111
ISPC 2018 BRISTOL CONFERENCE The International Society for the Philosophy of Chemistry Twenty-second annual conference 16-18 July 2018

References. - bristol.ac.uk€¦  · Web viewFoundations of Chemistry (FOCH) is the official journal of the International Society for the Philosophy of Chemistry. FOCH publishes

  • Upload
    lycong

  • View
    214

  • Download
    0

Embed Size (px)

Citation preview

ISPC 2018 BRISTOL CONFERENCEThe International Society for the Philosophy of Chemistry

Twenty-second annual conference

16-18 July 2018

Venue: Pugsley Lecture Theatre, Queen’s Building, University Walk, Bristol, BS8 1TR.

ISPC 2018

Contents1. ISPC2. Organiser and Scientific Committee3. Sponsors; 4. Keynote speakers.5. Practical information; 6. Map of location of chemistry building.7. Conference programme8. Abstracts

Page 2 3 4 4 5 6 7 10

1. 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. See: https://sites.google.com/site/socphilchem/

Philosophy of chemistry concerns both internal questions arising from the methods, concepts, and ontology unique to chemistry and chemical research, as well as traditional questions in the philosophy of science, addressed from a chemical perspective.

ISPC President: Rom Harré, Georgetown University, USA/ Linacre College, Oxford University, UK

The ISPC Executive Committee is: Michael Akeroyd, Bradford College, UK.Marina Banchetti-Robino, Florida Atlantic University, USA.Robin Hendry, University of Durham, UK.Olimpia Lombardi, Universidad de Buenos Aires, Argentina.Guillermo Restrepo, Max Planck Institute for Mathematics in the Sciences, Leipzig,

Germany and Interdisciplinary Center of Bioinformatics, Leipzig Klaus Ruthenberg, Hochschule Coburg, Germany.Eric Scerri, University of California, Los Angeles, USA.Brigitte Van Tiggelen, Chemical Heritage Foundation, USA, and Mémosciences,

Belgium.

Foundations of Chemistry (FOCH) is the official journal of the International Society for the Philosophy 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; modeling 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. See: https://link.springer.com/journal/10698;

2

ISPC 2018

2. Organiser and scientific committee.

The conference is organised by Dr. Geoffrey Blumenthal, Department of Philosophy, University of Bristol, and will be co-hosted by Vanessa Seifert, Department of Philosophy, University of Bristol.

Scientific Programme Committee:

Marina Banchetti-Robino, Florida Atlantic University, USA.José Antonio Chamizo, Universidad Nacional Autónoma de México, Mexico.Hasok Chang, University of Cambridge, UK.Grant Fisher, Korea Advanced Institute of Science and Technology, Korea.Michèle Friend, Georgetown University, USA.Elena Ghibaudi, Università di Torino, Italy.Clevis Headley, Florida Atlantic University, USA.Robin Hendry, University of Durham, UK.Jean-Pierre Llored, Linacre College, Oxford; Laboratoire Sphère, Université Paris 7. Olimpia Lombardi, Universidad de Buenos Aires, Argentina.Hirofumi Ochiai, Nagoya Bunri University, Japan.Guillermo Restrepo, Max Planck Institute for Mathematics in the Sciences, Leipzig,

Germany and Interdisciplinary Center of Bioinformatics, Leipzig Klaus Ruthenberg, Hochschule Coburg, Germany.Eric Scerri, University of California, Los Angeles, USA.Joachim Schummer, Editor of HYLE, Germany.Tami Spector, University of San Francisco, USA.

Queen’s Building.

3

ISPC 2018

Royal Fort House and gardens, University of Bristol.

4

3. Sponsors.

The Conference is sponsored by:The Centre for Science and Philosophy. The British Society for the Philosophy of Science.The British Society for the History of Science.Bristol Faculty of Arts Conference Support Programme.

4. Keynote speakers.

The keynote speakers are:Elena Ghibaudi, Università di Torino, Italy.Robin Hendry, University of Durham, UK.Jean-Pierre Llored, Linacre College, Oxford; Laboratoire

Sphère, Université Paris 7. Eric Scerri, University of California, Los Angeles, USA.

ISPC 2018

5. Practical Information.

The conference will be held in the Pugsley Lecture Theatre, Queen’s Building, University Walk, Clifton, Bristol, BS8 1TR.

All participants must register via the Bristol web-site. Details will be issued separately.This is so that we are able to plan for numbers, e.g. for refreshments.

Registration is free for keynote speakers, students and UoB staff. Other participants are required to pay a small registration fee, which will contribute to the refreshments and lunches provided each day.

Participants (other than keynote speakers) are asked to make their own arrangements for accommodation. Dinners are not included in the conference schedule, although there will be a conference dinner on Wednesday 18th, concerning which additional arrangements will be circulated.

Interior of Wills Building, University of Bristol.

Photo credits: Pages 3,4, 5: University of Bristol Gallery. Title page: Robert Cutts, 11 September 2010, CC BY-SA 2.0., cropped.

5

ISPC 2018

The Queens Building is building 20, accessed from University Walk or Tankards Close.

6

ISPC 2018

7. Programme.

Monday 16 July

8.45 Welcome.Chair: Jean-Pierre Llored

Causation, reasoning and theories in chemistry.

9.15 Vanessa Seifert,Geoffrey BlumenthalJames Ladyman

Welcome to Bristol. Details of the venue.The Centre for Science and Philosophy.

9.30 1

Eric Scerri (keynote)

Introduction to the Conference. Issues about causation in chemistry.

10.30 Coffee break10.45 2

Clevis Headley. Chemical Metaphors and Philosophical Analysis.

11.15 3

Michèle Friend A Formal Representation of Reasoning for Chemistry

11.45 4

Yafeng ShanJonathan Hricko

Davy on Analogical Reasoning.

12.15 5

David Glowacki Molecular dynamics in a multi-user virtual reality framework

12.45 Lunch break

Chair: Eric Scerri Chemistry and philosophical issues 13.45 6

Geoffrey Blumenthal Chemistry, Philosophy, and the Chemical Revolution.

14.45 7

Guillermo Restrepo Poorly explored chemical spaces and the challenges for chemistry

15.15 Coffee break15.30 8

Juan BautistaBengoetxea

Chemical modelling, fictions, and scientific evidence: The case of health claims.

16.00 9

José A. Chamizo C60 diffraction. The limits of the chemical substance,

16.30 10

Philip Stewart From Telluric Helix to Telluric Remix.

17.00 11

Mikhail Kurushkin 32-Column Periodic table and Left-step periodic table united.

7

ISPC 2018

Tuesday 17 July

8.30 WelcomeChair: Marina PaolaBanchetti-Robino

Chemistry and philosophical issues

9.00 12

Rom Harré,Jean-Pierre Llored(keynote)

Hinge Epistemology and Chemistry.

10.00 Coffee break10.15 13

Yona Siderer Philosophical Mediation in Japanese Chemistry Translation & Appropriation

10.45 14 Tami Spector Tracking Taxol: Aesthetics, The Natural, and Synthetic Organic Chemistry.

11.15 15

Klaus Ruthenberg,Juan Camilo MartinezGonzalez.

Radicals – Experiments, Existence, Electrons.

16.15 16 Filip Buyse Robert Boyle and Baruch Spinoza: The redintegration of saltpeter.

12.15 17

Roberto Barbosa de Castilho

The chemical element and the multilevel experience of chemistry

12.45 Lunch break.

Chair: Elena Ghibaudi Chemical elements. 13.45 18

Elena Ghibaudi(keynote)

Dual conception of chemical element: epistemic aspects and implications

14.15 19

Marina PaolaBanchetti-Robino

The Changing Relation between Atomicity and Elementarity: From Lavoisier to Dalton

14.45 20

Mark Leach Basic (Abstract) Substance vs Simple (Real) Substance

15.15 Coffee break15.30 21

W. H. Eugen Schwarz Properties and Systematics of Chemical Elements in Science and Humanities

16.00 22

Sarah Hijmans Chemical Elements and Chemical Substances: Rethinking Paneth’s Distinction

16.30 23

Karoliina Pulkkinen How Mendeleev’s valuing of completeness contributed to predictions

17.00 24

Farzad Mahootian Kant’s Conjoint Reconception of Chemistry and Science.

8

ISPC 2018

Wednesday 18 July

8.30 WelcomeChair: Robin Hendry Quantum chemistry and related topics.

9.00 25 Justin Price Landing Zones – a Case of Model Transfer in Chemistry.9.30 26

Juan Camilo Martinez Gonzalez, Olimpia Lombardi

Indulgence to quantum mechanics.

10.00 Coffee break10.15 27

Sebastian Fortin,Manuel Herrera,Jesus Alberto JaimesArriaga

About the ontological status of phonons.

10.45 28

Jesus Alberto JaimesArriaga, Sebastian Fortin

The Quantum Theory of Atoms in Molecules from a Bohmian perspective

11.15 29

Hirofumi Ochiai Overcoming skepticism about molecular structure by developing the concept of affordance.

11.45 30

James Ladyman On the significance of Heitler and London 1927.

12.15 Lunch break13.15 ISPC business meeting

Chair: James Ladyman

Emergence, chemical bond, identity.

14.15 31

Robin F. Hendry(keynote)

Emergence in Chemistry: Substance and Structure.

15.15 Coffee break15.30 32

Vanessa Seifert The Reality of the Chemical Bond: Revisiting the structural and energetic conceptions of the chemical bond.

16.00 33

Mariana Córdoba, Alfio Ariel Zambon

Old problems in new scientific domains: Identity and nanochemistry

16.30 34

Russell Helder A Bridge to Nowhere: Ensembles in Statistical Mechanics.

19.00 Conference dinner

9

ISPC 2018

Monday 16 July, morning. Causation, reasoning and theories in chemistry.

1. Introduction to the Conference. Issues about causation in chemistry.

Eric Scerri, Department of Chemistry & Biochemistry, [email protected];

I will begin with a brief introduction to the philosophy of chemistry for the benefit of newcomers.One of the main areas of research in the philosophy of science is the question of causation,

about which much has been written. Over the past 50 years or so there has been renewed interest in this topic that had been rendered somewhat obsolete by the logical positivist views on the nature of science.1 Nevertheless, recent work in the philosophy of physics, in particular, seems to suggest a return to the downgrading of causal explanations in favor of purely mathematical explanations.2

There has also been a certain amount of work on causation in chemistry and this seems to have followed the general post-positivist trend alluded to above in invoking causes.3 My paper will consist of a commentary on a recent article on causation in the context of the periodic table by the philosopher Lauren Ross.4

Ross claims that electronic configurations can be regarded as the causes of chemical and physical behavior of the elements on a Woodward-style interventionist account of causation. I respond by claiming that this is only the case in a rather weak sense of causation and that one can do a good deal better by appealing to quantum mechanics and the Schrödinger equation, since they provide a handle on the dynamics of the atom. Meanwhile electronic configurations, that are admittedly viewed as causes, especially in qualitative chemistry and chemical education, only provide a static rather than a dynamical picture. This is due to the fact that any electronic configuration is silent on the question of the nucleus. Although one regularly sees statements that chemistry is determined by the valence electrons, it is really determined by the dynamic interaction of electrons in the context of a particular nuclear charge. Electronic configurations per se only provide a weak form of causation in chemistry. Indeed it could even be argued that they fail to provide correlations given the need to invoke relativistic effects in atoms with heavy nuclei.

The presentation will also discuss causation in chemistry more generally, including speculations concerning the possible existence of downward causation and its relationship to the question of molecular structure and the Born-Oppenheimer approximation.5

1

references E. Sosa, M. Tooley (eds.), Causation, Oxford University Press, Oxford, 1993.2 M. Lange, What Makes a Scientific Explanation Distinctively Mathematical? British Journal for the Philosophy of Science, 64 (2013), 485–511.3 J. Stemwedel, Causes’ in chemical explanations.” In Joseph E. Earley, Sr. (ed.), Chemical Explanation: Characteristics, Development, Autonomy. Ann. N.Y. Acad. Sci., 998:217-226, 2003.4 L.N. Ross, Finding Causal Structure, Synthese, (in press)5 R. Hendry, 2010 Ontological reduction and molecular structure. Stud. Hist. Philos. Mod. Phys. 41, 183–191, (2010); E.R. Scerri, Top-down causation regarding the chemistry-physics interface: a skeptical view, Interface, 3, 20-25, (2012).

10

ISPC 2018

2. Chemical Metaphors and Philosophical Analysis.

Clevis Headley, Department of Philosophy, Florida Atlantic [email protected];

Jaap van Brakel writes that, “Until about 1960, English-language dominated philosophy of science mainly consisted of philosophy of physics. In the eighteen parts of the Foundations of the Unity of Science: Toward an International Encyclopedia of Unified Science, published between 1938 and 1970, the only references to chemistry can be found in Thomas Kuhn’s The Structure of Scientific Revolutions … and, significantly, a few pages on chemical bonding in Philipp Frank’s contribution on the foundation of physics….”1 However, although it is true that philosophy of chemistry was for all practical purposes excluded from the philosophy of science until quite recently, chemistry nevertheless had a discernible theoretical influence upon philosophy. In fact, there is a clear relationship between chemistry and philosophy with regard to various approaches to concept formation. More precisely, many philosophers have employed chemical metaphors to structure their competing accounts of concept creation, transformation, and clarification within philosophy.

Although the use of chemical metaphors is not limited to the analytic tradition in philosophy, this presentation will discuss the role chemical metaphors have played in shaping a specific notion of analysis. Indeed, a notion of analysis that subsequently occupied a formidable role in the origins of analytic philosophy. Bluntly put, the focus of this presentation is not on chemistry itself but on the role played by chemical metaphors in sustaining a particular conception of analysis within the history of analytic philosophy. One can isolate roughly three major models of analysis: The regressive model, the de-compositional model, and the transformative model.

My primary focus will be on Frege’s de-compositional model and on his use of chemical metaphors to achieve two goals: (1) To elucidate the fundamental elements of his logic and semantics, and (2) to frame his conception of philosophical analysis as de-compositional. This discussion is significant because Frege’s use of chemical metaphors has been hugely neglected. In particular, his claim that the basic elements of his logic and semantics were not amenable to further analysis and definition is entirely based on a chemical conception of analysis. Chemical metaphors accommodate Frege’s use of elucidations, hints, and clues to communicate his understanding of objects and concepts.

Bibliography:Beaney, M (2002) “Decompositions and Transformations: Conceptions of Analysis in the Early Analytic and

Phenomenological Traditions,” Southern Journal of Philosophy, 15: 53-99.Conant, J. (2002) “The Method of the Tractatus,” in From Frege to Wittgenstein: Perspectives on Early Analytic

Philosophy, ed. Erich Reck, Oxford: Oxford University Press: 374-462.Frege, G. (1970) Translation from the Philosophical Writings of Gottlob Frege, translated by Peter Geach and Max

Black. Oxford: Blackwell._______ (1979) Posthumous Writings, tr. Peter Long and Roger White, Oxford: Blackwell.Heck, R. and May, R. (2013) “The Function is Unsaturated,” M. Beaney (ed.) The Oxford

Handbook of Analytical Philosophy. Oxford: Oxford University Press: 825-850.Picardi, E. (1992) “ The Chemistry of Concepts,” in Language and Earth: Elective Affinities Between the Emerging

Sciences of linguistic and Geology, Bernd Naumann, Frans Plank, Gottfried Hofbauer (eds.). Amsterdam: Benjamins: 125-146.

Ruschig, U. (2001) “Logic and Chemistry in Hegel’s Philosophy,” HYLE - International Journal for Phhilosophy of Chemistry, Volume 7, Number 1: 5-22.

1 J van Brakel, “Philosophy of Science and Philosophy of Chemistry,” HYLE—International Journal for Philosophy of Chemistry, Volume 20, Number 1 (2014):11.

11

ISPC 2018

Schneider, H. (1990) “Syntactic Metaphor: Frege, Wittgenstein, and the Limits of a Theory of Meaning,” Philosophical Investigations, Volume 13, Number 2: 137-153.

Shaw, J. L. (1989) “‘Saturated and Unsaturated’: Frege and the Nyaya,” Synthese 80: 373-394.Textor, M. (2008) “Unsaturatedness; Wittgenstein’s Challenge, Frege’s Answer,” Proceedings of the Aristotelian

Society, Volume 109: 61-82Urquhart, A. (2008) “The Unnameable,” Canadian Journal of Philosophy, Volume 38, Supplement Volume 34: 119-

135.van Brakel, J. (2014) “Philosophy of Science and Philosophy of Chemistry,” HYLE— International Journal for

Philosophy of Chemistry, Volume 20, Number 1:11-57van Brakel, J. (2000) Philosophy of Chemistry: Between the Manifest and the Scientific Image (Leuven; Leuven

University Press)

12

ISPC 2018

3. A Formal Representation of Reasoning for Chemistry

Michèle Friend, Department of Philosophy, George Washington [email protected];

As we know, what is taken to be the underlying logic of an area of research guides and shapes that area of research. For example, there are pointed arguments to the effect that economics should be based on a constructive, or intuitionistic, logic rather than a classical logic.            The chemists studying: quantum chemistry, electro-chemistry, nano-chemistry, the chemistry of acids and bases, the chemistry of metals and so on; quite often are more concerned with properties, relations and sometimes functions, than they are with objects, or even elementary particles (qua only particles). The location boundary of a fluid or gas are vague, the purity of a substance is almost always non-existent, the mixing of fluids or gasses often yields emergent properties, chemical reactions depend on context, or ‘milieu’.             Standard formal representations of reasoning, assume, in: the grammar, the notation, the operations, that we begin with elements, or objects, or members of a domain. These formal systems representing reasoning are ill-suited to representing reasoning in chemistry.             I propose to give the beginnings of a formal representation of reasoning, departing from Lemmon’s 1960s development of “a formal logic (sic!) of attributes” and attempts at giving a semantics and syntax of mass expressions using metrology or set theory or both. In the formal representation I propose, the notion of ‘object’ or ‘element’ will be secondary and derivative, being merely a collection of properties vaguely located within a milieu. In contrast, the notions of ‘property’, ‘relation’ and ‘open operator’ will be central. The ‘open operator’ concept is new. They stand for ceteris paribus clauses and ‘milieu’.             Supplying chemists with a formal representation of reasoning tailored to chemist’s needs and practice, will bring a rigour of reasoning and a clarity of thought. Pluralists in logic, who think of formal representations of reasoning as various ways of regimenting thought in a subject area, might find that the formal system(s) developed for chemistry will be useful in other areas of thinking that are more concerned with properties and processes than with objects.

13

ISPC 2018

4. Davy on Analogical Reasoning.

Yafeng Shan, Department of Philosophy, University of Durham, and Jonathon Hricko, National Yang-Ming University.

[email protected];

When discussing forms of reasoning in scientific practice, philosophers of science usually focus on deductive reasoning (Popper 1959), inductive reasoning (Hempel 1946; Howson and Urbach 1989), and abductive reasoning (Peirce 1932; Lipton 2004). Deductive reasoning and inductive reasoning have been widely regarded as two main modes of reasoning in the history and practice of science. James Ladyman (2002), for example, characterises a crucial feature of the Scientific Revolution in the 17th century as a shift away from Aristotelian deductive reasoning to Baconian inductive reasoning. Abductive reasoning was first articulated by Charles S. Peirce (1932) as an alternative to both deductive and inductive reasoning. In contrast, though analogical reasoning, or analogy, is frequently employed in scientific practice, its significance has to some extent been overlooked by philosophers of science. Analogy has been regarded as a part of inductive reasoning (Keynes 1921; Hacking 1983; Copi, Cohen, and McMahon 2014). ‘Analogy’ was also sometimes construed as a synonym for ‘model’ (Achinstein 1964). Although its role in modelling and theory construction has been widely discussed (Hesse 1963; Agassi 1964; Carloye 1971; Girill 1972; Darden 1982; Harre 1988), there is arguably no comprehensivephilosophical analysis of analogical reasoning in scientific practice more generally. In particular, the role of analogical reasoning in experimentation has not been sufficiently examined.

