April 02, 2024
In the product \(L_1\times L_2\) of two Kripke complete consistent logics, local tabularity of \(L_1\) and \(L_2\) is necessary for local tabularity of \(L_1\times L_2\). However, it is not sufficient: the product of two locally tabular logics may not be locally tabular. We provide extra semantic and axiomatic conditions that give criteria of local tabularity of the product of two locally tabular logics, and apply them to identify new families of locally tabular products. We show that the product of two locally tabular logics may lack the product finite model property. We give an axiomatic criterion of local tabularity for all extensions of \({\mathrm{\small S4}}.1 [ 2 ]\times {\mathrm{\small S5}}\). Finally, we describe a new prelocally tabular extension of \({\mathrm{\small S4}}\times{\mathrm{\small S5}}\).
It is well-known that the operation of product of modal logics does not preserve the finite model property of the factors, see, e.g., [1],[2], or the monography [3]. In this paper we describe new families of modal products which have the finite model property and, in fact, satisfy a stronger property of local tabularity.
A logic is locally tabular, if each of its finite-variable fragments contains only a finite number of pairwise nonequivalent formulas. In particular, every locally tabular logic has the finite model property. It is well known that for unimodal logics above \({\mathrm{\small K4}}\), local tabularity is equivalent to finite height [4],[5]. In the non-transitive unimodal, and in the polymodal case, no axiomatic criterion of local tabularity is known.
It follows from [6] that the product of a locally tabular logic with a tabular one is locally tabular. However, these cases are not exhaustive: other families of locally tabular modal products were identified in [7], see also [8]. For close systems, intuitionistic modal logics and expanding products, locally tabular families were identified in [9], [10], [11], and a recent manuscript [12].
In the product \(L_1\times L_2\) of two Kripke complete consistent logics, local tabularity of \(L_1\) and \(L_2\) is necessary for local tabularity of \(L_1\times L_2\). However, it is not sufficient: the product of two locally tabular logics can be not locally tabular. The simplest example is the logic \({\mathrm{\small S5}}\times {\mathrm{\small S5}}\) [13]. We provide extra semantic (Theorem 26) and axiomatic (Corollary 3) conditions which give criteria of local tabularity of the product of two locally tabular logics: bounded cluster property of one of the factors; a condition we call product reducible path property; finiteness of the one-variable fragment of the product.
In Section 4, we apply the criteria to identify new families of locally tabular products. In particular, we generalize some results from [7], [8].
In Section 5, we discuss the product finite model property in the locally tabular case. A modal logic \(L\) has the product fmp, if \(L\) is the logic of a class of finite product frames. The product fmp is stronger than the fmp: for example, \({\mathrm{\small K4}}\times {\mathrm{\small S5}}\) has the fmp [14], but lacks the product fmp [3]. It is perhaps surprising that the local tabularity of a product logic does not imply the product fmp even in the case of height 3, as we discovered in Theorem 31.
In Section 6, we discuss local tabularity and prelocal tabularity in extensions of \({\mathrm{\small S4}}\times{\mathrm{\small S5}}\). The logic \({\mathrm{\small S5}}\) is known to be one of the five pretabular logics above \({\mathrm{\small S4}}\) [15], [16]. We observe that for another pretabular logic \({\mathrm{\small Tack}}\), the logic of the ordered sum of a countable cluster and a singleton, the product \({\mathrm{\small Tack}}\times {\mathrm{\small S5}}\) is not prelocally tabular. Then we consider a weaker than \({\mathrm{\small Tack}}\) logic \({\mathrm{\small S4}}.1 [ 2 ]\), the extension of \({\mathrm{\small S4.1}}\) with the axiom of height 2, and give an axiomatic criterion of local tabularity for all normal extensions of \({\mathrm{\small S4}}.1 [ 2 ]\times {\mathrm{\small S5}}\). Finally, we discuss prelocal tabularity. It is known that \({\mathrm{\small S5}}\times {\mathrm{\small S5}}\) is prelocally tabular [17]. We construct a frame of height two, which we call two-dimensional tack, and show that its logic is another example of a prelocally tabular logic above \({\mathrm{\small S4}}\times {\mathrm{\small S5}}\).
For basic notions in modal logic, see, e.g., [18] or [19].
Let \(\mathrm{A}\) be a finite set, an alphabet of modalities. Modal formulas over \(\mathrm{A}\), \({\mathrm{\small ML}}(\mathrm{A})\) in symbols, are constructed from a countable set of variables \({\mathrm{\small PV}}=\{p_0,p_1,\ldots\}\) using Boolean connectives \(\bot,\rightarrow\) and unary connectives \(\lozenge\in \mathrm{A}\). Other logical connectives are defined as abbreviations in the standard way, in particular \(\Box\varphi\) denotes \(\neg \lozenge\neg \varphi\).
We define the following abbreviations: \(\lozenge^0 \varphi=\varphi\), \(\lozenge^{i+1}\varphi=\lozenge^i\lozenge\varphi\), \(\lozenge^{\leq m} \varphi= \bigvee_{i\leq m} \lozenge^i \varphi\), \(\Box^{\leq m} \varphi=\neg \lozenge^{\leq m} \neg \varphi\). We write \(\lozenge_\mathrm{A}\varphi\) for \(\bigvee_{\lozenge\in\mathrm{A}}\lozenge\varphi\).
By an \(\mathrm{A}\)-logic we mean a normal modal logic whose alphabet of modalities is \(\mathrm{A}\), see, e.g. [19]. The terms unimodal and bimodal refers to the cases \(\mathrm{A}=\{\lozenge\}\) and \(\mathrm{A}= \{\lozenge_1,\lozenge_2\}\), respectively.
An \(\mathrm{A}\)-frame is a pair \(F = (X,\,(R_\lozenge)_{\lozenge\in \mathrm{A}})\), where \(X\) is a non-empty set and \(R_\lozenge\subseteq X\times X\) for any \(\lozenge\in \mathrm{A}\). We write \(\mathop{\mathrm{dom}}F\) for \(X\). By the cardinality \(|F|\) of \(F\) we mean the cardinality of \(X\). We put \(R_F = \bigcup_{\lozenge\in \mathrm{A}} R_\lozenge.\) For \(a\in X\), \(Y\subseteq X\), we put \(R_\lozenge(a)=\{b\mid aR_\lozenge b\}\), \(R_\lozenge[Y]=\bigcup_{a\in Y} R_\lozenge(a)\).
A model on \(F\) is a pair \((F,\,\theta)\), where \(\theta:\:PV \to \mathcal{P}(X)\), and \(\mathcal{P}(X)\) is the powerset of \(X\); \(\theta\) is called a valuation in \(F\). The truth-relation \((F,\theta),a\vDash\varphi\) is defined in the standard way; in particular, for \(\lozenge\in \mathrm{A}\), \((F,\,\theta),\,a\models \lozenge\varphi\) means that \((F,\,\theta),\,b\models \varphi\text{ for some }b\in R_\lozenge(a)\). We put \[\bar{\theta}(\varphi)=\{a\mid (F,\theta),a\vDash\varphi\}.\] A formula \(\varphi\in {\mathrm{\small ML}}(\mathrm{A})\) is valid in an \(\mathrm{A}\)-frame \(F\), in symbols \(F\vDash\varphi\), if \(X=\bar{\theta}(\varphi)\) for every model \((F,\theta)\) on \(F\). For a set of formulas \(\Gamma\subseteq {\mathrm{\small ML}}(\mathrm{A})\), \(F\vDash\Gamma\) means that \(F\vDash\varphi\) for all \(\varphi\in\Gamma\); in this case \(F\) is said to be a \(\Gamma\)-frame. For a class \(\mathcal{F}\) of \(\mathrm{A}\)-frames \(\mathcal{F}\), \(\mathcal{F}\vDash\Gamma\) means that \(F\vDash\Gamma\) for all \(F\in\mathcal{F}\).
The set of \(\mathrm{A}\)-formulas that are valid in \(F\) is denoted by \(\mathop{\mathrm{Log}}F\). For a class \(\mathcal{F}\) of \(\mathrm{A}\)-frames, \(\mathop{\mathrm{Log}}\mathcal{F}=\bigcap\{\mathop{\mathrm{Log}}F \mid F\in \mathcal{F}\}\); this set is a logic [19] and is called the logic of \(\mathcal{F}\). Such logics are said to be Kripke complete.
Let us recall that \({\mathrm{\small K4}}\) is the smallest unimodal logic that contains \(\lozenge\lozenge p \to \lozenge p\), \({\mathrm{\small S4}}\) extends \({\mathrm{\small K4}}\) with \(p\rightarrow\lozenge p\), and \({\mathrm{\small S5}}\) extends \({\mathrm{\small S4}}\) with \(p\rightarrow\Box\lozenge p\). These logics are Kripke complete; \(\lozenge\lozenge p \to \lozenge p\) expresses the transitivity, \(p\rightarrow\lozenge p\) reflexivity, and \(p\rightarrow\Box\lozenge p\) the symmetry of a binary relation, see, e.g., [19] or [18].
The notions of generated subframe and p-morphism (or bounded morphism) are defined in the standard way, see, e.g., [19]. In particular, for frames \(F = (X,\,(R_\lozenge)_{\lozenge\in \mathrm{A}})\) and \(G = (Y,\,(S_\lozenge)_{\lozenge\in \mathrm{A}})\), \(f:\:X\to Y\) is a p-morphism from \(F\) to \(G\), if
\(f\) is a homomorphism, that is for any \(a,\,b\in X\) and \(\lozenge\in \mathrm{A},\) \(a R_\lozenge b\) implies \(f(a) S_\lozenge f(b)\), and
\(f\) satisfies the back condition, that is for any \(a\in X,\,u\in Y\), and \(\lozenge\in \mathrm{A},\) \(f(a) S_\lozenge u\) implies that there exists \(b\in X\) such that \(a R_\lozenge b\) and \(f(b) = u.\)
If \(f\) is surjective, we write \(f:\:F\twoheadrightarrow G.\) If there exists a p-morphism from \(F\) onto \(G,\) we write \(F \twoheadrightarrow G;\) in this case, \(\mathop{\mathrm{Log}}{F} \subseteq \mathop{\mathrm{Log}}{G}\) [19].
For a frame \(F = (X,\,(R_\lozenge)_{\lozenge\in A})\) and an equivalence \(\approx\) on \(F\), the quotient frame \({F}{/}{\approx}\) is the frame \(({X}{/}{\approx},\,(R^\approx_\lozenge)_{\lozenge\in\mathrm{A}})\), where for \([a],[b]\in {X}{/}{\approx}\) and \(\lozenge\in \mathrm{A}\), \[[a] R^\approx_\lozenge[b] \text{ iff }\exists a'\in[a]\;\exists b'\in [b]\;\left(a' R_{\lozenge} b' \right).\]
Let \(\mathrm{A}\) and \(\mathrm{B}\) be two disjoint finite sets. For an \(\mathrm{A}\)-frame \(F=\left(X,\,(R_\lozenge)_{\lozenge\in \mathrm{A}}\right)\) and \(\mathrm{B}\)-frame \(G = \left(Y,\,(R_\lozenge)_{\lozenge\in \mathrm{B}}\right)\), the product frame \(F\times G\) is the frame \((X\times Y,\,(R^h_\lozenge)_{\lozenge\in \mathrm{A}}, (R^v_\lozenge)_{\lozenge\in \mathrm{B}})\), where \[\begin{align} R^h_\lozenge&=\{((a,\,b),\,(a',\,b))\mid a,\,a'\in X,\,b\in Y,\,a R_\lozenge a'\} &\text{ for }\lozenge\in \mathrm{A};\\ R^v_\lozenge&= \{((a,\,b),\,(a,\,b'))\mid a\in X,\,b,\,b'\in Y,\,b R_\lozenge b'\} &\text{ for }\lozenge\in \mathrm{B}. \end{align}\] We say that relations \(R^h_\lozenge\) are horizontal, and \(R^v_\lozenge\) are vertical. For a class \(\mathcal{F}\) of \(\mathrm{A}\)-frames and a class \(\mathcal{G}\) of \(\mathrm{B}\)-frames, we put \(\mathcal{F}\times \mathcal{G}= \{F\times G \mid F\in \mathcal{F},\,G\in\mathcal{G}\}.\)
For logics \(L_1,L_2\), the product logic \(L_1\times L_2\) is defined as \[L_1\times L_2 = \mathop{\mathrm{Log}}\{F\times G \mid F\models L_1,\,G\models L_2\}.\] If \(L_1\) and \(L_2\) are unimodal logics, we rename their modalities and follow the convention that the alphabet of \(L_1\times L_2\) is \(\{\lozenge_1,\lozenge_2\}\). We write \(L^2\) for \(L\times L\).
For finite \(k\), a logic \(L\) is \(k\)-finite, if \(L\) contains only a finite number of pairwise nonequivalent formulas in variables \(p_i\), \(i<k\). \(L\) is said to be locally tabular, if it is \(k\)-finite for all finite \(k\). The following fact is well known, see, e.g., [18].
Proposition 1. Let \(L\) be locally tabular. Then:
\(L\) has the finite model property, that is, \(L\) is the logic of a class of finite frames. In particular, \(L\) is Kripke complete.
Every extension of \(L\) (in the same modal alphabet) is locally tabular.
In algebraic terms, local tabularity of a logic means that its Lindenbaum algebra is locally finite, which, in turn, means that the variety of its algebras is locally finite (see, e.g., [20], [18] for corresponding notions).
A class \(\mathcal{C}\) of algebras of a finite signature is said to be uniformly locally finite, if there exists a function \(f:\omega\to \omega\) such that the cardinality of a subalgebra of any \(B\in \mathcal{C}\) generated by \(k<\omega\) elements does not exceed \(f(k)\).
Theorem 2. [21]. Local finiteness of the variety generated by a class \(\mathcal{C}\) is equivalent to uniform local finiteness of \(\mathcal{C}\).
For a binary relation \(R\) on \(X\) and a number \(m < \omega\), put \(R^{\leq m}= \bigcup_{i \leq m} R^i\), where \(R^{i+1}=R\circ R^i\), \(\circ\) is the composition of relations, \(R^0\) is the diagonal relation on \(X\). The transitive reflexive closure \(\bigcup_{i < \omega} R^i\) of \(R\) is denoted by \(R^*\).
A frame \(F\) is said to be \(m\)-transitive, if \(R_F^{\leq m}=R_F^*\).1 \(F\) is pretransitive, if it is \(m\)-transitive for some finite \(m\). The following fact is standard.
Proposition 3 (Jankov-Fine theorem for pretransitive frames). If \(\mathop{\mathrm{Log}}G\subseteq \mathop{\mathrm{Log}}F\), \(G\) is pretransitive, and \(F\) is finite and rooted, then there exists a point-generated subframe \(H\) of \(G\) such that \(H\twoheadrightarrow F\).
