquet-Bruhat and Geroch [1969]) is relevant to


whether or not GR is deterministic in the vacuum case:
Let hM; g; ki be a vacuum data set. Then there is a unique, up to
isometry, maximal vacuum Cauchy development (MVCD) of hM; g; ki.
To explain, a vacuum Cauchy development of hM; g; ki is a globally
hyperbolic 4-dimensional Lorentzian space-time in which hM; g; ki is
embeddable. Two space-times hM; g i and hM0 ; g0 i are isometric if
there is a diffeomorphism ’ : M!M0 such that ’ ðg Þ ¼ g0 (it is not

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


required that other objects be dragged along by ’).

Theorems similar to this hold for the Einstein equations with other data.32
From our perspective, the interesting point is that these theorems do not
prohibit a maximal Cauchy development of an initial data set from having
more than one non-isometric extension—the theorems only prohibit these
extensions from being globally hyperbolic. A case in point is provided by
the so-called polarized Gowdy space-time. This is a globally hyperbolic
space-time, defined for a restricted set of values of one coordinate, a
vacuum solution to the Einstein equations, and an MVCD of an appropriate
initial data set. When this space-time is extended for the full range of the
coordinate, some of its maximal extensions turn out not to be isometric
(Chruściel and Isenberg [1993], pp. 1623ff). These non-isometric extensions
might be viewed as possible histories of a faithful indeterministic model, yield-
ing the verdict that GR is indeterministic.33
The issue is, however, complicated. The mentioned non-isometric maximal
extensions of the polarized Gowdy space-time do not admit global time func-
tions and contain closed time-like curves (CTCs). This spells trouble for any
definition of determinism based on partial orderings, since there is no natural
antisymmetric ordering on a CTC. Accordingly, the notion of alternatives for
the future, which is basic to the core idea of indeterminism, makes sense only
locally, but not globally any more.34
32
Notably, (Ringström [2009], p. 147, Theorem 16.6) proves the existence of a maximal globally
hyperbolic development of the data for a specific model with matter, the so-called non-linear
scalar field model. In this case, the initial data sets are different from their counterparts in the
vacuum solution case, and embeddability applies to the matter field as well.
33
Non-isometric extensions of a maximal Cauchy development are deemed non-generic by the
strong cosmic censorship conjecture. In this spirit, (Chruściel and Isenberg [1993]) prove that
non-isometric extensions of a polarized Gowdy space-time are rare, in a measure-theoretical
sense, in the set of all extensions of that space-time. As there is little ground to equate ‘rare’ with
‘non-physical’, the example cannot be discounted easily. For a discussion of the strong cosmic
censorship conjecture in the context of polarized Gowdy space-times, see (Chruściel et al.
[1990]).
34
This calls for spelling out our Definition 2 of indeterminism in terms of modal forks: ‘A model is
indeterministic if and only if it contains at least one modal fork’; for the definition of modal
forks, see (Placek et al. [2014], p. 423). The two formulations coincide in the context of branch-
ing time, as any two histories in branching time form such a fork. The formulation in terms of
modal forks is, however, also applicable in the context of more complex branching theories in
which histories are not linearly ordered.
Defining Determinism 239

Certainly more work is needed to fully develop the core idea of indetermin-
ism with respect to space-times admitting CTCs. With respect to our approach
laid out in Section 4, we note the following: The main definition of the DBRN
approach, Definition 1, is still adequate due to its abstract nature. It will,
however, be necessary to extend the definition of an indeterministic model
(Definition 2) such that space-times admitting CTCs are covered as well, by
taking local alternatives into account.35 In parallel to this development, it will

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


be necessary to scrutinize the arguments of the practitioners. For example,
invoking the DEQN approach, it is claimed that ‘the fact that there are
inequivalent maximal extensions means that the initial data do not uniquely
determine a maximal development. In this sense, the general theory of rela-
tivity is not deterministic’ (Ringström [2009], p. 18). Depending on details of
the definition, DMAP and DBRN may here come to opposing verdicts; see
our discussion of the interrelation of the two approaches in Section 5.2 below.

