Vous êtes sur la page 1sur 8

c.

Ontological Reduction
In light of the above objections against the epistemological reduction of chemistry, there are philosophers who
have investigated whether it is possible to support chemistry’s ontological reduction to quantum mechanics in a
manner that is consistent with the failure of chemistry’s epistemological reduction. Most notable is Le Poidevin,
who formulated a detailed account for the ontological reduction of chemical properties which does not depend
on the success of an epistemic reduction of chemistry to quantum mechanics. In fact, Le Poidevin accepts that
chemistry has not been epistemically reduced to quantum mechanics and argues that, despite this, it can be
argued that chemical elements are ontologically reduced to physical properties. He claims that the argument for
the ontological reduction of chemical elements can be generalised to all chemical properties in the following
manner:@

Chemical properties reduce to those properties variation in which is discrete, and combinations of which
constitute the series of physically possible chemical properties. (Le Poidevin 2005: 132)

In particular, he takes that the discreteness of chemical elements as specified via the periodic table supports a
combinatorial argument for their ontological reduction. According to this argument, ‘a finite number of
fundamental entities combine together to give a discrete set of composite elements’ (Scerri 2007a: 929).

Le Poidevin’s argument is based on two premises. The first is the ‘combinatorial criterion for ontological
reduction’, which states that

a property type F is ontologically reducible to a more fundamental property type G if the possibility of
something’s being F is constituted by a recombination of actual instances of G, but the possibility of something’s
being G is not contributed by a recombination of actual instances of F. (Le Poidevin 2005: 132)

The second premise concerns the ‘discreteness of chemical ordering’: ‘between any two elements there is a finite
number of physically possible intermediate elements’ (Le Poidevin 2005: 132).

According to Le Poidevin, the combinatorial criterion for the ontological reduction of chemical properties is
preferable to existing physicalist accounts regarding the ontological reduction of special science properties
because it overcomes two insurmountable problems of physicalism. The first problem is the ‘vacuity problem’,
according to which physicalism is in danger of becoming a trivial thesis depending on what one takes to be
included in the domain of physics (Le Poidevin 2005: 121-122). The second problem is the ‘asymmetry
problem’, according to which the supervenience relation, as postulated by physicalism, does not necessitate an
asymmetric relation between higher and lower-level properties (Le Poidevin 2005: 122).

Scerri, Hendry and Needham are sympathetic towards Le Poidevin’s argument of the ontological reduction of
chemical elements (Scerri 2007b: 76; Hendry and Needham 2007: 340). As Hendry and Needham state, the
combinatorial argument establishes that ‘the discreteness of the elements is explained by the nomologically
required discrete variation in a physical quantity, namely nuclear charge’ (Hendry and Nedham 2007: 34).
However, all of them take that there are certain problematic features in Le Poidevin’s account.

First, the argument is allegedly not well-supported for all chemical properties. Scerri doubts that the
combinatorial argument can be generalised so as to apply to all chemical properties because, unlike chemical
elements, most chemical properties are not discreet (such as the solubility and acidity of elements) (Scerri
2007a: 929). Similarly, Hendry and Needham argue that the combinatorial argument is only investigated with
respect to chemical elements, thus disregarding a large part of chemistry. This is a central shortcoming of Le
Poidevin’s account because there are particular features of chemistry and of quantum mechanics which are
often regarded as posing unique challenges to chemistry’s reduction to quantum mechanics. For example, the
structure of molecules is a chemical property which some argue is not in principle derivable by quantum
mechanics (Hendry and Needham 2007: 341-342). This is regarded problematic for the reduction of chemistry
to quantum mechanics, whether epistemic or ontological. Another issue is how chemistry describes the rate of
chemical reactions. Kinetic theory and thermodynamics play a fundamental role in explaining and describing
the rate of chemical reactions, and thus need to be considered in the context of chemistry’s relation to quantum
mechanics (Hendry and Needham 2007: 343-344). These are problems that concern particular chemical
properties and which need to be tackled if any account of (ontological) reduction is to be well-supported for all
chemical properties.