For any \(\mathrm{A}\)-frame \(F\), we have the following equivalence [22]: \[\label{pretr:sem} \text{ R_F is m-transitive iff F\vDash\lozenge_\mathrm{A}^{m+1} p \rightarrow\lozenge_\mathrm{A}^{\leq m} p. }\tag{1}\] A logic \(L\) is said to be \(m\)-transitive, if \(L\) contains \(\lozenge_\mathrm{A}^{m+1} p \rightarrow\lozenge_\mathrm{A}^{\leq m} p\), and pretransitive, if it is \(m\)-transitive for some \(m\geq 0\).
For a frame \(F\), let \(F^*\) be the preorder \((\mathop{\mathrm{dom}}{F},R_F^*)\). For a class \(\mathcal{F}\) of frames, \(\mathcal{F}^*=\{F^*\mid F\in F\}\).
The following fact is straightforward: for an \(m\)-transitive frame \(F=(X,(R_\lozenge)_{\lozenge\in \mathrm{A}})\), and any unimodal formula \(\varphi\), we have \[\label{eq:pretrans} F\vDash{ [\varphi ]}^{m} \text{ iff } F^*\vDash\varphi.\tag{2}\] where \({ [\varphi ]}^{m}\) is the translation that is compatible with Boolean connectives (that is, \({ [\bot ]}^{m}=\bot\), \({ [p ]}^{m}=p\) for variables, \({ [\psi_1\rightarrow\psi_2 ]}^{m}={ [\psi_1 ]}^{m}\rightarrow{ [\psi_2 ]}^{m}\)), given by \({ [\lozenge\psi ]}^{m}= \lozenge_\mathrm{A}^{\leq m} { [\psi ]}^{m}\).
For a frame \(F = (X,\,(R_\lozenge)_{\lozenge\in \mathrm{A}}),\) let \(\sim_F\) be the equivalence \(R_F^*\cap (R_F^*)^{-1}\) on \(X\), and \(\mathop{\mathrm{Sk}}F={F^*}{/}{\sim_F}\). The quotient frame \(\mathop{\mathrm{Sk}}F\) is a poset; it is called the skeleton of \(F\). Equivalence classes modulo \(\sim_F\) are called clusters in \(F\).
We say that \(F\) has a finite height \(h\), in symbols \(h(F)=h\), if \(\mathop{\mathrm{Sk}}F\) contains a chain of \(h\) elements and no chains of more than \(h\) elements. Consider unimodal formulas \[B_0 = \bot,\quad B_{i+1} = p_{i+1}\rightarrow\Box \left (\lozenge p_{i+1} \lor B_i\right).\] For a unimodal frame \(G=(X,R)\) with a transitive \(R\), we have [4]: \[G\vDash B_h \text{ iff }h(G)\leq h.\] Consequently, if \(F\) is an \(m\)-transitive \(\mathrm{A}\)-frame, then \(F\models { [B_h ]}^{m}\) iff \(h(F)\le h\). Formulas \(B_h\) and their polymodal pretransitive generalizations \({ [B_h ]}^{m}\) are called formulas of finite height. For an \(m\)-transitive logic \(L\), we define the height of \(L\), denoted by \(h(L),\) as the smallest \(h < \omega\) such that \(L \vdash [B_h]^m.\)
Theorem 4. [23] Let \(L\) be an \(\mathrm{A}\)-logic. If \(L\) is 1-finite, then for some \(h,m<\omega\), \(L\) contains \(\lozenge_\mathrm{A}^{m+1} p \rightarrow\lozenge_\mathrm{A}^{\leq m} p\) and \({ [B_h ]}^{m}.\)
In [23], this fact was stated for the unimodal case; its polymodal generalization is straightforward, details can be found in [24].
A partition \(\mathcal{V}\) of a set \(X\) is a family of non-empty pairwise disjoint sets such that \(X=\bigcup \mathcal{V}\). For \(x\in X\), \([x]_\mathcal{V}\) denotes \(V\in\mathcal{V}\) that contains \(x\). A partition of \(X\) induced by \(h :X\to Y\) is the set of non-empty sets \(h^{-1}(y)\), \(y\in Y\). A partition \(\mathcal{U}\) refines \(\mathcal{V}\), if each element of \(\mathcal{V}\) is the union of some elements of \(\mathcal{U}\).
Definition 1. Let \(R\) be a binary relation on \(X\). A partition \(\mathcal{U}\) of \(X\) is said to be \(R\)-tuned, if for every \(U,V\in \mathcal{U}\), \[\label{eq:part} \exists a\in U \exists b\in V \, (aR b) \,\text{ implies }\, \forall a\in U \exists b\in V \, (aR b).\tag{3}\] Let \(F=(X,(R_\lozenge)_{\lozenge\in \mathrm{A}})\) be an \(\mathrm{A}\)-frame. A partition \(\mathcal{U}\) of \(X\) is said to be tuned in \(F\), if it is \(R_\lozenge\)-tuned for every \(\lozenge\in \mathrm{A}\). In other terms, the condition \(\eqref{eq:part}\) means that the map \(x\mapsto [x]_\mathcal{U}\) is a p-morphism \(F\twoheadrightarrow{F}{/}{\sim}\), where \(\mathcal{U}={X}{/}{\sim}\), see, e.g., [25] for details.
A class \(\mathcal{F}\) of \(\mathrm{A}\)-frames is said to be \(f\)-tunable for a function \(f:\omega\to\omega\), if for every \(F\in\mathcal{F}\), for every finite partition \(\mathcal{V}\) of \(F\) there exists a refinement \(\mathcal{U}\) of \(\mathcal{V}\) such that \(|\mathcal{U}|\leq f(|\mathcal{V}|)\) and \(\mathcal{U}\) is tuned in \(F\). A class \(\mathcal{F}\) is uniformly tunable, if it is \(f\)-tunable for some \(f:\omega\to\omega\).
The following theorem is a Kripke-style version of Malcev criterion:
Theorem 5. [23]
The logic of a class \(\mathcal{F}\) of \(\mathrm{A}\)-frames is locally tabular iff \(\mathcal{F}\) is uniformly tunable.
Every locally tabular logic is the logic of a uniformly tunable class.
This characterization makes many properties of locally tabular logics visible.
Example 1. [26] Let \(L\) be a locally tabular logic, and let \({L}_{\mathrm{\small u}}\) be the expansion of \(L\) with the universal modality [27]. Then \({L}_{\mathrm{\small u}}\) is locally tabular. Indeed, adding the universal modality to a pretransitive logic preserves Kripke-completeness [28]. Hence, \({L}_{\mathrm{\small u}}\) is characterized by its point-generated frames, which are \(L\)-frames with additional universal relation \(X\times X\). Clearly, \(X\times X\) is tuned with respect to any partition, and so this class inherits uniform tunability from \(L\)-frames.
For a frame \(F = (X,\,(R_\lozenge)_{\lozenge\in \mathrm{A}})\) and \(Y \subseteq X\), put \(F{\upharpoonright}Y = (Y,\,(R_\lozenge{\upharpoonright}Y)_{\lozenge\in\mathrm{A}})\), where \(R_\lozenge{\upharpoonright}Y= R_\lozenge\cap (Y\times Y)\). For a class \(\mathcal{F}\) of frames, put \[\mathop{\mathrm{Sub}}{\mathcal{F}}=\{F{\upharpoonright}Y\mid F\in \mathcal{F}\text{ and } \varnothing\neq Y\subseteq \mathop{\mathrm{dom}}{F}\}.\]
Proposition 6. [24]If \(\mathop{\mathrm{Log}}\mathcal{F}\) is \(k\)-finite for some positive \(k<\omega\), then the logic of \(\mathop{\mathrm{Sub}}{\mathcal{F}}\) is \({(k{-}1)}\)-finite.
A frame \(F\) is a cluster, if \(a R_F^* b\) for all \(a,b\) in \(F\). For a class \(\mathcal{F}\) of frames, \(Cl{(\mathcal{F})}\) is the class of cluster-frames which are restrictions on clusters (as sets) occurring in frames in \(\mathcal{F}\): \[Cl{(\mathcal{F})}=\{F{\upharpoonright}C\mid C \text{ is a cluster in } F\in\mathcal{F}\}.\] A class \(\mathcal{F}\) is of uniformly finite height, if for some finite \(h\), for every \(F\in \mathcal{F}\), we have \(h(F)\leq h\).
The following characterization is one of the main technical tools for our study.
Theorem 7. [23]A logic \(\mathop{\mathrm{Log}}\mathcal{F}\) is locally tabular iff \(\mathcal{F}\) is of uniformly finite height and \(\mathop{\mathrm{Log}}Cl{(\mathcal{F})}\) is locally tabular.
Remark. In [23], Theorems 5 and 7 were stated for the unimodal case. Their polymodal generalizations are straightforward. For Theorem 5, details are given in [24]. The polymodal version of Theorem 7 was not considered before; for the sake of rigor, we prove it in Appendix.
Throughout this section, \(\mathrm{A}\) and \(\mathrm{B}\) are finite disjoint alphabets, \(L_1\) is an \(\mathrm{A}\)-logic, \(L_2\) is a \(\mathrm{B}\)-logic, \(\mathcal{F}\) and \(\mathcal{G}\) are classes of \(\mathrm{A}\)- and \(\mathrm{B}\)-frames, respectively.
In the product \(L_1\times L_2\) of two Kripke complete consistent logic, local tabularity of the factors is a necessary condition for local tabularity of \(L_1\times L_2\): the product is conservative over the factors in this case. However, it is not sufficient: it is known that the product of two locally tabular logics can be not locally tabular. The simplest example is the logic \({\mathrm{\small S5}}^2\): that \({\mathrm{\small S5}}\) is locally tabular is well known (for example, one can observe that in the class of frames of this Kripke complete logic every partition is tuned); at the same time \({\mathrm{\small S5}}^2\) is not locally tabular [13]. We provide extra conditions which give criteria of local tabularity of the product of locally tabular logics.
The following classical fact about \({\mathrm{\small S5}}^2\) is important for our goals. For a set \(X\), let \({\boldsymbol{X}}\) be the frame \((X,X\times X)\). For sets \(X,\,Y\), the product frame \(\boldsymbol{X} \times \boldsymbol{Y}\) is called a rectangle.
Theorem 8. [29] If \(\mathcal{X}\) and \(\mathcal{Y}\) are families of non-empty sets and \(\sup\{|X| \mid X\in\mathcal{X}\}\) and \(\sup\{|Y| \mid X\in\mathcal{Y}\}\) are infinite, then \(\mathop{\mathrm{Log}}\{\boldsymbol{X} \times \boldsymbol{Y}\mid X\in \mathcal{X}\text{ and } Y\in\mathcal{Y}\}\) is \({\mathrm{\small S5}}^2\).
Remark 9. The system considered in [29] is a conservative extension of \({\mathrm{\small S5}}^2\). See [3] for other references.
Proposition 10. Let \(R_h\) and \(R_v\) be the unions of all horizontal and all vertical relations of \(F\times G\), respectively. Then \(R_{F\times G}^* = R_h^* \circ R_v^*=R_v^* \circ R_h^*\).
Proof. If \(R\) is a vertical and \(S\) is a horizontal relation in \(F\times G\), then they commute: \(R\circ S=S\circ R\). By a routine argument, \(R_v^i\) and \(R_h^j\) commute for all \(i,j\). It follows that \(R_h^* \circ R_v^*=R_v^* \circ R_h^*\).
Since \(R_h^*\) and \(R_v^*\) are contained in the transitive relation \(R_{F\times G}^*\), we have \(R_h^* \circ R_v^*\subseteq R_{F\times G}^*\). By induction on \(m\), one can easily check that \(R_{F\times G}^{\leq m} \subseteq R_h^* \circ R_v^*\) for all \(m\). Hence, \(R_{F\times G}^* \subseteq R_h^* \circ R_v^*\). ◻
Proposition 11.
If \(F\) is \(m\)-transitive and \(G\) is \(n\)-transitive, then \(F\times G\) is \((m+n)\)-transitive.
The product of two pretransitive logics is pretransitive.
Proof. The second statement is immediate from the first. To check the first statement, let \(R_h\) and \(R_v\) be the unions of all horizontal and all vertical relations of \(F\times G\), respectively. Since \(F\) is \(m\)-transitive, \(R_h\) is \(m\)-transitive; similarly, \(R_v\) is \(n\)-transitive. By Proposition 10, we have \(R^*_{F\times G}=R_h^* \circ R_v^*=R_h^{\leq m} \circ R_v^{\leq n}\subseteq R^{\leq m}_{F\times G}\circ R^{\leq n}_{F\times G} = R^{\leq m+n}_{F\times G}\). ◻
Example 2. \({\mathrm{\small S4}}^2\), \({\mathrm{\small S4}}\times {\mathrm{\small S5}}\), \({\mathrm{\small S5}}^2\) are 2-transitive logics.
Proposition 12. If \(h(F)=h_1\) and \(h(G)=h_2\) for some finite \(h_1, h_2\), then \(h(F\times G)=h_1+h_2-1\).
Proof. Consider a chain \([a_0] < [a_1] < \ldots < [a_{h_1}]\) in \(\mathop{\mathrm{Sk}}F\) and a chain \([b_0]< [b_1] < \ldots < [b_{h_2}]\) in \(\mathop{\mathrm{Sk}}G.\) Then there is a chain of \(h_1 + h_2 - 1\) elements in \(\mathop{\mathrm{Sk}}(F\times G)\): \[[(a_0,\,b_0)] < [(a_1,\,b_0)] < \ldots < [(a_{h_1},\,b_0)] < [(a_{h_1},\,b_1)] < \ldots < [(a_{h_1},\,b_{h_2})]\] Hence, \(h(F\times G) \geq h_1 + h_2 -1\).
If \(\Sigma\) is a chain of size \(l\) in \(\mathop{\mathrm{Sk}}(F\times G)\) with the least \([(a,b)]\) and the largest \([(c,d)]\), then there is a chain \(\Sigma_1\) in \(\mathop{\mathrm{Sk}}F\) with the least \([a]\) and the largest \([c]\), and a chain \(\Sigma_2\) in \(\mathop{\mathrm{Sk}}G\) with the least \([b]\) and the largest \([d]\) such that \(|\Sigma_1|+|\Sigma_2|=l+1\). This follows by induction on \(l\). Hence, every chain in \(\mathop{\mathrm{Sk}}(F\times G)\) has at most \(h_1+h_2-1\) elements. ◻
Example 3. \({\mathrm{\small S5}}^2\) is a logic of height 1.
For a formula \(\varphi\) in the alphabet \(\{\lozenge_1,\lozenge_2\}\), we define a the translation \({ [\varphi ]}^{m,n}\in {\mathrm{\small ML}}(\mathrm{A}\cup\mathrm{B})\). Let \({ [\varphi ]}^{m,n}\) be compatible with Boolean connectives, and let \({ [\lozenge_1 \psi ]}^{m,n}= \lozenge_\mathrm{A}^{\leq m} { [\psi ]}^{m,n}\), \({ [\lozenge_2 \psi ]}^{m,n}= \lozenge_\mathrm{B}^{\leq n} { [\psi ]}^{m,n}\). The following is straightforward:
Proposition 13. If \(F\) is an \(m\)-transitive \(\mathrm{A}\)-frame and \(G\) is an \(n\)-transitive \(\mathrm{B}\)-frame, then \(F\times G\vDash{ [\varphi ]}^{m,n}\) iff \(F^*\times G^*\vDash\varphi\).
