1 Introduction
Unlike classical and intuitionistic first-order and propositional logics, numerous modal logics, L, do not enjoy the Craig interpolation property (CIP): they contain valid implications
$\varphi \to \psi $
without an interpolant in L—a formula
$\iota $
in the shared signature of
$\varphi $
and
$\psi $
such that both
$\varphi \to \iota $
and
$\iota \to \psi $
are also valid in L. Typical examples of such L are first-order modal logics with constant domains between
$\mathsf {K}$
and
$\mathsf {S5}$
[Reference Fine13] and propositional modal logics with linear transitive Kripke frames of unbounded depth [Reference Gabbay and Maksimova14, Reference Wolter39]. There have been various attempts to classify propositional modal logics with the CIP, successful for extensions of
$\mathsf {S4}$
[Reference Gabbay and Maksimova14, Section 8] and unsuccessful for extensions of
$\mathsf {K4}$
or
$\mathsf {GL}$
, where the CIP turned out to be undecidable [Reference Chagrov and Zakharyaschev9, Sections 14 and 17].
While establishing the CIP of a logic L typically gives rise to further research problems—develop proof systems that admit efficient/elegant interpolant computation [Reference Benedikt, ten Cate and Van den Boom3, Reference Kuznets29], investigate the complexity of computing interpolants from proofs [Reference Krajicek25, Sections 17 and 18], consider restrictions on the shape of interpolants such as in, say, Lyndon’s interpolation [Reference Lyndon31], or employ the CIP to investigate related properties such as Beth definability [Reference Craig10, Reference Feferman11]—a counterexample to the CIP has usually terminated further research of Craig interpolants and their applications for the unfortunate logic in question.
In this article, we take a different, non-uniform view of Craig interpolation and aim to understand interpolants also for logics L without the CIP. We consider the following interpolant existence problem (IEP) for L: given formulas
$\varphi $
and
$\psi $
, decide whether
$\varphi \to \psi $
has an interpolant in L. For L without the CIP, the existence of an interpolant for
$\varphi $
and
$\psi $
does not follow from the validity of
$\varphi \rightarrow \psi $
in L, and so the IEP does not reduce to validity checking (which is reducible to the IEP). A first question then is whether the former problem is genuinely harder than the latter one. In fact, when the IEP was introduced [Reference Artale, Jung, Mazzullo, Ozaki and Wolter1, Reference Jung and Wolter24], this was shown to be the case for modal logics with nominals and for the two-variable and guarded fragments of first-order logic. Since then, this has also been confirmed for the one-variable fragment of first-order modal logic
$\mathsf {S5}$
and weak
$\mathsf {K4}$
[Reference Kurucz, Wolter, Zakharyaschev, Ciabattoni, Gabelaia and Sedlár26, Reference Kurucz, Wolter and Zakharyaschev27].
Here, we show that the opposite is true of propositional modal logics containing
$\mathsf {K4.3}$
, the logic of linear transitive frames: while none of these logics with frames of unbounded depth has the CIP [Reference Gabbay and Maksimova14, Reference Wolter39], interpolant existence is nevertheless decidable in coNP for finitely axiomatisable logics, and so is as hard as validity [Reference Litak and Wolter30]. This is the first general result on Craig interpolant existence covering a large family of modal logics and, potentially, a step towards a classification of modal logics according to the complexity of the IEP.
We proceed as follows. To begin with, we give a ‘folklore’ characterisation of interpolant existence via bisimulations between models based on descriptive frames:
$\varphi \to \psi $
does not have an interpolant in L iff
$\varphi $
and
$\neg \psi $
can be satisfied in
$\textit {sig}(\varphi )\cap \textit {sig}(\psi )$
-bisimilar models based on descriptive frames for L. If L had the CIP, we could merge these two models into a single one satisfying
$\varphi \land \neg \psi $
(using, say, bisimulation products [Reference Marx32]) or amalgamate the induced modal algebras [Reference Gabbay and Maksimova14], which is impossible in our case. Instead, we aim to understand the fine-grained structure of the required bisimilar models and use it to decide their existence. We show that, for some logics (such as first-order definable cofinal subframe logics), any pair of bisimilar models can be transformed into bisimilar models of polynomial size; in other words, such logics enjoy the polysize bisimilar model property. However, for other logics like
$\mathsf {GL.3}$
, not even models based on infinite Kripke frames are enough despite
$\mathsf {GL.3}$
having the finite model property (fmp).
