Вы находитесь на странице: 1из 7

a.

Epistemological, or Intertheoretic, Reduction


Discussion of epistemological, or intertheoretic, reduction primarily happens in the context of Nagel’s account
of reduction. In the philosophy of chemistry, a Nagelian reduction is understood as requiring at least, in
principle, the derivation or deduction of chemistry from quantum mechanics (Needham 2010: 164; Hettema
2017: 7). A Nagelian reduction consists of two ‘formal’ requirements, namely the ‘connectability and
derivability’ of the two theories ((Scerri 1994: 160), see also (Hettema 2017: 7)). Moreover, the reduction of
chemistry to quantum mechanics would fall under the cases of heterogeneous reductions. This is because ‘some
typically chemical terms cannot be found in the quantum mechanical language’, thus requiring the existence of
bridge laws (Scerri 1994: 160; see also Primas 1983: 5). A successful reduction would allegedly be sufficiently
supported if the chemical properties of atoms and molecules can, at least in principle, be calculated by quantum
mechanics ‘entirely from first principles, without recourse to any experimental input whatsoever’ (Scerri 1994:
162). Note that the latter form of quantum mechanics is often referred to as ‘ab initio quantum mechanics’
(Scerri 1994; Schwarz 2007).

In the philosophy of chemistry there has been debate on what the appropriate criteria are for a successful
Nagelian reduction of chemistry to physics (see for example Hettema 2012a; 2017; Needham 1999; 2010; Scerri
1994). For example, Hettema claims that the use of the term ‘Nagelian’ with reference to the aforementioned
understanding of reduction is to an extent misleading because Nagel was not so strict in his account of
reduction:

Reduction is too often conceived of as a straightforward derivation or deduction of the laws and concepts of the
theory to be reduced to a reducing theory, notwithstanding Nagel’s insistence that heterogeneous reduction
simply does not work that way. (Hettema 2017: 1-2; see also Hettema 2012b: 146; Dizadji-Bahmani, Frigg and
Hartmann 2010; Fazekas 2009; Klein 2009; Nagel 1979; van Riel 2011)

While Nagel’s account of reduction is the most widely discussed account in the philosophy of chemistry, there
are other accounts from philosophy. They include Oppenheim’s and Putnam’s account of micro-reduction
(Oppenheim and Putnam 1958; Hendry 2012: 368-369). Very briefly, according to this account of reduction, a
theory T1 micro-reduces a theory T 2 if (i) the phenomena that are explained by T 2 can be explained by T 1; and
(ii) T1 describes the parts of the entities, properties, and so forth that are postulated by T 2. According to
Hendry, if ‘the micro reductive explanation takes the form of a deduction’, then Oppenheim’s and Putnam’s
account is a kind of Nagelian reduction (Hendry 2012: 369).
Nagel, Oppenheim and Putnam take chemistry’s relation to physics to be a paradigmatic case of their respective
accounts of reduction (Hendry 2012: 369). A large, though not the entire, part of the philosophy of chemistry
literature discusses reduction by investigating whether these accounts of reduction correctly apply to
chemistry’s relation to quantum mechanics. Popper’s understanding of reduction has also been investigated in
the context of chemistry’s relation to quantum mechanics (Scerri 1998; Needham 1999).

The epistemological reduction of chemistry to quantum mechanics is primarily examined by looking at how
quantum mechanics, via the Schrödinger equation, describes the chemical properties of atoms and molecules.
Given this, it is useful to briefly present how quantum chemistry employs the Schrödinger equation in order to
describe the chemical properties of atoms and molecules. This sub-section henceforth focuses on the non-
relativistic Schrödinger equation since this is the one that is standardly employed for the description of atoms
and molecules and that is discussed with respect to chemistry’s relation to quantum mechanics.

