Sei sulla pagina 1di 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/225485717

Mereology and quantum chemistry: The approximation of molecular orbital

Article  in  Foundations of Chemistry · November 2010


DOI: 10.1007/s10698-010-9092-7

CITATIONS READS

12 31

1 author:

LLored Jean-Pierre
Linacre College
18 PUBLICATIONS   67 CITATIONS   

SEE PROFILE

All content following this page was uploaded by LLored Jean-Pierre on 16 February 2018.

The user has requested enhancement of the downloaded file.


Found Chem (2010) 12:203–221
DOI 10.1007/s10698-010-9092-7

Mereology and quantum chemistry: the approximation


of molecular orbital

Jean-Pierre Llored

Published online: 8 July 2010


 Springer Science+Business Media B.V. 2010

Abstract Mulliken proposed an Aufbauprinzip for the molecules on the basis of


molecular spectroscopy while establishing, point by point, his concept of molecular orbit.
It is the concept of electronic state which becomes the lever for his attribution of electronic
configurations to a molecule. In 1932, the concept of orbit was transmuted into that of the
molecular orbital to integrate the probabilistic approach of Born and to achieve quanti-
tative accuracy. On the basis of the quantum works of Hund, Wigner, Lennard-Jones and
group theory, he suggested the fragment method to establish the characteristics of
molecular orbital for polyatomic molecules. These developments make it possible to bring
elements of thought on the relation between a molecular ‘‘whole’’ and its ‘‘parts’’. An
operational realism combined with the second law of thermodynamics can pave the way
for interesting tracks in the mereological study of chemical systems.

Keywords Reduction  Emergence  Mereology  Molecule  Atom 


Mixed  Aggregate  Properties  Spectroscopic state  Valence  Quantum state

Introduction

The study of the relations between a unit described as a ‘‘whole’’ and its identified parts as
such is brought up to date around a relevant debate on the concept of ‘‘property’’. This
evolution is hardly surprising if we refer to Alfred North Whitehead (1967: 157) who
asserted: ‘‘For physics, the thing itself is what it does’’.
An epistemological shift is noted between the concepts of intrinsic properties
(belonging to the studied object itself independently of the rest of the world) and extrinsic
or relational properties and those of resulting or co-operative properties. The resulting
properties announce foreseeable properties of the whole which derive from the simple
combination of properties of its parts, for instance like the mass of an object in non
relativistic mechanics, the optical absorption of a hydrated complex of the divalent ion of

J.-P. Llored (&)


Ecole Polytechnique CREA/CNRS, 91128 Palaiseau Cedex, France
e-mail: jean-pierre.llored@polytechnique.edu

123
204 J.-P. Llored

copper in monochromatic light or the optical activity of a solution of saccharose. The non
reducible total properties with those of the parts are known as co-operative or emergent.
They can be for example the spectroscopic and chemical properties of some molecules
compared to the properties of the atoms present in the chemical formulas which identify
them or in another field, the rheological behavior of a mixture of polymers. According to
Joseph E. Earley (2003: 89), this slip, between concepts of properties more particularly
interests the philosophers of science:
Most philosophers have yet to recognize that, when components enter into chemical
combination, those components do not, in general, maintain the same identity that
they would have had absent that combination…Interactions of such insights with the
philosophical study of wholes and parts (mereology) is its initial stages. It would be
useful to develop a mereology adequate to deal with chemical systems, in order to
facilitate future progress in dealing with other and more complex problems.
We aim at showing in this paper how some works in quantum chemistry, such as the
approximation of molecular orbital suggested by Robert Sanderson Mulliken and his
collaborators, can provide a support for a mereological approach concerning chemical
systems. Indeed, Mulliken considered a molecule to be a composite in which the atoms lost
their singularity. In this context, a molecule has properties which the atoms do not express
and its decomposition gives again the separate atoms. Mulliken was at odds with the idea
of an aggregation of atoms which referred to the general properties of the masses and the
movements, i.e. to a mechanics whatever its form. This molecular composite creates new
homogeneous bodies starting from heterogeneous elements, which cannot be interpreted in
terms of simple spatial vicinity of particles. The energy approach of molecular orbital does
not consider valence as an intrinsic property of the atoms. This step recalls, to a certain
extent, that of Pierre Duhem (1902, 1985) when he undertook to retranslate the Aristotelian
concept of power in terms of thermodynamic potential. The electronic state, the ‘‘binding
capacity’’, the ‘‘promotion’’ of an electron, ‘‘the energy-bonding-power’’, ‘‘bonding and
anti-bonding orbitals’’ are the many concepts we will present which aim at explaining the
capacity of the electrons to be linked to nuclei to form a molecule seen as a whole.
Bernadette Bensaude-Vincent (2005: 146) points out1:
Much more characteristic, in our eyes, is the recurrence in the history of chemistry of
those two possible interpretations of what a phenomenon must be: a mixture or an
aggregate. The chemists were always confronted with this choice and, according to
the times, they chose one or other interpretation or attempted to reconcile both. But
the pluralism of possible interpretations does not stop living in chemistry.
Thus the composite challenges the mere sum of the parts (Bensaude-Vincent 2007: 157–
165) and we can understand the pleas of mereological adaptations put forward by Earley
and Needham (2003). We want to elucidate how the concept of molecular orbit transmuted
into molecular orbital in 1932 by Mulliken questioning the relation between a molecule
and atoms. Are there atoms in a molecule? Does this question admit an answer similar to
that brought by Earley (2005: 85–102) concerning salt in sea water?

1
I have translated from B. Bensaude-Vincent, Faut-il avoir peur de la chimie?, Les empêcheurs de penser
en rond éditeurs, Paris, 146 (2005). ‘‘Beaucoup plus caractéristique, à nos yeux, est la récurrence dans
l’histoire de la chimie de ce duel d’interprétations des phénomènes –le mixte ou l’agrégat-. Les chimistes
ont toujours été confrontés à ce choix et, suivant les époques, ils ont opté pour l’une ou l’autre interprétation
ou bien tenté de concilier les deux. Mais le pluralisme des interprétations possibles ne cesse d’habiter la
chimie’’.

