You are on page 1of 46

24th International Symposium of the ISPC

Buenos Aires, Argentina. July 2021

Organizing Committee: Buenos Aires Group for the Philosophy of Chemistry

Members of the organizing Committee: Olimpia Lombardi (Director of the group), Hernán
Accorinti, Fiorela Alassia, Sebastian Fortin, Jesús Jaimes Arriaga, Juan Camilo Martínez
González, and Alfio Zambon.

With the academic support of:


CONICET (Consejo Nacional de Investigaciones Científicas y Técnicas)
UBA (University of Buenos Aires)
UA (Austral University)
About the International Society for the Philosophy of Chemistry

The International Society for the Philosophy of Chemistry (ISPC) is devoted to the international
exchange of ideas concerning the philosophical foundations of the chemical sciences and related
areas. This exchange fosters discourse between chemists, biochemists, philosophers, historians,
sociologists and educators.
Philosophy of chemistry concerns both internal questions arising from the methods, concepts, and
ontology specific to chemistry and chemical research, as well as traditional questions in the
philosophy of science, addressed from a chemical perspective. Although philosophy of chemistry
is a reflecting discipline in the first place, it is also considered to contribute constructively to the
image of chemistry in science, education, and society.

ISPC President:
Rom Harré† (Honorary President), Georgetown University, USA/ Linacre College,
Oxford University, UK

ISPC Executive Committee:


Michael Akeroyd, Bradford College, UK
Marina Banchetti, Florida Atlantic University, USA
Elena Ghibaudi, University of Torino, Italy
Robin Hendry, University of Durham, UK
Olimpia Lombardi, Universidad de Buenos Aires, Argentina
Guillermo Restrepo, Max Planck Institute for Mathematics in the Sciences, Germany
Klaus Ruthenberg, Hochschule Coburg, Germany
Eric Scerri, University of California, Los Angeles, USA

Secretary:
Juan Camilo Martínez González, Universidad de Buenos Aires, CONICET Argentina
Foundations of Chemistry (Journal)

Foundations of Chemistry (FOCH) is the official journal of the International Society for the
Philosophy of Chemistry. FOCH is an international journal in which conceptual and fundamental
issues are discussed in their relation to the ‘central science’ of chemistry.
FOCH publishes peer-reviewed articles on a wide range of topics, including chemical models,
language, metaphors, and theoretical terms; chemical evolution and artificial self-replication; the
social and ethical aspects of chemistry's professionalism; modelling and instrumentation in
chemistry; the nature of explanation in the chemical sciences; theoretical chemistry, molecular
structure and chaos; ancient chemistry, medieval chemistry and alchemy; philosophical and
historical articles; and more. Foundations of Chemistry offers themed issues, and includes book
reviews and discussion notes.
INDEX of AUTHORS and TALKS

1. Alassia, Fiorela and Córdoba, Mariana: “A processual ontology approach in biochemistry and
the review of the «structure-activity-relationships» dogma. The case of drug receptors.”
2. Banchetti-Robino, Marina: “Robert Boyle and the dispositional and relational nature of
chemical properties.”
3. Chamizo, José Antonio: “Transformations. About novelty and persistence in the history of
chemistry.”
4. Deichmann, Ute: “The rule of simplicity and template theories in biology and chemistry - The
example of Pauling’s questionable research on antibody formation.”
5. Friend, Michèle: “Comparing production of biofuel to oil.”
6. García Zerecero, Gabriela: “The unitary and substantial character of molecular systems.”
7. Ghibaudi, Elena and Branchini, Federica: “How epistemology may inform chemical
education: the case of the notion of chemical bonding.”
8. Headley, Clevis: “Natural kinds, chemical practice, and interpretive communities”
9. Hendry, Robin Findlay: TBA
10. Hijmans, Sarah: “The many fathers of aluminium: Collaboration, credit and the construction
of discovery.”
11. Jaimes Arriaga, Jesus Alberto: “The powerfulness and qualitativeness of electron density.”
12. Llored, Jean-Pierre: “The elimination of the holism-reductionism dichotomy through the
analysis: of quantum chemistry.”
13. Lombardi, Olimpia and Fortin, Sebastian: “The problema of isomerism.”
14. Maeno, Yoshiteru: “3-dimensional atomic and nuclear periodic tables.”
15. Martínez González, Juan Camilo: “The magnetic theory of electrons and the cubic model of
the atom.”
16. Matta, Cherif, Anderson, James S. M., and Massa, Lou: “Non-nuclear maxima and the
universality of Bright Wilson’s justification of the First Hohenberg-Kohn Theorem.”
17. Mocellin, Ronei Clécio: “Thinking with your hands: On the mode of existence of chemical
objects in the Enlightenment century.”
18. Ochiai, Hirofumi: “«Phenomenal field» provides philosophical grounds for designing
invisible structure of molecules.”
19. Pastoriza, Bruno, Dutra da Silva, Fernanda Karolaine, da Silva Rodrigues, Tavane, Bardini,
Laura, and Brahm dos Santos, Guilherme: “Fundamental concepts in chemistry education: A
problematization.”
20. Pérgola, Martín and Galagovsky, Lydia: “Modeling cellular respiration. Educational
reflections.”
21. Renier, Mike L.: “History and philosophy of chemical modeling.”
22. Restrepo, Guillermo and Jost, Jürgen: “The evolution of chemical knowledge: How to study
it.”
23. Scerri, Eric. “Hasok Chang on acid”
24. Seifert, Vanessa and Franklin, Alexander: “The problem of molecular structure just is the
measurement problem.”
25. Srivastava, Alok: “Plotting syntheses of rubber and of DNA: How chemists use chemical
structures and reaction pathways to craft new chemistries.”
26. Sukumar, N.: “Mereology of chemical space and chemical libraries.”
27. Tandon, Hiteshi, Labarca, Martín and Chakraborty, Tanmoy: “Electronegativity,
polarizability and FSGO method.”
28. Thyssen, Pieter: “Particular symmetries: From elementary particles to chemical elements.”
29. Zambon, Alfio: “Chemical reactivity and substance as spectrum.”
24th International Symposium of the ISPC
Buenos Aires, Argentina. July 2021

Sessions and time zones

Argentina CET PDT


Session 1 1 PM 6 PM 9 AM
Session 2 1:30 PM 6:30 PM 9:30PM
Session 3 2 PM 7 PM 10 AM
Cofee-break 2:30 PM 7:30 PM 10:30 AM
Session 4 3 PM 8: PM 11 AM
Session 5 3:30 PM 8:30 PM 11:30 M
Session 6 4 PM 9 PM 12 AM
Short program
Opening Session: Monday 12, 12:30 PM (Argentina time)

Monday 12 Wednesday 14 Friday 16


Vanessa Seifert and
Eric Scerri Yoshiteru Maeno
Alexander Frankin
Olimpia Lombardi and
N. Sukumar Hirofumi Ochiai
Sebastian Fortin
Chérif Matta, James S. M. H. Tandon, M. Labarca
Robin FindlayHendry
Anderson, and Lou Massa and T. Chakraborty
Cofee break Cofee break Cofee break

Alok Srivastava Gabriela García Zerecero Clevis Headley

José A. Chamizo Pieter Thyssen Michèle Friend

Mike L. Renier Jesus Alberto Jaimes Arriaga Ronei Clécio Mocellin

Monday 19 Wednesday 21
Guillermo Restrepo Marina Banchetti-Robino
and Jürgen Jost
Alfio Zambon and Juan Camilo
Graciela Pinto Vitorino Martínez González

Jean-Pierre Llored Sarah Hijmans

Cofee break Cofee break

Elena Ghibaudi and Ute Deichmann


Federica Branchini
Fiorela Alassia and
Bruno Pastoriza et al.
Mariana Córdoba
Martín Pérgola and ISPC Meeting
Lydia Galagovsky
Full Program
Opening Session: Monday 12, 12:30 PM (Argentina time)
Monday 12
Session 1: Eric Scerri. “Hasok Chang on acid”
Session 2: N. Sukumar. “Mereology of chemical space and chemical libraries”
Session 3: Chérif Matta, James S. M. Anderson, and Lou Massa. “Non-nuclear maxima
and the universality of Bright Wilson’s justification of the First Hohenberg-
Kohn Theorem”

Coffee break

Session 4: Alok Srivastava. “Plotting syntheses of rubber and of DNA: How chemists use
chemical structures and reaction pathways to craft new chemistries”
Session 5: José Antonio Chamizo. “Transformations. About novelty and persistence in
the history of chemistry”
Session 6: Mike L. Renier. “History and philosophy of chemical modeling”

Wednesday 14
Session 1: Vanessa Seifert and Alexander Frankin. “The problem of molecular structure
just is the measurement problem”
Session 2: Olimpia Lombardi and Sebastian Fortin. “The problema of isomerism”
Session 3: Robin Findlay Hendry. TBA

Coffee break

Session 4: Gabriela García Zerecero. “The unitary and substantial character of molecular
systems”
Session 5: Pieter Thyssen. “Particular symmetries: From elementary particles to chemical
elements”
Session 6: Jesus Alberto Jaimes Arriaga. “The powerfulness and qualitativeness of
electron density”
Friday 16
Session 1: Yoshiteru Maeno. “3-dimensional atomic and nuclear periodic tables”
Session 2: Ochiai, Hirofumi. “«Phenomenal field» provides philosophical grounds for
designing invisible structure of molecules.”
Session 3: Hiteshi Tandon, M. Labarca and Tanmoy Chakraborty. “Electronegativity,
polarizability and FSGO method”

Coffee break

Session 4: Clevis Headley. “Natural kinds, chemical practice, and interpretive


communities”
Session 5: Michèle Friend. “Comparing production of biofuel to oil”
Session 6: Ronei Clécio Mocellin. “Thinking with your hands: On the mode of existence
of chemical objects in the Enlightenment century”

Monday 19
Session 1: Guillermo Restrepo and Jürgen Jost. “The evolution of chemical knowledge:
How to study it”
Session 2: Alfio Zambon and Graciela Pinto Vitorino. “Reflections on the concept of
«chemical space of pharmacological action»”
Session 3: Jean-Pierre Llored. “The elimination of the holism-reductionism dichotomy
through the analysis: of quantum chemistry”