5.1.4 Morals from the applications


We have illustrated above how the three approaches to determinism fare in
analysing the theories of Newtonian mechanics, non-relativistic quantum
mechanics, and GR. Our first observation is that it is DEQN that is mostly
used in the cases considered (and also in actual discussions among practi-
tioners). The existence of non-unique solutions to Newton’s equations in
Norton’s dome indicates indeterminism; in a similar vein, the fact that
Einstein’s field equations allow for non-isometric extensions to a MVCD
counts against determinism of GR. The behaviour of the Schrödinger equa-
tion, however, does not account for the general sentiment that quantum
mechanics is an indeterministic theory. In this case, the DEQN approach is
severely limited, as it does not accommodate the quantum measurement
algorithm.
Second, the DMAP characterization of determinism does not seem to be
really used. That is, in cases like that of Norton’s dome, the construction of
realizations needed by the DMAP analysis is straightforward and completely
relies on solutions to the theory’s defining equations. The construction does
not add any new value to what is achieved by the DEQN analysis. In the case
of quantum mechanics, DMAP looks promising, as it offers a chance to
account for measurement results, apart from the Schrödinger evolution.
However, that promise of a DMAP analysis of quantum mechanics has
never been fulfilled, as far as we know.
Third, approaching the question of determinism of GR from the perspective
of the initial-value problem refers one to the DEQN approach. Non-isometric
35
A theory of this sort is developed in (Placek [2014]).
240 Thomas Müller and Tomasz Placek

space-times that witness indeterminism in GR might then be viewed as diver-


ging realizations according to the DMAP definition. Since these space-times
contain CTCs, a full DBRN analysis will need to be based on an extended
notion of an indeterministic space-time model.
Finally, we have seen that the branching analysis of determinism comes
naturally—a DBRN-style representation is often quite literally out there.
For quantum mechanics, DBRN-style models are immediately available, in

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


the formalism of quantum histories. For other cases, such models need to be
constructed. We showed such an explicit construction for Norton’s dome.
These constructions are quite natural, and we conclude that the notion of
an indeterministic DBRN model supplies a useful representation of a theory’s
indeterminism.

5.2 Formal comparison of the DMAP and DBRN frameworks


We now turn to a formal comparison of the DMAP and DBRN frameworks,
in order to find out about their interrelation. The comparison will be at the
level of a single system falling under the theory in question, which means that
we are treating the set of the system’s states, S, as fixed and given.36 We will
compare formal mapping and branching representations of the system’s
dynamics, using the following data format: A mapping representation of the
system’s dynamics is a pair
M ¼ hðMj Þj2J ; Ai; Mj ¼ hTj ; <j ; fj i;

with J some index set. Here, the Mj are the realizations characterizing the
system; any hTj ; <j i is a linear ordering of times (typically, hR; <R i); and
fj : Tj ° S is a specification of system states for times t 2 Tj . Furthermore,
A is a class of isomorphisms between realizations, allowing for the fact that
different mathematical structures may represent the same physics (see our
discussion of gauge transformations in Section 3 above.). Technically,
each  2 A is a mapping between realizations that preserves their structure,
which means that it specifies an order-preserving bijection identifying the
times across different realizations, and it maps corresponding system states
onto physically equivalent system states. In line with typical considerations in
physics, we will assume that the set of isomorphisms, A, has the structure of a
group, that is, elements of A can be combined such that (i) there is a neutral
element (the identity mapping, id), (ii) each element  2 A has an inverse
1 2 A for which 1 ¼ 1  ¼ id, and (iii) composition of elements is
associative, that is, ðgÞ ¼ ðÞg. As we will see, for the DMAP approach
36
Given an assessment of single systems as deterministic or indeterministic, the verdict transfers
immediately to the theory itself: a theory is indeterministic if and only if there is at least one
indeterministic system falling under it.
Defining Determinism 241

it is crucial to identify the right set, A, of isomorphisms; the verdict as to a


theory’s determinism depends sensitively upon the choice of A.
A branching representation of the system’s possible developments has the
form
B ¼ hðBi Þi2I ; Ai; Bi ¼ hBi ; <i ; fi i;

with I some index set. The Bi are the faithful branching models of the system’s

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


development, in which hBi ; <i i is a branching (tree-like) partial ordering (see
Section 4.1) and fi : Bi ° S is a specification of system states for moments
m 2 Bi . The set A is a group of isomorphisms between branching models. If
such models Bk and Bl are connected by some  2 A, this means that they
represent the same physics. As in the DMAP case, such  2 A thus specifies
both an order-preserving bijection between the sets of moments Bk and Bl, and
a mapping of physically equivalent states. Observe, however, that in contrast
to the DMAP case,  2 A relates (typically) non-linear models. Derivatively,
 specifies isomorphisms between linear realizations as well, which can be
used to check a model’s faithfulness (see below). In this sense, two
realizations hh1 ; <jh1 ; S; fjh1 i and hh2 ; <jh2 ; S; fjh2 i that belong to branching
models B1 ¼ hB1 ; <1 ; f1 i and B2 ¼ hB2 ; <2 ; f2 i, respectively, are isomorphic
if, for some  2 A,
jh1 ðhh1 ; <jh1 ; S; fjh1 iÞ ¼ hh2 ; <jh2 ; S; fjh2 i;

