Skip to main content
Log in

Logic of Justified Beliefs Based on Argumentation

  • Original Research
  • Published:
Erkenntnis Aims and scope Submit manuscript

Abstract

This manuscript presents a topological argumentation framework for modelling notions of evidence-based (i.e., justified) belief. Our framework relies on so-called topological evidence models to represent the pieces of evidence that an agent has at her disposal, and it uses abstract argumentation theory to select the pieces of evidence that the agent will use to define her beliefs. The tools from abstract argumentation theory allow us to model agents who make decisions in the presence of contradictory information. Thanks to this, it is possible to define two new notions of beliefs, grounded beliefs and fully grounded beliefs. These notions are discussed in this paper, analysed and compared with the existing notion of topological justified belief. This comparison revolves around three main issues: closure under conjunction introduction, the level of consistency and their logical strength.

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. There are exceptions, as some earlier developments do represent doxastic notions and their supporting evidence. An example is the Dempster-Shafer theory of belief functions (Dempster 1968; Shafer 1976).

  2. Indeed, McKinsey (1941) already relates topology with modal logic. An overview of the work in this direction can be found in Aiello et al. (2007).

  3. Note how this idea of defining beliefs in terms of evidence and argumentation is different from, e.g., the proposal Grossi and van der Hoek (2014), which takes beliefs and arguments as two primitive and independent notions and studies their relationship.

  4. For a further philosophical discussion on the notion of information-as-range, we refer to Adriaans and van Benthem (2008) and Martinez and Sequoiah-Grayson (2019).

  5. Thus, a representation of Nora’s scenario includes neither the people of the village nor the newspaper: it only includes the information that these sources provide.

  6. In other words, we are stating that a piece of evidence \(E \in 2^W\) supports a proposition \(P \subseteq W\) if and only if P holds whenever E is truthful (i.e., whenever E contains the real world).

  7. For a philosophical discussion on the use of these concepts in the context of scientific inquiry, we point to Kelly’s work in Kelly (1996).

  8. Note that it is the idea of evidence as affirmative information what makes topology a suitable structure for modelling evidence. Other structures may be suitable as well (and even better suited) if some other aspects of evidence are taken into account. Here, no attempt is made to establish a topological structure as “the” structure for evidence.

  9. Another classical example of an assertion which is not affirmative is “all ravens are black”. It is clear that we should never take a piece of information which is impossible to be verified to be our evidence. Readers are referred to Vickers (1989, Chapter 2) for a deeper discussion.

  10. The topology generated by a given \({\mathcal {Y}}\subseteq 2^X\) is the smallest topology \(\tau _{{\mathcal {Y}}}\) over X such that \({\mathcal {Y}}\subseteq \tau _{{\mathcal {Y}}}\). When no confusion arises, \(\tau _{{\mathcal {E}}_0}\) will be denoted simply by \(\tau \).

  11. In fact, the argumentation framework is general enough to allow an argument to be another argument’s attacker and defender simultaneously. Still, there are proposals introducing new relations between arguments, including different forms of attack and support (e.g., Oren and Norman 2008; Cayrol and Lagasquie-Schiex 2009; Prakken 2010).

  12. Previous presentations of this structure (Shi 2018; Shi et al. 2017, 2018) asked for a further condition on the attack relation: for every \(T, T_1, T'_1 \in \tau \), if \(T_1 \leftarrowtail T\) and \(T'_1 \subseteq T_1\), then \(T'_1 \leftarrowtail T\). The definition provided here is a generalisation of the previous proposal. As it will be proved later, this does not change the resulting logic.

  13. The non-empty elements of \(\tau \) have already been called arguments in Baltag et al. (2016). For a deeper discussion on this, see Özgün (2017, Section 5.2).

  14. Thus, from the first condition, it follows that the empty set only attacks itself. In particular, this says that no subset of \(\tau \) containing \(\varnothing \) can be conflict-free.

  15. For more details and other alternatives, the reader is referred to Baroni and Giacomin (2009); Baroni et al. (2011).

  16. Thus, \(\varnothing \notin {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \), and hence \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) contains only arguments.

  17. Attack edges involving the empty set are not drawn.

  18. Alternatively, \(\{1,2,3\}\) is the only set supported by every element of \({{\,\mathrm{\mathsf {LFP}}\,}}\).

  19. This paper does not elaborate on the interpretation of \(\mathop {{\mathfrak {T}}}\) as grounded knowledge. Interested readers are referred to Shi (2020), which makes a fine-grained comparative analysis of three different notions of knowledge that arise within TA models.

  20. The completeness proof takes advantage of this fact, working only with \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\).

  21. When \(\psi \) is globally true, \(\mathop {{\text {U}}}\psi \) is also globally true, and thus is true in the factive argument that makes \(\mathop {{\mathfrak {T}}}\varphi \) true.

  22. Note that “the agent has the argument \(\varphi \)” is different from “the agent has an argument for \(\varphi \)”. The former is expressed by \(\mathop {{\text {U}}}(\varphi \rightarrow \mathop {\Box }\varphi )\), semantically stating that there is an argument T satisfying \(T = \llbracket \varphi \rrbracket \); the latter corresponds to \(\mathop {\widehat{{\text {U}}}}\mathop {\Box }\varphi \), semantically stating that there is an argument T satisfying \(T\subseteq \llbracket \varphi \rrbracket \).

  23. The completeness result proved in Appendix B is actually stronger than what Theorem 1 states (see Theorem 2).

  24. Note that Proposition 4 only provides two sufficient conditions.

  25. Note that we write P holds in w (i.e. \(w \in P\)) as \(M,w\models P\).

  26. See Bjorndahl and Özgün (2019) and Balbiani et al. (2019) for proposals modelling this process.

  27. In fact, fully grounded belief is exactly what emerges when the set of arguments on which grounded beliefs are based, \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \), is closed under intersections (Proposition 14).

  28. A similar strategy is used in Baltag et al. (2016): the authors first showed show that any consistent set of formulas is satisfiable in a quasi-model, and then transformed such model into a modally-equivalent topological evidence model.

  29. For \(\varvec{(\subseteq })\), suppose \(\varDelta \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\); since \(\leqslant \) is reflexive, \(\varDelta \in {\leqslant }[\varDelta ]\), and thus \(\varDelta \in \bigcup _{\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}} {\leqslant }[\varGamma ]\). For \(\varvec{(\supseteq })\), take \(\varDelta \in \bigcup _{\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}} {\leqslant }[\varGamma ]\); then, \(\varDelta \in {\leqslant }[\varGamma ]\) for some \(\varGamma \) in \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\), that is, for some \(\varGamma \) with \(\mathop {{\mathfrak {T}}}\varphi \in \varGamma \). But then, from axioms \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi \) and \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {\Box }\varphi \), we get \(\mathop {\Box }\mathop {{\mathfrak {T}}}\varphi \in \varGamma \). Then, from \(\leqslant \)’s definition, \(\varGamma \leqslant \varDelta \) implies \(\mathop {{\mathfrak {T}}}\varphi \in \varDelta \). Hence, \(\varDelta \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\).

  30. Indeed, \(T \subseteq \bigcup _{\varGamma \in T} {\leqslant }[\varGamma ]\) is immediate (\(\leqslant \) is reflexive), and \(T \supseteq \bigcup _{\varGamma \in T} {\leqslant }[\varGamma ]\) follows from the fact that if \(\varGamma \in T\) then \({\leqslant }[\varGamma ] \subseteq T\).

  31. This includes \(\varnothing \), as \(T \in {\mathcal {C}}_1\) implies \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \ne \varnothing \) for T’s corresponding \(\varphi \); hence, \(T \ne \varnothing \) and thus \(T \not \leftarrowtail \varnothing \).

References

  • Adriaans, P., & van Benthem, J. (Eds.). (2008). Philosophy of information. Elsevier.

  • Aiello, M., Pratt-Hartmann, I., & van Benthem, J. (Eds.). (2007). Handbook of spatial logics. Springer.