In contrast to philosophers of science, scientists have, for some time, recognised the role of analogy in aspects of scientific practice besides modelling and theorising. Chemists (e.g., Priestley 1775; Kirwan 1797) working in the late eighteenth and early nineteenth centuries were particularly explicit regarding the importance of analogy to scientific practice. Humphry Davy, for example, writes in his Elements of Chemical Philosophy:

The foundations of chemical philosophy, are observation, experiment, and analogy. By observation, facts are distinctly and minutely impressed on the mind. By analogy, similar facts are connected. By experiment, new facts are discovered; and, in the progression of knowledge, observation, guided by analogy, leads to experiment, and analogy, confirmed by experiment, becomes scientific truth. (Davy 1812, 1)Although he never explicitly defines ‘analogy’, Davy’s work on chemical decomposition in

his 1806 and 1807 Bakerian Lectures provides a number of illustrations of the kind of analogical reasoning that he has in mind. Analogical reasoning, for Davy, is a way of reasoning by looking for similarities to guide the enquiry in order to acquire generalised facts and knowledge in science . It should be noted that, on Davy’s understanding, analogy is not a simple process of theorising or modelling by accumulating similar observed facts and generalising to a universal statement. Analogy is involved in the process of both generating and justifying the hypothesis. In addition, analogy, observation, and experiment are not three independent activities. Rather they are mutually intertwined in practice. Observation is not only guided by analogy as Davy suggests; it also provides the foundation for analogical reasoning. Moreover, experiment is not merely undertaken to test the hypothesis proposed by analogy; it is also designed with the help of analogical reasoning.

The purpose of this paper is to gain a greater understanding of the nature and role of analogical reasoning in scientific practice by examining Davy’s use of analogical reasoning in his work in electrochemistry. We shall review Davy’s work on electrochemical decomposition reported in

14

ISPC 2018

his 1806 and 1807 Bakerian Lectures. Then we shall analyse and examine the nature and role of Davy’s analogical reasoning in his work on chemical decomposition. Finally we shall argue that Davy’s use of analogical reasoning in his work on electrochemical decomposition suggests a new account of analogical reasoning, which not only guides modelling and theorising, but also instructs the way of experimentation, and echoes the practical turn in contemporary philosophy of science.

6. Bibliography:Achinstein, Peter. 1964. “Models, Analogies, and Theories.” Philosophy of Science 31 (4): 328–50.Agassi, Joseph. 1964. “Analogies as Generalizations.” Philosophy of Science 31 (4): 351–56.Carloye, Jack C. 1971. “An Interpretation of Scientific Models Involving Analogies.” Philosophy of Science 38

(4): 562–69.Copi, Irving M., Carl Cohen, and Kenneth McMahon. 2014. Introduction to Logic. 14th ed. Harlow: Pearson.Darden, Lindley. 1982. “Artificial Intelligence and Philosophy of Science: Reasoning by Analogy in Theory

Construction.” PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1982: 147–65.

Davy, Humphry. 1807. “The Bakerian Lecture [for 1806]: On Some Chemical Agencies of Electricity.” Philosophical Transactions of the Royal Society of London 97: 1-56.

——— 1808. “The Bakerian Lecture [for 1807]: On Some New Phenomena of Chemical Changes Produced by Electricity, Particularly the Decomposition of the Fixed Alkalies, and the Exhibition of the New Substances Which Constitute Their Bases; and on the General Nature of Alkaline Bodies.” Philosophical Transactions of the Royal Society of London 98: 1-44.

——— 1812. Elements of Chemical Philosophy. Philadelphia, PA: Bradford and Inskeep.Girill, T R. 1972. “Analogies and Models Revisited.” Philosophy of Science 39 (2): 241–44.Hacking, Ian. 1983. Representing and Intervening: Introductory Topics in the Philosophy of Natural Science.

Cambridge University Press.Harre, R. 1988. “Where Models and Analogies Really Count.” International Studies in the Philosophy of

Science 2 (2): 118–33.Hempel, Carl Gustav. 1946. “Studies in the Logic of Confirmation.” Mind 54: 1–26, 97–121.Hesse, Mary B. 1963. Models and Analogies in Science. Notre Dame, IN: University of Notre Dame Press.Howson, Colin, and Peter Urbach. 1989. Scientific Reasoning: The Bayesian Approach. La Salle, IL: Open

Court.Keynes, John Maynard. 1921. A Treatise on Probability. London: Macmillan and Co., Limited.Kirwan, Richard. 1797. “Thoughts on Magnetism.” The Transactions of the Royal Irish Academy 6: 177–91.Ladyman, James. 2002. Understanding Philosophy of Science. London and New York: Routledge.Lipton, Peter. 2004. Inference to the Best Explanation. 2nd ed. London and New York: Routledge.Peirce, Charles Saunders. 1932. Collected Papers of Charles Sanders Peirce. Edited by C. Hartshorne, P.

Weiss, and A. Burks. Cambridge, MA: Harvard University Press.Popper, Karl. 1959. The Logic of Scientific Discovery. 1sted. London: Hutchinson & Co.Priestley, Joseph. 1775. The History and Present State of Electricity with Original Experiments Vol. II. London:

Printed for C. Bathurst, T.Lowndes, in Fleet Street; J. Rivington, and J. Johnson, in St. Paul’s, Church-Yard; S. Crowder, G. Robinson, and R. Baldwin, in Paternofter Fow; T. Bicket, and T. Cadell, in the Strand.

15

ISPC 2018

5. The phenomenology of molecular perception in virtual reality

David Glowacki1,2,3

1Centre for Computational Chemistry, School of Chemistry, University of Bristol; 2Department of Computer Science, University of Bristol, 3Pervasive Media Studio, Watershed, Bristol, UK

Over the past few years we have developed a framework for real-time interactive molecular dynamics in a multiuser virtual reality (VR) environment, combining rigorous atomistic physics simulations run on cloud-mounted supercomputers with commodity VR hardware (1-3). This system allows users to not only visualize, but to reach out touch, with surgical, atomic-level precision, the structures and dynamics of complex molecular structures ‘on the fly’, and also to interact with other users in the same virtual environment. A series of controlled studies, wherein participants were tasked with a range of molecular manipulation goals (threading methane through a nanotube, changing the screw-sense of a helical molecule, and tying a protein knot), quantitatively demonstrate that users within the interactive VR environment can complete sophisticated molecular modelling tasks more quickly than they can using more conventional interfaces. To date, we have put thousands of people into VR, allowing them to “touch and feel” the dynamics of a wide range of different molecular structures.

The use of sophisticated VR in the medical field has been widespread for nearly a decade now, with several studies having shown that VR-trained surgeons complete surgical procedures faster, with significantly lower error rates. However, the use of VR in surgical contexts, where it is intended to simulate a surgeon’s experience of manipulating human tissue, is rather distinct from the use of VR to manipulate molecular structure and dynamics. Whereas surgical applications have a well-defined and measurable design reference (that is, how does the VR simulation “feel” compared to an experience involving human tissue?), molecular applications have no similarly well-defined design reference (that is, what does [or should] a molecular system feel like?). This lack of reference is what makes developing a real-time molecular simulation and manipulation framework such a fascinating challenge, which must necessarily consider aesthetics, design, and user psychology to be effective. In this talk, I will reflect on some of some of the design questions we have faced in developing this system, and also comment on our user observations to date – particularly how users own phenomenological accounts of experiencing tangible nanoscale dynamics.

(1) M. O Connor, H. M. Deeks, E. Dawn, O. Metatla, A. Roudaut, M. Sutton, B. R. Glowacki, L. M. Thomas, R. Sage, P. Tew, M. Wonnacott, P. Bates, A. J. Mulholland, D. R. Glowacki, “Sampling molecular conformations and dynamics in a multi-user virtual reality framework”, 2018, arXiv:1801.02884, Science Advances, accepted

(2) https://vimeo.com/244670465 (3) https://vimeo.com/274862765

16

ISPC 2018

Monday 16 July, afternoon. Chemistry and philosophical issues.

6. Chemistry, Philosophy, and the Chemical Revolution.

Geoffrey Blumenthal, Department of Philosophy, University of Bristol. [email protected];

This paper builds on our previous work in arguing that a number of major questions in the history, meta-history, philosophy and social studies of chemistry can be resolved or illuminated via a far more detailed and analytical study of the primary literature, in this case of the period of the Chemical Revolution, than is usual in such studies. Some of these questions are as follows:

How are different theories compared? Kuhn’s (1970) view that different “paradigms” were incommensurable and untranslatable, so that they were compared in terms of epistemic virtues, can be comprehensively disconfirmed with regard to actual practice, with evidence from the Chemical Revolution (contrary to a large literature, e.g. Best 2015). In actuality, there was a very large number of aspects of the available theories of which differing sets were taken into account by different participants when comparing the theories (Blumenthal and Ladyman 2017b).

How are theories chosen from among those available? In practice, each main individual participant gave their own list of reasons for choosing and changing theory. There was no stage at which “the other side gave up the struggle” (Duhem) or “the profession agreed” (Kuhn 1970) fully, but general change occurred as an aggregate of individual choices (Blumenthal and Ladyman 2017b).

Under what general circumstances does the choice between theories become clearer? The events of the Chemical Revolution showed that the more experimental evidence that needed to be satisfactorily and consistently explained by a theory that was as experimentally-testable as practicable, then the more the potential contenders failed until the best available theory was the only remaining choice (Blumenthal and Ladyman 2017a).

When and in what circumstances is pluralism useful in chemistry? The failures of the phlogiston theories provide a fund of information about types of plurality that did not prove to be useful, while plurality was essential during the development of the new chemistry. Neither monism nor pluralism expresses what does or should happen in general in chemistry (Blumenthal, under review).

Can issues in science be decided on the basis of experimental evidence? Under some circumstances during the Chemical Revolution, some participants decided to take the evidence from one of two particular experiments as being crucial (e.g. Mauskopf 2002, 205; Blumenthal and Ladyman, under review).

Under what circumstances is it possible for participants to retain their favoured theories indefinitely, and what are the consequences? Any participant could do so provided that they concentrated on single experiments, did not require their theories to be consistent between experiments, and were prepared to hypothesise experimentally-untestable substances as constituents whenever this was necessary to protect a theory against disconfirmation. Yet if they did so, their theories were not usable by others, and so the price of these tactics was being left out of the progress of chemistry. Accordingly, theory choice for the future progress of chemistry could effectively be decided while disregarding theories that were the results of such tactics (Blumenthal, under review).

To what extent are social factors important in chemistry, and to what extent are they determinative of the chemistry itself? Evidence from the Chemical Revolution shows both that the development of chemistry had intensely social aspects, and that these did not determine the chemistry itself (e.g. Blumenthal and Ladyman, under review).

17

ISPC 2018

Was there a Chemical Revolution? The direct evidence is that approximately a score of participants including opponents stated that there was an ongoing revolution in chemistry, or had just been one. For example, Kirwan (1785, 337) stated that if Lavoisier’s theory “should be established, all that we have hitherto regarded as well-founded will be overturned”. The very high degree of concern about this was signalled by contemporary opponents’ use of ad hominem rhetoric. For example Crell (in Hagen 1850, 84) talked of “our phlogiston, which the French wish to steal from us”. Those writers in the secondary literature who have argued that there was no such thing as a Chemical Revolution (e.g. Klein 2015) have not taken this primary literature into account.

Do scientific revolutions involve “ruptures”, or do they involve a great deal of continuity and incorporation, as well as major changes? The second of these was the case in the Chemical Revolution, in which the resulting chemistry incorporated ideas that derived from every tradition that had contributed to chemistry, including alchemy, metallurgy and pharmacy.

What was revolutionary about the Chemical Revolution? Cuvier (1810, 78) noted that before Lavoisier chemistry had been a labyrinth which had “been explored by many industrious men” without overall comprehension, whereas Lavoisier knew how to “understand its layout with the eye of an eagle”. In general, this paper argues that the Chemical Revolution involved a consistent set of several changes, which will be explained in detail, and all of which were targeted at specific widespread general issues with the previous chemistry.

How do previous general theories of scientific revolutions compare with what occurred during the Chemical Revolution? Neither Kuhn’s nor any other general account of scientific revolutions or scientific change (e.g. Barseghyan 2015) is of considerable use in explaining the Chemical Revolution.

How do previous detailed proposals for the Chemical Revolution compare with this account? To take a few examples, Wurz (1869) inaccurately claimed that “Chemistry is a French science: it was constituted by Lavoisier”, but Volhard (1870) inaccurately claimed that Lavoisier was a dilettante plagiarist who only inverted Stahl’s theory, a view later followed in general by Gough (1988). Many accounts have assumed that the major changes were to do with atomism, (e.g. Siegfried 2002 and studies by Newman) which have been effectively answered by Chalmers (e.g. 2009). The questions concerning the sources of the changes in the Chemical Revolution, debated in a long series of papers by Newman (e.g. 2010), Principe (e.g. 2007), Klein (e.g. 2015), Chalmers (e.g. 2009), etc., will be discussed in the light of the primary literature during the Chemical Revolution. Chang (2015) and Klein (2015) have criticised each other’s views on the Chemical Revolution. This paper will differ from many arguments by Best (2016).

References.Barseghyan, H. The Laws of Scientific Change. Dordrecht, Netherlands: Springer (2015) Best, N. Meta-Incommensurability between Theories of Meaning: Chemical Evidence. Perspectives on Science,

23, 361-378 (2015)Best, N. What was Revolutionary about the Chemical Revolution? In: E. Scerri, G. Fisher (eds), Essays in the

Philosophy of Chemistry, 37-59. Oxford: Oxford University Press (2016) Blumenthal, G. Some late phlogiston theories, the debates on theory choice in Germany 1787-98 and

implications concerning pluralities and pluralisms in chemistry (Under review)Blumenthal, G., Ladyman, J. The development of problems within the phlogiston theories, 1766–1791. Found.

Chem, 19, 241–280 (2017a)Blumenthal, G., Ladyman, J. Theory comparison and choice in chemistry, 1766-91. Found. Chem. Online

(2017b)Blumenthal, G., Ladyman, J. On apparently “crucial” experiments in chemistry (Under review). Chalmers, A. The Scientist’s Atom and the Philosopher’s Stone. Dordrecht: Springer (2009) Chang, H. The Chemical Revolution revisited. Studies in History and Philosophy of Science, 49, 91-98 (2015)Cuvier, G. Rapport historique sur les progrès des sciences naturelles depuis 1789, et sur leur état actuel. Paris,

France: Imprimerie Impériale (1810)

18

ISPC 2018

Gough, J. Lavoisier and the Fulfilment of the Stahlian Revolution. Osiris 2(4), 15-33 (1988)Hagen, A. K. G. Hagen’s Leben und Wirkung. Neue preussiche Provincial-Blätter, 9, 46-86 (1850)Kirwan, R. Vom Hrn. R. Kirwan in London, Chemische Annalen, 1785:1, 335-337 (1785)Klein, U. A Revolution that never happened. Studies in History and Philosophy of Science, 49, 80-90 (2015) Kuhn, T.S. The Structure of Scientific Revolutions. 2nd edn. Chicago: University of Chicago Press. (1970).Mauskopf, S. Richard Kirwan’s Phlogiston Theory: its Success and Fate. Ambix, 49, 185-205 (2002)Newman, W. How not to integrate the history and philosophy of science: a reply to Chalmers. Studies in History

and Philosophy of Science, 41, 203-213 (2010)Principe, L. A Revolution Nobody Noticed? Changes in Early Eighteenth-Century Chymistry. In: L. Principe,

(ed.) New Narratives in Eighteenth-Century Chemistry, 1-22. Dordrecht: Springer (2007) Siegfried, R. From Elements to Atoms: a History of Chemical Composition. Transactions of the American

Philosophical Society, 92, nr. 4. (2002)Volhard, J. Die Begründung der Chemie durch Lavoisier. Journal für praktische Chemie, 110, 1-47 (1870)Wurtz, A. History of Chemical Theory: from the age of Lavoisier to the Present Time. London: Macmillan.

translation of (1869) Histoire des doctrines chimiques: depuis Lavoisier jusqu’a nos jours. Paris:

19

ISPC 2018

7. Poorly explored chemical spaces and the challenges for chemistry

Guillermo Restrepo,Max Planck Institute for Mathematics in the Sciences, Leipzig, GermanyInterdisciplinary Center of Bioinformatics, Leipzig University, Germany.

[email protected];

Mathematical and computational tools estimate that the number of energetically stable chemical compounds is greater than 1060 [1], which is called the chemical space. Despite the exponential growth of chemical substances [2-4], chemists have explored less than 0.1% of the space [5-7]. If synthesis of new substances is an end in itself for chemistry [8], the small exploration of the space is discouraging. This is even worse, for the narrow range of reactions conditions, the preferred use of a small number of reactants and of reactions has been suggested as the cause of the poorly explored chemical space [4, 7, 9-11]. The main problem this brings up touches the foundations of chemistry, for it is strongly related to the chemical way of thinking [12, 13] and the way chemistry is taught [13].

The question that arises is whether chemistry is bound to keep tinkering in such a narrow space, as suggested on the basis of statistical bounds for the growth of substances [3], or whether it is possible to free it by pushing the boundaries of the traditional way of chemical thinking [7, 13].

In this paper I will discuss these theses and their implications for the chemistry of the future, which includes questions on what and how to teach chemistry for the future generations [13].

1. Reymond, J-L. Acc. Chem. Res. 2015, 48, 722-730.2. Schummer, J. Scientometrics 1997, 39, 107-123.3. Fialkowski, M. et al. Angew. Chem. Int. Ed. 2005, 44, 7263-7269.4. Llanos, E. J. et al. Book of Abstracts, 254th American Chemical Society National Meeting & Exposition,

Washington, D. C., August 20-24, 2017; American Chemical Society: Washington, DC, 2017; CINF-14.

5. Cernak, T. Chem 2016, 1, 6-9.6. Pye, C. R. et al. PNAS 2017, 114, 5601-5606.7. Keserü, G. M. et al. Chem. Soc. Rev. 2014, 43, 5387-5399.8. Schummer, J. Scientometrics 1997, 39, 125-140.9. Bajorath, J. Expert Opin. Drug. Dis. 2016, 11, 825-829.10. Brown, D. G.; Boström, J. J. Med. Chem. 2016, 59, 4443-4458.11. Schneider, N. et al. J. Med. Chem. 2016, 59, 4385-4402.12. Bensaude-Vincent, B. Ber. Wissenschaftgesch. 2009, 32, 365-378.13. Schummer, J. Educ. Quím. 1999, 10, 92-101.

20

ISPC 2018

8. Chemical modelling, fictions, and scientific evidences: The case of health claims Juan Bautista Bengoetxea, University of the Balearic Islands. [email protected];

According to the fictionalism about scientific knowledge (partially derived from the fictionalism about literary works), models do not always account for real phenomena. Often, models are taken as ‘stand-ins’ or subrogates for systems that in fact have not been instantiated (Godfrey-Smith 2006). That is, because of idealisation and abstraction, model systems’ descriptions are not close to any complete description of real phenomena. This question is related to the notion of truth, about which I would like to present a brief pragmatic account of truth in order to understand a specific aspect of the practice in chemistry, i.e. the activity of idealisations. I discuss this activity through the analysis of the kinetic theory of ideal gases in which the law of van der Waals is to be taken as a concretization of Boyle’s law. This is actually a suggestion to replace idealisational models such as those of Nancy Cartwright by a dialectic process of idealisation-concretisation, which might tell us why fundamental laws are false but acceptable. This will allow us to incorporate the notion of dynamic model (Zeidler 2000), which can take various forms; among them, it is the form of mechanisms of action in chemistry.

The previous reflections allow us to qualify two pertinent points that could serve to see how chemical modelling could act as an element of the practice of obtaining evidences for, among other things, make regulatory decisions. On the one hand, modelling should eliminate, as far as possible, the vestiges of subjectivity attributed to it by the fictionalism of Kendall (1990): If a model is just a space for what is necessary to maintain in order to carry out the conceptual construction by an agent, it would important to understand how the same kind of props (fictional element) is a reliable source of the same kind of conceptual constructions. This qualification, on the other hand, permits us to see that in chemistry is usual to handle fictional propositions. Some are false about, for example, the particles and the collisions in the ideal gas modelled world (Toon 2012), but they belong to the model and thus they are just fictional, not merely false. But we can say that other ones are (approximatively) true. This is why some accounts (Toon’s, for instance) need to distinguish between fictional true and false propositions about a target, especially if an account intends to explain how a modeller may learn about targets. If we cannot offer a criterion for this distinction, it fails as an epistemology of models in science (chemistry) (Poznic 2016, Morrison 2009).

