You are on page 1of 50

Mechanisms in Science

First published Wed Nov 18, 2015


Around the turn of the twenty-first century, what has come to be called the new
mechanical philosophy (or, for brevity, the new mechanism) emerged as a framework
for thinking about the philosophical assumptions underlying many areas of science,
especially in sciences such as biology, neuroscience, and psychology. In this entry, we
introduce and summarize the distinctive features of this framework, and we discuss how
it addresses a range of classic issues in the philosophy of science, including
explanation, metaphysics, the relations between scientific disciplines, and the process of
scientific discovery. For each of these issues, we show how the mechanistic framework
has reoriented philosophical work, what the new mechanism has contributed to the
discussion, and what remains to be done.

 1. The Rise of the New Mechanism


 2. The Concept of a Mechanism
o 2.1 Phenomenon
 2.1.1 Producing, Underlying, and Maintaining
 2.1.2 Regularity
o 2.2 Parts
o 2.3 Causings
 2.3.1 Conserved Quantity Accounts
 2.3.2 Mechanistic Accounts
 2.3.3 Activity-Based Accounts
 2.3.4 Counterfactual Accounts
o 2.4 Organization
 2.4.1 Organization and Aggregativity
 2.4.2 Varieties of Organization
 2.4.3 Modularity
 2.4.4 Jointness
 2.4.5 Levels
 2.4.6 Stable and Ephemeral Mechanisms
o 2.5 What Mechanisms Are Not and What Are Not Mechanisms
 2.5.1 What Mechanisms Are Not
 2.5.2 What Are Not Mechanisms
o 2.6 Philosophical Work to Be Done
 3. Explanation: From Formal Analyses to Material Structures
o 3.1 Etiological and Constitutive Explanations
o 3.2 Constitutive Relevance
o 3.3 Mechanisms and Models
 3.3.1 Characterizing Completeness
 3.3.2 Abstraction and Idealization
o 3.4 Philosophical Work to Be Done
 4. Metaphysics of Mechanisms
o 4.1 Token Mechanisms, Type Mechanisms, and Laws
o 4.2 Mechanisms, Levels, and Emergence
o 4.3 Mechanisms and Realization
o 4.4 Mechanisms and Natural Kinds
o 4.5 Mechanisms and Functions
o 4.6 Philosophical Work to Be Done
 5. Relations between Scientific Disciplines: From Theory Reduction to
Mechanism Integration
o 5.1 Theory Reduction
o 5.2 Mechanism Integration
 5.2.1 Integrative Pluralism
o 5.3 Philosophical Work to Be Done
 6. Discovery: From A-ha Moments to Discovery Strategies
o 6.1 Discovery via A-ha Moments
o 6.2 Discovery via Strategies
o 6.3 Mechanistic Evidence in Medical Discoveries
o 6.4 Philosophical Work to Be Done
 7. Conclusion
 Bibliography
 Academic Tools
 Other Internet Resources
 Related Entries

1. The Rise of the New Mechanism


Twentieth century philosophy of science was largely dominated by logical empiricism.
More a framework for doing philosophy of science than any coherent set of doctrines,
logical empiricism addressed a range of issues in philosophy of science through the lens
of the logical and mathematical structures constitutive of scientific thought and practice
(see the entry on logical empiricism). Logical empiricism tended, by and large, to focus
on abstract, epistemic features of science, with little attention to scientific practice.
Physics was the dominant exemplar.
The new mechanical philosophy emerged around the turn of the twenty-first century as
a new framework for thinking about the philosophy of science. The philosophers who
developed this framework were, by comparison with the logical empiricists,
practitioners as well of the history of science and tended, by and large, to focus on the
biological, rather than physical, sciences. Many new mechanists developed their
framework explicitly as a successor to logical empiricist treatments of causation, levels,
explanation, laws of nature, reduction, and discovery.
As with logical empiricism, the new mechanical philosophy is less a systematic and
coherent set of doctrines than it is an orientation to the subject matter of the philosophy
of science. The approach emerged as philosophers and historians of science began to
break from the once-standard practice of reconstructing scientific inference with the
tools of logic and, instead, to embrace detailed investigation of actual episodes from the
history of science. The main tenets of logical empiricism had been under intense
criticism for decades, and a new era of historically informed philosophy of science had
taken hold through the works of, e.g., Kuhn (1962), Laudan (1977), and Lakatos (1977).
To many such scholars raised in this post-logical empiricist milieu, it appeared that
much of the practice of contemporary science (both in the laboratory and in print) was
driven by the search for mechanisms, that many of the grand achievements in the
history of science were discoveries of mechanisms, and that more traditional philosophy
of science, for whatever reason, had failed to appreciate this central feature of the
scientific worldview.
Aspects of the new mechanical philosophy began to emerge in the late 1960s. Fodor
(1968), for example, contrasted mechanistic explanations (dealing with parts and their
law-like interactions) with functional explanations in psychology. Wimsatt (1972a,
1976), building on the work of Simon (1962) and Kaufman (1971), argued repeatedly
that the abstract and idealized structures of logical empiricism were ill-suited to
understanding how scientists discover and explain complex systems at multiple levels
of organization. Cummins (1975) provided an account of functional analyses,
characterizing a function as a contribution a component part makes to the overall
capacity of some system that includes that component. Salmon (1984, 1989) argued that
empiricist views of scientific explanation are fundamentally flawed because they
neglect causal mechanisms. Cartwright (1989) argued that the logical empiricist
conception of a law of nature is, in fact, a philosophical fiction used to describe the
search for capacities and nomological machines.
These strands began to coalesce into an over-arching perspective in the 1990s. The
earliest clear statement of the new mechanism was Bechtel and Richardson’s (2010
[1993]) Discovering Complexity. They self-consciously put aside logical empiricist
concerns with theory reduction and focused instead on the process by which scientists
discover mechanisms (see Section 6 below). Soon after, Glennan argued that
mechanisms are the secret connexion Hume sought between cause and effect (1996), a
thesis related to and partly inspired by Cartwright’s focus on capacities and
nomological machines (Glennan 1997). Likewise, Thagard’s How Scientists Explain
Disease centered the search for causes and mechanisms in medicine (Thagard 2000; see
also Section 6 below). Machamer, Darden and Craver’s “Thinking about Mechanisms”
(Machamer, Darden, and Craver 2000; familiarly known as “MDC”) drew these strands
together and became for many the lightening rod of the new mechanist perspective.
MDC suggested that the philosophy of biology, and perhaps the philosophy of science
more generally, should be restructured around the fundamental idea that many scientists
organize their work around the search for mechanisms.

2. The Concept of a Mechanism


The term “mechanism” emerged in the seventeenth century and derived from Greek and
Latin terms for “machine” (Dijksterhuis 1961). Descartes understood mechanics as the
basic building block of the physical world; in Le Monde, he proposed to explain diverse
phenomena in the natural world (such as planetary motion, the tides, the motion of the
blood, and the properties of light) in terms of the conservation of inertial motion
through contact action (see the entry on René Descartes). Subsequently, the idea of
mechanism has been transformed many times to reflect an evolving understanding of
the basic causal forces in the world (besides conserved motion): e.g., attraction and
repulsion (du Bois Reymond), conservation of energy (Helmholtz), gravitational
attraction (Newton) (Boas 1952; Westfall 1971; see also entries on Hermann von
Helmholtz and Isaac Newton). The concept of mechanism has had an almost separate
evolution in the history of the life sciences (Allen 2005; Des Chene 2001, 2005;
Nicholson 2012), at times eschewing the metaphysical austerity embraced by Descartes
and many early mechanists.
The new mechanists inherit the word “mechanism” from these antecedents, but, in their
effort to capture how the term is used in contemporary science, have distanced
themselves both from the idea that mechanisms are machines and especially from the
austere metaphysical world picture in which all real change involves only one or a
limited set of fundamental activities or forces (cf. Andersen 2014a,b).
Mechanists have generally eschewed the effort to spell out necessary and sufficient
conditions for something to be a mechanism. Instead, they offer qualitative descriptions
designed to capture the way scientists use the term and deploy the concept in their
experimental and inferential practices.
Three characterizations are most commonly cited:

 MDC: “Mechanisms are entities and activities organized such that they are
productive of regular changes from start or set-up to finish or termination
conditions” (2000: 3).
 Glennan: “A mechanism for a behavior is a complex system that produces that
behavior by the interaction of a number of parts, where the interaction between
parts can be characterized by direct, invariant, change-relating generalizations”
(2002: S344).
 Bechtel and Abrahamsen: “A mechanism is a structure performing a function in
virtue of its component parts, component operations, and their organization. The
orchestrated functioning of the mechanism is responsible for one or more
phenomena” (2005: 423).
Each of these characterizations contains four basic features: (1) a phenomenon, (2)
parts, (3) causings, and (4) organization. We consider each of these in detail below.
A useful canonical visual representation of a mechanism underlying a phenomenon is
shown in Figure 1 (from Craver 2007). At the top is the phenomenon, some
system S engaged in behavior ψ. This is the behavior of the mechanism as a whole.
Beneath it are the parts (the Xs) and their activities (the φs) organized together. The
dotted roughly-vertical lines reflect the fact that the parts and activities are contained
within, are components of, the mechanism engaged in this behavior. Thus represented,
mechanisms are decompositional in the sense that the behavior of the system as a whole
can be broken down into organized interactions among the activities of the parts.
Figure 1. A visual representation of a mechanism (adapted from Craver 2007).

In the early literature, these different characterizations were often treated as


competitors. Tabery (2004) argued instead that they reflect different, and
complementary, emphases and intellectual orientations. Many mechanists have adopted
this ecumenical stance. For example, Illari and Williamson offer a “consensus concept”
of mechanism:
A mechanism for a phenomenon consists of entities and activities organized in such a
way that they are responsible for the phenomenon. (2012: 120)
Likewise, Glennan refers to “minimal mechanism”:
A mechanism for a phenomenon consists of entities (or parts) whose activities and
interactions are organized in such a way that they produce the phenomenon. (Glennan
forthcoming: Ch. 2)
These ecumenical characterizations each include the four basic elements and are
designed to make the characterization more inclusive. MDC’s insistence on the
regularity of mechanisms is abandoned, for example, to accommodate mechanisms that
work only once or that work irregularly (Skipper and Milstein 2005; Bogen 2005; see
also Section 2.1.2 below). Bechtel and Abrahamsen’s emphasis on the “functions” is
abandoned to accommodate mechanisms that serve no end and to distance mechanism
from this loaded term so often opposed to mechanism (although see Craver 2001a;
Garson 2013; Maley and Piccinini forthcoming; see also Section 4.5 below).
These ecumenical characterizations intentionally downplay the fact that the term
“mechanism” is used differently in different scientific and philosophical contexts (see
Levy 2013 and Andersen 2014a,b for alternative overviews of the differences). Indeed,
much of the progress in the early years involved learning to recognize the many ways
that the term “mechanism” can be used and the many commitments that can be
undertaken in its name. (For still other characterizations of mechanism, see Woodward
(2002), Fagan (2012), Nicholson (2012), and Garson (2013)). Taking these ecumenical
views as a starting point, we now consider the four basic components: 1) the
phenomenon, 2) parts, 3) causings, and 4) organization.

2.1 Phenomenon
The phenomenon is the behavior of the mechanism as a whole. All mechanisms are
mechanisms of some phenomenon (Kauffman 1971; Glennan 1996, 2002). The
mechanism of protein synthesis synthesizes proteins. The mechanism of the action
potential generates action potentials. The boundaries of a mechanism—what is in the
mechanism and what is not—are fixed by reference to the phenomenon that the
mechanism explains. The components in a mechanism are components in virtue of
being relevant to the phenomenon.
MDC (2000) describe mechanisms as working from start- or set-up conditions to
termination conditions. They insist that it is impoverished to describe the phenomenon
as an input-output relation because there are often many such inputs and outputs from a
mechanism and because central features of a phenomenon might be neither inputs nor
outputs (but rather details about how the phenomenon unfolds over time). Darden,
appealing to the example of protein synthesis, often associates the phenomenon with the
end-state: the protein (Darden 2006). Craver (2007), following Cummins (1975) and
Cartwright (1989), often speaks of the phenomenon roughly as a capacity or behavior of
the mechanism as a whole.
2.1.1 Producing, Underlying, and Maintaining
New mechanists speak variously of the mechanism as producing, underlying, or
maintaining the phenomenon (Craver and Darden 2013). The language of production is
best applied to mechanisms conceived as a causal sequence terminating in some end-
product: as when a virus produces symptoms via a disease mechanism or an enzyme
phosphorylates a substrate. In such cases, the phenomenon might be an object (the
production of a protein), a state of affairs (being phosphorylated), or an activity or event
(such as digestion). For many physiological mechanisms, in contrast, it is more
appropriate to say that the mechanism underlies the phenomenon. The mechanism of
the action potential or of working memory, for example, underlies the phenomenon,
here characteristically understood as a capacity or behavior of the mechanism as a
whole. Finally, a mechanism might maintain a phenomenon, as when homeostatic
mechanisms hold body temperature within tightly circumscribed boundaries. In such
cases, the phenomenon is a state of affairs, or perhaps a range of states of affairs, that is
held in place by the mechanism. These ways of talking can in many cases be inter-
translated (e.g., the product is produced, the production has an underlying mechanism,
and the state of affairs is maintained by an underlying mechanism). Yet clearly
confusion can arise from mixing these ways of talking.
2.1.2 Regularity
Must the relationship between the mechanism and the phenomenon be regular? This is
an area of active discussion (DesAutels 2011; Andersen 2011, 2014a,b; Krickel 2014).
MDC stipulate that mechanisms are regular in that they work “always or for the most
part in the same ways under the same conditions” (2000: 3). Some have understood this
(incorrectly in our view) as asserting that there are no mechanisms that work only once,
or that a mechanism must work significantly more than once in order to count as a
mechanism.
Some argue that mechanisms have to be regular in this factual sense (Andersen
2014a,b); i.e., repeated on many occasions (see Leuridan 2010). This view would seem
to require a somewhat arbitrary cut-off point in degree of regularity between things that
truly count as mechanisms and those that do not. Some mechanists (Bogen 2005;
Glennan 2009) argue that there is no difficulty applying the term “mechanism” to one-
off causal sequences, as when an historian speaks of the mechanism that gave rise to
World War I. Other mechanists argue that the type-token distinction is too crude a
dichotomy to capture the many levels of abstraction at which mechanism types and
tokens might be characterized (Darden 1991).
It is possible, however, to read the MDC statement as asserting, not a factual kind of
regularity, but as a counterfactual kind of near-determinism: were all the conditions the
same, then the mechanism would likely produce the same phenomenon, where “likely”
accommodates mechanisms with stochastic elements.
While the MDC account leaves open the possibility that some mechanisms are
stochastic, it clearly rules out mechanisms that usually fail to produce their phenomena.
Skipper and Millstein (2005) press this point to argue that the MDC account cannot
accommodate the idea that natural selection is a mechanism. If, as Gould (1990) argued,
one could not reproduce the history of life by rewinding the tapes and letting things
play forward again, then natural selection would not be an MDC mechanism (see
also Section 2.6 below). It is unclear why MDC would allow for the possibility of
stochastic mechanisms and rule out, by definition, the possibility that they might fail
more often than they work. Whether any biological mechanisms are truly irregular in
this sense (i.e., all the causally relevant factors are the same but the product of the
mechanism differs) is a separate question from whether they are mechanisms
simpliciter (see Bogen 2005; Machamer 2004; Steel 2008 develops a stochastic account
of mechanisms).
Krickel (2014) reviews the many different ways of unpacking the relevant notion of
regularity (see also Andersen 2012). Her favored solution, “reverse regularity,” holds
that there must be a generalization to the effect that, typically, when the phenomenon
occurs, the mechanism was acting.