Coffee break

Session 4: Elena Ghibaudi and Federica Branchini. “How epistemology may inform
chemical education: the case of the notion of chemical bonding”
Session 5: Bruno Pastoriza, Fernanda Karolaine Dutra da Silva, Tavane da Silva
Rodrigues, Laura Bardini, and Guilherme Brahm dos Santos. “Fundamental
concepts in chemistry education: A problematization”
Session 6: Martín Pérgola and Lydia Galagovsky. “Modeling cellular respiration.
Educational reflections”
Wednesday 21
Session 1: Marina Banchetti-Robino. “Robert Boyle and the dispositional and relational
nature of chemical properties”
Session 2: Juan Camilo Martínez González. “The magnetic theory of electrons and the
cubic model of the atom”
Session 3: Sarah Hijmans. “The many fathers of aluminium: Collaboration, credit and the
construction of discovery”

Coffee break

Session 4: Ute Deichmann. “The rule of simplicity and template theories in biology and
chemistry - The example of Pauling’s questionable research on antibody
formation”
Session 5: Fiorela Alassia and Mariana Córdoba. “A processual ontology approach in
biochemistry and the review of the «structure-activity-relationships» dogma.
The case of drug receptors”
Session 6: ISPC Meeting
A processual ontology approach in biochemistry and the review of the “structure-
activity-relationships” dogma. The case of drug receptors.
Fiorela Alassia* and Mariana Córdoba**
*
Universidad Nacional de la Patagonia San Juan Bosco, Argentina. fiorela.alassia@gmail.com
**
Conicet – UBA, Argentina. marianacordoba@conicet.gov.ar

In this work we will analyze the particularities of biochemical entities (biomolecules) under the
consideration of a processual ontology approach, according to which the living world is not
populated by “substantial” individuals but by a dynamic hierarchy of processes instead (Dupré
and Nicholson, 2018).
We will analyze whether a processual view of biomolecules is accurate, with particular
emphasis to a central dogma in some biosciences, such as biochemistry, medicinal chemistry or
pharmacology. That is, the deterministic unidirectional relationship between the chemical
structure and the biological activity of a molecule or, as it is reformulated in the field of protein
studies, the unidirectional relationship between primary structure (amino acid sequence) and
macromolecular function (Lacková, 2019; Santos, Vallejos & Vecchi, 2020). From an alternative
point of view based on a processual approach, one question arises about whether it is possible to
consider biomolecules, such as proteins, as relational, dynamic and context-dependent entities
(Guttinger 2018).
In order to anchor the analysis in a concrete example, we will consider the particular case of
drug receptors. These macromolecules are found in cells and act as mediators or transmitters of
internal or external chemical signals (for example, a hormone or a drug), which triggers a certain
biological response (Voet & Voet, 2005; Silverman, 2004). Taking as a basis current knowledge
about structures and functioning of drug receptors, we will analyze some benefits and limitations
that processual ontology can offer to elucidate the nature of these biochemical entities.

References
Dupré, J. and Nicholson, D. (2018). “A manifesto for a processual philosophy of biology.” Pp. 3-
45 in D. Nicholson and J. Dupré (eds.), Towards a Processual Philosophy of Biology. Oxford:
Oxford University Press.
Guttinger, S. (2018). “A process ontology for macromolecular biology.” Pp. 303-320 in D.
Nicholson and J. Dupré (eds.), Everything Flows: Towards a Processual Philosophy of
Biology. Oxford: Oxford University Press.
Lacková, L. (2019). “Towards a processual approach in protein studies.” Biosemiotics, 12(3):
469-480.
Santos, G., Vallejos, G., and Vecchi, D. (2020). “A relational-constructionist account of protein
macrostructure and function.” Foundations of Chemistry, 22(3): 363-382.
Silverman, R. B. and Holladay, M. W. (2004). The Organic Chemistry of Drug Design and Drug
Action. Cambridge MA: Academic Press.
Voet, D. and Voet, J. (2005). Bioquímica. Buenos Aires: Médica Panamericana.
Robert Boyle and the dispositional and relational nature
of chemical properties
Marina Banchetti-Robino
Florida Atlantic University, USA. banchett@fau.edu

This presentation establishes that Robert Boyle’s complex chemical ontology implies a non-
reductionistic conception of chemical qualities and, more specifically, a conception of chemical
qualities as being dispositional and relational. Though Peter Anstey has already shown that that
Boyle considered sensible qualities to be dispositional and relational, this moves beyond
Anstey’s work by extending his arguments to chemical properties. These arguments are,
however, merely a first step in establishing a non-reductionistic interpretation of Boyle’s
chemical ontology. A further argument will show that Boyle regards chemical and other higher-
level properties as being emergent and supervenient properties. These arguments are supported
by substantial textual evidence from Boyle’s writings, which show that he clearly conceived of
chemical substances as functional wholes whose properties emerge not only from the
microstructural ordering of their parts but also from their relationship with other chemical
substances within the context of experimental practice.

References
Anstey, Peter, The Philosophy of Robert Boyle (London: Routledge, 2000).
Boyle, Robert, Considerations and Experiments, Touching the Origin and Forms of Qualities, in
The Works of Robert Boyle, edited by Michael Hunter and Edward B. Davis, Vol. 13,
(London: Pickering and Chatto, 2000)
Boyle, Robert, History of Particular Qualities, in The Works of Robert Boyle, edited by Michael
Hunter and Edward B. Davis, Vol. 6, (London: Pickering and Chatto, 2000)
Boyle, Robert, The Origin of Forms and Qualities, in The Works of Robert Boyle, edited by
Michael Hunter and Edward B. Davis, Vol. 5, (London: Pickering and Chatto, 2000)
Boyle, Robert, The Producibleness of Chymical Principles, in The Works of Robert Boyle, edited
by Michael Hunter and Edward B. Davis, Vol. 13, (London: Pickering and Chatto, 2000)
Transformations.
About novelty and persistence in the history of chemistry
José A. Chamizo
FQ-IIF-UNAM, México. jchamizo@unam.mx

In a previous work (Chamizo 2019) I reported on the most adequate interpretation of the history
of chemistry through Hacking-type revolutions instead of Kuhn-type revolutions. In other words,
the result of a specific rupture, often the result of the incorporation of new instruments and new
practices corresponding to new sub-disciplines, there is continuity, not everything is completely
lost. The above said is accepted by a laminated interpretation of the history of chemistry, but it
requires the identification of the substrate that remains. That substrate is what we know as
chemistry, established as a science in the late 18th and early 19th centuries, and it has changed
through four great transformations characterized, among other indicators, by the appearance of
four new sub-disciplines related to chemistry method: analysis and synthesis (Table).
A transformation is, simultaneously, novelty and persistence. We transform what is already
there, what we have, and there is always something that remains. Transformations are not
absolute.
There was no chemical revolution, what happened a little less than three hundred years ago,
was the emergence of an independent scientific discipline: chemistry.

Table. Transformation of Chemistry through its subdisciplines


Transformation Subdisciplines Another name Reference

First Organic Chemistry Quiet revolution Rocke (1993)

Second Physical Chemistry Atomic number revolution Wray (2018)

Third Instrumental Chemistry Instrumental revolution Morris (2002)

Fourth Organometallic Chemistry Organometallic historical regime Restrepo (2019)

References
Chamizo J.A. (2019) “About continuity and rupture in the history of chemistry: The fourth
chemical revolution (1945-1966).” Foundations of Chemistry, 21: 11-29.
Morris, P.J.T. (ed.) (2002). From Classical to Modern Chemistry. The Instrumental Revolution.
London: RSC-Science Museum-CHF.
Restrepo G. et al (2019). “Exploration of the chemical space and its three historical regimes.”
PNAS, 116: 12660-12665.
Rocke A.J. (1993) The Quiet Revolution: Hermann Kolbe and the Science of Organic Chemistry,
Berkeley: University of California Berkeley.
Wray K.W. (2018). “The atomic number revolution in chemistry: a Kuhnian analysis.”
Foundations of Chemistry, 20: 209-217.
The rule of simplicity and template theories in biology and chemistry -
The example of Pauling's questionable research on antibody formation
Ute Deichmann
Jacques Loeb Centre for the History and Philosophy of the Life Sciences,
Ben-Gurion University of the Negev, Beer Sheva. uted@post.bgu.ac.il

In 1940, Linus Pauling proposed his template theory of antibody formation, one of many such
theories that rejected Paul Ehrlich's selective theory of preformed 'receptors' (antibodies),
assuming instead a direct moulding of antibody shapes onto that of the antigen. Pauling's theory
was informed by his pioneering work on protein structure, and it was inspired by the intuitive
“rule of parsimony” and simplicity. His subsequent claim of having produced specific artificial
antibodies through experiments based on his 1940 theory, could not be reproduced by prominent
immunochemists, and his theory was shown to be mistaken. A citation analysis shows that
Pauling's papers on antibody generation are cited until today, in particular as pioneering studies
of the chemical technology 'molecular imprinting.'
My talk shows that philosophical ideas (such as simplicity or parsimony), and the aesthetic
power of models strongly impacted Pauling's work and that of other immunochemists and that it
facilitated irreproducibility. It points to problems in applying the principle of simplicity in
biology and shows that contrasting the concepts of 'template' and 'selection' is misleading in
immunology and beyond.
Comparing production of biofuel to oil
Michèle Friend
UCCS Université Lille Nord Europe, France.
michle@univ-lille.fr
Department of Philosophy, George Washington University, USA.
michele@gwu.edu

The problems surrounding our use of fuel are not all, or only, moral problems, although there is a
strong moral element in negotiating resolutions. If we compare biofuel production to oil refining,
and if we place these in their geo-political and ecological contexts, then we discover different
suites of problems. Some of the problems are systemic, others scientific, some are political,
others are existential. I shall be comparing problems surrounding the biofuel industry to the oil
industry in the Santiago River basin in Mexico to the same industries in northern France.
I’ll be drawing on information from actual cases to focus on the problem of how to weigh
seemingly incomparable objectives, such as: creation of jobs versus pollution or loss of
biodiversity in an ecosystem. Using the actual information is important insofar as one
acknowledges that philosophy is sometimes tested and informed by the real political world, rather
than exclusively by the minds of other philosophers.
I argue that we cannot solve the problem of comparison of competing and very different
interests in the sense of finding an obvious optimum. The latter is what we practice now, and it is
what is proposed by environmental economists when they calculate the contribution of natural
systems to the market or when we do an environmental impact assessment. The common measure
is money. I argue that this way of comparing very different interests is too superficial and will
lead to more irreversible problems.
I propose that we do something else. Instead, we resolve the problem by using an institutional
compass. The resolution is a much more flexible and adaptive solution, but it requires
philosophical subtlety. As such, the exercise of developing a compass for evaluating the overall
effect of developing biofuel in different regions of the world, raises other philosophical
questions. The methodology of developing institutional compasses can be applied to any
institution, and is suitable for the development of any chemical process for the purpose of
deployment on a large scale.
The findings here will be of limited interest to philosophers of pure chemistry or fundamental
chemistry.
The unitary and substantial character of molecular systems
Gabriela García Zerecero
Universidad Panamericana. Departamento de Humanidades. Zapopan, Jalisco, México.
gagarcia@up.edu.mx

Since the middle of the 20th century, the problem of the complexity of the real has become one of
the main challenges for scientific knowledge. The same advances in science have contributed to
gradually forge an increasingly complex intelligibility of the real, a complexity far superior to
what ordinary knowledge allows us to verify. To understand this complexity, we are now talking
about systems.
Can it be said that molecular modeling corresponds to a material system? Molecular models
meet the very general definition of system, namely, an ordered complex of interacting elements.
In addition, in most cases, these molecular compounds can be isolated under normal physical
conditions and, therefore, can be described as relatively firm and stable systems.
In this paper, I shall show that molecules have a particular holistic character. If molecules
possess all the characteristics that define a central unitary system within the framework of the
general theory of systems, then it can be sayed that these same chemical entities are bodily
substances in a philosophical sense. Therefore, what the general theory of systems understands as
a whole has some correspondence with what traditional philosophy designates as a substantial
metaphysical unit.

