Skip to main content
Log in

Propositional discourse logic

  • Published:
Synthese Aims and scope Submit manuscript

Abstract

A novel normal form for propositional theories underlies the logic pdl, which captures some essential features of natural discourse, independent from any particular subject matter and related only to its referential structure. In particular, pdlallows to distinguish vicious circularity from the innocent one, and to reason in the presence of inconsistency using a minimal number of extraneous assumptions, beyond the classical ones. Several, formally equivalent decision problems are identified as potential applications: non-paradoxical character of discourses, admissibility of arguments in argumentation networks, propositional satisfiability, and the existence of kernels of directed graphs. Directed graphs provide the basis for the semantics of pdl and the paper concludes by an overview of relevant graph-theoretical results and their applications in diagnosing paradoxical character of natural discourses.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. Of course, we are not saying that this is what facts are, only that they can be treated in this way, without impairing correctness of the formal model. Our model works unchanged also when no such facts are available.

  2. When unfolded in time as a sequence of consecutive statements, such a holistic network of mutual dependencies gives rise to anaphoric and cataphoric references, yielding the non-monotonic character of the discourse. But since non-monotonicity appears thus only as a special, temporal view of mutual dependencies, we will not devote it separate treatment.

  3. We do not claim that decision determining the borders of the actually relevant totality is as arbitrary as suggested by the examples. But we do mean that in many cases, the actual totalities have unsharp borders, which may be adjusted in different ways.

  4. Many results that will be presented hold also for infinite discourses (theories) and infinitary logic (allowing infinite \(I_x\)’s), but we are addressing primarily the finite and finitary case.

  5. One can think of a propositional letter appearing on the left of an equivalence as naming the complex formula that appears on the right. The equivalences become then instances of Tarski’s T-schema, formulated in propositional logic.

  6. In Prakken and Vreeswijk (2002), p. 238, the authors note disappointingly little attention paid to the self-defeating arguments in the argumentation literature. Although psychologically very different from incoherent totalities of arguments, like \(a-b-c\), their formal role and effects are entirely analogous.

  7. Direction of the edges may be reversed, provided that it is done consistently throughout the whole development. Argumentation networks, or various derivative concepts, are typically formulated in the literature with edges going in the opposite direction.

       A particular consequence of the representation (2.1) and this graphical counterpart is that statements, the actual carriers of truth-values, correspond to the sentence tokens and not types. Saying the same sentence (type) at two different points may turn it into different statements. A token, or a statement, is in this context just a point in a network of cross-references, a node of the discourse graph.

  8. The equivalence of kernels and non-paradoxical discourses was first noted in Cook (2004), while of kernels and stable extensions of argumentation networks in Doutre (2002).

  9. In argumentation theory, this is referred to as credulous acceptance, inroduced in Dung (1995), whereby an argument is designated as acceptable when there is a local admissible set containing it. Our work demonstrates that this notion can be looked at as classical satisfiability of a special type of subdiscourse containing the argument, and that both these viewpoints are captured by the technical notion of a local kernel, which has been studied by graph-theorists since the 70ties Neumann-Lara (1971).

  10. All other connectives can be defined from \(\{{\lnot ,\wedge }\}\) in the classical manner. This choice does not in any way limit the expressivity of the language, and is made only for establishing an easy connection to graphs.

  11. This generalizes the notion of admissibility of arguments in argumentation theory, which considers only \(\Gamma \) consisting of a single propositional variable.

  12. The corresponding idea in Kripke’s theory of truth from Kripke (1975) would be to take as paradox only those sentences which are neither true nor false in any fixed-point. We do not claim that this is appropriate for a general theory of truth, which is not our object.

  13. For displaying proofs, it is convenient to write a graph as a list of sinks and edges, e.g., \(\langle x,x\rangle \) is the liar graph, while \(x\) the same graph with the loop removed. Some redundancy in notation may ease readability, e.g., \(a,\langle b,a\rangle \) and \(\langle b,a\rangle \) denote the same graph \(b\rightarrow a\).

  14. It is not clear whether the sequent calculus for \({{\L }3}\) presented in Béziau (1998) could be used in a similar way. This seems rather unlikely, in particular, as it has multiple rules with the same principal formula.