where hi are maximal chains in hBi ; <i i. We will say that an isomorphism, h,
restricted to a maximal chain, h, of some branching model is a linearization of
. There is a certain subtlety concerning the issue of where the linearizations
come from. Clearly, the set A contains automorphisms among branching
models (minimally, the identity), and these give rise to linearizations (one
takes B1 ¼ B2 in the formula above.) But the linearizations can also be
derived from isomorphisms reaching across different branching models
(take B1 6¼ B2 , above). (These linear mapping are partial automorphisms
within a branching model, but they are not derived from an automorphism,
but from an isomorphism between different models.) Faithfulness of a branch-
ing model is assessed by both sorts of isomorphisms between linear realiza-
tions. A branching model, Bi , will be declared unfaithful if it has two maximal
chains, h1 and h2, such that jh1 ðhh1 ; <jh1 ; S; fjh1 iÞ ¼ hh2 ; <jh2 ; S; fjh2 i for some
 2 A. Observe that this isomorphism-based assessment uses a weaker notion
of faithfulness, which does not require a difference immediately after branch-
ing—it only requires absence of total isomorphism of linear realizations (com-
pare with Section 4.2, above). This weaker constraint will make it easier to
meet the demand of deriving a branching representation from a mapping
representation (see below).
242 Thomas Müller and Tomasz Placek

Given the above data structures, our definitions of indeterminism (and


thereby, of determinism as indeterminism’s negation) take the forms outlined
in what follows.

5.2.1 DBRN
A system with faithful branching representation B is indeterministic if and

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


only if there is some Bi ; i 2 I, for which there are m; m0 2 Mi ; m 6¼ m0 , such
that neither m<i m0 nor m0 <i m. (That is, indeterminism corresponds to there
being a non-linear, branching structure among the Bi .)
Note that the set of isomorphisms, A, plays no direct role in the assessment
of a theory’s determinism according to this recipe. We did, however, have to
assume that the branching representation was faithful, and as discussed above,
linearizations of the isomorphisms in A provide a security check for
faithfulness.

5.2.2 DMAP
In terms of the DMAP approach, the core idea of determinism translates into
the thought that agreement of two realizations up to some time implies their
total agreement.37 Thus, indeterminism means that there are two realizations
that agree up to some time, but disagree later on. Here, ‘agreement’—both
with respect to states and with respect to times—has to be spelled out in terms
of isomorphisms. In line with (Butterfield [2005]), the definition is as follows:
A system with mapping representation M is indeterministic if and only if there
are realizations Mk ¼ hTk ; <k ; fk i; Ml ¼ hTl ; <l ; fl i; k; l 2 J, for which there
is some t0 2 Tk and some  2 A such that
. for all t  t0 ; fk ðtÞ ¼ ðfl ÞðtÞ, that is, the states on an initial segment can
be identified, but
. there is no  2 A mapping Mk wholly onto Ml , that is, no isomorphism
 2 A for which fk ¼ fl .38
It should be clear from the form of the definition that the choice of A matters
greatly. Minimally, the set of isomorphisms has to contain the identity, but it is
a difficult matter to decide which other mappings are to be included for a given
system. The verdict about determinism can depend on that choice.

37
This idea is somewhat more general than the idea that agreement at some time, or in some small
region around some time, should imply global agreement. Our choice makes the DMAP/
DBRN comparison somewhat more transparent.
38
Our notation, fl or fl , indicates that an isomorphism  2 A maps function fl to function fl ,
thereby taking care of two things in accordance with our discussion above: mapping of times
and mapping of states.
Defining Determinism 243

5.2.3 Comparing DMAP and DBRN


We now move to our main task, which is to establish whether the verdicts as to
a system’s determinism delivered by DMAP and by DBRN agree or not.
Technically, we will tackle these questions by describing how to derive a
mapping representation from a branching representation and vice versa,
and then checking the verdicts as to determinism.