References
Artigas, M. (1995). La inteligibilidad de la naturaleza. Pamplona: EUNSA.
Bertalanffy, L. von (1976). Teoría general de los sistemas. México D. F.: Fondo de Cultura
Económica.
Cruz Cruz, Juan (1967). La filosofía de la estructura. Pamplona. EUNSA.
Durán, R. and Espinoza, R (2016). “Zubiri and Prigogine: Considerations on Life, Dynamism and
Self-organization.” ARBOR Ciencia, Pensamiento y Cultura, Vol. 192-780, julio-agosto, a328,
doi: http://dx.doi.org/10.3989/arbor.2016.780n4004.
Juignet, P. (2015). “État actuel de la théorie des systèmes.” Philosophie, science et
société.https://philosciences.com/philosophie-generale/complexite-systeme-organisation-
emergence/44-etat-actuel-de-la-theorie-des-systemes.
Laszlo, E. (1972). Introduction to Systems Philosophy, Toward a New Paradigm of
Contemporary Thought. New York. Gordon and Breach.
Murad, M. (2011). “Models, scientific realism, the intelligibility of nature, and their cultural
significance.” Studies in History and Philosophy of Science, 42: 253-261.
Zeidler, P. (2000). “The epistemological status of theoretical models of molecular structure.”
HYLE - International Journal for the Philosophy of Chemistry, 6: 17-34.
How epistemology may inform chemical education:
the case of the notion of chemical bonding
Elena Ghibaudi and Federica Branchini
Dept. Chemistry, University of Torino, Torino, Italy; elena.ghibaudi@unito.it

The notion of chemical bonding has a foundational character for chemistry, as -together with
other related concepts, such as the notion of molecule, valence and molecular structure (whereof
chemical bonding is the necessary premise)- it contributes to structure the chemical discipline.
The history of the notion of chemical bonding highlights the fact that its theoretical reference
field has changed over time: this has influenced both the definition and the modelling of bonding.
Nevertheless, since the elaboration of the MO-LCAO theory, calculation methods have
significantly improved, whereas the definition of bonding does not seem to have been further
debated (Pauling 1960).
On the educational side, several investigations prove that students experience serious obstacles
in understanding this notion and they often prove unable to use it effectively. Several scholars
found recurrent misconceptions of chemical bonding in distinct teaching and cultural
environments (Levy Nahum et al. 2010): likely, these misconceptions are fostered by
inappropriate didactic transpositions, which may be due to the insufficient epistemological
awareness of many chemistry teachers.
Hence, the notion of chemical bonding is an optimal ground for discussing the two-way
relation between epistemology and chemical education. The former informs chemical education
by highlighting conceptual knots and inconsistencies. The latter raises questions to which
epistemology is called upon to answer.
A main feature of the notion of chemical bonding is its relational character, which justifies the
definition of chemistry as an inherently complex discipline (Villani, 2008). Other peculiar aspects
of this notion are the numerous dichotomies related to it: i) the chemical perspective vs. the
physical one, which reflects into the structural vs. energetic view of bonding. ii) the distinction
between intramolecular and intermolecular bonds. iii) the directional or non-directional character
of bonding. iv) the classical and quantum descriptions, that coexist in the chemical practice. v) its
use as discriminating factor for classifying the transformations of matter. And, above all, the
distinction between the concept of bonding and its descriptive models, which is rarely discussed
in the teaching practice. Some of these aspects will be discussed in the light of the
epistemological reflections on chemical bonding and the educational evidence available in the
literature (Hendry 2008; Del Re 1996; Shaik 2007)

References
T. Levy Nahum, R.Mamlok-Naaman, A.Hofstein, K. S. Taber, Studies in Science Education,
2010, 46(2): 179, DOI: 10.1080/03057267.2010.504548.
G. Villani, Complesso e organizzato, Editori Riuniti, 2008.
R. Hendry, Philosophy of Science, 2008, 75, 909.
G. Del Re, Chimica nella Scuola, 1996, 5, 155.
L. Pauling, The nature of the chemical bond, Cornell University Press, Ithaca, 1960.
S. Shaik, J. Comput. Chem., 2007, 28, 51.
Natural kinds, chemical practice, and interpretive communities
Clevis Headley
Florida Atlantic University, USA. headley@fau.edu

Many philosophers attribute extraordinary importance to the idea of natural kinds seemingly
intimating that the very possibility of certain kinds of activity are ontologically beholden to the
existence of kinds.
Specifically regarding chemistry, Ellis (2002) intimated that the success of any plausible
metaphysical essentialism depends upon its “reliance on examples from chemistry” (Vandewll
2018: 907). Ellis’s view is representative of a tradition in analytic philosophy that has utilized
chemical examples as paradigmatic natural kinds. In this regard, Kripke (1975) and Putnam
(1980) emerge as the architects of an entrenched research program dedicated to the chemical
tradition of natural kinds in analytic philosophy.
The emergence of a critical body of literature by philosophers of chemistry and others has
shattered the complacent reliance upon chemical examples as exemplary kinds (Bursten 2014;
LaPorte 1996; Leslie 2013; Mellor 1991; Needham , 2000, 2011; VandeWall, 2007; Van Brakel,
1986, 2000, 2005; Weisberg 2006). On the basis of this emerging critical literature, I will
critically explore the way in which chemical practice and inquiry affects the natural kind debate.
So, instead of the pretense that we simply carve nature at its joints, we need to become better
grounded in the practice of science, and especially with regard to the debate about natural kinds
in chemical practice. Consistent with this orientation, we need to make the practice turn, that is,
eradicate the fantasy of logical reconstruction and become involved with the interpretative and
historical challenges of understanding the nuances of practice. In this regard, Hasok Chang
recommends not viewing natural kinds as objective features of reality patiently awaiting
discovery; rather, he maintains that natural kinds “can only be discerned in the process of inquiry
in the natural sciences” (Chang 2012: 42). Chang adds that what qualifies as a natural kind
concept is one that performs a successful role in a system of practice, which is roughly another
way of saying that a natural kind concept is one that performs expeditiously, it contributes
constructively to the epistemic objective of a system of inquiry.
Those thinkers supporting the practice turn also support shifting “philosophical investigation
of the naturalness of natural kinds to how they are used, discovered, or made. This shifts
metaphysical inquiry of natural kinds from the contents of the world to the activities of
portioning, conceptualizing, comparing, and categorizing– that is to ontologizing practices”
(Kendig 2016: 3). The point here is quite clear, metaphysical questions regarding natural kind
should be imminent to scientific practice. Indeed, any legitimate metaphysics of natural kinds
should be appropriately informed and grounded in practice and not operate on the basis of a priori
sovereignty.
I will insert this critical discussion within the analytical context of the notion of interpretive
communities (Fish 1980), and make the case that philosophers should not assume that appeals to
the purity of philosophy can substitute for the complexity and practical orientation of chemical
practice.

References
Chang, Hasok (2012). Is Water H2O? Evidence, Realism and Pluralism. Boston Studies in the
History and Philosophy of Science, vol. 293. Berlin: Springer.
Ellis, Brian (2002). The Philosophy of Nature: A Guide to the New Essentialism. Montreal:
McGill-Queen’s University Press.
Fish, Stanley (1980). Is There a Text in This Class?, Cambridge, MA: Harvard University Press.
Kendig, Catherine (2016). Natural Kinds and the Classification of Scientific Practice. London:
Taylor & Francis.
Kripke, Saul (1980). Naming and Necessity. Cambridge, Massachusetts: Harvard University
Press.
LaPorte, Joe (1996). “Chemical Kind Term Reference and the Discovery of Essence.” Noûs,
30(1): 112-132.
Leslie, Sarah-Jane (2013). “Essence and Natural Kinds: When Science Meets Preschooler
Intuition.” Oxford Studies in the Epistemology, Volume 4, (eds.) Tamar Szabó and John
Hawthorne (eds.). Oxford: Oxford University Press.
Mellor, D. (1991). Matters of Metaphysics. Cambridge: Cambridge University Press.
Needham, Paul (2000). “What is Water?” Analysis, 60(1): 13-21.
_____________ (2011). “Microessentialism: ‘What Is the Argument?’” Noûs, 45(1): 1-21.
Putnam, Hillary (1975). “The Meaning of ‘meaning’.” Mind, Language, and Reality. Collected
Papers, Vol. 2. Cambridge: Cambridge University Press.
Van Brakel, Jaap (1986). “The Chemistry of Substances and the Philosophy of Mass Terms.”
Synthese, 69: 291-324.
______________ (2000). Philosophy of Chemistry: Between the Manifest and the Scientific
Image. Leuven: Leuven University Press.
______________ (2005). “On the Inventors of XYZ.” Foundations of Chemistry, 7: 57-84.
Van de Wall, Holly (2018). “ Why Water is Not H2O, and Other Critiques of Essentialist
Ontology from the Philosophy of Chemistry.” Philosophy of Science, 74: 906-919.
Weisberg, Michael (2006). “Water Is Not H2O.” Pp. 337-345 in Boston Studies in the Philosophy
of Science: Philosophy of Chemistry, Davies Baird, Eric Scerri, and Lee McIntyre (eds.).
Boston Studies in the Philosophy of Science, vol. 2422. Berlin: Springer.
TBA
Robin Findlay Hendry
Durham University, UK. R.F.Hendry@durham.ac.uk
The many fathers of aluminium:
Collaboration, credit and the construction of discovery
Sarah Hijmans
Laboratoire SPHERE, Université de Paris, France.
sarahnhijmans@gmail.com