  15. Curry’s paradox may be negation-free only if \(x \leftrightarrow (x\rightarrow y)\) does not abbreviate \(x \leftrightarrow \lnot (x\wedge \lnot y)\). In our case, it is exactly what it does, as the arrow \(\rightarrow \) on the right is defined as in strong Kleene logic.

  16. This may require a qualification. On the one hand, purely logical means are inherently inadequate, since the language of classical logic is designed exactly so as to prevent any direct self-reference. Typically, one is forced to step beyond first-order logic and apply intricate Gödelizations in order to express something as simple as the liar (as a matter of fact, only something which merely reminds of the liar). On the other hand, one may take a more semantic approach. A good example is the use of non-well-founded sets, that is, eventually arbitrary graphs in Barwise and Moss (1996), as the semantic basis for modeling circularity of discourses. Accepting the anti-foundation axiom is, however, a dramatic step, bringing us out of classical set theory. It may happen that a general solution to paradoxes of all kinds might require such a fundamental departure. We prefer to avoid it as long as possible, in particular, when it suffices to represent circularity in classical set theory and analyze it using essentially classical logic.

  17. This excludes any “infinite cycle” and makes Yablo’s paradox non-circular. (Infinite cycles can be introduced into infinite graphs, by topological means, using completions of infinite rays. They seem to have no relation to infinitary paradoxes, though, and Yablo remains acyclic also when such cycles are allowed.) Yablo’s circularity, suggested in Priest (1997), concerned its finitary formulation but not its actual referential structure. This is the relevant structure, captured by our graphs. On the other hand, since every person in the Yablo’s path says (*) “All my followers are lying”, Priests suggests that “one individuates the thought in such a way that all the people are thinking the same thought”. This is certainly possible, but asks us to ignore the crucial structure of the reference involved: the thought of the \(n\)-th person includes the \((n+1)\)-th person, while the thought of the \((n+1)\)-th person does not. As observed in footnote 7, one can plausibly ask also here to individuate the thought (*)—if one wants to insist on the singular form—at the level of tokens and not of its type. The isomorphism of every tail of the Yablo graph with the whole graph does not mean that they are identical.

  18. Incidentally, this form of discourse (a ring of size \(n \ge 3\) where each \(x_i\) claims falsity of \(x_{i+1}\) and \(x_{i-2}\)) is paradoxical even when the ring is even, but this follows from a particular argument concerning the impossibility of breaking the involved odd 3-cycles. (To see this, assume a solution, pick a node \(x_i\) that is \(\mathbf 1 \) and look for any \(\mathbf 1 \)-successor of its \(\mathbf 0 \)-successor \(x_{i+1}\) on the ring. No such can exist, since all successors of \(x_{i+1}\) are in- or out-neighbours of \(x_i\).)

  19. One can propose finer criteria for distinguishing statements, so that (1)–(3) come out as different, even to the point where (3) becomes a No-No paradox. But as far as their truth-conditions under the classical semantics are concerned, there is as little problem with their equivalence—and the absence of paradox—as with the fact that among two persons accusing each other of lying, only one is telling the truth, the symmetry of appearances notwithstanding. Which one it is, may vary between various tokens of truth-teller.

  20. One can certainly envision more detailed distinctions within such a non-classical semantics, separating paradoxical and non-paradoxical discourses, but such distinctions remain to be proposed and investigated. Given that argumentation semantics try to determine successful arguments in every possible context, they will tend to disregard any distinction between proper and malfunctioning discourses, founding the concept of paradox.

  21. It can also be seen as a successful continuation of the attempts to determine where to place the third value in pointer semantics for circular discourses, arising from Gaifman (1988).

  22. SCC-recursiveness must not be misunderstood for such rules. It allows for an iterative construction (here, of kernels) only between strongly connected components, but tells one nothing how to obtain kernels of any such component.

  23. The only exception is the lack of the explicit truth-operator in pdl. Possibility of including such an operator remains to be investigated.

References

  • Baroni, P., & Giacomin, M. (2003). Solving semantic problems with odd-length cycles in argumentation. In T. D. Nielsen & N. L. Zhang (Eds.), Symbolic and quantitative approaches to reasoning with uncertainty. Lecture notes in computer science (Vol. 2711, pp. 440–451). Berlin: Springer.

