Skip to main content

A Stit Logic of Intentionality

  • Conference paper
  • First Online:

Part of the book series: Lecture Notes in Computer Science ((LNCS,volume 13780))

Abstract

We extend epistemic stit theory with a modality \(I_\alpha \varphi \), meant to express that at some moment agent \(\alpha \) had a present-directed intention toward the realization of \(\varphi \). The semantics is based on the extension of stit frames with special topologies associated to agents. The open sets of the associated topology are interpreted as present-directed intentions, that support whether an agent had an intention of realizing a specific state of affairs when it chose one of its available actions and executed it. As an important application, we use \(I_\alpha \varphi \) to formalize intentional action and intentional responsibility. We present an axiom system for our logic of intentionality, and prove that it is sound and complete.

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

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   44.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   59.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

Notes

  1. 1.

    An open set is dense in a topology iff it is consistent with all the other non-empty open sets of the topology (see Definition 2).

  2. 2.

    Of course, this reading of the modality \(I_\alpha \varphi \) and of the conjunction \([\alpha ]\varphi \wedge I_\alpha [\alpha ]\varphi \) positions our proposal as belonging to a particular philosophical standpoint on the relation between intentions and intentional action—in the context of the discussion, on the trends in philosophy of intention, at the beginning of the present section. We address the details of such a standpoint in Subsect. 3.

  3. 3.

    Observe that \(\alpha \)’s ex ante knowledge, at a given index, is itself a \(\text{ p-d }\) intention of \(\alpha \), as witnessed by the fact that, since \(\tau _\alpha ^{\langle m,h \rangle }\) is a topology on \(\alpha \)’s ex ante information set, then such an information set must be an element of the topology.

  4. 4.

    As pointed out by [29] in his lecture notes for a course on neighborhood semantics, “[s]ets paired with a distinguished collections of subsets are ubiquitous in many areas of mathematics.”.

  5. 5.

    As implied by the discussion on the problems in philosophy of intention in the introduction, it is not easy to model the relation between the concepts of future-directed intentions, intentional action, and intention-with-which. In instantaneous stit theory—without temporal modalities—such a relation is even harder to address. The topological semantics for \(\text{ p-d }\) intentions, then, helped me establish one such relation in clear, tractable terms.

  6. 6.

    The reader might wonder why we did not opt for a relational semantics for intentionality. As mentioned here, we think that the topological representation is better suited to the notion of \(\text{ p-d }\) intentions than potential relational representations. To be sure, the relational paradigms that we explored before choosing topology involved a more convoluted truth condition for \(I_\alpha \varphi \), as well as a less clear-cut formulation of my models’ constraints. Moreover, topological semantics generalize standard relational semantics and are more expressive. In the words of [28], “topological spaces are equipped with well-studied basic operators such as the interior and closure operators which, alone or in combination with each other, succinctly interpret different modalities, giving a better understanding of their axiomatic properties.”.

  7. 7.

    According to [9], it is clear that an agent who intends to perform an action usually does not intend all the consequences of that action, or even all the consequences that the agent anticipates. Some of the consequences are indeed goals of the agent, while others are ‘side effects’ that the agent is not committed to. The typical example supporting this view involves an agent intending to go to get his tooth filled at the dentist. Being uninformed about anesthetics, the agent believes that the process of having his tooth filled will necessarily cause him much pain. Although the agent intends to ask the dentist to fill his tooth, and, believing what he does, he is willing to put up with pain, the agent would surely deny that he intends to be in pain [14].

  8. 8.

    Most versions of the side-effect problem (see, for instance [9, 13, 14], and [30]) only argue that intentions should not be closed under believed (or anticipated) side effects, in the sense that if an agent intends \(\varphi \) and also believes that \(\varphi \rightarrow \psi \), then one should not conclude that the agent also intends \(\psi \). Indeed, it is unclear to me why the side-effect problem is sometimes assumed to refer to closure of intentions under logical consequence [17, 23]. Logical consequence is a very strong assumption for most logic-based models, since it means that \(\varphi \rightarrow \psi \) holds at all states. For logically ideal agents as the ones here presupposed, this implies that at every possible configuration of the world every agent knows for sure (with absolute, indefeasible certainty) that \(\varphi \rightarrow \psi \) will hold (at every possible configuration of the world). Recall once again the dentist example, and suppose that we phrase it in terms of logical consequence. Now pain is a necessary consequence of getting one’s tooth filled at every possible configuration of the world, and the agent not only believes that the process of having his tooth filled will cause him pain, but he also knows for sure that there is no possible configuration of the world in which he will not feel pain by getting his tooth filled. Can we really say that the agent intended to get his tooth filled without intending pain? Most likely, this is why [25, 30], and [7], for instance, all disregard that intentions should be not closed under logical consequence. As for approaches that manage to yield logics of intending where intentions are not closed under logical consequence, it is either the case that non-closure is only possible when side effects are already known/believed [13, 14] or the case that the logics come with different problems arising from the use of non-normal operators [17, 23].

  9. 9.

    Philosophers might then advance the criticism that some side effects can be deliberative.

  10. 10.

    [17] explicitly states that there is a distinction between intending and intending to do. He writes: “[t]here are two different types of future-directed intentions: I can intend to perform a certain action, or I can intend to realize a certain state of affairs.”.

  11. 11.

    For \(A\subseteq X\) and \(\tau \) a topology on X, the subspace topology of \(\tau \) on A is the family \(\left\{ U\cap A \ | U\in \tau \right\} \).

References

  1. Ramírez Abarca, A.I., Broersen, J.: A logic of objective and subjective oughts. In: Calimeri, F., Leone, N., Manna, M. (eds.) JELIA 2019. LNCS (LNAI), vol. 11468, pp. 629–641. Springer, Cham (2019). https://doi.org/10.1007/978-3-030-19570-0_41

    Chapter  Google Scholar 

  2. Abarca, A.I.R., Broersen, J.: Stit semantics for epistemic notions based on information disclosure in interactive settings. In: Soares Barbosa, L., Baltag, A. (eds.) DALI 2019. LNCS, vol. 12005, pp. 171–189. Springer, Cham (2020). https://doi.org/10.1007/978-3-030-38808-9_11

    Chapter  Google Scholar 

  3. Anscombe, G.E.M.: Intention. Harvard University Press, Cambridge (1963)

    Google Scholar 

  4. Baltag, A., Bezhanishvili, N., Özgün, A., Smets, S.: The topology of full and weak belief. In: Hansen, H.H., Murray, S.E., Sadrzadeh, M., Zeevat, H. (eds.) TbiLLC 2015. LNCS, vol. 10148, pp. 205–228. Springer, Heidelberg (2017). https://doi.org/10.1007/978-3-662-54332-0_12

    Chapter  Google Scholar 

  5. Baltag, A., Bezhanishvili, N., Özgün, A., Smets, S.: Justified belief and the topology of evidence. In: Väänänen, J., Hirvonen, Å., de Queiroz, R. (eds.) WoLLIC 2016. LNCS, vol. 9803, pp. 83–103. Springer, Heidelberg (2016). https://doi.org/10.1007/978-3-662-52921-8_6

    Chapter  Google Scholar 

  6. Belnap, N., Perloff, M., Xu, M.: Facing the Future: Agents and Choices in Our Indeterminist World. Oxford University Press, Oxford (2001)

    Google Scholar 

  7. Bentzen, M.M.: Stit, Iit, and deontic logic for action types. Ph.D. thesis, University of Amsterdam (2012)

    Google Scholar 

  8. Bratman, M.: Two faces of intention. Philos. Rev. 93(3), 375–405 (1984)

    Article  Google Scholar 

  9. Bratman, M.: Intention, Plans, and Practical Reason. Harvard University Press, Cambridge (1987)

    Google Scholar 

  10. Broersen, J.: A complete STIT logic for knowledge and action, and some of its applications. In: Baldoni, M., Son, T.C., van Riemsdijk, M.B., Winikoff, M. (eds.) DALT 2008. LNCS (LNAI), vol. 5397, pp. 47–59. Springer, Heidelberg (2009). https://doi.org/10.1007/978-3-540-93920-7_4

    Chapter  Google Scholar 

  11. Broersen, J.: A stit-logic for extensive form group strategies. In: 2009 IEEE/WIC/ACM International Joint Conference on Web Intelligence and Intelligent Agent Technology, vol. 3, pp. 484–487. IEEE (2009)

    Google Scholar 

  12. Broersen, J.: Deontic epistemic stit logic distinguishing modes of mens rea. J. Appl. Log. 9(2), 137–152 (2011)

    Article  Google Scholar 

  13. Broersen, J.: Making a start with the stit logic analysis of intentional action. J. Philos. Log. 40(4), 499–530 (2011)

    Article  Google Scholar 

  14. Cohen, P.R., Levesque, H.J.: Intention is choice with commitment. Artif. Intell. 42(2–3), 213–261 (1990)

    Article  Google Scholar 

  15. Davidson, D.: Intending. In: Yovel, Y. (ed.) Philosophy of History and Action, pp. 41–60. Springer, Dordrecht (1978). https://doi.org/10.1007/978-94-009-9365-5_5

    Chapter  Google Scholar 

  16. Davidson, D.: Essays on Actions and Events. Clarendon Press, Oxford (1980)

    Google Scholar 

  17. Duijf, H.: Let’s do it!: collective responsibility, joint action, and participation. Ph.D. thesis, Utrecht University (2018)