In historical accounts or textbooks of chemistry, it is very common to find lists of discoveries


identifying a single discoverer and discovery date for each chemical element. However, the
notion of a scientific discovery as a clearly identifiable event is problematic. For example, it is
extremely difficult to answer the question « Who discovered aluminium? » (Kragh, 2019). In the
literature, this metal is said to be discovered either by Friedrich Wöhler in 1827 or Hans Christian
Ørsted in 1825, whereas Henri Sainte-Claire Deville enabled its production on an industrial scale
in 1754 (Ferrand, 1961). Despite the fact that the metal’s existence was considered to be proven
by Humphry Davy’s 1808 work on the decomposition of the earths, Davy’s experiments are often
referred to as ‘failed’ or ‘unsuccessful’ and excluded from the discovery process (for example
Weeks, 1956).
This paper will propose a historiographical reflection on the notion of scientific discovery
using aluminium as a case study. The isolation of aluminium was the result of a progressive
accumulation of knowledge by many different chemists, rather than the efforts of a single
chemist. Later, when aluminium became an object of scientific and economic prestige, the
attribution of priority to Wöhler and Deville erased earlier contributions from the discovery
narrative. Until the twentieth century, histories of aluminium remained centred on the question of
priority and this resulted in a focus on the comparison of the works of Ørsted and Wöhler, once
again erasing earlier efforts. I argue that in order to understand the birth of aluminium, we should
take into account all complexities of this story: not only the long and collaborative process that
led to its isolation but also the way in which its story has been constructed and reconstructed over
time.

References
Ferrand, L. (1961). Histoire de la Science et des Techniques de l’Aluminium (The Author,
France).
Kragh, H. (2019). “Controversial Elements: Priority Disputes and the Discovery of Chemical
Elements.” Substantia, 3: 79-90.
Weeks, M. E. (1956). Discovery of the Elements, 6th ed. Easton Pen: Journal of Chemical
Education.
The powerfulness and qualitativeness of electron density
Jesus Alberto Jaimes Arriaga
CONICET - University of Buenos Aires, Argentina. ja.jaimes@conicet.gov.ar

In recent years there has been a growing interest in the metaphysics of science. Within this field,
one of the main debates is that referred to the nature of the properties which scientists deal with.
This debate has developed mainly around the well-known distinction between categorical and
dispositional properties. Regarding this point there are two extreme positions. According to
categoricalism, there are only categorical properties or qualities (Armstrong 1997);
pandispositionalism asserts the opposite, that there are only properties that grant powers (Bird
2007, Molnar 2003). Orthodoxy in property metaphysics builds this distinction on the basis that
qualities are non-dispositional, while dispositions are non-qualitative, that is, dispositionality and
qualitativity are mutually exclusive attributes. However, in recent years this view has been
questioned by several authors, who have proposed to abandon the categorical/dispositional
distinction and to embrace the idea of the identity between quality and power, that is, that all (or
at least many) properties are both qualitative and dispositional. This position is known as the
“powerful qualities theory” (Martin 2007, Heil 2003, Jacobs 2011). In the light of this discussion,
the aim of this work is to inquire about the nature of the properties postulated by quantum
chemists. Particularly, I will analyze the property of electron density, central to quantum
chemistry, in order to argue that the theory of powerful qualities is the best candidate to account
for the nature of electron density. Following the idea put forward by Ingthorsson (2021), that
science “describes the qualities of objects identifying those qualities with what objects can do”
(p. 131), I will argue that electron density is a property whose qualitativeness is associated with
its topology, while its powerfulness is related to the causal powers it bestows on molecules,
which make them to behave differently under distinct circumstances.

References
Armstrong, D. M. (1997). A World of States of Affairs. Cambridge: Cambridge University Press.
Bird, A. (2007). Nature’s Metaphysics. Oxford: Clarendon Press.
Heil, J. (2003). From an Ontological Point of View. Oxford: Oxford University Press.
Ingthorsson, R. D. (2021). A Powerful Particulars View of Causation. New York: Routledge.
Jacobs, J. D. (2011). “Powerful qualities, not pure powers.” The Monist, 94: 81–102.
Martin, C. B. (2007). Mind in Nature. Oxford: Oxford University Press.
Molnar, G. (2003). Powers: A Study in Metaphysics. Oxford: Oxford University Press.
The elimination of the holism-reductionism dichotomy through
the analysis of quantum chemistry
Jean-Pierre Llored
Associate Professor in Humanities and Social Sciences
École Centrale Casablanca, Morocco.
Department of Humanities and Social Sciences, École CentraleSupélec, France.
jean-pierre.llored@centralesupelec.fr

The starting point of this paper is the investigation of quantum chemistry as it is practised in
research laboratories. To do so, we investigate how whole-parts strategies are used to perform
calculations by referring to different chemical quantum methods. This scrutiny enables us to
highlight: (1) the mutual dependence of whole and parts within calculation procedures, and (2)
that these procedures are not about whole and parts only, but also include a third term, that is to
say, the environment with which the whole in question interacts. To conclude, we assess the
extent to which the holism-reductionism dichotomy is relevant when addressing the specific
situation of quantum chemistry. We also ask the question as to whether or not this dichotomy
should be eliminated, at least in its current forms, not only with respect to quantum chemistry,
but also with regard to other domains of science for which other studies have underscored the
deep problems raised by this dichotomy.

References:
Ayala FJ and Dobzhansky T (eds) (1974) Studies in the philosophy of biology. Reduction and
related problems, University of California Press, Berkeley and Los Angeles
Bader RFW (1990) Atoms in Molecules: A Quantum Theory. Oxford University Press, Oxford
BaroneV, Improta R, Rega N (2004) “Computation of protein pK’s values by an integrated
density functional theory/polarizable continuum model approach.” Theor. Chem. Acc., 111:
237–245
Campbell DT (1974) “‘Downward causation’ in hierarchically organized biological systems.” Pp.
179–186 in Ayala FJ and Dobzhansky T (eds). Studies in the philosophy of biology,
MacMillan, London.
Carnap R (1959) “The Elimination of Metaphysics Through the Logical Analysis of Language.”
p. 60-81 in Ayer AJ (ed.) Logical Positivism. The Free Press, New York. The original paper
was published in German in 1932
El-Hani CN, Pereira AM (2000) “Higher-level descriptions: why should we preserve them?” Pp.
118–142 in Andersen PB, Emmeche C, Finnemann NO, Christiansen PV (eds) Downward
causation: minds, bodies and matter. Aarhus University Press, Aarhus.
Feibleman JK (1954) “Theory of integrative levels.” The British Journal for the Philosophy of
Science, 5(17): 59-66
Heitler W, London F (1927) “Wechselwirkung neutraler Atome und homoopolare Bindung nach
der Quantenmechanik.” Zeitschrift für Physik, 44: 455-472.
Hendry RF (2006) “Is there Downward Causation in Chemistry?” Pp. 173-190 in Bairs D, Scerri
ER and McIntyre L (eds) Philosophy of Chemistry: The Synthesis of a New Discipline.
Springer, Dordrecht.
Hendry RF (2010) “Ontological reduction and molecular structure.” Studies in History and
Philosophy of Modern Physics, 41(2): 183-191.
Kohn W, Becke AD, Parr RG (1996) “Density functional theory of electronic structure.” J. Phys.
Chem., 100: 12974–12980.
Langmuir I (1919) “The arrangement of electrons in atoms and molecules.” Journal of the
American Chemical Society, 41: 868-934.
Lewis GN (1923) Valence and the structure of molecules, The Chemical Catalog Company, New
York.
Llored JP (2014) “Wholes and Parts in Quantum Chemistry: Some Mereological and
Philosophical Consequences.” HYLE, International Journal for the Philosophy of Chemistry,
20: 141-163.
Llored JP (2012) “Emergence and Quantum Chemistry.” Foundations of Chemistry, 4(3): 245-
274.
Lombardi O and Labarca L (2005) “The Ontological Autonomy of the Chemical World.”
Foundations of Chemistry, 7: 125-148.
Lombardi O and Labarca L (2006) “The Ontological Autonomy of the Chemical World: A
Response to Needham.” Foundations of Chemistry, 8: 81-92.
Mulliken RS (1932a) “Interpretation of band spectra, part III. Electron quantum numbers and
states of molecules and their atoms.” Rev. Mod. Phys., 4(1): 1–86
Mulliken RS (1932b) “Electronic structures of polyatomic molecules and valence I.” Phys. Rev.,
40: 55–62.
Nagel T (1998) “Reductionism and antireductionism.” Pp. 3-14 in Bock GR and Goode JA (eds)
The Limits of Reductionism in Biology. Novartis Foundation Symposium 213, John Wiley,
Chichester.
Nagel E (1961) The structure of science: Problems in the logic of scientific explanation.
Routledge & K. Paul, London
Pauling L (1939) The Nature of the Chemical Bond, 1st edition. Cornell University Press, Ithaca,
New York
Pauling L, Sherman J (1933b) “The Nature of the Chemical Bond. VII. The calculation of
resonance energy in conjugated systems.” Journal of Chemical Physics, 1: 679-686
Popellier P (2000) Atoms in Molecules. An Introduction. Prentice Hall, London
Salthe SN (1985) Evolving hierarchical systems: Their structure and representation. Columbia
University Press, New York
Scerri E (2007) “The Ambiguity of Reduction.” HYLE, International Journal for the Philosophy
of Chemistry, 13(2): 67-81.
Scerri E (2000) “Realism, Reduction, and the ‘Intermediate Position’.” Pp. 51-72 in Bhushan N
and Rosenfeld S (eds) Of Minds and Molecules. New Philosophical perspectives on
Chemistry. Oxford University Press, New York.
Scerri E (1997) “Has the Periodic Table been Successfully Axiomatized.” Erkentnnis, 47: 229-
243.
Scerri E (1991a) “Electronic Configurations. Quantum Mechanics and Reduction.” The British
Journal for the Philosophy of Science, 42: 309-325.
Scerri E (1991b) “Chemistry, Spectroscopy, and the Question of Reduction.” Journal of
Chemical Education, 68: 122-126.
Schaffner KF (1967) “Approaches to reduction.” Philosophy of Science, 34: 137–147.
Suppe F (1988) The Semantic Conception of Theories and Scientific Realism, University of
Illinois Press, Urbana
Weininger SJ (2001) “Affinity, Additivity and the Reification of the Bond.” Pp. 237-251 in Klein
U (ed) Tools and Modes of Representation in the Laboratory Sciences. Boston Studies in the
Philosophy of Science, Vol. 222. Kluwer Academic Publishers, Amsterdam.
Wimsatt WC (1976) “Reductive explanation: A functional account.” Pp. 671-710 in Michalos
AC, Hooker CA, Pearce G, and Cohen RS (eds). PSA-1974, Boston Studies in the Philosophy
of Science, Vol. 30, Reidel, Dordrecht.
Wittgenstein L (1974) On Certainty. Anscombe GEM and von Wright GH (eds). Basil
Blackwell, Oxford.
Zhao Y, Truhlar DG (2008a) “Density functionals with broad applicability in chemistry.” Acc.
Chem. Res., 41(2): 157–167.
The problem of isomerism
Olimpia Lombardi* and Sebastian Fortin**
*
Conicet – UBA, Argentina. olimpiafilo@gmail.com
**
Conicet – UBA, Argentina. sfortin@conicet.gov.ar