The Schrödinger equation is the ‘equation of motion for the wave function’ which describes ‘the state of a
quantum-mechanical system, and (more generally) for the corresponding state-vector’ (Palgrave Macmillan Ltd
2004: 2029). The solutions of the time-dependent Schrödinger equation (Ψ(x,t)) are (potentially) the
wavefunctions of the system under examination (that is of an electron, atom, molecule and so forth).
The generic form of the time-dependent Schrödinger equation is the following:

iħ ∂Ψ(x,t)/ ∂t = – (ħ2/2m)(∂2Ψ(x,t)/∂x2) + VΨ(x,t),


where

∂: partial derivative

Ψ(x,t): a system’s wavefunction

ħ: Planck’s constant

m: the system’s mass

x: position

t: time

V: potential energy

i: imaginary unit (square root of negative one)

If one assumes that a system’s potential energy is independent of time, then it is possible to solve the
Schrödinger equation using the method of separation of variables (Griffiths 2005: 24). In this context, the
resulting solutions are wavefunctions of the following form (Griffiths 2005: 24):

Ψ(x,t) = ψ(x)φ(t),

where

ψ: a function of position

φ: a function of time

Based on the ability to separate the variables of the Schrödinger equation, it is possible to formulate the time-
independent Schrödinger equation, which is an equation independent of time and whose solutions are a
system’s time-independent wavefunctions, ψ(x). These wavefunctions correspond to the stationary states of the
system under examination.

The time-independent Schrödinger equation does not yield a unique solution (that is, one wavefunction)
(Griffiths 2005: 27). It yields an infinite number of solutions (ψ(x 1), ψ(x2), …), each of which corresponds to a
different state of the system under examination. In accordance with the superposition principle, any linear
combination of the solutions of the time-independent Schrödinger equation is also regarded as a wavefunction
that represents a possible state of the system (Griffiths 2005: 27).
The stationary state of a system, through its wavefunction ψ(x), provides useful information about the total
state of the system, Ψ(x,t). First, the probability density Ψ(x,t) equals ∣ψ(x) ∣2. This means that knowledge of just
the stationary state of a system, through the solution of the time-independent Schrödinger equation, provides
the probability of finding the system at a particular region in space. Secondly, it is possible to calculate the
expectation value of any dynamical variable of a state of the system through the stationary state of the system
alone (Griffiths 2005: 26). Stationary states are states of definite total energy, E (Griffiths 2005: 26). Each
solution to the time-independent Schrödinger equation is associated with a particular allowed total energy of
the system (E1, E2, …). The wavefunction that is associated with the minimum total energy corresponds to the
ground state of the system, whereas the wavefunctions whose total energies are larger correspond to the excited
states of the system.
The time-independent Schrödinger equation for an isolated molecule provides an infinite number of solutions
(that is, wavefunctions), each of which corresponds to different stationary states of the molecule. For example, a
stable isolated molecule, in virtue of being stable, is said to be in the ground state. From this, it follows that it is
represented by the wavefunction that is associated with the system’s ground state and that it has the minimum
total energy.

The Hamiltonian operator plays a central role in the solution of the time-independent Schrödinger equation for
quantum systems and isolated molecules in particular. When the system under examination is an isolated
molecule, the Hamiltonian operator corresponds to the total energy of the molecule (that is, its eigenvalues are
the total energy of each state of the molecule); hence it is called the molecular Hamiltonian. In principle, the
molecular Hamiltonian operator includes all the factors that determine the kinetic and dynamic energy of the
molecule. That is, it should take into account the kinetic energy of each nucleus and electron in the system, the
repulsion between each pair of electrons and between each pair of nuclei, and the attraction between each pair
of electron and nucleus.