    Google Scholar 

  18. Dunin-Keplicz, B., Verbrugge, R.: Collective intentions. Fund. Inform. 51(3), 271–295 (2002)

    Google Scholar 

  19. Engelking, R.: General Topology. Heldermann (1989)

    Google Scholar 

  20. Green, S.P.: Six senses of strict liability: a plea for formalism. Appraising Strict Liability 1 (2005)

    Google Scholar 

  21. Herzig, A., Longin, D.: C &L intention revisited. KR 4, 527–535 (2004)

    Google Scholar 

  22. Horty, J.F., Belnap, N.: The deliberative stit: a study of action, omission, ability, and obligation. J. Philos. Log. 24(6), 583–644 (1995)

    Article  Google Scholar 

  23. Konolige, K., Pollack, M.E.: A representationalist theory of intention. In: IJCAI, vol. 93, pp. 390–395 (1993)

    Google Scholar 

  24. Larkin, P.J., Jr.: Strict liability offenses, incarceration, and the cruel and unusual punishments clause. Harvard J. Law Public Policy 37(3), 1065 (2014)

    Google Scholar 

  25. Lorini, E., Herzig, A.: A logic of intention and attempt. Synthese 163(1), 45–77 (2008)

    Article  Google Scholar 

  26. Lorini, E., Longin, D., Mayor, E.: A logical analysis of responsibility attribution: emotions, individuals and collectives. J. Log. Comput. 24(6), 1313–1339 (2014)

    Article  Google Scholar 

  27. Moran, R., Stone, M.J.: Anscombe on expression of intention. In: Sandis, C. (ed.) New Essays on the Explanation of Action, pp. 132–168. Springer, London (2009). https://doi.org/10.1057/9780230582972_9

    Chapter  Google Scholar 

  28. Özgün, A.: Evidence in epistemic logic: a topological perspective. Ph.D. thesis, Université de Lorraine (2017)

    Google Scholar 

  29. Pacuit, E.: Neighborhood semantics for modal logic: an introduction (2007). Course Notes for ESSLLI

    Google Scholar 

  30. Rao, A.S., Georgeff, M.P.: Modeling rational agents within a BDI-architecture. KR, 91, 473–484 (1991)

    Google Scholar 

  31. Setiya, K.: Knowledge of intention. In: Essays on Anscombe’s Intention, pp. 170–197. Harvard University Press (2011)

    Google Scholar 

  32. Setiya, K.: Intention. In: Zalta, E.N. (ed.) The Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University, Fall 2018 edn. (2018)

    Google Scholar 

  33. Thompson, M.: Life and Action. Harvard University Press, Cambridge (2008)

    Book  Google Scholar 

  34. Willard, S.: General Topology. Courier Corporation (2004)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Aldo Iván Ramírez Abarca .

Editor information

Editors and Affiliations

Appendix A Proofs of Soundness and Completeness

Appendix A Proofs of Soundness and Completeness

1.1 A.1 Soundness

Proposition 1

The system \({\varLambda _I}\) is sound with respect to the class of iebt-models.

Proof

The proof of soundness is routine: the validity of (SET) and (IA) is standard from BST; the validity of (OAC) and \((Unif-H)\) is shown exactly as [1]; the validity of (InN) follows straightforwardly from Definitions 4 and 5; and the validity of (KI) follows from frame condition \((\texttt{KI})\).

1.2 A.2 Completeness

Definition 7

(Alexandrov spaces). A topological space \((X,\tau )\) is said to be an Alexandrov space iff the intersection of any collection of open sets of X is an open set as well.

Notice that a space is Alexandrov iff every point \(x\in X\) has a \(\subseteq \)-smallest open set including it, namely the intersection of all the open sets around x.

Definition 8

For a given frame (XR) such that R is reflexive and transitive, a set \(A\subseteq X\) is called upward-closed iff for \(x\in A\), if \(x\le y\) for some \(y\in X\), then \(y\in A\) as well. For \(x\in X\), \(x\uparrow _R\) denotes the set \(\left\{ y\in X \ | \ x R y\right\} \), which is clearly upward closed.

Remark 2

For a frame (XR) such that R is reflexive and transitive, the set of all R-upward-closed sets forms an Alexandrov topology on X, which will be denoted by \(\tau _R\). For \(x\in X\), the \(\subseteq \)-smallest open set including x is precisely \(x\uparrow _R\). This implies that \(\left\{ x\uparrow _\le \ | \ x\in X\right\} \) is a basis for the topology \(\tau _R\).

Definition 9

(Kripke- ies -frames & models).

A tuple

$$\begin{aligned} \left\langle W, Ags, R_\square , Ags, \texttt{Choice}, \left\{ \mathtt {\approx }_\alpha \right\} _{\alpha \in Ags},\left\{ R_\alpha ^I\right\} _{\alpha \in Ags}\right\rangle \end{aligned}$$

is called a Kripke-ies-frame (where the acronym ‘ies’ stands for ‘epistemic intentional stit’) iff

  • W is a set of possible worlds. \(R_\square \) is an equivalence relation over W. For \(w\in W\), the class of w under \(R_\square \) is denoted by \(\overline{w}\).

  • \(\texttt{Choice}\) is a function that assigns to each \(\alpha \in Ags\) and each \(\square \)-class \(\overline{w}\) a partition \(\texttt{Choice}_\alpha ^{\overline{w}}\) of \(\overline{w}\) given by an equivalence relation, denoted by \(R_\alpha ^{\overline{w}}\). \(\texttt{Choice}\) must satisfy the following constraint:

    • \(\mathtt {(IA)_K}\) For \(w\in W\), each function \(s:Ags\rightarrow 2^{\overline{w}}\) that maps \(\alpha \) to a member of \(\texttt{Choice}^{\overline{w}}_\alpha \) is such that \(\bigcap _{\alpha \in Ags} s(\alpha ) \ne \emptyset \).

    For \(\alpha \in Ags\), \(w\in W\), and \(v\in \overline{w}\), the class of v in the partition \(\texttt{Choice}^{\overline{w}}_\alpha \) is denoted by \(\texttt{Choice}^{\overline{w}}_\alpha (v)\).

  • For \(\alpha \in Ags\), \(\approx _\alpha \) is an (epistemic) equivalence relation on W that satisfies the following conditions:

    • \(\mathtt {(OAC)_{\texttt{K}}}\) For \(w\in W\) and \(v\in \overline{w}\), \(v\approx _\alpha u\) for every \(u\in \texttt{Choice}^{\overline{w}}_\alpha (v)\).

    • \(\mathtt {(Unif-H)_{\texttt{K}}}\) Let \(\alpha \in Ags\) and \(v, u\in W\) such that \(v\approx _\alpha u\). For \(v'\in \overline{v}\), there exists \(u'\in \overline{u}\) such that \(v'\approx _\alpha u'\).

    For \(w, v\in W\) such that \(w\approx _\alpha v\) and \(L\subseteq \overline{w}\), L’s epistemic cluster at \(\overline{v}\) is the set \([\![L]\!]_\alpha ^{\overline{v}}:=\left\{ u\in {\overline{v}} ; \text{ there } \text{ is } o\in L \ \text { such that }\ o \approx _\alpha u \right\} .\)

    For \(\alpha \in Ags\) and \(w\in W\), \(\alpha \)’s ex ante information set at w is defined as \(\pi _\alpha ^\square [w]:=\left\{ v; w\approx _\alpha \circ R_\square v \right\} \), which by frame condition \((\mathtt {Unif-H})_\texttt{K}\) coincides with the set \(\left\{ v; w R_\square \circ \approx _\alpha v \right\} \). To clarify, \((\mathtt {Unif-H})_\texttt{K}\) implies that \(R_\square \circ \approx _\alpha = \approx _\alpha \circ R_\square \). Thus \(\approx _\alpha \circ R_\square \) is an equivalence relation such that \(\pi _\alpha ^\square [w]=\pi _\alpha ^\square [v]\) for every \(w,v \in W\) such that \(w\approx _\alpha \circ R_\square v\).

  • For \(\alpha \in Ags\), \(R_\alpha ^I\) is a serial, transitive, and euclidean relation on W such that \(R_\alpha ^I\subseteq \approx _\alpha \circ R_\square \) and such that the following condition is satisfied:

    • \(\mathtt {(Den)_K}\) For \(v, u\in W\) such that \(v \approx _\alpha \circ R_\square u\), there exists \(u'\in W\) such that \(vR_\alpha ^I u'\) and \(uR_\alpha ^I u'\).

    For \(\alpha \in Ags\), \(R_\alpha ^{I+}\) denotes the reflexive closure of \(R_\alpha ^I\).

A Kripke-ies-model \(\mathcal {M}\) consists of the tuple that results from adding a valuation function \(\mathcal {V}\) to a Kripke-ies-frame, where \(\mathcal {V}: P\rightarrow 2^{W}\) assigns to each atomic proposition a set of worlds.

Kripke-ies-models allow us to evaluate the formulas of \(\mathcal L_{\textsf {I}}\) with semantics that are analogous to the ones provided for iebt-frames. The semantics for the formulas of \(\mathcal L_{\textsf {I}}\) are given in the definition below.