Proposition 14. Let \(\mathcal{F}\) and \(\mathcal{G}\) be classes of \(m\)- and \(n\)-transitive clusters, respectively. If \(\sup\{|F| \mid F\in\mathcal{F}\}\) and \(\sup\{|G | \mid G\in\mathcal{F}\}\) are infinite, then \(\{\varphi\in {\mathrm{\small ML}}(\lozenge_1,\lozenge_2) \mid \mathcal{F}\times \mathcal{G}\vDash{ [\varphi ]}^{m,n}\} ={\mathrm{\small S5}}^2\).
Proposition 15. Assume that a frame \(F\) is \(f\)-tunable, and a frame \(G\) is finite. Then \(F\times G\) is \(g\)-tunable for \(g(n)=f\left(n^{|G|}\right)|G|\).
Proof. Let \(\mathcal{V}\) be a finite partition of \(F\times G\) , \(n=|\mathcal{V}|\).
Define \(\alpha:\:\mathop{\mathrm{dom}}F \to \mathcal{V}^{\mathop{\mathrm{dom}}G}\) by \[\alpha(a) = ([(a,\,b)]_\mathcal{V})_{b\in\mathop{\mathrm{dom}}G},\quad a\in \mathop{\mathrm{dom}}F.\] Then \(\alpha\) induces a partition of \(F\) of cardinality at most \(n^{|G|}.\) Since \(F\) is \(f\)-tunable, there is a refinement \(\mathcal{U}\) of this partition such that \(|\mathcal{U}|\le f\left(n^{|G|}\right)\) and \(\mathcal{U}\) is tuned in \(F\).
For \((a,b)\) in \(F\times G\), put \(\beta(a,b) = ([a]_{\mathcal{U}},\,b)\). Then \(\beta\) induces a refinement \(\mathcal{S}\) of \(\mathcal{U}\) of cardinality at most \(f\left(n^{|G|}\right)| G|.\) We show that \(\mathcal{S}\) is tuned in \(F\times G\).
Let \(R_\lozenge\) be a relation of \(F,\) and suppose \((a_1,b) R_\lozenge^h (a_2,b)\) in \(F\times G\). Then \(a_1 R_\lozenge a_2.\) Let \((a_1',b')\in [(a_1,b)]_\mathcal{S}.\) Then \(b'=b\) and \([a_1']_{\mathcal{U}} = [a_1]_{\mathcal{U}}\). We have \(a_1' R_\lozenge a_2'\) for some \(a_2'\in [a_2]_\mathcal{U}\), since \(\mathcal{U}\) is tuned in \(F\). Then \((a_1',b') R_\lozenge^h (a_2',b)\) and \((a_2',b)\in [(a_2,b)]_\mathcal{S}\).
Now let \((a,b_1) R_\lozenge^v (a,b_2)\), where \(R_\lozenge\) is a relation of \(G.\) Let \((a',b_1') \in[(a,b_1)]_\mathcal{S}.\) Then \(b_1' = b_1\), so \(b_1' R_\lozenge b_2.\) Moreover, \([a']_\mathcal{U}\) = \([a]_\mathcal{U}\), so \(\beta(a',b_2) = \beta(a,b_2)\). Hence, we have \((a',b_1') R_\lozenge^v (a',b_2)\) and \((a',b_2)\in[(a,b_2)]_\mathcal{S}\). ◻
Corollary 1. If \(L_1\) is locally tabular and \(L_2\) is tabular, then \(L_1\times L_2\) is locally tabular.
Recall that for a class \(\mathcal{F}\), \(Cl{(\mathcal{F})}\) denotes the class of clusters occurring in frames in \(\mathcal{F}\). We say that \(\mathcal{F}\) has the bounded cluster property, if for some \(m<\omega\), \(|C|<m\) for all \(C\in Cl{(\mathcal{F})}\).
Theorem 16. Let \(\mathcal{F}\) and \(\mathcal{G}\) be non-empty classes of \(\mathrm{A}\)- and \(\mathrm{B}\)-frames, respectively. The logic \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is locally tabular iff the logics \(\mathop{\mathrm{Log}}\mathcal{F}\) and \(\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular, and at least one of the classes \(\mathcal{F}\), \(\mathcal{G}\) has the bounded cluster property.
Proof. Assume that \(L=\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is locally tabular. Then \(L_1=\mathop{\mathrm{Log}}\mathcal{F}\) and \(L_2=\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular, since \(L\) is a conservative extension of each of them. By Theorem 4, \(L_1\) and \(L_2\) are pretransitive. Let \(\mathcal{C}_1=Cl{(\mathcal{F})}\), \(\mathcal{C}_2=Cl{(\mathcal{G})}\), and \(\mathcal{C}=Cl{(\mathcal{F}\times \mathcal{G})}\). It follows from Proposition 10 that \(\mathcal{C}_1\times \mathcal{C}_2=\mathcal{C}\). By Theorem 7, the logic of \(\mathcal{C}\) is locally tabular. On the other hand, \({\mathrm{\small S5}}^2\) is not locally tabular. By Proposition 14, for some finite \(n\) we have \(|C|<n\) for all \(C\) in \(\mathcal{C}_1\) or for all \(C\) in \(\mathcal{C}_2\).
The other direction follows from the facts that locally tabular logics have finite height (Theorem 4), which is preserved under the product (Proposition 12). By Theorem 7, it is enough to show that clusters occurring in the underlying products form a uniformly tunable class. The required partitions exist according to the bounded cluster property, in view of Proposition 15. ◻
For \(m<\omega\), consider the following first-order property \(\mathrm{RP}_m\): \[\forall x_0,\ldots, x_{m+1} \; (x_0Rx_1R\ldots R x_{m+1}\rightarrow\bigvee\limits_{i<j\leq m+1} x_i = x_j \vee \bigvee\limits_{i< j\leq m} x_i R x_{j+1}).\] We say that a class of \(\mathcal{F}\) of frames has the reducible path property (for short, rpp), if for some fixed \(m\), for every \(F\in \mathcal{F}\), \(\mathrm{RP}_m\) holds in \((X,R_F)\).
Notice that \(\mathrm{RP}_m\) is stronger than \(m\)-transitivity. This property is another nesessary condition for local tabularity:
Theorem 17. [23] If a class of unimodal frames is uniformly tunable (equivalently, has a locally tabular logic), then it has the rpp.
Let \(\mathrm{rp}_m(\lozenge)\) be the modal formula \[p_0\wedge\lozenge(p_1\wedge\lozenge(p_2\wedge\ldots \wedge\lozenge p_{m+1})\ldots )\rightarrow \bigvee\limits_{i<j\leq m+1} \lozenge^i (p_i \wedge p_j) \vee \bigvee\limits_{i< j\leq m} \lozenge^i(p_i \wedge\lozenge p_{j+1}).\]
Proposition 18. For a frame \(F=(X,R)\), \(F\vDash\mathrm{rp}_m(\lozenge)\) iff \(F\) satisfies \(\mathrm{RP}_m\).
Proof. Straightforward. ◻
Since every locally tabular logic \(L\) is the logic of a uniformly tunable class, \(L\) contains \(\mathrm{rp}_m(\lozenge_\mathrm{A})\) for some \(m\).
If reducible path property is sufficient for local tabularity was left as an open question (in view of Theorem 7, it is enough to consider clusters, that is frames of height 1). Below we give a very simple example that shows that it is not sufficient; at the same time, we show that a special form of rpp is sufficient in the case of products.
Proposition 19 (Negative solution to Problem 8.1 in [23]). There is a unimodal cluster \(C\) such that \(\mathrm{RP}_2\) holds in \(C\), while its logic is not locally tabular.
Proof. Let \(C=(\omega,R)\), where \(m R n \text{ iff }m\neq n+1\). Clearly, \(C\) is a cluster: \(m R^2 n\) for all \(m,n\). Consider a path \(m_0 R m_1 R m_2 R m_3\) in \(C.\) If \(m_0,\,\ldots,\,m_3\) are distinct and \(m_0 = m_2 + 1\) then \(m_0 \ne m_3 +1,\) so \(m_0 R m_2\) or \(m_0 R m_3.\) In either case, the path \(m_0 R \ldots R m_3\) can be reduced. Therefore \(C\) satisfies \(\mathrm{RP}_2.\)
Consider a valuation \(\theta\) on \(\omega\) with \(\theta(p) = \{0\}.\) Let \(\varphi_0 = p\), \(\varphi_{n+1} = \lnot\lozenge\varphi_n\). A simple induction on \(n\) shows that \((F,\,\theta),\,n\models \varphi_m\) \(n=m.\) So the logic of \(F\) is not locally tabular (in fact, not 1-finite). ◻
Proposition 20. Assume that \(\mathcal{F}^*\times \mathcal{G}^*\) has the rpp. Then \(\mathcal{F}\) or \(\mathcal{G}\) has the bounded cluster property.
Proof. For the sake of contradiction, assume that for all \(l<\omega\) there are \(F\in \mathcal{F}\) and \(G\in\mathcal{G}\), a cluster \(C\) in \(F\) and a cluster \(D\) in \(G\) such that \(|C|,|D|>l\). Let \(H=F^*\times G^*\). Pick distinct \(x_0,\ldots, x_l\in C\) and distinct \(y_0,\ldots,y_l\in D\) and consider \(a_0,\ldots, a_{2l}\) in \(H\), where \(a_{2i}=(x_i,y_i)\), and \(a_{2i+1}=(x_i,y_{i+1})\), Then the zigzag path \[a_0R_H a_1 R_H a_2 R_H \ldots R_H a_{2l}\] falsifies \(\mathrm{RP}_{2l-1}\) in \(H\). It follows that \(\mathrm{RP}_m\) holds in \(\mathcal{F}\times \mathcal{G}\) for no \(m\), and so \(\mathcal{F}^*\times \mathcal{G}^*\) lacks the rpp. ◻
From Theorem 16 and Proposition 20, we obtain
Corollary 2. If the logics \(\mathop{\mathrm{Log}}\mathcal{F}\) and \(\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular, and \(\mathcal{F}^*\times \mathcal{G}^*\) has the rpp, then \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is locally tabular.
Let \(\mathrm{rp}_m(k,n)\) denote the formula \(\mathrm{rp}_m(\bar{\lozenge})\), where \(\bar{\lozenge}\varphi\) abbreviates \(\lozenge_\mathrm{A}^{\leq k}\varphi\vee \lozenge_\mathrm{B}^{\leq n}\varphi\). These formulas are called product rpp formulas. They give another necessary axiomatic condition for local tabularity. We illustrate it for the bimodal transitive case, while the case of more modalities and pretransitive logics is a straightforward generalization. For a formula \(\varphi\in {\mathrm{\small ML}}(\lozenge)\), let \(t(\varphi)\) be the translation of \(\varphi\) that is compatible with Boolean connectives and satisfies \(t(\lozenge\varphi) = \lozenge_1 \varphi \lor \lozenge_2 \varphi.\) Let \(L_0 = \{\varphi \mid t(\varphi) \in L\}.\) Then \(L_0\) is a normal logic, and it is locally tabular. So it contains an \(\mathrm{rp}_m(\lozenge)\) formula. Hence, \(L\) contains \(\mathrm{rp}_m(1,1) = t(\mathrm{rp}_m(\lozenge)) \in L.\) So we have
Proposition 21. If a bimodal logic \(L\) is locally tabular, then it contains a product rpp formula.
Proposition 22. For a frame \(F=(X,(R_\lozenge)_{\lozenge\in\mathrm{A}\cup\mathrm{B}})\), we have \(F \vDash\mathrm{rp}_m(k,n)\) iff \((X,(\bigcup_{\lozenge\in\mathrm{A}} R_\lozenge)^{\leq k})\cup (\bigcup_{\lozenge\in\mathrm{B}} R_\lozenge)^{\leq n})\) satisfies \(\mathrm{RP}_m\).
Proof. Follows from Proposition 18. ◻
Proposition 23. Formulas \(\mathrm{rp}_m(k,n)\) are canonical.
Proof. These formulas are Sahlqvist. ◻
Proposition 24. Let \(\mathcal{F}\) be a class of \(k\)-transitive \(\mathrm{A}\)-frames, and \(\mathcal{G}\) a class of \(n\)-transitive \(\mathrm{B}\)-frames. If \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is locally tabular, then \(\mathcal{F}^*\times \mathcal{G}^*\) has the rpp, and \(\mathcal{F}\times \mathcal{G}\vDash\mathrm{rp}_m(k,n)\) for some \(m\).
Proof. For a unimodal formula \(\varphi\), we define a translation \([\varphi]^\star\) by compatibility with Boolean connectives and \([\lozenge\psi]^\star=\lozenge_\mathrm{A}^{\leq k}[\varphi]^\star \vee \lozenge_\mathrm{B}^{\leq n}[\varphi]^\star\). Consider frames \(F\) and \(G\), and assume that \(F^*\times G^*\) is \((X\times Y, R_1,R_2)\); then put \(F \star G=(X\times Y,R_1\cup R_2)\).
It is straightforward that for \(k\)-transitive \(F\) and \(n\)-transitive \(G\) we have \[\label{eq:star-RPP} F\times G \vDash[\varphi]^\star \text{ iff }F\star G\vDash\varphi.\tag{4}\] It follows that \(L=\{ \varphi\in {\mathrm{\small ML}}(\lozenge) \mid F\times G \vDash[\varphi]^\star \}\) is a unimodal logic. Since \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is locally tabular, \(L\) is locally tabular as well. Hence the class \(\mathcal{H}=\{F \star G\mid F\in\mathcal{F}\text{ and } G\in \mathcal{G}\}\) has the rpp [23]; by Proposition 22, \(\mathcal{F}\times \mathcal{G}\vDash\mathrm{rp}_m(k,n)\). ◻
It is known that above \({\mathrm{\small S4}}\), 1-finiteness is sufficient for local tabularity [30]. In general, there are 1-finite logics that are not locally tabular [24], [31]. We show that for products of locally tabular logics, \(1\)-finiteness guarantees local tabularity.
Theorem 25. Let \(\mathcal{F},\,\mathcal{G}\) be non-empty classes of frames. If the logics \(\mathop{\mathrm{Log}}\mathcal{F}\) and \(\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular, and \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is \(1\)-finite, then \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is locally tabular.
Proof. Assume that \(\mathop{\mathrm{Log}}\mathcal{F}\) and \(\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular. Let \(\mathcal{C}_1=Cl{(\mathcal{F})}\), \(\mathcal{C}_2=Cl{(\mathcal{G})}\), and let \(\mathcal{R}\) be the class of rectangles \(\boldsymbol{X}_1 \times \boldsymbol{X}_2\) such that \(X_i\) is the domain of \(C_i\in \mathcal{C}_i\), \(i=1,2\).