    Chapter  Google Scholar 

  • Baroni, P., Giacomin, M., & Guida, G. (2005). Scc-recursiveness: A general schema for argumentation semantics. Artificial Intelligence, 168(1), 162–210.

    Article  Google Scholar 

  • Barwise, J., & Moss, L. (1996). Vicious circles: On the mathematics of non-wellfounded phenomena. Stanford: CSLI.

    Google Scholar 

  • Beall, J. C. (2001). Is Yablo’s paradox non-circular? Analysis, 63(1), 176–187.

    Google Scholar 

  • Bezem, M., Grabmayer, C., & Walicki, M. (2012). Expressive power of digraph solvability. Annals of Pure and Applied Logic, 162(3), 200–212.

    Google Scholar 

  • Béziau, J.-Y. (1998). A sequent calculus for Łukasiewicz’s three-valued logic based on suszko’s bivalent semantics. Bulletin of the Section of Logic, 28(2), 89–97.

    Google Scholar 

  • Boros, E., & Gurvich, V. (2006). Perfect graphs, kernels and cooperative games. Discrete Mathematics, 306, 2336–2354.

    Article  Google Scholar 

  • Chvátal, V. (1973). On the computational complexity of finding a kernel. Technical Report CRM-300, Centre de Recherches Mathématiques, Univeristé de Montréal. http://users.encs.concordia.ca/~chvatal

  • Cook, R. (2004). Patterns of paradox. The Journal of Symbolic Logic, 69(3), 767–774.

    Article  Google Scholar 

  • Cook, R. (2006). There are non-circular paradoxes (but Yablo’s isn’t one of them). The Monist, 89, 118–149.

    Article  Google Scholar 

  • Creignou, N. (1995). The class of problems that are linearly equivalent to satisfiability or a uniform method for proving np-completeness. Theoretical Computer Science, 145, 111–145.

    Article  Google Scholar 

  • Dimopoulos, Y., & Magirou, V. (1994). A graph theoretic approach to default logic. Information and Computation, 112, 239–256.

    Article  Google Scholar 

  • Dimopoulos, Y., Magirou, V., & Papadimitriou, C. H. (1997). On kernels, defaults and even graphs. Annals of Mathematics and Artificial Intelligence, 20, 1–12.

    Article  Google Scholar 

  • Dimopoulos, Y., & Torres, A. (1996). Graph theoretical structures in logic programs and default theories. Theoretical Computer Science, 170(1–2), 209–244.

    Article  Google Scholar 

  • Doutre, S. (2002). Autour de la sématique préférée des systèmes d’argumentation. PhD thesis, Université Paul Sabatier, Toulouse.

  • Duchet, P. (1980). Graphes noyau-parfaits, II. Annals of Discrete Mathematics, 9, 93–101.

    Article  Google Scholar 

  • Duchet, P., & Meyniel, H. (1983). Une généralisation du théorème de Richardson sur l’existence de noyaux dans les graphes orientés. Discrete Mathematics, 43(1), 21–27.

    Article  Google Scholar 

  • Dung, P. M. (1995). On the acceptability of arguments and its fundamental role in nonmonotonic reasoning, logic programming and \(n\)-person games. Artificial Intelligence, 77, 321–357.

    Article  Google Scholar 

  • Gabbay, D. (2009). Modal provability foundations for argumentation networks. Studia Logica, 93(2–3), 181–198.

    Article  Google Scholar 

  • Gabbay, D., & van der Torre, L. (Eds.) (2009). New ideas in argumentation theory [Special issue]. Studia Logica, 93(2–3) 357–381.

    Google Scholar 

  • Gaifman, H. (1988). Operational pointer semantics: Solution to self-referential puzzles. In M. Vardi (Ed.), Theoretical aspects of reasoning about knowledge (pp. 43–59). Los Atlos, CA: Morgan Kaufman.

  • Gaifman, H. (1992). Pointers to truth. The Journal of Philosophy, 89(5), 223–261.

    Article  Google Scholar 

  • Gaifman, H. (2000). Pointers to propositions. In A. Chapuis & A. Gupta (Eds.), Circularity, definition and truth (pp. 79–121). New Delhi: Indian Council of Philosophical Research.

    Google Scholar 

  • Galeana-Sánchez, H., & Neumann-Lara, V. (1984). On kernels and semikernels of digraphs. Discrete Mathematics, 48(1), 67–76.