Secondly, Scerri takes that Le Poidevin’s attempt to circumvent any talk about the epistemic reduction between
the two relevant theories is illusory. The latter takes that a ‘periodic ordering is a classification rather than a
theory’, thus rendering his account of ontological reduction ‘theory-neutral’ (Le Poidevin 2005: 131). However,
Scerri disagrees on this point as he takes reference to the periodic table to inevitably require the investigation of
how chemistry and quantum mechanics are epistemically related (Scerri 2007a: 929). Hendry and Needham
take this point a step further by suggesting that reference to a theory cannot be avoided when specifying the
micro-constituents of chemical elements (Hendry and Needham 2007: 344). In fact, they argue that there is ‘a
close evidential connection’ between epistemological and ontological reduction; one cannot entirely avoid the
investigation of inter-theoretic reduction when seeking to provide sufficient empirical support to ontological
reduction (Hendry and Needham 2007: 351).

Another objection to Le Poidevin’s account is that the combinatorial argument, even if correct, does not succeed
in establishing the ontological reduction of chemistry to physics. The asymmetric relation that Le Poidevin
allegedly establishes via his combinatorial argument establishes ‘only an asymmetrical relationship between the
(actual) physical and the (merely possible) chemical’ (Hendry and Needham 2007: 349). Given this, such a
relation does not preclude the possibility of chemical properties having novel causal powers, thus rendering Le
Poidevin’s account consistent with non-reductive (metaphysical) accounts (such as emergentist accounts)
(Hendry and Needham 2007: 350).

Hendry also offers independent support to the claim that chemistry fails to ontologically reduce to quantum
mechanics, outside his critique of Le Poidevin’s account. Specifically, he assumes that ontological reduction
involves the acceptance of the causal completeness of physics (Hendry 2010b: 187). Given this, it follows that
ontological reduction is committed to the claim that only physical entities, properties, and so forth possess
novel causal powers (Hendry 2010b: 187). Based on this understanding of ontological reduction, he argues that
what he calls the ‘symmetry problem’ undermines the tenability of ontological reduction. The symmetry
problem arises from the fact that, for any atom or molecule, the arbitrary solutions of the Schrödinger equation
are spherically symmetrical (Hendry 2010b: 186). This comes in contrast to the asymmetry exhibited by
polyatomic molecules and which chemistry invokes in order to explain many of their chemical properties, such
as the acidic behaviour and boiling point of the hydrogen chloride molecule (Hendry 2010b: 186). The
symmetry problem allegedly challenges the ontological reduction of chemistry because it undermines the
tenability of the causal completeness of physics, namely the principle that every physical effect has a physical
cause (Hendry 2010b: 187). This is because

•quantum mechanics is consistent with the view that the asymmetry of molecules ‘is not conferred by the
molecule’s physical basis according to physical laws’ (Hendry 2010b: 187); and
•the symmetry problem ‘removes much of the empirical support that is claimed for’ the causal
completeness of physics (Hendry 2010b: 187).
Lastly, it should be noted that there are positions which argue for the ontological autonomy of chemistry in a
manner that is implicitly or explicitly incompatible with the ontological reduction of chemistry to quantum
mechanics. This includes Lombardi and Labarca (2005) and Schummer (2014b) (see subsection 5b).
d. Alternative Forms of Reduction
Despite the arguments against chemistry’s epistemological and ontological reduction to quantum mechanics,
there are philosophers who attempt to establish reduction. For example, Hettema states that ‘the widespread
rejection of reduction by philosophers of chemistry might have been premature’ (Hettema 2012b: 147). Hettema
argues that, contrary to how Nagel’s account of reduction has been understood and argued against in the
philosophy of chemistry, Nagel was in fact not so strict about the requirements for reduction (Hettema 2014:
193; see also Hettema 2012a). In light of this, Hettema proposes ‘a suitable paraphrase of the Nagelian
reduction programme’ which is ‘reinforced by a modern notion of both connectibility and derivability’ (Hettema
2017: 24) (italics are in the original text). Hettema’s position is a reductive account which advocates the
existence of autonomous areas. Characterising Hettema’s account as a form of reduction is justified given the
quotes just mentioned. Nevertheless, it should be noted that Hettema often refers to his proposal as one that
advocates a form of unity (for example Hettema 2012b; 2017). In order to explicate his proposal, Hettema
analyses the development of the reaction rate theory and presents, among other things, Eyring’s theory of
absolute reaction rates (2017: 71-81; see also Hettema 2012b) (see also subsection 5a).
Needham has also investigated reduction and identified those aspects of Nagelian reduction which should be
amended for a more convincing defence of chemistry’s reduction to physics to be achieved. As Needham states:

Chemistry is, perhaps, so entwined with physics that what would be left after removal of physics is but a pale
shadow of modern chemistry. It is, perhaps, not even clear what the removal of physics from chemistry would
amount to. (Needham 2010: 163)

Needham identifies the weaknesses of Nagelian reduction and examines whether historical developments in
chemistry and physics are consonant with how reduction tells us that two theories are related (2010: 170).
Based on such an analysis, he argues that it is possible to understand Nagelian reduction in a way that permits
and takes into account the use of approximations in science (Needham 2010: 168-169).

