scientific explanation: towards a neo-deductivist account

166
Scientific Explanation: Towards a Neo-Deductivist Account By Martin King A Thesis presented to The University of Guelph In partial fulfilment of requirements for the degree of Doctor of Philosophy in Philosophy Guelph, Ontario, Canada © Martin King, May, 2016

Upload: others

Post on 29-Nov-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Scientific Explanation: Towards a Neo-Deductivist Account

By

Martin King

A Thesis

presented to

The University of Guelph

In partial fulfilment of requirements

for the degree of

Doctor of Philosophy

in

Philosophy

Guelph, Ontario, Canada

© Martin King, May, 2016

ABSTRACT

SCIENTIFIC EXPLANATION: TOWARDS A NEO-DEDUCTIVIST ACCOUNT

Martin King Advisor:

University of Guelph Andrew Wayne

This thesis is an investigation of philosophical accounts of scientific explanation. It is centered

around the question: how do scientific models explain? I argue that the model-based deductivist

account I propose in Chapter 5 is a viable and promising candidate for a successful account of

scientific explanation. This thesis shows that other structural, causal, and deductivist accounts

face significant challenges or are not reflective of the practice of scientific explanation.

In the first chapter, I introduce the concept of scientific explanation and review the goals

of a philosophical account of scientific explanation. In the second chapter, I explore the role of

idealized models in scientific explanation. In the remainder of Chapter 2, and in Chapters 3 and

4, I critically review literature on three of the main approaches to scientific explanation:

structural, causal, and deductivist respectively. Some of the main results of these investigations

are that: i) an account of explanation should include a broader range of models than a strictly

causal account, to more accurately reflect the explanatory practices of science; ii) a variety of

models can be explanatory in a given system, but that causal interpretations of these models can

be problematic; iii) and that highly-idealized models can be explanatory but that neither

structural nor causal accounts can fully capture why this is so.

Taking these results into account, in the final chapter, I present the most significant

contribution of the dissertation, which is the integrated model account of explanation. The

account relaxes local constraints on allowable features in explanatory models and instead

introduces a global constraint on the model’s relation to theory. It is a model-based deductivist

account that stipulates four criteria for explanation: the explanandum is deductively entailed by

the explanans; the statements in the explanans are true of a scientific model; the model shows on

what the explanandum depends; and the model is integrated with a global theory of science.

iii

To my father, who would have been so proud.

iv

ACKNOWLEDGMENTS

I express my sincerest thanks to my advisor, Dr. Andrew Wayne, for his continuous support and

encouragement over the years. Without his knowledge, insight, and enthusiastic dedication to his

students, this project would never have been completed. In fact, without him, I would have never

even undertaken this project. I owe him a great debt of gratitude for the effect his tutelage has

had on me, my research, and my future successes.

I would also like to thank the members of my advisory committee, Dr. Stefan Linquist and Dr.

Jessica Wilson, for their challenging comments and questions, as well as for their help and

guidance in the development of my project. I am also grateful for the expertise of Dr. Juha

Saatsi, whose input and careful reading of my work made for great discussion and will be

beneficial for my future work.

I am indebted to my fellow graduate student colleagues for the late-night philosophical

discussions over whisky and beer, and for their resolve and good spirits that made the last five

years of being overworked and underpaid, not only tolerable, but enjoyable.

Enfin et surtout, un grand merci à ma mère pour tous la patience et l’encouragement pendant ma

longue carrière d’étudient, and to David, Robin, and Kayla for the profound effect they have on

every aspect of my life.

v

Table of Contents

Acknowledgments ..................................................................................................................... iv

Chapter 1. The Nature of Scientific Explanation ............................................................. 1

Introduction .................................................................................................................. 1

Some Kinds of Explanation ......................................................................................... 4

1.2.1. Non-Explanations ........................................................................................... 7

1.2.2. Token and Type Explanations ........................................................................ 7

Explanation and Understanding ................................................................................... 8

1.3.1. Pragmatics, Explananda, and Why-Questions ............................................. 11

Goals of an Account of Explanation .......................................................................... 15

1.4.1. Descriptivism and Normativity .................................................................... 15

1.4.2. Threshold and Depth .................................................................................... 17

1.4.3. Scope ............................................................................................................ 18

1.4.4. Accurate Representation .............................................................................. 19

Conclusion ................................................................................................................. 20

Chapter 2. Idealization and Structural Accounts of Explanation ................................... 21

Introduction ................................................................................................................ 21

Models and Idealization ............................................................................................. 22

2.2.1. What is idealization? .................................................................................... 23

2.2.2. Laws and Idealized Models .......................................................................... 26

2.2.3. Idealization and Explanation ........................................................................ 31

Modelling and Explanation ........................................................................................ 33

2.3.1. Desiderata ..................................................................................................... 33

Idealization in Physics ............................................................................................... 37

vi

2.4.1. Structural Model Explanations ..................................................................... 38

2.4.2. Two Approaches for Assessing Structure .................................................... 47

2.4.3. E3: The Justificatory Step ............................................................................ 54

2.4.4. Heuristics and Explanations ......................................................................... 58

Conclusion and Discussion ........................................................................................ 60

Chapter 3. Causal Accounts of Explanation................................................................... 62

Introduction ................................................................................................................ 62

Scientific Pluralism .................................................................................................... 63

3.2.1. Age-Polytheism in Eusocial Insect Colonies ............................................... 65

3.2.2. Three Problems with the Integrative Pluralism Approach ........................... 68

Causal Interventionism .............................................................................................. 70

3.3.1. Invariance and Intervention .......................................................................... 72

3.3.2. Circularity ..................................................................................................... 74

3.3.3. Causal Realism ............................................................................................. 75

Emergence and Reductionism .................................................................................... 79

3.4.1. Physicalism and Supervenience ................................................................... 80

3.4.2. Exclusion Arguments ................................................................................... 81

3.4.3. Intervention and Emergent Causation .......................................................... 87

Kairetic Explanation .................................................................................................. 91

3.5.1. Abstracting and Optimizing ......................................................................... 92

3.5.2. Idealization and Causal Realism .................................................................. 93

3.5.3. Concerns about the Kairetic Account ........................................................... 95

Conclusion and Additional Concerns ........................................................................ 96

Chapter 4. Deductivist Explanation ............................................................................. 101

Introduction .............................................................................................................. 101

The D-N Account ..................................................................................................... 102

4.2.1. Is it Necessary? ........................................................................................... 104

4.2.2. Is it Sufficient? ........................................................................................... 107

vii

Unificationism .......................................................................................................... 109

4.3.1. Unificationist Solutions to D-N Problems ................................................. 112

4.3.2. Challenges to Unificationism ..................................................................... 115

Conclusion and the Current State of Deductivism ................................................... 116

Chapter 5. Model-based Deductivism .......................................................................... 119

Introduction .............................................................................................................. 119

A Model-based Deductivist Account ....................................................................... 120

What is a Model? ..................................................................................................... 123

5.3.1. A Simple Model of the Fixed-length Pendulum ........................................ 124

Counterfactuals ........................................................................................................ 126

5.4.1. Same-object Counterfactuals ...................................................................... 127

5.4.2. Truth Conditions for Counterfactuals ........................................................ 128

The Simple Pendulum Revisited .............................................................................. 129

A Global Constraint on Explanation ........................................................................ 131

5.6.1. Some Aspects of Theoretical Integration ................................................... 134

5.6.2. Prediction without Explanation .................................................................. 136

The Integrated Model Account of Explanation ........................................................ 139

Empiricism, Emergence, and Reduction .................................................................. 142

Conclusion and Limitations of the Account ............................................................ 144

Conclusion 148

References 152

1

Chapter 1. The Nature of Scientific Explanation

Introduction

Scientific explanation has been a central topic in the philosophy of science since Carl Hempel

and Paul Oppenheim published “Studies in the Logic of Explanation” in 1948. Explanation

became a large part of 20th century philosophy in part because it is intimately related to many

different issues and important questions, including “what is a law of nature?”, “what makes

something scientific?”, “what is the nature of causation?”, and many more. In the decades

following, many philosophers of science either promoted or criticized various aspects of the

Deductive-Nomological account of explanation that Hempel and Oppenheim originally proposed.

For Hempel and Oppenheim, an explanation is a deductive argument featuring two parts,

where that which is to be explained, the explanandum, is deductively entailed by a set of

sentences that do the explaining, the explanans. The explanatory information for Hempel and

Oppenheim comes from deriving a desired explanandum phenomenon or pattern from a set of

true sentences containing a law, or laws, of nature. The subsumption of a phenomenon under a

law of nature allows us to expect, predict, and explain its occurrence. This formulation of

explanation allows one to answer explanation-seeking why-questions, like “why is the sky

blue?” and also questions about regularities themselves, questions such as “why do Kepler’s and

Galileo’s laws hold?” In the same way, it is the derivational structure of the explanation that

allows the explanandum to be a law, that is derived as a special case of another more general

law, in this case Newton’s law of gravitation.

Many of the philosophers since Hempel and Oppenheim have focused on causal

strategies for capturing explanation (Anscombe, 1969; Lewis, 1973; Salmon, 1984; Dowe, 2000;

Woodward, 2003). These causal accounts all employ quite different strategies for identifying

genuine causal dependency relations, but they share the idea that explaining a phenomenon is

achieved by giving its causes. The most important task for a causal account of explanation then

is to conceptualize causation in such a way as to align causal judgments with our explanatory

judgments, which may otherwise be at odds. For instance, certain models that are not generally

2

thought of as causal may be thought of as explanatory. Causal approaches are popular in the

literature and considered to be very promising, though there are some issues of contention

surrounding the notion of causation. In Chapter 3, I present three causal accounts in some detail.

I first examine complexity, causal pluralism, and Sandra Mitchell’s account of explanation. The

pluralist solution to issues of complexity is to identify causes in models at various levels of

investigation, including high-level idealized models. I point out some shortcomings of taking this

approach and argue that her conclusions about the needs for causal pluralism are unwarranted.

The remaining accounts in Chapter 3 belong to Michael Strevens and Jim Woodward

respectively. The problems for Woodward’s account surround the fact that his manipulationist

criteria identify high-level causes and low-level causes in the same system, which invites

concerns about overdetermination and downward causation. I explore possible solutions via

some of the literature on non-reductive physicalism, which I review in 3.4. In the end, I argue

that Woodward’s position is properly emergentist and cannot benefit from defenses of non-

reductive physicalism. After this, I present Michael Strevens’ view of depth and his kairetic

account of explanation. Strevens attempts to show how a physicalist metaphysics can allow for

explanations at non-fundamental levels by preferring a degree of generality to total accuracy. I

argue that this approach is not subject to the same problems as Woodward’s, but is too

unrealistic to reflect explanatory practices or to be implemented.

Before I turn to causal explanation however, I take up structural accounts in the second

chapter, because it is in this context that I can best examine the role of idealization and modelling

in explanation. Proponents of structural explanation have taken approaches similar to causal

accounts but focused on the higher-level structural relations rather than the causal relations

(Worrall, 1989; McArthur, 2003; Bokulich, 2008). This allows these structural approaches to

identify a different class of model as explanatory, whose structure, it is argued, is the most

important factor in explaining the target system’s behaviour. The main benefit of structural

accounts is that they are able to count as explanatory models that do not accurately represent the

real causes of a system, and thus expand the scope of available explanations. I begin Chapter 2

by going over modelling and idealization and their roles in explanation. Idealization is a major

concern for explanation, and especially for causal and other representational accounts. The

immediate worry is that if an account identifies explanation by real causal dependency relations,

then it is not clear how to accommodate false assumptions, non-existing entities, abstractions,

3

and otherwise non-representing models. Of course, there are various proposed solutions to this,

and many places to draw the line for what counts as an explanation. In Chapter 2, I explore one

potential solution, which is to focus on structure. Alisa Bokulich proposes an account that

attempts to emphasize the explanatory role that idealizations play in explanation. I argue that the

structuralist account she proposes favours explanations at the fundamental level and thus is not

capable of capturing the way that non-representing models explain.

In the fourth chapter, I return in detail to the deductivist approaches of Hempel and

Oppenheim and Philip Kitcher (Kitcher, 1981, 1989). I present the traditional Deductive-

Nomological account of scientific explanation and some responses to it, including some work on

statistical explanation. I review some challenges to this account, and then present Kitcher’s

unificationism and its proposed solutions in 4.3. Ultimately Kitcher’s defense is not satisfactory,

but the problems for deductivism are not essential or insurmountable. By availing myself of the

resources and outlining some remaining problems for the deductivist approach, I can set the

stage to offer an account that is aimed at improving the current state of deductivist explanation.

These results of these investigations serve to inform the features of the neo-deductivist

account I propose in Chapter 5. The account takes explanations to be arguments that derive the

explanandum from statements about certain kinds of scientific model. The account makes use of

many insights from other accounts, including: the importance of prominently featuring

idealization and counterfactual information; allowing for non-representing models; having a

high-threshold for explanation; requiring the deductive derivation of the explanandum from the

explanans; and much else. The account is intended to circumvent some of the concerns raised

about structural and causal accounts, and to be more relevant to explanatory practice than the D-

N account and more tractable than Kitcher’s unificationism.

The remaining sections of this chapter will deal with the nature of scientific explanation

in order to prepare the ground for later discussions. I begin this chapter by first analyzing various

kinds of explanation and focusing the discussion in the following chapters on the right kinds of

explanation. Then, I examine the role of understanding in explanation, as well as the role of

pragmatics of question-asking and the activity of explaining. The last section of this chapter goes

over some possible goals of an account of explanation, in order to set the standards for what is

and is not to be achieved, what is inside or outside of its scope, and what is or is not desirable for

an account of explanation.

4

Some Kinds of Explanation

It is important to keep in mind that the notion of scientific explanation naturally contrasts itself

with three other classes: those that are scientific, but not explanations; explanations that are not

scientific, and those that are neither (see Table 1-1).

Table 1-1

Scientific & an Explanation Scientific & Not an Explanation

Not Scientific & an Explanation Not Scientific & Not an Explanation

The present work is only intended to focus on the top left cell. In order to better elucidate this

distinction, the following section will discuss some of the various kinds of explanation and make

clear what the appropriate kinds are.

Explanation is a very broad term. There are at least a few distinct kinds of explanation,

most of which are not scientific. Sometimes, what counts as an explanation can be little more

than providing a reasonable reason – a motivation to think that a statement is an answer to a why

question. For instance, in certain circumstances, the question “why is Jones not home?’ can be

answered satisfactorily by the statement “he was out of milk”. These kinds of explanations are

what we might call common-sense explanations, because the statement(s) alone do not

necessitate the explanandum, and some sort of common-sense “law” needs to be used to join the

two and motivate the explanation: milk is a staple, and you should replace it when it runs out. If

this “law” were not common sense, the answer would effectively cease to be explanatory.

Explanations such as this merely provide good reason for thinking that the answer given is the

true explanation. The success of these explanations is due in part to the commonality of these

kinds of inferences (say, from being out of milk to going to the store). These kinds of

explanation are doubtless very common, but they are not scientific explanations.

One reason that they are not scientific is that they often involve human motivations or

intentions. When it comes to these kinds of explanation, it is unreasonable to expect there to be a

single explanation or cause for the fact. If someone asks Mary why she went to university, there

are likely to be many reasons which could all serve as good explanations. Her answer to the

question might depend on who is asking, what she has recently been thinking about, or what she

considers most important in the decision at the time she is asked. What counts as a good

5

explanation here is almost entirely contextually dependent. I hope to avoid discussing

explanations that make use of motivations.

Another reason is that these common-sense answers do not necessitate the explanandum.

Because of this, one is never sure if they are correct explanations – there is an essential

ambiguity. Imagine a case where even if it is true that Jones is out of milk, the real reason Jones

is not home is because he is meeting a friend at the pub. The fact that he is out of milk is true,

and gives reason to expect his not being at home, but it is not in fact the correct explanation for

his not being at home. We can further complicate this scenario by adding instead that Jones did

go out to grab milk, but stopped at the pub on the way, making him late. In one sense, he is not

home because he went out to get milk, but in another sense, he is not home because he stopped at

the pub. In this new scenario, there are multiple factors that make it unclear that the answer

“because he went to get milk” is a good explanation. Many such examples have been raised in

the literature on explanation, but I think it is important to bracket off these explanations to focus

on those that are properly in the domain of science.

Other kinds of explanation have a closer relation to scientific explanation, and some

argue are a genuine part of it. There are what are known as teleological arguments or functional

arguments, which cite ends as explanations. Such an explanation would maintain, for instance,

that a trait’s function explains its existence or its continuation in a population. Whether these

arguments are distinct and ineliminable from other explanations in biology and other sciences is

a matter of some contention (Brandon, 1981; Sober, 1984; Reeve & Sherman, 1993; Godfrey-

Smith, 1994). These explanations fit well with human behaviour which has intentions that are

clearly future-oriented, but they are also popular in evolutionary biology, especially before

Darwin and often cited as evidence of intelligent design. Teleological explanations will not be

directly dealt with in this thesis.

In what follows, the kinds of explanation that will be focused on are scientific

explanations and not common-sense, or everyday, explanations. To explain something

scientifically is more than to make an everyday explanation about something that has scientific

content. This is something that is often passed over, and I think is something that is responsible

for many of the problematic counterexamples in the explanation literature. The connection from

explanans to explanandum must be more than merely an accepted inference. What I take a proper

scientific explanation to do is to demonstrate that a well-articulated explanandum is to be

6

expected on the basis of a certain kinds of scientific model. A full elaboration of this will have to

wait until further chapters.

Following in the deductivist tradition, I take explanations to be arguments. Other

accounts have taken explanations to be causal stories, explanation-acts, inferences, models, or

whatever. Treating explanations as arguments has some benefits, such as that one can formulate

arguments about models or featuring causal relations. For instance, on this view, to say that C

explains E because C causes E, is really to say that there is an explanatory derivation of E

featuring C and ‘C causes E’. A simple causal explanation can be recast in the form of an

argument. This deductivist approach is open to what are considered causal explanations, but is

not limited by this. The breadth of allowable varieties of explanation is a distinct benefit of the

deductivist approach that will be elaborated on in Chapters 4 and 5. Thus, on this picture, causal

explanations are just one type of explanation among many. It is worth noting that because

explanations are arguments, models themselves do not do the explaining. Rather, to say that a

model is explanatory is to say that it is capable of supporting explanatory derivations. The

relations between models and explanations and the role of models in explanations will be

explored in Chapters 2 and 5.

The purpose of distinguishing these kinds of explanation is not to deny that they are

explanations or to claim that they are not as explanatory as scientific explanations, but rather the

limit the scope of the present investigation to those explanations that concern scientific models.

These models are abstract objects that are constructed for the purposes of explaining some

behaviour, pattern, or property of a target system. They often feature quantitative generalizations

that serve to make reliable predictions about the system. It has been increasingly recognized that

accounts of explanation ought to focus on models, rather than merely on arguments, or diagrams,

or speech acts, though this is not uncontentious. Even among model-based accounts, others have

argued that a scientific explanation consists in identifying the relevant causes of a phenomenon,

or in revealing the structural properties of a system that are responsible for the manifestation of

the phenomenon. I will review some of these other accounts of explanation in the following

chapters. Most of the accounts that will be discussed in detail, including the account proposed in

Chapter 5, all take models to be important sources of information for scientific explanation.

Reasons why this is taken to be a central aspect of scientific explanation will be discussed in

even more detail in Chapter 2.

7

1.2.1. Non-Explanations

Conversely, there are also things that fit into the category of being scientific but not explanatory.

And so there is more to science than explaining, which even though it is perhaps obvious, bears

mentioning. Of course, it would beg the question to start an analysis of explanation by deciding

which cases are explanatory beforehand. While it is contentious whether certain models are

explanatory, there are some practices that do not aim to offer explanations. For instance, there

are aspects of science that primarily involve the collection and organization of data. Some of this

has been derogatively called “stamp-collecting”, but it can be an important aspect of scientific

research. Some research that falls into this might include the collection of astronomical

observations, the cataloguing of species, or the human genome project, which is aimed in part at

determining the sequence of base pairs of human DNA, but does not explain the emergence or

continuation of human traits or characteristics. This kind of information could potentially be used

to support explanation, but that is not necessarily the aim of these practices.

There is a different kind of scientific practice, and the last that I will mention, which aims

to create predictive models, like the models at work in meteorology, which make use of very

complicated algorithms that can serve, so legend goes, to predict the weather (Parker, 2011).

These are also known as data models, or phenomenological models. I maintain that to say that

these models of numerical weather prediction actually explain why, for instance, a storm

occurred, requires more than to simply say that it has predicted or forecasted it. The account I

propose in Chapter 5 aims to correctly identify why these predictive models are not explanatory,

unlike other accounts of explanation that have taken explanation and prediction to be two sides

of the same coin. The difficulty in identifying these non-explanatory models is that modelling is

often a very complicated process, and so a model merely built up from data and a model that is

idealized but genuinely explanatory can be difficult to distinguish.

1.2.2. Token and Type Explanations

The token-type distinction is an important one in many contexts, including in terms of scientific

explanation. Types refer to classes and tokens to the instantiations of those classes. A token

explanation is an explanation of a particular occurrence or phenomenon, while a type explanation

is an explanation of why a class of phenomena occurs. These explanations are often answers to

questions of the form why does/did this event happen? (known as why-questions). For instance,

a token explanation would explain why my plant died when I moved it from one part of a room

8

to another. A type explanation might explain why similar plants tend to die in similar situations.

Primarily, accounts of explanation have focused on token explanations, and handled type

explanations as a derivative. However, some have argued that type explanations are actually

more interesting, more important, and must be handled separately.

Robert Batterman focuses on universality as the object of explanation, rather than on

individual events (Batterman, 2002b). He hopes to capture a distinct kind of explanation that

relies not on mechanisms or difference making, but in the story about why a class of systems all

exhibit the same large-scale behaviour. This is done in showing that the details are irrelevant to

that behaviour by looking at the explanatory role of what are called minimal models. These

minimal models are caricatures of real systems and represent their physical aspects in almost no

accurate way. The models are used to answer questions concerning the universal behaviour, and

not the individual behaviour of a target system.

He separates why-questions into two distinct kinds: type 1 and type 2. Type 1 questions

ask about specific instances of patterns, and type 2 questions ask why patterns remain stable

under various changes. Answers to type 1 questions do not answer type 2 questions and vice

versa. While mechanistic explanations may suffice for type-1 questions, they will not answer

type-2 questions, and for this one needs minimal models. Minimal models work because the

large-scale behaviour is unaffected by the underlying microstructure (Batterman & Rice, 2014).

What the explanandum is constrains what model, or what kind of model, can explain it. These

distinctions are important to keep in mind when considering what kind of model would best

explain a given explanandum phenomenon.

Explanation and Understanding

One thing that most accounts of explanation often have in common is the notion that explanation

is that which grants understanding. Understanding, then, is what might be considered a goal or an

essential product of a successful explanation. In some accounts, the understanding generated by

explanation is a by-product, and not something that needs to be analyzed along with it (De Regt,

2013). While others have argued that explanation cannot be analyzed without it (Khalifa, 2012).

The idea that explanation grants understanding is an intuitive and appealing way to characterize

explanation, and one that cannot be abandoned completely. In the following the section, I briefly

explore this relation and some of the work that has been done on it. Understanding might be

9

intimately related to explanation, but I think that an analysis of explanation can be fruitful

without also being an analysis of understanding.

According to Hempel, in an explanation “the argument shows that, given the particular

circumstances and the laws in question, the occurrence of the phenomenon was to be expected;

and it is in this sense that the explanation enables us to understand why the phenomenon

occurred” (Hempel, 1965b, p. 337). The idea that being able to predict something because it

follows from certain conditions and a law of nature (covering law) is called nomic expectability.

The idea that this also provides understanding is an aspect of Hempel’s account that is

emphasized by Michael Friedman and by Kitcher, whose unificationism is looked at in detail in

4.3 (Friedman, 1974; Kitcher, 1981, 1985, 1989). The basic idea of unificationism, which is

quite compelling, is that the more unified is our knowledge of the world, the more we understand

it. Our understanding increases with a reduction in mystery and a reduction in brute facts. A

prime example of this is the unification of Galilean laws of free fall concerning objects on Earth,

and Kepler’s laws of celestial mechanics. When the two were theoretically unified in Newtonian

mechanics, and shown to be derivable from the same law of universal gravitation, our knowledge

of the world is increased. Aside from increasing the scope of our knowledge, the fact that two

disparate sciences were now united, increases our understanding of the world and is a testament

to the explanatory power of the unifying theory.

Some have denied the strong connection between unification and understanding, such as

Salmon (1984) and Barnes (1992), who argue that history is full of examples to the contrary.

They opt instead for a causal theory of understanding, which is discussed in Chapter 3. Salmon

seeks to modify Hempel’s idea of nomic systematicity, whereby a phenomenon or regularity is

explained if it is shown to fit into a nomic nexus. Salmon argues that a phenomenon is explained

if it can be fitted into a causal nexus (1984, p. 19). On this view, understanding why a

phenomenon is the case involves delineating its relevant causes. This view is shared by many,

and is also intuitively appealing. Proponents of the view also claim that it is able to corroborate

the intuitions about unification, because what is really happening on this picture as we move

from Galileo and Kepler to Newton is not a reduction in types of phenomena, but an

identification of their causal bases.

For Woodward, explanatory knowledge is knowledge that allows the manipulation of

causal systems (Woodward, 2003). Woodward argues that the reason we should care about and

10

have an account of causal explanation is largely rooted in the idea that we have an almost innate

drive to understand and manipulate the causal structure of the world; it is not mere intellectual

curiosity. One can identify real causal dependencies even without knowing laws, or being able to

systematize, or unify our knowledge. This, for Woodward shows that our aim is not merely to

predict with descriptive knowledge, but to be able to control the world with explanatory

knowledge.

Michael Strevens makes understanding a central part of his account of explanation

(Strevens, 2008). Understanding is needed to distinguish different depths of explanation and

accommodate higher-level explanations, which are more general. Only explanations that are

exhaustive and provide background information can confer full exhaustive understanding. Other

explanations featuring “black boxes” of bracketed-off mechanisms can only confer qualified

understanding. His account takes as a central problem how we understand and explain high-level

phenomena in a world of only fundamental-level causes.

Alisa Bokulich has yet another view on the relation between explanation and

understanding, which will be given in more detail in 2.4 (Bokulich, 2008). It is her view as well

that nomic expectability is not sufficient for understanding. Understanding is given in the

physical insight that one has into the behaviour of a system. This physical insight is intended to

highlight the graspability of employing classical methods in quantum systems, whose solutions

she claims are purely numerical. This physical insight can be determined on her account by the

amount of counterfactual information that a model can provide about the behaviour of a system.

The role of understanding on this account aims to prevent a notion of explanatory depth from

being reductive in order to preserve the explanatory autonomy of irreducible models.

Some have argued that explanation cannot be as distanced from understanding and

communication as others pretend (Bromberger, 1966; van Fraassen, 1980; Potochnik, 2011). On

these accounts, explanation is an act involving communicators and this is often problematically

ignored in accounts of explanation. Some of these issues will be addressed in the following

section on the pragmatics of explanation.

While understanding may be an integral part of explanation, it is not sufficient for an

account of explanation to simply say that that which explains is that which increases our

understanding. There are some concerns with reducing explanation to studies of understanding.

The most obvious worry is that the issues of understanding are at least as contentious as those of

11

explanation. To reframe problems in terms of understanding is unlikely to provide clarity and, I

think, more likely to obscure them. A project that serves to reduce explanation will miss out on

any insights that are gained in a study of explanation alone, which I think can be many. The most

promising strategy is to largely bracket issues of understanding to make up ground on the

epistemic issues of scientific explanation and explanatory knowledge. This coincides with what

Kareem Khalifa has argued about the relation between explanation and understanding: “Any

philosophically relevant ideas about scientific understanding can be captured by philosophical

ideas about the epistemology of scientific explanation without loss” (Khalifa, 2012, p. 17).

Khalifa looks at well-developed work on scientific understanding and finds that the idea that

understanding is distinct from and primary to explanation has not been successfully made and

further that the successes to be had are better solved by the explanation literature.

Separating explanation and understanding has the further implication that someone’s

ability to understand an explanation is not necessary. In a classroom setting, a professor may

lecture to students about how a particular model can be used to explain some interesting

phenomenon. It would seem absurd to claim that this is not an explanation if one of the students

cannot understand it. I claim that it ought to count as a genuine scientific explanation even if

none of the students are able to grasp it. Subjective notions of understanding should not limit

what counts as a scientific explanation. Accordingly, it can be decided that a model is capable of

supporting explanation independently of whether anyone understands or fully grasps the

explanation.

1.3.1. Pragmatics, Explananda, and Why-Questions

The common sense explanations seen in Section 1.2, involve fitting the facts to a reasonable

story in such a way that everything hangs together. By creating a narrative that is partly filled in

by the explainer and partly filled in by the explainee, a compelling justification of why the

explanandum is the case emerges. Fitting a fact into a reasonable story can also be considered as

explaining that fact. This, for some, points to the importance of understanding a network of

causes and effects. To others it points to the idea that the context of an explanation is an essential

part of explanation, and that a linguistic or syntactic analysis will always be incomplete.

The pragmatics of explanation are quite important to any actual explanation-act, but I do

not think that all the problems of explanation can reduced to problems of pragmatics and solved

that way. Much of the contemporary work on pragmatics draws on work from Bas van Fraassen

12

and others (van Fraassen, 1980; van Fraassen, Churchland, & Hooker, 1985). In order to expose

the importance of context in explanation, van Fraassen addresses the issues of why-questions. He

was not the first to work on this issue, but made substantial contributions.

Sylvain Bromberger proposes a linguistic approach to explanation that pays close

attention to the importance of why-questions and their effect on explanation (Bromberger, 1966,

1982). He notes that most philosophers dealing with explanation focus on a particular subfamily

or subfamilies of questions that can be considered explainable: some on what causes or what is

the mechanism for questions, others on how possible questions, or according to what law

questions. Each of these demands a different explanation, and so a successful explanation of one

is not necessarily a successful explanation of another. And further, not all why-questions admit

of explanations.

Van Fraassen offers an account of the pragmatics of explanation that is capable of further

distinguishing explanation-seeking why-questions (van Fraassen, 1980). He considers the

question “why did Adam eat the apple?” The very same why-question can be construed in

different ways, he argues, each asking for a different answer:

1. Why did Adam (and not someone else) eat the apple?

2. Why did Adam eat (instead of throw away) the apple?

3. Why did Adam eat the apple (instead of something else)?

This reveals something that Bromberger’s analysis was not fine-grained enough to capture.

When spelling out the question with emphasis one can see that there is a different contrast class

involved in each formulation, which I have taken the liberty of adding in parentheses. For van

Fraassen, this points to a serious problem in explanation which is that context determines

relevance. An explanation is then only an explanation in context and not independently. Even if

one has the information needed to answer a question, “that information is, in and by itself, not an

explanation: just as a person cannot be said to be older, or a neighbor, except in relation to

others” (p.130). Van Fraassen argues that an explanation is always given in context and a

question is always asked with respect to a contrast class. As such, explanation is not the same

kind of thing as description, but is rather a three term relation, between fact, theory, and context.

For van Fraassen, the right answer to an explanation-seeking why-question, cannot be

determined without context.

13

Another related view is given by Peter Achinstein (1984). Like Bromberger he places a

strong emphasis on linguistic analysis. He goes further into what he means by understanding and

offers an account of what kind of questions explanations answer. He first notes that an

explanation can be both a product and a process. It is, in Austin’s terms, a performative word

(Austin, 1955). Achinstein relies on a prior evidential relation between explanans and

explanandum for explanations to be true. So rather than only the explanans being required to be

true, the relation between the explanans and the explanandum must also be empirically true. This

is intended to be able to avoid problems of relevance that vex the D-N account (4.2.2). For

Achinstein, then, what counts as an explanation depends on the mechanisms in the world, and is

not determinable a priori. He also introduces the idea of a set of instructions for explanations that

specify relations to listeners. His view inevitably relies on empirical facts about psychological

conditions and senses of intellectual satisfaction, and because of this he finds no possible set of

instructions that satisfy specifying scientific explanations.

Robert Brandom champions a kind of inferentialism about language, but also about

explanation (Brandom, 1994, 2007). Broadly, it is the position that holds that inferences establish

the meaning of expressions. This is contrary to denotationalism, which holds denotative

meanings as primary. In terms of explanation, the conclusions one can draw is influenced not by

deductive logic, but by the available shared evidence, shared knowledge, and the intended

purposes. This is because what counts as reasons is context-dependent in just these ways. An

explanation can render the explanandum more credible, by either increasing its plausibility

(producing a warrant) or fruitfulness (increase ratio of successful to unsuccessful inferences, like

explanations by unification). This view has been taken up and expanded upon in recent literature

by others (Donato Rodríguez & Zamora Bonilla, 2009; Reiss, 2012).

Accounts such as these that focus on language and communication are not meant to be

accounts of specifically scientific explanations. And I believe this is a major reason why context

is seen to play such an important role. The situations in which one has only a why-question to

answer are different than those where one has a well-defined explanandum phenomenon and a

history and community of accepted explanatory practices and related previously explained

phenomena. In many ways, when one is looking to provide a scientific explanation, one already

has some knowledge of the context and an idea as to what will count as a good explanation; it is

not hopelessly contextual.

14

This discussion about pragmatics serves to illustrate the importance of disambiguating

why-questions for providing a successful explanation. I think the important lessons from van

Fraassen and others can be translated into lessons about explananda. A why-question is not

exactly the same as an explanandum, but it is related. It is more productive to recognize an

ambiguous why-question, like “why did Adam eat the apple?” and either ask for more

information, or admit that there is not enough information to provide a good explanation. If one

is sensitive enough to what makes a good and bad explanandum then much can be gained from

an analysis of explanation that does not entirely rely on the pragmatics and context. I think it is

very unlikely that one can reduce issues of scientific explanation to issues of communication and

pragmatics, but the latter can inform the former.

To see how this can work, consider the well-known case of Mayor John who develops

paresis (Scriven, 1959; Carroll, 1999). This example is intended as a counterexample to

Hempel’s account of statistical explanation, which will be reviewed in 4.2.1. Knowledge at the

time the example was put forth suggested that one can only develop this by having untreated

syphilis. Even if one has untreated syphilis it is still quite unlikely that one will develop paresis.

So when asked “why did the mayor come down with paresis?” a reasonable answer seems to be

that he had untreated syphilis. What this really is is a how-possible explanation with a poorly

worded why-question. The explanation is answering the question of how it was even possible

that the mayor developed paresis. In this case, it is an adequate explanation. However, if the

explanandum phenomenon were different and one was asked “why, given that the mayor had

untreated syphilis did he come down with paresis even though others who have the same do

not?” then this no longer serves as a good explanation. There simply is not enough information

provided to adequately answer the question. The same can be said for explanations of why

someone develops lung cancer given that they smoke. It is important to recognize that why-

questions can be ambiguous, and that only having an unambiguous why-question allows for a

satisfying analysis of explanation. My interpretation of this example is a sign of my commitment

to the deductive nature of explanation, which is expanded on in Chapters 4 and 5.

This raises a similar point about the answerability of explananda: complex explananda

may not support explanations. If one’s explanandum phenomenon is “the behaviour” of a large,

complex system, then it is not always clear what will count as a satisfactory explanation. If one’s

goal is to accurately represent the interactions of a certain class of micro-constituents, then a

15

representational model, or set of models, would suffice. If one wants to capture only the macro-

level behaviour of the system, then a tractable, non-representing model will be preferable. But it

is not fair to assess the success of an explanation irrespective of explanatory goals. One cannot

demand an explanation without clear expectations of what can count as a successful explanation.

Another important trait of explananda is that they are statements that are true or

approximately true of a real world system. This is an idea I return to in examining explanatory

derivations in 5.2. The truth or approximate truth of the explanandum can differentiate a good

from a bad explanation. If a robust scientific model makes a great deal of predictions about the

behaviour of a system, but is wildly inaccurate, then no matter what kind of restrictions are

placed on the model, it cannot be explanatory of the behaviour. A model needs to make fairly

accurate predictions in order to be said to support explanations. It is difficult to understand what

kind of explanation would result from a false or impossible explanandum, such as “why does the

Sun revolve around the Earth?” It may seem obvious but it bears remarking that if the

explanandum is false and not even approximately true, then there is no explanation because there

is no actual phenomenon or pattern to explain. Some explanations concern the behaviours of

models themselves. In such cases, the explananda are descriptions of the model and not a target

system, but what I have said still applies.

Goals of an Account of Explanation

In order to judge how successful an account of explanation is, it is important to establish what

the goals of an account should be. Outside of contributing to understanding, which was discussed

above, there are a few things that many have argued should be goals of an account of

explanation. There are other things that are not seen as necessary for explanation, but are

desirable. This section will speak to some of these.

1.4.1. Descriptivism and Normativity

One mark of a successful account of explanation is that it is informed by and largely agrees with

the consensus of the scientific community regarding explanation. It would be a failing for an

account of explanation to have vastly different judgments about explanation than scientists do. If

it did, there would not be much sense in calling it explanation, rather than an assessment of some

other scientific value. This means that good models from successful theories ought to be counted

16

as explanatory on a good account, but models from false theories or data models should be

debarred.

There seems to be two kinds of approaches one can take towards this aspect of

explanation: descriptive and normative. The obvious problem with descriptivism is that the

community could be mistaken and there are no independent assessments of the community’s

claims. There is a further concern that the community cannot be rigorously defined as a set and

surely not all scientists in a field agree. It is important not to rely exclusively on claims by

scientists about whether or not something is explanatory. Scientists are not always concerned

with providing rigorous explanatory derivations that we like to see in contemporary philosophy

of science. And of course, an account of explanation need not reflect every claim from scientists

who claim to explain some behaviour or phenomenon. I think that an account of explanation

should not simply agree with the judgments of scientists. To have an account of explanation that

says that what is explanatory is exactly what a community of scientists says is explanatory, is to

have no account of explanation at all; it essentially leaves the notion unanalyzed. This is

something I will come back to in 2.4.3 and 3.3.2. It is of course important to have an idea what

this consensus is, but it does not provide any independent reasons for claiming that something is

explanatory.

On the other hand, a normative account attempts to articulate independent standards

according to which valid explanatory judgments can be made. The main worry with this is that

this can be descriptively inadequate and epistemically suspect if it is only accountable to its own

independent evaluation. This becomes something of a balancing act as it is important to reflect

the practice of scientific explanation, but to not simply have a descriptive account. There is an

aspect of normativity involved in accounts of explanation. Perhaps a good approach is to attempt

to strike a balance; to have an account that captures the best explanatory practices, or to be

“partially revisionary” as Woodward calls it (3.3.2). This way one can have well-founded

reasons for excluding some claims about explanation, such as that they do not meet the necessary

criteria that are all met by prime and obvious examples of successful explanation.

This continuity with explanatory judgments also runs in the other dimension, in that an

account of scientific explanation ought to be continuous with an everyday sense of explanation.

It would be unreasonable to accept as successful an account of scientific explanation that made

no mention of knowledge or understanding, or which ran contrary to all our shared judgments,

17

insofar as we have them, about what explains what. I take from this is the idea that it is

legitimate and reasonable to rely on our judgments about explanation as guides, but not to solely

rely on them. It is also important not to automatically take conclusions about everyday

explanations as also applying to scientific explanations. There is no need to attempt to cover all

kinds of explanation in a single account. This can be done while also saying more than simply

that explanation is contextually determined. Independent criteria can and should be established.

Some might say that a significant marker of the success of an account of explanation is

that it can make sense of our past judgments about explanations. This is not normally seen as a

necessary goal of explanation, but it is perhaps a desideratum. One can be more confident in the

success of an account of explanation if it can also give insight into why past theories were

preferred over others. This kind of independent assessment of explanatory merit and theory

choice has of course been strongly problematized by Thomas Kuhn and others (Kuhn, 1962).

However, if an account of explanation can make sense of some past judgments about

explanation, then all the better.

1.4.2. Threshold and Depth

There are different methods of determining what counts as an explanation. Some accounts

construct a threshold to determine what is explanatory. This is often done by establishing a set of

necessary criteria for explanation, and allowing that any model (or theory, or argument,

depending on the account) that satisfies the criteria is considered explanatory. This method

makes room for a pluralist notion of explanation, where models from different levels of

investigation can all be considered explanatory if they meet the criteria.