  • Artemov, S. N. (2008). The logic of justification. The Review of Symbolic Logic, 1(4), 477–513. https://doi.org/10.1017/S1755020308090060.

    Article  Google Scholar 

  • Artemov, S. N., & Nogina, E. (2005). Introducing justification into epistemic logic. Journal of Logic and Computation, 15(6), 1059–1073. https://doi.org/10.1093/logcom/exi053.

    Article  Google Scholar 

  • Balbiani, P., Fernández-Duque, D., Herzig, A., & Lorini, E. (2019). Stratified evidence logics. In S. Kraus (Ed.), Proceedings of the 28th international joint conference on artificial intelligence, IJCAI 2019, Macao, China, August 10–16, 2019, ijcai.org (pp. 1523–1529). https://doi.org/10.24963/ijcai.2019/211.

  • Baltag, A., & Smets, S. (2008). A qualitative theory of dynamic interactive belief revision. Texts in Logic and Games, 3, 9–58.

    Google Scholar 

  • Baltag, A., Renne, B., & Smets, S. (2012). The logic of justified belief change, soft evidence and defeasible knowledge. In C. L. Ong, R. J. G. B. de Queiroz (Eds.), Logic, language, information and computation—19th international workshop, WoLLIC 2012, Buenos Aires, Argentina, September 3–6, 2012, Proceedings. Lecture Notes in Computer Science (Vol. 7456, pp. 168–190). Springer. https://doi.org/10.1007/978-3-642-32621-9_13.

  • Baltag, A., Renne, B., & Smets, S. (2014). The logic of justified belief, explicit knowledge and conclusive evidence. Annals of Pure and Applied Logic, 165(1), 49–81. https://doi.org/10.1016/j.apal.2013.07.005.

    Article  Google Scholar 

  • Baltag, A., Bezhanishvili, N., Özgün, A., & Smets, S. (2016). Justified belief and the topology of evidence. In: J. Väänänen, Å. Hirvonen, R. de Queiroz (Eds.), Logic, language, information, and computation: 23rd international workshop, WoLLIC 2016, Puebla, México, August 16–19th, 2016, Proceedings (pp. 83–103). Springer. https://doi.org/10.1007/978-3-662-52921-8_6.

  • Baroni, P., & Giacomin, M. (2009). Semantics of abstract argument systems. In: Simari and Rahwan (pp. 25–44). https://doi.org/10.1007/978-0-387-98197-0_2.

  • Baroni, P., Caminada, M., & Giacomin, M. (2011). An introduction to argumentation semantics. Knowledge Engineering Review, 26(4), 365–410. https://doi.org/10.1017/S0269888911000166.

    Article  Google Scholar 

  • van Benthem, J. (2011). Logical dynamics of information and interaction. Cambridge University Press.

  • van Benthem, J., & Pacuit, E. (2011). Dynamic logics of evidence-based beliefs. Studia Logica, 99(1–3), 61–92. https://doi.org/10.1007/s11225-011-9347-x.

    Article  Google Scholar 

  • van Benthem, J., Duque, D. F., & Pacuit, E. (2014). Evidence and plausibility in neighborhood structures. Annals of Pure and Applied Logic, 165(1), 106–133. https://doi.org/10.1016/j.apal.2013.07.007.

    Article  Google Scholar 

  • Bjorndahl, A., & Özgün, A. (2019). Uncertainty about evidence. In: L. S. Moss (Ed.), Proceedings 17th conference on theoretical aspects of rationality and knowledge, TARK 2019, Toulouse, France, 17–19 July 2019, EPTCS (Vol. 297, pp. 68–81). https://doi.org/10.4204/EPTCS.297.5.

  • Board, O. (2004). Dynamic interactive epistemology. Games and Economic Behaviour, 49(1), 49–80. https://doi.org/10.1016/j.geb.2003.10.006.

    Article  Google Scholar 

  • Cayrol, C., & Lagasquie-Schiex, M. (2009). Bipolar abstract argumentation systems. In: Simari and Rahwan (pp. 65–84). https://doi.org/10.1007/978-0-387-98197-0_4.

  • Dabrowski, A., Moss, L. S., & Parikh, R. (1996). Topological reasoning and the logic of knowledge. Annals of Pure and Applied Logic, 78, 73–110.

    Article  Google Scholar 

  • Dempster, A. P. (1968). A generalization of Bayesian inference. Journal of the Royal Statistical Society: Series B (Methodological), 30(2), 205–232. https://doi.org/10.1111/j.2517-6161.1968.tb00722.x.

    Article  Google Scholar 

  • van Ditmarsch, H., van der Hoek, W., & Kooi, B. (2008). Dynamic epistemic logic, synthese library series (Vol. 337). Springer. https://doi.org/10.1007/978-1-4020-5839-4.

  • 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. https://doi.org/10.1016/0004-3702(94)00041-X.

    Article  Google Scholar 

  • Easwaran, K. (2011). Bayesianism i: introduction and arguments in favor. Philosophy Compass, 6(5), 312–320. https://doi.org/10.1111/j.1747-9991.2011.00399.x.

    Article  Google Scholar 

  • Égré, P., Marty, P., & Renne, B. (2014). Knowledge, justification, and reason-based belief. CoRR. arXiv:1412.1862.

  • Fagin, R., Moses, Y., Vardi, M. Y., & Halpern, J. Y. (1995). Reasoning about knowledge. MIT Press.

  • Foley R (2009) Beliefs, degrees of belief, and the Lockean thesis. In: Huber and Schmidt-Petri (pp. 37–47). https://doi.org/10.1007/978-1-4020-9198-8_2.

  • Georgatos, K. (1992). Knowledge theoretic properties of topological spaces. In: M. Masuch, L. Pólos (Eds.), Knowledge representation and reasoning under uncertainty, logic at work [International conference logic at work, Amsterdam, The Netherlands, December 17–19, 1992]. Lecture Notes in Computer Science (Vol. 808, pp. 147–159). Springer. https://doi.org/10.1007/3-540-58095-6_11.

  • Grossi, D., & van der Hoek, W. (2014). Justified beliefs by justified arguments. In Proceedings of the 14th international conference on principles of knowledge representation and reasoning (pp. 131–140). AAAI Press.

  • Hintikka, J. (1962). Knowledge and belief. Cornell University Press.

  • Huber, F., Schmidt-Petri, C. (Eds.) (2009). Degrees of belief. No. 342 in Synthese Library. Springer. https://doi.org/10.1007/978-1-4020-9198-8_2.

  • Kelly, K. T. (1996). The logic of reliable inquiry. Oxford University Press.

  • Kelly, T. (2016). Evidence. In E. N. Zalta (Ed.), The Stanford encyclopedia of philosophy, winter (2016th ed.). Metaphysics Research Lab, Stanford University.

  • Knaster, B. (1928). Un théorème sur les fonctions d’ensembles. Annales de la Société Polonaise de Mathématiques, 6, 133–134.

    Google Scholar 

  • Kyburg, H. (1961). Probability and the logic of rational belief. Wesleyan University Press.

  • Leitgeb, H. (2017). The stability of belief—How rational belief coheres with probability. Oxford University Press.

  • Martinez, M., & Sequoiah-Grayson, S. (2019). Logic and information. In E. N. Zalta (Ed.), The stanford encyclopedia of philosophy, spring (2019th ed.). Metaphysics Research Lab, Stanford University.

  • McKinsey, J. C. C. (1941). A solution of the decision problem for the Lewis systems S2 and s4, with an application to topology. The Journal of Symbolic Logic, 6(4), 117–134. https://doi.org/10.2307/2267105.

    Article  Google Scholar 

  • Meyer, J. J. C., & van Der Hoek, W. (1995). Epistemic logic for AI and computer science. Cambridge University Press. https://doi.org/10.1017/CBO9780511569852.

  • Moss, L. S., & Parikh, R. (1992). Topological reasoning and the logic of knowledge. In Y. Moses (Ed.), Proceeding, 4th conf. on theoretical aspects of reasoning about knowledge (TARK 1992) (pp. 95–105). Morgan Kaufmann.

