Skip to main content

Béziau’s Contributions to the Logical Geometry of Modalities and Quantifiers

  • Chapter
The Road to Universal Logic

Part of the book series: Studies in Universal Logic ((SUL))

Abstract

The aim of this paper is to discuss and extend some of Béziau’s (published and unpublished) results on the logical geometry of the modal logic S5 and the subjective quantifiers many and few. After reviewing some of the basic notions of logical geometry, we discuss Béziau’s work on visualising the Aristotelian relations in S5 by means of two- and three-dimensional diagrams, such as hexagons and a stellar rhombic dodecahedron. We then argue that Béziau’s analysis is incomplete, and show that it can be completed by considering another three-dimensional Aristotelian diagram, viz. a rhombic dodecahedron. Next, we discuss Béziau’s proposal to transpose his results on the logical geometry of the modal logic S5 to that of the subjective quantifiers many and few. Finally, we propose an alternative analysis of many and few, and compare it with that of Béziau’s. While the two analyses seem to fare equally well from a strictly logical perspective, we argue that the new analysis is more in line with certain linguistic desiderata.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    It is well-known that ¬(¬φ∧¬ψ) is equivalent to φψ, but we choose to stick with the former notation, because it more clearly expresses the idea of φ and ψ being false together.

  2. 2.

    The system S is assumed to have connectives expressing classical negation (¬), conjunction (∧) and implication (→), and a model-theoretic semantics (⊨).

  3. 3.

    As limiting cases, the non-contingent bitstrings 0000 and 1111 can be called level 0 (L0) and level 4 (L4), respectively.

  4. 4.

    Formally, a diagram or set of formulas is said to be Boolean closed iff whenever it contains formulas φ,ψ, it also contains their contingent Boolean combinations (¬φ, φψ, φψ). Note that this definition is restricted to contingent Boolean combinations, so if □p and ¬□p occur in a diagram, then it is not required for the diagram to be Boolean closed that it also contains the contradiction □p∧¬□p and the tautology □p∨¬□p.

  5. 5.

    The process described above works not only for the S5 square (Fig. 2(b)), but also for the bitstring square (Fig. 2(c)). We thus obtain three bitstring JSB hexagons, the first one of which was already shown in Fig. 3(b). The two new ones are displayed simultaneously with their S5 counterparts in Fig. 4. For reasons of space, we will in the remainder of this paper no longer distinguish between S5 diagrams and bitstring diagrams, and decorate the Aristotelian diagrams sometimes with concrete formulas, sometimes with bitstrings, and sometimes with both simultaneously.

  6. 6.

    These hexagons also show that, from the perspective of subalternation, classical negation occupies a position that is intermediate between paracomplete and paraconsistent negation. In Fig. 4(b), the classical negation (¬p) is entailed by the paracomplete negation (¬◊p), while in Fig. 4(c), it entails the paraconsistent negation (¬□p).

  7. 7.

    Since there are 3 hexagons and each hexagon contains 6 formulas, one might expect the total number to be 3×6=18 formulas. However, this calculation ignores the fact that certain formulas occur in two distinct hexagons. In particular, the formulas ◊p and ¬◊p occur in hexagons (a) and (b), the formulas □p and ¬□p occur in hexagons (a) and (c), and the formulas p and ¬p occur in hexagons (b) and (c).

  8. 8.

    Somewhat confusingly, Béziau [1] talks about the ‘stellar dodecahedron’ instead of the ‘stellar rhombic dodecahedron’, thereby suggesting that the solid he had in mind (but never actually drew) is the stellation of the ‘ordinary’ pentagonal (i.e. Platonic) dodecahedron, rather than that of a rhombic dodecahedron. This confusion resurfaces in Moretti’s remarks that Béziau’s solid is “obtained by constructing a pentagonal pyramid or spike over each of the 12 pentagonal faces of a dodecahedron” [19, p. 75, our emphases]. Accordingly, the figure given by Moretti [19, p. 76] shows the stellation of a pentagonal dodecahedron, rather than that of a rhombic dodecahedron (although he still calls it ‘Escher’s solid’ and attributes it to Béziau). In a more recent paper, Béziau does provide a figure of the stellar rhombic dodecahedron (without its S5-decoration) [5, p. 13], but still calls it the ‘stellar dodecahedron’.