Other accounts have proposed means of evaluating the relative explanatory merits of

competing explanations, by means of something like explanatory depth. Strevens, for instance,

proposes a measure of depth that can be used to determine which explanations are the deepest,

but also not necessarily the most fundamental (Strevens, 2008). Philip Kitcher’s unificationist

account has winner-take-all conceptions of explanation, in which only the most unified theory is

considered explanatory (Kitcher, 1981). This allows Kitcher to debar certain non-explanations,

but has drawbacks of its own, which will be covered in 4.3.2.

These two methods are not exclusive. In fact, Woodward offers an account of explanation

that makes use of both (Woodward, 2003; Woodward & Hitchcock, 2003b). There is a threshold

above which a model is explanatory, but there is a range of relative explanatory depth above that

18

threshold. Having this hybrid approach helps to make sense of intuitions about explanation that

reasonably could find two models to be explanatory, but also show that one provides a better or

deeper explanation.

I do not think that having a method for comparing relative explanatory merit is necessary.

If an account is capable of supporting correct judgments about what counts as a genuine

explanation, then it ought to be considered successful. There is little reason to think that an

account that successfully characterizes explanation needs to also double as an account of

explanatory depth or explanatory power. Similarly, it seems reasonable to think that an account

of explanatory power or depth can be given independently of an account of explanation.

However, if an account has only a method of comparing relative explanatory merit, then it is

incapable of supporting claims that a model explains a certain phenomenon. Rather, it could only

support claims that a model explains a certain phenomenon better than another model does. This

is far less satisfying. The other obvious problem with this approach alone is that it will count all

models to be explanatory to some degree, which runs contrary to out intuitions about

explanation, and is likely just not true. This will be expanded on in 2.4.2.

It seems that a hybrid account is the most ambitious and if successful, would achieve

more. In the account I propose in Chapter 5, I argue that there should be a high threshold for

explanation. This will highlight the best explanatory practices and will establish normative

claims as to what a good explanation ought to do. I also allow that multiple models can be

explanatory and offer some suggestions as to how one could determine the relative explanatory

merit among those models that satisfy the necessary criteria.

1.4.3. Scope

The goals of what an account of explanation ought to accomplish have been growing steadily

since this was first analyzed by Hempel and Oppenheim (1965a). When Aspects of Scientific

Explanation was written, there was an effort to focus on a tractable project that could be

successfully analyzed. When the focus remains small, a sizeable contribution can be made to our

understanding of how science explains particular phenomena and generalities. The focus had

changed for Kitcher in Explanatory Unification (1981). What Kitcher undertook was an analysis

of the explanatory power of theories. He provided a formal system to capture the intuitive idea

that a global theory of science is explanatory to the degree that it unifies scientific knowledge. It

seems that problems arising from Kitcher’s account are due at least in part to the fact that the

19

goal he set for his account of explanation was very unlikely to succeed. The idea that a single

dimension can reflect what makes a theory explanatory across all scientific theories is

unfounded, and it is also not a necessary component of an account of explanation (Wayne, 2016).

There are good reasons to think that there is unlikely to be a completely general account of

explanation. Diez, Khalifa, and Leuridan have argued against Nickel’s stance that there can be

domain invariant constraints on explanation (Nickel, 2010). They argue that “the current

emphasis on domain-specific explanations seems to be justified, and philosophers interested in

explanation should feel little pressure to seek some underlying unity in explanations across

domains as disparate as physics and ethics” (Díez, Khalifa, & Leuridan, 2013, p. 395). I will

return to this in a discussion of contextual nature of explanation and explanatory judgments in

5.6 and 5.7.

It is important not to demand that an account of explanation do everything and cover all

domains. In particular, one thing the account I propose in Chapter 5 will not aim to do is to

capture everything that can be considered an explanation, such as everyday explanations,

teleological explanations, and explanations featuring motivations and intentions. It is an explicit

part of the strategy to focus on a tractable subtopic of explanation, viz. to investigate how certain

scientific models explain their target systems.

1.4.4. Accurate Representation

It is reasonable to consider a central goal of explanation to be the accurate representation of a

target system. This can be accomplished either by revealing its mechanism, capturing its causal

dependencies, or its structural relations. I will go over the benefits and limits of representational

accounts of explanation in 5.6. By contrast, the account presented in Chapter 5 does not require

that the model contain accurate representations of the target system. However, unlike the

statements in the explanans, the explanandum is a description of the phenomenon to be

explained. As such, the model is constrained in that the explanans must deductively entail

explanandum. I will expand on this in 5.7. This openness regarding the features of the model

allows for explanations to make use of models that do not accurately represent, even

approximately. The reason behind this is not simply to include more as explanatory, but more

accurately reflect what is considered explanatory in science, and to less problematically capture

how it is that those models explain. The increasing scope of goals of an account of explanation is

20

a problem in a different sense, but expanding the scope of explanatory idealization will help to

match an account of explanation to the explanatory practices of science.

Conclusion

There is much to say about the relation between explanation and understanding, but an account

of explanation need not fully analyze this. What is important is to draw a connection between a

good scientific explanation and an unanalyzed notion of understanding that it grants. There is a

subjective and psychological side to understanding that is not helpful in explicating explanation,

and so it is best to take on a tractable project and focus on one or the other. It is important for an

account of explanation to take pragmatic considerations into account, but pragmatic concerns do

not exhaust what can be said about scientific explanation. If an account is sensitive to the

concreteness and specificity of the explanans, then many problematic cases of explanation can be

dealt with.

I have outlined what I believe are reasonable goals for an account of explanation: an

account should have a high threshold and largely reflect the explanatory judgments of a scientific

community; it may have a measure of depth to further capture our judgments about which

models explain better or deeper; it should not be so broad as to be intractable, but should be

broad enough to capture a wide range of explanations accepted in practice.

At present, the literature on explanation is heavily focused on causal approaches.

Woodward’s account is very appealing to many. It lays out both a threshold for explanation and

allows for a range of explanations above that. It is well-motivated by manipulationist ideas about

how we understand causes, and what it means to explain something. Others, like Strevens, have

also favoured causal accounts, but of a different kind. Strevens is concerned with generality and

higher-level explanations. He develops his account in part to capture idealized models and

provide a measure of explanatory depth that does not favour explanations at the most

fundamental level. In response to these accounts and others, some like Bokulich, have proposed

structural approaches to explanation in order to capture the central role that idealization plays in

explanation and to allow for a wider range of explanatory models. The following chapter

examines the role of idealization in model-based explanations and assesses the success of

Bokulich’s structural account in capturing the way that highly-idealized models explain.

21

Chapter 2. Idealization and Structural Accounts of Explanation

Introduction

This chapter will first introduce the concept of idealization and the inevitable role it plays in the

practice of modelling in terms of realism and representation. I use this discussion of idealization

also as a means of introducing some concerns that have been raised about distinguishing laws

from accidental generalizations. These concerns have been used to problematize covering law

accounts of explanation. But I think that these potential problems can be circumvented with a

model-based account. I then present McMullin’s distinction between idealized models, which

approximately represent, and highly-idealized models, which do not. I follow others in

maintaining that highly-idealized models can be genuinely explanatory, and part of the

motivation of my account is to capture this. Section 2.3 focuses on idealization in the practice of

modelling and the different aims involved in model construction. Modelling, as a pragmatic and

goal-directed practice, has implications for the use of idealizations in explanation.

After introducing idealization, modelling, and highly-idealized models, I turn to examine

an account of explanation that is sensitive to the issues these raise for explanation. In 2.4, I

present Bokulich’s account of structural model explanations that is designed to capture the way

that highly-idealized models explain. On her account, some models of semiclassical mechanics

are explanatory because they capture the structural dynamics of quantum systems. However, I

argue that structure is incapable of capturing the explanatory role of highly-idealized models,

because it either counts all models as explanatory, or shows preference for models that

accurately represent (M. King, 2015). In either case, I aim to show that a structural approach to

explanation cannot capture the way that highly-idealized models explain. The problem is that her

account, even though structural, is still representational, and so the better the representation, the

deeper the explanation. This is not in itself problematic, but it does fail in the aim of showing

how models of semiclassical mechanics and other highly-idealized models are explanatory. This

leaves that task open to a non-representative account, like the one I present in Chapter 5.

22

Models and Idealization

The focus in the literature on scientific explanation in philosophy has shifted in recent years

towards model-based approaches. The idea that there are simple and true laws of nature has met

with considerable objections from philosophers such as Nancy Cartwright (1983), Paul Teller

(2001), and numerous others. This has made a strictly Hempelian D-N-style explanation largely

irrelevant to the explanatory practices of science (Hempel & Oppenheim, 1948). The practice of

explanation does not merely involve subsuming particular events under laws of nature. It is

increasingly recognized that science across the disciplines is to some degree a patchwork of

scientific models, with different methods, strategies, and with varying degrees of successful

prediction and explanation. And so accounts of scientific explanation have reflected this change

of perspective and model-based approaches have flourished in the explanation literature

(Woodward, 2003; Craver, 2006; Bokulich, 2008).

To talk about idealization, it is important to be clear on what is here meant by a model.

Paul Teller has argued in favour of a broad conception of a model (Teller, 2001). It is impossible,

he argues, to give necessary conditions for a particular representation to be considered a model.

A model is a model because it is chosen to represent something, and there simply are no intrinsic

features of a model. However, I think it is useful to consider at least some characteristics of

scientific models, even though there may be no general criteria for the similarity that a model

must bear to the system it represents. The models of interest to me (ones that feature in scientific

explanations) are not merely sets of propositions, or accurate representations of a system or

phenomenon, but quantitative and capable of reliably reproducing or predicting the desired

behaviour. Models are the result of a scientific modelling processes, involving experimentation,

measurement, testing against prediction, calculation, idealization, and often theory. There are

important explanatory differences between the various kinds of models used in science. A

narrower view of what constitutes a scientific model is, I think, advantageous.

A model may be either concrete as in the case of a physical object, like a model ship, or a

map, or it may be abstract like a set of statements, diagrams, and mathematical equations and

parameters. A single equation on its own is not enough to constitute a scientific model in the

sense employed here, though the term model is often employed this way. Historically, some

philosophers have looked at models as linguistic objects; as sets of statements. Recently others

have taken models to be artefacts, or tools of surrogative reasoning. Zamora-Bonilla and Donato-

23

Rodriguez follow Knuuttila in this approach (Knuuttila, 2005; Donato Rodríguez & Zamora

Bonilla, 2009). On their view, models are not only sources of knowledge about the world, but are

used for the inferences drawn from the systems they are about. I return to this idea later in the

section.

The models of interest to me in talking about scientific explanation are abstract objects

used to gain knowledge about the world. A model is an idealized version of a real-world system

(often called a target system) designed with a purpose; to bring to light some relation or capture

some behaviour. While it is not a perfectly accurate version of the real world, it is related to it in

various ways. This relation is often given in terms of approximation, abstraction, similarity, or

isomorphism. What this process of idealization consists in and how idealized models relate to the

real world is a matter of no small debate and is the focus of this section.

2.2.1. What is idealization?

Idealization in the context of modelling is the process of approximating a system in order to

facilitate calculation, to bring a certain relation to light, or to render it easier to understand.

Models are not only sources of knowledge about the world, but are used for the inferences drawn

regarding the systems they represent. Zamora-Bonilla and Donato-Rodriguez take an inferential

approach to models and explanation that I will not pursue very closely, however the process they

describe is largely representative of a variety of views (Donato Rodríguez & Zamora Bonilla,

2009). They separate the process they refer to as surrogative reasoning into three steps:

1) Interpret the physical system and make inferences about the model from propositions

about the empirical system.

2) Make formal inferences within the model, either deductive, inductive, counterfactual, etc.

3) Retranslate (interpret) the conclusions into language about the empirical system.

For them, the steps are successful if the inferences made are valid and what determines this is the

contextual standards of a discipline or community of competent language users, a notion they

take from Brandom and the inferentialist tradition (Brandom, 1994, 2007). They argue that there

is no algorithm for constructed models, not even in theory. They will always involve background

knowledge, analogy, intuition, metaphors, empirical facts, etc.

Idealizations are false assumptions about the target system, such as that there is no air

resistance, the gas has an infinite number of molecules, the carrying capacity is constant, that an

object is a perfect sphere, or a point particle, or that liquid in a pipe exhibits perfectly laminar

24

flow, and so on. They can be very helpful in providing mathematical representations of systems

that might otherwise be hopelessly intractable. This process might involve leaving details out of

a model in order to simplify it, or distorting existing data to eliminate noise that has no effect on

bringing about the phenomenon; a process often referred to as abstraction. In many cases, the

behaviour or phenomenon might be obscured without the idealization. But idealizations, even

though false, are not simply applicable to ideal worlds; they can generate knowledge about the

real world. Ernan McMullin, who wrote an important piece on idealization, reminds us that it is

not an “escape from the intractable irregularity of the real world into the intelligible order of the

Form, but to make use of this order in an attempt to grasp the real world from which the

idealization takes its origin,” (McMullin, 1985). Models are idealized in order to have real-world

application.

McMullin reaches back as far as Plato and Aristotle to make the point that the truth of the

mathematical representation of nature has always been an issue of debate. For Aristotle,

mathematics was a reliable mode of analysis of nature and not merely a study of a realm outside

of nature. One could be justified in using mathematics to study the quantitative aspects of the

real world. Plato on the other hand had argued that matter made a proper or complete realization

of geometry impossible.

When Galileo performed his experiments, he made assumptions that were false about

elements of the model. He assumed that a ball he was rolling was a perfect sphere, and that a

plane was perfectly flat, or exactly inclined. But he was not making claims only about the

behaviour of perfect spheres. In doing so, and applying the mathematical relations he derived to

the real world, he was in essence arguing that the difference between the perfect sphere of the

model and the ball he was using was negligible to the behaviour he was modelling. Of course,

when the difference is not sufficiently small the predictions of the model will begin to diverge

from actual measurements. McMullin also points out that models in contemporary science are far

more complex than justifying the application of a simplified geometrical scenario to a system

that closely resembles it. In the Galilean idealization that McMullin is characterizing, one can, at

least in principle, retrieve the original real-world system by de-idealizing the model; by adding

detail back into the idealized model one can generate a model arbitrarily close to the actual

system. These kinds of idealization are known as Galilean. This de-idealization helps to

25

demonstrate why the model, though idealized and not literally true of the world, applies in a

particular case. In order to see this, we will turn to the ideal gas law.

2.2.1.1. The Ideal Gas Law

In 1738, Bernoulli, following Boyle and others, had supposed that gases are composed of an

enormous number of very rapidly moving particles. The compressibility of gases and its relation

to the pressure a gas exerts on a vessel, led him to theorize that pressure was the result of the

particles’ collisions with the vessel walls in a mostly empty space. Boyle’s picture of gases as

particles that act as springs exerting pressure on each other was widely accepted in Bernoulli’s

day and Boyle’s equation,

(1) 𝑃 = 𝐹(𝑇)

𝑉,

for relating pressure, volume, and temperature, where F(T) is a function on temperature and the

amount of gas, was also known. But Bernoulli’s idea that pressure is a result of collisions was

new, and allowed important progress in the development of the kinetic theory of gases.

This idea easily leads to Boyle’s law. If the average speed of the particles remains

constant, then the number of collisions on the container walls would be proportional to the

density of the gas, and so pressure would be inversely proportional to volume: 𝑃 ∝ 1 𝑉⁄ . The

speed of the particles was assumed to increase with an increase in temperature, and so Boyle’s

law falls out of Bernoulli’s kinetic picture. But he was also able to add that an increase in the

number of particles would increase the number of collisions and so a constant was multiplied to

the temperature function to give

(2) 𝑃𝑉 = 𝑛𝑓(𝑇),

where the pressure P and volume V are equal to the mols of the gas multiplied by a function of

the temperature, f(T). An equation of state one immediately recognizes as being just a few steps

short of the ideal gas law, 𝑃𝑉 = 𝑛𝑅𝑇.

The kinetic theory of gases presents a picture of a fictitious gas in which the particles are

thinly dispersed point-masses that do not interact with one another and have elastic collisions

with vessel walls, which is measured as pressure. We know that this is quite false of any actual

gas. Gas molecules take up space, some quite a lot, and certainly interact with one another and

26

dissipate energy on vessel walls. But the idealization is still applicable to the real world. Within a

certain range of temperature and pressure, the ideal gas law is capable of providing results

accurate to those experimentally measured. The reason this only works well within a certain

range is due to the nature of the assumptions. For instance, the assumption that the particles are

point masses begins to be problematic when the molecular size becomes significant relative to

intermolecular distances, and real gases will begin to condense at low temperatures.

The idealizations involved in this model are just the Galilean idealizations mentioned by

McMullin. What this means is that details of the real gas can be built back into ideal model, and,

in a sense, one can retrieve the real-world system via de-idealization. In fact, collisions were

introduced to the kinetic theory of gases by Maxwell and Boltzman, by the use of the concept of

a statistical average. The kinetic theory of gases and its idealizations are a prime example of an

idealized model, and one can easily see that it is not a special case.

Many arguments have been made regarding the unavoidability of idealization and non-

optimal models given facts about complexity and irreducibility, and the growing acceptance of

the impossibility of completely accurately representing most real-world systems. But not all

idealizations are of the Galilean sort. Non-Galilean idealizations are incapable of being de-

idealized in this way. A model featuring non-Galilean idealizations is often called a highly-

idealized model. Robert Batterman has described these models as having “controllable”

idealizations, in that the idealizations of the system are justified theoretically (Batterman, 2005,

p. 235). Bokulich and others, such as Batterman, argue that these models can be genuinely

explanatory and offer accounts of explanation to capture how. I review such an attempt by

looking at Bokulich’s case study of semiclassical mechanics in 2.4.

2.2.2. Laws and Idealized Models

There are many reasons for maintaining that idealized models are capable of supporting

explanations, even though they do not contain laws of nature. As mentioned, idealized models

make use of false assumptions, but they are essential for scientific explanation. The D-N account

requires that the explanandum be a logical consequence of the explanans and contain a law of

nature. However, fundamental laws of nature, as Nancy Cartwright points out, are very rare

indeed (1983). And further, in some cases the fundamental theory is incapable of generating the

best explanation, as is argued by Sandra Mitchell (2003). This subsection will outline these

arguments for the need for idealized models, and the following subsection (2.2.3) will go over

27

arguments concerning not just the need for idealization and high-level, or non-fundamental,

models, but for the essential role of idealization in explanation.

2.2.2.1. Laws and Lies

Cartwright’s arguments regarding the status of laws of nature are well-rehearsed in the

philosophy of science, but her argument bears review here. This discussion will not support

Cartwright’s own conclusions that laws are false and non-explanatory, but rather that idealization

and compromise are integral parts of the practice of modelling and also the explanations in

which the models feature.

In How the Laws of Physics Lie, Cartwright argues that the entrenched, or traditional,

view that the laws of nature state facts is mistaken (1983). She calls this view the facticity view

of laws. She cites J.J.C. Smart as saying that biology has no genuine laws, but in fact, she argues,

neither does physics. Cartwright is a kind of anti-realist towards laws, but not in a

straightforward manner. Cartwright finds explanation and truth to be at odds everywhere. Her

argument revolves around the point that even laws of nature cannot accurately describe the

behaviour of systems to which they are applicable, because the application of each law assumes

that there are no other forces at play. Forces are ubiquitous in nature and so the actual movement

of two bodies is never fully given by laws of mechanics, for instance.

If one considers the universal law of gravitation, one will find that it never fully describes

the total force between two bodies. For charged bodies, this law and Coulomb’s law interact to

give the final resultant force. But Coulomb’s law does not work for massive bodies, and at very

small scales the electric charge far out-performs the gravitational force on massive bodies. So of

course the laws are to be considered ceteris paribus: if there are no other forces than gravity

acting on two bodies, they would be fully describable by the universal law of gravitation alone

(Hüttemann, 2014). Cartwright claims that this is no help in real-world scenarios, because the

antecedent is never satisfied. Curious, given that it is obviously well-known that this law and

other similar laws are in fact incredibly useful in real-world scenarios. When the other forces

acting on bodies are small, the law can accurately describe the behaviour of simple systems. One

can know why and when the law is reliable by looking at the assumptions involved: that the

masses are point particles, that other forces are negligible to the behaviour, and so on. When

28

these idealizations become very far from the truth, the regularities cease to hold and explanations

fail.

Much of Cartwright’s argument relies on the notion that an explanation is given in

subsumption under covering-laws. If one holds to a strict covering-law account of explanation

where all phenomena neatly fall out of the laws of nature, then her argument poses more

significant problems. Covering laws are rare in scientific explanation. As was mentioned, in

recent literature the focus has shifted from law-based explanation to model-based explanations.

The fact that the direct and simple application of laws of nature is perhaps never realized, serves

as a reminder that the models built in science are not exact and universally applicable, but are

rather designed with aims that make a compromise with other desiderata.

2.2.2.2. Complexity

Sandra Mitchell has argued in favour of scientific pluralism, largely on the grounds that the

world is simply too complex to be represented at a single level of study (2003). She argues for a

particular kind of pluralism, in which multiple levels of models can be integrated to form a single

explanation of a phenomenon: what she calls integrative pluralism. In this section, only her

arguments from complexity will be examined, as they make a strong case for the explanatory

power of idealized and non-fundamental, or higher-level, models.

Mitchell argues that complexity is a critical tool in understanding the nature and limits of

scientific diversity (Mitchell, 2003). Attention to different components and different degrees of

abstraction are appropriate to our various explanatory goals and conceptual and computational

abilities. An examination of complex systems, Mitchell argues, can inform us about the limits

scientific reduction and explanation. By the term complex system, she means a system in which

the micro-details are so many, or take place on such a long time scale, that the computation of

the evolution of the system becomes intractable. Mitchell distinguishes between three kinds of

complexity: a) constitutive complexity, where the sheer number and organization of the system’s

components demand a higher-level model. She has in mind here the case of eusocial insect

colony organization, where models of the lower-level interactions are insufficient to bring about

the behaviour observed at the colony level. In general, the problem is that the observed macro-

level properties of some systems cannot be determined from models of, and what is known

about, micro-level interactions; b) dynamic complexity, where the time scale and number of

29

interactions of the system’s components preclude full calculation. This is a familiar kind of

complexity featured in chaotic systems and A-Life simulations. Even simple deterministic

systems, left long enough, can become chaotic and unpredictable; and c) evolved diversity which

is a historical indeterminacy brought by the fact that one cannot know the particular history that

brought about the state of a complex system. Mitchell’s example here is the evolutionary history

of species. There is not enough information to determine the causal history of evolution. There is

no telling why certain traits were selected and others that could have succeeded as well or better

were not.

For Mitchell, all this goes to show that biology and the sciences in general ought to

represent the complexity and diversity of its subject matter by abandoning the idea of laws of

nature and recognizing the pragmatics of scientific modelling. Instead of merely claiming that

there are no laws, or that they are inapplicable in real scenarios, she plots all laws in a 3-

dimensional graph, measuring stability, strength, and degree of abstraction (Mitchell, 2000). The

values of scientific law are continuous and not binary. Mitchell makes this claim about laws as a

way to support the legitimacy of biological generalizations by placing on a spectrum,

generalizations such as the law of the conservation of mass-energy, Mendel’s laws of

inheritance, and the metallic composition of the coins in Nelson Goodman’s pockets. In earlier

work, she argued that the pragmatic function of the law was more important than whether it was

exceptionless (Mitchell, 1997). How and why the law applies under which circumstances is what

is important.

The search for quantitative generalizations in biology is not fruitless, but at the same

time, we must be cautious about claiming that all of these are laws of nature merely separated by

degrees. Mitchell has argued cogently that the binary view of laws as either contingent or

universal is either mistaken or unuseful, and that the generalizations employed in science are

often quite contingent and specialized. Scientific models feature just these generalizations, with

degrees of stability, strength, and abstraction. They are not universal, but designed for specific

purposes, limited in application, and restricted by tractability and complexity. Mitchell is right

that these generalizations are not perfect laws of nature.

30

2.2.2.3. Perfect Models

McMullin’s discussion of Galilean idealization involves the assumption that one can actually get

arbitrarily close to the real world by de-idealizing models. Consider an idealized model like the

simple pendulum, whose motion is partially described by the following equation,

(3) 𝑇 ∝ √𝐿

𝑔,

where T is the time, L the length of the string, and g the acceleration due to gravity. This

equation gets at the relation between the length of the string and period of oscillation (under

certain conditions), specifically by ignoring details like the damping effect of air resistance, the

mass of the string, elasticity, the minute change in gravity as the pendulum swings up and down,

and so on. It is an ideal pendulum that accurately describes no pendulum in the world. One can

take this model, or rather its modern version, and add corrections to account for the presence of

air, the mass of the string, shape of the bob, and so on. One can build detail back in and get

closer and closer to the real world pendulum, i.e. to a perfect model of the pendulum.

However, the idea that there even is a perfect model, toward which any scientific model

aims, has been criticized. Paul Teller has cogently argued in “Twilight of the Perfect Model

Model” that there is little reason to think that models could ever be perfect (Teller, 2001). The

main thrust of his argument comes from concerns about providing a general account of similarity

and approximate truth. The problem for many anti-realists is “to identify the relevant similarity

between situations, on the one hand the actual situation, and on the other some non-actual

idealized simplification of the way the world really is, what is being called a model” (Teller,

2001, pp. 403-404). The problem only arises when the idealized situation is considered in

linguistic terms. When theories are seen as a collection of models and the relevant interests of

similarity are fixed, then one can see that there is a sufficiently similar structure. The demand of

the anti-realist for a definition of closeness to truth simpliciter is misguided. Whether a model

exhibits similar behaviour to the system will depend on such things as the details of situation and

the aims of the modeller. The similarity between the model and the real world is largely

dependent on what feature or features of the system the scientist is aiming to model. There are no

general criteria for a perfect model; models are compromised from perfection and designed with

aims in mind. The aims of modelling will be looked at in more detail in section 2.3.

31

2.2.3. Idealization and Explanation

Idealization is unavoidable in science, but it is not merely unavoidable: it is an important aspect

of modelling and explanation. This subsection tries to highlight that idealization not just a

necessity that must be dealt with, but integral to what an explanation is. The organized,

simplified, and related information that idealized models provide is key to the model’s being

explanatory. Its importance is that it helps us understand by highlighting or capturing salient

features of systems, relative to our explanatory interests. The remainder of this section will be

spent showing that many of the most successful accounts of explanation attempt to get at higher-

level, and highly-idealized, explanations, but will not go into detail on how. Details about their

attempts and their success will have to wait until chapter 3.

2.2.3.1. Causal Processes

James Woodward criticises certain causal-mechanical approaches to explanation, such as

Salmon’s, that identify cause in physical processes and interactions (Woodward, 1989).

Salmon’s causal mechanical account of explanation states that explanation involves showing

how an occurrence fits into the causal network of the world (Salmon, 1984). For Salmon,

causation has three aspects: causal interactions, causal processes, and conjunctive forks. Causal

processes are characterized by the transmission of a mark or a structure, such as the movement of

a free particle. A causal interaction is when one causal process intersects and modifies another,

such as the collision of two particles. Conjunctive forks are correlations among spatio-temporally

separated effects, explained by separate causal processes deriving from a common cause.

Salmon’s conception of causation seems to fit best with simple physical systems governed by

classical mechanics, but complications arise when we move away from such systems. Woodward

points out that this would lead to denying the explanatory power of quantum mechanics and

other systems that make no appeal to underlying causal processes.

This also applies to any higher-level models. For example, when explaining the

expansion of a gas in a balloon when heated, one does not need to focus on the causal histories of

each molecule and say that the behaviour of the system is somehow the sum of these. Instead one

focuses on the general assumptions regarding the forces and then deriving and solving the

Maxwell-Boltzman equation for the system. One can then show the various behaviours of the

system that follow from this. An account of explanation that focuses on processes will inevitably

32

fail to include higher-level and highly idealized models. Woodward’s argument is that even if

this were possible, it would not be explanatory. The actual causal history of the molecules gives

no explanatory information on why the balloon expanded when the gas was heated. The best

kind of explanation is a higher-level model relating properties such as temperature, volume, and

pressure.

Michael Strevens also recognizes the need for higher-level explanations (Strevens, 2008).

He aims to provide an account of explanation that begins with the total causal picture of the

world, and abstracts away irrelevant details until only difference making causes are left. His

account of explanation has two aspects: lower-level causes and higher-level relevancy criteria.

Strevens’ aim is to use an aspect of unificationism to solve one of the major problems of causal

accounts: specifying the causes that are relevant to an explanandum phenomenon. He is using

unification in order to pick out not the most unified of theories, but which causes are the

difference makers. By using this process he will be able to count higher-level models as

explanatory by linking them through a process of optimization to the world’s fundamental causes

(Strevens, 2004).

As an example of how a higher-level, and non-veridical, model can be more explanatory

than a more fundamental model, Strevens compares possible explanations of Boyle’s law. He

begins with a textbook explanation that makes use of many idealizations, such as that the

collisions with container walls are completely elastic, and that molecules do not collide with one

another. It gives a decent explanation, but makes some false assumptions. He next goes over the

complete description from modern kinetic theory, including the influence of the molecules on

one another at a distance, and intermolecular collisions. He then goes over the process of

eliminating the aspects of the complete description that do not make a difference to the Boylean

behaviour of the gas. It is only in this way that one can understand Boyle’s law: the relations that

are highlighted in that law are the difference makers for the explanandum. Thus, the textbook

explanation is the best. It is not the most veridical, but it has only difference makers. He

supposes that the assumption that there are no collisions is actually the way of communicating

that it makes no difference to the behaviour to suppose that there are none. Idealization, for

Strevens, shows what is irrelevant. Idealized models are preferable to veridical models in a few

ways: they highlight the irrelevance of certain factors; they are much simpler; they can remain

effective predictors as long as the idealizations are reasonably faithful.

33

As was mentioned above, many have argued that non-Galilean idealizations featured in

highly-idealized models ought not to preclude explanation (Rueger, 2001; Batterman, 2002b;

Bokulich, 2008; Batterman, 2009; Wayne, 2011; Jansson, 2014a, 2014b; Reutlinger, 2014;

Saatsi, 2014; Wayne, 2014; Rice, 2015). The argument is often made on the basis of the

importance of these idealizations in the scientific practice of explanation. If highly-idealized

models are not explanatory, then scientists must be quite mistaken about explanation. It is widely

recognized that models featuring non-Galilean idealizations are pervasive and can support

explanation. These idealizations can range from two-dimensional agitators and lattice gas

automata to the perfect rationality of consumers and constant populations. There is a push for

philosophical accounts of explanation to handle this. Even though specific approaches that allow

for these are not without concerns. So far in this chapter we have focused on characterizing

idealized models in the context of an account of explanation. The following section will

elaborate on the aims of modelling mentioned earlier in order to preface the discussion of

Bokulich’s structural model-explanations.

Modelling and Explanation

This section aims to highlight the strong role that modelling plays in explanatory arguments.

Specifically, I look at the desiderata involved and show that they often pull in opposite

directions, and the model that one constructs is dependent on the aims of modeller, or

explanation seeker. This is to highlight the pragmatic and imperfect nature of models, and to

reinforce the idea that models can be constructed for specific explanatory purposes, which I

make use of in 5.4.

2.3.1. Desiderata

The aims of the modeller and the behaviour of interest in the system does much to determine

what information is included in the model. If precision and accuracy are the main desiderata for

the model, then maximizing details is a useful strategy. This is the kind of modelling that takes

place in highly detailed computer simulations of particular systems comprised of multiple

individual models. In biology, this can be used to model ecosystems while taking into account

models that reflect the changing effects of sunlight, temperature, air pollutants, water pollutants,

and various interacting species populations (Mitchell, 2003). Of course, maximizing details of a

34

particular system is likely to make it less generally applicable, and less tractable. And so for

instance, generality is compromised when maximizing realistic detail.

Sometimes, however, capturing the desired behaviour involves ignoring details. In order to

focus on a specific behaviour, the investigator might look at the system at a higher level of

inquiry. The term level has different meanings in various sciences. In biology, it might be used to

differentiate between the models of molecular biology, cellular biology, models of individual

behaviour, and models of group behaviour. In physics, it might mean the difference between the

study of a gas in fluid mechanics, and the same gas at a small enough scale that intermolecular

forces affect its behaviour. For instance, as Teller mentions, if one is interested in modelling

waves in water on the order of one meter in length or more, then one can safely ignore the effect

of surface tension (Teller, 2001). However, if one is looking for ripples only millimeters in

length, then surface tension will play a large role.

Physical scale alone is not the determiner of level either. Temporal scales can differ

enough to generate very different models, some effective at short time intervals, and some to

describe the long term behaviour. For instance, the equations governing the long-timescale

motion of an oscillator are quite different from those governing its short-timescale behaviour.

This is primarily because the model of the long-timescale behaviour does not exhibit the same

periodic motion. When the number of oscillations is small the model exhibits near-harmonic

oscillation, but when the number increases, the behaviour is dominated by the rate in the change

of amplitude (Wayne, 2012). The level at which the model is constructed is related to the

behaviour that is being modeled. Some models are designed to capture behaviour at scales or

values where the models of fundamental theories no longer apply. These higher-level models are

often said to capture the system’s emergent behaviour or properties, though when a behaviour or

property becomes properly emergent is a matter of no small debate (3.4).

But there are additional dimensions to modelling as well. Sometimes the aim of the

modeller is to generate a maximally simple model, or to capture the large-scale structure of quite

different systems. This model may retain very little of the system’s details, only enough to

generate the desired behaviour or to reveal a pattern it shares with different systems. This could

be favoured for its calculational tractability, or for revealing a bare-bones mechanism. Simple

models will lose out on accuracy, but many models could have the accuracy added back in, at

least in principle, as was seen in Section 2.2.1. Simple models have the added benefit of being

35

very general. Simple models that exhibit the same behaviour may apply to systems with radically

different components, and this is an interesting phenomenon in its own right, one looked at in

depth by Bob Batterman (Batterman, 2002b).

An example of such a minimal model might be the Lotka-Volterra model of predator-

prey interactions. This is the simplest model of predation. It can be written as the following pair

of equations describing the change in prey and predator populations:

(4) �̇� = 𝑏𝐻 − 𝑠𝐻𝑃; �̇� = −𝑑𝑃 + 𝑒𝑠𝐻𝑃,

where H and P are the prey and predator population, respectively, b is birthrate of prey, d the

deathrate of predators, s the searching efficiency, and e the efficiency with which extra food is

turned into predators. This model describes an ideal system with only two populations, and

ignores all realistic details about them. Not only this, but it makes clearly false assumptions

regarding the process of turning food into predators. This model as it stands does not describe

any particular population very accurately, but it can be refined with more details that are

particular to certain systems, and can then be used to approximate the near-cyclic fluctuations in

populations. This can be done by replacing exponential growth in the absence of predators with

two-term logistic growth expressions,

(5) �̇� = (𝑏𝐻 − 𝑟𝐻2) − 𝑠𝐻𝑃,

and the same can be done to represent alternative food sources,

(6) �̇� = (𝑘𝑃 − 𝑑𝑃2) + 𝑒𝑠𝐻𝑃.

A third population can be added to the food chain and the equation can be tailored in various

other ways. A model that favoured a more realistic representation or prediction of a particular

system would take more information into account, such as the competition between individuals

of the same species; the relation between predator consumption and prey density; and the

efficiency at which new food is turned into extra predators, and so on. The simplest way to take

some of this into account is to introduce carrying capacities for both species and a saturation

effect where the birthrate tends to a finite limit at high density. A more detailed model might

lend itself to a different interpretation of the observed cyclic behaviour: for instance, it could be a

result of a small habitat, one which is in the order of magnitude of the mean displacements of

individuals. Highly detailed and accurate models can be constructed, but this sometimes implies

36

that one must sacrifice the generality and tractability that comes with a simple model. (4) is very

minimal and very general, and attempts only to show the general behaviour of a large number of

possible predator-prey systems. Reasonably, even this most minimal model could support

explanations. It is reasonable to question whether this model can explain features of a population

that do not share its idealizations (Colyvan et al., 2009). It might not explain a particular

population’s changes very well, but it could be employed to explain the similarity of cyclic

population changes across very different systems. That is, it could reasonably be used to explain

why different systems exhibit the similar behaviours. Its aims and its explanatory potential are

very different than those of a model that attempts to maximize precision. It is not uncontentious

that this model ought to be considered explanatory, but I think it is reasonable that it could

explain certain explananda. I will return to this in 5.6.2.

Yet another strategy is to generate predictive accuracy, even at the cost of realistic

representation. These models may include what are called black boxes, which are mathematical

equations that are capable of approximating the desired quantitative result, but bear little

similarity to the system being modelled (Weisberg, 2007). These models can be very useful in

cases where high predictive accuracy is the main desiderata. Black boxes can stand in for

unknown mechanisms and processes and state only input and output. They are silent on the

causal or structural features of the system. Though Michael Strevens makes a case for certain

black boxes being explanatory when buttressed with the support of a framework, generally black

boxes are not considered explanatory (Strevens, 2008). Black boxes can be considered heuristic

devices that may simplify calculation or provide a numerical solution, with little to no

information about the process or mechanism at work.

As the previous sections showed, many models and strategies that are employed in

science are used only as heuristics. This means that the models are of practical use, but are not

intended to explain, or accurately represent the system. There are many models of particular

systems that do not aim to explain them. Some models focus on achieving high predictive

accuracy at the cost of realistic representation, employing black boxes. Some models are merely

calculational tools, whose use in the practice of science is entirely pragmatic, but others are

thought to actually explain the phenomena or system they are modelling. If one does not wish to

say that all models are explanatory, then one needs criteria for which models are going to count

as explanatory, and which are phenomenological, or merely predictively successful. I will return

37

to the relation between heuristics and explanation following a close examination of Bokulich’s

case study of semi-classical mechanics.

Idealization in Physics

This section will examine the potential for a structural model account to capture autonomy, the

idea that not the all good explanations are given from models of fundamental physics. This

section will look at Alisa Bokulich’s work on the central role that idealization has to play in

explanation (Bokulich, 2008, 2011, 2012). Bokulich claims that certain highly-idealized models

are explanatory, even though they are not considered explanatory by causal, mechanistic, or

covering law accounts of explanation. She calls these kinds of explanations structural model

explanations and argues that the structural similarity between the model and the target system

can play the role of determining whether or not that model is explanatory (Bokulich, 2008, p.

145). She formulates her account as structural in part to capture models that are not explanatory,

for instance, on Woodward’s manipulationist account (Woodward, 2003). She aims to expand

the store of explanatory models to include those that do not accurately represent - those that

model a physical system by means of fictitious entities or processes, what she calls explanatory

fictions.

This section examines Bokulich’s account to assess the success of her formulation of the

structural criterion, and also to make the more general claim that there are problems with

maintaining that a structural criterion can capture the way that highly-idealized models explain.

In order to argue this, I will first examine Bokulich’s account as given in (Bokulich, 2008, 2011,

2012) in 2.4.1. I then demonstrate the application of this account in explaining the phenomenon

of quantum wavefunction scarring, which is a pattern in probability distribution in certain

quantum systems. Bokulich argues that this quantum phenomenon is more deeply explained by

appropriating concepts and models from semiclassical periodic orbit theory, than it is by

employing quantum mechanical models. In order to evaluate this concern, I look at two

reasonable approaches to measuring or assessing a model’s structural information, and I argue

that neither approach is ultimately satisfactory. The simplest approach of just measuring

structure proves at worst impossible and at best arbitrary, and the comparative approach, while

succeeding in debarring the Ptolemaic explanation, fails to find the highly-idealized model

explanatory (M. King, 2015). The criterion either wrongly identifies all models as explanatory,

38

or prefers models from fundamental theory. Either way, it cannot capture the way that highly-

idealized models explain. Section 2.4.3, the non-structural aspects of the account are challenged

for their strong role in distinguishing explanatory from non-explanatory models. I close the

section by discussing heuristics and articulating what I believe to be the crux of the problem and

what can be taken away from the investigation.