For the sake of contradiction, assume that \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is not locally tabular. Then \(\mathop{\mathrm{Log}}\mathcal{R}={\mathrm{\small S5}}^2\) according to Theorem 16 and Proposition 14. It is known that \({\mathrm{\small S5}}^2\) is not 1-finite [13]. Hence, for each \(l\), there is a model \(M=(R,\theta)\) on \(R\in \mathcal{R}\) and formulas \(\psi_1, \ldots, \psi_l\) in the single variable \(p\) such that all sets \(\bar{\theta}(\psi_i)\) are pairwise distinct. Consider the frames \(F\) in \(\mathcal{F}\) and \(G\) in \(\mathcal{G}\) with clusters \(C_1\) in \(F\) and \(C_2\) in \(G\) that form the rectangle \(R\). Let \(H\) be the subframe of \(F\times G\) generated by a point in \(C_1\times C_2\). Consider a model \(M'=(H,\theta')\) with \(\theta'(p)=\theta(p)\).
Assume that \(l>2\). Then \(\theta(p)\neq \varnothing\): otherwise, every formula in the single variable \(p\) is equivalent to either \(\bot\) or \(\top\) on \(M\). By Theorem 4, for some \(m,n\), every frame in \(\mathcal{F}\) is \(m\)-transitive and every frame in \(\mathcal{G}\) is \(n\)-transitive. Let \(\bar{\lozenge} p\) abbreviate \(\lozenge_\mathrm{A}^{\leq m} \lozenge_\mathrm{B}^{\leq n} p\). By Proposition 10, we have: \[\text{M',a\vDash\bar{\lozenge} p iff a belongs to C_1\times C_2.}\] For bimodal formulas \(\varphi\) in the single variable \(p\), we define \([\bot]^\star=\bot\), \([p]^\star=p\), \([\psi_1\rightarrow\psi_2]^\star=\bar{\lozenge} p\wedge( [\psi_1]^\star\rightarrow[\psi_2]^\star)\), and \([\lozenge_1 \psi]^\star=\bar{\lozenge} p \wedge\lozenge_\mathrm{A}^{\leq m} [\psi]^\star\), \([\lozenge_2 \psi]^\star=\bar{\lozenge} p\wedge\lozenge_\mathrm{B}^{\leq n} [\psi]^\star\). Then we have \[\text{ M',a\vDash[\varphi]^\star iff (a is in M and M,a\vDash\varphi). }\] So formulas \([\psi_1]^\star, \ldots, [\psi_l]^\star\) are pairwise non-equivalent in \(M'\). It follows that \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is not 1-finite. ◻
We combine our previous observations in the following criteria.
Theorem 26. Let \(\mathcal{F}\) and \(\mathcal{G}\) be non-empty. TFAE:
\(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is locally tabular.
\(\mathop{\mathrm{Log}}\mathcal{F}\) and \(\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular and at least one of the classes \(\mathcal{F}\), \(\mathcal{G}\) has the bounded cluster property.
\(\mathop{\mathrm{Log}}\mathcal{F}\) and \(\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular and \(\mathcal{F}^*\times \mathcal{G}^*\) has the rpp.
\(\mathop{\mathrm{Log}}\mathcal{F}\) and \(\mathop{\mathrm{Log}}\mathcal{G}\) are locally tabular and \(\mathop{\mathrm{Log}}(\mathcal{F}\times \mathcal{G})\) is 1-finite.
Proof. Theorem 16 shows the equivalence between [cr-i0] and [cr-i1]. By Proposition 24, [cr-i0] implies [cr-i2]. By Proposition 20, [cr-i2] implies [cr-i1].
By Theorem 25, [cr-i4] implies [cr-i0]; the converse is trivial. ◻
Let \(F=(X,R)\) be a preorder whose skeleton is converse well-founded. Observe that \(F\) has no clusters of size greater than \(n\) iff \(F{\upharpoonright}Y\twoheadrightarrow\boldsymbol{n+1}\) for no \(Y\subseteq \mathop{\mathrm{dom}}F\). The latter property is modally definable [18]. It follows that in the unimodal transitive frames of finite height, the bounded cluster property can be expressed by modal formulas \(\mathrm{bc}_n\), where \(n\) bounds the size of clusters. Consequently, in the case of \(k\)-transitive \(\mathrm{A}\)-frames, the corresponding property is expressed by formulas \({ [\mathrm{bc}_n ]}^{k}\), which will be called bounded cluster formulas.
For a pretransitive logic \(L\), let \(\mathrm{tra}{(L)}\) be the least \(k\) such that \(L\) is \(k\)-transitive; \(k\) is called the pretransitivity index of \(L\). By Theorem 4, \(\mathrm{tra}{(L)}\) is defined for every locally tabular logic.
Corollary 3. Let \(L_1, L_2\) be Kripke complete consistent logics. TFAE:
\(L_1\times L_2\) is locally tabular.
\(L_1\) and \(L_2\) are locally tabular and at least one of them contains a bounded cluster formula \({ [\mathrm{bc}_n ]}^{k}\), where \(k\) is the pretransitivity index of this logic.
\(L_1\) and \(L_2\) are locally tabular and \(L_1\times L_2\) contains a product rpp formula \(\mathrm{rp}_m(k_1,k_2)\), where \(k_i=\mathrm{tra}{(L_i)}\).
It is well known that for unimodal logics above \({\mathrm{\small K4}}\), local tabularity is equivalent to finite height [4],[5]. This criterion was extended for weaker systems in [23]: it holds for logics containing \(\lozenge^{k+1} p\rightarrow\lozenge p\vee p\) with \(k>0\).
Corollary 4. Let \(L_1, L_2\) be Kripke complete consistent unimodal logics, and for some \(k,n>0\), \(L_1\) contains \(\lozenge^{k+1} p\rightarrow\lozenge p\vee p\) and \(L_2\) contains \(\lozenge^{n+1} p\rightarrow\lozenge p\vee p\). In this case, \(L_1\times L_2\) is locally tabular iff \(L_1\) and \(L_2\) contain formulas of finite height and for some \(m\), \(L_1\times L_2\) contains the product rpp formula \(\mathrm{rp}_m(k,n)\).
For a transitive logic \(L\), let \(L[h]\) be its extension with the axiom \(B_h\) of finite height.
Let \({\mathrm{\small GL}}\) and \({\mathrm{\small Grz}}\) be the logics of converse well-founded partial orders, strict and non-strict, respectively. Let \({\mathrm{\small GL}}.3\) and \({\mathrm{\small Grz}}.3\) be their extensions for the case of linear orders. Clearly, these logics have the bounded cluster properties.
Corollary 5. For any locally tabular \(L\), and any finite \(h\), the logics \({\mathrm{\small Grz}}[h]\times L\), \({\mathrm{\small GL}}[h]\times L\) are locally tabular. Consequently, \(L'\times L\) is locally tabular for any extension \(L'\) of \({\mathrm{\small Grz}}[h]\) or \({\mathrm{\small GL}}[h]\). In particular, \({\mathrm{\small Grz}}.3[h]\times L\), \({\mathrm{\small GL}}.3[h]\times L\) are locally tabular.
It is known [32] that if an \(\mathrm{A}\)-logic \(L\) contains a formula \(\Box_\mathrm{A}^m\bot\) for some \(m\), then this logic is locally tabular. In [7] (see also [8]), it was shown that \(L\times L\) is locally tabular. We notice that the bounded cluster property holds for \(L\)-frames: in this case, clusters are singletons. Hence, we have the following generalization of the above result.
Corollary 6. If \(L_1\) contains a formula \(\Box_\mathrm{A}^m\bot\) and \(L_2\) is locally tabular, then \(L_1\times L_2\) is locally tabular.
A modal logic \(L\) has the product fmp, if \(L\) is the logic of a class of finite product frames. The product fmp is stronger than the fmp: for example, \({\mathrm{\small K4}}\times {\mathrm{\small S5}}\) has the fmp [14], but lacks the product fmp [3]. Examples of logics with this property are also known: they include \({\mathrm{\small K}}\times {\mathrm{\small K}}\) (where \({\mathrm{\small K}}\) stands for the smallest unimodal logic) and \({\mathrm{\small K}}\times{\mathrm{\small S5}}\) [3], the logic \({\mathrm{\small S5}}\times {\mathrm{\small S5}}\) [29], or its extensions [17].
The product of a logic possesing the fmp with a tabular logic has the product fmp ([6], see Proposition 32 below). In this section, we show that the weaker property of local tabularity is not sufficient for the product fmp.
We define the saw frame \(F_S = (W,S,S_l,S_r)\) for the alphabet \(\mathrm{A}= \{\lozenge,\,\lozenge_l,\,\lozenge_r\}\) as follows: \[\begin{gather} W = \{u\} \sqcup \{v_i\}_{i < \omega} \sqcup \{w_i\}_{i < \omega}; \\ S = \{u\} \times \{v_i\}_{i < \omega}; \\ S_l = \{(v_i,w_i)\}_{i < \omega}; \\ S_r = \{(v_i,w_{i+1})\}_{i < \omega}. \end{gather}\] The saw frame is shown in Figure 1.
Let \({\mathrm{\small Saw}}= \mathop{\mathrm{Log}}(F_S).\) The next proposition follows from simple semantic observations:
Proposition 27. The following formulas are theorems of \({\mathrm{\small Saw}}\): \[\begin{gather} \Box_\mathrm{A}^3 \bot; \label{eq:saw95deadend} \\ \Box_l \Box_\mathrm{A}\bot; \label{eq:saw95l95deadend} \\ \Box_r \Box_\mathrm{A}\bot; \label{eq:saw95r95deadend} \\ \lozenge_l p \to \Box_l p; \label{eq:saw95l95functional} \\ \lozenge_r p \to \Box_r p; \label{eq:saw95r95functional} \\ \lozenge\lozenge_l p \to \lozenge\lozenge_r p; \label{eq:saw95expand} \end{gather}\] {#eq: sublabel=eq:eq:saw95deadend,eq:eq:saw95l95deadend,eq:eq:saw95r95deadend,eq:eq:saw95l95functional,eq:eq:saw95r95functional,eq:eq:saw95expand}
Proposition 28. \({\mathrm{\small Saw}}\times {\mathrm{\small S5}}\) is locally tabular.
Proof. Follows from Corollary 6 by ?? . ◻
We define the formula \(\varphi\in {\mathrm{\small ML}}(\mathrm{A}\cup \{\lozenge_{{\mathrm{\small S5}}}\})\) as \(\varphi = \varphi_1 \land \varphi_2 \land \varphi_3,\) where \[\begin{gather} \varphi_1 = \lozenge\top; \\ \varphi_2 = \Box \lozenge_{{\mathrm{\small S5}}} \left(\lozenge_l \lnot p \land \lozenge_r p\right); \\ \varphi_3 = \Box \Box_{{\mathrm{\small S5}}} \left(\lozenge_l p \to \lozenge_r p\right). \end{gather}\]
Proposition 29. The formula \(\varphi\) is consistent with \({\mathrm{\small Saw}}\times {\mathrm{\small S5}}.\)
Proof. It suffices to show that \(\varphi\) is satisfiable in \(F_S\times \boldsymbol{\omega}.\) Let \(\theta\) be a valuation in \(F_S\times \boldsymbol{\omega}\) given by \(\theta(p) = \{(w_i,j) \mid i < \omega,\,j < i\}\). Then a direct evaluation shows that \(F_S\times \boldsymbol{\omega},\theta,(u,0) \models \varphi.\) ◻
Proposition 30. If \(\varphi\) is satisfiable in a \({\mathrm{\small Saw}}\times {\mathrm{\small S5}}\)-frame \(F\times G\), then \(F\times G\) is infinite.
Proof. Let \({\mathrm{\small Saw}}\) be valid in a rooted frame \(F = (X,R,R_l,R_r)\) with a root \(r\) and let \(G = (Y,Y\times Y).\) Assuming that \(\varphi\) is true at \(F\times G,\theta,(r,s)\) for some valuation \(\theta\) and a point \(s \in Y,\) we will show that \(X\) is infinite.
We define the mapping \(V:\:X \to Y\) by \(V(a) = \{b\in Y \mid (a,b)\in \theta(p)\}.\) Let us show by induction that there exist countable subsets \(\{m_i\}_{i < \omega}\) and \(\{t_i\}_{i < \omega}\) of \(X\) such that the following are true for all \(i < \omega\): \[\begin{gather} r R m_i; \\ R_l(m_i) = \{t_i\}; \tag{5} \\ R_r(m_i) = \{t_{i+1}\}; \tag{6} \\ R_F(t_i) = \varnothing; \tag{7} \\ V(t_i) \subsetneq V(t_{i+1}) \tag{8}; \\ m_i \not\in \{m_j\}_{j < i}. \end{gather}\]
For the base case, observe that \(\varphi_1\) is true at \((r,s),\) so \((r,s) R^h (m_0,s)\) for some \(m_0 \in X.\) Then \(r R m_0.\)
For the transition, we assume that \(\{m_j\}_{j \le i}\) and \(\{t_j\}_{j < i}\) are already constructed. Then \((r,s) R^h (m_i,s)\), so \(\lozenge_{{\mathrm{\small S5}}} \left(\lozenge_l \lnot p \land \lozenge_r p\right)\) is true at \((m_i,s),\) so there exists \(b_i\in Y\) such that \((m_i,b_i) R_l^h (t_i,b_i)\) and \((m_i,b_i) R_r^h (t_{i+1},b_i)\) for some \(t_i\) and \(t_{i+1}\) in \(X\) such that \((t_i,b_i)\not\in \theta(p)\) and \((t_{i+1},b_i)\in \theta(p).\) By ?? and ?? , \(R_l(m_i) = \{t_i\}\) and \(R_r(m_i) = \{t_{i+1}\}\). Since \(\varphi_3\) is true at the root, \(V(t_i) \subseteq V(t_{i+1}).\) The inclusion is strict because \(b_i \in V(t_{i+1})\setminus V(t_i).\) By ?? and ?? we also have \(R_F(t_i) = R_F(t_{i+1}) = \varnothing.\) Notice that \(r R m_i R_r t_{i+1},\) so by ?? there exists \(m_{i+1}\in X\) such that \(r R m_{i+1} R_l t_{i+1}.\) By 8 , \(t_{i+1} \ne t_j\) for all \(j \le i.\) Then by 5 \(t_{i+1}\not\in R_l(m_j)\), so \(m_{i+1} \ne m_j\) for \(j \le i.\) The induction is complete.
It follows from 8 that \(t_i\) are pairwise distinct for all \(i < \omega.\) Then \(X\) contains an infinite subset \(\{t_i\}_{i < \omega}\), and therefore \(X\) is infinite. ◻
Theorem 31. The local tabularity of a product logic does not imply the product fmp.