The problem of isomerism can be expressed by the following question: Can quantum mechanics
account for the existence of isomers? The difficulty for a positive answer is that, when a molecule
is described exclusively in quantum terms, the description cannot distinguish between different
isomers: since they are quantum systems with the same number and types of elements, isomers
are described by the same Coulombian Hamiltonian. The strategy in chemistry is to appeal to
effective Hamiltonians, obtained by means of the Born-Oppenheimer approximation, which
embody the structure of the each particular isomer.
In this talk we will argue that, although the problem of isomerism can be conceived as a case of
the problem of symmetry (Can quantum mechanics account for the existence of certain molecular
asymmetries?), it cannot solved as other cases of asymmetry, by merely assuming an externally
induced symmetry breaking. In particular, the symmetric wavefunction of the molecule cannot be
reconstructed as a superposition of asymmetric wavefunctions corresponding to the different
isomers.

References
Fortin, S. and Lombardi, O. (2021). “Is the problem of molecular structure just the measurement
problema?” Foundations of Chemistry, on line first.
Fortin, S., Lombardi, O., and Martínez González, J. C. (2016). “Isomerism and decoherence.”
Foundations of Chemistry, 18: 225-240.
Fortin, S., Lombardi, O., and Martínez González, J. C. (2018). “A new application of the modal-
Hamiltonian interpretation of quantum mechanics: The problem of optical isomerism.” Studies
in History and Philosophy of Modern Physics, 62: 123-135.
Franklin, A. and Seifert. V. (2020). “The problem of molecular structure just is the measurement
problem.” The British Journal for the Philosophy of Science, forthcoming.
Hendry, R. F. (1998). “Models and approximations in quantum chemistry.” In N. Shanks (ed.),
Idealization in Contemporary Physics. Amsterdam-Atlanta: Rodopi.
Hendry, R. F. (2010). “Ontological reduction and molecular structure.” Studies in History and
Philosophy of Modern Physics, 41: 183-191.
Claverie, P. and Diner, S. (1980). “The concept of molecular structure in quantum theory:
interpretation problems.” Israel Journal of Chemistry, 19: 54-81.
Sutcliffe, B. T. and Woolley, R. G. (2012). “Atoms and molecules in classical chemistry and
quantum mechanics.” In R. F. Hendry and A. Woody (eds), Handbook of Philosophy of
Science. Vol. 6, Philosophy of Chemistry. Oxford: Elsevier.
Sutcliffe, B. T. and Woolley, R. G. (2021). “Is chemistry really founded in quantum mechanics?”
Forthcoming in O. Lombardi, S. Fortin, and J. C. Martínez González (eds.), Quantum
Chemistry: Philosophical Perspectives in Modern Chemistry. Berlin-Heidelberg: Springer.
Woolley, R. G. (1978). “Must a molecule have a shape?” Journal of the American Chemical
Society, 100: 1073-1078.
Woolley, R. G. and Sutcliffe, B. T. (1977). “Molecular structure and the Born-Oppenheimer
approximation.” Chemical Physics Letters, 45: 393-398.
Three-dimensional atomic and nuclear periodic tables
Yoshiteru Maeno
Department of Physics, Kyoto University, Japan.
maeno.yoshiterru.2e@kyoto-u.ac.jp

We introduce three-dimensional helical periodic tables of atoms and nuclei. With the paper
patterns provided, students can easily make these periodic tables in class to enhance
understanding of periodicities of the elements.
The periodic table currently used worldwide is of a long form pioneered by Werner in 1905. It
was Pfeiffer who refined Werner’s work and rearranged the rare-earth elements in a separate
table below the main table in 1920. Today’s widely used periodic table essentially inherits
Pfeiffer’s arrangements. Although long-form tables more precisely represent electron orbitals
around a nucleus, they lose some of the features of Mendeleev’s short-form table to express
similarities of chemical properties of elements when forming compounds.
Here we explain how the 3D helical periodic table “Elementouch”, which combines the s- and
p-blocks into one tube, can recover features of Mendeleev’s periodic law. We also introduce the
recently invented nuclear periodic table “Nucletouch”, based on the proton magic numbers. The
nuclear shell structure leads to a new arrangement of the elements with the proton magic number
nuclei treated like noble-gas atoms.

References
K. Hagino, K y Maeno, Y. (2020). “A nuclear periodic table.” Foundations of Chemistry, 22:
267–273. https://rdcu.be/b3F3W
Maeno, Y, Hagino, K., and Ishiguro, T. (2020). “Three related topics on the periodic tables of
elements.” Foundations of Chemistry, online first. https://doi.org/10.1007/s10698-020-
09387-z 09387-z.
The magnetic theory of electrons and the cubic model of the atom
Juan Camilo Martínez González
CONICET - University of Buenos Aires, Argentina. olimac62@hotmail.com

The year 1916 marked one of the most revolutionary years for theory in chemistry. The
influential works of G. N. Lewis settled our modern understanding for chemical bonds and
representations of molecular structures. Lewis’s program was expanded and developed in great
detail on his 1923 “Valence and Structure of Atoms and Molecules”. His ideas were influential
enough to appear highlighted on the “The Nature of the Chemical Bond” introduction (Pauling
1931). Linus Pauling himself on a 1939 letter confessed to Lewis his admiration: “You know, of
course, that I had you on my mind continually while it was being written, and I have been hoping
that my treatment would prove to be acceptable to you.” (Pauling, 1939).
The purpose of this article is to recover the influence of Alfred Lauck Parson (1889 Lucknow,
India - 1970 Allonby, England) works in the development of Lewis’s ideas concerning the
chemical aspects of Bohr’s theory, electron pairing and the reconciliation between two
conflicting theories about dynamics and structure of electrons in atoms and molecules. Parson
was a Harvard graduate student noted for his “magneton theory” of the atom. Between 1913 and
1915, Parson was a visiting student at the University of California, Berkeley, where coincidently
Gilbert N. Lewis was working as the chair of the department of chemistry. During these years,
Lewis had contact with Parson’s work, which argued that the electron, in the Bohr model, might
be a ring of negative electricity spinning (a Magneton) with a high velocity about its axis and that
a chemical bond results from two electrons being shared between two atoms.

References
Lewis, G. N. (1916) Lewis, G. N. (1923a). Valence and the Structure of Atoms and Molecules.
New York: The Chemical Catalog Company, reprinted by Dover Publications, 1966.
Parson, A. L. (1915). “Magneton theory of the structure of the atom.” Smithsonian Miscellaneous
Collection, Pub 2419, V65, N11, 1916; [Pub.no.2371, 80pgs, 2 plates, Nov, 29, 1915].
Pauling, L. (1931). The Nature of the Chemical Bond and the Structure of Atoms and Molecules.
An Introduction to Modern Structural Chemistry. New York: Cornell University Press.
Pauling (1939). Letter to Lewis. corr216.1-lp-lewis-19390829.
Non-nuclear maxima and the universality of Bright Wilson’s justification
of the First Hohenberg-Kohn Theorem
Chérif Matta*, James S. M. Anderson**, and Lou Massa***
*
Dept. of Chemistry and Physics, Mount Saint Vincent University, Halifax, Nova Scotia, Canada
Cherif.Matta@msvu.ca
**
Instituto de Química, Universidad Nacional Autónoma de México, Ciudad de México, Mexico
***
Dept. of Chemistry Hunter College, City University of New York, New York, USA

In 2009, Mikael Johansson sparked a thought provoking discussion on “Computational


Chemistry List” (Johansson, 2009) about the applicability of Kato’s cusp condition (Kato, 1957)
and E. Bright Wilson’s explanation of density functional theory (DFT) (Löwdin, 1985) in the
case of “non-nuclear maxima” (NNM) also known as “non-nuclear attractors” (NNA) (see e.g.
Bader and Platts, 1997; Cao et al., 1987; Gatti et al., 1987; Iversen et al., 1995; Martin-Pendas et
al., 1999; Taylor et al., 2007; Terrabuio et al., 2016).
A NNM behaves as a pseudo-atomic nucleus in the sense that it “attracts” the gradient vector
field lines of the electron density in its vicinity which leads to the formation of a basin delimited
from the rest of the system’s electron density by a surface of local zero-flux in the gradient field
of the electron density, that is, a surface satisfying the well-known Bader condition
 (r )  n(r ) = 0 (Bader, 1990). By ignoring the finite size of atomic nuclei, the wavefunction and
electron density exhibit cusps at the nuclear positions. Topologically, however, a maximum and a
cusp are equivalent as attractors of gradient vector fields.
Wilson has argued that, since the integral of the electron density over all space is just the total
number of electrons (N), it determines the upper limit of summations that appear in the
Hamiltonian that involve N. Further, since the density is dominated by nuclear cusps, it also
determines the positions and the number of the nuclei {RA}. Finally, the atomic number Z of a
given atomic nucleus can be obtained from Kato’s cusp condition (Steiner, 1963). In this way, the
electron density fixes the Hamiltonian completely, which is a rationalization of the Hohenberg-
Kohn first theorem (rigorously proved ad absurdum by Hohenberg and Kohn, 1964).
It will be shown that both Kato’s condition and Wilson argument can be extended to systems
with NNMs. It will be proved that the spherically-averaged derivative of the density with respect
to r from any well-behaved point is strictly zero. Since NNMs are not cusps but true maxima, this
result means that the “extended Kato’s condition” also applies in the case of NNMs. This result is
also supported by some simple numerical examples