123
Mereology and quantum chemistry: the approximation of molecular orbital 205

The link between atoms and molecule studied by spectroscopy: molecular


Aufbauprinzip and correlation diagrams

Analogies between atomic and molecular spectra: empirical bases

At the end of the 1920s, the American researchers studying the spectra of bands made their
observations in the infra-red, the near ultraviolet and the visible. The rotation and vibration
of the molecules were studied via the molecular spectra while the electronic configuration
of the molecules was studied independently by the chemists—Gilbert Newton Lewis for
example—without a formal link being established between the theory of valence and the
study of the spectra. The objective of many researchers was to create a molecular model
which accounted for the various experimental facts of spectroscopy and chemistry.
In 1923, Lewis published his famous Valence and the structure of molecules which
produced a deep repercussion in the community of the chemists. At the same time, The
Faraday Society organized at the university of Cambridge, a congress on the concept of
valence where Irving Langmuir helped Mulliken become aware of chemical analogies
between some molecules and atoms. A node between atom and molecule was formed
around molecular spectroscopy. Mulliken was one of the craftsmen setting up a bridge
between the description of the atom using the quantum numbers and that of the molecules.
For that, he used the cement of spectroscopy. It was a question of connecting the spectral
terms of the isoelectronic series of molecules (molecules having the same number of
electrons) to those of the corresponding atoms. Mulliken thought of indexing the molecular
electronic states by quantum numbers describing the angular momentum, like those used
for atoms.
Mulliken (1930: 60–70) made a meticulous study of each kind of spectrum (pure
rotation, rotation-vibration and electronic) to describe many molecules. He used the
principle of combination and the tables of Henri Deslandres to determine typical param-
eters describing the internal vibrations of the molecule and making it possible to charac-
terize each energy level of vibration. He took account of the coupling between the
movements of vibration and rotation which constituted one of the principal differences
between the spectra of rotation-vibration and electronic spectra. The use of the parabolas of
Fortrat enabled him to determine the distance between two atoms in a diatomic molecule as
well as other parameters related to the deformation of the molecule because of the coupling
between rotation and vibration (Mulliken 1931a: 123). He also determined the strength of
the chemical bond while following the evolution of the internuclear distances and the
binding strength in the various bands of the spectrum. He used, with some hesitation, the
terms ‘‘binding capacity’’ of the electrons and ‘‘binding states’’ to explain the evolution of
the spectroscopic parameters of the fundamental state towards certain excited states.
The comparison of the spectra suggested the possibility of distributing the first eight
electrons of BO and CN around the two nuclei using two molecular orbits, the ninth
electron, more slightly bound, belonged to an orbit similar to that of the valence electron of
sodium. Mulliken compared his results with those of Sommerfeld (1934) and Langmuir
(1919) on the N2 and CO molecules. His assumption was that similar electronic distri-
butions must have corresponded to similar energy levels and vice versa. This assumption
enabled him to foresee in 1925, the existence of bands whose reality would soon thereafter
be confirmed by experiments (Mulliken 1926: 158–162). These results converged with
those obtained by Fowler in 1915 in connection with the molecule of He2 and with those of
Mecke, Bonn and Birge concerning H2. The analysis of the analogies made it possible for
Mulliken to classify the spectra according to the number of valence electrons of the

123
206 J.-P. Llored

molecules and the concept of electronic configuration became of the utmost importance to
describe molecules.
Inversions between electronic levels and other divergences as regards intensity and
shape of bands came to sow the disorder in the mind of Mulliken. He became aware at the
time of the need of another approach making it possible to explain the molecular electronic
structure. Thanks to this analogy however, Mulliken conceived molecular orbits in which
the electrons belonged to the whole of the molecule and not to a particular atom of the
molecule. Mulliken (1967: 13–24) wrote: ‘‘I regard each molecule as a self-sufficient unit
and not as a mere composite of atom’’.

Bond between atoms and molecules: the key role of Hund

Mulliken studied Friedrich Hund’s work (1926: 657–674) on the role of the spin of the
electron in the establishment of a chemical bond. In 1927, Hund suggested an approach
radically different from the work developed by Walter Heitler and Fritz London and
generalized the study of Oyvind Burrau to diatomic molecules. Rather than build a
molecular wave function from those describing isolated atoms, he proposed to describe
each electron in the total molecular electric field of the nuclei and other electrons. This
approach consisted in considering the molecule as an entity instead of describing it as an
aggregate of atoms. The electrons of a molecule were submitted to the attraction of all the
nuclei and the repulsions of the other electrons. Hund carried his reflexion on the evolution
of electronic energy during the transfer of an orbit around the joined nuclei to an orbit
around the separate atoms isolated from each other. On the basis of works developed by
Erwin Schrödinger, Pascual Jordan and Max Born, Hund (1927, 1974) was able to
describe, at the same time, the exact stationary states of the two subsystems knowing those
of the system by using linear combination. Thus, the new quantum theory allowed him to
explain the adiabatic passage between two stationary states of the same system. Hund
adapted this result for the study of the molecules and proposed an interpolation between
the quantum states of the isolated atoms, the united atom (a fictitious atom obtained by the
coalescence of the two atoms such as helium He for two hydrogen H atoms) and the
molecule. This approach highly stimulated Mulliken and reinforced his idea to work out a
molecular ‘‘Aufbauprinzip’’.

Molecule and united atom

Mulliken resumed work on the coupling between the angular orbital momentum and the
spin suggested by Hund. According to the nature of the coupling, he deduced the fine
structure of the spectra as well as the selection rules which govern the transitions. The
Stark effect involves the loss of atomic spherical symmetry and explains why the orbital
angular momentum ceased being a conservative parameter. On the other hand, projections
mL of this momentum on the axis of the electric field were preserved, provided that axial
symmetry remained.
The compounds hydrides of the type AH where ‘‘A’’ was an atom different from
hydrogen verified Hund’s assumptions very well. Mulliken (1932a) wished to go further,
he aimed at assigning to each electron of the molecule a quantum number characterizing it.
He supposed that the value of quantum number mL was preserved during the formation of
the molecule and carried out tables to indicate the orbits of each molecular electron. In fact,

123
Mereology and quantum chemistry: the approximation of molecular orbital 207

he was interested in the energy transformation of each electron during the formation of a
molecule, he made a systematic empirical study for the whole of the possible electronic
configurations on the basis of the principle of Pauli. He classified the spectral states and
made an analysis of the spectral terms, he determined electronic configurations corre-
sponding in each term. This quantitative approach enabled him to assume the existence of
electronic states, the experimental confirmations came with the improvement from the
resolution of the apparatuses.
A systematic classification arose in which the relevant quantum numbers and molecular
spectroscopic terms were rigorously identified (Mulliken, 1930, 1931). Mulliken could thus
explain the multiplicity of the bands which he had observed and was very quickly rec-
ognized as a specialist in molecular bands at the international level (Bloch 1930: 319).
However, he had to cope with difficulties when he treated the case of nonhydrogenated
molecules. The order of the electronic terms within the framework of Hund’s work was
very different from those observed on the molecular spectrum. Mulliken (1932a: 12–13)
contrasted this result to that which he obtained for the CH molecule, in which the order of
the molecular orbits was similar to that of the atomic orbits. He located the divergence of
outcomes at the level of the spilt of the core of the united atom. The question was then to
establish the context of use of the model of the united atom. The relation between molecule
and atoms seemed to become all the more complicated as Mulliken was also intrigued by
the study of the spectral analogies of molecules of halogens.
On the basis of empirical measurement of the energy of dissociation and the internu-
clear distance when equilibrium was reached, he proposed a series of new curves of
potential energy using the work of Philip Morse. The spectral terms and the identical
shapes of curves run up against the direct application of the method of the united atom. The
united atom of the molecule of chorine Cl2 is selenium which has 34 electrons and a
fundamental electronic configuration in KLM4s24p4 whereas Argon is the united atom of
the fluorine F2. Argon has 18 electrons and a fundamental electronic configuration in
KL3s23p6. Two atomic configurations of 4s24p4 type and 3s23p6 cannot lead to the same
molecular electronic terms. Mulliken proposed an explanation using the experimental data.
He compared energies of dissociation ‘‘D’’ and the internuclear distances ‘‘re’’ of F2, Cl2
and Br2 on the one hand and those of CO and N2 on the other hand. He noticed that F2, Cl2
and Br2 had a high internuclear distance and a weak energy of dissociation. CO and N2 by
comparison had a short internuclear distance and a high energy of dissociation. Because of
a weak energy of dissociation, F2 is closer to the two separate fluorine atoms than CO is to
carbon and oxygen. The spectrum of Cl2 is similar to that of F2 because the separate
chlorine atoms look like more the atoms of fluorine and this in spite of the differences
between the united argon or selenium atoms.
By comparing the data, Mulliken worked out a relative empirical scale thus gathering
the molecules formed by the elements of the first two rows of the periodic table. The
position of a molecule on this scale made it possible to estimate if it was closer to the
separate atoms than to the united atom. Mulliken worked to establish a wide range which
spread out between these two extremes. He defined a molecule as an interpolation between
the two extreme cases of the united atom on the one hand and the isolated constituent
atoms on the other one. The problems of the link between molecule and atoms moved from
the united atom to the separate atoms of the molecules. In addition, Buhm Soon Park (Park
2001: 185–189) notices the dissension of John Edward Lennard-Jones who highlighted the
limits of the work of Mulliken and Hund. Mulliken carried out a second analysis then by
considering the link between separate atoms and the molecule.