2.2 Parts
Mechanists have struggled to find a concise way to express the idea of parthood
required of the components in a mechanism. The project is to develop an account that is
both sufficiently permissive to include the paradigmatic mechanisms from diverse areas
of science and yet not vacuous.
Formal mereologies are difficult to apply to the material parts of biological
mechanisms. Axioms of mereology, such as reflexivity (everything is a part of itself)
and unrestricted composition (any two things form a whole) do not apply in standard
biological uses of the “part” concept.
Glennan (1996) recognized the difficulty of defining parthood very early on. His
proposal:
The parts of mechanisms must have a kind of robustness and reality apart from their
place within that mechanism. It should in principle be possible to take the part out of
the mechanism and consider its properties in another context. (1996: 53)
Yet even this is perhaps too strong, given that some parts of a mechanism might
become unstable when removed from their mechanistic context. Later, Glennan (2002:
S345) says that the properties of a part must be stable in the absence of interventions, or
that parts must be stable enough to be called objects. This notion is perhaps too strong
to accommodate the more ephemeral parts of some biochemical mechanisms or of the
mechanisms of natural selection (Skipper and Millstein 2005; but see Illari and
Williamson 2010).

2.3 Causings
Mechanists have disagreed with one another about how to understand the cause in
causal mechanism. New mechanists have in general been at pains both (1) to liberate
the relevant causal notion from any overly austere view that restricts causation to only a
small class of phenomena (such as collisions, attraction/repulsion, or energy
conservation), and (2) to distance themselves from the Humean, regularist conception
of causation common among logical empiricists (see also the entry on the the
metaphysics of causation). Four ways of unpacking the cause in causal mechanism have
been discussed: conserved quantity accounts, mechanistic accounts, activities accounts,
and counterfactual accounts. (It should be noted that some mechanists have evolved in
their thinking about causation.)
2.3.1 Conserved Quantity Accounts
According to transmission accounts, causation involves the transmission and
propagation of marks or conserved quantities (Salmon 1984, 1994; Dowe 1992). The
most influential form of this view holds that two causal processes causally interact
when they intersect in space-time and exchange some amount of a conserved quantity,
such as mass. On this view, causation is local (the processes must intersect) and
singular (it is fully instantiated in particular causal processes), though the account relies
upon laws of conservation (Hitchcock 1995). Although this view inspired many of the
new mechanists, and although it shares their commitment to looking toward science for
an account of causation, it has generally been rejected by new mechanists (though see
Millstein 2006; Roe 2014).
This view has been unpopular in part because it has little direct application in
nonfundamental sciences, such as biology. The causal claims biologists make usually
don’t involve explicit reference to conserved quantities (even if they presuppose such
notions fundamentally) (Glennan 2002; Craver 2007). Furthermore, biological
mechanisms often involve causation by omission, prevention, and double prevention
(that is, when a mechanism works by removing a cause, preventing a cause, or
inhibiting an inhibitor) (Schaffer 2000, 2004). Such forms of causal disconnection are
ubiquitous in the special sciences.
2.3.2 Mechanistic Accounts
Glennan (1996, 2009) sees causation (at least non-fundamental causation) as derivative
from the concept of mechanism: causal claims are claims about the existence of a
mechanism. The truth-maker for a causal claim at one level of organization is a
mechanism at a lower level. In short, mechanisms are the hidden connexion Hume
sought between cause and effect. Like the Salmon-Dowe account, Glennan’s view is
singular: particular mechanisms link particular causes and particular effects (Glennan
forthcoming)
This view has been charged with circularity: the concept of mechanism ineliminably
contains a causal element. However, Glennan replies that many accounts of causation
(such as Woodward’s 2003 account, see Section 2.3.4 below) share this flaw.
Furthermore, he argues that for at least all non-fundamental causes, a mechanisms
clearly explains how a given cause produces its effect.
Whether the analysis succeeds depends on how one deals with the resulting regress
(Craver 2007). As Glennan (2009) notes, the decomposition of causes into mechanisms
might continue infinitely, in which case there is no point arguing about which notion is
more fundamental, or the decomposition might ground out in some basic, lowest-level
causal notion that is primitive and so not analyzable into other causal mechanisms. The
latter option must confront the widely touted absence of causation in the theories of
fundamental physics (Russell 1913); at very small size scales, classical conceptions of
objects and properties no longer seem to apply, making it difficult to see what content is
left to the idea that there are mechanisms at work (see also Teller 2010; Kuhlman and
Glennan 2014).
2.3.3 Activity-Based Accounts
Still other mechanists, such as Bogen (2005, 2008a) and Machamer (Machamer 2004),
embrace an Anscombian, non-reductive view that causation should be understood in
terms of productive activities (see also the entry on G.E.M. Anscombe). Activities are
kinds of causing, such as magnetic attraction and repulsion or hydrogen bonding.
Defenders of activity-based accounts eschew the need to define the concept, relying on
science to say what activities are and what features they might have. This view is a kind
of causal minimalism (Godfrey-Smith 2010). Whether an activity occurs is not a matter
of how frequently it occurs or whether it would occur always or for the most part in the
same conditions (Bogen 2005).
This account has been criticized as vacuous because it fails to say what activities are
(Psillos 2004), to account for the relationship of causal and explanatory relevance
(Woodward 2002), and to mark an adequate distinction between activities and
correlations (Psillos 2004), though see Bogen (2005, 2008a) for a response. Glennan
(forthcoming) argues that these problems can be addressed by recognizing that
activities in a mechanism at one level depend on lower-level mechanisms. (See also
Persson 2010 for a criticism of activities based on their inability to handle cases of
polygenic effects.)
2.3.4 Counterfactual Accounts
Lastly, some new mechanists, particularly those interested in providing an account of
scientific explanation, have gravitated toward a counterfactual view of causal relevance,
and in particular, to the manipulationist view expressed in Woodward (2001, 2003)
(see, e.g., Glennan 2002; Craver 2007). The central commitment of this view is that
models of mechanisms describe variables that make a difference to the values of other
variables in the model and to the phenomenon. Difference-making in this
manipulationist sense is understood as a relationship between variables in which
interventions on cause variables can be used to change the value of effect variables (see
the entry on causation and manipulability).
Unlike the views discussed above, this way of thinking about causation provides a
ready analysis of explanatory relevance that comports well with the methods for testing
causal claims. Roughly, one variable is causally relevant to a second when there exists
an ideal intervention on the first that changes the value of the second via the change
induced on the first. The view readily accommodates omissions, preventions, and
double preventions—situations that have traditionally proven troublesome for
production-type accounts of causation. In short, the claim that C causes E requires only
that ideal interventions on C can be used to change the value of E, not that C and E are
physically connected to one another. Finally, this view provides some tools for
accommodating higher-level causal relations and the non-accidental laws of biology.
On the other hand, the counterfactual account is non-reductive (like the mechanistic
view), and it inherits challenges faced by other counterfactual views, such as pre-
emption and over-determination which are common in biological mechanisms (see the
entry on counterfactual theories of causation).

2.4 Organization
The characteristic organization of mechanisms is itself the subject of considerable
discussion.
2.4.1 Organization and Aggregativity
Wimsatt (1997) contrasts mechanistic organization with aggregation, a distinction that
mechanists have used to articulate how the parts of a mechanism are organized together
to form a whole (see Craver 2001b). Aggregate properties are properties of wholes that
are simple sums of the properties of their parts. In aggregates, the parts can be
rearranged and intersubstituted for one another without changing the property or
behavior of the whole, the whole can be taken apart and put back together without
disrupting the property or behavior of the whole, and the property of the whole changes
only linearly with the addition and removal of parts. These features of aggregates hold
because organization is irrelevant to the property of the whole. Wimsatt thus conceives
organization as non-aggregativity. He also describes it as a mechanistic form of
emergence (see Section 4.2 below).
Mechanistic emergence is ubiquitous—truly aggregative properties are rare. Thus
mechanists have tended to recognize a spectrum of organization, with aggregates at one
end and highly organized mechanisms on the other. Indeed, many mechanisms studied
by biologists involve parts and causings all across this spectrum. (For further discussion
of mechanistic emergence in relationship to other varieties, see Richardson and Stephan
2007.)
2.4.2 Varieties of Organization
Following Wimsatt, mechanists have detailed several kinds of organization
characteristic of mechanisms. A canonical list includes both spatial and temporal
organization. Spatial organization includes location, size, shape, position, and
orientation; temporal organization includes the order, rate, and duration of the
component activities. More recently, mechanists have emphasized organizational
patterns in mechanisms as a whole. Bechtel, for example, discusses how mathematical
models, and dynamical models in particular, are used to reveal complex temporal
organization in interactive mechanisms (Bechtel 2006, 2011, 2013b). Some argue that
dynamical models push beyond the limits of the mechanistic framework (e.g., Chemero
and Silbestein 2008 and, at times, Bechtel himself; see Kaplan and Bechtel 2011).
Others argue that dynamical models are, in fact, often merely descriptive (i.e., non-
explanatory models) or, alternatively, that they are used to describe the temporal
organization of mechanisms (Kaplan and Bechtel 2011; Kaplan 2012).
Mechanists have also recently borrowed from Alon’s (2006; Milo et al. 2002) work on
network motifs, repeated patterns in causal networks, to expand the vocabulary for
thinking about abstract patterns of organization (Levy 2014; Levy and Bechtel 2012).
Understanding how parts compose wholes is likely to be a growth area in the future of
the mechanistic framework. (For some other recent additions, see Kuorikoski and
Ylikoski 2013; Kuhlmann 2011; Glennan forthcoming.)
2.4.3 Modularity
Woodward’s (2001, 2002, 2011, 2014) counterfactual definition of a mechanism (which
is indirectly specified via an account of mechanistic models), as well as a descendant
elaborated by Menzies (2012), require that models of mechanisms be modular. This
means, roughly, that it should be physically possible to intervene on a putative cause
variable in a mechanism without disrupting the functional relationships among the other
variables in the mechanism. In terms of structural equation models in particular, this
means that one should be able to replace the right-hand side of an equation in the model
with a particular value (i.e., set the left-hand variable to a value) without needing to
change any of the other equations in the model. This is intended to formally capture the
sense in which mechanism is composed of separable, interacting parts. For arguments in
favor of a modularity condition on mechanistic models see Menzies (2012).
Steel (2008) appeals to a somewhat weaker form of modularity in his probabilistic
analysis of mechanisms—one that follows directly from Simon’s (1996 [1962]) idea of
nearly decomposable systems. On Simon’s view, the parts of a mechanism have more
and stronger causal relations with other components in the mechanism than they do
with items outside the mechanism. This gives mechanisms (and parts of mechanisms) a
kind of “independence” or “objecthood” defined ultimately in terms of the intensity of
interaction among components. Grush (2003), following Haugeland (1998), develops an
idea of modularity in terms of the bandwidth of interaction, where modules are high-
bandwidth in their internal interactions and low-bandwidth in their external interactions.
On this view, modularity is not an all-or-none proposition but a matter of degree;
mechanisms are only nearly decomposable. Craver (2007) argues that such a generic
notion fails to account for the relevance of different causal interactions for different
mechanistic decompositions; what counts as a part of a mechanism can only be defined
relative to some prior decision about what one takes the mechanism to be doing. For
criticisms of modularity, see Mitchell (2005) and Cartwright (2001, 2002).
2.4.4 Jointness
Fagan (2012, 2013) emphasizes the interdependent relationship between parts of a
mechanism. Components in a mechanism, she points out, often form a more complex
unit by virtue of the individual properties that unite them—their “meshing properties”;
the complex unit then figures into the mechanism’s behavior. This interdependent
relationship—jointness—is exemplified by the lock-and-key model of enzyme action.
Fagan applies this notion to research on stem cells (Fagan 2013) but argues that it is a
general feature of experimental biology (Fagan 2012).
2.4.5 Levels
Many mechanists emphasize the hierarchical organization of mechanisms and the
multilevel structure of theories in the special sciences (see especially Craver 2007, Ch.
5). Antecedents of the new mechanism focused almost exclusively on etiological,
causal relations. However, the new emphasis on mechanisms in biology and the special
sciences demanded an analysis of mechanistic relations across levels of organization.
From a mechanistic perspective, levels are not monolithic divides in the furniture of the
universe (as represented by Oppenheim and Putnam 1958), nor are they fundamentally
a matter of size or the exclusivity of causal interactions within a level (Wimsatt 1976).
Rather, levels of mechanisms are defined locally within a multilevel mechanism: one
item is at a lower level of mechanisms than another when the first item is a part of the
second and when the first item is organized (spatially, temporally, and actively) with
the other components such that together they realize the second item. Thus, the
mechanism of spatial memory has multiple levels, some of which include organs such
as the hippocampus generating a spatial map, some of which involve the cellular
interactions that underlie map generation, and some of which involve the molecular
mechanisms that underlie those cellular interactions (Craver 2007). For more on levels,
see Section 4.2 below.
2.4.6 Stable and Ephemeral Mechanisms
Finally, mechanists have found it necessary to distinguish between stable mechanisms,
which rely fundamentally upon the more or less fixed arrangement of parts and
activities, and ephemeral mechanisms, which involve a process evolving through time
without fixed spatial and temporal arrangement (Glennan 2009). The time-keeping
mechanism in a clock, for example, is a relatively stable assemblage of components in
relatively fixed locations that work the same way, with the same organizational
features, each time it works. Ephemeral mechanisms, in contrast, involve a much looser
kind of organization: items still interact in space and time, but they do not do so in
virtue of robust, stable structures. Many chemical mechanisms in a cell are like that
(Richardson and Stephan 2007). Ephemeral mechanisms are surely a primary focus of
historical sciences, such as archaeology, history, and evolutionary biology (Glennan
2009).