Definition 10

(Evaluation rules on Kripke models). Let \(\mathcal {M}\) be a Kripke-ies-model. The semantics on \(\mathcal {M}\) for the formulas of \(\mathcal {L}_{\textsf {I}}\) are defined recursively by the following truth conditions, evaluated at a given world w:

$$ \begin{array}{lll} \mathcal {M}, w \models p &{} \text{ iff } &{} w \in \mathcal {V}(p) \\ \mathcal {M}, w \models \lnot \varphi &{} \text{ iff } &{} \mathcal {M}, w \not \models \varphi \\ \mathcal {M}, w \models \varphi \wedge \psi &{} \text{ iff } &{} \mathcal {M}, w \models \varphi \text{ and } \mathcal {M}, w \models \psi \\ \mathcal {M}, w \models \Box \varphi &{} \text{ iff } &{} \text{ for } v\in \overline{w},\mathcal {M}, v \models \varphi \\ \mathcal {M}, w \models [\alpha ] \varphi &{} \text{ iff } &{} \text{ for } v\in \texttt{Choice}^{\overline{w}}_\alpha (w), \mathcal {M}, v \models \varphi \\ \mathcal {M}, w \models K_{\alpha } \varphi &{} \text{ iff } &{} \text{ for } v \text{ s.t. } w \approx _{\alpha }v, \mathcal {M}, v \models \varphi \\ \mathcal {M}, w \models I_\alpha \varphi &{} \text{ iff } &{} \text{ there } \text{ exists } x \in \pi _\alpha ^\square [w] \text{ s.t. } x\uparrow _{R_\alpha ^{I+}} \subseteq |\varphi |. \end{array} $$

where we write \(|\varphi |\) to refer to the set \(\left\{ w\in W;\mathcal {M},w \models \varphi \right\} \). Satisfiability, validity on a frame, and general validity are defined as usual.

Definition 11

(Associated iebt -frame).

Let

$$\mathcal {F}=\left\langle W, Ags, R_\square , \texttt{Choice}, \left\{ \mathtt {\approx }_\alpha \right\} _{\alpha \in Ags}, \left\{ R_\alpha ^I\right\} _{\alpha \in Ags} \right\rangle $$

be a Kripke-ies-frame. Then \( \mathcal {F}^T:=\left\langle M_W, \sqsubset , Ags, \textbf{Choice},\left\{ \sim _\alpha \right\} _{\alpha \in Ags}, \tau \right\rangle \) is called the iebt-frame associated to \(\mathcal {F}\) iff

  • \(M_W:={W}\cup \left\{ \overline{w} ; w\in W \right\} \cup \left\{ W\right\} \), and \(\sqsubset \) is a relation on \(M_W\) such that \(\sqsubset \) is defined as the transitive closure of the union \(\left\{ (\overline{w},v) ; w\in W \text{ and } v\in \overline{w} \right\} \cup \left\{ (W,\overline{w}) ; w\in W\right\} \).

    It is clear that \(\sqsubset \) is a strict partial order on \(M_W\) that satisfies “no backward branching” straightforwardly. Since the tuple \(\left\langle M_W, \sqsubset \right\rangle \) is thus a tree, let us refer to the maximal \(\sqsubset \)-chains in \(M_W\) as histories, and let us denote by \(H_W\) the set of all histories of \(M_W\). Observe that the definition of \(\sqsubset \) yields that there is a bijective correspondence W and \(H_W\). For \(v\in W\), let \(h_v\) be the history \(\left\{ W, \overline{v},v\right\} \). For \(o\in W\), it is clear that \(o\in h_v\) iff \(o=v\). Therefore, each history in \(H_W\) can be identified using the world at its terminal node. Consequently, for \(w\in W\), if \(H_{\overline{w}}\) denotes the set of histories passing through \(\overline{w}\), then \( H_{\overline{w}}=\left\{ h_v ; v\in \overline{w}\right\} \)—since \(\overline{w}\in h_v\) iff \(v\in \overline{w}\). Observe, then, that \(H_W=\left\{ h_v ; v\in W\right\} \).

  • For \(B\in 2^{W}\), let \(B^T\) denote the set \(\left\{ h_v ; \ v\in B\right\} \). With such a terminology, we define \(\textbf{Choice}\) as a function on \(Ags\times M_W\) given by the rules:

    • For \(\alpha \in Ags\) and \(v\in W\), \(\textbf{Choice}(\alpha ,v)=\left\{ \left\{ h_v\right\} \right\} \).

    • For \(\alpha \in Ags\) and \(w\in W\), \(\textbf{Choice}(\alpha , \overline{w})=\left\{ C_\alpha ^T ; C_\alpha \in \texttt{Choice}_\alpha ^{\overline{w}}\right\} \).

    • For \(\alpha \in Ags\), \(\textbf{Choice}(\alpha , W)=\left\{ H_W\right\} \).

    To keep notation consistent, the sets of the form \(\textbf{Choice}(\alpha , \overline{w})\) will be denoted by \(\textbf{Choice}_\alpha ^{\overline{w}}\). The choice-cell of a given \(h_v\) in \(\textbf{Choice}_\alpha ^{\overline{w}}\) is denoted by \(\textbf{Choice}_\alpha ^{\overline{w}}(h_v)\).

  • For \(\alpha \in Ags\), \(\thicksim _\alpha \) is a relation on \(I\left( M_W\times H_W\right) \) defined as follows:

    $$\begin{array}{lll} \sim _\alpha &{}:= &{} \left\{ \left( \left\langle \overline{w},h_v\right\rangle , \left\langle \overline{w'},h_{v'}\right\rangle \right) ; w, w'\in W \text{ and } v\approx _\alpha v'\right\} \cup \\ &{}&{} \left\{ \left( \left\langle z,h_z\right\rangle ,\left\langle z,h_z\right\rangle \right) ; z\in W\right\} \cup \\ &{}&{}\left\{ \left( \left\langle W,h_v\right\rangle ,\left\langle W,h_{v'}\right\rangle \right) ; v,v'\in W\right\} . \end{array}$$

    It is clear that this definition entails that \(\sim _\alpha \) is an equivalence relation for every \(\alpha \in Ags\) and that, for \(w\in W\) and \(L\in \texttt{Choice}_\alpha ^{\overline{w}}\), \(v\in [\![L]\!]_\alpha ^{\overline{w}}\) iff \(h_v\in \left[ L^T\right] ^{\overline{w}}_\alpha \).

  • \(\tau \) is a function defined as follows:

    • For \(\alpha \in Ags\) and \(z\in W\), \(\tau _\alpha ^{\left\langle z, h_z \right\rangle }= \left\{ \emptyset , \pi _\alpha ^\square \left[ \left\langle z, h_z \right\rangle \right] \right\} \)

    • For \(\alpha \in Ags\), we first define a relation \(R_\alpha ^{IT}\) on \(\left\{ \left\langle \overline{w},h_v\right\rangle ; w\in W \text{ and } v\in \overline{w}\right\} \) by the rule: \(\left\langle \overline{w},h_v\right\rangle R_\alpha ^{IT} \left\langle \overline{w'},h_{v'}\right\rangle \) iff \(vR_\alpha ^I v'\). For \(\alpha \in Ags\), \(w\in W\), and \(v\in \overline{w}\), then, we define \(\tau _\alpha ^{\left\langle \overline{w}, h_v \right\rangle }\) as the subspace topology of \(\tau _{R_\alpha ^{IT+}}\) on \(\pi _\alpha ^\square \left[ \left\langle \overline{w}, h_v \right\rangle \right] \).Footnote 11 Observe that, for \(\alpha \in Ags\), \(w\in W\), and \(v\in \overline{w}\), \(\pi _\alpha ^\square \left[ \left\langle \overline{w}, h_v\right\rangle \right] =\left\{ \left\langle \overline{v'}, h_{v'}\right\rangle ; v'\in \pi _\alpha ^\square [v]\right\} \). Thus, the fact that \(R_\alpha ^I\subseteq \approx _\alpha \circ R_\square \) implies that, for \(\left\langle \overline{x}, h_{x}\right\rangle \in \pi _\alpha ^\square \left[ \left\langle \overline{w}, h_v \right\rangle \right] \), \(\left\langle \overline{x}, h_{x}\right\rangle \uparrow _{R_\alpha ^{IT+}}\subseteq \pi _\alpha ^\square \left[ \left\langle \overline{w}, h_v \right\rangle \right] \), so that \(\pi _\alpha ^\square \left[ \left\langle \overline{w}, h_v \right\rangle \right] \) is open in \(\tau _{R_\alpha ^{IT+}}\).

    • For \(\alpha \in Ags\) and \(v\in W\), \(\tau _\alpha ^{\left\langle W, h_v \right\rangle }= \left\{ \emptyset , \pi _\alpha ^\square \left[ \left\langle W, h_v \right\rangle \right] \right\} \).

Proposition 2

Let \(\mathcal {F}\) be a Kripke-ies-frame. Then \(\mathcal {F}^T\) is an iebt-frame.

Proof

It amounts to showing that \(\sqsubset \) is a strict partial order that satisfies no backward branching, that \(\textbf{Choice}\) is a function that satisfies frame conditions \((\texttt{NC})\) and \((\texttt{IA})\), that \(\left\{ \sim _\alpha \right\} _{\alpha \in Ags}\) is such that \(\approx _\alpha \) is an equivalence relation for every \(\alpha \in Ags\) and frame conditions \((\texttt{OAC})\) and \((\mathtt {Unif-H})\) are met, and that \(\tau \) is a function that meets the requirements of Definition 4.

  • As mentioned in Definition 11, it is straightforward to show that \(\sqsubset \) is a strict partial order that satisfies no backward branching. It is also clear from Definition 11 that \(\sim _\alpha \) is an equivalence relation for every \(\alpha \in Ags\).

  • \((\texttt{NC})\) is vacuously validated at moment W. It is validated in moments of the form \(\overline{w}\) (\(w\in W\)), since two different histories never intersect in a moment later than \(\overline{w}\). Finally, it is also validated in moments of the form v such that \(v\in W\) (since there are no moments above v).

  • For \((\texttt{IA})\), we reason by cases:

    1. (a)

      At moment W, \((\texttt{IA})\) is validated straightforwardly, since \(\textbf{Choice}(\alpha , W)=\left\{ H\right\} \) for each \(\alpha \in Ags\).