123
208 J.-P. Llored

Separate atoms and molecule: refusal of the aggregative concept of valence

Mulliken drew up an assessment on the distribution of the electrons in the separate atoms
and the molecule. In order to check the principle of Pauli and the conservation of a
quantum number appropriate to the study of the diatomic molecules, he proposed the
concept of electronic promotion. The quantum number of some electrons increases by a
unit during the formation of the molecule. Mulliken (1928: 186–189) focused his effort on
the understanding of the energy-bonding-power and introduced two terms, an electric term
of repulsion between the nuclei, noted N.E for nuclear energy and a term related to the
potential energy of an electron in the electric field of the nuclei and other electrons, noted
B.E for binding energy. The average internuclear distance ‘‘re’’ for which the chemical
bond takes place corresponds to a minimum of the total energy of the molecule.
Mulliken characterized the behaviour of each kind of electron—promoted or not pro-
moted—when the internuclear distance decreased. He proposed that the bonding energy of
an unpromoted electron must have increased in absolute value as the molecule was formed.
Indeed, the nuclear load perceived by the electron concentrated in a restricted volume
when the nuclei approached. The situation would have to be moderated for a promoted
electron, as a matter of fact if the nuclear density increased while nuclei were coming
together, the increase in the electric attraction which resulted between nuclei and electron
was partly compensated by the change of orbit. The increase in the principal quantum
number which characterized electronic promotion, caused the distance between the elec-
tron and the nuclei to increase. The bonding-power-energy of a promoted electron could
decrease or increase in absolute value. The form and the size of the orbit were decisive in
the evolution of the bonding-power-energy of the promoted electron. Generally, Mulliken
noted in experiments that if the initial principal quantum number of the promoted electron
was higher than one, bonding energy increased in absolute value if the orbit was known to
be more penetrating i.e. if the electron was closer to the nuclear zone.
Mulliken was not convinced by the concept of valence regarded as an intrinsic property
of the atom. He preferred to emphasize the binding capacity of an electron in a given
molecular orbit. He construed a continuous approach of the chemical bond by connecting
the concept of energy state deduced from the spectra to that of electronic configuration, i.e.
with the distribution of the molecular electrons in various orbits. In this description, each
orbit is delocalized over all the nuclei and could contribute, depending on each specific
case, a stabilizing or destabilizing energy contribution to the total energy of the molecule.
The sum of the energy contributions of each electron in its orbit determined whether the
electronic configuration allowed for the existence of a stable molecule, i.e., whether its
energy was stabilizing overall. For Mulliken, the atom did not exist anymore in a molecule.
His concept of molecular state suggested molecular variability of energy and geometry that
could not be considered within the approaches of Gilbert N. Lewis and Langmuir.
The idea of the binding capacity of an electron in a given orbit around the nuclei was
crucial in the evolution of Mulliken’s thought. Indeed, it made possible an energy approach
to the formation of a molecule understood as a whole and not as an aggregation of atoms.
The step of Mulliken made it possible to include the dynamics of the physical charac-
teristics of the molecule at the time of an electronic transition. An electron, populating after
transition an orbit in which its binding capacity became negative, could weaken or destroy
the molecule. This idea of succession of orbits and modulation of the binding capacity of
the electrons which belonged to them allowed one forecast the emergence of a molecular
‘‘Aufbauprinzip’’ which connected molecular electronic configuration and considerations
of energy.

123
Mereology and quantum chemistry: the approximation of molecular orbital 209

Towards the concept of a molecular ‘‘Aufbauprinzip’’: birth of a molecular holism?

The spectral analogies corresponded to identical molecular electronic configurations.


Mulliken (1932a: 78–83) made a very wide study of diatomic molecules and provided data
tables where the molecular experimental parameters and associated electronic configura-
tion appeared. He wrote (1932a: 13):
The fact that the four isoelectronic molecules BO, CN, CO?, N2? have the same type
of normal state is an example of the truth of the statement that the order of binding of
electrons is fairly definite within a set of similar molecules. The same statement
applies to the molecule CO, N2, NO?, and to NO, O2?. The fact that the normal state
of N2 and CO is obtained from that of NO?, merely by adding one more electron in
higher orbit, without making any changes in the quantum numbers of those already
present, is a further example of the applicability of the Aufbauprinzip in molecules.
The close similarities should also be noticed. These give further strong evidence that
the electronic structures of the molecules belonging to such a group (e.g., N2? and
CO?) can be understood in terms of a molecular Aufbauprinzip which is largely
independent of the particular atoms (e.g., N ? N? or C??O) which go to form them.
Mulliken studied a huge number of molecules gathered in various different series to
establish the general interest in his molecular ‘‘Aufbauprinzip’’. In doing so, he generalized
the use of the potential energy curves and pointed out that molecular electronic configu-
rations made it possible to give an account of experimental observations. Like Bohr for
atoms, he succeeded in finding laws to enlighten the electronic molecular distribution.
Another important point, the analogies between the molecules S2 and O2 enabled him to
explain the similar chemical properties of molecules composed by atoms of the same
group. He carried out his reflexion on the concept of valence of the molecules and com-
pared the molecular configurations of these two molecules deduced from the study of the
spectral states. He checked that the configurations were similar. This strengthened the
results which he had obtained for the molecules of halogens. The comparative study
between atoms and molecules allowed him to explain the inversion of some electronic
energy levels which had remained misunderstood since its discovery in 1927 by Jenkins, a
collaborator of Mulliken in Harvard. Mulliken reached an additional stage by studying the
bond between separate atoms and united atom thanks to new diagrams.

A new description of the molecule: composite and interpolation

Mulliken’s study was very well adapted to the hydrogenated diatomic molecules but was
irrelevant for the molecules of halogens and the heteronuclear diatomic molecules made up
of atoms belonging to the first two periods of periodic classification. The reflexion on the
promotion of some electrons and the concept of binding capacity were thus deepened. This
close connection between the quantum theory and spectral studies gave birth to the cor-
relation diagrams in 1932. Those diagrams made it possible to consider the degree of
likeness between a molecule and its separated atoms or its united atom thanks, in partic-
ular, to the empirical knowledge of the inter-nuclear distances and of the charges of the
nuclei. Mulliken specified their, primarily experimental, origin and his recourse to some
theoretical considerations in particular of symmetry and spin. He stressed that it was about
an approximate description insofar as the repulsions between electrons, as well as the spin-
orbital coupling between, were considered as negligible (Fig. 1).