As follows from Theorem 4 and Corollary 3, it is necessary for any locally tabular product of two logics that both have a finite height and one of them contains a bounded cluster formula. We can classify the locally tabular logics \(L_1 \times L_2\) with respect to the height of \(L_1\) and \(L_2\). For this section, we will assume that \(L_1 \vdash [bc_n]^{\mathrm{tra}{(L_1)}}\) for some \(n < \omega\). We will describe how the product fmp depends on the height of \(L_1\) and \(L_2.\)
By Proposition 31, the product fmp fails even when \(h(L_2) = 1\). Thus we will focus on the height of \(L_1.\)
If \(L_1\) has unit height, then the product fmp follows from the following result. A logic is called tabular, if it is the logic of a single finite frame.
Proposition 32. [6]If \(L\) is a tabular logic and \(L'\) has the fmp, then \(L\times L'\) has the product fmp.
Observe that \(L_1\) is Kripke complete by 1. Since \(L_1\) is locally tabular, it is the logic of its class of frames. Every \(L_1\)-frame has unit height, so it consists of one cluster, which has size at most \(n\) by the frame condition of \([bc_n]^{\mathrm{tra}{(L)}}\). Then \(L_1\) is tabular by Proposition 32.
Corollary 7. If \(h(L_1) = 1,\) then \(L_1 \times L_2\) has the product fmp.
Proposition 31 shows that the product fmp can fail when \(h(L_1)=3.\) The tabularity of \(L_1\) or \(L_2\) is a sufficient condition for the product fmp in this case, as well as for greater \(h(L_1)\). Describing weaker sufficient conditions is an open problem.
Finally, the case when \(L_1\) has height \(2\) is the most intriguing. Neither a proof, nor a counterexample for the product fmp of such logics is known yet.
Recall that a logic is tabular, if it is characterized by a single finite frame. A non-tabular logic is pretabular, if all of its proper extensions are tabular. There are exactly five pretabular logics above \({\mathrm{\small S4}}\), see [15], [16], or [18]. One of them is the logic \({\mathrm{\small S5}}\). Another is the logic \({\mathrm{\small Tack}}\) defined below.
An important fact about \({\mathrm{\small S5}}^2\) was established in [17]: this logic is prelocally tabular, that is, the following theorem holds.
Theorem 33. [17]All proper extensions of \({\mathrm{\small S5}}^2\) are locally tabular.
While \({\mathrm{\small S5}}^2\) is prelocally tabular, it turns out that products of pretabular logics can be not prelocally tabular even for the case of height 2.
The (one-dimensional) tack frame \(\mathrm{\small T}\) is the ordered sum \((\omega+1,R)\) of a countable cluster and a singleton, that is \(\alpha R \beta\) iff \(\alpha<\omega\) or \(\beta=\omega\). Let \({\mathrm{\small Tack}}\) be the logic of its frame. \({\mathrm{\small Tack}}\) is one of exactly five pretabular logics above \({\mathrm{\small S4}}\).
In this subsection we show that \({\mathrm{\small Tack}}\times {\mathrm{\small S5}}\) is not prelocally tabular, and give an axiomatic criterion of local tabularity for a family of logics that contains \({\mathrm{\small Tack}}\times {\mathrm{\small S5}}\) and its extensions.
Proposition 34. \({\mathrm{\small Tack}}\times {\mathrm{\small S5}}\) is not prelocally tabular.
Proof. Let \(H=T\times \boldsymbol{\omega}\). The formula \(\varphi=\lozenge_2\lozenge_1\Box_1 p\rightarrow\Box_1\lozenge_1 p\) is not valid in \(H\): consider a model \(M=(H,\theta)\) with \(\theta(p)=\{(\omega,1)\}\); then \(\varphi\) is falsified in \(M\) at \((0,0)\).
Consider the equivalence \(\equiv\) on \(\mathop{\mathrm{dom}}H\) that extends the diagonal by the set of pairs of form \(((\omega,m),(\omega,n))\). It is easy to check that \(H \twoheadrightarrow{H}{/}{\equiv}\), hence the logic \(L\) of the latter frame extends \({\mathrm{\small Tack}}\). It is also straightforward that \(\varphi\) is valid in \({H}{/}{\equiv}\), so \(L\) is a proper extension of \({\mathrm{\small Tack}}\). Finally, we observe that \(L\) is not locally tabular: \({H}{/}{\equiv}\) contains \(\boldsymbol{\omega} \times \boldsymbol{\omega}\) as a subframe, and the logic \({\mathrm{\small S5}}^2\) of this subframe is not locally tabular. ◻
We say that the product rpp criterion holds for a logic \(L\), if for any normal extension \(L'\) of \(L\), we have:
\(L'\) contains a product rpp formula iff \(L'\) is locally tabular.
Recall that \(L[h]\) denotes the extension of \(L\) with the axiom \(B_h\) of finite height.
The unimodal logic \({\mathrm{\small S4.1}}\) is the extension of \({\mathrm{\small S4}}\) with the axiom \(\Box\lozenge p\rightarrow\lozenge\Box p\). In transitive frames, \(\Box\lozenge p\rightarrow\lozenge\Box p\) defines the first-order property \[\label{eq:s4461-condition} \forall x\,\exists y\,\left(xRy \land \forall z \left(yRz \to z = y\right)\right),\tag{9}\] see, e.g., [19].
Hence, in our notation, the unimodal logic \({\mathrm{\small S4.1}}[2]\) is the extension of \({\mathrm{\small S4}}\) with the axioms \(\Box\lozenge p\rightarrow\lozenge\Box p\) and \(B_2\). Now we are going to show that the product rpp criterion holds for \({\mathrm{\small S4.1}}[2] \times {\mathrm{\small S5}}\).
Lemma 1. Let \(F=(X,R,S)\) validate a product \(L_1 \times L_2\) of transitive logics, where \(L_1\) has finite height. Then \(R\cap S^{-1}\subseteq R^{-1}.\)
Proof. Let \(\lozenge\varphi\) abbreviate \(\lozenge_1 \varphi \lor \lozenge_2 \varphi,\) and let \(\varphi\) be \[p \land \Box^{\leq 2} (p\to \lozenge_1 q ) \land \Box^{\leq 2}(q \to \lozenge_2 p) \to \lozenge^{\le 2} (q \land \lozenge_1 p).\]
We claim that \(\varphi\in L_1\times L_2.\) Indeed, let the antecedent of \(\varphi\) be satisfied at a point\((a_0,b_0)\) of a model \((G\times H,\theta),\) where \(G\models L_1\) and \(H\models L_2.\) Then \(G\times H\) is \(m\)-transitive, so \(p\to \lozenge_1 q\) and \(q\to \lozenge_2 p\) are true at any point in \(R_{G\times H}^*(a_0,b_0).\) Consequently, there exists a sequence \(\{(a_i,b_i)\}_{i < \omega}\) of points in \(G\times H\) such that \((a_i,b_i) \in \theta(p),\) \((a_{i+1},b_i) \in \theta(q)\) and \((a_i,b_i) R^h (a_{i+1},b_i) R^v (a_{i+1},b_{i+1})\) for all \(i < \omega.\) Then \(a_i R_G a_{i+1}\) for all \(i.\) Since \(L_1\) has finite height, \(a_{i+1} R_G a_i\) for some \(i < \omega.\) Then \((a_{i+1},b_i) R^h (a_i,b_i),\) thus \(q \land \lozenge_1 p\) is true at \((a_{i+1},b_i).\) Then the consequent of \(\varphi\) is true at \((a_0,b_0).\)
Now let \(a R b\) and \(b S a\) for some \(a\) and \(b\) in \(F.\) Let \(\theta\) be a valuation on \(F\) such that \(\theta(p) = \{a\}\) and \(\theta(q) = \{b\}.\) Then \((F,\theta),a\models \varphi\) since \(\varphi \in L_1 \times L_2\), and therefore \(b R a.\) ◻
Remark 35. The formula \(\varphi\) from the proof of this lemma shows that the pairs of logics \(L_1,\,L_2\), where both \(L_1\) and \(L_2\) contain \({\mathrm{\small S4}}\), \(1 < h(L_1) < \omega\) and \(p \to \Box p \not \in L_2\), are not product-matching in the sense of [3]. Indeed, consider \(F_1 = (\{0,1\},\le,\ge)\) and \(F_2 = (\{0,1\},\le,\{0,1\}^2).\) Then \(\varphi\) is refuted in both \(F_1\) and \(F_2\). This observation extends the one of [32] (or cf. [3]), where a similar reasoning describes the case where \({\mathrm{\small Grz}}\subseteq L_1\).
A point \(a\in X\) is called \(R\)-terminal in a transitive frame \((X,R)\), if \(R(a) = \{a\}\).
::: {#lem:S4.1[h]xS5_terminal .lemma} Lemma 2. Let \(F = (X, R, \approx)\models{{\mathrm{\small S4.1}}[h]\times {\mathrm{\small S5}}}\). If \(a\in X\) is \(R\)-terminal, then \([a]_\approx\) contains only \(R\)-terminal elements. :::
Proof. Let \(a\) be \(R\)-terminal and consider \(b\in [a]_\approx.\) By the frame condition of \({\mathrm{\small S4.1}}[h],\) there exists an \(R\)-terminal element \(c\in R(b)\). Then \(a \approx b R c\), so by commutativity \(a R d \approx c\) for some \(d\in X.\) But since \(a\) is \(R\)-terminal, \(a=d\), and so \(a \approx c.\) Hence, \(b\approx c\). Then \((b,c)\in R \cap {\approx},\) so \(c R b\) by Lemma 1. Finally, \(b = c\) because \(c\) is \(R\)-terminal. ◻
Let \(\mathcal{F}_h\) be the class of all \({\mathrm{\small S4.1}}[h]\times {\mathrm{\small S5}}\)-frames that do not validate \({\mathrm{\small S5}}^2.\) For any \(F = (X,R,\approx)\in \mathcal{F}_h,\) let \(Z_F\) be the set of all \(R\)-terminal elements in \(F.\) We define \(\alpha(F)\) to be the subframe \(F {\upharpoonright}\left(X \setminus Z_F\right).\) Notice that if \(Z_F = X,\) then \(R\) is an equivalence relation, so \(F \models {\mathrm{\small S5}}^2.\) It follows that \(\alpha(F)\) is not empty for any \(F\in \mathcal{F}_h.\)
Let \(L_1\) be an \(\mathrm{A}\)-logic and \(L_2\) be an \(\mathrm{B}\)-logic. The fusion \(L_1*L_2\) is the smallest \((\mathrm{A}\cup \mathrm{B})\)-logic that contains \(L_1 \cup L_2.\) The commutator \(\left[L_1,L_2\right]\) is the smallest logic that contains \(L_1*L_2\) and the axioms \(\mathrm{com}(a,b),\,\mathrm{com}(b,a)\) and \(\mathrm{chr}(a,b)\) for all \(a\in \mathrm{A}\) and \(b\in\mathrm{B},\) where \[\begin{align} \mathrm{com}(a,b) &= \lozenge_a \lozenge_b p \to \lozenge_b \lozenge_a p; \\ \mathrm{chr}(a,b) &= \lozenge_a \Box_b p \to \Box_b \lozenge_a p. \end{align}\] The validity of \(\mathrm{com}(a,b)\) in an \((\mathrm{A}\cup\mathrm{B})\)-frame \((X,(R_a)_{a\in \mathrm{A}}, (R_b)_{a\in \mathrm{B}})\) is equivalent to the commutativity \(R_a\circ R_b = R_b \circ R_a.\) The formula \(\mathrm{chr}(a,b)\) defines the Church-Rosser property \[\forall x\,\forall y\,\forall z\,(x R_a y\,\&\,x R_b z \to \exists u\,(y R_b u\,\&\,z R_a u))\] For any \(L_1\) and \(L_2,\) the product logic \(L_1\times L_2\) is a normal extension of \(\left[L_1,L_2\right]\) [3].
Proposition 36. For any \(F\in \mathcal{F}_h,\) where \(h < \omega,\) the logic \(\left[{\mathrm{\small S4}}[h-1],{\mathrm{\small S5}}\right]\) is valid in \(\alpha(F)\).
Proof. The validity of \({\mathrm{\small S4}}\) and \({\mathrm{\small S5}}\) is defined by universal sentences, so \(\alpha(F)\) validates the fusion \({\mathrm{\small S4}}*{\mathrm{\small S5}}\), since \(\alpha(F)\) is a subframe of \(F.\)
Let \(S\) be an \(R\)-chain in \(\alpha(F).\) Then \(S\) is also a chain in \(F.\) Observe that \(F\models B_h,\) then \(F\) contains \(R\)-chains of size at most \(h.\) We claim that \(|S|<h.\) Since \(F\models {\mathrm{\small S4.1}},\) any chain of size \(h\) in \(F\) contains an \(R\)-terminal element, which belongs to \(Z_F\). But \(S \subseteq W \setminus Z_F,\) so \(|S| < h.\) Since \(S\) was arbitrary, \(\alpha(F) \models B_{h-1}.\)
Let us show the commutativity. Let \(a R b \approx c\) in \(\alpha(F).\) Then the same holds in \(F\), so there exists \(d\) in \(F\) such that \(a \approx d R c.\) Then \(d\) is not \(R\)-terminal since \(d R c\) and \(c\not \in Z_F.\) Then \(d\) belongs to \(\alpha(F).\)
Let \(a \approx b R c\) in \(\alpha(F).\) There exists \(d\) in \(F\) such that \(a R d \approx c.\) By Lemma 2, \(d\not\in Z_F\), so \(d\) is in \(\alpha(F).\)
Finally, we show the Church-Rosser property. If \(a R b\) and \(a \approx c\) in \(\alpha(F),\) then \(b \approx d\) and \(c R d\) for some \(d\) in \(F.\) Then \(d \not\in Z_F\) since \(b\not\in Z_F\) while \(b \in [d]_\approx.\) Thus \(d\) belongs to \(\alpha(F)\) by Lemma 2. ◻
Proposition 37. Let \(h < \omega.\) If \(F \subseteq \mathcal{F}_h,\) then \(F\) validates a product rpp formula iff \(\alpha(F)\) does.
Proof. For the ‘if’ direction, observe that for any \(a,\,b,\,c\in Z_F\) that form a path we have \(a \approx c.\) Assume \(\alpha(F)\models \mathrm{rp}_m(1,1).\) It is straightforward that \(F\models \mathrm{rp}_{m+2}(1,1)\). The ‘only if’ direction is true since \(\mathrm{RP}_n\) is a universal sentence for any \(n.\) ◻
Proposition 38. For any class of frames \(\mathcal{F}\subseteq \mathcal{F}_h,\) \(\mathop{\mathrm{Log}}\mathcal{F}\) is locally tabular iff \(\mathop{\mathrm{Log}}\alpha[\mathcal{F}]\) is locally tabular.
Proof. The ‘only if’ direction follows from Proposition 6.