4. Emergence in Chemistry
The emergence of chemistry was first discussed and defended by British Emergentists. British Emergentism
defended the emergence of chemistry before the advent of quantum mechanics. With the development of
quantum mechanics and quantum chemistry, the emergence of chemistry, as it was advocated by British
emergentists, was mostly rejected in philosophy. However, in the contemporary literature the emergence of
chemistry from quantum mechanics has been reformulated and supported on new grounds. Perhaps the most
detailed and widely discussed account of emergence with respect to chemistry is Robin Hendry’s account of the
strong emergence of molecular structure. However, there are also alternative understandings of emergence
within the philosophy of chemistry.

a. British Emergentism in Chemistry


British Emergentism refers to a group of philosophers in the 19th and 20th centuries which is regarded as the
first to provide a detailed and coherent philosophical account of emergence. Among the examples that British
Emergentists invoked in order to support the existence of emergence is that of chemistry and in particular of
chemical bonding. In particular, J. S. Mill argued that ‘the different actions of a chemical compound will never,
undoubtedly, be found to be the sums of the actions of its separate elements’ (quote in McLaughlin 1992: 28;
see also Mill 1930). C. D. Broad also advocated the emergence of chemistry on the grounds that it is not
‘theoretically possible to deduce the characteristic behaviour of any element from an adequate knowledge of the
number and arrangement of the particles in its atom, without needing to observe a sample of that substance’
(Broad 1925: 70; see also McLaughlin 1992: 47; Hendry 2006: 176-180; Hendry 2010a: 210; Hendry 2010b:
185).
The putative empirical evidence that emergentists invoked for the support of the emergence of chemical
bonding is the failure to deduce the chemical behaviour of elements from the entities and properties that
constitute those chemical elements. Since one does not describe and predict how chemical elements are bonded
to each other only with reference to the entities that compose them, then this suffices to support that chemical
bonding is an emergent chemical property which exerts downward causal powers to the entities that constitute
the relevant chemical elements (Scerri 2007a: 921).

The British Emergentists’ argument for the emergence of chemical bonding was formulated before the advent of
quantum mechanics. According to McLaughlin, once quantum mechanics contributed to the understanding of
atomic and molecular properties, including the chemical bond, the emergence of chemical bonding was no
longer justified in the manner that British Emergentism advocated:

Quantum mechanical explanations of chemical bonding suffice to refute central aspects of Broad’s Chemical
Emergentism: Chemical bonding can be explained by properties of electrons, and there are no fundamental
chemical forces. (Mclaughlin 1992: 49; see also Scerri 2007a)

On the other hand, Scerri argues that McLaughlin is mistaken to reject the emergence of chemistry and rejects
McLaughlin’s claims that

•there was no complete or adequate theory of chemical bonding before the advent of quantum
mechanics; and
•quantum mechanics provided a complete theory of chemical bonding (Scerri 2007a: 922-923; see also
Scerri 2012a).
In fact, Scerri claims that the quantum mechanical theory of chemical bonding should be viewed as continuous
and as enhancing Lewis’s theory of chemical bonding (Scerri 2007a: 922-923). The advent of quantum
mechanics does not refute pre-quantum, chemical theories of bonding, but rather offers a deeper understanding
of chemical bonding. Chemistry remains vital in the description and explanation of the chemical behaviour of
elements because quantum mechanics cannot offer by itself a complete account of chemical bonding and of the
overall chemical behaviour of elements. While quantum mechanics provides quantitative information regarding
particular chemical properties of elements and compounds, it ‘cannot predict what compounds will actually
form’ (Scerri 2007a: 924). Quantum mechanics can neither provide an explanation of how atoms and molecules
evolve in time, nor can it provide a complete explanation of their overall chemical behaviour (Scerri 2007b: 78).
These two characteristics of quantum mechanics, apart from blocking the possibility of a ‘complete’ reduction of
chemistry, also allegedly support the claim that chemical entities and properties emerge at a level ‘over and
above what one would expect from the constituents of the system’ (Scerri 2007b: 77; see also Llored 2012: 254).
What Scerri means by emergence is, however, unclear since he only specifies this notion contrary to physicalism
and does not provide a detailed account of the emergence of chemistry.