Because of the mathematical complexity involved in the formulation of the Hamiltonian operator, atomic and
molecular systems are examined within the framework of the Born-Oppenheimer approximation (henceforth
BO approximation; also referred to as the adiabatic approximation). The BO approximation is a
‘(r)epresentation of the complete wavefunction as a product of an electronic and a nuclear part Ψ(r,R) =
Ψe( r,R) ΨN(R)’ (IUPAC 2014: 179). The validity of the BO approximation is ‘founded on the fact that the ratio
of electronic to nuclear mass […] is sufficiently small and the nuclei, as compared to the rapidly moving
electrons, appear to be fixed’ (IUPAC 2014: 179).
Within the BO approximation, one can in principle formulate the Hamiltonian operator by positioning the
nuclei at all the possible fixed positions. Each set of nucleonic positions corresponds to different quantum states
of the system (hence to different wavefunctions) and to different values of the total energy, E, of the atom or
molecule. However, in practice this process is not followed. By having prior knowledge of the quantum system
that is under examination—for example, by knowing the chemical and structural properties of the examined
molecule—only particular nucleonic conformations are considered when constructing the Hamiltonian
operator.

The BO approximation is a feature of quantum mechanics which plays a central role in the investigation of
chemistry’s relation to quantum mechanics (Bishop 2010: 173; van Brakel 2014: 31-33; Woolley 1976; 1978;
1991; 1998; Woolley and Sutcliffe 1977; Sutcliffe and Woolley 2012). It has often been invoked as putative
empirical evidence for the rejection of chemistry’s reduction to quantum mechanics as well as for the support of
the emergence of chemistry (see next sections). Solving the equation outside the BO approximation in order to
describe atomic and molecular properties is currently investigated in chemistry and quantum chemistry (for
example Tapia 2006). This implies that there are features of quantum mechanics which may further contribute
to our understanding of chemistry’s relation to quantum mechanics (for example Woolley 1991).

Note that even when the nucleonic conformation is fixed in the manner represented by the BO approximation,
calculating the solution of the Schrödinger equation remains a complicated task. Each nucleonic conformation
is compatible with different quantum states of the system (and thus different wavefunctions). This is compatible
with chemistry’s understanding of atoms and molecules because, even if the nuclei are fixed at particular
positions, the electrons may behave in more than one possible way within an atom or molecule.
In light of the above, the Schrödinger equation is not solved analytically for all atoms and molecules. As Hendry
states:

There is an exact analytical solution to the non-relativistic Schrödinger equation for the hydrogen atom and
other one-electron systems, but these are special cases on account of their simplicity and symmetry properties.
(Hendry 2010a: 212)

Instead, researchers have developed various approximate methods in order to solve it, most of which employ
the BO approximation. In general, the development of computation has led to the proliferation of complex
computational methods that solve the equation by following different mathematical strategies and by making
different assumptions. These methods include the Valence Bond Approach, the Molecular Orbital Approach, the
Hartree-Fock Method and Configuration Interaction.

Based on the above, there are philosophers who argue in favour of the epistemological reduction of chemistry to
quantum mechanics. For example, Schwarz argues that ab initio quantum mechanics can in principle derive all
‘well-defined numerical properties’ of the chemical elements (Schwarz 2007: 168). Ab initio quantum
mechanics refers to quantum mechanical methods that are ‘independent of any experiment other than the
determination of fundamental constants. The methods are based on the use of the full Schrödinger equation to
treat all the electrons of a chemical system’ (IUPAC 2014: 5).
While Schwarz does not examine chemistry’s relation to quantum mechanics in terms of a particular
philosophical account of reduction (such as Nagel’s account of reduction), he advocates some sort of reductive
relation between chemistry and quantum mechanics. He claims that the ‘difficulty’ of ab initio quantum
mechanics to (in practice) derive certain chemical properties is due to the fact that ‘basic qualitative chemical
concepts are so vaguely defined’ and ‘fuzzy’ (Schwarz 2007: 172, 174). Given the above, he believes that the
periodic system is in a ‘transition phase’ from a primarily ‘empirical model of chemistry’ to ‘an understandable
model based in physical theory’ (Schwarz 2007: 173).

The epistemological reduction of chemistry to quantum mechanics is alternatively supported by Bader’s


Quantum Theory of Atoms in Molecules (QTAIM) (Bader 1990; Bader and Matta 2013; Matta and Boyd 2007;
Matta 2013). The QTAIM provides a topological analysis of electron density through which one derives
information regarding atomic and bonding properties. The QTAIM provides experimentally verifiable
information regarding the properties of large molecules, by reconstructing their properties from ‘smaller
fragments’ (Matta 2013). It is a scientific theory which ‘demonstrates that every measurable property of a
system, finite or periodic, can be equated to a sum of contributions from its composite atoms’ (Bader 1990).