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


5.2.4 DBRN ! DMAP
Starting with a given branching model B ¼ hðBi Þi2I ; Ai, we derive the corre-
sponding mapping representation in three steps. For each Bi ; i 2 I, that is, for
each individual branching model, we extract the linear realizations, lump these
together, and derive the appropriate set of isomorphisms between realizations
from the given A. In more detail, for i 2 I, the branching model Bi ¼ hBi ; <i
; fi i has histories hki ; k 2 Ji , where Ji is an index set enumerating the histories in
branching model Bi . We individuate these histories as a set of realizations by
restriction:

Ci ¼ fhhki ; <i jhk ; fi jhk ijk 2 Ji g:


i i

We construct the set AM of isomorphisms by restricting the isomorphisms


 2 A to linear realizations:
AM ¼ f0 j0 a linearization of some  2 Ag:

Finally, we collect all the linear realizations in one set [i2I Ci , arriving at the
mapping structure
M ¼ h[i2I Ci ; AM i:

Let us now consider how the verdict of determinism or indeterminism for


the branching structure B fares with respect to the derived mapping
structure M.
If B is indeterministic, then M will be diagnosed as indeterministic as
well. This can be seen as follows: Indeterminism of B means that there is
some Bi containing histories hki and hli that are not globally isomorphic
(by faithfulness). These two histories reappear, by construction, as realiza-
tions Mk ¼ hTk ; <k ; fk i and Ml ¼ hTl ; <l ; fl i in M. We can show that these
realizations provide a witness of indeterminism in the mapping sense. As (by
the definition of a branching model) hki \ hli 6¼ ;, there is some t0 2 Tk such
that fk ðt0 Þ ¼ fl ðt0 Þ for all t0  t0 . So, the realizations Mk and Ml have an
isomorphic initial segment (using the identity as isomorphism, which belongs
to AM by construction). But these realizations are not globally isomorphic,
244 Thomas Müller and Tomasz Placek

since that would contradict the faithfulness of the original branching repre-
sentation B.
If B is deterministic, the verdict as to the derived M need not coincide,
depending on how much information about isomorphisms is given through
A. As stated above, for branching structures the set A plays a double role.
On the one hand, A can provide information about the global fact that
two different branching models picture the same physical situation, because

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


(for instance) they correspond to a different choice of gauge. On the other
hand, A gives rise to a set of linearized isomorphisms. This latter set provides a
local criterion for faithfulness of an individual branching model Bi from B: no
two histories within one such branching model may be globally isomorphic.
(Note that any two of them are, by overlap of histories, isomorphic on an
initial segment, with the isomorphism provided by the identity.)
A deterministic B means that all branching models Bi are in fact linear, that
is, contain just a single history. Thus, each Bi already has the mathematical
structure of a realization Mi . Following exactly the same procedure as in the
indeterministic case described above, we arrive at the mapping structure
M ¼ hðBi Þi2I ; AM i:

Whether this M is judged to be deterministic or indeterministic now depends on


whether AM gives rise to partial but not global isomorphisms. To illustrate how
the verdicts could diverge, consider the case of electromagnetism described in
Section 3, above. A branching representation for a system falling under that
theory will contain only linear models. The different models correspond to
different initial conditions, or to different choices of gauge, or both. The verdict
on this representation will be determinism—after all, all branching models are
linear, there is no case of branching. This holds even if the gauge transforma-
tions are excluded from the set of isomorphisms (for instance, if A contains just
the identity). In that case, the derived mapping representation will, however, be
judged to be indeterministic; some realizations will agree initially, but diverge
later, due to a difference in gauge. The set AM will be too small to capture the
fact that these realizations picture the same physics by different mathematical
means. This is exactly the dialectics of diagnosing gauge freedom via spurious
indeterminism described in Section 3, above. So we see that in order to have a
reliable verdict of determinism in the mapping representation, care needs to be
taken to correctly identify all the physical isomorphisms for the system in
question. There is no formal procedure for that step.