b. Strong Emergence
Hendry formulates one of the most detailed and widely discussed accounts of emergence regarding chemistry.
Hendry’s account focuses on a metaphysical understanding of emergence that has direct implications on the
metaphysical relation between chemical and quantum mechanical entities and properties, as well as on the
nature of molecular structure. His account of strong emergence is formulated in terms of downward causation,
and the putative empirical evidence that supports his position is drawn from the manner in which quantum
mechanics and chemistry each describe molecular structure.

According to Hendry, the structure of a molecule strongly emerges from its quantum mechanical entities in the
sense that it exhibits downward causal powers. Specifically, ‘the emergent behaviour of complex systems must
be viewed as determining, but not being fully determined by, the behaviour of their constituent parts’ (Hendry
2006: 180).

Strong emergence is supported by the ‘counternomic criterion for downward causation’ (Hendry 2010b: 189).
According to this criterion, ‘a system exhibits downward causation if its behavior would be different were it
determined by the more basic laws governing the stuff of which it is made’ (Hendry 2010b: 189). The manner in
which quantum mechanics describes a molecule’s structure allegedly satisfies the counternomic criterion and
thus supports the view that molecular structure strongly emerges.

In order to support this claim, Hendry makes a distinction between ‘resultant’ and ‘configurational’
Hamiltonians. A molecule’s resultant Hamiltonian takes into account all the intra-molecular interactions and is
constructed using as input only fundamental physical interactions and the value of the physical properties of the
entities (such as masses, charges, and so forth) (Hendry 2010a: 210-211). Given the resultant Hamiltonian, the
so-called ‘Coulombic Schrödinger equation’ is constructed, which is a complete and exact description of the
relevant molecule. However, the resultant Hamiltonian is in practice never used for the solution of the
Schrödinger equation. This is primarily due to the equation’s mathematical complexity. Nevertheless, if the
Coulombic Schrödinger equation were to be solved, it would not distinguish between different molecular
structures (specifically that of isomers), and it would not explain the symmetry properties of a molecule.
Instead, quantum explanations of molecular structure are based on the construction of ‘configurational
Hamiltonians’ for the solution of the Schrödinger equation of a molecule (Hendry 2010a: 210-211).
Configurational Hamiltonians are constructed on the basis of ad hoc assumptions which impose on the
Schrödinger equation the molecular structure that is supposed to be derived from that equation. This situation
satisfies the counternomic criterion because we did not recover a molecule’s ‘structure from the “resultant”
Hamiltonian, given the charges and masses of the various electrons and nuclei; rather we viewed the motions of
those electrons and nuclei as constrained by the molecule of which they are part’ (Hendry 2006: 183).

Hendry presents two examples that illustrate that the counternomic criterion is satisfied with respect to
molecular structure. The first example concerns isomers (see also Bishop 2010: 172-173). Isomers are sets of
molecules that contain the same number and kind of atoms, but whose atoms are arranged differently. This
means that isomers differ only in terms of their structure. Isomers have distinct chemical descriptions and they
are invoked for the explanation of a variety of physical and chemical phenomena. If one is to describe an isomer
via the use of its resultant Hamiltonian, then the Coulombic Schrödinger equation is identical with the
Coulombic Schrödinger equations that describe the other relevant isomers (Hendry 2017: 153). On the other
hand, if one is to describe an isomer via the use of its configurational Hamiltonian, then the Schrödinger
equation that is subsequently constructed, is not identical to those that describe the other relevant isomers.
According to Hendry, this means that this example satisfies the counternomic criterion. He thinks it illustrates
that the molecule’s behaviour, as this is described ‘by the more basic laws governing the stuff of which it is
made’ (that is, via the resultant Hamiltonian) is different from its behaviour, as this is described by assuming
certain chemical properties (namely, its structure) via the configurational Hamiltonian.