2.5 What Mechanisms Are Not and What Are Not


Mechanisms
The term “mechanism” has been used in many different ways to express many different
ideas. The new mechanists’ appropriation of the term is thus likely to cause unhelpful
associations, and their liberalization of the term is likely to raise worries that the notion
of mechanism has thereby been trivialized (see, e.g., Moss 2012 and Nicholson 2012).
Here, we first distinguish the new mechanism from other doctrines with which it shares
both name and family resemblance. We then discuss some things to which the new use
of the term “mechanism” does not apply.
2.5.1 What Mechanisms Are Not
New mechanists have explicitly eschewed the following associations with the term
“mechanism”:

1. Mechanisms are not necessarily deterministic. Mechanisms might be stochastic


if, for example, they are composed of stochastic activities (Bogen 2005, 2008a),
or, in a more mundane sense (i.e., one consistent with determinism), because it is
always possible for one or more factors to interfere with the working of a
mechanism; one of the parts might be broken, or an unexpected preventer might
interfere with the operation of a mechanism. The truth or falsity of determinism,
and its relevance to understanding the special sciences, is an independent issue
from the question of whether something is a mechanism.
2. Appeal to mechanisms is not necessarily reductionistic. Mechanisms are often
described as multi-level, with activities at different levels being equally essential
to how a mechanism works. Mechanistic explanations might look, up, down, or
around depending on the choice of an explanandum and the presuppositions of
the explanatory context (Bechtel 2009a). Mechanists might be reductionists or
anti-reductionists. That said, many mechanists opt for some form of explanatory
anti-reductionism, emphasizing the importance of multilevel and upward-
looking explanations, without rejecting the central ideas that motivate a broadly
physicalist world-picture (e.g., McCauley and Bechtel 2001; Craver 2007). (For
further discussion, see Theurer 2013; see also Sections 3.1 and 5 and the entry
on reductionism in biology.)
3. Not all mechanisms are machines. Machines are human-made contrivances with
each part added and organized by a designer to perform a function; biological
and social mechanisms, in contrast, are products of evolution, broadly construed
(Darden 2006), and so display ornate forms of organization in comparison with
contrivances. One machine might contain multiple mechanisms (a car, for
example, has mechanisms for braking, propulsion, playing music and climate
control). Machines are also capable of being both active and passive (a stopped
clock is still a machine); mechanisms, in contrast, have a productive aspect and
are always doing something.
4. Mechanisms are not necessarily sequential or linear. Mechanisms can have
feedback loops and cycles wherein the output of the mechanism or components
in turn influences the input of the mechanism or components in a subsequent
iteration (Bechtel 2011). Also, the interactions among components in a
mechanism need not be describable by a linear equation.
5. Mechanisms are not necessarily localizable (Bechtel and Richardson 2010
[1993]). Components of mechanisms might be widely distributed (as are many
brain mechanisms) and might violate our intuitive or tutored sense of the
boundaries of objects (as an action potential violates the cell boundary). The
assumption of localization is often an important heuristic in the search for
mechanisms; however, this heuristic often must be abandoned as the
mechanism’s organization reveals itself.
6. Mechanisms are not limited to push-pull dynamics. Descartes’ mechanism had
this feature, but (as noted above) the new mechanism explicitly liberalizes the
notion to account for other kinds of causing.
7. Mechanisms are not just fictions/metaphors. When a scientist says that there is a
mechanism that makes proteins in living organisms, she is not just using a
machine metaphor; rather, she is saying that there are in fact parts and activities
organized in living organisms such that they produce proteins.
2.5.2 What Are Not Mechanisms
One might object that there’s nothing left of mechanism once it sheds these historical
associations. One might suspect that it has been trivialized (Dupré 2013).
The idea of mechanism is a central part of the explanatory ideal of understanding the
world by learning its causal structure. The history of science contains many other
conceptions of scientific explanation and understanding that are at odds with this
commitment. Some have held that the world should be understood in terms of divine
motives. Some have held that natural phenomena should be understood teleologically.
Others have been convinced that understanding the natural world is nothing more than
being able to predict its behavior. Commitment to mechanism as a framework concept
is commitment to something distinct from and, for many, exclusive of, these alternative
conceptions. If this appears trivial, rather than a central achievement in the history of
science, it is because the mechanistic perspective now so thoroughly dominates our
scientific worldview.
Yet there are many ways of organizing phenomena besides revealing mechanisms.
Some scientists are concerned with physical structures and their spatial relations
without regard to how they work: an anatomist might be interested in the spatial
organization of parts within the body with minimal interest in how those parts articulate
together to do something. Many scientists build predictive models of systems without
any pretense that these models in fact reveal the causal structures by which the systems
work. Some scientists are concerned with taxonomy, sorting like with like without
regard to how the sorted items came about or how they work. Finally, in many areas of
science, there is a widely recognized and practically significant distinction between
knowing that C (e.g., smoking) is a cause of E (lung cancer) and knowing
how C causes E. This is not so much an ontological difference as it is a difference in the
grain with which one seeks to understand a system’s causal structure. In short, there are
many framework concepts in science, and not all of them can be assimilated to
mechanisms.
But what, the critic might push further, does not count as a mechanism? Here are some
contrast classes:

1. Entities (or objects) are not mechanisms. Mechanisms do things. If an object is


not doing anything (i.e., if there is no phenomenon), then it is not a mechanism.
2. Correlations are not mechanisms. Mechanisms explain at least many
correlations, and many correlations can be used to characterize causal or
mechanistic relations, but correlations themselves are not mechanistic. The same
can be said of mere temporal sequences of events.
3. Inferences, reasons, and arguments are not mechanisms. Though there are
mechanisms of inference and reasoning, what makes something an inference or a
reason is logical relation and not (merely) a causal relationship between premise
and conclusion.
4. Symmetries are not mechanisms. Many kind of symmetry are of fundamental
importance in different areas of physics (e.g., translational symmetries,
rotational symmetries). These are features of physical systems that are highly
general facts or assumptions, not mechanisms.
5. Fundamental laws and fundamental causal relations are not mechanisms. If a
law or causal relation is fundamental, then (by definition) there is no mechanism
for it.
6. Relations of logical and mathematical necessity are not mechanisms. Such truths
hold in all possible worlds and so do not depend for their truth on facts about the
causal structure of this world.
This is not an exhaustive list of non-mechanisms or non-mechanistic framework
concepts. Yet it demonstrates that even the liberalized concept of mechanism is neither
vacuous nor trivial.

2.6 Philosophical Work to Be Done


Much of the early new mechanical philosophy has focused on the special sciences, such
as neuroscience and molecular biology. In the years since, philosophers have extended
the mechanistic framework to other scientific disciplines, such as cell biology (Bechtel
2006), cognitive science (Bechtel 2008; Thagard 2006), neuroeconomics (Craver and
Alexandrova 2008), organic chemistry (Ramsey 2008), physics (Teller 2010),
astrophysics (Illari and Williamson 2012), behavior genetics (Tabery 2014a), and
phylogenetics (Matthews forthcoming). Philosophers continue to test the limits of this
framework, with the expectation that alternative organizing frameworks might play
central roles in other sciences. For example, a debate has emerged in the philosophy of
biology over whether or not natural selection is a mechanism (see, for example, Skipper
and Millstein 2005; Baker 2005; Barros 2008; Illari and Williamson 2012; Havstad
2011; and Matthewson and Calcott 2011). Similar debates have emerged concerning
mechanistic explanation in cognitive science (Bechtel 2008; Piccinini and Craver 2011;
Weiskopf 2011; Povich forthcoming).
One area that has received particular attention is the effort to understand computational
mechanisms. On some accounts, computational mechanisms form a proper subclass of
mechanisms that can be defined explicitly in terms of the kinds of entities, properties,
and activities involved in mechanisms in that class (Piccinini 2007; Milkowski 2013).
According to this view, computational mechanisms are mechanisms that have the
function to manipulate medium independent vehicles in accordance with a general rule
that applies to all vehicles and depends on the inputs for its application (Piccinini and
Scarantino 2011). Digital computers are distinctive in that their vehicles are digits
(Piccinini 2007). Proponents of this account hope to demarcate computing mechanisms
from non-computing mechanisms by appeal to the distinctive components proprietary to
computing mechanisms. This view contrasts both with a semantic view, according to
which computation is essentially a matter of manipulating symbols or representations,
and with perspectivalist views, according to which whether a mechanism counts as
computing is a matter of whether it is so described (Churchland 1986; Churchland and
Sejnowski 1992; Shagrir 2010).
Philosophers of the social sciences have also emphasized and debated the importance of
mechanistic knowledge (e.g., Elster 1989; for a useful review of these connections, see
Hedström and Ylikoski 2010). In that context, appeals to mechanisms are intended to
remedy the relative uninformativeness of social (or macro-level) explanations of social
phenomena (such as widespread norms, persistent inequalities, network and
institutional structures) by insisting that these explanations ultimately be grounded in
mechanistic details about individual agents and actors, their desires and motivations,
and, importantly, their relations to one another. The emphasis on relations among actors
distances this mechanistic view from methodological individualism (see the entry
on methodological individualism). Mechanists in the social sciences have also tended to
shy away from grand, overarching theories and toward more local explanations:
scientific knowledge grows by adding items to a toolbox of mechanisms and showing
how items from that toolbox can be combined to provide an explanation for a particular
phenomenon. Frederica Russo (2009) discusses a number of strategies for modeling
social mechanisms (see also Little 1991, 1998; Hedström 2005; Hedström and
Swedberg 1998).

3. Explanation: From Formal Analyses to


Material Structures
The covering-law model of explanation was a centerpiece of the logical empiricist
conception of science. According to that model, explanations are arguments showing
that the event to be explained (the explanandum event) was to have been expected on
the basis of laws of nature and the antecedent and boundary conditions (the explanans).
For advocates of the covering-law model, the philosophical problem of explanation is
thus largely a matter of analyzing the formal structure of explanatory arguments
(Hempel and Oppenheim 1948; Hempel 1965). A rainbow, for example, is explained
under the covering-law model by reference to laws of reflection and refraction
alongside conditions concerning the position of the sun and the nature of light, the
position of the raindrops, and the position of the person seeing the rainbow. The
description of the rainbow is the conclusion of a deductive argument with law
statements and descriptions of conditions as premises, and so the rainbow was to be
expected in light of knowledge of the laws and conditions.
Mechanists, in contrast, insist explanation is a matter of elucidating the causal structures
that produce, underlie, or maintain the phenomenon of interest. For mechanists, the
philosophical problem is largely about characterizing or describing the worldly or ontic
structures to which explanatory models (including arguments) must refer if they are to
count as genuinely explanatory. A rainbow, for the mechanist, is explained by situating
that phenomenon in the causal structure of the world; the explanation is an account of
how the phenomenon was produced by entities (like rain drops and eyeballs) with
particular properties (like shapes and refractive indices) that causally interact with light
propagating from the sun. Mechanists typically distinguish several ways of situating a
phenomenon within the causal structure of the world.

3.1 Etiological and Constitutive Explanations


Most mechanists recognize two main aspects of mechanistic explanation: etiological
and constitutive. Salmon (1984) describes them as two different ways of situating an
explanandum phenomenon in the causal nexus (see also Craver 2001b; Glennan 2009).
Etiological explanations reveal the causal history of the explanandum phenomenon, as
when one says a virus explains a disease. Constitutive explanations, in contrast, explain
a phenomenon by describing the mechanism that underlies it, as when one says brain
regions, muscles, and joints explain reaching.
Philosophical arguments against the covering law model often focused on its inability to
deal with causal, etiological explanations. The model failed to deliver the right verdict
on a variety of problem cases precisely because it attempted to provide an account of
explanation without any explicit mention of causation (Bromberger 1966; Salmon 1984;
Scriven 1959)
New mechanists extend these kinds of criticism to the covering law model of
intertheoretic, micro-reduction. According to the covering law model of reductive
explanation, a theory about parts reduces, and so explains, a theory about wholes when
it is possible to derive the second from the first given bridge laws to connect the two
(see Nagel 1961; Schaffner 1993).
Some mechanists argue that the covering law model of constitutive explanation has
problems analogous to those that beset the covering-law model of etiological
explanations. Action potentials cannot be explained by mere temporal sequences of
events utterly irrelevant to the phenomenon, but one can derive a description of the
action potential from descriptions of such irrelevant phenomena. Action potentials
cannot be explained by mere patterns of correlation that are not indicative of an
underlying causal relation. Irrelevant byproducts of a mechanism might be correlated
with the behavior of the mechanism, even perfectly correlated such that one could form
bridge laws between levels, but would not thereby explain the relationship. Merely
finding a neural correlate of consciousness, for example, would not, and is not taken by
anyone to, constitute an explanation of consciousness. So mechanists argue that micro-
reductive explanations must satisfy causal constraints just as surely as etiological
explanations must (Craver 2007).
The covering law model also fails to distinguish models that merely re-describe the
phenomenon in general terms from explanations that, in addition to predicting aspects
of the phenomenon, reveal the mechanisms that produce it (Craver 2006; Kaplan and
Craver 2011; but see Weiskopf 2011). For example, Snell’s law allows one to predict
how light bends when passing from one medium to another, but it does not explain why
the light bends. New mechanists also argue that the covering law model fails to
distinguish predictively adequate but fictional models from explanatory models.
Finally, mechanists argue that the intertheoretic model of reduction fails to capture an
important dimension of explanatory quality: depth. An implication of the covering law
model is that any true law statements that allow one to derive the explanandum law
(with suitable corrections and assumptions) will count as a complete explanation. Yet it
seems one can deepen an explanation by opening black boxes and revealing how things
work down to whatever level one takes as relatively fundamental for the purposes at
hand. Such criticisms suggest that the covering-law model of constitutive explanation is
too weak to capture the norms of explanation in the special sciences.
Other mechanists have argued that the covering law model is too strong. Philosophers
of biology have long argued that there are no laws of the sort the logical empiricist
described in biology and other special sciences (Beatty 1995; Mitchell 1997, 2000;
Woodward 2001). One might conclude from this that there are no explanations in
biology (Rosenberg 1985), but such a radical conclusion is difficult to square with
obvious advances in understanding, e.g., protein synthesis, action potentials, cell
signaling, and a host of other biological phenomena. In such cases, one finds that
scientists appeal to mechanisms to do the explanatory work, even in cases where
nothing resembling a law appears to be available.