  • Oren, N., & Norman, T. J. (2008). Semantics for evidence-based argumentation. In P. Besnard, S. Doutre, A. Hunter (Eds.), Computational models of argument: proceedings of COMMA 2008, Toulouse, France, May 28–30, 2008, Frontiers in artificial intelligence and applications (Vol. 172, pp. 276–284). IOS Press.

  • Özgün, A. (2017). Evidence in epistemic logic: a topological perspective. Ph.D thesis, Institute for Logic, Language and Computation, University of Amsterdam.

  • Prakken, H. (2010). An abstract framework for argumentation with structured arguments. Argument & Computation, 1(2), 93–124. https://doi.org/10.1080/19462160903564592.

    Article  Google Scholar 

  • Shafer, G. (1976). A mathematical theory of evidence. Princeton University Press.

  • Shi, C. (2018). Reason to believe. University of Amsterdam.

  • Shi, C. (2020). No false grounds and topology of argumentation. Journal of Logic and Computation.https://doi.org/10.1093/logcom/exaa057.

  • Shi, C., Smets, S., & Velázquez-Quesada, F. R. (2017). Argument-based belief in topological structures. In: EPTCS 251: Theoretical Aspects of Rationality and Knowledge (pp 489–503). https://doi.org/10.4204/EPTCS.251.36.

  • Shi, C., Smets, S., & Velázquez-Quesada, F. R. (2018). Beliefs based on evidence and argumentation. In Proceedings of WOLLIC 2018 - Workshop on Logic, Language, Information, and Computation (pp. 289–306). https://doi.org/10.1007/978-3-662-57669-4_17.

  • Spohn, W. (1988). Ordinal conditional functions: A dynamic theory of epistemic states. In W. L. Harper, B. Skyrms (Eds.), Causation in decision, belief Hange, and statistics (pp. 105–134). Kluwer. https://doi.org/10.1007/978-94-009-2865-7_6.

  • Tarski, A. (1955). A lattice-theoretical theorem and its applications. Pacific Journal of Mathematics, 5(2), 285–309.

    Article  Google Scholar 

  • Vickers, S. (1989). Topology via logic. Cambridge University Press.

  • Williamson, T. (2000). Knowledge and its limits. Oxford University Press.

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chenwei Shi.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

A Proofs

1.1 A.1 Proof of Proposition 4

For the first sufficient condition, assume that \(\leftarrowtail \) is symmetric on the set of arguments. Observe that, then, \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau = \{T \in \tau \mid \forall x \in \tau {\setminus } \{\varnothing \}: x \cap T \ne \varnothing \}\), which is closed under conjunction.

The second sufficient condition requires more details, and the following lemma will be useful.

Lemma 2

Let \(M= (W, E_0, \tau _{E_0}, \leftarrowtail , V)\) be a TA model. Then, for all \(F_1, F_2 \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \),

$$\begin{aligned} \begin{array}{l} F_1, F_2 \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \text { implies } F_1 \cap F_2 \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \\ \text { if and only if } \\ \text {for all } T \in \tau , \text { if } F_1 \cap F_2 \leftarrowtail T \text { then there is } F \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \text { such that } T \cap F = \varnothing . \end{array} \end{aligned}$$

Proof

Take arbitrary \(F_1,F_2\in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \). From left to right, consider the contrapositive, and suppose there is an open \(T \in \tau \) such that T attacks \(F_1\cap F_2\) but is not in conflict with anybody in \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \). From the latter it follows that nobody in \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) attacks T, and thus the attacked \(F_1 \cap F_2\) is not defended by \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \); therefore, \(F_1 \cap F_2\) is not in \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \).

From right to left, take an argument \(T \in \tau \) such that \(F_1 \cap F_2\leftarrowtail T\). Then, there is \(F' \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) such that \(T \cap F' = \varnothing \), and thus either \(T \leftarrowtail F'\) or else \(F' \leftarrowtail T\). The first case gives us an argument in \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) attacking T, namely \(F'\); the second case does that too, as \(F'\) is in \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \), and thus there should be \(F'' \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) such that \(T \leftarrowtail F''\). Hence, for all \(T \in \tau \) such that \(F_1\cap F_2\leftarrowtail T\) there is \(F \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) such that \(T \leftarrowtail F\): the set \(F_1\cap F_2\) is defended by \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \). Thus, \(F_1\cap F_2\in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \). \(\square \)

Now, take any \(F_1, F_2 \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) and, for the contrapositive, suppose \(F_1 \cap F_2\) is not in \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \). By Lemma 2, there is an open \(T \in \tau \) which attacks \(F_1 \cap F_2\) (i.e., \(F_1 \cap F_2 \leftarrowtail T\)) and which is not in conflict with any elements of \({{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) (i.e., \(F \in {{\,\mathrm{\mathsf {LFP}}\,}}_\tau \) implies \(T \cap F \ne \varnothing \)). The goal is to show that \(\leftarrowtail \) is not unambiguous, and in order to achieve that define the following \(T_1, T_2\) and \(T_3\):

$$\begin{aligned} T_1 := F_1 \cap \ T, \quad T_2 := F_2 \cap \ T, \quad T_3 := F_1 \cap F_2. \end{aligned}$$

Note how none of the sets are empty. Note also how, due to the fact that T attacks \(F_1 \cap F_2\), T must be in conflict with \(F_1 \cap F_2\) (that is, \((F_1\cap F_2)\cap T = \varnothing \)); hence, \(T_1 \cap T_2 = T_2 \cap T_3 = T_3 \cap T_1 = \varnothing \): the sets are in conflict with one another, and thus there should be pairwise attacks in at least one direction. Consider two cases, \(T_1 \leftarrowtail T_2\) or \(T_2\leftarrowtail T_1\).

  • Suppose \(T_1 \leftarrowtail T_2\). If \(T_2\leftarrowtail T_3\) is also the case, then the fact that either \(T_1\leftarrowtail T_3\) or else \(T_3\leftarrowtail T_1\) should hold make \(\leftarrowtail \) not unambiguous. Otherwise, \(T_3\leftarrowtail T_2\) should hold and then, similarly, the fact that either \(T_1\leftarrowtail T_3\) or else \(T_3\leftarrowtail T_1\) hold make \(\leftarrowtail \) not unambiguous.

  • Suppose \(T_2\leftarrowtail T_1\). By an analogous argument, \(\leftarrowtail \) is not unambiguous.

Thus, the attack relation is not unambiguous.

B Proof of Theorem 1

The text has already argued for the soundness of the system of Table 1 within topological argumentation models (the validity of some of the interesting axioms is proved in Propositions 8, 9 and 10).

For completeness, It will be shown that any \({\mathsf {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-consistent set of \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-formulas is satisfiable. Satisfiability will be proved in an Alexandroff qTA model (see below), which is \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-equivalent to its corresponding TA model.Footnote 28 In fact, we will prove a stronger completeness result than that stated in Theorem 1.

Theorem 2

The axiom system of Table 1 is strongly complete for the language \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\) w.r.t. topological argumentation models which satisfies the following extra condition: for every \(T, T_1, T'_1 \in \tau \): if \(T_1 \leftarrowtail T\) and \(T'_1 \subseteq T_1\), then \(T'_1 \leftarrowtail T\).

Next we prove the theorem.

Definition 11

(qTA model) A quasi-topological argumentation model (qTA) is a tuple in which \((W, {\mathcal {E}}_0, \tau , \leftarrowtail , V)\) is a TA model (with \(\tau \) generated by \({\mathcal {E}}_0\), as before) and \({\leqslant } \subseteq (W \times W)\) a preorder such that, for every \(E \in {\mathcal {E}}_0\), if \(u \in E\) and \(u \leqslant v\), then \(v \in E\).