    Article  Google Scholar 

  • Grossi, D. (2010a). Argumentation in the view of modal logic. In P. McBurney, I. Rahwan, & S. Parsons (Eds.), ArgMAS. Lecture notes in computer science (Vol. 6614, pp. 190–208). Berlin-Heidelberg: Springer.

  • Grossi, D. (2010b). On the logic of argumentation theory. In W. van der Hoek, G. A. Kaminka, Y. Lespérance, M. Luck, & S. Sen (Eds.), 9th AAMAS (pp. 409–416). Toronto, Canada.

  • Kripke, S. (1975). Outline of a theory of truth. The Journal of Philosophy, 72(19), 690–716.

    Article  Google Scholar 

  • Neumann-Lara, V. (1971). Seminúcleos de una digráfica. Technical report Anales del Instituto de Matemáticas II, Universidad Nacional Autónoma México.

  • Prakken, H., & Vreeswijk, G. (2002). Logics for deafeasible argumentation. In D. Gabbay & F. Guenthner (Eds.), Handbook of philosophical logic (Vol. 4, pp. 219–318). Dordrecht: Kluwer Academic Publishers.

    Google Scholar 

  • Priest, G. (1997). Yablo’s paradox. Analysis, 57, 236–242.

    Article  Google Scholar 

  • Richardson, M. (1953). Solutions of irreflexive relations. The Annals of Mathematics, Second Series, 58(3), 573–590.

    Article  Google Scholar 

  • Sorensen, R. (1998). Yablo’s paradox and kindred infinite liars. Mind, 107, 137–155.

    Article  Google Scholar 

  • von Neumann, J., & Morgenstern, O. (1947). Theory of games and economic behavior. Princeton: Princeton University Press.

    Google Scholar 

  • Walicki, M. (2009). Reference, paradoxes and truth. Synthese, 171, 195–226.

    Article  Google Scholar 

  • Walicki, M., & Dyrkolbotn, S. (2012). Finding kernels or solving SAT. Journal of Discrete Algorithms, 10, 146–164.

    Google Scholar 

  • Yablo, S. (1993). Paradox without self-reference. Analysis, 53(4), 251–252.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michał Walicki.

Appendix: Proofs

Appendix: Proofs

Proposition 3.4

Given a graph \(\mathsf{G}\), we have:

  1. (a)

    If \(L \in Lk(\mathsf{G})\) then \(\alpha _{\overline{L}} \models _{{{\L }}}\mathcal{D}(\mathsf{G})\) and;

  2. (b)

    If \(\alpha \models _{{{\L }}}\mathcal{D}(\mathsf{G})\) for \(\alpha : G{\rightarrow }\{\mathbf{1 , \mathbf 0 , \bot }\}\) then \(\overline{\varnothing }\subseteq \alpha ^\mathbf 1 \in Lk(\mathsf{G})\)

Proof

  1. (a)

    Assume \(L \in Lk(\mathsf{G})\) and consider arbitrary \(x \leftrightarrow \bigwedge _{y \in E(x)} \lnot y \in \mathcal{D}(\mathsf{G})\). Assume towards contradiction that \(\alpha _{\overline{L}} \not \models _{{{\L }}}x \leftrightarrow \bigwedge _{y \in E(x)} \lnot y\). We let \(\overline{\alpha }_{\overline{L}}\) denote the evaluation of complex formulae obtained from \(\alpha _{\overline{L}}\) according to tables (3.3). Then we have \(\alpha _{\overline{L}}(x) \not = \overline{\alpha }_{\overline{L}}(\bigwedge _{y \in E(x)} \lnot y)\). If \(\alpha _{\overline{L}}(x) = \mathbf 1 \) this inequality means that there is one \(y \in E(x)\) such that \(\alpha _{\overline{L}}(y) \in \{\mathbf{1 ,\bot }\}\), impossible by the fact that \(\overline{L}\) is a local kernel (which requires, for all \(y \in E(x)\), \(y \in E^\smallsmile \!(\overline{L})\), i.e. \(\alpha _{\overline{L}}(y) = \mathbf 0 \)). If \(\alpha _{\overline{L}}(x) = \mathbf 0 \) we must then have, for every \(y \in E(x)\), \(\alpha _{\overline{L}}(y) \in \{\mathbf{0 ,\bot }\}\) but this is also ruled out by the fact that \(\overline{L}\) is a local kernel (which requires existence of some \(y \in E(x)\) such that \(\alpha _{\overline{L}}(y) = \mathbf 1 \)). The last possibility is that \(\alpha _{\overline{L}}(x) = \bot \) in which case there are two possibilities. (1) We have some \(y \in E(x)\) such that \(\alpha _{\overline{L}}(y) = \mathbf 1 \). This contradicts \(x \not \in E^\smallsmile \!(L)\) (required since we have \(\alpha _{\overline{L}}(x) = \bot \)). (2) For all \(y \in E(x)\) we have \(\alpha _{\overline{L}}(y) = \mathbf 0 \). This means \(x \in sinks(\mathsf{G}\setminus (\overline{L} \cup E^\smallsmile \!(\overline{L}))\), impossible by Definition 2.2 of \(\overline{L}\).