3.2 Constitutive Relevance


With increased attention to constitutive explanation, mechanists realized the need for an
account of constitutive relevance, a principal for sorting relevant from irrelevant factors
in a mechanism (Craver 2007; Ylikoski 2013). A system (S) exhibiting phenomenon (ψ)
is composed of many different entities (x), with various properties, engaging in myriad
activities (φ) organized together (see Figure 1 above in Section 2). One central research
problem is to say which of these entities, activities, and organizational features
contribute to the phenomenon and which do not. In a sense, this is a challenge of
defining the boundaries of a mechanism: of saying what is and is not in the mechanism.
Three proposals have been considered. The first, the mutual manipulability account,
understands constitutive relevance in terms of the experimental manipulations used to
test interlevel relations. According to this account, if it can be shown (i) that the
putative components are contained within S, (ii) that some ideal interventions on the
putative component (x’s φ-ing) change the phenomenon (S’s ψ-ing), and (iii) that some
ideal interventions on S’s ψ-ing change x’s φ-ing, that is sufficient to establish that x is
a component in the mechanism. The notion of an ideal intervention in this account is
explicitly indebted to and a proposed extension of Woodward’s theory of causal
relevance to the constitutive domain (see Craver 2007; see also Kaplan 2012). A
concern with the mutual manipulability account, though, is that it is best an epistemic
guide to constitutive relevance, not an account of what constitutive relevance is (Couch
2011). The account offers, at best, a sufficient condition of relevance. Also, the notion
of an “ideal” intervention, borrowed from Woodward’s account of causal relevance,
cannot be applied straightforwardly to constitutive explanations. An ideal intervention
on a system cannot intervene on both the independent and the dependent variable at the
same time. However, when one intervenes to make S ψ (or prevent S from ψ-ing), one
invariably also intervenes on the components of S’s ψ-ing. And when one intervenes on
the components of S’s ψ-ing, one often intervenes on S’s ψ-ing. Because x’s φ-ing
and S’s ψ-ing are related as part to whole, they are not independent, and so require
another way to think about ideal interventions (see Baumgartner 2010; 2013; Leuridan
2011; yet see Menzies 2012; Woodward 2014.
A second proposal offers a regularity account of constitutive relevance modeled on
Mackie’s notion of understanding a cause as an INUS condition: an Insufficient but
Non-redundant part of an Unnecessary but Sufficient condition for the effect in question
(Mackie 1974; see also Cummins 1983). On this account, a constitutively relevant
component is an insufficient but non-redundant part of an unnecessary but sufficient
mechanism for a given phenomenon (Couch 2011; see also Harbecke 2010, 2014).
Allow that any number of mechanisms might suffice to bring about S’s ψ-ing; each
possible sufficient mechanism is then unnecessary for ψ-ing. Each of these mechanisms
is made of components, none of which is alone sufficient to produce the behavior of the
mechanism as a whole, but each of which is necessary in the context of the mechanism
for S to ψ. This account presupposes the idea of being necessary in context, and one
might reasonably worry about sorting accidentally correlated Xs from Xs that in fact
make a difference to S’s ψ-ing.
A third approach to constitutive relevance dispenses with the interlevel framing
enforced by the mutual manipulability account and attempts to analyze relevance using
causal notions only. According to accounts of this sort, constitutive relevance is a kind
of causal between-ness. If S’s ψ-ing is understood as an input-output relationship of
some sort, then mechanistic relevance could be understood as being a necessary link in
the causal chain between the input and the output (see Harinen forthcoming; Menzies
2012). The putatively interlevel experiments in the mutual manipulability account can
then be recast as different kinds of unilevel causal experiments. Romero (forthcoming)
provides a helpful framing of these issues and offers the novel suggestion that
putatively high-level interventions are in fact fat-handed interventions relative to their
lower-level counterparts.

3.3 Mechanisms and Models


The philosophical literature on mechanisms also overlaps with the philosophical
literature on scientific models (see the entry on models in science). Here we distinguish
mechanical models from models of mechanisms and we discuss varieties of non-
mechanical models.
3.3.1 Characterizing Completeness
Glennan (2005) proposed a definition of a mechanical model as follows:
(MM) A mechanical model consists of (i) a description of the mechanism’s behavior
(the behavioral description); and (ii) a description of the mechanism that accounts for
that behavior (the mechanical description). (446)
Such models can be represented in many different ways (see also Giere 2004). They are
evaluated in terms of their ability to predict the features of the phenomenon and in
terms of the mapping between items in the model and the entities, activities, and
organizational features in the mechanism (Glennan 2005: 17; Kaplan and Craver 2011).
Glennan emphasizes that there is no hard line between complete and incomplete
models; rather models are continually in the process of articulation and refinement.
Whether a model is complete enough is determined by pragmatic considerations.
This last point is related to Darden’s distinction between mechanism
schemas and mechanism sketches (Darden and Cain 1989; Darden 2002). In discovering
a mechanism, it is often crucial to identify gaps that have to be filled in one’s model.
While no model is ever complete in the absolute sense, some models have lacunae that
must be filled before the model is complete enough
Mechanism schemas are abstract descriptions of mechanisms that can be filled in with
details to yield a specific type or token mechanism. Thus, the schema:
DNA → RNA → Protein
can be filled in with a specific sequence of bases in DNA, its complement in RNA, and
a corresponding amino acid sequence in the protein. The arrows can be filled in,
showing how transcription and translation work. A mechanism sketch is an incomplete
representation of a mechanism that specifies some of the relevant entities, activities, and
organizational features but leaves gaps that cannot yet be filled. Black boxes, question
marks, and filler-terms (such as “activate”, “cause”, or “inhibitor”) hold the place for
some entity, activity or process yet to be discovered. The distinction between sketches
and schemas is a matter of completeness: schemas are more complete than sketches in
the sense that a sketch omits one or more stages of the mechanism that have to be
understood if one really wants to solve one’s discovery problem.
Mechanists also emphasize the distinction between a how-possibly schema and a how-
actually-enough schema (Craver and Darden 2013). A how-possibly schema describes
how entities and activities might be organized to produce a phenomenon. A how
possibly model is n hypothesis about how the mechanism works. Such models might be
true (enough) or false. A true (enough) how-possibly model is (though we may not
know it) also a how-actually (enough) model. A how-actually-enough schema describes
how entities and activities are in fact organized to produce the phenomenon. The term
“how-actually-enough” captures the idea that the requisite “accuracy” of a mechanistic
model can vary considerably from one pragmatic context to another (Weisberg 2013).A
false how possibly model is merely a how possibly model; just-so-stories are merely
how possibly models (Dray 1957; Brandon 1985). Used in this way, the term “how
possibly model” is similar to the term “hypothesis”: it is entertained as a possibility but
not necessarily endorsed.
3.3.2 Abstraction and Idealization
In contrast to mechanism schemas and sketches, some models of mechanisms work not
by describing all of the parts, causal interactions, and organizational features, but rather
by abstracting away from such potentially obfuscating details (Craver and Darden 2013;
Strevens 2008; Levy and Bechtel 2012). In such cases, idealizing assumptions can be
introduced to bring the relevant feature of the mechanism most clearly into view:
infinite populations, frictionless planes, perfect geometrical shapes are presumed in
order to strip the model of detail that does not matter for, or would only obstruct, the
intended purposes of model.
Critics of the new mechanical philosophy have pushed on the importance of abstraction
in science, drawing attention to the above discussions of completeness. The goals of
completeness and accuracy are taken to conflict with the common practice of being
satisfied with models that sacrifice detail and truth for clarity and generality (Strevens
2008; Woodward 2014). The normative distinction between a schema and a sketch, for
example, seems to suggest that science progresses by moving from incomplete to
complete models. And the distinction between how-possibly and how-actually-enough
likewise seems to privilege accuracy over other goals of modeling, which often require
distortion and falsity (see Wimsatt 2007; Weisberg 2007; Levy and Bechtel 2012;
Batterman and Rice 2014; Chirimuuta 2014; Levy 2014).
Yet mechanists can surely allow that not all models of mechanisms are mechanical
models or mechanism schemas. Often other sorts of model are useful for isolating
central aspects of a mechanism’s functioning. Dynamical models, for example, can be
used to characterize the temporal dynamics of a mechanism (Bechtel 2013a,b; Kaplan
and Bechtel 2011). Network models can be used to characterize patterns of connectivity
regardless of what units are connected and regardless of what kinds of connections one
is particularly interested in characterizing (Hunneman 2010). Minimal models can be
used to capture something fundamental about the dynamics of a broad class of
mechanisms that share no entities and activities in common (Batterman 2002). A model
of a mechanism is a model that describes a mechanism. It need not be a mechanical
model or a mechanism schema, in the above sense, to play that role.
Some mechanists reserve the term “mechanical models” for models that describe the
entities, activities, and organizational features of a system. According to Glennan’s
(2005) account, a mechanical model that leaves out some relevant features is, ipso
facto, incomplete and sketchy. One specific instantiation of this debate concerns the
explanatory force of functional models in psychology. Piccinini and Craver (2011)
argue that such models should be understood as mechanistic sketches, black-box
models to be evaluated and filled in as details about the underlying mechanism are
discovered. Black box models are incomplete in virtue of leaving out details about
underling mechanisms and that those models ultimately depend for their explanatory
force on the promise that the functional models do, in fact, correspond to how the
mechanism works. (See Weiskopf 2011 for a criticism of this account and Povich
forthcoming for a response)
One might talk about mechanistic explanation in a way that abstracts from the kind of
model used to describe the mechanism: the commitment to mechanistic explanation is
not a commitment about the form of the model but rather a commitment about what
such models must represent: namely, causal and mechanistic structures. Models are
explanatory in virtue of the fact that they represent the causal/mechanistic structures
that produce, underlie, or maintain the phenomenon. They are non-mechanistic if they
refer to some non-causal, non-mechanistic kind of relation (Salmon 1984; Craver
2014).

3.4 Philosophical Work to Be Done


To date, much of the work on mechanistic explanation has been driven by the goal of
providing a descriptively and normatively adequate theory of mechanistic explanation.
Some claim there are kinds of explanation that rely very little on a precise
understanding of the mechanistic details of a system (Woodward 2014; Weiskopf 2011)
or that work fundamentally by removing all such details from one’s model (Batterman
and Rice 2014). Resolving such debates will require being very clear about precisely
what one expects out of a philosophical theory of scientific explanation and what one
takes a scientific explanation to be (Strevens 2008; Craver 2014). Research is required
to understand the diverse representational forms that scientists use to represent
mechanisms (Burnston et al. forthcoming), and to understand the role of idealization in
mechanistic explanation (Levy and Bechtel 2012; Huneman 2010). Further work is also
required to limn the boundaries between mechanistic explanation and other putative
varieties of explanation and to say, as perspicuously as Hempel or the causal-
mechanical theory, what a model must do to count as explanatory and precisely how
good explanations are to be distinguished from bad.

4. Metaphysics of Mechanisms
In this section, we review some of the ways that the concept of mechanism has been
used in diverse areas of metaphysics. Of all the areas we have discussed, this is likely
the most in need of future development. Here we discuss the relationship between
mechanisms and laws, emergence, realization, natural kinds, and functions.

4.1 Token Mechanisms, Type Mechanisms, and Laws


In much of the early literature on mechanisms, mechanisms are contrasted explicitly
with laws of nature (Bechtel 1988; Bechtel and Abrahamsen 2005; MDC 2000). This
contrast clearly grew out of an emerging consensus in philosophy that there are few, or
perhaps no, laws of biology (see Section 3.1 above). The empirical generalizations one
finds in biology tend to be hedged by ceteris paribus clauses; whether they hold or not
depends on background conditions that might not hold and on conditions internal to the
mechanism that might fail. These generalizations, in short, are mechanistically
explicable; what necessity they have derives from a mechanism (Cummins 2000;
Glennan 1996). Mechanisms thus seem to play the role of laws in the biological
sciences: we seek mechanisms to explain, predict, and control phenomena in nature
even if mechanisms lack many of the characteristics definitive of laws in the logical
empiricist framework (such as universality, inviolable necessity, or unrestricted scope).
One specific strand of this discussion emerged from consideration of Weber’s (2005)
argument that biology is heteronymous, i.e., that it ultimately borrows its explanatory
power from the laws of physics and chemistry. Weber uses Hodgkin and Huxley’s
model of the action potential as an exemplar of the reducing biological phenomena to
physical laws (such as Ohm’s law and the Nernst equation). Craver (2006) responds
that the explanatory force of Hodgkin and Huxley’s model, in fact, requires a grasp of
the distinctly biological properties of ion channels, which properties were black-boxed
in Hodgkin and Huxley’s total current equation (see also Craver 2007; Bogen 2008b;
Weber 2008).
Yet the contrast between laws and mechanisms has not always been entirely clear.
Some, such as Bogen (2005), Machamer (2004), and Glennan (forthcoming) emphasize
that causes and mechanisms are, at base, singular, not general or universal. Leuridan
(2010), building on the work of Mitchell (2000), objects that mechanisms cannot
replace laws of nature in our conceptual understanding of explanation and the
metaphysics of science. Scientists rarely investigate token mechanisms, one might
think, but are much more interested in types. And once one starts talking about types of
mechanisms, one is back in the business of formulating general regularities about how
mechanisms work. So it would appear that the concept of mechanism cannot supplant
the work that generalization was supposed to do, but requires the idea of regularity, and
so something akin to laws, if it is to do that explanatory work (see Andersen 2011,
2012, 2014a,b; Krickel 2014). For a reply to Leuridan, see Kaiser and Craver (2013).

4.2 Mechanisms, Levels, and Emergence


Work on mechanisms has also helped to clarify the idea of levels of organization and its
relation to other forms of organization and non-mechanistic forms of emergence.
Many mechanists, following Simon (1996 [1962]), emphasize that biological systems
are hierarchically organized into near-decomposable structures: mechanisms within
mechanisms, within mechanism. Using the parable of Tempus and Hora, Simon (1962)
argued that a watchmaker who builds hierarchically decomposable watches (Tempus)
will make more watches than one who builds holistic watches (Hora). This parable led
Simon to the conclusion that evolved structures are more likely to be nearly
decomposable into hierarchically organized, more or less stable structures and sub-
structures. Some have objected that the story is misleading because evolution does not
construct organisms from scratch, piece by piece (Bechtel 2009b). Steel (2008),
building on the work of others (Schlosser and Wagner 2004), therefore attempts to
reconstruct this argument as a way of showing that evolved systems are more likely to
be modular: systems made of independently manipulable parts can quarantine the
effects of changes to specific parts, giving them added flexibility to make local changes
without causing catastrophic side-effects.
The near decomposability of mechanisms is directly related to the idea that mechanisms
span multiple levels of organization. The behavior of the whole is explained in terms of
the activities and interactions among the component parts. These activities and
interactions are themselves sustained by underlying activities and interactions among
component parts, and so on (see Bechtel and Richardson 2010 [1993]). Craver (2007)
defines levels of mechanisms in terms of a relationship between the behavior (ψ)
exhibited by a system (S) and the activity (φ) of some component part (X) of that
system. On this account, X’s φ-ing is at a lower level of mechanistic organization
than S’s ψ-ing if and only if (i) X is a part of S, and (ii) X’s φ-ing is a component in S’s
ψ-ing. In short, to say that something is at a lower mechanistic level than the
mechanism as a whole is to say that it is a working part of the mechanism. Though the
term “level” is used in many legitimate ways, levels of mechanisms seem to play a
central role in structuring the relations among many different models in contemporary
biology (e.g., between Mendelian and molecular genetics (Darden 2006), between
learning and memory and channel physiology (Craver 2007), and between population-
level variation and developmental mechanisms (Tabery 2009, 2014a)).
One implication of this view of levels, combined with certain familiar assumptions
about causal relations, is that there can be no causal relationships between items at
different levels of mechanisms. There can be causal relationships between things of
different sizes, and there can be causal relationships between things described in very
different vocabularies; but (again, conjoined with certain assumptions about the
temporal asymmetry of cause and effect and the independence of cause and effect) there
cannot be causal relationships between the behavior of a mechanism and the activities
of the parts that jointly constitute that behavior. Claims about interlevel causation,
which are ubiquitous in the scientific literature, are best understood either as targeting a
different sense of levels or, concerning levels of mechanisms, as expressing hybrid
claims combining constitutive claims about the relationship between the behavior of the
mechanism as a whole and the activities of its parts, and causal claims concerning
relationships between things not related as part and whole (Craver and Bechtel 2007).
For alternative interpretations of levels, see Fehr 2004; Leuridan 2011; Thalos 2013;
Eronen 2013; 2015; Baumgartner and Gebhardter forthcoming; Romero forthcoming.
For reflections on the metaphysical status of higher-level phenomena and higher-level
causes, see Baumgartner 2010; Glennan 2010 a, b; Hoffman-Kolss 2014, as well as the
entries on causation and manipulability, physicalism, and scientific reduction.
As noted above, the fact that phenomena at higher levels of mechanisms depend upon
the organization of component parts entails that the properties/activities of wholes are
not simple sums of the properties/activities of the parts. Levels of mechanisms can thus
be contrasted with levels of mere aggregation. Because the whole is greater (in this
sense) than the sum of the parts, some (such as Wimsatt) have found it appropriate to
describe this as a kind of emergence. Mechanistic (or organizational) emergence thus
understood is ubiquitous and banal but extremely important for understanding how
scientists explain things.
Also familiar is epistemic emergence, the inability to predict the properties or behaviors
of wholes from properties and behaviors of the parts. Epistemic emergence can arise as
a result of ignorance, such as failing to recognize a relevant variable, or from failing to
know how different variables interact in complex networks. It might also result from
limitations on human cognitive abilities or in current-generation representational tools
(Bedau 1997; Boogerd 2005; Richardson and Stephan 2007). The practical necessity of
studying mechanisms by decomposing them into component parts raises the epistemic
challenge of putting the parts back together again in a way that actually works (Bechtel
2013a).
The mechanists’ emphasis on mechanistic/organizational and epistemic emergence
contrasts with their desire to distance themselves from spooky emergence (Richardson
and Stephan 2007). Spooky emergence would involve the appearance of new properties
with no sufficient basis in mechanisms. It is not clear that emergent properties are
properly said to be properties of the necessary mechanisms; and it is not clear in what
since the emergent property is “emergent” rather than a fundamental feature of the
causal structure of the world.
In short, such forms of emergence are altogether distinct from, and so gain no
plausibility from, verbal association with organizational/mechanistic and epistemic
emergence.