2.4.1. Structural Model Explanations

This subsection examines Bokulich’s structural model account of explanation as laid out in

(Bokulich, 2008) and incorporates the responses and clarifications made in (Bokulich, 2011,

2012). Bokulich’s account of explanation relies in part on work done by James Woodward, in

particular the notion of depth he develops along with Christopher Hitchcock, which I go over

first (Woodward & Hitchcock, 2003a, 2003b). This notion is rather important for Bokulich as it

forms the basis of the structural criterion of her account. I then show how her structural account

aims to capture the explanatory value of semiclassical models, which are essentially classical

models that can be used to approximate quantum systems. I do this by looking at the

phenomenon of quantum wavefunction scarring and demonstrating how it satisfies the account’s

criteria. This shows both how a successful structural explanation is intended to proceed and

motivate some of the intuitions we have about highly-idealized models being explanatory.

On Woodward’s account, causality is framed in terms of manipulability relations rather

than in terms of causal mechanisms or physical interactions (Woodward, 2003; Woodward &

Hitchcock, 2003b). Causal relationships, he claims, are out there in the world, but they can be

well described in the reliable variable dependency relations of models. Explanation is the activity

of gaining information about these causal relations by discovering through intervention which

dependency relations are strongly invariant. The counterfactual dependency of these relations

gives us important information that provides explanatory depth. This is information that answers

what-if-things-had-been-different questions, or w-questions. Thus, the range of questions that

counterfactual dependence answers is related to the explanatory power of that causal relation.

This is because it is important to see “what sort of difference it would have made for the

explanandum if the factors cited in the explanans had been different in various possible ways”

(Woodward, 2003, p. 11).

Bokulich rejects the causal approach and favours a structural interpretation of

counterfactual information, or depth. In fact, she aims to give an account that can capture the

39

explanatory value of the highly-idealized, non-causal models that are not captured by

Woodward’s account. She claims that semiclassical models are not deemed explanatory on a

causal account because the entities involved (electron trajectories) are fictional and have can no

real causal power. As will be shown in more detail below, the morphologies (scarring patterns)

of the quantum systems of interest correlate strongly with the particular periodic, or repeating,

orbits of semiclassical mechanics, but the orbits cannot be said to cause the wavefunction

distributions, even though there is a reliable dependency relation. Bokulich argues that none of

the three main types of accounts of explanation (causal, covering law, or mechanistic) can

capture the way semiclassical models explain quantum phenomena, and offers her own account

of structural model explanations as a supplement. This account highlights the structural

similarities between the real world system and the idealized or fictional model. Bokulich argues

that structural model explanations are ones in which there is a pattern of counterfactual

dependence among the variables of the model, which can be measured in terms of w-questions,

and that this dependence is a consequence of the structural features of the target system

(Bokulich, 2008, p. 145).

In developing her account, Bokulich draws on a suggestion made by Margaret Morrison

that explanation has to do with structural dependencies (Morrison, 1999). Similar ideas have

been developed by John Worrall, James Ladyman, and others (Worrall, 1989; Ladyman, 1998;

French & Ladyman, 2003; Esfeld & Lam, 2008), but Bokulich’s account does not draw heavily

on these. Rather, Bokulich offers three general requirements for a structural model explanation,

which I have paraphrased and enumerated as follows:

E1. The explanation makes reference to a scientific model, M.

This first criterion specifies that the explanation is a model explanation and not a covering law or

mechanistic explanation. The structural aspect of the structural model explanation comes from

the second criterion, which says:

E2. M is explanatory by showing how there is a pattern of counterfactual dependence of

the relevant features of the target system on the structures represented in M.

This is intended to determine which models are genuinely explanatory by ensuring that an

explanatory model bears a close structural similarity to the counterfactual structure of the target

40

system. This structural ‘isomorphism’, as she calls it, is given an objective measure in terms of

w-questions (Bokulich, 2008, p. 145). The final criterion states:

E3. There must be a justification that specifies the domain of the application of M.

This is what she refers to as the justificatory step, intended to specify “where and to what extent

the model can be treated as an adequate representation of the world” (Bokulich, 2008, p. 146).

The bulk of what follows will assess the success of the structural criterion, but some concerns

about E3 will be raised in 2.4.3.

The worry for E2 is that all models exhibit a pattern of counterfactual dependency, and

that therefore E2, satisfied by all models, is an idle wheel. Bokulich herself does not think that

structure can distinguish explanatory from non-explanatory models, and instead relies on E3.

But, if E2 is an idle wheel and E3 is little more than a judgment about what is an adequate

representation in contemporary science, then the notion of explanation is unanalyzed – what is

explanatory is what is considered explanatory by the current state of science. However, I think

that there is a more promising avenue for structure: I argue that the criterion can in fact show

preference for certain models over others when employed in a comparative approach. However, I

also argue that this is unhelpful for Bokulich as the criterion shows preference for models of

fundamental theory over those of highly-idealized models. I return to these approaches in 2.4.2,

but now let us turn to the case study of structural model explanations in practice.

Bokulich applies her criteria for explanation to some cases of semiclassical mechanics as

part of a larger project of reconceiving the intertheoretic division between the quantum and the

classical. She argues that semiclassical mechanics can be genuinely explanatory of certain

quantum systems even though they are either deemed non-explanatory or fall outside of the

range of other accounts of explanation. The reason seems to be that the models of semiclassical

mechanics are non-Galilean, or highly idealized. To reiterate, idealizations that are non-Galilean

cannot be de-idealized to smoothly approach the real-world system. Many of these systems

identified by Batterman have singular limits which preclude this possibility. These models then

lack the approximate representation that is traditionally thought to justify their use in

explanation. However, as mentioned above, Batterman, Bokulich, and others argue that this does

not preclude explanation.

41

Models of semiclassical mechanics are highly idealized in that it is not possible to

recover the quantum models by removing approximations and de-idealizing. If the semiclassical

models have explanatory power, it cannot be due to an underlying causal mechanism of which

they are an approximate representation recoverable via de-idealization. Bokulich thinks that

semiclassical models of quantum wavefunction scarring are precisely the kinds of structural

explanations that Woodward’s account cannot consider explanatory. This is why she allows that

the justification of the application of the model to quantum phenomena (E3) can be top-down

from theory, rather than bottom-up where it would be smoothly recovered in Galilean

idealization. For semiclassical mechanics there is no smooth approximation, but there are top-

down theories that can be used to model quantum features in classical terms.

Batterman was the first to argue that semiclassical appeals to classical structures in

quantum phenomena at the asymptotic limit between the two can be explanatorily important

(Batterman, 1992, 2002b). Bokulich claims that structural explanations are actually quite popular

in mechanics, where appeals to structural restrictions can account for certain aspects of systems.

She argues that semiclassical mechanics can be an important interpretive and explanatory tool

for certain quantum phenomena, specifically in the subfield of quantum chaos. Classical chaos is

found in a great number of systems in which there is an extreme sensitivity to initial conditions,

such that an immeasurably small difference in two initial conditions may result in an exponential

divergence between them. Of course, this kind of extreme sensitivity to initial position and

velocity has no part in quantum theory, but quantum models that also describe these systems are

expected to exhibit something like chaos themselves. According to Bokulich and Batterman, one

expects to find a correlate of classical chaos in quantum systems. Due to the agreement between

quantum and classical predictions at the appropriate limit, there ought to be quantum systems

that underlie classically chaotic systems as well (Batterman, 1992, pp. 51-52; Bokulich, 2008).

One of Bokulich’s strongest examples is that of quantum wavefunction scarring in

systems known as quantum billiards. These are systems where the wavefunction of a particle

inside a stadium-shaped enclosure exhibits unusual patterns, which are called scars. Studies of

these quantum billiard systems by means of semiclassical mechanics and cellular automata have

revealed that there is an unusual accumulation of the wavefunction density along the trajectories

that would be periodic orbits in a classical system (C. King, 2009). The strong correlation

between the classical orbits and the observed quantum phenomenon makes these systems useful

42

for studying the quantum systems underlying classical chaotic systems. Bokulich argues that

these semiclassical approaches can be genuinely explanatory of the quantum scarring

phenomenon.

Work on these stadium billiards was introduced by Leonid Bunimovich (Bunimovich,

1974). In the classical billiard systems, a stadium shaped enclosed space is inhabited by a free-

moving particle whose trajectory is mapped. The boundary is defined by two semicircles

connected by parallel lines, and the particle suffers specular reflections with the boundary. The

shape of the enclosure has been since shown to be chaotic no matter how short its parallel

segments are and to exhibit other interesting properties (Bunimovich, 1979; Bleher, Ott, &

Grebogi, 1989). One implication of these results is that a mapped trajectory of the particle

generally displays an irregular pattern (Figure 1). This irregular pattern eventually leads to a

uniform distribution of trajectories throughout the space.

Figure 1 A typical example of a classical chaotic trajectory of a particle in a stadium shaped enclosure

(Stöckmann, 2010).

However, the chaos of this system is intermittent, as there are certain special initial

conditions that lead to periodic orbits in which the motion of the particle constantly repeats itself.

There are certain starting positions and velocities that will not result in a uniformly distributed

stadium, but exhibit a simple pattern of repeated motion. This pattern can occur in different

shapes including a rectangle, a vee, and a bow tie, among others, and are called periodic orbits

(Figure 2).

43

Figure 2 Examples of periodic orbits in a stadium enclosure (Heller, 1986).

These periodic, or bouncing ball, orbits exhibit a “stickiness” on nearby chaotic

trajectories. The nearby trajectories, though not exactly periodic, exhibit quasi-regular behaviour

for long periods of time. When mapped out, this stickiness results in an accumulation of

trajectories in the near vicinity of the periodic orbits, due to the vanishingly small local

Lyapunov exponent, which gives the strength of the scar (Dettman & Georgiou, 2011). This

property of stickiness is also exhibited in other shapes of dynamical billiards such as in drive belt

shapes, where the semicircles are of different sizes, and in various mushroom shapes, such as the

one seen in Figure 3.

Figure 3 A mushroom shaped billiard system that clearly exhibits intermittent chaos. There is stickiness in

the vicinity of the periodic orbits and a region of diffuse chaotic trajectories elsewhere (Dettman & Georgiou,

2010).

What is very interesting about these systems is that because there are no trajectories in

the quantum analogs of these systems, one would expect to be unable to distinguish the chaotic

trajectories and the periodic orbits. Without orbit theory and sensitivity to position and velocity,

there is no obvious reason to expect these particular strong patterns in the quantum wavefunction

44

density. The interesting fact is that the quantum scarring phenomena actually converge on the

classically stable periodic orbits. It is almost as if the quantum system is sensitive to which

classical trajectories are periodic orbits and which are chaotic – a distinction that relies on the

details of the particular trajectory.

The structural model explanation claim is that the phenomenon of quantum wavefunction

scarring is best explained with the semiclassical model of the particle’s behaviour. On the

semiclassical account, the shape and size of the enclosure leads to certain periodic orbits being

favoured and exhibiting stickiness on nearby trajectories. And so as one changes the shape, the

allowed periodic orbits change in predictable ways, and the measured quantum wavefunction

distribution also changes accordingly. The semiclassical model is highly idealized and non-

representing. The use of electron trajectories, which are false of the quantum system, is justified

for Bokulich by means of Gutzwiller’s periodic orbit theory, which is a method of approximating

the density of quantum states from classical periodic trajectories. Gutzwiller’s theory specifies

how the behaviour of a Gaussian wavepacket (x,0) can serve as accurate solutions to the time-

dependent Schrödinger equation, and thus how the allowed classical periodic orbits correspond

to the accumulation of wavefunction density (phase interference) observed as the scarring

phenomenon (Heller, 1984). By considering the autocorrelation function of a Gaussian

wavepacket, (t)|(0), on a phase space point associated with a periodic orbit, one can see an

increase as the wavepacket overlaps with its initial state. The Fourier transform of this function

can be used to calculate the quantum spectrum. If the wavepacket is not initially on a phase space

point associated with periodic orbit there will be no pattern of increase in the autocorrelation as it

propagates, and hence no significant accumulation of the wavefunction in that region (Bokulich,

2008, pp. 128-129).

Bokulich does not argue that quantum mechanics cannot predict the scarring

phenomenon. In the quantum analog, the scarring phenomena can be reproduced by means of the

dynamics of the time-dependent Schrödinger equation with Dirichlet boundary conditions, so the

function vanishes at the walls. Rather than a classical bouncing ball in a stadium, the model

features a wave packet propagated through an infinite potential well with a stadium shaped

boundary. In a similar manner to the simpler rectangular infinite well, one can show that the

system will lead to ordered wavefunctions, and exhibit a phase interference pattern observed as

the scarring phenomenon (Figure 4). As the system evolves, the wavefunction settles on a

45

periodic solution. The scarring occurs when the probability amplitudes overlap in certain areas as

the wavepacket propagates throughout the space and reflects off the boundary (C. King, 2009).1

Figure 4 Amplitude contour maps of eigenstates that display a strong correspondence with the periodic orbits

(Heller, 1986).

Bokulich is primarily interested in the explanatory potential of the semiclassical model.

She argues that classical periodic orbits, though fictions – false of the quantum system – are

explanatorily relevant to the phenomenon of quantum wavefunction scarring. By falsely

assuming that the particle travels along a classical continuous space and time trajectory, one

correctly expects to find certain scarring patterns in quantum billiard systems, which one would

not expect prima facie in a quantum system. She argues that this example is a case of bona fide

structural model explanation. This example is not an outlier case, but one of many in which

Bokulich reaches the same conclusion about explanation; including the conductance peaks of

ballistic quantum dots, the orbits of Bohr’s model of the atom, and the resonance peaks of the

Rydberg electrons (Bokulich, 2008).

For Bokulich, these examples suggest that there is a “dynamical structural continuity”

between the classical and quantum theories. Because of this she argues that semiclassical

1 For more information about the quantum models, simulations of the scarring phenomenon, and ergodic and unique

ergodic properties of classical and quantum billiards, see (McDonald & Kaufman, 1979; Heller, 1984, 1986;

Gutzwiller, 1990; Tomsovic & Heller, 1993; L. Kaplan & Heller, 1999; Tao, 2007; C. King, 2009; Dettman &

Georgiou, 2010)

46

fictions, in this case the classical periodic orbit theory applied to a quantum particle, can serve to

give counterfactual information about the quantum system. The closed orbits are not real, in the

sense that the particle is not actually travelling in a classically defined orbit with position and

velocity. Bokulich does not want to argue that the trajectories are real, but rather that they are a

special kind of fiction that is also explanatory: “These closed and periodic classical orbits can be

said to explain features of the spectral resonances and scarring insofar as they provide a

semiclassical model of these phenomena” (Bokulich, 2008, p. 140). The dependence of the

scarring phenomenon on the classical orbits conveys physical insight, or structural information,

on the underlying quantum dynamics.

Of course, in order for Bokulich to claim that there is a genuine explanation here, the

semiclassical model must satisfy her three criteria E1-3. And it can be quickly shown that it does.

The explanation makes reference to a scientific model, Gutzwiller’s periodic orbit theory, and so

it satisfies E1. The semiclassical model exhibits a strong counterfactual dependence of the

probability density in the billiard systems on the particular periodic orbits. It satisfies E2, the

structural criterion, because one is able to say how the wavefunction morphology inside the

stadium would change if the periodic orbit had been different, or if the shape of the stadium had

been changed. This semiclassical explanation is also justified in being applied to this domain

(E3) because Gutzwiller’s periodic orbit theory specifies how to model features of quantum

dynamics with classical trajectories – how to get real-world information from the information in

the model. So for Bokulich, this semiclassical model qualifies as explanatory.

As we have seen, Bokulich does not claim that quantum mechanics alone cannot predict

these scarring phenomena. Rather, her claim is that its explanations are deficient because they do

not provide as much counterfactual information about the system, which gives us physical

insight into the system and grants understanding. In order to get a more concrete sense of the

counterfactual information a model gives about the system, she makes use of w-questions and

Woodward and Hitchcock’s notion of explanatory depth, mentioned earlier. The more w-

questions a model answers, the more structural information it gives, the deeper the explanation it

provides (Bokulich, 2008, p. 152). Bokulich argues not only that the semiclassical models are

explanatory, but that the semiclassical models actually provide deeper explanations than the fully

quantum ones: “More importantly, the semiclassical models provide more information about the

structure of the quantum dynamics than do the fully quantum calculations. That is, the

47

semiclassical model allows one to answer a wider variety of w-questions, about how the system

would behave if certain parameters were changed…” (Bokulich, 2008, p. 154). Reference to the

classical structures is eliminable in that one can make simulations that exhibit scarring

phenomena from a quantum model. Nonetheless, Bokulich argues that “without knowledge of

the classical orbits, our understanding of the quantum spectra and wavefunction morphologies is

incomplete” (Bokulich, 2008, p. 154).

For Woodward and Hitchcock, the range of w-questions that a model can reliably answer

about a phenomenon is directly related to the model’s range of invariance. Brad Westlake has

identified many distinct measures of invariance in their account (Weslake, 2010). For instance, a

model can be more invariant if it is more accurate within a certain range; invariant under a wider,

or more continuous range of interventions; invariant under a wider range of different ways the

interventions may be performed; or under a wider range of background conditions. “What they

have in common is that they provide the resources to describe a greater range of true

counterfactuals concerning the possible changes to the system in question – that is, to answer

more w-questions…” (Weslake, 2010, p. 278). All of these different kinds of invariance are

relevant to the structural information Bokulich is looking for in E2.

Bokulich claims that the semiclassical model of wavefunction scarring gives counterfactual

information about the quantum system, and further that “there can be situations in which less

fundamental theories can provide deeper explanations than more fundamental theories”

(Bokulich, 2008, p. 153). Given that there are local models framed entirely in quantum terms that

can predict these scarring phenomena, if Bokulich wants to argue that the semiclassical model

provides deeper explanations than the fully quantum one, then she has to show that the

semiclassical model can answer a wider range of w-questions, in the full sense described above.

2.4.2. Two Approaches for Assessing Structure

I have shown how Bokulich’s account aims to capture the way highly-idealized models like

those of semiclassical mechanics explain, and I now turn to examine a serious concern about the

claim. In order to make the best case for a structural criterion, I present two possible ways of

assessing structure: a direct and straightforward measure of the number of answers to w-

questions (or w-answers) with a certain threshold for explanation, and an indirect measure for

comparing the different classes of w-answers that two competing models provide. I aim to show

that both avenues for assessing structure are unsatisfactory. The first approach proves impossible

48

and leaves E2 an idle wheel. The second, when it is possible, does not side in favour of the

highly-idealized model, which it was intended to do. Further, I suggest that this is not something

particular to w-answers but applies generally to any similar measure for structure. Now, let me

first turn to the original worry about structure, before presenting the more promising view.

In a review of (Bokulich, 2008), Belot and Jansson express a concern that once the

account of structural model explanations allows for such fictions as classical trajectories in

quantum systems, it will be unable to reject models that are widely considered non-explanatory,

such as the models of Ptolemaic astronomy. They ask, “what is to stop you from viewing the

Ptolemaic model of the solar system as giving an adequate structural model explanation of this

phenomenon? Indeed, an appeal to the Ptolemaic model on this question would appear to satisfy

all four requirements for a structural model explanation” (Belot & Jansson, 2010, pp. 82, 83).

The worry is that once she opens the door up to explanatory fictions her criteria are not strong

enough to debar non-explanatory fictions, such as planetary epicycles. Bokulich (2012) is

explicit in wanting to allow for the fictitious electron trajectories in quantum billiard systems, but

not the fictitious epicycles of Ptolemaic astronomy. As is well known, the Ptolemaic model of

the solar system makes use of epicycles, on which there is a strong counterfactual dependence of

the apparent retrograde motion of the planets across the night sky as seen from Earth. At first

glance, Belot and Jansson are right to worry that epicycles and electron trajectories both satisfy

E2, but Bokulich argues that her account is capable of admitting one and not the other.

Both of the models satisfy E1, insofar as they reference scientific models: the geocentric

model of the solar system and the semiclassical model of the quantum billiards. The models also

seem to satisfy E2. They are counterfactually reliable under a range of conditions. The Ptolemaic

system has trigonometric tables of chords used for calculations, and these give us counterfactual

information about the visible solar system, and as Bokulich showed, the semiclassical model

gives counterfactual information about the scarring patterns. In fact, according to Bokulich, w-

questions are unlikely to be able to distinguish semiclassical explanations from Ptolemaic ones.

What she argues instead is that “the difference between explanatory and nonexplanatory models

is determined by something like a contextual relevance relation set by the current state of

scientific knowledge,” (Bokulich, 2012, p. 733). Thus, it is only on E3 that a distinction can be

made. In order to examine this claim, I will first assess the models according to the structural

criterion E2, and move to the third criterion E3, in 2.4.3.

49

The structural aspect of Bokulich’s account (E2) requires that the “model must explain

the explanandum by showing how there is a pattern of counterfactual dependence of the relevant

features of the target system on the structures represented in the model” (Bokulich, 2008, p.

146). Unlike how it is framed in Woodward and Hitchcock, for Bokulich there is no need to

insist that the counterfactual dependency is representative of causal relations. Rather, she claims

what is being exhibited is the structural dynamics of the system. In order to most accurately

assess the satisfaction of E2, one needs to actually measure the w-answers that a model can

provide.

Unfortunately, obtaining a measure of the number of w-questions a certain model can

answer is not straightforward. Let us examine one approach, which is simply to count the

number of w-questions answered by a model and if the number is sufficiently large then the

model is explanatory. However, two problems with this immediately arise. First, the Ptolemaic

system, for instance, has methods of calculating the positions of the bodies of the solar system

for any given day, for any place on Earth, including not just positions in the night sky, but

eclipses, solstices, equinoxes, and so on. Importantly, these bodies have cycles and epicycles that

are geometrically continuous. This means that the model provides an infinite number of w-

answers, as there are an infinite number of points on the lines of the spherical trigonometry. And

so the number of w-questions answered by even a non-explanatory model is infinite. One could

make a stronger case for this approach by articulating a method which looks at the size of the

parameter space, the domain of values allowed by the model. This might allow one to get more

varied information about the amount of counterfactual information a model provides. This is an

interesting approach, however, even if one were equipped with a good quantitative measure of

the structural information a model provides, the second problem remains: there is no principled

way of determining exactly how many w-answers or what size of measure would count as

“sufficient” for explanation. Without some larger framework for determining exactly the

minimum size, a proposed threshold seems arbitrary.

And so this approach fails, as all models would equally satisfy E2, and thus the entirety

of judging whether something is explanatory falls on E1 and E3, which without E2, amounts to

nothing more than reflecting current judgments about explanation. In this case, there is no

analysis of explanation and no reason to think of these as structural explanations. However, I

think there is a more promising comparative approach. While Bokulich does not explicitly frame

50

w-questions in a comparative way, I suggest that the Ptolemaic model does answer fewer w-

questions than, for example, the Newtonian model, and it can be debarred in that fashion. As we

have seen, a quantitative method for counting w-answers is not possible, and so a simple

quantitative comparison cannot be made either. Bokulich suggests examining the classes of w-

questions, and I believe that there is a sense in which an intuitive comparison of the classes of w-

questions is possible. Consider a relative formulation of the structural criterion E2*, which states

that M1 provides deeper explanations than M2 if M1 gives more counterfactual information than

M2, which is given in terms of classes of w-questions.

Because there is a lot of overlap in the information one gets from the Ptolemaic and

Newtonian models of the solar system, an argument could be made that the Ptolemaic model has

a narrower scope. In fact, because the Newtonian model can give all the w-answers that the

Ptolemaic model provides, but also a lot of additional w-answers as well, the Ptolemaic w-

answers are a subset. For instance, one can get reliable w-answers about how the orbits would be

different if the planets or the Sun had different masses, or if the planets were different distances

from the Sun, and so on. But as Woodward and Hitchcock were careful to point out, there is

more to explanatory depth than scope. In this case, the Newtonian model is dynamic, which

many argue gives a richer, deeper account of the planetary motions. It is not simply a kinematic

model that describes behaviours and matches observations, it makes use of forces and torques, to

give the causes of things. It seems that intuitively, the Newtonian model answers more different

kinds of w-questions more accurately. It gives stable and robust counterfactual information. The

Ptolemaic model does not fare well in this kind of comparison, so perhaps a comparative

structural criterion can be capable of distinguishing explanatory from non-explanatory models.

This could provide a defense for this structural approach from criticisms like those raised by

Belot and Jansson.

If this comparative framework works for Ptolemaic astronomy, is it also capable of

showing that, as Bokulich argues, the semiclassical model provides a deeper explanation of the

scarring phenomenon than the fully quantum model featuring the Schrödinger equation? Well,

when one returns to the semiclassical model and attempts to compare the w-answers with those

provided by the local quantum model, the comparison does not seem to lead to the conclusion

that the highly-idealized model provides deeper explanations.

51

The semiclassical model can give counterfactual information about the distribution of

probability densities in the enclosure in straightforward way. There is a certain range of

questions that can be answered about the dependence of the wavefunction morphology on the

classical orbits. There is a striking correlation between the trajectories predicted by the classical

theory and the observed phenomenon. The semiclassical model is indifferent to the details and

particulars of the quantum dynamics and allows certain features like scarring phenomena to be

highlighted. Information about why particular scarring patterns, as seen in Figure 4, occur is

given by the semiclassical model, so the argument would go, because it is easily capable of

accounting for the chaotic and the particular periodic trajectories, and can show how the

quantum scarring would change if things (the periodic orbits in Figure 3) had been different. It

does seem that here too, “rather than obscuring the genuine mechanisms at work, this

idealization actually brings them into focus” (Batterman, 1992, p. 64). So it seems plausible that

there could be a class of w-questions that are more deeply, or at least more intuitively, answered

by the highly-idealized model.

Bokulich argues that the non-fundamental model can provider deeper explanations. But

the semiclassical model is not the only model that accounts for the scarring phenomena. The

Schrödinger equation can be used in a quantum model to achieve all the results that are obtained

in the semiclassical model. In fact, that is how we know the semiclassical model is successful.

As Bokulich freely admits: “one can “deduce” the phenomenon of wavefunction scarring by

numerically solving the Schrödinger equation,” the problem, she argues, is that “such an

explanation fails to provide adequate understanding of this phenomenon” (Bokulich, 2008, p.

151). Understanding, for Bokulich, is given by the physical insight, or information in terms of w-

questions that a model answers. And so, this means that the non-fundamental model provides

more w-answers. However, Weslake points out that the w-question notion of explanatory depth

favours more fundamental generalizations: “The fundamental laws are those generalizations that

are maximally accurate, robust, continuous, stabile, insensitive, and portable” (Weslake, 2010, p.

278). This is what many assume to be the case and is what Woodward and Hitchcock assume

when they explored the notions of depth and counterfactual information (Woodward &

Hitchcock, 2003b).

It is worth unpacking this assumption. What is crucial is that the quantum model can

account for all the same scarring phenomena as the semiclassical model. Of course, the quantum

52

model cannot answer specific, classically-framed w-questions about trajectories or Lyapunov

exponents, because the concepts are unavailable to the quantum model. But, while the

fundamental theory loses w-questions in terms of trajectories, it gains “corresponding” w-

answers framed in terms of wave packet propagation. The quantum model is capable of giving

counterfactual information (w-answers) about the morphology of the scarring patterns.

But the local quantum model is also able to provide very deep explanations of these

phenomena, because as a model of a fundamental theory it can provide so much counterfactual

information. My contention is that the semiclassical to quantum relation of w-answers is the

same subset relation as that between Ptolemaic and Newtonian w-answers. Though this

comparative approach is not strictly quantitative, intuitively, the quantum model can be seen to

give at least as many w-answers as the semiclassical model, given that whenever the

semiclassical model can give information about wavefunction morphology, so can the quantum

model. However, the quantum model can also provide additional counterfactual information

about a variety of w-questions that can be answered by the quantum dynamics of the system in

terms unavailable to the semiclassical model. The semiclassical method is an approximation,

which is poor in certain conditions and fails in others. While a quantitative analysis of the

closeness of the semiclassical approximation is more properly suited to a physics paper, it is

clear the equations of the fundamental model that are more accurate and more strongly invariant,

in the many senses outlined above. An approximation can be very useful, but it does not contain

more information.

However, this comparative approach has an interesting corollary. If there were a domain of

phenomena in which the more fundamental theory could not derive the desired results, or

reproduce the phenomena, then the best available explanation would be given by the less

fundamental theory – which is to say that hope is not lost for this approach in capturing the way

some highly-idealized models explain. In this particular case however, Bokulich admits that a

quantum model can account for the scarring phenomena described by the semiclassical model.

And so it turns out that even though the classical trajectories can answer interesting w-questions

about the particular morphologies of the wavefunction scarring, models from the more

fundamental theory will always win out in terms of w-questions when they can account for the

same phenomena, because they can answer at least as many classes of w-questions.

53

Someone working from Woodward’s account might accept that these highly-idealized

models are explanatory, but less explanatory than models of more fundamental theories that offer

much deeper explanations, as long as they satisfy his criteria for explanation by exhibiting a

strong enough degree of invariance under a range of interventions. However, this will not work

for Bokulich in this case, because it will not favour the models of semiclassical mechanics over

those of quantum mechanics because they are both capable of accounting for the scarring

phenomena. It is important for Bokulich, not only that semiclassical models be explanatory, but

that they actually provide deeper explanations of some quantum phenomena than the fully

quantum explanations.

While this measure of structure sides in favour of models of more fundamental theories

when they can predict the same phenomenon, when they cannot, an intuitive sense of which

model answers more classes of w-questions does not seem to have any bearing on the

explanatory depth of one model or the other. For example, if one compares a semiclassical model

with a Ptolemaic one, regardless of how the w-answer balance tips, the structural criterion still

has no real bearing on whether the models of semiclassical mechanics are themselves

explanatory. When there is no overlap in the domain of the models, the comparison is not

helpful. I contend that even if there were a quantitative way to measure the structural similarity

using something other than classes of w-questions, these problems remain. Imagine it were

possible to give a compressed scalar rating of all the complex representations of structural

similarity, given by a complicated process of calculations and perhaps insights from measure

theory. Now suppose that the Ptolemaic model of the solar system was given a rating of 4 and

the semiclassical model of quantum wavefunction scarring a generous 8. Even though it received

a higher rating than the Ptolemaic model, it is still reasonable to ask “does the semiclassical

model explain quantum wavefunction scarring?” And so it does not seem that there can be any

way that such a comparative framework can provide a general criterion for explanation. This

problem remains whether one uses w-answers or some other representational measure of

structure. It is only when both models can reproduce the same phenomenon that a meaningful

comparison can be made, but when it can be made, it does not favour the highly-idealized model.

The main worry for a structural criterion for explanation is that some measure of

structural similarity can be given to almost any model, no matter how inaccurately it represents,

or how little structural information it provides. And so if one wishes to debar the worst of these

54

then either a threshold must be drawn, or a comparison must be made. For these semiclassical

models, a threshold proves unworkable. The comparison, when possible, sides in the favour of

the models of the more fundamental theory and not the highly-idealized model. Further, this

comparison is only helpful for models with overlapping domains, and leaves unanswered the

question of whether a particular model explains its target phenomenon. It does not seem possible

to conclude that the semiclassical model provides deeper explanations.

This does not mean that depth is incompatible with autonomy. Weslake, for instance,

provides a different abstractive account of depth, which he argues is more promising for

preserving autonomy (Weslake 2010). An assessment of Weslake’s arguments or of the

possibility of incorporating different notions of depth into a structural account, fall outside the

scope of this paper. It is also possible that one not choose w-questions to form the basis of a

measure of structure. Others have attempted to capture explanatory depth by other means

(Schupbach and Sprenger 2011; Strevens 2008; Weslake 2010), but few if any have offered

concrete measures of structure. One of the aims of offering a structural account is to preserve

autonomy of higher level explanations, and to allow some highly-idealized and non-causal

models, such as those of semiclassical mechanics, to be considered explanatory. I admit that

some notion of depth might be able to preserve explanatory autonomy, but a notion of structure

that depends on isomorphism or counterfactual information will favour fundamental models and

confront the difficulties of the approaches I have presented.

2.4.3. E3: The Justificatory Step

Bokulich’s solution to the problem posed by Belot and Jansson was to debar Ptolemaic

epicycles, not with the structural criterion E2, but with the justificatory criterion E3. And so in

this section I will analyze this criterion of Bokulich’s account and raise some concerns about it

playing the major role in distinguishing explanatory from non-explanatory fictions.

This justificatory step has three aspects, which I have taken the liberty of enumerating as

J1-3. J1 is a contextual relevance relation set by the contemporary state of science, which

ensures that scenarios like falling barometers causing storms are simply not even candidate

explanations. The justification also involves an articulation of the domain of applicability, J2,

wherein it is an adequate representation of the system. To satisfy this there must be either a top-

down or bottom-up justification of the model’s use, as I described above. Lastly, and closely

related is J3, a translation key of sorts that allows information about the model to be translated

55

into conclusions about the real system. There must be some reason why information gained in

the model is applicable to conclusions about the world. For example, Gutzwiller’s periodic orbit

theory specifies how the semiclassical trace formula is related to the actual observed

morphologies in the quantum stadium billiard (Bokulich, 2012, p. 736). E3 taken together is

something like the explanatory standards of current science. It ensures that the model makes

reference to the right kinds of accepted entities, states, and processes, and that the relation

between the model and the real system is not merely accidental. For Bokulich, it is on E3 that the

distinction between explanatory and non-explanatory lies.

It is only on the third criterion, E3, that the Ptolemaic models will be debarred according

to Bokulich. The geocentric model and its epicycles are not at all adequate representations of the

real structure of solar system, as determined by the current state of scientific knowledge, and so

not deemed relevant to the explanation of planetary motion (Bokulich, 2012, p. 735).

Explanatory fictions “represent real entities, processes, or structures in the world, while [non-

explanatory ones] represent nothing at all (Bokulich, 2012, p. 734). She wants to allow for

fictions to be explanatory, but only fictions that count as adequate representations. In the context

of these two examples, the Ptolemaic model is non-explanatory because the orbits are not

adequately representative of the real structure of planetary motion: “given the relevance relation

set by the state of contemporary science, epicycles are irrelevant to the explanation of retrograde

motion. This is not simply because they are fictional but, rather, because they fail to be an

adequate fictional representation of the real structure of our solar system,” whereas “the classical

periodic orbits are able to capture, in their fictional representation, real features of the quantum

dynamics…” (Bokulich, 2012, p. 735). So the representational inadequacy is what debars

Ptolemaic epicycles.

In her response to Belot and Jansson, she says: “although the range of w-questions that a

phenomenological model can answer will typically be more limited, scope alone cannot

distinguish between explanatory and phenomenological models” (Bokulich, 2012, p. 733). She

offers instead the idea that the current state of scientific knowledge precludes the possibility of

Ptolemaic epicycles being counted as explanatory, in the same way that it ought to preclude

falling barometers causing storms – neither satisfy J1. It was shown in the previous section that

E2* also debars both Ptolemaic epicycles and semiclassical models. Now it can be seen that E3

debars Ptolemaic epicycles but not, according to Bokulich, semiclassical models. Semiclassical

56

models are not so obviously inadequate as to be excluded from the explanatory store, like

shadows causing flagpole heights, and so they satisfy J1. For J2 and J3, the semiclassical models

of interest employ Gutzwiller’s periodic orbit theory to justify their application and provide a

means of getting real-world information from the fictional model. And so the semiclassical

models seem to satisfy E3 as a whole and are justified in being used in these systems exhibiting

quantum chaos, even though it was shown that they did not satisfy the comparative formulation

of E2.

In the remainder of this section, I will raise three worries about E3 and about this kind of

criterion as the main deciding factor for explanation. The first worry is that even though she

insists that electron trajectories in semiclassical models are explanatory and Ptolemaic epicycles

are not, it is not clear that, in distinction from epicycles, classical electron trajectories are

representative of the true electron dynamics, of the real structure of the quantum systems, as she

claims (Bokulich, 2012). Part of the requirements of E3 is that entities and processes of the

model are considered by scientists to be potentially relevant to the explanation (J1). Earlier, I

conceded that the semiclassical models should not be dismissed from potential explanations

outright, but this does not imply the positive claim that they do capture real quantum structures.

Consider the fact that the predictive success of semiclassical models is rather unexpected. This is

so precisely because they are not true descriptions of the systems. It may be that there is a certain

range of counterfactual information about the systems’ morphologies that can be gathered, but it

is not readily understood why it is that the dependency relation holds. Given only the full

semiclassical explanation, it is still a bit mysterious why the quantum effect would be dependent

on the classical trajectories. If one is able to derive this phenomenon and render it expectable on

a fully quantum picture, that mystery would disappear. This seems to suggest that the real

structure of the system is only given in a fully quantum picture, in the same way that the

numerical coincidences of Ptolemaic calculations are revealed by more fundamental theories.

The second worry is that because this is supposedly a structural explanation, a lot should

depend on the satisfaction or degree of satisfaction of E2, as Bokulich’s own formulation

implies. But, this does not seem to be the case. E2 is not capable of doing the work of showing

how a model is explanatory, since on one interpretation all models are explanatory and on the

other it debarred both Ptolemaic models and those of semiclassical mechanics in favour of

models with broader scopes from more fundamental theories. Due to this, E3 has to do most of

57

the heavy lifting. However, if E3 is largely responsible for maintaining a threshold for

explanation, then there is not much of a sense in which these three criteria taken together are

independent criteria for explanation. The deciding factor is what satisfies E3, i.e. what is

consistent with that currently considered to be explanatory in science. The structural criterion

that was intended to pick out which models were genuinely explanatory by showing whether the

models exhibited the relevant structural properties of the system failed to do so. In order to make

a strong case that semiclassical mechanics can provide structural model explanations, the

structure that allegedly links the models to the real-world system should determine that. In a

structural explanation, the structural criterion should not be an idle wheel.

The last related concern is not only that E2 should distinguish explanatory from non-

explanatory in a structural explanation, but that E3 is too context sensitive to do this. It seems as

though E3 could be determined, or estimated, with structural information. If one wanted to assess

the adequacy of a model’s depiction of reality, determine whether its relation was numerological,

or correlational, and know if the model’s information is applicable to the real world system, then

its ability to give a wide range of reliable counterfactual information about that system seems a

reasonable measure. This information is something that the model can provide on Woodward’s

account, because it is explicitly manipulationist. But because Bokulich does away with the causal

interventions and only imports the notion of explanatory depth, this must be added on as a

separate criterion and loses objectivity. On Bokulich’s account, there can be no interventions to

separate the correlational from the causal, instead it falls on the scientific community to decide if

it is adequate. E3 is not meant to employ the measure of w-questions – it is not a measure of

structural similarity, but a criterion for ensuring that the model is not known to be

phenomenological. The criterion is context sensitive and particular to the details of the model

and the current views in science regarding what explains and what accurately represents. What

counts as an adequate fictional representation (J1) is a moving target, and not likely to be

unanimously agreed on across a discipline.

But even if this were widely agreed upon, there is something missing in this kind of

justification – a degree of normativity. And that even if some scientists, or a majority, do find

these models to be explanatory, there is more to a philosophical account of explanation than

capturing that. The judgments of scientists regarding explanation is important, but it should not

be the only aspect of an account of explanation. An account of explanation should not be merely

58

descriptive, but provide independent criteria capable of assessing putative explanations, as I

argued in 1.4.1. There are benefits to maintaining a high threshold for explanation. It allows the

account to not simply capture the current use of explanation, but to highlight the best explanatory

practices of science. Having an account of explanation that claims that what is explanatory is

what is considered explanatory by a relevant scientific community says very little precisely

because it lacks normativity.

2.4.4. Heuristics and Explanations

Heuristics are powerful tools of science. A model that is heuristic is one that makes calculations

easier or provides a short-cut kind of reasoning. In psychology, a heuristic can be a kind of rule

of thumb or mnemonic. A model that is heuristically valuable can be one that is simple enough to

provide approximate answers with little computational load. Or it can be a model that can reveal

new avenues for research (Hartmann, 1996). A heuristic model, might allow researchers to learn

more about the behaviour or system at hand and discover new, related phenomena as well.

Heuristics are strategies or methods that can be very useful, generally accessible, and widely,

though loosely, applicable. The importance of heuristics in philosophy and scientific discovery

has been noted since at least Popper (1959).

Throughout her work, Bokulich argues for two points that should be considered

separately: 1) classical approaches in quantum mechanics are not only fruitful, but expose a

complex relation between the two that is more than a simple reduction of one to the other.