  2. (b)

    Assume \(\alpha \models _{{{\L }}}\mathcal{D}(\mathsf{G})\). We show that \(\alpha ^\mathbf 1 \) is a local kernel. We show first that \(\alpha ^\mathbf 1 \) is independent. Assume towards contradiction that it is not. Then there are \(x,y \in \alpha ^\mathbf 1 \) with \(y \in E(x)\). So we have \(\alpha (x) = \alpha (y) = \mathbf 1 \) and from inspecting the tables (3.3) we see that \(\overline{\alpha }(\bigwedge _{y \in E(x)} \lnot y) = \mathbf 0 \). In particular, we have \(\alpha (x) \not = \overline{\alpha }(\bigwedge _{y \in E(x)} \lnot y)\), contrary to hypothesis. Assume towards contradiction that \(\alpha ^\mathbf 1 \) is not locally absorbing. Then there is some \(x \in \alpha ^\mathbf 1 \) with \(y \in E(x)\) such that \(E(y) \cap \alpha ^\mathbf 1 = \varnothing \). Since \(\alpha (z) \not = \mathbf 1 \) for all \(z \in E(x)\), by \(\alpha ^\mathbf 1 \) being independent, this means that \(\alpha (y) = \bot \) and that \(\overline{\alpha }(\bigwedge _{y \in E(x)} \lnot y) = \bot \not = \alpha (x) = \mathbf 1 \), contrary to hypothesis. To show that \(\overline{\varnothing }\subseteq \alpha ^\mathbf{1 }\) is a simple proof by induction over definition 2.2. For the basis, if \(x \in sinks(\mathsf{G})\), i.e. \(x \in \varnothing _1\), then \(x \leftrightarrow \mathbf 1 \in \mathcal{D}(\mathsf{G})\). Then is is clear that \(x \in \alpha ^\mathbf 1 \). The inductive step is also trivial.

\(\square \)

Theorem 3.5

\(\models \langle \Gamma : \mathsf{G}\rangle \) iff there is some \(\alpha : G{\rightarrow }\{\mathbf{1 , \mathbf 0 , \bot }\}\) such that \(\alpha \models _{{{\L }}}\mathcal{D}(\mathsf{G})\) and \(\alpha \models _{{{\L }}}\Gamma \).