This is where Zeidler's notion of a dynamic model (2000) can play a relevant role. In particular, I am interested in both re-situating his 'anti-truth' picture of models in terms of empirical adequacy (van Fraassen 1980) and showing that dynamic models can recover the meaning of the notion of mechanism of action (in a sense similar to that of the symbolic models of Tomasi (1999). This partiality may be articulated through the steps incorporated to mechanisms of action, which use representational elements of either empirical fields (here researchers would have some referent (Francoeur 2000) or simulated fields (computational modelling without material referents) (Fisher 2017, Mainzer 1999, Morrison 2015).

In science, the transition from black boxes (phenomenological hypotheses) to translucent ones (mechanism-based hypotheses) is something really significative (we can say that we move from a descriptive modelling to a modelling closer to explanation (Bunge 2012) and might help us adequately situate the role fiction plays in chemical modelling. To see this, we propose the case of the obtention of scientific evidence for health claims linked to anti-oxidants. As De Boer et al. (2014) point out, after several Europe-wide food scares in the 1990s, there was call to reform European food law and to a General Food Law (GFL), which entered into force in 2002. In addition to the GFL, the EU adopted a great number of specific rules (Nutrition and Health Claim Regulation 2006 (NHCR)) dealing with various aspects of the food chain and specific food components, a regulation that required the information on the label provided to consumers to be based on scientific evidence. The

21

ISPC 2018

NHCR aims to ensure a high level of protection for consumers and to facilitate their choice by making sure claims are scientifically substantiated. The substantiation is directly related to the use of mechanisms of action, which here we adopt as models that use fictional ingredients.

New insights in the mechanism of action of antioxidants are important for the substantiation of health claims on antioxidants. Our proposal will present (very briefly) the mechanism of action of antioxidants as a case of a dynamic model that incorporates, as has been said, idealising elements. In the benefit assessment of foods, it is assumed that the main source of evidence required for authorization of a health claim are human intervention studies, usually randomized controlled trials (RCT), because of their capacity for identifying causal relationships between intake for an ingredient and a specific outcome. In this respect, John Worrall (2002) developed a critique to the supposition that RCT (at least, in medicine) discards all the confounding factors. This consideration is correct: No experimental trial can show that a strong statistical correlation proves that there is a causal relationship. In order to prove the existence of such a relationship, statistical «evidence» should be reinforced with the help of laboratory studies; that is, there may be causality if and only if, in addition to obtaining «true positives», we can prove the existence of a mechanism of action. This kind of chemical modelling is one of the means of scientific substantiation for health claims based on a type of dynamic modelling that has what fictionalists would call 'non-referential' elements.

References.

Bunge, Mario (2012): Filosofía para médicos. Barcelona: Gedisa.De Boer, Alie; Vos, Ellen; Bast, Aalt (2014): Implementation of the nutrition and health claim regulation – The

case of antioxidants. Regulatory Toxicology and Pharmacology 68: 475-487.Fisher, Grant (2017): Content, design, and representation in chemistry. Foundations of Chemistry 19: 17-28.Francoeur, Eric (2000): Beyond dematerialization and inscription. HYLE-International Journal for Philosophy

of Chemistry 6: 63-84.Gelfert, Axel (2016): How to Do Science with Models: A Philosophical Primer. Dordrecht: Springer.Godfrey-Smith, Peter (2006): The strategy of model-based science. Biology & Philosophy 21(5): 725–740.Heaney, Robert P. (2008): Nutrients, Endpoints, and the Problem of Proof. The Journal of Nutrition 138: 1591-

1595.Kendall, Walton (1990): Mimesis as Make-Believe: On the Foundations of Representational Arts. Cambridge,

Mass.: Harvard University Press.Mainzer, Klaus (1999): Computational Models and Virtual Reality: New Perspectives of Research in Chemistry.

HYLE-International Journal for Philosophy of Chemistry 5: 135-144.Morrison, Margaret (2009): Fictions, Representations, and Reality. In Suárez (ed.), pp. 110-135.Morrison, Margaret (2015): Reconstructing Reality: Models, Mathematics, and Simulations. Oxford: Oxford

University Press.Parker, Wendy S. (2009): Does matter really matter? Computer simulations, experiments, and materiality.

Synthese 169: 483-496.Poznic, Michael (2016): Make-Believe and Model-Based Representation in Science: The Epistemology of

Frigg’s and Toon’s Fictionalist Views of Modeling. Teorema XXXV/3: 201-218.Richardson, David P. (2011): Nutrition and health claims: hel or hindrance. Preparing dossiers: strenght of the

evidence and problems of proof. Proceedings of the Nutrition Society, 71: 127-140.Tomasi, Jacopo (1999): Towards ‘chemical congruence’ of the models in theoretical chemistry. HYLE-

International Journal for Philosophy of Chemistry 5: 79-115.Toon, Adam (2012): Models as Make-Believe, pp. 108-130.Van Fraasen, Bas C. (1980): The Scientific Image. Oxford: Oxford University Press.Worrall, John (2002): What evidence in evidence-based medicine? Philosophy of Science 69: S316-S330.

Zeidler, Paweł (2000): The Epistemological Status of Theoretical Models of Molecular Structure. HYLE-International Journal for Philosophy of Chemistry 6: 17-34.

22

ISPC 2018

9. C60 diffraction. The limits of the chemical substance.

José A. Chamizo, Facultad de Química, UNAM, Mé[email protected];

Beyond the different ways of characterizing chemical substances (Schummer, 2002) and the conflicting definitions established by the American Chemical Society (through Chemical Abstracts) and the IUPAC, substances always have a limited existence. Two examples of this are:

Related to energy, water dissociates in hydrogen and oxygen at temperatures higher to 500 oC. Related to time, argon dichloride dissociates in argon and chorine in 10-12 s.

For the last reason, Bachelard’s 1940 words, the old conception of a substance that is by definition outside time can no longer be maintained, sounded prophetic. These are some of the reasons why van Brakel (van Brakel, 2012 p. 222) indicated: Perhaps we should limit the notion of substance to what can exist independently “in bottles”. That is to say, the chemical substance exists in a particular context with particular limits, in which specific pressure and temperature conditions are required.

To these limits of chemical substance, we should add another one identified with an extraordinary experiment carried out in 1999, at the end of the fifth chemical revolution (Chamizo, 2017). That year a team of Austrian physicists took, from a bottle of Dynamic Enterprises Ltd, a British commercial company, the solid and black substance C60 (99.5% pure as determined by HPCL with CAS number 99685-96-8). No doubt, C60 is a substance in a bottle. Subsequently, they sublimated it, collimated the emerging molecular beam and diffracted it through a double slit. In that experiment, a multitude of those molecules, one by one, coming from that substance, were the most massive and complex object in which wave behaviour has been observed (Arndt et al, 1999). Thus, from an unequivocal and material chemical substance with well-characterized physical and chemical properties (for example, density 1.65g/cm3, insoluble in water), a multitude of chemical species are extracted, mainly the C60 molecule which, in a certain context, behaves like a wave.

It is possible to indicate then, that the description of certain chemical substance, that is to say its limits, are therefore the description of the contexts in which that certain chemical substance appears.

References:Arndt M., Nairz O., Vos-Andreae J., Keller C., van der Zouw G. and Zeilinger A. (1999). Wave–particle

duality of C60 molecules, Nature 401, 680–682.Chamizo J.A. (2017). The fifth chemical revolution: 1973–1999, Foundations of Chemistry 19, 157-179. Schummer, J. (2002). The Impact of Instrumentation on Chemical Species Identity, From Chemical Substances

to Molecular Species. In: Morris P.J.T. (ed) From Classical to Modern Chemistry. The Instrumental Revolution, London: Royal Society of Chemistry-Science Museum.

van Brakel J. (2012). Substances: The Ontology of Chemistry in Woody A.I., Hendry, R.F. and Needham P. (eds.), Philosophy of Chemistry, Oxford: Elsevier.

23

ISPC 2018

10. From Telluric Helix to Telluric Remix

Philip J Stewart, University of Oxford.

[email protected];

The first attempt to represent the Periodic system graphically was the Telluric Helix (Vis Tellurique) of Alexandre-Emile de Chancourtois, 1862, in which the sequence of elements was wound round a cylinder1. This has hardly been attempted since, because the periods vary in length from 2 to 32 elements, but Charles Janet presented a model wound round four nested cylinders, one for each block2. Janet’s periods are defined by a constant sum of the first two quantum numbers, n and l, so that they end with the s block, at the head of which he placed H and He. By combining Janet’s periods, Edward Mazurs’ table in which each row represents an electron shell3 and Valery Tsimmerman’s use of a half square for each element4, I have produced a representation that can be printed out and wound round to make a cylinder with manageable dimensions. As a table it has the originality of placing the s block in the middle, to emphasise its pivotal nature, since it both ends each Janet period and lends its electrons to the valence shell of the f, d and p block elements in the subsequent period. This makes it clear that it belongs neither (or both) on the left or the right of the table. The downward arrows that link subshells within each period graphically illustrate the Janet [or Madelung] Rule. To acknowledge those whose ideas I incorporate, I have called my design ‘Telluric Remix’.

1 Described in van Spronsen, J, The Periodic System of Chemical Elements, Elsevier, 1969, pp.97-1022 Janet, C, ‘Concordance de l’arrangement quantique de base des électrons planétaires des atomes avec la classification scalariforme. hélicoïdale des elements chimiques. Beauvais Imprimerie Départementalede l’Oise, Beauvais (1930), passim.3 Mazurs, E, Graphic Representations of the Periodic System During one hundred Years, University of Alabama Press, 2nd editition, 1974, figure 136, p. 134.4 Tsimmerman, V, https://www.meta-synthesis.com/webbook/35_pt/pt_database.php?PT_id=32 accessed 22/5/2018.

24

ISPC 2018

11. 32-Column Periodic table and Left-step periodic table united.

Mikhail V. Kurushkin, International Engineering Educator (IGIP), Associate Professor, Laboratory of Solution Chemistry of Advanced Materials and Technologies, ITMO

University, 9 Lomonosova Str., Saint Petersburg, Russian Federation, [email protected];

The pursuit of optimal representation of the Periodic Table has been a central topic of interest for chemists, physicists, philosophers and historians of science for decades (Leigh, 2009; Scerri, 2009). Should the Periodic Table of Chemical Elements first and foremost serve the needs of chemists as implied by its name? Or should it start from considerations of quantum mechanics and thus be more appealing to physicists (Scerri, 2010, 2012b)? is there a representation which overcomes this problem?

The Periodic Table is from a fundamental point of view a graphic representation of periodicity as a phenomenon of nature. While the 32-column Periodic Table, popularized by Glenn T. Seaborg, is considered by chemists the most scientifically correct representation (Scerri, 2012a), physicists apparently prefer the Left-step Periodic Table above all (Scerri, 2005; Stewart, 2010). Alternatively, it is suggested that a rigorously fundamental representation of periodicity could only take the form of a spiral as, evidently, the abrupt periods of 2-D Periodic Tables contradict the gradual increase of atomic number, and the spiral representation reconciles this debate (Imyanitov, 2016).

An optimal representation is eagerly sought after both for the needs of philosophy of chemistry and chemical education as their never-ending dialogue secures a thorough methodology of chemistry. The aim of the present work is to show that the 32-column Periodic Table and the Left-step Periodic Table can co-exist in mutual tolerance in a form of what Philip Stewart has already called Kurushkin’s Periodic Table (Kurushkin, 2017), Figure 1:

Figure 1 Kurushkin’s Periodic Table (2017)

Addition of another s-block to the left of the Left-step Periodic Table is conditional as, obviously, the number of s-elements is not doubled, but the Periodic Table in Figure 1 is to be rolled into a spiral so that the left and right s-blocks are merged together and the number of elements is exactly 118. Hence, the suggested Periodic Table is a 2-D spiral Periodic Table.

If the chemical relationships among the elements are to be taken into the account, the s-block should appear on the far left side of the Left-step Periodic Table, thus uniting the two outstanding versions into one.

25

ISPC 2018

12. ReferencesImyanitov, N. S. (2016). Spiral as the fundamental graphic representation of the Periodic Law. Blocks of

elements as the autonomic parts of the Periodic System. Foundations of Chemistry, 18(2). https://doi.org/10.1007/s10698-015-9246-8

Kurushkin, M. (2017). Building the Periodic Table Based on the Atomic Structure. Journal of Chemical Education, 94(7), 976–979. https://doi.org/10.1021/acs.jchemed.7b00242

Leigh, G. J. (2009). Periodic Tables and IUPAC. Chemistry International, 31(1), 4–6.Scerri, E. R. (2005). Presenting the left-step Periodic Table. Education in Chemistry, 42(5), 135–136.Scerri, E. R. (2009). The dual sense of the term “element,” attempts to derive the madelung rule, and the optimal

form of the periodic table, if any. International Journal of Quantum Chemistry, 109(5), 959–971. https://doi.org/10.1002/qua.21914

Scerri, E. R. (2010). Explaining the periodic table, and the role of chemical triads. Foundations of Chemistry, 12(1), 69–83. https://doi.org/10.1007/s10698-010-9082-9

Scerri, E. R. (2012a). Mendeleev’s Periodic Table Is Finally Completed and What To Do about Group 3? Chemistry International, 34(4), 28–31.

Scerri, E. R. (2012b). What is an element? What is the periodic table? And what does quantum mechanics contribute to the question? Foundations of Chemistry, 14(1). https://doi.org/10.1007/s10698-011-9124-y

Stewart, P. J. (2010). Charles Janet: Unrecognized genius of the periodic system. Foundations of Chemistry, 12(1). https://doi.org/10.1007/s10698-008-9062-5

26

ISPC 2018

Tuesday 17 July, morning. Chemistry and philosophical issues.

12. Hinge Epistemology and Chemistry.

Rom Harré1 and Jean-Pierre Llored2.1Linacre College, Oxford,

2Linacre College, Oxford and Laboratoire Sphère, Université Paris [email protected] ;

In several recent publications in the philosophy of chemistry, we have made use of a novel conceptual repertoire for assessing the intelligibility of a discourse and the trustworthiness of associated practices. Our proposal partly draws on a development of Wittgenstein’s ‘hinge’ concept, as interpreted and developed in recent studies of his later writings, in particular On Certainty. In trying to understand a scientific discourse, from a philosophical point of view, that is with respect to the concepts in use, we look for propositions which have gone unformulated and so unchallenged and once formulated seem germane to the assessment of the intelligibility of habitual procedures and practices in a certain field, that is, in the present case, that of chemistry.

We try out hinge-pairs that might shape our importation of content into the bare bones of a formal presentation of an explanation, a practical manual describing skills and relevant know-how, or into a relevant cluster of discipline defining propositions as well. Hinges are implicit in our activities, but they do not have an independent existence. We know them only as expressed in a proposition-procedure pair, doppelgängers of one another as Moyal-Sharrock has it. For instance, once made explicit, propositions like ‘Life on earth has existed for millions of years’ can be examined and sometimes tested as putative matters of fact. Hinges exist only as expressed in taken for granted pairs; hinge propositions and their doppelgängers, hinge procedures or practices. The relevant propositions are unformulated and so not examined empirically, and the paired procedures are habitual and more or less skillful, usually not guided by an actor paying attention to explicit rules.

We take the role of philosophy of science, not as critical commentary on ways of proposing ontological presuppositions or highest level theoretical premises, nor as formulating laws of nature and their relations to particular instances of phenomena, but as a digging out of the ‘hinges’, that are the tacit elements of a discipline. In this lecture, we will develop our line of reasoning further. To do so, our starting point will be the result of our scrutiny of different chemical practices, that is, quality control in analytical chemistry, green chemical engineering, quantum chemistry, nanoprecipitations of polymers, combinatorial chemistry in drug discovery, and formulation and product development in cosmetics. The aim of our philosophical analysis will be to reveal at least some of the proposition-practice doppelgängers. According to this prescription neither propositions nor practices can be treated independently. In turn, this procedure will allow for the possibility for hitherto unexamined propositions to be assessed empirically, and hitherto taken-for-granted practices to be tested for efficacy.

27

ISPC 2018

13. Bibliography:A. Coliva, Extended Rationality: A Hinge Epistemology (Palgrave Macmillan, 2015).A. Coliva, Moore and Wittgenstein: Scepticism, Certainty and Common Sense (Palgrave Macmillan, 2010).J. Bouveresse, La Force de la règle : Wittgenstein et l'invention de la nécessité (Paris: Editions de minuit, 1987).J. Bouveresse, Wittgenstein : la rime et la raison. Science, éthique et esthétique (Paris: Editions de minuit,

1973).C. Chauviré, Le grand miroir  : essais sur Peirce et sur Wittgenstein (Presses Universitaires de Franche-Conté,

2004).D. Moyal-Sharrock and A. Coliva (eds.), Hinge Epistemology (London: Brill, 2016).D. Moyal-Sharrock, (ed.) The Third Wittgenstein : The Post-Investigations Works (London: Routledge, 2004).D. Moyal-Sharrock, Understanding Wittgenstein's On Certainty (Palgrave Macmillan, 2004).R. Harré, ‘Affordances and hinges: New tools in the Philosophy of Chemistry’, in The Philosophy of Chemistry:

Practices, Methodologies, and Concepts, J.-P Llored (Ed.) (Newcastle upon Tyne: Cambridge Scholars Publishing, 2013, 580-596).

R. Harré & J.-P. Llored, ‘Procedures, Products, and Pictures’, Philosophy, The Royal Institute of Philosophy, 93, 2018, 167-186.

R. Harré & J.-P. Llored, ‘Mereologies and Molecules’, Foundations of Chemistry 15 (2), 2013, 127–144.R. Harré & J.-P. Llored, ‘Mereologies as the Grammars of Chemical Discourses’, Foundations of Chemistry 13

(1), 2011, 63–76.D. Pritchard, Epistemic Angst – Radical Skepticism and the Groundlessness of Our Believing (Princeton

University Press, 2016).L. Wittgenstein, On Certainty (Oxford: Blackwell, 1979).

28

ISPC 2018

13. Philosophical Mediation in Japanese Chemistry Translation and Appropriation.

Yona Siderer.Edelstein Center for the History and Philosophy of Science, Technology and Medicine,

The Hebrew University of Jerusalem, [email protected];

Japan had its traditional chemical processes like food and drinks fermentation, metallurgical swords production, textile dyeing and ceramics manufacture for many generations. Those were based on trial and error work. Coming to learn the fundamentals of Western chemistry thought, Japanese scholars in late 18th century and in the 19th century had to understand and then choose the specific Chinese/Japanese characters, kanji, to translate and invent new chemical terms in Japanese. In this study a) an example of the thought to compose suitable Japanese chemical terms, and b) a Japanese scholar's meditation comparing scientific thought in the East and West are briefly presented.