Regarding the influence of heuristics in the development of quantum mechanics she says that the

correspondences between the two theories “…are continuing to play a role in modern

semiclassical research. Understanding the heuristic ground of these correspondences can also

help us recognize that intertheory relations are not static, but rather are evolving, and are

continuing to be developed and extended in new ways” (Bokulich, 2008, p. 172). This

conclusion is in support of her interstructuralist position on the relation of quantum and classical

theory, and well summed up here:

Unlike Heisenberg and Pauli who thought new insights into physics would come only

through working with our most fundamental physical theory, semiclassical theories –

along with Dirac – think that classical mechanics still has something more to teach us.

More specifically, one of the most important insights of semiclassical mechanics is that

many new discoveries about quantum mechanics can be made by exploring its relation to

classical mechanics.

59

(Bokulich, 2008, pp. 174-175)

Semiclassical mechanics has proved useful and fruitful in recent physics and might still,

but I maintain that this is not enough for it to count as explanatory. Bokulich is also sensitive to

this distinction, which is why she makes a second point directly related to the topic of

explanation which is that: 2) semiclassical mechanics and their fictitious electron orbits can in

fact explain certain quantum phenomena. The first may be convincingly argued for, but I have

shown that in the end, the second relies on nothing other than the models’ current accepted use in

science, as evidenced by testimonials. However, even though scientists may make claims of

explanation in proposing a new method, discovering a mechanism, or more accurately predicting

the behaviour of a system, this does not mean that their collected use of the term is coherent,

useful, or philosophically rigorous. I contend that often when scientists make claims about

explanation and explanatory value, they are actually talking about heuristic value; these

testimonial speak largely to the fruitfulness of semiclassical mechanics.

And indeed, semiclassical mechanics is a fruitful research avenue and it is intuitively

powerful. It allows us to picture and grasp systems that we should not be able to picture, and

frame them in familiar terms. And quite astonishingly, it can give us simple and reliable

counterfactual information about certain quantum systems. On the contrary, Ptolemaic epicycles

are no longer a fruitful avenue of research and their ability to predict seems quite accidental,

almost numerological. When Bokulich argues for the explanatory power of semiclassical

mechanics she concludes from the work of Wintgen, Richter, and Tanner (1992), as well as

others, that it is more than a tool or a method for generating simple, reliable predictions.

Bokulich cites physicists as saying that semiclassical descriptions are desirable because the full

quantum mechanical calculations are cumbersome and elaborate and that the “simple

interpretation of classical and semiclassical methods assists in illuminating the structure of

solutions” (Wintgen et al., 1992, p. 19). Here we see that it is in getting “the structure of

solutions” that the semiclassical methods are most useful. Batterman has argued along similar

lines citing the work of W.H. Miller: “Semiclassical theory plays an interpretive role; that is, it

provides an understanding of the nature of quantum effects in chemical phenomena, such as

interference effects in product state distributions and tunnelling corrections to rate constants for

chemical reactions” (Miller, 1986). These quotations are clearly in favour of the cognitive value

of semiclassical mechanics, and some explicitly state that semiclassical mechanics are

60

explanatory. However, it is important to remember that scientists are unlikely to have in mind a

rigorous and philosophically robust notion explanation, complete with independent criteria.

There are a few reasons for interpreting these testimonials as actually being in favour of

the heuristic value of semiclassical mechanics. First, many of the reasons cited for considering it

explanatory, such as fruitfulness, simplicity, and understanding, more directly support its being

heuristic. The explanatory interpretation is also no longer supported by independent arguments

about the semiclassical models’ structural isomorphism, since I have shown that E2 is inert. It is

not as though these quotes corroborate the findings of a structural model account; they are now

its sole evidence. It is a consequence of the account that I present in Chapter 5 that these

semiclassical models lack a necessary component for explanation, which makes them at best

heuristic.

Bokulich hopes that by expanding on her criteria for explanation and introducing an

appeal to the current state of science she can maintain this distinction. It is true that current

science has turned its back on geocentric models of the solar system, and the Ptolemaic model

has been found to be empirically wanting. Semiclassical mechanics is new by any standards and

new results are published in reputable scientific journals. But more than this is needed is make

the case that one is explanatory and the other is not.

Bokulich wants to place semiclassical mechanics in a special place that is neither full-

blown realism about semiclassical mechanics, nor mere instrumentalism. Semiclassical

mechanics are certainly more relevant to the current state of science than Ptolemaic epicycles

because they are heuristically valuable in providing frameworks for investigating and calculating

quantum systems. I maintain that the value that semiclassical mechanics has is merely heuristic

and not explanatory. Where semiclassical explanations fit her criteria, quantum explanations fit

better, and thus she has not provided an account that can fully capture the way that highly-

idealized models explain.

Conclusion and Discussion

Bokulich has taken bold steps forward in arguing that w-questions can be used to measure

structural similarity. However, this measure proves rather difficult to determine. Neither of the

two approaches I outlined is capable of concluding that the highly-idealized semiclassical model

is explanatory. Even non-explanatory models provide an infinite number of w-answers, and even

61

when employing measure theory, a given threshold seems arbitrary. When comparing, one finds

that a fundamental model wins out in terms of classes of w-answers. Either way, the account

does not capture the way that highly-idealized models explain.

Because the contribution of E2 is negligible to whether a highly-idealized model is

explanatory, the entirety of judging explanation falls on E1 and E3. Indeed, Bokulich herself

does not appeal to structure in order to debar the Ptolemaic explanation, rather she argues that

the model fails to be explanatory because it simply does not qualify as an adequate fictional

representation of the solar system in contemporary science (it does not satisfy E3). The relevance

relation set by the contemporary state of science has precluded explanations of planetary motion

featuring epicycles, barometers causing storms, and shadows causing flagpole heights, but not

necessarily the models of semiclassical mechanics. E3 only reflects our current judgments about

whether a model is explanatory and does not make any claims about why the model ought to be

considered explanatory. E3’s descriptive nature takes away the normative aspect that an account

of explanation ought to have, and has traditionally aimed for. Rather, it is that E3 has to play

such a strong role in judgments about which models are explanatory in light of what I have

shown about E2. I do not think that E1 and E3 alone can provide an adequate account of how

highly-idealized models explain.

Highly-idealized models are common in science and as other have argued (Batterman,

2002a; Wayne, 2011; Rice, 2012, 2013; Batterman & Rice, 2014), there is reason to consider

them explanatory. Developing an account of explanation that more accurately reflects the

explanatory practices of science is a next major step in the philosophy of science. Bokulich tries

to do so by providing a quantitative measure for a structural criterion but it ends up not working

in her favour. She attempts to preserve the autonomy of semiclassical, and perhaps other highly-

idealized explanations, but her notion of structure is representational and cashed out in a measure

of explanatory depth that prefers more fundamental generalizations with strongly invariant

mechanical laws.

62

Chapter 3. Causal Accounts of Explanation

Introduction

This chapter turns to examine causal accounts of explanation. In 2.2, I made the case that higher-

level models can be explanatory by citing arguments made by Woodward against the causal-

mechanical account provided by Salmon and the arguments made by Mitchell regarding the

complexity of biological phenomena (Salmon, 1984; Woodward, 1989; Mitchell, 2003). This

third chapter examines how the particular accounts of Woodward, Strevens, and Mitchell

propose to treat these high-level explanations. In 3.2, I present Mitchell’s account of integrative

pluralism and its proposed solution for explanation given the complexity of biology. Her view of

explanation is that multiple models of a complex system are combined, or integrated, in order to

form an explanation. According to Mitchell, there are facts about the sheer complexity of biology

that preclude any traditional kind of explanation. She makes a case for her integrative pluralism

in part by looking at the self-organization of eusocial insect colonies. I examine her case study to

demonstrate her arguments against other forms of pluralism. In the end, I argue that Mitchell

provides no real framework for performing the integrations she proposes and that the detail and

particularity required for modelling the kinds of complex phenomena she examines actually

preclude the possibility of explanation. Her example of the detailed pluralistic model of the Lake

Erie ecosystem focuses on simulating its behaviour and not explaining why it behaves the way it

does.

In 3.3, I lay out how Woodward’s manipulationist account proposes to handle explanation at

multiple levels. I argue that the possibility of there being models on multiple levels that pass his

criteria for manipulationist causation implies emergent causation, and that therefore there can be

many incompatible causal models all operating in the same real world system. In 3.4, I present

the worry that having multiple models representing the real causal dependency relations brings

with it all the problems of emergentism, including downward causation and overdetermination.

These problems are examined by looking at Kim’s causal exclusion argument and its application

63

to interventionist causation. I then briefly look at the reception of Kim’s arguments and defenses

of non-reductive physicalism (NRP) by Jessica Wilson and others.

An alternative strategy to emergent causation is to favour explanatory reductionism or a kind

of non-reductive physicalism. In 3.5, I review Michael Strevens’ kairetic account of explanation.

Strevens offers an account of depth that is capable of counting high-level models as explanatory

in certain cases, while maintaining that the real causes are all operating at a fundamental level. I

argue that this method avoids any problems of emergentism generated by other causal accounts

but has other drawbacks. I claim that Strevens’ account does not reflect the actual practice of

scientific explanation and is for all but toy examples not implementable.

I finish this chapter by recapitulating the results of this investigation and reviewing

additional concerns that have been raised for causal accounts of explanation in general. The

central aim of this chapter is to present a variety of popular accounts of causal explanation and

show that they face serious challenges. This does not entail that causal approaches cannot

succeed or that the problems are insurmountable. However, exposing these concerns about causal

accounts is a key result in the overall dissertation, which aims to show that a neo-deductivist

approach is more promising.

Scientific Pluralism

Sandra Mitchell explicitly states that pluralism in science is an unsurprising fact (Mitchell,

2003). It is reflected everywhere in the models and methods that are being advanced in science.

It is not unreasonable to ask why explanations of one world would be so diverse. Kuhn said that

it was a mark of the adolescence of science, but Mitchell notes that the diversity and

specialization of science is only increasing. She attributes the diversity, not to the immaturity of

the study, but to the complexity of the subject matter. This is partly what led Feyerabend to

advocate his epistemological anarchy (Feyerabend, 1975). His extreme position has opened up a

more modest middle ground for many kinds of pluralism or disunity accounts of science, such as

those provided by Cartwright, Longino, Harding, Dupré and others (Cartwright, 1983; Dupré,

1983; Harding, 1986; Longino, 1990; Davies, 1996).

Arguments from complexity are common in the literature of the philosophy of biology on

scientific pluralism. As was mentioned in 2.2.2, Cartwright developed a view of scientific anti-

realism based on arguments that it is impossible for laws to perfectly represent any real-world

64

scenario (Cartwright, 1983, 1999). Cartwright concludes that laws are not true and exceptionless

as we ordinarily think of them, but function more like simulacra. Explanations on her account

make use of models, which employ laws that range over abstract objects in the model, but not the

actual phenomena. The objects in the model do not have stable properties, but rather capacities to

act in certain conditions. This is what counterfactual cases of the model represent. For

Cartwright, explanations involve a complexity of causal interactions which simulate real-world

scenarios.

Mitchell forms her integrative or critical pluralism specifically in response to this

scientific pluralism, in particular to account for the complexity of biological phenomena. She

argues that her integrative form of pluralism can explain the behaviours of complex systems of

biology better than other pluralist accounts. By the term ‘complex system’ she means a system in

which the micro-details are so many, or take place on such a long time-scale, that the

computation of the evolution of the system becomes intractable. In pluralist arguments from

complexity, there is an assumption that the impossibility of a complete lower-level description

requires a multi-level causal pluralism. Many have argued that this would preclude the

possibility of any deductive-nomological or base-level causal explanation of certain complex

phenomena (Dupré, 1983; Hüttemann, 2004; Love, 2012). I accept that multiple models can be

explanatory of a given target system, but deny the causal interpretation of these models, because

it is both unhelpful, and potentially problematic.

Mitchell argues that science, and biology in particular, is riddled with multilevel and

multicomponent systems that are incapable of being captured by anything but causal pluralism.

An account of scientific explanation, she says, must represent that fact. She argues that models at

different levels are compatible and can be integrated to form a single pluralistic explanation of a

complex phenomenon or system. She argues that the sheer number and variety of biological

models and methods belies the assumption that we live in a world that can be completely unified

and reduced. She claims that the diversity and variety of life is devalued in reductionism which

seeks to reduce the diversity of explanations, to a privileged set of laws. For reductionists, the

unity is found in a metaphysical and methodological monism. But, she points out that the

reductions of chemistry to physics, or biology to chemistry has never been fully realized.

Mitchell does not think that reductionism or global unification is even a desirable goal of

science:

65

The complexity of the various subjects studied by science and the limitations of our

representations of acquired knowledge jointly entail an integrative, pluralistic model of

science. It is a model that allows for but does not require reduction. It is a model that

recognizes that pluralist ontologies and methodologies are required to give an accurate

picture of science as currently understood and practiced.

(Mitchell, 2003, p. 2)

Thus, she begins by assuming the fact of pluralism, then asking what kind it is, and what kind of

pluralist account best represents it.

In her earlier work, she marks the distinction between compatible and competitive views of

scientific pluralism (Mitchell, 2002b). Competitive views of pluralism hold that different

explanations conflict and a better one will win out and gradually move science towards unity.

Compatible views hold that the pluralism in science is not just a way to unity, but a benefit in

itself. This includes such views as the classification of levels of analysis and involves the

division of questions according to their own framework, as argued by Sherman (1988). Mitchell

thinks this latter view is better, but not enough, as it does not represent the interaction between

levels. She argues for this by means of a case study of eusocial insects, the conclusions of which

lead her to formulate a third integrative kind of pluralism. Next, I will present the case study and

critically examine Mitchell’s conclusions. I find that her approach stays true to the practice of

scientific modelling, but presents little in the way of an account of explanation. Further, the case

study does not support her conclusions as strongly as she claims.

3.2.1. Age-Polytheism in Eusocial Insect Colonies

Page and Mitchell (1991) run simulations of bee colonies using individual units with simulated

genetic variation and which interact with one another. When they were given stimulus they

would self-organize and specialization emerged spontaneously and a response to colony needs

was observed. This self-organization is what is called “age polytheism,” which divides tasks

non-randomly among individuals in the colony. The colony adjusts the proportion of workers in

each caste in correspondence to both internal and external factors. The standard interpretation of

this phenomenon has been the adaptationist one, which looks at past fitness of this strategy

compared to alternatives in order to explain its emergence (Bourke & Franks, 1995). According

to Mitchell, this ignores the mechanism by which the strategy is first adopted, something central

to the explanation of its emergence as an adaptation.

66

The levels at which the bee colony can be analyzed are the cellular level, the individual

level, and the colony level. The debate among philosophers of biology has revolved around

which level of explanans should be prioritized in explaining evolved traits and behaviours. Some

have favoured reductive pictures of evolution, centered on the gene as the major player

(Dawkins, 1989), while others have resisted this, claiming that some explananda require group-

level selection to be explained (Brandon, 1978; Sober & Wilson, 1998; Bersgstrom, 2002; Boyd

& Richerson, 2002). Sherman has argued for a ‘levels of analysis’ approach which places the

appropriate level of selection at the level of the explanandum phenomenon (Sherman, 1988).

Here colony-level explanations are given for colony-level phenomena, and individual-level

explanations for individual level phenomena, and so on. Mitchell wants to resist even this. She

claims that there is only one true explanation of the phenomenon of age-polytheism in honey bee

colonies and it is one which allows for the interplay of models at all the various levels. The same

phenomenon in an ant colony or in some other complex system, will have a different single

integrative explanation. This is because “there is only one causal history that, in fact, has

generated the phenomenon to be explained,” which is itself “a combination of genetic, learning,

and architectural causal components,” (Mitchell, 2003, p. 216). The true explanation is local,

piecemeal, and an integration of partial solutions.

She says that three models have surfaced that challenge any explanation of this

phenomenon which does not cite multiple levels of causes. Along with her own work with

Robert Page (Page & Mitchell, 1991), she cites the work of Tofts and Franks, who propose a

separate algorithm in which foraging for work is sufficient to form a basis for a division of

labour (Tofts & Franks, 1992). In their study, a distribution of workers was organized to satisfy

the colony needs given simple rules regarding workload. Deneubourg et al. offered yet another

model for self-organization where the individual learning and forgetting is sufficient to generate

castes (1987). Mitchell says that these models suggest the phenomenon of colony level

organization is not necessarily determined by the genetic blueprint of individuals. Models of

genetic diversity, uniformity of work, and learning algorithms are all capable of generating self-

organization in these simulations. She claims that these studies promote her account in particular

because of its ability to treat these models explaining the phenomenon as compatible. On her

integrative pluralism, the models all pick out partial causes actually involved in the emergence of

67

age polytheism. The explanation of its emergence then is a pluralistic one involving all these

models.

In response to a possible reductive explanation, Mitchell argues that if the pattern

generated in self-organization arises necessarily from properties of individuals, then this pattern

would have emerged everywhere and there would have been no variation between colonies to

conflict with each other. “Yet,” she says, “the evolutionary explanation of the origin of division

of labour appeals to colony-level selection for energy efficiency and thus must, by definition,

presuppose a history of heritable variation between colonies with such a pattern and colonies

without it” (Mitchell, 2003, p. 214). She concludes from this that evolutionary and

developmental levels interact. Developmental explanations might limit the possible solutions and

the variations for natural selection to choose between, or they might discover structural

necessities or universals, in which case natural selection is not the sole agent in bringing about

the present trait.

She also argues that if Sherman is correct in his levels of analysis approach to explaining

this adaptation of self-organization, then the different models of self-organization should be

mutually exclusive, but Mitchell does not think this is the case. She thinks this because each

model is an abstraction of a particular cause at the expense of others that, therefore, all models

are in fact acting in a real world scenario. For her, the separate models do not compete because

they describe “nonoverlapping ideal worlds” (Mitchell, 2003, p. 216); the worlds in which the

idealized models are true. In the case of the self-organization models she examines, each focuses

on particular idealized aspects of system: one models only genetic diversity, another only

learning diversity, and another only architectural diversity. She maintains that there is only one

causal history, and one causal explanation, that is true of any concrete situation, but it is a

pluralistic one composed the various causes identified in each of the specific idealized models.

Each complete explanation is piecemeal, local, and unique. She takes her advantage over

Sherman to be that the ideal worlds described by these models do not conflict and so they can be

integrated into a pluralistic explanation. By claiming that the models describe non-overlapping

ideal worlds, her account aims to make room for colony-level, individual-level, and genetic-level

selection in the explanation of age-polytheism.

68

3.2.2. Three Problems with the Integrative Pluralism Approach

The first problem with this approach is that contrary to her conclusions, the studies she cites

actually seem to show the possibility of self-organization from individuals without colony-level

causes. These simulations and others that can be easily performed with artificial life (A-life)

games show that nothing more than information at the individual level is required to entail the

emergence of self-organization. Each of the factors in the studies are sufficient to generate age-

polytheism. What Mitchell and Page should have tried to show in order to argue for colony-level

selection is to demonstrate that no amount of genetic diversity was sufficient to reproduce the

observed caste system, and that what was actually needed was a colony-level directive. If age-

polytheism required rules about worker distribution this might have evidenced the claim that the

base-level models are insufficient to produce the observed behaviour. It would have been better

still if the simulation of the desired behaviour had required that many directives operate

simultaneously at the levels of genetic variation, individual behaviour, and colony distribution

requirements.

The second problem is that her integrative solution to the difficulties of modelling

complex phenomena does not provide much, if any, explanatory information. This problem can

be seen in a different example she uses involving the ecosystem behaviour of Lake Erie. In a

case such as this, she claims that an explanation of the whole system’s behaviour has to take into

account many different models and spatial simulations describing the advance of zebra mussels,

the levels of phosphorous in sediment, fish harvesting, the change in solar radiation, temperature,

predation, and so on (Mitchell, 2003). All these models and simulations are constructed using

data specific to that particular lake. The particular models used will be informed by research into

that particular ecosystem and are unlikely to accurately describe any other. The final product of

an integrative explanation will be a collection of such models and simulations and be quite

distinct from any other integrative explanation. The scope or generality of such an explanation is

essentially one system. The set of models, or pluralistic model, that Mitchell aims at ranges only

over a single scenario.

The tension between accuracy and generality reflects the difference between models and

simulations. The value of simulations in scientific practice is clear, but lies mainly in the analysis

of particular cases, and not necessarily concerned with explaining phenomena. What is missing

in simulations is information regarding why the ecosystem, for instance, behaves the way it does.

69

Her account seems to suggest that highly detailed simulations are the best possible explanations

of these complex systems. But these simulations are made for accurate predictions, and do not

contribute to our understanding of either the system in question or other similar systems. There is

no information about why it works. The ways in which all the various factors are integrated are

unlikely to be global or algorithmic and more likely to be particular and concrete; because the

models of these complex phenomena are so specialized, there are not likely to be general rules

for explaining complex phenomena. It is not clear that her integrative pluralism is actually

providing an account of explanation rather than merely describing the challenges of modelling

complex systems.

The real problem this points to is that all models are explanatory on her account.

Integrative pluralism is so inclusive that no scientific model or simulation is non-explanatory. It

is part of her aim to remain true to the practice of science, but as I have argued, one aspect of a

successful account of explanation is that it ought to set some reasonably high threshold for what

counts as an explanation. Proposing an account of scientific explanation that makes room for all

models of a system to be incorporated into a pluralist explanation is too inclusive. Some models

are not explanatory, and this account fails to reveal this.

The third concern is that she claims that all three models for age-polytheism can be

integrated to explain the real concrete case, but provides no framework or method for doing so.

One of the merits of her account compared to other pluralist accounts was that it did not treat the

various levels of models at play as competing for priority of explanatory relevance. All the

models were considered partial, picking out aspects of a single pluralistic causal explanation. But

simply saying that the models can be integrated does not guarantee that they can or that this is

generally the case. The particular models describing aspects of an ecosystem’s behaviour may

not be compatible, in the sense that their predictions could conceivably differ. That the many

models capable of describing a system can all be brought together without conflict is not

something that can be decided a priori.

Mitchell’s work in this area is mostly focused on making her pluralism distinct from

others’, not on arguing for causal pluralism as a whole. That causal pluralism is a fact is an

assumption she starts off with. She takes for granted that explanation is causal and so concludes

from her investigations into the complexity of biological phenomena that some kind of pluralist

causation is correct. In fact, because of the problems with her approach, as well as those that will

70

be examined more closely in the following sections, I maintain that it is better to conclude that

science makes use of multiple explanatory models, rather than multiple levels of causes. Mitchell

provides no actual framework for performing the kind of integrations she proposes, and there are

reasons to doubt the possibility of this seamless integration. It may be that for complex systems,

such as Great Lake ecosystems, the most accurate modelling techniques will run multiple

simulations. It is very unlikely that a small set of equations will be discovered to govern its

behaviour, but this does not imply that a loosely grouped set of simulations and models is

explanatory.

Causal Interventionism

In this section, I will present James Woodward’s account of scientific explanation and review

some worries I have about this account concerning its potentially circular notion of causation and

an ambiguity with respect to commitments of causal realism. Causal realism is at the heart of

many of the issues with causal accounts of explanation and discussion of potential problems here

will transition to a review of emergence and reduction in the following section.

Woodward rejects law-based accounts of explanation in favour of a manipulationist

account. While he notes that Hempel’s D-N account captures many explanatory relations, there

are many explanations that do not feature laws, but are nonetheless explanatory. This is because,

he argues, they give counterfactual information about a system which allows us to understand

how to manipulate it. The reason the D-N account is as successful as it is, is because good D-N

explanations also give us this counterfactual information in lawlike relations. Thus, his account

attempts to demonstrate why a broader range of models ought to be considered explanatory.

Woodward asks us to consider the explanation of the magnitude of the electric field

created by a long wire with a positive charge uniformly distributed along its length. A standard

textbook explanation “proceeds by modelling the wire as divided into a large number of small

segments, each of which acts as a point charge of magnitude dq. Each makes a contribution dE to

the total field E in accord with a differential form of Coulomb’s law:

(7) 𝑑𝐸 = (1

4𝜋𝜀0) (

𝑑𝑞

𝑠2 ),

71

where s is the distance from the charge to an arbitrary point in the field. Integrating over these

contributions yields the result that the field is at right angles to the wire and its intensity is given

by

(8) 𝐸 = (1

2𝜋𝜀0) (

𝜆

𝑟),

where r is the perpendicular distance to the wire and 𝜆 the charge density along the wire”

(Woodward & Hitchcock, 2003a, pp. 3-4). Here Coulomb’s law plays a vital role in the

deductive derivation. But it can also give w-answers, and this is the real reason that the law is

explanatory, though it is not terribly clear in this case. Instead imagine a case from the structural

equations literature. We want to determine the extent to which the amount of water X1 and

fertilizer X2 influences plant height Y. We can use this linear regression equation:

(9) 𝑌 = 𝑎1𝑋1 + 𝑎2𝑋2 + 𝑈,

where 𝑎1 and 𝑎2 are fixed coefficients and U is an error term. Here, even if this gives

information about general causal relations, it falls short of the requirements of laws. It will fail to

hold at large values of 𝑋1 and 𝑋2, it does not account for background conditions that may cause it

to fail, and even in perfect conditions it may never perfectly describe the system. Here, because

the model presents relevant counterfactual information about changes to the system, the model

can be explanatory even if there are no laws of nature. For Woodward, a derivation in the D-N

account only plays an explanatory role insofar as it also gives counterfactual information.

Rather than physical interactions, or causal mechanisms, Woodward favours a causal

approach to explanation that focuses on counterfactual information derived from reliable

manipulations. The ability to explain comes not from tracing and delineating causal histories, but

from the ability of the generalization to answer w-questions, as seen in 2.4.1. Woodward

criticizes Wesley Salmon’s causal-mechanical approach to causation, in which causes are

identified by the transmission of information, or marks, because it does not account for

explanations at higher levels (Woodward, 1989). He uses as an example an explanation of the

changes in pressure given changes in temperature, which feature the ideal gas law. He argues

that Salmon’s account requires tracing the trajectories of individual molecules. Not only is this

impossible, but even if it were done, it still would not give us a good explanation of why the

phenomenon occurred the way it did; it would give us no information about why this happens in

72

general and why other changes occurred in the gas’s temperature and pressure. The standard

explanation, and the one proposed by Woodward, proceeds by making assumptions about the

distribution of the molecules, their forces, and collisions, in order to describe the general

behaviour of the gas. For Woodward, physical interactions will not suffice as a basis for

causation, though some doubt that its rejection is this simple (Ney, 2009).

3.3.1. Invariance and Intervention

Woodward’s proposal is to identify causes as evidenced by certain reliable changes in the

variables of a model (Woodward, 2003). This account of causation involves the notions of

invariance and intervention, notions that stem from the work of Judea Pearl and others (Spirtes,

Glymour, & Scheines, 1993; Meek & Glymour, 1994; Pearl, 2000). Suppose that X and Y are

variables. If an intervention I is performed on X for the purpose of changing the value of Y in

such a way that the change in Y is due only to the change in X, then X can be said to be a cause of

Y. The relation between X and Y is invariant if the relation holds under a range of interventions

on X. This domain of invariance then specifies the strength of the causal relation and explains

the causal dependence of Y on X. Any domain of invariance implies explanatory power. This

power comes from the ability to derive information regarding what would have happened if

things had been different. In this way, invariance under interventions serves to distinguish causal

relations from accidental generalizations, which would not display this invariance.

Woodward introduces the notion of a testing variable, to make sure that there is a range

of possible values for Y. This ensures that it must be possible for the explanandum to be different

than it is. It must be possible to turn off a light bulb by flicking a switch, so that it is not the case

that one explains a light being off by the fact that the switch is down while the light is also

broken. One can contrast this with the D-N account, and notice that the problem in certain

derivations is that the generalization gives no information about on what the explanandum

depends. Testing interventions allow one to debar unhelpful generalizations, which give no w-

answers. This prevents those generalizations such as “all men who regularly take birth control

pills will not become pregnant” and “all hexed salt dissolves in water,” because the other testing

interventions, where the salt is not hexed and Jones does not take birth control, do not have

different results.

Woodward’s manipulationist theory holds that to make the claim that X causes Y, where

X and Y are variables, is to say that there is some possible manipulation of X that can be used as a

73

strategy for changing the value of Y. “Causal claims tell us not that one property is associated or

necessitates another, but rather that certain changes in the value of a variable will produce

associated changes in the value of another variable” (Woodward, 2003, p. 112). An intervention

on one variable is used to discover its relation to another variable. Formally, this is given as

follows:

I1. I causes X

I2. I acts as a switch for all the other variables that cause X.

I3. Any directed path from I to Y goes through X

I4. I is independent of any variable Z that causes Y and that is on a directed path

that does not go through X.

I5. I does not alter the relationship between Y and any of its causes Z that are not on

any directed path from X to Y.

(Woodward, 2003, pp. 98-99)

It is important to note that he explicitly makes use of the concept of cause in I1 in formulating

his concept of intervention, which itself is intended to discover causes between X and Y. He

acknowledges this and thinks that it is unproblematic and necessary, as it would be problematic

to give a fully reductive, or eliminative, account of causation. This results in a kind of circularity,

which I will speak to in 3.3.2. His account is rather a way to discover which dependency

relations are truly causal in virtue of the underlying metaphysics of causation, by looking at the

counterfactuals that generalizations support.

Of course not all counterfactuals are going to describe causal relations. He distinguishes

between interventionist and non-interventionist counterfactuals. Traditionally, counterfactuals

are seen to be what he calls other-object counterfactuals, which give information about what

would be the case for a different object. His same-object, or intervention, counterfactuals refer to

hypothetical changes to the same object. Basically, an explanatory counterfactual will tell us

what would happen to the value of Y, say where 𝑋 = 𝑥1, 𝑌 = 𝑦1, if X were manipulated, rather

than what value Y would have in some other system where 𝑋 = 𝑥2. This distinguishes

information about how a system responds to manipulations, from information given by a mere

regularity. Only the relations which are invariant under the changes of these same-object

counterfactuals support explanations.

Knowledge about these counterfactuals comes from brute facts about the real causes in

the world. Explanation is the activity of gaining information about these causal relations by

74

discovering through intervention which dependency relations are invariant. “No causal

difference without a difference in manipulability relations, and no difference in manipulability

relations without a causal difference” (p. 61). Woodward is committed to the idea that if there is

an invariant dependency relation between two variables then they are causally related, even if

this is contrary to popular belief.

The notion of invariance is that there are true counterfactuals about interventions on X,

and the ensuing values of Y. Invariance is a notion tied to explanatory depth for Woodward

(Woodward & Hitchcock, 2003b). As discussed earlier (2.4.2), Brad Weslake has outlined the

various ways depth, or relative invariance can be established in the interventionist framework

(Weslake, 2010). He notes that a regularity is more invariant than another if it is: more accurate

within a specific range; more robust, in being more accurate under a wider range of

interventions; invariant under a more continuous range of interventions; invariant under a wider

range of ways in which interventions are performed; and invariant under a wider range of

background conditions. Essentially, when there is a stronger the connection between the

variables, the generality can provide a deeper explanation.

3.3.2. Circularity

As mentioned above, Woodward’s account of causation is non-reductive, as he calls it. His

account does not attempt to eliminate the notion of causation by replacing it with the concepts of

invariance and intervention: “Because the notion of an intervention is already a causal notion, it

follows that one cannot use it to explain what it is for a relationship to be causal in terms of

concepts that are themselves noncausal” (p. 104).

The intervention requirement I1 involves knowledge that I causes X, but it is used in

determining whether there is a causal relation between X and Y. This sounds suspicious, but

Woodward does not think that this is problematic. He notes that the causal information needed to

characterize the notion of an intervention is only information about the relation between I and X,

but not X and Y. It is information about other causal relationships than that between X and Y

that is evidence for the claim that X causes Y. He acknowledges the fact that he has imported a

notion of causation into his concept of intervention. However, he thinks it would be problematic

to give a fully reductive account of cause, where causes are stipulated in the method by which

one comes to know them. He claims that if a reductive account were attempted one would be

attempting to derive causal claims from correlational information, something notoriously

75

problematic, if not impossible. Causal claims contain more information than there is in

correlational information (Woodward, 2003, p. 106). Woodward remains silent on what this

information really is, on the metaphysical truth-makers for causal claims.

Woodward claims that his account is neither viciously circular, nor trivial. It attempts to

elucidate the relations between these interconnected terms in the context of scientific

explanation, which is the practice of gaining information about these genuine causal relations.

He maintains that if his definitions constrain the properties, then they are not viciously circular.

His project is epistemological, then, rather than metaphysical, insofar as it remains silent on the

metaphysics of causal connections and focuses on elucidating what information causal claims are

giving us.

Even if it is acknowledged by Woodward, I believe there is a cause for concern. It is

difficult to see how one can gain any causal knowledge if one begins with knowledge of causal

relations. In order to make the claim that X causes Y, one must know that I causes X. But the

way that one gets information about causal relations, say between I and X, is by knowing that

there are interventions I' on I that make reliable changes in the value of X. This in turn is

garnered by knowing that there are interventions I'' on I' that make reliable changes in the value

of I, and so on. There is an infinite regress that results in never getting new information about

causal dependency relations. The worry is that the account only works if one assumes one has

genuine knowledge of causes, say between I and X, to begin with. In order to set up a directed

graph, one must have knowledge of all the causal relations among the variables. In which case,

the purpose of gaining knowledge about causal relations is trivialized. This point ties in to a

more general worry about his commitments to causal realism and the normativity of the project,

which I turn to next.

3.3.3. Causal Realism

Behind all causal accounts of explanation is the idea that giving an explanation involves

identifying the relevant causes. Citing real causes is an effective method for debarring non-

explanatory relations, like backtracking counterfactuals, cases of symmetry, and other non-

explanatory generalizations or derivations. Because of this, it is important that such an account

be committed to the metaphysical stance of causal realism, even accounts that focus on

epistemological issues, like Woodward’s.

76

Woodward distinguishes his own manipulationist account from regularity accounts of

causation. Regularity accounts are focused on establishing regular connections for the purposes

of prediction, but say little to nothing about why the regularity itself holds. They fail to

distinguish between cases where the regularity is causal and cases where it is not, according to

Woodward. Manipulationist accounts are designed specifically to handle such cases. It is built

into the framework of manipulationist causation that there is a stronger realism about causes that

allows for this distinction. Further, unlike other manipulationist accounts, Woodward does not

think there is a serious way in which these causal claims are merely projections of our agency.

Causal relations exist independently.

It is a presupposition of her deliberation that if it is possible to change Y by intervening

on X, then there must be an independently existing, invariant relationship between X and

Y that the agent makes use of when she changes X and, in doing so, changes Y – a

relationship that would exist and have whatever characteristics it has even if the agent

were unable to manipulate X or chose not to manipulate X or did not exist.

(Woodward, 2003, p. 119)

For Woodward, the fact that the manipulation is invariant is due to the real causal dependency

relation, whatever it may be, that underlies the reliable manipulation. The truth-makers for causal

claims are the prior, independently existing, objective differences between causal and

correlational relations. Even though his manipulationist account does not focus on specifying the

metaphysical relations between cause and effect, it does rely on there being such relations.

However, Woodward claims that a benefit of the manipulationist account is that it can

focus on the instrumental success of manipulations in making its causal claims. Woodward

maintains that he can be more or less uncommitted to any particular interpretation of the

processes and mechanisms involved in causal models. He does not require that there be a

continuous causal process – his theory “assigns a more limited significance to correctness at the

level of fundamental ontology.” As he puts it, “a theory might, for example, correctly capture the

dependency relations between a certain set of measured quantities and, hence, qualify as

explanatory, even if it says nothing about or makes mistaken claims about intervening processes

or mechanisms” (Woodward, 2003, pp. 223-224). This is reflected in his argument against

Salmon, which is not simply that conserved quantity accounts of causation do not capture what

77

we consider to be causal relations, but that they fail to include the explanatory phenomenological

generalizations such as those of thermodynamics.

Even though Woodward maintains that “what matters for purposes of causal explanation

is what the real dependency relations in the world actually are” (Woodward, 2003, p. 202), it has

been argued that the causal aspect comes apart from his counterfactual account of explanation.

Saatsi and Pexton argue that if counterfactual information alone can provide a basis for

explanation, then there is little explanatory role for the causal interpretation of this information

(Saatsi & Pexton, 2012). I suggest that this exposes an ambiguity in Woodward’s account

between the requirement of causal realism and the ability of his counterfactual account to

misrepresent or be mistaken about causal claims. Causal realism is necessary when, for instance,

Woodward puts his account to the task of debunking the case of explaining the length of the

pendulum by its period. The way that his account does this is by tracing “this explanatory

asymmetry to an underlying physical asymmetry in the roles played by the length and the

period” (2003, p. 197). It is true that the length in this model is related to the period via a

generalization, but the fact remains that there are no physical manipulations on the period that

will change the length. Because of this, it cannot be featured this way in an explanation. This

difference is reflected in the common sense judgment about not being able to change the length

through manipulations of period. Our knowledge of causes debars certain relations. I think that it

is not possible to simply jettison the causal interpretation of the account; it is needed to debar

certain non-explanatory cases, such as this symmetry.

This raises an interesting issue about normativity for Woodward. Woodward makes a

strong case against Salmon and others that high-level models, such as those of thermodynamics,

are explanatory in their own right. There are many models that are capable of reliably capturing

counterfactual dependencies, but many of these are not commonly held to be causal. These are

the explanatory models that make mistaken or false claims about intervening processes or

mechanisms. Woodward’s account is “partially revisionary.” It aims to “make recommendations

about what one ought to mean by various causal and explanatory claims, rather than just

attempting to describe how we use those claims” (Woodward, 2003, p. 7). And so, his account

will find certain phenomenological generalizations to be causal and explanatory, even contrary to

popular belief. One might be concerned that the revisionary aspect of the project will misidentify

certain counterfactual relations as causal. In fact, it seems as though if any equations are reliably

78

invariant according to the manipulationist criteria, then they would be given a causal

interpretation. This would include quantum models that support manipulationist counterfactuals

about the reliable changes to variables and other highly-idealized models that are non-causal like

those of semiclassical mechanics.

However, because the project is only partially revisionary, it is not clear that one needs to

be committed to any particular revision. One could use the very same project to declare that

quantum models are outside the scope of the account. This is because in order to demonstrate

that a model satisfies the manipulability criteria, one needs to know I1, that ‘I causes X’ (3.3.1).

Because I does not cause X in the non-causal case, the account makes no misidentification.

There is a drawback to this in that it precludes a wide range of models that make explanatory use

of principles, rules, or are otherwise thought to be non-causal. For example, the use of entropy in

thermodynamics, stable states in dynamical systems, and any laws of coexistence would fall

outside the scope of the account. Woodward is quite clear that there are non-causal explanations

and other limitations to his account, and so perhaps the most charitable way to understand this is

to see the account as having a more limited scope and being less revisionary. Some might think

that this is a small price to pay, but one of the main motivations behind the account in Chapter 5

is to broaden the scope of possible explanations to precisely these kinds of models and

idealizations.

What the circularity demonstrates is that there is an ambiguity with respect to

Woodward’s commitment to causal realism. The concern is that it is not clear when the project is

describing or recommending. The project as a whole is rather ambiguous concerning about this,

but in certain passages it is clear that models of macroeconomics and thermodynamics are found

to be causal and explanatory. In which case, perhaps the model is more revisionary than

descriptive and really only precludes models from satisfying I1 when they are obviously non-

causal. In fact, this seems to be what Woodward claims. Given a single real-world system,

Woodward’s account would identify as explanatory and causal a number of models that describe

its various behaviours at various levels. Again, consider the ideal gas law. One can use the

formula to approximate the behaviour of gases under certain conditions. Woodward’s account

identifies the best explanation of some high-level phenomena as being at a higher level than the

molecular level, but it would also identify causes at the molecular level, and as described by

models of fluid mechanics, given that these too would satisfy manipulationist criteria. Woodward

79

does not claim that all causes are actually fundamental. As mentioned above, he is explicit that

there are models that are explanatory and causal that are non-fundamental, like the ideal gas law.

There is an important sense in which models at high levels of generalization are describing

different causes than models at a base level. High-level generalizations reflect independent

causal facts that are not simply parasitic on base-level causal stories. As such, Woodward seems

to be committed to a kind of emergent causation. There are some concerns that have been raised

about the metaphysical implications of emergent causes. And it is to this which I next turn.