The second example that Hendry takes as empirical support for downward causation involves the symmetry
properties of molecules. Similarly to the case of isomers, one cannot derive the different chemical symmetry
properties from the relevant resultant Hamiltonian because ‘the only force appearing in molecular Schrödinger
equations is the electrostatic or Coulomb force: other forces are negligible at the relevant scales. But the
Coulomb force has spherical symmetry’ (Hendry 2017: 154).
As is the case with other accounts of strong emergence in philosophy of science, Hendry’s account of strong
emergence overcomes the overdetermination problem by postulating that there are certain quantum
mechanical effects which do not have purely quantum mechanical causes (Wilson 2015: 353). That is, accounts
of strong emergence deny the causal completeness of the physical (CCP), which states that ‘every lower-level
physically acceptable effect has a purely lower-level physically acceptable cause’ (Wilson 2015: 352). Instead of
the CCP, Hendry proposes an alternative principle; namely the ‘ubiquity of physics’ (UP):

Under the ubiquity of physics, physical principles constrain the motions of particular systems though they may
not fully determine them. (Hendry 2010b: 188)

This principle acts as a substitute for the causal completeness of the physical (CCP) which Hendry rejects and
which is incompatible with his notion of strong emergence. UP allows for the physical principles (as these are
formulated via the physical laws and theories) to ‘apply universally without accepting that they fully determine
the motions of the systems they govern’ (Hendry 2010b: 188). According to Hendry, unlike UP, the CCP is not
well supported by physics itself, and he follows Bishop in thinking it ‘does not imply its own causal closure’
(Bishop 2006: 45). Note that, given the rejection of the CCP, strong emergence, as understood by Hendry, is
incompatible with not only some form of epistemic reduction but also with reductive and non-reductive
physicalism.

A critique of Hendry’s account in the philosophy of chemistry literature is provided by Scerri, who argues that
the putative empirical evidence invoked by the former for the support of strong emergence is merely a
‘theoretical rather than ontological issue’ (Scerri 2012a: 25).

c. Alternative Forms of Emergence


There are alternative accounts of emergence with respect to chemistry. These are mostly accounts which focus
on the unique epistemological features of chemistry and propose an understanding of emergence that is
primarily epistemic, rather than metaphysical. For example, Bishop and Atmanspacher (2006) formulate an
account of ‘contextual emergence’ which they take to successfully apply in two separate cases: namely to the
case of molecular structure and to that of temperature (see also Bishop 2010). With respect to molecular
structure, they argue that quantum mechanics provides necessary but not sufficient conditions for the
description of molecular structure. This implies that reduction is not the appropriate account to correctly
specify the relation between the two relevant descriptions. In order to derive a lower-level (that is, quantum
mechanical) description of molecular structure, one introduces sufficient conditions by specifying the particular
context in which the relevant lower-level system is considered. This allegedly supports the claim that molecular
structure is a novel property which is not derivable by the quantum mechanical description alone but rather
emerges from it (Bishop and Atmanspacher 2006: 1774; see also Bishop 2010: 176-177; Llored 2012: 248).

Furthermore, Llored presents ‘a relational form of emergence which pays attention to the constitutive role of
the modes of intervention and to the co-definition of the levels of organization’ (Llored 2012: 245). This is not a
metaphysical account of emergence; as Llored states, his proposed account is ‘agnostic’ with respect to the
ontology of chemistry and rather focuses on ‘what chemists do in their daily work’ (Llored 2012: 245). In
particular, Llored looks at how ‘from the Twenties to nowadays, quantum chemical methods have
been constitutively concerned with the links between the molecule and its parts’ (2012: 257) (italics are in the
original text). Among other things, he presents and analyses the debate between Linus Pauling and Robert
Mulliken who both ‘focused on the description and the understanding of the molecule, its reactivity, and thus its
transformations’ (Llored 2012: 257). Llored argues that his proposed account of emergence is not one which
advocates an asymmetric relation between higher and lower-level properties. Rather, both chemical and
quantum mechanical properties ‘co-emerge’ (Llored 2014: 156). Chemical phenomena are understood ‘as
relative to a certain experimental context, with no possibility of separating them from this context’ (Llored
2014: 156; see also Llored and Harré 2014).