5.2.5 DMAP ! DBRN


Let us now consider how the verdict of determinism or indeterminism for a
mapping structure M fares with respect to a derived branching structure B. It
Defining Determinism 245

turns out that the construction is somewhat involved and not unique, but the
verdicts agree.
We will derive a branching representation from a mapping representation
by successively constructing branching models from appropriate sets of reali-
zations. In such a set, any two realizations must be partially isomorphic, but
not globally isomorphic. More formally, let us call a subset fMk jk 2 Kg of the
set of realizations ‘good’ if and only if for any two k; l 2 K; k 6¼ l, there is

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


some t0 2 Tk and some  2 A such that
. for all t0  t0 ; fk ðt0 Þ ¼ ðfl Þðt0 Þ, that is, the states on an initial segment can
be identified, but
. there is no  2 A mapping Mk wholly onto Ml , that is, no isomorphism
 2 A for which fk ¼ fl .
By Zorn’s lemma, there are maximal good sets (note that a singleton set is
good), and the whole set of realizations of the given mapping representation
can be partitioned into good sets: repeatedly take out a maximal good set and
identify the next maximal good set in what remains. Note, however, that this
partitioning is not unique.39
For any resulting good set fMk jk 2 Kg, we can derive a branching model in
the following way40: We pick one realization M0 ¼ hT0 ; <0 ; f0 i, which will
form a reference history in the resulting branching model. By goodness, for
every other realization Mk we can pick an isomorphism k 2 A that identifies
M0 and Mk up to some tk 2 T0 (but not for all times). Call the mapped
realization
M0k ¼ hT 0k ; < 0k ; f 0k i :¼ hk ðTk Þ; k ð<k Þ; k ðfk Þi:

So, for all k 2 K  f0g, we have


f0 ðt0 Þ ¼ f 0k ðt0 Þ for all t0  tk :

Exactly as in the construction for Norton’s dome in Section 5.1.1 above, we


now define a base set and derive the branching model by dividing out an
equivalence relation. To save some ink, we set T 00 :¼ T0 , < 00 :¼ <0 ; f 00 :¼ f0 .
We set
~ :¼
B [ ðT
k2K
0
k
fkgÞ:

For the equivalence relation &, we set


ht; ni&hs; mi

39
This is connected to the fact that the relation of being partially, but not globally, isomorphic is
not transitive, that is, it is not an equivalence relation.
40
The following construction is not unique, and it is not guaranteed to deliver branching models
that are intuitively satisfying. The construction does, however, fulfil all formal requirements.
246 Thomas Müller and Tomasz Placek

if and only if t ¼ s and for all t0  t we have

f 0n ðt0 Þ ¼ f 0m ðt0 Þ:

Next, our set of moments (the base set for the partial branching order) is
~
B :¼ B=&;

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


and we define the state-assignment function f on B to be
f ð½ht; niÞ ¼ f 0n ðtÞ:

Note that the definition of & guarantees well-definedness, meaning that for
hs; mi 2 ½ht; ni, we have f 0n ðtÞ ¼ f 0m ðsÞ. It remains to define the partial ordering
< on the set of moments B:
½ht; ni < ½hs; mi

if and only if t < 0n s and; for all t0  0n t; we have

f 0n ðt0 Þ ¼ f 0m ðt0 Þ:
Pulling things together, adding all these branching models for all the good sets
into which the given set of realizations was partitioned, will give a full branching
representation for the system in question. It remains to specify the set AB of
isomorphisms. For the verdict of determinism or indeterminism, this subtle issue
is, however, not important, so that we can set AB to contain just the identity.
It is easy to see that by the given construction, a verdict as to determinism or
indeterminism of the mapping structure is retained in the branching structure.
In the case of determinism, all good sets are singletons, giving rise to only
linear, deterministic branching models. In the case of indeterminism, there will
be at least one non-trivial good set, giving rise to a faithful branching model
with at least two histories. Such a non-linear structure triggers the verdict of
indeterminism.

5.2.6 Summing up
As we saw, the relations between two representations of determinism, as
offered by DMAP and DBRN, are somewhat intricate. Their verdicts with
respect to a system’s determinism usually agree, but not always. That is, if
DBRN diagnoses a system as indeterministic, DMAP will concur. However, if
DBRN’s verdict is ‘determinism’, a DMAP analysis might disagree. In the
opposite direction verdicts agree, that is, if DMAP deems a system as deter-
ministic/indeterministic, DBRN will come with the same diagnosis. Although
the two approaches agree on the verdict in such cases, they might view the
underlying details differently, as the DMAP representation of determinism (or
Defining Determinism 247

indeterminism) does not translate uniquely into a DBRN representation of


determinism (or indeterminism). This last subtlety and the possible divergence
of verdicts about determinism derives from the different roles the set of iso-
morphisms plays in the two approaches, and from the status of the faithfulness
assumption.
The dialectics here is as follows: Given a physical system in whose determin-
ism or indeterminism we are interested, we need to construct a mathematical