Emergence and Reductionism

This section will present and review some of the literature on reduction and emergence and

attempt to apply it to discussions of causation and scientific explanation. A full treatment of this

is not feasible in this context, but I hope to clarify some issues about Woodward concerning

emergent causation. After introducing emergence and physicalism, I present Jaegwon Kim’s

argument against emergence and non-reductive physicalism as well as Jessica Wilson’s analysis

of it in 3.4.2. Wilson shows that Kim’s conclusion is not necessary, and in fact leaves open the

possibility of a kind of non-reductive physicalism. In 3.4.3, I look at List and Menzies’ defence

of high-level causation in the context of a counterfactual account of causation. I aim to show that

neither of these defences are applicable to Woodward’s account because his account is

emergentist. An account that is explicitly non-emergentist, like Michael Strevens’, is then

reviewed in the following section, 3.5.

The literature on emergence goes back at least as far as C.D. Broad and Samuel

Alexander, but this discussion will involve more contemporary contributions (Alexander, 1920;

Broad, 1925). There have been many articulations of what constitutes emergence and whether

emergence is best understood as epistemological or ontological. Epistemological views see

emergence as describing the limits on human knowledge of complex systems and come in

various forms: predictive, which says that emergent properties are features which cannot be

predicted from the pre-emergent stage even given complete knowledge of laws and features; and

irreducible-pattern, which says that emergent properties and laws are true laws which govern

special sciences which are irreducible to physical theory for conceptual reasons (Fodor, 1974).

Some theorists focus on diachronic relations between pre- and post-complexity, and others study

synchronic patterns at different levels (Bedau & Humphreys, 2008).

80

Popper offers an argument for the existence of emergent properties by focusing on the

unpredictability of systems. For him unpredictability is a sign of emergence (Popper & Eccles,

1977). Mark Bedeau’s weak emergence defines a weakly emergent state as one that could be

derived from knowledge of the system’s microdynamics and external conditions, but only by

simulating all the interactions (Bedau, 1997). He formulates this in order to deal with chaos,

which relies on very tiny differences generating unpredictability. This leads to a kind of

emergence in principle for any realistic observer. Andy Clark also takes into account dynamical

systems theory, but focuses more on cognitive science, and argues that emergent phenomena are

those best understood by changing values of collective variables (Clark, 1997). Batterman by

contrast connects emergence to inter-theoretic reduction. He shows that these reductions are

rarely smooth in actual science. What often happens is that there are singular limits for reduction

wherein one model is reduced by taking elements of the other. These special cases are emergent

for Batterman. However, Andrew Wayne argues that there are systems like the van der Pol

nonlinear oscillator whose high-level behaviour can be adequately explained across these

singular limits which suggests that the behaviour is not genuinely emergent (Wayne, 2012).

Ontological views see the physical world as entirely constituted by physical structures,

but composite structures are not always mere aggregates of simples. At each stratum there is a

new kind of property, complete with new causal powers, novel entities, or laws which connect

the complex physical structure to the emergent features. This kind of view has numerous

positions2. Paul Humphreys argues in favour of a metaphysical approach he calls fusion, wherein

entities can become fused and cease to function as separate entities, and gain novel properties

and causal powers (Humphreys, 1997). Jessica Wilson argues that emergence comes from

novelty in terms of a sets of causal powers (Wilson, 2011b). This allows her to argue that higher

levels do not compete with lower levels when their sets of powers are proper subsets of

fundamental casual powers. I will return to this in 3.4.2 when I turn to exclusion arguments.

3.4.1. Physicalism and Supervenience

Physicalism is a position that essentially denies that there is any genuine emergence. A popular

way to look at physicalism is to talk about supervenience. Supervenience is a relation that holds

2 For a comprehensive taxonomy of positions in the emergence literature, see (Wilson, 2011a)

81

between two properties, such that changes in one necessitate changes in the other, but not vice

versa. Imagine the properties of being red, and that of being a specific hue of red, scarlet. An

object that is scarlet is red. However, it can also be non-scarlet and still be red, but if it is non-

red, then it must be non-scarlet. In this case, the property of being red supervenes on the property

of being scarlet.

David Lewis introduces the idea of physicalism by analogy to a printer (Lewis, 1986).

The physical matter, he says, is like the dots of ink and the social and political aspects of the

world are the patterns in the ink. The patterns supervene on the dots – the image cannot be

different unless the physical arrangement of dots is changed, while the reverse is not true. Lewis’

physicalist claim is that a physical duplicate of our world is a duplicate simpliciter. What it

means to say that everything is physical is to say that any physical duplicate of our world would

be identical to it in every respect, because all non-physical properties supervene on the physical.

This is a position that has been taken up by Frank Jackson (Jackson, 1998).

Physicalism comes in two distinct varieties, reductive and non-reductive. Reductionism

has also been popular in the philosophy of science with people like Nagel (Nagel, 1961). Nagel

argued that a theory was reducible to another if one could be deduced from the other with bridge

laws. There have been strong reactions against this kind of approach, and since his time,

physicalists have moved away from it, because it seems incapable of handling the problems of

multiple realizability, the idea that the same high-level property (or behaviour, etc.) can be

instantiated in different physical properties, states, or events (Fodor, 1975; Putnam, 1988). Many

hold that a kind of non-reductive physicalism is the least problematic on the scale from strong

emergence and substance dualism to reductive physicalism.

3.4.2. Exclusion Arguments

I stated earlier that Mitchell’s and Woodward’s positions entail realist commitments to multiple

levels of causes and that there were concerns about this emergent causation. The concern, as

raised by Kim, is that emergent properties are identified by their novel causal powers, but that

because they must always be instantiated by their realizers, they are better explained by the (non-

novel) causes of their realizers. Thus, because they both require novel causal powers, yet cannot

have them, emergence is problematic.

Kim notes that in a supervenience relation, an emergent E occurs only when the base-

level realizers of E are instantiated. If one knows that neural state N occurs, and that emergent

82

pain state E occurs, and that N is the realizer of E, then the presence of N explains the occurrence

of E. One can also predict the occurrence or non-occurrence of E given knowledge of its

emergent base. Because he is working from an assumption that realizers are known he is quite

confident that “all our explanatory demands can be met” (Kim, 1999, p. 13). However, this claim

is not on trail.

There is a further ontological consideration. Kim puts forward what he has called the

causal inheritance principle, which says that the causal powers of an emergent that has been

functionally reduced are identical with the causal powers of the realizer. If this is accepted then

one must ask what the status of E is, and why it should be considered distinct. He finds three

options. First, one can maintain that the emergents are irreducible to multiple realizers. Second,

one can identify the emergent with a disjunction of realizers. Third, one can simply give up the

idea that E is a genuine property and only recognize it as a useful concept. The second two are

both reductions of E. To the first, he responds that the realizers themselves must have distinct

causal powers and so the multiply realizable properties must be distinct. All this, he says, points

to the fact that E is unfit to play a role in science. No one would insist on the existence of

emergents if they had no role to play in explanation or prediction in science. Their causal

efficacy is their most important feature. For Woodward, as will be shown in the following

subsection, the right kind of invariable dependency relations at higher levels identify causes at

that level. He is not a reductionist about causes or entities.

Kim shows that there are three kinds of causation that could occur among emergents and

realizers: same level, downward, and upward (Kim, 1999, p. 19). He argues that upward and

same level causation among emergents implies the possibility of downward causation. The

argument proceeds as follows.

Suppose there is a property M at level L which causes property M+ at L+. Given that M+

is an emergent property, it has an emergent base at L, call it M*. The question of what caused the

occurrence of M+ has two solutions and only one reasonable one, viz. 2.

1. L+ M+ 2. L+ M+

L M M* L M M*

83

Here the slim dark arrows represent causal relations and the black arrows represent the

emergence relation. The original scenario in 1 is far better explained by 2, since every

occurrence of M+ must be instantiated by its realizer. The only way M causes M+ is by causing

its realizer, M* which also must be caused by M. Upward causation is only possible if same level

causation is possible. The same can be argued for same level causation presupposing downward

causation.

To see this, suppose again there is a property M at L which causes M*. M* itself has a

basal realizer M- at L-. If we ask again how M* was brought about, there are two answers and

only one reasonable one, viz. 4.

3. L M M* 4. L M M*

L- M- L- M-

Again, M can only cause M* by causing its realizer M-. This can all be generalized such that to

cause a property one must cause its basal realizer – what Kim calls ‘the principle of downward

causation’. The next step in the argument is to problematize downward causation, because he has

shown that if emergents cannot have downward causation, they cannot have causal powers at all.

Kim asks why the putative cause of M cannot always be the cause of its emergent base P.

Consider case 4 above. His argument is that it can always be better explained by a cause of its

emergent base as shown below (5.).

5. L M M*

L- P P*

Here we can see that the emergent property M is entirely dispensable in bringing about M*. “If

emergent properties exist, they are causally, and hence explanatorily, inert and therefore largely

useless for the purpose of causal/explanatory theories” (p. 33). This points to a problem that has

far-ranging consequences for any kind of non-reductive position. However, it may not be

applicable to Woodward’s account, which makes use of an entirely different notion of causation.

On Woodward’s account, in order to discover an invariant relation between X and Y, one

must fix the other causal pathways such that the only effect on Y is X. If X and Y are emergents,

84

irreducible higher-level variables (like M and M* above), then there can be no manipulation on X

in order to see if it is a cause of Y, because one cannot fix the emergent base of X due to the very

nature of the supervenience relation. The Kim diagrams are part causal mapping/part

supervenience mapping, but the relation between the realizer and emergent is not causal, but an

“automatic change”. Woodward argues that there is a meaningful sense in which one should

draw a causal arrow between supervening properties M and M* (Woodward, 2011). This is

because it is possible that P causes P*, and yet their emergent properties are not causally linked.

Thus, there is additional causal information that can be gained by saying that M causes M* over

and above saying that P causes P*. Contrary to Kim’s conclusion, Woodward thinks that

manipulationism is open to those emergent causes.

There might be additional information in high-level causes but it does not necessarily

mean that that is unproblematic. And, of course, not everyone is convinced by Kim’s argument.

Some have argued that the argument does not lead to Kim’s conclusion, and others that it only

applies to certain interpretations of causation. To show how one might argue that it is

unproblematic for counterfactual accounts of causation, I will turn to a defence mounted by List

and Menzies in 3.4.3. First, I look at an examination of Kim’s argument by Jessica Wilson,

which effects a defence of non-reductive physicalism by means of a metaphysical constraint on

causal powers.

Kim’s argument has been challenged by many philosophers who advocate some kind of

non-reductive physicalism or want to defend high-level causation. According to Jessica Wilson,

the problem articulated by Kim is that there is no satisfying answer to the question: how can

special science entities have real causal powers given their dependence on lower-level entities

(Wilson, 2011a). She identifies six premises that lead to the problem:

1. Dependence. Special science features depend on low-level physically acceptable

features.

2. Reality. Both features are real

3. Efficacy. Special science features are efficacious.

4. Distinctness. Special science features are distinct.

5. Causal Closure. Every low-level effect has a low-level cause.

6. Non-overdetermination. Effects are not causally overdetermined.

Kim’s argument shows that these premises are inconsistent. Based on Kim’s commitments and

assumptions, he takes this to entail the denial of P4, and thus takes the argument to lead to

reductive physicalism. However, Wilson notes that other premises may be rejected in its place,

85

since all that was shown was that together they are incompatible. One could deny P1 and support

substance dualism; deny P2 and support eliminativism; deny P3 and support epiphenomenalism;

deny P5 and support strong emergence; or deny P6 and support non-reductive physicalism.

Wilson focuses on the latter two positions. Strong emergentists deny the causal closure of

the physical, which means that at least some high-level entities have the power to produce an

effect which low-level entities do not – a position taken, for instance, by Sandra Mitchell. The

powers of the high-level feature must satisfy this new power condition. Thus the movement of

reaching for a water glass when thirsty would not have a purely lower-level physical cause, but

come in part from the property of being thirsty. The emergent property of being thirsty can do

one thing that the physical realizers cannot.

The NRP claim is that special science features are real and distinct, but stand in an

intimate relation that, while not identity, is close enough to avoid overdetermination. A number

of such relations have been proposed, such as functional realization, part-whole relation, and

determinable-determinate relation. These all rely on what Wilson calls “the subset condition on

powers: Token higher-level feature S has, on a given occasion, a non-empty proper subset of the

token powers of the token lower-level feature P on which S synchronically depends, on that

occasion” (Wilson, 2011a, p. 263). This condition both avoids overdetermination and conforms

to physicalism. Wilson argues that all forms of emergent dependence conform to either strong or

weak emergence. According to strong emergence, the higher-level entities have at least one

causal power that the lower-level entities do not have. NRP is associated rather with a kind of

weak emergence, where the upper-level entity has a proper subset of the powers of the lower-

level, such as that it can be counted a distinct entity but still physically acceptable.

The ability to explain a higher-level property in terms of a physically acceptable property

is not sufficient to demonstrate its physical acceptability, because this may not speak to its

having independent causal powers, which is the distinguishing feature of an emergent property

(Wilson, 1999, p. 42). What Wilson proposes is a constraint on causal powers, CCP, that a causal

power associated with a supervenient property is numerically identical with a causal power

associated with its base property. Higher-level entities need not have a novel causal power in

order to be autonomous, they only need have a distinct power set, even if it is a proper subset.

This is because an entity is determined by the set of causal powers it possesses. What this allows

is that there can be distinct entities (as determined by non-identical sets of powers) without

86

maintaining that there is a novel (emergent) causal power. In this way, a strong case can be made

that there are physically acceptable higher-level entities and that NRP does not necessarily slide

into problematic emergentism or reductionism.

However, there is a worry that the discourse of metaphysics and that of model-based

scientific explanation are quite distinct. The way that explanatory models are is not necessarily

the way that the world is. This is something that I showed in Chapter 2. In order for such a

defense to work for Woodward, he would need to be committed to a claim that the causal powers

of entities featured in explanatory models, as identified by manipulationism, stand in this proper

subset relation to each other. Another way to put this would be that physicalism and the

constraint on causal powers must hold for features of explanatory models. No argument to this

effect is given by Woodward or Wilson, and it is not obviously true.

Further, there is at least one reason to suggest that it is not always the case: some

explanatory models make use of entities that do not properly exist and employ properties that are

nowhere instantiated, e.g. perfect spheres, frictionless planes, infinite populations, etc. For

instance, Mitchell maintains that the various explanatory models at work in a complex system

describe “non-overlapping ideal worlds.” It is not clear in what sense the causal powers of

entities in a group-level model could be a subset of those in an individual-level model, and that

the powers of entities of an individual-level model are subsets of those at the genetic-level.

As we also saw in the case of semiclassical mechanics in Chapter 2, classical properties

cannot be de-idealized to quantum properties. Properties like having velocity and position are not

subsets of quantum properties like having a spin quantum number. Wilson’s concern is not with

properties in general but with causal powers. The constraint on causal powers means that the

powers belonging to entities in explanatory classical models would have to be proper subsets of

the powers of entities in explanatory quantum models. The lack of causal powers at the quantum

level shows that features of semiclassical models at least do not satisfy the constraint on causal

powers.

Generally, it is difficult to see how there could be known or demonstrable proper subset

relations between features of models where there are no known regular limits or other mapping

relations. Because Woodward’s account holds that manipulation relations among the variables of

high-level models describe causes that are non-fundamental, he is unlikely benefit from this

defense of NRP. This is something I expand upon in the following section as well. It is important

87

to note, that NRP has not been problematized here, and may well turn out to be true of the world.

However, the entities and dependency relations of explanatory models need not reflect a

consistent metaphysical position – that requires an additional claim of physicalism about models.

The claim that the real entities of the world are physically acceptable and that the claim that the

entities featured in explanatory models are physically acceptable are distinct. If the constraint on

causal powers was shown to be true about explanatory models, then it seems that problems of

high-level causation would disappear. But a convincing argument for this has not been offered.

There is no obvious reason why physicalism and the constraint on causal powers ought to hold

for the features of explanatory models, and good reason to think that in some cases it does not. In

the following subsection, I turn to a different kind of defense that is geared towards

counterfactual accounts of explanation and may turn out to be more useful for Woodward.

3.4.3. Intervention and Emergent Causation

In order to see how a counterfactual account could be defended against Kim’s argument, I will

look at one such defence mounted by List and Menzies (List & Menzies, 2009). Woodward

distances his notion of causation from the classic causal-mechanical view of Salmon. But

because of this, it is not clear that Kim’s causal exclusion argument applies to his interventionist

notion of causation, which is quite distinct from the one Kim problematizes. List and Menzies

attempt to make use of a concept of causation as difference making in order to show that the

causal exclusion principle only works for some formulations of cause. They argue that causal

exclusion actually supports the causal autonomy of certain higher-level properties. However,

similar to the application of subsets in the previous subsection, I think that it is not obvious that

the features of explanatory models always stand in the right supervenience relation to one

another. I argue that the defense against the exclusion argument fails to save Woodward’s

account because it too does not apply.

List and Menzies interpret Kim’s argument slightly differently than Wilson. They claim

that non-reductive physicalism has three theses: The properties of the special sciences are not

reducible to physics, but multiply realizable by them; these properties supervene on physical

properties; these properties are the causes and effects of other properties. Kim’s argument

proceeds by showing how the first two of these contradict the third. It makes use of the exclusion

principle: if a property F is causally sufficient for some effect G then no distinct property F* that

supervenes on F can be a cause of the effect G. Simply put, problems of overdetermination are

88

avoided because only supervening properties are excluded. List and Menzies have two main

claims in this 2009 paper. First, that the truth of this principle is contingent, and second, that

when true, can actually support causal autonomy at the higher levels. They opt for an account of

cause as different making, which aims to include contrastive, counterfactual, and interventionist

accounts of causation, such as Woodward’s.

A central assumption in this argument is that a cause should be proportional to its effect,

something they draw from Stephen Yablo (1992). They argue that in many cases the difference

maker in the occurrence of a phenomenon is the supervening property and not the realizer. Given

the case of a pigeon who is trained to peck at red targets for food, the redness of a target is the

important factor in pigeon pecking, rather than the particular hue of red (Yablo, 1992). In fact,

the hue is not sufficient to count as a cause, or the pigeon would not peck at any red colour, but

only a particular hue, which it does not.

List and Menzies look at two counterfactuals to apply the idea of causal exclusion to a

difference making account of causation.

1a Target is red □→ pigeon pecks

1b Target is not red □→ pigeon does not peck

1a and 1b are true, as outlined by the description of the thought experiment. But, consider the

following two counterfactuals about specific hues of red:

2a Target is crimson □→ pigeon pecks

2b Target is not crimson □→ pigeon does not peck

2a and 2b are not both true despite the closest world being where the target is still some shade of

red where the pigeon would peck. The supervening property is clearly the difference maker. The

counterfactual that demonstrates the difference is the one that identifies the relevant cause. And

so, on this account of causation, there can be high-level causes without an underlying causal

relation. In effect, the excluded cause is the realizer. Kim did not consider causation to be cashed

out in counterfactual terms of any kind, but in a primitive notion of production or generation, and

this, they argue is why high-level causation seemed to be problematic.

On Woodward’s account, wherever there are the manipulability relations, there are causal

relations and vice versa. What this means is that if one can reliably manipulate M in order to

change M*, then M is a cause of M*. It is also true that if one can reliably manipulate P in order

89

to change P*, then P is a cause of P*. Thus, the Kim-style diagram for Woodward would take the

following shape (assuming the properties are related via supervenience relations).

6. L M M*

L- P P*

What this implies is that unlike the strategy taken by List and Menzies (or Sherman (1988)), for

Woodward, there can be genuinely emergent causes.

As was shown in 3.3, Woodward’s notion of causation is different than Kim’s, but it is

also different from List and Menzies notion. It does not operate by selecting the difference

maker. It can allow for the higher-level generalization to be explanatory, while not having to

deny that there are causes also at the system’s micro-level. He puts forward no principle of the

proportionality of causes that selects a single level of causal operation. Because manipulationism

identifies both high-level and low-level generalizations as causal and explanatory, then

Woodward, like Mitchell, is committed to emergent causation. This position makes realist claims

about both the lower and higher levels; for Woodward, any level where a generalization passes

his criteria for invariance under interventions makes claims about real causes

On Woodward’s account, and on Mitchell’s, there are many models and levels of models

which exhibit causal dependencies. The relations between features of models may not be that of

supervenience. The models might be competing models on the same level, or irreducible in terms

of time scale or spatial scale. Just as in the case of the proper subset relation, there is no reason

given to think that the entities and causal properties of a model at higher levels will necessarily

supervene on those of models at lower levels. There are surely some cases where this relation is

clear, as in the simple examples often cited. But it has been argued that in other cases there are

no such relations, for example concerning entangled particles, and other non-classical properties

(Karakostas, 2009). I am not committed to this claim, but only wish to point out that in the case

of explanatory models, the argument has not been presented that they will stand in the

appropriate supervenience relations. And further, that it is far from obvious that the properties of

various idealized, and highly-idealized, models will stand in a determinant-determinate relation

or a relation of parts and wholes.

90

Nonetheless, Woodward’s and Mitchell’s accounts identify explanations and causes at

multiple levels. I think that there are some reasons to be concerned about emergent causation that

implicitly claims that models track causes. I assume that there are explanations at multiple levels

and that this is not itself problematic. It seems like there are a few options to make sense of this:

1) Claim they are all real causes; 2) Privilege a certain level as real causes; or 3) remain silent on

whether they exhibit real causal dependency relations or not. Each of these has some

implications, some more undesirable and problematic than others. List and Menzies opt for the

second position, as does Strevens, which will be shown in the following section. Woodward’s

and Mitchell’s strategy is to take the first option, but this involves demonstrating that emergence

is unproblematic, or that some other defence is applicable. I advocate the third option to remain

silent on whether the dependency relations among the variables of explanatory models exhibit

causal dependencies. As was discussed in 2.1-2.3, different models can be used for different

explanatory purposes, but there is nothing in this alone that necessitates that there also be

different causal dependencies in the world that correspond to the relations among the variables of

the models. Giving explanatory models at multiple levels a causal interpretation is at best

unnecessary and at worst problematic. But, if one is to remain silent on whether the dependency

relations in explanatory models are causal or not, then one has to provide another means of

identifying genuine explanations, which is what I offer in Chapter 5. There are predictive

divergences and conceptual inconsistencies among models, but this need not preclude their being

explanatory. By favouring a non-representative account of explanation, it is much easier to make

sense of the fact that explanatory models make use of entities and relations that may not be

physically acceptable.

None of this is to demonstrate that causal accounts cannot succeed or that non-reductive

physicalism is untenable. Rather the aim is to demonstrate that there are some challenges that

make a causal approach less attractive. And nor is this meant to be an exhaustive treatment.

There remains a great deal of work to be done regarding supervenience relations and scientific

models, which is outside the aims and scope of this dissertation. Woodward’s account requires

knowledge of causes in order get further information about causal dependency relations, and may

never get off the ground. It is also committed to causal emergentism, which invites problems of

overdetermination and downward causation. Defenses mounted to preserve either NRP or high-

level explanations are not going to work for Woodward, whose account identifies genuinely

91

emergent causes. Wilson’s and List and Menzies’ defences of NRP may be strong defenses of a

physicalist conception of the world, but it is not obviously true of explanatory models. Let me

now turn to Strevens’ account which may be more compatible with these defenses because of its

explicit rejection of emergent causal facts.

Kairetic Explanation

This section looks at an account of causal explanation that is compatible with non-reductive

physicalism and specifically geared towards capturing the explanatory role of non-fundamental

models. Michael Strevens’ kairetic account of explanation begins with a basis in causal

explanation, but proposes an account of depth that can prefer higher-level models (Strevens,

2004). Strevens’ aim in this project is to use unificationism to accomplish one of the major goals

of causal accounts, which is to specify the relevant causes of a given effect, such as to constitute

an explanation of that effect. He is using unification in order to pick out not the most unified

theories, but which causes are the difference makers. In this, it is similar to the position presented

by List and Menzies, though the focus is not on counterfactuals.

Strevens claims that any of the common accounts of causation are sufficient to give

causal asymmetry, and so he does not want to specify the details of the metaphysics of causal

relations and will instead focus on relevancy criteria. He thinks this approach can ignore the

problems (and potential benefits) of metaphysical realism. He does not even argue that causation

is either reducible or not, whether explanations feature laws or not, whether all causation is local,

or forward in time. The only strong claim he thinks he needs to make is that there is nothing over

and above fundamental-level causation: “Causal emergentism has no place in the causal

approach to explanation” (Strevens, 2008, p. 35). There are no independent high-level causal

facts. As such, fundamental-level causal influences can explain everything that can be explained

causally. Because of this, his account faces different problems than Woodward’s or Mitchell’s

does. In 3.5.1 and 3.5.2, I present his take on idealization and abstraction and how a notion of

depth is supposed to preserve high-level explanation on this picture. In 3.5.3, I present some

objections that have been raised to this account concerning how to weigh his criteria, the process

of abstraction, and the unrealistic process of forming an explanation.

On Strevens’ account, explanation is a two-step process of isolating difference makers.

The first step is to start from the causal web of an occurrence E and remove all causal influences

92

that are not necessary to bring about E. The second step is to remove all things that are not

relevant to the logical entailment of E by the set of circumstances and laws. What is left is what

he considers to be the causal difference makers for E. The role of logical entailment here is only

to represent the causal factors, which is why it is the second step. If it were the first step, non-

causal dependencies would be allowed. He finds that in this way one can debar claims about the

gravitational influence of Mars being causally relevant, for instance, to Strevens’ much-loved

case of the death of Rasputin. The goal is to separate what makes a difference to the physical

system from what makes a difference to the explanandum.

His account can, he claims, identify all the difference makers for a particular event. The

eliminative procedure is to remove as many parts of a causal model M that entails event E

without invalidating the entailment of E. This is followed by an optimizing procedure which

constructs sets of statements, or models, to form a standalone explanation (more on this in 3.5.1).

He takes a causal model to be a set of veridical and deterministic causal statements about

the world that entails E. An atomic model may be thought of as picking out a single link or a

length in a long causal chain. A composite model contains two or more atomic models strung

together. For Strevens, explanations are still deductive entailments, and what makes them causal

is that they make use of causal laws whose content is determined by the metaphysics of causal

influence. There are no independent high-level causal facts, and so fundamental-level causal

influences can explain everything that can be explained causally.

3.5.1. Abstracting and Optimizing

The first step of his account is to eliminate causal influences that are irrelevant. Picking out

difference makers is the process of identifying which causal factors are relevant to the

occurrence of E, by looking at what plays an essential role in the causal entailment of E. If C

cannot be removed from the causal model without eliminating E, then C is a difference maker.

He claims that the kairetic account can solve many of the problems known to face causal

accounts.

The process he refers to as abstracting is that of ignoring the details of a model. This is

the process by which a mechanism is substituted for a black box making no reference to the

internal causal process. His example involves explaining why a window broke when a

cannonball was thrown at it. The fact that the projectile weighed exactly 10kg is not important to

the explanation of why the window broke. What is important is that the ball was rather heavy,

93

say more than half a kg. Strevens gives us the conditions under which he considers a model to be

an abstraction of another:

A. One model M is an abstraction of another M' if its propositions are entailed by the

detailed model M' and if all the causal influences of M are described by M'.

In this case, it follows that if the cannonball is 10 kg, that it is more than half a kilogram. What

makes a difference is that the cannonball is either above or below a certain threshold, not that it

is or is not exactly 10kg. By abstracting in this manner one can get to the difference making

cause, what Strevens calls a kernel.

A model can also be too abstract. In order to prevent the optimization from favouring

radically and uselessly disjunctive models, he imposes a cohesion requirement on the process. A

model is cohesive only if all of its realizers possess the same causal elements; if it is causally

contiguous at the fundamental level. The kairetic account is constrained on both sides by

ignoring causal detail and requiring cohesion among the model’s realizers. This is what Strevens

finds so promising in this approach, the balance between abstractions and causal realism. An

abstract model is best unless it violates the cohesion requirement. This happens in cases where

the model is radically multiply-realizable.

3.5.2. Idealization and Causal Realism

As was mentioned in the last section, causal realism plays a strong role in causal accounts of

explanation, and the kairetic account is no exception: “no causal account of explanation –

certainly not the Kairetic account – allows nonveridical models to explain” (p. 297). Thus,

contrary to Bokulich, myself, and many others, he explicitly rejects that highly-idealized models

can support explanations. Models may intentionally misrepresent elements of the causal

mechanism, but Strevens’ kairetic account “demands that factors be omitted in a way that does

not compromise the veridicality of the model” (p. 298). He examines the case of Boyle’s Law,

which assumes that molecules in a gas do not collide, even though they surely do. Strevens must

justify this falsity.

Strevens does not take the stance that idealizations are better on pragmatic grounds. For

this strategy, idealizations are compromises of a perfect explanation, which would be entirely

veridical. This downplays the importance of idealization in explanation. For Strevens,

idealization makes an explanation better by conveying only essential information. He goes as far

94

as to claim that an idealizing explanation is always better than its veridical counterpart. It is

better because it does not present all causes as equal, but rather highlights the difference makers

that are key to good explanations. He characterizes an idealization as the following: It is false;

the false claim fills out details left unspecified; and the details are filled in with zeros or infinites

in order to eliminate them, because they do not matter (p. 318). He is a physicalist, but he is not

reductive with respect to explanation.

Take for instance Boyle’s law relating pressure and volume,

(10) 𝑃 ∝1

𝑉.

A textbook explanation of gas behaviour that makes use of the law features many idealizations,

such as that the collisions with container walls are completely elastic, and that molecules do not

collide with one another. It gives a decent explanation, but makes mostly false assumptions about

the nature of the gas. There is another common explanation involving a more complete

description from modern kinetic theory, including the influence of the molecules on one another

at a distance, and allowing for intermolecular collisions. Strevens’ claim is that one can eliminate

details of the kinetic model that make no difference to the Boylean behavior of the gas and

recover the relation found in (10). It is only in this way that one can understand Boyle’s law and

how it can be explanatory; the relations that are highlighted in that law are the difference makers

for Boylean behaviour. Thus, the textbook explanation is the best; it is not the most veridical, but

it has only difference makers.

Idealized models are favourable to veridical models in a few ways: they highlight the

irrelevance of certain factors; they are much simpler; they are effective predictors as long as the

idealization is reasonably faithful. That a kairetic explanation is always at bottom a physical

explanation does not imply that kairetic explanatory models must describe the trajectories of

particles. Idealized models can be effectively employed where the omissions they make coincide

with the non-difference makers of the kairetic explanation. Strevens justifies the explanatory

value of idealizations by claiming that they are only distortions of non-difference makers. If the

range of values a variable can take make no difference to the explanatory target, then the

idealization is justified.

95

3.5.3. Concerns about the Kairetic Account

The kairetic account attempts to strike a balance between the fundamental causes in nature and

the explanatory value of idealized models. Strevens’ account is complex and nuanced and worthy

of in-depth discussion on a variety of topics. I will restrict this to a few concerns that seem most

worrisome for the account. For Strevens, the first step to an explanation is to begin with total

causal information. One must ask, in what circumstance is this the first step to forming an

explanation? One can use Strevens’ account to justify judgments about higher-level explanations

if they are abstractions of a complete picture of the base-level causes of a system. But this

ignores the fact that when modelling a high-level behavior, a scientist will not begin by looking

at the fundamental level of interactions and idealize away non-difference making processes. His

account is essentially a justification of textbook explanations and singular event causal

explanations in the face of more accurate competitors, but it does not reflect the practice of

scientific explanation or modelling outside of this pedagogical context.

Echoing concerns mentioned in 3.4, the scope of this account is quite limited. Very few

models are just abstractions of base-level causes. He maintains that it is consistent to allow for

the ideal gas law, because it is an abstraction of physical laws described in the molecular model

of a gas. But his notion of abstraction (A) is quite stringent. It involves a requirement that an

abstract model is entailed by a lower-level model and that all of its causal powers are described

by the lower-level model. This precludes any case in which this kind of reduction has not been

performed, including highly-idealized models and models with multiply realizable properties. In

order for a high-level model to be explanatory, it must be able to be mapped somehow onto the

real causal mechanism, or “distilled” from it, as Strevens says. These models may be

explanatory, but I and others maintain that this is a small subset of the set of explanatory models.

What I view as the most serious problem for the kairetic account is that the process is

very detached from the scientific practice of explaining. This is an objection raised by others

who have noted that his oft-cited examples are far from those considered scientifically adequate

(Hartmann & Schupbach, 2010; Levy, 2011). He focuses on what I have called common sense

explanations, and on maintaining an account of causation that is strongly continuous with an

everyday sense of what it means to explain. He refers throughout the book to an anecdotal case

study of the death Rasputin. He uses this to explore issues such as overdetermination, pre-

emption, and more. He takes these results as having direct implications for his account of

96

scientific explanation. The accounts limited scope and distance from the explanatory practice

science make it a bit too ideal.

It is Strevens’ aim to formulate an account that allows for trade-offs between the

generality gained in the optimizing procedure and the cost to cohesion, but Potochnik argues that

there is no reason to always prefer the most general model (Potochnik, 2011). He finds it

unproblematic to trade off cohesion and accuracy for generality across the board. If the cost is

low, he says it is mandated to make the trade-off. But Potochnik notes that it is not always the

case that the most general explanation is the deepest. On some occasions a scientist might prefer

models that highlight subtle perturbations and explore fine-grained dynamics. There is no

objective standard for determining these trade-offs. Hartmann and Schupbach also argue that a

great deal of work remains to be done to fill out the concepts of accuracy, cohesion, and

generality, such that they and the account might be meaningfully applied (2010). They also

mention that the limited scope of Strevens’ account is slightly ironic. There are much more

general accounts of explanation and given its preference for trading off accuracy in favour of

generality, its limited scope almost seems to “mandate its own rejection.”

Strevens attempts to formulate an account that captures how high-level models can be

explanatory. In the end, it is very limited in scope, provides very few measures for implementing

and trading off its desiderata, and is not reflective of explanatory practice.

Conclusion and Additional Concerns

Mitchell, Woodward, and Strevens recognize the need for higher-level explanations and have

formulated causal accounts of explanation that attempt to include some higher-level models as

explanatory. Mitchell’s solution of integrative pluralism is hardly a solution at all. It provides no

framework for performing the integration of multiple levels of models into a singular multi-level

causal explanation. And further, even if it did, it advocates that a system-specific explanation is

going to be the best explanation of a complex system, but such system-specific models tell us

very little about why the system behaves similarly or dissimilarly to other systems. By including

all models that can be used to describe a system as explanatory and veridically tracking partial

causes, there is no actual threshold for explanation: all models qualify. If we want to maintain

that some models are mere phenomenological generalizations, then some reasonable threshold

for which models count as explanatory ought to be set.

97

Causes on Woodward’s account are determined by the reliable variable dependencies of

models. This allows for models at levels other than that of fundamental physics to be considered

explanatory. This strategy relies on a commitment to causal realism that provides underlying

metaphysical truth-makers to support causal claims and debar non-causal ones. But his account is

at least partially revisionary and so determines to an extent what is considered causal. As such,

he has an ambiguous stance towards causal metaphysics. His account does find many high-level

generalizations to be causal and explanatory, which threatens to invite the problems of

emergentism. Unfortunately, defenses of NRP and high-level explanations, such as those given

by Wilson and List and Menzies, cannot be used to rescue Woodward’s account. This is because

Woodward is committed to independent high-level causal facts. Further, the literature on

emergence and physicalism and the literature on models and scientific explanation talk past one

another to a large extent. Deciding which models are genuinely explanatory need not coincide

with physicalism, for instance, when one takes a non-representative approach, the properties and

features of explanatory models are under no obligation to be physically acceptable, even for a

physicalist. A lot of work remains to be done concerning where supervenience relations might

hold between features of models, and where not. The present chapter could only survey some of

the various positions and possibilities.

I have expressed suspicions that supervenience relations and proper subsets are not

ubiquitous among explanatory models. However, it is interesting to note that in a case where

there are no supervenience relations, then the problem as articulated by Kim does not even apply.

The argument is aimed at exposing problems of overdetermination and downward causation for

non-reductive or emergent accounts that employ supervenience. So, in the cases where there are

no such relations, this is not an issue. It is possible that this could present a viable way of

circumventing Kim-style arguments. I was not able to explore this avenue here. What this

suggests is that perhaps the features of explanatory models are not best understood as exhibiting

genuine causal relations, which reflects the limited scope of causal accounts.

The strategy taken by Strevens is to focus on the explanatory relevance relation among all

the fundamental causes in a system, thus denying any high-level causal facts. By beginning with

a complete causal story and swapping out processes for black boxes, only relevant causal

difference makers remain, and thus abstractions and certain kinds of idealization play the role of

highlighting important explanatory information. But the two-step process for explanation is

98

largely fictional and does not reflect any actual explanatory practice. The account seems to have

a different aim than to capture scientific explanation. It shows how one could justify preferring

abstract models to more detailed models, and how harmless idealizations allow important

information to be highlighted.

It is worth reiterating my concern with the limited scope of causal accounts, which has

been emphasized throughout this dissertation and by many others. Deductivist approaches to

explanation promise to permit a wider variety of models as explanatory, including those that

feature non-Galilean idealizations, laws and principles, and other causally unacceptable features.

For instance, non-reductive systems where long time-scale behaviour cannot be smoothly

approached by models of short time-scale behaviour, and systems where a high-level

explanandum behaviour cannot be explained with the use of low-level models would not be

precluded. There are many cases of idealized models which are not representative of target

physical systems. All of Batterman’s asymptotic explanations of universal behaviour do not

qualify as causal explanations (Batterman, 2002b, 2002a). Further, cases for non-causal

explanation will be presented in reviewing Hempel and Kitcher in the following chapter.

There are some additional reasons that I think warrant being skeptical about the promise

of causal accounts of explanation. These are not arguments that are meant to seriously

problematize the causal accounts examined in this chapter. I am merely raising some flags on

issues that I think would benefit from further investigation and development. My main concerns

stem from commitments to causal realism. This has implications for the kinds of idealizations,

features, and relations that can be permitted in explanatory models. This in turn severely limits

the scope of such accounts and is potentially problematic.

It seems that causal accounts of model-based explanation must be committed to some

degree of realism about explanatory models. The relations in explanatory models need to

describe real causes or encode their causal information in some way. Let us assume that

explanatory models track real causes and that propositions about these models are approximately

true. I have two reservations about this being an accurate way to look at explanatory models. The

first is that the various models that operate at various levels in a real-world system diverge in

their predictions of its behaviours. Some are more accurate, others more general, some are only

accurate at short time-scales, others only at long, and so on. They may focus on different

explananda, but if they concern the same real-world system, then they are all tracking real causes

99

in the system. This works fine where the predictions are all in line. But it seems that if the

predictions of two or models diverge significantly, then they cannot both be tracking real causes.

This realism about explanatory models also has implications for the features of these

models. And so my second concern is that the features and the entities of explanatory models are

incompatible, in the sense that propositions describing the models are contradictory. This can be

seen in the explanations of light diffraction in particle and wave theories of light, and the

statistical mechanical and kinetic explanations of the behaviour of gases and fluids. It seems odd

to maintain that the entities in this real-world system are both fluids with no discrete part and are

mere particles, or that the air has a damping effect and does not. Again, I find these two

scenarios puzzling, but I can develop them no further here. It is also worth mentioning that this

would not be puzzling at all if one’s account maintains only fundamental-level causes, because

there is no need for realism about high-level causal models, but this is not the position taken by

Woodward and Mitchell.

Strevens’ account in particular highlights a different concern about the prospects of

causal explanation in general. Strevens relies on a known web of causal influence, among which

difference makers are selected for a particular explanandum, but this starting point is in serious

trouble if there are no causes at the fundamental level. Fundamental physics (i.e. quantum

mechanics) is non-causal and does not entail the propositions of classical physics, nor describe

its causal influences. Taking this seriously would preclude all classical explanations, because no

classical model counts as an abstraction of a fundamental model. In which case, if one cannot

abstract smoothly up from the base level then there are no high-level explanations. This goes

beyond the epistemological and methodological concern raised above. It is widely understood

that at the level of fundamental physics there are no classical causal processes. If one takes

Strevens’ requirement on abstraction literally, it is only applicable in a classical world where the

lowest causes are kinetic interactions between atoms or molecules. And thus, there is no place

for the fundamental causes to enter.