Assume that \(\mathop{\mathrm{Log}}\alpha[\mathcal{F}]\) is locally tabular. The class \(\mathcal{F}\) has uniformly finite height. By Theorem 7, we only need to show that \(Cl{(\mathcal{F})}\) is uniformly tunable. If \(C\in Cl{(\mathcal{F})}\), then either \(C\in Cl{(\alpha[\mathcal{F}])}\) or \(C\cong \boldsymbol{1} \times \boldsymbol{S}\) for a set \(S\). Trivially, the class of rectangle frames of the form \(\boldsymbol{1} \times \boldsymbol{S}\) is uniformly tunable. Hence, \(Cl{(\mathcal{F})}\) is the union of two uniformly tunable classes, and hence is uniformly tunable as well.2 ◻
Lemma 3. If the product rpp criterion holds for \(\left[{\mathrm{\small S4}}[h],{\mathrm{\small S5}}\right]\), then it also holds for \({\mathrm{\small S4.1}}[h+1] \times {\mathrm{\small S5}},\) for any \(h < \omega.\)
Proof. Let \(h < \omega\). Assuming the product rpp criterion for \({\mathrm{\small S4}}[h]\times {\mathrm{\small S5}}\), we will show it for \({\mathrm{\small S4.1}}[h+1]\times {\mathrm{\small S5}}.\)
If \(L\) is a locally tabular normal extension of \({\mathrm{\small S4.1}}[h+1]\times {\mathrm{\small S5}},\) then it contains a product rpp formula by Proposition 21.
Conversely, let \(L\) be a normal extension of \({\mathrm{\small S4.1}}[h+1]\times {\mathrm{\small S5}}\) containing a product rpp formula. If \(L\) is an extension of \({\mathrm{\small S5}}^2\), then it is a proper extension of \({\mathrm{\small S5}}^2\), and so is locally tabular by Theorem 33.
Otherwise the canonical frame \(F_L\) of \(L\) does not validate \({\mathrm{\small S5}}^2\), since \({\mathrm{\small S5}}^2\) is canonical [29]. By Proposition 23, \(F_L\) validates a product rpp formula. By Proposition 37, \(\alpha(F_L)\) also validates a product rpp formula. Then by the assumption \(\mathop{\mathrm{Log}}\alpha(F_L)\) is locally tabular. By Proposition 38, so is \(\mathop{\mathrm{Log}}F_L \subseteq L.\) Then \(L\) is locally tabular. ◻
The product rpp criterion holds for \({\mathrm{\small S5}} \times {\mathrm{\small S5}}\). Indeed, \({\mathrm{\small S5}}\times {\mathrm{\small S5}}\) has no product rpp, and is not locally tabular. All extensions of \({\mathrm{\small S5}}\times {\mathrm{\small S5}}\) are locally tabular, and hence they have the product rpp by Proposition 21. It is well known that \({\mathrm{\small S5}} \times {\mathrm{\small S5}}=\left[{\mathrm{\small S5}},{\mathrm{\small S5}}\right]\) [29]. Hence, we have
Theorem 39. The product rpp criterion holds for \({\mathrm{\small S4.1}}[2]\times {\mathrm{\small S5}}.\)
A prelocally tabular logic \({\mathrm{\small S5}}\times {\mathrm{\small S5}}\) has height \(1\). In this subsection, we give another example of a prelocally tabular logic above \({\mathrm{\small S4}}\times {\mathrm{\small S5}}\), which has height \(2\).
We define the semi-ordered sum \(F \oplus G\) of disjoint bimodal frames \(F = (X,R_1,R_2)\) and \(G = (Y,S_1,S_2)\) as \((X\cup Y, R_1 \cup S_1 \cup (X \times Y), R_2 \cup S_2).\) Let \(\circ\) denote the reflexive, with respect to both relations, singleton \((\{0\},R,S)\) with \(R = S = \{(0,0)\}.\)
Let \(\mathrm{\small T}(X,Y)={{(\boldsymbol{X} \times \boldsymbol{Y})}\,\oplus\, \circ}\). The two dimensional tack is the frame \(\mathrm{\small T}(\omega,\omega)\). Let \({\mathrm{\small Tack}}_2 = \mathop{\mathrm{Log}}{\mathrm{\small T}(\omega,\omega)}.\) We show that \({\mathrm{\small Tack}}_2\) is canonical, has the finite model property, and is prelocally tabular.
Firstly, we identify some axioms of \({\mathrm{\small Tack}}_2\), study its abstract frames, and then give a complete axiomatization.
Notice that the composition of relations \(R=R_1\circ R_2\) in \(\mathrm{\small T}(\omega,\omega)\) is a preorder. Let \(\lozenge\) denote the compound modality \(\lozenge_1\lozenge_2\).
Observe that the skeleton of \(\mathrm{\small T}(\omega,\omega)\) is a two-element chain. Hence, \({\mathrm{\small Tack}}_2\) contains the axiom of height two for the compound modality \(\lozenge\). Moreover, the height of the first relation in \(\mathrm{\small T}(\omega,\omega)\) is 2, so the fragment of \({\mathrm{\small Tack}}_2\) in the first modality contains \({\mathrm{\small S4}}[2]\). Clearly, its fragment in the second modality is \({\mathrm{\small S5}}\).
Moreover, our observation about the skeleton being the two-element chain implies the following modally definable properties. Clearly, the composition \(R_1\circ R_2\) in \(\mathrm{\small T}(\omega,\omega)\) satisfies the Church-Rosser property, and hence, the formula \(\lozenge\Box p\rightarrow\Box \lozenge p\), which is an abbreviation for \[\label{eq:TL-CR} \lozenge_1\lozenge_2 \Box_1\Box_2 p\rightarrow\Box_1\Box_2 \lozenge_1\lozenge_2 p,\tag{10}\] is a valid principle of \({\mathrm{\small Tack}}_2\). In the combination with height 2, it follows that in every rooted \({\mathrm{\small Tack}}_2\)-frame, there is a top cluster. Next, observe that \(R\) satisfies the condition 9 . It follows that the formula \(\Box\lozenge p\rightarrow\lozenge\Box p\), which is an abbreviation for \[\label{eq:TackMC} \Box_1\Box_2\lozenge_1\lozenge_2 p\rightarrow\lozenge_1\lozenge_2\Box_1\Box_2 p,\tag{11}\] belongs to the logic \({\mathrm{\small Tack}}_2\). In particular, it follows that in every rooted \({\mathrm{\small Tack}}_2\)-frame, the top cluster is a singleton.
Our next step aims to express the symmetry property in the bottom cluster of an arbitrary rooted \({\mathrm{\small Tack}}_2\)-frame (notice that in clusters of \({\mathrm{\small S4}}\times {\mathrm{\small S5}}\)-frames, the symmetry of the first relation is not guaranteed). Consider the following relativized version of the symmetry axiom: \[\label{eq:relativized-sym} \lozenge_1\Box_1 q \wedge\neg q\wedge p\rightarrow\Box_1 (\neg q\rightarrow\lozenge_1 p)\tag{12}\] Readily, this formula is valid in \(\mathrm{\small T}(\omega,\omega)\). Let \(F\) be a point-generated \({\mathrm{\small Tack}}_2\)-frame of height 2. We claim that its rooted cluster \(C\), considered as a frame-restriction, validates \({\mathrm{\small S5}}\times {\mathrm{\small S5}}\). Indeed, assume that \(x,y\in C\) are connected by the first relation \(R_1\). Then consider a valuation \(\theta\) in \(F\) with \(\theta(q)\) being the top cluster, and \(\theta(p)=\{x\}\). The validity of 12 implies that \(yR_1 x\), which proves the claim.
Let \(L\) be the extension of the commutator \([{\mathrm{\small S4}}[2], {\mathrm{\small S5}}]\) with the axioms 10 , 11 , and 12 . Putting our observations together, we obtain
Lemma 4. Let \(F\) be a rooted \(L\)-frame. Then \(F\) is isomorphic to \(C\oplus \circ\) for an \({\mathrm{\small S5}}^2\)-frame \(C\), or to \(\circ\).
Our next goal is to prove completeness.
Lemma 5. Assume that \(\mathop{\mathrm{Log}}(\mathcal{C})=\mathop{\mathrm{Log}}(\mathcal{D})={\mathrm{\small S5}}^2\) for classes \(\mathcal{C}\), \(\mathcal{D}\) of frames. Then \(\mathop{\mathrm{Log}}\{{{C}\,\oplus\, \circ}\mid C\in \mathcal{C}\} =\mathop{\mathrm{Log}}\{{{D}\,\oplus\, \circ}\mid D\in \mathcal{D}\}\).
This lemma follows from [33]. The semi-ordered sum \({{F}\,\oplus\, \circ}\) is the sum of frames \(F\) and the singleton \(\circ\) over the indexing frame \((2,\leq,\varnothing)\) in the sense of [33]. Two classes are interchangeable, if they have the same logic in the language enriched with the universal modality. It was shown that in a class of sums, replacing interchangeable classes of summands does not change the logic of the sums. Classes \(\mathcal{C}\) and \(\mathcal{D}\) are interchangeable, since the universal modality is expressible as \(\lozenge_1\lozenge_2\) in \({\mathrm{\small S5}}^2\)-frames. Hence \(\mathop{\mathrm{Log}}\{{{C}\,\oplus\, \circ}\mid C\in \mathcal{C}\} =\mathop{\mathrm{Log}}\{{{D}\,\oplus\, \circ}\mid D\in \mathcal{D}\}\). See Appendix, Section 9.2, for details.
It follows from Lemma 5 that \({\mathrm{\small Tack}}_2\) has an analog of the product fmp:
Corollary 8. \({\mathrm{\small Tack}}_2\) is the logic of \(\{\mathrm{\small T}(m,m) \mid m < \omega\}\).
Theorem 40.
Proof. Let \(L\) be the mentioned axiomatic extension of \([{\mathrm{\small S4[2]}}, {\mathrm{\small S5}}]\).
Firstly, we show that \(L\) is canonical. The axioms of \([{\mathrm{\small S4}}[2], {\mathrm{\small S5}}]\) are known to be canonical, in particular, canonicity of the finite height formula is given in [4]. It is immediate that 12 is equivalent to the Sahlqvist formula \[\lozenge_1\Box_1 q \wedge p\rightarrow\Box_1( q \vee \lozenge_1 p)\vee p,\] and so is canonical. The formula \(\lozenge\Box p \rightarrow\Box \lozenge p\) is also Sahlqvist.
The formula \(\Box\lozenge p \rightarrow\lozenge\Box p\) is not a Sahlqvist formula. However, it is well-known that it is canonical in the transitive case [34], which, in our setting, pertains to the composition \(R_1\circ R_2\). Hence, we have a singleton cluster above every point in the canonical frame of \(L\).
It follows that \(L\) is canonical.
Let us show that \(L={\mathrm{\small Tack}}_2\). Since \(L\) is canonical, it is Kripke complete, and so, in view of Lemma 4, \(L\) is the logic of the class \(\{{{C}\,\oplus\, \circ}\mid C\vDash{\mathrm{\small S5}}^2\}\). From Lemma 5, the logic of this class is \({\mathrm{\small Tack}}_2\). ◻
Lemma 6. \({\mathrm{\small Tack}}_2\) is not locally tabular.
Proof. The frame \(\mathrm{\small T}(\omega,\omega)\) contains \(\boldsymbol{\omega} \times \boldsymbol{\omega}\) as a subframe. Now the statement follows from Proposition 6. ◻
Lemma 7. Let \(\mathcal{F}\) be a class of bimodal frames such that \(\mathop{\mathrm{Log}}\mathcal{F}\) is a proper extension of \({\mathrm{\small Tack}}_2\). Then \(\mathop{\mathrm{Log}}\mathcal{F}\) is locally tabular.
Proof. Let \(\mathcal{G}\) be the class of point-generated subframes of frames in \(\mathcal{F}\). By Lemma 4, every frame in \(\mathcal{G}\) is isomorphic to \(C\oplus \circ\) for an \({\mathrm{\small S5}}^2\)-cluster \(C\). Let \(\mathcal{C}\) be the class of all such frames. Since \(\mathop{\mathrm{Log}}\mathcal{F}=\mathop{\mathrm{Log}}\mathcal{G}\) is a proper extension of \({\mathrm{\small Tack}}_2\), then, in view of Lemma 5, \(\mathop{\mathrm{Log}}(\mathcal{C})\) is a proper extension of \({\mathrm{\small S5}}^2\). By Theorem 33, the logic \(L\) of this class is locally finite. So \(\mathcal{C}\) is a uniformly tunable class. Clearly, the top clusters in \({\mathrm{\small Tack}}\)-frames also form a uniformly tunable class, since they are singletons. Now the statement follows from Theorem 7. ◻
Remark 41. For a Kripke complete proper extension \(L\) of \({\mathrm{\small S5}}^2\), local tabularity can be obtained directly, as a corollary of Theorem 26. See Appendix, Section 9.3.
Theorem 42. \({\mathrm{\small Tack}}_2\) is prelocally tabular.
Proof. Let \(L\) be a non-locally tabular extension of \({\mathrm{\small Tack}}_2\). We show that \(L={\mathrm{\small Tack}}_2\).
For some finite \(k\), the \(k\)-generated canonical Kripke frame \(F_L\) of \(L\) is infinite.
Let us list some properties of \(F_L\). Since \({\mathrm{\small Tack}}_2\) is canonical (Theorem 40), and \(L\) contains \({\mathrm{\small Tack}}_2\), \(F_L\) is a \({\mathrm{\small Tack}}_2\)-frame. The top clusters in \(F_L\) are singletons. Let \(\mathcal{C}\) be the family of non-top clusters in \(F_L\). In view of Lemma 4, \(\mathcal{C}\) are \({\mathrm{\small S5}}^2\)-frames. We also have: \[\label{eq:tack-clusterD} \text{For C\in \mathcal{C}, there is a unique cluster D above C, and D is a singleton. }\tag{13}\]
Now we claim that \(\mathcal{C}\) contains an infinite cluster. For the sake of contradiction, suppose that all clusters in \(\mathcal{C}\) are finite. In this case, all point-generated subframes if \(F_L\) are finite, and it is straightforward that \(\mathop{\mathrm{Log}}(F_L)=L\). But in this case, \(L\) is locally tabular by Lemma 7. This contradiction proves that \(\mathcal{C}\) contains an infinite cluster.
Let \(C\) be an infinite cluster in \(\mathcal{C}\), and let \(D\) be the cluster above \(C\). Put \(X=C\cup D\). Clearly, the restriction of \(F=F_L{\upharpoonright}X\) is a generated subframe of \(F_L\).
Consider the extension \({\mathrm{\small Tack}}_2[1]\) of \({\mathrm{\small Tack}}_2\) with the axiom of height 1, which makes the composite relation \(R_1\circ R_2\) equivalence. Clearly, \({\mathrm{\small Tack}}_2[1]\) is locally tabular (even tabular, since this is the logic of the singleton \(\circ\)). It follows that the set \(T\) of points in top clusters in \(F_L\) is definable [31]: for a formula \(\psi_T\), \[\psi_T\in x \text{ iff } x\in T\]
Hence, for every \(x\) in \(X\): \[\label{eq:tack:defTop} x\in C \text{ iff } \psi_T\notin x\tag{14}\]
Let \(\Psi\) be the set of all bimodal formulas in \(k\) variables. For a formula \(\psi\in \Psi\), let \(\theta(\psi)=\{x \in X \mid \psi\in x\}\), \(\eta(\psi)=\{x \in C \mid \psi\in x\}\).