Formulas in \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\) are interpreted in qTA models just as in TA models. The only difference is \(\mathop {\Box }\), which becomes a normal universal modality for \(\leqslant \). More precisely, iff for all \(v \in W\), if \(w \leqslant v\) then . Now, two topological definitions, a refined qTA model, and the connection.

Definition 12

(Specification preorder) Let \((X,\tau )\) be a topological space. Its specification preorder \({\sqsubseteq _\tau } \subseteq (X \times X) \) is defined, for any \(x, y \in X\), as \(x \sqsubseteq _\tau y\) iff for all \(T \in \tau \), \(x \in T\) implies \(y \in T\).

Definition 13

(Alexandroff space) A topological space \((X, \tau )\) is Alexandroff iff \(\tau \) is closed under arbitrary intersections (i.e., \(\bigcap T \in \tau \) for any \(T \subseteq \tau \)).

Definition 14

(Alexandroff qTA model) A qTA-model is called Alexandroff iff (i) \((W, \tau _{{\mathcal {E}}_0})\) is Alexandroff, and (ii) \({\leqslant } = {\sqsubseteq _\tau }\).

Proposition 16

Given an Alexandroff qTA model , take \(M= (W, {\mathcal {E}}_0, \tau , \leftarrowtail , V)\). Then, for every \(\varphi \in {{\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}}\).

Proof

Exactly as that of Özgün (2017, Prop. 5.6.14) for topological evidence models and \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}}\), as \(\mathop {{\mathfrak {T}}}\) has the same truth condition in qTA and TA models. \(\square \)

For notation, define \(\varGamma ^\bigcirc = \{ \varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \mid \bigcirc \varphi \in \varGamma \}\) for \(\varGamma \subseteq {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) and \(\bigcirc \in \{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}\}\). For the proof, let \(\varPhi _0\) be a \({\mathsf {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-consistent set of \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-formulas. A slightly modified version of Lindenbaum Lemma shows that it can be extended to a maximal consistent one. Let \({\text {MCS}}\) be the family of all maximally \({\mathsf {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-consistent sets of \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-formulas; let \(\varPhi \) be an element of \({\text {MCS}}\) extending \(\varPhi _0\).

Definition 15

(Canonical qTA model) The canonical qTA model for \(\varPhi \), , is defined as follows.

  • \(W^\varPhi := \{ \varGamma \in {\text {MCS}}\mid \varGamma ^{\mathop {{\text {U}}}} = \varPhi ^{\mathop {{\text {U}}}} \}\) and \(V^\varPhi (p) := \{ \varGamma \in W^\varPhi \mid p \in \varGamma \}\).

  • For \(\varGamma , \varDelta \in W^\varPhi \;\), \(\varGamma \leqslant ^\varPhi \varDelta \quad iff _ def \quad \text {for any } \varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \;, \mathop {\Box }\varphi \in \varGamma \text { implies } \varphi \in \varDelta . \)

  • For any \(\varGamma \in W^\varPhi \), define the set \({\leqslant ^{\varPhi }}[\varGamma ] := \{ \varOmega \in W^\varPhi \mid \varGamma \leqslant ^\varPhi \varOmega \}\). Then, let \( {\mathcal {E}}_0^\varPhi := \{ \bigcup _{\varGamma \in U} {\leqslant ^{\varPhi }}[\varGamma ] \mid U \subseteq W^\varPhi \} {\setminus } \{\varnothing \}. \)

While \(\leqslant ^\varPhi \) and \(V^{\varPhi }\) are standard (recall: \(\mathop {\Box }\) is a normal universal modality for \(\leqslant \)), each \(E \in {\mathcal {E}}_0^{\varPhi }\) is a non-empty union of the \(\leqslant ^{\varPhi }\)-upwards closure of the elements of some subset of \(W^{\varPhi }\). The last component, the attack relation \(\leftarrowtail ^{\varPhi }\), is the novel one in this model, and it requires more care. First, define \( \{\!\vert \varphi \vert \!\}_M:= \{ \varGamma \in W^{\varPhi } \mid \varphi \in \varGamma \}. \) Then, by taking \(\tau ^\varPhi \) to be the topology generated by \({\mathcal {E}}_0^\varPhi \) define, for any \(T, T' \in \tau ^\varPhi \),

$$\begin{aligned} \bullet \;\; T \leftarrowtail ^\varPhi T' \quad iff _ def \quad \left\{ \begin{array}{l@{\quad }l@{}} T = \varnothing &{} \text {if } T' = \varnothing \\ T \cap T' = \varnothing \text { and there is no } \varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \text { s.t. } &{} \text {otherwise } \\ \qquad \qquad \qquad \quad \text {both } \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq T \text { and } \mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \\ \end{array} \right. \end{aligned}$$

When no confusion arises, the superscript \(\varPhi \) will be omitted.

Note how is indeed a qTA model (Definition 11). First, it is clear that \(\varnothing \notin {\mathcal {E}}_0\) and \(W \in {\mathcal {E}}_0\). Moreover, \(\leqslant \) is indeed a preorder (see its axioms) satisfying the extra condition. Finally, it can be proved that \(\leftarrowtail \) satisfies the three conditions.

Lemma 3

Let be the model of Definition 15. Then,

  1. (i)

    for every \(T_1, T_2 \in \tau \): \(T_1 \cap T_2 = \varnothing \) if and only if \(T_1 \leftarrowtail T_2\) or \(T_2 \leftarrowtail T_1\);

  2. (ii)

    for every \(T, T_1, T'_1 \in \tau \): if \(T_1 \leftarrowtail T\) and \(T'_1 \subseteq T_1\), then \(T'_1 \leftarrowtail T\);

  3. (iii)

    for every \(T \in \tau {\setminus } \{\varnothing \}\): \(\varnothing \leftarrowtail T\) and \(T \not \leftarrowtail \varnothing \).

Proof

  1. (i)

    The right-to-left direction is immediate. From left to right, assume \(T_1 \cap T_2 = \varnothing \); moreover, for a contradiction, suppose both \(T_1 \not \leftarrowtail T_2\) and \(T_2 \not \leftarrowtail T_1\). Then, there is \(\varphi _1 \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \subseteq T_1\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _1 \in \varPhi \), and there is \(\varphi _2 \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} \subseteq T_2\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _2 \in \varPhi \). It follows that \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} = \varnothing \); to finish the proof, it is enough to show

Lemma 4

For any \(\varphi _1, \varphi _2 \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \), having \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} = \varnothing \) and both \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _1 \in \varPhi \) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _2 \in \varPhi \) leads to a contradiction.

Proof

From \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} = \varnothing \) and theorem \(\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {{\mathfrak {T}}}\varphi \) it follows that \(\{\!\vert \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \cap \{\!\vert \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} = \varnothing \). Moreover: \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _1 \in \varPhi \) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _2 \in \varPhi \) imply, respectively, \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \ne \varnothing \) and \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} \ne \varnothing \) (Proposition 17 below); this, together with axiom \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi \), yields both \(\{\!\vert \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \ne \varnothing \) and \(\{\!\vert \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} \ne \varnothing \), which imply (Proposition 17) \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _1 \in \varPhi \) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _2 \in \varPhi \).

Observe how \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi _2 \vert \!\} = \varnothing \) also implies \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _1 \vert \!\} \subseteq \{\!\vert \lnot \mathop {{\mathfrak {T}}}\varphi _2 \vert \!\}\), so \(\mathop {{\text {U}}}(\mathop {{\mathfrak {T}}}\varphi _1 \rightarrow \lnot \mathop {{\mathfrak {T}}}\varphi _2) \in \varPhi \); then, from Proposition 7 it follows that \(\mathop {{\text {U}}}(\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _1 \rightarrow \mathop {{\mathfrak {T}}}\lnot \mathop {{\mathfrak {T}}}\varphi _2) \in \varPhi \). From the latter and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _1 \in \varPhi \) we get \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\lnot \mathop {{\mathfrak {T}}}\varphi _2 \in \varPhi \); but then, axiom \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \rightarrow \lnot \mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\lnot \varphi \) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _2 \in \varPhi \) imply \(\lnot \mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\lnot \mathop {{\mathfrak {T}}}\varphi _2 \in \varPhi \). Thus, we have a contradiction, as \(\varPhi \) is consistent.