    2. (b)

      For a moment of the form \(\overline{w}\) (with \(w\in W\)), let s be a function that assigns to each agent \(\alpha \) a member of \(\textbf{Choice}^{\overline{w}}_\alpha =\left\{ (C_\alpha )^T; C_\alpha \in \texttt{Choice}^{\overline{w}}_\alpha \right\} \). Let \(s_k:Ags \rightarrow \bigcup _{\alpha \in Ags}\texttt{Choice}^{\overline{w}}_\alpha \) be a function such that \(s_k(\alpha )=C_\alpha \) iff \(s(\alpha )=(C_\alpha )^T\). Since \(\mathcal {M} \) satisfies condition \((\texttt{IA})_{\texttt{K}}\), then \(\bigcap _{\alpha \in Ags} s_k(\alpha ) \ne \emptyset \). Take \(v\in \bigcap _{\alpha \in Ags} s_k(\alpha )\). Then \(v\in C_\alpha \) for every \(\alpha \in Ags\). This implies that \(h_v\in (C_\alpha )^T\) for every \(\alpha \in Ags\), so \(\bigcap _{\alpha \in Ags} s(\alpha ) \ne \emptyset \).

    3. (c)

      At moments of the form v such that \(v\in W\), if s is a function that assigns to each agent \(\alpha \) a member of \(\textbf{Choice}(v,\alpha )\), s must be constant and \(\bigcap _{\alpha \in Ags} s(\alpha ) = \left\{ h_v\right\} \).

  • For \((\texttt{OAC})\), again we reason by cases:

    1. (a)

      Assume that \(\left\langle \overline{w}, h_v \right\rangle \sim _\alpha \left\langle \overline{w'}, h_{v'} \right\rangle \) (for \(w, w' \in W\)). This means that \(v\approx _\alpha v'\). We want to show that, for every \(h_u\in \textbf{Choice}^{\overline{w}}_\alpha \) such that \(h_u\in \textbf{Choice}^{\overline{w}}_\alpha (h_v)\), \(\left\langle \overline{w}, h_u \right\rangle \sim _\alpha \left\langle \overline{w'}, h_{v'} \right\rangle \). Therefore, let \(h_u\in \textbf{Choice}^{\overline{w}}_\alpha (h_v)\). By definition, this means that \(u\in \texttt{Choice}^{\overline{w}}_\alpha (v)\). Since \(\mathcal {M} \) satisfies condition \((\texttt{OAC})_{\texttt{K}}\), this last fact implies, with \(v\approx _\alpha v'\), that \(u\approx _\alpha v'\), which in turn yields that \(\left\langle \overline{w}, h_u \right\rangle \sim _\alpha \left\langle \overline{w'}, h_v \right\rangle \).

    2. (b)

      For indices based on moments of the form v such that \(v\in W\), \((\texttt{OAC})\) is met straightforwardly, since for \(h_v\) the choice-cell in \(\textbf{Choice}(\alpha ,v)\) to which \(h_v\) belongs is just \(\left\{ h_v\right\} \).

    3. (c)

      For indices based on moment W, \((\texttt{OAC})\) is also met straightforwardly, since for every \(\alpha \in Ags\) \(\sim _\alpha \) is defined such that \(\left\langle W,h_v\right\rangle \sim _\alpha \left\langle W, h_{v'}\right\rangle \) for every pair of histories \(h_v, h_{v'}\) in H.

  • For \((\mathtt {Unif-H})\), again we reason by cases:

    1. (a)

      Assume that \(\left\langle \overline{w}, h_v \right\rangle \sim _\alpha \left\langle \overline{w'}, h_{v'} \right\rangle \) (for \(w, w' \in W\)). This means that \(v\in \overline{w}\), \(v'\in \overline{w'}\), and \(v\approx _\alpha v'\). Let \(h_{z}\in H_{\overline{w}}\) (which means that \(z\in \overline{w}\)). We want to show that there exists \(h\in H_{\overline{w'}}\) such that \(\left\langle \overline{w}, h_z \right\rangle \sim _\alpha \left\langle \overline{w'}, h \right\rangle \). Condition \((Unif-H)_{\texttt{K}}\) gives us that there exists \(z'\in \overline{w'}\) such that \(z\approx _\alpha z'\), which by definition of \(\sim _\alpha \) means that \(\left\langle \overline{w}, h_z \right\rangle \sim _\alpha \left\langle \overline{w'}, h_{z'} \right\rangle \).

    2. (b)

      For indices based on moments of the form v such that \(v\in W\), \(\left\langle v, h_v \right\rangle \) is \(\sim _\alpha \)-related only to itself, so \((\mathtt {Unif-H})\) is met straightforwardly.

    3. (c)

      For indices based on moment W, \((\mathtt {Unif-H})\) is also met straightforwardly, since \(\left\langle W,h_v\right\rangle \sim _\alpha \left\langle W,h_{v'}\right\rangle \) for every \(v,v'\in W\).

  • As for \(\tau \), it is clear that, for \(\alpha \in Ags\) and index \(\langle m, h\rangle \) either of the form \(\left\langle z, h_z \right\rangle (z\in W)\) or of the form \(\left\langle W, h_v \right\rangle (v\in W)\), \(\tau _\alpha ^{\langle m, h \rangle }\) is a topology on \(\pi _\alpha ^\square [\left\langle m, h \right\rangle ]\) that satisfies frame conditions \((\texttt{CI})\) and \((\texttt{KI})\).

    Assume, then, that \(\langle m, h\rangle \) is of the form \(\left\langle \overline{w}, h_v \right\rangle \) such that \(v\in \overline{w}\). Let \(\alpha \in Ags\). By Definition 11, \(\tau _\alpha ^{\left\langle \overline{w}, h_v \right\rangle }\) is the subspace topology of \(\tau _{R_\alpha ^{IT+}}\) on \(\pi _\alpha ^\square [\left\langle \overline{w}, h_v \right\rangle ]\). Thus, it is clear that \(\tau _\alpha ^{\left\langle \overline{w}, h_v \right\rangle }\) is a topology on \(\pi _\alpha ^\square [\left\langle \overline{w}, h_v \right\rangle ]\), so that \(\tau \) straightforwardly satisfies \((\texttt{KI})\). Let us show that condition \((\texttt{CI})\) is also satisfied: let \(U,V\in \tau _\alpha ^{\left\langle \overline{w}, h_v \right\rangle } \) such that U and V are non-empty. Let \(\left\langle \overline{u}, h_u \right\rangle \in U\) and \( \left\langle \overline{x}, h_x \right\rangle \in V\). Definition 11 implies that \(u\approx _\alpha \circ R_\square x\). \(\mathcal {F}\)’s condition \(\mathtt {(Den)_K}\) implies that there exists \(u'\in W\) such that \(uR_\alpha ^I u'\) and \(xR_\alpha ^I u'\), which implies that \(\left\langle \overline{u}, h_u\right\rangle R_\alpha ^{IT} \left\langle \overline{u'}, h_{u'}\right\rangle \) and \(\left\langle \overline{x}, h_x\right\rangle R_\alpha ^{IT} \left\langle \overline{u'}, h_{u'}\right\rangle \), by definition of \(R_\alpha ^{IT}\). This means that \(\left\langle \overline{u'}, h_{u'}\right\rangle \in \left\langle \overline{u}, h_{u}\right\rangle \uparrow _{R_\alpha ^{IT+}}\) and \(\left\langle \overline{u'}, h_{u'}\right\rangle \in \left\langle \overline{x}, h_{x}\right\rangle \uparrow _{R_\alpha ^{IT+}}\). Since \(\pi _\alpha ^\square \left[ \left\langle \overline{w}, h_v \right\rangle \right] \) is open in \(\tau _{R_\alpha ^{IT+}}\), we know that \(\left\langle \overline{u}, h_{u}\right\rangle \uparrow _{R_\alpha ^{IT+}}\subseteq U\) and that \(\left\langle \overline{x}, h_{x}\right\rangle \uparrow _{R_\alpha ^{IT+}}\subseteq V\). Thus, \(\left\langle \overline{u'}, h\right\rangle \in U\cap V\), so that U and V are \(\tau _\alpha ^{\left\langle \overline{w}, h_v \right\rangle }\)-dense.

Let \(\mathcal {M} \) be a Kripke-ies-model with valuation function \(\mathcal {V}\). The frame upon which \(\mathcal {M} \) is based has an associated iebt-frame. If to the tuple of this iebt-frame one adds a valuation function \(\mathcal {V}^t\) such that \(\mathcal {V}^t(p)=\left\{ \left\langle \overline{w},h_w\right\rangle ; w\in \mathcal {V}(p)\right\} \), the resulting model is called the iebt-model associated to \(\mathcal {M}\).

Proposition 3

Let \(\mathcal {M}\) be a Kripke-ies-model, and let \(\mathcal {M}^T\) denote its associated iebt-model. For \(\varphi \) of \(\mathcal {L}_{\textsf {I}}\) and \(w\in W\), \(\mathcal {M} ,w\models \varphi \) iff \(\mathcal {M} ^T,\left\langle \overline{w},h_w\right\rangle \models \varphi \).

Proof

We proceed by induction on \(\varphi \). For the base case, take a propositional letter p and an arbitrary \(w\in W\). Then \(\mathcal {M} ,w\models p\) iff \(w\in \mathcal {V}(p)\) iff \(\left\langle \overline{w},h_w\right\rangle \in \mathcal {V}^T(p)\) iff \(\mathcal {M} ^T,\left\langle \overline{w},h_w\right\rangle \models p\). The cases of Boolean connectives are standard, so let us deal with the modal operators. Let \(w\in W\) and \(\alpha \in Ags\).