References
Bader RFW (1990). Atoms in Molecules: A Quantum Theory. Oxford, U.K.: Oxford University
Press.
Bader RFW, Platts JA (1997). “Characterization of an F-center in an alkali halide cluster.” J.
Chem. Phys., 107: 8545-8553.
Cao WL, Gatti C, MacDougall PJ, Bader RFW (1987). “On the presence of non-nuclear
attractors in the charge distributions of Li and Na clusters.” Chem. Phys. Lett., 141: 380-385.
Gatti C, Fantucci P, Pacchioni G (1987). “Charge density topological study of bonding in lithium
clusters. Part I: Planar Li n clusters (n=4, 5, 6).” Theor. Chem. Acc. (Formerly, Theoret. Chim.
Acta), 72: 433-458.
Hohenberg P, Kohn W (1964). “Inhomogeneous electron gas.” Phys. Rev. B, 136: 864-871.
Iversen BB, Larsen FK, Souhassou M, Takata M (1995). “Experimental evidence for the
existence of non-nuclear maxima in the electron-density distribution of metallic beryllium. A
comparative study of the maximum entropy method and the multipole refinement method.”
Acta Cryst. B, 51: 580-591.
Johansson M (2009). “Summary: Total electron density cusps outside nuclei.” CCL
(Computational Chemistry List), Message:
http://www.ccl.net/chemistry/resources/messages/2009/10/18.001-dir/index.html.
Kato WA (1957). “On the eigenfunctions of many-particle systems in quantum mechanics.”
Commun. Pure Appl. Math., 10: 151-177.
Löwdin P-O (1985). “Twenty-five years of Sanibel Symposia: A brief historic and scientific
survey.” Int. J. Quantum Chem., 28: 19-37.
Martin-Pendas A, Blanco MA, Costales A, Mori-Sanchez P, Luana V (1999). “Non-nuclear
maxima of the electron density.” Phys. Rev. Lett.; 83: 1930-1933.
Steiner E (1963). “Charge densities in atoms.“. J. Chem. Phys., 39: 2365-2366.
Taylor A, Matta CF, Boyd RJ (2007). “The hydrated electron as a pseudo-atom in cavity-bound
water clusters.” J. Chem. Theor. Comput., 3: 1054-1063.
Terrabuio LA, Teodoro TQ, Matta CF, Haiduke RLA (2016). “An investigation of non-nuclear
attractors in heteronuclear diatomic systems.” J. Phys. Chem. A, 120: 1168-1174.
Thinking with your hands: On the mode of existence of chemical objects
in the Enlightenment century.
Ronei Clécio Mocellin
Department of Philosophy, Federal University of Paraná, Brazil.
roneimocellin@ufpr.br

Thinking and acting on the materials has always been the very work of chemists. During the 18th
century, this overlap between theory and practice was articulated in order to offer a cognitive
identity to a chemical way of knowing and intervening in the natural world. The article “Chymie”
from the Diderot and d’Alembert Encyclopedia presented a true “chemical manifesto”, which
claimed the ontological and epistemic originality of knowledge produced in a specific physical
space, the chemical laboratory. In these places, materials for different purposes were manipulated
and the knowledge of their properties resulted from an experimental investigation. The objective
of the chemists was to explain certain macroscopic properties from microscopic entities,
invisible, which manifested themselves under certain operating conditions. However, the entities
imagined by them only gained significance if they were part of a network of causal relationships
that not only explained a particular case, but also guided future investigations. The Geoffroy’s
Table of Relations was the representational model of this dual track between the macro and the
micro material, it was a true “paper tool”. Chemists were not willing to demonstrate what the
ultimate elements of matter were, nor to apply a theoretical model a priori to explain their
properties, but were limited to manipulating materials and imagining concepts that engendered
new operations. The concepts forged by them to denote chemical objects constituted a
“provisional ontology”, which did not seek to unveil the ultimate essence of materiality, but to
explain what could be done with them. The chemical operations carried out in the laboratory
were attempts to reproduce and even improve those carried out in a much larger one, the
laboratory of Nature. This analogy implied not only a dynamic conception of Nature, but also
that, in order to know it, its experimental reconstruction was necessary, which made chemistry a
model science to “interpret Nature” from a chemical perspective. This was the path taken by the
philosophers Diderot and d’Holbach. By using a paraphrase of Gilbert Simondon’s
complementary thesis, we want to highlight what the chemists of the Enlightenment were doing
in the laboratory and in manufacturing with their concepts and practices and some philosophical
consequences. Like the blind Saunderson, Diderot’s “conceptual character”, chemists knew the
world through touch, but it was a touch that went beyond sensitive perception, reaching their own
“material sensitivity” thanks to physical and chemical instruments. In the hands of chemists,
thinking and operative action converged, producing a “conceptual mix”, so that they effectively
thought through them.
‘Phenomenal field’ provides philosophical grounds
for designing invisible structure of molecules.
Hirofumi, Ochiai
Nagoya Bunri University, Japan. ochiai.hirofumi@nagoya-bunri.ac.jp

Structure is the term whose proper use is exemplified by an expression like ‘the structure of a
diesel-engine,’ in which what is referred to is accessible to immediate observation. It is also used
figuratively to represent a kind of pattern or relationship among constituents as is shown by ‘a
database structure,’ ‘social structure’ and so on: though invisible, what is referred to is
empirically accessible. Then, what philosophical grounds enable us to refer to invisible structure
of molecules and argue about designing molecules?
Our cognition of objects becomes realized as phenomena when objects are given to our
phenomenal fields. (Ochiai 2020, pp.77-86; 2021, pp.147-174) ‘Phenomenal field’ is a pictorial
representation to show the mind’s self-transcending character and so the relation between ‘self’
and ‘world.’ Molecular structure, or to be more precise, what we assume to be molecular
structure, is composed of chemical elements, bonds, functional groups and so on. Hence, it is not
the mere physical arrangement of atoms but, rather, an appearance of molecules realized in a
particular phenomenal field proper to organic chemists. This implies there are contexts in which
molecular structure makes sense: molecular structure is a context-sensitive dispositional attribute
of an {organic chemist-world} complex, i.e., an affordance. Our intention to observe molecules
from the viewpoint of organic chemistry produces molecular structure as an affordance.
The act of designing molecules aims to create or modify molecules in order to realize certain
chemical/physical properties of compounds. Hence, although it presupposes the concept of
molecular structure, the final arbiter is empirical success: whether it provides intended
compounds and properties. All this implies that the act of designing molecules, too, is an
affordance.
Thus, we can say it is our intention to see ‘structure’ of molecules that enables us to design
molecules. In other words, a phenomenal field in which molecular structure makes sense provides
the act of designing molecules with ontological as well as epistemological grounds.

References
Ochiai, H. (2020). “Overcoming scepticism about molecular structure by developing the concept
of affordance.” Foundations of Chemistry, 22: 77-86.
Ochiai, H. (2021). A Philosophical Essay on Molecular Structure. Cambridge: Cambridge
Scholars Publishing.
Fundamental concepts in chemistry education: A problematization
Bruno Pastoriza, Fernanda Karolaine Dutra da Silva, Tavane da Silva Rodrigues, Laura
Bardini, and Guilherme Brahm dos Santos
Universidade Federal de Pelotas, Brazil. bspastoriza@gmail.com

Studies in General Science (Latour & Woolgar, 1997) and Science Education (Simon, 2013;
Vilani, Dias, & Valadares, 2010) points to the production of a school knowledge that
differentiates in complexity, level, fundamentals and other characteristics from knowledge
produced in scientific research. However, school education has the scientific knowledge as
reference to its work (Pastoriza, Mazzotti, & Loguercio, 2014) and to its formal process of
organizing knowledges. From this, the importance to discuss articulation between scientific and
school fields and its knowledges are put on focus in this research. In this sense, this study,
inserted in a broader investigation, purposes to map and to analyze chemistry fundamental
concepts in Chemistry Education (CE). From a documental analysis into knowledge repositories,
as well as Portal de Periódicos da CAPES, Google Scholar, and Web of Knowledge, 693 peer-
reviewed scientific articles were found. After a preliminary analysis, 460 texts centred in CE
concepts were selected to the study. Each text was analyzed by Contrastive Analysis with focus
in five topics: i) central concept of discussion; ii) area of discussion; iii) context of development;
iv) objectives of the study; v) conclusions of the study and its relation to the concept. From the
analysis, 80 Chemistry concepts were identified in CE texts of which 53 texts (12%) covered
Chemical Bond, 42 (9%) Matter and Substance discussions, 35 (8%) Structure of Matter, 34 (7%)
Chemical Reactions, 32 (7%) Atomism, as well as same number to Thermodynamics and
Thermochemistry. Other concepts were discussed with smaller percentages. Throughout this
study, instead of determining ‘the fundamental’ concept or concepts in CE, we could highlight
some recurrent concepts whom organize stabilities in CE and deserve attention about its
implications and effects in CE. In this sense, we are sure that discuss this centrality or
fundamentality through epistemological and philosophical view could help to expand and to
deepen our knowledge of CE and its contents and framework.

References
Latour, B., & Woolgar, S. (1997). Vida de laboratório: a produção dos fatos científicos. Rio de
Janeiro: Relume Dumará.
Simon, J. (2013). “Cross-National and comparative history of science education: an
introduction.” Science and Education, 22(4): 763-768.
Vilani, A., Dias, V., & Valadares, J. (2010). “The development of science education research in
Brazil and contributions from the history and philosophy of science.” International Journal of
Science Education, 32(7): 907–937.
Pastoriza, B., Mazzotti, T., & Loguercio, R. (2014). “A delimitação do conceito de
representações escolares aplicada à educação em ciências.” Acta Scientiae, 16: 153-163.
Modeling cellular respiration. Educational reflections
Martín Pérgola and Lydia Galagovsky
Instituto CEFIEC, CCPEMS, Facultad de Ciencias Exactas y Naturales,
Universidad de Buenos Aires, Argentina. martinpergola@gmail.com

Scientific models act as mediators between theory and reality (Adúriz-Bravo, Labarca, and
Lombardi, 2014). Scientists, as well as teachers, need to communicate models by using
languages, which are rich in specific and characteristic codes and syntactic formats. Scientific
models' adaptation to different audiences has been called “didactic transposition” (Chevallard,
1998). Novice students often find it difficult to understand scientific models as they lack a
background of specific knowledge.
Problems on Cell Respiration (RC) model understandings could be unveiled by finding some
theoretical elements underpinning traditional teaching discourse. The present work searches for
epistemological roots in modeling the global chemical reaction of RC as a chemical combustion
analogy (Pérgola and Galagovsky, 2020).