123
210 J.-P. Llored

Fig. 1 Diagram of correlation in the case of homonuclear diatomic molecule (Mulliken: 1932a). Binding
scheme of electron orbits for molecules whose nuclei are of the same element. The abscissas (distance
between nuclei divide by mean diameter of electron orbit) and the ordinates (negative energy of ionization
of the electron) are not plotted on a uniform scale. The ordinates on the right are proportional to the
logarithms of the binding energies of electrons in the N atom, except for the 2s level which has been
arbitrarily shifted downward. Those on the left, except 2s, are proportional to a function of the logarithms of
the binding energies of electrons in the Si atom, which is the united-atom corresponding to N2. Thus the
diagram is specially suited to the N2 molecule, but is also satisfactory in a somewhat more qualitative way
for other molecules. The thin lines going upward and to the left indicate roughly for various molecules the
actual n values of various orbits, for stable molecular states. From these lines, the qualitative order of
binding of orbit-types in each molecule can be seen, and at least a rough idea can often be obtained of the
relative energy-differences between orbit-types (those comments are tailored from Mulliken (Mulliken:
1932a))

A preliminary idea of the relative variation of energy between orbits could be extracted
from these diagrams. Knowing the experimental characteristics of a molecule, it was
possible to position its curve in the diagram and to make predictions on energetic
parameters and the binding capacity of the electrons in these orbits. The molecule was a
composite understood as an interpolation between spectroscopic states regarded as limit
states. Mulliken (1931b: 350) wrote:
Although quantum mechanics has not yet reached the point of accounting in detail
for all the facts of valence, it does very definitely give the solution of the more
general problem of why it is that atoms are capable of forming molecules at all.
Quantum theory, following experiment, demands the existence of discrete stationary
states of energy, for molecules as well as for atoms. It shows further that in each such
stationary state the electrons may be thought of as moving in what used to and may
still with reservations, be called orbits. And finally it shows, in outline at least, how
when two or more atoms come together, the orbits of their electrons can be altered in
a perfectly continuous manner to give the appropriate electron orbits of the molecule.

123
Mereology and quantum chemistry: the approximation of molecular orbital 211

Buhm Soon Park (2001: 193) writes:


The correlation diagrams—visual representations of the Aufbau principle for mol-
ecules—allowed theory and experiment to be mediated, as tools for interpreting band
spectra and as ways of refining or redefining such concepts as molecule and valence.
The qualitative aspect of this study provided limited predictions, in spite of the sig-
nificant numbers of empirical data available. It appeared necessary to have an adequate
analytical representation of these orbits in order to determine their energies. In addition,
Mulliken wished to widen his work to polyatomic molecules. These two objectives led him
to conceive the concept of molecular orbital.

Molecular orbital and mereology

The criticism of the quantum work based on resonance: the composite against
the aggregate

For Mulliken, valence was first and foremost an energy problem, contrary to Heitler and
London who focused their study on electronic pairing. On March 31, 1931, Mulliken read a
communication during the 18th congress of American Chemical Society in Indianapolis,
which was the subject of a publication intended for chemists in Chemical Review. In this
paper, he supplemented the approach of Lewis for whom the electrons of the molecule
could be binding or not binding when they remained localized on the atoms. Mulliken
introduced the concept of anti-bonding electrons on the basis of his spectral concept of
electronic promotion. For Mulliken, electronic pairing, even if it was important, was not of
primary importance, the proof was the existence of the H2? molecule. According to him,
the concept of valence was arbitrary, the molecule was perfectly understood in terms of
molecular or atomic electronic configurations. He specified (Mulliken 1931b: 383) that:
‘‘we should regard a single bonding electron as the natural unit of bonding, an anti-bonding
electron as a negative unit’’.
For Heitler and London, the valence of atom corresponded to the number of unpaired
electrons which it contained. From this point of view, an atom such as the helium He
couldn’t form any molecule because its electronic shells were all saturated. The experiment
showed however that the molecules HeH and He2? existed. Mulliken explained this for-
mation by studying the molecular spectrum and proposed the assessment: He (1s2) ? He?
(1s) ? He2? (1sr22pr). The two unpromoted electrons ‘‘1sr2’’ had a bonding power more
important than the destabilizing effect of the electron promoted ‘‘2pr’’. He concluded
(1931b : 369) then:
In the ‘molecular’ point of view advanced here, the existence of the molecule as a
distinct individual built up of nuclei and electrons is emphasized, whereas according
to the usual atomic point of view the molecule is regarded as composed of atoms or
of ions held together by valence bonds. From the molecular point of view, it is a
matter of secondary importance to determine through what intermediate mechanism
(union of atoms or ions) the finished molecule is most conveniently reached. It is
really not necessary to think of valence bonds as existing in the molecule.
The composite exceeded the aggregate to explain the empirical facts related to observable
spectroscopic.

123
212 J.-P. Llored

This heuristic approach enabled Mulliken to be in the context of the time. He needed to
reach the analytical expression of the molecular wave function to determine molecular
energies. The need for a quantitative description made it possible for Mulliken to clarify its
attachment with quantum work and to use calculations related to the wave functions. This
work was completed in the series ‘‘Electronic structure of polyatomic molecules’’ in which
he introduced his concept of molecular orbital.

Concept of molecular orbital: Probabilistic approach of the composite

In the first paper of his series, Mulliken (1932b: 55) described some molecules and radicals
in terms of one-electron wave functions:
The electronic structures of polyatomic molecules can probably best be understood
by expressing them in terms of one-electron wave functions. The forms of these are
conditioned by the symmetry of the molecule, which is that given by the arrangement
of the nuclei.
This approach recalled Condon’s work which studied the molecular wave function of H2
starting from the mono electronic wave function of the ion H2?. According to Mulliken
(1932b: 57), the existence of stable molecules was related to energy considerations and the
existence of saturated shells:
Every nucleus in a molecule tends to be surrounded by an electron density distri-
bution corresponding to some stable electron configuration having a total charge
approximately equal to or somewhat exceeding the charge of the nucleus; the
electron density distribution as a whole, and the individual wave functions, have
symmetries adapted to the configuration of nuclei surrounding the given nucleus. By
‘stable configuration’ is meant a set of wave function completely occupied by
electrons (i.e., a set of closed shells) and of such type that further electrons could go
only into wave functions of distinctly higher energy,—usually of higher quantum
number, from the point of view of the central nucleus.
Mulliken explicitly referred to the concept of wave function and thought that the presence
of saturated shells within the molecules was only one quantum generalization of the usual
rules of Lewis and Langmuir. He added (1932b: 57):
Attempts to regard a molecule as consisting of specific atomic or ionic units held
together by discrete numbers of bonding electrons or electron-pairs are considered as
more or less meaningless, except as an approximation in special cases, or as a
method of calculation […]. A molecule is here regarded as a set of nuclei, around
each of which is grouped an electron configuration closely similar to that of a free
atom in an external field, except that the outer parts of the electron configurations
surrounding each nucleus usually belong, in part, jointly to two or more nuclei.
The change from the concept of molecular orbit to that of molecular orbital occurred in
1932. The concept of orbital took all its sense in Max Born’s probabilistic interpretation
that the square of a molecular orbital corresponded to the probability density of finding this
electron in space. In the second paper of this series, Mulliken (1932c: 50) wrote:
By an atomic orbital is meant an orbital corresponding to the motion of an electron in
the field of a single nucleus plus other electrons, while a molecular orbital corre-
sponds to the motion of an electron in the field of two or more nuclei plus other

123
Mereology and quantum chemistry: the approximation of molecular orbital 213

electrons. Both atomic and molecular orbitals may be thought of as defined in


accordance with the Hartree method of the self-consistent field, in order to allow so
far as possible for the effects of other electrons than the one whose orbital is under
consideration.
Mulliken immediately used this type of approximation to which Hund did not refer
explicitly. He took another degree of freedom compared to Hund’s work and added (1932c:
51):
In the present method, molecular orbitals are conceived of as entities quite inde-
pendent of atomic orbitals. Nevertheless in practice molecular orbital can usually be
conveniently approximated by building up linear combinations of orbitals of the
atomic type. The present method of thinking in terms of the finished molecule, used
already by Lewis in his valence theory, avoid the disputes and ambiguities, or the
necessity of using complicated linear combination, which arise if one thinks of
molecules as composed of definite atoms or ions.
He referred to Lennard-Jones’s work and aimed at developing it in the case of high
internuclear distances.
It is interesting to notice that at the time when Mulliken integrated more abstract
considerations into his language, he expressed the need to assert that he had never been as
close to Lewis and Langmuir. He thought that this probabilistic molecular orbital reading
explained their approach fully. Mulliken then considered the study of the polyatomic
molecules that group theory and the work of Lennard-Jones had made easier.