4.3 Mechanisms and Realization


Because the framework concept of a mechanism is so useful for thinking about levels
and explanation in the sciences, some scholars have sought in the notion of mechanism
a way of fleshing out the ontological relationship of realization.
According to the “flat view” (Gillett 2002) realization is a relationship between
different properties of one and the same thing (Kim 1998; Shapiro 2000; Shoemaker
2003, 2007; Polger 2007). The subset view, which holds that a property P1 (e.g., mean
kinetic energy of the gas) realizes property P2 (e.g., temperature of the gas) when the
causal powers distinctive of P2 (temperature) are a subset of the causal powers
distinctive of P2 (mean kinetic energy), is an example of the flat view. P1 and P2 are
both attributed to the same thing, the gas (Gillett 2002, 2003). The dimensioned view
describes realization as a relationship holding between the properties of wholes and the
properties of the parts and their organization. This view of realization comports with the
explanatory aims of the special sciences and fits nicely with the evidential base on
which interlevel claims are grounded (see Aizawa and Gillett 2011). Gillett has since
expanded this notion to handle the realization of objects, properties, and processes
(Gillett 2013); for criticism and alternatives, see Polger 2010; Melnyk 2003; Melnyk
2010)

4.4 Mechanisms and Natural Kinds


Mechanistic theories of natural kinds develop out of Boyd’s homeostatic property
cluster (HPC) view. The HPC view is a theory of natural kinds designed to work in
domains with high individual variability. The HPC view is offered as a third way
between essentialism and nominalism about kinds in the special sciences (Boyd 1991,
1997, 1999; Kornblith 1993; Wilson 1999, 2005; see also the entry on natural kinds).
According to this view, a natural kind is characterized by i) a cluster of properties that
regularly co-occur, and ii) a similarity generating mechanism that explains why the
properties in (i) tend to co-occur. In short, kinds are property clusters explained by
mechanisms.
This view of natural kinds has been deployed to argue for taxonomic revision in, for
example, the biology of human emotion (Griffiths 1997), the structure of concepts
(Machery 2009), and the taxonomy of psychiatry (Kendler, Zachar, and Craver 2010;
see Craver 2009): a putatively single kind is split into multiple kinds because it is
discovered that distinct properties in a property cluster are explained by distinct
mechanisms. This view of kinds can also be used to make sense of kinds that are
historically transient and, in some ways, the product of human attitudes and so socially
constructed in this straightforward sense; perhaps race is like this (see Kuorikoski and
Pöyhönin 2012; Khalidi 2013).

4.5 Mechanisms and Functions


Emphasis on the importance of mechanisms is historically associated with the rejection
of teleology and formal causes (e.g., Westfall 1971). Yet contemporary biology and
many other special sciences, despite the widespread acceptance of the mechanism
framework concept, continue to make use of the concept of function, a teleological
notion (see the entry on teleological notions in biology). How is the notion of function
at play in contemporary science related to the concept of mechanism? Craver (2001a),
following Cummins (1975), argues that functional description is a perspectival means
of situating some part within a higher-level mechanism. According to this view,
teleology is not a feature of the world so much as it is imposed upon it by an intentional
describer (see also Machamer 1977). Garson (2011, 2012), following Wimsatt (1972b),
Wright (1973) and Neander (1991a,b), argues that functions are effects of an item that
are part of the etiological explanation (through selection, learning, or reinforcement) for
why the item is present; as such, functions are reduced to causal histories. In a third
view, Maley and Piccinini (forthcoming) argue that the (teleological) function of an
item is its contribution to the goals of organisms, which may be objective or subjective
goals. Like Garson, Maley and Piccinini hold that functions are objective (i.e., not
observer relative). Unlike Garson, however, they are not grounded in the etiology of the
item but in their current contribution to survival or reproduction (the objective goals of
organisms) or to what the organism itself desires (the subjective goals of organisms).

4.6 Philosophical Work to Be Done


What must the world be like for this mechanistic perspective to be accurate? Clearly,
there are many ways of answering this question from different metaphysical starting
assumptions. And clearly, many metaphysical starting assumptions rule this world
picture out as illegitimate. The clearest path forward, it would seem, is to work out
precisely what one must be committed to in holding that the world is composed of a
hierarchy of mechanisms and precisely what of that can be recovered on the basis of
different starting assumptions.
That said, not all applications of the mechanism framework require a fully articulated
metaphysics. Work on discovery and explanation might proceed perfectly well without
embracing any particular metaphysical world picture. Philosophers with different
interests (discovery, explanation, testing, reduction, emergence, and so) are likely to
elaborate the concept in different ways. There is every reason to doubt, that the idea of
mechanism can be given a one-size fits all metaphysical analysis that will adequately
address the diverse philosophical ends to which the concept is being deployed.

5. Relations between Scientific Disciplines:


From Theory Reduction to Mechanism
Integration
5.1 Theory Reduction
According to Nagel (1961), reduction is a species of covering-law explanation: one
theory is reduced to another when it is possible to identify the theoretical terms of the
first with those of the second and to literally derive the first from the second. On the
assumption that scientific disciplines and theories correspond to one another, reduction
serves as a model of interdisciplinary integration as well. Mechanist’s objections to the
covering-law model of constitutive explanation (i.e., micro-reduction) are discussed
above (see Section 3.1); here the focus is rather on how distinct disciplines of science
are integrated.
On the Nagel view, reduction is an interlevel relationship. It is also a relationship
between theories. Theories about phenomena at a higher level (e.g., gases, lightning,
and life) are reduced to (i.e., derived from) theories about phenomena at lower levels
(e.g., molecules, electrons, and physiological systems). Finally, the relationship is
formally specified and has little to do with either the content of the theories or the
material structures those theories describe. From the mechanistic perspective, each of
these features of the Nagel model is problematic.

5.2 Mechanism Integration


First, mechanists criticize the idea that reduction should be understood primarily as a
relationship between theories. In integrating their results, scientists are not simply
building theories simpliciter; they are building theories about mechanisms. Mechanisms
can perhaps be described using formal accounts of theories—perhaps they can be
axiomatized in predicate logic or reconstructed as set theoretic predicates. But such
formal accounts of the structures of scientific theories gloss over the mechanistic
structures crucial for understanding how these theories are constructed and evaluated
(Craver 2001b).
Mechanists also challenge the idea that disciplines are related by way of the
relationship between their theories. Darden and Maull (1977) argued that disciplinary
fields often integrate their findings through the construction of interfield theories,
appealing to diverse material relationships between items in the different fields’
domains: relations such as cause and effect, part and whole, or structure and function.
Darden and Maull did not offer a general account of interfield theories, but Bechtel
presciently suggested that such theories often take the form of descriptions of
mechanisms (Bechtel 1988: 101–102).
The mechanistic approach also has been claimed to have many advantages over
reduction for thinking about interlevel forms of interdisciplinary integration. First, it
provides a straightforward way to interpret the talk of levels (see
Sections 2.4.5 and 4.2). Second, it offers significantly more insight into what interlevel
integration is, into the evidential constraints by which interlevel bridges are evaluated,
and into the forces driving the co-evolution of work at different levels. Constraints on
the parts, their causal interactions, and their spatial, temporal, and hierarchical
organization all help to flesh out an interlevel integration. Finally, mechanists
repeatedly recognize the need to not only look down to the constitutive mechanisms
responsible for a given phenomenon (emphasized by classical reduction models), but
also to look up and around to the context within which the phenomenon is embedded:
interlevel integration is an effort to see how phenomena at many different levels are
related to one another (Bechtel 2009a; Craver 2007).
Mechanists have developed several extended examples of the many forms of
mechanism integration pursued in mechanistic research programs. Darden (2005), for
example, suggests that philosophers in the grip of classical reduction fundamentally
misunderstood the relationship between Mendelian and molecular genetics. While
reductionists see it as an instance of interlevel explanation, she argues, it is in fact a
case in which different scientists worked on different parts of a mechanism that are
etiologically (not constitutively) related to one another. Mendelian genetics did not
reduce to molecular biology; rather, classical geneticists and molecular biologists
integrated their work by focusing on different working entities in the sequentially
operating chromosomal and molecular hereditary mechanisms. Examples have also
been drawn from the discovery of the mechanisms of protein synthesis (Darden 2006)
and cell biology (Bechtel 2006). Craver (2007) uses examples from the neuroscience of
memory to explore how multilevel integration does and ought to proceed. In each case,
the search for mechanisms serves as an abstract scaffold onto and around which the
findings of diverse scientists converge.
5.2.1 Integrative Pluralism
The mechanistic perspective tends to emphasize integrative pluralism in scientific
research (Mitchell 2003, 2009). The goal is not to explain the less fundamental in terms
of the more fundamental in a step-wise relating of monolithic theories at one level to
monolithic theories at another. Rather, such scientific achievements are collaborative
and piecemeal, adding incremental constraints to an emerging picture of how a
mechanism works both at a level and across levels. The many scientific disciplines that
investigate a phenomenon pluralistically co-exist and co-inform one another by
integratively contributing to the etiological, constitutive, and contextual mechanistic
explanations of that phenomenon (Bechtel 2009a; Tabery 2014a).

5.3 Philosophical Work to Be Done


The Nagel model of theory reduction offers a clear vision of the “unity of science” (see
the entry on unity of science). According to the model, the unity among scientific
disciplines is achieved by reducing theories of higher-level disciplines to the theories of
lower-level disciplines. Integration, on that vision, is understood as progress toward a
grand, unified body of scientific knowledge. For mechanists, in contrast, integration is
piecemeal, local, and pluralistic. What sort of unity could such an “integration” sustain?
This question plays out in a back-and-forth between Longino and Tabery concerning
disciplinary relationships in the behavioral sciences. Tabery argues that disciplines as
disparate as neurobiology and quantitative genetics could pluralistically co-exist and co-
inform one another’s causal explanations of complex behaviors by way of mechanism
integration. Longino counters that Tabery’s pluralism is only a “moderate” sort because
the push for integration ultimately is a push for unification (Longino 2013, 2014;
Tabery 2014a,b). Sullivan (2009) also challenges the push for mechanism integration;
she argues that there are significant barriers to the kind of integration mechanists
envision. Different laboratories use different experimental protocols to study what they
assume to be the same phenomenon; however, these different protocols often in fact
target different phenomena, so the integration achieved by combining results is only
illusory. These discussions are symptomatic of more general philosophical questions
faced by mechanists: How are mechanism integrations actually achieved (as opposed to
just asserted)? And what is the relationship between mechanism integration and
unification? The new mechanical philosophy stands to benefit from future efforts to
situate mechanistic integration into more general philosophical views of integration and
pluralism.

6. Discovery: From A-ha Moments to Discovery


Strategies
6.1 Discovery via A-ha Moments
What can philosophers say about scientific discovery? Many logical empiricists had a
simple answer: Nothing. According to Popper, for example, philosophers can illuminate
the epistemology of testing, but they can say nothing of substance about how scientists
generate the ideas to be tested (Popper 1959). Such “A-ha!” moments of creativity are
in the province of psychology, not philosophy. Reichenbach distinguished the context
of discovery from the context of justification (the “context distinction”) (Reichenbach
1938; but see the entry on Hans Reichenbach for an alternate interpretation of this
distinction). The process of scientific discovery was thus largely off limits to
philosophers.
Not all philosophers of science agreed. Hanson, for example, articulated a logic of
discovery involving abductive inferences from anomalous data to new hypotheses
designed to account for them (Hanson 1958). Others focused on methodologies of
discovery that could either allow one to rationally reconstruct why something was
discoverable at a given time (Nickels 1985) or to explain why a new hypothesis is
considered promising and worthy of further investigation (Schaffner 1993). Early
contributions to the new mechanical philosophy followed this path and characterized
investigative strategies scientists use to discover mechanisms (see the entry on scientific
discovery).