Proof

\(\Rightarrow \)) Assume that \(\models \langle \Gamma : \mathsf{G}\rangle \). Then by Definition 3.1 there is some local kernel \(L \in Lk(\mathsf{G})\) such that \(L \models \langle \Gamma : \mathsf{G}\rangle \). We have from Proposition 3.4.(a) that \(\alpha _{\overline{L}} \models _{{{\L }}}\mathcal{D}(\mathsf{G})\). We show \(\alpha _{\overline{L}} \models _{{{\L }}}\Gamma \) by induction on the complexity of \(\Gamma \). We take its complexity to be the sum of the complexity of its formulae divided by \(|\Gamma |\). The basis is for \(\Gamma \) a collection of literals. Then \(\Gamma ^+ \subseteq L\) and \(\Gamma ^- \subseteq E^\smallsmile \!(L)\), so for all \(x \in \Gamma ^+\) we have \(\alpha _{\overline{L}}(x) = \mathbf 1 \) and for all \(y \in \Gamma ^-\) we have \(\alpha _{\overline{L}}(y) = \mathbf 0 \) (remember that \(L \subseteq \overline{L}\)). It follows from inspecting tables (3.3) that \(\alpha _{\overline{L}} \models \Gamma \). The inductive steps are easy. For instance, if \(\models \langle \Gamma : \mathsf{G}\rangle \) and there is \(\lnot \lnot A \in \Gamma \) that has maximal complexity among formulae of \(\Gamma \), then we form \(\Gamma ^{\prime }\) which is like \(\Gamma \) except that \(A\) replaces \(\lnot \lnot A\). \(\Gamma ^{\prime }\) has lower complexity than \(\Gamma \) and, obviously from Definition 3.1, \(\models \langle \Gamma ^{\prime } : \mathsf{G}\rangle \). So by IH we get \(\models _{{{\L }}}\Gamma ^{\prime }\). Consulting tables (3.3) we see that this gives us \(\models _{{{\L }}}\Gamma \) so we are done. The cases for \(\lnot (A \wedge B)\) and \(A \wedge B\) are equally easy.

\(\Leftarrow \)) Assume \(\alpha \models _{{{\L }}}\mathcal{D}(\mathsf{G})\) and \(\alpha \models _{{{\L }}}\Gamma \). We have \(\alpha ^\mathbf 1 \in Lk(\mathsf{G})\) from 3.4.(b) and obtain \(\alpha ^\mathbf 1 \models \langle \Gamma : \mathsf{G}\rangle \) by induction on the complexity of \(\Gamma \), measured as in the proof of \(\Rightarrow \)). From this \(\models \langle \Gamma : \mathsf{G}\rangle \) follows by Definition 3.1. The basis is for \(\Gamma \) a collection of literals. Consulting tables (3.3), we see that for all \(x \in \Gamma ^+\), we have \(\alpha (x) = \mathbf 1 \) so \(x \in \alpha ^\mathbf 1 \). For all \(y \in \Gamma ^-\), on the other hand, we have \(\alpha (y) = \mathbf 0 \). Since \(\alpha \models _{{{\L }}}\mathcal{D}(\mathsf{G})\), we have \(\alpha \models _{{{\L }}}y \leftrightarrow \bigwedge _{z \in E(y)} \lnot z\), meaning \(\alpha (y) = \overline{\alpha }(\bigwedge _{z \in E(y)} \lnot z)\) (where \(\overline{\alpha }\) is the evaluation of \(\alpha \) according to tables (3.3)). From tables (3.3) we see that there must then be some \(z \in E(y)\) such that \(\alpha (z) = \mathbf 1 \), meaning \(y \in E^\smallsmile \!(\alpha ^\mathbf 1 )\). It follows from Definition 3.1 that \(\alpha ^\mathbf 1 \models \langle \Gamma : \mathsf{G}\rangle \). The inductive steps are straightforward. For instance, if there is some \(A \wedge B \in \Gamma \) that has maximal complexity among formulae of \(\Gamma \), we form \(\Gamma ^{\prime }\) which is like \(\Gamma \) except that we replace \(A \wedge B\) by \(A\) and \(B\). Then \(\models _{{{\L }}}\Gamma ^{\prime }\) and \(\Gamma ^{\prime }\) has smaller complexity than \(\Gamma \) so by IH \(\alpha ^\mathbf 1 \models \langle \Gamma ^{\prime } : \mathsf{G}\rangle \). It follows immediately from Definition 3.1 that \(\alpha ^\mathbf 1 \models \langle \Gamma : \mathsf{G}\rangle \) and we are done. The cases of \(\lnot \lnot A\) and \(\lnot (A \wedge B)\) are equally easy. \(\square \)

1.1 Soundness and completeness of pdl

Soundness and completeness follow easily from the following simple lemma giving us the compositionality we need with respect to admissibility in graphs.