Let \(A\) denote the powerset modal algebra of the cluster-frame \(F{\upharpoonright}C\), and let \(B\) be its subalgebra generated by the sets \(\eta(p_i)\), \(i<k\).
By a straightforward induction on the structure of formulas, we have: \[\label{eq:tack:defSets} \text{ For every b\in B, there is \psi\in\Psi such that b=\theta(\psi)\cap C.}\tag{15}\]
Since \(C\) is infinite, \(B\) is infinite as well: any two distinct points in \(C\) are separable by some formula. Since \(C\) is an \({\mathrm{\small S5}}^2\)-frame, \(B\) is an \({\mathrm{\small S5}}^2\)-algebra. Let \(A_m\) denote the powerset modal algebra of the rectangle \(\boldsymbol{m} \times \boldsymbol{m}\). It follows from [17] that all finite \(A_m\) are embeddable in \(B\). Let \(A_m\cong B_m\subset A\), and let \(f_m\) be the corresponding p-morphism \(F{\upharpoonright}C\twoheadrightarrow\boldsymbol{m} \times \boldsymbol{m}\).
Let \(b_{ij}\), \(i,j<m\) be the atoms of \(B_m\). By 15 , there are formulas \(\psi_{ij}\) such that \(b_{ij}=\theta(\psi_{ij})\cap C\). By 14 , \(b_{ij}\) is definable in \(F\) by the formula \(\psi_{ij}\wedge \neg \psi_T\). Let \(g_m:F\to \mathrm{\small T}{(m,m)}\) be the extension of \(f_m\) that maps the top to the top. Clearly, \(g_m\) is a p-morphism. Since all preimages \(g_m^{-1}(i,j)\) are definable in \(F\) by \(\psi_{ij}\wedge \neg \psi_T\), the algebra of \(\mathrm{\small T}(m,m)\) is embeddable in the algebra of sets \(\theta(\psi)\), \(\psi\in\Psi\). The latter algebra is an \(L\)-algebra, since \(F\) is a generated subframe of \(F_L\).
So the algebras of all frames \(\mathrm{\small T}(m,m)\) are \(L\)-algebras, and due to Corollary 8, \(L\subseteq {\mathrm{\small Tack}}_2\). Consequently, \(L={\mathrm{\small Tack}}_2\). ◻
The main result of this paper is the criteria obtained in Section 3.
In Section 4, the criteria are applied to describe new families of locally tabular products. In particular, we generalized results from [7], [8].
In Section 5, we discussed the product finite model property. In particular, we showed that the local tabularity of a product logic does not imply the product fmp, even in the case of height 3.
We showed in Theorem 39 that the product rpp formula gives an axiomatic criterion of local tabularity for all extensions of \({\mathrm{\small S4}}.1 [ 2 ]\times {\mathrm{\small S5}}\).
Finally, we described a prelocally tabular extension of \({\mathrm{\small S4}}\times{\mathrm{\small S5}}\) of height 2.
According to Theorem 26, local tabularity of the factors and a product rpp formula are equivalent to the local tabularity of the product. In particular, in the case of transitive logics of finite height, local tabularity of the product is equivalent to the product rpp due to Corollary 4. These facts do not give a criterion for normal extensions of a product. Unlike these facts, Theorem 39 gives an axiomatic criterion for all extensions of \({\mathrm{\small S4}}.1 [ 2 ]\times {\mathrm{\small S5}}\). We conjecture that Theorem 39 can be generalized for a wider class of logics.3
Problem. Let \(S4[h]\) be the extension of \({\mathrm{\small S4}}\) with the axiom of finite height \(B_h\). Does the product rpp criterion hold for every extension \(L\) of \({\mathrm{\small S4}}[h]\times {\mathrm{\small S5}}\)?
If not, what is the largest such \(h\)?
It is a well-known open problem whether every non-locally tabular unimodal logic is contained in a prelocally tabular [18]. For the case of logics above \({\mathrm{\small S4}}\times {\mathrm{\small S4}}\), this question is also open. We conjecture that this is true for logics above \({\mathrm{\small S4}}\times {\mathrm{\small S5}}\).
One of the central tools for our results was the criterion given in Theorem 7 [23], which, informally, says, that under the necessary conditions of pretransitivity and finite height, only local tabularity on clusters matters. While this criterion has many applications, there is an obvious limitation for this approach. Namely, even in the unimodal case, if the relation is symmetric, then every point-generated frame consists of one cluster, and no additional structure is given by the skeleton construction. Hence, it is of definite interest to study local finiteness of symmetric relations. Under the necessary conditions of local finiteness, this means that we are interested in graphs of finite diameters. No axiomatic criterion is known for this case, and it is not guaranteed that an explicit axiomatic characterization of local finiteness is possible in this case.
We are grateful to Guram Bezhanishvili and Valentin Shehtman for valuable discussions.
Lemma 8. Let \(\mathrm{A}\) be finite, \(F\) an \(\mathrm{A}\)-frame such that:
(a) \(\mathop{\mathrm{dom}}F = X_1\cup X_2\) for disjoint \(X_1\) and \(X_2\), and
(b) \(F{\upharpoonright}X_1\) is a generated subframe of \(F,\) and
(c) \(F{\upharpoonright}X_i\) is \(f_i\)-tunable for some monotone \(f_i:\:\omega\to \omega,\) \(i = 1, 2.\)
Then \(F\) is \(g\)-tunable for \[g(n) = f_1(n) + f_2\left(n\cdot2^{f_1(n)\cdot |\mathrm{A}|} \right).\]
Proof. Let \(\mathcal{V}\) be a partition of \(\mathop{\mathrm{dom}}F,\,|\mathcal{V}| = n < \omega.\) Then \(\mathcal{V}\) induces partitions \(\mathcal{V}_1,\,\mathcal{V}_2\) of \(X_1\) and \(X_2,\) respectively. There exists a refinement \(\mathcal{U}_1\) of \(\mathcal{V}_1\) such that \(\mathcal{U}_1\) is tuned in \(F_1\) and \[\label{eq:clusters-1} |\mathcal{U}_1| \le f_1(|\mathcal{V}_1|) \le f_1(n).\tag{16}\]
For \(\lozenge\in\mathrm{A}\), define \(\alpha_\lozenge:\:X_2 \to 2^{\mathcal{U}_1}\) by letting \[\alpha_\lozenge(x) = \{U\in \mathcal{U}_1 \mid x\in R_\lozenge^{-1}[U]\}.\] Let \(\sim_\lozenge\) be the equivalence induced by \(\alpha_\lozenge\); it has at most \(2^{|\mathcal{U}_1|}\) classes. Hence, the equivalence \(\sim\; = \;\bigcap\{ \sim_\lozenge\mid \lozenge\in \mathrm{A}\}\) has at most \(2^{f_1(n)\cdot |\mathrm{A}|}\) classes. Let \(\equiv\) be the equivalence on \(X_2\) whose quotient is \(\mathcal{V}_2\). Put \(\mathcal{S}={X_2}{/}{(\sim\cap \equiv)}\). We have \[\label{eq:clusters-2} |\mathcal{S}| \le |\mathcal{V}_2| \cdot 2^{|\mathcal{U}_1|\cdot |\mathrm{A}|} \le n \cdot 2^{f_1(n)\cdot |\mathrm{A}|},\tag{17}\] where the latter inequality holds by 16 . Hence, there exists a refinement \(\mathcal{T}\) of \(\mathcal{S}\) of size at most \(f_2\left(n\cdot 2^{f_1(n)\cdot |\mathrm{A}|}\right)\), which is tuned in \(F_2\).
Consider the partition \(\mathcal{U}=\mathcal{U}_1 \cup \mathcal{T}\) of \(F\). Clearly, it refines \(\mathcal{V}\), and it has at most \(g(n)\) elements in view of 16 and 17 . So it remains to check that \(\mathcal{U}\) is tuned in \(F\).
Let \(xR_{\lozenge} y\) for some \(x,\,y\in X\) and \(\lozenge\in \mathrm{A}\), and let \(x'\in [x]_\mathcal{U}.\) We will show that \(x' R_\lozenge y'\) for some \(y'\in [y]_\mathcal{U}\).
Assume that \(x\in X_1\). Then \(y\in X_1\), since \(F{\upharpoonright}X_1\) is a generated subframe of \(F\). In this case, \(x\) and \(x'\) belong to the same element of \(\mathcal{U}_1.\) Since \(\mathcal{U}_1\) is tuned in \(F_1\), \(x' R_{\lozenge} y'\) for some \(y'\in [y]_{\mathcal{U}_1}=[y]_\mathcal{U}\), as desired.
Now assume \(x\in X_2\). Then \(x'\in [x]_\mathcal{T}\), so \(\alpha_\lozenge(x) = \alpha_\lozenge(x').\) If \(y\in X_1\), then \([y]_{\mathcal{U}_1}\in \alpha_\lozenge(x) = \alpha_\lozenge(x')\), so \(x' R_\lozenge y'\) for some \(y'\in [y]_{\mathcal{U}_1}=[y]_\mathcal{T}\), as desired. If \(y\in X_2\), then \(x' R_{\lozenge} y'\) for some \(y'\in [y]_\mathcal{T}=[y]_\mathcal{U}\), since \(\mathcal{T}\) is tuned in \(F_2\). ◻
Proof of Theorem 7. ‘Only if’. Finite height follows from Theorem 4. The logic of subframes of frames in \(\mathcal{F}\) is contained in the logic of the clusters \(Cl{(\mathcal{F})}\) and is locally tabular by Proposition 6. Hence, the logic of \(Cl{(\mathcal{F})}\) is locally tabular.
‘If’. Let \(\mathcal{F}_h=\{F\in \mathcal{F}\mid h(F)\leq h\}\), \(1\leq h<\omega\). By induction on \(h\) we show that \(\mathcal{F}_h\) is \(g_h\)-tunable for some monotone \(g_h:\omega\to\omega\). Let \(h=1\). The class \(\mathcal{F}_1\) consists of disjoint sums of frames in \(Cl{(\mathcal{F})}\), so \(\mathop{\mathrm{Log}}\mathcal{F}_1=\mathop{\mathrm{Log}}Cl{(\mathcal{F})}\). Since \(\mathop{\mathrm{Log}}\mathcal{F}_1\) is locally tabular, the class \(\mathcal{F}_1\) is \(f\)-tunable for some \(f\) by Theorem 5. Put \(g_1(n)=\max\{f(i)\mid i\leq n\}\); clearly, \(\mathcal{F}_1\) is \(g_1\)-tunable. Let \(h>1\). By induction hypothesis, \(\mathcal{F}_{h-1}\) is \(g_{h-1}\)-tunable for some monotone \(g_{h-1}\). Put \(g_h(n) = g_1(n) + g_{h-1}\left(n\cdot2^{g_1(n)\cdot |\mathrm{A}|} \right).\) Consider \(F\in \mathcal{F}_{h}\). Let \(F_1\) be the restriction of \(F\) on its maximal clusters, \(X_1=\mathop{\mathrm{dom}}F_1\), \(X_2=(\mathop{\mathrm{dom}}F){\setminus}X_1\). If \(X_2 =\varnothing\), then \(F\in\mathcal{F}_1\), and so is \(g_1\)-tunable; hence, \(F\) is \(g_h\)-tunable. If \(X_2 \neq \varnothing\), \(F\) is \(g_h\)-tunable by Lemma 8.
Since \(\mathcal{F}=\mathcal{F}_h\) for some \(h\), \(\mathcal{F}\) is uniformly tunable, and hence its logic is locally tabular by Theorem 5. 0◻
Definition 2. Consider an \(\mathrm{A}\)-frame \(I=(Y,(S_\lozenge)_{\lozenge\in \mathrm{A}})\) and a family \((F_i)_{i\in Y}\) of \(\mathrm{A}\)-frames \(F_i=(X_i,(R_{i,\lozenge})_{\lozenge\in \mathrm{A}})\). The sum \({\textstyle \sum_{I}{F_i}}\) of the family \((F_i)_{i\in Y}\) over \(I\) is the \(\mathrm{A}\)-frame \((\bigsqcup_{i \in Y} X_i, (R^\Sigma_\lozenge)_{\lozenge\in \mathrm{A}})\), where \(\bigsqcup_{i\in Y}{X_i}=\bigcup_{i\in Y}(\{i\}\times X_i)\), and \[(i,a)R^\Sigma_\lozenge(j,b) \quad \text{ iff }\quad (i = j \text{ and }a R_{i,\lozenge} b) \text{ or } (i\neq j \text{ and } iS_\lozenge j).\] For classes \(\mathcal{I}\), \(\mathcal{F}\) of \(\mathrm{A}\)-frames, let \({\textstyle \sum_{\mathcal{I}}{\mathcal{F}}}\) be the class of all sums \({\textstyle \sum_{I}{F_i}}\) such that \(I\in \mathcal{I}\) and \(F_i\in \mathcal{F}\) for every \(i\) in \(I\).
Theorem 43. [24]Let \(\mathcal{F}\) and \(\mathcal{I}\) be classes of \(\mathrm{A}\)-frames. If the logics \(\mathop{\mathrm{Log}}{\mathcal{F}}\) and \(\mathop{\mathrm{Log}}{\mathcal{I}}\) are locally tabular, then \(\mathop{\mathrm{Log}}{{\textstyle \sum_{\mathcal{I}}{\mathcal{F}}}}\) is locally tabular as well.
Definition 3. Two classes of frames are said to be interchangeable, if they have the same logic in the language enriched with the universal modality.
The following fact follows from [33].
Lemma 9. Let \(I\) be an \(\mathrm{A}\)-frame, and for each \(i\in I\), let \(\mathcal{F}_i\) and \(\mathcal{G}_i\) be two interchangeable families of \(\mathrm{A}\)-frames. Let \(\mathcal{K}_1\) be the class of sums \({\textstyle \sum_{I}{F_i}}\) with \(F_i\in \mathcal{F}_i\) for each \(i\in I\), and let \(\mathcal{K}_2\) be the class of sums \({\textstyle \sum_{I}{G_i}}\) with \(G_i\in \mathcal{G}_i\) for each \(i\in I\). Then \(\mathop{\mathrm{Log}}\mathcal{K}_1 =\mathop{\mathrm{Log}}\mathcal{K}_2\).