6.2 Discovery via Strategies


Bechtel and Richardson’s Discovering Complexity (2010 [1993]) is organized around a
flowchart representing choice-points in the discovery of a mechanism. The process of
searching for mechanisms begins with a provisional characterization of the
phenomenon. Then follow strategies of localizing the mechanism within the system,
and decomposing the phenomenon into distinct sub-functions. Localization of function
involves determining which of these sub-functions of the system is performed by which
parts. Bechtel and Richardson further characterize the use of excitatory and inhibitory
experiments to obtain these kinds of information. Bechtel and Abrahamsen (2013) add a
subsequent stage, in which scientists recompose what they have learned about the
functional parts by putting them back together to produce the phenomenon in question
(perhaps using simulations).
Darden also emphasized mechanisms as an important framework concept in scientific
discovery (Darden 1980, 1982, 1986, 1991). In the discovery of protein synthesis
(jointly investigated by molecular biologists and biochemists in the 1950s and 1960s),
scientists didn’t simply have an “A-ha” moment. Rather, they deployed strategies for
revealing how a mechanism works (Darden 2006; Craver and Darden 2013). Darden
characterizes the process of mechanism discovery as an “extended, piecemeal process
with hypotheses undergoing iterative refinement”; that process occurs via the
construction, evaluation, and revision of mechanism schemas in light of observational
and experimental constraints (Darden 2006: 272).
Darden’s construction strategies are strategies for generating new hypotheses about a
mechanism. In addition to decomposition and localization, Darden shows that scientists
often borrow a schema type from another area of science, as when selection-type
mechanisms were borrowed to understand how the immune system works, or assemble
a mechanism from known modules of functional activity (modular sub-assembly), as is
common in biochemistry and molecular biology. Sometimes, scientists know one part
of the mechanism and attempt to work forward or backward through to the other parts
and activities. In the discovery of the mechanism of protein synthesis, for example,
molecular biologists worked forward from the structure of DNA to figure out what
molecules could interact with it (forward chaining), and biochemists worked backward
from proteins to figure out what chemical reactions would be necessary to create them
(backward chaining). They met in the middle at RNA. Protein synthesis is now
understood to involve transcribing DNA into RNA and then translating RNA into
proteins. Far from being philosophically inscrutable, Darden points out that scientists
used what they knew about the working entities and activities in the mechanism to infer
what could come next or before in the mechanism of protein synthesis (Darden 2006;
see also the entry on molecular biology).
Evaluation strategies, for Darden, involve constraint-based reasoning to limn the
contours of the space of possible mechanisms for a given phenomenon. Often scientists
reason about how a mechanism works by building off basic findings concerning the
spatial and temporal organization of its parts. Harvey, for example, reasoned his way to
the circulation of the blood by considering the locations of the valves of the veins and
their orientation with respect to the heart. These organizational constraints, and many
others, combined to narrow the space of possible mechanisms to a small region
containing a model in which the blood completes a circuit of the body (Craver and
Darden 2013).
Darden and Craver also discuss experimental strategies for learning how a mechanism
works. These strategies reveal how different entities and activities in a mechanism act,
interact, and are organized together. For example, one might intervene to remove a
putative component to see if and how the mechanism functions in its absence
(inhibitory experiments). Or one might stimulate that component to see if it can drive
the mechanism or modulate its behavior. Or one might activate a mechanism by placing
it in the precipitating conditions for the phenomenon and observe how the entity or
activity changes as the mechanism works. Craver (2002) discusses these under the
heading of “interlevel experiments” (see also Harinen forthcoming). Craver and Darden
(2013) also discuss more complex kinds of experiments for learning what sort of entity
or activity contributes to a process and for learning more complex features of a
mechanism’s organization.
Datteri (2009; Datteri and Tamburrini 2007), explores the use of robotic simulations for
the purposes of testing mechanisms. They discuss both how assumptions are built into
robotic models and how experiments can be designed to reveal how mechanisms work.
This work extends the mechanistic framework into the area of bio-robotics and reveals
a set of strategies distinct from those explored in Darden’s work.
Rather than focusing on the process by which mechanism schemas are constructed,
evaluated, and revised, Steele focuses on the question of how one extrapolates from a
sample population or a model organism to the structure of a mechanism in the target.
Will a treatment proven to suppress tumors in mice (a model organism) also suppress
tumors in humans (the target population)? After developing a probabilistic account of
mechanisms, Steele considers how researchers get around what he calls the
extrapolator’s circle: determining
how we could know that the model and the target are similar in causally relevant
respects without already knowing the causal relationship in the target. (Steel 2008: 78)
Steel breaks the extrapolator’s circle by developing a mechanisms-based extrapolation
strategy—the strategy of comparative process tracing. Once a mechanism for some
phenomenon has been elucidated in a model (such as a particular process of
carcinogenesis in rats), scientists (toxicologists in this case) then compare key stages
(particularly downstream stages) of the model with the stages in the target, paying
particular attention to points in the process where differences are most likely to arise.
The greater the similarities of the entities, activities, and organization of the
mechanisms in both populations, the stronger is the basis for extrapolation; the greater
the differences, the weaker the basis (but see Howick et al. 2013; see also the sections
on extrapolation in the entries on molecular biology and experiment in biology).

6.3 Mechanistic Evidence in Medical Discoveries


Discovery in medicine is another domain where the mechanical philosophy has been
applied. Thagard draws on the case of H. pylori as a cause of ulcers to provide an
account of how investigating mechanisms contributes to scientific discovery.
Thagard draws attention to both statistical evidence that suggests ulcers are somehow
associated with H. pylori, as well as mechanistic evidence that can explain how the
agent of infection could persist in a hostile environment long enough to cause an ulcer.
More recently, philosophers interested in evidence-based medicine have probed the
relationship between these two types of evidence in the health sciences. Russo and
Williamson argue that both types of evidence are necessary to justify causal inference;
the correlational evidence establishes that there is a difference-making relation between
some cause and some effect, while the mechanistic evidence establishes how exactly
the cause produces its effect—the “Russo-Williamson Thesis” (Russo and Williamson
2007). Philosophers have since refined the Russo-Williamson Thesis, pointing out, for
instance, that “type of evidence” could refer to different methodologies for gathering
evidence or to different objects of evidence. Difference-making methodologies include
observational studies and randomized controlled trials, while mechanistic
methodologies include interventionist experiments such as those described above;
likewise, the object of evidence could be the evidence of an associated difference or it
could be the evidence concerning the mechanism linking the cause and effect (Illari
2011; see also Campaner 2011). Evidence-based medicine hierarchies, which rank
different kinds of evidence in terms of its epistemic strength, tend to prioritize evidence
from difference-making methodologies (such as randomized controlled trials and meta-
analyses) over mechanistic evidence; in reply, these philosophers argue that the
different types of evidence are on a par (each with its own strengths and weaknesses)
and advocate for integrating difference-making and mechanistic evidence, a sentiment
which aligns with the emphasis on mechanism integration discussed in Section
5.2 above (Clarke et al. 2013, 2014).

6.4 Philosophical Work to Be Done


Many mechanists have explored the strategies that scientists use in discovery. Bechtel
and Richardson attended to decomposition and localization; Darden and Craver
highlighted forward and backward chaining; Russo and Williamson emphasized
drawing on both difference-making and mechanistic evidence. These strategies were
found in specific, experimental sciences, such as neuroscience and molecular biology.
So one task for philosophers moving forward is to assess whether or not similar
strategies exist in other sciences, especially those that operate outside the traditional
laboratory, both in the human sciences (such as sociology and economics) and in the
physical sciences (such as cosmology).
We also expect tremendous development to come from bridging the gap between the
qualitative accounts of mechanisms and mechanistic explanation developed in the new
mechanism and quantitative theories of discovery from the discipline of machine
learning and causal modeling (Spirtes et al. 2000; Pearl 2009). The latter offer tools to
mine correlational data for causal dependencies. Such tools might escape more
qualitative, historical approaches and might, in fact, go beyond the common strategies
that scientists traditionally use. Such tools also offer a means to assess discovery
strategies by exploring the conditions under which they succeed and fail and the
efficiency with which they deliver verdicts on causal hypotheses.

7. Conclusion
The new mechanical philosophy and, more generally, attention to the framework
concept of “mechanism” has expanded rapidly over the last two decades bringing with
it new orientations toward a wide range of issues in the philosophy of science. Yet it is
clear that many of the major topics are only beginning to develop, leaving a lot of work
for scholars to elaborate the basic commitments of this framework and to consider what
it means to do science outside of that framework. The near future is likely to see
continued discussion of the implications and limits of this framework for thinking about
science and scientific practice.