One last concern is a very general epistemological one. I have mentioned it as a drawback

that Woodward’s and Strevens’ accounts begin from correct causal knowledge: knowing that I

causes X; knowing how to arrange a directed graph; and beginning from a complete picture of

fundamental causes. The reservation about beginning from this point is that it arouses a

perennial, epistemological concern stemming from Hume about how one gets there in the first

100

place. Kitcher has noted that philosophers concerned with explanation once strongly avoided

invoking causation. To borrow his words, appeals to causal knowledge were seen as appeals

“that would make life so much easier if only they could be made” (1989, p. 460). In the wake of

logical empiricism, invoking causes has become commonplace. But, the concern that remains is

not that it offends empiricist sensibilities. The concern is that we cannot justify our inferences to

causal claims and so we begin with them. I sometimes see these starting points as a kind of

conditional: if one has knowledge of causes, then many of the issues surrounding causal

explanation disappear. Indeed.

101

Chapter 4. Deductivist Explanation

Introduction

This chapter will review the tradition of deductivist accounts of explanation to survey their

successes and failures in order to inform the account I propose in Chapter 5. Hempel and

Oppenheim began the tradition of analyzing scientific explanation and offering a metatheory

about what an explanation does and what it ought to do (Hempel & Oppenheim, 1948). Their

deductive-nomological (D-N) account has met with many objections. Some argue that the

account is not necessary, because there are many kinds of important scientific explanations that

are outside the scope of the account, such as explanations that make use of statistical regularities.

Others have argued that the account is not sufficient for explanation, because many non-

explanatory derivations meet the D-N criteria. These criticisms led philosophers to turn to a

stronger notion of causation as a mean of debarring certain of these now well-known

counterexamples. The D-N account was formed around the idea that explanation was derivation

– that an explanation should take the form of an argument which derives the explanandum from a

set of sentences that contain at least one general law (Hempel & Oppenheim, 1948).

An explanation functions like a syllogism. One has a law or set of laws as the major; as

the minor, there are sentences of particular facts about the antecedent conditions that are

subsumed by the law or laws; and we find the explanandum as the deductive conclusion. These

explanations are intended to be strongly relevant to the explanandum, since they do more than

simply give good grounds for phenomena, they logically entail them. Questions of the sort ‘how

did this phenomenon happen?’ are regarded as ‘according to what general laws and by what

antecedent conditions does the phenomenon occur?” This account applies not only to particular

phenomena, but can also be applied to explain regularities. This occurs in the same way, i.e., by

subsuming one law under a more general one. It is in this way that one can account for the truth

of Galileo’s laws, since they can be deduced from Newton’s laws of motion.

Section 4.2 begins by reconstructing Hempel and Oppenheim’s D-N account and then

reviewing some of the challenges that have been issued regarding its necessity and sufficiency.

102

This involves reviewing the role of statistical generalizations in explanation, and reviewing some

counterexample cases. In 4.3, I turn to Kitcher’s account of explanation, which sought to

supplement the D-N account with a formal account of the explanatory power of theories by

unification. After looking at Kitcher’s proposed solutions to the D-N’s original problems in

section 4.3.1, I review some objections that have been raised for unificationism. In 4.4, I then

present an overview of the current state of deductivist accounts, looking at what challenges

remain to be adequately dealt with and in which direction some solutions may lie. The results of

this directly inform the model-based deductivist account I propose in Chapter 5.

The D-N Account

Hempel and Oppenheim begin by distinguishing explanation from description by distinguishing

knowledge how from knowledge why. When put in these terms it is not unforeseeable that some

would question whether science can indeed offer explanations over and above descriptions.

Hempel and Oppenheim set out to characterize the kinds of deductive arguments that could be

said to do such explaining. The D-N model they outlined was the first serious attempt at showing

that science is concerned with explanations.

Hempel and Oppenheim analyse explanation into the explanans and the explanandum.

The latter is the phenomenon to be explained and the former is what does the explaining.

Explanans are subdivided into two: the antecedent conditions and the law statements. Together,

these jointly entail the explanandum. The derivation serves to give explanatory information

about the explanandum’s occurrence by showing under what conditions and according to which

scientific law or laws it was to be expected. In order to qualify as an explanation, the derivation

must satisfy certain conditions of adequacy R1-4 as follows (Hempel & Oppenheim, 1948, pp.

137-138):

R1. The explanandum must be a logical consequence of the explanans.

R2. The explanans must contain general laws.

R3. The explanans must have empirical content.

R4. The sentences of the explanans must be true.

They give the following explication, where T is the law statement, and C is the antecedent

conditions:

103

<T, C> is a potential explanans of singular sentence E (explanandum) only if:

1) T is essentially general and C is singular, and

2) E is derivable from T and C jointly, but not from C alone.

They are careful to note that this is not a definition but a necessary condition. It lacks

sufficiency, which would allow any statement to be joined to a theory and constitute an

explanation. In order to prevent some false explanations via the quirks of truth-functional logic,

they add the following restriction that

3) T must be compatible with some basic class of sentences which has C but not

E as a consequent, i.e., T must be verifiable without reference to E, in order to

avoid circularity.

(Salmon, 1989, pp. 20-21)

The explanandum also need not be a single occurrence but could be a kind, thus giving the

explanation universality. This would take the form of ‘in all cases of kind F, conditions of kind G

are realized.’ But not all statements of this form are true laws of nature. For example, Kepler’s

and Galileo’s laws are only considered approximations. Statements concerning restricted cases

would also display the same form but fail to be true universals, because they are only accidental

generalizations. The difference between accidental and true generalizations is not easy to

articulate. Goodman finds it in the ability of laws to support counterfactual and subjunctive

conditionals about potential instances (Goodman, 1973). But any universal statement can only be

counted as a law if it is implied by the accepted scientific theories at the time, and will not

qualify as law if it precludes hypothetical occurrences that an accepted theory finds possible.

In order for explanation, only laws of nature and not accidental generalizations can be

featured in the derivation. True laws are those that express real empirical regularities. It might be

true that all the coins in my pockets are quarters, but it is a law that all gases expand when

heated. Distinguishing between true laws and accidental generalizations has been problematized,

as reviewed in 2.2.2. Hempel and Oppenheim introduced a formal language in which to

formulate laws. It is essentially a standard first order calculus with no open statements. Hempel

resisted formulating a general account of laws. He did not think it was necessary as long as we

can recognize and agree on whether a generalization is a law. We need not know why a

generalization is a law, as long as it is one. Even though Hempel would not provide a definition

of a law, he outlined four characteristics. A law must be (1) universal, (2) unrestricted in scope,

104

(3) not referring to particulars, and (4) contain only purely qualitative predicates. This

characterization of law has been criticized as too stringent and not necessary for explanation by

Woodward, Mitchell, and others (Achinstein, 1971; Mitchell, 1997; Lawton, 1999; Mitchell,

2000; Woodward, 2000; Frisch, 2014). A laxer characterization of explanatory regularities would

open the door to more kinds of explanation, including those citing probabilistic causes, such as

‘smoking causes cancer’. The literature since Hempel has been moving in the direction of a more

inclusive conception of law or explanatory generalization.

Hempel argued that causal explanations can be formulated in the character of a D-N

explanation, but that not all D-N explanations are causal. For instance, Kepler’s laws of motion

can be explanatorily derived from Newtonian mechanics, but this is not in virtue of Kepler’s

laws being caused by Newtonian mechanics. Thus, the D-N account is capable of supporting

causal explanations, but is more general. This empiricism implied a regularity account of

causation, which enabled Hempel to respond to certain criticisms about the D-N but also

generated others, as will be shown in the following subsections.

4.2.1. Is it Necessary?

The D-N account has met with numerous concerns that chiefly fall into two categories: concerns

that it is not necessary for explanation (it is too narrow), and concerns that it is not sufficient for

explanation (it is too broad). Together these are taken to show that satisfying the D-N criteria is

not really relevant to capturing explanations. It is a self-acknowledged limitation of the D-N

account that it is not necessary for explanation; there are many kinds of explanation that do not

meet these criteria. However, some hold that this is a serious limitation of the account, because

there are important kinds of explanation that ought to be included, like causal explanations that

make no reference to laws, and statistical explanations.

Michael Scriven argued that many explanations are causal and make no reference to laws

(Scriven, 1962). For instance, his claim is that a statement like ‘the impact of my knee on the

desk caused the inkwell to tip over’ is an explanation of the tipped-over inkwell. This is often

referred to as a singular causal explanation, and notably, it makes no explicit reference to a law.

How is the D-N account to handle such simple and common cases? Hempel’s response is to

argue that the use of ‘cause’ in that sentence is indicative of a causal regularity that links knee

impacts on desks under certain conditions with the tipping over of inkwells (1965a, p. 423).

105

Thus, what makes the single sentence causal statement and explanation of the event is that it, at

least implicitly, shares the D-N form in virtue of a causal regularity.

Explanations featuring statistical regularities are another such example. Statistical

explanation is outside of the scope of the D-N account, because it is not deductive. It is

nonetheless prevalent and important in science, such as in the predictive use of radioactive decay

and in the basic laws of genetics. Because statistical regularities are not universal and necessary

laws, it is always possible that they be undermined by new information. When the statistical law

fails to predict, it has no explanatory significance for the case at hand. The fact that a statistical

explanation could have the same true premises and yet would yield both favourable and

unfavourable predictions is what Hempel calls “the ambiguity of statistical explanation”

(Hempel, 1965a, p. 382). Hempel was pressured by critics into extending the D-N account to

include statistical or probabilistic explanations.

4.2.1.1. Statistical Explanation

Hempel distinguishes two types of statistical explanation, the deductive and the inductive. The

deductive-statistical (D-S) performs much like the D-N, but it deduces one statistical uniformity

from a more general statistical law. The inductive-statistical (I-S) involves the subsumption of

events under statistical laws. This kind of inference offers explanation where it finds a high

degree of probability to an event. For instance, if one wants to know why they failed to roll three

sixes on three dice, the high probability of this failure is explanatorily relevant. Thus, while not

guaranteeing the explanandum, a degree of rational expectability can still be conferred to the

explanandum, given the high probability in the explanans. In certain cases, it definitely seems as

though there are genuine statistical explanations.

Nicholas Rescher was one who made a plea for the inclusion of statistical explanations

(M. King, 2014). Hempel responded the same year with an inclusive account (M. King, 2015).

These explanations also take the covering law form of following from laws, but from statistical

laws. They, therefore, cannot be arrived at with deductive certainty, but with at most high

probability. It usually takes the form of:

P(G|F) = r

Fb

====== [r]

Gb

106

The idea behind these explanations is that if one can give a statistical law that says that an event

G is highly likely given another event F, and one looks at an example of an F event, then one has

an explanation of the G event (Hempel, 1965a, pp. 383-384). While this does grant predictive

value, it fails to necessitate G’s occurrence. This is simply because the same explanans could be

true even if the G event had not taken place. The problem is that one can generate an acceptable

derivation with either outcome.

Inductive logic has no weakening principle like deductive logic. The addition of a new

and contradictory piece of evidence can conclusively refute a previous strongly favored inductive

argument. Thus, there is a principle of total evidence, which makes the claims into true

accidental general statements, which cannot be refuted by new evidence. But if the conclusion of

such an induction is included among the premises, then the argument is not inductive at all, but

deductive. So Hempel had to look for a less stringent requirement than that of total evidence.

This he called the requirement of maximal specificity, which requires that the conditions reflect

all the relevant information about the specific situation. Hempel holds that it is high-probability

regularities (𝑟 > .5), are the ones that can support explanations. Hempel maintains that

explanations confer nomic expectability, even given that the covering law can either be statistical

or universal.

The ambiguity of I-S and other problems led philosophers to offer alternative accounts of

statistical explanation. Salmon raises several criticisms of the I-S account. He argues that it is

unnecessary for explanation, because it is unable to handle cases where the probability is low. If

the probability of an event occurring is 1%, then for Hempel it is not explanatory, even if it is the

only known explanation. This is what was seen in the case of the mayor who develops paresis in

1.3.1. Salmon also argues that it is not sufficient, because derivations meet the I-S requirements,

but are not explanations. For example, consider the case where John has a cold and is taking

vitamin C. There is a statistical relation between the taking of vitamin C and a cold’s

disappearance in a week. Because colds generally clear up in a week anyways, the fact that he is

taking vitamin C ought not count as an explanation of the cold’s disappearance. Salmon’s

solution to the problem is to argue that what is important is a change in the probability of one’s

getting well in a week. Salmon provides an account of causation which replaces high probability

of the explanandum with the statistical relevance of the explanans on the explanandum. Salmon’s

107

strategy required that he compare a prior and posterior probability, and thus establish that the

result is due to the intervening cause.

Kitcher’s strategy is to reject the idea of statistical explanation altogether. On Kitcher’s

view, it is a mistake to think that individual occurrences are explained by statistical regularities.

Instead one merely offers a deductive explanation of the probability itself, like a D-S

explanation. Kitcher’s position is that all explanations are deductive; what he acknowledges as

deductive chauvinism. This position will be examined more closely in 4.3.1, when considering

further counterexamples to the D-N and explanations in quantum mechanics.

4.2.2. Is it Sufficient?

Some have argued that the D-N account is not sufficient for explanation, which is to say that a

derivation might meet all requirements of a D-N explanation, yet still not be an explanation. This

is more problematic then there being good explanations that are outside the scope of the D-N

account. The worries about the sufficiency of the D-N account come in two main varieties. There

are those that expose the symmetrical nature of derivations to generate non-explanation. These

cases run an explanatory derivation in reverse, while still satisfying Hempel’s criteria. There are

also cases that expose the irrelevance of the generalization in necessitating the explanandum.

Together the explanans are sufficient to guarantee the explanandum, but are irrelevant to its

actually being the case.

Many argue that problems of symmetry arise when the explanans and explanandum do

not stand in the right causal relation to each other. Generally, the problem is that one can derive

the explanandum E, by means of a general law L, and initial conditions C, and meet the

requirements of the D-N account. However, in some instances, one can also derive C from L and

E and meet the requirements of the D-N account, yet this is not a good explanation. Let us briefly

consider the counterexample of the flagpole, as raised by Bromberger (1966).

In this example, there is a flagpole that is casting a shadow on the ground (Fig. 4.1). With

the D-N account, one can explain the particular length of the shadow, s, by the initial conditions

of the height of the flagpole, h, the angle of the Sun θ, and electromagnetic laws about the

straight-line propagation of light L. However, one can use the same laws coupled with the length

of a flagpole’s shadow to explain its height. This is an incorrect explanation, because clearly this

is not the reason why the flagpole is the height it is.

108

(Fig. 4.1)

This is not a particularly unusual case. Another example that have been raised concerns the

derivation of a simple pendulum’s length from its period. The mathematical regularity allows us

to know the length if we know the period, but this is not an explanation of why it is the length

that it is. The regularity alone does not possess information about the causal structure, or give

counterfactual information, which could potentially debar these kinds of cases. These cases

reflect the symmetric nature of law-like generalizations and the covering-law nature of D-N

explanation. The D-N account is not sensitive to the asymmetry that a good explanation

sometimes demands.

Hempel maintains that no general account of laws is necessary for the success of the D-N

account, but this relies on clarity about what counts as a scientific law. A brand of

counterexample is to employ generalizations that are ostensively laws, but that fail to support

explanations. Some argue that one of the most vexing problems for the D-N account is this

characterization of law sentences. A particular problem arises when the generalization used in

the derivation is true but irrelevant. One ends up with a sound argument, but one that does not

explain why the conclusion is true. One can think of this as a kind of epistemic luck, reminiscent

of Gettier cases of knowledge. One ends up with good grounds for believing an explanation, and

the explanans and explanandum are true, but the explanandum is not true in virtue of the

explanans. This happens where the generalization is true, but it is not relevant to the occurrence

of the explanandum.

A famous counterexample to demonstrate the irrelevancy that can obtain in a D-N

derivation is the hexed salt example (Kyburg, 1965). It is presumably a true law that all table salt

that has been hexed with the wand of a witch will dissolve in water. Thus, one can putatively

𝜃 s

s

h

109

explain the dissolution of a sample of hexed salt by citing the law that all hexed salt dissolves in

water. Citing that the reason that some salt dissolved was because it was hexed and all hexed salt

dissolves satisfies the D-N criteria, but is clearly not an explanation of why the salt dissolves.

Another example is that Jones takes birth control pills and it is true that taking birth control pills

will prevent pregnancy. This ought to imply that, on the D-N account, the reason why Jones does

not get pregnant is because he is taking birth control pills. This is obviously not a good

explanation of why the explanandum is the case.

These examples strike me as cases of common place explanations that are not intended to

be covered by an account of scientific explanation. Yet this is not a response that Hempel gives.

This is reflective of the idea that there is strong continuity between ordinary and scientific

explanation, such that results like this have consequences for accounts of scientific explanation. I

stated in Chapter 1 that covering common-sense explanations is not a condition for a successful

account of scientific explanation. However, these cases of irrelevancy seem to demonstrate that

nomic expectability is at most an answer about how we know something to be the case, but not

why it actually is the case. Many see the problem as the account’s inability to track causation.

The D-N’s regularity account of causation does not strictly lead through sound argument to the

fact that P together with Q caused r, but only states that it conforms to a regularity. If one finds

that the satisfaction of the D-N criteria is not sufficient to generate an explanation, then one

might offer an alternative causal account, or one might ask what additional criteria might help.

Philip Kitcher’s holds that nomic expectability plus unification is the answer.

Unificationism

Along with logical empiricism’s official account of explanation (D-N/I-S) was the idea that

scientific explanation has been achieved with the goal of unification in mind. Science aims at

maximum explanations with the minimum possible theoretical concepts and assumptions. For

Hempel and Oppenheim, an explanation is given in subsuming particular phenomena under

general theories, because theories have systematic power. For Hempel, it was important that the

theory featuring in the explanation can make “systematic connections among the data of our

experiences, so as to make possible the derivation of some of that data from others” (Hempel &

Oppenheim, 1948, p. 164). The ability of some theories to derive large amounts of data from a

small amount of initial information speaks to their explanatory power. The official account, as it

110

is given by Hempel, is criticized as being too stringent, and far too liberal in accommodating

intuitively non-explanatory derivations. What is worse, according to Kitcher, our ability to derive

a “description of a phenomenon from a set of premises containing a law seems quite tangential to

our understanding of the phenomenon” (Kitcher, 1981, pp. 508-509). Nomic expectability is not

always clearly connected to our understanding of why; it is not enough to guarantee explanation.

Hempel’s account is fraught with difficulty, but there is an ‘unofficial account’ of logical

empiricism that makes use of this concept of unification, and it is this that Kitcher will develop.

As Hempel himself says, “what scientific explanation, especially theoretical explanation, aims at

is... an objective kind of insight that is achieved by a systematic unification, by exhibiting the

phenomena as manifestations of common, underlying structures and processes that conform to

specific, testable, basic principles” (Hempel, 1966, p. 83). This story involving unification,

Kitcher argues, can be much more easily connected with our understanding than mere nomic

expectability. Kitcher’s approach, which follows in this tradition, is to assess the worth of

explanations by their unification with in a systematic picture of the order of nature (Kitcher,

1981). To explain is to show that a sentence is appropriately related with the explanatory store of

scientific knowledge. Much of Kitcher’s account involves specifying precisely what this means.

Unificationism really started with Michael Friedman, who was the first to look deeply at

explaining not only single events, but regularities (Friedman, 1974). According to Friedman, the

subsumption of a regularity under another is what makes it explanatory. Reducing a regularity

replaces two facts or more with one. It is the unification, or integration, of one regularity under

another that provides understanding. The reduction of the total number of disparate phenomena

is a primary goal of science and this account of explanation stresses that. What this involves is

reducing the number of theories and regularities that are needed to account for the facts. Theories

that are the most unified in this way are explanatory.

Kitcher’s task is to show what it means to say that an explanatory theory is one which

best unifies knowledge. Friedman conceives of it as a trade-off between the minimization of

theses and the maximization of conclusions reached. For Kitcher, the degree to which a theory is

unified is determined by three criteria. The first is similar to Friedman’s idea that unification

involves a reduction in regularities. Instead of this, Kitcher proposes the idea of the reduction of

what he calls argument patterns. An argument pattern is a triplet consisting of a schematic

sentence with dummy letters that can be filled in, a set of filling instructions that specify how the

111

dummy letters are to be replaced, and a classification that determines how a set of schematic

sentences can be arranged as premise and conclusion in the form of an argument. On Kitcher’s

unification, what is minimized is the number of patterns that a theory employs for deriving

conclusions. The second mark of unification is a theory’s empirical scope; the greater the

empirical scope, the more unified it is. A theory unifies our beliefs if a small number of

argument patterns can be used to derive a great range of conclusions. This makes room for

argument patterns to be very general, in fact, so general that clearly non-explanatory derivations

would qualify. Thus, he introduces the last central aspect of unification, viz., the stringency of

the argument patterns used. An argument pattern is said to be more stringent than another if there

are more restrictions on the arguments that instantiate it. Kitcher proposes that scientists are

concerned with stringent patterns that place restrictions on the substitution conditions for dummy

letters and on the logical structure imposed by classification. If one relaxes both conditions, the

pattern admits of more and more arguments, and conversely if one tightens restrictions on both,

then it admits of fewer and fewer. It can be seen that these criteria pull away from each other.

The fewer and the less stringent the argument patterns, the more likely the theory is to lead to a

wider range of conclusions, and vice versa.

For Kitcher, good explanations are instances of patterns that rank better along these

criteria than derivations that we consider bad explanations. These derivations are available

explanations, and so they are in what he calls the explanatory store. The set of argument patterns

that most unify a set of accepted sentences, K, is the explanatory store over K, which he denotes

as E(K).

Kitcher points us to two prime cases where we can see the explanatory power of

unification at work: in the reception of Darwin’s theory of evolution, and in the wake of

Newton’s mechanical theory. Newton’s successes prompted others to take on an even more

ambitious enterprise called dynamic corpuscularianism, which sought to unify all the phenomena

of nature in a single framework. It encouraged Newtonians to construct corpuscular theories of

everything, including light, even in the absence of evidence. Its main appeal was the promise of

unification. The hope was that one kind of force law would suffice to describe all interactions in

the same way that gravitation was ruled by a single law. A small number of general patterns of

argument were sought to explain all of nature – this was the Newtonian ideal.

112

Similarly, the attraction of Darwin’s theory was its ability to unify a host of biological

phenomena in a small number of common patterns. The theory presents a pattern to describe

imaginary examples that can be used to explain the features of any existing species. It offers

explanation-sketches by describing how particular traits may be advantageous and contribute to

the survival of the species. Kitcher claims that it was the argument pattern that made the theory

compelling and that is precisely what characterizes unification and explanation.

Kitcher addresses the problem of what he calls spurious unification. The problem is that

it is possible to derive for example, a law L, from L and an arbitrary conjunction with another

law B: L & B, therefore L. This self-explanation is clearly not what Kitcher has in mind, and

opens up the door to unifying our beliefs completely by this one simple argument pattern. His

answer to this is that such derivation may win in least number of argument patterns, but fails

when it comes to stringency, since it would allow any vocabulary to fill the dummy letters. Even

though one can strengthen this problem by artificially introducing restrictions on the pattern, the

accidental quality of the restrictions will never fail to provide argument patterns as one changes

the filling instructions for the pattern. By contrast, if one considers the Newtonian pattern, the

constraints are essential to it, and cannot be amended without destroying its stringency.

So after having found a way to distinguish the genuine from the spurious unification, he

makes this requirement explicit. If the filling instructions can be replaced to yield a new pattern

which allows the derivation of any sentence, then the unification is spurious. This new

requirement will also be able to decide against the unification of doctrinal arguments, which may

make claims to unification by such laws as ‘What God wants to be the case is the case”. Because

the restrictions of the filling instructions in this case are so liberal as to bar almost nothing, if

anything at all, the unification is clearly spurious. He assures us that this requirement is not out

of place, and is closely tied with the idea that explanations should be testable. If the argument

pattern unifies incredibly well, but makes no restrictions on its possible filling instructions and

which conclusions it is capable of accommodating, then it is spurious.

4.3.1. Unificationist Solutions to D-N Problems

Kitcher proposes that the unofficial account is capable of solving the three most challenging

problems that remain for the D-N account. Kitcher defends a position that is independent of his

account of unification, but which would help circumvent problems that arise for statistical

explanation. This position has been called “deductive chauvinism” by Salmon and Coffa. For

113

Kitcher, though, it is a virtue and not a defect of his account. Kitcher says that “in a certain

sense, all explanation is deductive” (1989, p. 448). He maintains that the explanatory store

contains only deductive arguments, and so it actually prohibits the possibility of inductive

explanation. The objection might be raised that if D-S were sufficient to characterize

explanation, then Hempel would never have needed to talk about I-S explanation in the first

place. Kitcher thinks that the idea that there is a deductive explanation that is being replaced by

an epistemic probability is not sufficient to account for all cases of probabilistic explanation.

This can be clearly seen in the cases where there is no, or no likely, deductive story of particular

occurrences – for instance, in quantum mechanics.

Kitcher argues that there are two senses of ideal explanation. In the first sense, it is a

deductive derivation, but in the other, it is the best explanation the phenomena will admit.

Quantum mechanical explanations are ideal in the second, but not the first, sense. Consider a

case where there is a 0.9 probability that a barrier will reflect a particle, and given two instances,

one where a particle 𝑒1 is reflected and one 𝑒2 where it penetrates, an important question must be

asked to whether it is possible that there is an ideal explanation of both cases. Kitcher argues that

there is not. There can be no ideal explanatory account, because there is no information to

distinguish these cases. There is no account of why 𝑒1 was reflected and 𝑒2 tunnelled through.

Kitcher thinks that we mistake what it is that quantum mechanics explains. Quantum

mechanics can explain how things are possible by allowing them to be possible results, but it

does not explain particular instances. Probabilistic explanations in quantum mechanics take on

the form of D-S explanations – they explain probabilities about individual outcomes. So what is

explained is not why they occur, but why they occur with a certain probability. One has an

explanation when one has facts about barrier penetration for example, that are derived from a

generalization like the Schrodinger equation. It is easy to confuse a why-question with the how-

possible questions that quantum mechanics seems to address. The why-question it answers is

about why the probability is 0.9, it is a question of the general occurrence. Kitcher argues that the

need for I-S explanations comes from a confusion of the kinds of questions asked about quantum

mechanics. Kitcher’s solution is simple and radical: all statistical explanations of individual

events need to be spelled out in terms of deductive arguments about the probabilities, rather than

inductive arguments about the event itself.

114

Unificationism proposes to solve the problems of asymmetry and irrelevance that

troubled the D-N account. Kitcher’s strategy is to compare the putative explanations that fit some

deductive argument pattern but are not explanatory to another possible explanation that also

derives the explanandum, and showing that the former does not best unify our set of accepted

statements, K. Thus, it is not explanatory; not a member of E(K) and not an explanation at all.

This marks an important aspect of unificationism, which is that only the theory that best unifies

our knowledge is explanatory. This is known as the winner-take-all conception of explanation,

which was mentioned in 1.4.2.

Even though one can derive the length of a simple pendulum by looking at its period, the

problem as Kitcher sees it, is that this is not what we normally take to be an explanation of the

dimensions of manmade bodies. Those we normally take to be explanations of artifacts, he calls

the “origin and development” type of explanations, as Kitcher calls them. If we consider the case

of the flagpole, it is surely true that some objects do not have shadows or cannot always lend

themselves to deriving facts about the object’s dimensions. What is one to do about explaining

the dimensions of these kinds of bodies? Adopt a separate and quite different argument pattern,

or pick the origin and development explanation that is capable of explaining both? The best

explanation for unificationism is not the counterexample case. It is the origin and development

pattern, which is more widely applicable in explanations of the dimensions of manmade bodies,

and thus more unifying.

The unificationist solution to the case of irrelevance is to once again compare the

troublesome case to a more reasonable explanation and show that it is not the most unifying.

Given the explanation that employs the fact that all hexed salt dissolves in water, what is one to

do about instances of the dissolution of unhexed salt? One could either maintain two separate

explanations, one for each case, or one could simply choose the explanation which is capable of

covering both cases, viz. that all salt is water soluble. The second option is clearly more unifying

and instantiates a pattern that is much more generally applied. He concludes then that

“unificationism has the resources to solve some traditional difficulties for theories of

explanation” (Kitcher, 1981, p. 526). While the counterexample cases might meet deductivist

criteria for nomic expectability, it does not instantiate the most unifying argument pattern.

Kitcher made several contributions to deductivist explanation. He developed an account

of explanation with different aims than Hempel’s. He sought to provide an account of how laws

115

confer explanatory power, something on which D-N account remained silent. By developing the

unofficial view of logical empiricist explanation he was able to propose solutions to vexing

problems. In denying that inductive statistical explanations exist he is able to avoid the problems

of the ambiguity of I-S, without making use of a high-probability requirement or maximal

specificity. The comparative nature of unificationist explanation helps Kitcher determine that the

counterexample cases ought not be considered explanatory and thus defend deductivist

approaches to explanation. However, many have concerns about unificationism and these

proposed solutions, and it is to this that I turn next.

4.3.2. Challenges to Unificationism

There are a number of serious challenges to unificationism and to the proposed solutions to the

D-N counterexamples. I will bring up a few that I find to be most troubling and relevant to what

has been, and will be, discussed.

Some concede that Kitcher’s defense of the flagpole case works well enough, but that it is

not a general result. Barnes has noted that when one considers a time-symmetrical system like

the Newtonian mechanical description of the solar system, one finds that there are as many

retrodictive as predictive derivations (Barnes, 1992). The argument patterns for the retrodictions

are as unified as those of the predictions, and thus contrary to our judgments, they are equally

explanatory on the unificationist picture. Woodward has reinforced this criticism by showing that

the more general problem is that there are many kinds of unification and not all of them are

relevant to explanation (Woodward, 2003). Some unifications that count as explanatory for

Kitcher are no more than the application of a common mathematical formalism to different sorts

of phenomena. Woodward argues that “the mere fact that we can describe both the behavior of a

system of gravitating masses and the operation of an electric circuit by means of Lagrange’s

equations does not mean that we have “unified” gravity and electricity in any physically

interesting sense” (Woodward, 2003, p. 363). It could be argued that the argument patterns

involve more than merely making use of a set of equations, such that the patterns are only the

same for very non-stringent characterizations.

Woodward notes that this raises the question of how one is to trade off the criteria of

stringency, paucity, and scope, against one another. One theory is more explanatory than another

if it can derive a wider range of phenomena with fewer, more stringent argument patterns. But,

these criteria pull apart, and there’s no rule or procedure for how to weigh them against each

116

other, even though this weighing is quite important. Kitcher’s solution to the flagpole

counterexample relied on a single origin and development pattern for explaining the dimensions

of objects, but this is a very non-stringent pattern. It is quite different from patterns used to

explain the dimensions of objects with biological or geological origins. In order to debar

counterexamples, a lot turns on “correctly” counting argument patterns and assessing stringency.

A different concern for unificationism is the winner-take-all conception of explanation.

Recall that according to Kitcher, only the most unifying pattern that derives the explanandum is

explanatory. This is an essential feature of the account that enables him to debar the

counterexample cases. However, it leads to counterintuitive judgments about certain derivations

being non-explanatory. One compares the degree of unification of two competing theories and

the one that unifies more is explanatory, and the other not at all. Many have argued that this does

not follow. I stated in Chapter 1 that it seems reasonable to hold that two derivations of a

phenomenon can be counted as explanatory. However, this is not tenable on the unificationist

view, because then the case of the hexed salt and the flagpole’s shadow pattern are also

explanatory. However, I do not believe that these problems are ineliminable for deductivist

accounts, as I will show in the following chapter.

Conclusion and the Current State of Deductivism

There are two main motivators for thinking that the D-N/I-S account can cover most scientific

explanations. First, because of its deductive structure, Hempel says that it provides an answer to

why the particular phenomenon occurred. It says why the result was expected and allows us to

understand its occurrence. The I-S model does not show expectation with certainty, but high

probability – they both share in conferring nomic expectability. Secondly, it only requires a

regularity theory of causation, namely, the laws it uses. Following in the empiricist tradition,

Hempel construed causation as the obtaining of regularities, but spoke no more about the

metaphysics of causation. A benefit of the account is that one could avoid talking about the

metaphysics of causation and scientific realism. Instead one could talk about entities and

dependency relations in terms of scientific explanation. This also more accurately reflects the

fact that many kinds of scientific explanations are outside the scope of causal accounts.

Kitcher proposes supplementing the D-N with a formal account of the explanatory power

of theories. The hope is that the extra criterion that theories that feature the deductive argument

117

patterns in the explanatory store can debar three of the main objections to deductivist explanation

(statistical explanation, irrelevance, and asymmetry). Deductive chauvinism seems to provide a

means of circumventing the myriad problems that inductive explanations generate, many of

which I passed over. Moreover, making use of a comparative, or winner-take-all conception of

explanation, allows Kitcher to prefer standard explanations to the counterfactual cases, which are

less unifying. However, some serious problems remain for unificationism, and so it is not clear

that the current state of deductivism is very promising. However, I aim to show that prospects are

better than is commonly thought.

There are several lessons for deductivist approaches that can be taken from this

investigation. One is that counterexample cases seem to demonstrate that there is a difference

between explanation and prediction. Being able to make statistical inferences is not the same as

giving a statistical explanation, and perhaps making deductive inferences is not the same as

giving deductive explanations either. What is required is that if the explanation proceeds by

covering law, then there needs to be something about the regularity that guarantees its relevance

to the explanandum. This should avoid the problem of being able to derive the explanandum by

accident, so to speak. While nomic expectability is not enough, nomic expectability plus

unificationism is a step in the wrong direction.

Providing an account that is capable of handling counterexamples cases will also require

something beyond nomic expectability plus some other criteria. It ought to be possible to

preclude problematic symmetrical derivations and to constrain the relevance of the explanans for

the occurrence of the explanandum. The symmetry problem follows from the symmetrical nature

of the logical form of deduction and the problem of irrelevance follows from the covering law

conception. What these problems seem to require is to make use of facts about the empirical

content of the explanation. Purely syntactic restrictions are unlikely to be able to reflect the

asymmetry and the close relevance relations that some explanations require. This might go a

long way towards showing that a deductivist account can be sufficient for explanation.

I think that Kitcher was correct in his claim that deductive chauvinism is a virtue of his

account. It is also something that Hempel endorsed, to a lesser degree, about deductive

explanations. It is also important to note that this position is independent from unificationism.

One of the motivations for a deductivist account is that it opens up the range of possible

explanations to include those that are non-causal. Causal explanation is merely one kind of

118

explanation. Being able to accommodate singular causal explanations and denying that there are

real statistical explanations removes perceived limitations on the scope of a deductivist account.

With deductive chauvinism, one can hopefully mitigate the counterexamples that showed that

deduction is unnecessary for explanation.

A pervasive criticism of deductivist explanation is that it is largely irrelevant to the

actual practice of scientific explanation. The way that explanations proceed is not by subsuming

particular instances under general laws. Even for Kitcher, individual explanations are still D-N

derivations. There is a trend in the philosophy of science to move towards explanations that

feature models. The importance of this is something I discussed in Chapters 1 and 2, and in the

following chapter, I hope to incorporate that scheme into a deductivist approach. The plan is to

synthesize the results of the previous investigations and show that there can be a promising and

relevant account of deductivist explanation.

119

Chapter 5. Model-based Deductivism

Introduction

Having taken stock of the resources and the shortcomings of deductivist approaches, I am now in

a position to present a novel, model-based deductivist account of explanation. I start from the

position that takes explanations to be deductive arguments. On this view, the explanans is a set of

statements about a scientific model that is used to derive the explanandum, which is a statement

about, or a description of, a target phenomenon, pattern, or behaviour. I take scientific models to

be the objects of explanation, in part because I and many others feel that covering law accounts

are largely irrelevant to the actual explanatory practices of science. Many explanations offered

do not simply derive explananda from laws of nature. This is a very restrictive selection of the

kinds of explanations scientists are actually offering.

As was shown in Chapter 2, the recent literature on explanation has focused on model-

based accounts of explanation. Some accounts require that explanatory models reflect the real

causal or structural relations of a target system (Woodward, 2003; Bokulich, 2008; Strevens,

2008). In place of a causal or structural restriction, I propose that models that support

explanations are those that give appropriate counterfactual information and are integrated with an

established scientific theory. This integration is what gives the model its unifying power and

demonstrates its ability to explain. This requirement has the added benefit that it does not place

any representative or metaphysical restrictions (causal, structural, or otherwise) on the relations

of the model itself. This expands the scope of explanation to include non-representing, or highly-

idealized, models; a goal that has been argued for by many, mentioned in Chapter 2.

The goal of the chapter is to present an updated, model-based version of a deductivist

account and demonstrate its promise. A model-based approach allows the explanation to give

counterfactual information about the system, but does not require that the relations among its

variables represent real causal dependency relations. As a deductivist account, it both confers the

nomic expectability that Hempel desired and remains open to causal and non-causal

explanations. The incorporation of counterfactual information about the facts of the models helps

120

to establish relevancy and asymmetry. The criteria as presented at this stage are still preliminary

and in need of being tested and refined in the crucible of detailed case studies.

The following section begins to build up an account from the basic deductivist criterion

that the explanans entails the explanandum. In 5.3, I revisit what a model is in this context and

begin to introduce some restrictions on what kinds of models can support explanations. In 5.4

and 5.5, I articulate the local criteria for explanation. I argue that an explanatory model must

support same-object counterfactuals about changes to the model. In 5.6, I discuss the strategy of

placing a global constraint on explanatory models such that they must be appropriately related to

a global theory of science. In 5.7, I present the account as it currently stands, and in 5.8, I put the

account to work in reviewing some of the case studies that have already been brought up in light

of the account I am proposing. I hope to motivate the deductivist approach, demonstrate its wide

scope, its ability to clarify thorny issues of explanation, and its agreement with our judgments

concerning explanation.

A Model-based Deductivist Account

Let me now propose the formal criteria for this account of scientific explanation. Let us begin

with the simplest and most central criterion of a deductivist account. This is the criterion of

deductive entailment, which is as follows:

D1. The explanandum must be a deductive consequence of the explanans.

This criterion is the backbone of deductivist explanation. This alone is enough to preclude many

putative explanations. Just like Kitcher’s deductive chauvinist position, I maintain that all

genuine scientific explanations are deductive. Many reasonable explanations are not deductive.

They may cite single-event causal stories, or accepted inferences, say from lightning to thunder.

My claim is not that these are not explanations, but that are not scientific explanations. This

suggests some degree of discontinuity between the two, which Hempel and Kitcher were hesitant

to admit. However, I maintain that many common sense explanations merely provide a good

reason to think that the explanans is the right explanation. Being out of milk does not necessitate

being out of the house, and therefore Jones’ being out of milk fails to scientifically explain why

he is not at home.

121

I think that it would be difficult to have a successful account of scientific explanation that

allows the explanans to be true and the explanandum to be false. This is essentially the ambiguity

of I-S and echoes concerns about ambiguous explananda (4.2.1; 1.3.1). To see why I think that

inductive-statistical derivations are not explanations, let us take for example that we want to

know why a particular smoker develops cancer. A reasonable explanation is that she was a

smoker and smoking causes cancer. We can formulate this in a deductive argument:

1. Mary is a smoker.

2. Smoking causes cancer.

3. Therefore, Mary develops cancer.

However, this explanation may fail to hold for Bob, who is a smoker, but does not have cancer:

1. Bob is a smoker.

2. Smoking causes cancer.

3. Therefore, Bob develops cancer.

Here, the premises are true, as in the previous explanation, but the explanandum is false. A

reason for this is that the premises are not informative enough. It is not clear what it means to be

a smoker or how long or how much Bob and Mary smoked. Further, the second premise could

more precisely be stated as:

2*. Smoking increases the risk of developing cancer.

However, the argument, when substituted with this premise, now fails to hold deductively. The

explanandum that deductively follows is:

3*. Bob/Mary has an increased risk of developing cancer.

This explanandum now fails to establish the difference between Bob and Mary, both have an

increased risk. But this is quite reasonable, as there is no information provided to distinguish

between the two. There is no explanation of both cases. Based on this criterion, if one wants to

know why Mary develops cancer and Bob does not, one would need more information. Just as

Kitcher argued, I too claim that the single occurrence of this statistical relation cannot be

explained given the available information. My contention is that this probabilistic causal story is

not scientifically explanatory.