  • (\(\square \)) \(\mathcal {M} ,w\models \square \varphi \) iff for every \(v\in \overline{w}\), \(\mathcal {M},v\models \varphi \), which by induction hypothesis happens iff \(\mathcal {M} ^T,\left\langle \overline{v},h_v\right\rangle \models \varphi \) for every \(v\in \overline{w}\), which happens iff \(\mathcal {M} ^T,\left\langle \overline{w},h_w\right\rangle \models \square \varphi \), since it is the case that \(h_v\in H_{\overline{w}}\) iff \(v\in \overline{w}\).

  • (\([\alpha ]\)) \(\mathcal {M} ,w\models [\alpha ]\varphi \) iff for every \(v\in W\) such that \(wR_\alpha ^{\overline{w}} v\), \(\mathcal {M} ,v\models \varphi \), which by induction hypothesis happens iff \(\mathcal {M} ^T,\left\langle \overline{w},h_v\right\rangle \models \varphi \) for every \(h_v\in \textbf{Choice}^{\overline{w}}_\alpha (h_w)\), which in turn happens iff \(\mathcal {M} ^T,\left\langle \overline{w},h_w\right\rangle \models [\alpha ]\varphi \).

  • (\(K_\alpha \)) \(\mathcal {M} ,w\models K_\alpha \varphi \) iff for every \(v\in W\) such that \(w\approx _\alpha v\), \(\mathcal {M} ,v\models \varphi \), which by induction hypothesis occurs iff \(\mathcal {M} ^T,\left\langle \overline{v},h_v\right\rangle \models \varphi \) for every \(h_v\in H\) such that \(\left\langle \overline{w},h_w\right\rangle \sim _\alpha \left\langle \overline{v},h_v\right\rangle \), which happens iff \(\mathcal {M} ^T,\left\langle \overline{w},h_w\right\rangle \models K_\alpha \varphi \).

  • (\(I_\alpha \)) First, observe that the induction hypothesis implies that \(\Vert \varphi \Vert =\left\{ \left\langle \overline{w}, h_w \right\rangle ; w\in |\varphi |\right\} \). Therefore, \(\mathcal {M} ,w\models I_\alpha \varphi \) iff there exists \(x\in \pi _\alpha ^\square [w]\) such that \(x\uparrow _{R_\alpha ^{I+}}\subseteq |\varphi |\) iff \(\left\langle \overline{x}, h_{x}\right\rangle \uparrow _{R_\alpha ^{IT+}}\subseteq \Vert \varphi \Vert \) iff there exists \(U\in \tau ^{\left\langle \overline{w}, h_w \right\rangle }_\alpha \) such that \(U\subseteq \Vert \varphi \Vert \) iff \(\mathcal {M} ^T,\left\langle \overline{w},h_w\right\rangle \models I_\alpha \varphi \).

1.3 A.3 Canonical Kripke-ies-structure

We show that the proof system \({\varLambda _I}\) is complete with respect to the class of Kripke-ies-models. For each \({\varLambda _I}\)-consistent formula \(\varphi \), we build a canonical structure from the syntax that satisfies \(\varphi \).

Definition 12

(Canonical Structure). The tuple

$$\mathcal {M}=\left\langle W^{\varLambda _I}, R_\square , \texttt{Choice}, \left\{ \approx _\alpha \right\} _{\alpha \in Ags}, \left\{ R_\alpha ^I\right\} _{\alpha \in Ags} ,\mathcal {V} \right\rangle $$

is called a canonical structure for \({\varLambda _I}\) iff

  • \(W^{\varLambda _I}=\left\{ w ;w \text{ is } \text{ a } {\varLambda _I}\text {-MCS}\right\} \). \(R_\square \) is a relation on \(W^{\varLambda _I}\) defined by ther rule: for \(w,v\in W^{\varLambda _I}\), \(wR_{\square }v\) iff \(\square \varphi \in w\Rightarrow \varphi \in v\) for every \(\varphi \) of \(\mathcal {L}_{\textsf {I}}\). For \(w\in W^{\varLambda _I}\), the set \(\left\{ v\in W^{\varLambda _I} ;wR_\square v \right\} \) is denoted by \(\overline{w}\).

  • \(\texttt{Choice}\) is a function that assigns to each \(\alpha \) and \(\overline{w}\) a subset of \(2^{\overline{w}}\), denoted by \(\texttt{Choice}_\alpha ^{\overline{w}}\), and defined as follows: let \(R_\alpha ^{\overline{w}}\) be a relation on \({\overline{w}}\) such that, for \(w,v\in W^{\varLambda _I}\), \(wR_{\alpha }^{\overline{w}}v\) iff \([\alpha ]\varphi \in w\Rightarrow \varphi \in v\) for every \(\varphi \) of \(\mathcal {L}_{\textsf {I}}\); if \(\texttt{Choice}^{\overline{w}}_\alpha (v):=\left\{ u\in \overline{w} ; vR_\alpha ^{\overline{w}} u\right\} \), then \(\texttt{Choice}^{\overline{w}}_\alpha :=\left\{ \texttt{Choice}^{\overline{w}}_\alpha (v); v\in \overline{w}\right\} \).

  • For \(\alpha \in Ags\), \(\approx _\alpha \) is an epistemic relation on \(W^{\varLambda _I}\) given by the rule: for \(w,v\in W^{\varLambda _I}\), \(w\approx _{\alpha }v\) iff \(K_\alpha \varphi \in w\Rightarrow \varphi \in v\) for every \(\varphi \) of \(\mathcal {L}_{\textsf {I}}\).

  • For \(\alpha \in Ags\), \(R_\alpha ^I\) is a relation on \(W^{\varLambda _I}\) given by the rule: for \(w,v\in W^{\varLambda _I}\), \(wR_{\alpha }^Iv\) iff \(I_\alpha \varphi \in w\Rightarrow \varphi \in v\) for every \(\varphi \) of \(\mathcal {L}_{\textsf {I}}\).

  • \(\mathcal {V}\) is the canonical valuation, defined such that \(w\in \mathcal {V}(p)\) iff \(p\in w\).

Proposition 4

The canonical structure \(\mathcal {M}\) for \({\varLambda _I}\) is a Kripke-ies-model.

Proof

We want to show that the tuple

\(\left\langle W^{\varLambda _I}, R_\square , \texttt{Choice}, \left\{ \mathtt {\approx }_\alpha \right\} _{\alpha \in Ags}, \left\{ \texttt{R}_\alpha ^I\right\} _{\alpha \in Ags}\right\rangle \) is a Kripke-ies-frame, which amounts to showing that the tuple satisfies the items in the definition of Kripke-ies-frames (Definition 9).

  • It is clear that \(R_\square \) is an equivalence relation, since \({\varLambda _I}\) includes the \(\textbf{S5}\) axioms for \(\square \).

  • Since \({\varLambda _I}\) includes the \(\textbf{S5}\) schemata for \([\alpha ]\) (\(\alpha \in Ags\)), \(R_\alpha ^{\overline{w}}\) is an equivalence relation for \(\alpha \in Ags\) and \(w\in W^{\varLambda _I}\). Moreover, since \({\varLambda _I}\) includes schema (SET), \(R_\alpha ^{\overline{w}}\subseteq \overline{w}\times \overline{w}\) for every \(w\in W^{\varLambda _I}\). Thus, \(\texttt{Choice}\) indeed assigns to each \(\alpha \) and \(\overline{w}\) a partition of \(\overline{w}\).

    To show that frame condition \(\mathtt {(IA)_K}\) is satisfied, we first prove two intermediate results:

    1. (a)

      For \(w_*\in W^{\varLambda _I}\), \(w\in \overline{w_*}\) iff \(\left\{ \square \psi ; \square \psi \in w_*\right\} \subseteq w\). \((\Rightarrow )\) Let \(w\in \overline{w_*}\) (which means that \(w_*R_\square w\)). Take \(\varphi \) of \(\mathcal {L}_{\textsf {Kx}}\) such that \(\square \varphi \in w_*\). Since \(w_*\) is closed under Modus Ponens, axiom (4) for \(\square \) implies that \(\square \square \varphi \in w_*\). By definition of \(R_\square \), \(\square \varphi \in w\). \((\Leftarrow )\) Assume that \(\left\{ \square \psi ; \square \psi \in w_*\right\} \subseteq w\). Take \(\varphi \) of \(\mathcal {L}_{\textsf {KX}}\) such that \(\square \varphi \in w_*\). By assumption, \(\square \varphi \in w\). Since w is closed under Modus Ponens, axiom (T) for \(\square \) implies that \(\varphi \in w\). Thus, \(w_*R_\square w\) and \(w\in \overline{w_*}\).

    2. (b)

      If \(w_*\in W^{\varLambda _I}\) and \(s:Ags\rightarrow 2^{\overline{w_*}}\) maps \(\alpha \) to a member of \(\texttt{Choice}^{\overline{w_*}}_\alpha \) such that \(v_{s(\alpha )}\in s(\alpha )\), then \(w\in s(\alpha )\) iff \(\varDelta _{s(\alpha )}=\left\{ [\alpha ] \psi ;[\alpha ]\psi \in v_{s(\alpha )}\right\} \subseteq w\). \((\Rightarrow )\) Let \(w\in s(\alpha )\) (which means that \(v_{s(\alpha )}R_\alpha w\)). Take \(\varphi \) of \(\mathcal {L}_{\textsf {KX}}\) such that \([\alpha ]\varphi \in v_{s(\alpha )} \). Since \(v_{s(\alpha )}\) is closed under Modus Ponens, schema (4) for \([\alpha ]\) implies that \([\alpha ][\alpha ]\varphi \in v_{s(\alpha )}\). Therefore, by definition of \(R_\alpha \), \([\alpha ]\varphi \in w\). \((\Leftarrow )\) Assume that \(\varDelta _{s(\alpha )}=\left\{ [\alpha ] \psi ; [\alpha ]\psi \in v_{s(\alpha )}\right\} \subseteq w\). Take \(\varphi \) of \(\mathcal {L}_{\textsf {KX}}\) such that \([\alpha ]\varphi \in v_{s(\alpha )}\). By assumption, \([\alpha ]\varphi \in w\). Since w is closed under Modus Ponens, axiom (T) for \([\alpha ]\) implies that \(\varphi \in w\). Thus, \(v_{s(\alpha )}R_\alpha ^{\overline{w_*}} w\) and \(w\in s(\alpha )\).