  1. (ii)

    Take \(T, T_1, T'_1 \in \tau \) with \(T_1 \leftarrowtail T\) and \(T'_1 \subseteq T_1\). If \(T = \varnothing \), the case is trivial (\(T_1\) should be \(\varnothing \), and so \(T'_1\)), so suppose \(T \ne \varnothing \). From \(T_1 \leftarrowtail T\), there is no \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq T_1\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \). But \(T'_1 \subseteq T_1\), so no such \(\varphi \) exists for \(T'_1\) either. Moreover, \(T_1 \cap T = \varnothing \) so \(T'_1 \cap T = \varnothing \); hence, \(T'_1 \leftarrowtail T\).

  2. (iii)

    Immediate.\(\square \)

Thus, is a qTA model. The next proposition (standard proof) provides existence lemmas for the standard modality \(\mathop {\Box }\) and the global modality \(\mathop {\widehat{{\text {U}}}}\).

Proposition 17

For any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) and any \(\varGamma \in W\):

  • \(\mathop {\Diamond }\varphi \in \varGamma \) iff there is \(\varDelta \in W\) s.t. \(\varGamma \leqslant \varDelta \) and \(\varphi \in \varDelta \).

  • \(\mathop {\widehat{{\text {U}}}}\varphi \in \varGamma \) iff there is \(\varDelta \in W\) s.t. \(\varphi \in \varDelta \);

Now, tools to prove a similar result for the operator \(\mathop {{\mathfrak {T}}}\), whose truth clause relies on \({{\,\mathrm{\mathsf {LFP}}\,}}\), given by \(\leftarrowtail \). First, some useful properties of the model.

Fact 2

(i) \(\tau = {\mathcal {E}}_0\cup \{\varnothing \}\). (ii) If \(\mathop {\widehat{{\text {U}}}}\mathop {\Box }\varphi \in \varPhi \), then \(\{\!\vert \mathop {\Box }\varphi \vert \!\} \in \tau \). (iii) If \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \), then \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \in \tau \). (iv) For any \(T \in \tau \) and any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \): if \(T \subseteq \{\!\vert \varphi \vert \!\}\), then \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\).

Proof

  1. (i)

    Immediate, as \(\{ \bigcup _{\varGamma \in U} {\leqslant ^{\varPhi }}[\varGamma ] \mid U \subseteq W^\varPhi \}\) is closed under intersections and unions.

  2. (ii)

    Analogous to the next one.

  3. (iii)

    Assume \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \); then, there is at least one \(\varGamma \in W\) such that \(\mathop {{\mathfrak {T}}}\varphi \in \varGamma \), that is, \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \ne \varnothing \). Now, note how \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} = \bigcup _{\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}} {\leqslant }[\varGamma ]\).Footnote 29 Moreover, \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq W\) so, by \({\mathcal {E}}_0\)’s definition, \(\bigcup _{\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}} {\leqslant }[\varGamma ] = \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \in {\mathcal {E}}_0\); then, by the first item, \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \in \tau \).

  4. (iv)

    First, note how \(T = \bigcup _{\varGamma \in T} {\leqslant }[\varGamma ]\).Footnote 30 Then, from \(T \subseteq \{\!\vert \varphi \vert \!\}\) it follows that \(\bigcup _{\varGamma \in T} {\leqslant }[\varGamma ] \subseteq \{\!\vert \varphi \vert \!\}\), that is, \({\leqslant }[\varGamma ] \subseteq \{\!\vert \varphi \vert \!\}\) for every \(\varGamma \in T\); therefore, \(\mathop {\Box }\varphi \in \varGamma \) (Proposition 17) for every such \(\varGamma \), and hence \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\).

\(\square \)

Here are the first steps towards locating \({{\,\mathrm{\mathsf {LFP}}\,}}\).

Definition 16

(Semi-acceptable and acceptable) Define \({\mathcal {C}}_1\) as

$$\begin{aligned} {\mathcal {C}}_1= \{ T \in \tau \mid \text {there exists } \varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \text { such that } \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq T \text { and } \mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \} \end{aligned}$$
  • An open \(T \in \tau \) is semi-acceptable if and only if, for any \(\psi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(T \subseteq \{\!\vert \mathop {\Box }\psi \vert \!\}\), there is \(\xi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\xi \vert \!\} \subseteq \{\!\vert \mathop {\Box }\psi \vert \!\}\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\xi \in \varPhi \).

  • An open \(T \in \tau \) is acceptable if and only if T is semi-acceptable and there is no \(T' \in \tau \) such that \(T \cap T' = \varnothing \) and \(T' \cap T'' \ne \varnothing \) for all \(T'' \in {\mathcal {C}}_1\).

Define \({\mathcal {C}}_2\) as \({\mathcal {C}}_2= \{T \in \tau {\setminus } {\mathcal {C}}_1\mid T \text { is acceptable} \}\).

Note that no element of \({\mathcal {C}}_1\) is attacked by elements of \(\tau \). Moreover,

Fact 3

(i) For any \(T \in \tau \), if \(T \in {\mathcal {C}}_1\), then T is acceptable. (ii) If \(T \in \tau \) is semi-acceptable, then \(T \cap T' \ne \varnothing \) for all \(T' \in {\mathcal {C}}_1\).

Proof

  1. (i)

    If \(T \in {\mathcal {C}}_1\), then there is \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq T\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \). For semi-acceptability, take any \(\psi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(T \subseteq \{\!\vert \mathop {\Box }\psi \vert \!\}\); the initial \(\varphi \) satisfies both \(\{\!\vert \mathop {{\mathfrak {T}}}\xi \vert \!\} \subseteq \{\!\vert \mathop {\Box }\psi \vert \!\}\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\xi \in \varPhi \). For the second condition of acceptability, \(T\in {\mathcal {C}}_1\), so for any \(T'\in \tau \) such that \(T'\cap T'' \ne \varnothing \) for all \(T''\in {\mathcal {C}}_1\), it is the case that \(T'\cap T \ne \varnothing \).

  2. (ii)

    It will be proved that, for any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \), any semi-acceptable open \(T \in \tau \) satisfies that \(T \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \ne \varnothing \). Fact 3.(ii) follows, because every open in \({\mathcal {C}}_1\) should be a superset of \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\) for some \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \).

    For a contradiction, suppose there is \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \) and \(T \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} = \varnothing \) for a semi-acceptable T. Now take any \(\psi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(T \subseteq \{\!\vert \mathop {\Box }\psi \vert \!\}\); then, \(T\subseteq \{\!\vert \mathop {\Box }\psi \wedge \lnot \mathop {{\mathfrak {T}}}\varphi \vert \!\}\). By Item (iv) of Fact 2, \(T \subseteq \{\!\vert \mathop {\Box }(\mathop {\Box }\psi \wedge \lnot \mathop {{\mathfrak {T}}}\varphi ) \vert \!\}\). But T is semi-acceptable, so there is \(\xi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\xi \vert \!\} \subseteq \{\!\vert \mathop {\Box }(\mathop {\Box }\psi \wedge \lnot \mathop {{\mathfrak {T}}}\varphi ) \vert \!\}\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\xi \in \varPhi \). Now, \(\vdash \mathop {\Box }(\mathop {\Box }\psi \wedge \lnot \mathop {{\mathfrak {T}}}\varphi ) \rightarrow \lnot \mathop {{\mathfrak {T}}}\varphi \), so \(\{\!\vert \mathop {{\mathfrak {T}}}\xi \vert \!\} \subseteq \{\!\vert \lnot \mathop {{\mathfrak {T}}}\varphi \vert \!\}\), that is, \(\{\!\vert \mathop {{\mathfrak {T}}}\xi \vert \!\} \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} = \varnothing \). The contradiction then follows from Lemma 4.