  9. 9.

    Also recall Footnote 7.

  10. 10.

    Note, trivially perhaps, that these two formulas do not occur together in any of the four JSB hexagons considered above. After all, if a diagram contains these two formulas, then it cannot be Boolean closed.

  11. 11.

    Pellissier [20] and Moretti [18] distinguish between a strong and a weak kind of JSB hexagon (the strong kind is the kind we have considered up till now), and note that the 14 formulas not only yield 6 strong JSB hexagons, but also 4 weak JSB hexagons.

  12. 12.

    The differences between these three visualisations are discussed in more detail in [27, Sect. 2].

  13. 13.

    Going beyond the rhombic dodecahedron would thus require us to introduce bitstrings of length 5. This can certainly be done, but the Aristotelian diagrams will become exponentially larger. For example, as far as Boolean closed diagrams are concerned, we move from the rhombic dodecahedron (which has 24−2=14 vertices) to a diagram that has 25−2=30 vertices. (Recall Footnotes 3 and 4.)

  14. 14.

    We will henceforth add a ‘1’ in subscript to the expressions many and few to refer to the semantic interpretation they receive in Béziau’s analysis. Similarly, in the next sections, we will add a ‘2’ in subscript to refer to our alternative analysis.

  15. 15.

    In terms of bitstrings, the one-sided reading corresponds to a bitstring that has one transition in bit values, i.e. from 1 to 0 or vice versa (e.g. 1110), whereas the two-sided reading corresponds to a bitstring having two transitions in bit values (e.g. 0110).

  16. 16.

    This distinction also applies to the modal operators, where one-sided possibility (◊p) is compatible with necessity, but two-sided possibility (◊p∧¬□p, usually called ‘contingency’) is not.

  17. 17.

    Strictly speaking, these two quantifier expressions did not feature in Béziau’s presentation [3], but as was argued in Sect. 4, they can straightforwardly be added by taking the Boolean closure of the original set of 12 expressions.

  18. 18.

    We only look at the minimal number of binary connectives, because every semantic value (bitstring) can be expressed in a number of syntactically different ways. For example, the bitstring 0010 can be expressed as ¬p∧◊p [0011 ∧ 1110] with one binary connective, but also as ¬p∧(□p∨(¬p∧◊p)) [0011 ∧ 1010] with three binary connectives.

  19. 19.

    From a linguistic point of view, the semantics of many and few is notoriously complex. For example, Keenan writes: “we shall largely exclude many and few from the generalisations we propose since our judgements regarding their interpretations are variable and often unclear” [16, pp. 47–48].

  20. 20.

    If we were to work with bitstrings of length 3, ignoring the subjective quantifiers and focussing on the six quantifier expressions in the top part of Table 2, some would be the L1 bitstring 010, whereas at least one would be the L2 bitstring 110.

  21. 21.

    More specifically, three types of mismatches can be distinguished: (i) there is both a semantic and a lexical arrow but they point in opposite directions—e.g. between 1100 and 0100, (ii) there is a semantic arrow but no lexical arrow at all—e.g. between 0100 and 1101, and (iii) there is a lexical arrow but no semantic arrow at all—e.g. between 1100 and 0101.

  22. 22.

    Notice, incidentally, that the difference in subscripts between Béziau’s many 1 and our many 2 nicely reflects this contrast between the one-sided and the two-sided readings.

  23. 23.

    Recall Footnote 15.

  24. 24.

    Using simple propositional reasoning, the expression many if not all can be shown to be equivalent to the expression many or all. Intuitions differ as to whether an expression of the form p if not q should be read as the conditional ¬pq or rather as ¬qp, but both readings are equivalent to the disjunction pq.