Proof of Lemma 5. Any frame \({{F}\,\oplus\, \circ}\) is the sum of \(F\) and the singleton \(\circ\) over the indexing frame \(I=(2,\leq,\varnothing)\) in the sense of Definition 2. Classes \(\mathcal{C}\) and \(\mathcal{D}\) are interchangeable, since the universal modality is expressible as \(\lozenge_1\lozenge_2\) in \({\mathrm{\small S5}}^2\)-frames. Now the statement follows from Lemma 9. ◻
Definition 4. Let \(F = (X,\,(R_\lozenge)_{\lozenge\in A})\) be an \(\mathrm{A}\)-frame, where all \(R_\lozenge\) are preorders. Define the equivalence relation \({\sim}=\bigcap_{\lozenge\in A} (R_\lozenge\cap R_\lozenge^{-1}).\) The quotient frame \({F}{/}{\sim}\) is denoted \(\widetilde{F}\). Given a class \(\mathcal{F}\) of \(\mathrm{A}\)-frames, let \(\widetilde{\mathcal{F}} = \{\widetilde{F} \mid F\in \mathcal{F}\}.\)
Lemma 10. If the relations of \(F\) are preorders, then \(F\twoheadrightarrow\widetilde{F}.\)
Proof. For any frame, the quotient map is a surjective homomorphism. To check the back condition, assume that \([a]\widetilde{R_\lozenge}[b]\) for some \(a,\,b\in X\) and \(\lozenge\in \mathrm{A}.\) Then \(a' R_\lozenge b'\) for some \(a'\sim a\) and \(b'\sim b.\) By the definition, \(a R_\lozenge a'\) and \(b' R_\lozenge b,\) so \(a R_\lozenge a' R_\lozenge b' R_\lozenge b ,\) hence \(a R_\lozenge b\) by transitivity. ◻
Lemma 11. Let \(\mathcal{F}\) be a class of frames such that in every \(F\in\mathcal{F}\), all relations are preorders. Then \(\mathop{\mathrm{Log}}\widetilde{\mathcal{F}}\) is locally tabular iff \(\mathop{\mathrm{Log}}\mathcal{F}\) is locally tabular.
Proof. The ‘if’ direction is trivial, since \(\mathop{\mathrm{Log}}\mathcal{F}\subseteq \mathop{\mathrm{Log}}\widetilde{\mathcal{F}}\) by Lemma 10.
‘Only if’. Consider \(F=(X,\,(R_\lozenge)_{\lozenge\in \mathrm{A}})\in \mathcal{F}.\) Let \(\sim\) be defined as in Definition 4, and denote \(Y = X/{\sim}.\) It is straightforward that \(F\) is isomorphic to the sum \({\textstyle \sum_{\widetilde{F}}{(}}{F{\upharpoonright}U})\), where \(U\subseteq X\) ranges over the equivalence classes in \(Y\). It follows that \(\mathop{\mathrm{Log}}\mathcal{F}\) is included in \(\mathop{\mathrm{Log}}{{\textstyle \sum_{\widetilde{\mathcal{F}}}{\mathcal{G}}}}\), where \(\mathcal{G}\) is the class of all frames \(F{\upharpoonright}U\) where \(F\in \mathcal{F}\) and \(U \in \mathop{\mathrm{dom}}F/{\sim}.\) In frames in \(\mathcal{G}\), every relation is universal, hence every partition is tuned, and so its logic is locally tabular by Theorem 5. Since \(\mathop{\mathrm{Log}}\widetilde{\mathcal{F}}\) is also locally tabular, \(\mathop{\mathrm{Log}}\mathcal{F}\) is locally tabular by Theorem 43. ◻
Lemma 12. Let \(F=(X,R_1,R_2)\) be a point-generated \({\mathrm{\small S5}}^2\)-frame, \(|{X}{/}{R_1}|=\kappa\), \(|{X}{/}{R_2}|=\mu\). Then \(\widetilde{F}\) is isomorphic to \(\boldsymbol{\kappa} \times \boldsymbol{\mu}\).
Proof. Let \(\{H_\alpha\}_{\alpha < \kappa}\) and \(\{V_\beta\}_{\beta < \mu }\) be the elements of \({X}{/}{R_1}\) and \({X}{/}{R_2},\) respectively. We claim that \(H_\alpha\cap V_\beta\ne \varnothing\) for any \(\alpha< \kappa,\,\beta < \mu.\) Let \(a \in H_\alpha\) and \(b\in V_\beta.\) Since \(F\) is point-generated, \(b\in R_2[R_1(a)],\) so there exists \(c\in X\) such that \(a R_1 c\) and \(c R_2 b.\) Then \(c \in R_1(a) = H_\alpha\) and \(c\in R_2(b) = V_\beta,\) so \(c\in H_\alpha\cap V_\beta.\)
Define \(f:\:\kappa\times \mu \to {X}{/}{\sim}\) by \(f(\alpha,\,\beta) = H_\alpha \cap V_\beta.\) We claim that \(f\) is an isomorphism between \(\boldsymbol{\kappa} \times \boldsymbol{\mu}\) and \(\widetilde{F}.\) For any \([a]\in {X}{/}{\sim}\) there exist unique \(H_\alpha,\,V_\beta\) such that \(a\in H_\alpha \cap V_\beta,\) so \([a] = H_\alpha\cap V_\beta,\) thus \(f\) is a bijection. Let \(R_H\) and \(R_V\) denote the horizontal and vertical relations in \(\boldsymbol{\kappa} \times \boldsymbol{\mu}\). If \((\alpha,\beta) R_H (\alpha',\beta')\) then \(\alpha = \alpha'.\) Consider \(a\in H_\alpha\cap V_\beta\) and \(b\in H_\alpha\cap V_\beta'.\) Observe that \(H_\alpha = R_1(a),\) so \(a R_1 b,\) thus \(f(\alpha,\beta) = [a] \widetilde{R_1}[b] = f(\alpha',\beta').\) Conversely, if \(f(\alpha,\beta) \widetilde{R_1} f(\alpha',\beta'),\) there exist \(a\in H_\alpha\cap V_\beta\) and \(b\in H_\alpha'\cap V_\beta'\) such that \(a R_1 b.\) Then \(H_\alpha = R_1(a) = R_1(b) = H_\alpha',\) so \(\alpha = \alpha'\) and \((\alpha,\beta) R_H (\alpha,\beta') = (\alpha',\beta').\) An analogous argument applies for \(R_V.\) ◻
By Theorem 33 [17], all proper extensions of \({\mathrm{\small S5}}^2\) are locally tabular. Assuming Kripke completeness of a proper extension \(L\) of \({\mathrm{\small S5}}^2\), local tabularity is a simple corollary of Theorem 26 and our previous observations. Indeed, let \(\mathcal{F}\) be the class of point-generated frames of \(L\). By Lemma 10, \(L \subseteq \mathop{\mathrm{Log}}{\widetilde{\mathcal{F}}}\). So \(\mathop{\mathrm{Log}}{\widetilde{\mathcal{F}}}\) is a proper extension of \({\mathrm{\small S5}}^2\) as well, and hence \(n = \sup\{m \mid \exists F\in\mathcal{F}\; \,(\widetilde{F}\twoheadrightarrow\boldsymbol{m} \times \boldsymbol{m} )\}\) is finite. For \(i = 1,\,2\), let \(\mathcal{G}_i\) be the class of frames of form \(\boldsymbol{X}_1 \times \boldsymbol{X}_2\) with \(|X_i|\leq n\). By Lemma 12, each frame in \(\widetilde{\mathcal{F}}\) is isomorphic to a frame in \(\mathcal{G}_1 \cup \mathcal{G}_2\). Both classes \(\mathcal{G}_i\) are uniformly tunable by Theorem 26, and so is the class \(\widetilde{\mathcal{F}}\).4 Then \(L\) is locally tabular by Lemma 11.
This does not give a complete proof of Theorem 33: we need to exclude the case of incomplete logics. For a logic \(L\), let \(F_L\) be its canonical frame [19]. The following reasoning is a variant of the proof given in [17];5 we provide it for the self-containment of the text. Together with the previous reasoning, it completes the proof of Theorem 33.
Lemma 13. [17]If \(L\) is a proper extension of \({\mathrm{\small S5}}^2\), then so is \(\mathop{\mathrm{Log}}{F_L}\).
Proof. For the sake of contradiction, assume that \(\mathop{\mathrm{Log}}{F_L} = {\mathrm{\small S5}}^2\). We show that in this case \(L={\mathrm{\small S5}}^2\).
Let \(m<\omega\). By Proposition 3, \(f:F\twoheadrightarrow\boldsymbol{m} \times \boldsymbol{m}\) for a point-generated subframe \(F=(X,R_1,R_2)\) of \(F_L\) and a p-morphism \(f\).
The logic \({\mathrm{\small S5}}^2\) is canonical [29], [36]. It follows that \(R_1\) and \(R_2\) are equivalences, and for all \(a,b\in X\), we have \[\label{eq:cone} a (R_1\circ R_2) b \text{ and } a (R_2\circ R_1) b.\tag{18}\]
For \(i<m\), choose \(d_i\in f^{-1}(i,i)\). For distinct \(i,j<m\), we have \((d_j,d_i)\notin R_1\cup R_2;\) hence, there is a formula \(\delta_{ij}\in d_i\) such that \(\lozenge_1\delta_{ij}\vee \lozenge_2\delta_{ij}\notin d_j\). For \(i>0\), put \(\delta_i=\bigwedge_{i\neq j<m} \delta_{ij}\); put \(\delta_0=\bigwedge_{0<k<m} \neg (\lozenge_1\delta_{k}\vee \lozenge_2\delta_{k})\). We claim that \[\label{eq:dij} \delta_i\in d_i.\tag{19}\] For \(i>0\), this is trivial. To show that \(\delta_0\in d_0\), assume that \(\delta_{k}\in a\) for some \(a\in X\) and \(k>0\). Then \(\delta_{k0}\in a\) and \(\neg (\lozenge_1\delta_{k0}\vee \lozenge_2\delta_{k0})\in d_0\). Hence, \((d_0,a)\notin R_1\cup R_2\). Hence, \(\neg (\lozenge_1\delta_{k}\vee \lozenge_2\delta_{k})\in d_0\).
Define horizontal and vertical sets: \(H_i=\{a\in X\mid \lozenge_1 \delta_i\in a\}\), \(V_i=\{a\in X\mid \lozenge_2 \delta_i\in a\}\). We have \[\label{eq:lift} R_1(d_i) \subseteq H_i \text{ and }R_2(d_i) \subseteq V_i.\tag{20}\] Since \(R_1\) and \(R_2\) are equivalence relations, we also have \[\label{eq:HVmon} \text{R_1[H_i]=H_i and R_2[V_i]=V_i.}\tag{21}\] All sets \(H_i\) and \(V_i\) are non-empty, since \(d_i\in H_i\cap V_i\).
Let us check that \(\bigcup_{i<m} H_i=\bigcup_{i<m} V_i=X\). For this, assume that \(a\in X\), and \(a\notin \bigcup_{0<i<m} V_i\). By 18 , \(d_0R_1bR_2a\) for some \(b\). Let \(0<k<m\). Hence: \(\neg \lozenge_1\delta_{k}\in b\), since \(d_0 R_1b\); \(\neg \lozenge_2\delta_{k}\in b\), since \(bR_2a\); hence, \(\delta_0\in b\), and so \(a\in H_0\). Likewise, \(\bigcup_{i<m} H_i=X\).
Let us check that \(H_i \cap H_j =\varnothing=V_i \cap V_j\) for distinct \(i,j\). Indeed, if \(a\in H_i \cap H_j\), then \(\lozenge_1 \delta_i\wedge\lozenge_1 \delta_j\in a\), and so \(\delta_i\wedge\lozenge_1 \delta_j\in b\) for some \(b\in R_1(a)\), which implies \(i=j\). Similarly for vertical sets. It follows that \(\{H_i\}_{i<m}\) and \(\{V_i\}_{i<m}\) are \(m\)-element partitions of \(X\).
Put \(C_{ij}=V_i \cap H_j\). By 18 , we have \(d_i R_2 a R_1 d_j\) for some \(a\), and so each \(C_{ij}\) is non-empty according to 20 . It follows that \(\{C_{ij}\}_{i,j<m}\) is an \(m^2\)-element partition of \(X\). For \(a\in C_{ij}\), put \(g(a)=(i,j)\). Hence, \(g\) maps \(X\) onto \(m\times m\).
We claim that \(g\) is a p-morphism. By 21 , \(g\) is a homomorphism. It remains to check the back property. Let \(a\in C_{ij}\), and assume that in \(\boldsymbol{m} \times \boldsymbol{m}\), \(\tau\) is related to \((i,j)\) by the horizontal relation. Then \(\tau=(k,j)\) for some \(k<m\). We have \(a R_1 b R_2 d_{k}\) for some \(b\) due to 20 . Hence, \(b\in C_{kj}\) and so \(g(b)=(k,j)\). The proof for \(R_2\) is completely analogous.
Assume that a formula \(\varphi\) is satisfiable in a model \((\boldsymbol{m} \times \boldsymbol{m},\theta)\). Define the following valuation \(\eta\) on \(F\): put \(\eta(p)=\bigcup\{C_{ij}\mid (i,j)\in \theta(p)\}\). Then \(g\) is a p-morphism of \((F,\eta)\) onto \((\boldsymbol{m} \times \boldsymbol{m},\theta)\). So \(\varphi\) is satisfiable in \((F,\eta)\).
We have \((F,\theta_L)\vDash L\), where \(\theta_L\) is the restriction of the canonical valuation on \(F\). Let \(\psi_p\) be the formula \(\bigvee \{\lozenge_1 \delta_i\wedge\lozenge_2 \delta_j \mid (i,j)\in \theta(p) \}\). For each \(a\in X\), we have \(a\in \eta(p) \text{ iff } \psi_p \in a,\) so \((F,\eta)\vDash L\). It follows that \(L={\mathrm{\small S5}}^2\). ◻
Now Theorem 33 follows: if \(L\) is a proper extension of \({\mathrm{\small S5}}^2\), then so is the logic \(\mathop{\mathrm{Log}}F_{L}\) of the canonical frame of \(L\); the latter logic is Kripke complete and so is locally tabular; since \(\mathop{\mathrm{Log}}F_{L}\subseteq L\), \(L\) is locally tabular.
To avoid any ambiguity, we remark that we write \(R_F^{\leq m}\) for \((R_F)^{\leq m}\), and \(R_F^*\) for \((R_F)^*\).↩︎
This is a particular case of the following general fact. If the varieties generated by classes \(\mathcal{C}_1\) and \(\mathcal{C}_2\) are locally finite, then the variety generated by the class \(\mathcal{C}_1 \cup \mathcal{C}_2\) is locally finite as well: \(\mathcal{C}_1 \cup \mathcal{C}_2\) is uniformly locally finite by Malcev criterion given in Theorem 2.↩︎
In a very recent preprint [35], it was announced that Theorem 39 extends to logics above \({\mathrm{\small S4}}[2]\times {\mathrm{\small S5}}\).↩︎
This is a particular case of the following general fact. If the varieties generated by classes \(\mathcal{C}_1\) and \(\mathcal{C}_2\) are locally finite, then the variety generated by the class \(\mathcal{C}_1 \cup \mathcal{C}_2\) is locally finite as well: \(\mathcal{C}_1 \cup \mathcal{C}_2\) is uniformly locally finite by Malcev criterion given in Theorem 2.↩︎
We are grateful for an anonymous reviewer for providing the reference to this claim.↩︎