    Next we show that, for \(w_*\in W^{\varLambda _I}\) and \(s:Ags\rightarrow 2^{\overline{w_*}}\) just as in item b above, \(\bigcup _{\alpha \in Ags}\varDelta _{s(\alpha )}\cup \left\{ \square \psi ; \square \psi \in w_*\right\} \) is \(\varLambda _I\)-consistent: first we show that \(\bigcup _{\alpha \in Ags} \varDelta _{s(\alpha )}\) is consistent. Suppose that this is not the case. Then there exists a set \(\left\{ \varphi _1,\dots ,\varphi _n\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \([\alpha _{i} ]\varphi _i\in v_{s(\alpha _i)}\) for every \(1\le i \le n\) and

    $$\begin{aligned} \vdash _{\varLambda _I}[\alpha _{1} ]\varphi _1\wedge \dots \wedge [\alpha _{n} ]\varphi _n\rightarrow \bot . \end{aligned}$$
    (1)

    Without loss of generality, assume that \(\alpha _i\ne \alpha _j\) for all \(j\ne i\) such that \(j,i\in \left\{ 1,\dots ,n\right\} \)—this assumption hinges on the fact that any stit operator distributes over conjunction. Notice that the fact that \([\alpha _{i} ]\varphi _i\in v_{s(\alpha _i)}\) for every \(1\le i \le n\) implies that \(\Diamond [\alpha _{i} ]\varphi _i \in w_*\) for every \(1\le i \le n\). Since \(w_*\) is closed under conjunction, \(\Diamond [\alpha _{1} ]\varphi _1\wedge \dots \wedge \Diamond [\alpha _{n} ]\varphi _n\in w_*\).

    Axiom (IA) then implies that

    $$\begin{aligned} \vdash _{\varLambda _I}\Diamond [\alpha _{1} ]\varphi _1\wedge \dots \wedge \Diamond [\alpha _{n} ]\varphi _n\rightarrow \Diamond \left( [\alpha _{1} ]\varphi _1\wedge \dots \wedge [\alpha _{n} ]\varphi _n\right) . \end{aligned}$$
    (2)

    Therefore, Eqs. (2) and (1), imply that

    $$\begin{aligned} \vdash _{\varLambda _I}\Diamond [\alpha _{1} ]\varphi _1\wedge \dots \wedge \Diamond [\alpha _{n} ]\varphi _n\rightarrow \Diamond \bot . \end{aligned}$$
    (3)

    But this is a contradiction, since \(\Diamond [\alpha _{1} ]\varphi _1\wedge \dots \wedge \Diamond [\alpha _{n} ]\varphi _n\in w_*\), and \(w_*\) is a \(\varLambda _I\)-MCS. Therefore, \(\bigcup _{\alpha \in Ags} \varDelta _{s(\alpha )}\) is consistent. Secondly, we show that the union \(\bigcup _{\alpha \in Ags} \varDelta _{s(\alpha )}\cup \left\{ \square \psi ; \square \psi \in w_*\right\} \) is also consistent. Suppose that this is not the case. Since \(\bigcup _{\alpha \in Ags} \varDelta _{s(\alpha )}\) is consistent, there must exist sets \(\left\{ \varphi _1,\dots ,\varphi _n\right\} \) and \(\left\{ \theta _1,\dots ,\theta _m\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \([\alpha _i ]\varphi _i\in v_{s(\alpha _i)}\) for every \(1\le i \le n\), \(\square \theta _i\in w_*\) for every \(1\le i\le m\), and

    $$\begin{aligned} \vdash _{\varLambda _I}[\alpha _1 ]\varphi _1\wedge \dots \wedge [\alpha _n]\varphi _n\wedge \square \theta _1\wedge \dots \wedge \square \theta _m\rightarrow \bot . \end{aligned}$$
    (4)

    Let \(\theta =\theta _1\wedge \dots \wedge \theta _m\). Since \(\square \) distributes over conjunction, \(\vdash _{\varLambda _I}\square \theta \leftrightarrow \square \theta _1\wedge \dots \wedge \square \theta _m\), where it is important to mention that, since \(w_*\) is a \(\varLambda _I\)-MCS closed under logical equivalence, \(\square \theta \in w_*\). Thus, (4) implies that

    $$\begin{aligned} \vdash _{\varLambda _I}([\alpha _1 ]\varphi _1\wedge \dots \wedge [\alpha _n ]\varphi _n)\rightarrow \lnot \square \theta .\end{aligned}$$
    (5)

    Once again, assume without loss of generality that \(\alpha _i\ne \alpha _j\) for all \(j\ne i\) such that \(j,i\in \left\{ 1,\dots ,n\right\} \). By an argument analogous to the one used to show that \(\bigcup _{\alpha \in Ags} \varDelta _{s(\alpha )}\) is consistent, (5) implies that

    $$\begin{aligned} \vdash _{\varLambda _I}\Diamond [\alpha _{1} ]\varphi _1\wedge \dots \wedge \Diamond [\alpha _{n} ]\varphi _n\rightarrow \Diamond \lnot \square \theta . \end{aligned}$$
    (6)

    This entails that \(\Diamond \lnot \square \theta \in w_*\), but this is a contradiction, since the fact that \(\square \theta \in w_*\) implies with axiom (4) for \(\square \) that \(\square \square \theta \in w_*\). Now, let \(u_*\) be the \(\varLambda _I\)-MCS that includes \(\bigcup _{\alpha \in Ags} \varDelta _{s(\alpha )}\cup \left\{ \square \psi ; \square \psi \in w_*\right\} \). By intermediate result a, it is clear that \(u_*\in \overline{w_*}\). By intermediate result b, it is clear that \(u_*\in s(\alpha )\) for every \(\alpha \in Ags\). Therefore, we have shown that, for \(w_*\in W\), each function \(s:Ags\rightarrow 2^{\overline{w_*}}\) that maps \(\alpha \) to a member of \(\texttt{Choice}^{\overline{w_*}}_\alpha \) is such that \(\bigcap _{\alpha \in Ags} s(\alpha ) \ne \emptyset \), which means that \(\mathcal {M}\) satisfies \(\mathtt {(IA)_K}\).

  • Since the proof system \({\varLambda _I}\) includes the \(\textbf{S5}\) axioms for \(K_\alpha \) (\(\alpha \in Ags\)), \(\approx _\alpha \) is an equivalence relation for \(\alpha \in Ags\). We verify that \(\mathcal {M}\) satisfies conditions \((\texttt{OAC})_{\texttt{K}}\) and \((\mathtt {Unif-H})_{\texttt{K}}\).

    For \((\texttt{OAC})_{\texttt{K}}\), let \(w_*\in W^{\varLambda _I}\) and \(\alpha \in Ags\). Assume that \(v,u\in \overline{w_*}\) are such that \(v\approx _\alpha u\). Let \(v'\in \texttt{Choice}^{\overline{w_*}}_\alpha (v)\). This means that \(vR_\alpha v'\). We want to show that \(v'\approx _\alpha u\), so let \(\varphi \) be a formula of \(\mathcal {L}_{\textsf {KO}}\) such that \(K_\alpha \varphi \in v'\). By schema (4) for \(K_\alpha \), \(K_\alpha K_\alpha \varphi \in v'\). Similarly, since all substitutions of axiom (OAC) lie within \(v'\) and it is closed under Modus Ponens, \([\alpha ]K_\alpha \varphi \) also lies in \(v'\). Since \(v'R_\alpha v\), this implies that \(K_\alpha \varphi \in v\). Therefore, our assumption that \(v\approx _\alpha u\) entails that \(\varphi \in u\). With this, we have shown that the fact that \(K_\alpha \varphi \in v'\) implies that \(\varphi \in u\), which means that \(v'\approx _\alpha u\).