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


representation with respect to which we can study the question of determinism
in a formally precise way—that is the overarching framing of the determinism
issue in philosophy of science. The three approaches under discussion here
differ with respect to their mathematical representation of a system. We have
pointed out that DEQN makes good sense in most, but not all, applications,
quantum mechanics being a notable exception. Here we have considered the
interrelation between the two remaining approaches, DMAP and DBRN. In
our discussion of case studies in Section 5.1, we have seen that the actual
construction of DBRN models typically leads to models whose faithfulness
is guaranteed. In the case of Norton’s dome, the behaviour of the differential
equation secured the necessary difference in physical state (here, position of
the particle); in quantum mechanics, the assignment of different projectors in
different histories did the same. A set of isomorphisms between branching
models, which we have included in our formal description in this section, can
be helpful in providing a more adequate picture, showing that two different
branching models may depict the same physical situation. This, however, does
not affect the verdict as to determinism or indeterminism, which is based on
the ordering structure of the individual branching models whose faithfulness is
assured beforehand. Our discussion of the DMAP approach, on the other
hand, shows that the choice of the set of isomorphisms is crucial for the
assessment of determinism or indeterminism. Trouble can arise if too few
isomorphisms are identified, since then a spurious assessment of indetermin-
ism threatens. There is nothing in the construction of a DMAP representation
of a system that secures the identification of a physically adequate group of
isomorphisms.
Our diagnosis of this state of things is as follows: Branching is the natural
representation of indeterminism—we directly understand a non-linear
branching structure as representing an indeterministic scenario. Typically,
an investigated problem contains information about the grouping of realiza-
tions into sets of alternative mutually possible developments. This informa-
tion comes in a statement of an initial-value problem, or a system’s
symmetries, or even similarities of processes considered. This information,
and the partition of realizations it affords, is lost, or is not being used, in
the mapping-based account. Instead, a lot of mathematical surplus structure
needs to be considered. The worry is that, in the end, getting this surplus
248 Thomas Müller and Tomasz Placek

structure right will only be possible in case we have a different underlying


representation (for example, a branching-based representation) that anchors
our assessment in the first place.

6 Conclusions
Our aim in this article was to elaborate formally the core idea of determinism,

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


according to which a deterministic system has no alternative future develop-
ments, and to apply the resulting framework to theories of physics. That
framework is based on branching theories that are well known in tense
logic; a novelty of this article consists in showing how to construct branching
models for theories of physics. A salient feature of branching models is their
modal thickness: a single model of that sort has resources to represent alter-
native possible evolutions of a system. Technically speaking, a branching
model may contain more than one history (maximal chain), in which case
we call it ‘indeterministic’; otherwise we call it ‘deterministic’. We say that a
theory is indeterministic if it has at least one faithful indeterministic model;
otherwise, the theory is deterministic. This topic of a branching-style analysis
of determinism of theories was discussed in Section 4.
To locate our analysis with respect to extant debates on determinism, in
Section 2 we singled out three styles of thinking about determinism: DEQN
characterizes determinism in terms of solutions to a theory’s defining equa-
tions; DMAP proceeds in terms of mappings between linear temporal realiza-
tions admitted by the theory; and DBRN uses the concept of alternative
possible future continuations. In Section 3, we focused on the dominant
approach to determinism in current philosophy of science, as exemplified in
writings of Earman and Butterfield. We pointed out that both these authors
advocate a position that combines the DMAP and DEQN approaches; this
combination seems, however, to call for further elaboration, as no proof of the
equivalence of (or of other logical relations between) DMAP and DEQN is
known. Another critical issue is the orthodoxy’s appeal to the notion of iso-
morphism, which (we claim) is used rather loosely, since a theory’s models are
not required to be models in the logical sense. Two further troubling issues,
which arise in a similar way for other approaches, are the identification of
times across different realizations, and the possible difference between math-
ematical and physical states.
Having provided a detailed exposition of the DBRN framework in Section
4, in Section 5.1 we compared the three approaches with respect to how they
apply to particular cases. Focusing on Norton’s dome and quantum
mechanics, we found (perhaps surprisingly) that the DMAP analysis is not
used in the literature. In contrast, the formalism of consistent histories for
quantum mechanics is immediately translatable into branching models.
Defining Determinism 249

Branching models also provide a natural representation of Norton’s dome. By


adding some extra structure, these branching models can be transformed into
DMAP models, but no extra value appears to be provided by such a move.
For GR, the core idea of indeterminism becomes problematic in the absence of
a global time function. All three approaches need to be based on the notion of
a local alternative for the future. This is natural for the DEQN approach;
DMAP and DBRN may come to diverging conclusions. This highlights the