References
Adúriz-Bravo, A., Labarca, M., and Lombardi, O. (2014). “Una noción de modelo útil para la
formación del profesorado de química.” Pp. 37-49 in C. Merino, M. Arellano, and A. Adúriz-
Bravo (eds.), Avances en Didáctica de la Química: Modelos y Lenguajes. Valparaíso:
Ediciones Universitarias de Valparaíso, Pontificia Universidad Católica de Valparaíso.
Chevallard, Y. (1998). “La transposición didáctica.” La Transposición Didáctica - Del Saber
Sabio Al Saber Enseñado, 1–67.
Pérgola, M. and Galagovsky, L. (2020). “Estudio didáctico-epistemológico sobre la relación entre
los modelos de respiración celular y de combustión.” Revista de Educación en Biología, 23(1):
49–63.
History and philosophy of chemical modeling
Mike L Renier
Chemistrymeds, Michigan, USA. mlrenier@yahoo.com

The history and philosophy of chemical modeling started in 1968 in England and Israel higher
education institutions. The first computer called Golem-1 for chemical modeling was home built
at the Weizmann Institute in Israel. The language utilized back in the late 1960s was fortran, a
language for math and science purposes. The first chemical modeling project in 1968 was done
with a series of nalkane and cycloalkane molecules and performing conformational,vibrational
spectra and enthalpies calculations on computers (Lifson and Warshel, 1968). The method
utilized was molecular mechanics based rather than quantum mechanics. The workup of the data
used least squares optimization to look at best fit between theoretical calculations and
experimental values. Current chemical modeling studies rarely utilize statistical methods
anymore like this study has or they don’t consider experimental data and just present theoretical
based data.
The second major chemical modelling project back in 1969 was comparing the theoretical
molecular mechanics based structure to x ray crystal structures of lysozyme and myoglobin
proteins (Levitt and Lifson 1969). The molecular mechanics force field consisted of bond lengths,
bond angles, dihedral angles,non bonded interactions and fifth term for restraint purposes to
compensate for the differences between theoretical and experimental x ray structure. Current state
of the art molecular mechanics force fields utilize the same four terms but drop the fifth term.
The math for the dihedral angles and non-bonded interactions is also more complex in current
molecular mechanic functions and another factor is the new polarization term and models. The
theoretical derived structure was also energy minimized to find a lower energy conformation
which is common practice with chemical modeling simulations today. Surprisingly in 1969 the
theoretical derived structures only vary slightly from x ray structures, root mean square deviation
was 0.22 angstroms for lysozyme and root mean square deviation was 0.086 angstroms for
myoglobin. The computing time on the Golem-1 computer for lysozyme (964 atoms) was 18
minutes without including non bonded forces while computing time for myoglobin was much
longer. Its still common in simulations today that molecular mechanics calculations have faster
computing times than quantum mechanics calculations with more complex math functions.
Quantum mechanic energy calculation times are dependent on the number of electrons called size
extensivity in the model molecule system (Boyd and Lipkowitz 2000). One important point about
both quantum mechanics and molecular mechanics methods is that researchers studying sets of
homologous molecules is very common practice today because the methods aren’t suitable for
every molecule (Morozov et al 2004). For example,the general standard quantum methods have
to utilize more recent basis sets to study molecules like free radicals.

References
Boyd, D. and Lipkowitz, K. (eds.) (2000). Reviews in Computational Chemistry, Volume 14.
Weinheim: Wiley-VCH.
Lifson, S. and Warshel, A (1968). “Consistent force field for calculations of conformations,
vibrational spectra, and enthalpies of cycloalkane and n‐alkane molecules.” The Journal of
Chemical Physics, 49: 5116.
Levitt, M. and Lifson, S. (1969). “Refinement of protein conformations using a macromolecular
energy minimization procedure.” Journal of Molecular Biology, 46: 269-279.
Morozov, A. V., Misura, K. M. S., Tsemekhman, K., and Baker. D. (2004). “Comparison of
quantum mechanics and molecular mechanics dimerization energy landscapes for pairs of
ring-containing amino acids in proteins.” The Journal of Physical Chemistry B, 108 (24):
8489-8496.
The evolution of chemical knowledge: How to study it
Guillermo Restrepo and Jürgen Jost
Max Planck Institute for Mathematics in the Sciences, Germany.
restrepo@mis.mpg.de

Chemistry shapes and creates the disposition of the world resources and exponentially provides
new substances for the welfare and hazard of our civilisation. Over the history chemists -- driven
by social, semiotic and material factors -- have shaped the discipline, while creating a colossal
corpus of information and knowledge. Historians, sociologists and philosophers, in turn, have
devised causal narratives and hypotheses to explain major events in chemistry as well as its
current status. In this contribution we discuss the approaches to the evolution of the social,
semiotic and material systems of chemistry. We critically analyse their reaches and challenge
them by putting forward the need of a more holistic and formal setting to modelling the evolution
of chemical knowledge. We indicate the advantages for chemistry of considering chemical
knowledge as a complex dynamical system, which, besides casting light on the past and present
of chemistry, allows for estimating its future, as well as the effects of hypothetical past events.
We describe how this approach turns instrumental for forecasting the effects of material, semiotic
and social perturbations upon chemical knowledge. Available data and the most relevant
formalisms to analyse the different facets of chemical knowledge are discussed.
Hasok Chang on acid
Eric Scerri
Department of Chemistry & Biochemistry, University of California Los Angeles (UCLA), USA.
scerri@chem.ucla.edu

For a period of several years the philosopher of science Hasok Chang has promoted several inter-
related views including pluralism, pragmatism, and an associated view of natural kinds. He has
also argued for what he calls the persistence of everyday terms in the scientific view. Chang
claims that terms like phlogiston were never truly abandoned but became transformed into
different concepts that remain useful. On the other hand, Chang argues that some scientific terms
such as acidity have suffered a form of “rupture”, especially in the case of the modern Lewis
definition of acids. Chang also complains that the degree of acidity of a Lewis acid cannot be
measured using a pH meter and seems to regard this as a serious problem.
My paper will examine all of these views beginning with what Chang claims to be a rupture in
the definition of acidity. I will suggest that there has been no such rupture but a genuine
generalization, on moving from the Bronsted-Lowry theory to the Lewis theory. I will explain
how Chang is taking Lewis’ theory out of context by focusing solely on how it defines acids. In
doing so Chang is negating all the benefits of Lewis’ approach to chemical structure and bonding
in terms of shared pairs of electrons, a concept that it essentially retained in the quantum
mechanical approaches to bonding (molecular orbital theory and valence bond/hybridization). I
will also show how the quantification and measurement of Lewis acidity can easily be realized
through the use of equilibrium theory and the use of stability constants for the formation of
transition metal complexes for example.

References
H. Chang, “The rising of chemical natural kinds through epistemic iteration,” in C. Kendig, ed.,
Natural Kinds and Classification on Scientific Practice, Routledge, London, 2016.
H. Chang, “Acidity, The Persistence of the Everyday in the Scientific,” Philosophy of Science, 79
(December 2012), 690-670.
H. Chang, “The Persistence of Epistemic Objects Through Scientific Change,” Erkenntnis, 75,
413-429, 2011.
E. Scerri, The Periodic Table, Its Story and Its Significance, 2nd edn., Oxford University Press,
2020.
The problem of molecular structure just is the measurement problem
Vanessa Seifert* and Alexander Franklin**
*
University of Bristol, UK. vs14902@bristol.ac.uk
**
King’s College London, UK. alexander.r.franklin@kcl.ac.uk

Molecular structure is central to chemistry; identifying the structure of a molecule is required for
the explanation of chemical, physical, and biological phenomena. Given this, philosophers who
are interested in understanding the relation of chemistry to quantum physics have examined in
detail whether and how successfully quantum physics describes molecular structure from the
bottom up. In this context, three problems have been identified –we refer to these as the problems
of molecular structure. We argue that all these problems are just alternative formulations of a
single problem– the measurement problem of quantum mechanics. Therefore, insofar as the
measurement problem is solved, the problems of molecular structure are resolved as well.
In the first part of the talk, we argue that the problems of molecular structure not only have the
same form but also the same array of putative solutions as the measurement problem. In order to
support this, we present the problems of molecular structure together with the responses to these
problems in the current literature. We then present Maudlin’s (1995) trilemma as a way of
expressing the measurement problem of quantum mechanics, and show how each of the problems
of molecular structure reproduces a version of each lemma, thus recovering the inconsistency of
the trilemma. In the second part, we consider how each solution to the measurement problem can
inform our understanding of molecular structure. We suggest that any philosophical claim about
molecular structure is at least contingent on (if not determined by) the choice of a particular
resolution to the measurement problem.
Overall, we demonstrate that three problems generally distinguished in the philosophy of
chemistry literature ought to be regarded as instances of the same problem –one that is central to
debates in the philosophy of physics. It is a consequence of our argument that the different
interpretations of quantum mechanics have implications throughout the philosophy of chemistry.
Our work should, thus, open the way for a mutually informative interchange between the
foundations of quantum mechanics and the foundations of chemistry communities. While work in
the former has clear implications for the latter discipline, it is also our belief that insights from
chemistry and its foundations could lead to a deeper understanding of the nature of quantum
mechanics. As such, while we are advocating the conflation of the central problems of two
disciplines, this ought not to undermine the value of any of the research programmes discussed.