    For \(\mathtt {(Unif-H)_K}\), let \(v, u\in W^{\varLambda _I}\) such that \(v\approx _\alpha u\). Take \(v'\in \overline{v}\). We want to show that there exists \(u'\in \overline{u}\) such that \(v'\approx _\alpha u'\). We show that \(u''=\left\{ \psi ; K_\alpha \psi \in v'\right\} \cup \left\{ \square \psi ; \square \psi \in u \right\} \) is consistent. To do so, we first show that \(\left\{ \psi ; K_\alpha \psi \in v'\right\} \) is consistent. Suppose for a contradiction that it is not consistent. Then there exists a set \(\left\{ \psi _1,\dots ,\psi _n\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \(K_\alpha \psi _i\in v'\) for every \(1\le i \le n\) and (a) \(\vdash _{\varLambda _I}\psi _1\wedge \dots \wedge \psi _n\rightarrow \bot \). By Necessitation for \(K_\alpha \) and its distributivity over conjunction, (a) implies that \(\vdash _{\varLambda _I}K_\alpha \psi _1\wedge \dots \wedge K_\alpha \psi _n\rightarrow K_\alpha \bot \), but this is a contradiction, since \( v'\) is a \(\varLambda _I\)-MCS and it includes \(K_\alpha \psi _1\wedge \dots \wedge K_\alpha \psi _n\). Next we show that \(u''=\left\{ \psi ; K_\alpha \psi \in v'\right\} \cup \left\{ \square \psi ; \square \psi \in u \right\} \) is also consistent. Suppose for a contradiction that it is not consistent. Since \(\left\{ \psi ; K_\alpha \psi \in v'\right\} \) and \(\left\{ \square \psi ; \square \psi \in u \right\} \) are consistent, there must exist sets \(\left\{ \varphi _1,\dots ,\varphi _n\right\} \) and \(\left\{ \theta _1,\dots ,\theta _m\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \(K_\alpha \varphi _i\in v'\) for every \(1\le i \le n\), \(\square \theta _i\in w_2\) for every \(1\le i\le m\), and (b) \(\vdash _{\varLambda _I}\varphi _1\wedge \dots \wedge \varphi _n\wedge \square \theta _1\wedge \dots \wedge \square \theta _m\rightarrow \bot \). Let \(\theta =\theta _1\wedge \dots \wedge \theta _m\) and \(\varphi =\varphi _1\wedge \dots \wedge \varphi _n\). Since \(\square \) distributes over conjunction, \(\vdash _{\varLambda _I}\square \theta \leftrightarrow \square \theta _1\wedge \dots \wedge \square \theta _m\), where it is important to mention that, since u is a \(\varLambda _I\)-MCS, then \(\square \theta \in u\) and (\(\star \)) \(\square \square \theta \in u\) as well. In this way, (b) implies that \(\vdash _{\varLambda _I}\varphi \rightarrow \lnot \square \theta \) and thus that (c) \(\vdash _{\varLambda _I}\Diamond \varphi \rightarrow \Diamond \lnot \square \theta \). Notice that the facts that \(K_\alpha \varphi _i\in v'\) for every \(1\le i\le n\), that \(K_\alpha \) distributes over conjunction, and that \(v'\) is a \(\varLambda _I\)-MCS imply that \(K_\alpha \varphi \in v'\). The fact that \(v'\in \overline{v}\) implies that \(\Diamond K_\alpha \varphi \in v\), so that \((Unif-H)\) entails that \(K_\alpha \Diamond \varphi \in v\). Now, this last inclusion implies, with our assumption that \(v\approx _\alpha u\), that \(\Diamond \varphi \in u\), which by (c) in turn yields that \(\Diamond \lnot \square \theta \in u\), contradicting \((\star )\). Therefore, \(u''\) is consistent. Let \(u'\) be the \(\varLambda _I\)-MCS that includes \(u''\). It is clear from its construction that \(u'\in \overline{u}\) and that \(v'\approx _\alpha u'\). With this, we have shown that \(\mathcal {M}\) satisfies condition \(\mathtt {(Unif-H)_K}\).

  • Since \({\varLambda _I}\) includes the \(\textbf{KD45}\) schemata for \(I_\alpha \) (\(\alpha \in Ags\)), then \(R_\alpha ^I \) is a serial, transitive, and euclidean relation on W, for \(\alpha \in Ags\). Since \({\varLambda _I}\) includes schema (InN), then \(R_\alpha ^I \subseteq \approx _\alpha \circ R_\square \) for \(\alpha \in Ags\).

    We now verify that frame condition \(\mathtt {(Den)_K}\) is satisfied. Let \(v, u\in W^{\varLambda _{I}}\) such that \(v \approx _\alpha \circ R_\square u\). This means that there exists w such that \(v\in \overline{w}\) and \(w\approx _\alpha u\). We want to show that there exists \(uR_\alpha ^I u'\) such that \(vR_\alpha ^I u'\). We show that \(u''=\left\{ \psi ; I_\alpha \psi \in v\right\} \cup \left\{ \psi ; I_\alpha \psi \in u \right\} \) is consistent. To do so, we first show that \(\left\{ \psi ; I_\alpha \psi \in v\right\} \) is consistent. Suppose for a contradiction that it is not consistent. Then there exists a set \(\left\{ \psi _1,\dots ,\psi _n\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \(I_\alpha \psi _i\in v\) for every \(1\le i \le n\) and (a) \(\vdash _{\varLambda _{I}}\psi _1\wedge \dots \wedge \psi _n\rightarrow \bot \). By Necessitation for \(I_\alpha \) and its distributivity over conjunction, (a) implies that \(\vdash _{\varLambda _{I}}I_\alpha \psi _1\wedge \dots \wedge I_\alpha \psi _n\rightarrow I_\alpha \bot \), but this is a contradiction, since v is a \(\varLambda _{I}\)-MCS and it includes \(I_\alpha \psi _1\wedge \dots \wedge I_\alpha \psi _n\). Next we show that \(u''=\left\{ \psi ; I_\alpha \psi \in v\right\} \cup \left\{ \psi ; I_\alpha \psi \in u \right\} \) is also consistent. Suppose for a contradiction that it is not consistent. Since \(\left\{ \psi ; I_\alpha \psi \in v'\right\} \) and \(\left\{ \psi ; I_\alpha \psi \in u \right\} \) are consistent, there must exist sets \(\left\{ \varphi _1,\dots ,\varphi _n\right\} \) and \(\left\{ \theta _1,\dots ,\theta _m\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \(I_\alpha \varphi _i\in v\) for every \(1\le i \le n\), \(I_\alpha \theta _i\in w_2\) for every \(1\le i\le m\), and (b) \(\vdash _{\varLambda _{I}}\varphi _1\wedge \dots \wedge \varphi _n\wedge \theta _1\wedge \dots \wedge \square \theta _m\rightarrow \bot \). Let \(\theta =\theta _1\wedge \dots \wedge \theta _m\) and \(\varphi =\varphi _1\wedge \dots \wedge \varphi _n\). Thus, (b) implies that \(\vdash _{\varLambda _{I}}\varphi \rightarrow \lnot \theta \) and thus that (c) \(\vdash _{\varLambda _{I}}\langle I_\alpha \rangle \varphi \rightarrow \langle I_\alpha \rangle \lnot \theta \). Notice that the facts that \(I_\alpha \varphi _i\in v\) for every \(1\le i\le n\), that \(I_\alpha \) distributes over conjunction, and that v is a \(\varLambda _{I}\)-MCS imply that \(I_\alpha \varphi \in v\). Analogously, one has that (\(\star \)) \(I_\alpha \theta \in u\). The fact that \(v\in \overline{w}\) implies that \(\Diamond I_\alpha \varphi \in w\), so that (Den) entails that \(K_\alpha \langle I_\alpha \rangle \varphi \in w\). Now, this last inclusion implies, with the fact that \(w\approx _\alpha u\), that \(\langle I_\alpha \rangle \varphi \in u\), which by (c) in turn yields that \(\langle I_\alpha \rangle \lnot \theta \in u\), contradicting \((\star )\). Therefore, \(u''\) is consistent. Let \(u'\) be the \(\varLambda _{I}\)-MCS that includes \(u''\). It is clear from its construction that \(uR_\alpha ^I u'\) and that \(vR_\alpha ^I u'\). With this, we have shown that \(\mathcal {M}\) satisfies \(\mathtt {(Den)_K}\).

Lemma 1

(Existence for non-intentional operators). Let \(\mathcal {M}\) be the canonical Kripke-ies-model for \({\varLambda _I}\). Let \(w\in W^{\varLambda _I}\). For \(\varphi \) of \(\mathcal L_{\textsf {I}}\), the following items hold:

  1. 1.

    \(\square \varphi \in w\) iff \(\varphi \in v\) for every \(v\in \overline{w}\).

  2. 2.

    \([\alpha ]\varphi \in w\) iff \(\varphi \in v\) for every \(v\in \overline{w}\) such that \(wR_\alpha ^{\overline{w}} v\).

  3. 3.

    \(K_\alpha \varphi \in w\) iff \(\varphi \in v\) for every \(v\in W^{\varLambda _I}\) such that \(w\approx _\alpha v\).

Proof

Let \(w\in W^{\varLambda _I}\), and take \(\varphi \) of \(\mathcal L_{\textsf {I}}\). All items are shown in the same way. Let \(\triangle \in \left\{ \square , [\alpha ], K_\alpha \right\} \), and let \(R_\triangle \) stand for the relation upon which the semantics of \(\triangle \varphi \) is defined. We show that \(\triangle \varphi \in w\) iff \(\varphi \in v\) for every \(v\in W^{\varLambda _I}\) such that \(wR_\triangle v\).

\((\Rightarrow )\) Assume that \(\triangle \varphi \in w\). Let \(v\in W^{\varLambda _I}\) such that \(wR_\triangle v\). The definition of \(R_\triangle \) straightforwardly gives that \(\varphi \in v\).