An interesting example (a) is the choice and change of the word "chemistry" itself. European chemistry books that arrived to Japan were mainly in their Dutch translation due to the ruling restrictions on foreign books. The medical doctor, Dutch Studies scholar and translator Udagawa Youan (1789-1846) invented the Japanese term seimi as a phonetic sound imitating the Dutch word Chemie for "chemistry"; seimi is written 舎密 . The meaning of the character 密 is secrecy. From many Chinese characters that sound "mi" Youan chose the one pointing at the secrecy of the chemical process. For chemistry, the Chinese wrote 化学 .  化 ka meaning“change, take the form of”;学gaku meaning "study", and the two characters put together mean "study of change". Since around 1860 the Japanese adopted this term, pronouncing it kagaku, different from the Chinese reading hua-hsὒeh.1

So moving from Youan's writing in the 1830s, the phonetic term seimi, the Japanese scholars in the 1860 changed the term for chemistry to kagaku that carries the meaning of this discipline, "study of change", change of materials, adopting the Chinese term with its Japanese reading. Choosing the right Chinese/Japanese character to express the object in the best way takes some philosophical thought and deep understanding of the entity under consideration and a good knowledge of the Chinese characters. During the years, Japanese experts chose and then changed some terms, while Chinese used others, sometimes borrowing from each other.2-4

Example (b) suggests the following quotation written by the 20th century Japanese author Junichirō Tanizaki (1886-1965) in his essay In praise of Shadows (1933):

"…[But] it is on occasion like this that I always think how different everything would be if we in the Orient had developed our own science. Suppose for instance that we had developed our own physics and chemistry: would not the techniques and industry base on them have taken a different form, would not our myriads of everyday gadgets, our medicines, the products of our industrial arts – would they not have suited our national temper better than they do? In fact our conceptions of physics itself, and even of chemistry, would probably differ from that of Westerners; and the facts we are now taught concerning the nature and function of light, electricity, and atoms might well have presented themselves in different form."5

Do we accept Tanizaki's assumptions? To answer this question the following points should be considered: the fundamental approach of the Japanese scholars in the Tokugawa period could be taken as the intellectual search for knowledge while the ruling regime aim was to use Western chemistry for practical purposes. That is, merging the Western knowledge with Japan's knowhow in order to achieve economic, military and national status among the developed nations, yet keeping traditional

29

ISPC 2018

Japanese spirit. It can be said that the Japanese thought and writing on chemistry combined theoretical concepts with practical application. As Togo Tsukahara explained on the relation between theory and practice of chemistry: "…Philosophical discussions and practical demand were interrelated; they were interwoven into a new pattern of theory and practice, slightly different from that of the West… Scientific theory and technical practice were merged in Rangaku (Dutch Studies). This tradition was a remarkable feature of science in Japan."4

More examples and further discussion will be presented.

References 1 Sugawara Kunika article Seimi kara kagaku e 舎密から化学へ "From Seimi to Chemistry" Chapter

4 in Kindai Nihon, Sono Kagaku to Gijutsu: Genten eno Shotai, Kogaku Shuppan "Science and Technology in Modern Japan: An Introduction to the Original Texts", Sadaaki Shito and Ichiro Yabe, editors.1990. pp. 93-125.

2 James Reardon-Anderson. The study of Change: Chemistry in China, 1840-1949. Cambridge University Press 1991 pp. 85-86 Tables 4.2, 4.3

3 Scott L. Montgomery. Science in Translation Movements of Knowledge Through Culture and Time. The University of Chicago Press, Chicago and London 2007

4 Togo Tsukahara. Affinity and Shinwa Ryoku, Introduction of Western Chemical Concepts in Early Nineteenth–Century Japan, J.C. Gieben, Publishers Amsterdam. Ph.D. dissertation 1993

5 Junichirō Tanizaki: In Praise of Shadows (1933). Original title: 陰翳 礼讃 (In'ei Raisan). Translated by Thomas J. Harper and Edward G. Seidensticker. Vintage Books London 2001. pp. 13-14.

30

ISPC 2018

14. Tracking Taxol: Aesthetics and The Natural in Synthetic Organic Chemistry

Tami I. Spector, University of San Francisco. [email protected];

The oxymoronic subdiscipline, Natural Product Synthesis, captures the essence of the entangled philosophical conundrum of the natural and the synthetic in chemistry.1 In this paper I employ Taxol (paclitaxel) as a paradigmatic example to investigate this conundrum. I choose Taxol because it is one of the most important chemotherapeutics and revered illustrations of artful ingenuity in the total synthesis of a complex natural product. Unlike most “molecules of living nature” that synthetic organic chemists have replicated in the laboratory, Taxol’s historical path, from its discovery in and isolation from the bark of the Pacific yew tree to its recapitulation as a significant chemical and medicinal product with a precise molecular representation, has been well documented.2 In addition, Taxol’s history is consistently placed within a parallel narrative of natural scarcity, thus providing us with an overt and detailed scaffold for examining notions of the natural and synthetic in chemistry, including how Taxol, or any natural product, once loosed from its origins in Nature, is understood as Nature/not Nature.

For this paper I borrow from the arts for an understanding of the perceived cleft that divides the natural and the synthetic. My method is strategic: aesthetic queries elicited by the visual arts have already probed the interplay, at times conflicted, between the natural and the manmade. In landscape painting, for example, theorists have called attention to the fact that this division is highlighted and reinforced by the constructed nature of the landscape within such work – though it is not always clear how conscious most landscape artists are of the cultural and philosophical implications of their work in relation to nature. In the latter half of the 20th century, however, some artists began to overtly critique “the ways we frame nature through representation as well as through science and technology . . . forging an aesthetic awareness of how nature exceeds these discourses and representations.”3 My aim in this paper, therefore, is to highlight contemporary artists who have purposely mined these aethetics.4 These include, Bob Verschueren, whose “vegetal art” brings nature as found object into the built environment; Robert Smithson, who used – and some might say abused – nature to create works of “earth art”; and Arte Povera artist Piero Gilardi, whose polyurethane Nature Carpets precisely replicate natural environments.

I employ Verschueren’s work to probe how nature and place intersect as we transport source materials, like Pacific yew bark, from their ecological habitats into the lab; Smithson’s dialectic of site (the geographic location of his earth art) and non-site (the textual and visual representations of his work presented in a gallery) to examine how the molecular representations of Taxol reference, but are contradictory to, their originating site – the pristine forests of the Pacific Northwest, while simultaneously pointing to the environmental degradation of these landscapes;5 and Gilardi’s polymeric carpets, which, like synthetic Taxol, purport to amalgamate technology and nature for the greater good, yielding “chemically manufactured substitutes . . . that contribute to the conservation and protection of nature.”6

In sum, by applying the aesthetic logic of art works such as these to the historically and scientifically rich narrative of Taxol, I will track how our perception of its natural status alters from its origins in the bark of the Pacific yew tree; its isolation, characterization and representation in the laboratory; and its reconstitution as a synthetic product. In doing so, I explore how the chemical extraction and replication of Taxol inform our concept of it apart from its origin.

31

ISPC 2018

11. References. 1. For discussion of the natural and synthetic in chemistry see, for example: Joachim Schummer, “The Notion of Nature in Chemistry,” Stud. Hist. Phil. Sci. 34 (2003): 705-736 and Roald Hoffmann, The Same and Not the Same (New York: Columbia University Press, 1995): p. 111-115. 2. K. C. Nicolaou, “Organic Synthesis: The Art and Science of Replicating the Molecules of Living Nature and Creating Others Like Them in the Laboratory,” Proc. R. Soc. A 470 (2014): 20130690; Jordan Goodman and Vivian Walsh, The Story of Taxol: Nature and Politics in the Pursuit of an Anti-Cancer Drug (Cambridge: Cambridge University Press, 2001).3. Amanda Boetzkes, The Ethics of Earth Art, (Minneapolis: University of Minnesota Press): p. 2.4. See for example: John K. Grande, Art Nature Dialogues: Interviews with Environmental Artists, (Albany: SUNY Press, 2004) and Ben Tufnell, Land Art (London: Tate Publishing, 2006).5. A. Boetzkes, The Ethics of Earth Art, (Minneapolis: University of Minnesota Press): p. 68-76.6. Bernadette Bensaude-Vincent “Reconfiguring Nature Through Syntheses: From Plastics to Biomimetics,” in The Artificial and the Natural: An Evolving Polarity, ed. Bernadette Bensaude-Vincent and William R. Newman (Cambridge: MIT Press, 2007): 293.

32

ISPC 2018

15. Radicals – Experiments, Existence, Electrons.

Juan Camilo Martinez Gonzalez, CONICET-Universidad de Buenos Aires,& Klaus Ruthenberg, Coburg University of Applied Sciences and Arts.

[email protected]; [email protected];

Most chemists think of the development of physical organic chemistry in terms of the "electronic theory" and the elaboration of the ideas of the importance of electron supply and withdrawal in the interaction of electrophilic and nucleophilic reagents. However, our knowledge of the free radical concept has had a fascinating synthetic history (Walden 1924), the importance of which for the philosophy of chemistry is one main subject of the present contribution.

From very early on the question whether or not radicals could be isolated (or synthesized) to yield manifest stuff portions was fascinating the chemists. However, the first successful “preparation” of such a “free radical” was reported by Moses Gomberg (1866 – 1947) only in 1900. He came up with results which eventually – that is after many years of critical discussion – convinced the scientific community that certain free radicals can be assigned substantial existence (Gomberg 1914). About 30 years later Friedrich Paneth (1887 – 1958) and his co-workers published empirical evidence for the existence of the quite unstable methyl- and ethyl radicals in sophisticated, ingenious experiments (s. Paneth and Hofeditz 1929).

Although their definition has been shifted in modern chemistry, the contents of these early concepts of radicals are still vivid in chemistry, for example referring to the use of molecular fragments like the “methyl group” or the “hydroxyl group” in the talk about reactions and structures. Since the 1930s, radicals are widely defined as chemical species with at least one unpaired electron in the outer atomic shell (Pauling 1933), beginning perhaps with Lewis´ “odd molecules” (Lewis 1923). Hence, for a rough overview the history of chemical radicals might be put into three parts: the speculative, the synthetic, and the electronic period (Ruthenberg 2015). The present contribution will mainly refer to the first developments of the electronic period and its relations to the “making” of radicals.

The historical reconstruction of this phase will be subsidiary firstly to some questions relevant to the philosophy of chemistry related to the status of chemical entities and their theoretical description. Also it will be useful to elucidate different views about the uses and methodological commitments of quantum theory at work in solving chemical problems regarding the stability, structure and reactivity of radicals during the first days of the “in between” discipline called quantum chemistry.

References : Gavroglu, K. y Simões, A. (2012). Neither Physics nor Chemistry: a History of Quantum Chemistry. Cambridge

MA: MIT Press.Gomberg, M. (1914) The existence of free radicals, Journal of the American Chemical Society 36, 1144-1170.Lewis, G. (1923) Valence and the Structure of Atoms and Molecules. New York: The Chemical Catalog

Company Inc.Paneth, F., Hofeditz, W. (1929) Über die Darstellung von freiem Methyl, Berichte der Deutschen Chemischen

Gesellschaft 62, 1335-1347.Pauling, L. (1933) “The Nature of the Chemical Bond. V. The Quantum-Mechanical Calculation of the

Resonance Energy of Benzene and Naphthalene and the Hydrocarbon Free Radicals”, Journal of chemical physics. (1): 362-374.

Ruthenberg, K. (2015) Radicals, Reactions, Realism, in: E. Scerri an L. McIntyre, eds., Philosophy of Chemistry (Boston Studies in the Philosophy an History of Science 306), Dordrecht: Springer, 183-199.

33

ISPC 2018

Walden, P. (1924) Chemie der freien Radikale. Leipzig: Verlag von S.Hirzel.

16. Robert Boyle and Baruch Spinoza: The redintegration of saltpetre.A reply to Antonio Clericuzio

Filip Buyse, Descartes Centre – Utrecht [email protected];

In 1661, Henry Oldenburg (1618-1677) sent the Latin translation of “Certain Physiological Essays”, written “by an English nobleman”, to Spinoza with the request that he read and comment especially on the experiments explained in the book. In his Tentamina quædam physiologica diversis temporibus & occasionibus conscripta, Robert Boyle (1627-1691) had defined for the first time his new philosophy, the Corpuscular Philosophy (his preferred synonym for Mechanical Philosophy). In this work, the co-founder of the Royal Society attempted to convince his readers of the great value of his new philosophy, as an alternative to the Peripatetic philosophy of qualities of bodies and the views of those “Chymists” influenced by Paracelsus (1494-1541)

In the same period, Spinoza (1632-1677) was busy developing his own new philosophy. However, the philosopher of Rijnsburg was not impressed by Boyle’s work. He found it unoriginal and appears to have rejected it. Instead of discussing the new philosophy with Boyle, Spinoza engaged in a discussion about the nature of saltpeter. Consequently, both seem to hold their own parallel discussions. The philosopher Spinoza starts a scientific discussion about saltpeter, while “The father of chemistry” wants to discuss his new philosophy. Nevertheless, there are passages were Boyle and Spinoza come to real confrontation.

Firstly, this paper discusses the historical context of the correspondence. Secondly, it analyses in depth the so-called ‘redintegration of nitre experiment’ [experimento de redintegratione nitri] and gives a modern interpretation of this process which is essential in Boyle’s demonstration. Thirdly, it opposes Boyle’s and Spinoza’s views and compares the different interpretations by C.A. Crommelin, (1939), H. Daudi (1949), E. Yakira (1988), A.M. & M.B. Boas (1964), P. Macherey (1995), A. Gabbey (1996), S. Nadler (1999), S. Duffy (2006) and Marina Paola Banchetti-Robino (2012). However, the main aim of the paper is to criticize Antonio’s Clericuzio’s interpretation (2000 & 1990) and to develop an alternative view. At least four questions will be addressed.

This paper argues, firstly, that Clericuzio’s categorization of Spinoza’s philosophy as strictly mechanical and reductionist is misleading. The author of “Elements, principles and corpuscles” (2000) put Boyle’s interpretation in a broader context referring to works that Boyle (whose ideas were evolving) would write much later. The comments on Spinoza’s interpretation, by contrast, are exclusively based on what the Dutch philosopher writes in his letters belonging to the Boyle/Spinoza correspondence. This paper put also Spinoza’s words in a larger context and shows that is problematic to say that Spinoza is a radical mechanical philosopher. (Buyse, 2013)

Secondly, in his discussion of the Boyle/Spinoza controversy, A. Clericuzio does not say a word of R.S. Glauber. Amazingly, the name of the German chemist and alchemist is simply missing in his comments on the Boyle/Spinoza controversy. This paper argues that Glauber’s work should for several reasons be included in the discussion of the controversy. First of all, because it was Glauber who introduced the neologism ‘redintegratio” into Latin. Second, because Boyle mentions Glauber two times in his work in order to avoid being accused of plagiarism. Third, because it was Glauber who did the redintegration experiment for the first time. (Newman & Prinzipe, 2002) Fourth, because there is historical evidence that - as a hartlibian - Boyle knew about Glauber’s work via Benjamin Worsley (Newman & Prinzipe, 2002). And fifth, because it is likely that also Spinoza knew about

34

ISPC 2018

Glauber’s work on saltpeter which took place in his impressive laboratory in Amsterdam which was located near Spinoza’s birthplace. (Buyse, 2013)

Thirdly, Clericuzio argues that with the redintegration experiment, Boyle wanted to give “chemical theories an independent status”. This paper however defends the idea that Boyle used this experiment to promote the Mechanical Philosophy. As Michael Hunter put it: Boyle saw his explanation of the redintegration experiment “as demonstrating that a chemical substance is composed of distinct parts, which can be taken apart and put back together mechanically” (Hunter, 2009). Indeed, Boyle insisted - in each of his letters belonging to the Boyle/Spinoza - that Spinoza should have in mind the true aim of the book which was at the bases of another text and the publication of “The Origin of Forms and Qualities (According to the Corpuscular Philosophy,) Illustrated by Considerations and Experiments” (1665/66). Moreover, Boyle insisted that the Dutch philosopher read the preface of his book wherein this aim was revealed. He made clear Spinoza clear that he was missing the point. Even in book published much later, A Free Enquiry into the Vulgarly Received Notion of Nature, which was indeed mainly written in the mid – 1660s ( E.B. Davis & M. Hunter, 1996) but later adapted and only published in 1686, Boyle still refers to “certain paper, occasioned by ‘ A Chemico-Physical Essay about Salt-petre. ’, against the pretended origin and inexplicable nature of imaginary forms of the Peripatetics.”

Finally, the author of “A redefinition of Boyle’s Chemistry and Corpuscular Philosophy” (1990) also argues that the contrast between Boyle and Spinoza was not “that of a rationalist versus the experimental philosopher, since in fact Spinoza never denied the role of experiments”. The contrast, he continues, was on “the role of mechanical explanations in chemistry”, contesting Hall’s interpretation. However, this paper shows that, besides an ontological dispute, there was also a difference in epistemological stance.

References: Buyse, F., Spinoza, Boyle, Galileo: Was Spinoza a Strict Mechanical Philosopher? Intellectual History Review,

23, 1, 2013. Buyse, F., Boyle, Spinoza and the Hartlib Circle: The Correspondence which never took place. Society and

Politics, 7(2), 2013, 34-53. Clericuzio, A., Elements, Principles and Corpuscles: A Study of Atomism and Chemistry in the Seventeenth

Century. Dordrecht: Kluwer, 2000. Clericuzio, A., A redefinition of Boyle’s Chemistry and Corpuscular Philosophy. Annals of Science, 47, 1990,

561-589.Hunter, M., Boyle: Between God and Science. New Haven and London: Yale University Press, 2009.Newman, W.N. and Principe, L.M., Alchemy tried in the fire: Starkey, Boyle, and the fate of helmontian

chymistry. Chicago & London: The University of Chicago Press, 2002.

35

ISPC 2018

17. The chemical element and the multilevel experience of Chemistry.

Roberto Barbosa de Castilho, Department of Chemistry, Institute of Exact Sciences,Federal Rural University of Rio de Janeiro (UFRRJ), Rio de Janeiro,

[email protected];

The history of the concept of chemical element is neither linear nor simple, as it reveals some interesting aspects of the development of chemistry. In physics, the mathematization of variables paved the way to the scientific revolution, since modern science did not depend on the generalization of the empirical observations, but rather on an analysis capable of abstraction [Rossi, 2001]. Modern chemistry was a product of the scientific revolution, but it took a longer time to be consolidated as a distinct science. Chemical knowledge transcends the immediate perception and has always demanded a metaphysical effort to go beyond appearances. Chemistry does not depend solely on the primary qualities such as Physics. There is a gap between the perceptual experience of events and their conceptual representation, which is wider than for any other of the basic sciences [Siegfreid, 2002]. Therefore, learning chemistry is a multilevel experience in which the physical, metaphysical and symbolic are the aspects of nature needed to understand the entire chemical phenomenon. J.van Brakel argues there is a conflict between the manifest aspect of reality (the macroscopic) and its scientific image (the microscopic) and mankind does not have a unique criterion to decide which one is more important or fundamental to understand nature. [Brakel, 2013]. I intend to show in this paper that the dual nature of the chemical element is a fundamental part of understanding chemistry, and discuss the multilevel experience of learning and teaching.

In the history of philosophy, there were propositions to explain the nature of reality and cosmologies. Some were based on principles of what could be seen and experienced by the senses, while others postulated that reality might be related to what could be realized by thought. Interestingly, Heraclitus proposed that everything is changing while Parmenides stated that nothing can really change but that being always remains the same. Considering chemistry, does everything change or is there something that remains the same during the changes? Empirical knowledge had an important role in the genesis of the concept of chemical element, as in the empirical works of Boyle and Lavoisier – the chemical element was seen as a material concept associated with practice. In the eighteenth century, there was a materialization of the qualities, culminating in the individualization of matter as inert corpuscles, which retain macroscopic properties. Consequently, the chemical elements might be considered unobservable basic substances characterized by one observable property, namely the atomic weight.

The stoichiometry laws were based on the conservation of mass – a macroscopically measurable property – and also on the conservation of the chemical element – an abstract concept in the sense that it is not observed macroscopically. Therefore, the chemical elements were arranged in the periodic table with respect to their atomic weight. D. Mendeleev insisted that the periodic table display the abstract elements, hence the chemical revolution did not eliminate the metaphysical view of matter. F. Paneth differentiated the simple substances from the basic substances: the former are characterized by atomic weight while the latter are characterized by atomic number – an abstraction. The atomic number, unlike mass, is not directly observable, but it can be measured by spectrometers. Indeed, both the physical and metaphysical views may help the understanding of chemistry [Scerri, 2005, 2008; Paneth, 1963].

