Next Article in Journal
The Effect of Strain on the Aromatic Character of Infinitene
Previous Article in Journal
Recent Progress on Natural Rubber-Based Materials Containing Metallic and Metal Oxide Nanoparticles: State of the Art and Biomedical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Category Theory in Chemistry

Department of Chemistry, Faculty of Science, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan
Compounds 2023, 3(2), 334-335; https://doi.org/10.3390/compounds3020024
Submission received: 6 April 2023 / Accepted: 12 April 2023 / Published: 14 April 2023
“Mathematical chemistry” as an academic field is said to have been proposed by Weyl in the 20th century as a way of thinking that abstracts and expresses “variables”, “symbols” and “functions” [1]. Mathematics in the sense of arithmetic and statistics used to calculate chemical data is excluded here for the time being.
However, “group theory”, as well known as the mathematics that describes the three-dimensional shape of compounds, the electronic structure of atoms and molecules, and the geometric structure and symmetry of crystals, is widely used in various fields of material science [2]. Macromolecular substances, which are soft matter, are also objects of symmetry, such as surface arguments, tiling, and non-crystallographic symmetry (such as quasicrystals), even if they do not have crystal-like order [3]. Starting from regular polyhedra, surface curvature is also of mathematical interest [4]. Positive and negative curvature is also connected to rudimentary differential geometry. Regarding shapes, we often hear the word “topology” in chemistry [5]. In addition to using the properties of mathematical topology, there are occasional implications for describing three-dimensional shapes, but with the rise of supramolecular chemistry, it is clear that it is useful for describing shapes of soft molecules and assembly modes. Topology, which is the way the atoms that make up a molecule are connected, can be described and classified by graphs [6].
Furthermore, there are also examples of using more abstract set theory to express and relate chemical content [7]. By introducing a stoichiometric “chemical formula” in set theory notation and patterning it, material transformation was treated mathematically. Attempts to describe the logical structure of set theory in terms of the classification of chemical elements (elements and compounds) and their correlation with physical property values have long been known [8]. Recently, set theory has also been used to formulate ligand-target datasets in drug discovery [9]. Inspired by data science and data-driven research, it should be seen that mathematical methods were used as a means of judgment.
While the application of artificial intelligence to various sciences (such as computing, social networks, and so on) has accumulated achievements, the application of “category theory” to chemistry is regarded as being worthy of attention. Conventionally, a number of differential equations are used to describe chemical reactions (reaction rates). It is a recent mathematical achievement that their topological relationships as networks have been treated by the “cospan” of category theory as open reaction networks [10]. However, theories defining chemical reaction networks as homomorphisms between finitely generated free abelian groups have been developed, and a theorem to know the number of reactions has been discovered [11]. Furthermore, research on the relationships within chemical reaction networks has been further refined in recent years, such as simplification using topology [12].
Besides chemical reactions, composite design may be a notable application of categorical chemistry. There is the problem of determining the conditions necessary to replace one or more basic building blocks with others while preserving overall functionality, which is called the “building block replacement problem” [13]. The mathematical branch of category theory has a formal language at the heart of its methodology. Thus, a theoretical description of hierarchical material categories is possible: (1) a method of providing a mathematical tool that can replace one component with another to achieve substantially the same function; (2) how to use them to model and design seemingly different physical systems in a consistent mathematical framework. The utility of such approaches was demonstrated by using algebraic techniques to predict the specific conditions required for the exchange of building blocks.