\(\square \)

Lemma 5

Let \({\mathcal {C}}= {\mathcal {C}}_1\cup {\mathcal {C}}_2\). Then, \({{\,\mathrm{\mathsf {LFP}}\,}}= {\mathcal {C}}\).

Proof

\(\varvec{(\supseteq })\) The proof of this direction consists of proving two cases. (i) If \(T \in {\mathcal {C}}_1\), there is \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq T\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \). Hence, by \(\leftarrowtail \)’s definition, there is no \(T' \in \tau \) with \(T \cap T' = \varnothing \) such that \(T \leftarrowtail T'\);Footnote 31 thus, \(T \in {{\,\mathrm{\mathsf {LFP}}\,}}\). Therefore, \({\mathcal {C}}_1\subseteq {{\,\mathrm{\mathsf {LFP}}\,}}\). (ii) Otherwise, \(T \in {\mathcal {C}}_2\). Take any \(T' \in \tau \) such that \(T \leftarrowtail T'\). From \(\leftarrowtail \)’s definition, \(T' \cap T = \varnothing \). From \({\mathcal {C}}_2\)’s definition and \(T \cap T' = \varnothing \), there is \(T'' \in {\mathcal {C}}_1\) such that \(T' \cap T'' = \varnothing \). From the previous case, \(T''\) cannot be attacked, i.e., \(T'' \not \leftarrowtail T'\); then, by the first (iii) on Lemma 3, \(T' \leftarrowtail T''\). Summarising, for any \(T' \in \tau \) attacking T, there is \(T'' \in {\mathcal {C}}_1\) attacking \(T'\); hence, \(T \in d({\mathcal {C}}_1) \subseteq d({{\,\mathrm{\mathsf {LFP}}\,}}) = {{\,\mathrm{\mathsf {LFP}}\,}}\), that is, \(T \in {{\,\mathrm{\mathsf {LFP}}\,}}\). Therefore, \({\mathcal {C}}_2\subseteq {{\,\mathrm{\mathsf {LFP}}\,}}\).

\(\varvec{(\subseteq })\) Take now \(T \in \tau \) such that \(T \notin {\mathcal {C}}\); it will be shown that \(T \notin {{\,\mathrm{\mathsf {LFP}}\,}}\). The case with \(T = \varnothing \) is immediate, as \(\varnothing \leftarrowtail \varnothing \). Thus, suppose \(T \ne \varnothing \).

From \(T \notin {\mathcal {C}}\) it follows that \(T \notin {\mathcal {C}}_1\), so there is no \(\phi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\phi \vert \!\} \subseteq T\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\phi \in \varPhi \); hence, from \(\leftarrowtail \)’s definition, every \(T' \in \tau \) with \(T \cap T' = \varnothing \) satisfies that \(T \leftarrowtail T'\).

Note that there is at least one \(T' \in \tau \) with \(T \cap T' = \varnothing \), for suppose otherwise, i.e., suppose \(T' \in \tau \) implies \(T \cap T' \ne \varnothing \). Now, take any \(\varGamma \in W\); since \({\leqslant }[\varGamma ] \in \tau \) (from \({\leqslant }[\varGamma ] \in {\mathcal {E}}_0\) and Item (i) of Fact 2), it follows that \({\leqslant }[\varGamma ] \cap T \ne \varnothing \), i.e., there is \(\varDelta \) with \(\varGamma \leqslant \varDelta \) and \(\varDelta \in T\). Furthermore, take any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(T \subseteq \{\!\vert \varphi \vert \!\}\); by Item (iv) of Fact 2, \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\) and then, from \(\varGamma \leqslant \varDelta \) and \(\varDelta \in \{\!\vert \mathop {\Box }\varphi \vert \!\}\), Proposition 17 gives us \(\varGamma \in \{\!\vert \mathop {\Diamond }\mathop {\Box }\varphi \vert \!\}\), i.e., \(\mathop {\Diamond }\mathop {\Box }\varphi \in \varGamma \). Thus, it has been shown that for any \(\varGamma \in W\) and any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(T \subseteq \{\!\vert \varphi \vert \!\}\), we have \(\mathop {\Diamond }\mathop {\Box }\varphi \in \varGamma \). Then, from Proposition 17 it follows that any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(T \subseteq \{\!\vert \varphi \vert \!\}\) satisfies that \(\mathop {{\text {U}}}\mathop {\Diamond }\mathop {\Box }\varphi \in \varPhi \); thus, axiom \(\mathop {{\text {U}}}\mathop {\Diamond }\mathop {\Box }\varphi \rightarrow \mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \) implies \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \). Moreover, axiom \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {\Box }\varphi \) implies \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\). Thus, for any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\) we have found a formula in \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \), \(\varphi \) itself, such that \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \) and \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\); hence, T is semi-acceptable. But \(T \cap T' \ne \varnothing \) for all \(T'\in \tau \), so there is no \(T' \in \tau \) such that both \(T \cap T' = \varnothing \) and, for any \(T'' \in {\mathcal {C}}_1\), \(T \cap T'' \ne \varnothing \): in other words, T is acceptable. Hence, \(T \in {\mathcal {C}}_2\), i.e., \(T \in {\mathcal {C}}\): a contradiction. Thus, indeed there must be \(T' \in \tau \) such that \(T \cap T' = \varnothing \).

The rest of the proof is divided into two cases: either there is \(T' \in \tau \) with \(T \cap T' = \varnothing \) and \(T' \in {\mathcal {C}}\) (at least one \(T'\) contradicting T is in \({\mathcal {C}}\)), or else for any \(T'\in \tau \) with \(T \cap T' = \varnothing \) we have \(T'\notin {\mathcal {C}}\) (no \(T'\) contradicting T is in \({\mathcal {C}}\)). In both cases, it will be shown that \(T \notin {{\,\mathrm{\mathsf {LFP}}\,}}\).

  1. (i)

    The first case is the simple one: take any \(T' \in \tau \) such that \(T \cap T' = \varnothing \) and \(T' \in {\mathcal {C}}\). Then, as it has been argued, \(T \leftarrowtail T'\); moreover, as it has been proved, \({\mathcal {C}}\subseteq {{\,\mathrm{\mathsf {LFP}}\,}}\). Thus, \(T \notin {{\,\mathrm{\mathsf {LFP}}\,}}\), as \({{\,\mathrm{\mathsf {LFP}}\,}}\) has to be conflict-free.

  2. (ii)

    The second case requires more care. Since no element of \({\mathcal {C}}\) contradicts T, it follows that \(C \in {\mathcal {C}}\) implies \(T \cap C \ne \varnothing \). Now, consider the following two sub-cases: either T is semi-acceptable, or it is not. We will prove that, in both cases, \(T\notin d({\mathcal {C}})\).

    If T is semi-acceptable, recall the initial assumption \(T \notin C = {\mathcal {C}}_1\cup {\mathcal {C}}_2\). Then T cannot be acceptable, that is, there must be \(T'' \in \tau \) such that \(T \cap T'' = \varnothing \) and, for any \(C_1 \in {\mathcal {C}}_1\), we have \(T'' \cap C_1 \ne \varnothing \). Now, take any such \(T''\). From \(T \notin C\), \(T \cap T'' = \varnothing \) and \(T'' \ne \varnothing \) (as its intersection with \(C_1\) is non-empty), it follows that \(T \leftarrowtail T''\): this \(T''\) attacks T. Since \(T'' \cap C_1 \ne \varnothing \) for any \(C_1 \in {\mathcal {C}}_1\), there is no \(C_1 \in {\mathcal {C}}_1\) such that \(T'' \leftarrowtail C_1\). Moreover, there is no \(C_2 \in {\mathcal {C}}_2\) such that \(T'' \leftarrowtail C_2\) – otherwise, \(T'' \cap C_2 = \varnothing \) would imply \(C_2 \notin {\mathcal {C}}_2\). So, in summary, if T is semi-acceptable, then there is \(T'' \in \tau \) such that \(T \leftarrowtail T''\); there is no \(C \in {\mathcal {C}}\) such that \(T'' \leftarrowtail C\). Therefore, \(T \notin d({\mathcal {C}})\).