The chemistry is an interesting and complex science because it requires the integration of three levels of representation: macroscopic, submicroscopic or particulate and symbolic [Johnstone, 1982]. Macroscopically, there is no immediate sensory way to differentiate element from compound. One can perform a chemical analysis to decompose substances and to verify that the elementary bits

36

ISPC 2018

are the chemical elements [Johnstone, 1991]. There are sciences like classical thermodynamics that were formulated only using two levels of representation: the macro and the symbolic. Nevertheless, learning chemistry requires the integration of the three levels, although educators have shown that teaching all the three levels simultaneously might not be a good strategy (logically and psychologically), because students might be tempted to create misconceptions to simplify things [Johnstone, 2000]. Beyond the macro and the symbolic levels, there is the submicroscopic level, which cannot be directly experienced, but is essential to modeling and explaining the chemical phenomena. Indeed, an abstract construction (beyond observation) can help to promote a multilevel experience of learning chemistry. Considering the concept of the chemical element, a teacher has to emphasize the three forms of perceiving reality and be aware of the manifest or occult conflict between the observable and the metaphysical or abstract realities. Chemists and other scientists over the centuries have proposed theories based on empirical or metaphysical issues and I propose to associate the delay of the consolidation of the chemical element as a distinct concept to the difficulties of students in learning chemistry, as addressed by the three levels of representation – Johnstone’s chemical triangle.

In the realm of chemistry, the macroscopic experience and the metaphysical (or abstract) constructions might be relevant to make progress in scientific investigation and learning and teaching. One of the emblematic examples is the concept of chemical element. The metaphysical view states that the chemical elements are the basic substances or bearers of properties. On the other hand, the empirical definition of the chemical elements points to the experimental limit of chemical analysis. Moreover, at the microscopic level, the chemical elements, the atoms, are the basic substances which have properties quite different from bulk substances. Models of teaching chemistry created to understand the behavior of atoms in chemical reactions might help the comprehension of what is happening deeper inside matter. Indeed, there is no fundamental level of experiencing chemistry, as this involves construction and evolution of the concept of chemical element. In conclusion, there is nothing wrong with the dual nature of the concept of chemical element, physical or macroscopic and metaphysical or abstract, since chemistry is a multilevel experience.

References:Brakel, J.V. Philosophy of Chemistry: Between the Manifest and the Scientific Image. Leuven University

Press, 2013.Johnstone, A.H. Macro and micro chemistry. School Sc. Rev. 64, 227, 377-79, 1982.Johnstone, A.H. Why science is difficult to learn? Things are seldom what they seem. J.Comp.Assist. Learning.

7, 75-83, 1991.Johnstone, A.H. Teaching of chemistry: logical or psychological? Chem.Educ. Res.Pract.Eur. 1, 9-15, 2000.Paneth, F.A. ‘The epistemological status of the concept of element’. Foundations of Chemistry 5, 113-145,

2003.Rossi, P. La nascita della scienza moderna in Europa. Ed. Laterza, 2000.Scerri. E. Some aspects of the metaphysics of chemistry and the nature of the elements. HYLE - International

Journal for Philosophy of Chemistry 11, 2, 127-145, 2005.Scerri. E. Realism, Reduction, and the “Intermediate Position”. In “Of minds and molecules”, Bhushan,N.,

Rosenfeld,S. (eds.), Oxford University Press, New York, 2000, Scerri, E. R. Collected Papers on Philosophy of Chemistry, 121-142, Imperial College Press, 2008.

Siegfried, R. From elements to atoms: a history of chemical composition. Trans. Am. Philos. Soc. New Series, Vol. 92, No 4, Philadelphia, 2002.

37

ISPC 2018

38

ISPC 2018

Tuesday 17 July 2018, afternoon. Chemical elements.

18. The dual conception of chemical element: epistemic aspects and implications for chemistry teaching.

Elena Ghibaudi, Dip. Chimica, Università di Torino, Via Giuria 7, I-10125 Torino, [email protected];

The dual conception of chemical element, sanctioned by the IUPAC definition [1], has roots in the Greek philosophy and in the empirical perspective proposed by Lavoisier [2]. The epistemic implications of such dualism are the object of a wide discussion within the community of chemists [3-8]. Many authors maintain that Mendeleev’s Periodic Table is based on an idea of chemical element that is closer to the philosophical concept of substantia than to the final term of chemical analysis [9]. A fundamental contribution on this issue comes from F.Paneth, who authored a seminal work on the epistemological status of the chemical concept of element [10]. Paneth’s position stems from its being a radiochemist, as the discovery of isotopes represented a further challenge to the formulation of a univocal definition of chemical element.The epistemic problem of defining what persists unchanged in a chemical transformation is especially relevant in chemical education, where the concept of element allows highlighting the factor of continuity in a system that undergoes a chemical change. From this standpoint, the identification of the element with the simple substance, i.e. the final term of chemical analysis, leads to logical contradictions and inconsistencies. The identification of the element with a kind of atoms or atomic nucleus is troublesome for similar reasons.I will propose a critical analysis of various definitions of chemical element put forth by epistemologists and chemical education scholars since Paneth on [11-15]. I will discuss their implications for chemistry teaching and argue on the importance of attaining an unambiguous definition of element, capable of highlighting the main purpose of this concept, that is to designate what remains unchanged in a chemical transformation.

References:IUPAC. Compendium of Chemical Terminology, 2nd ed., the “Gold Book”, McNaught, A., Wilkinson, A.,

compilers; Blackwell Scientific Publications: Oxford, U.K., 1997Lavoisier, Traité Eleméntaire de Chimie, I, Discours Preliminaire; Cuchet: Paris, 1789B. Bensaude-Vincent. British J. Hist. Sci. 19: 3–17, 1986E. Scerri, Found.Chemistry 14: 69-81, 2012R. Hendry, Found.Chem. 7: 31–48, 2005J. Earley, Foundations of Chemistry 11: 65-77, 2009R.Siegfried and B.J. Dobbs. Annals of Science 24: 275–293, 1968 C.E. Perrin. Lavoisier’s Table of the Elements: A Reappraisal. Ambix 20: 95–105, 1973D. Mendeleev The Principles of Chemistry, London, 1891F. Paneth Schriften der Königsberger Gelehrten Gesellschaft, Naturwissenschaftliche Klasse, 1931 reprinted in

Found. Chem. 5: 113−145, 2003P. Nelson, Chem. Educ. Res. Pract. 7: 288-289, 2006 W. H. Roundy, J. Chem. Educ. 66: 729-730, 1989.R. Luft, Dictionnaire des Corps Purs Simples de la Chimie, 1997W. Jensen, J. Chem. Educ. 75: 817-828, 1998E.Ghibaudi, A.Regis, E.Roletto, J. Chem. Educ. 90: 1626-1631, 2013

39

ISPC 2018

19. The Changing Relation Between Atomicity and Elementarity: From Lavoisier to Dalton

Marina Paola Banchetti-Robino, Florida Atlantic [email protected];

Although both Antoine Lavoisier and John Dalton define atoms as particles that cannot be further divided in principle, they hold radically views on the relationship between atomicity and elementarity. This essay proposes to examine these distinct views in relation to the epistemic and methodological concerns underlying Lavoisier’s and Dalton’s chemical work.

The essay first examines Lavoisier’s rejection of atomicity, which he considers to be a suspect metaphysical concept. He rejects the epistemic value of positing indivisible atoms as the most simple and fundamental particles of matter and considers such positing to be mere metaphysical speculation. His emphasis on both empirical data and quantitative analysis greatly influences his position on this issue. In his view, since atoms have no empirically determinable or quantifiable properties, they contribute nothing to actual experimental work or to the chemist’s understanding of chemical elements. As an alternative to such speculation, Lavoisier adopts a ‘negative-empirical’ conceptualization of elementarity as defined by the limits of analysis. Thus, for Lavoisier, the term ‘element’ should not be applied to atoms or to fundamental particles, which he considers to be suspect metaphysical entities. Instead, by ‘element’, Lavoisier means those substances that remain as the last product of chemical analysis.

John Dalton, on the other hand, seeks to establish an empirical link between ‘elements’ and ‘atoms’, through the notion of ‘chemical atom’. The challenge for Dalton, however, is to avoid any suspect metaphysical implications and he does so by employing the experimental and quantitative criteria advanced by Lavoisier. Dalton purports to demonstrate that chemical atoms are empirical entities, that is, that they have empirical and quantifiable features that can be experimentally determined. For him, the primary determinable feature of chemical atoms is their relative weight. Dalton’s chemical atoms are not the atoms of classical or even Newtonian atomism, which all have equal shape, size, and impenetrability. Instead Dalton claims that, since elements are composed of atoms and since there are differences between elements, there must also be differences between the atoms that compose those elements. Through his chemical atomic theory, Dalton seeks to understand how the atoms of different elements combine to form compounds.

Dalton concludes that atoms of different weights combine differently, according to specific laws of proportion, to form the different elements. Thus, rather than defining elementarity in relation to the limits of analysis, Dalton defines an element as a substance composed entirely of atoms with identical properties. The difference in the properties of the elements is explained by the difference in the weights of the atoms that compose those elements. With the aid of experiment, analysis, and quantitative data regarding relative weights, Dalton concludes that chemical atoms have a demonstrable empirical status and are not, as Lavoisier had believed, suspect metaphysical entities.

References:Bensaude-Vincent, Bernadette and Isabelle Stengers, Histoire de la chimie (La Découverte, 2001).Boas Hall, Marie, “The History of the Concept of Element”, in John Dalton & the Progress of Science

(Manchester University Press, 1968).Dalton, John, A New System of Chemical Philosophy, introduction by Alexander Joseph (Manchester: George

Wilson, 1827).Dalton, John, “On the Absorption of Gases by Water and Other Liquids,” Memoirs of the Literary and

Philosophical Society of Manchester (1805), Second Series, Volume 1

40

ISPC 2018

Hendry, Robin Findlay, “Elements, Compounds, and Other Chemical Kinds”, Philosophy of Science 73 (December 2006).

Kedrov, B. M., “Dalton’s Atomic Theory and Its Philosophical Significance”, Philosophy and Phenomenological Research, Vol. 9, No. 4, pp. 644-662.

Klein, Ursula (ed.), Tools and Modes of Representation in the Laboratory Science, Series: Boston Studies in the Philosophy of Science (Springer, 2001).

Lavoisier, Antoine, Traité Élémentaire de Chimie (Cuchet Libraire, 1789).Leicester, Henry M. The Historical Background of Chemistry (New York: John Wiley & Sons, 1961).Levere, Trevor H., Transforming Matter: A History of Chemistry from Alchemy to the Buckyball (Baltimore:

The Johns Hopkins University Press, 2001).Rocke, Alan, Chemical Atomism in the Nineteenth Century: Dalton to Cannizzaro (Ohio State University Press,

1984).Scerri, Eric, “Some Aspects of the Metaphysics of Chemistry and the Nature of the Elements”, HYLE:

International Journal for Philosophy of Chemistry, Vol. 11, No. 2, pp. 127-145.Siegfried, Robert, From Elements to Atoms: A History of Chemical Composition (American Philosophical Society,

2002).

41

ISPC 2018

20. Basic (Abstract) Substance vs Simple (Real) Substance

Mark [email protected];

The idea of, and distinction between, the 118 chemical elements existing as basic (abstract) substances and simple (real) substances has importance because the dual concepts are routinely extended to all chemical substances and leads to ambiguity:

Many periodic tables include elemental data, and will often state that “oxygen has the atomic number 8, a mass of 15.999 and a boiling point of –182.962°C”. However, the mass refers to average relative atomic mass of an atom of oxygen and the boiling point refers to the diatomic molecular substance, O2, at 1.00 atm pressure.

All elements can be made to exist as gas phase atoms, X(g), the simplest possible state of the the simple (real) elements substance.

Sodium's metallic properties and ‘chlorine, the green gas’ do not exist in the ionic material, sodium chloride, NaCl. So, which ‘parts' of sodium and the chlorine are present in sodium chloride?

A doctor may prescribe “diamorphine” even though the pharmaceutical agent will likely be present as the hydrochloride salt in aqueous solution. Likewise, what is the substance “concrete"?

The paper will explore how the notions of the basic (abstract) substance and the simple (real) substance arise and how they impact science and society.

42

ISPC 2018

21. On the Diversity of Chemical Elements in Science and Humanities.

W. H. Eugen Schwarz, Theoretical Chemistry Groups of University Siegen (Germany) and Tsinghua Daxue Beijing (China)

[email protected]

The rational distinction between different analytical concepts appears recommendable for a better understanding of our objective material world. We here discuss some aspects concerning the chemical elements.

CHEMICAL ELEMENTS and ELEMENTARY SUBSTANCES: In Eastern and Western cultures, the same word ‘element’ has always been applied since antiquity to label two different macroscopic concepts, a real substance and an effective principle (Crossland 1971, Simonyi 2012, Wußing 1987). Nowadays IUPAC’s Gold-book defines ‘element’ twofold: A chemical element is 1) a species of atoms, and 2) a pure elementary substance (Gold 2014). Yet teachers tell that distinguishing the two concepts is a challenge for many students as well as for culturally educated adults, causing problems in life and politics (Ghibaudi et al. 2013).

MACROSCOPIC and MICROSCOPIC DEFINITION OF ELEMENTS: Since Boyle and Lavoisier, the chemical elements were operationally defined as the endpoints of chemical analysis of macroscopic chemical substances (Crossland 1971, Simonyi 2012, Wußing 1987). Nowadays they are theoretically defined at the microscopic level by their atomic nuclei of charge +Z∙e, being basically neutralized by surrounding electronic clouds of charge near to −e∙Z (Gold 2014, Ghibaudi et al. 2013). Nuclei carrying more or less electrons with them make up the outer shells of stars, the planets, the asteroids, the comets and the interstellar dust clouds (Herbst 2001).

INHUMAN and HUMAN ASPECTS OF THE ELEMENTS: The elements exist in the cosmos independent of human inquiry (Del Re 1998, Herbst 2001). On the background of the success of physical theories, it appears as a most reasonable assumption that the chemical elements and their relations are given by nature, independent of space and time. However, the properties and behavior of the elements, as human chemists know them, depend on the depth of their inquiry and on the ‘human’ environmental conditions, i.e. at T typically around room temperature, at pressures p around an atmosphere, for typical times t around an hour (around meaning × and : various powers of ten) (Grochala et al. 2007).

PHYSICAL NUCLEI and CHEMICAL ATOMS: Some scientists distinguish between physical nuclei, and elemental atoms as nuclei with electronic core and valence shells. The nuclear lifetime should at least be in the order of ns to form electronic valence shells, otherwise there would be no chemistry. At present, chemical specification of short-lived atoms still requires times of the order of a second. At the current end of the periodic system of chemical elements at Z=118, the nuclear lifetimes drop to the order of ms. Theoretical speculations of very long-lived super-heavy nuclei are weakening in the course of recent decades. (Chowdhury et al. 2008, Düllmann 2017, Düllmann et al. 2015, Kragh 2017, Wapstra 1991) Periodic tables up to Z around 170 or 1000 are ‘inspiring’ narratives (Karol 2002).

NATURAL PERIODIC SYSTEM and HUMAN PERIODIC TABLES: In the course of time we get the system of elements, as given by nature, better known. The best approximate representation of a chosen piece of some known aspects of the elements in the form of some type of graphical table may vary with one’s more or less general viewpoint in the course of time. (Scerri 2007, Van Spronsen 1969, Schwarz 2010, Schwarz & Rich 2010, Wang & Schwarz 2009)

SELF-CONTAINED and RELATIONAL PROPERTIES: Objects such as atoms or elements have properties ‘by themselves’ (nuclear charge, molar mass) and properties or potentials describing possible relations and interactions to other objects (common valence, electronegativity, ionic radii, bio-activities). The ‘whole’ of a chemical substance can be represented, in an extended sense, by a chemical structural formula of interconnected atoms, its ‘parts’. The types, the numbers and the

43

ISPC 2018

topology of the atoms are in principle sufficient to deduce the summative and the emergent properties of the compounds, fully exploiting the known physical theories. Also the various reactivities belong to the set of properties of parts and wholes. In this sense, the sentence ‘the whole is more than its parts’ is misleading. First, a whole belongs to a different category than its parts, and second a part is incompletely described if its interaction potentialities are suppressed, such as the Coulomb force of an electron, or the reactivities of compounds depending on their possible reaction partners. (Schummer 1998)

ELEMENTARY SUBSTANCES HAVE NO SIMPLE SET OF PROPERTIES: A macroscopic elementary substance, consisting of atoms of same charge number, e.g. 6C carbon, means any of its different thermodynamic phases and allotropes. The whole set may contain a few up to innumerable many different members (Bruchfield et al. 2017). Different portions of each one (depending on particle size and surface structure) may have different biological, chemical, catalytic, optical, or even thermodynamic properties. Specifying the properties of any sample of an elementary substance requires a very detailed specification of the reference specimen. There are no general attributes of an elementary substance. Also note that many properties of really existing chemical compounds, such as the density or the various reaction velocity constants are experimentally obtainable only within a very few digits.

CHEMICAL ELEMENTS CARRY MANY ACCURATE AND FUZZY PROPERTIES: The basis of some of the modern sciences are their specific conservation laws for their basic entities. In chemistry that are the conserved elements in reacting substances, at the macroscopic and microscopic levels. An element is uniquely defined and sharply identified by its integer nuclear charge number Z between 1 and a bit over 100. Many attributes are uniquely coupled to Z, such as the integer number of valence electrons, the quantum numbers of the valence-active orbitals, and the rather accurate decimal number of the common molar mass of the isotopic mixture as historically developed on our earth through previous star generations. Further there are various more fuzzy attributes of the elements such as electronegativity, atomic effective radii for several different interaction types, etc.

References

Burchfield, L. A., Al Fahim, M., Wittman, R. S., Delodovici, F., Manini, N: Novamene: A New Class of Carbon Allotropes. Heliyon 3, e00242 (2017).

Chowdhury, P. R., Samanta, C., Basu, D. N.: Search for Long Lived Heaviest Nuclei Beyond the Valley of Stability. Phys. Rev. C 77, 044603 (2008).

Crosland, M., Ed.: The Science of Matter – A Historical Survey. Penguin Books, Harmondsworth GB (1971). Del Re, G.: Ontological Status of Molecular Structure. HYLE 4, 81-103 (1998).Düllmann, C. E.: Studying Chemical Properties of the Heaviest Elements: One Atom at a Time. Nucl. Phys.

News 27(2), 14-20 (2017).Düllmann, C. E., Herzberg, R. D., Nazarewicz, W., Oganessian, Y., Eds.: Nucl. Phys. A 944, 1-690 (2015).Ghibaudi, E., Regis, A., Roletto, E.: What Do Chemists Mean When They Talk about Elements? J. Chem. Educ.

90, 1626-1631 (2013).Grochala, W., Hoffmann,R., Feng,J., Ashcroft, N. W.: The Chemical Imagination at Work in Very Tight Places.

Angew. Chem. Int. Ed. 46, 3620-3642 (2007).Gold, V.: International Union of Pure and Applied Chemistry, Ed.: Compendium of Chemical Terminology -

Gold Book, Version 2.3.3. (2014). http://goldbook.iupac.org/pdf/goldbook.pdfHerbst, E.: The chemistry of interstellar space. Chem. Soc. Rev. 30, 168-176 (2001).Karol, P. J.: The Mendeleev–Seaborg Periodic Table: Through Z = 1138 and Beyond. J. Chem. Educ. 79, 60-63

(2002).Kragh, H.: On the Ontology of Superheavy Elements. Substantia 1(2): 7-17 (2017).Scerri, E. R.: The Periodic Table – Its Story and Its Significance. University Press, Oxford GB (2007).Schummer, J.: The Chemical Core of Chemistry - A Conceptual Approach. HYLE 4, 129-162 (1998).

44

ISPC 2018

Schwarz, W. H. E.: The Full Story of the Electron Configurations of the Transition Elements. J. Chem. Educ.87, 444-448 (2010).

Schwarz, W. H. E., Rich, R. L.: Theoretical Basis and Correct Explanation of the Periodic System: Review and Update. J. Chem. Educ. 87, 435-443 (2010).

Simonyi, K.: A Cultural History of Physics. CRC Press, Bota Raton FL (2012).Van Spronsen, J. W.: The Periodic System of Chemical Elements – A History of the First Hundred Years.