Bader takes the QTAIM to provide correct descriptions, explanations and predictions of the chemical properties
of matter ((Bader 1990: vi), see also (Bader and Matta 2013), (Causá et al. 2014), (Hettema 2012a) and
(Hettema 2013)). While Bader does not explicitly talk about the reduction of chemistry to quantum mechanics
in philosophical terms, his account is regarded in the philosophy of chemistry as representing ‘a proper,
(reductionist) basis for chemistry’ (Hettema 2013: 311). This is because, according to Bader and Matta, the
QTAIM allegedly supports the claim that ‘chemistry is physics’ (Bader and Matta 2013: 254). However, Hettema
argues that while Bader’s view of the QTAIM suggests that the QTAIM is related to chemistry in a manner that
closely resembles Kemeny and Oppenheims’ reductive eliminativist account, the QTAIM fails to be a reductive
theory of this sort (Hettema 2013). Moreover, Arriaga, Fortin and Lombardi argue that while the QTAIM
manages to ‘provide a rigorous definition of the chemical bond and of atoms in a molecule, it appeals to
concepts that are unacceptable in the quantum-chemical context’, thus failing to sufficiently support the
reduction of chemistry to quantum mechanics (Arriaga et al. 2019: 125). Van Brakel makes a similar point,
arguing that the QTAIM works only after postulating facts from chemistry (van Brakel 2014: 32), thus rendering
it insufficient for the support of chemistry’s reduction to quantum mechanics.
b. Antireductionism with Respect to Chemistry
Many members of the philosophy of chemistry community reject the epistemological reduction of chemistry to
quantum mechanics, as understood in terms of the aforementioned accounts. As Hettema states:

The idea that chemistry stands in a reductive relationship to physics still is a somewhat unfashionable doctrine
in the philosophy of chemistry. (2017: 1)

Indeed, there are alternative and often incompatible positions in the philosophy of chemistry which argue,
either explicitly or implicitly, against the reduction of chemistry to quantum mechanics. These antireductionist
views can be divided into two main camps (Scerri 2007b). First are those positions which reject the reduction of
chemistry tout court (Schummer 1998; Schummer 2014b; van Brakel 2000). That is, they ‘deny the whole
enterprise’ of reducing chemistry to quantum mechanics on grounds that have to do with the unique
methodological, classificatory or other epistemological features of chemistry (Scerri 2007b: 70). Philosophers
that follow this antireductionist approach support, either implicitly or explicitly, the irreducibility of chemistry
by arguing that chemistry, in virtue of being a science of substances which employs unique classificatory tools
and concepts, cannot be reduced to a science which looks at the micro-constituents of those substances and
which disregards the classificatory or methodological tools and concepts that are of interest to chemists.
In the second camp are those positions which examine in detail how quantum mechanics describes, predicts,
and explains particular chemical entities, properties, and so forth (such as the chemical bond, molecular
structure, orbitals and the periodic system). They consider how quantum mechanics describes particular
chemical properties and through this analysis they implicitly or explicitly argue against the reduction of
chemistry to quantum mechanics (Bogaard 1978; González et al. 2019; Hendry 1998; 1999; 2010a; 2012;
Ramsey 1997; Scerri 1994; 1998; Woolley 1976; 1978; 1985; 1998; Woolley and Sutcliffe 1977; Weininger 1984;
Woody 2000).