5. Beyond Reduction and Emergence


Very few accounts consider the relation of chemistry to quantum mechanics without invoking some form of
reduction or emergence. In fact, if we are to understand epistemic reduction and strong emergence as the two
extremes of a spectrum of inter-theoretic accounts, then there is a variety of positions that have remained to
this day relatively unexplored with respect to chemistry. Nevertheless, there are some philosophers who
consider the possibility of understanding chemistry’s relation to quantum mechanics without reference to
reduction or emergence. This section distinguishes between two main camps. First are those accounts which
consider unity without reduction. Secondly, there are accounts which support the autonomy of chemistry
without reference to some form of emergence.

a. Unity without Reduction


Two philosophers of chemistry have primarily examined chemistry’s relation to quantum mechanics in terms of
unity without reduction. First, Needham examines unity without reduction by presenting Pierre Duhem’s
‘scheme’ of ‘unity without reduction’ (Needham 2010: 166). He states that

unity surely does not require reduction, intuitively understood as the incorporation of one theory within
another. […] Consistency, requiring the absence of contradiction, and more generally in the sense of the absence
of conflicts, tensions and barriers within scientific theory, would provide weaker, though apparently adequate,
grounds for unity. (Needham 2010: 163)

According to Duhem’s scheme of unity, ‘(m)icroscopic principles complement macroscopic theory in an


integrated whole, with no presumption of primacy of the one over the other’ (Needham 2010: 167). This implies
that Duhem’s understanding of unity is incompatible with reductionism in the sense that it rejects that physics
is the most fundamental science.

Moreover, Needham argues that positions on unity can be distinguished into four groups:

(i) Unity in virtue of reduction, with no autonomous areas,

(ii) unity in virtue of consistency and not reduction, but still no autonomy because of interconnections,

(iii) unity in virtue of consistency and not reduction, with no autonomous areas, and

(iv) disunity. (Needham 2010: 163-164)

Hettema engages in the discussion of unity with respect to chemistry and evaluates Needham’s scheme of unity
(2017). In particular, Hettema takes that the first form of unity assumes a form of ‘reductionism in which
derivation is strict and reduction postulates are identities’ (Hettema 2017: 277). Regarding the second form of
unity, Hettema argues that it faces certain challenges. For example, in this form of unity ‘the nature of the
“interconnections” is (..) not well specified in Needham’s scheme’ (Hettema 2017: 277). Moreover, ‘the theories
of chemistry and physics are not as strongly dependent on each other as implied (though not stated) in position
(ii) in the scheme’ (Hettema 2017: 277-278). Hettema rejects the third form of unity because it allegedly
disregards the ‘idea that one science may fruitfully explain aspects of another’ (Hettema 2017: 278).
As already mentioned, Hettema proposes a novel account of reduction regarding the relation between chemistry
and quantum mechanics (see subsection 3d). In the broader context of unity, Hettema takes his account to
propose a form of unity that Needham’s scheme does not capture. Specifically, Hettema’s account does not
support ‘a form of unity in virtue of reduction with no autonomous areas’ (in line with (i)) because, unlike (i), it
does not require strict derivation nor the existence of identity relations between the reduced and reducing
theory. Moreover, Hettema’s account does not advocate unity without reduction either. While he acknowledges
that his account shares common features with non-reductive accounts of unity in the philosophy of science
literature, he maintains that his account proposes a ‘naturalised Nagelian reduction’ (Hettema 2012b: 143).

Interestingly, there are two features that his account allegedly shares with certain non-reductive accounts of
unity. First, Hettema takes his account of reduction to be compatible with an understanding of theories as
‘interfield theories’ which ‘use concepts and data from neighbouring fields’ (in line with Darden and Maull 1977)
(Hettema 2012b: 160). In this context, absolute reaction rate theory is characterised as an interfield theory
‘where the theories comprising the interfield are in turn reductively connected’ (Hettema 2012b:168). There is
no one-to-one relation between the reduced and reducing theory; rather there is a ‘net of theories’ where
‘connective and derivative links of a Nagelian sort exist between all these theoretical approaches’ (Hettema
2012b:168). As a result, the overall reduction of chemistry is specified in terms of a network of different theories
that are reductively connected between them (Hettema 2012b:171). Secondly, Hettema takes his account to be
compatible with Bokulich’s non-reductive account of ‘interstructuralism’, according to which two theories are
related in virtue of the ‘structural continuities and correspondences’ between them (Bokulich 2008: 173;
Hettema 2012b: 163). Indeed, Hettema identifies structural continuities in the case of the absolute reaction rate
theory (Hettema 2012b: 171).

Lastly, Seifert (2017) advocates unity without reduction, arguing that chemistry and quantum mechanics are
unified in a non-reductive manner because they exhibit particular epistemic and metaphysical inter-
connections.

Vous aimerez peut-être aussi