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


fact that these approaches are closely related, but not equivalent.
Accordingly, in Section 5.2 we investigated the formal interrelations
between the DMAP and DBRN definitions of determinism. The comparison
highlights that DMAP models need some extra structure as compared to
DBRN models. This extra structure (coded in a set of isomorphisms) reflects
two decisions: which times across realizations should be identified and which
mathematical states represent the same physical state. We described the repre-
sentations of determinism offered by DMAP and DBRN, respectively, and
showed how one representation can be derived from the other, noting that
there is no uniqueness in the construction of a DBRN representation out of a
DMAP representation. We also showed that although in most cases the ver-
dicts of the two representations of determinism agree, a divergence is possible:
a DBRN verdict of determinism might be rejected in the DMAP approach.
These discrepancies are a consequence of the different role isomorphisms
play in the two representations of determinism, and we believe these different
roles cut to the bone of the controversy between the two approaches.
Informally speaking, whether a system is deterministic or not depends on
whether it has a possible development to which there is a true alternative.
For instance, multiple developments from shared initial conditions are true
alternatives, but multiple developments produced by exercising gauge free-
dom, or freedom of coordinate choice, are not true alternatives.
Accordingly, any good analysis of determinism requires a way of partitioning
a set of realizations into subsets of ‘truly alternative’ realizations. In DBRN,
this is achieved by lumping subsets of realizations into tree-like branching
models, whereas in DMAP a similar effect is simulated via a set of isomorph-
isms. Now, a typical question of determinism contains information about
partitioning a set of realizations into subsets of truly alternative realizations.
This kind of information is directly used in the construction of a DBRN
model; in contrast, it is not directly used in the mapping-based account.
Instead, a lot of mathematical surplus structure is postulated to derive a
partition of realizations mimicking that of the DBRN approach. Whether
that surplus structure is of the right sort appears to be decided by consulting
the information utilized in the construction of branching models. This, we
believe, tells strongly in favour of the greater simplicity and conceptual pri-
macy of the DBRN approach relative to the DMAP approach.
250 Thomas Müller and Tomasz Placek

The final message of this article is that branching is a natural representation


of a theory’s indeterminism, which moreover is rendered mathematically rig-
orous by the definitions we proposed. It is naturally used in the particular
cases we considered. Branching represents exactly the kind of structure that is
needed to assess a theory’s determinism or indeterminism.

Acknowledgments

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


Müller’s research was funded by the European Research Council under the
European Community’s Seventh Framework Programme (FP7/2007-2013)/
ERC Grant agreement no. 263227, and by the Dutch Organization for
Scientific Research, grant no. NWO VIDI 276-20-013. Placek’s research was
funded by research grant Mistrz 2011, Foundation for Polish Science, contract
no. 5/2011.

Thomas Müller
Fachbereich Philosophie
Universität Konstanz
Konstanz, Germany
Thomas.Mueller@uni-konstanz.de

Tomasz Placek
Department of Philosophy
Jagiellonian University
Krakow, Poland
Tomasz.Placek@uj.edu.pl

References
Arnol’d, V. I. [1992]: Ordinary Differential Equations, Berlin: Springer.
Belnap, N. [1992]: ‘Branching Space-Time’, Synthese, 92, pp. 385–434.
Belnap, N. [2012]: ‘Newtonian Determinism to Branching Space-Times Indeterminism
in Two Moves’, Synthese, 188, pp. 5–21.
Belnap, N., Perloff, M. and Xu, M. [2001]: Facing the Future, Oxford: Oxford
University Press.
Bishop, R. [2006]: ‘Determinism and Indeterminism’, in D. Borchert (ed.), Encyclopedia
of Philosophy, Volume 3, Farmington Hills: Thompson Gale, pp. 29–35.
Butterfield, J. [1989]: ‘The Hole Truth’, British Journal for the Philosophy of Science, 40,
pp. 1–28.
Butterfield, J. [2005]: ‘Determinism and Indeterminism’, in E. Craig (ed.), Routledge
Encyclopedia of Philosophy, Volume 3, London: Routledge.
Choquet-Bruhat, Y. and Geroch, R. [1969]: ‘Global Aspects of the Cauchy Problem in
General Relativity’, Communications in Mathematical Physics, 14, pp. 329–35.
Chruściel, P. and Isenberg, J. [1993]: ‘Nonisometric Vacuum Extensions of Vacuum
Maximal Globally Hyperbolic Spacetimes’, Physical Review D, 48, pp. 1616–28.
Defining Determinism 251
Chruściel, P., Isenberg, J. and Moncrief, V. [1990]: ‘Strong Cosmic Censorship in
Polarized Gowdy Spacetimes’, Classical and Quantum Gravity, 7, pp. 1671–80.
Earman, J. [1986]: A Primer on Determinism, Reidel: Dordrecht.
Earman, J. [2006]: ‘Aspects of Determinism in Modern Physics’, in J. N. Butterfield and
J. Earman (eds), Philosophy of Physics, Amsterdam: Elsevier, pp. 1369–434.
Earman, J. [2008]: ‘Pruning Some Branches from Branching Spacetimes’, in D. Dieks
(ed.), The Ontology of Spacetime II, Amsterdam: Elsevier, pp. 187–206.
Earman, J. and Norton, J. [1987]: ‘What Price Spacetime Substantivalism? The Hole