  25. 25.

    Since any can be seen as the negation of no, the expression few if any is semantically equivalent to few if not no, and can thus also be shown to boil down to the disjunction few or no (recall Footnote 24). Additional linguistic evidence for this equivalence comes from translational equivalents such as Dutch weinig of geen and French peu ou pas. Furthermore, even in English the disjunctive semantics of few or no is lexicalised, albeit only for abstract and mass nouns, viz. as little or no.

  26. 26.

    Note that there is no horizontal semantic arrow between the bitstrings 1100 and 0101, since neither of them entails the other one. There is no horizontal lexical arrow between them either, since their corresponding quantifier expressions have the same degree of lexical complexity (viz. a single Boolean operator). Because of this twofold absence, the correlation between semantic and lexical complexity is preserved in this case as well. Similar remarks apply to the case of 1010 and 0011.

  27. 27.

    It should be emphasised that we take the term ‘Buridan octagon’ to refer to (the visualisation of) a certain constellation of Aristotelian relations, rather than to the particular form or content matter of the formulas involved. As a matter of fact, Buridan’s own use of his octagon was in describing the logical geometry of first-order modal logic, rather than that of the subjective quantifiers.

  28. 28.

    I.e. the set is essentially a Boolean algebra with its top and bottom elements left out (recall Footnotes 3 and 4).

References

  1. Béziau, J.-Y.: New light on the square of oppositions and its nameless corner. Log. Invest. 10, 218–232 (2003)

    Google Scholar 

  2. Béziau, J.-Y.: Paraconsistent logic from a modal viewpoint. J. Appl. Log. 3, 7–14 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  3. Béziau, J.-Y.: Some—geometrical representations of the quantifiers. Talk presented at Logic Now and Then (Brussels) (2008)

    Google Scholar 

  4. Béziau, J.-Y.: The power of the hexagon. Log. Univers. 6, 1–43 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  5. Béziau, J.-Y.: The new rising of the square of opposition. In: Béziau, J.-Y., Jacquette, D. (eds.) Around and Beyond the Square of Opposition, pp. 6–23. Springer/Birkhäuser, Berlin (2012)

    Chapter  Google Scholar 

  6. Béziau, J.-Y.: The metalogical hexagon of opposition. Log. Invest. 10, 111–122 (2013)

    Google Scholar 

  7. Blanché, R.: Quantity, modality, and other kindred systems of categories. Mind 61, 369–375 (1952)

    Article  Google Scholar 

  8. Blanché, R.: Structures Intellectuelles. Essai sur l’Organisation Systématique des Concepts. Vrin, Paris (1969)

    Google Scholar 

  9. Coxeter, H.: Regular Polytopes. Courier Dover, New York (1973)

    Google Scholar 

  10. Demey, L.: Structures of oppositions for public announcement logic. In: Béziau, J.-Y., Jacquette, D. (eds.) Around and Beyond the Square of Opposition, pp. 313–339. Springer/Birkhäuser, Berlin (2012)

    Chapter  Google Scholar 

  11. Demey, L.: Believing in logic and philosophy. PhD thesis, KU Leuven (2014)

    Google Scholar 

  12. Demey, L., Smessaert, H.: The relationship between Aristotelian and Hasse diagrams. In: Dwyer, T., Purchase, H., Delaney, A. (eds.) Diagrammatic Representation and Inference. Lecture Notes in Computer Science, vol. 8578, pp. 213–227. Springer, Berlin (2014)

    Chapter  Google Scholar 

  13. Horn, L.: A Natural History of Negation. University of Chicago Press, Chicago (1989)

    Google Scholar 

  14. Hughes, G.: The modal logic of John Buridan. In: Corsi, G., Mangione, C., Mugnai, M. (eds.) Atti del Convegno Internzionale di Storia della Logica, le Teorie delle Modalità, pp. 93–111. CLUEB (1987)