Elsevier, Amsterdam (1969).Wang, S. G., Schwarz, W. H. E.: Icon of Chemistry: The Periodic System of Chemical Elements in the New

Century. Angew. Chem. Int. Ed. 48, 3404-3415 (2009).Wapstra, A. H.: Criteria that Must be Satisfied for the Discovery of a New Chemical Element to be Recognized.

Pure Appl. Chem. 63, 879-886 (1991).Wußing, H., Ed.: Geschichte der Naturwissenschaften. Aulis, Köln (1987).

45

ISPC 2018

22. Chemical Elements and Chemical Substances: Rethinking Paneth’s Distinction

Sarah Hijmans, Université Paris – [email protected];

In 1931, Paneth identified a dual meaning of the term chemical element and since then multiple philosophers of chemistry have also warned for ambiguities surrounding this concept. Today the IUPAC still holds a double definition, which defines the element as a type of atom on the one hand and an elementary or simple substance on the other. The aim of this article is to clarify the meaning of the concept of chemical element by reinforcing the distinction between elements and simple substances based on the idea that the simple substance is a chemical substance whereas the element isn’t. Some would say that chemical analysis shows us that simple substances are the components of matter and therefore they should be seen as instantiations of the elements. Nevertheless, we must not forget that laboratory analysis is a process which transforms matter: though they are linked, theoretical analysis (which is performed on symbols and shows composition in terms of elements) is not to be confused with operational analysis (which is performed on substances and chemically transforms compounds into elementary substances). Furthermore, the fact that elementary substances are composed of only one type of element is not sufficient for them to be called elements. On the contrary, elementary substances are chemical substances that are determined not only by their composition but also by their structure and chemical environment, whereas elements are abstract concepts that cannot be chemically transformed and that do not exhibit any macroscopic properties.

References:Bernal, A. and Daza, E. 2010. On the Epistemological and Ontological Status of Chemical Relations. HYLE –

An International Journal for the Philosophy of Chemistry 16(2): 80-103.Earley, J. E. Sr. 2006. Chemical “Substances” That Are Not “Chemical Substances”. Philosophy of Science

73(5): 841-852.Earley, J. E. Sr. 2009. How Chemistry Shifts Horizons: Element, Substance and the Essential. Foundations of

Chemistry 11: 65-77.Ghibaudi, E., Regis, A. and Roletto, E. 2013. What Do Chemists Mean When They Talk About Elements?

Journal of Chemical Education 90: 1626-1631.Hendry, R. F. 2006a. Elements, Compounds and Other Chemical Kinds. Philosophy of Science 73: 864-875.Hendry, R. F. 2006b. Substantial Confusion. Studies in the History and Philosophy of Science 37: 322-336. Jacob, C. 2001. Analysis and Synthesis: Interdependent Operations in Chemical Language and Practice. HYLE

– An International Journal for the Philosophy of Chemistry 7(1): 31-50. Paneth, F. A. 2003. The Epistemological Status of the Chemical Concept of Element. Foundations of Chemistry

5(2): 113-145. Copyright Oxford University Press 1962. Ruthenberg, K. 2009. Paneth, Kant and the Philosophy of Chemistry. Foundations of Chemistry 11: 79-91. Ruthenberg, K. 2015. Radicals, Reactions, Realism. In Philosophy of Chemistry: Growth of a New Discipline,

edited by E. Scerri and L. McIntyre. Dordrecht: Springer. Schummer, J. 1998. The Chemical Core of Chemistry I : A Conceptual Approach. HYLE – An International

Journal for the Philosophy of Chemistry 4: 129-162. Schummer, J. 2008. Matter versus Form, and Beyond. In Stuff: The Nature of Chemical Substances, edited by

K. Ruthenberg and J. van Brakel. Würzburg: Koningshausen & Neumann.Van Brakel, J. 2012. Substances: the Ontology of Chemistry. In Handbook for the Philosophy of Science, vol. 6:

Philosophy of Chemistry, edited by R. J. Hendry, P. Needham and A. Woody. Boston: Elsevier.

46

ISPC 2018

23. Values and uses: how Mendeleev’s valuing of completeness contributed to predictions.

Karoliina Pulkkinen, University of Cambridge. [email protected];

Recently, Hasok Chang has explicated the differences between Antoine Lavoisier’s and Joseph Priestley’s theories of combustion in terms of diverging epistemic values (Chang 2012, 22). Chang argues that Priestley’s account is characterised by its emphasis on completeness – a feature that Chang defines as “wanting to account for all the observed phenomena in a given problem-area and for all the observed aspects of those phenomena” (ibid.) In this paper, I argue that a similar emphasis on completeness can be found in Dmitri Mendeleev’s approach to classifying the chemical elements. While epistemic values were originally popularised in the context of theory comparison (esp. Kuhn 1977), this paper takes another approach: it examines the relationship between values guiding the development of scientific representation and the uses of those scientific representations. In particular, this paper seeks to establish a relationship between Mendeleev’s valuing of completeness (polnost’) and his detailed qualitative predictions.

This project was prompted by a particular quote from Mendeleev’s article Periodicheskaya Zakonnost’ (1871) where Mendeleev argues that keeping all the qualities of the elements in sight is important for making generalisations and for drawing “practical conclusions and chemical predictions.” (Mendeleev 1958, 104). This suggests a relationship between valuing of completeness and prediction. In particular, it appeared necessary for Mendeleev to attend to both shared and individual properties of elements for making predictions.

After contextualising Chang’s definition of completeness to make it fit with the historical situation of classifying the elements in the 1860s, I show how Mendeleev’s approach to classification can be characterised by its emphasis on completeness. For example, valuing of completeness is suggested by Mendeleev’s consideration of a large variety physical and chemical features in relation to the periodic system. A particularly important role is played by oxides and hydrogen compounds, which Mendeleev visually included in his periodic system (thus marking a difference to the classifications of other chemists).

I will then show how Mendeleev used his knowledge on oxides and hydrogen compounds to predict the properties of little-known elements like indium and unknown elements like eka-boron (scandium). After contrasting these predictions with Mendeleev’s later unsuccessful ones, it will become clear that Mendeleev’s treatment of oxides and hydrogen compounds contributed to his successful extrapolations of the properties of unknown elements. Eric Scerri has argued Mendeleev’s later unsuccessful predictions were characterised by their reliance on the numerical factors (Scerri 2007, 142). This suggests that valuing of completeness in accounting for chemical findings was necessary for successful predictions.

Finally, I argue that it would be fruitful to examine the relationship between values guiding the development of scientific representation and the uses of those scientific representations. Such approaches would widen the discussions concerning the roles of values in science beyond a focus on evaluation and appraisal of scientific representations.

Chang, Hasok. 2012. Is Water H2O? Dordrecht: Springer.Kuhn, Thomas. 1977. ‘Objectivity, Value Judgment, and Theory Choice’. In Essential Tension, 320–39.

Chicago: University of Chicago Press.Mendeleev, D.I. 1958. Periodichiskij Zakon: Redakzija, Statja I Primechanija B.M. Kedrova. Edited by B.M

Kedrov. Moscow: Akademija Nauk CCCR.Scerri, Eric R. 2007. The Periodic Table: Its Story and Its Significance. New York: Oxford University Press.

47

ISPC 2018

24. Kant’s Conjoint Reconception of Chemistry and Science.

Farzad Mahootian, New York [email protected];

In response to the chemical theories of Lavoisier and Newton, and new findings from chemical laboratories throughout Europe, Kant shifted his attention from relatively abstract “matter” to relatively concrete “elements”1. His interest in chemistry became increasingly intense from about 1780 onward, though he already favored chemistry as a metaphor for practical reasoning and analysis in other writings2. His late work focused on a new issue: the significance of the radical diversity and specificity of chemical elements. The status of chemistry as a science is signalled in Kant’s second edition of his first Critique(CPR)3, culminating in several late writings, but especially in his Opus Postumum (OP).

This new approach involved radical shifts in his idea of chemistry and also, as van Brakel has suggested4, in his idea of science. I follow up this suggestion by investigating how reconfiguring the metaphysical foundations of natural science in OP’s “Transition” project5 has epistemological consequences for the element concept. In his attempt to account for newly encountered chemical properties and interactions, Kant explicitly suspended the CPR’s key distinction between regulative and constitutive reason6. He did so in at least one case: the aether, a term used in Kant’s time for the all-pervasive, pluripotent medium of heat and electricity. Suspending the constitutive/regulative distinction affects all four of Kant’s “antinomies of pure reason,” but the 2nd antinomy (the “transcendental atomistic”) is affected in a manner especially relevant to chemists. The focus shifts from mereology to the core of Kant’s concern with chemistry, namely: questions about the identity, integrity and interactions of elements.

Kant passed away before he could resolve the problematic he had set out in his Transition project. In Cassirer’s neo-Kantian philosophy of chemistry7,8 we may find a further articulation of Kant’s late work and an interesting bridge to some contemporary mathematical 9 approaches to the concept of chemical elements and periodic tables.

1 Carrier, M., (2001) “Kant's Theory of Matter and His Views on Chemistry” in Watkins, E. (ed.), Kant and the Sciences. (NY: Oxford University Press)

2 van Brakel, J. (2006) “Kant’s Legacy for the Philosophy of Chemistry” in D. Baird et al. (eds.), Philosophy of Chemistry, 73. (NL: Springer).

3 Kant, I, (1929/1965) Critique of Pure Reason, trans. Norman Kemp Smith. (NY: St. Martin’s)4 van Brakel, 75, n.405 Übergang, the Transition from the Metaphysical Foundations of Natural Science to Physics.6 Kant, I, A 644-5.7 Cassirer, E., (1923/1953) Substance and Function (NY: Dover), 214.8 Schummer, J., (1998) “The Chemical Core of Chemistry” in HYLE – An International Journal for the

Philosophy of Chemistry, Vol. 4, 156-7.9 Restrepo, G., Llanos, E.J. & Mesa, H. (2006) “Topological Space of the Chemical Elements”.

48

ISPC 2018

Wednesday 18 July 2018, Morning. Quantum chemistry and related topics.

25. Landing Zones – A Case of Model Transfer in Chemistry.

Justin Price, University of South [email protected];

Recent literature in philosophy of science targets model transfer as a phenomenon in need of analysis. A characteristic feature of modeling is the recycling of mathematical tools. These tools may see use in domains much different from the ones where they originate. In order to explain how some models have cross-domain applicability, philosophical discussion identifies the properties that drive their transfer. What is it about certain models - the Ising model, Lotka-Volterra model - that make them transferable to different domains? There have been two approaches to answering this question so far.

Paul Humphreys, in the book Extending Ourselves (2004), introduces the notion of a computational template for analysis of the transferable mathematical tools. At theheart of a well-travelled model is a computational template - an equation, function or computational method. According to this approach, computational templates are transferable because they are tractable and generalizable, features they owe to a particular syntactic form. For example, in physics, scientists use the Ising model to study ferromagnetism. Computational methods at the heart of this model, called Percolation theory, now comprise the mathematical core of models used to study neural networks and fungal infections in orchards. According to Humphreys the shared computational template draws together the models of these disparate phenomena. (Humphreys 2004, p. 70).

Knuuttila and Loetgers’ model templates constitute the second approach in identifying the factors important to explaining model transfer, building on Humphrey’s notion of template. According to them, the analysis of syntactic form does not exhaust the factors involved in model transference. They write: “However, the model transfer between the Ising model, the Sherrington-Kirkpatrick model and the Hopfield model cannot be accounted for by tractability alone. In this case, apart from mathematical forms and methods, a very general conceptual idea of the kind of mechanism of interaction involved was also transferred.” (Knuuttila and Loetgers 2014) This general idea is embedded in the mathematical form. Model templates, they argue, facilitate the transference of models because they sensitize us to patterns shared by a variety of empirical systems. Thus, both computational and conceptual features of a particular mathematical tool contribute to its transference.

Both of these approaches analyze features that transferred templates possess in order to explain model transfer. I seek to bring to light a new question in this discussion, and provide a preliminary answer. What is it about some modelling contexts that facilitate the transfer of templates? In my view, the successful application of a computational or model template requires the preparing of a ground, a landing zone for transferred elements. Using a template in a model has formal conditions. Since templates are generalizable, and able to be applied in many types of domains, they require transformation in order to be a representation of a particular target system. Along with formal conditions, I argue that there are conceptual conditions for this specification of a template to a particular domain, conceptual conditions that require reference to notions about ontology. The notion of a landing zone identifies features within a model that satisfy these conditions for specification of a general template. A landing zone consists of an ontology - the kind of objects represented by the model, the nature of the properties and relations of these kinds of objects, and mechanisms of interaction between these objects. This ontology mediates the application of a template to a particular target system in the following three ways.

49

ISPC 2018

To demonstrate that the notion of landing zones is an important contribution to the discussion on model transfer, I will be providing a case study on a model in chemistry that uses computational templates transferred from physics. This model is the Quantum Theory of Atoms in Molecules (QTAIM) and the transferred templates it uses are the virial theorem and the wave function. First, I will be examining QTAIM’s use of the virial theorem as a technique to make calculating certain chemical properties more tractable. The virial theorem in this case is both a model template and a computational template. There is an ontology carried by it, the virial relation, and it sees uses in QTAIM as a computational method for making calculation of molecular energy from a wave function tractable. The virial theorem has a well-known conceptual condition for its use as a technique that the landing zone in QTAIM satisfies; it requires a system is in equilibrium. I will also show that the transfer of the wave function from quantum mechanics has conditions that the landing zone in QTAIM satisfies. The wave function requires approximations for its use to be tractable in chemistry. Choices concerning how to

approximate the wave functions require justification. The landing zone provides ground for these approximations

QTAIM’s landing zone for the use of mathematical tools transferred from physics consists of the topological atom. This landing zone satisfies three conditions for the transfer of the virial theorem and the wave function to chemistry. First, it satisfies the virial theorem’s condition for risk-free use. In other words, misuse of the virial theorem has the potential to conflict with empirical data. Owing to the topological atom’s zero-flux boundary definition, the virial relation is built into the notion of a topological atom removing risk in the use of the virial theorem as a computational method. The ground is prepared so that the virial theorem can land.

Computational templates also require transformation in order to be useful in the new domain. The landing zone satisfies this condition by providing a ground that justifies transformation of transferred mathematical templates with respect to approximations, constraints and parameterization. For example, the notion of the topological atom guides development of a distribution function in real space from the wave function, allowing QTAIM to represent the wave function as a distribution of charge density in three dimensions (real space) instead of using the multi-dimensional Hilbert space. This transformation of the wave function is necessary for a tractable description of chemical properties.

Its landing zone provides ground for the transformation of computational templates – the virial theorem and wave function - towards this representation of chemical properties as quantum mechanical, and grounds inferences made about chemical properties from prediction and mathematical description produced by QTAIM.

BibliographyBader, R.F.W., Beddall, P.M. 1972: “Virial field relationship for molecular charge distributions and the spatial

partitioning of molecular properties”. J. Chem. Phys. 56, 3320–3328 Bader, Richard F. W. 1975. “Molecular Fragments or Chemical Bonds.” Accounts of Chemical Research 8 (1):

34–40.Bader, Richard. 1985. “Atoms in Molecules.” Accounts of Chemical Research 18 (1): 9–15.Bader, Richard F W. 2011. “On the Non-Existence of Parallel Universes in Chemistry.” Foundations of

Chemistry: Philosophical, Historical and Interdisciplinary Studies of Chemistry 13 (1): 11–37.Bader, Richard F W, and Chérif F Matta. 2013. “Atoms in Molecules as Non-Overlapping, Bounded, Space-

Filling Open Quantum Systems.” Foundations of Chemistry: Philosophical, Historical and Interdisciplinary Studies of Chemistry 15 (3): 253–76.

Burdett, J. 1997. Chemical Bonds – A Dialogue. Wiley: New York.

50

ISPC 2018

Haaland, A., Helgaker, T.U., Ruud, K., Shorokhov, D.J. 2000: “Should gaseous BF3 and SiF4 be described as ionic compounds?” J. Chem. Edu. 77, 1076 

Hiberty, P., Shaik, S. 2014. “Bridging Cultures.” The Chemical Bond: Fundamental Aspect of Chemical Bonding edited by Gernot Frenking and Sason Shaik. Wiley-VCH.

Hohenberg, P., Kohn, W. 1964. “Inhomogeneous Electron Gas.” Physical Review 136: 864-871 Humphreys, Paul. 2002. “Computational Methods.” Philosophy of Science, 69 (3): 1-11Humphreys, Paul. 2004. Extending Ourselves: Computational Science, Empiricism, and Scientific Method.

Oxford University Press.Fisher, R.A. 1918. "The Correlation between Relatives on the Supposition of Mendelian Inheritance."

Philosophical Transactions of the volume 52, pages 399–433.Kohn, W., Sham, J. 1965. “Self-Consistent Equations Including Exchange and Correlation Effects.” Physical

Review 140 (4a): 1133-1138Knuuttila, Tarja, and Andrea Loettgers. 2011. “The Productive Tension : Mechanisms Vs. Templates in

Modeling the Phenomena.” In Models, Simulations, and Representations, edited by Paul Humphreys and Cyrille Imbert. Routledge.

Knuuttila, Tarja, and Andrea Loettgers. 2014. “Model Templates Within and Between Disciplines: From Magnets to Gases – and Socio-Economic Systems.” European Journal for Philosophy of Science, 1–24.

Morrison, M. 1996. “Physical Models and Biological Contexts.” Philosophy of Science vol. 64 Parr, Robert G., Paul W. Ayers, and Roman F. Nalewajski. 2005. “What Is an Atom in a Molecule?” The

Journal of Physical Chemistry A 109 (17): 3957–59.Pauling, L. 1939. The Nature of the Chemical Bond. Cornell Univ. Press, Ithsea, NY.Proft, F., Ayers, P., Geerlings, P. 2014. “The Conceptual Density Functional Theory Perspective of Bonding.”

The Chemical Bond: Fundamental Aspects of Chemical Bonding edited by Gernot Frenking and Sason Shaik. Wiley-VCH.

Poater, Jordi, Miquel Solà, and F. Matthias Bickelhaupt. 2006. “Hydrogen–Hydrogen Bonding in Planar Biphenyl, Predicted by Atoms-In-Molecules Theory, Does Not Exist.”Chemistry – A European Journal 12 (10): 2889–95. doi:.

Schmidt, M., Ivanic, J., Reudenberg, K. 2014. “The Physical Origin of Covalent Bonding.” The Chemical Bond: Fundamental Aspects of Chemical Bonding edited by Gernot Frenking and Sason Shaik. Wiley-VCH.

Sukumar, N. 2013. “The Atom in a Molecule as a Mereological Construct in Chemistry.” Foundations of Chemistry: Philosophical, Historical and Interdisciplinary Studies of Chemistry 15 (3): 303–9.

von Oertzen, W., Martin Freer, and Yoshiko Kanada-En’yo. 2006. “Nuclear Clusters and Nuclear Molecules.” Physics Reports 432 (2): 43–113. doi:. 

51

ISPC 2018

26. Indulgence to quantum mechanics.

Juan Camilo Martínez González and Olimpia Lombardi, Buenos Aires.CONICET – Universidad de Buenos Aires.

[email protected]; [email protected];

If we want to do Chemistry and think QM could help, then we have to be as indulgent to QM as we would be to an aged parent or a reluctant lover. We must tell her that she looks younger/more beautiful without parity, that we think her more fitter/seductive now she has given up nuclear permutation symmetry and so on.Brian Sutcliffe, personal communication, 2016.

Chemistry, without molecular structure, would be unintelligible. Perhaps precisely for this reason the discussion about the limitations of quantum mechanics to explain molecular structure has a long history. As Robin Hendry claims, “molecular structure is so central to chemical explanation that to explain molecular structure is pretty much to explain the whole of chemistry” (Hendry 2010, p. 183). In the literature, many different arguments have been put forward to stress the obstacles to that explanation.

Usually it is noticed that, through the Born-Oppenheimer approximation, electrons and nuclei are treated differently: whereas the Schrödinger equation is used to describe the behavior of electrons, nuclei play the role of classical entities, whose position in space supplies the Coulomb potential where the electrons move. However, this strategy introduces the molecular structure by hand: from a theoretical viewpoint, those models “simply assume the facts about molecular structure that ought to be explained” (Hendry 2010, p. 186). But this is not the strongest argument to reject interpreting the Born-Oppenheimer approximation as a reductive tool. Conceiving nuclei as systems at rest in definite positions leads to a contradiction with the Heisenberg principle, according to which no quantum system can simultaneously have definite values of position and of momentum. As Hasok Chang clearly states: “In this «clamping-down» approximation, the atomic nuclei are treated essentially as classical particles; […] this picture is non-quantum in a very fundamental way as the simultaneous assignment of fixed positions and fixed momenta (namely, zero) to them violates the Heisenberg uncertainty principle.” (Chang 2015, p. 198).

Despite the above remarks, from a merely instrumentalist viewpoint it can be argued that the conceptual weirdness of the Born-Oppenheimer approximation is innocuous, since the approximation leads to almost the same results that would be obtained with no approximation. Hendry (1998, 2010) calls this strategy the “proxy” defense of Born-Oppenheimer models, which is based on the assumption that using them instead of the exact solution makes only a small difference to the energy. It is at this point of the argumentation that optical isomers enter the scene: the molecular structures of the members of a pair of optical isomers are non-superposable mirror images. However, Coulomb Hamiltonians are invariant under spatial reflection; therefore, nothing in quantum mechanics can account for chirality.

The problem of isomerism is a particular case of what can be called, following Woolley and Sutcliffe (1977), the symmetry problem: if the interactions embodied in the Hamiltonian of the molecule are Coulombic, the solutions of the Schrödinger equation are spherically symmetrical; however, the asymmetry of polyatomic molecules is essential in the explanation of their chemical behavior. As Hendry (1998) stresses, in the quantum theoretical domain no directional properties can be assigned to an isolated molecule in a general energy eigenstate. This fact seems to be the silver bullet for the reductionist view of the relationship between chemistry and physics.

52

ISPC 2018

In this presentation we want to stress that, although optical isomerism makes the impossibility of reduction more explicit, when quantum mechanics in its standard version is taken seriously, the same obstacle is found in the case of structural isomers, since they have the same Hamiltonians. The idea that the case of optical isomerism is more serious is based on the idea that the definite positions of the nuclei introduce a difference in the Hamiltonians of distinct structural isomers because the distances between those positions are not the same in the different isomers. But this idea inadmissibly introduces classical assumptions into quantum mechanics. From an exclusively quantum viewpoint, once the components of the molecule are given, the Hamiltonian is fixed and is the same for all the isomers. In Woolley’s words: “we return to the example of C3H4; in the present terms this represents a collection of 3 carbon nuclei, 4 protons and 22 electrons. […] Suppose we apply quantum mechanics to all the particles in one go, what do we get? It is easier to say what we have never found so far –no suggestion of three distinct isomers for the molecules of allene, cyclopropene and methyl acetylene.” (Woolley 1998, p. 11).

But, what is the appropriate treatment for this diagnosis? The most reductionist temperaments assume that the idea of a neatly defined molecular structure is a “metaphor.” Others suppose that the obstacles will be overcome and the concept of molecular structure will arise when the simplifications will be removed and the quantum description will be completed with the interactions with other molecules, or with decoherence, or with more detailed Hamiltonians, etc. Independently of the dubious conclusions about the success of those strategies, it is interesting to notice that almost nobody seriously considers the possibility of modifying quantum mechanics itself. But, if the concept of molecular structure is so fruitful in chemistry and the shape of the molecules is so clearly detected by experimental means, why not think that molecules do have geometrical shape as described by chemistry?, why not consider that we must be indulgent with quantum mechanics, since it is a mere calculation tool that must be replaced in the future by a new theory that really describes microscopic reality?

References.Chang, H. (2015). “Reductionism and the relation between chemistry and physics.” Pp. 193-209, in T.

Arabatzis, J. Renn, and A. Simões (eds.), Relocating the History of Science: Essays in Honor of Kostas Gavroglu. Dordrecht: Springer.

Hendry, R. F. (1998). “Models and approximations in quantum chemistry.” Pp. 123-142, 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.

Woolley, R. G. (1998). “Is there a quantum definition of a molecule?” Journal of Mathematical Chemistry, 23: 3-12.

Woolley, R. G. and Sutcliffe, B. T. (1977). “Molecular structure and the Born-Oppenheimer approximation.” Chemical Physics Letters, 45: 393-398.

53

ISPC 2018

27. About the ontological status of phonons.

Sebastian Fortin1, Manuel Herrera2 and Jesus Alberto Jaimes Arriaga2

1 CONICET-University of Buenos Aires and Department of Physics, FCEyN, UBA.2 CONICET- University of Buenos Aires.

[email protected]; [email protected];

Recent discussions on the structure of matter generally focus on the ontological aspects of the molecular structure (Sutcliffe and Wolley 2012, Fortin et al. 2017). However, the diversity of chemical substances is not exhausted by a description based on atoms bound through the covalent bond. There are other types of substances that are not included in this category, for example, salts. In general, crystalline solids are described in terms of a periodic structure; so, the study of their chemical properties requires a completely different formalism that that used for molecules.

According to the standard formalism, it is assumed that N identical monatomic units compose the crystal. These units interact with each other through a force that is proportional to the distance. In this way they form a network. By means of a suitable change of variables (the so-called ‘phononic variables’), the network can be interpreted as a sum of harmonic oscillators (Kittel 1998.) In a later step, the quantum formalism is applied to this network of harmonic oscillators. As a result, the total

energy U of the solid is quantized and can be expressed as:

where and are constants. In this way, the solid is conceived as a network of atoms that can vibrate around its equilibrium position generating propagating waves. The fact that the energy of these waves is quantized suggests that the nature of these waves is similar to that of the electromagnetic field, which in quantum mechanics is represented by photons.

Within this formalism comes the idea of the phonon. A phonon is a quantum of vibration (or

sound), a quantum particle with an energy equal to , so that n phonons have an energy equal to

. In this way, the model of the solid based on the oscillating monatomic units can be left aside in favor of a model according to which the atoms of the network are fixed and the vibrations are represented by particles called phonons. Thus the vibrational energy of the solid is interpreted as the sum of the energies of the phonons in it. This model allows us, for example, to think heat as a flow of phonons through the solid and thus to reach an expression for the specific heat capacity that corresponds to experimental observations.

Scientists generally conceive the phonon as a pseudo-particle, that is, a mathematical device necessary to perform calculations but that does not have a “real” existence. However, some philosophers of science have questioned this position and have proposed understanding phonons as emerging entities (Franklin and Knox 2017). In this paper, we will study the possibility of conceiving the ontological status of phonon from a pluralist perspective (Lombardi and Labarca 2005). With this purpose, we will analyze the result of applying to the phonons the same criteria as that used to say that, for example, the Higgs boson is a fundamental particle.

54

ISPC 2018

28. References:Fortin, S., Lombardi, O. and Martínez González, J. C. (2017). “The relationship between chemistry and physics

from the perspective of Bohmian mechanics”, Foundations of Chemistry 19: 43-59.Franklin, A. and Knox, E. (2017). “Emergence without limits: the case of phonons”, PhilSci-Archive: Preprints

in Philosophy of Science, http://philsci-archive.pitt.edu/13397.Kittel, C. (1998). Introduction to Solid State Physics. Reverté: Barcelona.Lombardi, O. and Labarca, M. (2005). “The ontological autonomy of the chemical world”, Foundations of

Chemistry 7: 125-148.Sutcliffe, B.T. and Woolley, R.G. (2012). “Atoms and molecules in classical chemistry and quantum

mechanics”, in: Hendry, R. F., Woody, A. (eds.), Handbook of Philosophy of Science. Vol. 6, Philosophy of Chemistry. Elsevier: Oxford.

55

ISPC 2018

28. The Quantum Theory of Atoms in Molecules from a Bohmian perspective

Jesus Alberto Jaimes Arriaga1 and Sebastian Fortin2, 2CONICET-University of Buenos Aires and Department of Physics, FCEyN, UBA.

1 CONICET- University of Buenos [email protected]; [email protected];

One of the strongest arguments against the reduction of chemistry to physics is the one concerning the impossibility of recovering molecular structure from quantum mechanics. In this context, the Quantum Theory of Atoms in Molecules (QTAIM) (Bader 1990) has begun to attract the attention of chemists and philosophers of chemistry (see, e.g., Matta et al. 2011; Matta 2013), since it seems to be a rigorous formal proposal consistent with the principles and laws of quantum mechanics, which offers a theoretical foundation to the concept of molecular structure. This theory supports the idea that the physical manifestation of matter is given by the spatial distribution of the electron density and it is through this topology that the chemical concepts of atom, bond and structure can be defined. In particular, an atom in a molecule is bounded by a zero-flux surface in the gradient vector of electron density. This suggests that no electron can crossed such a surface and, hence, the electron density associated to each atom remains unchanged over time. In other words, it can be said that each particular electron belongs to a particular atom. However, this QTAIM view is completely alien to the conceptual world of standard quantum mechanics (SQM), where the picture of individual and distinguishable particles with defined though unknowable trajectories finds no place.

The idea of obtaining the electron density through the wave function was proposed early in the begin of quantum mechanics by the very Schrödinger, who advocated avoiding the use of wave function in the direct interpretation of physical phenomenon, since this is a function in configurational space and not in real –physical- space (Schrödinger 1926a, b). Then Bader introduced the electron density view into his theory as a fundamental quantity, which contains all the physical information about chemical systems (Bader 2011). This vision is far away of the proposal made by Born, who argued that the quantity |Ψ|2 represents a density of probability according to the SQM. The mere fact of interpreting the wave function in a different way that the SQM does discards many problems posed by the standard view. For instance, the principle of superposition has no sense in the framework of QTAIM, since the wave function does not represent quantum states but is only the mean for obtaining the electron density.

In this point is important to note that the SQM perspective has prevailed to face the problem of the relationship between chemistry and physics, leaving aside other formalisms as the Bohmian Quantum Mechanics (BQM). A surprising fact when it is considered that the conceptual scheme of BQM is more consistent with the assumptions raised by the QTAIM. To prove this, let us consider the important characteristic of the individuality of the systems under study in each theory. According to the SQM, the quantum systems do not possess a criterion of individuality due to the phenomenon of quantum indistinguishability. Moreover, when two quantum particles interact with each other they form a new system that is treated holistically and non-locally, where the original quantum particles have lost their individuality. In the case of BQM the particles possess well-defined positions and momenta while describe trajectories, whence they are clearly identified over time and can be considered individuals in the classical way. With respect to the QTAIM, the atoms in the molecule can also be considered individuals, because they are well identified within the molecule and retain

56

ISPC 2018

their individuality while their electron density associated is constant through time. In such a case, the criterion of individuality is a feature that the QTAIM and the BQM share.

On the other hand, by means of the conceptual basis posed by BQM is possible to interpret in a better way the concepts proposed by the QTAIM. Let us take the zero-flux surface as an example. According to Bader this surface confined the electron density associated to each atom, because cannot be traversed by any electron, that is, by any Bohmian particle. Thus, it can be said that each Bohmian electron belongs to a specific atom. Additionally, the QTAIM proposed the concept of bond path as a universal indicator of bonded interactions (Bader 1998) and this one being an accumulation of electron density between the nuclei that maintain the atoms together, an idea already suggested by Heitler and London in 1927. The interesting fact is that this vision is consistent to the explanation accounted by the BQM (Fortin et. al. 2017).

A final point to emphasize is the argument raised by Shan Gao in recent works (Gao 2013, 2017) in which he defends that the electron density can be seen as an effective distribution originated by the ergodic motion of a charged particle, that is, an electron. According to Gao, in a one-charged particle system, the ergodic motion of the electron forms a “cloud” of charge whose density is proportional to the square of the wave function. This argument reinforces the vision of QTAIM and is allied to the picture proposed by BQM, which can offer an adequate underlying ontology to recover charge density for quantum chemistry by appealing to Gao’s argument.

In summary, unveiling the role played by the QTAIM in the intertheoretical relationships between chemistry and physics is of particular relevance. The possible connections or conceptual ruptures between the QTAIM and both the SQM and the BQM deserve to be analyzed. This analysis paves the way toward a possible explanation of the electron density as used by the QTAIM in terms of the fundamental dynamics of Bohmian particles.

ReferencesBader, R. (1990). Atoms in Molecules. A Quantum Theory. Oxford: Oxford University Press.Bader, R. (1998). “A bond path: a universal indicator of bonded interactions”, Journal of Physical Chemistry A,

102: 7314-7323.Bader, R. (2011). “On the non-existence of parallel universes in chemistry”. Foundations of Chemistry, 13: 11-

37.Fortin, S., Lombardi, O. and Martinez Gonzalez, J. C. (2017) “The relationship between chemistry and physics

from the perspective of Bohmian mechanics”, Foundations of Chemistry, 19: 43-59.Gao, S. (2013). “Is an electron a charge cloud? An reexamination of Schrödinger´s charge density hypothesis”.

http://philsci-archive.pitt.edu/id/eprint/9696Gao, S. (2017) The meaning of the wave function. Cambridge and New York: Cambridge University Press.Heitler, W. and London, F. (1927). “Interaction between neutral atoms and homopolar binding according to

quantum mechanics”. Zeitschrit für Physik, 44: 455-472.Matta, C. F. (2013). “Special issue: Philosophical aspects and implications of the quantum theory of atoms in

molecule (QTAIM)”. Foundations of Chemistry, 15: 245-251.Matta, C. F., Massa, L., and Keith, T. A. (2011). “Richard F. W. Bader: a true pioneer”. Journal of Physical

Chemistry A, 115: 12427-12431.Schrödinger, E. (1926a). “Quatisierung als Eigenwertproblem (Vierte mitteilung)”. Annalen der Physik, 81:

437-490. Traducción al inglés en E. Schrödinger, Collected Papers on Wave Mechanics 1928. London and Glasgow: Blackie & Son Limited.

Schrödinger, E. (1926b). “An undulatory theory of the mechanics of atoms and molecules”. Physical Review, 28: 1049-1070.

57

ISPC 2018

29. Overcoming skepticism about molecular structure by developing the concept of affordance

Hirofumi Ochiai, Nagoya Bunri University, [email protected];

The fact that chemical syntheses based on designing molecules have been successfully performed seems to suggest molecular structure is real.1) On the other hand quantum mechanical calculations prove the existence of neither chemical bonds nor molecular structure. Is molecular structure real or not? One way to answer this question is to take the different outcomes not as a contradiction but as different affordances.2) Since it often happens that one and the same object gives birth to different affordances, the attributes of material objects should be determined by taking the possible affordance into consideration. Such is the case with the particle-wave duality of an electron, for instance. Niels Bohr claims that evidence obtained under different experimental conditions cannot be comprehended within a single picture, but must be regarded as complementary in the sense that only the totality of the phenomena exhausts the possible information about the objects.3) In fact, because we are involved in phenomena, we are liable to miss possible affordances and commit mereological fallacies. 4) The theory of phenomenal field which I propose in this study will help us see a whole picture of possible affordances. Phenomenal fields are conditions, under which phenomena to be observed become realized as affordances in such a way that the phenomena observed are necessary and inevitable to the conditions. Because phenomenal fields reflect observers’ interests, affordances arising from one and the same object may well be distinct from one another. This fact suggests that molecular structure is neither an emergent nor a supervenient property of the physical system composing the molecule. This theory sheds light on the philosophical aspect of chemical element as well.

References1. Hacking, I., 2008: Representing and Intervening, 22nd edn. Cambridge Univ. Press, Cambridge.2. Harre, R., 2014: New Tools for Philosophy of Chemistry, Hyle, pp77-91 (2014).3. Primass, H., 1983: Chemistry, Quantum Mechanics and Reductionis, Springer-Verlag, p.31.4. Bennett, M.R. and Hacker, P.M.S., 2014: Philosophical Foundations of Neuroscience, Blackwell, p.73.

58

ISPC 2018

30. On the significance of Heitler and London 1927

James Ladyman, Department of Philosophy, University of [email protected];

The valence bond theory of Lewis 1916 was based on chemical facts and could not be explained with the physics of the time. When in 1927 Heitler and London applied quantum mechanics to the problem of homopolar bonding of the neutral atoms, quantum chemistry was founded, and at the same time the general idea of covalent bonding was vindicated, and some of the details of Lewis’s ideas were confirmed. Yet while current chemical education and practice still involves covalent bonds, most bonds are best treated as at least partially ionic. The physical understanding of molecules involves more or less delocalised electrons, and no absolute distinction between the extreme cases of covalent and ionic bonding. This case is instructive for history and philosophy of science. It is an excellent example of empirical success and a type of unification in so far as the theory of bonding led to many successful empirical predictions and models, and to an understanding of chemistry in terms of quantum physics respectively. However, there is no reduction of chemistry (in the sense usually meant by the term in philosophy). Quantum chemistry has three pillars, namely the Schrödinger equation, the Pauli exclusion principle and group theory. The latter is essentially involved because of the need to simplify and constrain problems by reference to physical symmetries. Such considerations in quantum chemistry often involve putting aspects of molecular structure by hand. For example, Heitler and London fix the distance of the nuclei in their calculation for H2. This illustrates how quantum mechanics is used in scientific practice without a complete interpretation, or an explicit solution to the measurement problem. Chemical models are like semi-classical models in physics in supposing spatial and other structures as a framework for quantum mechanical calculations. In 1931 Majorana made quantitative predictions of the dissociation energy of Helium ions that have only been experimentally confirmed recently. The very early application of quantum mechanics to chemistry and the role that played in the acceptance of quantum mechanics, poses a challenge for those who argue that hidden variable theory has all the empirical support that quantum mechanics does. I argue that unless there is a viable counterfactual history in which quantum chemistry is developed in the context of a pilot wave model, that there is no interesting sense in which there is underdetemination of theory choice in this case. More generally, the development of the idea of the chemical bond through quantum physics teaches us a lot about ontology in science, since there is a clear sense in which chemical bonds exist, and yet they also clearly dissolve in electron density when we look closely at them.

References:Ballhausen, C.J. Quantum Mechanics and Chemical Bonding in Inorganic Complexes. J. Chem. Ed. 6, 215-218

(1979)Esposito, S. and Naddeo, A. (2015), Majorana, Pauling and the quantum theory of the chemical bond. Annalen

der Physik, 527: A29-A33. doi:10.1002/andp.201400154Heitler, R., and London, F. Z. Phys. 44 (1927) 455.Hettema, H. The Union of Chemistry and Physics. Dordrecht: Springer (2017)Pauling, L. Chem. Rev.5 (1928) 173.

59

ISPC 2018

Wednesday 18 July, Afternoon. Emergence, molecular structure, and radicals.

31. Emergence in Chemistry: Substance and Structure.

Robin F. Hendry, Department of Philosophy, University of [email protected];

Chemistry has a history in the emergence debate. Even before the term ‘emergence’ acquired its modern philosophical usage, John Stuart Mill cited chemical compounds as the bearers of emergent properties that could not be predicted from those of their constituent elements. The same is true of Mill’s successors in the philosophical tradition that Brian McLaughlin has called ‘British Emergentism’, who included C.D. Broad (McLaughlin 1992). Chemistry is not now widely cited as a rich source of candidate examples of emergence (although see Humphreys 2016). Firstly the emergence debate has moved beyond the epistemic criteria for emergence, such as predictability, which were applied by the British Emergentists. Secondly, many philosophers and scientists would no doubt agree with McLaughlin’s judgement that the advent of quantum mechanics in the 1920s, and the explanatory advances that came in its wake, rendered the British Emergentist’s central claims about chemistry and biology ‘enormously implausible’.