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021


Story’, British Journal for the Philosophy of Science, 38, pp. 515–25.
Gell-Mann, M. and Hartle, J. [1993]: ‘Classical Equations for Quantum Systems’,
Physical Review D, 47, p. 3345.
Griffiths, R. B. [2002]: Consistent Quantum Theory, Cambridge: Cambridge University
Press.
Halvorson, H. [2012]: ‘What Scientific Theories Could Not Be’, Philosophy of Science,
79, pp. 183–206.
Hodges, W. [1993]: Model Theory, Cambridge: Cambridge University Press.
Hoefer, C. [2010]: ‘Causal Determinism’, in E. N. Zalta (ed.), The Stanford Encyclopedia
of Philosophy, <http://plato.stanford.edu/archives/spr2010/entries/determinism-
causal/>
Horty, J. F. [2001]: Agency and Deontic Logic, Oxford: Oxford University Press.
Laplace, P. [1951]: A Philosophical Essay on Probabilities, New York: Dover
Publications.
Lewis, D. K. [1986]: On the Plurality of Worlds, Oxford: Blackwell.
Montague, R. [1974]: ‘Deterministic Theories’, in his Formal Philosophy: Selected
Papers of Richard Montague, New Haven, CT: Yale University Press.
Müller, T. [2007]: ‘Branch Dependence in the “Consistent Histories” Approach to
Quantum Mechanics’, Foundations of Physics, 37, pp. 253–76.
Müller, T. [2012]: ‘Branching in the Landscape of Possibilities’, Synthese, 188, pp. 41–65.
Müller, T. [2013]: ‘A Generalized Manifold Topology for Branching Space-Times’,
Philosophy of Science, 80, pp. 1089–100.
Munkres, J. R. [2000]: Topology, New York: Prentice Hall.
Norton, J. D. [2008]: ‘The Dome: An Unexpectedly Simple Failure of Determinism’,
Philosophy of Science, 75, pp. 786–99.
Peres, A. [2000]: ‘Classical Interventions in Quantum Systems, I: The Measuring
Process’, Physical Review A, 61, p. 22116.
Placek, T. [2014]: ‘Branching for General Relativists’, in T. Müller (ed.), Nuel Belnap on
Indeterminism and Free Action, Berlin: Springer, pp. 191–221.
Placek, T. and Belnap, N. [2012]: ‘Indeterminism Is a Modal Notion: Branching Space-
Times and Earman’s Pruning’, Synthese, 187, pp. 441–69.
Placek, T., Belnap, N. and Kishida, K. [2014]: ‘On Topological Aspects of
Indeterminism’, Erkenntnis, 79, pp. 403–36.
Prior, A. [1967]: Past, Present, and Future, Oxford: Oxford University Press.
Ringström, H. [2009]: The Cauchy Problem in General Relativity, Zürich: European
Mathematical Society.
252 Thomas Müller and Tomasz Placek

Thomason, R. [1970]: ‘Indeterminist Time and Truth-Value Gaps’, Theoria, 36, pp.
264–81.
van Kampen, N. G. [2007]: Stochastic Processes in Physics and Chemistry, Amsterdam:
Elsevier.
Wald, R. M. [1984]: General Relativity, Chicago, IL: University of Chicago Press.
Wilson, M. [1989]: ‘Critical Notice: John Earman’s A Primer on Determinism’,
Philosophy of Science, 56, pp. 502–32.

Downloaded from https://academic.oup.com/bjps/article/69/1/215/2669643 by guest on 23 January 2021

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy