This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\lmcsdoi

1522 \lmcsheadingLABEL:LastPageJan. 09, 2019Apr. 11, 2019 \ACMCCSTheory of computation \to Modal and temporal logics; Complexity theory and logic;

\titlecomment\lsuper

*This article is an extension of the conference publication [Lüc18b].

Canonical Models and the
Complexity of Modal Team Logic

Martin Lück Leibniz Universität Hannover, Institut für Theoretische Informatik, Appelstraße 4, 30167 Hannover lueck@thi.uni-hannover.de
Abstract.

We study modal team logic MTL, the team-semantical extension of modal logic ML closed under Boolean negation. Its fragments, such as modal dependence, independence, and inclusion logic, are well-understood. However, due to the unrestricted Boolean negation, the satisfiability problem of full MTL has been notoriously resistant to a complexity theoretical classification.

In our approach, we introduce the notion of canonical models into the team-semantical setting. By construction of such a model, we reduce the satisfiability problem of MTL to simple model checking. Afterwards, we show that this approach is optimal in the sense that MTL-formulas can efficiently enforce canonicity.

Furthermore, to capture these results in terms of complexity, we introduce a non-elementary complexity class, TOWER(poly), and prove that it contains satisfiability and validity of MTL as complete problems. We also prove that the fragments of MTL with bounded modal depth are complete for the levels of the elementary hierarchy (with polynomially many alternations). The respective hardness results hold for both strict or lax semantics of the modal operators and the splitting disjunction, and also over the class of reflexive and transitive frames.

Key words and phrases:
team semantics, modal logic, complexity, satisfiability

1. Introduction

It is well-known that non-linear quantifier dependencies, such as ww depending only on zz in the sentence xyzwφ\forall x\,\exists y\,\forall z\,\exists w\,\varphi, cannot be expressed in first-order logic. To overcome this restriction, logics of incomplete information such as independence-friendly logic [HS89] have been studied. Later, Hodges [Hod97] introduced team semantics to provide these logics with a compositional interpretation. The fundamental idea is to not consider single assignments to free variables, but instead whole sets of assignments, called teams.

In this vein, Väänänen [Vää07] expressed non-linear quantifier dependencies by the dependence atom =(x1,,xn,y){=\!\!(x_{1},\ldots,x_{n},y)}, which intuitively states that the values of yy in the team functionally depend on those of x1,,xnx_{1},\ldots,x_{n}. Logics with numerous other non-classical atoms such as independence \perp [GV13], inclusion \subseteq and exclusion \mid [Gal12] have been studied since, and manifold connections to scientific areas such as statistics, database theory, physics, cryptography and social choice theory have emerged (see also Abramsky et al. [AKVV16]).

Team semantics have also been adapted to a range of propositional [YV16, HKLV16], modal [Vää08], and temporal logics [KMV15, KMVZ18]. Besides propositional dependence logic 𝖯𝖣𝖫\xspace\mathsf{PDL}\xspace [YV16] and modal dependence logic 𝖬𝖣𝖫\xspace\mathsf{MDL}\xspace [Vää08], also propositional and modal logics of independence and inclusion have been studied [KMSV17, HKVV15, HS15, Han17]. Unlike in the first-order setting, the atoms such as the dependence atom range over flat formulas. For example, the instance =(p1,,pn,𝗎𝗇𝗌𝖺𝖿𝖾){=\!\!(p_{1},\ldots,p_{n},\Diamond\mathsf{unsafe})} of a modal dependence atom may specify that the reachability of an unsafe state is a function of p1pnp_{1}\cdots p_{n}, but instead of exhibiting the explicit function, the atom only stipulates its existence.

Most team logics lack the Boolean negation, and adding it as a connective {\sim} usually increases both the expressive power and the complexity tremendously. The respective extensions of propositional and modal logic are called propositional team logic 𝖯𝖳𝖫\xspace\mathsf{PTL}\xspace [HKLV16, YV17, HKVV18] and modal team logic 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace [Mül14, KMSV15]. With {\sim}, these logics can express all the non-classical atoms mentioned above, and in fact are expressively complete for their respective class of models [KMSV15, YV17]. For these reasons, they are both interesting and natural logics.

The expressive power of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is well-understood [KMSV15], and a complete axiomatization was presented by the author [Lüc18a]. Yet the complexity of the satisfiability problem has been an open question [Mül14, KMSV15, DKV16, HKMV17]. Recently, certain fragments of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace with restricted negation were shown ATIME\xspace-ALT\xspace(exp,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp,\mathrm{poly})-complete using the well-known filtration method [Lüc17]. In the same paper, however, it was shown that no elementary upper bound for full 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace can be established by the same approach, whereas the best known lower bound is ATIME\xspace-ALT\xspace(exp,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp,\mathrm{poly})-hardness, inherited from propositional team logic [HKVV18].

Logic Satisfiability Validity References
𝖯𝖣𝖫\xspace\mathsf{PDL}\xspace NP\xspace\mathrm{NP}\xspace NEXPTIME\xspace\mathrm{NEXPTIME}\xspace [LV13, Vir17]
𝖬𝖣𝖫\xspace\mathsf{MDL}\xspace NEXPTIME\xspace\mathrm{NEXPTIME}\xspace NEXPTIME\xspace\mathrm{NEXPTIME}\xspace [Sev09, Han17]
𝖯𝖨𝖫\xspace\mathsf{PIL}\xspace NP\xspace\mathrm{NP}\xspace NEXPTIME\xspace-hard,in Π2E\mathrm{NEXPTIME}\xspace\text{-hard},\text{in }\Uppi^{E}_{2} [HKVV15]
𝖬𝖨𝖫\xspace\mathsf{MIL}\xspace NEXPTIME\xspace\mathrm{NEXPTIME}\xspace Π2E-hard\Uppi^{E}_{2}\text{-hard} [KMSV17, Han16]
𝖯𝖨𝗇𝖼\xspace\mathsf{PInc}\xspace EXPTIME\xspace\mathrm{EXPTIME}\xspace co-NP\xspace\text{co-}\mathrm{NP}\xspace [HKVV15]
𝖬𝖨𝗇𝖼\xspace\mathsf{MInc}\xspace EXPTIME\xspace\mathrm{EXPTIME}\xspace co-NEXPTIME\xspace-hard\text{co-}\mathrm{NEXPTIME}\xspace\text{-hard} [HKMV15]
𝖯𝖳𝖫\xspace\mathsf{PTL}\xspace ATIME\xspace-ALT\xspace(exp,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp,\mathrm{poly}) ATIME\xspace-ALT\xspace(exp,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp,\mathrm{poly}) [HKLV16, HKVV18]
𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}) ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}) Theorem 31
𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) Theorem 31

Table 1. Complexity landscape of propositional and modal logics of dependence (𝖣𝖫*\mathsf{D}\mathsf{L}), independence (𝖨𝖫*\mathsf{IL}), inclusion (𝖨𝗇𝖼*\mathsf{Inc}) and team logic (𝖳𝖫*\mathsf{TL}). Entries are completeness results unless stated otherwise.

Contribution. We show that 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is complete for a non-elementary class we call TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}), which contains the problems decidable in a runtime that is a tower of nested exponentials of polynomial height. Likewise, we show that the fragments 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} of bounded modal depth kk are complete for classes we call ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}) and which corresponds to (k+1)(k+1)-fold exponential runtime and polynomially many alternations. These results fill a long-standing gap in the active field of propositional and modal team logics (see Table 1).

In our approach, we consider so-called canonical models. Loosely speaking, a canonical model satisfies every satisfiable formula in some of its submodels, and such models have been long known for, e.g.\xspace, many systems of modal logic [BRV01]. In Section 4, we adapt this notion for modal logics with team semantics, and prove that such models exist for 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace. This enables us to reduce the satisfiability problem to simple model checking, albeit on models that are of non-elementary size with respect to |Φ|+k{|\nobreak\Phi\nobreak|}+k, where Φ\Phi are the available propositional variables and kk is a bound on the modal depth.

Nonetheless, this approach is essentially optimal: In Section 5 and 6, we show that 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace can, in a certain sense, efficiently enforce canonical models, that is, with formulas that are of size polynomial in |Φ|+k{|\nobreak\Phi\nobreak|}+k. In this vein, we then obtain the matching complexity lower bounds in Section 7 and 8, where we encode computations of non-elementary length in such large models.

Finally, in Section 9 we extend the preliminary version of this paper [Lüc18b] and consider restrictions of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace to specific frame classes, and to so-called strict team-semantical connectives.

2. Preliminaries

The length of (the encoding of) xx is denoted by |x|{|\nobreak x\nobreak|}. We assume the reader to be familiar with alternating Turing machines [CKS81] and basic complexity theory. When a problem is hard or complete for a complexity class, in this paper we are always referring to logspace reductions.

The class ATIME\xspace-ALT\xspace(exp,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp,\mathrm{poly}) (also known as AEXPTIME\xspace(poly)\mathrm{AEXPTIME}\xspace(\mathrm{poly})) contains the problems decidable by an alternating Turing machine in time 2p(n)2^{p(n)} with p(n)p(n) alternations, where pp is a polynomial. We generalize it to capture the elementary hierarchy as follows.

Let exp0(n):=n\exp_{0}(n)\vcentcolon=n and expk+1(n):=2expk(n)\exp_{k+1}(n)\vcentcolon=2^{\exp_{k}(n)}. A function f:\xspace\xspacef\colon\mathbb{N}\xspace\to\mathbb{N}\xspace is elementary if it is computable in time 𝒪(expk(n)){\mathcal{O}(\exp_{k}(n))} for some fixed kk. In this paper, we consider the elementary hierarchy with polynomially many alternations:

{defi}

For k0k\geq 0, ATIME\xspace-ALT\xspace(expk,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k},\mathrm{poly}) is the class of problems decidable by an alternating Turing machine with at most p(n)p(n) alternations and runtime at most expk(p(n))\exp_{k}(p(n)), for a polynomial pp. Note that setting k=0k=0 or k=1k=1 yields the classes PSPACE\xspace\mathrm{PSPACE}\xspace and ATIME\xspace-ALT\xspace(exp,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp,\mathrm{poly}), respectively [CKS81]. Schmitz [Sch16] proposed the following non-elementary class that contains ATIME\xspace-ALT\xspace(expk,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k},\mathrm{poly}) for all kk. {defiC}[[Sch16]] TOWER\xspace\mathrm{TOWER}\xspace is the class of problems decidable by a deterministic Turing machine in time (or equivalently, space) expf(n)(1)\exp_{f(n)}(1) for an elementary function ff.

A suitable notion of reduction for this class is the following: An elementary reduction from AA to BB is an elementary function ff such that xAf(x)Bx\in A\Leftrightarrow f(x)\in B. AmelemBA\leq^{\mathrm{elem}}_{\mathrm{m}}B means that there exists an elementary reduction from AA to BB.

{propC}

[[Sch16]] TOWER\xspace\mathrm{TOWER}\xspace is closed under melem\leq^{\mathrm{elem}}_{\mathrm{m}}.

The next class results from imposing a polynomial bound on the number of exponentials in the definition of TOWER\xspace\mathrm{TOWER}\xspace, which leads to a strict subclass. {defi} TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) is the class of problems that are decided by a deterministic Turing machine in time (or equivalently, space) expp(n)(1)\exp_{p(n)}(1) for some polynomial pp.

The reader may verify that both ATIME\xspace-ALT\xspace(expk,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k},\mathrm{poly}) and TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) are closed under mP\leq^{\mathrm{P}}_{\mathrm{m}} and mlog\leq^{\log}_{\mathrm{m}}. Furthermore, by the time hierarchy theorem, TOWER\xspace(poly)TOWER\xspace\mathrm{TOWER}\xspace(\mathrm{poly})\subsetneq\mathrm{TOWER}\xspace.

To the author’s best knowledge, neither has been explicitly considered before. However, candidates for natural complete problems exist. Although not proved complete, several problems in TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) are provably non-elementary, such as the satisfiability problem of separated first-order logic [Voi17], the equivalence problem for star-free expressions [SM73], or the first-order theory of finite trees [CH90], to only name a few. We refer the reader also to the survey of Meyer [Mey74].

Another example is the two-variable fragment of first-order team logic, 𝖥𝖮\xspace2()\mathsf{FO}\xspace^{2}({\sim}). It is related to 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace in the same fashion as classical two-variable logic 𝖥𝖮\xspace2\mathsf{FO}\xspace^{2} to 𝖬𝖫\xspace\mathsf{ML}\xspace. By reduction from 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace to 𝖥𝖮\xspace2()\mathsf{FO}\xspace^{2}({\sim}), the satisfiability problem of 𝖥𝖮\xspace2()\mathsf{FO}\xspace^{2}({\sim}) is TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly})-complete problems as a corollary of our main result, Theorem 31, while its fragments 𝖥𝖮\xspacek2()\mathsf{FO}\xspace^{2}_{k}({\sim}) of bounded quantifier rank kk are ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly})-hard [Lüc18c].

Next, we justify why we use only mlog\leq^{\log}_{\mathrm{m}}-reductions (or polynomial time reductions in general) in this paper instead of melem\leq^{\mathrm{elem}}_{\mathrm{m}}.

Proposition 1.

Every problem that is melem\leq^{\mathrm{elem}}_{\mathrm{m}}-complete for TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) is also melem\leq^{\mathrm{elem}}_{\mathrm{m}}-complete for TOWER\xspace\mathrm{TOWER}\xspace.

Proof 2.1.

Clearly, TOWER\xspace(poly)TOWER\xspace\mathrm{TOWER}\xspace(\mathrm{poly})\subseteq\mathrm{TOWER}\xspace. For the lower bound, let AA be melem\leq^{\mathrm{elem}}_{\mathrm{m}}-complete for TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}), and let BTOWER\xspaceB\in\mathrm{TOWER}\xspace be arbitrary. BB is decidable in time expr(n)(1)\exp_{r(n)}(1) for some elementary rr. Define the set C:={x#0r(|x|)xB}C\vcentcolon=\{x\#0^{r({|\nobreak x\nobreak|})}\mid x\in B\}. First, we show that CTOWER\xspace(poly)C\in\mathrm{TOWER}\xspace(\mathrm{poly}). Consider the algorithm that first checks if the input zz is of the form x#0x\#0^{*}, computes r(|x|)r({|\nobreak x\nobreak|}) in elementary time, checks whether z=x#0r(|x|)z=x\#0^{r({|\nobreak x\nobreak|})}, and then whether xBx\in B. The first two steps clearly take elementary time in nn, where n:=|x#0r(|x|)|n\vcentcolon={|\nobreak x\#0^{r({|\nobreak x\nobreak|})}\nobreak|}, and the final step runs in time expr(|x|)(1)expn(1)\exp_{r({|\nobreak x\nobreak|})}(1)\leq\exp_{n}(1).

By assumption, CmelemAC\leq^{\mathrm{elem}}_{\mathrm{m}}A via an elementary reduction ff. But clearly also BmelemCB\leq^{\mathrm{elem}}_{\mathrm{m}}C by the elementary reduction g:xx#0r(|x|)g\colon x\mapsto x\#0^{r({|\nobreak x\nobreak|})}. As a consequence, the function h:=fgh\vcentcolon=f\circ g is a reduction from BB to AA. hh is computable in time expk1(expk2(n))=expk1+k2(n)\exp_{k_{1}}(\exp_{k_{2}}(n))=\exp_{k_{1}+k_{2}}(n) for fixed k1,k20k_{1},k_{2}\geq 0 depending on ff and gg, and hence again elementary.

Corollary 2.

TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) is not closed under melem\leq^{\mathrm{elem}}_{\mathrm{m}}-reductions.

Proof 2.2.

Suppose TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) is closed under melem\leq^{\mathrm{elem}}_{\mathrm{m}}-reductions, and let AA be any problem complete for TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) (such AA exists; see also our main result, Theorem 31). By the previous proposition, then TOWER\xspaceTOWER\xspace(poly)\mathrm{TOWER}\xspace\subseteq\mathrm{TOWER}\xspace(\mathrm{poly}), contradiction.

3. Modal team logic

We fix a countably infinite set 𝒫𝒮\mathcal{PS} of propositional symbols. Modal team logic 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace, introduced by Müller [Mül14], extends classical modal logic 𝖬𝖫\xspace\mathsf{ML}\xspace. Formulas of classical 𝖬𝖫\xspace\mathsf{ML}\xspace are built following the grammar

α\displaystyle\alpha ::=¬αααααααp,\displaystyle\vcentcolon\vcentcolon=\neg\alpha\mid\alpha\land\alpha\mid\alpha\lor\alpha\mid\Box\alpha\mid\Diamond\alpha\mid p\mid\top\text{,}

where p𝒫𝒮p\in\mathcal{PS} and \top is constant truth. 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace extends 𝖬𝖫\xspace\mathsf{ML}\xspace by the grammar

φ\displaystyle\varphi ::=φφφφφφφα,\displaystyle\vcentcolon\vcentcolon={\sim}\varphi\mid\varphi\land\varphi\mid\varphi\lor\varphi\mid\Box\varphi\mid\Diamond\varphi\mid\alpha\text{,}

where α\alpha denotes an 𝖬𝖫\xspace\mathsf{ML}\xspace-formula.

The set of propositional variables occurring in a formula φ𝖬𝖳𝖫\xspace\varphi\in\mathsf{MTL}\xspace is 𝖯𝗋𝗈𝗉(φ)\mathsf{Prop}(\varphi). We use the common abbreviations :=¬\bot\vcentcolon=\neg\top, αβ:=¬αβ\alpha\rightarrow\beta\vcentcolon=\neg\alpha\lor\beta and αβ:=(αβ)(¬α¬β)\alpha\leftrightarrow\beta\vcentcolon=(\alpha\land\beta)\lor(\neg\alpha\land\neg\beta). For easier distinction, we have classical 𝖬𝖫\xspace\mathsf{ML}\xspace-formulas denoted by α,β,γ,\alpha,\beta,\gamma,\ldots and reserve φ,ψ,ϑ,\varphi,\psi,\vartheta,\ldots for general 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace-formulas.

The modal depth 𝗆𝖽(φ)\mathsf{md}(\varphi) of a formula φ\varphi is recursively defined:

𝗆𝖽(p)\displaystyle\mathsf{md}(p) :=𝗆𝖽()\displaystyle\vcentcolon=\mathsf{md}(\top) :=0\displaystyle\vcentcolon=0
𝗆𝖽(φ)\displaystyle\mathsf{md}({\sim}\varphi) :=𝗆𝖽(¬φ)\displaystyle\vcentcolon=\mathsf{md}(\neg\varphi) :=𝗆𝖽(φ)\displaystyle\vcentcolon=\mathsf{md}(\varphi)
𝗆𝖽(φψ)\displaystyle\mathsf{md}(\varphi\land\psi) :=𝗆𝖽(φψ)\displaystyle\vcentcolon=\mathsf{md}(\varphi\lor\psi) :=max{𝗆𝖽(φ),𝗆𝖽(ψ)}\displaystyle\vcentcolon=\mathrm{max}\{\mathsf{md}(\varphi),\mathsf{md}(\psi)\}
𝗆𝖽(φ)\displaystyle\mathsf{md}(\Diamond\varphi) :=𝗆𝖽(φ)\displaystyle\vcentcolon=\mathsf{md}(\Box\varphi) :=𝗆𝖽(φ)+1\displaystyle\vcentcolon=\mathsf{md}(\varphi)+1

𝖬𝖫\xspacek\mathsf{ML}\xspace_{k} and 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} are the fragments of 𝖬𝖫\xspace\mathsf{ML}\xspace and 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace with modal depth k\leq k, respectively. If the propositions are restricted to a fixed set Φ𝒫𝒮\Phi\subseteq\mathcal{PS} as well, then the fragment is denoted by 𝖬𝖫\xspacekΦ\mathsf{ML}\xspace_{k}^{\Phi}, or 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k}, respectively.

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} be finite. A Kripke structure (over Φ\Phi) is a tuple 𝒦=(W,R,V)\mathcal{K}=(W,R,V), where WW is a set of worlds or points, (W,R)(W,R) is a directed graph called frame, and V:Φ𝔓(W)V\colon\Phi\to\mathfrak{P}(W) is the valuation, with 𝔓(X)\mathfrak{P}(X) being the power set of XX.

Occasionally, by slight abuse of notation, we use the inverse mapping V1:W𝔓(Φ)V^{-1}\colon W\to\mathfrak{P}(\Phi) defined by V1(w):={pΦwV(p)}V^{-1}(w)\vcentcolon=\{p\in\Phi\mid w\in V(p)\} instead of VV, i.e.\xspace, the set of propositions that are true in a given world. If wWw\in W, then (𝒦,w)(\mathcal{K},w) is called pointed structure. 𝖬𝖫\xspace\mathsf{ML}\xspace is evaluated on pointed structures in the classical Kripke semantics.

By contrast, 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is evaluated on pairs (𝒦,T)(\mathcal{K},T) called structures with teams, where 𝒦=(W,R,V)\mathcal{K}=(W,R,V) is a Kripke structure and TWT\subseteq W is called team (in 𝒦\mathcal{K}). Every team TT has an image RT:={vwT,(w,v)R}RT\vcentcolon=\{v\mid w\in T,(w,v)\in R\}, and for wWw\in W, we simply write RwRw instead of R{w}R\{w\}. RiTR^{i}T is inductively defined as R0T:=TR^{0}T\vcentcolon=T and Ri+1T:=RRiTR^{i+1}T\vcentcolon=RR^{i}T. An RR-successor team (or simply successor team) of TT is a team SS such that SRTS\subseteq RT and TR1ST\subseteq R^{-1}S, where R1:={(v,w)(w,v)R}R^{-1}\vcentcolon=\{(v,w)\mid(w,v)\in R\}. Intuitively, SS is formed by picking at least one RR-successor of every world in TT. The semantics of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace can now be defined as follows. 111Often, the “atoms” of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace are restricted to literals p,¬pp,\neg p instead of 𝖬𝖫\xspace\mathsf{ML}\xspace-formulas α\alpha. However, this implies a restriction to formulas in negation normal form, and both definitions are equivalent due to the flatness property of 𝖬𝖫\xspace\mathsf{ML}\xspace (cf. [KMSV15, Proposition 2.2]).

(𝒦,T)α\displaystyle(\mathcal{K},T)\vDash\alpha wT:(𝒦,w)α if α𝖬𝖫\xspace, and otherwise as\displaystyle\Leftrightarrow\;\forall w\in T\colon(\mathcal{K},w)\vDash\alpha\;\text{ if }\alpha\in\mathsf{ML}\xspace\text{, and otherwise as}
(𝒦,T)ψ\displaystyle(\mathcal{K},T)\vDash{\sim}\psi (𝒦,T)ψ,\displaystyle\Leftrightarrow\;(\mathcal{K},T)\nvDash\psi\text{,}
(𝒦,T)ψθ\displaystyle(\mathcal{K},T)\vDash\psi\land\theta (𝒦,T)ψ and (𝒦,T)θ,\displaystyle\Leftrightarrow\;(\mathcal{K},T)\vDash\psi\text{ and }(\mathcal{K},T)\vDash\theta\text{,}
(𝒦,T)ψθ\displaystyle(\mathcal{K},T)\vDash\psi\lor\theta S,UT such that T=SU(𝒦,S)ψ, and (𝒦,U)θ,\displaystyle\Leftrightarrow\;\exists S,U\subseteq T\text{ such that }T=S\cup U\text{, }(\mathcal{K},S)\vDash\psi\text{, and }(\mathcal{K},U)\vDash\theta\text{,}
(𝒦,T)ψ\displaystyle(\mathcal{K},T)\vDash\Diamond\psi (𝒦,S)ψ for some successor team S of T,\displaystyle\Leftrightarrow\;(\mathcal{K},S)\vDash\psi\text{ for some successor team }S\text{ of }T\text{,}
(𝒦,T)ψ\displaystyle(\mathcal{K},T)\vDash\Box\psi (𝒦,RT)ψ.\displaystyle\Leftrightarrow\;(\mathcal{K},RT)\vDash\psi\text{.}

We often omit 𝒦\mathcal{K} and write only TφT\vDash\varphi (for team semantics) or wαw\vDash\alpha (for Kripke semantics).

An 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace-formula φ\varphi is satisfiable if it is true in some structure with team over 𝖯𝗋𝗈𝗉(φ)\mathsf{Prop}(\varphi), which is then called a model of φ\varphi. Analogously, φ\varphi is valid if it is true in every structure with team (over 𝖯𝗋𝗈𝗉(φ)\mathsf{Prop}(\varphi)). For a logic \mathcal{L}, the sets of all satisfiable resp. valid formulas of \mathcal{L} are 𝖲𝖠𝖳()\mathsf{SAT}(\mathcal{L}) and 𝖵𝖠𝖫()\mathsf{VAL}(\mathcal{L}), respectively.

In the literature on team semantics, the empty team is usually excluded in the above definition, since most {\sim}-free logics with team semantics have the empty team property, i.e.\xspace, the empty team satisfies every formula [Vää08, KMSV17, HS15]. However, this distinction is unnecessary for 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace: φ\varphi is satisfiable iff φ\top\lor\varphi is satisfied by some non-empty team222Note that φ\top\lor\varphi is not a tautology in general, since \lor is not the Boolean disjunction. Rather, φ\top\lor\varphi existentially quantifies a subteam where φ\varphi holds. In fact, φ\top\lor\varphi is a tautology if and only if φ\varphi holds in the empty team., and φ\varphi is satisfied by some non-empty team iff φ{\sim}\bot\land\varphi is satisfiable.

The modality-free fragment 𝖬𝖳𝖫\xspace0\mathsf{MTL}\xspace_{0} syntactically coincides with propositional team logic 𝖯𝖳𝖫\xspace\mathsf{PTL}\xspace [HKLV16, HKVV18, YV17]. The usual interpretations of the latter, i.e.\xspace, sets of Boolean assignments, can easily be represented as teams in Kripke structures. For this reason, we treat 𝖯𝖳𝖫\xspace\mathsf{PTL}\xspace and 𝖬𝖳𝖫\xspace0\mathsf{MTL}\xspace_{0} as identical in this article.

Note that the connectives \lor, \rightarrow and ¬\neg are not the Boolean disjunction, implication and negation, except on singleton teams, which correspond to Kripke semantics. Using \land and {\sim} however, we can define team-wide Boolean disjunction φ1∨⃝φ2:=(φ1φ2)\varphi_{1}\ovee\varphi_{2}\vcentcolon={\sim}({\sim}\varphi_{1}\land{\sim}\varphi_{2}) and material implication φ1φ2:=φ1∨⃝φ2\varphi_{1}\rightarrowtriangle\varphi_{2}\vcentcolon={\sim}\varphi_{1}\ovee\varphi_{2}.

The notation iφ\Box^{i}\varphi is defined via 0φ:=φ\Box^{0}\varphi\vcentcolon=\varphi and i+1φ:=iφ\Box^{i+1}\varphi\vcentcolon=\Box\Box^{i}\varphi, and analogously for iφ\Diamond^{i}\varphi. To express that at least one element of a team satisfies α𝖬𝖫\xspace\alpha\in\mathsf{ML}\xspace, we use 𝖤α:=¬α\mathsf{E}\alpha\vcentcolon={\sim}\neg\alpha.

𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace can express the (extended) dependence atom =(α1,,αn1,αn){=\!\!(\alpha_{1},\ldots,\alpha_{n-1},\alpha_{n})} of (extended) modal dependence logic [Vää08, EHM+13], which states that the truth value of αn\alpha_{n} is a function of the truth values of α1,,αn1\alpha_{1},\ldots,\alpha_{n-1}, where α1,,αn𝖬𝖫\xspace\alpha_{1},\ldots,\alpha_{n}\in\mathsf{ML}\xspace. It is definable in 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace as [(i=1n1=(αi)=(αn))]{\sim}\left[\top\lor{\sim}\left(\bigwedge_{i=1}^{n-1}{=\!\!(\alpha_{i})}\rightarrowtriangle{=\!\!(\alpha_{n})}\right)\right], where =(α):=α∨⃝¬α{=\!\!(\alpha)}\vcentcolon=\alpha\ovee\neg\alpha is the constancy atom, stating that the truth value of α𝖬𝖫\xspace\alpha\in\mathsf{ML}\xspace is constant throughout the team.

The well-known bisimulation relation kΦ\rightleftharpoons^{\Phi}_{k} fundamentally characterizes the expressive power of modal logic [BRV01] and plays a key role in our results.

{defi}

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} and k0k\geq 0. For i{1,2}i\in\{1,2\}, let (𝒦i,wi)(\mathcal{K}_{i},w_{i}) be a pointed structure, where 𝒦i=(Wi,Ri,Vi)\mathcal{K}_{i}=(W_{i},R_{i},V_{i}). Then (𝒦1,w1)(\mathcal{K}_{1},w_{1}) and (𝒦2,w2)(\mathcal{K}_{2},w_{2}) are (Φ,k)(\Phi,k)-bisimilar, in symbols (𝒦1,w1)kΦ(𝒦2,w2)(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},w_{2}), if

  • pΦ:w1V1(p)w2V2(p)\forall p\in\Phi\colon w_{1}\in V_{1}(p)\Leftrightarrow w_{2}\in V_{2}(p),

  • and if k>0k>0,

    • v1R1w1:v2R2w2:(𝒦1,v1)k1Φ(𝒦2,v2)\forall v_{1}\in R_{1}w_{1}\colon\exists v_{2}\in R_{2}w_{2}\colon(\mathcal{K}_{1},v_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},v_{2}) (forward condition),

    • v2R2w2:v1R1w1:(𝒦1,v1)k1Φ(𝒦2,v2)\forall v_{2}\in R_{2}w_{2}\colon\exists v_{1}\in R_{1}w_{1}\colon(\mathcal{K}_{1},v_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},v_{2}) (backward condition).

So-called characteristic formulas or Hintikka formulas capture the essence of the bisimulation relation in the following sense:

{propC}

[[GO07, Theorem 32]] Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} be finite, k0k\geq 0, and let (𝒦,w)(\mathcal{K},w) be a pointed structure. Then there is a formula ζ𝖬𝖫\xspacekΦ\zeta\in\mathsf{ML}\xspace^{\Phi}_{k} such that for all pointed structures (𝒦,w)(\mathcal{K}^{\prime},w^{\prime}) we have (𝒦,w)ζ(\mathcal{K}^{\prime},w^{\prime})\vDash\zeta if and only if (𝒦,w)kΦ(𝒦,w)(\mathcal{K},w)\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime},w^{\prime}).

The notion of bisimulation was lifted to team semantics by Hella et al. [HLSV14, KMSV17, KMSV15]:

{defi}

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} and k0k\geq 0. For i{1,2}i\in\{1,2\}, let (𝒦i,Ti)(\mathcal{K}_{i},T_{i}) be a structure with team. Then (𝒦1,T1)(\mathcal{K}_{1},T_{1}) and (𝒦2,T2)(\mathcal{K}_{2},T_{2}) are (Φ,k)(\Phi,k)-team-bisimilar, written (𝒦1,T1)kΦ(𝒦2,T2)(\mathcal{K}_{1},T_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},T_{2}), if

  • w1T1:w2T2:(𝒦1,w1)kΦ(𝒦2,w2)\forall w_{1}\in T_{1}\colon\exists w_{2}\in T_{2}\colon(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},w_{2}),

  • w2T2:w1T1:(𝒦1,w1)kΦ(𝒦2,w2)\forall w_{2}\in T_{2}\colon\exists w_{1}\in T_{1}\colon(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},w_{2}).

If no confusion can arise, we will also refer to teams T1,T2T_{1},T_{2} that are (Φ,k)(\Phi,k)-team-bisimilar simply as (Φ,k)(\Phi,k)-bisimilar. Throughout the paper, we will make use of the following characterizations of bisimilarity.

Proposition 3.

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} be finite, and k0k\geq 0. For i{1,2}i\in\{1,2\}, let (𝒦i,wi)(\mathcal{K}_{i},w_{i}) be a pointed structure, where 𝒦i=(Wi,Ri,Vi)\mathcal{K}_{i}=(W_{i},R_{i},V_{i}). The following statements are equivalent:

  1. (1)

    α𝖬𝖫\xspacekΦ:(𝒦1,w1)α(𝒦2,w2)α\forall\alpha\in\mathsf{ML}\xspace^{\Phi}_{k}\colon(\mathcal{K}_{1},w_{1})\vDash\alpha\Leftrightarrow(\mathcal{K}_{2},w_{2})\vDash\alpha,

  2. (2)

    (𝒦1,w1)kΦ(𝒦2,w2)(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},w_{2}),

  3. (3)

    (𝒦1,{w1})kΦ(𝒦2,{w2})(\mathcal{K}_{1},\{w_{1}\})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},\{w_{2}\}),

and if k>0k>0,

  1. (4)

    (𝒦1,w1)0Φ(𝒦2,w2)(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{0}(\mathcal{K}_{2},w_{2}) and (𝒦1,R1w1)k1Φ(𝒦2,R2w2)(\mathcal{K}_{1},R_{1}w_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},R_{2}w_{2}).

Proof 3.1.

(1) \Leftrightarrow (2) is a standard result ([GO07, Theorem 32]). (2) \Leftrightarrow (3) follows from Definition 3. For k>0k>0, we show that (2) + (3) implies (4). Clearly, (𝒦1,w1)0Φ(𝒦2,w2)(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{0}(\mathcal{K}_{2},w_{2}) follows from (2). Due to Hella et al. [HLSV14, Lemma 3.3], (3) implies (𝒦1,R1w1)k1Φ(𝒦2,R2w2)(\mathcal{K}_{1},R_{1}w_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},R_{2}w_{2}).

Finally, we show (4) \Rightarrow (2). Suppose (𝒦1,w1)0Φ(𝒦2,w2)(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{0}(\mathcal{K}_{2},w_{2}) and (𝒦1,R1w1)k1Φ(𝒦2,R2w2)(\mathcal{K}_{1},R_{1}w_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},R_{2}w_{2}). Then to show (𝒦1,w1)kΦ(𝒦2,w2)(\mathcal{K}_{1},w_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},w_{2}), it is sufficient to prove the forward and backward conditions of Definition 3. Suppose v1R1w1v_{1}\in R_{1}w_{1}. Since (𝒦1,R1w1)k1Φ(𝒦2,R2w2)(\mathcal{K}_{1},R_{1}w_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},R_{2}w_{2}), by Definition 3 there exists v2R2w2v_{2}\in R_{2}w_{2} such that (𝒦1,v1)k1Φ(𝒦2,v2)(\mathcal{K}_{1},v_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},v_{2}), proving the forward condition. The backward condition is symmetric.

As a consequence, the forward and backward condition from Definition 3 can be equivalently stated in terms of team-bisimilarity of the respective image teams. A similar characterization exists for team-bisimilarity:

Proposition 4.

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} be finite, and k0k\geq 0. Let (𝒦i,Ti)(\mathcal{K}_{i},T_{i}) be a structure with team for i{1,2}i\in\{1,2\}. Then the following statements are equivalent:

  1. (1)

    α𝖬𝖫\xspacekΦ:(𝒦1,T1)α(𝒦2,T2)α\forall\alpha\in\mathsf{ML}\xspace^{\Phi}_{k}\colon(\mathcal{K}_{1},T_{1})\vDash\alpha\Leftrightarrow(\mathcal{K}_{2},T_{2})\vDash\alpha,

  2. (2)

    φ𝖬𝖳𝖫\xspacekΦ:(𝒦1,T1)φ(𝒦2,T2)φ\forall\varphi\in\mathsf{MTL}\xspace^{\Phi}_{k}\colon(\mathcal{K}_{1},T_{1})\vDash\varphi\Leftrightarrow(\mathcal{K}_{2},T_{2})\vDash\varphi,

  3. (3)

    (𝒦1,T1)kΦ(𝒦2,T2)(\mathcal{K}_{1},T_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},T_{2}).

Proof 3.2.

The above statements are all true if T1=T2=T_{1}=T_{2}=\emptyset, and they are all false if exactly one of the teams is empty, since a team TT satisfies the 𝖬𝖫\xspace\mathsf{ML}\xspace-formula \bot precisely if T=T=\emptyset. For this reason, we can assume that both T1T_{1} and T2T_{2} are non-empty.