Bibliography
 Aizawa, K. and C. Gillett, 2011, “The Autonomy of Psychology in the Age of
Neuroscience”, in Causality in the Sciences, in Illari et al. 2011: 202–23.
 Allen, G.E., 2005, “Mechanism, Vitalism and Organicism in Late Nineteenth
and Twentieth Century Biology: The Importance of Historical Context”, in
Craver and Darden 2005: 261–283.
 Alon, U., 2006, An Introduction to Systems Biology, Boca Raton: Chapman and
Hall/CRC Press.
 Andersen, Holly, 2011, “Mechanisms, Laws, and Regularities”, Philosophy of
Science, 78(2): 325–331.
 –––, 2012, “The Case for Regularity in Mechanistic Causal
Explanation”, Synthese, 189: 415–432.
 –––, 2014a, “A Field Guide to Mechanisms: Part I”, Philosophy Compass, 4:
274–283.
 –––, 2014b, “A Field Guide to Mechanisms: Part II”, Philosophy Compass, 4:
283–297.
 Baker [now Byron], J.M., 2005, “Adaptive Speciation: The Role of Natural
Selection in Mechanisms of Geographic and Non-geographic Speciation”, in
Craver and Darden 2005: 303–326.
 Barros, D.B., 2008, “Natural Selection as a Mechanism”, Philosophy of Science,
75: 306–322.
 Batterman, R., 2002, “Asymptotics and the Role of Minimal Models”, The
British Journal for the Philosophy of Science, 53: 21–38.
 Batterman, R. and C. Rice, 2014, “Minimal Model Explanations”, Philosophy of
Science, 81: 349–376.
 Baumgartner, M., 2010, “Interventionism and Epiphenomenalism”, Canadian
Journal of Philosophy, 40: 359–384.
 –––, 2013, “Rendering Interventionism and Non-Reductive Physicalism
Compatible”, Dialectica, 67: 1–27.
 Baumgartner, M. and A. Gebharter, forthcoming, “Constitutive Relevance,
Mutual Manipulability, and Fat-Handedness”, The British Journal for the
Philosophy of Science. doi:10.1093/bjps/axv003 [Baumgartner and Gebharter
forthcoming available online]
 Beatty, J., 1995, “The Evolutionary Contingency Thesis”, in James G. Lennox
and Gereon Wolters (eds), Concepts, Theories, and Rationality in the Biological
Sciences, Pittsburgh, PA: University of Pittsburgh Press, pp. 45–81.
 Bechtel, W., 1988, Philosophy of science: An overview for cognitive science,
Hillsdale, NJ: Erlbaum. Italian translation Filosofia della scienza e scienza
cognitiva, Gius. Laterza & Figli, 1995. Second edition in preparation. [some of
Bechtel 1988 available online]
 –––, 2006, Discovering Cell Mechanisms: The Creation of Modern Cell Biology,
Cambridge: Cambridge University Press. [some of Bechtel 2006 available
online]
 –––, 2008, Mental Mechanisms: Philosophical Perspectives on Cognitive
Neuroscience, London: Routledge.
 –––, 2009a, “Looking Down, Around, and Up: Mechanistic Explanation in
Psychology”, Philosophical Psychology, 22: 543–564. [Bechtel 2009a available
online]
 –––, 2009b, “Explanation: Mechanism, Modularity, and Situated Cognition”, in
P. Robbins and M. Aydede (eds). Cambridge handbook of situated cognition,
Cambridge: Cambridge University Press, pp. 155–170. [Bechtel 2009b available
online]
 –––, 2011, “Mechanism and Biological Explanation”, Philosophy of Science, 78:
533–557.
 –––, 2013a, “Addressing the Vitalist’s Challenge to Mechanistic Science:
Dynamic Mechanistic Explanation”, in S. Normandin & C.T. Wolfe
(eds), Vitalism and the Scientific Image in Post-Enlightenment Life Science,
1800–2010, Dordrecht: Springer, pp. 345–370.
 –––, 2013b, “From Molecules to Behavior and the Clinic: Integration in
Chronobiology”, Studies in History and Philosophy of Biological and
Biomedical Sciences, 44: 493–502.
 Bechtel, W. and A. Abrahamsen, 2005, “Explanation: A Mechanistic
Alternative”, Studies in History and Philosophy of the Biological and
Biomedical Sciences, 36: 421–441. [Bechtel and Abrahamsen 2005 available
online]
 –––, 2013, “Thinking Dynamically about Biological Mechanisms: Networks of
Coupled Oscillators”, Foundations of Science, 18: 707–723. [Bechtel and
Abrahamsen 2013 available online]
 Bechtel, W. and R.C. Richardson, 2010 [1993], Discovering Complexity:
Decomposition and Localization as Strategies in Scientific Research, Second
Edition. Cambridge, MA: MIT Press/Bradford Books.
 Bedau, M., 1997, “Weak Emergence”, Philosophical Perspectives, 11: 375–399.
 Beebee, H., C. Hitchcock, and P. Menzies, (eds), 2010, The Oxford Handbook of
Causation, Oxford: Oxford University Press.
 Boas, M., 1952, “The Establishment of the Mechanical Philosophy”, Osiris, 10:
412–541.
 Bogen, J., 2005, “Regularities and Causality; Generalizations and Causal
Explanations”, in Craver and Darden 2005: 397–420.
 –––, 2008a, “Causally Productive Activities”, Studies in History and Philosophy
of Science, 39: 112–123.
 –––, 2008b, “The Hodgkin-Huxley Equations and the Concrete Model:
Comments on Craver, Schaffner, and Weber”, Philosophy of Science, 75: 1034–
1046.
 Boogerd, F.C., F.J. Bruggeman, R.C. Richardson A. Stephan, and H. V.
Westerhoff, 2005, “Emergence and Its Place in Nature: A Case Study of
Biochemical Networks”, Synthese, 145: 131–164.
 Boyd, R., 1991, “Realism, Anti-Foundationalism and the Enthusiasm for Natural
Kinds”, Philosophical Studies, 61: 127–148.
 –––, 1997, “Kinds as the ‘Workmanship of Men’: Realism, Constructivism, and
Natural Kinds”, in J. Nida-Rumelin (ed.), Rationality, Realism, Revision:
Proceedings of the 3rd International Congress of the Society for Analytical
Philosophy, New York: Walter de Gruyter, pp. 52–89.
 –––, 1999, “Homeostasis, Species, and Higher Taxa”, in R.A. Wilson
(ed.), Species, MIT Press: Cambridge, 141-185.
 Brandon, R., 1985, “Grene on Mechanism and Reductionism: More Than Just a
Side Issue”, in Peter Asquith and Philip Kitcher (eds), PSA 1984, v. 2. East
Lansing, MI: Philosophy of Science Association, pp. 345–353.
 Bromberger, S., 1966, “Why Questions”, in R.G. Colodny (ed.), Mind and
Cosmos, Pittsburgh: University of Pittsburgh Press, pp. 86–111.
 Burnston, D.C., B. Sheredos, A. Abrahamsen, and W. Bechtel, in press,
“Scientists’ Use of Diagrams in Developing Mechanistic Explanations: A Case
Study from Chronobiology”, Pragmatics and Cognition.
 Campaner, R., 2011, “Understanding Mechanisms in the Health
Sciences”, Theoretical Medicine and Bioethics, 32: 5–17.
 Cartwright, N.D., 1989, Nature’s Capacities and their Measurement, New York:
Oxford University Press. [Cartwright 1989 available online]
 –––, 1999, The Dappled World: A Study of the Boundaries of Science,
Cambridge: Cambridge University Press.
 –––, 2001, “Modularity: It Can—and Generally Does—Fail”, in Stochastic
Dependence and Causality, D. Costantini, M.C. Galavotti, and P. Suppes (eds),
Stanford: CSLI Publications, pp 65–84.
 –––, 2002, “Against Modularity, the Causal Markov Condition and Any Link
Between the Two: Comments on Hausman and Woodward”, British Journal for
the Philosophy of Science, 53: 411–453.
 Chemero A. and M. Silberstein, 2008, “After the Philosophy of Mind: Replacing
Scholasticism with Science”, Philosophy of Science, 75: 1–27.
 Churchland, P.S., 1986, Neurophilosophy: Toward a Unified Science of the
Mind/Brain, Cambridge, MA: MIT Press.
 Churchland, P.S. and T.J. Sejnowski, 1992, The Computational Brain,
Cambridge, MA: MIT Press.
 Chirimuuta, M., 2014, “Minimal Models and Canonical Neural Computations:
The Distinctness of Computational Explanation in Neuroscience”. Synthese,
191: 127–153.
 Clarke, B., D. Gillies, P. Illari, F. Russo, and J. Williamson, 2013, “The
Evidence that Evidence-Based Medicine Omits”, Preventive Medicine, 57: 745–
747.
 –––, 2014, “Mechanisms and the Evidence Hierarchy”, Topoi, 33: 339–360.
 Couch, M.B., 2011, “Mechanisms and Constitutive Relevance”, Synthese, 183:
375–88.
 Craver, C.F., 2001a, “Role Functions, Mechanisms and Hierarchy”, Philosophy
of Science, 68: 31–55.
 –––, 2001b, “Structures of Scientific Theories”, in P.K. Machamer and M.
Silberstein (eds), Blackwell Guide to the Philosophy of Science, Blackwell:
Oxford, pp. 55–79.
 –––, 2006, “When Mechanistic Models Explain”, Synthese, 153: 355–376.
 –––, 2007, Explaining the Brain: Mechanisms and the Mosaic Unity of
Neuroscience, Oxford: Clarendon Press.
 –––, 2009, “Mechanisms and Natural Kinds”, Philosophical Psychology, 22:
575–594.
 –––, 2013, “Functions and Mechanisms: A Perspectivalist Account”, in P.
Huneman (ed.), Functions: Selection and Mechanisms, Dordrecht: Springer, pp.
133–158.
 –––, 2014, “The Ontic Conception of Scientific Explanation”, in Andreas
Hütteman and Marie Kaiser (eds), Explanation in the Special Sciences: The
Cases of Biology and History, Dordrecht: Springer, pp. 27–52.
 Craver, C.F. and A. Alexandrova, 2008, “No Revolution Necessary: Neural
Mechanisms for Economics”, Economics and Philosophy, 24: 381–406.
 Craver, C.F. and W.M. Bechtel, 2007, “Top-down Causation without Top-down
Causes”, Biology and Philosophy, 22: 547–563.
 Craver, C.F. and L. Darden, 2013, In Search of Mechanisms: Discoveries Across
the Life Sciences, Chicago: University of Chicago Press.
 Cummins, R., 1975, “Functional Analysis”, Journal of Philosophy, 72: 741–764.
 –––, 1983, The Nature of Psychological Explanation, Cambridge, MA:
Bradford/MIT Press.
 –––, 2000, “‘How Does It Work?’ Vs. ‘What Are The Laws?’ Two Conceptions
of Psychological Explanation”, in F. Keil and R. Wilson (eds), Explanation and
Cognition, Cambridge, MA: MIT Press, pp. 117–144.
 Darden, L., 1986, “Reasoning in Theory Construction: Analogies, Interfield
Connections, and Levels of Organization”, in P. Weingartner and G. Dorn
(eds), Foundations of Biology, Vienna: Holder-Pichler-Tempsky, pp. 99–107.
 –––, 1991, Theory Change in Science: Strategies from Mendelian Genetics, New
York: Oxford University Press.
 –––, 2002, “Strategies for Discovering Mechanisms: Schema Instantiation,
Modular Subassembly, Forward/Backward Chaining”, Philosophy of Science,
69: S354–S365.
 –––, 2005, “Relations Among Fields: Mendelian, Cytological and Molecular
Mechanisms”, Studies in History and Philosophy of Biological and Biomedical
Sciences, 36: 349–371.
 –––, 2006, Reasoning in Biological Discoveries: Mechanism, Interfield
Relations, and Anomaly Resolution, New York: Cambridge University Press.
 Darden, L. and J. Cain, 1989, “Selection Type Theories”, Philosophy of Science,
56:106–129.
 Darden, L. and N. Maull, 1977, “Interfield Theories”, Philosophy of Science, 44:
43–64.
 Datteri, E., 2009, “Simulation Experiments in Bionics: A Regulative
Methodological Perspective”, Biology and Philosophy, 24: 301–324.
 Datteri, E. and G. Tamburrini, 2007, “Biorobotic Experiments for the Discovery
of Biological Mechanisms”, Philosophy of Science, 74: 409–430.
 DesAutels, L., 2011, “Against Regular and Irregular Characterizations of
Mechanisms”, Philosophy of Science, 78: 914–925.
 Des Chene, D., 2001, Spirits & Clocks: Machine & Organism in Descartes,
Ithaca, NY: Cornell University Press.
 –––, 2005, “Mechanisms of Life in the Seventeenth Century: Borelli, Perrault,
Régis”, Studies in the History and Philosophy of Biological and Biomedical
Sciences, 36: 245–260.
 Dijksterhuis, E.J., 1961, The Mechanization of the World Picture, New York:
Oxford University Press.
 Dowe, P., 1992, “Wesley Salmon’s Process Theory of Causality and the
Conserved Quantity Theory”, Philosophy of Science, 59: 195–216.
 –––, 2011, “The Causal-Process-Model Theory of Mechanisms”, in Illari et al.
2011: 865–879.
 Dray, W., 1957, Laws and Explanation in History, London: Oxford University
Press.
 Dupré, J., 1993, The Disorder of Things: Metaphysical Foundations of the
Disunity of Science, Cambridge, MA: Harvard University Press.
 –––, 2013, “Living Causes”, Proceedings of the Aristotelian Society, 87: 19–35.
 Elster, J., 1989, Nuts and Bolts for the Social Sciences, Cambridge, UK:
Cambridge University Press.
 Eronen, M.I., 2013, “No Levels, No Problems: Downward Causation in
Neuroscience”, Philosophy of Science, 80: 1042–1052.
 –––, 2015, “ Levels of Organization: A Deflationary Account”, Biology and
Philosophy, 30: 39–58.
 Fagan, M.B., 2012, “The Joint Account of Mechanistic
Explanation”, Philosophy of Science, 79: 448–472.
 –––, 2013, Philosophy of Stem Cell Biology: Knowledge in Flesh and Blood,
London: Palgrave MacMillan.
 Fehr, C., 2004, “Feminism and Science: Mechanism without
Weductionism”, National Women’s Studies Association Journal, 16: 136–156.
 Fodor, J., 1968, Psychological Explanation, New York: Random House.
 Garson, J., 2011, “Selected Effects Functions and Causal Role Functions in the
Brain: The Case for an Etiological Approach to Neuroscience”, Biology &
Philosophy, 26: 547–565.
 –––, 2012, “Function, Selection, and Construction in the Brain”, Synthese, 189:
451–481.
 –––, 2013, “The Functional Sense of Mechanism”, Philosophy of Science, 80:
317–333.
 Giere, R.N., 2004, “How Models Are Used to Represent Reality”, Philosophy of
Science, 71: 742–752.
 Gillett, C., 2002, “The Dimensions of Realization: A Critique of the Standard
View”, Analysis, 62: 316–323.
 –––, 2003, “The Metaphysics of Realization, Multiple Realizability, and the
Special Sciences”, The Journal of Philosophy, 100: 591–603.
 –––, 2013, “Constitution, and Multiple Constitution, in the Sciences: Using the
Neuron to Construct a Starting Framework”, Minds and Machines, 23: 309–37.
 Glennan, S.S., 1996, “Mechanisms and The Nature of Causation”, Erkenntnis,
44: 49–71.
 –––, 1997, “Capacities, Universality and Singularity”, Philosophy of Science,
64: 605–626.
 –––, 2002, “Rethinking Mechanistic Explanation”, Philosophy of Science, 69:
S342–S353.
 –––, 2005, “Modeling Mechanisms”, Studies in the History and Philosophy of
the Biological and Biomedical Sciences, 36: 375–388.
 –––, 2009, “Productivity, Relevance and Natural Selection”, Biology and
Philosophy, 24: 325–339.
 –––, 2010a, “Mechanisms”, in Beebee et al. 2010: 315–325.
 –––, 2010b, “Mechanisms, Causes, and the Layered Model of the
World”, Philosophy and Phenomenological Research, 81: 362–381.
 –––, forthcoming, The New Mechanical Philosophy, Oxford: Oxford University
Press.
 Godfrey-Smith, P., 2010, “Causal Pluralism”, in Beebee et al. 2010: 326–337.
[Godfrey-Smith 2010 available online]
 Gould, S.J., 1990, Wonderful Life: Burgess Shale and the Nature of History,
New York: W.W. Norton and Company.
 Griffiths, P.E., 1997, What Emotions Really Are: The Problem of Psychological
Categories, Chicago: University of Chicago Press.
 Grush, R., 2003, “In Defense of Some ‘Cartesian’ Assumptions Concerning the
Brain and its Operation”, Biology and Philosophy, 18: 53–93.
 Hanson, N.R., 1958, Patterns of Discovery, Cambridge: Cambridge University
Press.
 Harbecke, J., 2010, “Mechanistic Constitution in Neurobiological
Explanations”, International Studies in the Philosophy of Science, 24: 267–285.
 –––, 2014 [onlinefirst], “Regularity Constitution and the Location of
Mechanistic Levels”, Foundations of Science, 19. doi: 10.1007/s10699–014–
9371–1
 Harinen, T., forthcoming, “Mutual Manipulability and Causal
Betweenness”, Synthese, doi:10.1007/s11229-014-0564-5
 Haugeland, J., 1998, Having Thought, Cambridge, MA: Harvard University
Press
 Havstad, J.C., 2011, “Discussion: Problems for Natural Selection as a
Mechanism”, Philosophy of Science, 78: 512–523.
 Hedström, P., 2005, Dissecting the Social: On the Principles of Analytical
Sociology, Cambridge: Cambridge University Press.
 Hedström, P. and R. Swedberg, 1998, Social Mechanisms: An Analytical
Approach to Social Theory, Cambridge: Cambridge University Press.
 Hedström, P. and P. Ylikoski, 2010, “Causal Mechanisms in the Social
Sciences”, Annual Review of Sociology, 36: 49–67.
 Heil, J. and A. Mele, 1993, Mental Causation, Oxford: Clarendon Press.
 Hempel, C.G., 1965, Aspects of Scientific Explanation and Other Essays in the
Philosophy of Science, New York: Free Press.
 Hempel, C.G. and P. Oppenheim, 1948, “Studies in the Logic of
Explanation”, Philosophy of Science, 15: 135–175.
 Hitchcock, C.R., 1995, “Discussion: Salmon on Explanatory
Relevance”, Philosophy of Science, 62: 304–20.
 Hoffmann-Kloss, V., 2014, “Interventionism and Higher-Level
Causation”, International Studies in the Philosophy of Science, 28: 49–64.