Molecular orbital and group theory: Is the relation between the whole and its parts
arbitrary?

Mulliken used the principle of Wigner to determine a classification of the values of


molecular energy and adapted the method of symmetry developed by Bethe within the
framework of solids to that of the molecules. He worked out a method making it possible to
characterize the various electronic states (1933: 279–280):
For a molecule with fixed nuclei, the complete electronic wave function w is
restricted to one of certain types which depend on the symmetry of the nuclear
skeleton. In the language of group theory, w must conform to an irreducible repre-
sentation of the symmetry group of the corresponding Schrödinger equation,—which
contains a potential energy whose symmetry is that of the nuclear skeleton—. Or
more briefly, one may say that every w must belong to an irreducible representation
of the symmetry group of the nuclear skeleton. Corresponding statements apply to
every molecular orbital u. In nature w is of course further restricted, in accordance
with the Pauli principle, to form antisymmetrical in the electrons.
He construed an application of the theory by drawing up irreducible representations for
thirty-two specific symmetry groups. He developed the fragment method in 1933, two
fragments could interact provided they had the same kind of symmetry and that the energy
gap, measured by spectroscopy, was not too high. For the ethylene molecule ‘‘C2H4’’
Mulliken considered two fragments ‘‘CH2’’ and determined suitable molecular orbital by
using the irreducible representations of ethylene. He could thus propose a representation of
molecular orbital of ethylene by increasing order of energy as well as its correlation
diagram thanks to those of the two fragments. In doing so, he grasped all the characteristics

123
214 J.-P. Llored

of molecular orbital diagram of the ethylene molecule. Mulliken (1932d: 754) was thus
able to understand the molecule in terms of a description using molecular orbital. The
correlation diagrams enabled him to make the synthesis between spectral data and the types
of suitable molecular orbital, it was a powerful method to predict molecular reactivity
(Jean and Volatron 1991; Walton 1998; Rivail 1989; Cotton 1971).
Mulliken generalized his study with aldehydes, ketones, alkanes, olefinic hydrocarbons
and alkyne compounds like with the borated molecules such as B2H6 that the method of
Pauling and Slater could not explain. He thus succeeded in explaining chemical bonds
without using any pairing of electrons. In the articles V, VII and VIII of his series, he made
predictions on the ionization potentials starting from the correlation diagrams which were
in agreement with experiments. Whereas he worked on the molecules of halogens, Mul-
liken compared the experimental values of the ionization potentials with those envisaged
by the molecular orbital approach. The ionization potential of a bonding electron was
higher than the atomic ionization potentials. The experiment confirmed his predicts and
also showed that the ionization potential of an anti-bonding electron was lower than the
atomic potentials. This is an example of what Earley calls a co-operative property because
the ionization potential of the molecule does not result linearly from those of the isolated
atoms.
Mulliken wrote two papers (1942a, 1942b) and, proposed with Walsh, new diagrams
which made it possible to envisage the most stable geometry of a molecule. Those dia-
grams established the correlation between the molecular orbital built starting from two
extreme geometries and classified by order of increasing energy. The global solution
enabled them to explain the solid geometry of the molecular whole without referring to an
assembly of local geometries.
Mulliken was interested in cyclic molecules having an alternation of simple and double
chemical bonds. Calculations of spectral radiant intensities and the structural predictions
that he proposed starting from group theory were confirmed by experiment. He built
molecular orbitals for these cyclic molecules starting from fragments such as a polyene and
a grouping of the type CH2, O, NH, S. The measurements of index of refraction and
spectral radiant intensities rooted deeply the relevance of his method. The approach of
molecular orbital made it possible to corroborate the colouring of cyclic or linear alternate
polyenes, which was not explained by any colour of their ‘‘parts’’. It really seems that
chemistry is ‘‘the embodiment of Emergence’’ as Pier Luigi Luisi (2002: 183–200) asserts
it.
Brown, Riecke and Mulliken explained the effects of orientation of each substituent
during chemical reactions on aromatic compounds called ‘‘halides of benzyl’’. They
showed that a grouping methyl ‘‘CH3’’ had a molecular orbital of close energy and suitable
symmetry to interact with the molecular orbital of benzene. The replacement of one
substituent of the aromatic nucleus which did not have this property by the grouping
methyl modified the electronic structure of the building and the chemical reactivity of the
latter. This method called hyperconjugation allowed a simple explanation with a strong
heuristic nature. To some extent, it was an approximate method that allowed researchers to
have simple molecular outcomes while avoiding the use of methods of calculation, more
satisfactory on the theoretical level, but which either diluted the experimental observation
in a multitude of ways, or proposed an approximate result whose only comparison with the
experiment remained possible. The hyperconjugation explained the energy stabilization of
a molecule starting from the interactions between the various fragments which made it up.
This stabilization is a co-operative property in the sense that it is not linearly foreseeable
starting from the properties of the various parts.

123
Mereology and quantum chemistry: the approximation of molecular orbital 215

The use of Wigner’s principle led to the creation of diagrams of molecular energy
corroborated by the practical analysis of electronic molecular spectra. Molecular spec-
troscopy crossed quantum theory to give access to a molecule understood as a whole. The
possibility of an experimental support was all the more important as the nature of the initial
fragments could change depending on each specific case. To model the molecule C2H2,
Mulliken could just as easily have considered a fragment ‘‘C2’’ and another ‘‘H4’’ of
adapted symmetries. The relation on whole ‘‘C2H2’’ with its parts was of secondary
interest. G.K. Vemulapalli (2003: 97) writes: ‘‘Because different initial states may be
chosen in building molecular states and the initial states need not to be atomic states, there
is arbitrariness in designating the parts’’. This example shows that the choice of the parts is
arbitrary whereas the energy diagrams of the ‘‘molecular whole’’ are in agreement with the
experiment, the holist approach is thus reinforced. Mulliken generalized it thanks to the
work of Lennard-Jones in order to envisage the molecular properties more simply.

Orbital molecular and LCAO approach: a reduction of the whole to its parts?