For example, Scerri evaluates the manner in which the Schrödinger equation is solved so as to yield accurate
results about the properties of atoms and molecules. He claims that ab initio quantum mechanics has yielded
relatively accurate results regarding the ground-state energy of particular atoms and has acknowledged the
success of quantum mechanics in providing a mathematical analysis of chemical phenomena and in generating
sufficiently accurate quantitative values of chemical properties such as bond strength and dipole moments
(2007b; 2012). However, he takes that this does not sufficiently support the reduction of chemistry to quantum
mechanics (Scerri 1994: 164). Specifically, the approximate methods that are employed for the solution of the
Schrödinger equation—and without which a solution cannot be provided—involve the use of ad hoc
assumptions which, in virtue of being ad hoc and reliant ‘on experimental data’, undermine the thesis that
chemistry is reduced in a Nagelian manner to quantum mechanics (Scerri 1994: 165-168; see also Scerri 1991:
320-321). Note that Hofmann (1990) presents how models and approximations have been employed
throughout the history of quantum mechanics for the description of chemical properties; see also Gavroglu and
Simões (2012).

Scerri invokes the periodic table and the electronic configuration model as examples that support the failure of
chemistry’s reduction to quantum mechanics (Scerri 2007b: 74; Scerri 2012b: 79-80; Scerri 1991).

Before presenting Scerri’s argument, it is useful to briefly define the chemical terms that his and subsequent
analyses invoke. The electronic configuration is ‘a distribution of the electrons of an atom or a molecular entity
over a set of one-electron wavefunctions called orbitals, according to the Pauli principle’ (IUPAC 2014: 317). An
orbital, whether atomic or molecular, is a ‘(w)avefunction depending explicitly on the spatial coordinates of only
one electron’ (IUPAC 2014: 1034). An atomic orbital is a ‘(o)ne-electron wavefunction obtained as a solution of
the Schrödinger equation for an atom’ (IUPAC 2014: 124). Given that orbitals depend on the spatial coordinates
of electrons, the electronic configuration of an atom provides a representation of the distribution of electrons in
the atom. This is particularly important in chemistry because it serves as a basis for the explanation and
prediction of the type of bonds that are formulated between atoms.

With respect to the periodic table then, Scerri’s claim is broadly the following. The manner in which chemical
elements are ordered in the periodic table is partially explained and could be regarded as derived by quantum
mechanics because quantum mechanics specifies the electronic configuration of the atoms of each element
(Scerri 2012b: 75). However, there are certain features of the periodic table, such as the length of its periods,
which are not deducible from quantum mechanics (Scerri 2012b: 77-78). Therefore, the derivation of the
periodic table from quantum mechanics, and thus the reduction of chemistry, cannot be sufficiently supported.

Moreover, a Nagelian reduction ‘requires axiomatised versions of the theory to be reduced as well as the
reducing theory’, which at least with respect to chemistry cannot possibly be argued for (Scerri 2006: 124). A
similar point is made by Hettema regarding Nagelian reduction: ’chemistry is a field, whereas reduction tends to
be a relation between individual theories, or between laws and theories’ (Hettema 2017: 1). Furthermore,
quantum mechanics does not provide on its own ‘a conceptual understanding of chemical phenomena’ (Scerri
2007b: 74). Instead, chemists employ chemical models and theories in order to formulate sufficient
descriptions, explanations, and predictions of chemical phenomena and properties. Another problem for the
reduction of chemistry is that quantum mechanics is symmetric under time inversion, and thus cannot provide
an explanation of why chemical entities evolve in time the way they do. It can only provide a ‘reductive
description’ of chemical properties independent of time (Scerri 2007b: 78). In fact, while quantum mechanics
provides numerical values to particular chemical properties, it does not provide a complete explanation of a
system’s chemical behaviour (Scerri 2007b: 78).
Scerri also rejects the success of an approximate reduction of chemistry to quantum mechanics (1994; 1998). By
approximate reduction, Scerri refers to Putnam’s analysis of reduction, which permits the reducing theory to be
approximately and not exactly true (Scerri 1994: 161). That is, ‘the relationships postulated by the theory hold
not exactly, but with a certain specifiable degree of error’ (Putnam 1965: 206-207). In this context, reduction is
not undermined if ab initio quantum mechanics provides only approximate results of the value of atomic and
molecular properties, as long as these results are accompanied by a specifiable degree of error. However, Scerri
rejects approximate reduction as the errors ‘are seldom computed by independent ab initio criteria’ (Scerri
1994: 168). Scerri also examines approximate reduction in relation to Popper’s analysis of the reduction of
chemistry. In this context, Scerri draws a very similar conclusion with respect to the approximate reduction of
chemistry (Scerri 1998: 42).