122

While I believe this criterion can circumvent problems associated with causal and

inductive explanations, D1 alone will not guarantee explanation. Hempel proposed that only

certain kinds of regularities feature in explanatory derivations. Rather than follow suit with a

covering law account and perhaps attempt to specify lawlikeness, I will turn to a model-based

approach. Let me propose the second criterion:

D2. The statements in the explanans are true of a relevant scientific model, M.

This criterion refocuses the explanans to truth about models, rather than truth about the world. So

rather than requiring true laws of nature, this account requires that the generalization in the

explanans be true of a model. The truth of the explanans is considered to be necessary on many

accounts, but some, such as van Fraassen, have argued otherwise (van Fraassen, 1980).

There are some advantages to not requiring accurate representations. First, unlike other

deductivist accounts, lawlikeness is no longer an issue. As we saw in Chapter 4, one of the

criticisms levelled against the D-N account is that it relies on lawlike generalizations in the

explanans, but tells us little about how we can decide if a generalization is lawlike. Some who

have examined laws on causal grounds, focus instead on a spectrum of generalizations ranging

from accidental truths (statements about the coins in Goodman’s pockets) to truly universal and

exceptionless laws of nature (law of the conservation of mass-energy) (Mitchell, 2000, 2002a;

Woodward & Hitchcock, 2003b). Mitchell frames this in terms of realigning our concept of laws

with the practice of biology, while Woodward focuses on the degree of invariance at the heart of

explanatory generalizations. Both move away from the traditional concept of law and this

account also does, albeit in a different manner. This account requires that the generalization in

the explanans be true of a certain kind of model, and thus avoids having to distinguish between

laws and accidental generalizations, which has been problematized (Scriven, 1962; Goodman,

1973; Cartwright, 1994, 1997; Lange, 2002, 2004; Craver, 2006).

Second, idealization is no longer an obstacle to explanation. In fact, this allows for such

things as idealizations and abstractions to be central components of explanation. An issue for

many accounts of explanation is the reconciliation of idealized models with truth about the

world. As we saw in Chapter 2, when one recognizes that all models are to some degree

idealized, the requirement that the statements in explanans be strictly true cannot be satisfied.

Others claim that approximate representation is enough to justify a model’s application and

123

satisfy any weak commitments to realism, but even that is not a factor here. The statements in the

explanans are true of the model, but no explicit limitations are placed on the representative

accuracy of the features of the model. Not requiring that explanatory models accurately represent

also expands of the scope of explanatory idealizations, by opening it up to non-Galilean, or

highly-idealized, models. As we saw in Chapters 1 and 2, there are persuasive reasons to think

that highly-idealized models can support explanations, but this is very difficult to reconcile with

representational accounts of explanation. With D2, non-representing models are not necessarily

debarred from supporting explanations.

It is important to mention again that, while the explanans need not accurately represent

the target system and need only be true of a scientific model, the explanandum must be an

approximately true statement about the target system. Unless the explanandum concerns the

behaviour of the model itself, in which case, it must be an approximately true statement about the

model. The explanandum is a description of the phenomenon to be explained and it cannot be

false. It is important that this indirectly restricts the explanans. This is because if the explanans is

not capable of deductively leading to a statement that accurately describes the explanandum

phenomenon, then there is no explanation.

Lastly, it employs no particular account of causation and makes no commitment of causal

realism. This avoids any problems or issues that may come with the metaphysics of emergent

causation. Many accounts hold that explanatory relations are those that describe or capture the

real causal relations in the world. This is a useful way to debar some non-explanatory relations,

like backtracking counterfactuals, but has implications for the accuracy of the statements in the

explanans and consequently limits the scope of explanatory models. Just as Hempel’s account

was designed to capture explanations that are causal as well as those that are not, this account

aims to be more broadly applicable than one with causal criteria.

What is a Model?

One of the most important developments in the literature on explanation since Hempel, is the

recognition that explanations deal with models. This change has come from the desire to reflect

the explanatory practices of science, which largely proceed from pragmatically-oriented models

rather than laws of nature.

124

Many other accounts of explanation have focused on models, but it is either assumed or

argued that the models ought to be causal models. Some have downplayed the role of causation

itself and focused instead on mechanisms (Machamer, Darden, & Craver, 2000; Cat, 2005;

Darden, 2006; Weber, 2006; Craver, 2007, 2009, 2010; D. M. Kaplan & Craver, 2011; Knuuttila

& Loettgers, 2012; Schindler, 2013), or on structural models (Worrall, 1989; French &

Ladyman, 2003; McArthur, 2003; French & Saatsi, 2006; Bokulich, 2008). However, this

account will focus on models in a deductivist framework.

As was mentioned in 2.2, my approach treats models as objects. Some are physical

constructions, like a road map, a model ship, or a pendulum that sits on a desk, but the kinds

referenced in scientific explanations are abstract objects. This kind of model is not simply a set

of statements, but can be described by statements. Some of these statements feature idealizations

that are false of the world. Such as: “the plane is frictionless,” “there is no damping due to air

resistance,” “there is an infinite population,” and so on. These can be true of the model but false

of the target system. The model is that about which these idealizations are true.

A target system does not have a single corresponding model. Models are constructed and

constructed with a purpose. Usually this purpose is to bring to light some particular relation, the

effect of one variable change on another, for instance, between string length and period of

oscillation of a pendulum. The model will be designed to bring to light this relation, but not

necessarily any other (Weisberg, 2007). In constructing a model, the modeller must make trade-

offs. Some models are made to be as accurate as possible to all the actual components and

processes, to function as simulations; others to be maximally tractable, or have high predictive

accuracy within a limited range; others to be very general and widely applicable to different

systems to show why different materials exhibit the same behaviour, and so on. Which desiderata

are considered most important and how a balance is struck, will change from case to case.

Modelling desiderata are inevitably contextual. But in all cases, there is no single perfect model.

5.3.1. A Simple Model of the Fixed-length Pendulum

In order to clarify exactly what kinds of models are at issue, let us examine a simple example. If

we want to build a model to explain the period of the oscillation of a simple fixed-length

pendulum, we might construct it like this:

125

1. There is a bob suspended from the end of a string fixed to a pivot.

2. The bob is a point mass that moves in two dimensions.

3. The length of the string is L.

4. The string is massless, fixed in length, and taut.

5. There is no loss of energy via air resistance or friction.

6. The only forces on the bob are the tensile strength of the string and

gravity.

7. The acceleration due to gravity is fixed at g = 9.8m/s2

8. The period of the oscillation of the pendulum, T, depends on L and g, such

that 𝑇 = 2𝜋√𝐿/𝑔.

9. The oscillation amplitude is small enough that sin 𝜃 ≈ 𝜃.

This is an ideal pendulum that accurately describes no real-world pendulum. Its motion is

actually given by this model. The motion of this pendulum can be related to a physical one in

approximate conditions, but that is not important at the moment. We are speaking of this abstract

object and these statements are true of this particular model. The model is very simple. Note that

there are no equations for angular acceleration, no Lagrangian, and no differential equations.

This is because this model is constructed to capture the dependency of the period of a

pendulum’s oscillation on its length and not, for example, to represent its movement.

If we want to explain a particular oscillation period phenomenon we can formulate a

deductive derivation of this, with explanans statements that are true of this model. For instance,

we might ask “Why is the period of oscillation of this simple pendulum 1s?” The obvious answer

is because the length of the string is a particular length, viz. 25cm. An outline of the explanatory

derivation might look something like this:

1. 𝑇 = 2𝜋√𝐿/𝑔

2. 𝐿 = 0.25𝑚

3. 𝑇 = 1𝑠

This is not only a derivation of a particular phenomenon. The model also gives us counterfactual

information about changes to the system, such as what the period would be if the length were

1m, or 10cm, or whatever. What this model tells you is that changing the length of the string will

126

change the period, and it tells you by how much. This counterfactual information is essential to

an explanation-supporting model.

Counterfactuals

The D-N account makes use of causal laws as regularities, which can be exploited because the

derivation is allowed to run both ways. It is important to note that regularities alone do not

express counterfactual information. Without this information there is nothing to prevent

backwards-running non-explanations. And so, some restrictions outside of the derivation itself

are required for explanation. Counterfactual information about the dependency relations of the

model is relevant to explanation, something argued by Woodward and Hitchcock and others

(Lewis, 1979; Tooley, 2003). My proposal is that this counterfactual information is available in

the details of the model. Woodward and Hitchcock argue that an explanatory generalization “not

only shows that the explanandum was to be expected, given the initial conditions that actually

obtained, but it can also be used to show how explanandum would change if the initial and

boundary conditions were to change in various ways,” (2003a, p. 4).

The model referenced in the explanans must provide information about how the values of

the explanandum variable would change if there were changes to the values of the explanans.

Formally, the criterion can be stated as follows:

D3. The model M referenced by the explanans gives counterfactual information that

shows on what the explanandum depends.

This criterion distinguishes a model-based account from a covering law in an important way. The

regularities of the D-N account were able to be exploited precisely because there was no

requirement on counterfactual information.

This requirement also ensures that the explanandum phenomenon can actually change.

Consider again a case introduced by Salmon (Salmon, 1971). Jones can avoid becoming

pregnant by regularly taking birth control pills, and every man who regularly takes birth control

pills avoids pregnancy. The problem with an derivation such as this is that the generalization is

irrelevant to the reason why the explanandum is the case. However, such cases are debarred by

this criterion, since that kind of regularity gives incorrect counterfactual information about what

the explanandum depends on. In effect, this criterion is able to serve as a requirement for

127

testability. If it is not possible for the explanandum to change (Jones becoming pregnant), then

the criterion is not met.

One could ask why it is that the ability to answer counterfactuals is so crucial for

explanation. Woodward ties the importance to knowledge of changes made by manipulations. On

this account, the importance comes in confidence in the reliability of explanatory models.

Models that are capable of supporting explanations are able to provide accurate, reliable

information about changes to the target system. Because the explanandum is approximately true

of the target system, and the counterfactual information in the model demonstrates on what the

explanandum depends, then the counterfactual information in the model has implications for its

application to the target system.

5.4.1. Same-object Counterfactuals

The account I propose borrows the notion of a same-object counterfactual from Woodward and

Hitchcock. This kind of counterfactuals is useful, in that it can provide a means to debar

backtracking counterfactuals. Requiring this kind of counterfactual information precludes the

possibility of exploiting the symmetry of a deductive derivation.

To reiterate from 3.3.1, Woodward and Hitchcock state that traditionally counterfactuals

are seen to be what they call other-object counterfactuals, which give information about what

would be the case for a different object to have different values for its variables. Same-object

counterfactuals refer to hypothetical changes on the same object. To illustrate the difference,

consider the following example of Galileo’s pendulum law, which explains why a pendulum of

length x has period y. According to the law, the following counterfactual is true: if this laptop

were a pendulum with length x, its period would be y. The law is sufficient to support such

other-object counterfactuals that involve changes in identity. This counterfactual however, tells

us nothing about how changes in the values of x would affect the values of y. If this

counterfactual concerned the very same object’s hypothetical values it might be framed as

follows: if a pendulum with length x and period y had its length adjusted to x', then its period

would be y'. This counterfactual gives information about what would happen to other features of

the model given hypothetical changes to the very same object. It is this information about the

possible changes to a model that satisfies the conditions of D3. It is a generalization’s ability to

support same-object counterfactuals that is relevant to its being explanatory.

128

Of course, more than one model can support same-object counterfactuals for a single

target system. It is useful, but not necessary, for my account to provide or at least be compatible

with a method for comparing relative explanatory depth. Some philosophers want more than a

threshold in an account of explanation in order to match up with our intuitions about which

models are more explanatory than others. There is a further way that one can rank, or compare

different models that have already qualified as explanatory by meeting minimum requirements.

This could be done by the depth of the counterfactual relation, but I will not explore this

possibility here outside of what has already been said about w-questions and depth in Chapter 2.

However, it seems likely to me that there are many measures of explanatory value, not all

of which will be compatible. It is very reasonable to assume that certain pragmatic issues, such

as levels of abstraction and the context of knowledge and communication skills, that can have a

great deal to do the with the explanatory value of an explanation-act. And furthermore, these

likely will not all pull in the same direction. In different circumstances, some explanations can be

of more value if they are simpler and others if they are more detailed. A notion of explanatory

value then is unlikely to be a single unanimous measure.

Before looking in more depth, so to speak, at the pragmatics of explanation with respect

to models and model construction, it is worth saying a few words about the truth conditions of

counterfactuals and what exactly is supporting the counterfactuals that underwrite these model-

explanations.

5.4.2. Truth Conditions for Counterfactuals

Counterfactuals are different from conditionals in an important way. If one wants to know

whether a conditional is true, one can simply test for it. However, because counterfactuals are by

definition contrary to fact, there is no simple truth-functional way of determining whether they

are true or not. And yet, we have strong intuitions about the truth of some counterfactuals, and

use them in everyday reasoning.

A central tradition in the literature on counterfactuals stems from Robert Stalnaker’s and

David Lewis’ accounts of counterfactuals. This approach make use of possible world semantics

and work done by Saul Kripke (Stalnaker, 1968; Kripke, 1972; Lewis, 1973). Very briefly, this

approach is to look at the closeness of possible worlds to determine which counterfactuals are

true. By contrast, Woodward and Hitchcock look to the metaphysics of causation. According to

this view, there are brute facts about the causal relations in the world that can be used to

129

determine the truth value of counterfactuals. We gain knowledge about these causal facts by

discovering relations that are invariant over a range of interventions.

A model is an abstract object. And in a similar vein, there are brute facts about a model.

There are objective facts about relevant counterfactual circumstances in a model. Statements

about a model can either be true or false or indeterminate. An explanatory model should be

described adequately enough to support same-object counterfactuals. What changes can be made

and what boundary conditions and parameters there are in place may need to be spelled out. The

truth of counterfactuals come from brute facts about this abstract object.

Because models are constructed purposefully and involve only certain aspects or relations

of a system, there are only certain relevant w-questions that can be answered. It is obvious that in

a model of sunspot activity there is no information about what would happen given changes in

the population of lemmings in Norway. In other cases, it is less obvious when there is no

information. In a simple model focused on the changes in barometer readings as storms

approach, there is no information about what would happen storm-wise if one fiddles with the

barometer. While a regularity alone might support this counterfactual, it does not support the

right kinds of counterfactuals about changes to the objects in the model.

The Simple Pendulum Revisited

Now that we have an understanding of same-object counterfactuals and the work they will be

doing in this account, let us return to the model of the simple pendulum. Recall that we were able

to explain what would happen to the period of a pendulum if its length were changed. Now let us

consider the symmetrical case, and say a particular string length is our explanandum

phenomenon. We might ask “Why is the length of the string in this fixed-length pendulum

25cm?” Here, the model has no obvious answer. There are likely many reasonable pragmatic and

circumstantial reasons for the exact length of string, but none that come from the model we

outlined above. There is nothing in this particular model to say what the length of the string

depends on. It might be possible to use the equation provided in the model to derive the length

given a period, but this is not enough for explanation. This does not show counterfactual

information about what the explanandum actually depends on.

What we have is a statement about the dependency relation of the period on the length.

In the model we have built above, there is no information about what would happen to the string

130

length if we changed the period. Further, changing the period would involve adding a driving

force, pushing on the bob, changing the acceleration due to gravity, or something else which is

outside the parameters of the model. The model is not equipped to handle this kind of

counterfactual.

Of course, one can build a very general model in which the explanation runs both ways.

And one can also build a model in which this explanandum phenomenon seems perfectly

reasonable. But it will likely look different than this model. Grandfather clocks, like other

mechanical time-keeping devices, need to be calibrated. You can adjust the time that the clock

keeps by adjusting a small screw at the bottom of the bob, which raises or lowers it along the

rod. The height of the bob along the rod effectively changes the length of the rod, since the

portion of the rod that dangles below is massless in our model. If you want to build a model to

figure out what length of string (to keep the language consistent) you should have to keep proper

time with the right period, you might describe the model this way:

1. There is a bob suspended from the end of a string fixed to a pivot.

2. The bob is a point mass that moves in two dimensions.

3. There is no loss of energy via air resistance or friction.

4. The only forces on the bob are the tensile strength of the string and

gravity.

5. The acceleration due to gravity is fixed at g = 9.8m/s2.

6. The period of the oscillation of the pendulum is T.

7. The string is massless and its length is variable.

8. The length of the string L depends on T and g, such that 𝐿 = 𝑇2𝑔/𝜋2.

9. The oscillation amplitude is small enough that sin 𝜃 ≈ 𝜃.

The key difference in this model is that the dependency relation is spelled out in terms of

changing the length to fit a desired time. Now there is an explanation, an answer to our why

question. The length of the string is 25cm, because the desired period of oscillation is 1s. The

length of the string depends on what one desires as a value for T. We know what the length

depends on, T, and we know how, quantitatively, the dependency works, viz. by the

generalization 𝐿 = 𝑇2𝑔/4𝜋2. This is more than a simply changing the equation around: it

reflects the dependency relations stipulated by the model. What this model tells you, is that

131

changing desired period will the change the length of the string, and by how much. Here,

whatever the length of the string is, it depends on what the period of its oscillation is. This

explanation no longer seems backwards.

What this shows is that there are brute facts about the model and about hypothetical

changes to the features of the model that make it different from a mere regularity that might

allow D-N-style counterexample derivations. The counterfactuals that would support

counterexample cases are not stipulated in the model. This is one of the main benefits of taking a

model-based, rather than covering law, approach to deductivism. However, it seems reasonable

to worry that if one can simply construct any model with certain counterfactuals but not others,

that the victory is rather hollow. In order to prevent this from seeming so arbitrary, I will next

introduce a global constraint on which same-object-counterfactual-supporting models can be

explanatory.

A Global Constraint on Explanation

Some models are phenomenological; they are merely representations, or systematized

collections, of data, like in the statistical modelling of regression analysis. These can be used to

examine the relations between sets of variables, make predictions, and can have considerable

heuristic value, but I maintain that they are not explanatory. This is a consequence of the

account, but also reflects a disagreement with Hempel about the similarity of prediction and

explanation and the adequacy of nomic expectability. One of the aims of the account is to be

sensitive to the difference between showing that we know something to be the case and

explaining why it is the case.

An obvious problem arises with the criterion that we have established so far. It is too easy

to just build dependency relations into a model and use it to support only the explanations you

want. Without further restrictions, a model can be built to support all manner of counterfactuals,

and of course, not all of these are explanatory. In order to maintain a high threshold for

explanation, there must be further restrictions on the kinds of models that support explanation.

We are now in a place to put in place the last component of this account of explanation. There

are a few avenues one might take to impose the right restrictions on explanatory models in order

to reflect this distinction between models that predict and models that also explain.

132

One could require that the model be approximately accurate to the real world system,

either in terms of causal processes, causal mechanisms, or structure. These strategies all require

accurate representation and allow for idealizations only when they are not harmful, i.e., when

they are Galilean (2.2.1). It is not impossible to reconcile representational approaches with

explanatory autonomy, but it has been agued that they favour reductive explanations given in

terms of fundamental theory, since they are the most accurate and invariant and robust and stable

and so on – they give the most accurate representations (Weslake, 2010). These are the kinds of

account that Batterman and Rice refer to as common features accounts (Batterman & Rice,

2014). Bokulich cogently argues that idealizations need to play a central role in explanation, and

not merely be tolerated when mostly harmless (approximately representative). Something

Bokulich makes very clear is that a representational approach loses out on a way to capture how

highly-idealized models explain. Her requirement for which idealizations, or fictions, can be

explanatory ultimately relies on a kind of representational view of structural realism: explanatory

models must be isomorphic to the systems they describe. Following Belot and others, I maintain

that a stronger criterion is needed.

Rather than imposing an additional local constraint on explanatory models, this account

introduces a global constraint. This means that instead of relying on structural or causal realism

in order to debar such models as being counted as explanatory, this theoretical approach requires

that explanatory models be integrated with a scientific theory. There needs to be information, a

set of assumptions, other models, or laws, justifying the idealizations in order to show why the

relations the model describes hold; why the system exhibits this behaviour. This information

explains why the model works when it does and why it fails when it does, which is vital to

understanding why the behaviour occurs the way it does. This also does the work of Alisa

Bokulich’s E3 criterion of specifying the domain of applicability (2.4.3), but also gives a

stronger requirement that helps to only capture explanatory models. This is the global

requirement of theoretical integration.

D4. M must be integrated with an independently explanatory scientific theory, T.

It is not enough that the model can faithfully reproduce the explanandum and give counterfactual

information, the model must also be appropriately related to a global scientific theory, such as

General Relativity, Quantum Mechanics, or the Theory of Evolution. The theory must have large

133

empirical scope, which serves to unify otherwise disparate phenomena. The theory must also be

considered true, or mostly true, and explanatory by a relevant scientific community.

Instead of restricting the model on representational grounds, I propose that explanatory

models are a part of, or able to be integrated with, a successful and well-established scientific

theory. A motivation for this approach is the idea that an explanation is that which joins the

derivation of a new phenomenon to that which we already understand; it simplifies, organizes,

and relates phenomena and regularities, and so contributes to our understanding. An explanation

presents a new or unexplained phenomenon in a similar manner to already understood

phenomena, or in the context of an established body of scientific theory. Connecting a model to a

theory by performing an integration can broaden and deepen our understanding of an already

partially understood field.

Of course, not all theories are explanatory and not all theories are scientific. The theories

that give explanatory power to models are global scientific theories that feature models, make

use of shared assumptions, contain laws (or law-like statements), and are able to account for and

accurately predict a wide range of phenomena. They are global theories in that they are well-

established in science, broad in empirical scope, and widely believed by a relevant scientific

community to be explanatory. Prime examples are theories such as general relativity, Newtonian

mechanics, cellular biology, the kinetic theory of gases, the theory of evolution, and many

others.

A theoretical, rather than representational, approach allows for idealizations in a way that

a representational approach cannot. A representational approach is limited by requiring that

explanatory models feature only real entities and approximately real dependency relations. This

precludes highly-idealized models outright. Many representational models are of course

explanatory, but they are not explanatory because they accurately represent. It is rather because

they satisfy the criteria of this account, including because they are theoretically integrated, even

if this relation is not explicit. I contend that systems are modeled by theories and successful

explanatory models that seem to accurately represent causal relations are smuggling the

justifications and idealizations from some (likely Newtonian) theory.

As an example, Andrew Wayne (2015), echoing Hempel, argues that Galileo’s ideal

pendulum model is not explanatory in itself. Rather, it seems that it was mostly important

because it was a particular instance of Galileo’s general laws of motion. Historically, Galileo’s

134

pendulum model served to base his studies of free-fall and is essentially a phenomenological

model. It gives some counterfactual information, and expresses a regularity. Wayne argues that

this suggests that its actual explanatory power comes from its integration into Newton’s general

mechanics, which could provide explanations of forces and the motions of bodies. The model’s

ability to satisfy Woodward’s requirements is not sufficient to guarantee its being explanatory,

regardless of whether it accurately maps onto the target system and answers counterfactuals.

Rather, it is the integration with a global theory with independent explanatory power that makes

the local model explanatory.

5.6.1. Some Aspects of Theoretical Integration

The relation between a local model and global theory can be very straightforward. In some cases,

the model features a generalization that is a central component of the theory and is capable of

deriving the explanandum phenomenon. This is seen in cases of textbook explanations, in which

the generalization or law is applied to a system and dictates its behaviour. The simplest case of

integration then is where the equations of the global theory are directly applicable. In these cases,

the systems themselves are probably rather idealized. Let us look briefly at a simple example.

Imagine we want to explain the rise in temperature when pressure is increased in a gas.

We might build a model containing the ideal gas law. One can use the formula to approximate

the behaviour of gases under certain conditions. The formula works most accurately at high

temperatures, low pressures, and for monatomic as opposed to molecular gases. The formula can

describe how an ideal gas behaves under certain changes. But it is only when it is seen in the

context of the kinetic theory of gases or statistical mechanics that it can be explanatory. This

comes forward in the idealizations included in the model. The theory is able to provide

justification for the idealizations that the particles are point masses, have elastic collisions, and

so on. The model is built according to the theory. The model’s idealizations show why the

regularity works best with monatomic gases at low pressure, for instance: the assumption that the

particles are point masses begins to be problematic when the molecular size becomes significant

relative to intermolecular distances. One can constrain the range of phenomena based on the

parameters included in the model. The theoretical background is what allows the formula to

provide explanatory information on the behaviour of the gas; information on why it behaves this

way when it does. The model featuring this generalization can be explanatory, but it is not in

virtue of the fact that it accurately represents or because it supervenes on the real fundamental-

135

level causes. It is because the system can be modeled according to a theory. Again, a regularity

on its own is not explanatory.

However, in many cases, when the explanandum or the target system demands more

accurate models, the laws or fundamental equations of the scientific theory will no longer serve

to derive the explanandum phenomenon. In yet other cases, the theory’s equations may not be

directly solvable. In such cases, the integration becomes rather complex. I take it as a reasonable

assumption that the integration relation is quite different in various scientific disciplines. There is

a range of integration, from the straightforward pedagogical cases, like that shown above, to very

complex relations. In cases where the integration is complex, a modeller may need to employ

various mathematical methods for constructing a local model that is related to and consistent

with the global theory. What this means precisely will depend on the model and the theory. As

such, I cannot give a complete formal account of integration, but I can elaborate on some of the

methods and procedures that might be employed. One can think of integration as a process that

may feature one, a few, or several steps. Some of these steps may involve single perturbation

methods, renormalization groups, other mathematical techniques, and otherwise theoretically

justifying idealizations.

Wayne has argued that one of the factors that can differentiate a successful integration

from a failed one is that it makes certain idealizing assumptions unproblematic for the

underpinning theory (2016). He notes that in two competing putative explanations of the

phenomenon of gravitational waves, there is the same point-particle idealization, which is false

of the fundamental theory of general relativity. However, the post-Newtonian model has recently

been able to accommodate extended bodies, and thus partially discharge the assumption. The

competing derivation cannot discharge this assumption. In this case it is this theoretically-based

justification for the idealization that marks a distinction between an explanatory model and one

that is not.

A full exposition of all the various kinds of complex integration that might allow a model

to support explanations in different disciplines and subdisciplines is far beyond the scope of what

I can to accomplish here. However, since it is easy enough to see what kinds of models are

simply or deductively integrated with theory, it would be fruitful to see what kinds of models fail

to be integrated and thus fail to support explanations.

136

5.6.2. Prediction without Explanation

Part of the support for the claim that integration is essential to explanation is that cases of non-

explanation lack theoretical integration. The hope is that the account of explanation can show

that models appropriately connected with these theories are explanatory and that two kinds of

models, even ones that are predictively successful, are not explanatory: those that are not

appropriately connected with theory; and those that are connected with theories that are not

explanatory. Phenomenological models, forecasting models, and others are of the former; models

of phlogiston theory, Ptolemaic astronomy and others are of the latter.

It will be useful to consider again the semiclassical model of quantum wavefunction

scarring from 2.4. It was shown that because it was still possible to derive the scarring

phenomena with a semiclassical model that there are good grounds for it being explanatory. The

model was counterfactually robust, though not as much as the local quantum model, and it fit

Bokulich’s criteria for explanation. The real problem, I claim, is not only that there is a better

explanation, but that the models of semiclassical mechanics are not integrated with a theory that

has independent explanatory power.

If one wants to claim that semiclassical mechanics is explanatory there are two

possibilities for showing this. Either semiclassical mechanics is just such an explanatory global

theory, or it is a method of integrating semiclassical models with quantum mechanics. If one

argues the former, however, the only way semiclassical mechanics can be explanatory is if

semiclassical mechanics were a well-established global theory of science. The two related

theories mentioned by Bokulich, closed-orbit theory and periodic-orbit theory, are not widely

held to be true, and have rather limited empirical scope. They make false claims about the

contributions of orbits to the quantum spectrum and are really only applicable in certain special

cases of quantum chaos. They are methods for approximating quantum calculations, but do not

have any real theoretical components. It is important to debar models of non-explanatory

theories. To use a rather extreme and unlikely example, if a model is appropriately connected to

astrology or the phlogiston theory of combustion, it is not therefore explanatory. There is an

important role to play by the relevant scientific community in determining what is an acceptable

and explanatory theory.

Perhaps it is best understood as attempting to integrate with quantum mechanics. What

the periodic orbit theory specifies is a way to approximate the quantum wavefunction density by

137

taking the Fourier transform of the autocorrelation function of a Gaussian wavepacket. What it

essentially shows is that under certain conditions, its solutions are reliably approximate to the

full quantum calculations. The semiclassical approximations attempt to match the calculations of

quantum mechanics, and in this way show that the scarring phenomenon can be expected. This,

granted, is no small feat. However, there is no justification from the theory that justifies the

contrary to fact assumptions about the classical trajectories in the model and their effect on the

quantum spectrum. This is unlike the case examined by Wayne where the explanatory model can

discharge the assumptions that are false according to the theory. The false assumptions of the

semiclassical model cannot as yet be discharged. The quantum model on the other hand is

capable of showing that the phenomenon is to be expected without implementing assumptions

about classical trajectories and the effects of periodic orbits. It is not enough that there is reliable

prediction from the semiclassical model; justification needs to be given from the theory. It is

possible that there will one day be a way to refine semiclassical methods to discharge these

assumptions, but it would probably require more research about how the quantum to classical

transition takes place. This marks an important aspect of integration: it is relative to the state of

science.

While Hempel and Oppenheim argue that explanation entails prediction and vice versa, I

maintain that only the former holds. The reason for this follows from the idea that nomic

expectability is not sufficient for explanation. This can be seen by looking at cases where models

are predictively accurate, but not explanatory. Consider also the numerical models of circadian

rhythms. Circadian rhythms are biological processes that follow a 24-hour period. They are often

modeled as oscillators. It is possible to derive the right explanandum of say, a sleeping pattern,

with a mathematical model that fits the data. In fact, even some story can be developed about

exactly why this particular mathematical relation holds, but it is nonetheless non-explanatory.

Here there is no connection to a theory. It is a consequence of this criterion that such

phenomenological models are not explanatory.

The case of a Ptolemaic model of the solar system explaining planetary motion (2.4.2)

points to a different issue regarding the role of scientific theories in explanation. Systems are

modeled according to a theory, and it is certainly possible to model planetary motion in this way,

but this is not how a competent scientist today would actually proceed. It is important to have an

account that reflects how scientific explanation actually happens. However, if a scientist were

138

committed to the truth or approximate truth of the Ptolemaic system and constructed such a

model to derive the explanandum, the derivation would still not count as explanatory on this

account. The reason this will not work out is that the Ptolemaic theory does not have independent

explanatory power. It is only predictively adequate insofar as it can be used to approximately

derive a limited range of phenomena. For reasons already mentioned, the conditions that make a

theory acceptable and explanatory is not something I want to focus on, but it can be clearly seen

that this theory does not have the empirical accuracy, nor the unifying and systematic value that

Newtonian mechanics or general relativity has. Further, while the debate between Ptolemy and

Copernicus was famously underdetermined, by contemporary standards the theory is now known

to be completely false and as a result has a seriously limited range of true counterfactuals that it

can support.

Mitchell’s case study of Lake Erie (3.2.2) has a rather different problem: it has no

specific explanandum. What is being asked for here in explaining the “behaviour” of an

ecosystem is unclear. This is what leads Mitchell to conclude that no single model, and

especially none at the fundamental level, can capture the target system’s behaviour. Had the

explanandum been a more precisely formulated phenomenon, such as the effect of zebra mussels

on the aquatic plant life, then a specific model could possibly have been built that derived the

result and provided counterfactual information. In the case of Lake Erie, as provided by Mitchell,

there is little sense as to what would constitute a good explanation of the behaviour of the

ecosystem. What this seems to be aimed at is a detailed simulation of various interacting

elements, and that is precisely what Mitchell claims provides the best explanation. But how this

constitutes an explanation on her account is not clear. In light of what has been said of this

account, that does not provide an explanation, but is rather phenomenological. Explanations are

to be had in the target system for various well-articulated explanandum phenomena.

The idea that combining multiple models can be explanatory something favoured by

Michael Weisberg, who also claims that it can be compatible with causal realism (Weisberg,

2007). This strategy is popular in dealing with highly complex phenomena. Each of the

assumptions made in particular models will be different and often incompatible. Practical trade-

offs are made to favour certain desiderata: generality, precision, simplicity, or accuracy. Levins

says the following about the conflicting assumptions of such models: “These conflicts are

irreconcilable. Therefore, the alternative approaches even of contending schools are part of a

139

larger mixed strategy. But the conflict is about method, not nature, for the individual models,

while they are essential for understanding reality, should not be confused with that reality itself”

(Levins, 1968, p. 431). Wimsatt thinks that such highly-idealized models help us build truer

theories (Wimsatt, 1987). Some of these multiple-model idealizations, however are employed for

the sole purpose of predicting, such as the practice of weather forecasting. Here the accuracy or

realism of the idealizations is hardly a concern compared to their predictive abilities. If one looks

at the literature on complex models there is little in the way of a search for a single model, rather

what one finds is that there are many models with different desiderata. Levins and others are

realists with respect to these multiple model idealization, since they lead us to true theories “Our

truth is at the intersection of independent lies” (Levins, 1966, p. 423).

I maintain that the multiple-models solution is not really an explanation, but it does

reinforce something I have said about potentially explanatory models, viz. there is more than one

for any system. It is not reasonable to expect that there is only one explanatory model of a real-

world system. No model is perfect; models are constructed for certain purposes. With a clearly

defined and individuated explanandum phenomenon a good explanation featuring a particular

model can be explanatory (provided it meets certain criteria). The benefit of this account is that it

makes no demands that the relations described in the explanatory model be real causes. As such,

it is possible to consider multiple models to be explanatory, while avoiding commitments to

causal realism. In the end, Mitchell is right about the importance and explanatory value of

multiple models, but the assumption that these models are causal is unwarranted and potentially

problematic.

The Integrated Model Account of Explanation

Let us now assemble the criteria that we have established. We can take the criteria that have been

gathered to be sufficient to constitute an explanation, but not necessary for an explanation. As

mentioned as early as Chapter 1, there are many kinds of explanation and many senses of the

word and there is little hope that one account of explanation can point to some criteria that are

necessary for any explanation whatsoever. Rather, what I argue is that if the following criteria

are satisfied by a derivation, then it constitutes a scientific explanation.

D1. The explanandum must be a deductive consequence of the explanans.

D2. The statements in the explanans are true of a relevant scientific model, M.

140

Representation

n

Deduction Idealization

D3. The model M referenced by the explanans gives counterfactual information that

shows on what the explanandum depends.

D4. M must be integrated with an independently explanatory scientific theory, T.

These criteria taken together are sufficient to capture good scientific explanations, though of

course, there are many other explanations and kinds of explanations which do not meet these

criteria. What we have in this account is a model-based deductivist explanation, whose models

are constrained by their integration with an accepted theory of science.

There are two aspects to this kind of explanation: local and global. The local aspect of

explanation concerns the local idealized model, the statements in the explanans that are true of

the local model, and the explanandum statement. As mentioned, sometimes the explanandum

phenomenon concerns the behaviour of the model itself. This is seen in cases of pedagogical

examples where only the behaviour of an ideal and abstract object is being accounted for. Where

there is a real-world target system, the explanandum is true, or approximately true of that system.

The following chart shows the relations that hold between the aspects of explanation.

Figure 5.1. A caricatured mapping of the relations between an explanation, the world, models, and a global

scientific theory.

Gobal Scientific Theory

Local Model Explanans

Target System Explanandum

Data Model

Description

Description

Integration

141

Here we see the global scientific theory with systematic explanatory power at the top. Below this

is the local idealized model, which is related to it via the large arrow of integration. This

indicates that the model was constructed on the basis of the theory, or in a way that its

idealizations are theoretically acceptable. The construction of the model will depend on the

explanandum, and on various pragmatic concerns of the modeller, as well as on available data

from the target system. By contrast, we see that a data model is not connected via integration to

theory and primarily built from data. Realistically, the data model is not constructed from mere

raw data points, and is also refined over time by testing predictions and being modified. On the

right we see an explanation, separated into its two parts: the explanans and the explanandum.

The statements in the explanans are descriptions of the local model, but not necessarily of the

target system. This is a key difference in a non-representational approach. The statements in the

explanans deductively entail the explanandum statement, which itself is true, or approximately

true of the target system. Recall that, under this framework, saying that a model is explanatory is

to say that the model is capable of supporting explanatory arguments. I believe this account and

this simplified diagram reflect the way a range of scientific explanations proceed. An explanation

uses a model constructed according to a global theory to derive an explanandum phenomenon.

The explanatory power of a theory T is relevant to whether a model of that theory MT explains

its target phenomenon.

It is a reasonable question at this point to ask what it is that makes a global theory

explanatory. Unfortunately, the criteria for a good scientific theory can probably not be stated in

exact terms. What precisely constitutes a global explanatory theory is, I think, impossible to fully

articulate, because of the contextual, domain-specific, and paradigmatic aspects of scientific

theories. What counts as a successful theory changes over time and with respect to different

scientific communities, subdisciplines, and schools of thought. However, one characteristic of an

explanatory theory has been suggested above, and that is that it systematizes knowledge. This is

something that deductivists have always maintained, from Hempel to Feigl to Kitcher, and a

great many others besides. Kitcher attempted to elaborate on and formalize this unofficial

account of logical empiricism to explain how global theories are explanatory. Wayne has argued

that this is an aim that cannot be, and does not need to be, answered on a deductivist account

(Wayne, 2016). Following Hempel, and Kuhn to a degree, I think that it must be sufficient to

142

point to examples of theories that are known to be explanatory, and examples of theories that are

known to not be (more on this in 5.6.2).

In terms of explanation, what counts as a theory with independent explanatory power is

determined not by strict criteria but by shared practices and standards of a scientific community.

It is only reasonable to allow that standards vary between communities, disciplines, fields, and at

different times. However, this is not mere descriptivism or relativism with respect to explanation,

as described in 1.4.1. There can still be normativity to claims about explanation. Rather, only the

kind of independent, syntactic justification that Kitcher was seeking is given up. What there is is

a network of mutually enforcing justifications that arguably give it more normative power than

the syntactic assessment of argument patterns on the unificationist view. The contextual

justification of explanatory judgments depends on known evidence, accepted standards, the goals

of investigation, and more.

One might be concerned that this is no different than Bokulich’s third criterion. Like

Bokulich, I also place some importance on the explanatory judgments and practices of a relevant

community of scientists, and I think to an extent, this is unavoidable. The role that this

contextual element plays in my account is quite different. Bokulich sees the structural criterion

as incapable of determining whether a model is explanatory. The role of distinguishing

explanatory from non-explanatory models is something she places solely on E3. I argued that

this makes the account descriptive, and rather uninformative. What I have offered is a high

threshold for explanation which can be assessed independently of the assessments of the

scientific community. That what counts as an global explanatory theory is contextually

determined reflects a fact. It also allows the account to appeal to a wider range of disciplines,

because it allows that the standards are, to a large degree, relative.

Empiricism, Emergence, and Reduction

This account, as it is not specifically causal, has the added benefit of being compatible with anti-

realism with respect to high-level causes and science itself. This account allows that properties

and entities in a model can be genuinely explanatory whether or not they accurately represent

features in the target system. The account I propose remains silent on the causal nature of the

relations of models that support same-object counterfactuals. The truth-makers for the

counterfactuals stem from brute facts about the model, but no further commitments are made.

143

The account is still open to accepting as explanatory models that are widely considered to

accurately capture the causal (and/or structural) dependency relations in the world, as long as

they also satisfy its criteria. Importantly, on this account, an empiricist can say something about

explanation beyond issues of pragmatics.

Many accounts of explanation have focused on making high-level explanations palatable

to physicalist intuitions, and this often involves either showing how they can be consistent, or

showing how they can be reduced. It is often said that scientific explanations effect a reduction

of the puzzling to the familiar, but Hempel does not think this stands up (Hempel, 1966, p. 83). It

is not merely concerned with the psychological aspects of understanding that may help one grasp

what is going on. In fact, the at-homeness of explanation can lead to very unscientific theories

(anthropomorphic nature, e.g.). Science should not hesitate to do the opposite and go against

intuition if it is necessary. It is not the aim of scientific explanation to explain away the

everyday. Rather, it is to account for it. The kinetic theory of gases does not say that there are not

swarms of gases that change volumes; and atomic theory does not show that there are no tables

and chairs. They are not explained away, but remain objects and entities that can be legitimately

be used in scientific explanations.