In addition, the structure of the “ontology log” [14] is based on a branch of mathematics called category theory, which is a category that models a particular real-world situation. The authors of [15] stated that the data from sciences should be organized in such a way that this work is reusable, transferable, and comparable to the work of other scientists. In combination with category theory, which is strong in describing relationships, there are studies that also apply graph theory to chemical reactions [16] and molecular structures [17].
The mainstream of chemical research is the accumulation of experimental facts by synthesizing new compounds and reporting their properties, and this will continue to be the case. However, today, when methods such as theoretical calculation and data science are mature and available, the information derived from experimental facts and its use will also increase in importance. In the midst of this trend, it is to be expected that category theory has been attracting attention recently.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Restrepo, G.; Villaveces, J. Mathematical Thinking in Chemistry. HYLE Int. J. Philos. Chem. 2012, 18, 3–22. [Google Scholar]
  2. Matsuzawa, J. Symmetry and Group Theory. Kobunshi 2008, 57, 66–70. [Google Scholar] [CrossRef]
  3. Matsushita, Y.; Dotera, T. Symmetry in Polymers-Tricontinuous Cubic and Tiling Structures. Kobunshi 2008, 57, 71–75. [Google Scholar] [CrossRef]
  4. King, R.B. Negative Curvature Surface in Chemical Structures. J. Chem. Inf. Comput. Sci. 1998, 38, 180–188. [Google Scholar] [CrossRef]
  5. Tezuka, Y. Topological Polymer Chemistry: An Insight with Poincare into Nonlinear Macromolecular Constructions. Kobunshi 2008, 57, 81–85. [Google Scholar] [CrossRef]
  6. Rouvray, D.H. Predicting Chemistry from Topology. Sci. Am. 1986, 255, 40–47. [Google Scholar] [CrossRef] [PubMed]
  7. Bartholomay, A.F. Molecular set theory: A mathematical representation for chemical reaction mechanisms. Bull. Math. Phys. 1960, 22, 285–307. [Google Scholar] [CrossRef]
  8. Rouvray, D.H. Chemistry and Logical Structures. 2. The Role of Set Theory. Int. J. Math. Educ. Sci. Technol. 1974, 5, 173–789. [Google Scholar] [CrossRef]
  9. Maggiora, G.; Vogt, M. Set-Theoretic Formalism for Treating Ligand-Target Datasets. Molecules 2021, 26, 7419. [Google Scholar] [CrossRef] [PubMed]
  10. Baez, J.; Cho, S.; Cicala, D.; Otter, N.; de Paiva, V. Applied Category Theory in Chemistry, Computing, and Social Networks. Not. Am. Math. Soc. 2022, 69, 292–297. [Google Scholar] [CrossRef]
  11. Masavetas, K.A.; Roumpani-Kalantzopoulou, F. Categories and functors which characterize chemical reactions, their kinetics and mechanism. Math. Comput. Model. 1998, 10, 731–738. [Google Scholar] [CrossRef]
  12. Hirono, Y.; Okada, T.; Miyazaki, H.; Hidaka, Y. Structural reduction of chemical reaction networks based on topology. Phys. Rev. Res. 2021, 3, 043123. [Google Scholar] [CrossRef]
  13. Giesa, T.; Spivak, D.I.; Buehler, M.J. Category Theory Based Solution for the Building Block Replacement Problem in Materials Design. Adv. Eng. Mater. 2012, 14, 810–817. [Google Scholar] [CrossRef] [Green Version]
  14. Andersen, J.L.; Flamm, C.; Merkle, D.; Stadler, P.F. Inferring chemical reaction patterns using rule composition in graph grammars. J. Syst. Chem. 2013, 4, 4. [Google Scholar] [CrossRef] [Green Version]
  15. Spivak, D.I.; Kent, R.E. Ologs: A Categorical Framework for Knowledge Representation. PLoS ONE 2012, 7, e24274. [Google Scholar] [CrossRef]
  16. Andersen, J.L.; Flamm, C.; Merkle, D.; Stadler, P.F. Rule Composition in Graph Transformation Models of Chemical Reactions. MATCH Commun. Math. Comput. Chem. 2018, 80, 661–704. [Google Scholar]
  17. Andersen, L.J.; Flamm, C.; Merkle, D.; Stadler, P.F. An intermediate level of abstraction for computational systems chemistry. Phylos. Trans. A 2017, 375, 20160354. [Google Scholar] [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Akitsu, T. Category Theory in Chemistry. Compounds 2023, 3, 334-335. https://doi.org/10.3390/compounds3020024

AMA Style

Akitsu T. Category Theory in Chemistry. Compounds. 2023; 3(2):334-335. https://doi.org/10.3390/compounds3020024

Chicago/Turabian Style

Akitsu, Takashiro. 2023. "Category Theory in Chemistry" Compounds 3, no. 2: 334-335. https://doi.org/10.3390/compounds3020024

Article Metrics

Back to TopTop