    If T is not semi-acceptable, there is \(\varphi _T \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(T \subseteq \{\!\vert \mathop {\Box }\varphi _T \vert \!\}\) and there is no \(\psi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that both \(\{\!\vert \mathop {{\mathfrak {T}}}\psi \vert \!\} \subseteq \{\!\vert \mathop {\Box }\varphi _T \vert \!\}\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\psi \in \varPhi \). In particular, \(\varphi _T\) itself cannot be such \(\psi \), so either \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _T \vert \!\} \not \subseteq \{\!\vert \mathop {\Box }\varphi _T \vert \!\}\) or else \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _T \not \in \varPhi \). But axiom \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {\Box }\varphi \) implies \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _T \vert \!\}\subseteq \{\!\vert \mathop {\Box }\varphi _T \vert \!\}\), so \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _T \notin \varPhi \). Now, take any \(C \in {\mathcal {C}}_1\); let \(\varphi _C \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) be one of the formulas satisfying both \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _C \vert \!\} \subseteq C\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _C \in \varPhi \) (by \({\mathcal {C}}\)’s definition, there is at least one). From theorem \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {\Box }\mathop {{\mathfrak {T}}}\varphi \), it follows that \((\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C) \rightarrow (\mathop {\Box }\varphi _T \wedge \mathop {\Box }\mathop {{\mathfrak {T}}}\varphi _C)\) is a theorem too, and thus so are \((\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C) \rightarrow (\mathop {\Box }\mathop {\Box }\varphi _T \wedge \mathop {\Box }\mathop {{\mathfrak {T}}}\varphi _C)\) (by axiom \(\mathop {\Box }\varphi \rightarrow \mathop {\Box }\mathop {\Box }\varphi \)) and \((\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C) \rightarrow \mathop {\Box }(\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C)\) (axiom K for \(\mathop {\Box }\)). Hence, by Proposition 17, \(\mathop {{\text {U}}}\big ( (\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C) \rightarrow \mathop {\Box }(\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C) \big ) \in \varPhi \).

    So far we have \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _T \notin \varPhi \) and, for every \(C \in {\mathcal {C}}_1\), not only \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi _C \in \varPhi \) but also \(\mathop {{\text {U}}}\big ( (\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C) \rightarrow \mathop {\Box }(\mathop {\Box }\varphi _T \wedge \mathop {{\mathfrak {T}}}\varphi _C) \big ) \in \varPhi \). The first and theorem \(\mathop {{\mathfrak {T}}}\varphi \leftrightarrow \mathop {{\mathfrak {T}}}\mathop {\Box }\varphi \) imply \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\mathop {\Box }\varphi _T \notin \varPhi \); the second and axiom \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi \) imply \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\varphi _C \in \varPhi \). These two, the third, and axiom \(\big (\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \wedge \lnot \mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\psi \wedge \mathop {{\text {U}}}((\varphi \wedge \psi ) \rightarrow \mathop {\Box }(\varphi \wedge \psi ))\big ) \rightarrow \mathop {\widehat{{\text {U}}}}\mathop {\Box }(\varphi \wedge \lnot \psi )\) imply \(\mathop {\widehat{{\text {U}}}}\mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _C \wedge \lnot \mathop {\Box }\varphi _T) \in \varPhi \). For the final part, take S to be the union of \(\{\!\vert \mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _C \wedge \lnot \mathop {\Box }\varphi _T) \vert \!\}\) for all \(C \in {\mathcal {C}}_1\), that is,

    $$\begin{aligned} S := \bigcup _{C \in {\mathcal {C}}_1} \{\!\vert \mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _C \wedge \lnot \mathop {\Box }\varphi _T) \vert \!\} \end{aligned}$$

The following two facts about S are crucial for proving \(T \notin d({\mathcal {C}})\):

  1. (a)

    \(S \cap T = \varnothing \). For its proof, from \(T \subseteq \{\!\vert \mathop {\Box }\varphi _T \vert \!\}\) and \(\{\!\vert \mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _C \wedge \lnot \mathop {\Box }\varphi _T) \vert \!\} \subseteq \{\!\vert \lnot \mathop {\Box }\varphi _T \vert \!\}\) for all \(C \in {\mathcal {C}}_1\) (for the latter, use axiom \(\mathop {\Box }\varphi \rightarrow \varphi \)), it follows that \(T \cap \{\!\vert \mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _C \wedge \lnot \mathop {\Box }\varphi _T) \vert \!\} = \varnothing \) for all such C. Therefore, \(S \cap T = \varnothing \).

  2. (b)

    For any \(C' \in {\mathcal {C}}\), we have \(C' \cap S \ne \varnothing \). For its proof, take any \(C' \in {\mathcal {C}}\). If \(C' \in {\mathcal {C}}_1\) then, as \(\{\!\vert \mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _{C'} \wedge \lnot \mathop {\Box }\varphi _T) \vert \!\} \subseteq \{\!\vert \mathop {{\mathfrak {T}}}\varphi _{C'} \vert \!\}\) (same as before) and \(\{\!\vert \mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _{C'} \wedge \lnot \mathop {\Box }\varphi _T) \vert \!\}\ne \varnothing \) (by \(\mathop {\widehat{{\text {U}}}}\mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _C \wedge \lnot \mathop {\Box }\varphi _T) \in \varPhi \)), it follows that \(\{\!\vert \mathop {\Box }(\mathop {{\mathfrak {T}}}\varphi _{C'} \wedge \lnot \mathop {\Box }\varphi _T) \vert \!\} \cap \{\!\vert \mathop {{\mathfrak {T}}}\varphi _{C'} \vert \!\} \ne \varnothing \). But \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi _{C'} \vert \!\} \subseteq C'\) for any \(C' \in {\mathcal {C}}_1\); hence, \(S \cap C' \ne \varnothing \). Otherwise, \(C' \in {\mathcal {C}}_2\) and, for a contradiction, assume \(S \cap C' = \varnothing \). Then, since \(S\cap C''\ne \varnothing \) for all \(C''\in {\mathcal {C}}_1\) as we have proved, there is an open in \(\tau \), S, such that \(S\cap C' = \varnothing \) and \(S\cap C''\ne \varnothing \) for all \(C''\in {\mathcal {C}}_1\). Thus, \(C'\) violates the second condition of acceptance, and hence \(C'\notin {\mathcal {C}}_2\): contradiction. So \(S\cap C' \ne \varnothing \) for any \(C'\in {\mathcal {C}}_2\). Therefore, in sum, \(C' \in {\mathcal {C}}\) implies \(C' \cap S \ne \varnothing \)

Now, since \(S \cap T = \varnothing \) and T is not semi-acceptable (so there is no \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) s.t. both \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq T\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \)), we have found an open S in \(\tau \) with \(T \leftarrowtail S\), according to the definition of \(\leftarrowtail \). But \(S \cap C \ne \varnothing \) for all \(C \in {\mathcal {C}}\), so \(S \not \leftarrowtail C\) for all \(C \in {\mathcal {C}}\): no open in \({\mathcal {C}}\) attacks S. Hence, \(T \notin d({\mathcal {C}})\).

Therefore, regardless of whether T is semi-acceptable or not, we have proved that \(T\notin {\mathcal {C}}\) implies that \(T \notin d({\mathcal {C}})\).

By the fact that \({\mathcal {C}}_1\) is not attacked and \({\mathcal {C}}_2\subseteq d({\mathcal {C}}_1)\), which we have proved in our proof of \({\mathcal {C}}\subseteq {{\,\mathrm{\mathsf {LFP}}\,}}\), it follows that \({\mathcal {C}}\subseteq d({\mathcal {C}})\). Together with \(d({\mathcal {C}})\subseteq {\mathcal {C}}\), it follows that \({\mathcal {C}}= d({\mathcal {C}})\). By the fact that \({\mathcal {C}}\subseteq {{\,\mathrm{\mathsf {LFP}}\,}}\) and \({{\,\mathrm{\mathsf {LFP}}\,}}\) is the least fixed point, it follows that \({\mathcal {C}}= {{\,\mathrm{\mathsf {LFP}}\,}}\).