Mulliken (1933: 281) wrote:


Following a method first used by Bloch for metals and later used by Hückel, Hund,
and others, molecular orbitals will as a matter of convenience usually be approxi-
mated here by linear combinations of atomics orbitals, although eventually we may
hope to obtain forms which are better approximations.
The underlying idea of Lennard-Jones was that when an electron in the molecule
approached a given nucleus, the forces which acted on it were mainly originating in this
nucleus. It was thus possible to consider that the molecular orbital, near a given nucleus,
was very close to the atomic orbital which described the behaviour of the electron around
this nucleus. The molecular orbital could, in this context of internuclear distance, be
approximated by a linear combination of atomic orbitals. The wave function was thus
written:
U ¼ k1 w þ k2 v
k1 and k2 were coefficients determined by the variation method. U was the molecular
orbital, w and v were atomic orbitals.
In 1935, Mulliken extended the method of Lennard-Jones to the polyatomic molecules
giving way to the LCAO approximation i.e. ‘‘Linear Combinaison of Atomic Orbitals’’.
Within the framework of this approximation, the concepts of bonding and anti-bonding
electrons became much more convenient to grasp for the chemists. In the case of diatomic
molecules, the bonding electrons were described by a molecular orbital resulting from the
sum of the two atomic wave functions while the difference in these wave functions made it
possible to obtain anti-bonding molecular orbital. In fact, the ‘‘positive or negative’’
binding capacity of molecular orbitals depended on the additive or subtractive form of the
linear combinations on atomic orbitals and not of the state of promotion of the electrons.
A problem arose for the choice of molecular orbitals. As a matter of fact, Hund showed
that two-centered orbitals could also be used in the case of polyatomic molecules while
Mulliken used delocalized orbitals. For Mulliken, ‘‘the best molecular orbital’’ was that
which described with the most possible accuracy an electron influenced by the average
electric field of the other electrons. Within the context of diatomic molecules, the
molecular orbital spanned the entire extension of the two nuclei. In the case of the

123
216 J.-P. Llored

polyatomic molecules, the molecular orbitals were more or less localized but as Mulliken
(1935: 375) underlined:
Also, while fully non-localised or ‘best’ MOs which spread at least to some slight
extent over all atoms, give the most accurate electronic structure description, we can
arbitrarily impose various kinds of transformations and constraints to obtain useful
approximate localised MO descriptions which correlate instructively with the older
valence theory.
Hund (1932) established the equivalence of localized and delocalized orbitals. Under
certain conditions, some cases could be described as localized and/or delocalized, others
only localized—carbon diamond—, and others only delocalized such as metals and
aromatic compounds. Mulliken (1970) added: ‘‘I believe that the placing of two electrons
in such a localized MO represents the best simple quantum-mechanical counterpart for
Lewis electron pair bond’’. This descriptive plurality of the molecular wave functions
qualified the relation between the ‘‘molecular whole’’ and its parts.
Mulliken brought a new contribution to LCAO approximation in a new series entitled
‘‘Electronic population analysis one LCAO-MO molecular wave functions’’ published in
1955 in Chemical Physics. He developed in the four articles of this series, a method
making it possible to determine overlapping integrals, the covalent bonding energy, the
order of connection—the number of connections between two given atoms—and the
electronic populations on a given atom. The scale of electro negativity which he worked
out in 1935 enabled him to provide that the molecule of lithium hydride had a higher
electronic density around the nucleus of lithium. By reasoning for an electron pertaining to
bonding molecular orbitals, he showed that a fraction equal to 0, 111 of this load belonged
to hydrogen whereas a fraction equal to 0, 641 belonged to lithium, this outcome was
coherent insofar as lithium attracted more the electrons than hydrogen. The last fraction
equal to 0, 248 belonged to the two atoms simultaneously. This holist approach to the
molecule was consolidated by this total distribution of the electrons in keeping with
experiments on molecular reactivity.
At the same time, Mulliken studied some load transfer complexes by means of LCAO
calculations. He showed that in its electronic fundamental state, the complex could be
described by a wave function based on the two nondependent species but that there was
however a connection between the donor and the acceptor in an excited state. As a result,
the wave function of the whole was written as a linear combination of terms dealing with
those two different quantum states. This work was confirmed quantitatively by the chemist
J.N. Murrel (1963) in 1949 thanks to a study of the intensity of the absorption bands.
Indeed, Murrel connected this spectral radiant intensity to the stability of the complex, i.e.,
with the strength of the connection between the donor and the acceptor. This spectral
property of the complex is also co-operative insofar as it is not expressed by the electrons
of the donor and the acceptor and unpredictable from them.
This linear combination can suggest that a molecule is reducible to its constitutive
atoms. G. K. Vemulapalli (2003: 95) notices that:
While properties of the whole are not the sums or products of the properties of
parts—in our case different spectral properties—, the states of the system can be
obtained by adding the states of parts. Because properties in turn can be derived from
the states, it appears that we have shown that properties of wholes are completely
determined by parts. But there are two problems here: (1) It is true that the states of
the system are composed of states of the parts, but there are also weighting factors in

123
Mereology and quantum chemistry: the approximation of molecular orbital 217

the composition. There are the constants k in the linear combination. What factors
determine these constants? (2) Just as in the molecular wave function, an atomic
wave function may also be represented by a sum of an arbitrary set of functions. Thus
one may claim that an atomic function is a linear combination of molecular functions
or atomic states (parts) reduced to molecular states (wholes!).
The arbitrary character of the relation between the whole and its parts is once again
highlighted and remain more than ever present in current semi-empirical or ab initio
methods of molecular orbital calculation that depend on the choice of atomic or molecular
orbital used. The fundamental choice relates to the nature and the extent of the basis.
Vemulapalli clarifies the role of the coefficient weighting appearing in front of the orbitals
of the key basis. These coefficients determined by the variation method are those which
minimize the molecular potential energy. How is this minimum of energy justified?
Vemulapalli refers to the second law of thermodynamics to explain why the studied
molecular system continuously eliminates its excess energy by interactions with its envi-
ronment. An energy transformation into local entropy returns to legitimate the use of the
variation method. Vemulapalli (2003: 97) adds:
Thus we are led to conclude that it doesn’t matter what the states of the parts are, but
it does matter the surroundings soak up the excess energy of the molecule, increasing
entropy, and make the molecule settle down into the lowest energy state. It is that
part of the universe coupled to the system, and the varieties of interactions between
the system (molecules) and the surroundings that determines the structure of the
molecule. Holism thus appears as the root of the apparent reduction of properties of a
molecule to its parts through coupling states. We are able to follow a reductionist
program in calculating molecular properties, but what we are able to do is a gift of
holism.
This approach considerably qualifies the relation between a whole and its parts within
the framework of LCAO approach. This fine analysis avoids the often inappropriate
alternative between a reductionist or an emergentist interpretation and enters the state of
mind of a search for an ecology of the practices such as Isabelle Stengers (2003) develops.
In addition, the analysis in terms of electronic populations shows that each atom has lost its
identity insofar as the distribution of the electrons does not correspond to those of the
atoms taken separately. The conclusion that a molecule is reducible to its atoms because of
linear combinations is therefore too hasty. Moreover, understanding of the relation between
a whole and its parts is reached if we consider the practices of approximation, the chemical
practical knowledge which gives rise to those approximations and the use of principles
which do not belong to Quantum theory such as the Pauli Principle, the second law of
thermodynamics and others.

Conclusion

Did Mulliken believe or not believe in the existence of the atoms? This is not the question.
The atoms of Mulliken are used to connect some principles with phenomenal relations
expressed in the form of laws—empirical spectroscopic laws, selection rules, principle of
Pauli—and to make them understandable by means of an overall diagram, of a ‘‘conceptual
scheme’’ according to his turn of phrase.