The account I propose, along with many others, does not favour reductionism to a

problematic degree. Some have focused on levels of selection, others have made abstraction or

proportionality and explicit requirement (Sherman, 1988; Yablo, 1992; Strevens, 2008; List &

Menzies, 2009). Some even object to any preference for reductive explanations (Mitchell, 2003;

Grantham, 2004; Brigandt, 2011). However, I suggest that having a criterion of derivation (D1)

can also help to prevent problematic reductionism, of which many accuse those like Bickle

(Bickle, 1998, 2003; Churchland, 2004). If one allows explanations featuring highly-idealized

models, then there is little reason to focus on proportionality as a distinct criterion, as some do.

Consider a case where one wants to know why Joe sold his house. It is reasonable to think that a

high-level explanation is perfectly adequate. In fact, an explanation from fundamental physics is

not likely to ever be formulated. Recall that if one cannot derive the explanandum from the

explanans, then there is no explanation. In this sense, a deductive criterion can actually help

prevent harmful reductionism. Successful explanation of a macro-level phenomenon requires

either that the explanans contains macro-level terms, or has micro-level terms as well as

144

information about the relationship between the micro- and macro-level terms. If it does not, then

it simply does not derive the explanandum and does not answer the why-question.

I think that this can prevent the account from being harmfully fundamentalist. I think that

this account and its approach to high-level explanation can be consistent with a wide variety of

metaphysical views. Unlike certain approaches that require emergent causation, this account

remains open to non-reductive physicalism. It also does not require any particular theory of

emergence to account for non-fundamental entities, such as Humphreys (1997). High-level

models are explanatory when they meet the requirements of the account, regardless of whether

there is another explanation featuring a model from a more fundamental theory. For instance, it

is possible to maintain that the ideal gas law can be explanatory when the explanandum regards,

say temperature. The ideal gas law can be de-idealized, and so in a straightforward way, one

knows why it is accurate within certain ranges of the values of its variables. One might argue that

the statistical mechanical explanation is better, or deeper, but on this account, the high-level

model is explanatory if it meets the criteria.

Lastly, it bears mentioning that the account remains open to being supplemented by a

measure of explanatory depth. The incorporation of counterfactual information into the account

likely makes it amenable to an account of w-questions, but possibly also some independent

measure. However, it is important to note that even if this account exhibits a preference for

reductive explanations, a high-level model can be not only explanatory, but also the best

available explanation when there is a high-level explanandum and no way to derive the

explanandum from a more fundamental level, as was mentioned in the case of semiclassical

mechanics in 2.4.2.

Conclusion and Limitations of the Account

This account is intended to reflect the results of the investigations in previous chapters. Firstly, it

aims to reflect the model-based approach to explanation in practice. This allows it to avoid

relying on laws of nature and opens it up to more than covering law explanations. Secondly, it is

deductivist and non-representational. This allows it to avoid any problems arising from the

metaphysics of emergent causation, while remaining compatible with metaphysical realism about

causation as well as empiricism and non-reductive physicalism. This also expands the scope of

explanatory idealization beyond the merely causal. This opens up the range of explanatory

145

models to include high-level models that may or may not be causal, which better reflects

explanatory practices. However, the scope is not increased too broadly, because, thirdly, it

imposes both a local and global constraint to ensure that explanatory models are those that can

provide counterfactual information and are not merely predictively successful phenomenological

models. This allows one to set a high threshold for explanation, while still allowing the range of

possibly explanatory models to exceed that of other accounts.

The account is still largely conceptual at this stage. A lot of work remains to be done to

test it against scientific practice. No doubt this requires in-depth case studies. Future work will

attempt to explore how integration actually proceeds for different kinds of models. Case studies

will also show how the application of the account fares in various disciplines and with respect to

changes. As of now, it is an assumption that the restrictions on models are sufficient and that

they track our judgments about explanation reasonably well. It remains to be seen how the

criteria hold up against detailed case studies. I would like to acknowledge some possible

concerns and limitations to the account and to what was accomplished here.

There is a concern some might have that a theoretical approach is better suited to some

fields like physics, which have well established laws and theories. Not all disciplines are equally

theory-oriented. However, this account is aimed at increasing the scope of possible explanations

and I think the case can be made that this approach is general enough to capture explanations in

various sciences. One reason is that this approach is capable of capturing the way highly-

idealized models explain. Many have brought to light the fact that various branches of science

make extensive use of highly-idealized models that we would want to claim are explanatory,

even though they may not fit into a causal-explanation scheme. So while the emphasis on theory

may seem to limit disciplines, I think it can be understood as including causal and non-causal

explanations.

Another reason is that unlike other deductivist accounts of explanation, it does not focus

on universal and exceptionless laws of nature. This allows models from ecology and other

biological sciences to be capable of supporting explanations. Much as the work of Mitchell and

Woodward focused on degrees of invariance, stability, exceptionlessness, and universality, this

account can make use of models that support same-object counterfactuals. The generalizations in

other sciences are not precluded from explanation because they are not exceptionless and

universal. Further, the fact that there are discipline-relative standards of what counts as an

146

explanation is recognized. These standards are unlikely to be the same across the sciences, and

on this account there is no independent measure of the explanatory power of theories.

Explanations in physics are not necessarily better than those in other disciplines. Some will find

it unsatisfying to simply leave the explanatory power of theories to the judgments of a relevant

scientific community, but I think that formally analysing that power may well be intractable.

Many have argued that explanations in the social sciences are necessarily local and

piecemeal, and that there is not likely to be any global constraints on explanation in areas as

diverse as psychology and economics (Kincaid, 1986, 2004). However, as we have seen,

explanation in the hard sciences are also local and piecemeal. In many cases, one uses the

equations of a fundamental theory as a kind of base with theoretical constraints in order to

construct a local model of limited applicability. Something very similar happens in population

biology. Here too, one never finds a system that is perfectly captured by a simple base model of

predator-prey interaction. What is done to explain the behaviour of a particular population is to

construct a local model based on a very highly-idealized base model, featuring something like

the Lotka-Volterra predator-prey equation, but heavily modified to the target.

In some disciplines, it is not the case that theories are successively replaced with better

ones, in which case it is difficult to see how integration with a theory is necessarily tied to

explanation. However, I think it still is, but what counts as theory here is often implicit, but also

involved in the model construction. I think it is possible that in some cases there is a theory

operating in the background, according to which the model is being constructed. It is little more

than conjecture at this point, but this account could reasonably provide insight into explanation

in non-fundamental sciences, and even the social sciences. Its high threshold can serve as a guide

for formulating better explanations. If one considers a non-explanatory model in higher-level

science, one might see that an aspect that makes it merely phenomenological is not its accuracy,

but the fact that it is built up from data with no regard to an explanatory theory. An explanatory

model in this context might be constrained on both sides by accurately matching observed

phenomena and reflecting global theories about the subject matter.

The thesis was also not able to deal with mechanistic explanations in any capacity. The

popularity of mechanistic accounts and mechanistic explanations in practice demands that this be

treated in detail. A further study of this would inform the fit of this account in disciplines outside

of physics.

147

More is also needed in terms of data about scientific judgments concerning explanation.

In particular, the account makes use of scientists’ judgments about which models are considered

explanatory and what is generally thought of as causal. This dissertation acknowledges the

limitations in not beginning with this kind of data, and not also being a study of the sociology of

explanatory practices in science. This no doubt limits the precision and strength of claims about

explanation.

Outside of concerns about its general applicability, there is a genuine worry that a lot of

details remain to be filled in. The scope of the project at this early stage means that there are

quite a few promissory notes written in. For instance, there is much more that can be done to

determine more precisely what a global explanatory theory is. A number of questions remain

unanswered: are there certain necessary features of explanatory theories, or is it entirely

contextually determined? Where is the cut-off for exactly how broad in scope a theory has to be?

There are doubtless many other important question that will have to be relegated for future work.

Though the account is embryonic, untested, and in need to fleshing out, I hope to have

demonstrated in this chapter that the model-based deductivist approach is a promising one.

148

Conclusion

I will here briefly review the results of this thesis and speak to the prospects of this account. In

the first chapter, I introduce the idea of scientific explanation and the tradition in the philosophy

of science going back to Hempel and Oppenheim. I begin by establishing what is the proper

domain of an account of scientific explanation. I do this by contrasting scientific explanations

with non-scientific explanations and scientific practices that are not explanatory. In 1.3, I argue

that pragmatic issues concerning understanding and communication are an important part of

explanation, but do not exhaust it. In 1.4, I review the goals and desiderata of an account of

explanation and argue that an account should maintain a reasonably high threshold for

explanation. And it should largely reflect the explanatory judgments of a scientific community,

but not be merely descriptive. I maintain that in order to remain tractable, an account of scientific

explanation should not attempt to also be an account of explanation in general, and also that an

account of scientific explanation can succeed without also being an account of explanatory

power or explanatory depth.

In Chapter 2, I review structural accounts of explanation, in particular the one given by

Alisa Bokulich. Structural accounts are intended in part to expand the scope of explanation by

allowing non-representing, or highly-idealized, models to be considered explanatory. I begin by

taking a stance on what scientific models are, viz. that they abstract objects, and introducing the

notions of idealization and idealized models and the problems they might present for

explanation. I find that arguments problematizing laws of nature and the representational nature

of explanatory models to be persuasive. I take this as a strong indication that a model-based

account of explanation is much more likely to be successful and reflective of the practice of

explanation. The remainder of the chapter is dedicated to reviewing Alisa Bokulich’s account of

structural model explanations A structural account of explanation maintains that an explanatory

model is one that captures the structure of the system it represents, regardless of whether it

accurately represents the entities or causes. I argue that structural accounts of explanation similar

to hers are faced with two possible choices: either they stipulate a threshold above which a model

149

is explanatory, or they take a comparative approach for deciding the best structural explanation. I

conclude that neither is an attractive option: The first is arbitrary and the second shows

preference for representational models. Thus, I conclude that even though structuralists have

argued for expanding explanation to non-representing models, structural accounts of explanation

like Bokulich’s do not succeed in showing how non-representing models can be explanatory. I

take accounts like Bokulich’s to have demonstrated very cogently the need for an account of

explanation that can capture the way that highly-idealized models explain. This is a worthy goal

of an account of explanation, but the structural approach is at base representational and this gives

it preferences for accurate, representative models.

In Chapter 3, I turn my focus to causal accounts of explanation. I first review arguments

from Sandra Mitchell regarding the inherent complexity of biological systems and the

implications this has for explanation. Mitchell argues that the sheer scale, heterogeneity, and

time scales involved in biology gives rise to complexity and chaos which prevents a single causal

explanation from succeeding. She argues that explanation in biology is piecemeal and local and

not theoretical and law-based like physics. I critically review her arguments and find that her

case studies do not show what she takes them to show; that her integrative pluralism, while less

problematic than other pluralist accounts, offers little information about why these models are

explanatory; and lastly, that her account finds all models to be explanatory. I take this to show

that there are multiple models that can describe aspects of a system, which could all be

explanatory of various explananda. Where I part with Mitchell is in giving all these models a

causal interpretation.

I then present James Woodward’s account of manipulationist explanation in 3.3.

Woodward identifies causes as exhibited in the relations between the variables of models that are

invariant under a range of interventions. I note that this implies that there can be multiple models

at various levels of a system that all represent real causes. This is a metaphysical position similar

to Mitchell’s wherein there are high-level causal facts, a position referred to as emergent

causation. A worry emerges that this will fall prey to the arguments from Kim and others

proposing that emergence is inherently problematic. I then review these exclusion arguments and

their application to Woodward’s account. I find that while Wilson and List and Menzies can

mount defenses against exclusion arguments in favour of non-reductive physicalism, this fails to

apply to Woodward. Woodward maintains that there are high-level causal facts that are not

150

reducible to a fundamental level. They are described by the invariant relations between variables

of a model, of which there are many in a physical system. This is in contrast to List and Menzies

who take causation to be fully counterfactual and proportional to its effects, and to Michael

Strevens who denies that there are high-level causal facts.

Strevens’ account is concerned with how to show preference for non-fundamental

explanation given that all causal facts are at the fundamental level. What he introduces is a

measure of explanatory depth that is capable of preferring high-level explanations. Like List and

Menzies, he too looks to difference making to establish this. Strevens’ account is a two-step

process of identifying all the causes of the explanandum and eliminating those that make no

difference to it. The second step is to remove all factors that do not logically entail the

explanandum. Strevens argues that this should provide fundamental-level causes for explanations

at multiple levels. I argue that this approach is very distinct from the practice of scientific

explanation which never simply begins from knowledge of total causes. I also take issue with the

idea that all high-level models are simply abstractions and idealizations of fundamental level

causes. This is an assumption that was problematized in Chapter 2 and would preclude all

highly-idealized models. Lastly, I mention that it is not clear what fundamental level causes are,

since the fundamental level of physics is quantum.

Chapter 4 presents the traditional deductivist accounts of explanation to inform the

model-based deductivism presented in Chapter 5. I begin by looking at Hempel and

Oppenheim’s account of D-N explanation, according to which explanations are deductions of

phenomena by means of initial conditions and laws of nature. I then explore the limitations of the

D-N account before I turn to Kitcher’s unificationism, which attempts to add unification to the

D-N account in order to solve many of the counterexamples that were raised. Kitcher’s idea was

that only the most unified set of argument patterns is explanatory. By demonstrating that the

counterexample cases could be debarred as not the most unified, he was able to propose a

solution that might defend deductivism. Unfortunately, there are concerns about Kitcher’s

unificationism. One concern is that Kitcher’s criteria are to be weighed against one another, but

there is framework for determining how to do this. Another concern is that the solutions to the D-

N counterexamples do not work as well as he hopes. Lastly, the winner-take-all conception of

explanation is necessary for Kitcher, but is highly unintuitive and does not follow from the

comparative assessment of theories. I conclude the chapter reviewing the current state of

151

deductivism and presenting what the investigation of the deductivist account has provided me in

formulating a neo-deductivist account.

In the final chapter, I offer an account of scientific explanation that aims at capturing

explanatory models regardless of whether they accurately represent causes or structure. It further

does not require articulating laws of nature or restricting explanation to derivations therefrom. I

begin from the idea that explanations are arguments that feature certain scientific models, viz.

those that provide counterfactual information about the system and are related to a global theory

of science. I maintain that this account is more reflective of explanatory practice than other

accounts and that it is broadly compatible with metaphysical stances on causation as well as with

empiricism and non-reductive physicalism. It allows for the high-level models to be explanatory

independently of low-level models and fundamental causes, but does not require a theory of

emergence or physical acceptability. As such, there is a hope that this account will be broadly

applicable and largely reflective of the methods, practices, and scope of explanation in the

sciences and also palatable to a wide range of philosophical positions.

At this stage the account is rather conceptual and the criteria are still preliminary.

Although, it remains to be tested against various in-depth case studies, I hope to have

demonstrated that the prospects for a neo-deductivist account are better than the past and current

literature indicates. I hope that the work I have done motivates the idea that this kind of non-

representational approach to explanation reflects explanatory practice and that the particular

account I have outlined is a promising alternative to other popular accounts of explanation.

152

References

Achinstein, P. (1971). Law and Explanation: An Essay in the Philosophy of Science. Oxford:

Oxford University Press.

Achinstein, P. (1984). The Pragmatic Character of Explanation. Paper presented at the Biennial

Meeting of the Philosophy of Science Association.

Alexander, S. (1920). Space, time, and deity. London: Macmillan.

Anscombe, G. E. M. (1969). Causality and Extensionality. Journal of Philosophy, 66(6), 8.

Austin, J. L. (1955). How to do Things with Words. Paper presented at the William James

Lectures, Harvard University.

Barnes, E. (1992). Explanatory Unification and Scientific Understanding. Proceedings from the

Philosophy of Science Association.

Batterman, R. W. (1992). Quantum Chaos and Semiclassical Mechanics. Proceedings of the

Biennial Meetings of the Philosophy of Science Association, 1992(2), 15.

Batterman, R. W. (2002a). Asymptotics and the Role of Minimal Models. British Journal for the

Philosophy of Science, 53, 17.

Batterman, R. W. (2002b). The Devil in the Details. Oxford: Oxford University Press,.

Batterman, R. W. (2005). Critical Phenomena and Breaking Drops: Infinite Idealizations in

Physics. Studies in History and Philosophy of Modern Physics, 36, 225-244.

Batterman, R. W. (2009). Idealization and Modelling. Synthese, 169(3), 19.

Batterman, R. W., & Rice, C. C. (2014). Minimal Model Explanations. Philosophy of Science,

81(3), 349-376. doi: 10.1086/676677

Bedau, M. (1997). Weak Emergence Philosophical Perspectives 11. Cambridge: MIT Press.

Bedau, M., & Humphreys, P. (2008). Emergence : contemporary readings in philosophy and

science. Cambridge, Mass.: MIT Press.

Belot, G., & Jansson, L. (2010). Review of Reexamining the Quantum-Clasical Relation:

Beyond Reductionism and Pluralism, by A. Bokulich. Studies in History and Philosophy

of Modern Physics, 41, 3.

Bersgstrom, T. (2002). Evolution of Social Behavior: Individual and Group Selection. Journal of

Economic Perspectives, 16(2), 21.

Bickle, J. (1998). Psychoneural reduction : the new wave. Cambridge, Mass. ; London: MIT

Press.

Bickle, J. (2003). Philosophy and neuroscience : a ruthlessly reductive account. Dordrecht ;

Boston: Kluwer Academic Publishers.

Bleher, S., Ott, E., & Grebogi, C. (1989). Routes to chaotic scattering. Physical Review Letters,

63(9), 919-922.

Bokulich, A. (2008). Reexamining the Quantum-Classical Relation: Beyond Reductionsm and

Pluralism New York: Cambridge University Press.

Bokulich, A. (2011). How Scientific Models Can Explain. Synthese, 180(1), 13.

Bokulich, A. (2012). Distinguishing Explanatory from Nonexplanatory Fictions. Philosophy of

Science, 79(5), 725-737.

153

Bourke, A. F. G., & Franks, N. R. (1995). Social Evolution in Ants. Princeton: Princeton

University Press.

Boyd, R., & Richerson, P. J. (2002). Group Beneficial Norms Can Spread Rapidly in a

Structured Population. Journal of Theoretical Biology, 215, 9.

Brandom, R. (1994). Making it explicit : reasoning, representing, and discursive commitment.

Cambridge, Mass.: Harvard University Press.

Brandom, R. (2007). Inferentialism and Some of Its Challenges. Philosophy and

Phenomenological Research, 74(3), 651-676. doi: 10.1111/j.1933-1592.2007.00044.x

Brandon, R. (1978). Adaptation and Evolutionary Theory. Studies in History and Philosophy of

Science, 9(3), 25.

Brandon, R. (1981). Biological teleology: Questions and explanations. Studies in History and

Philosophy of Science Part A, 12(2), 91-105. doi: http://dx.doi.org/10.1016/0039-

3681(81)90015-7

Brigandt, I. (2011). Explanation in Biology: Reduction, Pluralism, and Explanatory Aims.

Science and Education.

Broad, C. D. (1925). The mind and its place in nature. Paterson, N. J.: Littlefield, Adams.

Bromberger, S. (1966). Why-questions. In R. G. Colodny (Ed.), Mind and Cosmos. Pittsburgh:

University of Pittsburgh Press.

Bromberger, S. (1982). On what we know we don't know. Chicago: University of Chicago Press.

Bunimovich, L. (1974). The ergodic properties of certain billiards. Functional Analysis and its

Applications, 8, 73-74.

Bunimovich, L. (1979). On the Ergodic Properties of Nowhere Dispersing Billiards.

Communications in Mathematical Physics, 65, 295-312.

Carroll, J. W. (1999). The Two Dams and That Damned Paresis. The British Journal for the

Philosophy of Science, 50(1), 65-81. doi: 10.2307/40072208

Cartwright, N. (1983). How the Laws of Physics Lie. Oxford: Oxford University Press.

Cartwright, N. (1994). Fundamentalism vs. the patchwork of laws. Proceedings of the

Aristotelian Society, 94, 279-292.

Cartwright, N. (1997). Where Do the Laws of Nature Come From? Dialectica, 51(1), 65 - 78.

Cartwright, N. (1999). The Dappled World. Cambridge, UK: Cambridge University Press.

Cat, J. (2005). Modelling cracks and cracking models: structures, mechanisms, boundary

conditions, constraints, inconsistencies and the proper domains of natural laws. Synthese,

146(3), 441-481.

Churchland, P. M. (2004). Functionalism at Forty. NYU.

Clark, A. (1997). Being there : putting brain, body, and world together again. Cambridge,

Mass.: MIT Press.

Colyvan, M., Linquist, S., Grey, W., Griffiths, P. E., Odenbaugh, J., & Possingham, H. P.

(2009). Philosophical issues in ecology: Recent trends and future directions. Ecology and

Society, 14(2).

Craver, C. (2006). Physical Law and Mechanistic Explanation in the Hodgkin and Huxley Model

of the Action Potential. Philosophy of Science, 75(5), 11.

Craver, C. (2007). Top-Down Causation without Top-down Causes. Biology and Philosophy, 22,

16.

Craver, C. (2009). Mechanisms and Natural Kinds. Philosophical Psychology, 22(5), 19.

Craver, C. (2010). Prosthetic Models. Philosophy of Science, 77(5), 11.

154

Darden, L. (2006). Reasoning in biological discoveries : essays on mechanisms, interfield

relations, and anomaly resolution. Cambridge ; New York: Cambridge University Press.

Davies, D. (1996). Explanatory Disunities and the Unity of Science. International Studies in the

Philosophy of Science, 10(1), 5-21.

Dawkins, R. (1989). The Selfish Gene. Oxford: Oxford University Press.

De Regt, H. (2013). Understanding and Explanation: Living Apart Together? Studies in History

and Philosophy of Science, 44, 4.

Deneubourg, J. e. a. (1987). Self-Organization Mechanisms in Ant Societies (II): Learning in

Foraging and Division of Labour. Experientia Supplementum, 54, 177-196.

Dettman, C. P., & Georgiou, O. (2010). Open Intermittent Billiards: A dynamical window.

Retrieved Feb 3, 2015, from http://iopscience.iop.org/1751-8121/labtalk-article/46000

Dettman, C. P., & Georgiou, O. (2011). Open Mushrooms: Stickiness Revisited. Journal of

Physics: Mathematical and Theoretical, 44, 30pp.

Díez, J., Khalifa, K., & Leuridan, B. (2013). General theories of explanation: buyer beware.

Synthese, 190(3), 379-396. doi: 10.1007/s11229-011-0020-8

Donato Rodríguez, X., & Zamora Bonilla, J. (2009). Credibility, Idealisation, and Model

Building: An Inferential Approach. Erkenntnis, 70(1), 101-118. doi: 10.1007/s10670-

008-9139-5

Dowe, P. (2000). Physical Causation. Cambridge ; New York: Cambridge University Press.

Dupré, J. (1983). The Disunity of Science. Mind: A Quarterly Review of Philosophy, 92(0026-

4423, 0026-4423), 321-346.

Esfeld, M., & Lam, V. (2008). Moderate structural realism about space-time. Synthese, 160, 27-

46.

Feyerabend, P. (1975). Against method (3rd ed.). London: Verso.

Fodor, J. (1974). Special Sciences. Synthese, 28, 97-115.

Fodor, J. (1975). The Language of Thought. Harvard, MA: Harvard University Press.

French, S., & Ladyman, J. (2003). Remodelling structural realism: Quantum physics and the

metaphysics of structure. Synthese, 136, 31-56.

French, S., & Saatsi, J. (2006). Realism about Structure: The Semantic View and Nonlinguistic

Representations. Philosophy of Science, 73(5), 548-559.

Friedman, M. (1974). Explanation and Scientific Understanding. Journal of Philosophy, 71(5).

Frisch, M. (2014). Laws in Physics. European Review, 22(S1), S33-S49. doi:

10.1017/S1062798713000768

Godfrey-Smith, P. (1994). A Modern History Theory of Functions. Nous, 28(3), 344-362. doi:

10.2307/2216063

Goodman, N. (1973). Fact, fiction, and forecast (3d ed.). Indianapolis: Bobbs-Merrill.

Grantham, T. A. (2004). Conceptualizing the (Dis)unity of Science. Philosophy of Science,

71(2), 133-155.

Gutzwiller, M. C. (1990). Chaos in classical and quantum mechanics. New York: Springer

Verlag.

Harding, S. (1986). The Science Question in Feminism: Cornell University Press.

Hartmann, S. (1996). The World as a Process: Simulations in the Natural and Social Sciences. In

R. Hegselman, U. Muller & K. G. Troitzsch (Eds.), Modeling and Simulation in the

Social Sciences from the Philosophy of Science Point of View. Dortrecht: Kluwer.

Hartmann, S., & Schupbach, J. (2010). Notre Dame Philosophical Reviews, 6(38).

155

Heller, E. J. (1984). Bound-State Eigenfunctions of Classically Chaotic Hamiltonian Systems:

Scars of Periodic Orbits. Physical Review Letters, 53(16), 1515-1518.

Heller, E. J. (1986). Qualitative properties of eigenfunctions of classically chaotic Hamiltonian

Quantum Chaos and Statistical Nuclear Physics, 263, 162 - 181.

Hempel, C. G. (1965a). Aspects of Scientific Explanation and Other Essays. New York: The Free

Press.

Hempel, C. G. (1965b). Studies in the Logic of Confirmation Aspects of Scientific Explanation.

New York: Free Press.

Hempel, C. G. (1966). Philosophy of natural science. Englewood Cliffs, N.J.: Prentice-Hall, Inc.

Hempel, C. G., & Oppenheim, P. (1948). Studies in the Logic of Explanation. Philosophy of

Science, 15(2), 40.

Humphreys, P. (1997). How Properties Emerge. Philosophy of Science, 64, 1-17.

Hüttemann, A. (2004). What's wrong with microphysicalism? London ; New York: Routledge.

Hüttemann, A. (2014). Ceteris Paribus Laws in Physics. Erkenntnis, 79(10), 1715-1728. doi:

10.1007/s10670-014-9637-6

Jackson, F. (1998). From Metaphysics to Ethics: A Defence of Conceptual Analysis

Jansson, L. (2014a). Causal Theories of Explanation and the Challenge of Explanatory

Disagreement. Philosophy of Science, 81.

Jansson, L. (2014b). Nomological Explanation Without Causation. Paper presented at the

Explanation Beyond Causation, LMU Munich.

Kaplan, D. M., & Craver, C. (2011). The Explanatory Force of Dynamical and Mathematical

Models in Neuroscience: A Mechanistic Perspective. Philosophy of Science, 78(4), 26.

Kaplan, L., & Heller, E. J. (1999). Measuring scars of periodic orbits. Physical Review E, 59(6),

6609-6628.

Karakostas, V. (2009). From Atomism to Holism: The Primacy of Non-Supervenient Relations.

NeuroQuantology, 7(4). doi: http://dx.doi.org/10.14704/nq.2009.7.4.258

Khalifa, K. (2012). Inaugurating Understanding or Repackaging Explanation? Philosophy of

Science, 79(1), 15-37.

Kim, J. (1999). Making Sense of Emergence. Philosophical Studies, 95(1), 33.

Kincaid, H. (1986). Reduction, Explanation, and Individualism. Philosophy of Science, 53(4),

492-513.

Kincaid, H. (2004). Contextualism, Explanation, and the Social Sciences. Philosophical

Explorations, 7(3), 201-218.

King, C. (2009). Exploring Quantum, Classical and Semiclassical Chaos in the Stadium Billiard.

Quanta, 3(1), 16-31.

King, M. (2014). Idealization and Structural Explanation in Physics. Paper presented at the

Biennial Meeting of Philosophy of Science Association, Chicago, IL.

King, M. (2015). On structural accounts of model-explanations. Synthese. doi: 10.1007/s11229-

015-0885-z

Kitcher, P. (1981). Explanatory unification. Philosophy of Science, 48.

Kitcher, P. (1985). Two Approaches to Explanation. Philosophy of Science, 48, 507-531.

Kitcher, P. (1989). Explanatory Unification and the Causal Structure of the World. In P. Kitcher

& W. Salmon (Eds.), Scientific Explanation. Minneapolis: University of Minnesota Press.

Knuuttila, T. (2005). Models, representation, and mediation. Philosophy of Science, 72, 1260-

1271.

156

Knuuttila, T., & Loettgers, A. (2012). Synthetic Modeling and Mechanistic Account: Material

Recombination and Beyond. Proceedings from the Philosophy of Science Association.

Kripke, S. (1972). Naming and Necessity. In D. D. & H. G. (Eds.), Semantics and Natural

Language. Boston: Reidel.

Kuhn, T. S. (1962). The Structure of Scientific Revolution. Chicago, IL: University of Chicago

Press.

Kyburg, H. E. (1965). Comment. Philosophy of Science, 32, 147-151.

Ladyman, J. (1998). What is structural realism? Studies in History and Philosophy of Modern

Science, 29, 409-424.

Lange, M. (2002). Who's Afraid of Ceteris Paribus Laws? Erkenntnis, 57, 407-423.

Lange, M. (2004). A Note on Scientific Essentialism, Laws of Nature, and Counterfactual

Conditionals. Australasian Journal of Philosophy, 82(2), 227-241.

Lawton, J. H. (1999). Are There General Laws in Ecology? Nordic Society Oikos, 84(2), 15.

Levins, R. (1966). The Strategy of Model Building in Population Biology. American Scientist,

54(4), 421-431.

Levins, R. (1968). Evolution in changing environments : some theoretical explorations.

Princeton, N.J.: Princeton University Press.

Levy, A. (2011). Makes a Difference. Biology & Philosophy.

Lewis, D. (1973). Counterfactuals. Cambridge: Harvard University Press.

Lewis, D. (1979). Counterfactual Dependence and Time's Arrow. Nous, 13, 455 - 476.

Lewis, D. (1986). On the plurality of worlds. Oxford, UK ; New York, NY, USA: B. Blackwell.

List, C., & Menzies, P. (2009). Nonreductive Physicalism and the Limits of the Exlcusion

Principle. Journal of Philosophy, 106, 27.

Longino, H. (1990). Science as Social Knowledge. Princeton, NJ: Princeton University Press.

Love, A. C. (2012). Hierarchy, Causation and Explanation: Ubiquity, Locality and Pluralism.

Interface Focus, 2, 155-125.

Machamer, P., Darden, L., & Craver, C. (2000). Thinking About Mechanisms. Philosophy of

Science, 67(1).

McArthur, D. (2003). Reconsidering Structural Realism. Canadian Journal of Philosophy, 33(4),

517 - 536.

McDonald, S. W., & Kaufman, A. N. (1979). Spectrum and Eigenfunctions for a Hamiltonian

with Stochastic Trajectories. Physical Review Letters, 42(18), 1189-1191.

McMullin, E. (1985). Galilean Idealization. Studies in History and Philosophy of Science, 16(3),

26.

Meek, C., & Glymour, C. (1994). Conditioning and Intervening. The British Journal for the

Philosophy of Science, 45(4), 1001-1021. doi: 10.1093/bjps/45.4.1001

Miller, W. H. (1986). Semiclassical Methods in Chhemical Physics. Science, 23, 171-177.

Mitchell, S. (1997). Pragmatic Laws. Proceedings from the Philosophy of Science Association,

64, 11.

Mitchell, S. (2000). Dimensions of Scientific Law. Philosophy of Science, 67, 23.

Mitchell, S. (2002a). Ceteris Paribus – An Inadequate Representation for Biological

Contingency. Erkenntnis, 57, 21.

Mitchell, S. (2002b). Integrative Pluralism. Biology and Philosophy, 17, 55-70.

Mitchell, S. (2003). Biological Complexity and Integrative Pluralism. New York: Cambridge

University Press.

157

Morrison, M. (1999). Models as Autonomous Agents. In M. Morrison & M. Morgan (Eds.),

Models as Mediators: Perspectives on Natural and Social Science (pp. 38 - 65 ).

Cambridge: Cumbridge University Press.

Nagel, E. (1961). The Structure of Science : Problems in the Logic of Scientific Explanation.

New York: Harcourt, Brace & World.

Ney, A. (2009). Physical Causation and Difference-Making. British Journal for the Philosophy

of Science, 60, 27.

Nickel, B. (2010). How General do Theories of Explanation Need to be? Nous, 44(2), 305-328.

Page, R., & Mitchell, S. (1991). Self Organization and Adaptation in Insect Societies. Philosophy

of Science, 1990(2), 9.

Parker, W. S. (2011). When Climate Models Agree: The Significance of Robust Model

Predictions. Philosophy of Science, 78(4), 21.

Pearl, J. (2000). Causality: Models, Reasoning, and Inference. Cambridge: Cambridge

University Press.

Popper, K. R. (1959). The Logic of Scientific Discovery. New York: Routledge.

Popper, K. R., & Eccles, J. C. (1977). The self and its brain. Berlin: Springer International.

Potochnik, A. (2011). Explanation and understanding. European Journal for Philosophy of

Science, 1(1), 29-38. doi: 10.1007/s13194-010-0002-6

Putnam, H. (1988). Representation and Reality. Cambridge, MA: MIT Press.

Reeve, H. K., & Sherman, P. W. (1993). Adaptation and the goals of evolutionary research.

Quarterly Review of Biology, 62, 1-32.

Reiss, J. (2012). Causation in the sciences: An inferentialist account. Studies in History and

Philosophy of Biol & Biomed Sci, 43(4), 769-777. doi: 10.1016/j.shpsc.2012.05.005

Reutlinger, A. (2014). Are causal facts really explanatorily emergent? Ladyman and Ross on

higher-level causal facts and renormalization group explanation. Synthese, 1-15. doi:

10.1007/s11229-014-0530-2

Rice, C. (2012). Optimality explanations: a plea for an alternative approach. Biology and

Philosophy, 27, 18.

Rice, C. (2013). Moving Beyond Causes: Optimality Models and Scientific Explanation. Nous,

49(2).

Rice, C. (2015). Moving Beyond Causes: Optimality Models and Scientific Explanation. Nous,

49(3), 589-615.

Rueger, A. (2001). Explanations at Multiple Levels. Minds and Machines, 11, 503–520.

Saatsi, J. (2014). Worthwhile Distinctions: Kinematic, Dynamic and Non-causal explanations.

Paper presented at the Explanation Beyond Causation, LMU Munich.

Saatsi, J., & Pexton, M. (2012). Reassessing Woodward’s account of explanation: regularities,

counterfactuals, and non-causal explanations. Paper presented at the Biennial Meeting of

the Philosophy of Science Association.

Salmon, W. C. (1971). Statistical Explanation. In W. C. Salmon, R. C. Jeffrey & J. G. Greeno

(Eds.), Statistical Explanation & Statistical Relevance (pp. 29-87). (Pittsburgh)

University of Pittsburgh Press (C1971).

Salmon, W. C. (1984). Scientific explanation and the causal structure of the world. Princeton,

NJ: Princeton University Press.

Salmon, W. C. (1989). Four decades of scientific explanation. Minneapolis: University of

Minnesota Press.

158

Schindler, S. (2013). Mechanistic Explanation: Asymmetry Lost. In Karakostas & Dieks (Eds.),

Recent Progress in Philosophy of Science: Perspectives and Foundational Problems.

Scriven, M. (1959). Explanation and Prediction in Evolutionary Theory. Science, 130(3374), 5.

Scriven, M. (1962). Explanations, Predictions, and Laws. Minnesota Studies in the Philosophy of

Science, 3, 170-230.

Sherman, P. W. (1988). The Levels of Analysis. Animal Behaviour, 36, 4.

Sober, E. (1984). The Nature of Selection. Cambridge, MA: MIT Press.

Sober, E., & Wilson, D. S. (1998). Unto others : the evolution and psychology of unselfish

behavior. Cambridge, Mass. ; London: Harvard University Press.

Spirtes, P., Glymour, C. N., & Scheines, R. (1993). Causation, prediction, and search (2nd ed.).

Cambridge, Mass.: MIT Press.

Stalnaker, R. (1968). A Theory of Conditionals. In N. Rescher (Ed.), Studies in Logical Theory

(pp. 98-112). Oxford: Blackwell.

Stöckmann, H.-J. (2010). Stoe Billiards. In stoe_billiards.jpeg (Ed.). Sholarpedia.

Strevens, M. (2004). Causal and Unificationist Approaches Unified-Causally. Nous, 38(1), 154-

176.

Strevens, M. (2008). Depth: An Account of Scientific Explanation. Harvard, MA: Harvard

University Press.

Tao, T. (2007). Open question: scarring for the Bunimovich stadium. Retrieved from

http://terrytao.wordpress.com/2007/03/28/open-question-scarring-for-the-bunimovich-

stadium/

Teller, P. (2001). Twilight of the Perfect Model Model. Erkenntnis, 55(3).

Tofts, C., & Franks, N. R. (1992). Doing the Right Thing: Ants, Honeybeed and Naked Mole-

Rats. Trends in Ecology and Evolution, 7, 4.

Tomsovic, S., & Heller, E. J. (1993). Long-time semiclassical dynamics of chaos: The stadium

billiard. Physical Review E, 47(1), 282-299.

Tooley, M. (2003). The Stalnaker-Lewis Approach to Counterfactuals. The Journal of

Philosophy, 100(7), 371-377. doi: 10.2307/3655784

van Fraassen, B. C. (1980). The Scientific Image. New York: Clarendon Press.

van Fraassen, B. C., Churchland, P. M., & Hooker, C. A. (1985). Images of science : essays on

realism and empiricism, with a reply from Bas C. van Fraassen. Chicago: University of

Chicago Press.

Wayne, A. (2011). Extending the Scope of Explanatory idealization. Philosophy of Science,

78(5), 11.

Wayne, A. (2012). Emergence and Singular Limits. Synthese, 184, 15.

Wayne, A. (2014). Idealization and Explanation: A Deductivist Approach. Paper presented at the

Explanation Beyond Causation, LMU Munich.

Wayne, A. (2015). Causal Relations and Explanatory Strategies in Physics. International Studies

in the Philosophy of Science, 29, 75-89.

Wayne, A. (2016). Deductivist Explanation Redux. Manuscript submitted for publication.

Weber, M. (2006). Causes without Mechanisms: Experimental Regularities, Physical Laws, and

Neuroscientific Explanation. Philosophy of Science, 75(5), 995 - 1007.

Weisberg, M. (2007). Three Kinds of idealization. The Journal of Philosophy, 104(12), 20.

Weslake, B. (2010). Explanatory Depth. Philosophy of Science, 77(2), 273-294.

Wilson, J. (1999). How Superduper does a Physicalist Supervenience need to be? Philosophical

Quarterly, 49(194).

159

Wilson, J. (2011a). Metaphysical emergence: Weak and Strong. In T. Bigaj & C. Wuthrich

(Eds.), Metaphysics in Contemporary Physics: Poznan Studies in the Philosophy of the

Sciences and the Humanities.

Wilson, J. (2011b). Non-reductive Realization and the Powers-based Subset Strategy. The

Monist, 94(1), 121-154.

Wimsatt, W. C. (1987). False Models as Means to Truer Theories. In M. H. H. Nitecki, A. (Ed.),

Neural Models in Biology. Oxford, UK: Oxford University Press.

Wintgen, D., Richter, K., & Tanner, G. (1992). The semiclassical helium atom. Chaos: An

Interdisciplinary Journal of Nonlinear Science, 2(1), 19-33.

Woodward, J. (1989). A Causal-Mechanical Model of Explanation. In P. Kitcher & W. Salmon

(Eds.), Scientific Explanation. Minneapolis: University of Minnesota Press.

Woodward, J. (2000). Explanation and Invariance in the Special Sciences. British Journal for the

Philosophy of Science, 51, 197-254.

Woodward, J. (2003). Making Things Happen: A Theory of Causal Explanation. Oxford: Oxford

University Press.

Woodward, J. (2011). Interventionism and Causal Exclusion. University of Pittsburgh.

Woodward, J., & Hitchcock, C. (2003a). Explanatory Generalizations, part I: A Counterfactual

Account. Nous, 37(1), 1-24.

Woodward, J., & Hitchcock, C. (2003b). Explanatory Generalizations, Part II: Plumbing

Explanatory Depth. Nous, 37(2), 181-199.

Worrall, J. (1989). Structural Realism: The Best of Both Worlds? Dialectica, 43(1-2).

Yablo, S. (1992). Mental Causation. Philosophical Review, 101(2), 35.