\((\Leftarrow )\) We work by contraposition. Assume that \(\triangle \varphi \notin w\). We show that there is a world v in \(W^{\varLambda _I}\) such that \(wR_\triangle v\) and such that \(\varphi \) does not lie within it. For this, let \(v'=\left\{ \psi ;\triangle \psi \in w\right\} \), which is shown to be consistent as follows: suppose for a contradiction that \(v'\) is not consistent; then there exists a set \(\left\{ \psi _1,\dots ,\psi _n\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \(\left\{ \psi _1,\dots ,\psi _n\right\} \subseteq v'\) and (a) \(\vdash _{\varLambda _I}\psi _1\wedge \dots \wedge \psi _n\rightarrow \bot \); now, the fact that \(\left\{ \psi _1,\dots ,\psi _n\right\} \subseteq v'\) means that \(\triangle \psi _i\in w\) for every \(1\le i \le n\); Necessitation for \(\triangle \) and its distributivity over conjunction yield that (a) implies that \(\vdash _{\varLambda _I}\triangle \psi _1\wedge \dots \wedge \triangle \psi _n\rightarrow \triangle \bot \), but this is a contradiction, since w is a \(\varLambda _I\)-MCS which includes \(\triangle \psi _1\wedge \dots \wedge \triangle \psi _n\). Now, we define \(v':'=v'\cup \left\{ \lnot \varphi \right\} \), and we show that it is also consistent: suppose for a contradiction that it is not consistent; since \(v'\) is consistent, there exists a set \(\left\{ \psi _1,\dots ,\psi _n\right\} \) of formulas of \(\mathcal {L}_{\textsf {KX}}\) such that \(\left\{ \psi _1,\dots ,\psi _n\right\} \subseteq v'\) and \(\vdash _{\varLambda _I}\psi _1\wedge \dots \wedge \psi _n \wedge \lnot \varphi \rightarrow \bot \), which then implies that (b) \(\vdash _{\varLambda _I}\psi _1\wedge \dots \wedge \psi _n \rightarrow \varphi \); By Necessitation for \(\triangle \) and its distributivity over conjunction, (b) implies that \(\vdash _{\varLambda _I}\triangle \psi _1\wedge \dots \wedge \triangle \psi _n\rightarrow \triangle \varphi \); but, since w is a \(\varLambda _I\)-MCS, then \(\triangle \psi _1\wedge \dots \wedge \triangle \psi _n\in w\), so that (b) and the fact that w is closed under Modus Ponens entail that \(\triangle \varphi \in w\), contradicting the initial assumption that \(\triangle \varphi \notin w\). Let v be the \(\varLambda _I\)-MCS that includes \(v''\). It is clear from its construction that \(\varphi \notin v\) and that \(wR_\triangle v\), by definition of \(R_\triangle \).

Lemma 2

(Truth Lemma). Let \(\mathcal {M}\) be the canonical Kripke-ies-model for \({\varLambda _I}\). For \(\varphi \) of \(\mathcal {L}_{\textsf {I}}\) and \(w\in W^{\varLambda _I}\), \(\mathcal {M},w\models \varphi \ \text {iff} \ \varphi \in w.\)

Proof

We proceed by induction on \(\varphi \). The cases of Boolean connectives are standard. For formulas involving \(\square , [\alpha ]\), and \(K_\alpha \), both directions follow straightforwardly from Lemma 1 (items 1, 2, and 3, respectively). As for \(I_\alpha \), we have the following arguments:

  • (“\(I_\alpha \)”)

    (\(\Rightarrow \)) We work by contraposition. Suppose that \(I_\alpha \varphi \notin w\). Take \(x\in \pi _\alpha ^\square [w]\). The assumption that \(\lnot I_\alpha \varphi \in w\) implies, by schema (KI) and closure of w under Modus Ponens, that \(\square K_\alpha \lnot I_\alpha \varphi \in w\). Since \(x\in \pi _\alpha ^\square [w]\), this implies that \(\lnot I_\alpha \varphi \in x\). By an argument analogous to the one used in Proposition 4 to show that the canonical model satisfies \((\texttt{Den})_{\texttt{K} }\), the set \(\left\{ \psi ; I_\alpha \psi \in x\right\} \) is consistent. Next, observe that \(\left\{ \psi ; I_\alpha \psi \in w\right\} \cup \left\{ \lnot \varphi \right\} \) is consistent. Suppose it is not consistent. Since \(\left\{ \psi ; I_\alpha \psi \in w\right\} \) is consistent, there must exist a set \(\left\{ \varphi _1,\dots ,\varphi _n\right\} \) such that \(I_\alpha \varphi _i\in w\) for every \(1\le i \le n\) and \( \vdash _{\varLambda _I}(\varphi _1\wedge \dots \wedge \varphi _n)\wedge \lnot \varphi \rightarrow \bot \) Now, this \({\varLambda _I}\)-theorem implies that \( \vdash _{\varLambda _I}(\varphi _1\wedge \dots \wedge \varphi _n)\rightarrow \varphi \). By Necessitation of \(I_\alpha \), its schema (K), and its distributivity over conjunction, one then has that (\(\star \)) \( \vdash _{\varLambda _I}(I_\alpha \varphi _1\wedge \dots \wedge I_\alpha \varphi _n)\rightarrow I_\alpha \varphi \). Now, closure of w under conjunction then implies that \(\left( \bigwedge _{1\le i\le n}I_\alpha \varphi _i\right) \in x\), so that the antecedent in \({\varLambda _I}\)-theorem \((\star )\) lies in x. Closure of x under Modus Ponens then implies that \(I_\alpha \varphi \in x\), but this contradicts the previously shown fact that \(I_\alpha \varphi \notin x\). Therefore, \(\left\{ \psi ; I_\alpha \psi \in w\right\} \cup \left\{ \lnot \varphi \right\} \) is in fact consistent. Let u be the \({\varLambda _I}\)-MCS that includes \(\left\{ \psi ; I_\alpha \psi \in w\right\} \cup \left\{ \lnot \varphi \right\} \). It is clear from construction that \(x R_\alpha ^I u\), so that \(u\in x\uparrow _{R_\alpha ^{+}}\). It also clear from construction that \(\lnot \varphi \in x\), so that the induction hypothesis yields that \(\mathcal {M}, x\models \lnot \varphi \). Thus, x is such that \(x\in \pi _\alpha ^\square [w]\) and such that \(x\uparrow _{R_\alpha ^{+}}\not \subseteq |\varphi |\), which implies that \(\mathcal {M}, w\not \models \varphi \).

    (\(\Leftarrow \)) Assume that \(I_\alpha \varphi \in w\). Suppose for a contradiction that \(\mathcal {M}, w\not \models I_\alpha \varphi \). This means that, for \(x\in \pi _\alpha ^\square [w]\), there exists y such that \(xR_\alpha ^{I+} y\) and \(\mathcal {M}, y\not \models \varphi \). Now, we have two cases. If for every \(x\in \pi _\alpha ^\square [w]\) the y such that \(xR_\alpha ^{I+} y\) and \(\mathcal {M}, y\not \models \varphi \) is actually x itself, then \(\mathcal {M}, x\not \models \varphi \) for every \(x\in \pi _\alpha ^\square [w]\). By induction hypothesis, this implies that \(\lnot \varphi \in x\) for every \(x\in \pi _\alpha ^\square [w]\), which, by items 1 and 3 of Lemma 1, implies that \(\square K_\alpha \lnot \varphi \in x\) for every \(x\in \pi _\alpha ^\square [w]\). In particular, \(\square K_\alpha \lnot \varphi \in w\). Schema (InN) and closure of w under Modus Ponens then imply that \(I\lnot \varphi \in w\), but this is a contradiction, since the fact that \(I_\alpha \varphi \in w\), with schema (D) for \(I_\alpha \) and closure of w under Modus Ponens, implies that \(\lnot I\lnot \varphi \in w\). The other case is that there exist \(x, y\in \pi _\alpha ^\square [w]\) such that \(xR_\alpha ^{I+} y\), \(\mathcal {M}, y\not \models \varphi \), and \(y\ne x\). By induction hypothesis, \(\varphi \notin y\). Since \(xR_\alpha ^{I+} y\) and \(y\ne x\), then \(xR_\alpha ^{I} y\), so the definition of \(R_\alpha ^{I}\) implies that \(I_\alpha \varphi \not \in x\). As such, \(\lnot I_\alpha \varphi \in x\), which, by schema (KI) and closure of x under Modus Ponens, implies that \(\square K_\alpha \lnot I_\alpha \varphi \in x\). Since \(x\in \pi _\alpha ^\square [w]\), this implies that \(\lnot I_\alpha \varphi \in w\), but this is a contradiction to the initial assumption.

Theorem 2

(Completeness w.r.t. Kripke-ies-models). The proof system \({\varLambda _I}\) is complete with respect to the class of Kripke-ies-models.

Proof

Let \(\varphi \) be a \({\varLambda _I}\)-consistent formula of \(\mathcal {L}_{\textsf {I}}\). Let w be the \({\varLambda _I}\)-MCS including \(\varphi \). The canonical Kripke-ies-model \(\mathcal {M}\) for \(\varLambda _I\) is such that \(\mathcal {M}, w\models \varphi \).

Back to Branching-Time Models

Theorem 3

(Completeness w.r.t. iebt-models). The proof system \({\varLambda _I}\) is complete with respect to the class of iebt-models.

Proof

Let \(\varphi \) be a \({\varLambda _I}\)-consistent formula of \(\mathcal {L}_{\textsf {I}}\). Theorem 2 implies that there exists a Kripke-ies-model \(\mathcal {M}\) and a world w in its domain such that \(\mathcal {M} , w \models \varphi \). Proposition 3 then ensures that the iebt-model \(\mathcal {M}^T\) associated to \(\mathcal {M}\) is such that \(\mathcal {M}^T, \left\langle \overline{w},h_w \right\rangle \models \varphi \).

Therefore, the following result, appearing in the main body of the paper, has been shown:

Theorem 1

The proof system \({\varLambda _I}\) is sound and complete with respect to the class of iebt-models.

   \(\square \)

Rights and permissions

Reprints and permissions

Copyright information

© 2023 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this paper

Check for updates. Verify currency and authenticity via CrossMark

Cite this paper

Ramírez Abarca, A.I., Broersen, J. (2023). A Stit Logic of Intentionality. In: Areces, C., Costa, D. (eds) Dynamic Logic. New Trends and Applications. DaLí 2022. Lecture Notes in Computer Science, vol 13780. Springer, Cham. https://doi.org/10.1007/978-3-031-26622-5_8

Download citation

  • DOI: https://doi.org/10.1007/978-3-031-26622-5_8

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-031-26621-8

  • Online ISBN: 978-3-031-26622-5

  • eBook Packages: Computer ScienceComputer Science (R0)

Publish with us

Policies and ethics