References
Anderson, P. W. (1972). “More Is Different”. Science 177(4047): 393–396. DOI:
10.1126/science.177.4047.393.
Bahrami, M., Shafiee, A., and Bassi, A. (2012). “Decoherence effects on superpositions of chiral
states in a chiral molecule”. Physical Chemistry Chemical Physics, 14(25): 9214–9218.
Hendry, Robin Findlay (2010). “Emergence vs. reduction in chemistry.” Pp. 205–221 in C.
Macdonald and G. Macdonald (eds.), Emergence in Mind. Oxford: Oxford University Press.
Maudlin, T. (1995). “Three measurement problems.” Topoi, 14(1): 7–15.
Plotting Syntheses of rubber and of DNA: How chemists use
chemical structures and reaction pathways to craft new chemistries
Alok Srivastava
Independent Scholar, USA. alok.srivastava@gmail.com

The abundant variety of robust DNA chemistries currently enabling the research worlds of
Biotechnology and Synthetic Biology got their start in 1955 in the work of Michelson & Todd
who demonstrated the laboratory synthesis of a dimer of DNA - a dinucleotide, two-unit polymer
of DNA - through formation of a single inter-nucleotide linkage. This work demonstrated three
inter-related scientific objects: (i) the specific chemical structure of the inter-nucleotide linkage
from other alternative linkages, (ii) an actual chemical synthetic method - a recipe - for driving
the preferential formation of this linkage from other possible reaction pathways, and (iii) the
chemical pathway - the sequence of changes in chemical structure -for this preferential formation
of the specific linkage between DNA nucleotides. In a few years this triad of chemical structure,
chemical method and chemical pathway had been used, versioned, and elaborated in multiple
laboratories so that some labs were making 20-mer polymers of well-defined sequences. The
world of DNA chemistries was set into vigorous growth and development. Over the next 50
years, the world of DNA chemistries had grown to provide the DNA Synthesis economy of
today’s research and biomedical worlds.
This history of the growth and development of the chemist’s world of DNA chemistries
recapitulates the earlier history of the growth and development of Rubber chemistries and
polymer chemistries. In 1884 the work of Tilden showed the dimerization for isoprene - an
abundant chemical found in rubber sap - into dipentene. This work demonstrated (i) the chemical
formula of the dimer dipentene, (ii) the chemical method of synthesis, and (iii) the chemical
pathway of dimerization. This triad became the basis of many projects of chemists exploring the
making of synthetic rubber. Over the next 50 years these synthetic rubber chemists developed the
worlds of polymerization chemistry, rubber chemistry and chemistries of plastics.
In this paper I use the framework of narrative science being developed by Mary Morgan and
colleagues and apply some principles from literary theory to explore the ways of world-making
of chemists, i.e. how chemists use chemical structure and chemical methods and pathways to
craft new chemistries. In this view, chemical structures function as narrative-plots and chemical
methods (or recipes), and pathways, function as chronicles. Together they work in an
interdependent and interactive manner like an engine for the elaboration of new ‘versions and
visions’ of the world i.e. new chemistries.
Mereology of chemical space and chemical libraries
N. Sukumar
Shiv Nadar University, India. n.sukumar@snu.edu.in

The generation of chemical libraries, their screening and analysis are increasingly important for
drug discovery and materials design. A chemical library is a finite collection of discrete
molecules, assembled by chemists for some specific purpose. So one might naively expect that
chemical libraries would have simple, classical mereology, where the whole is simply the sum of
its parts. However, different chemical libraries have distinct properties (Prabhu, et al, 2016)
which are not identical to those of their sub-libraries or to that of their composite. A chemical
library is thus better represented as a graph than as a set. In this view, chemical libraries are
graphs whose nodes are distinct molecules and whose edges are some pair-wise relation between
molecules. The relation can be one of similarity (leading to a similarity network), of interaction
(giving rise to a molecular interaction network) or of transformation (generating a reaction
network). This graphical representation of chemical libraries makes it easy to understand their
non-classical mereology, in the same way that a molecule (consisting of atoms connected to each
other by bonds) is clearly more than just the sum of its individual atoms (Sukumar, 2013). This
non-trivial mereology gives rise to a distinct eigenvalue spectrum that can be used to quantify the
degree of randomness or otherwise of a chemical library. We will illustrate these concepts with
some real examples.
The concepts of chemical space and chemical libraries are intimately related, but unlike a
chemical library, chemical space need not be finite, discrete, nor consist of an enumerable
number of real molecules. However, distance relationships can still be defined between
molecules in chemical space. The existence of activity cliffs (Maggiora, 2006) in chemical space
(very similar molecules exhibiting a huge difference in their biological activities) again indicates
the possibility of a non-trivial network graph, and thus a non-classical mereology.

References
Prabhu, G., Bhattacharya, S., Krein, M. P., and Sukumar, N. (2016). “Investigation of similarity
and diversity threshold networks generated from diversity-oriented and focused chemical
libraries.” Journal of Mathematical Chemistry, 54(10): 1916-1941.
Sukumar, N. (2013). “The atom in a molecule as a mereological construct in chemistry.”
Foundations of Chemistry, 15(3): 303-309.
Maggiora, G. M. (2006). “On Outliers and Activity Cliffss - Why QSAR Often Disappoints.”
Journal of Chemical Information and Modeling, 46: 1535.
Electronegativity, polarizability and FSGO method
Hiteshi Tandon*, Martín Labarca**, and Tanmoy Chakraborty***
*
Department of Chemistry, Manipal University Jaipur, India. hiteshitandon@yahoo.co.in
**
CONICET – Universidad de Buenos Aires, Argentina. mglabarca@gmail.com
***
Department of Chemistry and Biochemistry, School of Basic Sciences and Research,
Sharda University, India. tanmoychem@gmail.com

Electronegativity is a widely accepted notion in the field of science which assists in describing
relations between species and their probable behaviour or reactions. It is a means through which
electronic arrangement in any atom/molecule/ion, whether static or dynamic, can be understood.
Such information is very valuable in predicting and explaining physicochemical as well as
biochemical properties of a species. So far, this quantity has been commonly employed for
correlating several atomic/molecular properties. Since electronegativity is not a directly
measurable quantity, it cannot be determined through experiments. This has led to the
development of a wide variety of formulations coexisting in the scientific praxis.
In this work, we will introduce an electronegativity model based on polarizability within the
context of Floating Spherical Gaussian Orbital (FSGO) approach. The model provides atomic
electronegativity scale for 120 elements. The calculated electronegativity presents marked
periodicity and relates well with other electronegativity scales. Further, Electronegativity
Equalization Principle is validated by the computed electronegativity data. Some advantages
compared to the Simons scale are also shown. In accordance with the literature, this is the first
FSGO-polarizability based model of electronegativity.

References
Frost, A. A. (1977). “The floating spherical gaussian method.” Pp. 29−49 in Schaefer III, H. F.
(ed.), Methods of Electronic Structure Theory. New York: Plenum Press.
Nagle, J. K. (1990). “Atomic polarizability and electronegativity.” Journal of the American
Chemical Society, 112(12): 4741−4747.
Simons, G., Zandler, M. E., and Talaty, E. R. (1976). “Nonempirical electronegativity scale.”
Journal of the American Chemical Society, 98(24): 7869−7870.
Particular symmetries: From elementary particles to chemical elements
Pieter Thyssen
KU Leuven, Belgium. pieter.thyssen@kuleuven.be

Since Mendeleev’s discovery in 1869, the Periodic System of Chemical Elements has figured as
the undisputed cornerstone of modern chemistry. And yet, Mendeleev’s iconic chart has
remained something of a mystery till the present day. Despite the quantum revolution, a century
ago, the overall structure of the Periodic System remains unaccounted for in quantum-mechanical
terms.
In the early seventies of the previous century, several research groups almost simultaneously
embarked on a group-theoretical, rather than quantum-mechanical, study of the Periodic System.
Their hope was that symmetry might provide a key to the System’s secrets. They found their
inspiration in the recent developments in elementary particle physics, and in particular in the
work of the German physicist Werner Karl Heisenberg. It was the success of Heisenberg’s
approach in elementary particle physics that sparked and fuelled the group-theoretical approach
to the Periodic System.
In this talk, I retrace the intellectual history of ideas that led to these novel symmetry-based
attempts at explaining the Periodic System. The aim of my talk is threefold:
First and foremost, I argue that the symmetry program required a peculiar paradigm shift in
the ontology of the chemical elements. The idea, in essence, consisted in treating the chemical
elements, not as particles, but as distinct states of a superparticle: the baruton. I retrace the origin
of this idea back to elementary particle physics, and in particular to Heisenberg’s suggestion to
treat the proton and neutron, not as particles, but as distinct states of a superparticle: the nucleon.
Second, I show that Heisenberg’s suggestion gave rise to an important paradigm shift. The
materialistic interpretation of the world as consisting of particles gave way to a new
understanding which views the particles as representations of symmetry groups. Symmetries are
ontologically prior to particles. Symmetries, not particles, represent the fundamental level of
reality. Heisenberg depicted this as the confrontation between the atomism of Democritus versus
the idealism of Plato.
Third, I critically examine the prospects and limits of the symmetry program, and the move
from materialism to idealism, from Democritus to Plato. By treating the elements as states of a
superparticle, the Periodic System could be studied as a whole, and the traditional quantum-
mechanical challenges could be circumvented. But I also highlight the inevitable tension that the
symmetry program created between the formal mathematical treatment and the underlying
physical reality.

References
Thyssen, P. and Ceulemans, A. (2020). “Particular symmetries.” Substantia: 4(1): 7-22.
Chemical reactivity and substance as spectrum
Alfio Zambon
Universidad Nacional de la Patagonia San Juan Bosco, Argentina.
alfiozambon@gmail.com

Paneth (1931) introduced the dual nature of the concept of element by distinguishing between
elements as simple substances, according to their phenomenological manifestations, and elements
considered in the abstract sense as basic substances, characterized by a single property, their
atomic number (for more details, see Scerri 2005). For Paneth, they are not two descriptions of
the same entity, product of an epistemic limitation that could be overcome in the future due to
better knowledge. According to the author, the concept of chemical element in itself is endowed
with a double nature (Paneth 1931).
Instead of Paneth’s distinction, I will propose that the concept of substance involves a
phenomenon that unfolds in time and space. The concept does not point to a single kind of
entities, but refers to a spectrum of infinite ways of conceiving chemical phenomena between two
extreme forms. One of them is change manifested in time, abstracted from space; the other
extreme is the geometric relationship between entities, as manifested in space, abstracted from
time. The first case corresponds to the “macroscopic temporal substance” (MTS), as conceived in
macroscopic chemistry; the second case corresponds to what can be called “geometric spatial
substance” (GSS). These two extreme forms of substance bear some relation to Paneth’s
distinction: MTS, which is involved in chemical reactions developed in time, is analogous to
simple substance, whereas GSS has points of contact with basic substance. MTS and GSS are
extreme forms of manifestation of the “substance in itself”.
In this paper, I will introduce a representation of chemical reactivity from this perspective.
First, I will distinguish between reactions involving MTS’s in time (material domain) and
reactions involving GSS’s in geometric space (formal domain). The former kind of reactions can
be represented, as usual, by a graph of energy versus the reaction coordinate. Regarding the
representation of reactions in the geometric space, I will describe a model that can be called
“reaction diagram”, which consists in placing the different substances in concentric circles with
cylindrical projections. I will explain the model and discuss the scope and limitations of the
proposal. Finally I will consider the relationships the two extreme forms of substance.

References
Paneth, F. (1931). “The epistemological status of the chemical concept of element.” [reprinted in
Foundations of Chemistry. 2003, 5: 113-145].
Scerri, E. (2005). “Some aspects of the metaphysics of chemistry and the nature of the elements.”
HYLE – International Journal for Philosophy of Chemistry, 11: 127-145.

You might also like