By Kontinen et al. [KMSV15, Proposition 3.10], for non-empty T1,T2T_{1},T_{2} there exists an 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k}-formula φ\varphi that is true in (𝒦1,T1)(\mathcal{K}_{1},T_{1}), but holds in (𝒦2,T2)(\mathcal{K}_{2},T_{2}) if and only if (𝒦1,T1)kΦ(𝒦2,T2)(\mathcal{K}_{1},T_{1})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},T_{2}). This immediately proves (2) \Rightarrow (3). The direction (3) \Rightarrow (2) is due to Kontinen et al. [KMSV15, Proposition 2.8] as well.

Finally, (1) \Leftrightarrow (2) follows from the fact that 𝖬𝖫\xspacekΦ𝖬𝖳𝖫\xspacekΦ\mathsf{ML}\xspace^{\Phi}_{k}\subseteq\mathsf{MTL}\xspace^{\Phi}_{k}, and that conversely every 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k}-formula is equivalent to a formula of the form

\scalerel∨⃝i=1n(αij=1mi𝖤βi,j),\operatorname*{\scalerel*{\ovee}{\sum}}_{i=1}^{n}\Big{(}\alpha_{i}\land\bigwedge_{j=1}^{m_{i}}\mathsf{E}\beta_{i,j}\Big{)}\text{,}

where {α1,,αn,β1,1,,βn,mn}𝖬𝖫\xspacekΦ\{\alpha_{1},\ldots,\alpha_{n},\beta_{1,1},\ldots,\beta_{n,m_{n}}\}\subseteq\mathsf{ML}\xspace^{\Phi}_{k} (see [Lüc18a, Theorem 5.2] or [KMSV15, p. 11]).

Note that the analog of condition 4 in Proposition 3 for team bisimulation is not equivalent: It is possible that (𝒦1,T1)0Φ(𝒦2,T2)(\mathcal{K}_{1},T_{1})\rightleftharpoons^{\Phi}_{0}(\mathcal{K}_{2},T_{2}) and (𝒦1,R1T1)k1Φ(𝒦2,R2T2)(\mathcal{K}_{1},R_{1}T_{1})\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}_{2},R_{2}T_{2}), but (𝒦1,T1)⇌̸kΦ(𝒦2,T2)(\mathcal{K}_{1},T_{1})\not\rightleftharpoons^{\Phi}_{k}(\mathcal{K}_{2},T_{2}).

4. Types and canonical models

Many modal logics admit a “universal” model, also called canonical model. The defining property of a canonical model is that it simultaneously witnesses all satisfiable (sets of) formulas in some of its points. These models are a popular tool for proving the completeness of manifold systems of modal logics; for the explicit construction of such a model for 𝖬𝖫\xspace\mathsf{ML}\xspace, consult, e.g.\xspace, Blackburn et al. [BRV01, Section 4.2].

Unfortunately, any canonical model for 𝖬𝖫\xspace\mathsf{ML}\xspace is necessarily infinite, and consequently impractical for complexity theoretic considerations. Instead, we use so-called (Φ,k)(\Phi,k)-canonical models for finite Φ𝒫𝒮\Phi\subseteq\mathcal{PS} and k\xspacek\in\mathbb{N}\xspace; as the name suggests they are canonical for the fragment 𝖬𝖫\xspacekΦ\mathsf{ML}\xspace^{\Phi}_{k}. While these models are finite, by Proposition 3 their size is at least the number of equivalence classes of kΦ\rightleftharpoons^{\Phi}_{k}. We call the equivalence classes of kΦ\rightleftharpoons^{\Phi}_{k} types.

A first issue arises since types are then proper classes, and in team semantics, we need to speak about sets of types. For this reason, we begin this section by defining types on proper set-theoretic grounds, by indentifying the type of a point with the set of formulas that are true in it, which is a standard approach in first-order model theory.

4.1. Types

{defi}

A set τ𝖬𝖫\xspacekΦ\tau\subseteq\mathsf{ML}\xspace^{\Phi}_{k} is a (Φ,k)(\Phi,k)-type if it is satisfiable and for all α𝖬𝖫\xspacekΦ\alpha\in\mathsf{ML}\xspace^{\Phi}_{k} contains either α\alpha or ¬α\neg\alpha. The (Φ,k)(\Phi,k)-type of a pointed structure (𝒦,w)(\mathcal{K},w) is

𝒦,w:=kΦ{α𝖬𝖫\xspacekΦ(𝒦,w)α}.\llbracket{}\mathcal{K},w\rrbracket{}^{\Phi}_{k}\vcentcolon=\big{\{}\alpha\in\mathsf{ML}\xspace^{\Phi}_{k}\mid(\mathcal{K},w)\vDash\alpha\big{\}}.

The set of all (Φ,k)(\Phi,k)-types is ΔkΦ\Delta^{\Phi}_{k}. Given a team TT in 𝒦\mathcal{K}, the types in TT are

𝒦,T:=kΦ{𝒦,wkΦwT}.\llbracket{}\mathcal{K},T\rrbracket{}^{\Phi}_{k}\vcentcolon=\big{\{}\llbracket{}\mathcal{K},w\rrbracket{}^{\Phi}_{k}\mid w\in T\big{\}}.

The following assertions ascertain that the above definition of types properly reflects the bisimulation relation.

Proposition 5.

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} and k0k\geq 0. Then

  1. (1)

    The unique (Φ,k)(\Phi,k)-type satisfied by (𝒦,w)(\mathcal{K},w) is 𝒦,wkΦ\llbracket{}\mathcal{K},w\rrbracket{}^{\Phi}_{k}.

  2. (2)

    (𝒦,w)kΦ(𝒦,w)(\mathcal{K},w)\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime},w^{\prime}) if and only if 𝒦,w=kΦ𝒦,wkΦ\llbracket{}\mathcal{K},w\rrbracket{}^{\Phi}_{k}=\llbracket{}\mathcal{K}^{\prime},w^{\prime}\rrbracket{}^{\Phi}_{k}.

  3. (3)

    (𝒦,T)kΦ(𝒦,T)(\mathcal{K},T)\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime},T^{\prime}) if and only if 𝒦,T=kΦ𝒦,TkΦ\llbracket{}\mathcal{K},T\rrbracket{}^{\Phi}_{k}=\llbracket{}\mathcal{K}^{\prime},T^{\prime}\rrbracket{}^{\Phi}_{k}.

Proof 4.1.

Property (1) is straightforward: two distinct types τ,τ\tau,\tau^{\prime} satisfied by (𝒦,w)(\mathcal{K},w) differ in some α𝖬𝖫\xspacekΦ\alpha\in\mathsf{ML}\xspace^{\Phi}_{k}. But then (𝒦,w)α,¬α(\mathcal{K},w)\vDash\alpha,\neg\alpha, contradiction. Property (2) immediately follows from Proposition 3. For (3), first consider \Rightarrow. Due to symmetry, we only show that (𝒦,T)kΦ(𝒦,T)(\mathcal{K},T)\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime},T^{\prime}) implies 𝒦,TkΦ𝒦,TkΦ\llbracket{}\mathcal{K},T\rrbracket{}^{\Phi}_{k}\subseteq\llbracket{}\mathcal{K}^{\prime},T^{\prime}\rrbracket{}^{\Phi}_{k}. Hence suppose τ𝒦,TkΦ\tau\in\llbracket{}\mathcal{K},T\rrbracket{}^{\Phi}_{k}. Then there exists wTw\in T of type 𝒦,w=kΦτ\llbracket{}\mathcal{K},w\rrbracket{}^{\Phi}_{k}=\tau. By Definition 3, there is wTw^{\prime}\in T^{\prime} with (𝒦,w)kΦ(𝒦,w)(\mathcal{K},w)\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime},w^{\prime}). Then 𝒦,w=kΦτ𝒦,TkΦ\llbracket{}\mathcal{K}^{\prime},w^{\prime}\rrbracket{}^{\Phi}_{k}=\tau\in\llbracket{}\mathcal{K}^{\prime},T^{\prime}\rrbracket{}^{\Phi}_{k} by property (2). The direction \Leftarrow of (3) is shown analogously.

It is unsurprising that the type of a point ww is determined solely by the propositions in ww and the types in the image RwRw. In other words, all pointed structures of type τ\tau satisfy the same propositions in their roots, viz. τΦ\tau\cap\Phi, and have the same types contained in their image teams. Regarding the latter, we define τ:={τΔkΦ{αατ}τ}\mathcal{R}\tau\vcentcolon=\big{\{}\tau^{\prime}\in\Delta^{\Phi}_{k}\mid\{\alpha\mid\Box\alpha\in\tau\}\subseteq\tau^{\prime}\big{\}}, given a (Φ,k+1)(\Phi,k+1)-type τ\tau. Intuitively, τ\mathcal{R}\tau is the set of (Φ,k)(\Phi,k)-types that occur in the image team of a world of type τ\tau.

The following proposition shows that types are indeed uniquely determined by the above constituents:

Proposition 6.

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} be finite and k0k\geq 0.

  1. (1)

    wkΦΦ=V1(w)Φ\llbracket{}w\rrbracket{}^{\Phi}_{k}\cap\Phi=V^{-1}(w)\cap\Phi and Rw=kΦwk+1Φ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}=\mathcal{R}\llbracket{}w\rrbracket{}^{\Phi}_{k+1}, for all pointed structures (W,R,V,w)(W,R,V,w).

  2. (2)

    The mapping h:ττΦh\colon\tau\mapsto\tau\cap\Phi is a bijection from Δ0Φ\Delta^{\Phi}_{0} to 𝔓(Φ)\mathfrak{P}(\Phi).

  3. (3)

    The mapping h:τ(τΦ,τ)h\colon\tau\mapsto(\tau\cap\Phi,\mathcal{R}\tau) is a bijection from Δk+1Φ\Delta^{\Phi}_{k+1} to 𝔓(Φ)×𝔓(ΔkΦ)\mathfrak{P}(\Phi)\times\mathfrak{P}(\Delta^{\Phi}_{k}).

Proof 4.2.

See the appendix.

Lemma 7.

Let (W,R,V,w)(W,R,V,w) be a pointed structure.

  1. (1)

    If τΔ0Φ\tau\in\Delta^{\Phi}_{0}, then w=0Φτ\llbracket{}w\rrbracket{}^{\Phi}_{0}=\tau if and only if V1(w)=τΦV^{-1}(w)=\tau\cap\Phi.

  2. (2)

    If τΔk+1Φ\tau\in\Delta^{\Phi}_{k+1}, then w=k+1Φτ\llbracket{}w\rrbracket{}^{\Phi}_{k+1}=\tau if and only if V1(w)=τΦV^{-1}(w)=\tau\cap\Phi and Rw=kΦτ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}=\mathcal{R}\tau.

Proof 4.3.

The direction \Rightarrow of 1. and 2. follows directly from Proposition 6. Moreover, we prove \Leftarrow only for statement 2., as the proof is analogous for 1.

Suppose that there are τ,τΔk+1Φ\tau,\tau^{\prime}\in\Delta^{\Phi}_{k+1} such that V1(w)=τΦV^{-1}(w)=\tau\cap\Phi and Rw=kΦτ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}=\mathcal{R}\tau, but w=k+1Φτ\llbracket{}w\rrbracket{}^{\Phi}_{k+1}=\tau^{\prime}. Then, by \Rightarrow, we have V1(w)=τΦV^{-1}(w)=\tau^{\prime}\cap\Phi and Rw=kΦτ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}=\mathcal{R}\tau^{\prime} as well. In other words, τΦ=τΦ\tau\cap\Phi=\tau^{\prime}\cap\Phi and τ=τ\mathcal{R}\tau=\mathcal{R}\tau^{\prime}. However, since the mapping h:τ(τΦ,τ)h\colon\tau\mapsto(\tau\cap\Phi,\mathcal{R}\tau) is bijective according to Proposition 6, we have τ=τ=wk+1Φ\tau=\tau^{\prime}=\llbracket{}w\rrbracket{}^{\Phi}_{k+1}.

We are now ready to state the formal definition of canonicity by the notion of types:

{defi}

A structure with team (𝒦,T)(\mathcal{K},T) is (Φ,k)(\Phi,k)-canonical if 𝒦,T=kΦΔkΦ\llbracket{}\mathcal{K},T\rrbracket{}^{\Phi}_{k}=\Delta^{\Phi}_{k}.

In the following, we often omit Φ\Phi and 𝒦\mathcal{K} and instead write wk\llbracket{}w\rrbracket{}_{k} and Tk\llbracket{}T\rrbracket{}_{k}, respectively, and simply say that TT is (Φ,k)(\Phi,k)-canonical if 𝒦\mathcal{K} is clear.

4.2. Canonical models in team semantics

It is a standard result that for every Φ\Phi and k0k\geq 0 there exists a (Φ,k)(\Phi,k)-canonical model [BRV01], or in other words, that the logic 𝖬𝖫\xspacekΦ\mathsf{ML}\xspace^{\Phi}_{k} admits canonical models.

We will show that, given a (Φ,k)(\Phi,k)-canonical model 𝒦\mathcal{K}, every satisfiable 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k}-formula can be satisfied in some team of 𝒦\mathcal{K} as well, despite 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace being significantly more expressive than 𝖬𝖫\xspace\mathsf{ML}\xspace [KMSV15]. In other words, the canonical models for 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k} and 𝖬𝖫\xspacekΦ\mathsf{ML}\xspace^{\Phi}_{k} coincide:

Theorem 8.

Let (𝒦,T)(\mathcal{K},T) be (Φ,k)(\Phi,k)-canonical and φ𝖬𝖳𝖫\xspacekΦ\varphi\in\mathsf{MTL}\xspace^{\Phi}_{k}. Then φ\varphi is satisfiable if and only if (𝒦,T)φ(\mathcal{K},T^{\prime})\vDash\varphi for some TTT^{\prime}\subseteq T.

Proof 4.4.

Assume (𝒦,T)(\mathcal{K},T) and φ\varphi are as above. As the direction from right to left is trivial, suppose that φ\varphi is satisfiable, i.e.\xspace, has a model (𝒦^,T^)(\hat{\mathcal{K}},\hat{T}). As a team in 𝒦\mathcal{K} that satisfies φ\varphi, we define

T:={wT|𝒦,wkΦ𝒦^,T^kΦ}.T^{\prime}\vcentcolon=\Set{w\in T}{\llbracket{}\mathcal{K},w\rrbracket{}^{\Phi}_{k}\in\llbracket{}\hat{\mathcal{K}},\hat{T}\rrbracket{}^{\Phi}_{k}}\text{.}

By Proposition 4 and 5, it suffices to prove 𝒦^,T^=kΦ𝒦,TkΦ\llbracket{}\hat{\mathcal{K}},\hat{T}\rrbracket{}^{\Phi}_{k}=\llbracket{}\mathcal{K},T^{\prime}\rrbracket{}^{\Phi}_{k}. Moreover, the direction \supseteq is clear by definition. As TT is (Φ,k)(\Phi,k)-canonical, for every τ𝒦^,T^kΦ\tau\in\llbracket{}\hat{\mathcal{K}},\hat{T}\rrbracket{}^{\Phi}_{k} there exists a world wTw\in T of type τ\tau. Consequently, 𝒦^,T^kΦ𝒦,TkΦ\llbracket{}\hat{\mathcal{K}},\hat{T}\rrbracket{}^{\Phi}_{k}\subseteq\llbracket{}\mathcal{K},T^{\prime}\rrbracket{}^{\Phi}_{k}.

How large is a (Φ,k)(\Phi,k)-canonical model at least? The number of types is captured by the function expk\exp^{*}_{k}, defined by

exp0(n):=nexpk+1(n):=n2expk(n).\exp^{*}_{0}(n)\vcentcolon=n\qquad\qquad\exp^{*}_{k+1}(n)\vcentcolon=n\cdot 2^{\exp^{*}_{k}(n)}\text{.}
Proposition 9.

|ΔkΦ|=expk(2|Φ|){|\nobreak\Delta^{\Phi}_{k}\nobreak|}=\exp^{*}_{k}\big{(}2^{|\nobreak\Phi\nobreak|}\big{)} for all k0k\geq 0 and finite Φ𝒫𝒮\Phi\subseteq\mathcal{PS}.

Proof 4.5.

By induction on kk. For the base case k=0k=0, this follows from Proposition 6, as there is a bijection between Δ0Φ\Delta^{\Phi}_{0} and 𝔓(Φ)\mathfrak{P}(\Phi) and exp0(2|Φ|)=2|Φ|=|Δ0Φ|\exp^{*}_{0}\big{(}2^{|\nobreak\Phi\nobreak|}\big{)}=2^{|\nobreak\Phi\nobreak|}={|\nobreak\Delta^{\Phi}_{0}\nobreak|}.

We proceed with the inductive step, i.e.\xspace, k+1k+1. First note that by induction hypothesis

expk+1(2|Φ|)=2|Φ|2expk(2|Φ|)=|𝔓(Φ)×𝔓(ΔkΦ)|.\exp^{*}_{k+1}\big{(}2^{|\nobreak\Phi\nobreak|}\big{)}=2^{|\nobreak\Phi\nobreak|}\cdot 2^{\exp^{*}_{k}(2^{|\nobreak\Phi\nobreak|})}={|\nobreak\mathfrak{P}(\Phi)\times\mathfrak{P}(\Delta^{\Phi}_{k})\nobreak|}\text{.}

Again, there exists a bijection from Δk+1Φ\Delta^{\Phi}_{k+1} to 𝔓(Φ)×𝔓(ΔkΦ)\mathfrak{P}(\Phi)\times\mathfrak{P}(\Delta^{\Phi}_{k}) by Proposition 6.

Next, we present an algorithm that solves the satisfiability and validity problems of 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} by computing a canonical model. Let us first explicate this construction in a lemma.

Lemma 10.

There is an algorithm that, given Φ𝒫𝒮\Phi\subseteq\mathcal{PS} and k0k\geq 0, computes a (Φ,k)(\Phi,k)-canonical model in time polynomial in |ΔkΦ|{|\nobreak\Delta^{\Phi}_{k}\nobreak|}.

Proof 4.6.

The idea is to construct sets L0L1LkL_{0}\cup L_{1}\cup\cdots\cup L_{k} of worlds in stage-wise manner such that LiL_{i} is (Φ,i)(\Phi,i)-canonical. For L0L_{0}, we simply add a world ww for each Φ𝔓(Φ)\Phi^{\prime}\in\mathfrak{P}(\Phi) such that V1(w)=ΦV^{-1}(w)=\Phi^{\prime}. For i>0i>0, we iterate over all L𝔓(Li1)L^{\prime}\in\mathfrak{P}(L_{i-1}) and Φ𝔓(Φ)\Phi^{\prime}\in\mathfrak{P}(\Phi) and insert a new world ww into LiL_{i} such that LL^{\prime} is the image of ww and such that again V1(w)=ΦV^{-1}(w)=\Phi^{\prime}. An inductive argument based on Proposition 4 and 6 shows that LiL_{i} is (Φ,i)(\Phi,i)-canonical for all i{0,,k}i\in\{0,\ldots,k\}. As k|ΔkΦ|k\leq{|\nobreak\Delta^{\Phi}_{k}\nobreak|}, and each LiL_{i} is constructed in time polynomial in |ΔiΦ||ΔkΦ|{|\nobreak\Delta^{\Phi}_{i}\nobreak|}\leq{|\nobreak\Delta^{\Phi}_{k}\nobreak|}, the overall runtime is polynomial in |ΔkΦ|{|\nobreak\Delta^{\Phi}_{k}\nobreak|}.

With the help of a small lemma, we conclude the upper bound for the satisfiability and validity problem of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace and its fragments.

Lemma 11.

For every polynomial pp there is a polynomial qq such that

p(expk(n))expk(q((k+1)n))p(\exp^{*}_{k}(n))\leq\exp_{k}(q((k+1)\cdot n))

for all k0k\geq 0 and n1n\geq 1.

Proof 4.7.

See the appendix.

Theorem 12.

𝖲𝖠𝖳(𝖬𝖳𝖫\xspacek)\mathsf{SAT}(\mathsf{MTL}\xspace_{k}) and 𝖵𝖠𝖫(𝖬𝖳𝖫\xspacek)\mathsf{VAL}(\mathsf{MTL}\xspace_{k}) are in ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}).

Proof 4.8.

Consider the following algorithm. Let φ𝖬𝖳𝖫\xspacek\varphi\in\mathsf{MTL}\xspace_{k} be the input, n:=|φ|n\vcentcolon={|\nobreak\varphi\nobreak|}, and Φ:=𝖯𝗋𝗈𝗉(φ)\Phi\vcentcolon=\mathsf{Prop}(\varphi). Construct deterministically, as in Lemma 10, a (Φ,k)(\Phi,k)-canonical structure 𝒦=(W,R,V)\mathcal{K}=(W,R,V) in time p(|ΔkΦ|)p({|\nobreak\Delta^{\Phi}_{k}\nobreak|}) for a polynomial pp.

By a result of Müller [Mül14], the model checking problem of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is solvable by an alternating Turing machine that has runtime polynomial in |φ|+|𝒦|{|\nobreak\varphi\nobreak|}+{|\nobreak\mathcal{K}\nobreak|}, and alternations polynomial in |φ|{|\nobreak\varphi\nobreak|}. We call this algorithm as a subroutine: by Theorem 8, φ\varphi is satisfiable (resp. valid) if and only if for at least one subteam (resp. all subteams) TWT\subseteq W we have (𝒦,T)φ(\mathcal{K},T)\vDash\varphi. Equivalently, this is the case if and only if (𝒦,W)(\mathcal{K},W) satisfies φ\top\lor\varphi (resp. (φ){\sim}(\top\lor{\sim}\varphi)).

Let us turn to the overall runtime. 𝒦\mathcal{K} is constructed in time polynomial in |ΔkΦ|=expk(2|Φ|)expk+1(|Φ|)expk+1(n){|\nobreak\Delta^{\Phi}_{k}\nobreak|}=\exp^{*}_{k}(2^{|\nobreak\Phi\nobreak|})\leq\exp^{*}_{k+1}({|\nobreak\Phi\nobreak|})\leq\exp^{*}_{k+1}(n). The subsequent model checking runs in time polynomial in |𝒦|+n{|\nobreak\mathcal{K}\nobreak|}+n, and hence polynomial in expk+1(n)\exp^{*}_{k+1}(n) as well. By Lemma 11, we obtain a total runtime of expk+1(q((k+2)n))\exp_{k+1}(q((k+2)\cdot n)) for a polynomial qq.

The upper bound for 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is proved identically, since k:=𝗆𝖽(φ)k\vcentcolon=\mathsf{md}(\varphi) is polynomial in |φ|{|\nobreak\varphi\nobreak|}.

Corollary 13.

𝖲𝖠𝖳(𝖬𝖳𝖫\xspace)\mathsf{SAT}(\mathsf{MTL}\xspace) and 𝖵𝖠𝖫(𝖬𝖳𝖫\xspace)\mathsf{VAL}(\mathsf{MTL}\xspace) are in TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}).

The usual definition of a canonical model is a structure that has all (infinite) maximal consistent subsets of a certain class of modal formulas as worlds (see virtually any textbook on modal logic, e.g.\xspace[BRV01]). This indeed results in a finite number of worlds in the case of, say, 𝖬𝖫\xspacekΦ\mathsf{ML}\xspace^{\Phi}_{k} (cf. [Cre83, CH96]). Truly finitary constructions of canonical models can be traced back to Fine [Fin75], whose work has been extended towards various other modal systems (e.g.\xspace, by Moss [Mos07]). Furthermore, Cresswell and Hughes [CH96] used mini canonical models, models that are “canonical” only with respect to all subformulas of a fixed 𝖬𝖫\xspace\mathsf{ML}\xspace-formula, which allows them to be finite models with finite sets of formulas as worlds.

All these approaches have in common that they still are non-constructive and intended for completeness proofs. Even computing a “mini canonical model” would not be guaranteed to be feasible enough for 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace: This would require an explicit translation of a given input 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k}-formula to a Boolean combination of 𝖬𝖫\xspacekΦ\mathsf{ML}\xspace^{\Phi}_{k}-formulas first (see the proof of Proposition 4), and it is open whether there is an elementary translation for every fixed kk (cf. [Lüc18a]).

In this light, our approach yields a purely constructive definition of a canonical model (in Lemma 10), which can easily be plugged into the algorithms used for the above results, and has optimal runtime up to a polynomial.

5. Scopes and Subteam Quantifiers

Kontinen et al. [KMSV15] proved that 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is expressively complete up to bisimulation: it can define every property of teams that is (Φ,k)(\Phi,k)-bisimulation invariant, that is closed under kΦ\rightleftharpoons^{\Phi}_{k}, for some finite Φ\Phi and kk. Two team properties that fall into this category are in fact (Φ,k)(\Phi,k)-bisimilarity itself—in the sense that all worlds in a team have the same (Φ,k)(\Phi,k)-type—as well as (Φ,k)(\Phi,k)-canonicity. Consequently, these properties are definable by 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k}-formulas. However, by a simple counting argument, formulas defining arbitrary team properties require non-elementary size w. r. t.\xspaceΦ\Phi and kk.

In this section, we consider a special class of structures, and on these, define kk-bisimilarity by a formula χk\chi_{k} of polynomial size in Φ\Phi and kk. (From now on, we always assume some finite Φ𝒫𝒮\Phi\subseteq\mathcal{PS} and omit it in the notation, i.e.\xspace, we write kk-canonicity, kk-bisimilarity, k\rightleftharpoons_{k}, and so on.) Afterwards, in Section 6 we devise a formula 𝖼𝖺𝗇𝗈𝗇k\mathsf{canon}_{k} of polynomial size that expresses kk-canonicity.

5.1. Scopes

It is natural to implement kk-bisimilarity by mutual recursion in the spirit of Proposition 3: the (k+1)(k+1)-bisimilarity of two points w,vw,v is expressed in terms of kk-team-bisimilarity of RwRw and RvRv, and conversely, to verify kk-team-bisimilarity of RwRw and RvRv, we proceed analogously to the forward and backward conditions of Definition 3 and reduce the problem to checking kk-bisimilarity of pairs of points in RwRw and RvRv.

𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace-formulas define team properties, but we want to express a relation between teams such as RwRw and RvRv. For this reason, we consider the “marked union” of RwRw and RvRv as a single team using the following tool. Formally, if α𝖬𝖫\xspace\alpha\in\mathsf{ML}\xspace, then the “conditioned” subteam TαTT_{\alpha}\subseteq T is defined as

Tα:={wT|wα}.T_{\alpha}\vcentcolon=\Set{w\in T}{w\vDash\alpha}\text{.}

In the literature, TαT_{\alpha} is also written TαT\upharpoonright\alpha [Gal15, Gal16, Gal18]. The corresponding “decoding” operator

αφ:=¬α(αφ)\alpha\hookrightarrow\varphi\vcentcolon=\neg\alpha\lor(\alpha\land\varphi)

was introduced by Galliani [Gal15, Gal16, Gal18] as well: αφ\alpha\hookrightarrow\varphi is true in TT if and only if TαφT_{\alpha}\vDash\varphi.

Now, instead of defining an nn-ary relation on teams, a formula φ\varphi can define a unary relation—a team property—parameterized by formulas α1,,αn𝖬𝖫\xspace\alpha_{1},\ldots,\alpha_{n}\in\mathsf{ML}\xspace. We emphasize this by writing φ(α1,,αn)\varphi(\alpha_{1},\ldots,\alpha_{n}).

It will be useful if the “markers” of the constituent teams are invariant under traversing edges in the structure. In that case, we call these formulas scopes:

{defi}

Let 𝒦=(W,R,V)\mathcal{K}=(W,R,V) be a Kripke structure. A formula α𝖬𝖫\xspace\alpha\in\mathsf{ML}\xspace is called a scope (in 𝒦\mathcal{K}) if (w,v)R(w,v)\in R implies wαvαw\vDash\alpha\Leftrightarrow v\vDash\alpha. Two scopes α,β\alpha,\beta are called disjoint (in 𝒦\mathcal{K}) if WαW_{\alpha} and WβW_{\beta} are disjoint.

To avoid interference, we always assume that scopes are formulas in 𝖬𝖫\xspace0𝒫𝒮Φ\mathsf{ML}\xspace^{\mathcal{PS}\setminus\Phi}_{0}, i.e.\xspace, they are always purely propositional and do not contain propositions from Φ\Phi.

TTSSα1\alpha_{1}α3\alpha_{3}α2\alpha_{2}\Rightarrow α1\alpha_{1}α3\alpha_{3}α2\alpha_{2}SSTSα2T^{\alpha_{2}}_{\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}S}
Figure 1. Example of subteam selection in the scope α2\alpha_{2}

It is desirable to be able to speak about subteams in a specific scope. If SS is a team, let TSα:=T¬α(TαS)T^{\alpha}_{S}\vcentcolon=T_{\neg\alpha}\cup(T_{\alpha}\cap S). For singletons {w}\{w\}, we simply write TwαT^{\alpha}_{w} instead of T{w}αT^{\alpha}_{\{w\}}. Intuitively, TSαT^{\alpha}_{S} is obtained from TT by “shrinking” the subteam TαT_{\alpha} down to SS without impairing TTαT\setminus T_{\alpha} (see Figure 1 for an example). Scopes have several desirable properties:

Proposition 14.

Let α,β\alpha,\beta be disjoint scopes and S,U,TS,U,T teams in a Kripke structure 𝒦=(W,R,V)\mathcal{K}=(W,R,V). Then the following laws hold:

  1. (1)

    Distributive laws: (TS)α=TαS=TSα=TαSα{(T\cap S)}_{\alpha}=T_{\alpha}\cap S=T\cap S_{\alpha}=T_{\alpha}\cap S_{\alpha} and (TS)α=TαSα{(T\cup S)}_{\alpha}=T_{\alpha}\cup S_{\alpha}.

  2. (2)

    Disjoint selection commutes: (TSα)Uβ=(TUβ)Sα{\big{(}T^{\alpha}_{S}\big{)}}^{\beta}_{U}={\big{(}T^{\beta}_{U}\big{)}}^{\alpha}_{S}.

  3. (3)

    Disjoint selection is independent: ((TSα)Uβ)α=TαS{\big{(}{(T^{\alpha}_{S})}^{\beta}_{U}\big{)}}_{\alpha}=T_{\alpha}\cap S.

  4. (4)

    Image and selection commute: (RT)α=(R(Tα))α=R(Tα){(RT)}_{\alpha}={\big{(}R(T_{\alpha})\big{)}}_{\alpha}=R(T_{\alpha})

  5. (5)

    Selection propagates: If STS\subseteq T, then R(TSα)=(RT)RSαR{\big{(}T^{\alpha}_{S}\big{)}}={(RT)}^{\alpha}_{RS}.

Proof 5.1.

Straightforward; see the appendix.

Accordingly, we write RiTαR^{i}T_{\alpha} instead of (RiT)α{(R^{i}T)}_{\alpha} or Ri(Tα)R^{i}(T_{\alpha}) and TS1,S2α1,α2T^{\alpha_{1},\alpha_{2}}_{S_{1},S_{2}} for (TS1α1)S2α2{(T^{\alpha_{1}}_{S_{1}})}^{\alpha_{2}}_{S_{2}}.

5.2. Subteam quantifiers

We refer to the following abbreviations as subteam quantifiers, where α𝖬𝖫\xspace\alpha\in\mathsf{ML}\xspace:

αφ\displaystyle\exists^{\subseteq}_{\alpha}\;\varphi :=αφ\displaystyle\vcentcolon=\alpha\lor\varphi αφ\displaystyle\forall^{\subseteq}_{\alpha}\;\varphi :=αφ\displaystyle\vcentcolon={\sim}\exists^{\subseteq}_{\alpha}{\sim}\varphi
α1φ\displaystyle\exists^{1}_{\alpha}\;\varphi :=α[𝖤αα(𝖤αφ)]\displaystyle\vcentcolon=\exists^{\subseteq}_{\alpha}\left[\mathsf{E}\alpha\land\forall^{\subseteq}_{\alpha}(\mathsf{E}\alpha\rightarrowtriangle\varphi)\right] α1φ\displaystyle\forall^{1}_{\alpha}\;\varphi :=α1φ\displaystyle\vcentcolon={\sim}\exists^{1}_{\alpha}{\sim}\varphi

Intuitively, they quantify over subteams STαS\subseteq T_{\alpha} or worlds wTαw\in T_{\alpha} such that TSαT^{\alpha}_{S} resp. TwαT^{\alpha}_{w} satisfies φ\varphi.

Proposition 15.

The subteam quantifiers have the following semantics:

T\displaystyle\qquad T\;\vDash\;\; αφ\displaystyle\exists^{\subseteq}_{\alpha}\varphi STα\displaystyle\;\Leftrightarrow\;\exists S\subseteq T_{\alpha} :TSαφ\displaystyle\colon T^{\alpha}_{S}\vDash\varphi T\displaystyle\qquad\qquad T\;\vDash\;\; α1φ\displaystyle\exists^{1}_{\alpha}\varphi wTα\displaystyle\;\Leftrightarrow\;\exists w\in T_{\alpha} :Twαφ\displaystyle\colon T^{\alpha}_{w}\vDash\varphi
T\displaystyle\qquad T\;\vDash\;\; αφ\displaystyle\forall^{\subseteq}_{\alpha}\varphi STα\displaystyle\;\Leftrightarrow\;\forall S\subseteq T_{\alpha} :TSαφ\displaystyle\colon T^{\alpha}_{S}\vDash\varphi T\displaystyle\qquad\qquad T\;\vDash\;\; α1φ\displaystyle\forall^{1}_{\alpha}\varphi wTα\displaystyle\;\Leftrightarrow\;\forall w\in T_{\alpha} :Twαφ\displaystyle\colon T^{\alpha}_{w}\vDash\varphi
Proof 5.2.

We prove the existential cases, as the other ones work dually.

Let us first consider the \Rightarrow direction for α\exists^{\subseteq}_{\alpha}. Accordingly, suppose TαφT\vDash\exists^{\subseteq}_{\alpha}\,\varphi, i.e.\xspace, TαφT\vDash\alpha\lor\varphi. Then there exist STS\subseteq T and UTαU\subseteq T_{\alpha} such that SφS\vDash\varphi and T=SUT=S\cup U. Since UT¬α=U\cap T_{\neg\alpha}=\emptyset, it holds T¬αST_{\neg\alpha}\subseteq S. Moreover, S=(STα)(ST¬α)=((STα)Tα)T¬α=TSTααS=(S\cap T_{\alpha})\cup(S\cap T_{\neg\alpha})=((S\cap T_{\alpha})\cap T_{\alpha})\cup T_{\neg\alpha}=T^{\alpha}_{S\cap T_{\alpha}}. Consequently, TSTααφT^{\alpha}_{S\cap T_{\alpha}}\vDash\varphi for some set STαTαS\cap T_{\alpha}\subseteq T_{\alpha}.

For \Leftarrow, suppose TSαφT^{\alpha}_{S}\vDash\varphi for some STαS\subseteq T_{\alpha}. Then TSαT^{\alpha}_{S} and TTSαT\setminus T^{\alpha}_{S} form a division of TT. Since TTSα=T(T¬α(TαS))TT¬α=TαT\setminus T^{\alpha}_{S}=T\setminus\left(T_{\neg\alpha}\cup(T_{\alpha}\cap S)\right)\subseteq T\setminus T_{\neg\alpha}=T_{\alpha}, it holds TTSααT\setminus T^{\alpha}_{S}\vDash\alpha. As a consequence, TαφT\vDash\alpha\lor\varphi.

We proceed with α1\exists^{1}_{\alpha}. For \Rightarrow, suppose that Tα1φT\vDash\exists^{1}_{\alpha}\varphi. Then there exists STαS\subseteq T_{\alpha} such that TSα𝖤αα(𝖤αφ)T^{\alpha}_{S}\vDash\mathsf{E}\alpha\land\forall^{\subseteq}_{\alpha}(\mathsf{E}\alpha\rightarrowtriangle\varphi). Since TSα𝖤αT^{\alpha}_{S}\vDash\mathsf{E}\alpha, there exists w(TSα)αw\in{(T^{\alpha}_{S})}_{\alpha}. As α\forall^{\subseteq}_{\alpha} now applies to (TSα){w}α=Twα{(T^{\alpha}_{S})}^{\alpha}_{\{w\}}=T^{\alpha}_{w} as well, it follows Twα𝖤αφT^{\alpha}_{w}\vDash\mathsf{E}\alpha\rightarrowtriangle\varphi, and consequently TwαφT^{\alpha}_{w}\vDash\varphi.