123
218 J.-P. Llored

Ernst Cassirer, neo-Kantian philosopher of the school of Marburg, asserted that the
atom is never a starting postulate but rather a point of arrival of the statements of the
chemists2: ‘‘It is seemingly only that such or such property is then attached to the atom as
with its ‘‘absolute support’’, within a unit which appears to control and solidify the whole
of the relation’’. The concept of atom is a tool which concentrates and makes it possible to
unify multiple relations between empirical data, the Mulliken’s correlation diagrams are a
fine example of this unifying capacity. The atom is not a ‘‘thing’’, it just makes it possible
to federate knowledge (Cassirer 1977: 242)3:
The positive benefit that chemical knowledge gains there consists in the systematic
articulation of the ratios themselves—here the relationship between quantum num-
bers according to selection rules or between the various spectral phenomenological
laws—. The initially dispersed facts, start from now on to be organized; instead of
coexisting in indifference, they are ordered around a center of precise reference.
But Cassirer went further, Cassirer (1977: 242–243) claimed4:
It is clear however that this ‘‘subject’’—the atom—does not have only logical value
to provide afterwards the description and the convergence of the recorded experi-
ments. The unification thus construed plays an immediately productive role; it sets up
a total diagram applicable to the future observations and assigns to them a given
direction.
The correlation diagrams of Mulliken carry out the synthesis of the experimental data and
theoretical models thanks to the mediator concept of atom. These diagrams allow a great
number of predictions not only about spectral molecular states but also as regards their
physical properties. They make it possible moreover to study the formation of molecules
without alluding to a hypothetical ‘‘intrinsic’’ valence of atoms.
Mulliken went further with his concept of ‘‘binding capacity’’ of electrons. He pointed
out the process of molecular dynamics and tried to rationalize molecular reactivity. The
heuristic character of the explanations which he proposed is undeniable but this is not all.
For Mulliken, electrons existed because a scientist can act on them via the electromagnetic
waves. He reflected on the relational capacity which saw an electron interacting with
various nuclei in a molecular orbit. Ian Hacking’s work makes it possible to clarify the step
of Mulliken. As a matter of fact, Hacking (1983: 242) asserts:
The entities of which we suppose reality must be capable of a causal effect on the
real things took prima facie, i.e. over the material things of ordinary size. […] We
should count as real all that we can use to intervene in the world so as to affect
something, or all that the world can use to affect us.

2
I have translated from: E. Cassirer, ‘‘Substance et fonction. E´le´ments pour une the´orie du concept’’,
traduction française de Pierre Caussat, Les éditions de minuit, Paris, 236–254 (1977). ‘‘C’est en apparence
seulement que telle ou telle propriété se trouve alors rattachée à l’atome comme à son «support» absolu, au
sein d’un ensemble qui paraı̂t asservir et figer le tout de la relation’’.
3
I have translated: ‘‘Le bénéfice positif que la connaissance chimique y gagne consiste dans l’articulation
systématique des rapports eux-mêmes. Les faits d’abord dispersés, commencent désormais à être organisés;
au lieu de coexister dans l’indifférence, ils s’ordonnent autour d’un centre de référence précis’’.
4
I have translated: ‘‘Il est clair toutefois que ce «sujet» n’a pas pour seule valeur logique de pourvoir après
coup à la description et à la convergence des expériences enregistrées. L’unification ainsi instituée joue un
rôle immédiatement productif; elle met en place un schéma global applicable aux observations futures et
leur assigne une direction déterminée’’.

123
Mereology and quantum chemistry: the approximation of molecular orbital 219

If Mulliken designed the atom as a unifying concept, he acted on the electrons by the
means of the spectroscopy. He did not believe in electrons because he was looking for a
theory of the structure of the matter but because he acted on them and because electrons
caused actions on the phenomena—lengthening of the internuclear distances, change of the
angles of connection, evolution of energies of dissociation and so on. Mulliken sought
causal capacities at the origin of the molecular phenomena. He tried to quantify this
binding capacity via many spectral studies. We stressed that he proposed many quantitative
diagrams and data tables to explain and envisage molecular properties and the formation of
molecules.
With the thread of his work, Mulliken construed new concepts and tools like promotion
of electrons, the binding capacity, correlation diagrams, the molecular orbital and others.
These notions resulted from the blending of the experimental data obtained by the chemists
with the quantum theory. These concepts are deprived of any ontological claim, they make
it less possible to describe the real architecture of the molecules, to classify them, analyze
and envisage their formation and their physical and chemical properties. These concepts
are ‘‘paper tools’’, according to the expression of Ursula Klein (2001), which enabled
Mulliken to propose an energy approach of the molecule. This heuristic approach can be
enlightened by the comments of the French chemist of the nineteenth century, Jean-
baptiste Dumas (1837, 1972)5:
There exists between the current chemists and the former chemists something of
common run; it is the method. And which is this method, old like our science itself,
and which is characterized as of his cradle? It is the most complete faith in the
testimony of senses; it is a belief without terminal granted to the experiment; it is a
blind submission with the power of the facts. Old or modern, the chemists want to
see with the eyes of the body before using those of the spirit: they want to make
theories for the facts, and not to seek facts for the preconceived theories.
Mulliken introduced the concept of ‘‘bonding power’’ to indicate the binding ability of
an electron in a molecular orbital. This concept of being able is not without recalling that
of ‘‘power’’ developed by Aristotle or that of ‘‘affinity’’ proposed by the energetists in the
nineteenth century. Mulliken sought a causal capacity which worked in the chemical
phenomena and suspected at that time that this ‘‘binding capacity’’ was related to the
stability and the reactivity of molecules i.e. on their capacity to act on other molecules. He
tried to measure the capacity of the electrons to be put in relation with nuclei or electrons
of other molecules to produce chemical phenomena. In this respect, his approach was
fundamentally non-essentialist. Nancy Cartwright’s work (1989) on the reality of the
capacities of Nature allows a second reading of the scientific adventure of Mulliken in the
light of an operational realism. The spectroscopy makes it possible for Mulliken to eval-
uate this reactive capacity of electrons and to propose tables to foresee the properties of
molecules starting from analogies made between atomic and molecular spectral states. In
addition, this capacity to interact with the external world authorizes a reduction within the
framework of approximation LCAO but a possible reduction on the basis of holism. It is

5
I have translated from: J.B. Dumas (1837), ‘‘Leçons sur la philosophie chimique’’, Paris, rééd. Bruxelles,
Culture et civilisation (1972). ‘‘Il existe entre les chimistes actuels et les anciens chimistes quelque chose de
commun; c’est la me´thode. Et quelle est cette me´thode, vieille comme notre science elle-meˆme, et qui se
caracte´rise de`s son berceau? C’est la foi la plus comple`te dans le te´moignage des sens; c’est une croyance
sans borne accorde´e à l’expérience; c’est une aveugle soumission à la puissance des faits. Anciens ou
modernes, les chimistes veulent voir avec les yeux du corps avant d’employer ceux de l’esprit: ils veulent
faire des the´ories pour les faits, et non chercher des faits pour les the´ories pre´conçues’’.

123
220 J.-P. Llored

necessary furthermore to consider the practice of research from which those concepts
originate, i.e., the technoscientific seam which combines quantum mechanics, approxi-
mations (Scerri 2007), chemical knowledge, instrumental and algorithmic techniques. The
predictive capacity of these approaches does not only depend on the molecular wave
function but also on a host of approximations and compromises that make it possible for
numerical properties and molecular landscapes to be calculated.
A mereological study must take account of the interaction of the whole with its envi-
ronment and should not artificially isolate it from the external world. The study of a
mereology of the chemical systems that Earley wishes derives all its importance from the
chemical entities which are defined by their capacities to act on their surroundings. In his
book Rameaux, the French philosopher Michel Serres (2004) seeking to establish a phi-
losophy of relationality accounts: ‘‘The relation precedes existence’’. For even thinking of
the relation between a molecule and its parts, it is necessary to take account of the
capacities of this molecule to act on the external world. This capacity to enter in relation
makes possible the study of the chemical properties but also accounts for the molecular
form i.e. of the internal relation, subjected to randomness, between the molecule and its
‘‘parts’’. In addition, the striking analogy between chemical properties and quantum
observables is a means to pave the way of a new mereological approach of chemical
systems. In our introduction, we put the questions: Are there atoms in a molecule? Does
this question admit an answer similar to that brought by Earley concerning salt in sea
water? In the light of our arguments, it seems to us that Mulliken and Earley speak in the
same way.