Lemma 7.1

For any graph \(\mathsf{G}\) and \(a \in G\) we have:

  1. (1)

    \(\models \langle \Gamma , a : \mathsf{G}\rangle \) iff \(\models \langle \Gamma , \{{\lnot b \mid b \in E(a)}\} : \mathsf{G}\setminus out(a)\rangle \)

  2. (2)

    \(\models \langle \Gamma , \lnot a : \mathsf{G}\rangle \) iff for some \(b \in E(a)\), \(\models \langle \Gamma , b : \mathsf{G}\rangle \)

Proof

(1) \(\Rightarrow \)) Assume \(\Gamma , a\) is admissible in \(\mathsf{G}\) and let \(L \subseteq G\) be a local kernel witnessing to \(\Gamma \) and containing \(a\). Clearly, \(L\) is a local kernel also in \(\mathsf{G}\setminus out(a)\). Now, since \(a \in L\) it follows that \(E(a) \subseteq E^\smallsmile \!(L)\), so \(\Gamma \cup \{{\lnot b \mid b \in E(a)}\}\) is indeed admissible (in both \(\mathsf{G}\) and \(\mathsf{G}\setminus out(a)\))

\(\Leftarrow \)) Assume \(\Gamma \cup \{{\lnot b \mid b \in E(a)}\}\) is admissible in \(\mathsf{G}\setminus out(a)\) and let \(L \subseteq G\) be an arbitrary local kernel in \(\mathsf{G}\setminus out(a)\) witnessing to this fact. Then for every \(b \in E(a)\) we have \(E(b) \cap L \not = \varnothing \) so \(L \cup \{a\}\) is a local kernel in \(\mathsf{G}\) (as well as in \(\mathsf{G}\setminus out(a)\))

(2) \(\Rightarrow \)) Let \(L \subseteq G\) be a local kernel witnessing to the admissibility of \(\Gamma , \lnot a\) in \(\mathsf{G}\). Then, for some \(b \in E(a)\), we have \(b \in L\). So \(\Gamma , b\) is admissible in \(\mathsf{G}\).

\(\Leftarrow \)) Assume that there is some \(b \in E(a)\) such that \(\Gamma , b\) is admissible. Let \(L \subseteq G\) be a witness. Then \(L\) also witness to the admissibility of \(\Gamma , \lnot a\) in \(\mathsf{G}\). \(\square \)

This lemma establishes soundness and invertibility of the only rules from pdl that are not essentially classical. The rest is easily verified, yielding

Theorem 7.2

pdl is sound and all its rules are invertible.

Proof

The standard sequent rules for the composite formulae in \(\Theta \vdash \Phi \) are trivially invertible, as are the rules for non-atomic basic \(\langle \Gamma :\mathsf{G}\rangle \) (which form a one-sided sequent system for propositional logic). Lemma 7.1 established soundness and invertibility of the four rules for literals in \(\Gamma \). We only have to show that the two axiom schemata are valid:

  1. (1)

    \(\Theta , \langle \Gamma , \lnot a : \mathsf{G}\rangle \vdash \Phi \) for some \(a \in sinks(\mathsf{G})\).

To show \(\Theta , \langle \Gamma , \lnot a : \mathsf{G}\rangle \models \Phi \), it suffices to show that \(\not \models \langle \lnot a : \mathsf{G}\rangle \), by Definition 3.1. By Definition 3.1, this amounts to the nonexistence of a local kernel \(L\) of \(\mathsf{G}\) containing a successor of \(a\). But since \(a\) is a sink in \(\mathsf{G}\), no such \(L\) exists.

  1. (2)

    \(\Theta \vdash \langle \Gamma : \mathsf{G}\rangle , \Phi \) for some \(\Gamma \subseteq sinks(\mathsf{G})\).

To show \(\Theta \models \langle \Gamma : \mathsf{G}\rangle , \Phi \), it suffices to show \(\models \langle \Gamma : \mathsf{G}\rangle \). Since \(\Gamma \) is a collection of atomic expressions this amounts to showing that there is a local kernel \(L\) in \(\mathsf{G}\) such that \(\Gamma \subseteq L\). But \(sinks(\mathsf{G})\) is such a local kernel in \(\mathsf{G}\) so the claim follows. \(\square \)

Completeness of pdl follows now by the standard line of reasoning, demonstrating invalidity of any unprovable sequent. We say that a sequent \(\Theta \vdash \Phi \) is reduced when \(\Theta \) and \(\Phi \) contain only atomic formulae, i.e., every \(\langle \Gamma : \mathsf{G}\rangle \in \Theta \cup \Phi \) contains only literals and, moreover, literals over sinks of \(\mathsf{G}\), i.e.,

$$\begin{aligned} \Gamma = \{a\mid a\in \Gamma ^+\subseteq sinks(\mathsf{G})\} \cup \{{\lnot b\mid b \in \Gamma ^-\subseteq sinks(\mathsf{G})}\}. \end{aligned}$$

We first argue that, for any sequent, the rules suffice to create a proof-tree with all leafs reduced.