Suppose for \Leftarrow that TwαφT^{\alpha}_{w}\vDash\varphi for some wTαw\in T_{\alpha}. Let STαS\subseteq T_{\alpha} be arbitrary. If wSw\notin S, then (Twα)Sα=Tα𝖤α{(T^{\alpha}_{w})}^{\alpha}_{S}=T^{\alpha}_{\emptyset}\nvDash\mathsf{E}\alpha, and if wSw\in S, then (Twα)Sα=Twαφ{(T^{\alpha}_{w})}^{\alpha}_{S}=T^{\alpha}_{w}\vDash\varphi. Therefore, for any STαS\subseteq T_{\alpha} it holds (Twα)Sα(𝖤αφ){(T^{\alpha}_{w})}^{\alpha}_{S}\vDash(\mathsf{E}\alpha\rightarrowtriangle\varphi), so Twαα(𝖤αφ)T^{\alpha}_{w}\vDash\forall^{\subseteq}_{\alpha}(\mathsf{E}\alpha\rightarrowtriangle\varphi). Since also Twα𝖤αT^{\alpha}_{w}\vDash\mathsf{E}\alpha, it follows Tα[𝖤αα(𝖤αφ)]T\vDash\exists^{\subseteq}_{\alpha}\left[\mathsf{E}\alpha\land\forall^{\subseteq}_{\alpha}(\mathsf{E}\alpha\rightarrowtriangle\varphi)\right].

5.3. Implementing bisimulation

With scopes and subteam quantifiers at our hands, we have all ingredients to implement kk-bisimulation.

χ0(α,β)\displaystyle\chi_{0}(\alpha,\beta) :=(αβ)pΦ=(p)\displaystyle\vcentcolon=(\alpha\lor\beta)\hookrightarrow\bigwedge_{p\in\Phi}{=\!\!(p)}
χk+1(α,β)\displaystyle\chi_{k+1}(\alpha,\beta) :=χ0(α,β)χk(α,β)\displaystyle\vcentcolon=\chi_{0}(\alpha,\beta)\land\Box\chi^{*}_{k}(\alpha,\beta)
χk(α,β)\displaystyle\chi^{*}_{k}(\alpha,\beta) :=(¬α¬β)∨⃝(𝖤α𝖤β[(α∨⃝β)(𝖤α𝖤βα1β1χk(α,β))])\displaystyle\vcentcolon=(\neg\alpha\land\neg\beta)\ovee\Big{(}\mathsf{E}\alpha\land\mathsf{E}\beta\land{\sim}\big{[}(\alpha\ovee\beta)\lor(\mathsf{E}\alpha\land\mathsf{E}\beta\land{\sim}\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta))\big{]}\Big{)}

Note that a literal translation of the forward and backward condition would rather result in the formula χk(α,β):=α1β1χk(α,β)β1α1χk(α,β)\chi^{*}_{k}(\alpha,\beta)\vcentcolon=\forall^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta)\land\forall^{1}_{\beta}\exists^{1}_{\alpha}\chi_{k}(\alpha,\beta). The more complicated formula shown above however avoids the exponential size that would come with two recursive calls.

Theorem 16.

Let k0k\geq 0. For all Kripke structures 𝒦\mathcal{K}, teams TT and disjoint scopes α,β\alpha,\beta in 𝒦\mathcal{K}, and points wTαw\in T_{\alpha} and vTβv\in T_{\beta} it holds:

Tw,vα,β\displaystyle T^{\alpha,\beta}_{w,v} χk(α,β)\displaystyle\;\vDash\;\chi_{k}(\alpha,\beta)\; if and only if wk\displaystyle w\rightleftharpoons_{k}\, v,\displaystyle v\text{,}
T\displaystyle T χk(α,β)\displaystyle\;\vDash\;\chi^{*}_{k}(\alpha,\beta)\; if and only if Tαk\displaystyle T_{\alpha}\rightleftharpoons_{k}\, Tβ.\displaystyle T_{\beta}\text{.}

Moreover, both χk(α,β)\chi_{k}(\alpha,\beta) and χk(α,β)\chi^{*}_{k}(\alpha,\beta) are 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formulas that are constructible in space 𝒪(log(k+|Φ|+|α|+|β|)){\mathcal{O}(\log(k+{|\nobreak\Phi\nobreak|}+{|\nobreak\alpha\nobreak|}+{|\nobreak\beta\nobreak|}))}.

α\alphaβ\betaTTzz0\rightleftharpoons_{0}\!0\rightleftharpoons_{0}\!0\rightleftharpoons_{0} 1\rightleftharpoons_{1}? \Rightarrow 0\rightleftharpoons_{0}0\rightleftharpoons_{0}0\rightleftharpoons_{0}\RightarrowRTRTzz zzRTzβRT^{\beta}_{z}
Figure 2. As zz violates the backward condition, shrinking RTβRT_{\beta} leads to a 0\rightleftharpoons_{0}-free subteam, falsifying α1β1χ0(α,β)\exists^{1}_{\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\alpha}\exists^{1}_{\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\beta}\color[rgb]{0,0.5,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,0}\chi_{0}(\alpha,\beta).
Proof 5.3.

The idea is to isolate a single point in zTαTβz\in T_{\alpha}\cup T_{\beta} that serves as a counter-example against Tα=kTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}=\llbracket{}T_{\beta}\rrbracket{}_{k} by, say, zkTβkTαk\llbracket{}z\rrbracket{}_{k}\in\llbracket{}T_{\beta}\rrbracket{}_{k}\setminus\llbracket{}T_{\alpha}\rrbracket{}_{k}. We erase Tβ{z}T_{\beta}\setminus\{z\} from TT using the disjunction \lor, as Tβ{z}α∨⃝βT_{\beta}\setminus\{z\}\vDash\alpha\ovee\beta. The remaining team is exactly TzβT^{\beta}_{z}, in which α1β1χk(α,β)\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta) fails (see Figure 2). The case zkTαkTβk\llbracket{}z\rrbracket{}_{k}\in\llbracket{}T_{\alpha}\rrbracket{}_{k}\setminus\llbracket{}T_{\beta}\rrbracket{}_{k} is detected analogously.

We proceed with a formal correctness proof by induction on kk. Let 𝒦=(W,R,V)\mathcal{K}=(W,R,V) as in the theorem. The base case k=0k=0 is straightforward, as no proposition pΦp\in\Phi occurs in α\alpha or β\beta. The induction step is split into two parts.

χkχk\chi_{k}\Rightarrow\chi^{*}_{k}: Let TT be a team and α,β\alpha,\beta disjoint scopes. Observe that χk\chi^{*}_{k} is always true if TαT_{\alpha} and TβT_{\beta} are both empty (then Tα=kTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}=\llbracket{}T_{\beta}\rrbracket{}_{k}), and that it is always false if exactly one of them is empty (then TαkTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}\neq\llbracket{}T_{\beta}\rrbracket{}_{k}). Therefore, let TαT_{\alpha}\neq\emptyset and TβT_{\beta}\neq\emptyset. Then χk(α,β)\chi^{*}_{k}(\alpha,\beta) boils down to ((α∨⃝β)𝖤α𝖤βα1β1χk(α,β)){\sim}((\alpha\ovee\beta)\lor\mathsf{E}\alpha\land\mathsf{E}\beta\land{\sim}\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta)), which we prove equivalent to Tα=kTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}=\llbracket{}T_{\beta}\rrbracket{}_{k}.

The first direction is proved by contradiction. Suppose Tα=kTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}=\llbracket{}T_{\beta}\rrbracket{}_{k} but T(α∨⃝β)𝖤α𝖤βα1β1χk(α,β)T\vDash(\alpha\ovee\beta)\lor\mathsf{E}\alpha\land\mathsf{E}\beta\land{\sim}\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta). The disjunction is witnessed by some division T=SUT=S\cup U, where w.l.o.g.\xspaceSTαS\subseteq T_{\alpha} satisfies α∨⃝β\alpha\ovee\beta, (if STβS\subseteq T_{\beta}, the proof is symmetric), and U𝖤α𝖤βα1β1χk(α,β)U\vDash\mathsf{E}\alpha\land\mathsf{E}\beta\land{\sim}\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta). Since TαTβ=T_{\alpha}\cap T_{\beta}=\emptyset, then TβUT_{\beta}\subseteq U, and clearly TβUβT_{\beta}\subseteq U_{\beta}. By the formula, some wUαw\in U_{\alpha} exists. By assumption that Tα=kTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}=\llbracket{}T_{\beta}\rrbracket{}_{k}, UβU_{\beta} must contain a world vv of type wk\llbracket{}w\rrbracket{}_{k} as well. But then Uw,vα,βχk(α,β)U^{\alpha,\beta}_{w,v}\vDash\chi_{k}(\alpha,\beta) by induction hypothesis, contradiction to Uα1β1χk(α,β)U\vDash{\sim}\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta).

For the other direction, suppose TαkTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}\neq\llbracket{}T_{\beta}\rrbracket{}_{k}. W.l.o.g.\xspacethere exists wTαw\in T_{\alpha} such that wkTβk\llbracket{}w\rrbracket{}_{k}\notin\llbracket{}T_{\beta}\rrbracket{}_{k}. (For wTβw\in T_{\beta}, the proof is again symmetric.) Consider S:=Tα{w}S\vcentcolon=T_{\alpha}\setminus\{w\} and U:=TwαU\vcentcolon=T^{\alpha}_{w} as a division of TT. Then Sα∨⃝βS\vDash\alpha\ovee\beta and U𝖤α𝖤βU\vDash\mathsf{E}\alpha\land\mathsf{E}\beta. It remains to show Uα1β1χk(α,β)U\vDash{\sim}\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta). However, this is easy to see: Uα1β1χk(α,β)U\vDash\exists^{1}_{\alpha}\exists^{1}_{\beta}\chi_{k}(\alpha,\beta) if and only if Uβ1χk(α,β)U\vDash\exists^{1}_{\beta}\chi_{k}(\alpha,\beta), but TβT_{\beta} and hence UβU_{\beta} contains no world of type wk\llbracket{}w\rrbracket{}_{k}, so by induction hypothesis UU cannot satisfy β1χk(α,β)\exists^{1}_{\beta}\chi_{k}(\alpha,\beta).

χkχk+1\chi^{*}_{k}\Rightarrow\chi_{k+1}: We follow Definition 3 and Proposition 3.

Tw,vα,βχk+1(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\chi_{k+1}(\alpha,\beta)
\displaystyle\Leftrightarrow\; Tw,vα,βχ0(α,β)χk(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\chi_{0}(\alpha,\beta)\land\Box\chi^{*}_{k}(\alpha,\beta) (Definition of χk+1\chi_{k+1})
\displaystyle\Leftrightarrow\; w0v and Tw,vα,βχk(α,β)\displaystyle w\rightleftharpoons_{0}v\text{ and }T^{\alpha,\beta}_{w,v}\vDash\Box\chi^{*}_{k}(\alpha,\beta) (Induction hypothesis)
\displaystyle\Leftrightarrow\; w0v and RTRw,Rvα,βχk(α,β)\displaystyle w\rightleftharpoons_{0}v\text{ and }RT^{\alpha,\beta}_{Rw,Rv}\vDash\chi^{*}_{k}(\alpha,\beta) (Proposition 14)
\displaystyle\Leftrightarrow\; w0v and RwkRv\displaystyle w\rightleftharpoons_{0}v\text{ and }Rw\rightleftharpoons_{k}Rv (Induction hypothesis)
\displaystyle\Leftrightarrow\; wk+1v.\displaystyle w\rightleftharpoons_{k+1}v\text{.} (Proposition 3)

It is routine to check that the formulas are constructible in logarithmic space from α\alpha, β\beta, Φ\Phi and kk, and that 𝗆𝖽(χk)=𝗆𝖽(χk)=k\mathsf{md}(\chi_{k})=\mathsf{md}(\chi^{*}_{k})=k.

Let us stress that χk\chi_{k} relies on disjoint scopes to be present in the structure, and it is open whether the property |T|k1{|\nobreak\llbracket{}T\rrbracket{}_{k}\nobreak|}\leq 1 is polynomially definable without these. Incidentally, the related property |Rw|k1{|\nobreak\llbracket{}Rw\rrbracket{}_{k}\nobreak|}\leq 1 of points ww was recently studied by Hella and Vilander [HV16], and was proven to be expressible in 𝖬𝖫\xspace\mathsf{ML}\xspace, but only by formulas of non-elementary size. However, they proved that it is definable in exponential size in 2-dimensional modal logic 𝖬𝖫\xspace2\mathsf{ML}\xspace^{2} (for an introduction to 𝖬𝖫\xspace2\mathsf{ML}\xspace^{2}, see Marx and Venema [MV97]). Roughly speaking, 𝖬𝖫\xspace2\mathsf{ML}\xspace^{2} is evaluated by traversing over pairs of points independently. The relationship between 𝖬𝖫\xspace2\mathsf{ML}\xspace^{2} and 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is unclear: Arguably, pairs of points are a special case of teams. But on the other hand, the modalities in 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace do not act on the points in a team independently, as in 𝖬𝖫\xspace2\mathsf{ML}\xspace^{2}, but instead always proceed to a successor team “synchronously”. As a consequence, it is also open whether 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace can define one of the above properties by a formula of elementary size.

6. Enforcing a canonical model

In this section, we approach the canonical models of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace from a lower bound perspective. Here, we devise an 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formula that is satisfiable but permits only kk-canonical models.

For k=0k=0, that is propositional team logic, Hannula et al. [HKVV15] defined the 𝖯𝖳𝖫\xspace\mathsf{PTL}\xspace-formula

𝗆𝖺𝗑(X):=xX=(x)\displaystyle\mathsf{max}(X)\vcentcolon={\sim}\bigvee_{\mathclap{x\in X}}{=\!\!(x)}

and proved that T𝗆𝖺𝗑(Φ)T\vDash\mathsf{max}(\Phi) if and only if TT is 0-canonical, i.e.\xspace, contains all Boolean assignment over Φ\Phi. We generalize this for all kk, i.e.\xspace, construct a satisfiable formula 𝖼𝖺𝗇𝗈𝗇k\mathsf{canon}_{k} that has only kk-canonical models.

6.1. Staircase models

Our approach is to express kk-canonicity by inductively enforcing ii-canonical sets of worlds for i=0,,ki=0,\ldots,k located in different “height” inside the model. For this purpose, we employ distinct scopes 𝔰0,,𝔰k\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} (stairs), and introduce a specific class of models:

{defi}

Let k,i0k,i\geq 0 and let (𝒦,T)(\mathcal{K},T) be a Kripke structure with team, 𝒦=(W,R,V)\mathcal{K}=(W,R,V). Then TT is kk-canonical with offset ii if for every τΔk\tau\in\Delta_{k} there exists wTw\in T with Riw=k{τ}\llbracket{}R^{i}w\rrbracket{}_{k}=\{\tau\}. (𝒦,T)(\mathcal{K},T) is called kk-staircase if for all i{0,,k}i\in\{0,\ldots,k\} we have that T𝔰iT_{\mathfrak{s}_{i}} is ii-canonical with offset kik-i.

𝔰0\mathfrak{s}_{0}𝔰1\mathfrak{s}_{1}𝔰2\mathfrak{s}_{2}𝔰3\mathfrak{s}_{3}, 2222|Φ|=16=|Δ3|2^{2^{2^{2^{|\nobreak\Phi\nobreak|}}}}=16={|\nobreak\Delta_{3}\nobreak|} elements \cdots\cdots 3-canonical2-canonical1-canonical0-c.OffsetScope: TT
Figure 3. Visualization of the 33-staircase for Φ=\Phi=\emptyset, where the subteam T𝔰iT_{\mathfrak{s}_{i}} is ii-canonical with offset 3i3-i.

As an example, a 33-staircase for Φ=\Phi=\emptyset is depicted in Figure 3. Observe that it is a directed forest, i.e.\xspace, it is acyclic and all worlds are either roots (i.e.\xspace, without predecessor) or have exactly one predecessor. Moreover, it has bounded height, where the height of a directed forest is the greatest number hh such that every path traverses at most hh edges.

Proposition 17.

For each k0k\geq 0, there is a finite kk-staircase (𝒦,T)(\mathcal{K},T) such that 𝔰0,,𝔰k\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} are disjoint scopes in 𝒦\mathcal{K}, and 𝒦\mathcal{K} is a directed forest with height at most kk and its set of roots being exactly TT.

Proof 6.1.

See Figure 3.

Observe that in such a model, T𝔰kT_{\mathfrak{s}_{k}} is kk-canonical with offset 0, which is simply kk-canonical:

Corollary 18 (Finite tree model property of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace).

Every satisfiable 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formula has a finite model (𝒦,T)(\mathcal{K},T) such that 𝒦\mathcal{K} is a directed forest with height at most kk and its set of roots being exactly TT.

6.2. Enforcing canonicity

In the rest of the section, we illustrate how a kk-staircase can be enforced in 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace inductively.

For Φ=\Phi=\emptyset, the inductive step—obtaining (k+1)(k+1)-canonicity from kk-canonicity—is done by the formula αβ1χk(α,β)\forall^{\subseteq}_{\alpha}\,\exists^{1}_{\beta}\,\Box\chi^{*}_{k}(\alpha,\beta). The idea is that this formula states that for every subteam TTαT^{\prime}\subseteq T_{\alpha} there exists a point wTβw\in T_{\beta} such that RT=kRwk\llbracket{}RT^{\prime}\rrbracket{}_{k}=\llbracket{}Rw\rrbracket{}_{k}. Intuitively, every possible set of types is captured as the image of some point in TβT_{\beta}. As a consequence, if TαT_{\alpha} is kk-canonical with offset 11, then TβT_{\beta} will be (k+1)(k+1)-canonical.

Note that the simpler formula k𝗆𝖺𝗑(Φ)\Box^{k}\mathsf{max}(\Phi) expresses 0-canonicity of RkTR^{k}T, but not 0-canonicity of TT with offset kk (consider, e.g.\xspace, a singleton TT). Instead, we use the formula

𝗆𝖺𝗑i:=\displaystyle\mathsf{max}_{i}\vcentcolon=\; (ipΦ(ip∨⃝i¬p)).\displaystyle\top\lor(\Diamond^{i}\top\land{\sim}\bigvee_{\mathclap{p\in\Phi}}(\Diamond^{i}p\ovee\Diamond^{i}\neg p))\text{.}

It states not only that RiTR^{i}T is 0-canonical, but also that RiwR^{i}w contains exactly one propositional assignment for each wTw\in T, which together yields 0-canonicity with offset ii.

Lemma 19.

T𝗆𝖺𝗑iT\vDash\mathsf{max}_{i} iff TT is 0-canonical with offset ii.

Proof 6.2.

By the distributive law φ(ψ1∨⃝ψ2)(φψ1)∨⃝(φψ2)\varphi\lor(\psi_{1}\ovee\psi_{2})\equiv(\varphi\lor\psi_{1})\ovee(\varphi\lor\psi_{2}), the duality (ψ1∨⃝ψ2)ψ1ψ2{\sim}(\psi_{1}\ovee\psi_{2})\equiv{\sim}\psi_{1}\land{\sim}\psi_{2}, and the definition 𝖤ψ=¬ψ\mathsf{E}\psi={\sim}\neg\psi,

pΦ(ip∨⃝i¬p)\scalerel∨⃝PΦ(pPippΦPi¬p)PΦ𝖤(pPi¬ppΦPip).{\sim}\bigvee_{\mathclap{p\in\Phi}}(\Diamond^{i}p\ovee\Diamond^{i}\neg p)\equiv{\sim}\operatorname*{\scalerel*{\ovee}{\sum}}_{P\subseteq\Phi}\Big{(}\bigvee_{p\in P}\Diamond^{i}p\lor\bigvee_{\mathclap{p\in\Phi\setminus P}}\Diamond^{i}\neg p\Big{)}\equiv\bigwedge_{P\subseteq\Phi}\!\!\mathsf{E}\Big{(}\bigwedge_{p\in P}\Box^{i}\neg p\land\bigwedge_{\mathclap{p\in\Phi\setminus P}}\Box^{i}p\Big{)}\text{.}

The rightmost formula now states that for all types τΔ0\tau\in\Delta_{0} (each represented by a subset of Φ\Phi, cf. Proposition 6), there exists a world wTw\in T such that Riw0Φ{τ}\llbracket{}R^{i}w\rrbracket{}^{\Phi}_{0}\subseteq\{\tau\}. Likewise, TiT\vDash\Diamond^{i}\top iff RiwR^{i}w\neq\emptyset for every wTw\in T.

Based on this, kk-canonicity with offset ii is now recursively defined as ρki\rho^{i}_{k}:

ρ0i(β)\displaystyle\rho^{i}_{0}(\beta)\, :=β𝗆𝖺𝗑i\displaystyle\vcentcolon=\;\beta\hookrightarrow\mathsf{max}_{i}
ρk+1i(α,β)\displaystyle\rho^{i}_{k+1}(\alpha,\beta) :=αβ(ρ0i(β)iβ1χk(α,β))\displaystyle\,\vcentcolon=\;\forall^{\subseteq}_{\alpha}\,\exists^{\subseteq}_{\beta}\left(\rho^{i}_{0}(\beta)\land\Box^{i}\forall^{1}_{\beta}\;\Box\chi^{*}_{k}(\alpha,\beta)\right)
𝖼𝖺𝗇𝗈𝗇k\displaystyle\mathsf{canon}_{k}\, :=ρ0k(𝔰0)m=1kρmkm(𝔰m1,𝔰m)\displaystyle\vcentcolon=\;\rho^{k}_{0}(\mathfrak{s}_{0})\land\bigwedge_{m=1}^{k}\rho^{k-m}_{m}(\mathfrak{s}_{m-1},\mathfrak{s}_{m})
Theorem 20.

Let k0k\geq 0 and 𝒦\mathcal{K} be a structure with disjoint scopes 𝔰0,,𝔰k\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k}. Then (𝒦,T)𝖼𝖺𝗇𝗈𝗇k(\mathcal{K},T)\vDash\mathsf{canon}_{k} if and only if (𝒦,T)(\mathcal{K},T) is a kk-staircase. Moreover, 𝖼𝖺𝗇𝗈𝗇k\mathsf{canon}_{k} is an 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formula constructible in space 𝒪(log(|Φ|+k)){\mathcal{O}(\log({|\nobreak\Phi\nobreak|}+k))}.

Proof 6.3.

Similar to Theorem 16, the construction of the above formula in logspace is straightforward. We proceed with the correctness of the formula. Suppose that 𝔰0,,𝔰k\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} are disjoint scopes in 𝒦\mathcal{K}. We show the following by induction on 0ik0\leq i\leq k: Assuming that TαT_{\alpha} is kk-canonical with offset i+1i+1, it holds that TβT_{\beta} is (k+1)(k+1)-canonical with offset ii if and only if Tρk+1i(α,β)T\vDash\rho^{i}_{k+1}(\alpha,\beta). With the induction basis done in Lemma 19, the inductive step is proved by the following equivalence:

Tβ is (k+1)-canonical with offset i\displaystyle T_{\beta}\text{ is }(k+1)\text{-canonical with offset }i
\displaystyle\Leftrightarrow\; τΔk+1:wTβ:Riw=k+1{τ}\displaystyle\forall\tau\in\Delta_{k+1}\colon\exists w\in T_{\beta}\colon\llbracket{}R^{i}w\rrbracket{}_{k+1}=\{\tau\}
Using the inverse of the bijection h:τ(τΦ,τ)h\colon\tau\mapsto(\tau\cap\Phi,\mathcal{R}\tau) from Proposition 6, we can equivalently quantify over 𝔓(Δk)\mathfrak{P}(\Delta_{k}) and 𝔓(Φ)\mathfrak{P}(\Phi):
\displaystyle\Leftrightarrow\; ΔΔk:ΦΦ:wTβ:Riw=k+1{h1(Φ,Δ)}\displaystyle\forall\Delta^{\prime}\subseteq\Delta_{k}\colon\forall\Phi^{\prime}\subseteq\Phi\colon\exists w\in T_{\beta}\colon\llbracket{}R^{i}w\rrbracket{}_{k+1}=\{h^{-1}(\Phi^{\prime},\Delta^{\prime})\}
\displaystyle\Leftrightarrow\; ΔΔk:ΦΦ:wTβ:Riw and vRiw:v=k+1h1(Φ,Δ)\displaystyle\forall\Delta^{\prime}\subseteq\Delta_{k}\colon\forall\Phi^{\prime}\subseteq\Phi\colon\exists w\in T_{\beta}\colon R^{i}w\neq\emptyset\text{ and }\forall v\in R^{i}w\colon\llbracket{}v\rrbracket{}_{k+1}=h^{-1}(\Phi^{\prime},\Delta^{\prime})
By Lemma 7, V1(v)=Φ and Rv=kΔV^{-1}(v)=\Phi^{\prime}\text{ and }\llbracket{}Rv\rrbracket{}_{k}=\Delta^{\prime} is equivalent to v=k+1h1(Φ,Δ)\llbracket{}v\rrbracket{}_{k+1}=h^{-1}(\Phi^{\prime},\Delta^{\prime}):
\displaystyle\Leftrightarrow\; ΔΔk:ΦΦ:wTβ:Riw\displaystyle\forall\Delta^{\prime}\subseteq\Delta_{k}\colon\forall\Phi^{\prime}\subseteq\Phi\colon\exists w\in T_{\beta}\colon R^{i}w\neq\emptyset
and vRiw:V1(v)=Φ and Rv=kΔ\displaystyle\qquad\text{and }\forall v\in R^{i}w\colon V^{-1}(v)=\Phi^{\prime}\text{ and }\llbracket{}Rv\rrbracket{}_{k}=\Delta^{\prime}
Again by Proposition 6, h:ττΦh\colon\tau\mapsto\tau\cap\Phi is a bijection from Δ0\Delta_{0} to 𝔓(Φ)\mathfrak{P}(\Phi):
\displaystyle\Leftrightarrow\; ΔΔk:τ0Δ0:wTβ:Riw\displaystyle\forall\Delta^{\prime}\subseteq\Delta_{k}\colon\forall\tau_{0}\in\Delta_{0}\colon\exists w\in T_{\beta}\colon R^{i}w\neq\emptyset
and vRiw:V1(v)=τ0Φ and Rv=kΔ\displaystyle\qquad\text{and }\forall v\in R^{i}w\colon V^{-1}(v)=\tau_{0}\cap\Phi\text{ and }\llbracket{}Rv\rrbracket{}_{k}=\Delta^{\prime}
Once more by Lemma 7:
\displaystyle\Leftrightarrow\; ΔΔk:τ0Δ0:wTβ:Riw\displaystyle\forall\Delta^{\prime}\subseteq\Delta_{k}\colon\forall\tau_{0}\in\Delta_{0}\colon\exists w\in T_{\beta}\colon R^{i}w\neq\emptyset
and vRiw:v=0τ0 and Rv=kΔ\displaystyle\qquad\text{and }\forall v\in R^{i}w\colon\llbracket{}v\rrbracket{}_{0}=\tau_{0}\text{ and }\llbracket{}Rv\rrbracket{}_{k}=\Delta^{\prime}
\displaystyle\Leftrightarrow\; ΔΔk:τ0Δ0:wTβ:Riw=0{τ0} and vRiw:Rv=kΔ\displaystyle\forall\Delta^{\prime}\subseteq\Delta_{k}\colon\forall\tau_{0}\in\Delta_{0}\colon\exists w\in T_{\beta}\colon\llbracket{}R^{i}w\rrbracket{}_{0}=\{\tau_{0}\}\text{ and }\forall v\in R^{i}w\colon\llbracket{}Rv\rrbracket{}_{k}=\Delta^{\prime}
Since TαT_{\alpha} is assumed kk-canonical with offset i+1i+1, for every τΔk\tau^{\prime}\in\Delta_{k} there exists uTαu\in T_{\alpha} such that Ri+1u=k{τ}\llbracket{}R^{i+1}u\rrbracket{}_{k}=\{\tau^{\prime}\}. Accordingly, for every set ΔΔk\Delta^{\prime}\subseteq\Delta_{k} there exists STαS\subseteq T_{\alpha} such that Ri+1S=kΔ\llbracket{}R^{i+1}S\rrbracket{}_{k}=\Delta^{\prime}:
\displaystyle\Leftrightarrow\; STα:τ0Δ0:wTβ:Riw=0{τ0} and vRiw:Rv=kRi+1Sk\displaystyle\forall S\subseteq T_{\alpha}\colon\forall\tau_{0}\in\Delta_{0}\colon\exists w\in T_{\beta}\colon\llbracket{}R^{i}w\rrbracket{}_{0}=\{\tau_{0}\}\text{ and }\forall v\in R^{i}w\colon\llbracket{}Rv\rrbracket{}_{k}=\llbracket{}R^{i+1}S\rrbracket{}_{k}
For each SS, gather the respective ww in a team UTβU\subseteq T_{\beta}:
\displaystyle\Leftrightarrow\; STα:UTβ:(τ0Δ0:wU:Riw=0{τ0})\displaystyle\forall S\subseteq T_{\alpha}\colon\exists U\subseteq T_{\beta}\colon\left(\forall\tau_{0}\in\Delta_{0}\colon\exists w\in U\colon\llbracket{}R^{i}w\rrbracket{}_{0}=\{\tau_{0}\}\right)
 and vRiU:Rv=kRi+1Sk\displaystyle\qquad\qquad\text{ and }\forall v\in R^{i}U\colon\llbracket{}Rv\rrbracket{}_{k}=\llbracket{}R^{i+1}S\rrbracket{}_{k}
\displaystyle\Leftrightarrow\; STα:UTβ:U is 0-canonical with offset i\displaystyle\forall S\subseteq T_{\alpha}:\exists U\subseteq T_{\beta}\colon U\text{ is $0$-canonical with offset }i
 and vRiU:Rv=kRi+1Sk\displaystyle\qquad\qquad\text{ and }\forall v\in R^{i}U\colon\llbracket{}Rv\rrbracket{}_{k}=\llbracket{}R^{i+1}S\rrbracket{}_{k}
By the base case k=0k=0, and since U=(TS,Uα,β)βU={(T^{\alpha,\beta}_{S,U})}_{\beta}:
\displaystyle\Leftrightarrow\; STα:UTβ:TS,Uα,βρ0i(β) and vRiU:Rv=kRi+1Sk\displaystyle\forall S\subseteq T_{\alpha}\colon\exists U\subseteq T_{\beta}\colon T^{\alpha,\beta}_{S,U}\vDash\rho^{i}_{0}(\beta)\text{ and }\forall v\in R^{i}U\colon\llbracket{}Rv\rrbracket{}_{k}=\llbracket{}R^{i+1}S\rrbracket{}_{k}
By Theorem 16:
\displaystyle\Leftrightarrow\; STα:UTβ:TS,Uα,βρ0i(β) and vRiU:(Ri+1T)Ri+1S,Rvα,βχk(α,β)\displaystyle\forall S\subseteq T_{\alpha}\colon\exists U\subseteq T_{\beta}\colon T^{\alpha,\beta}_{S,U}\vDash\rho^{i}_{0}(\beta)\text{ and }\forall v\in R^{i}U\colon{(R^{i+1}T)}^{\alpha,\beta}_{R^{i+1}S,Rv}\vDash\chi^{*}_{k}(\alpha,\beta)
By Proposition 14 (5.):
\displaystyle\Leftrightarrow\; STα:UTβ:TS,Uα,βρ0i(β) and vRiU:(RiT)RiS,vα,βχk(α,β)\displaystyle\forall S\subseteq T_{\alpha}\colon\exists U\subseteq T_{\beta}\colon T^{\alpha,\beta}_{S,U}\vDash\rho^{i}_{0}(\beta)\text{ and }\forall v\in R^{i}U\colon{(R^{i}T)}^{\alpha,\beta}_{R^{i}S,v}\vDash\Box\chi^{*}_{k}(\alpha,\beta)
By Proposition 15 applied to (RiT)RiS,RiUα,β{(R^{i}T)}^{\alpha,\beta}_{R^{i}S,R^{i}U}:
\displaystyle\Leftrightarrow\; STα:UTβ:TS,Uα,βρ0i(β) and (RiT)RiS,RiUα,ββ1χk(α,β)\displaystyle\forall S\subseteq T_{\alpha}\colon\exists U\subseteq T_{\beta}\colon T^{\alpha,\beta}_{S,U}\vDash\rho^{i}_{0}(\beta)\text{ and }{(R^{i}T)}^{\alpha,\beta}_{R^{i}S,R^{i}U}\vDash\forall^{1}_{\beta}\Box\chi^{*}_{k}(\alpha,\beta)
Again by Proposition 14 (5.) and Proposition 15:
\displaystyle\Leftrightarrow\; STα:UTβ:TS,Uα,βρ0i(β) and Ri(TS,Uα,β)β1χk(α,β)\displaystyle\forall S\subseteq T_{\alpha}\colon\exists U\subseteq T_{\beta}\colon T^{\alpha,\beta}_{S,U}\vDash\rho^{i}_{0}(\beta)\text{ and }R^{i}\left(T^{\alpha,\beta}_{S,U}\right)\vDash\forall^{1}_{\beta}\Box\chi^{*}_{k}(\alpha,\beta)
\displaystyle\Leftrightarrow\; STα:UTβ:TS,Uα,βρ0i(β)iβ1χk(α,β)\displaystyle\forall S\subseteq T_{\alpha}\colon\exists U\subseteq T_{\beta}\colon T^{\alpha,\beta}_{S,U}\vDash\rho^{i}_{0}(\beta)\land\Box^{i}\forall^{1}_{\beta}\Box\chi^{*}_{k}(\alpha,\beta)
\displaystyle\Leftrightarrow\; Tαβ(ρ0i(β)iβ1χk(α,β))\displaystyle T\vDash\forall^{\subseteq}_{\alpha}\,\exists^{\subseteq}_{\beta}\,(\rho^{i}_{0}(\beta)\land\Box^{i}\forall^{1}_{\beta}\Box\chi^{*}_{k}(\alpha,\beta))
\displaystyle\Leftrightarrow\; Tρk+1i(α,β).\displaystyle T\vDash\rho^{i}_{k+1}(\alpha,\beta)\text{.}

6.3. Enforcing scopes

As the next step, we lift the restriction of the 𝔰i\mathfrak{s}_{i} being scopes a priori. In a sense, this condition is definable in 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace as well. For this, let Ψ𝒫𝒮\Psi\subseteq\mathcal{PS} be disjoint from Φ\Phi. Then the formula below ensures that Ψ\Psi is a set of disjoint scopes “up to height kk.

𝗌𝖼𝗈𝗉𝖾𝗌k(Ψ)\displaystyle\mathsf{scopes}_{k}(\Psi) :=x,yΨxy¬(xy)i=1k((xix)(¬xi¬x)).\displaystyle\vcentcolon=\bigwedge_{\begin{subarray}{c}x,y\in\Psi\\ x\neq y\end{subarray}}\neg(x\land y)\land\bigwedge_{i=1}^{k}\Big{(}(x\land\Box^{i}x)\lor(\neg x\land\Box^{i}\neg x)\Big{)}\text{.}

The definition up to height kk is sufficient for our purposes, which follows from the next lemma.

Lemma 21.

If φ𝖬𝖳𝖫\xspacek\varphi\in\mathsf{MTL}\xspace_{k}, then φ\varphi is satisfiable if and only if φk+1\varphi\land\Box^{k+1}\bot is satisfiable.

Proof 6.4.

As the direction from right to left is trivial, suppose φ\varphi is satisfiable. By Corollary 18, it then has a model (𝒦,T)(\mathcal{K},T) that is a directed forest of height at most kk. But then (𝒦,T)k+1(\mathcal{K},T)\vDash\Box^{k+1}\bot, since Rk+1T=R^{k+1}T=\emptyset and (𝒦,)(\mathcal{K},\emptyset) satisfies all 𝖬𝖫\xspace\mathsf{ML}\xspace-formulas, including \bot.

Theorem 22.