 Howick, J.P. Glasziou, and J.K. Aronson, 2013, “Problems with Using
Mechanisms to Solve the Problem of Extrapolation”, Theoretical Medicine and
Bioethics, 34: 275–291.
 Huneman, P., 2010, “Topological Explanations and Robustness in Biological
Sciences”, Synthese, 177: 213–245.
 Illari, P.M., 2011, “Disambiguating the Russo-Williamson
Thesis”, International Studies in the Philosophy of Science, 25: 139–157.
 Illari, P.M., F. Russo, and J. Williamson (eds), 2011, Causality in the Sciences,
Oxford: Oxford University Press.
 Illari, P.M. and J. Williamson, 2010, “Function and Organization: Comparing
the Mechanisms of Protein Synthesis and Natural Selection”, Studies in the
History and Philosophy of the Biological and Biomedical Sciences, 41: 279–291.
 –––, 2012, “What is a Mechanism?: Thinking about Mechanisms Across the
Sciences”, European Journal for Philosophy of Science, 2: 119–135.
 Kaiser, M. and C.F. Craver, 2013, “Mechanisms and Laws: Clarifying the
Debate”, in Hsiang-Ke Chao, Szu-Ting Chen and Roberta L. Millstein
(eds), Mechanism and Causality in Biology and Economics, Dordrecth:
Springer, pp. 125–145. [Kaiser and Craver 2013 available online]
 Kaplan, D.M., 2012, “How to Demarcate the Boundaries of Cognition”, Biology
and Philosophy, 27: 545–570.
 Kaplan, D.M. and W. Bechtel, W, 2011, “Dynamical Models: An Alternative or
Complement to Mechanistic Explanations”, Topics, in Cognitive Science, 3:,
438–444. [Kaplan and Bechtel 2011 available online]
 Kaplan, D.M. and C.F. Craver, 2011, “The Explanatory Force of Dynamical
Models”, Philosophy of Science, 78: 601–627.
 Kauffman, S.A., 1971, “Articulation of Parts Explanation in Biology and the
Rational Search for Them”, in Roger C. Buck and Robert S. Cohen (eds), PSA
1970, Boston Studies in the Philosophy of Science, volume 8. Dordrecht: Reidel,
pp. 257–272. Reprinted in Marjorie Grene and Everett Mendelsohn (eds),
1976, Topics in the Philosophy of Biology, Dordrecht: Reidel, pp. 245–263.
 Kendler, K., P. Zachar, and C.F. Craver, 2010, “What Kinds of Things are
Psychiatric Disorders?”, Psychological Medicine, 41: 1143–1150.
 Khalidi, M.A., 2013, Natural Categories and Human Kinds, Cambridge:
Cambridge University Press.
 Kim, J., 1998, Mind in a Physical World, Cambridge, MA: The MIT Press.
 Krickel, B., 2014, The Metaphysics of Mechanism, PhD Dissertation. Humboldt-
Universität zu Berlin.
 Kornblith, H., 1993, Inductive Inference and Its Natural Ground, Cambridge,
MA: MIT Press.
 Kuhlmann, M., 2011, “Mechanisms in Dynamically Complex Systems”, in Illari
et al. 2011: 880–906.
 Kuhlmann, M. and S. Glennan, 2014, “On the Relation between Quantum
Mechanical and Neo-Mechanistic Ontologies and Explanatory
Strategies”, European Journal for Philosophy of Science, 4: 337–359.
 Kuhn, T.S., 1962, The Structure of Scientific Revolutions, Chicago: University
of Chicago Press.
 Kuorikoski, J. and S. Pöyhönen, 2012, “Looping Kinds and Social
Mechanisms”, Sociological Theory, 30: 187–205.
 Kuorikoski, J. and P.K. Ylikoski, 2013, EPSA11 Perspectives and Foundational
Problems in Philosophy of Science, V. Karakostas & D. Dieks (eds), Heidelberg:
Springer, pp. 69–80.
 Lakatos, I., 1977, The Methodology of Scientific Research Programmes:
Philosophical Papers, vol. 1. Cambridge; Cambridge University Press.
 Laudan, L., 1977, Progress and Its Problems, Berkeley: University of California
Press.
 Leuridan, B., 2010, “Can Mechanisms Really Replace Laws of
Nature?”, Philosophy of Science, 77: 317–340.
 –––, 2011, “Three Problems for the Mutual Manipulability Account of
Constitutive Relevance in Mechanisms”, The British Journal for the Philosophy
of Science, 63: 399–427.
 Levy, A., 2013, “Three Kinds of ‘New Mechanism’”, Biology and Philosophy,
28: 99–114.
 –––, 2014, “What was Hodgkin and Huxley’s Achievement?”, British Journal
for the Philosophy of Science, 65: 469–492. [Levy 2014 available online]
 Levy, A. and W. Bechtel, 2012, “Abstraction and the Organization of
Mechanisms”, Philosophy of Science, 80: 241–261.
 Little, D., 1991, Varieties of Social Explanation: An Introduction to the
Philosophy of Social Science, Boulder, CO: Westview.
 –––, 1998, Microfoundations, Method, and Causation: On the Philosophy of the
Social Science, New Brunswick, NJ: Transaction.
 Longino, H., 2013, Studying Human Behavior: How Scientists Investigate
Aggression and Sexuality, Chicago: University of Chicago Press.
 –––, 2014, “Pluralism, Social Action, and the Causal Space of Human
Behavior”, Metascience, 23: 443–459.
 Machamer, P.K., 1977, “Teleology and Selective Processes”, in R. Colodny
(ed.), Logic, Laws, and Life: Some Philosophical Complications, Pittsburgh, PA:
University of Pittsburgh Press, pp. 129–142.
 Machamer, P., 2004, “Activities and Causation: The Metaphysics and
Epistemology of Mechanisms”, International Studies in the Philosophy of
Science, 18: 27–39.
 Machamer, P.K., L. Darden, and C.F. Craver, 2000 [MDC], “Thinking about
Mechanisms”, Philosophy of Science, 67:1–25.
 Machery, E., 2009, Doing Without Concepts, New York: Oxford University
Press.
 Mackie, J.L., 1974, The Cement of the Universe, Oxford: Clarendon Press.
 Maley, C.J. and G. Piccinini, forthcoming, “A Unified Mechanistic Account of
Teleological Functions for Psychology and Neuroscience”, in David Kaplan
(ed.), Integrating Mind and Brain Science: Mechanistic Perspectives and
Beyond, Oxford: Oxford University Press.
 Matthews, L., forthcoming, “Embedded Mechanisms and
Phylogenetics”, Philosophy of Science.
 Matthewson, J. and B. Calcott, 2011, “Mechanistic Models of Population-Level
Phenomena”, Biology and Philosophy, 26: 737–756.
 McCauley, R.N. and W. Bechtel, 2001, “Explanatory Pluralism and the
Heuristic Identity Theory”, Theory and Psychology, 11: 736–760.
 Melnyk, A., 2003, A Physicalist Manifesto: Thoroughly Modern Materialism,
Cambridge: Cambridge University Press.
 Melnyk, A., 2010, “Comments on Sydney Shoemaker’s Physical
Realization”, Philosophical Studies, 148: 113–123.
 Menzies, P., 2012, “The Causal Structure of Mechanisms”, Studies in History
and Philosophy of Biological and Biomedical Sciences, 43: 796–805.
 Milkowski, M., 2013, Explaining the Computational Mind, Cambridge, MA:
MIT Press.
 Millstein, R.L., 2006, “Natural Selection as a Population-Level Causal
Process”, The British Journal for the Philosophy of Science, 57: 627 -653.
 Milo, R, S. Shen-Orr, S. Itzkovitz, N. Kashtan, D. Chklovskii, and U. Alon,
2002, “Network Motifs: Simple Building Blocks of Complex
Networks”, Science, 298(5594): 824–27.
 Mitchell, S.D., 1997, “Pragmatic Laws”, Philosophy of Science, 64: S468–S479.
 –––, 2000, “Dimensions of Scientific Law”, Philosophy of Science 67: 242–265.
 –––, 2003, Biological Complexity and Integrative Pluralism, Cambridge:
Cambridge University Press.
 –––, 2005, “Modularity: More than a Buzzword? Essay Review, Biological
Theory, 1: 98–101.
 –––, 2009, Unsimple Truths: Science, Complexity, and Policy, Chicago:
University of Chicago Press.
 Moss, L., 2012, “Is the Philosophy of Mechanism Philosophy Enough?” Studies
in the History and Philosophy of Science: C, 43: 164–72.
 Nagel, E., 1961, The Structure of Science: Problems in the Logic of Scientific
Explanation, New York: Harcourt, Brace and World.
 Neander, K., 1991a, “The Teleological Notion of ‘Function’”, Australasian
Journal of Philosophy, 69: 454–468.
 –––, 1991b, “Functions as Selected Effects: the Conceptual Analyst’s
Defence”, Philosophy of Science, 58: 168–184.
 Nicholson, D.J., 2012, “The Concept of Mechanism in Biology”, Studies in
History and Philosophy of Biological and Biomedical Sciences, 43: 152–163.
 Nickels, T., 1985, “Beyond Divorce: Current Status of the Discovery
Debate”, Philosophy of Science, 52: 177–206.
 Oppenheim, P. and H. Putnam, 1958, “The Unity of Science as a Working
Hypothesis”, in H. Feigl, M. Scriven and G. Maxwell (eds), Concepts, Theories
and the Mind-Body Problem, (Minnesota Studies in the Philosophy of Science,
v. 2), Minneapolis: University of Minnesota Press, pp. 3–36.
 Pearl, J., 2009, Causality: Models, Reasoning, and Inference, second edition,
New York: Cambridge University Press.
 Persson, Johannes, 2010, “Activity-Based Accounts of Mechanism and the
Threat of Polygenic Effects”, Erkenntnis, 72: 135–149.
 Piccinini, G., 2007, “Computing Mechanisms”, Philosophy of Science, 74: 501–
526.
 Piccinini, G. and C.F. Craver, 2011, “Integrating Psychology and Neuroscience:
Functional Analyses as Mechanism Sketches”, Synthese, 183: 283–3
 Piccinini, G. and A. Scarantino, 2011, “Information Processing, Computation,
and Cognition”, Journal of Biological Physics, 37: 1–38.
 Polger, T.W., 2007, “Realization and the Metaphysics of Mind”, Australasian
Journal of Philosophy, 85: 233–59.
 –––, 2010, “Mechanisms and Explanatory Realization Relations”, Synthese, 177:
193–212.
 Popper, K., 1959, The Logic of Scientific Discovery, London: Hutchinson and
Co.
 Povich, M., forthcoming, “Mechanisms and Model-based fMRI”, Philosophy of
Science.
 Psillos, S., 2004, “A Glimpse of the Secret Connexion: Harmonizing Mechanism
with Counterfactuals”, Perspectives on Science, 12: 288–319.
 Ramsey, J.L., 2008, “Mechanisms and Their Explanatory Challenges in Organic
Chemistry”, Philosophy of Science, 75: 970–982.
 Reichenbach, H., 1938, Experience and Prediction: An Analysis of the
Foundations and the Structure of Knowledge, Chicago: The University of
Chicago Press.
 Richardson, R.C. and A. Stephan, 2007, “Emergence”, Biological Theory, 2: 91–
96.
 Roe, S., 2014, The Salmon-Roe Approach to Mechanistic Explanations, Doctoral
Dissertation, UC Davis.
 Romero, F., forthcoming, “Why there isn’t interlevel causation in
mechanisms”, Synthese. doi:10.1007/s11229-015-0718-0
 Rosenberg, A., 1985, The Structure of Biological Science, Cambridge:
Cambridge University Press.
 Russell, B., 1913, “On the Notion of Cause”, Proceedings of the Aristotelian
Society, 13: 1–26.
 Russo, F., 2009, Causality and Causal Modeling in the Social Sciences:
Measuring Variation, Dordrecht: Springer.
 Russo, F. and J. Williamson, 2007, “Interpreting Causality in the Health
Sciences”, International Studies in the Philosophy of Science, 21: 157–170.
 Salmon, W.C., 1984, Scientific Explanation and the Causal Structure of the
World, Princeton, NJ: Princeton University Press.
 –––, 1989, Four Decades of Scientific Explanation, Minneapolis, MN:
University of Minnesota Press.
 –––, 1994, “Causality Without Counterfactuals”, Philosophy of Science, 61:
297–312.
 –––, 1997, “Causality and Explanation: A Reply to Two Critiques”, Philosophy
of Science, 64: 461–77.
 Schaffer, J., 2000, “Causation by Disconnection”, Philosophy of Science, 67:
285–300.
 –––, 2004, “Causes Need Not be Physically Connected to Their Effects: The
Case for Negative Causation”, in C. Hitchcock (ed.), Contemporary Debates in
Philosophy of Science, Malden, MA: Blackwell Publishing. Pp. 197–216.
 Schaffner, K., 1993, Discovery and Explanation in Biology and Medicine,
Chicago, IL: University of Chicago Press.
 Schlosser, G. and G.P. Wagner, 2004, Modularity in Development and
Evolution, Chicago, IL: University of Chicago Press.
 Scriven, M., 1959, “Explanation and Prediction in Evolutionary
Theory”, Science, 130: 477–482.
 Shagrir, O., 2010, “Computation, San Diego Style”, Philosophy of Science, 77:
862–874.
 Shapiro, L., 2000, “Multiple Realizations”, Journal of Philosophy, 97: 635–654.
 Shoemaker, S., 2003, “Realization, Micro-Realization, and
Coincidence”, Philosophy and Phenomenological Research, 67: 1–23.
 –––, 2007, Physical Realization, Oxford: Oxford University Press.
 Simon, H.A., 1996 [1962], The Sciences of the Artificial, 3rd ed., Cambridge,
MA: MIT Press.
 Skipper Jr., R.A. and R.L. Millstein, 2005, “Thinking about Evolutionary
Mechanisms: Natural Selection”, in Craver and Darden 2005: 327–347.
 Spirtes, P., C. Glymour, and R. Scheines, 2000, Causation, Prediction, and
Search, 2nd ed., Cambridge, MA: MIT Press.
 Steel, D.P., 2008, Across the Boundaries: Extrapolation in Biology and Social
Science, New York: Oxford University Press.
 Strevens, M., 2008, Depth: An Account of Scientific Explanation, Cambridge,
MA: Harvard University Press.
 Sullivan, Jacqueline, 2009, “The Multiplicity of Experimental Protocols: A
Challenge to Reductionist and Non-Reductionist Models of the Unity of
Neuroscience”, Synthese, 167: 511–539.
 Tabery, J., 2004, “Synthesizing Activities and Interactions in the Concept of a
Mechanism”, Philosophy of Science, 71: 1–15.
 –––, 2009, “Difference Mechanisms: Explaining Variation with
Mechanisms”, Biology and Philosophy, 24: 645–664.
 –––, 2014a, Beyond Versus: The Struggle to Understand the Interaction of
Nature and Nurture, Cambridge, MA: The MIT Press.
 –––, 2014b, “Pluralism, Social Action, and the Causal Space of Human
Behavior”, Metascience, 23: 443–459.
 Teller, P., 2010, “Mechanism, Reduction, and Emergence in Two Stories of the
Human Epistemic Enterprise”, Erkenntnis, 73: 413–425.
 Thagard, P., 1998, “Explaining Disease: Causes, Correlations, and
Mechanisms”, Minds and Machines, 8: 61–78.
 –––, 2000, How Scientists Explain Disease, Princeton, NJ: Princeton University
Press.
 –––, 2006, Hot Thought: Mechanisms and Applications of Emotional Cognition,
Cambridge, MA: The MIT Press.
 Thalos, M., 2013, Without Hierarchy: An Essay on the Scale Freedom of the
Universe, Oxford: Oxford University Press.
 Theurer, K.L., 2013, “Compositional Explanatory Relations and Mechanistic
Reduction”, Minds and Machines, 23: 287–307.
 Weber, M., 2005, Philosophy of Experimental Biology, Cambridge: Cambridge
University Press.
 –––, 2008, “Causes without Mechanisms: Experimental Regularities, Physical
Laws, and Neuroscientific Explanation”, Philosophy of Science, 75: 995–1007.
 Weisberg, M., 2007, “Three Kinds of Idealization”, The Journal of Philosophy,
104: 639–659.
 –––, 2013, Simulation and Similarity: Using Models to Understand the World,
Oxford: Oxford University Press.
 Weiskopf, D., 2011, “Models and Mechanisms in Psychological
Explanation”, Synthese, 183: 313–338.
 Westfall, R., 1971, The Construction of Modern Science, Cambridge: Cambridge
University Press.
 –––, 2005, Genes and the Agents of Life, Cambridge: Cambridge University
Press.
 Wimsatt, W.C., 1972a, “Complexity and Organization”, in Kenneth F. Schaffner
and Robert S. Cohen (eds), PSA 1972, Proceedings of the Philosophy of Science
Association, Dordrecht: Reidel, pp. 67–86.
 –––, 1972b, “Teleology and the Logical Structure of Function
Statements”, Studies in the History and Philosophy of Science, 3: 1–80.
 –––, 1976, “Reductionism, Levels of Organization, and the Mind–Body
Problem”, in G. Globus, I. Savodnik, and G. Maxwell (eds), Consciousness and
the Brain, New York: Plenum Press, pp. 199–267.
 –––, 1997, “Aggregativity: Reductive Heuristics for Finding
Emergence”, Philosophy of Science, 64: S372–S384.
 –––, 2007, Re-Engineering Philosophy for Limited Beings: Piecewise
Approximations to Reality, Cambridge, MA: Harvard University Press.
 Woodward, J., 1989, “The Causal/Mechanical Model of Explanation”, in P.
Kitcher and W.C. Salmon (eds), Scientific Explanation, (Minnesota Studies in
the Philosophy of Science 13), pp. 357–383.
 –––, 2001, “Law and Explanation in Biology: Invariance Is the Kind of Stability
that Matters”, Philosophy of Science, 68: 1–20.
 –––, 2002, “What Is a Mechanism?: A Counterfactual Account”, Philosophy of
Science, 69: S366–S377.
 –––, 2003, Making Things Happen: A Theory of Causal Explanation, New York:
Oxford University Press.
 –––, 2011, “Mechanisms Revisited”, Synthese, 183: 409–427.
 –––, 2014, “A Functional Account of Causation; or, A Defense of the
Legitimacy of Causal Thinking by Reference to the Only Standard That Matters
—Usefulness (as Opposed to Metaphysics or Agreement with Intuitive
Judgment)”, Philosophy of Science, 81: 691–713.
 Wright, L., 1973, “Functions”, Philosophical Review, 82: 139–168.
 Ylikoski, P., 2013, “Causal and Constitutive Explanation
Compared”, Erkenntnis, 78: 277–97.

You might also like