    Google Scholar 

  15. Jacoby, P.: A triangle of opposites for types of propositions in Aristotelian logic. New Scholast. 24, 32–56 (1950)

    Article  Google Scholar 

  16. Keenan, E.: The semantics of determiners. In: Lappin, S. (ed.) The Handbook of Contemporary Semantic Theory, pp. 41–63. Blackwell, Oxford (1997)

    Google Scholar 

  17. Luke, D.: Stellations of the rhombic dodecahedron. Math. Gaz. 41, 189–194 (1957)

    Article  MathSciNet  MATH  Google Scholar 

  18. Moretti, A.: The geometry of logical opposition. PhD thesis, University of Neuchâtel (2009)

    Google Scholar 

  19. Moretti, A.: The critics of paraconsistency and of many-valuedness and the geometry of oppositions. Log. Log. Philos. 19, 63–94 (2010)

    MathSciNet  MATH  Google Scholar 

  20. Pellissier, R.: “Setting” n-opposition. Log. Univers. 2, 235–263 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  21. Read, S.: John Buridan’s theory of consequence and his octagons of opposition. In: Béziau, J.-Y., Jacquette, D. (eds.) Around and Beyond the Square of Opposition, pp. 93–110. Springer/Birkhäuser, Berlin (2012)

    Chapter  Google Scholar 

  22. Sauriol, P.: Remarques sur la théorie de l’hexagone logique de Blanché. Dialogue 7, 374–390 (1968)

    Article  Google Scholar 

  23. Sesmat, A.: Logique II. Les Raisonnements. Hermann, Paris (1951)

    Google Scholar 

  24. Seuren, P.: The Logic of Language. Language from Within, vol. II. Oxford University Press, Oxford (2010)

    Google Scholar 

  25. Smessaert, H.: On the 3D visualisation of logical relations. Log. Univers. 3, 303–332 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  26. Smessaert, H.: Boolean differences between two hexagonal extensions of the logical square of oppositions. In: Cox, P., Plimmer, B., Rodgers, P. (eds.) Diagrammatic Representation and Inference. Lecture Notes in Computer Science, vol. 7352, pp. 193–199. Springer, Berlin (2012)

    Chapter  Google Scholar 

  27. Smessaert, H., Demey, L.: Logical and geometrical complementarities between Aristotelian diagrams. In: Dwyer, T., Purchase, H., Delaney, A. (eds.) Diagrammatic Representation and Inference. Lecture Notes in Computer Science, vol. 8578, pp. 246–260. Springer, Berlin (2014)

    Chapter  Google Scholar 

  28. Smessaert, H., Demey, L.: The unreasonable usefulness of bitstrings in logical geometry. Talk presented at the Fourth World Congress on the Square of Opposition (Vatican) (2014)

    Google Scholar 

  29. Smessaert, H., Demey, L.: La géometrie logique du dodécaèdre rhombique des oppositions. In: Ben Aziza, H., Chatti, S. (eds.) Le Carré et ses Extensions: Aspects Théoriques, Pratiques et Historiques. Presses Universitaires, Tunis (2014)

    Google Scholar 

  30. Smessaert, H., Demey, L.: The Logical Geometry of the Aristotelian Rhombic Dodecahedron. Book Manuscript (2015)

    Google Scholar 

Download references

Acknowledgement

We thank Dany Jaspers, Alessio Moretti, Fabien Schang and Margaux Smets for their comments on earlier versions of this paper. The second author gratefully acknowledges financial support from the Research Foundation—Flanders (FWO).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hans Smessaert .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Smessaert, H., Demey, L. (2015). Béziau’s Contributions to the Logical Geometry of Modalities and Quantifiers. In: Koslow, A., Buchsbaum, A. (eds) The Road to Universal Logic. Studies in Universal Logic. Birkhäuser, Cham. https://doi.org/10.1007/978-3-319-10193-4_23

Download citation

Publish with us

Policies and ethics