Trivially, the top level rules and rules for composite \(\Gamma \) suffice to create a proof-tree where all leafs have the form \(\Theta \vdash \Phi \) with \(\Theta \) and \(\Phi \) being collections of atomic expressions \(\langle \Gamma : \mathsf{G}\rangle \), i.e., each \(\Gamma \) being a collection of literals. Now, we employ the rules for literals, as long as there is some \(a \in \Gamma \) or \(\lnot a \in \Gamma \) with \(E(a) \not = \varnothing \), i.e. as long as the sequent is not reduced. For any finite graph \(\mathsf{G}\), it is clear that by employing these rules we will eventually reach a stage where all sequents have been reduced. If (i) \(a\in \Gamma \) is not a sink, an application of the rule \((\vdash {\!a})\), resp., \((a\!\vdash )\), makes it a sink. If (ii) \(\lnot a\in \Gamma \) is not a sink, then an application of the rule \((\vdash {\!\lnot })\), resp., \((\lnot \!\vdash )\), replaces it by all its out-neighbours with positive polarity, for which case (i) applies in the next round.

Theorem 3.10

System pdl is sound and complete: \(\Theta \vdash \Phi \) iff \(\Theta \models \Phi \).

Proof

We show that reduced, non-axiomatic \(\Theta \vdash \Phi \), is invalid. We have:

(1) \(\forall \langle \Gamma _T : \mathsf{G}_T\rangle \in \Theta : \Gamma _T^- = \varnothing \) and (2) \(\forall \langle \Gamma _F : \mathsf{G}_F\rangle \in \Phi : \Gamma _F^- \not = \varnothing .\)

(1) follows since \(\Theta \vdash \Phi \) is reduced, so for all \(\langle \Gamma _T : \mathsf{G}_T\rangle \in \Theta : \Gamma _T^+ \cup \Gamma _T^- \subseteq sinks(\mathsf{G}_T)\). Since the sequent is not axiomatic, we must have \(\Gamma _T^- = \varnothing \). Consequently, \(\Gamma _T \subseteq sinks(\mathsf{G}_T)\) and, since \(sinks(\mathsf{G}_T)\in Lk(\mathsf{G}_T)\), so \(\models \langle \Gamma _T : \mathsf{G}_T\rangle \).

(2) follows since, as before, \(\Gamma _F^+ \cup \Gamma _F^- \subseteq sinks(\mathsf{G}_F)\) and, since the sequent is not axiomatic, \(\Gamma _F \not \subseteq sinks(\mathsf{G}_F)\). Consequently, \(\Gamma _F^- \not = \varnothing \) (since atoms from this set are negated in \(\Gamma _F\)). So there is some \(a \in \Gamma _F^- \subseteq sinks(\mathsf{G}_F)\), i.e. \(\lnot a \in \Gamma _F\) while \(a \in sinks(\mathsf{G}_F)\). It follows that \(\not \models \langle \Gamma _F : \mathsf{G}_F\rangle \).

Having obtained \(\models \langle \Gamma _T : \mathsf{G}_T\rangle \) for all \(\langle \Gamma _T : \mathsf{G}_T\rangle \in \Theta \) and \(\not \models \langle \Gamma _F : \mathsf{G}_F\rangle \) for all \(\langle \Gamma _F : \mathsf{G}_F\rangle \in \Phi \), we conclude by Definition 3.1 that \(\Theta \not \models \Phi \). Invertibility of all the rules ensures that if such a reduced sequent is obtained as a leaf in a proof tree from some initial sequent \(S\), then also \(S\) is invalid. Invertibility was shown in Theorem 7.2 and here we also established soundness of the system. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Dyrkolbotn, S., Walicki, M. Propositional discourse logic. Synthese 191, 863–899 (2014). https://doi.org/10.1007/s11229-013-0297-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-013-0297-x

Keywords

Navigation