There should be no doubt about how significant are the scientific advances that unified chemistry and physics for understanding the relationship between the two sciences. But they are much more complicated than McLaughlin’s account allows. During the nineteenth century, chemists established that chemical substances are composed of a finite number of more basic substances: the chemical elements. To account for the existence of isomers—distinct chemical substances which are composed of the same elements in the same proportions—chemists developed theories of structure. Differences between isomers were accounted for by the different ways in which the same elements are combined within them, that is, by their different structures. In the early twentieth century, investigations of atomic structure offered a physical account of the nature of the elements: they are essentially classes of atoms which are alike in respect of their nuclear charge, but which may differ in respect of their mass. G.N. Lewis offered an account of how bonds are realised by shared electrons: although this identification produced many novel insights into the mechanisms of chemical reactions, especially in organic chemistry (see Hendry 2017), Lewis’ work was in its turn superseded by the advent of quantum mechanics.

Leaving aside the plausibility or otherwise of the British Emergentists’ particular conception of emergence, in this paper I argue that chemistry is a far more fertile ground for emergence than many contemporary scientists and philosophers allow. I focus on two key issues—chemical substances and their structures—arguing that it is perfectly plausible to see the unification of chemistry and physics as a synthesis of chemical and physical theory, rather than a derivation of one from the other.

Substance Chemistry studies substances, accounting for their chemical and spectroscopic behaviour in terms of their structure. In a number of papers I have defended the idea that chemistry individuates substances by their structures at the molecular scale (2006, 2010, 2016a), and would argue further that this information provides the basis for a strong argument for microstructural essentialism, the more robustly metaphysical thesis that a chemical substance is the particular substance it is in virtue of its microstructure. Many philosophers see claims of this sort as providing a

60

ISPC 2018

quick argument for the reducibility of substances to their microstructures. I argue in contrast that it is plausible to see substances as emergent, and bearing emergent causal powers, with respect to the populations of molecular species from which they arise.

Structure The physical sciences, including chemistry, condensed matter physics and crystallography, deploy two distinct conceptions of structure: geometrical structure and bond structure. They do not (I argue) treat either as being more fundamental than the other (Hendry 2016b). Geometrical structures involve spatial relationships between species at the molecular scale. I argue that they are emergent from dynamical processes at the molecular scale: different structures emerge at different scales of energy, time and length, and there is no reason to privilege processes at higher energy scales, or shorter length- or time-scales. Bond structures are the networks of bonds connecting atoms or ions, familiar from the long-established structural formulae of organic chemistry. Quantum-mechanical explanations of chemical bonding have launched many philosophical arguments for reduction, but the emergence of quantum chemistry—the application of quantum-mechanical principles to systems of electrons and nuclei—was not a deductive affair, and was seen by some of its founders as a synthesis of chemistry and physics. In keeping with this view, I argue that quantum-mechanical explanations of bond structures assume that they arise and persist within quite specific dynamical conditions involving the adiabatic separability of nuclear and electronic motions and the localisation of nuclei. Quantum mechanics provides not a derivation of structure, but an explanation of how it is possible for it to exist. I therefore argue that there are no purely scientific reasons to believe that bond structures are determined to exist by the quantum mechanics of systems of electrons and nuclei alone.

References.Hendry, R.F. 2006 ‘Elements, compounds and other chemical kinds’ Philosophy of Science 73, 864-875.Hendry, R.F. 2010 ‘The elements and conceptual change’ in Helen Beebee & Nigel Sabbarton-Leary (eds.) The

Semantics and Metaphysics of Natural Kinds (London: Routledge), 137-158.Hendry, R.F. 2016a ‘Natural kinds in chemistry’ in Grant Fisher and Eric Scerri (eds.) Chapters in the

Philosophy of Chemistry (Oxford: Oxford University Press), 253-75.Hendry, R.F. 2016b ‘Structure as abstraction’ Philosophy of Science 83, 1070-1081.Hendry, R.F. 2017 ‘Mechanisms and reduction in organic chemistry’ in M. Massimi, J.W. Romeijn and G.

Schurz (eds.) EPSA15 Selected Papers: The 5th Conference of the European Philosophy of Science Association in Düsseldorf. (Berlin: Springer), 111-124.

Humphreys, Paul 2016 Emergence: A Philosophical Account (Oxford: Oxford University Press)McLaughlin, B. 1992 ‘The rise and fall of British Emergentism’ in A. Beckermann, H. Flohr and J. Kim, (eds.)

Emergence or Reduction? Essays on the Prospects for Non-Reductive Physicalism (Berlin: Walter de Gruyter), 49–93.

61

ISPC 2018

32. The Reality of the Chemical Bond: Revisiting the structural and energetic conceptions of the chemical bond

Vanessa Seifert, Department of Philosophy, University of [email protected];

Molecular structure is a collective term; it doesn't refer to one particular and empirically measurable property, but rather to a collection of empirically measurable properties of the molecule. Therefore, in order to examine whether the structure of a molecule is a real property, one needs to examine whether the set of empirically measurable properties which are sufficient for the specification of molecular structure, are real. 

A central concept which is standardly invoked in order to specify the structure of a molecule is the chemical bond. In this context, an important question that one needs to address when examining the reality of molecular structure is whether chemical bonds exist. In order to answer whether the world is occupied by chemical bonds, it is useful to specify what exactly we are looking for; that is, what is a chemical bond? Is it a property, an entity, an interaction, or a state of the molecule? What is the scale to which a chemical bond is relevant? To what extent is our understanding of a chemical bond an idealisation? Another pressing question regarding the reality of chemical bonds concerns the chemical bond’s relation to the entities, properties, etc. that are invoked by quantum mechanics in order to describe a molecule. Is the chemical bond identical with some set of quantum mechanical entities, etc., and if so what sort of identity is it; namely type-type, type-token, token-token identity? Does the chemical bond merely supervene on some set of quantum mechanical entities, etc. or does it emerge in a higher scale of ontology?

This paper primarily investigates the nature of the chemical bond.  In this context, it presents and examines two distinct conceptions of the chemical bond which were proposed by Robin Hendry; the structural and the energetic conception. According to the structural conception, chemical bonds are “material parts of the molecule that are responsible for the spatially localized sub molecular relationships between individual atomic centers” (p.917, Hendry 2006a). It encompasses chemistry’s understanding of the chemical bond, and is compatible with chemistry’s classification of chemical bonds. On the other hand, the energetic conception explains why a bond is formed in terms of the energetic stabilisation of the molecule, rather than defines what the chemical bond is. In fact, the energetic conception does not even require that there are chemical bonds within a molecule, since it merely identifies the energetic states in which an entity (i.e. the molecule) is more than an aggregate of its parts. 

Hendry argues that, while the energetic conception is compatible with all types of bonding and with quantum mechanics (unlike the structural conception), its choice over the structural conception leads to explanatory loss regarding chemical phenomena. Moreover, the structural conception faces several objections regarding (i) its compatibility with the quantum mechanical description of molecules, and (ii) its explanatory relevance to particular sets of substances. Michael Weisberg places further tension to the structural conception as he claims that central aspects of the structural conception are not ‘robust’ across different available

62

ISPC 2018

quantum models that describe molecular structure. This makes him skeptical towards the reality of the chemical bond, as this is described by the structural conception.

This paper argues that the aforementioned objections as well as the postulated tension between the energetic and the structural conceptions, can be partially overcome if we distinguish between the entity itself (i.e. the chemical bond) and the conceptions utilised to describe it (i.e. the “chemical bonds”). This is based on Quine’s distinction between intension and extension. In this context, the two conceptions can be understood as corresponding to two distinct, incomplete, yet complementary descriptions of the chemical bond; i.e. the “structural chemical bond” and the “energetic chemical bond”. As long as the two “chemical bonds” do not postulate incompatible features of the chemical bond, each can be taken to correspond to a distinct intension that describes one and the same extension; namely the chemical bond. In this context, it is possible to understand the shortcomings of each conception merely as epistemic disadvantages (in terms of explanatory, predictive, or descriptive incompleteness) of the respective intension, and not as metaphysical problems regarding the nature of the chemical bond. 

Within the context of such an understanding of the two conceptions, it is possible to specify the nature of the chemical bond (i.e. of the referent), but also to support its reality, by drawing information from both conceptions.

ReferencesAtkins P., Jones L. (2009), Chemical Principles, W.H. Freeman and Company: New YorkFrege G., (1980) , ‘On Sense and Reference’ in Translations from the Philosophical Writings of Gottlob Frege,

P. Geach and M. Black (eds. and trans.), Oxford: BlackwellFrench St., Saatsi J. (editors) (2011), The Continuum Companion to the Philosophy of Science, Bloomsbury

Companions: LondonHendry R.F., 2006, ‘Two Conceptions of the Chemical Bond’, Philosophy of Science, Vol. 75, No. 5, pp. 909-

920Hendry R.F., Needham P., Woody A.I. (2012), “Philosophy of Chemistry”, Handbook of the Philosophy of

Science Vol.6, General Editors: Gabbay D.M., Thagard P. & Woods J.IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book”), Compiled by A. D. McNaught and

A. Wilkinson, Blackwell Scientific Publications, Oxford, 1997. XML on-line corrected version: http://goldbook.iupac.org (2006-) created by M. Nic, J. Jirat, B. Kosata; updates compiled by A. Jenkins. ISBN 0-9678550-9-8. https://doi.org/10.1351/goldbook.

Ladyman J., Ross D., 2007, Every Thing Must Go, Oxford: OUPNeedham P., 2013, ‘Hydrogen bonding: Homing in on a tricky chemical concept’, Studies in History and

Philosophy of Science, 44: 51–65 Needham P., 2014, ‘The source of Chemical Bonding’, Studies in History and Philosophy of Science, 45: 1–13 Quine W.V.O, 1951, ‘Two Dogmas of Empiricism’, The Philosophical Review, 60: 20-43 Quine W.V.O, 1960, Word and Object, Cambridge: The MIT PressStroll A., 1998, Sketches of Landscapes: Philosophy by Example, Cambridge, MA: MIT PressWeisberg M., ‘Challenges to the Structural Conceptions of Chemical Bonding’, Philosophy of Science, 75

(5):932-946 (2008)

63

ISPC 2018

33. Old problems in new scientific domains: Identity and nanochemistryMariana Córdoba1 and Alfio Zambon2

1 CONICET – Universidad de Buenos Aires, Argentina2 Universidad Nacional de la Patagonia San Juan Bosco, Argentina

[email protected] - [email protected]

The traditional categories related to the ontology of physics and chemistry are challenged by the analysis of the ontological nature of nanomaterials. In this sense, we have argued that, given that an ontology of individuals and an ontology of stuff are inadequate to account for nanomaterials, a third ontological category must be considered: the category of nanoindividuals (Córdoba and Zambon 2017).

Following this line, paying especial attention to the ontological issues that arises from the consideration of nano-items, in this paper we will analyze the philosophical problem of identity, individuation and distinguishability regarding nanomaterials. The classical problem of identity is proving particularly relevant in this context, since nanomaterials traditionally characterized by means of their longitude scale are the result of the reduction of particles of a chemical substance till the nanometric scale, in which the materials shows odd properties, i.e., very different properties from the properties showed by the substance at the macro-level (Córdoba and Zambon 2017). Given this situation, it is important to wonder in which sense we are referring to the same items, being so radical the differences of properties. In order to answer to this question, we should elucidate, in the first place, which is the nature of nanomaterial’s identity: what the identity of nanomaterials rests on?

It will be necessary to evaluate the traditional theory of identity according to which identity is a reflexive relation that every single thing maintains with itself; a relation of substitutability (Noonan y Curtis 2014). On the other hand, it will be necessary to discuss if, regarding the case of nanomaterials, the principle of individuation is related to a “beginning of existence” in space and time, or if it depends on a spatio-temporal continuity (Noonan 2003). But if we are thinking of the properties of materials, we may also discuss if nanomaterial’s qualitative identity is at stake or, even more, if quantitative identity can only be defined in terms of (extreme) qualitative identity.

We will argue that this classical way of considering identity makes complete sense in order to evaluate the individuation and distinguishability of individuals (in which a kind of substratum with properties can be identified). But is it suitable to think beyond a classical ontology of individuals and properties?

We have also argued that from a basal micro-molecular level, macrochemistry and nanochemistry emerge in parallel (Zambon and Córdoba 2017). We have defended the idea that while the emergence from nanochemistry is an inter-domain emergence, the emergence of the nanolevel is an inter-domain one. If this is the case, which are the consequences regarding the identity of the items involved in every level? In other words, if we accept a picture of two chemical domains and three levels, related by different kinds of emergence, are we referring to the same or to different entities in every level?

ReferencesCórdoba, M. and Zambon, A. (2017). “How to handle nanomaterials? The reentry of individuals into the

philosophy of chemistry”. Found. Chem. 19, 185–196. Noonan, Harold. 2003. Personal Identity, second edition, London: Routledge.Noonan, Harold and Curtis, Ben. (2014) "Identity", The Stanford Encyclopedia of Philosophy (Summer 2014

Edition), Edward N. Zalta (ed.), URL = <https://plato.stanford.edu/archives/sum2014/entries/identity/>.

64

ISPC 2018

Zambon, A. and Córdoba, M. (2017). “Nanomaterials and intertheoretical relations: macro and nanochemistry as emergent levels” (communication in ISPC 2017).

65

ISPC 2018

34. A Bridge to Nowhere: Ensembles in Statistical Mechanics.Russell Helder, Georgia State University. [email protected];

For the reductionist, statistical mechanics has long been the presumptive heir of thermodynamics (e.g., Nagel 1961). By examining the ensemble approach, the means by which statistical mechanics allegedly reduces thermodynamics, I will show on the one hand how statistical mechanics itself is so elegantly built to accommodate this ‘reduction’, but on the other hand how there are serious circularities in the ensemble approach that undermine statistical mechanics’ claim to have succeeded thermodynamics. I conclude by considering statistical mechanics’ relationship to chemical thermodynamics and to chemistry as a whole.

The problem with typical microscopic and mechanical rationalizations of macroscopic thermodynamic properties given in classrooms (e.g., temperature is mean molecular kinetic energy of a gas) is that the constancy of the values of thermodynamic properties at equilibrium and even the exceptionless nature of the second law turn out to just be consequences of the large number of molecules in a macroscopic system (Oxtoby, Gills, and Campion 2012, 577). In a small enough system or on a small enough time scale, we would actually expect the mechanical property to fluctuate (Sklar 1993). An approach is needed that allows us to account for the random variations inherent to the micro level while still giving results that are totally consistent with our results on the macro level.

Statistical mechanics’ solution is, of course, statistical. To bridge the gap from the macro to the micro level, one posits an ensemble of systems, each of which is a duplicate of some macroscopic system, except that each specifies a different mechanical microstate compatible with that macroscopic system. Because all of the microstates through which the macroscopic system could possibly fluctuate are represented in the ensemble, the fluctuations mentioned before are ‘built in’ to the ensemble.

Ensembles thus offer a way to relate constant thermodynamic properties to fluctuating mechanical properties and vice-versa by the calculation of ‘ensemble averages’, which essentially give the (overwhelmingly) most probable value of the mechanical property. Gibbs’ postulate states that the ensemble average is the value of the corresponding thermodynamic property and takes the form φ ↔φ, (where φ is the ensemble average and φ is the value of the thermodynamic property). So Gibbs’ postulate, in combination with ensemble averages, resembles a bridge law by which mechanical and thermodynamic properties can be interderived.

Temperature is the real villain, which makes this entire delicate edifice constructed by statistical mechanics come tumbling down. The problem is that the Maxwell-Boltzmann function, which must describe the kinetic energy of a gas’s molecules for that gas to have a temperature at all, includes temperature (Oxtoby, Gills, and Campion, 414-17); so if temperature has been reduced to mean molecular kinetic energy of a gas, it is a circular reduction. If this circularity means that temperature is irreducible, the consequences are devastating for the reduction of thermodynamics to statistical mechanics, because temperature is a part of the calculation of every ensemble average. Therefore, ensemble averages cannot be used to reduce thermodynamics to statistical mechanics without the same circularity that precludes the reduction of temperature.

One question is why temperature is implicated in ensemble averages at all, and the answer exposes further circularity in the ‘reduction’. All ensemble averages are functions of a partition function, and all of these partition functions are themselves functions of the undetermined multiplier β. So each ensemble average is ultimately a function of β. It turns out that it is possible in principle to solve for β just as a function of the ensemble average of energy, but not in practice. One alternative method is to derive an equation relating the partial derivative of the ensemble average of energy to the partial derivative of the ensemble average of pressure. By relating that equation to the purely

66

ISPC 2018

thermodynamic equation for pressure, it can be deduced that β is equal to 1

kB T , where kB is

Boltzmann’s constant and T is temperature. Another method, which involves relating β to the purely thermodynamic equation for entropy, yields the same result. In this way, temperature is introduced into every ensemble average, because solving for β requires thermodynamic equations that contain temperature (McQuarrie 2000, 40-44).

One way or another, the cause of the first circularity, namely the mere presence of temperature, is an even deeper circularity. Ensemble averages, the means with which we interderive thermodynamic properties and mechanical properties, themselves take equations from chemical thermodynamics as premises; therefore, any ab initio derivation of thermodynamics from statistical mechanics is bound to be circular, because the means with which the derivation is effected are themselves derived, in part, from thermodynamics. To put it another way, statistical mechanics itself is apparently not fully reduced, since it relies on chemical thermodynamics, and so it cannot well serve as a reducing theory for chemical thermodyamics.

Although thermodynamics seems to not reduce to statistical mechanics without circularity, statistical mechanics is more than just a reducing theory. Rather than contributing to a reduction of chemistry to physics, statistical mechanics unifies chemistry; it explains how and why the laws of chemical thermodynamics work in virtue of one of chemistry’s other signature achievments, atomic theory. Statistical mechanics has explanatory power here first of all because it can mathematically demonstrate the link between chemical thermodynamics and atomic theory. The fact that it would be impossible for statistical mechanics to do so if it did not avail itself of purely thermodynamic resources shows that reduction is not a desirable goal here. Second, statistical mechanics has explanatory power because it makes the link between chemical thermodynamics and atomic theory intelligible by means of the ensemble approach, which makes it possible to understand and make calculations about microscopic systems that would otherwise be too complex to comprehend. By offering such an explanation of how the micro and macro levels are related in chemistry, statistical mechanics strengthens the case for chemistry as an autonomous field capable of accounting for a wide range of phenomena.

Bibliography:Albert, David Z. ‘Statistical Mechanics’. Chap. 3 in Time and Chance. Cambridge, MA: Harvard University

Press, 2009.Chang, Hasok. ‘Reductionism and the Relation Between Chemistry and Physics’. In Relocating the History of

Science: Essays in Honor of Kostas Gavroglu, edited by Theodore Arabatzis, Jürgen Renn, and Ana Simões, 193-209. Boston Studies in the Philosophy and History of Science 312. Switzerland: Springer, 2015.

Kim, Jaegwon. ‘Making Sense of Emergence’. Philosophical Studies: An International Journal for Philosophy in the Analytic Tradition 95, no. 1/2 (August 1999): 3-36.

Kitcher, Philip. ‘Explanatory Unification and the Causal Structure of the World’. In Scientific Explanation, edited by Philip Kitcher and Wesley Salmon, 410–505. Minneapolis: University of Minnesota Press, 1989.

Kripke, Saul. Naming and Necessity. Cambridge, MA: Harvard University Press, 1980.McQuarrie, Donald A. Statistical Mechanics. Sausalito, CA: University Science Books, 2000.Nagel, Ernest. ‘The Reduction of Theories’. Chap. 11 in The Structure of Science: Problems in the Logic of

Scientific Explanation. New York: Harcourt, Brace, and World, 1961.Needham, Paul. ‘Reduction and emergence: a critique of Kim’. Philosophical Studies: An International Journal

for Philosophy in the Analytic Tradition 146, no. 1 (October 2009): 93-116.Oxtoby, David W., H.P. Gillis, and Alan Campion. Principles of Modern Chemistry. Belmont, CA:

Brooks/Cole, Cengage Learning, 2012.

67

ISPC 2018

Sklar, Lawrence. Physics and chance: Philosophical issues in the foundations of statistical mechanics. Cambridge: Cambridge University Press, 1993.

Woody, Andrea. ‘Telltale Signs: What Common Explanatory Strategies in Chemistry Reveal About Explanation Itself’. Foundations of Chemistry 6, no. 1 (January 2004): 13-43.

68