𝖼𝖺𝗇𝗈𝗇k𝗌𝖼𝗈𝗉𝖾𝗌k({𝔰0,,𝔰k})k+1\mathsf{canon}_{k}\land\mathsf{scopes}_{k}(\{\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k}\})\land\Box^{k+1}\bot is satisfiable, but has only kk-staircases as models.

Proof 6.5.

By combining Proposition 17, Theorem 20 and Lemma 21, the formula is satisfiable. Since in every model (𝒦,T)(\mathcal{K},T) the propositions 𝔰0,,𝔰k\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} must be disjoint scopes due to k+1\Box^{k+1}\bot and 𝗌𝖼𝗈𝗉𝖾𝗌k\mathsf{scopes}_{k}, we can apply Theorem 20.

As for bisimilarity, it is open whether (Φ,k)(\Phi,k)-canonicity can be defined in 𝖬𝖳𝖫\xspacekΦ\mathsf{MTL}\xspace^{\Phi}_{k} efficiently without restricting the models to those with scopes. Note that the results of this section alone do not imply that the brute force algorithm given in Theorem 12 is optimal, as there could possibly be a satisfiability algorithm that does not need to construct a model. To show proper complexity theoretic hardness, we need to encode non-elementary computations in such models, to which we will proceed in the next sections.

7. Defining an order on types

In the previous section, we enforced kk-canonicity with a formula, i.e.\xspace, such that |Δk|{|\nobreak\Delta_{k}\nobreak|} different types are contained in the team. In order to encode computations of length |Δk|{|\nobreak\Delta_{k}\nobreak|}, we additionally need to be able to talk about an ordering of Δk\Delta_{k}.

Let us call any finite strict linear ordering simply an order. We specify an order k\prec_{k} on Δk\Delta_{k}, and analogously to team bisimilarity, an order k\prec^{*}_{k} on 𝔓(Δk)\mathfrak{P}(\Delta_{k}). To begin with, let us first agree on some arbitrary order << on Φ\Phi, say, p1<p2<<p|Φ|p_{1}<p_{2}<\cdots<p_{{|\nobreak\Phi\nobreak|}}. Furthermore, if \sqsubset is some order on XX, then the lexicographic order \sqsubset^{*} on 𝔓(X)\mathfrak{P}(X) is defined by

X1X2 iff xX2X1 such that xX:(xx)(xX1xX2).\displaystyle X_{1}\sqsubset^{*}X_{2}\text{ iff }\exists x\in X_{2}\setminus X_{1}\text{ such that }\forall x^{\prime}\in X\colon(x\sqsubset x^{\prime})\Rightarrow(x^{\prime}\in X_{1}\Leftrightarrow x^{\prime}\in X_{2})\text{.}

For example, let X={0,1}X=\{0,1\} and 010\sqsubset 1. Then {0}{1}{0,1}\emptyset\sqsubset^{*}\{0\}\sqsubset^{*}\{1\}\sqsubset^{*}\{0,1\}. The order k\prec_{k} depends on the propositions true in a world, and otherwise recursively on the lexicographic order of the image team:

τ0τ\displaystyle\tau\prec_{0}\tau^{\prime}\; τΦ<τΦ,\displaystyle\Leftrightarrow\;\tau\cap\Phi<^{*}\tau^{\prime}\cap\Phi\text{,}
τk+1τ\displaystyle\tau\prec_{k+1}\tau^{\prime}\; τΦ<τΦ or (τΦ=τΦ and τkτ).\displaystyle\Leftrightarrow\;\tau\cap\Phi<^{*}\tau^{\prime}\cap\Phi\text{ or }(\tau\cap\Phi=\tau^{\prime}\cap\Phi\text{ and }\mathcal{R}\tau\;\prec_{k}^{*}\;\mathcal{R}\tau^{\prime})\text{.}

It is easy to verify by induction that k\prec_{k} and k\prec^{*}_{k} are orders on Δk\Delta_{k} and 𝔓(Δk)\mathfrak{P}(\Delta_{k}), respectively.

The next step is to prove that k\prec_{k} and k\prec^{*}_{k} are (efficiently) definable in 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}. For this, we pursue the same approach as for χk\chi_{k} and χk\chi^{*}_{k} in Section 5, and show that k\prec_{k} and k\prec^{*}_{k} are definable in formulas ζk\zeta_{k} and ζk\zeta^{*}_{k} in a mutually recursive fashion. Since order is a binary relation, the formulas below are once more parameterized by two scopes.

ζ0(α,β)\displaystyle\zeta_{0}(\alpha,\beta) :=pΦ[(α¬p)(βp)qΦq<p(αβ)=(q)]\displaystyle\vcentcolon=\bigvee_{p\in\Phi}\Big{[}(\alpha\hookrightarrow\neg p)\land(\beta\hookrightarrow p)\land\bigwedge_{\begin{subarray}{c}q\in\Phi\\ q<p\end{subarray}}(\alpha\lor\beta)\hookrightarrow{=\!\!(q)}\Big{]}
ζk+1(α,β)\displaystyle\zeta_{k+1}(\alpha,\beta) :=ζ0(α,β)∨⃝χ0(α,β)ζk(α,β)\displaystyle\vcentcolon=\zeta_{0}(\alpha,\beta)\,\ovee\,\chi_{0}(\alpha,\beta)\land\,\Box\zeta^{*}_{k}(\alpha,\beta)
ζk(α,β)\displaystyle\zeta^{*}_{k}(\alpha,\beta) :=𝔰k1(β1χk(𝔰k,β))(α1χk(𝔰k,α))\displaystyle\vcentcolon=\exists^{1}_{\mathfrak{s}_{k}}\left(\exists^{1}_{\beta}\chi_{k}(\mathfrak{s}_{k},\beta)\right)\land\left({\sim}\exists^{1}_{\alpha}\chi_{k}(\mathfrak{s}_{k},\alpha)\right)
((χk(α,β)(αβ))(αβ1ζk(𝔰k,αβ)))\displaystyle\qquad\land\Big{(}\big{(}\chi^{*}_{k}(\alpha,\beta)\land(\alpha\lor\beta)\big{)}\,\lor\,\big{(}\forall^{1}_{\alpha\lor\beta}{\sim}\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta)\big{)}\Big{)}

Note that we make use of the scopes 𝔰0,,𝔰k\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} in the formula, and in the following we restrict ourselves to kk-staircase models. Moreover, in the subformula ζk(𝔰k,αβ)\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta), we use the fact that αβ\alpha\lor\beta is a scope whenever α,β\alpha,\beta are scopes.

We require the next lemma for the correctness of ζk\zeta_{k} and ζk\zeta^{*}_{k}. Intuitively, it states that 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} is invariant under substitution of “locally equivalent” 𝖬𝖫\xspace\mathsf{ML}\xspace-formulas.

Lemma 23.

Let α,β𝖬𝖫\xspace\alpha,\beta\in\mathsf{ML}\xspace and φ𝖬𝖳𝖫\xspacek\varphi\in\mathsf{MTL}\xspace_{k}. Let TT be a team such that RiTαβR^{i}T\vDash\alpha\leftrightarrow\beta for all i{0,,k}i\in\{0,\ldots,k\}. Then TφT\vDash\varphi if and only if TSub(φ,α,β)T\vDash\mathrm{Sub}(\varphi,\alpha,\beta), where Sub(φ,α,β)\mathrm{Sub}(\varphi,\alpha,\beta) is the formula obtained from φ\varphi by substituting every occurrence of α\alpha with β\beta.

Proof 7.1.

By straightforward induction; see the appendix.

The following theorem states that in the class of kk-staircase models (see the previous section) ζk\zeta_{k} and ζk\zeta^{*}_{k} define the required orders.

Theorem 24.

Let k0k\geq 0, and let (𝒦,T)(\mathcal{K},T) be a kk-staircase with disjoint scopes α,β,𝔰0,,𝔰k\alpha,\beta,\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k}. If wTαw\in T_{\alpha} and vTβv\in T_{\beta}, then

Tw,vα,β\displaystyle T^{\alpha,\beta}_{w,v} ζk(α,β)\displaystyle\;\vDash\;\zeta_{k}(\alpha,\beta)\; if and only if wk\displaystyle\llbracket{}w\rrbracket{}_{k} kv,k\displaystyle\prec_{k}\llbracket{}v\rrbracket{}_{k}\text{,}
T\displaystyle T ζk(α,β)\displaystyle\;\vDash\;\zeta^{*}_{k}(\alpha,\beta)\; if and only if Tαk\displaystyle\llbracket{}T_{\alpha}\rrbracket{}_{k} kTβ.k\displaystyle\prec^{*}_{k}\llbracket{}T_{\beta}\rrbracket{}_{k}\text{.}

Furthermore, both ζk(α,β)\zeta_{k}(\alpha,\beta) and ζk(α,β)\zeta^{*}_{k}(\alpha,\beta) are 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formulas that are constructible in space 𝒪(log(k+|Φ|+|α|+|β|)){\mathcal{O}(\log(k+{|\nobreak\Phi\nobreak|}+{|\nobreak\alpha\nobreak|}+{|\nobreak\beta\nobreak|}))}.

We first give a rough idea of the proof, and after a series of required lemmas fully prove the theorem. The definition of ζk+1\zeta_{k+1} simply follows the definition of k+1\prec_{k+1}. Furthermore, the formula ζk\zeta^{*}_{k} implements the lexicographic order k\prec^{*}_{k} as follows. As shown in Figure 4, we first choose some zT𝔰kz\in T_{\mathfrak{s}_{k}} that acts as an pivot to determine if TαkkTβ\llbracket{}T_{\alpha}\rrbracket{}_{k}\prec^{*}_{k}\llbracket{}T_{\beta}\rrbracket{}, in the sense that it is the k\prec_{k}-maximal type in which TαT_{\alpha} and TβT_{\beta} differ.333Since the pivot is selected from T𝔰kT_{\mathfrak{s}_{k}}, at this point it is crucial that the underlying structure is a kk-staircase. The first line of ζk\zeta^{*}_{k} indeed expresses that zkTβkTαk\llbracket{}z\rrbracket{}_{k}\in\llbracket{}T_{\beta}\rrbracket{}_{k}\setminus\llbracket{}T_{\alpha}\rrbracket{}_{k}.

The disjunction in the second line intuitively states that we then can “split off” the subteam of TαTβT_{\alpha}\cup T_{\beta} consisting of the elements k\prec_{k}-greater than zz (the solid green area in Figure 4), while χk\chi^{*}_{k} ensures that they agree on the contained types (this reflects the part after the quantifier in the definition of \sqsubset^{*}). To achieve this, the subformula αβ1ζk(𝔰k,αβ)\forall^{1}_{\alpha\lor\beta}{\sim}\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta) stipulates that any “remaining” elements from TαTβT_{\alpha}\cup T_{\beta} possess only types not k\prec_{k}-greater than zk\llbracket{}z\rrbracket{}_{k} (the dashed green area in the figure).

Here, Lemma 23 is applied, as it ensures that after processing αβ1\forall^{1}_{\alpha\lor\beta} the formula ζk(𝔰k,αβ)\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta) in fact behaves as either ζk(𝔰k,α)\zeta_{k}(\mathfrak{s}_{k},\alpha) or ζk(𝔰k,β)\zeta_{k}(\mathfrak{s}_{k},\beta); and hence behaves correctly by induction hypothesis.

k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}Tβ=^ 10𝟏¯000\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}T_{\beta}\;\hat{=}\;\mathbf{10\underline{1}}000k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}k\scriptstyle\succ_{k}Tα=^ 10𝟎¯110\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}T_{\alpha}\;\hat{=}\;\mathbf{10\underline{0}}110 T𝔰kT_{\mathfrak{s}_{k}}zzk\prec^{*}_{k} ? k\succ_{k}k\rightleftharpoons_{k}\;k\quad\preceq_{k}
Figure 4. The pivot zT𝔰kz\in T_{\mathfrak{s}_{k}} determines that TαkkTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}\prec^{*}_{k}\llbracket{}T_{\beta}\rrbracket{}_{k}. The subteam of TαβT_{\alpha\lor\beta} of worlds k\prec_{k}-greater than zz must satisfy χk(α,β)\chi^{*}_{k}(\alpha,\beta).
{defi}

Let k0k\geq 0. Let α,β\alpha,\beta be disjoint scopes and TT a team in a Kripke structure. Then α\alpha and β\beta are called k\prec_{k}-comparable in TT if for all wTα,vTβw\in T_{\alpha},v\in T_{\beta}

Tw,vα,β\displaystyle T^{\alpha,\beta}_{w,v} ζk(α,β) iff wkkv andk\displaystyle\vDash\zeta_{k}(\alpha,\beta)\text{ iff }\llbracket{}w\rrbracket{}_{k}\prec_{k}\llbracket{}v\rrbracket{}_{k}\text{ and}
Tw,vα,β\displaystyle T^{\alpha,\beta}_{w,v} ζk(β,α) iff vkkw.k\displaystyle\vDash\zeta_{k}(\beta,\alpha)\text{ iff }\llbracket{}v\rrbracket{}_{k}\prec_{k}\llbracket{}w\rrbracket{}_{k}\text{.}

Likewise, α\alpha and β\beta are k\prec^{*}_{k}-comparable in TT if

T\displaystyle T ζk(α,β) iff TαkkTβ andk\displaystyle\vDash\zeta^{*}_{k}(\alpha,\beta)\text{ iff }\llbracket{}T_{\alpha}\rrbracket{}_{k}\prec^{*}_{k}\llbracket{}T_{\beta}\rrbracket{}_{k}\text{ and}
T\displaystyle T ζk(β,α) iff TβkkTα.k\displaystyle\vDash\zeta^{*}_{k}(\beta,\alpha)\text{ iff }\llbracket{}T_{\beta}\rrbracket{}_{k}\prec^{*}_{k}\llbracket{}T_{\alpha}\rrbracket{}_{k}\text{.}

The next lemma shows that the correctness of k\prec^{*}_{k} follows from that of k\prec_{k}.

Lemma 25.

Suppose that (𝒦,T)(\mathcal{K},T) is a kk-staircase with disjoint scopes α,β,𝔰0,,𝔰k\alpha,\beta,\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k}. If both α\alpha and β\beta are k\prec_{k}-comparable to 𝔰k\mathfrak{s}_{k} in all subteams SS of the form T𝔰0T𝔰k1STT_{\mathfrak{s}_{0}}\cup\cdots\cup T_{\mathfrak{s}_{k-1}}\subseteq S\subseteq T, then α\alpha and β\beta are k\prec_{k}^{*}-comparable in TT.

Proof 7.2.

Assuming 𝒦,T,α,β,𝔰0,,𝔰k\mathcal{K},T,\alpha,\beta,\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} as above, the proof is split into the following claims.

Claim 26 (a).

The disjoint scopes αβ\alpha\lor\beta and 𝔰k\mathfrak{s}_{k} are k\prec_{k}-comparable in any team SS that satisfies T𝔰0T𝔰k1STT_{\mathfrak{s}_{0}}\cup\cdots\cup T_{\mathfrak{s}_{k-1}}\subseteq S\subseteq T.

Proof 7.3 (Proof of claim.).

Let wSαβw\in S_{\alpha\lor\beta} and vS𝔰kv\in S_{\mathfrak{s}_{k}}. W.l.o.g.\xspacewSαw\in S_{\alpha} (the case wSβw\in S_{\beta} works analogously). Then

Sw,vαβ,𝔰kζk(αβ,𝔰k)\displaystyle S^{\alpha\lor\beta,\mathfrak{s}_{k}}_{w,v}\vDash\zeta_{k}(\alpha\lor\beta,\mathfrak{s}_{k})
\displaystyle\Leftrightarrow\; Sw,,vα,β,𝔰kζk(αβ,𝔰k)\displaystyle S^{\alpha,\beta,\mathfrak{s}_{k}}_{w,\emptyset,v}\vDash\zeta_{k}(\alpha\lor\beta,\mathfrak{s}_{k}) (Since Sw,vαβ,𝔰k=Sw,,vα,β,𝔰kS^{\alpha\lor\beta,\mathfrak{s}_{k}}_{w,v}=S^{\alpha,\beta,\mathfrak{s}_{k}}_{w,\emptyset,v})
\displaystyle\Leftrightarrow\; Sw,,vα,β,𝔰kζk(α,𝔰k)\displaystyle S^{\alpha,\beta,\mathfrak{s}_{k}}_{w,\emptyset,v}\vDash\zeta_{k}(\alpha,\mathfrak{s}_{k}) (By Lemma 23, as i=0kRiSw,,vα,β,𝔰kα(αβ)\bigcup_{i=0}^{k}R^{i}S^{\alpha,\beta,\mathfrak{s}_{k}}_{w,\emptyset,v}\vDash\alpha\leftrightarrow(\alpha\lor\beta))
\displaystyle\Leftrightarrow\; wkkv.k\displaystyle\llbracket{}w\rrbracket{}_{k}\prec_{k}\llbracket{}v\rrbracket{}_{k}\text{.} (By assumption of the lemma)

The case ζk(𝔰k,αβ)\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta) is symmetric. \triangleleft

For the remaining proof, we omit the subscript kk when referring to types and \prec. Furthermore, for all τΔk\tau\in\Delta_{k}, let Tτ\llbracket{}T\rrbracket{}^{\tau} denote the restriction of T\llbracket{}T\rrbracket{} to types τ\tau^{\prime} such that ττ\tau^{\prime}\succ\tau. Intuitively, these types are the “more significant positions” for the lexicographic ordering. In the next claim, we essentially show that the second line in the definition of ζk(α,β)\zeta^{*}_{k}(\alpha,\beta) can be expressed as a statement of the form Tα=τTβτ\llbracket{}T_{\alpha}\rrbracket{}^{\tau}=\llbracket{}T_{\beta}\rrbracket{}^{\tau}.

Claim 27 (b).

Let TT be a team and τΔk\tau\in\Delta_{k}. Then Tα=τTβτ\llbracket{}T_{\alpha}\rrbracket{}^{\tau}=\llbracket{}T_{\beta}\rrbracket{}^{\tau} if and only if there exists STαβS\subseteq T_{\alpha\lor\beta} such that Sα=Sβ\llbracket{}S_{\alpha}\rrbracket{}=\llbracket{}S_{\beta}\rrbracket{} and wτ\llbracket{}w\rrbracket{}\nsucc\tau for all wTαβSw\in T_{\alpha\lor\beta}\setminus S.

Proof 7.4 (Proof of claim.).

\Rightarrow: Let S:={vTαβvτ}S\vcentcolon=\{v\in T_{\alpha\lor\beta}\mid\llbracket{}v\rrbracket{}\succ\tau\}. Then Sα=Tα=τTβ=τSβ\llbracket{}S_{\alpha}\rrbracket{}=\llbracket{}T_{\alpha}\rrbracket{}^{\tau}=\llbracket{}T_{\beta}\rrbracket{}^{\tau}=\llbracket{}S_{\beta}\rrbracket{}. Moreover, for every wTαβSw\in T_{\alpha\lor\beta}\setminus S clearly wτ\llbracket{}w\rrbracket{}\nsucc\tau holds.

\Leftarrow: Assume that SS exists as stated in the claim. By symmetry, we only prove TατTβτ\llbracket{}T_{\alpha}\rrbracket{}^{\tau}\subseteq\llbracket{}T_{\beta}\rrbracket{}^{\tau}. Consequently, let wTαw\in T_{\alpha} such that wTατ\llbracket{}w\rrbracket{}\in\llbracket{}T_{\alpha}\rrbracket{}^{\tau}. Then wτ\llbracket{}w\rrbracket{}\succ\tau by definition. But then wTαβSw\notin T_{\alpha\lor\beta}\setminus S. However, we have wTαw\in T_{\alpha}, hence wTαβw\in T_{\alpha\lor\beta}, which only leaves the possibility wSw\in S. Combining wSw\in S and wTαw\in T_{\alpha} yields wSαw\in S_{\alpha}, which by assumption also implies wSβ\llbracket{}w\rrbracket{}\in\llbracket{}S_{\beta}\rrbracket{}. As SβTβ\llbracket{}S_{\beta}\rrbracket{}\subseteq\llbracket{}T_{\beta}\rrbracket{} and wτ\llbracket{}w\rrbracket{}\succ\tau, the membership wTβτ\llbracket{}w\rrbracket{}\in\llbracket{}T_{\beta}\rrbracket{}^{\tau} follows. \triangleleft

Claim 28 (c).

α\alpha and β\beta are k\prec^{*}_{k}-comparable in TT.

Proof 7.5 (Proof of claim.).

Due to symmetry, we prove only that Tζk(α,β)T\vDash\zeta^{*}_{k}(\alpha,\beta) iff TαkkTβk\llbracket{}T_{\alpha}\rrbracket{}_{k}\prec^{*}_{k}\llbracket{}T_{\beta}\rrbracket{}_{k}.

TαTβ\displaystyle\llbracket{}T_{\alpha}\rrbracket{}\prec^{*}\llbracket{}T_{\beta}\rrbracket{}
\displaystyle\Leftrightarrow\; τTβTα:τΔ,ττ:τTατTβ\displaystyle\exists\tau\in\llbracket{}T_{\beta}\rrbracket{}\setminus\llbracket{}T_{\alpha}\rrbracket{}\colon\forall\tau^{\prime}\in\Delta,\tau\prec\tau^{\prime}\colon\tau^{\prime}\in\llbracket{}T_{\alpha}\rrbracket{}\Leftrightarrow\tau^{\prime}\in\llbracket{}T_{\beta}\rrbracket{} (Definition of k\prec^{*}_{k})
\displaystyle\Leftrightarrow\; τTβTα:Tα=τTβτ\displaystyle\exists\tau\in\llbracket{}T_{\beta}\rrbracket{}\setminus\llbracket{}T_{\alpha}\rrbracket{}\colon\llbracket{}T_{\alpha}\rrbracket{}^{\tau}=\llbracket{}T_{\beta}\rrbracket{}^{\tau} (Definition of τ\llbracket{}\cdot\rrbracket{}^{\tau})
Since T𝔰kT_{\mathfrak{s}_{k}} is kk-canonical, for every τΔ\tau\in\Delta there exists zT𝔰kz\in T_{\mathfrak{s}_{k}} of type τ\tau:
\displaystyle\Leftrightarrow\; zT𝔰k:Tα=zTβ and zzTβTα\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon\llbracket{}T_{\alpha}\rrbracket{}^{\llbracket{}z\rrbracket{}}=\llbracket{}T_{\beta}\rrbracket{}^{\llbracket{}z\rrbracket{}}\text{ and }\llbracket{}z\rrbracket{}\in\llbracket{}T_{\beta}\rrbracket{}\setminus\llbracket{}T_{\alpha}\rrbracket{}
\displaystyle\Leftrightarrow\; zT𝔰k:Tα=zTβ and zxTβ:z=x and yTα:z=y\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon\llbracket{}T_{\alpha}\rrbracket{}^{\llbracket{}z\rrbracket{}}=\llbracket{}T_{\beta}\rrbracket{}^{\llbracket{}z\rrbracket{}}\text{ and }\exists x\in T_{\beta}\colon\llbracket{}z\rrbracket{}=\llbracket{}x\rrbracket{}\text{ and }\nexists y\in T_{\alpha}\colon\llbracket{}z\rrbracket{}=\llbracket{}y\rrbracket{}
As α,β\alpha,\beta and 𝔰k\mathfrak{s}_{k} are disjoint, we have Tα=OαT_{\alpha}=O_{\alpha}, where O:=Tz𝔰kO\vcentcolon=T^{\mathfrak{s}_{k}}_{z}, and likewise Tβ=OβT_{\beta}=O_{\beta}:
\displaystyle\Leftrightarrow\; zT𝔰k:Oα=zOβ and zxOβ:z=x and yOα:z=y\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon\llbracket{}O_{\alpha}\rrbracket{}^{\llbracket{}z\rrbracket{}}=\llbracket{}O_{\beta}\rrbracket{}^{\llbracket{}z\rrbracket{}}\text{ and }\exists x\in O_{\beta}\colon\llbracket{}z\rrbracket{}=\llbracket{}x\rrbracket{}\text{ and }\nexists y\in O_{\alpha}\colon\llbracket{}z\rrbracket{}=\llbracket{}y\rrbracket{}
\displaystyle\Leftrightarrow\; zT𝔰k:xOβ:z=x and yOα:z=y\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon\exists x\in O_{\beta}\colon\llbracket{}z\rrbracket{}=\llbracket{}x\rrbracket{}\text{ and }\nexists y\in O_{\alpha}\colon\llbracket{}z\rrbracket{}=\llbracket{}y\rrbracket{}
and SOαβ:Sα=Sβ and wOαβS:zw\displaystyle\text{ and }\exists S\subseteq O_{\alpha\lor\beta}\colon\llbracket{}S_{\alpha}\rrbracket{}=\llbracket{}S_{\beta}\rrbracket{}\text{ and }\forall w\in O_{\alpha\lor\beta}\setminus S\colon\llbracket{}z\rrbracket{}\nprec\llbracket{}w\rrbracket{} (by Claim (b))
Clearly SS is a subteam of OαβO_{\alpha\lor\beta} if and only if it is a subteam of OO and satisfies αβ\alpha\lor\beta:
\displaystyle\Leftrightarrow\; zT𝔰k:xOβ:z=x and yOα:z=y\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon\exists x\in O_{\beta}\colon\llbracket{}z\rrbracket{}=\llbracket{}x\rrbracket{}\text{ and }\nexists y\in O_{\alpha}\colon\llbracket{}z\rrbracket{}=\llbracket{}y\rrbracket{}
and SO:Sα=Sβ and Sαβ and wOαβS:zw\displaystyle\text{ and }\exists S\subseteq O\colon\llbracket{}S_{\alpha}\rrbracket{}=\llbracket{}S_{\beta}\rrbracket{}\text{ and }S\vDash\alpha\lor\beta\text{ and }\forall w\in O_{\alpha\lor\beta}\setminus S\colon\llbracket{}z\rrbracket{}\nprec\llbracket{}w\rrbracket{}
Letting U=OSU=O\setminus S, we have OαβS=UαβO_{\alpha\lor\beta}\setminus S=U_{\alpha\lor\beta}:
\displaystyle\Leftrightarrow\; zT𝔰k:xOβ:z=x and yOα:z=y and SO:\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon\exists x\in O_{\beta}\colon\llbracket{}z\rrbracket{}=\llbracket{}x\rrbracket{}\text{ and }\nexists y\in O_{\alpha}\colon\llbracket{}z\rrbracket{}=\llbracket{}y\rrbracket{}\text{ and }\exists S\subseteq O\colon
Sα=Sβ and Sαβ and UO:U=OS and wUαβ:zw\displaystyle\llbracket{}S_{\alpha}\rrbracket{}=\llbracket{}S_{\beta}\rrbracket{}\text{ and }S\vDash\alpha\lor\beta\text{ and }\exists U\subseteq O\colon U=O\setminus S\text{ and }\forall w\in U_{\alpha\lor\beta}\colon\llbracket{}z\rrbracket{}\nprec\llbracket{}w\rrbracket{}
Clearly, the property wUαβ:zw\forall w\in U_{\alpha\lor\beta}:\llbracket{}z\rrbracket{}\nprec\llbracket{}w\rrbracket{} is preserved when taking subteams of UU. Hence, U=OSU=O\setminus S satisfies it if and only if some (not necessarily proper) superteam UU^{\prime} of OSO\setminus S does:
\displaystyle\Leftrightarrow\; zT𝔰k:xOβ:z=x and yOα:z=y\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon\exists x\in O_{\beta}\colon\llbracket{}z\rrbracket{}=\llbracket{}x\rrbracket{}\text{ and }\nexists y\in O_{\alpha}\colon\llbracket{}z\rrbracket{}=\llbracket{}y\rrbracket{}
and SO:Sα=Sβ and Sαβ\displaystyle\text{ and }\exists S\subseteq O\colon\llbracket{}S_{\alpha}\rrbracket{}=\llbracket{}S_{\beta}\rrbracket{}\text{ and }S\vDash\alpha\lor\beta
 and UO:UOS and wUαβ:zw\displaystyle\qquad\text{ and }\exists U^{\prime}\subseteq O\colon U^{\prime}\supseteq O\setminus S\text{ and }\forall w\in U^{\prime}_{\alpha\lor\beta}\colon\llbracket{}z\rrbracket{}\nprec\llbracket{}w\rrbracket{}
By Theorem 16:
\displaystyle\Leftrightarrow\; zT𝔰k:O(β1χk(𝔰,β))(α1χk(𝔰,α)) and SO:\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon O\vDash(\exists^{1}_{\beta}\chi_{k}(\mathfrak{s},\beta))\land({\sim}\exists^{1}_{\alpha}\chi_{k}(\mathfrak{s},\alpha))\text{ and }\exists S\subseteq O\colon
S(αβ)χk(α,β) and UO:UOS and wUαβ:zw\displaystyle S\vDash(\alpha\lor\beta)\land\chi^{*}_{k}(\alpha,\beta)\text{ and }\exists U^{\prime}\subseteq O\colon U^{\prime}\supseteq O\setminus S\text{ and }\forall w\in U^{\prime}_{\alpha\lor\beta}\colon\llbracket{}z\rrbracket{}\nprec\llbracket{}w\rrbracket{}
Note that T𝔰0,,T𝔰k1T_{\mathfrak{s}_{0}},\ldots,T_{\mathfrak{s}_{k-1}} are retained in OO. Moreover, SOαβS\subseteq O_{\alpha\lor\beta}, which implies that they are still subteams of OSO\setminus S and hence of UU^{\prime}. But by Claim (a), αβ\alpha\lor\beta and 𝔰k\mathfrak{s}_{k} are then k\prec_{k}-comparable scopes in UU^{\prime} and we can replace zw\llbracket{}z\rrbracket{}\nprec\llbracket{}w\rrbracket{}:
\displaystyle\Leftrightarrow\; zT𝔰k:O(β1χk(𝔰,β))(α1χk(𝔰,α))\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon O\vDash(\exists^{1}_{\beta}\chi_{k}(\mathfrak{s},\beta))\land({\sim}\exists^{1}_{\alpha}\chi_{k}(\mathfrak{s},\alpha))
and SO:S(αβ)χk(α,β)\displaystyle\text{ and }\exists S\subseteq O\colon S\vDash(\alpha\lor\beta)\land\chi^{*}_{k}(\alpha,\beta)
 and UO:UOS and wUαβ:(U)wαβζk(𝔰k,αβ)\displaystyle\qquad\text{ and }\exists U^{\prime}\subseteq O\colon U^{\prime}\supseteq O\setminus S\text{ and }\forall w\in U^{\prime}_{\alpha\lor\beta}\colon{(U^{\prime})}^{\alpha\lor\beta}_{w}\vDash{\sim}\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta)
Recalling that O=Tz𝔰kO=T^{\mathfrak{s}_{k}}_{z}, and by Proposition 15, we obtain:
\displaystyle\Leftrightarrow\; zT𝔰k:Tz𝔰k(β1χk(𝔰,β))(α1χk(𝔰,α))\displaystyle\exists z\in T_{\mathfrak{s}_{k}}\colon T^{\mathfrak{s}_{k}}_{z}\vDash(\exists^{1}_{\beta}\chi_{k}(\mathfrak{s},\beta))\land({\sim}\exists^{1}_{\alpha}\chi_{k}(\mathfrak{s},\alpha))
and STz𝔰k:S(αβ)χk(α,β)\displaystyle\text{ and }\exists S\subseteq T^{\mathfrak{s}_{k}}_{z}\colon S\vDash(\alpha\lor\beta)\land\chi^{*}_{k}(\alpha,\beta)
 and UTz𝔰k:UTz𝔰kS and Uαβ1ζk(𝔰k,αβ)\displaystyle\qquad\text{ and }\exists U^{\prime}\subseteq T^{\mathfrak{s}_{k}}_{z}\colon U^{\prime}\supseteq T^{\mathfrak{s}_{k}}_{z}\setminus S\text{ and }U^{\prime}\vDash\forall^{1}_{\alpha\lor\beta}{\sim}\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta)
\displaystyle\Leftrightarrow\; T𝔰k1(β1χk(𝔰,β))(α1χk(𝔰,α))\displaystyle T\vDash\exists^{1}_{\mathfrak{s}_{k}}(\exists^{1}_{\beta}\chi_{k}(\mathfrak{s},\beta))\land({\sim}\exists^{1}_{\alpha}\chi_{k}(\mathfrak{s},\alpha))
((αβ)χk(α,β))(αβ1ζk(𝔰k,αβ))\displaystyle\land\big{(}(\alpha\lor\beta)\land\chi^{*}_{k}(\alpha,\beta)\big{)}\lor\big{(}\forall^{1}_{\alpha\lor\beta}{\sim}\zeta_{k}(\mathfrak{s}_{k},\alpha\lor\beta)\big{)}
\displaystyle\Leftrightarrow\; Tζ(α,β).\displaystyle T\vDash\zeta^{*}(\alpha,\beta)\text{.} \triangleleft

In the next lemma, we prove the converse direction of Lemma 25.

Lemma 29.

Let k>0k>0, and let (𝒦,T)(\mathcal{K},T) be a kk-staircase with disjoint scopes α,β,𝔰0,,𝔰k1\alpha,\beta,\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k-1}. Then α\alpha and β\beta are k\prec_{k}-comparable in every subteam SS of TT that contains T𝔰0T𝔰k1T_{\mathfrak{s}_{0}}\cup\cdots\cup T_{\mathfrak{s}_{k-1}}.

Proof 7.6.

The proof is by induction on kk. Disjoint scopes α\alpha and β\beta are always 0\prec_{0}-comparable, which can be easily seen in ζ0\zeta_{0}. For the inductive step to k+1k+1, assume (𝒦,T)(\mathcal{K},T) and SS as above, and let 𝒦=(W,R,V)\mathcal{K}=(W,R,V). Let O:=Sw,vα,βO\vcentcolon=S^{\alpha,\beta}_{w,v} with wSα,vSβw\in S_{\alpha},v\in S_{\beta} arbitrary.

Claim 30 (a).

α\alpha and β\beta are k\prec^{*}_{k}-comparable in RORO.

Proof 7.7 (Proof of claim.).

In the inductive step, now 𝔰0,,𝔰k,α,β\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k},\alpha,\beta are disjoint scopes. Additionally, (𝒦,RT)(\mathcal{K},RT) is a kk-staircase. In particular, in the induction step α\alpha and β\beta are disjoint from 𝔰k\mathfrak{s}_{k}. For this reason, (𝒦,RO)(\mathcal{K},RO) is a kk-staircase as well, as (RO)𝔰0𝔰k=(RT)𝔰0𝔰k{(RO)}_{\mathfrak{s}_{0}\lor\cdots\lor\mathfrak{s}_{k}}={(RT)}_{\mathfrak{s}_{0}\lor\cdots\lor\mathfrak{s}_{k}}.

Hence, by induction hypothesis, for every team UU such that RO𝔰0RO𝔰k1URORO_{\mathfrak{s}_{0}}\cup\cdots\cup RO_{\mathfrak{s}_{k-1}}\subseteq U\subseteq RO, we obtain that 𝔰k\mathfrak{s}_{k} and α\alpha are k\prec_{k}-comparable in UU, as well as 𝔰k\mathfrak{s}_{k} and β\beta. Consequently, we can apply Lemma 25, which proves the claim. \triangleleft

We proceed with the induction step. Again by symmetry, we only show that Oζk+1(α,β)O\vDash\zeta_{k+1}(\alpha,\beta) iff wk+1k+1vk+1\llbracket{}w\rrbracket{}_{k+1}\prec_{k+1}\llbracket{}v\rrbracket{}_{k+1}. We distinguish three cases w. r. t.\xspace0\prec_{0}:

  • If w00v0\llbracket{}w\rrbracket{}_{0}\prec_{0}\llbracket{}v\rrbracket{}_{0}, then Oζ0(α,β)O\vDash\zeta_{0}(\alpha,\beta) by the induction basis. As the former implies wk+1k+1vk+1\llbracket{}w\rrbracket{}_{k+1}\prec_{k+1}\llbracket{}v\rrbracket{}_{k+1} and the latter Oζk+1(α,β)O\vDash\zeta_{k+1}(\alpha,\beta), the equivalence holds.