Thus, in both cases \(T \notin {\mathcal {C}}\) implies \(T \notin {{\,\mathrm{\mathsf {LFP}}\,}}\). This completes the proof. \(\square \)

Proposition 18

(Truth lemma) For any \(\varphi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) and any \(\varGamma \in W\),

Proof

The proof proceeds by induction, with the cases for atomic propositions and Boolean connectives being routine, and those for and \(\mathop {\Box }\) and \(\mathop {{\text {U}}}\) relying on Proposition 17. Here we focus on the case for \(\mathop {{\mathfrak {T}}}\).

From left to right, suppose \(\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\). Then, \(\mathop {{\mathfrak {T}}}\varphi \in \varGamma \) so, by Proposition 17, \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \) which, by Item (iii) of Fact 2, implies \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \in \tau \). Now, let \(T = \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\). Then, (i) from \(\{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\} \subseteq T\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\varphi \in \varPhi \), it follows that \(T \in {\mathcal {C}}_1\) which, by Lemma 5, implies \(T \in {{\,\mathrm{\mathsf {LFP}}\,}}\); (ii) \(\varGamma \in T\), as \(\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\); (iii) from axiom \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \varphi \) it follows that \(T \subseteq \{\!\vert \varphi \vert \!\}\) which, by inductive hypothesis \(\{\!\vert \varphi \vert \!\} = \llbracket \varphi \rrbracket \), implies \(T \subseteq \llbracket \varphi \rrbracket \). Hence, by \(\mathop {{\mathfrak {T}}}\)’s truth condition, \(\varGamma \in \llbracket \mathop {{\mathfrak {T}}}\varphi \rrbracket \).

From right to left, suppose \(\varGamma \in \llbracket \mathop {{\mathfrak {T}}}\varphi \rrbracket \). Then, by \(\mathop {{\mathfrak {T}}}\)’s truth condition, there is \(T \in {{\,\mathrm{\mathsf {LFP}}\,}}\) with \(\varGamma \in T\) and \(T \subseteq \llbracket \varphi \rrbracket \).

The inductive hypothesis implies \(\llbracket \varphi \rrbracket = \{\!\vert \varphi \vert \!\}\). By Item (iv) of Fact 2 and \(T \subseteq \{\!\vert \varphi \vert \!\}\), we have \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\). So we have proved that \(\varGamma \in T\) and \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\).

By Lemma 5, \({{\,\mathrm{\mathsf {LFP}}\,}}= {\mathcal {C}}_1\cup {\mathcal {C}}_2\); thus, \(T \in {\mathcal {C}}_1\cup {\mathcal {C}}_2\). Suppose \(T \in {\mathcal {C}}_1\); then there is \(\psi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(\{\!\vert \mathop {{\mathfrak {T}}}\psi \vert \!\} \subseteq T\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\psi \in \varPhi \). Thus, \(\{\!\vert \mathop {{\mathfrak {T}}}\psi \vert \!\} \subseteq T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\), so \(\mathop {{\text {U}}}(\mathop {{\mathfrak {T}}}\psi \rightarrow \mathop {\Box }\varphi ) \in \varPhi \). Now, take any \(\varDelta \in \{\!\vert \mathop {{\mathfrak {T}}}\psi \vert \!\}\). The fact that \(\mathop {{\text {U}}}(\mathop {{\mathfrak {T}}}\psi \rightarrow \mathop {\Box }\varphi ) \in \varDelta \), together with theorem \(\mathop {{\text {U}}}(\varphi \rightarrow \psi ) \rightarrow (\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {{\mathfrak {T}}}\psi )\) (Proposition 7), implies \(\mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\psi \rightarrow \mathop {{\mathfrak {T}}}\mathop {\Box }\varphi \in \varDelta \). Moreover: \(\varDelta \in \{\!\vert \mathop {{\mathfrak {T}}}\psi \vert \!\}\) implies \(\varDelta \in \{\!\vert \mathop {{\mathfrak {T}}}\mathop {{\mathfrak {T}}}\psi \vert \!\}\), so \(\varDelta \in \{\!\vert \mathop {{\mathfrak {T}}}\mathop {\Box }\varphi \vert \!\}\), that is, \(\mathop {{\mathfrak {T}}}\mathop {\Box }\varphi \in \varDelta \). The latter, together with theorem \(\mathop {{\mathfrak {T}}}\varphi \leftrightarrow \mathop {{\mathfrak {T}}}\mathop {\Box }\varphi \) and axiom \(\mathop {{\mathfrak {T}}}\varphi \rightarrow \mathop {{\text {U}}}(\mathop {\Box }\varphi \rightarrow \mathop {{\mathfrak {T}}}\varphi )\), imply \(\mathop {{\text {U}}}(\mathop {\Box }\varphi \rightarrow \mathop {{\mathfrak {T}}}\varphi ) \in \varDelta \), and thus \(\mathop {{\text {U}}}(\mathop {\Box }\varphi \rightarrow \mathop {{\mathfrak {T}}}\varphi ) \in \varPhi \). Hence, \(\{\!\vert \mathop {\Box }\varphi \vert \!\} \subseteq \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\) and thus, since \(\varGamma \in T\) and \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\), we have \(\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\). Otherwise, \(T \in {\mathcal {C}}_2\), and hence for any \(\psi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(T \subseteq \{\!\vert \mathop {\Box }\psi \vert \!\}\) there is \(\xi \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) with \(\{\!\vert \mathop {{\mathfrak {T}}}\xi \vert \!\} \subseteq \{\!\vert \mathop {\Box }\psi \vert \!\}\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\xi \in \varPhi \). Thus, since \(\varphi \) satisfies that \(T \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\), there is \(\eta \in {\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}} \) such that \(\{\!\vert \mathop {{\mathfrak {T}}}\eta \vert \!\} \subseteq \{\!\vert \mathop {\Box }\varphi \vert \!\}\) and \(\mathop {\widehat{{\text {U}}}}\mathop {{\mathfrak {T}}}\eta \in \varPhi \). From here we can repeat the argument used in the case of \(T \in {\mathcal {C}}_1\) in order to get \(\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\) again. Thus, in both cases, \(\varGamma \in \{\!\vert \mathop {{\mathfrak {T}}}\varphi \vert \!\}\), which completes the proof. \(\square \)

Lemma 6

is Alexandroff.

Proof

Whether is Alexandroff has nothing to do with \(\leftarrowtail \); thus, we can apply Prop. 5.6.15 in Özgün (2017), which states that if \(\tau = \{\bigcup _{\varGamma \in U} {\leqslant }[\varGamma ] \mid U \subseteq W\}\) then is Alexandroff. But Item (i) of Fact 2 and the definition of \({\mathcal {E}}_0\) imply the required condition; then, is Alexandroff. \(\square \)

Since is Alexandroff, Proposition 16 tells us it has a modally equivalent topological argumentation model. Hence, the \({\mathsf {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-consistent set of \({\mathcal {L}}_{\mathop {\Box }, \mathop {{\text {U}}}, \mathop {{\mathfrak {T}}}}\)-formulas \(\varPhi _0\) is satisfiable in a topological argumentation model which satisfies the following condition: for every \(T, T_1, T'_1 \in \tau \): if \(T_1 \leftarrowtail T\) and \(T'_1 \subseteq T_1\), then \(T'_1 \leftarrowtail T\) (see Lemma 3.2).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shi, C., Smets, S. & Velázquez-Quesada, F.R. Logic of Justified Beliefs Based on Argumentation. Erkenn 88, 1207–1243 (2023). https://doi.org/10.1007/s10670-021-00399-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10670-021-00399-5

Keywords

Navigation