Acknowledgments I warmly thank Isabelle Stengers, Rom Harré, Eric Scerri and Michel Bitbol for their
advice, their second readings of this work and above all for their deep humanism.

References

Bensaude-Vincent B.: Faut-il avoir peur de la chimie?, Les empêcheurs de penser en rond éditeurs, Paris
(2005)
Bensaude-Vincent B.: Le mixte: un défi au tout comme somme des parties In: Le tout et les parties dans les
systèmes naturels. Under the direction of Thierry Martin, Collection Philosophie des sciences, Vuibert,
Paris (2007)
Bloch L.: Introduction à l’étude des spectres de bandes et de la constitution des molécules, Annals of the
Institut Henri Poincaré, vol 319, Issue 4, Paris (1930)
Cartwright, N.: Nature’s Capacities and Their Measurement. Oxford University Press, Oxford (1989)
Cassirer E.: Substance et fonction. Éléments pour une théorie du concept, traduction française de Pierre
Caussat, pp. 236–254. Les éditions de minuit, Paris (1977)
Cotton, A.: Chemical Applications of Group Theory. Wiley-Interscience, New York (1971)
Duhem, P.: Le mixte et la combinaison chimique. Essai sur l’évolution d’une idée, Fayard, Paris (1985)
Dumas J.B. Leçons sur la philosophie chimique, Paris, rééd. Bruxelles, Culture et civilisation (1972) (1837)
Earley J.E.: Varieties of properties. An alternative distinction among qualities In: Earley J.E. (ed.) Chemical
Explanation. Characteristics, Development, Autonomy. Annals of the New York Academy of sciences
vol 988, Issue 1, pp. 80–89. The New York Academy of Science Publisher, New York (2003)
Earley, J.E.: Why there is no salt in the sea. Found. Chem. 7, 85–102 (2005)
Hacking, I.: Representing and Intervening. Cambridge University Press, Cambridge (1983)
Hund, F.: Zur Deutung einiger Erscheinungen in den Molekelspektren. Zeitschrift für Physik 36, 657–674
(1926)
Hund F.: On the Interpretation of Molecular Spectra I, Zeitschrift für Physik 40. (1927) (Reproduced in
(2000): H. Hettema, Quantum Chemistry: Classic Scientific Papers, World Scientific Publishing, 226)
Hund F.: Zeitschrift für Physik. 74, 1 (1932)
Hund F.: The History of Quantum Theory. Harrap London. Geschichte der Quantentheorie (1967), trans-
lated by Georges G. Harrap & Co. Ltd, London, 187 (1974)

123
Mereology and quantum chemistry: the approximation of molecular orbital 221

Jean Y., Volatron F.: Les orbitales moléculaires en chimie. McGraw-Hill, Paris. (1991)
Klein, U.: Tools and Modes of Representation in the Laboratory Sciences. Kluwer Academic Publishers,
Dordrecht (2001)
Langmuir, I.: The arrangement of electrons in atoms and molecules. J. Am. Chem. Soc. 41, 868–934 (1919)
Mulliken, R.S.: Systematic relations between electronic structure and band-spectrum structure in diatomic
molecules I. Proc. Nat. Acad. Sci. USA. 12, 158–162 (1926)
Mulliken R.S.: The assigment of quantum numbers for electrons in molecules. I. Rev. Mod. Phys. 32,
186–189 (1928)
Mulliken R.S.: Interpretation of Band Spectra, Part I, IIa, IIb. Rev. Mod. Phys. 2, 60–70 (1930)
Mulliken R.S.: Interpretation of band spectra, part IIc, empirical band types. Rev. Mod. Phys. 1, 123 (1931a)
Mulliken R.S.: Bonding power of electrons and theory of valence. Chem. Rev. 9, 350 (1931b)
Mulliken R.S. Interpretation of band spectra, part III. Electron quantum numbers and states of molecules and
their atoms. Rev. Mod. Phys. 4, 1–86 (1932a)
Mulliken R.S.: Electronic structures of polyatomic molecules and valence I. Rev. Mod. Phys. 40, 55–57
(1932b)
Mulliken R.S.: Electronic structures of polyatomic molecules and valence II. General consideration. Phys.
Rev. 41, 49–71 (1932c)
Mulliken R.S.: Electronic structures of polyatomic molecules and valence III. Quantum theory of the double
bond. Rev. Mod. Phys. 41, 754 (1932d)
Mulliken R.S.: Electronic structures of polyatomic molecules and valence IV. Electronic states, quantum
theory of the double bond. Rev. Mod. Phys. 43, 279–280 (1933)
Mulliken R. S.: Electronic structures of polyatomic molecules and valence VI. On the method of molecular
orbitals. J. Chem. Phys. 3, 375 (1935)
Mulliken, R.S.: Structure and ultraviolet spectra of ethylene, butadiene, and their alkyl derivatives. Rev.
Mod. Phys. 14, 265–274 (1942a)
Mulliken, R.S.: Electronic structures and spectra of triatomic oxide molecules. Rev. Mod. Phys. 14, 204–215
(1942b)
Mulliken R.S.: Spectroscopy, molecular orbital and chemical bonding (Nobel lecture), Science 157 (1967)
Mulliken, R.S.: The path to molecular orbital. Pure Appl. Chem. 24, 203–215 (1970)
Murrell, J.N.: The Theory of the Electronic Spectra of Organic Molecules, pp. 270–283. Chapman and Hall,
London (1963)
Needham P.: Chemical substances and intensive properties. In: Earley J.E. (ed.) Chemical Explanation.
Characteristics, Development, Autonomy. Annals of the New York Academy of sciences 988, 1 (2003)
Park B.S.: A principle written in diagrams: The Aufbau principle for molecules and its visual representa-
tions. In: Klein U. (ed.) Tools and Modes of Representation in the Laboratory Sciences. Kluwer
Academic Publishers, Dordrecht (2001)
Luisi P.L.: Emergence in chemistry: Chemistry as the embodiment of Emergence. Found. Chem. 2, 183–200
(2002)
Rivail J.-L.: Eléments de chimie quantique à l’usage des chimistes. Savoirs Actuels, InterEditions/Editions
CNRS, Paris (1989)
Scerri, E.R.: The ambiguity of reduction. HYLE 13(2), 67–81 (2007)
Serres M.: Rameaux. Le Pommier, Paris (2004)
Stengers I.: La vie et l’artifice: visages de l’émergence In: Cosmopolitiques II, La découverte, Paris (2003)
Sommerfeld A.: Atomic Structure and Spectral Lines, translated by H. L. Brose, Methuen, London (1934)
Vemulapalli G.K.: Property reduction in chemistry. Some lessons In: Earley J.E. (ed.) Chemical Explana-
tion. Characteristics, Development, Autonomy. Annals of the New York Academy of sciences, 988
(2003)
Walton, P.H.: Beginning Group Theory for Chemists, Workbooks in Chemistry. Oxford University Press,
Oxford (1998)
Alfred North Whitehead: Adventures of Ideas. Macmillan, New York (1967)

123
View publication stats

Potrebbero piacerti anche