  • If w00v0\llbracket{}w\rrbracket{}_{0}\succ_{0}\llbracket{}v\rrbracket{}_{0}, then wk+1k+1vk+1\llbracket{}w\rrbracket{}_{k+1}\nprec_{k+1}\llbracket{}v\rrbracket{}_{k+1}. Moreover, Oζ0(α,β)O\nvDash\zeta_{0}(\alpha,\beta) by induction basis. Additionally, Oχ0(α,β)O\nvDash\chi_{0}(\alpha,\beta) by Theorem 16. Consequently, both sides of the equivalence are false.

  • If w=0v0\llbracket{}w\rrbracket{}_{0}=\llbracket{}v\rrbracket{}_{0}, then Oχ0(α,β)O\vDash\chi_{0}(\alpha,\beta) by Theorem 16, but Oζ0(α,β)O\nvDash\zeta_{0}(\alpha,\beta) by induction basis. Consequently, Oζk+1(α,β)O\vDash\zeta_{k+1}(\alpha,\beta) iff Oζk(α,β)O\vDash\Box\zeta^{*}_{k}(\alpha,\beta). Also, wk+1k+1vk+1\llbracket{}w\rrbracket{}_{k+1}\prec_{k+1}\llbracket{}v\rrbracket{}_{k+1} iff wkk+1vk+1\mathcal{R}\llbracket{}w\rrbracket{}_{k+1}\prec^{*}_{k}\mathcal{R}\llbracket{}v\rrbracket{}_{k+1}. The following equivalence concludes the proof:

    wkk+1vk+1\displaystyle\mathcal{R}\llbracket{}w\rrbracket{}_{k+1}\prec^{*}_{k}\mathcal{R}\llbracket{}v\rrbracket{}_{k+1}
    \displaystyle\Leftrightarrow\; RwkkRvk\displaystyle\llbracket{}Rw\rrbracket{}_{k}\prec^{*}_{k}\llbracket{}Rv\rrbracket{}_{k} (By Proposition 6)
    \displaystyle\Leftrightarrow\; ROζk(α,β)\displaystyle RO\vDash\zeta^{*}_{k}(\alpha,\beta) (By Claim (a))
    \displaystyle\Leftrightarrow\; Oζk(α,β).\displaystyle O\vDash\Box\zeta^{*}_{k}(\alpha,\beta)\text{.}

With the above lemmas we are now in the position to prove Theorem 24:

Proof 7.8 (Proof of Theorem 24).

First, it is straightforward to construct ζk\zeta_{k} and ζk\zeta^{*}_{k} in logarithmic space. For the correctness, let (𝒦,T)(\mathcal{K},T) be a model with disjoint scopes α,β,𝔰0,,𝔰k\alpha,\beta,\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} as in the theorem. By Lemma 29 it immediately follows that α\alpha and β\beta are k\prec_{k}-comparable in TT. The second part, that α\alpha and β\beta are k\prec^{*}_{k}-comparable in TT, follows from the combination of Lemma 25 and 29.

8. Encoding non-elementary computations

We combine all the previous sections and extend Theorem 12 and Corollary 13 by their matching lower bounds:

Theorem 31.
  • 𝖲𝖠𝖳(𝖬𝖳𝖫\xspace)\mathsf{SAT}(\mathsf{MTL}\xspace) and 𝖵𝖠𝖫(𝖬𝖳𝖫\xspace)\mathsf{VAL}(\mathsf{MTL}\xspace) are complete for TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}).

  • If k0k\geq 0, then 𝖲𝖠𝖳(𝖬𝖳𝖫\xspacek)\mathsf{SAT}(\mathsf{MTL}\xspace_{k}) and 𝖵𝖠𝖫(𝖬𝖳𝖫\xspacek)\mathsf{VAL}(\mathsf{MTL}\xspace_{k}) are complete for ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}).

The above complexity classes are complement-closed, and additionally 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace and 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} are syntactically closed under negation. For this reason, it suffices to prove the hardness of 𝖲𝖠𝖳(𝖬𝖳𝖫\xspace)\mathsf{SAT}(\mathsf{MTL}\xspace) and 𝖲𝖠𝖳(𝖬𝖳𝖫\xspacek)\mathsf{SAT}(\mathsf{MTL}\xspace_{k}), respectively. Moreover, the case k=0k=0 is equivalent to 𝖲𝖠𝖳(𝖯𝖳𝖫\xspace)\mathsf{SAT}(\mathsf{PTL}\xspace) being ATIME\xspace-ALT\xspace(exp,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp,\mathrm{poly})-hard, which was proven by Hannula et al. [HKVV18]. Their reduction also works in logarithmic space. Consequently, the result boils down to the following lemma:

Lemma 32.
  • If LTOWER\xspace(poly)L\in\mathrm{TOWER}\xspace(\mathrm{poly}), then Lmlog𝖲𝖠𝖳(𝖬𝖳𝖫\xspace)L\leq^{\log}_{\mathrm{m}}\mathsf{SAT}(\mathsf{MTL}\xspace).

  • If k1k\geq 1 and LATIME\xspace-ALT\xspace(expk+1,poly)L\in\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}), then Lmlog𝖲𝖠𝖳(𝖬𝖳𝖫\xspacek)L\leq^{\log}_{\mathrm{m}}\mathsf{SAT}(\mathsf{MTL}\xspace_{k}).

We devise for each LL a reduction xφxx\mapsto\varphi_{x} such that φx\varphi_{x} is a formula that is satisfiable if and only if xLx\in L. By assumption, there exists a single-tape alternating Turing machine MM that decides LL (for LTOWER\xspace(poly)L\in\mathrm{TOWER}\xspace(\mathrm{poly}), w.l.o.g.\xspaceMM is alternating as well).

Let MM have states QQ, which is the disjoint union of QQ_{\exists} (existential states), QQ_{\forall} (universal states), QaccQ_{\text{acc}} (accepting states) and QrejQ_{\text{rej}} (rejecting states). Also, QQ contains some initial state q0q_{0}. Let MM have a finite tape alphabet Γ\Gamma with blank symbol Γ\flat\in\Gamma, and a transition relation δ\delta.

We design φx\varphi_{x} in a fashion that forces its models (𝒦,T)(\mathcal{K},T) to encode an accepting computation of MM on xx. Let us call any legal sequence of configurations of MM (not necessarily starting with the initial configuration) a run. Then, similarly as in Cook’s theorem [Coo71], we encode runs as square “grids” with a vertical “time” coordinate and a horizontal “space” coordinate in the model, i.e.\xspace, each row of the grid represents a configuration of MM.

W.l.o.g.\xspaceMM never leaves the input to the left, and there exists NN that is an upper bound on both the length of a configuration and the runtime of MM. Formally, a run of MM is then a function C:{1,,N}2Γ(Q×Γ)C\colon{\{1,\ldots,N\}}^{2}\to\Gamma\cup(Q\times\Gamma), Here, C(i,j)=cC(i,j)=c for cΓc\in\Gamma means that the ii-th configuration (i.e.\xspace, after MM performed i1i-1 transitions) contains the symbol cc in its jj-th cell. The same holds if C(i,j)=(q,c)C(i,j)=(q,c) for (q,c)Q×Γ(q,c)\in Q\times\Gamma, but then additionally the machine is in the state qq with its head visiting the jj-th cell in the ii-th configuration. As an example, for a run CC from MM’s initial configuration we have C(1,1)=(q0,x1)C(1,1)=(q_{0},x_{1}), C(1,i)=xiC(1,i)=x_{i} for 2in2\leq i\leq n, and C(1,i)=C(1,i)=\flat for n<iNn<i\leq N.

Due to the semantics of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace, such a run must be encoded in (𝒦,T)(\mathcal{K},T) very carefully. We let the team TT contain N2N^{2} worlds wi,jw_{i,j} in which the respective value of C(i,j)C(i,j) is encoded as a propositional assignment. However, we cannot simply pursue the standard approach of assembling a large N×NN\times N-grid in the edge relation RR in order to compare successive configurations; by Corollary 18, we cannot force the model to contain RR-paths longer than |φx|{|\nobreak\varphi_{x}\nobreak|}. Instead, to define grid neighborship, we let wi,jw_{i,j} encode ii and jj in its type. More precisely, we use the linear order k\prec_{k} on Δk\Delta_{k} we defined with the 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formula ζk\zeta_{k} in the previous section. Then, instead of using \Box and \Diamond, we examine the grid by letting ζk\zeta_{k} judge whether a given pair of worlds is deemed (horizontally or vertically) adjacent.

8.1. Encoding runs in a team

Next, we discuss how runs C:{1,,N}2Γ(Q×Γ)C\colon{\{1,\ldots,N\}}^{2}\to\Gamma\cup(Q\times\Gamma) are encoded in TT. Given a world wTw\in T, we partition the image RwRw with two special propositions 𝔱Φ\mathfrak{t}\notin\Phi (“timestep”) and 𝔭Φ\mathfrak{p}\notin\Phi (“position”). Then we assign to ww the pair (w):=(i,j)\ell(w)\vcentcolon=(i,j) such that (Rw)𝔱k1\llbracket{}{(Rw)}_{\mathfrak{t}}\rrbracket{}_{k-1} is the ii-th element, and (Rw)𝔭k1\llbracket{}{(Rw)}_{\mathfrak{p}}\rrbracket{}_{k-1} is the jj-th element in the order k1\prec^{*}_{k-1}. We call the pair (w)\ell(w) the location of ww (in the grid).

Accordingly, we fix N:=|𝔓(Δk1Φ)|N\vcentcolon={|\nobreak\mathfrak{P}(\Delta^{\Phi}_{k-1})\nobreak|}. For the case of fixed kk, MM has runtime bounded by expk+1(g(n))\exp_{k+1}(g(n)) for a polynomial gg. Then taking Φ:={p1,,pg(n)}\Phi\vcentcolon=\{p_{1},\ldots,p_{g(n)}\} yields a sufficiently large coordinate space, as

expk+1(g(n))=expk+1\displaystyle\exp_{k+1}(g(n))=\exp_{k+1} (|Φ|)=2expk1(2|Φ|)2expk1(2|Φ|)=2|Δk1Φ|=N\displaystyle({|\nobreak\Phi\nobreak|})=2^{\exp_{k-1}\left(2^{|\nobreak\Phi\nobreak|}\right)}\leq 2^{\exp^{*}_{k-1}\left(2^{|\nobreak\Phi\nobreak|}\right)}=2^{{|\nobreak\Delta^{\Phi}_{k-1}\nobreak|}}=N

by Proposition 9. For runtime expg(n)(1)\exp_{g(n)}(1) of MM, we let Φ:=\Phi\vcentcolon=\emptyset and precompute k:=g(|x|)+1k\vcentcolon=g({|\nobreak x\nobreak|})+1, but otherwise proceed identically.

Next, let Ξ\Xi be a constant set of propositions disjoint from Φ\Phi that encodes the range of CC via some bijection c:ΞΓ(Q×Γ)c\colon\Xi\to\Gamma\cup(Q\times\Gamma). If a world ww satisfies exactly one proposition pp of those in Ξ\Xi, then by slight abuse of notation we write c(w)c(w) instead of c(p)c(p). Intuitively, c(w)Γ(Q×Γ)c(w)\in\Gamma\cup(Q\times\Gamma) is the content of the grid cell represented by ww.

Using \ell and cc, the function CC can be encoded into a team TT as follows. First, a team TT is called grid if every point in TT satisfies exactly one proposition in Ξ\Xi, and if every location (i,j){1,,N}2(i,j)\in{\{1,\ldots,N\}}^{2} occurs as (w)\ell(w) for some point wTw\in T. Moreover, a grid TT is called pre-tableau if for every location (i,j)(i,j) and every element pΞp\in\Xi there is some world wTw\in T such that (w)=(i,j)\ell(w)=(i,j) and wpw\vDash p. Finally, a grid TT is a tableau if any two elements w,wTw,w^{\prime}\in T with (w)=(w)\ell(w)=\ell(w^{\prime}) also agree on Ξ\Xi, i.e.\xspace, c(w)=c(w)c(w)=c(w^{\prime}).

Let us motivate the above definitions. Clearly, the definition of a grid TT means that TT captures the whole domain of CC, and that cc is well-defined on the level of points. If TT is additionally a tableau, then cc is also well-defined on the level of locations. In other words, a tableau TT induces a function CT:{1,,N}2Γ(Q×Γ)C_{T}\colon{\{1,\ldots,N\}}^{2}\to\Gamma\cup(Q\times\Gamma) via Cα(i,j):=c(w)C_{\alpha}(i,j)\vcentcolon=c(w), where wTw\in T is arbitrary such that (w)=(i,j)\ell(w)=(i,j).

A pre-tableau can be seen as the union of all possible CC. In particular, given any pre-tableau, the definition ensures that arbitrary tableaus can be obtained from it by the means of subteam quantification \exists^{\subseteq} (cf. p. 5.2).

A tableau TT is legal if CTC_{T} is a run of MM, i.e.\xspace, if every row is a configuration of MM, and if every pair of two successive rows represents a valid δ\delta-transition.

The idea of the reduction is now to capture the alternating computation of MM by nesting polynomially many quantifications (via \exists^{\subseteq} and \forall^{\subseteq}) of legal tableaus, of which each one is the continuation of the computation of the previous one.

8.2. Accessing two components of locations

An discussed earlier, we choose to represent a location (i,j)(i,j) in a point ww as a pair (Δ,Δ′′)(\Delta^{\prime},\Delta^{\prime\prime}) by stipulating that Δ=(Rw)𝔱k1\Delta^{\prime}=\llbracket{}{(Rw)}_{\mathfrak{t}}\rrbracket{}_{k-1} and Δ′′=(Rw)𝔭k1\Delta^{\prime\prime}=\llbracket{}{(Rw)}_{\mathfrak{p}}\rrbracket{}_{k-1}. To access the two components of a encoded location independently, we introduce the operator

|𝔮αψ:=(α¬𝔮)((α𝔮)ψ),|^{\alpha}_{\mathfrak{q}}\,\psi\vcentcolon=(\alpha\land\neg\mathfrak{q})\lor((\alpha\hookrightarrow\mathfrak{q})\land\psi)\text{,}

where 𝔮{𝔱,𝔭}\mathfrak{q}\in\{\mathfrak{t},\mathfrak{p}\} and α𝖬𝖫\xspace\alpha\in\mathsf{ML}\xspace. It is easy to check that T|𝔮αψT\vDash|^{\alpha}_{\mathfrak{q}}\,\psi iff TT𝔮αψT^{\alpha}_{T_{\mathfrak{q}}}\vDash\psi.

In order to compare the locations of grid cells, for each component 𝔮{𝔱,𝔭}\mathfrak{q}\in\{\mathfrak{t},\mathfrak{p}\} we define the following formulas: ψ𝔮(α,β)\psi^{\mathfrak{q}}_{\prec}(\alpha,\beta) tests whether the location in TαT_{\alpha} is less than the one in TβT_{\beta} w. r. t.\xspaceits 𝔮\mathfrak{q}-component (assuming singleton teams TαT_{\alpha} and TβT_{\beta}). Analogously, ψ𝔮(α,β)\psi^{\mathfrak{q}}_{\equiv}(\alpha,\beta) checks for equality of the respective component:

ψ𝔮(α,β):=\displaystyle\psi^{\mathfrak{q}}_{\prec}(\alpha,\beta)\vcentcolon=\; |𝔮α|𝔮βζk1(α,β)\displaystyle\Box\,|^{\alpha}_{\mathfrak{q}}|^{\beta}_{\mathfrak{q}}\zeta_{k-1}^{*}(\alpha,\beta)
ψ𝔮(α,β):=\displaystyle\psi^{\mathfrak{q}}_{\equiv}(\alpha,\beta)\vcentcolon=\; |𝔮α|𝔮βχk1(α,β)\displaystyle\Box\,|^{\alpha}_{\mathfrak{q}}|^{\beta}_{\mathfrak{q}}\chi^{*}_{k-1}(\alpha,\beta)

For this purpose, ψ𝔮\psi^{\mathfrak{q}}_{\prec} is built upon the formula ζk1\zeta^{*}_{k-1} from Theorem 24, while ψ𝔮\psi^{\mathfrak{q}}_{\equiv} checks for equality with the help of χk1\chi^{*}_{k-1} from Theorem 16.

Claim 33 (a).

Let 𝒦\mathcal{K} be a structure with a team TT and disjoint scopes α\alpha and β\beta. Suppose wTαw\in T_{\alpha} and vTβv\in T_{\beta}, where (w)=(iw,jw)\ell(w)=(i_{w},j_{w}) and (v)=(iv,jv)\ell(v)=(i_{v},j_{v}). Then:

Tw,vα,βψ𝔱(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\psi^{\mathfrak{t}}_{\equiv}(\alpha,\beta) iw=iv\displaystyle\;\Leftrightarrow\;i_{w}=i_{v}
Tw,vα,βψ𝔭(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\psi^{\mathfrak{p}}_{\equiv}(\alpha,\beta) jw=jv\displaystyle\;\Leftrightarrow\;j_{w}=j_{v}

Moreover, if α,β,𝔰0,,𝔰k\alpha,\beta,\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k} are disjoint scopes in 𝒦\mathcal{K} and (𝒦,T)(\mathcal{K},T) is a kk-staircase, then:

Tw,vα,βψ𝔱(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\psi^{\mathfrak{t}}_{\prec}(\alpha,\beta) iw<iv\displaystyle\;\Leftrightarrow\;i_{w}<i_{v}
Tw,vα,βψ𝔭(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\psi^{\mathfrak{p}}_{\prec}(\alpha,\beta) jw<jv\displaystyle\;\Leftrightarrow\;j_{w}<j_{v}
Proof 8.1 (Proof of claim.).

Let us begin with ψ𝔱\psi^{\mathfrak{t}}_{\equiv} (ψ𝔭\psi^{\mathfrak{p}}_{\equiv} works identically):

iw=iv\displaystyle i_{w}=i_{v}\;\Leftrightarrow\; (Rw)𝔱=k1(Rv)𝔱k1\displaystyle\llbracket{}{(Rw)}_{\mathfrak{t}}\rrbracket{}_{k-1}=\llbracket{}{(Rv)}_{\mathfrak{t}}\rrbracket{}_{k-1} (By Definition)
\displaystyle\Leftrightarrow\; RT(Rw)𝔱,(Rv)𝔱α,βχk1(α,β)\displaystyle RT^{\alpha,\beta}_{{(Rw)}_{\,\mathfrak{t}},{(Rv)}_{\,\mathfrak{t}}}\vDash\chi^{*}_{k-1}(\alpha,\beta) (By Theorem 16)
\displaystyle\Leftrightarrow\; (RTRw,Rvα,β)RT𝔱,RT𝔱α,βχk1(α,β)\displaystyle{\Big{(}RT^{\alpha,\beta}_{Rw,Rv}\Big{)}}^{\alpha,\beta}_{RT_{\mathfrak{t}},RT_{\mathfrak{t}}}\vDash\chi^{*}_{k-1}(\alpha,\beta)
\displaystyle\Leftrightarrow\; RTRw,Rvα,β|𝔱α|𝔱βχk1(α,β)\displaystyle RT^{\alpha,\beta}_{Rw,Rv}\vDash|^{\alpha}_{\mathfrak{t}}|^{\beta}_{\mathfrak{t}}\chi^{*}_{k-1}(\alpha,\beta)
\displaystyle\Leftrightarrow\; Tw,vα,β|𝔱α|𝔱βχk1(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\Box\,|^{\alpha}_{\mathfrak{t}}|^{\beta}_{\mathfrak{t}}\chi^{*}_{k-1}(\alpha,\beta) (Proposition 14)

Similarly for ψ𝔱\psi^{\mathfrak{t}}_{\prec} (ψ𝔭\psi^{\mathfrak{p}}_{\prec} again works identically):

iw<iv\displaystyle i_{w}<i_{v}\;\Leftrightarrow\; (Rw)𝔱k1k1(Rv)𝔱k1\displaystyle\llbracket{}{(Rw)}_{\mathfrak{t}}\rrbracket{}_{k-1}\prec^{*}_{k-1}\llbracket{}{(Rv)}_{\mathfrak{t}}\rrbracket{}_{k-1} (By Definition)
\displaystyle\Leftrightarrow\; RT(Rw)𝔱,(Rv)𝔱α,βζk1(α,β)\displaystyle RT^{\alpha,\beta}_{{(Rw)}_{\,\mathfrak{t}},{(Rv)}_{\,\mathfrak{t}}}\vDash\zeta^{*}_{k-1}(\alpha,\beta) (By Theorem 24)
\displaystyle\Leftrightarrow\; (RTRw,Rvα,β)T𝔱,T𝔱α,βζk1(α,β)\displaystyle{\Big{(}RT^{\alpha,\beta}_{Rw,Rv}\Big{)}}^{\alpha,\beta}_{T_{\mathfrak{t}},T_{\mathfrak{t}}}\vDash\zeta^{*}_{k-1}(\alpha,\beta)
\displaystyle\Leftrightarrow\; RTRw,Rvα,β|𝔱α|𝔱βζk1(α,β)\displaystyle RT^{\alpha,\beta}_{Rw,Rv}\vDash|^{\alpha}_{\mathfrak{t}}|^{\beta}_{\mathfrak{t}}\zeta^{*}_{k-1}(\alpha,\beta)
\displaystyle\Leftrightarrow\; Tw,vα,β|𝔱α|𝔱βζk1(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\Box\,|^{\alpha}_{\mathfrak{t}}|^{\beta}_{\mathfrak{t}}\zeta^{*}_{k-1}(\alpha,\beta) (Proposition 14)  \triangleleft

8.3. Defining grids, pre-tableaus, and tableaus

Next, we aim at constructing formulas that check whether a given team is a grid, pre-tableau, or a tableau, respectively.

First, to check that every location (i,j){1,,N}2(i,j)\in{\{1,\ldots,N\}}^{2} of the grid occurs as (w)\ell(w) of some wTw\in T, we quantify over all corresponding pairs (Δ,Δ′′)𝔓(Δk1)2(\Delta^{\prime},\Delta^{\prime\prime})\in\mathfrak{P}(\Delta_{k-1})^{2}. To cover all these sets of types we can quantify, for instance, over the images of all points of T𝔰kT_{\mathfrak{s}_{k}}. However, as subteam quantifiers ,1,,1\exists^{\subseteq},\exists^{1},\forall^{\subseteq},\forall^{1} cannot pick two subteams from the same scope, we enforce a kk-canonical copy 𝔰k\mathfrak{s}^{\prime}_{k} of 𝔰k\mathfrak{s}_{k} in the spirit of Theorem 20:

𝖼𝖺𝗇𝗈𝗇:=\displaystyle\mathsf{canon}^{\prime}\vcentcolon=\; ρ0k(𝔰0)m=1kρmkm(𝔰m1,𝔰m)ρk0(𝔰k1,𝔰k)\displaystyle\rho^{k}_{0}(\mathfrak{s}_{0})\land\bigwedge_{m=1}^{k}\rho^{k-m}_{m}(\mathfrak{s}_{m-1},\mathfrak{s}_{m})\land\rho^{0}_{k}(\mathfrak{s}_{k-1},\mathfrak{s}^{\prime}_{k})
Claim 34 (b).

If 𝔰0,,𝔰k,𝔰k\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k} are disjoint scopes in 𝒦\mathcal{K}, then (𝒦,T)𝖼𝖺𝗇𝗈𝗇(\mathcal{K},T)\vDash\mathsf{canon}^{\prime} if and only if (𝒦,T)(\mathcal{K},T) is a kk-staircase and T𝔰kT_{\mathfrak{s}^{\prime}_{k}} is kk-canonical. Moreover, 𝖼𝖺𝗇𝗈𝗇𝗌𝖼𝗈𝗉𝖾𝗌k({𝔰0,,𝔰k,𝔰k})k+1\mathsf{canon}^{\prime}\land\mathsf{scopes}_{k}(\{\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}\})\land\Box^{k+1}\bot is satisfiable, but is only satisfied by kk-staircases (𝒦,T)(\mathcal{K},T) in which both T𝔰kT_{\mathfrak{s}_{k}} and T𝔰kT_{\mathfrak{s}^{\prime}_{k}} are kk-canonical. Furthermore, both formulas are constructible in space 𝒪(log(|Φ|+k)){\mathcal{O}(\log({|\nobreak\Phi\nobreak|}+k))}.

Proof 8.2 (Proof of claim.).

Proven similarly to Theorem 20 and 22. \triangleleft

The next formula checks whether a given team is a grid. More precisely, the subformula ψpair\psi_{\text{pair}} compares the 𝔱\mathfrak{t}-component of the selected location in α\alpha to the image of the world quantified in 𝔰k\mathfrak{s}_{k}, and its 𝔭\mathfrak{p}-component to 𝔰k\mathfrak{s}^{\prime}_{k}, respectively. That every world satisfies exactly one element of Ξ\Xi is guaranteed by ψgrid\psi_{\text{grid}} as well.

ψgrid(α):=\displaystyle\psi_{\text{grid}}(\alpha)\vcentcolon=\; (αeΞeeΞee¬e)𝔰k1𝔰k1α1ψpair(α)\displaystyle\Big{(}\alpha\hookrightarrow\bigvee_{e\in\Xi}e\land\bigwedge_{\begin{subarray}{c}e^{\prime}\in\Xi\\ e^{\prime}\neq e\end{subarray}}\neg e^{\prime}\Big{)}\land\forall^{1}_{\mathfrak{s}_{k}}\mkern 2.0mu\forall^{1}_{\mathfrak{s}^{\prime}_{k}}\mkern 2.0mu\exists^{1}_{\alpha}\,\psi_{\text{pair}}(\alpha)
ψpair(α):=\displaystyle\psi_{\text{pair}}(\alpha)\vcentcolon=\; [(|𝔱αχk1(𝔰k,α))(|𝔭αχk1(𝔰k,α))]\displaystyle\Box\left[\big{(}\,|^{\alpha}_{\mathfrak{t}}\;\chi^{*}_{k-1}(\mathfrak{s}_{k},\alpha)\big{)}\land\big{(}\,|^{\alpha}_{\mathfrak{p}}\;\chi^{*}_{k-1}(\mathfrak{s}^{\prime}_{k},\alpha)\big{)}\right]

In the following and all subsequent claims, we always assume that TT is a team in a Kripke structure 𝒦\mathcal{K} such that (𝒦,T)(\mathcal{K},T) satisfies 𝖼𝖺𝗇𝗈𝗇k+1\mathsf{canon}^{\prime}\land\Box^{k+1}\bot. Moreover, all stated scopes are always assumed pairwise disjoint in 𝒦\mathcal{K} (as we can enforce this later in the reduction with 𝗌𝖼𝗈𝗉𝖾𝗌k()\mathsf{scopes}_{k}(\cdots)).

Claim 35 (c).

Tψgrid(α)T\vDash\psi_{\text{grid}}(\alpha) if and only if TαT_{\alpha} is a grid.

Proof 8.3 (Proof of claim.).

Clearly TαeΞeeΞ,ee¬eT\vDash\alpha\hookrightarrow\bigvee_{e\in\Xi}e\land\bigwedge_{e^{\prime}\in\Xi,e^{\prime}\neq e}\neg e^{\prime} if and only if every world wTαw\in T_{\alpha} satisfies exactly one element of Ξ\Xi. Consequently, for the proof it remains to show the following equivalence:

(i,j){1,,N}2:wTα:(w)=(i,j)\displaystyle\forall(i,j)\in{\{1,\ldots,N\}}^{2}\colon\exists w\in T_{\alpha}\colon\ell(w)=(i,j)
\displaystyle\Leftrightarrow\; Δ,Δ′′Δk1:wTα:(Rw)𝔱=k1Δ and (Rw)𝔭=k1Δ′′\displaystyle\forall\Delta^{\prime},\Delta^{\prime\prime}\subseteq\Delta_{k-1}\colon\exists w\in T_{\alpha}:\llbracket{}{(Rw)}_{\mathfrak{t}}\rrbracket{}_{k-1}=\Delta^{\prime}\text{ and }\llbracket{}{(Rw)}_{\mathfrak{p}}\rrbracket{}_{k-1}=\Delta^{\prime\prime}
By kk-canonicity of 𝔰k,𝔰k\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k} due to Claim (b):
\displaystyle\Leftrightarrow\; vT𝔰k,vT𝔰k:wTα:(Rw)𝔱=k1Rv and k1(Rw)𝔭=k1Rvk1\displaystyle\forall v\in T_{\mathfrak{s}_{k}},v^{\prime}\in T_{\mathfrak{s}^{\prime}_{k}}\colon\exists w\in T_{\alpha}\colon\llbracket{}{(Rw)}_{\mathfrak{t}}\rrbracket{}_{k-1}=\llbracket{}Rv\rrbracket{}_{k-1}\text{ and }\llbracket{}{(Rw)}_{\mathfrak{p}}\rrbracket{}_{k-1}=\llbracket{}Rv^{\prime}\rrbracket{}_{k-1}
By Theorem 16:
\displaystyle\Leftrightarrow\; vT𝔰k,vT𝔰k:wTα:RT(Rw)𝔱,Rv,Rvα,𝔰k,𝔰kχk1(𝔰k,α)\displaystyle\forall v\in T_{\mathfrak{s}_{k}},v^{\prime}\in T_{\mathfrak{s}^{\prime}_{k}}\colon\exists w\in T_{\alpha}\colon RT^{\alpha,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}}_{{(Rw)}_{\mathfrak{t}},Rv,Rv^{\prime}}\vDash\chi^{*}_{k-1}(\mathfrak{s}_{k},\alpha)
 and RT(Rw)𝔭,Rv,Rvα,𝔰k,𝔰kχk1(𝔰k,α)\displaystyle\qquad\qquad\text{ and }RT^{\alpha,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}}_{{(Rw)}_{\mathfrak{p}},Rv,Rv^{\prime}}\vDash\chi^{*}_{k-1}(\mathfrak{s}^{\prime}_{k},\alpha)
\displaystyle\Leftrightarrow\; vT𝔰k,vT𝔰k:wTα:(RTRw,Rv,Rvα,𝔰k,𝔰k)RT𝔱αχk1(𝔰k,α)\displaystyle\forall v\in T_{\mathfrak{s}_{k}},v^{\prime}\in T_{\mathfrak{s}^{\prime}_{k}}\colon\exists w\in T_{\alpha}:{\big{(}RT^{\alpha,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}}_{Rw,Rv,Rv^{\prime}}\big{)}}^{\alpha}_{RT_{\mathfrak{t}}}\vDash\chi^{*}_{k-1}(\mathfrak{s}_{k},\alpha)
 and (RTRw,Rv,Rvα,𝔰k,𝔰k)RT𝔭αχk1(𝔰k,α)\displaystyle\qquad\qquad\text{ and }{\big{(}RT^{\alpha,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}}_{Rw,Rv,Rv^{\prime}}\big{)}}^{\alpha}_{RT_{\mathfrak{p}}}\vDash\chi^{*}_{k-1}(\mathfrak{s}^{\prime}_{k},\alpha)
\displaystyle\Leftrightarrow\; vT𝔰k,vT𝔰k:wTα:RTRw,Rv,Rvα,𝔰k,𝔰k|𝔱αχk1(𝔰k,α)|𝔭αχk1(𝔰k,α)\displaystyle\forall v\in T_{\mathfrak{s}_{k}},v^{\prime}\in T_{\mathfrak{s}^{\prime}_{k}}\colon\exists w\in T_{\alpha}\colon RT^{\alpha,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}}_{Rw,Rv,Rv^{\prime}}\vDash\,|^{\alpha}_{\mathfrak{t}}\chi^{*}_{k-1}(\mathfrak{s}_{k},\alpha)\land\,|^{\alpha}_{\mathfrak{p}}\chi^{*}_{k-1}(\mathfrak{s}^{\prime}_{k},\alpha)
By Proposition 14:
\displaystyle\Leftrightarrow\; vT𝔰k,vT𝔰k:wTα:Tw,v,vα,𝔰k,𝔰k|𝔱αχk1(𝔰k,α)|𝔭αχk1(𝔰k,α)\displaystyle\forall v\in T_{\mathfrak{s}_{k}},v^{\prime}\in T_{\mathfrak{s}^{\prime}_{k}}\colon\exists w\in T_{\alpha}\colon T^{\alpha,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}}_{w,v,v^{\prime}}\vDash\Box\,|^{\alpha}_{\mathfrak{t}}\chi^{*}_{k-1}(\mathfrak{s}_{k},\alpha)\land\,|^{\alpha}_{\mathfrak{p}}\chi^{*}_{k-1}(\mathfrak{s}^{\prime}_{k},\alpha)
By Proposition 15:
\displaystyle\Leftrightarrow\; T𝔰k1𝔰k1α1|𝔱αχk1(𝔰k,α)|𝔭αχk1(𝔰k,α)\displaystyle T\vDash\forall^{1}_{\mathfrak{s}_{k}}\forall^{1}_{\mathfrak{s}^{\prime}_{k}}\exists^{1}_{\alpha}\Box\,|^{\alpha}_{\mathfrak{t}}\chi^{*}_{k-1}(\mathfrak{s}_{k},\alpha)\land\,|^{\alpha}_{\mathfrak{p}}\chi^{*}_{k-1}(\mathfrak{s}^{\prime}_{k},\alpha)
\displaystyle\Leftrightarrow\; T𝔰k1𝔰k1α1ψpair(α)\displaystyle T\vDash\forall^{1}_{\mathfrak{s}_{k}}\forall^{1}_{\mathfrak{s}^{\prime}_{k}}\exists^{1}_{\alpha}\,\psi_{\text{pair}}(\alpha) \triangleleft

With slight modifications it is straightforward to define pre-tableaus:

ψpre-tableau(α):=\displaystyle\psi_{\text{pre-tableau}}(\alpha)\vcentcolon=\; ψgrid(α)𝔰k1𝔰k1eΞα1(ψpair(α)(αe))\displaystyle\psi_{\text{grid}}(\alpha)\land\forall^{1}_{\mathfrak{s}_{k}}\mkern 2.0mu\forall^{1}_{\mathfrak{s}^{\prime}_{k}}\mkern 2.0mu\bigwedge_{e\in\Xi}\exists^{1}_{\alpha}\big{(}\psi_{\text{pair}}(\alpha)\land(\alpha\hookrightarrow e)\big{)}
Claim 36 (d).

Tψpre-tableau(α)T\vDash\psi_{\text{pre-tableau}}(\alpha) if and only if TαT_{\alpha} is a pre-tableau.

Proof 8.4 (Proof of claim.).

Proven similarly to Claim (c). \triangleleft

The other special case of a grid, that is, a tableau, requires a more elaborate approach to define in 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace. The difference to a grid or pre-tableau is that we have to quantify over all pairs (w,w)(w,w^{\prime}) of points in TT, and check that they agree on Ξ\Xi if (w)=(w)\ell(w)=\ell(w^{\prime}). However, as discussed before, while 1\forall^{1} can quantify over all points in a team, it cannot quantify over pairs.

As a workaround, we consider not only a tableau TαT_{\alpha}, but also a second tableau that acts as a copy of TαT_{\alpha}. Formally, for grids Tα,TβT_{\alpha},T_{\beta}, let TαTβT_{\alpha}\approx T_{\beta} denote that for all pairs (w,w)Tα×Tβ(w,w^{\prime})\in T_{\alpha}\times T_{\beta} it holds that (w)=(w)\ell(w)=\ell(w^{\prime}) implies c(w)=c(w)c(w)=c(w^{\prime}).

As \approx is symmetric and transitive, TαTβT_{\alpha}\approx T_{\beta} in fact implies both TαTαT_{\alpha}\approx T_{\alpha} and TβTβT_{\beta}\approx T_{\beta}, and hence that both TαT_{\alpha} and TβT_{\beta} are tableaus such that CTα=CTβC_{T_{\alpha}}=C_{T_{\beta}}, where CTα,CTβ:{1,,N}2Γ(Q×Γ)C_{T_{\alpha}},C_{T_{\beta}}\colon{\{1,\ldots,N\}}^{2}\to\Gamma\cup(Q\times\Gamma) are the induced runs as discussed on p. 8.1.