Based on all the above, Scerri concludes that the reduction of chemistry is ambiguous since, depending on what
the set criteria for a successful reduction are, chemistry’s reduction to quantum mechanics ‘is both successful
and unsuccessful’ (Scerri 2007b: 76; Scerri 2012b: 80).

Other philosophers also argue that chemistry has failed to epistemically reduce to quantum mechanics by
pointing out similar issues with respect to the quantum mechanical description of chemical phenomena (see
Bogaard 1978; Hendry 1998; Hendry 2010b: 183; Primas 1983; Woolley 1976; 1998; Woolley and Sutcliffe
1977). For example, Primas argues that quantum mechanics is ‘incorrect and should be revised, partly because
[it] seems incapable of rendering a robust account of concepts such as molecular shape’ (Hettema 2017: 53, see
also Primas 1983). Bogaard points out that chemists disregard a number of features of the behaviour of
subatomic particles when specifying an atom’s or molecule’s Schrödinger equation. These features include (a)
the behaviour of subatomic particles (namely protons and neutrons); (b) the energetic contribution of the
movement of the nuclei; and, (c) relativistic effects (Bogaard 1978: 346). Moreover, the fact that the
Schrödinger equation is ‘adapted’ so as to provide an accurate description of each particular system challenges
the view that quantum mechanics can, even in principle, deduce complete explanations of chemical phenomena
(Bogaard 1978).
González et al. (2019) argue that there is a tension between the theoretical postulates of quantum mechanics
and how molecular structure is understood in chemistry. In particular, Heisenberg’s uncertainty principle
implies that a ‘quantum “particle” is not an individual in the traditional sense, since it has properties—those
represented by its observables—that have no definite value’ (González et al. 2019: 36). Such a metaphysical
understanding of quantum particles comes in contrast to chemistry’s understanding of molecular structure,
which is defined ‘in terms of the spatial relation of the nuclei conceived as individual localised objects’
(González et al. 2019: 43). The failure of chemistry’s reduction is further supported by the fact that the
Schrödinger equation cannot be solved analytically without the use of approximations and models (for example
Bogaard 1978: 347; González et al. 2019; Hendry 2010b). These approximations and models are based on
‘theoretical assumptions drawn from chemistry’, thus rendering the quantum chemical description of complex
atoms and molecules in a ‘loose relationship to exact atomic and molecular Schrödinger equations’ (Hendry
2010b: 183).

Lastly, Chang argues that since its advent, quantum chemistry was practiced in a manner that required the use
of pre-quantum, chemical knowledge (Chang 2015; 2017). The views of Linus Pauling, one of the main founders
of quantum chemistry, allegedly corroborate this argument, as Pauling took quantum chemistry to be ‘a direct
continuation of nineteenth-century organic structural chemistry’ (Chang 2015: 197-198). Chang also claims that
physics consists of many different branches and that the relation of those branches with more fundamental
physical theories has not been decisively shown to be reductionist. In light of this, and given that chemistry’s
relation to physics is examined in the context of a physical theory (that is, quantum mechanics) which is not the
most fundamental theory in physics, one should not assume chemistry to be unproblematically reduced to
physics (Chang 2015: 200; Chang 2017: 365). Thirdly, Chang looks at how chemistry is done in practice and
claims from this that chemistry is very far from being ‘submitted’ to physics (Chang 2015: 201). This claim
allegedly undermines the reduction of chemistry to quantum mechanics since quantum mechanics has never in
practice been sufficient for the description, explanation or prediction of phenomena that are within the purview
of chemistry (Chang 2015: 201-202).

Вам также может понравиться