ψtableau(α):=\displaystyle\psi_{\text{tableau}}(\alpha)\vcentcolon=\; ψgrid(α)γ0ψgrid(γ0)ψ(α,γ0)\displaystyle\psi_{\text{grid}}(\alpha)\land\,\exists^{\subseteq}_{\gamma_{0}}\,\psi_{\text{grid}}(\gamma_{0})\land\psi_{\approx}(\alpha,\gamma_{0})
ψ(α,β):=\displaystyle\psi_{\approx}(\alpha,\beta)\vcentcolon= α1β1((ψ𝔱(α,β)ψ𝔭(α,β))\scalerel∨⃝eΞ((αβ)e))\displaystyle\forall^{1}_{\alpha}\forall^{1}_{\beta}\;\Big{(}\big{(}\psi^{\mathfrak{t}}_{\equiv}(\alpha,\beta)\land\psi^{\mathfrak{p}}_{\equiv}(\alpha,\beta)\big{)}\rightarrowtriangle\operatorname*{\scalerel*{\ovee}{\sum}}_{e\in\Xi}((\alpha\lor\beta)\hookrightarrow e)\Big{)}

In the following claim (and in the subsequent ones), we use the scopes γ0,γ1,γ2,\gamma_{0},\gamma_{1},\gamma_{2},\ldots as “auxiliary pre-tableaus”. Later, we will also use them as domains to quantify extra locations or rows from. (The index of γi\gamma_{i} is incremented whenever necessary to avoid quantifying from the same scope twice.) For this reason, from now on we always assume, for sufficiently large ii, that TγiT_{\gamma_{i}} is a pre-tableau. This can be later enforced in the reduction with ψpre-tableau(γi)\psi_{\text{pre-tableau}}(\gamma_{i}).

Claim 37 (e).
  1. (1)

    Tψtableau(α)T\vDash\psi_{\text{tableau}}(\alpha) if and only if TαT_{\alpha} is a tableau.

  2. (2)

    For grids Tα,TβT_{\alpha},T_{\beta}, it holds Tψ(α,β)T\vDash\psi_{\approx}(\alpha,\beta) if and only if TαTβT_{\alpha}\approx T_{\beta}.

Proof 8.5 (Proof of claim.).

(2) follows straightforwardly from Claim (a). Let us consider (1). As ψtableau\psi_{\text{tableau}} implies ψgrid\psi_{\text{grid}}, and by Claim (c), we can assume that TαT_{\alpha} is a grid.

Suppose that the formula is true. Then there exists STγ0S\subseteq T_{\gamma_{0}} such that TSαψgrid(γ0)T^{\alpha}_{S}\vDash\psi_{\text{grid}}(\gamma_{0}). By Claim (c), then SS is a grid as well. Moreover, TαST_{\alpha}\approx S by (2). As argued above, this implies that TαT_{\alpha} (and SS) is a tableau.

For the other direction, suppose that TαT_{\alpha} is a tableau. Then it defines a function CTαC_{T_{\alpha}}. Since Tγ0T_{\gamma_{0}} is a pre-tableau, we can pick a subteam SS of it that contains for each (i,j){1,,N}2(i,j)\in{\{1,\ldots,N\}}^{2} exactly those worlds ww with (w)=(i,j)\ell(w)=(i,j) such that c(w)=CTα(i,j)c(w)=C_{T_{\alpha}}(i,j). Then TαST_{\alpha}\approx S, and ψtableau\psi_{\text{tableau}} is true, with the quantifier γ0\exists^{\subseteq}_{\gamma_{0}} witnessed by SS. \triangleleft

8.4. From tableaus to runs

To ascertain that a tableau contains a run of MM, we have to check whether each row indeed is a configuration of MM—in other words, exactly one cell of each row contains a pair (q,a)Q×Γ(q,a)\in Q\times\Gamma—and whether consecutive configurations obey the transition relation δ\delta of MM.

For this, in the spirit of Cook’s theorem [Coo71] it suffices to consider all legal windows in the grid, i.e.\xspace, cells that are adjacent as follows, where e1,,e6Γ(Q×Γ)e_{1},\ldots,e_{6}\in\Gamma\cup(Q\times\Gamma):

e1e_{1} e2e_{2} e3e_{3}
e4e_{4} e5e_{5} e6e_{6}

If, say, (q,a,q,a,R)δ(q,a,q^{\prime},a^{\prime},R)\in\deltaMM switches to state qq^{\prime} from qq, replacing aa on the tape by aa^{\prime}, and moves to the right—then the windows obtained by setting e1=e4=be_{1}=e_{4}=b, e2=(q,a)e_{2}=(q,a), e5=ae_{5}=a^{\prime}, e3=be_{3}=b^{\prime}, e6=(q,b)e_{6}=(q^{\prime},b^{\prime}) are legal for all b,bΓb,b^{\prime}\in\Gamma. Using this scheme, δ\delta is completely represented by a constant finite set 𝗐𝗂𝗇Ξ6\mathsf{win}\subseteq\Xi^{6} of tuples (e1,,e6)(e_{1},\ldots,e_{6}) that represent the allowed windows in a run of MM.

Let us next explain how adjacency of cells is expressed. Suppose that two points wTαw\in T_{\alpha} and vTβv\in T_{\beta} are given. That vv is the immediate (𝔱\mathfrak{t}- or 𝔭\mathfrak{p}-)successor of ww then means that no element of the order exists between them. Simultaneously, ww and vv have to agree on the other component of their location, which is expressed by the first conjunct below. If 𝔮{𝔱,𝔭}\mathfrak{q}\in\{\mathfrak{t},\mathfrak{p}\} and 𝔮¯{𝔱,𝔭}{𝔮}\overline{\mathfrak{q}}\in\{\mathfrak{t},\mathfrak{p}\}\setminus\{\mathfrak{q}\}, we define:

ψsucc𝔮(α,β):=\displaystyle\psi^{\mathfrak{q}}_{\text{succ}}(\alpha,\beta)\vcentcolon=\; ψ𝔮¯(α,β)ψ𝔮(α,β)γ01(ψ𝔮(α,γ0)ψ𝔮(γ0,β))\displaystyle\psi^{\overline{\mathfrak{q}}}_{\equiv}(\alpha,\beta)\land\psi^{\mathfrak{q}}_{\prec}(\alpha,\beta)\land{\sim}\exists^{1}_{\gamma_{0}}\left(\psi^{\mathfrak{q}}_{\prec}(\alpha,\gamma_{0})\land\psi^{\mathfrak{q}}_{\prec}(\gamma_{0},\beta)\right)
Claim 38 (f).

If wTαw\in T_{\alpha} and vTβv\in T_{\beta}, then:

Tw,vα,βψsucc𝔱(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\psi^{\mathfrak{t}}_{\text{succ}}(\alpha,\beta) i,j{1,,N}:(w)=(i,j) and (v)=(i+1,j)\displaystyle\Leftrightarrow\exists i,j\in\{1,\ldots,N\}\colon\ell(w)=(i,j)\text{ and }\ell(v)=(i+1,j)
Tw,vα,βψsucc𝔭(α,β)\displaystyle T^{\alpha,\beta}_{w,v}\vDash\psi^{\mathfrak{p}}_{\text{succ}}(\alpha,\beta) i,j{1,,N}:(w)=(i,j) and (v)=(i,j+1)\displaystyle\Leftrightarrow\exists i,j\in\{1,\ldots,N\}\colon\ell(w)=(i,j)\text{ and }\ell(v)=(i,j+1)
Proof 8.6 (Proof of claim.).

Let us consider only 𝔮=𝔱\mathfrak{q}=\mathfrak{t}, as the case 𝔮=𝔭\mathfrak{q}=\mathfrak{p} is proven analogously. Assume that the formula ψsucc𝔱(α,β)\psi^{\mathfrak{t}}_{\text{succ}}(\alpha,\beta) is true in Tw,vα,βT^{\alpha,\beta}_{w,v}. By Claim (a), ψ𝔭\psi^{\mathfrak{p}}_{\equiv} holds if and only if there is a unique jj such that (w)=(i,j)\ell(w)=(i,j) and (v)=(i,j)\ell(v)=(i^{\prime},j), for some i,ii,i^{\prime}; in other words, if ww and vv agree on their 𝔭\mathfrak{p}-component.

Next, consider the sets Δw:=(Rw)𝔱k1\Delta_{w}\vcentcolon=\llbracket{}{(Rw)}_{\mathfrak{t}}\rrbracket{}_{k-1} and Δv:=(Rv)𝔱k1\Delta_{v}\vcentcolon=\llbracket{}{(Rv)}_{\mathfrak{t}}\rrbracket{}_{k-1} which correspond to the 𝔱\mathfrak{t}-components of (w)\ell(w) and (v)\ell(v). Suppose that Δw\Delta_{w} is the ii-th element of k1\prec^{*}_{k-1}. By ψ𝔱\psi^{\mathfrak{t}}_{\prec} and Claim (a), then clearly Δv\Delta_{v} is the ii^{\prime}-th element for some i>ii^{\prime}>i.

Suppose for the sake of contradiction that also i>i+1i^{\prime}>i+1, and let then instead ΔΔk1\Delta^{\prime}\subseteq\Delta_{k-1} be the (i+1)(i+1)-th element of k1\prec^{*}_{k-1}. As Tγ0T_{\gamma_{0}} is a pre-tableau, it contains a world zz such that (z)=(i+1,j)\ell(z)=(i+1,j). But then ψ𝔮(α,γ0)ψ𝔮(γ0,β)\psi^{\mathfrak{q}}_{\prec}(\alpha,\gamma_{0})\land\psi^{\mathfrak{q}}_{\prec}(\gamma_{0},\beta) is true in Tw,v,zα,β,γ0T^{\alpha,\beta,\gamma_{0}}_{w,v,z}, contradiction to ψsucc𝔱\psi^{\mathfrak{t}}_{\text{succ}}. Consequently, i=i+1i^{\prime}=i+1. The direction from right to left is shown similarly. \triangleleft

To check all windows in the tableau TαT_{\alpha}, we need to simultaneously quantify elements from six tableaus Tγ1,,Tγ6T_{\gamma_{1}},\ldots,T_{\gamma_{6}} that are copies of TαT_{\alpha}. For this purpose, we define

γiαφ:=\displaystyle\exists^{\approx\alpha}_{\gamma_{i}}\,\varphi\vcentcolon=\; γiψgrid(γi)ψ(α,γi)φ.\displaystyle\exists^{\subseteq}_{\gamma_{i}}\;\psi_{\text{grid}}(\gamma_{i})\land\psi_{\approx}(\alpha,\gamma_{i})\land\varphi\text{.}

Intuitively, under the premise that TγiT_{\gamma_{i}} is a pre-tableau and TαT_{\alpha} is a tableau, it “copies” the tableau TαT_{\alpha} into TγiT_{\gamma_{i}} by shrinking TγiT_{\gamma_{i}} accordingly. This is proven analogously to Claim (e). The next formula states that the picked points are arranged as in the picture:

ψwindow(γ1,,γ6):=\displaystyle\psi_{\text{window}}(\gamma_{1},\ldots,\gamma_{6})\vcentcolon=\; ψsucc𝔱(γ1,γ4)ψsucc𝔱(γ2,γ5)ψsucc𝔱(γ3,γ6)\displaystyle\psi^{\mathfrak{t}}_{\text{succ}}(\gamma_{1},\gamma_{4})\land\psi^{\mathfrak{t}}_{\text{succ}}(\gamma_{2},\gamma_{5})\land\psi^{\mathfrak{t}}_{\text{succ}}(\gamma_{3},\gamma_{6})\;\land
ψsucc𝔭(γ1,γ2)ψsucc𝔭(γ2,γ3)\displaystyle\psi^{\mathfrak{p}}_{\text{succ}}(\gamma_{1},\gamma_{2})\land\psi^{\mathfrak{p}}_{\text{succ}}(\gamma_{2},\gamma_{3})
Tγ1T_{\gamma_{1}} Tγ2T_{\gamma_{2}} Tγ3T_{\gamma_{3}}
Tγ4T_{\gamma_{4}} Tγ5T_{\gamma_{5}} Tγ6T_{\gamma_{6}}

The formula defining legal tableaus follows.

ψlegal(α)\displaystyle\psi_{\text{legal}}(\alpha) :=ψtableau(α)γ1αγ6αϑ1ϑ2ϑ3\displaystyle\vcentcolon=\psi_{\text{tableau}}(\alpha)\land\exists^{\approx\alpha}_{\gamma_{1}}\cdots\exists^{\approx\alpha}_{\gamma_{6}}\;\vartheta_{1}\land\vartheta_{2}\land\vartheta_{3}
We check that at most cell per row contains a state of MM:
ϑ1\displaystyle\vartheta_{1} :=γ11γ21(ψ𝔱(γ1,γ2)ψ𝔭(γ1,γ2))\displaystyle\vcentcolon=\forall^{1}_{\gamma_{1}}\forall^{1}_{\gamma_{2}}\Big{(}\psi^{\mathfrak{t}}_{\equiv}(\gamma_{1},\gamma_{2})\land\psi^{\mathfrak{p}}_{\prec}(\gamma_{1},\gamma_{2})\big{)}\rightarrowtriangle
(q1,a1),(q2,a2)Q×Γ((γ1c1(q1,a1))(γ2c1(q2,a2)))\displaystyle\bigwedge_{\mathclap{(q_{1},a_{1}),(q_{2},a_{2})\in Q\times\Gamma}}{\sim}\big{(}(\gamma_{1}\hookrightarrow c^{-1}(q_{1},a_{1}))\land(\gamma_{2}\hookrightarrow c^{-1}(q_{2},a_{2})\big{)}\Big{)}
We also check that every row contains some state. For this, γ11\forall^{1}_{\gamma_{1}} fixes some row and γ21ψ𝔱(γ1,γ2)\exists^{1}_{\gamma_{2}}\psi^{\mathfrak{t}}_{\equiv}(\gamma_{1},\gamma_{2}) searches that particular row for a state:
ϑ2\displaystyle\vartheta_{2} :=γ11γ21ψ𝔱(γ1,γ2)\scalerel∨⃝(q,a)Q×Γ(γ2c1(q,a))\displaystyle\vcentcolon=\forall^{1}_{\gamma_{1}}\exists^{1}_{\gamma_{2}}\;\psi^{\mathfrak{t}}_{\equiv}(\gamma_{1},\gamma_{2})\land\operatorname*{\scalerel*{\ovee}{\sum}}_{\mathclap{(q,a)\in Q\times\Gamma}}\;(\gamma_{2}\hookrightarrow c^{-1}(q,a))
Finally, every window must obey the transition relation:
ϑ3\displaystyle\vartheta_{3} :=γ11γ61(ψwindow(γ1,,γ6)\scalerel∨⃝(e1,,e6)wini=16(γiei))\displaystyle\vcentcolon=\forall^{1}_{\gamma_{1}}\cdots\forall^{1}_{\gamma_{6}}\;\Big{(}\psi_{\text{window}}(\gamma_{1},\ldots,\gamma_{6})\rightarrowtriangle\operatorname*{\scalerel*{\ovee}{\sum}}_{(e_{1},\ldots,e_{6})\in\mathrm{win}}\;\bigwedge_{i=1}^{6}(\gamma_{i}\hookrightarrow e_{i})\Big{)}
Claim 39 (g).

Tψlegal(α)T\vDash\psi_{\text{legal}}(\alpha) iff TαT_{\alpha} is a legal tableau, i.e.\xspace, iff CTαC_{T_{\alpha}} exists and is a run of MM.

Proof 8.7 (Proof of claim.).

Suppose that the formula holds. We show that TαT_{\alpha} is a legal tableau; the other direction is proven similarly.

Due to Claim (e), there are tableaus S1Tγ1S_{1}\subseteq T_{\gamma_{1}}, …, S6Tγ6S_{6}\subseteq T_{\gamma_{6}} that are copies of TαT_{\alpha} such that ϑ1ϑ2ϑ2\vartheta_{1}\land\vartheta_{2}\land\vartheta_{2} holds in TS1,,S6γ1,,γ6T^{\gamma_{1},\ldots,\gamma_{6}}_{S_{1},\ldots,S_{6}}.

Due to Claim (a), the subformula ϑ1\vartheta_{1} ensures the following: For all wS1,wS2w\in S_{1},w^{\prime}\in S_{2}, (w)=(i,j)\ell(w)=(i,j), (w)=(i,j)\ell(w^{\prime})=(i^{\prime},j^{\prime}), if i=ii=i^{\prime} and j<jj<j^{\prime} hold, then it is not the case that both c(w)=(q,a)c(w)=(q,a) and c(w)=(q,a)c(w^{\prime})=(q^{\prime},a^{\prime}) for any state symbols q,qQq,q^{\prime}\in Q. Since CS1=CS2=CTαC_{S_{1}}=C_{S_{2}}=C_{T_{\alpha}}, this is precisely the case if each row of CTαC_{T_{\alpha}} contains at most one state symbol.

Conversely, again by Claim (a), the subformula ϑ2\vartheta_{2} states that for every cell wS1w\in S_{1} there is another cell wS2w^{\prime}\in S_{2} in the same row that carries a state symbol: in other words, every row of CTαC_{T_{\alpha}} contains at least one state symbol.

Finally, ϑ3\vartheta_{3} relies on Claim (f) and states for every choice of singletons w1,,w6w_{1},\ldots,w_{6} in S1,,S6S_{1},\ldots,S_{6}, assuming that they are arranged as a window, that there exists a tuple (e1,,e6)win(e_{1},\ldots,e_{6})\in\mathrm{win} such that wiSiw_{i}\in S_{i} satisfies c(wi)=eic(w_{i})=e_{i}. As we showed that CTαC_{T_{\alpha}} contains in each row a configuration of MM, this implies that CTαC_{T_{\alpha}} exists and is a run of MM. \triangleleft

8.5. From runs to a computation

To encode the initial configuration on input x=x1xnx=x_{1}\cdots x_{n} in a tableau, we access the first nn cells of the first row and assign the respective letter of xx, as well as the initial state, to the first cell. Moreover, we assign \flat to all other cells in that row. For each 𝔮{𝔱,𝔭}\mathfrak{q}\in\{\mathfrak{t},\mathfrak{p}\}, we can check whether the location of a point in TαT_{\alpha} is minimal in its 𝔮\mathfrak{q}-component:

ψmin𝔮(α):=\displaystyle\psi^{\mathfrak{q}}_{\text{min}}(\alpha)\vcentcolon=\; γ01ψ𝔮(γ0,α)\displaystyle{\sim}\exists^{1}_{\gamma_{0}}\psi^{\mathfrak{q}}_{\prec}(\gamma_{0},\alpha)
This enables us to fix the first row of the configuration:
ψinput(α):=\displaystyle\psi_{\text{input}}(\alpha)\vcentcolon=\; γ1αγn+1αγ11γn1ψmin𝔱(γ1)ψmin𝔭(γ1)(γ1c1(q0,x1))\displaystyle\exists^{\approx\alpha}_{\gamma_{1}}\cdots\exists^{\approx\alpha}_{\gamma_{n+1}}\;\exists^{1}_{\gamma_{1}}\cdots\exists^{1}_{\gamma_{n}}\,\psi^{\mathfrak{t}}_{\text{min}}(\gamma_{1})\land\psi^{\mathfrak{p}}_{\text{min}}(\gamma_{1})\land\big{(}\gamma_{1}\hookrightarrow c^{-1}(q_{0},x_{1})\big{)}
i=2nψsucc𝔭(γi1,γi)(γic1(xi))\displaystyle\quad\bigwedge_{i=2}^{n}\psi^{\mathfrak{p}}_{\text{succ}}(\gamma_{i-1},\gamma_{i})\;\land\big{(}\gamma_{i}\hookrightarrow c^{-1}(x_{i})\big{)}
γn+11((ψ𝔱(γn,γn+1)ψ𝔭(γn,γn+1))(γn+1c1()))\displaystyle\qquad\land\forall^{1}_{\gamma_{n+1}}\Big{(}\big{(}\psi^{\mathfrak{t}}_{\equiv}(\gamma_{n},\gamma_{n+1})\land\psi^{\mathfrak{p}}_{\prec}(\gamma_{n},\gamma_{n+1})\big{)}\rightarrowtriangle\big{(}\gamma_{n+1}\hookrightarrow c^{-1}(\flat)\big{)}\Big{)}
Claim 40 (h).

Let TαT_{\alpha} be a tableau. Then Tψinput(α)T\vDash\psi_{\text{input}}(\alpha) if and only if

  1. (1)

    CTα(1,1)=(q0,x1)C_{T_{\alpha}}(1,1)=(q_{0},x_{1}),

  2. (2)

    CTα(1,i)=xiC_{T_{\alpha}}(1,i)=x_{i} for 2in2\leq i\leq n,

  3. (3)

    CTα(1,i)=C_{T_{\alpha}}(1,i)=\flat for n<iNn<i\leq N.

Proof 8.8 (Proof of claim.).

Suppose that the formula holds. After processing the quantifiers γ1αγn+1α\exists^{\approx\alpha}_{\gamma_{1}}\cdots\exists^{\approx\alpha}_{\gamma_{n+1}}, for all m{1,,n+1}m\in\{1,\ldots,n+1\} the team TγmT_{\gamma_{m}} is a tableau such that CTγm=CTαC_{T_{\gamma_{m}}}=C_{T_{\alpha}}. (Obviously this requires these teams to be pre-tableaus beforehand.) For this reason, we can freely replace CTα(i,j)C_{T_{\alpha}}(i,j) with CTγm(i,j)C_{T_{\gamma_{m}}}(i,j) when proving the properties (1)–(3).

In the second line of the formula, we make sure that c(w)=(q0,x1)c(w)=(q_{0},x_{1}) holds for least one point wCTγ1w\in C_{T_{\gamma_{1}}} of location (w)=(1,1)\ell(w)=(1,1). That (w)=(1,1)\ell(w)=(1,1) holds follows from Claim (a), ψmin𝔮\psi^{\mathfrak{q}}_{\text{min}}, and the assumption that Tγ0T_{\gamma_{0}} is a pre-tableau (which it still is after processing γ1αγn+1α\exists^{\approx\alpha}_{\gamma_{1}}\cdots\exists^{\approx\alpha}_{\gamma_{n+1}}). In particular, CTγ1(1,1)=(q0,x1)C_{T_{\gamma_{1}}}(1,1)=(q_{0},x_{1}).

The third line works similarly: for 2in2\leq i\leq n, it assigns xix_{i} to CTγi(1,i)C_{T_{\gamma_{i}}}(1,i) and hence to CTα(1,i)C_{T_{\alpha}}(1,i). Note that ψsucc𝔭\psi^{\mathfrak{p}}_{\text{succ}} also preserves the position in 𝔭\mathfrak{p}-direction”, i.e.\xspace, it is not necessary to repeat it for every cell of the first row. Finally, the last two lines state that every other location (1,j)(1,j^{\prime}) with j>nj^{\prime}>n contains \flat. The other direction is again similar. \triangleleft

Until now, we ignored the fact that MM (polynomially often) alternates. To simulate this, we alternatingly quantify polynomially many tableaus, each containing a part of the computation of MM. Each of these tableaus possesses a tail configuration, which is the configuration where MM either accepts, rejects, or alternates. Formally, a number i{1,,N}i\in\{1,\ldots,N\} is a tail index of CC if there exists jj such that either

  1. (1)

    C(i,j)C(i,j) has an accepting or rejecting state,

  2. (2)

    or C(i,j)C(i,j) has an existential state and and there are i<ii^{\prime}<i and jj^{\prime} with a universal state in C(i,j)C(i^{\prime},j^{\prime}),

  3. (3)

    or C(i,j)C(i,j) has a universal state and there are i<ii^{\prime}<i and jj^{\prime} with an existential state in C(i,j)C(i^{\prime},j^{\prime}).

The least such ii is called first tail index, and the corresponding configuration is the first tail configuration. The idea is that we can split the computation of MM into multiple tableaus if any tableau (except the initial one) contains a run that continues from the previous tableau’s first tail configuration.

We formalize the above as follows. Assume that TαT_{\alpha} is a tableau, and that TβT_{\beta} marks a single row ii by being a singleton {w}\{w\} with (w)=(i,j)\ell(w)=(i,j) for some jj. Then the formula ψtail(α,β)\psi_{\text{tail}}(\alpha,\beta) below will be true if and only if the ii-th row of CTαC_{T_{\alpha}} is a tail configuration. With

Q-𝗌𝗍𝖺𝗍𝖾(β)\displaystyle Q^{\prime}\text{-}\mathsf{state}(\beta) :=\scalerel∨⃝(q,a)Q×Γ(βc1(q,a)),\displaystyle\vcentcolon=\operatorname*{\scalerel*{\ovee}{\sum}}_{\mathclap{(q,a)\in Q^{\prime}\times\Gamma}}(\beta\hookrightarrow c^{-1}(q,a))\text{,}

we check if a given singleton Tβ={w}T_{\beta}=\{w\} encodes an accepting, rejecting, existential, universal, or any state by setting QQ^{\prime} to QaccQ_{\mathrm{acc}}, QrejQ_{\mathrm{rej}}, QQ_{\exists}, QQ_{\forall} or QQ, respectively. We define ψtail\psi_{\text{tail}}:

ψtail(α,β)\displaystyle\psi_{\text{tail}}(\alpha,\beta)\; :=γ0αα1ψ𝔱(α,β)Q-𝗌𝗍𝖺𝗍𝖾(α)[Qacc-𝗌𝗍𝖺𝗍𝖾(α)∨⃝Qrej-𝗌𝗍𝖺𝗍𝖾(α)∨⃝\displaystyle\vcentcolon=\exists^{\approx\alpha}_{\gamma_{0}}\,\exists^{1}_{\alpha}\psi^{\mathfrak{t}}_{\equiv}(\alpha,\beta)\land Q\text{-}\mathsf{state}(\alpha)\land\Big{[}Q_{\mathrm{acc}}\text{-}\mathsf{state}(\alpha)\ovee Q_{\mathrm{rej}}\text{-}\mathsf{state}(\alpha)\;\ovee
γ01\displaystyle\exists^{1}_{\gamma_{0}} (ψ𝔱(γ0,α)(Q-𝗌𝗍𝖺𝗍𝖾(α)Q-𝗌𝗍𝖺𝗍𝖾(γ0))∨⃝(Q-𝗌𝗍𝖺𝗍𝖾(α)Q-𝗌𝗍𝖺𝗍𝖾(γ0)))]\displaystyle\Big{(}\psi^{\mathfrak{t}}_{\prec}(\gamma_{0},\alpha)\land\big{(}Q_{\exists}\text{-}\mathsf{state}(\alpha)\land Q_{\forall}\text{-}\mathsf{state}(\gamma_{0}))\ovee(Q_{\forall}\text{-}\mathsf{state}(\alpha)\land Q_{\exists}\text{-}\mathsf{state}(\gamma_{0})\big{)}\Big{)}\Big{]}
ψfirst-tail(α,β)\displaystyle\psi_{\text{first-tail}}(\alpha,\beta)\; :=ψtail(α,β)γ11(ψ𝔱(γ1,β)ψtail(α,γ1))\displaystyle\vcentcolon=\psi_{\text{tail}}(\alpha,\beta)\land{\sim}\exists^{1}_{\gamma_{1}}\Big{(}\psi^{\mathfrak{t}}_{\prec}(\gamma_{1},\beta)\land\psi_{\text{tail}}(\alpha,\gamma_{1})\Big{)}
Claim 41 (i).

Suppose that TαT_{\alpha} is a tableau, Tβ={w}T_{\beta}=\{w\}, and (w)=(i,j)\ell(w)=(i,j). Then Tψtail(α,β)T\vDash\psi_{\text{tail}}(\alpha,\beta) if and only if ii is a tail index of CTαC_{T_{\alpha}}. Moreover, Tψfirst-tail(α,β)T\vDash\psi_{\text{first-tail}}(\alpha,\beta) if and only if ii is the first tail index of CTαC_{T_{\alpha}}.

Proof 8.9 (Proof of claim.).

Since Tγ1T_{\gamma_{1}} is a pre-tableau and hence contains all locations in rows i<ii^{\prime}<i, it is easy to see that the proof for ψfirst-tail\psi_{\text{first-tail}} boils down to that of ψtail\psi_{\text{tail}}. Consequently, let us consider ψtail\psi_{\text{tail}}.

First, due to γ0α\exists^{\approx\alpha}_{\gamma_{0}}, we can assume that Tγ0T_{\gamma_{0}} is a tableau that is a copy of TαT_{\alpha}, i.e.\xspace, CTα=CTγ0C_{T_{\alpha}}=C_{T_{\gamma_{0}}}. Here, it is required for the inner quantification in the definition of a tail index.

The first line of the formula reduces TαT_{\alpha} to a singleton that is (due to ψ𝔱\psi^{\mathfrak{t}}_{\equiv}) in row ii. Furthermore, it carries a state qq of MM due to Q-𝗌𝗍𝖺𝗍𝖾(α)Q\text{-}\mathsf{state}(\alpha). The further examination of this state will determine if ii is a tail index. Now, qq is exactly one of accepting, rejecting, existential, or universal. If qQaccQrejq\in Q_{\mathrm{acc}}\cup Q_{\mathrm{rej}}, then ii is a tail index by definition.

Otherwise we quantify over the states qq^{\prime} of all (copies of) earlier rows in TαT_{\alpha}, using γ01ψ𝔱(γ0,α)\exists^{1}_{\gamma_{0}}\psi^{\mathfrak{t}}_{\prec}(\gamma_{0},\alpha), and search for a universal state if qq is existential and vice versa, which as well, if it exists, proves by definition that ii is a tail index. \triangleleft

Formally, given a run CC of MM that has a tail configuration, CC accepts if the state qq in its first tail configuration is in QaccQ_{\mathrm{acc}}, CC rejects if that qq is in QrejQ_{\mathrm{rej}}, and CC alternates otherwise. That a run of the form CTαC_{T_{\alpha}} accepts or rejects is expressed by

ψacc(α)\displaystyle\psi_{\text{acc}}(\alpha) :=γ2αγ21Qacc-𝗌𝗍𝖺𝗍𝖾(γ2)ψfirst-tail(α,γ2),\displaystyle\vcentcolon=\exists^{\approx\alpha}_{\gamma_{2}}\;\exists^{1}_{\gamma_{2}}\;Q_{\mathrm{acc}}\text{-}\mathsf{state}(\gamma_{2})\land\psi_{\text{first-tail}}(\alpha,\gamma_{2})\text{,}
ψrej(α)\displaystyle\psi_{\text{rej}}(\alpha) :=γ2αγ21Qrej-𝗌𝗍𝖺𝗍𝖾(γ2)ψfirst-tail(α,γ2).\displaystyle\vcentcolon=\exists^{\approx\alpha}_{\gamma_{2}}\;\exists^{1}_{\gamma_{2}}\;Q_{\mathrm{rej}}\text{-}\mathsf{state}(\gamma_{2})\land\psi_{\text{first-tail}}(\alpha,\gamma_{2})\text{.}

In this formula, first the tableau TαT_{\alpha} is copied to Tγ2T_{\gamma_{2}} to extract with γ21\exists^{1}_{\gamma_{2}} the world carrying an accepting/rejecting state, while ψfirst-tail(α,γ2)\psi_{\text{first-tail}}(\alpha,\gamma_{2}) ensures that no alternation or rejecting/accepting state occurs at some earlier point in CTαC_{T_{\alpha}}.

If the first tail configuration of the run contains an alternation, and if the run was existentially quantified, then it should be continued in a universally quantified tableau, and vice versa. The following formula expresses, given two tableaus Tα,TβT_{\alpha},T_{\beta}, that CTβC_{T_{\beta}} is a continuation of CTαC_{T_{\alpha}}, i.e.\xspace, that the first configuration of CTβC_{T_{\beta}} equals the first tail configuration of CTαC_{T_{\alpha}}. In other words, if ii is the first tail index of CTαC_{T_{\alpha}}, then CTα(i,j)=CTβ(1,j)C_{T_{\alpha}}(i,j)=C_{T_{\beta}}(1,j) for all j{1,,N}j\in\{1,\ldots,N\}.

ψcont(α,β):=\displaystyle\psi_{\text{cont}}(\alpha,\beta)\vcentcolon=\; γ21ψfirst-tail(α,γ2)α1β1\displaystyle\exists^{1}_{\gamma_{2}}\,\psi_{\text{first-tail}}(\alpha,\gamma_{2})\land\forall^{1}_{\alpha}\forall^{1}_{\beta}
[(ψmin𝔱(β)ψ𝔱(α,γ2)ψ𝔭(α,β))(\scalerel∨⃝eΞ(αβ)e)]\displaystyle\quad\Big{[}\Big{(}\psi^{\mathfrak{t}}_{\text{min}}(\beta)\land\psi^{\mathfrak{t}}_{\equiv}(\alpha,\gamma_{2})\land\psi^{\mathfrak{p}}_{\equiv}(\alpha,\beta)\Big{)}\rightarrowtriangle\Big{(}\operatorname*{\scalerel*{\ovee}{\sum}}_{e\in\Xi}(\alpha\lor\beta)\hookrightarrow e\Big{)}\Big{]}

The above formula first obtains the first tail index ii of CTαC_{T_{\alpha}} and stores it in a singleton yTγ2y\in T_{\gamma_{2}}. Then for all worlds wTαw\in T_{\alpha} and vTβv\in T_{\beta}, where vv is 𝔱\mathfrak{t}-minimal (i.e.\xspace, in the first row) and ww is in the same row as yy, and which additionally agree on their 𝔭\mathfrak{p}-component, the third line states that ww and vv agree on Ξ\Xi. Altogether, the ii-th row of CTαC_{T_{\alpha}} and the first row of CTβC_{T_{\beta}} then have to coincide.

MM performs at most r(n)1r(n)-1 alternations for some polynomial rr. Then we require r=r(n)r=r(n) tableaus, which we call α1,,αr\alpha_{1},\ldots,\alpha_{r}. In the following, the formula ψrun,i\psi_{\text{run},i} describes the behaviour of the ii-th run, i.e.\xspace, the part of the computation after i1i-1 alternations. W.l.o.g.\xspacerr is even and q0Qq_{0}\in Q_{\exists}. We may then define the final run by

ψrun,r\displaystyle\psi_{\text{run},r} :=αr[(ψlegal(αr)ψcont(αr1,αr))(ψrej(αr)ψacc(αr))].\displaystyle\vcentcolon=\forall^{\subseteq}_{\alpha_{r}}\Big{[}\Big{(}\psi_{\text{legal}}(\alpha_{r})\land\psi_{\text{cont}}(\alpha_{r-1},\alpha_{r})\Big{)}\rightarrowtriangle\Big{(}{\sim}\psi_{\text{rej}}(\alpha_{r})\land\psi_{\text{acc}}(\alpha_{r})\Big{)}\Big{]}\text{.}

For 1<i<r1<i<r and even ii, let

ψrun,i:=\displaystyle\psi_{\text{run},i}\vcentcolon=\; αi[(ψlegal(αi)ψcont(αi1,αi))(ψrej(αi)(ψacc(αi)∨⃝ψrun,i+1))]\displaystyle\forall^{\subseteq}_{\alpha_{i}}\;\Big{[}\Big{(}\psi_{\text{legal}}(\alpha_{i})\land\psi_{\text{cont}}(\alpha_{i-1},\alpha_{i})\Big{)}\rightarrowtriangle\Big{(}{\sim}\psi_{\text{rej}}(\alpha_{i})\land\big{(}\psi_{\text{acc}}(\alpha_{i})\ovee\psi_{\text{run},{i+1}}\big{)}\Big{)}\Big{]}
and for 1<i<r1<i<r and odd ii
ψrun,i:=\displaystyle\psi_{\text{run},i}\vcentcolon=\; αi[ψlegal(αi)ψcont(αi1,αi)ψrej(αi)(ψacc(αi)∨⃝ψrun,i+1)].\displaystyle\exists^{\subseteq}_{\alpha_{i}}\Big{[}\psi_{\text{legal}}(\alpha_{i})\land\psi_{\text{cont}}(\alpha_{i-1},\alpha_{i})\land{\sim}\psi_{\text{rej}}(\alpha_{i})\land\Big{(}\psi_{\text{acc}}(\alpha_{i})\ovee\psi_{\text{run},{i+1}}\Big{)}\Big{]}\text{.}

Analogously, the initial run is described by

ψrun,1:=\displaystyle\psi_{\text{run},1}\vcentcolon= α1(ψlegal(α1)ψinput(α1)ψrej(α1)(ψacc(α1)∨⃝ψrun,2))\displaystyle\;\exists^{\subseteq}_{\alpha_{1}}\Big{(}\psi_{\text{legal}}(\alpha_{1})\land\psi_{\text{input}}(\alpha_{1})\land{\sim}\psi_{\text{rej}}(\alpha_{1})\land\Big{(}\psi_{\text{acc}}(\alpha_{1})\ovee\psi_{\text{run},{2}}\Big{)}\Big{)}

We are now in the position to state the full reduction. Let us gather all relevant scopes in the set Ψ𝒫𝒮\Psi\subseteq\mathcal{PS}:

Ψ:=\displaystyle\Psi\vcentcolon=\; {𝔰i0ik}{𝔰k}{γi0in+1}{αi1ir}\displaystyle\{\mathfrak{s}_{i}\mid 0\leq i\leq k\}\cup\{\mathfrak{s}^{\prime}_{k}\}\;\cup\;\{\gamma_{i}\mid 0\leq i\leq n+1\}\;\cup\;\{\alpha_{i}\mid 1\leq i\leq r\}

The scopes that accommodate pre-tableaus are

Ψ:=\displaystyle\Psi^{\prime}\vcentcolon=\; {γi0in+1}{αi1ir}.\displaystyle\{\gamma_{i}\mid 0\leq i\leq n+1\}\cup\{\alpha_{i}\mid 1\leq i\leq r\}\text{.}

W.l.o.g.\xspacen5n\geq 5, as γ1,,γ6\gamma_{1},\ldots,\gamma_{6} are always required in the construction. The reduction now maps xx to

φx:=\displaystyle\varphi_{x}\vcentcolon=\; 𝖼𝖺𝗇𝗈𝗇𝗌𝖼𝗈𝗉𝖾𝗌k(Ψ)pΨψpre-tableau(p)ψrun,1.\displaystyle\mathsf{canon}^{\prime}\land\mathsf{scopes}_{k}(\Psi)\land\bigwedge_{p\in\Psi^{\prime}}\psi_{\text{pre-tableau}}(p)\land\psi_{\text{run,1}}\text{.}

It is easy to see that this formula is an 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formula that is logspace-constructible from xx and kk, where kk itself is either constant or a polynomial in |x|{|\nobreak x\nobreak|} and hence logspace-computable. By Lemma 21, φx\varphi_{x} is satisfiable if and only if φxk+1\varphi_{x}\land\Box^{k+1}\bot is satisfiable. For this reason, we conclude the reduction with the following proof.

Proof 8.10 (Proof of Lemma 32).

It remains to argue that φxk+1\varphi_{x}\land\Box^{k+1}\bot is satisfiable if and only if MM accepts xx. For the sake of simplicity, assume r=2r=2. The cases r>2r>2 are proven analogously.

\Rightarrow: Suppose (𝒦,T)φxk+1(\mathcal{K},T)\vDash\varphi_{x}\land\Box^{k+1}\bot. Similarly as in Theorem 22, the pΨp\in\Psi are disjoint scopes due to 𝗌𝖼𝗈𝗉𝖾𝗌k(Ψ)\mathsf{scopes}_{k}(\Psi). Moreover, by 𝖼𝖺𝗇𝗈𝗇\mathsf{canon}^{\prime} and Claim (b), (𝒦,T)(\mathcal{K},T) is then a kk-staircase in which T𝔰kT_{\mathfrak{s}_{k}} and T𝔰kT_{\mathfrak{s}^{\prime}_{k}} both are kk-canonical teams. Due to Claim (d) and the large conjunction in φx\varphi_{x}, Tα1,Tα2,Tγ1,,Tγn+1T_{\alpha_{1}},T_{\alpha_{2}},T_{\gamma_{1}},\ldots,T_{\gamma_{n+1}} are then pre-tableaus.

As the formula ψrun,1\psi_{\text{run,1}} holds, by Claim (g) and (h), Tα1T_{\alpha_{1}} has a subteam S1S_{1} that is a legal tableau and starts with MM’s initial configuration on xx. In particular, CS1C_{S_{1}} exists. Moreover, either ψacc\psi_{\text{acc}} holds (i.e.\xspace, CS1C_{S_{1}} and hence MM is accepting) or ψrun,2\psi_{\text{run,2}} holds (i.e.\xspace, if CS1C_{S_{1}} alternates). Consider the latter case. Then for all legal tableaus S2Tα2S_{2}\subseteq T_{\alpha_{2}} such that CS2C_{S_{2}} is a continuation of CS1C_{S_{1}} it holds that CS2C_{S_{2}} is accepting. However, as Tα2T_{\alpha_{2}} is a pre-tableau, every run is of the form CS2C_{S_{2}} for some S2Tα2S_{2}\subseteq T_{\alpha_{2}}. Consequently, MM accepts xx.

\Leftarrow: Suppose MM accepts xx. First of all, due to Claim (b), the formula 𝖼𝖺𝗇𝗈𝗇𝗌𝖼𝗈𝗉𝖾𝗌k({𝔰0,,𝔰k,𝔰k})k+1\mathsf{canon}^{\prime}\land\mathsf{scopes}_{k}(\{\mathfrak{s}_{0},\ldots,\mathfrak{s}_{k},\mathfrak{s}^{\prime}_{k}\})\land\Box^{k+1}\bot has a model (𝒦,T)(\mathcal{K},T). Moreover, we can freely add a pre-tableau TpT_{p} for each pΨp\in\Psi to satisfy the large conjunction in φx\varphi_{x}. By labeling the propositions in Ψ\Psi correctly (as disjoint scopes), we ensure that 𝗌𝖼𝗈𝗉𝖾𝗌k(Ψ)\mathsf{scopes}_{k}(\Psi) holds as well.

It remains to demonstrate Tψrun,1T\vDash\psi_{\text{run,1}}. As MM accepts xx, there exists a run C1C_{1} starting from MM’s initial configuration such that either C1C_{1} accepts, or, for all runs C2C_{2} continuing C1C_{1}, C2C_{2} accepts.

Since Tα1T_{\alpha_{1}} is a pre-tableau, it also contains a subteam S1S_{1} such that S1S_{1} is a legal tableau and CS1=C1C_{S_{1}}=C_{1}. We choose S1S_{1} as witness for α1\exists^{\subseteq}_{\alpha_{1}}. If C1C_{1} itself accepts, then ψacc(α1)\psi_{\text{acc}}(\alpha_{1}) and hence ψrun,1\psi_{\text{run,1}} is satisfied. Otherwise we consider ψrun,2\psi_{\text{run,2}}. Suppose that S2Tα2S_{2}\subseteq T_{\alpha_{2}} is picked as a subteam by α2\forall^{\subseteq}_{\alpha_{2}}. If it forms a legal tableau and CS2C_{S_{2}} is a continuation of C1C_{1}, then C2C_{2} must be accepting since MM accepts xx by assumption. But this implies that ψacc(α2)\psi_{\text{acc}}(\alpha_{2}) is true for any such S2S_{2}. Consequently, ψrun,2\psi_{\text{run,2}} and hence ψrun,1\psi_{\text{run,1}} is true.

9. Hardness under strict semantics and on restricted frame classes

9.1. Lax and strict semantics

In this section, we further generalize the hardness result of the previous section.

Team-semantical connectives can be evaluated either in so-called standard or lax semantics, or alternatively in strict semantics. In Section 3, we defined 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace with lax semantics. In strict semantics, the connectives \lor and \Diamond are replaced by their counterparts s\lor_{s} and s\Diamond_{s}:

(𝒦,T)ψsθ\displaystyle(\mathcal{K},T)\vDash\psi\lor_{s}\theta S,UT such that T=SUSU=(𝒦,S)ψ, and (𝒦,U)θ,\displaystyle\Leftrightarrow\;\exists S,U\subseteq T\text{ such that }T=S\cup U\text{, }S\cap U=\emptyset\text{, }(\mathcal{K},S)\vDash\psi\text{, and }(\mathcal{K},U)\vDash\theta\text{,}
(𝒦,T)sψ\displaystyle(\mathcal{K},T)\vDash\Diamond_{s}\psi (𝒦,S)ψ for some strict successor team S of T,\displaystyle\Leftrightarrow\;(\mathcal{K},S)\vDash\psi\text{ for some strict successor team }S\text{ of }T\text{,}

where a strict successor team of TT is a successor team SRTS\subseteq RT for which there exists a surjective f:TSf\colon T\to S satisfying f(w)Rwf(w)\in Rw for all wTw\in T. Intuitively, in the lax disjunction the teams of the splitting may overlap, while in the strict disjunction they are disjoint. Likewise, a lax successor team may contain multiple successor of any wTw\in T, while in a strict successor team we pick exactly one successor for each wTw\in T.

An 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace-formula φ\varphi is downward closed if (𝒦,T)φ(\mathcal{K},T)\vDash\varphi implies (𝒦,S)φ(\mathcal{K},S)\vDash\varphi for all STS\subseteq T. For example, every 𝖬𝖫\xspace\mathsf{ML}\xspace-formula is downward closed, as is the constancy atom =(α)=α∨⃝¬α{=\!\!(\alpha)}=\alpha\ovee\neg\alpha or generally any monotone Boolean combination of 𝖬𝖫\xspace\mathsf{ML}\xspace-formulas. On such formulas, strict and lax semantics are equivalent:

Proposition 42.

Let φ,ψ𝖬𝖳𝖫\xspace\varphi,\psi\in\mathsf{MTL}\xspace such that φ\varphi is downward closed. Then φψφsψ\varphi\lor\psi\equiv\varphi\lor_{s}\psi and φsφ\Diamond\varphi\equiv\Diamond_{s}\varphi.

Proof 9.1.

Clearly φsψ\varphi\lor_{s}\psi entails φψ\varphi\lor\psi and sφ\Diamond_{s}\varphi entails φ\Diamond\varphi. If conversely TφψT\vDash\varphi\lor\psi via subteams S,UTS,U\subseteq T such that SU=TS\cup U=T, SφS\vDash\varphi and UψU\vDash\psi, then we instead split TT into the subteams UU and TUT\setminus U. Since TUST\setminus U\subseteq S and φ\varphi is downward closed, this proves TφsψT\vDash\varphi\lor_{s}\psi.

Likewise, suppose TφT\vDash\Diamond\varphi via some successor team SS of TT. Assuming the axiom of choice, there is some function f:TSf\colon T\to S such that f(w)Rwf(w)\in Rw for each wTw\in T. The team {f(w)wT}S\{f(w)\mid w\in T\}\subseteq S is now a strict successor team of TT and satisfies φ\varphi due to downward closure.

Due to Proposition 42, the distinction between strict and lax semantics was traditionally unnecessary for many team logics such as the original dependence logic [Vää07, Vää08], as it has only downward closed formulas. The distinction between strict and lax semantics was first made in the context of first-order team logic by Galliani [Gal12]. It has some interesting consequences, for instance first-order inclusion logic in strict semantics is as expressive as existential second-order logic [GHK13] (see also Hannula and Kontinen [HK15]).

With modal team logic, strict semantics was studied, e.g.\xspace, by Hella et al. [HS15, HKMV15, HKMV17]. In the works that explicitly study strict semantics, the underlying (first-order or modal) team logic was enriched by not downward closed constructs such as the inclusion atom \subseteq or exclusion atom \mid, or the independence atom \perp.

In this article, where we consider team-wide negation {\sim} as part of the logic, the distinction between strict and lax semantics becomes apparent already for simple formulas such as 𝖤𝖤𝖤s𝖤\mathsf{E}\top\lor\mathsf{E}\top\not\equiv\mathsf{E}\top\lor_{s}\mathsf{E}\top, where the former defines non-emptiness, but the latter means that the team contains at least two points.

We prove that our hardness results also hold in strict semantics. Let the logics 𝖬𝖳𝖫\xspace(s,)\mathsf{MTL}\xspace(\lor_{s},\Box) and 𝖬𝖳𝖫\xspacek(s,)\mathsf{MTL}\xspace_{k}(\lor_{s},\Box) be defined like 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace and 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}, but with s\lor_{s} instead of \lor and without \Diamond and s\Diamond_{s} (i.e.\xspace, only using the modality \Box).

Theorem 43.

𝖲𝖠𝖳()\mathsf{SAT}(\mathcal{L}) and 𝖵𝖠𝖫()\mathsf{VAL}(\mathcal{L}) are hard for TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}) if =𝖬𝖳𝖫\xspace(s,)\mathcal{L}=\mathsf{MTL}\xspace(\lor_{s},\Box), and hard for ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}) if =𝖬𝖳𝖫\xspacek(s,)\mathcal{L}=\mathsf{MTL}\xspace_{k}(\lor_{s},\Box) and k0k\geq 0.

Proof 9.2.

An analysis of the proof of Lemma 32 yields that the 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace-formula φx\varphi_{x} produced in the reduction can be easily adapted to strict semantics. First, observe that \Diamond occurs only in the subformula 𝗆𝖺𝗑i\mathsf{max}_{i}, which is by Proposition 42 equivalent to

s(¬ispΦ(¬ip∨⃝¬i¬p)),\top\lor_{s}\Bigl{(}\neg\Box^{i}\bot\land{\sim}\mathop{{\bigvee}_{\!\!s}}\limits_{\!p\in\Phi}(\neg\Box^{i}p\ovee\neg\Box^{i}\neg p)\Bigr{)}\text{,}

since ᬬα\Diamond\alpha\equiv\neg\Box\neg\alpha and ¬ip∨⃝¬i¬p\neg\Box^{i}p\ovee\neg\Box^{i}\neg p is a downward closed formula. A quick check reveals that all other instances of \lor in φx\varphi_{x} are subject to Proposition 42 as well, except of the occurrence in the second line of ζk\zeta^{*}_{k}. Here, the critical part of the correctness proof is the choice of the subteam UU^{\prime} in Claim (c) of Lemma 25. In strict semantics, the only possibility becomes U=U=OSU^{\prime}=U=O\setminus S, for which the proof works identically. Finally, for the case k=0k=0, a similar check of the proof for 𝖯𝖳𝖫\xspace\mathsf{PTL}\xspace [HKVV18, Theorem 4.9] reveals that there also every \lor can be replaced by s\lor_{s} due to Proposition 42.

Note that the corresponding upper bound via the construction of a canonical model (viz. Theorem 8) does not apply to strict semantics. The reason for this is the failure of Proposition 4: In strict semantics, 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formulas are not invariant under kk-team-bisimulation in general.

As an example, consider the formula φ:=𝖤s𝖤\varphi\vcentcolon=\mathsf{E}\top\lor_{s}\mathsf{E}\top. It states that the team contains at least two points. However, for every finite Φ𝒫𝒮\Phi\subseteq\mathcal{PS} and k0k\geq 0 it is easy to find a team TT of two points and a singleton SS that is (Φ,k)(\Phi,k)-bisimilar to it, while TφT\vDash\varphi and SφS\nvDash\varphi.

A possible approach could be to define a bisimulation relation that respects the multiplicity of types in a team, and to define a corresponding canonical model, but this is beyond the scope of this paper.

9.2. Restricted frame classes

A natural restriction in the context of modal logic is to focus on a specific subclass of Kripke frames, which is useful for instance for modeling belief or temporal systems. (For an introduction to frame classes, consider, e.g.\xspace, Fitting [Fit07].) Let F=(W,R)F=(W,R) denote a frame. Prominent frame classes include

𝖪\mathsf{K}:

all frames,

𝖣\mathsf{D}:

serial frames (wWRww\in W\Rightarrow Rw\neq\emptyset),

𝖳\mathsf{T}:

reflexive frames (wWwRww\in W\Rightarrow w\in Rw),

𝖪𝟦\mathsf{K4}:

transitive frames (uRv,vRw,wWuRwu\in Rv,v\in Rw,w\in W\Rightarrow u\in Rw),

𝖣𝟦\mathsf{D4}:

serial and transitive frames,

𝖲𝟦\mathsf{S4}:

reflexive and transitive frames.

In this section, we consider these classes from a complexity theoretic perspective, and show that the lower bounds of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace hold when restricted to these classes. Given a frame class \mathcal{F} and a fragment \mathcal{L} of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace, let 𝖲𝖠𝖳(,)\mathsf{SAT}(\mathcal{L},\mathcal{F}) denote the set of all \mathcal{L}-formulas that are satisfied in a model (W,R,V,T)(W,R,V,T) where (W,R)(W,R) is a frame in \mathcal{F}. Define 𝖵𝖠𝖫(,)\mathsf{VAL}(\mathcal{L},\mathcal{F}) analogously.

We prove the team-semantical analog of Ladner’s theorem, which states that classical modal satisfiability and validity are PSPACE\xspace\mathrm{PSPACE}\xspace-hard problem for any frame class between 𝖲𝟦\mathsf{S4} and 𝖪\mathsf{K} [Lad77, Theorem 3.1]. Note that this includes all the frame classes stated above.

Theorem 44.

Let \mathcal{F} be a frame class such that 𝖲𝟦𝖪\mathsf{S4}\subseteq\mathcal{F}\subseteq\mathsf{K}. Then 𝖲𝖠𝖳(𝖬𝖳𝖫\xspace,)\mathsf{SAT}(\mathsf{MTL}\xspace,\mathcal{F}) and 𝖵𝖠𝖫(𝖬𝖳𝖫\xspace,)\mathsf{VAL}(\mathsf{MTL}\xspace,\mathcal{F}) are hard for TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}), and 𝖲𝖠𝖳(𝖬𝖳𝖫\xspacek,)\mathsf{SAT}(\mathsf{MTL}\xspace_{k},\mathcal{F}) and 𝖵𝖠𝖫(𝖬𝖳𝖫\xspacek,)\mathsf{VAL}(\mathsf{MTL}\xspace_{k},\mathcal{F}) are hard for ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}), for k0k\geq 0.

Proof 9.3.

We give the proof for 𝖲𝖠𝖳(𝖬𝖳𝖫\xspacek)mlog𝖲𝖠𝖳(𝖬𝖳𝖫\xspacek,)\mathsf{SAT}(\mathsf{MTL}\xspace_{k})\leq^{\log}_{\mathrm{m}}\mathsf{SAT}(\mathsf{MTL}\xspace_{k},\mathcal{F}). Let φ𝖬𝖳𝖫\xspacek\varphi\in\mathsf{MTL}\xspace_{k}. The idea is to introduce new propositions 0,,k𝖯𝗋𝗈𝗉(φ)\ell_{0},\ldots,\ell_{k}\notin\mathsf{Prop}(\varphi) that mark the layers of different height in a structure, and to modify the formula such that all edges except between consecutive layers ii and i+1i+1 are ignored. (Here, we make the assumption that 𝒦\mathcal{K} is a acyclic, which relies on Corollary 18 and hence indirectly on Proposition 4).

Given a Φ{0,,k}\Phi\cup\{\ell_{0},\ldots,\ell_{k}\}-structure 𝒦=(W,R,V)\mathcal{K}=(W,R,V), let 𝒦:=(W,R,V)\mathcal{K}^{\circ}\vcentcolon=(W,R^{\circ},V) be the structure where only such edges are retained, i.e.\xspace,

R=Ri=0k1(V(i)×V(i+1)).R^{\circ}=R\cap\bigcup_{i=0}^{k-1}(V(\ell_{i})\times V(\ell_{i+1}))\text{.}

On the side of formulas, the reduction is φ0φ0\varphi\mapsto\ell_{0}\land\varphi^{0}, where φi\varphi^{i} is inductively as follows. The non-modal connectives are ignored, i.e.\xspace, pi:=pp^{i}\vcentcolon=p for pΦp\in\Phi, (ψθ)i:=ψiθi{(\psi\land\theta)}^{i}\vcentcolon=\psi^{i}\land\theta^{i}, (ψ)i:=ψi{({\sim}\psi)}^{i}\vcentcolon={\sim}\psi^{i}, (ψθ)i:=ψiθi{(\psi\lor\theta)}^{i}\vcentcolon=\psi^{i}\lor\theta^{i}. For the modalities, let (ψ)i:=(i+1ψi+1){(\Diamond\psi)}^{i}\vcentcolon=\Diamond(\ell_{i+1}\land\psi^{i+1}) and (ψi):=(i+1ψi+1)(\Box\psi^{i})\vcentcolon=\Box(\ell_{i+1}\hookrightarrow\psi^{i+1}). Intuitively, φi\varphi^{i} is meant to be evaluated in layer ii, and we make sure that successor teams always are contained in the next layer i+1i+1.

For the correctness of the reduction, we will first show the following claim.

Claim 45.

For all i{0,,k}i\in\{0,\ldots,k\} and TV(i)T\subseteq V(\ell_{i}), it holds that (𝒦,T)φi(\mathcal{K},T)\vDash\varphi^{i} iff (𝒦,T)φ(\mathcal{K}^{\circ},T)\vDash\varphi.

Proof 9.4 (Proof of claim.).

This is proved by a straightforward induction on the formula size:

  • Atomic propositions are clear. The Boolean connectives and splitting follow straightforwardly from the induction hypothesis (as subteams of TT are again in V(i)V(\ell_{i})).

  • Let φ=ψ\varphi=\Diamond\psi. Suppose (𝒦,T)φi(\mathcal{K},T)\vDash\varphi^{i}, i.e.\xspace, (𝒦,S)i+1ψi+1(\mathcal{K},S)\vDash\ell_{i+1}\land\psi^{i+1} for some RR-successor team SS of TT. Then by induction hypothesis (𝒦,S)ψ(\mathcal{K}^{\circ},S)\vDash\psi, as SV(i+1)S\subseteq V(\ell_{i+1}). SS is an RR^{\circ}-successor team of TT as well, since (w,v)R(w,v)R(w,v)\in R\Leftrightarrow(w,v)\in R^{\circ} for every (w,v)V(i)×V(i+1)(w,v)\in V(\ell_{i})\times V(\ell_{i+1}). This proves (𝒦,T)φ(\mathcal{K}^{\circ},T)\vDash\varphi.

    Conversely, if (𝒦,T)φ(\mathcal{K}^{\circ},T)\vDash\varphi, then (𝒦,S)ψ(\mathcal{K}^{\circ},S)\vDash\psi for some RR^{\circ}-successor team SS of TT. However, any RR^{\circ}-successor team of TT is a subset of V(i+1)V(\ell_{i+1}). As a consequence, (𝒦,S)i+1(\mathcal{K},S)\vDash\ell_{i+1}. Moreover, by induction hypothesis, (𝒦,S)ψi+1(\mathcal{K},S)\vDash\psi^{i+1}. This yields (𝒦,T)φi(\mathcal{K},T)\vDash\varphi^{i}, since SS is trivially also a RR-successor team of TT.

  • Let φ=ψ\varphi=\Box\psi. Then (𝒦,T)φi(\mathcal{K},T)\vDash\varphi^{i} iff (𝒦,RT)(i+1ψi+1)(\mathcal{K},RT)\vDash(\ell_{i+1}\hookrightarrow\psi^{i+1}) iff (𝒦,RTV(i+1))ψi+1(\mathcal{K},RT\cap V(\ell_{i+1}))\vDash\psi^{i+1} iff (𝒦,RTV(i+1)ψ(\mathcal{K}^{\circ},RT\cap V(\ell_{i+1})\vDash\psi by induction hypothesis. It remains to show that RT=RTV(i+1)R^{\circ}T=RT\cap V(\ell_{i+1}). Clearly, RTRTR^{\circ}T\subseteq RT and RTV(i+1)R^{\circ}T\subseteq V(\ell_{i+1}), since RRR^{\circ}\subseteq R, RV(i)V(i+1)R^{\circ}V(\ell_{i})\subseteq V(\ell_{i+1}), and TV(i)T\subseteq V(\ell_{i}). Conversely, if wRTV(i+1)w\in RT\cap V(\ell_{i+1}), then (v,w)R(v,w)\in R for some vTv\in T. As (v,w)V(i)×V(i+1)(v,w)\in V(\ell_{i})\times V(\ell_{i+1}), then (v,w)R(v,w)\in R^{\circ}, hence wRTw\in R^{\circ}T. \triangleleft

Now, due to the above claim, if 0φ0\ell_{0}\land\varphi^{0} is satisfiable, then clearly φ\varphi is as well. It remains to show that 0φ0\ell_{0}\land\varphi^{0} has a reflexive and transitive model if φ\varphi is satisfiable. Suppose that the latter is satisfied in a Φ\Phi-structure (𝒦,T)(\mathcal{K},T). By Corollary 18, we may assume that (𝒦,T)(\mathcal{K},T) is a forest of height kk with the set of roots being TT. Then we label the new propositions i\ell_{i} such that V(i)=RiTV(\ell_{i})=R^{i}T, i.e.\xspace, V(0)=TV(\ell_{0})=T, V(1)=RTV(\ell_{1})=RT and so on. As 𝒦\mathcal{K} is a forest, note that the sets T,RT,R2T,T,RT,R^{2}T,\ldots are pairwise disjoint. In other words, every world in 𝒦\mathcal{K} has a unique distance 0ik0\leq i\leq k from TT and hence exactly one i\ell_{i} labeled. This is required for the next part of the proof.

Let now RR^{*} be the reflexive transitive closure of RR. It remains to show (R)=R{(R^{*})}^{\circ}=R, since then we can again apply the previously proved claim and are done. It is easy to see that R(R)R\subseteq{(R^{*})}^{\circ}, since for every (w,v)R(w,v)\in R there is some ii such that wRiT=V(i)w\in R^{i}T=V(\ell_{i}), consequently (w,v)RiT×Ri+1T=V(i)×V(i+1)(w,v)\in R^{i}T\times R^{i+1}T=V(\ell_{i})\times V(\ell_{i+1}). For the other direction, suppose (w,v)(R)(w,v)\in{(R^{*})}^{\circ}. By definition of (R){(R^{*})}^{\circ}, there is ii such that wV(i)w\in V(\ell_{i}), vV(i+1)v\in V(\ell_{i+1}), and vv is reachable from ww by some RR-path (u0,,un)(u_{0},\ldots,u_{n}) where w=u0w=u_{0} and v=unv=u_{n}. But since u0RiTu_{0}\in R^{i}T, for all mm it holds umRi+mT=V(i+m)u_{m}\in R^{i+m}T=V(\ell_{i+m}). As V(i+n)V(i+1)=V(\ell_{i+n})\cap V(\ell_{i+1})=\emptyset for n1n\neq 1, we conclude n=1n=1, so (w,v)R(w,v)\in R.

10. Conclusion

Theorem 31 settles the complexity of 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace and proves that its satisfiability and validity problems are complete for the non-elementary complexity class TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly}). Moreover, the fragments 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} are proved complete for ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly}), the levels of the elementary hierarchy with polynomially many alternations.

In our approach, we developed a notion of (kk-)canonical models for modal logics with team semantics. We showed that such models exist for 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace and 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}, and that logspace-computable 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}-formulas exist that are satisfiable, but only have kk-canonical models.

Our lower bounds carry over to two-variable first-order team logic 𝖥𝖮\xspace2()\mathsf{FO}\xspace^{2}({\sim}) and its fragment 𝖥𝖮\xspacek2()\mathsf{FO}\xspace^{2}_{k}({\sim}) of bounded quantifier rank kk as well [Lüc18c]. While the former is TOWER\xspace(poly)\mathrm{TOWER}\xspace(\mathrm{poly})-complete, the latter is ATIME\xspace-ALT\xspace(expk+1,poly)\mathrm{ATIME}\xspace\text{-}\mathrm{ALT}\xspace(\exp_{k+1},\mathrm{poly})-hard. However, no matching upper bound for the satisfiability problem of 𝖥𝖮\xspacek2()\mathsf{FO}\xspace^{2}_{k}({\sim}) exists.

In the final section, we considered variants of the satisfiability problem for 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace. We showed that it is as hard as the original problem when 𝖬𝖳𝖫\xspace\mathsf{MTL}\xspace is interpreted in strict semantics, and in fact for \Diamond-free formulas with \lor being interpreted either lax or strict. Also, any restriction of the satisfiability problem to a frame class that includes at least the reflexive-transitive frames is as hard as the original problem.

In future research, it could be useful to further generalize the concept of canonical models to other logics with team semantics. Do logics such as 𝖥𝖮\xspacek2()\mathsf{FO}\xspace^{2}_{k}({\sim}) permit a canonical model in the spirit of kk-canonical models for 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k}, and does this yield a tight upper bound on the complexity of their satisfiability problem? How do 𝖬𝖳𝖫\xspacek\mathsf{MTL}\xspace_{k} and 𝖥𝖮\xspacek2()\mathsf{FO}\xspace^{2}_{k}({\sim}) differ in terms of succinctness?

Other obvious open questions are the upper bounds for Theorem 43 and 44, and also the combination of the above aspects, e.g.\xspace, does the lower bound still hold in strict semantics on reflexive-transitive frames? To solve these issues, the model theory of modal team logic has to be refined. For example, what is the analog of Proposition 4 for strict semantics?

Acknowledgements

The author wishes to thank Heribert Vollmer, Irena Schindler and Arne Meier, as well as the anonymous referee, for numerous helpful comments and suggestions.

References

  • [AKVV16] Samson Abramsky, Juha Kontinen, Jouko Väänänen, and Heribert Vollmer, editors. Dependence Logic, Theory and Applications. Springer, 2016.
  • [BRV01] Patrick Blackburn, Maarten de Rijke, and Yde Venema. Modal Logic. Cambridge University Press, New York, NY, USA, 2001.
  • [CH90] Kevin J. Compton and C. Ward Henson. A Uniform Method for Proving Lower Bounds on the Computational Complexity of Logical Theories. Ann. Pure Appl. Logic, 48(1):1–79, 1990.
  • [CH96] M. J. Cresswell and G. E. Hughes. A New Introduction to Modal Logic. Routledge, 1996.
  • [CKS81] Ashok K. Chandra, Dexter C. Kozen, and Larry J. Stockmeyer. Alternation. J. ACM, 28(1):114–133, January 1981.
  • [Coo71] Stephen A. Cook. The Complexity of Theorem-Proving Procedures. In Proc. of the 3rd Annual ACM Sym. on Theory of Computing, pages 151–158, 1971.
  • [Cre83] M. J. Cresswell. KM and the finite model property. Notre Dame Journal of Formal Logic, 24(3):323–327, 1983.
  • [DKV16] Arnaud Durand, Juha Kontinen, and Heribert Vollmer. Expressivity and Complexity of Dependence Logic. In Dependence Logic, Theory and Applications, pages 5–32. Springer, 2016.
  • [EHM+13] Johannes Ebbing, Lauri Hella, Arne Meier, Julian-Steffen Müller, Jonni Virtema, and Heribert Vollmer. Extended modal dependence logic. In WoLLIC, volume 8071 of Lecture Notes in Computer Science, pages 126–137. Springer, 2013.
  • [Fin75] Kit Fine. Normal forms in modal logic. Notre Dame Journal of Formal Logic, 16(2):229–237, 1975.
  • [Fit07] Melvin Fitting. 2 Modal proof theory. In Johan Van Benthem Patrick Blackburn and Frank Wolter, editors, Handbook of Modal Logic, volume 3 of Studies in Logic and Practical Reasoning, pages 85 – 138. Elsevier, 2007.
  • [Gal12] Pietro Galliani. Inclusion and exclusion dependencies in team semantics - On some logics of imperfect information. Ann. Pure Appl. Logic, 163(1):68–84, 2012.
  • [Gal15] Pietro Galliani. Upwards closed dependencies in team semantics. Information and Computation, 245:124–135, December 2015.
  • [Gal16] Pietro Galliani. On strongly first-order dependencies. In Dependence Logic, pages 53–71. Springer, 2016.
  • [Gal18] Pietro Galliani. Safe Dependency Atoms and Possibility Operators in Team Semantics. In GandALF, volume 277 of EPTCS, pages 58–72, 2018.
  • [GHK13] Pietro Galliani, Miika Hannula, and Juha Kontinen. Hierarchies in independence logic. In CSL, volume 23 of LIPIcs, pages 263–280. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2013.
  • [GO07] Valentin Goranko and Martin Otto. 5 Model theory of modal logic. In Johan Van Benthem Patrick Blackburn and Frank Wolter, editors, Handbook of Modal Logic, volume 3 of Studies in Logic and Practical Reasoning, pages 249 – 329. Elsevier, 2007.
  • [GV13] Erich Grädel and Jouko A. Väänänen. Dependence and Independence. Studia Logica, 101(2):399–410, 2013.
  • [Han16] Miika Hannula. Validity and Entailment in Modal and Propositional Dependence Logics. CoRR, abs/1608.04301, 2016.
  • [Han17] Miika Hannula. Validity and Entailment in Modal and Propositional Dependence Logics. In CSL, volume 82 of LIPIcs, pages 28:1–28:17. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2017.
  • [HK15] Miika Hannula and Juha Kontinen. Hierarchies in independence and inclusion logic with strict semantics. J. Log. Comput., 25(3):879–897, 2015.
  • [HKLV16] Miika Hannula, Juha Kontinen, Martin Lück, and Jonni Virtema. On Quantified Propositional Logics and the Exponential Time Hierarchy. In GandALF, volume 226 of EPTCS, pages 198–212, 2016.
  • [HKMV15] Lauri Hella, Antti Kuusisto, Arne Meier, and Heribert Vollmer. Modal Inclusion Logic: Being Lax is Simpler than Being Strict. In MFCS, volume 9234 of LNCS, pages 281–292. Springer, 2015.
  • [HKMV17] Lauri Hella, Antti Kuusisto, Arne Meier, and Jonni Virtema. Model Checking and Validity in Propositional and Modal Inclusion Logics. In MFCS, volume 83 of LIPIcs, pages 32:1–32:14. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2017.
  • [HKVV15] Miika Hannula, Juha Kontinen, Jonni Virtema, and Heribert Vollmer. Complexity of Propositional Independence and Inclusion Logic. In MFCS, volume 9234 of LNCS, pages 269–280. Springer, 2015.
  • [HKVV18] Miika Hannula, Juha Kontinen, Jonni Virtema, and Heribert Vollmer. Complexity of Propositional Logics in Team Semantic. ACM Trans. Comput. Log., 19(1):2:1–2:14, 2018.
  • [HLSV14] Lauri Hella, Kerkko Luosto, Katsuhiko Sano, and Jonni Virtema. The Expressive Power of Modal Dependence Logic. In Advances in Modal Logic, pages 294–312. College Publications, 2014.
  • [Hod97] Wilfrid Hodges. Compositional Semantics for a Language of Imperfect Information. Logic Journal of the IGPL, 5(4):539–563, 1997.
  • [HS89] Jaakko Hintikka and Gabriel Sandu. Informational Independence as a Semantical Phenomenon. In Jens Erik Fenstad, Ivan T. Frolov, and Risto Hilpinen, editors, Logic, Methodology and Philosophy of Science VIII, volume 126 of Studies in Logic and the Foundations of Mathematics, pages 571 – 589. Elsevier, 1989.
  • [HS15] Lauri Hella and Johanna Stumpf. The expressive power of modal logic with inclusion atoms. In GandALF, volume 193 of EPTCS, pages 129–143, 2015.
  • [HV16] Lauri Hella and Miikka Vilander. The succinctness of first-order logic over modal logic via a formula size game. In Advances in Modal Logic, pages 401–419. College Publications, 2016.
  • [KMSV15] Juha Kontinen, Julian-Steffen Müller, Henning Schnoor, and Heribert Vollmer. A Van Benthem Theorem for Modal Team Semantics. In CSL, volume 41 of LIPIcs, pages 277–291. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2015.
  • [KMSV17] Juha Kontinen, Julian-Steffen Müller, Henning Schnoor, and Heribert Vollmer. Modal independence logic. J. Log. Comput., 27(5):1333–1352, 2017.
  • [KMV15] Andreas Krebs, Arne Meier, and Jonni Virtema. A Team Based Variant of CTL. In TIME, pages 140–149. IEEE Computer Society, 2015.
  • [KMVZ18] Andreas Krebs, Arne Meier, Jonni Virtema, and Martin Zimmermann. Team semantics for the specification and verification of hyperproperties. In MFCS, volume 117 of LIPIcs, pages 10:1–10:16. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2018.
  • [Lad77] Richard E. Ladner. The computational complexity of provability in systems of modal propositional logic. SIAM J. Comput., 6(3):467–480, 1977.
  • [Lüc17] Martin Lück. The Power of the Filtration Technique for Modal Logics with Team Semantics. In CSL, volume 82 of LIPIcs, pages 31:1–31:20. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2017.
  • [Lüc18a] Martin Lück. Axiomatizations of team logics. Ann. Pure Appl. Logic, 169(9):928–969, 2018.
  • [Lüc18b] Martin Lück. Canonical Models and the Complexity of Modal Team Logic. In CSL, volume 119 of LIPIcs, pages 30:1–30:23. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2018.
  • [Lüc18c] Martin Lück. On the Complexity of Team Logic and Its Two-Variable Fragment. In MFCS, volume 117 of LIPIcs, pages 27:1–27:22. Schloss Dagstuhl - Leibniz-Zentrum fuer Informatik, 2018.
  • [LV13] Peter Lohmann and Heribert Vollmer. Complexity Results for Modal Dependence Logic. Studia Logica, 101(2):343–366, 2013.
  • [Mey74] Albert R Meyer. The inherent computational complexity of theories of ordered sets. In Proc. Int. Congress of Mathematicians, volume 19722, page 481, 1974.
  • [Mos07] Lawrence S. Moss. Finite Models Constructed from Canonical Formulars. J. Philosophical Logic, 36(6):605–640, 2007.
  • [Mül14] Julian-Steffen Müller. Satisfiability and model checking in team based logics. PhD thesis, University of Hanover, 2014.
  • [MV97] Maarten Marx and Yde Venema. Multi-dimensional modal logic. In Multi-Dimensional Modal Logic, pages 1–9. Springer, 1997.
  • [Sch16] Sylvain Schmitz. Complexity Hierarchies beyond Elementary. TOCT, 8(1):3:1–3:36, 2016.
  • [Sev09] Merlijn Sevenster. Model-theoretic and Computational Properties of Modal Dependence Logic. J. Log. Comput., 19(6):1157–1173, 2009.
  • [SM73] Larry J. Stockmeyer and Albert R. Meyer. Word Problems Requiring Exponential Time: Preliminary Report. In Proc. of the 5th Annual ACM Sym. on Theory of Computing, pages 1–9, 1973.
  • [Vää07] Jouko Väänänen. Dependence logic: A New Approach to Independence Friendly Logic. Number 70 in London Mathematical Society student texts. Cambridge University Press, Cambridge ; New York, 2007.
  • [Vää08] Jouko Väänänen. Modal dependence logic. New perspectives on games and interaction, 4:237–254, 2008.
  • [Vir17] Jonni Virtema. Complexity of validity for propositional dependence logics. Inf. Comput., 253:224–236, 2017.
  • [Voi17] Marco Voigt. A fine-grained hierarchy of hard problems in the separated fragment. In LICS, pages 1–12. IEEE Computer Society, 2017.
  • [YV16] Fan Yang and Jouko Väänänen. Propositional Logics of Dependence. Ann. Pure Appl. Logic, 167(7):557–589, 2016.
  • [YV17] Fan Yang and Jouko Väänänen. Propositional team logics. Ann. Pure Appl. Logic, 168(7):1406–1441, 2017.

Appendix A Proof details

In the appendix, we include several propositions that have straightforward but lengthy proofs.

Proofs of Section 4

Proposition 6.

Let Φ𝒫𝒮\Phi\subseteq\mathcal{PS} be finite and k0k\geq 0.

  1. (1)

    wkΦΦ=V1(w)Φ\llbracket{}w\rrbracket{}^{\Phi}_{k}\cap\Phi=V^{-1}(w)\cap\Phi and Rw=kΦwk+1Φ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}=\mathcal{R}\llbracket{}w\rrbracket{}^{\Phi}_{k+1}, for all pointed structures (W,R,V,w)(W,R,V,w).

  2. (2)

    The mapping h:ττΦh\colon\tau\mapsto\tau\cap\Phi is a bijection from Δ0Φ\Delta^{\Phi}_{0} to 𝔓(Φ)\mathfrak{P}(\Phi).

  3. (3)

    The mapping h:τ(τΦ,τ)h\colon\tau\mapsto(\tau\cap\Phi,\mathcal{R}\tau) is a bijection from Δk+1Φ\Delta^{\Phi}_{k+1} to 𝔓(Φ)×𝔓(ΔkΦ)\mathfrak{P}(\Phi)\times\mathfrak{P}(\Delta^{\Phi}_{k}).

Proof A.1 ().
  • Proof of (1). Assume (W,R,V,v),Φ𝒫𝒮(W,R,V,v),\Phi\subseteq\mathcal{PS} and k0k\geq 0 as above. For all pΦp\in\Phi, clearly pwkΦp\in\llbracket{}w\rrbracket{}^{\Phi}_{k} iff wpw\vDash p iff pV1(w)p\in V^{-1}(w). Next, we show that Rw=kΦwk+1Φ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}=\mathcal{R}\llbracket{}w\rrbracket{}^{\Phi}_{k+1}. Let τ=wk+1Φ\tau=\llbracket{}w\rrbracket{}^{\Phi}_{k+1}, and recall that τ={τΔkΦ{αατ}τ}\mathcal{R}\tau=\{\tau^{\prime}\in\Delta^{\Phi}_{k}\mid\{\alpha\mid\Box\alpha\in\tau\}\subseteq\tau^{\prime}\}. To prove RwkΦτ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}\subseteq\mathcal{R}\tau, let τRwkΦ\tau^{\prime}\in\llbracket{}Rw\rrbracket{}^{\Phi}_{k} be arbitrary. Then v=kΦτ\llbracket{}v\rrbracket{}^{\Phi}_{k}=\tau^{\prime} for some vRwv\in Rw. Now, for all α𝖬𝖫\xspacekΦ\alpha\in\mathsf{ML}\xspace^{\Phi}_{k}, ατ\Box\alpha\in\tau implies wαw\vDash\Box\alpha. In particular, vαv\vDash\alpha, i.e.\xspace, ατ\alpha\in\tau^{\prime}. Hence, {αατ}τ\{\alpha\mid\Box\alpha\in\tau\}\subseteq\tau^{\prime}, which implies ττ\tau^{\prime}\in\mathcal{R}\tau.

    For the converse direction, τRwkΦ\mathcal{R}\tau\subseteq\llbracket{}Rw\rrbracket{}^{\Phi}_{k}, let ττ\tau^{\prime}\in\mathcal{R}\tau be arbitrary. By definition, {αατ}τ\{\alpha\mid\Box\alpha\in\tau\}\subseteq\tau^{\prime}. Since τ\tau^{\prime} is a kk-type, it has a model (𝒦,v)(\mathcal{K}^{\prime},v^{\prime}), and due to Proposition 5, 𝒦,v=kΦτ\llbracket{}\mathcal{K}^{\prime},v^{\prime}\rrbracket{}^{\Phi}_{k}=\tau^{\prime}. By Proposition 3, there is a formula ζ𝖬𝖫\xspacekΦ\zeta\in\mathsf{ML}\xspace^{\Phi}_{k} such that (𝒦′′,v′′)ζ(\mathcal{K}^{\prime\prime},v^{\prime\prime})\vDash\zeta if and only if (𝒦,v)kΦ(𝒦′′,v′′)(\mathcal{K}^{\prime},v^{\prime})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime\prime},v^{\prime\prime}). As τ\tau is a (k+1)(k+1)-type, either ζτ\Diamond\zeta\in\tau or ¬ζτ\neg\Diamond\zeta\in\tau.

    First, suppose ¬ζτ\neg\Diamond\zeta\in\tau. Then ¬ζτ\Box\neg\zeta\in\tau, hence ¬ζτ\neg\zeta\in\tau^{\prime} by definition of τ\tau^{\prime}. But as (𝒦,v)τ(\mathcal{K}^{\prime},v^{\prime})\vDash\tau^{\prime}, then both (𝒦,v)ζ(\mathcal{K}^{\prime},v^{\prime})\nvDash\zeta and (𝒦,v)ζ(\mathcal{K}^{\prime},v^{\prime})\vDash\zeta, as (𝒦,v)kΦ(𝒦,v)(\mathcal{K}^{\prime},v^{\prime})\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime},v^{\prime}). Contradiction, therefore ζτ\Diamond\zeta\in\tau. Consequently, ww has an RR-successor vv such that vζv\vDash\zeta, i.e.\xspace, τ=vkΦRwkΦ\tau^{\prime}=\llbracket{}v\rrbracket{}^{\Phi}_{k}\in\llbracket{}Rw\rrbracket{}^{\Phi}_{k}.

  • Proof that hh in (2) and (3) is injective. Let τ,τΔkΦ\tau,\tau^{\prime}\in\Delta^{\Phi}_{k} be arbitrary. Let (𝒦,w)=(W,R,V,w)(\mathcal{K},w)=(W,R,V,w) be of type τ\tau, and (𝒦,w)=(W,R,V,w)(\mathcal{K}^{\prime},w^{\prime})=(W^{\prime},R^{\prime},V^{\prime},w^{\prime}) of type τ\tau^{\prime}. We first consider (2) and demonstrate that h:ττΦh\colon\tau\mapsto\tau\cap\Phi injective. This follows from (1), as τΦ=τΦ\tau\cap\Phi=\tau^{\prime}\cap\Phi implies V1(w)=τΦ=τΦ=V1(w)V^{-1}(w)=\tau\cap\Phi=\tau^{\prime}\cap\Phi=V^{\prime-1}(w^{\prime}), i.e.\xspace, (𝒦,w)0Φ(𝒦,w)(\mathcal{K},w)\rightleftharpoons^{\Phi}_{0}(\mathcal{K}^{\prime},w^{\prime}). By Proposition 5, then τ=τ\tau=\tau^{\prime}.

    For (3), let k>0k>0, and additionally suppose τ=τ\mathcal{R}\tau=\mathcal{R}\tau^{\prime}. Again by (1), we have 𝒦,Rw=k1Φτ=τ=𝒦,Rwk1Φ\llbracket{}\mathcal{K},Rw\rrbracket{}^{\Phi}_{k-1}=\mathcal{R}\tau=\mathcal{R}\tau^{\prime}=\llbracket{}\mathcal{K}^{\prime},R^{\prime}w^{\prime}\rrbracket{}^{\Phi}_{k-1}. By Proposition 5, (𝒦,Rw)k1Φ(𝒦,Rw)(\mathcal{K},Rw)\rightleftharpoons^{\Phi}_{k-1}(\mathcal{K}^{\prime},R^{\prime}w^{\prime}) follows. Since (𝒦,w)0Φ(𝒦,w)(\mathcal{K},w)\rightleftharpoons^{\Phi}_{0}(\mathcal{K}^{\prime},w^{\prime}) holds as before, (𝒦,w)kΦ(𝒦,w)(\mathcal{K},w)\rightleftharpoons^{\Phi}_{k}(\mathcal{K}^{\prime},w^{\prime}) by Proposition 3. By Proposition 5, τ=𝒦,w=kΦ𝒦,w=kΦτ\tau=\llbracket{}\mathcal{K},w\rrbracket{}^{\Phi}_{k}=\llbracket{}\mathcal{K}^{\prime},w^{\prime}\rrbracket{}^{\Phi}_{k}=\tau^{\prime}.

  • Proof that hh in (2) and (3) is surjective. First, consider (2). We have to show that, for all ΦΦ\Phi^{\prime}\subseteq\Phi, there exists a type τΔ0Φ\tau\in\Delta^{\Phi}_{0} such that τΦ=Φ\tau\cap\Phi=\Phi^{\prime}. Likewise, for (3) we have to show that for all k0k\geq 0, ΦΦ\Phi^{\prime}\subseteq\Phi and ΔΔkΦ\Delta^{\prime}\subseteq\Delta^{\Phi}_{k}, there exists a type τΔk+1Φ\tau\in\Delta^{\Phi}_{k+1} such that τΦ=Φ\tau\cap\Phi=\Phi^{\prime} and τ=Δ\mathcal{R}\tau=\Delta^{\prime}. We show the second statement, as the first one is shown analogously. The following model (𝒦,w)(\mathcal{K},w) witnesses that there exists τΔk+1\tau\in\Delta_{k+1} such that τΦ=Φ\tau\cap\Phi=\Phi^{\prime} and τ=Δ\mathcal{R}\tau=\Delta^{\prime}. First, recall that each τΔ\tau^{\prime}\in\Delta^{\prime} has a model (𝒩τ,vτ)(\mathcal{N}_{\tau^{\prime}},v_{\tau^{\prime}}) such that, by Proposition 5, 𝒩τ,vτ=kΦτ\llbracket{}\mathcal{N}_{\tau^{\prime}},v_{\tau^{\prime}}\rrbracket{}^{\Phi}_{k}=\tau^{\prime}. Define 𝒦\mathcal{K} as the disjoint union of all 𝒩τ\mathcal{N}_{\tau} and of a distinct point ww, and let V1(w)=ΦV^{-1}(w)=\Phi^{\prime}. By (1), then wk+1ΦΦ=Φ\llbracket{}w\rrbracket{}^{\Phi}_{k+1}\cap\Phi=\Phi^{\prime}. Moreover, let Rw={vττΔ}Rw=\{v_{\tau^{\prime}}\mid\tau^{\prime}\in\Delta^{\prime}\}. Again due to (1), w=k+1ΦRwkΦ\mathcal{R}\llbracket{}w\rrbracket{}^{\Phi}_{k+1}=\llbracket{}Rw\rrbracket{}^{\Phi}_{k}. By definition, Rw=kΦ{vττΔ}=kΦ{vτkΦτΔ}=Δ\llbracket{}Rw\rrbracket{}^{\Phi}_{k}=\llbracket{}\{v_{\tau^{\prime}}\mid\tau^{\prime}\in\Delta^{\prime}\}\rrbracket{}^{\Phi}_{k}=\{\llbracket{}v_{\tau^{\prime}}\rrbracket{}^{\Phi}_{k}\mid\tau^{\prime}\in\Delta^{\prime}\}=\Delta^{\prime}.

Lemma 11.

For every polynomial pp there is a polynomial qq such that

p(expk(n))expk(q((k+1)n))p(\exp^{*}_{k}(n))\leq\exp_{k}(q((k+1)\cdot n))

for all k0k\geq 0 and n1n\geq 1.

We require the following inequalities.

Lemma 46.

Let n,k,c0n,k,c\geq 0. Then c+expk(n)expk(c+n)c+\exp_{k}(n)\leq\exp_{k}(c+n). If also n1n\geq 1, then cexpk(n)expk(cn)c\cdot\exp_{k}(n)\leq\exp_{k}(cn).

Proof A.2.

Induction on kk, where k=0k=0 is trivial. For k1k\geq 1,

c+expk+1(n)\displaystyle c+\exp_{k+1}(n) =c+2expk(n)2c2expk(n)\displaystyle=c+2^{\exp_{k}(n)}\leq 2^{c}\cdot 2^{\exp_{k}(n)} (As c+a2cac+a\leq 2^{c}\cdot a for c0,a1c\geq 0,a\geq 1)
=2c+expk(n)2expk(c+n)\displaystyle=2^{c+\exp_{k}(n)}\leq 2^{\exp_{k}(c+n)} (Induction hypothesis)
=expk+1(c+n).\displaystyle=\exp_{k+1}(c+n)\text{.}

For the product, the cases c=0,1c=0,1 are trivial. For c2c\geq 2,

cexpk+1(n)\displaystyle c\cdot\exp_{k+1}(n) 2c12expk(n)\displaystyle\leq 2^{c-1}\cdot 2^{\exp_{k}(n)} (Since c2c\geq 2 implies c2c1c\leq 2^{c-1})
=2c1+expk(n)2expk(c1+n)\displaystyle=2^{c-1+\exp_{k}(n)}\leq 2^{\exp_{k}(c-1+n)} (By ++ case)
2expk(cn)=expk+1(cn).\displaystyle\leq 2^{\exp_{k}(cn)}=\exp_{k+1}(cn)\text{.} (As (c1)+ncn(c-1)+n\leq cn for c,n1c,n\geq 1)

Recall that exp0(n):=n\exp^{*}_{0}(n)\vcentcolon=n and expk+1(n):=n2expk(n)\exp^{*}_{k+1}(n)\vcentcolon=n\cdot 2^{\exp^{*}_{k}(n)}.

Lemma 47.

Let n,k0n,k\geq 0. Then expk(n)expk((k+1)n)\exp_{k}^{*}(n)\leq\exp_{k}((k+1)\cdot n).

Proof A.3.

Induction on kk. For k=0k=0, exp0(n)=n=exp0((0+1)n)\exp_{0}^{*}(n)=n=\exp_{0}((0+1)\cdot n). For the inductive step,

expk+1(n)\displaystyle\exp^{*}_{k+1}(n) =n2expk(n)2n2expk(n)=2n+expk(n)\displaystyle=n\cdot 2^{\exp^{*}_{k}(n)}\leq 2^{n}\cdot 2^{\exp^{*}_{k}(n)}=2^{n+\exp^{*}_{k}(n)}
2n+expk((k+1)n)\displaystyle\leq 2^{n+\exp_{k}((k+1)n)} (Induction hypothesis)
2expk(n+(k+1)n)=expk+1((k+2)n)\displaystyle\leq 2^{\exp_{k}(n+(k+1)n)}=\exp_{k+1}((k+2)n) (Lemma 46)

The next inequality states that a polynomial can be “pulled inside” expk\exp_{k}:

Lemma 48.

For every polynomial pp there is a polynomial qq such that p(expk(n))expk(q(n)))p(\exp_{k}(n))\leq\exp_{k}(q(n))) for all k0,n1k\geq 0,n\geq 1.

Proof A.4.

For every polynomial pp there are integers c,d1c,d\geq 1 such that p(n)cndp(n)\leq cn^{d} for all n1n\geq 1. Let q(n):=cdnd+cq(n)\vcentcolon=cdn^{d}+c. Then the case k=0k=0 is clear. For k1k\geq 1 and n1n\geq 1,

p(expk(n))\displaystyle p(\exp_{k}(n)) cexpk(n)d2c(2expk1(n))d=2c+dexpk1(n)\displaystyle\leq c\cdot{\exp_{k}(n)}^{d}\leq 2^{c}\cdot{(2^{\exp_{k-1}(n)})}^{d}=2^{c+d\cdot\exp_{k-1}(n)}
2q(expk1(n))\displaystyle\leq 2^{q(\exp_{k-1}(n))} (As q(n)c+dnq(n)\geq c+dn)
2expk1(q(n))=expk(q(n)).\displaystyle\leq 2^{\exp_{k-1}(q(n))}=\exp_{k}(q(n))\text{.} (Lemma 46)

Finally, we combine both lemmas:

Proof A.5 (Proof of 11).

Let pp be a polynomial as above. W.l.o.g.\xspacepp is non-decreasing. Then by Lemma 47, p(expk(n))p(expk((k+1)n))p(\exp^{*}_{k}(n))\leq p\big{(}\exp_{k}((k+1)\cdot n)\big{)}. Moreover, due to Lemma 48, there is a polynomial qq such that p(expk((k+1)n))expk(q((k+1)n))p\big{(}\exp_{k}((k+1)\cdot n)\big{)}\leq\exp_{k}\big{(}q((k+1)\cdot n)\big{)}.

Proofs of Section 5

Proposition 14.

Let α,β\alpha,\beta be disjoint scopes and S,U,TS,U,T teams in a Kripke structure 𝒦=(W,R,V)\mathcal{K}=(W,R,V). Then the following laws hold:

  1. (1)

    Distributive laws: (TS)α=TαS=TSα=TαSα{(T\cap S)}_{\alpha}=T_{\alpha}\cap S=T\cap S_{\alpha}=T_{\alpha}\cap S_{\alpha} and (TS)α=TαSα{(T\cup S)}_{\alpha}=T_{\alpha}\cup S_{\alpha}.

  2. (2)

    Disjoint selection commutes: (TSα)Uβ=(TUβ)Sα{\big{(}T^{\alpha}_{S}\big{)}}^{\beta}_{U}={\big{(}T^{\beta}_{U}\big{)}}^{\alpha}_{S}.

  3. (3)

    Disjoint selection is independent: ((TSα)Uβ)α=TαS{\big{(}{(T^{\alpha}_{S})}^{\beta}_{U}\big{)}}_{\alpha}=T_{\alpha}\cap S.

  4. (4)

    Image and selection commute: (RT)α=(R(Tα))α=R(Tα){(RT)}_{\alpha}={\big{(}R(T_{\alpha})\big{)}}_{\alpha}=R(T_{\alpha})

  5. (5)

    Selection propagates: If STS\subseteq T, then R(TSα)=(RT)RSαR\big{(}T^{\alpha}_{S}\big{)}={(RT)}^{\alpha}_{RS}.

Proof A.6.
  1. (1)

    Observe that Xα=XWαX_{\alpha}=X\cap W_{\alpha}. Hence, for the union (TS)α=(TS)Wα=(TWα)(SWα)=TαSα{(T\cup S)}_{\alpha}=(T\cup S)\cap W_{\alpha}=(T\cap W_{\alpha})\cup(S\cap W_{\alpha})=T_{\alpha}\cup S_{\alpha} holds. For the intersection, likewise (TS)Wα=(TWα)S=T(WαS)=(TWα)(SWα)(T\cap S)\cap W_{\alpha}=(T\cap W_{\alpha})\cap S=T\cap(W_{\alpha}\cap S)=(T\cap W_{\alpha})\cap(S\cap W_{\alpha}).

  2. (2)

    Proved in the following equation. We use the fact that Xγγ=(Xγ)γ=(Xγ)γ=XγγX_{\gamma\land\gamma^{\prime}}={(X_{\gamma})}_{\gamma^{\prime}}={(X_{\gamma^{\prime}})}_{\gamma}=X_{\gamma^{\prime}\land\gamma} for all teams XX and scopes γ,γ\gamma,\gamma.

    (TSα)Uβ\displaystyle{\big{(}T^{\alpha}_{S}\big{)}}^{\beta}_{U}
    =\displaystyle=\; (T¬α(TαS))¬β((T¬α(TαS))βU)\displaystyle{\big{(}T_{\neg\alpha}\cup(T_{\alpha}\cap S)\big{)}}_{\neg\beta}\cup\Big{(}{\big{(}T_{\neg\alpha}\cup(T_{\alpha}\cap S)\big{)}}_{\beta}\cap U\Big{)}
    Distributing all scopes according to (1):
    =\displaystyle=\; T¬α¬β(Tα¬βS¬β)(T¬αβU)(TαβSβU)\displaystyle T_{\neg\alpha\land\neg\beta}\cup\big{(}T_{\alpha\land\neg\beta}\cap S_{\neg\beta}\big{)}\cup\big{(}T_{\neg\alpha\land\beta}\cap U\big{)}\cup\big{(}T_{\alpha\land\beta}\cap S_{\beta}\cap U\big{)}
    Replace UU by U¬α/UαU_{\neg\alpha}/U_{\alpha} due to the intersection law of (1):
    =\displaystyle=\; T¬α¬β(Tα¬βS¬β)(T¬αβU¬α)(TαβSβUα)\displaystyle T_{\neg\alpha\land\neg\beta}\cup\big{(}T_{\alpha\land\neg\beta}\cap S_{\neg\beta}\big{)}\cup\big{(}T_{\neg\alpha\land\beta}\cap U_{\neg\alpha}\big{)}\cup\big{(}T_{\alpha\land\beta}\cap S_{\beta}\cap U_{\alpha}\big{)}
    Likewise, replace S¬β/SβS_{\neg\beta}/S_{\beta} by SS:
    =\displaystyle=\; T¬α¬β(Tα¬βS)(T¬αβU¬α)(TαβSUα)\displaystyle T_{\neg\alpha\land\neg\beta}\cup\big{(}T_{\alpha\land\neg\beta}\cap S\big{)}\cup\big{(}T_{\neg\alpha\land\beta}\cap U_{\neg\alpha}\big{)}\cup\big{(}T_{\alpha\land\beta}\cap S\cap U_{\alpha}\big{)}
    Reverse distribution of scopes:
    =\displaystyle=\; (T¬β(TβU))¬α((T¬β(TβU))αS)\displaystyle{\big{(}T_{\neg\beta}\cup(T_{\beta}\cap U)\big{)}}_{\neg\alpha}\cup\Big{(}{\big{(}T_{\neg\beta}\cup(T_{\beta}\cap U)\big{)}}_{\alpha}\cap S\Big{)}
    =\displaystyle=\; (TUβ)Sα.\displaystyle{\big{(}T^{\beta}_{U}\big{)}}^{\alpha}_{S}\text{.}
  3. (3)

    By definition and application of (2), ((TSα)Uβ)α{\big{(}{(T^{\alpha}_{S})}^{\beta}_{U}\big{)}}_{\alpha} equals

    [(T¬β(TβU))¬α((T¬β(TβU))αS)]α\displaystyle{\Big{[}{\big{(}T_{\neg\beta}\cup(T_{\beta}\cap U)\big{)}}_{\neg\alpha}\cup\Big{(}{\big{(}T_{\neg\beta}\cup(T_{\beta}\cap U)\big{)}}_{\alpha}\cap S\Big{)}\Big{]}}_{\alpha}
    =\displaystyle=\; (T¬β(TβU))¬αα((T¬β(TβU))αS)α\displaystyle{\big{(}T_{\neg\beta}\cup(T_{\beta}\cap U)\big{)}}_{\neg\alpha\land\alpha}\cup{\Big{(}{\big{(}T_{\neg\beta}\cup(T_{\beta}\cap U)\big{)}}_{\alpha}\cap S\Big{)}}_{\alpha}
    =\displaystyle=\; ((T¬β(TβU))αSα)\displaystyle\;\emptyset\cup\Big{(}{\big{(}T_{\neg\beta}\cup(T_{\beta}\cap U)\big{)}}_{\alpha}\cap S_{\alpha}\Big{)}
    =\displaystyle=\; (T¬βαSα)(TβαUαSα)\displaystyle\big{(}T_{\neg\beta\land\alpha}\cap S_{\alpha}\big{)}\cup\big{(}T_{\beta\land\alpha}\cap U_{\alpha}\cap S_{\alpha}\big{)}
    Since α\alpha and β\beta are disjoint:
    =\displaystyle=\; (TαSα)(UαSα)=TαS.\displaystyle\big{(}T_{\alpha}\cap S_{\alpha}\big{)}\cup(\emptyset\cap U_{\alpha}\cap S_{\alpha})=T_{\alpha}\cap S\text{.}
  4. (4)

    (RT)α(R(Tα))α{(RT)}_{\alpha}\subseteq{\big{(}R(T_{\alpha})\big{)}}_{\alpha}: Suppose v(RT)αv\in{(RT)}_{\alpha}. Then vRwv\in Rw for some wTw\in T. Moreover, wTαw\in T_{\alpha}, since α\alpha is a scope. Hence vR(Tα)v\in R(T_{\alpha}). As vαv\vDash\alpha, v(R(Tα))αv\in{\big{(}R(T_{\alpha})\big{)}}_{\alpha} follows.

    (R(Tα))αR(Tα){\big{(}R(T_{\alpha})\big{)}}_{\alpha}\subseteq R(T_{\alpha}): Obvious.

    R(Tα)(RT)αR(T_{\alpha})\subseteq{(RT)}_{\alpha}: Again, let vR(Tα)v\in R(T_{\alpha}) be arbitrary. Then vRwv\in Rw for some wTαw\in T_{\alpha}. Hence vRTv\in RT. Since vαv\vDash\alpha follows from wαw\vDash\alpha, we conclude v(RT)αv\in{(RT)}_{\alpha}.

  5. (5)

    For \subseteq, suppose vR(TSα)v\in R(T^{\alpha}_{S}), i.e.\xspace, vRwv\in Rw for some wTSαw\in T^{\alpha}_{S}. In particular, vRTv\in RT. If wαw\nvDash\alpha, then vRT¬αv\in RT_{\neg\alpha} and trivially v(RT)RSαv\in{(RT)}^{\alpha}_{RS}. If wαw\vDash\alpha, then necessarily wSw\in S. Moreover, vαv\vDash\alpha. Consequently, vRSαRTαv\in{RS}_{\alpha}\cap{RT}_{\alpha}, hence v(RT)RSαv\in{(RT)}^{\alpha}_{RS}.

    For \supseteq, suppose v(RT)RSα=RT¬α(RTαRS)v\in{(RT)}^{\alpha}_{RS}=RT_{\neg\alpha}\cup(RT_{\alpha}\cap RS).

    If vRT¬αv\in RT_{\neg\alpha}, then by (4) vRwv\in Rw for some wT¬αw\in T_{\neg\alpha}. In particular, wTSαw\in T^{\alpha}_{S}, hence vR(TSα)v\in R\big{(}T^{\alpha}_{S}\big{)}.

    If vRTαRSv\in RT_{\alpha}\cap RS, then by (1) vRSαv\in RS_{\alpha}. By (4) vR(Sα)v\in R(S_{\alpha}), in other words, vRwv\in Rw for some wSαw\in S_{\alpha}. As STS\subseteq T, then wSαTw\in S_{\alpha}\cap T, and in fact wTαSw\in T_{\alpha}\cap S due to (1) Consequently, wTSαw\in T^{\alpha}_{S} and vR(TSα)v\in R(T^{\alpha}_{S}).

Proofs of Section 7

Lemma 23.

Let α,β𝖬𝖫\xspace\alpha,\beta\in\mathsf{ML}\xspace and φ𝖬𝖳𝖫\xspacek\varphi\in\mathsf{MTL}\xspace_{k}. Let TT be a team such that RiTαβR^{i}T\vDash\alpha\leftrightarrow\beta for all i{0,,k}i\in\{0,\ldots,k\}. Then TφT\vDash\varphi if and only if TSub(φ,α,β)T\vDash\mathrm{Sub}(\varphi,\alpha,\beta), where Sub(φ,α,β)\mathrm{Sub}(\varphi,\alpha,\beta) is the formula obtained from φ\varphi by substituting every occurrence of α\alpha with β\beta.

Proof A.7.

Proof by induction on kk and the syntax on φ\varphi. W.l.o.g.\xspaceα\alpha occurs in φ\varphi. If φ=α\varphi=\alpha, then Sub(φ,α,β)=β\mathrm{Sub}(\varphi,\alpha,\beta)=\beta, in which case the proof boils down to showing TαTβT\vDash\alpha\Leftrightarrow T\vDash\beta. However, this easily follows from TαβT\vDash\alpha\leftrightarrow\beta by the semantics for classical 𝖬𝖫\xspace\mathsf{ML}\xspace-formulas.

Otherwise, α\alpha is a proper subformula of φ\varphi. We distinguish the following cases.

  • φ=¬γ\varphi=\neg\gamma: Then Sub(¬γ,α,β)=¬Sub(γ,α,β)\mathrm{Sub}(\neg\gamma,\alpha,\beta)=\neg\mathrm{Sub}(\gamma,\alpha,\beta), and

    TSub(φ,α,β)\displaystyle T\vDash\mathrm{Sub}(\varphi,\alpha,\beta)
    \displaystyle\Leftrightarrow\; T¬Sub(γ,α,β)\displaystyle T\vDash\neg\mathrm{Sub}(\gamma,\alpha,\beta)
    \displaystyle\Leftrightarrow\; wT:{w}¬Sub(γ,α,β)\displaystyle\forall w\in T\colon\{w\}\vDash\neg\mathrm{Sub}(\gamma,\alpha,\beta)
    \displaystyle\Leftrightarrow\; wT:{w}¬γ\displaystyle\forall w\in T\colon\{w\}\vDash\neg\gamma (Induction hypothesis, as {w},Rw,αβ\{w\},Rw,\ldots\vDash\alpha\leftrightarrow\beta)
    \displaystyle\Leftrightarrow\; T¬γ\displaystyle T\vDash\neg\gamma
    \displaystyle\Leftrightarrow\; Tφ\displaystyle T\vDash\varphi
  • φ=ψ\varphi={\sim}\psi: By induction hypothesis, TSub(φ,α,β)T\vDash\mathrm{Sub}(\varphi,\alpha,\beta) iff TSub(ψ,α,β)T\vDash{\sim}\mathrm{Sub}(\psi,\alpha,\beta) iff TψT\vDash{\sim}\psi.

  • φ=ψθ\varphi=\psi\land\theta: Proved similarly to {\sim}.

  • φ=ψθ\varphi=\psi\lor\theta: First note that Sub(ψθ,α,β)=Sub(ψ,α,β)Sub(θ,α,β)\mathrm{Sub}(\psi\lor\theta,\alpha,\beta)=\mathrm{Sub}(\psi,\alpha,\beta)\lor\mathrm{Sub}(\theta,\alpha,\beta). Then:

    TSub(φ,α,β)\displaystyle T\vDash\mathrm{Sub}(\varphi,\alpha,\beta)
    \displaystyle\Leftrightarrow\; TSub(ψ,α,β)Sub(θ,α,β)\displaystyle T\vDash\mathrm{Sub}(\psi,\alpha,\beta)\lor\mathrm{Sub}(\theta,\alpha,\beta)
    \displaystyle\Leftrightarrow\; S,U:T=SU,SSub(ψ,α,β),USub(θ,α,β)\displaystyle\exists S,U\colon T=S\cup U,S\vDash\mathrm{Sub}(\psi,\alpha,\beta),U\vDash\mathrm{Sub}(\theta,\alpha,\beta)
    By induction hypothesis, since S,U,RS,RU,αβS,U,RS,RU,\ldots\vDash\alpha\leftrightarrow\beta:
    \displaystyle\Leftrightarrow\; S,U:T=SU,Sψ,Uθ\displaystyle\exists S,U\colon T=S\cup U,S\vDash\psi,U\vDash\theta
    \displaystyle\Leftrightarrow\; Tφ\displaystyle T\vDash\varphi
  • φ=ψ\varphi=\Box\psi: We have Sub(ψ,α,β)=Sub(ψ,α,β)\mathrm{Sub}(\Box\psi,\alpha,\beta)=\Box\mathrm{Sub}(\psi,\alpha,\beta), hence

    TSub(φ,α,β)\displaystyle T\vDash\mathrm{Sub}(\varphi,\alpha,\beta)
    \displaystyle\Leftrightarrow\; TSub(ψ,α,β)\displaystyle T\vDash\Box\mathrm{Sub}(\psi,\alpha,\beta)
    \displaystyle\Leftrightarrow\; RTSub(ψ,α,β).\displaystyle RT\vDash\mathrm{Sub}(\psi,\alpha,\beta)\text{.}
    However, since ψ𝖬𝖳𝖫\xspacek1\psi\in\mathsf{MTL}\xspace_{k-1} and RT,,Rk1(RT)αβRT,\ldots,R^{k-1}(RT)\vDash\alpha\leftrightarrow\beta holds by assumption, we obtain by induction hypothesis:
    \displaystyle\Leftrightarrow\; RTψ\displaystyle RT\vDash\psi
    \displaystyle\Leftrightarrow\; Tφ\displaystyle T\vDash\varphi
  • φ=ψ\varphi=\Diamond\psi: As before, Sub(ψ,α,β)=Sub(ψ,α,β)\mathrm{Sub}(\Diamond\psi,\alpha,\beta)=\Diamond\mathrm{Sub}(\psi,\alpha,\beta). Then:

    TSub(φ,α,β)\displaystyle T\vDash\mathrm{Sub}(\varphi,\alpha,\beta)
    \displaystyle\Leftrightarrow\; TSub(ψ,α,β)\displaystyle T\vDash\Diamond\mathrm{Sub}(\psi,\alpha,\beta)
    \displaystyle\Leftrightarrow\; SRT,TR1S:SSub(ψ,α,β)\displaystyle\exists S\subseteq RT,T\subseteq R^{-1}S\colon S\vDash\mathrm{Sub}(\psi,\alpha,\beta)
    Note that S,RS,,Rk1SS,RS,\ldots,R^{k-1}S are subteams of RT,,RkTRT,\ldots,R^{k}T, respectively. For this reason, the teams S,RS,,Rk1SS,RS,\ldots,R^{k-1}S satisfy αβ\alpha\leftrightarrow\beta as well. As also ψ𝖬𝖳𝖫\xspacek1\psi\in\mathsf{MTL}\xspace_{k-1} holds, we obtain by induction hypothesis:
    \displaystyle\Leftrightarrow\; SRT,TR1S:Sψ\displaystyle\exists S\subseteq RT,T\subseteq R^{-1}S\colon S\vDash\psi
    \displaystyle\Leftrightarrow\; Tφ\displaystyle T\vDash\varphi