We prove, nevertheless, that every pair of bisimilar models satisfying
$\varphi $
and
$\neg \psi $
and based on descriptive frames for a finitely axiomatisable
$L \supseteq \mathsf {K4.3}$
can be converted to a pair of such models with an understandable structure. In a nutshell, their underlying frames look like a polynomial-size chain of polynomial-size clusters and tadpole-like descriptive frames that comprise a non-degenerate cluster
$\{a_0,\dots , a_{k-1}\}$
, for some polynomial-size
$k>0$
, followed by an infinite descending chain of points
$b_n$
,
$n < \omega $
, which are all irreflexive or all reflexive, with the internal sets (restricting possible valuations) generated as a modal algebra by the singletons
$\{b_n\}$
and the k-many pairwise disjoint infinite sets
$X_i = \{a_i\} \cup \{b_n \mid n \equiv i \ (\text {mod}\ k) \}$
. The picture below illustrates the underlying Kripke frame and the generators of the tadpole descriptive frame with
$k=2$
.

Because of this, we say that all finitely axiomatisable
$L\supseteq \mathsf {K4.3}$
have the quasi-polysize bisimilar model property. We show that the existence of such quasi-polysize bisimilar models can be checked in NP in the size of
$\varphi $
and
$\psi $
, for any finitely axiomatisable L.
Finally, we extend the developed techniques to analyse the IEP for a few Priorean temporal logics with past and future modal operators: the logic
$\mathsf {Lin}$
of all linear frames, the logic
$\mathsf {Lin}_{<\omega }$
of all finite strict linear orders, and the logics
$\mathsf {Lin}_{\mathbb {Q}}$
of the rationals,
$\mathsf {Lin}_{\mathbb {R}}$
of the reals, and
$\mathsf {Lin}_{\mathbb {Z}}$
of the integers. We prove that
$\mathsf {Lin}$
,
$\mathsf {Lin}_{\mathbb {Q}}$
, and
$\mathsf {Lin}_{\mathbb {R}}$
have the polysize bisimilar model property, while
$\mathsf {Lin}_{<\omega }$
and
$\mathsf {Lin}_{\mathbb {Z}}$
have the quasi-polysize one, with the IEP being coNP-complete. The proofs can be regarded as applications of the general method, which works for all extensions of
$\mathsf {K4.3}$
, to a few concrete logics with transparent frames. In fact, one could read the Priorean case in parallel with the full general proof, using the former as an illustration of the latter.
The remainder of the article is organised as follows. The introduction is concluded with a brief discussion of related work. Section 2 contains the necessary modal logic preliminaries. Section 3 gives the bisimulation-based criterion of interpolant existence and applies it to first-order definable cofinal subframe logics above
$\mathsf {K4.3}$
. It also provides illustrative examples explaining why the same method does not work in general and what kind of descriptive frames might be needed. Section 4 establishes the quasi-finite bisimilar model property of all logics above
$\mathsf {K4.3}$
and the quasi-polysize bisimilar model property of all finitely axiomatisable ones; for the latter, it gives a coNP-algorithm for deciding the IEP. Section 5 extends the developed techniques to the Priorean temporal logics mentioned above.
1.1 Related work
The IEP for some logics of linear frames turns out to be closely related to separability of regular languages by first-order definable languages. Formally, the separability problem is to decide whether two input regular languages
$L_1$
and
$L_2$
can be separated by some language L in a given class
$\mathcal {L}$
in the sense that
$L_{1} \subseteq L$
and
$L \cap L_{2} = \emptyset $
. If
$\mathcal {L}$
is the class of first-order definable languages over finite words, the separability problem is equivalent to the IEP for the linear temporal logic
$\mathsf {LTL}$
extending modal logic with the operators ‘next’ and ‘until’ over finite strict linear orders. For regular languages of infinite words, the separability problem is equivalent to the IEP for
$\mathsf {LTL}$
over the natural numbers (see [Reference Kurucz, Wolter, Zakharyaschev, ten Cate, Jung, Koopmann, Wernhard and Wolter28] for details). It was shown in [Reference Henkell21, Reference Henkell, Rhodes and Steinberg22, Reference Place and Zeitoun34] that both of these separability problems are decidable in 2ExpTime in the size of NFAs defining
$L_1$
and
$L_2$
. It follows that the corresponding IEPs are decidable in 3ExpTime in the size of
$\mathsf {LTL}$
-formulas. (Separability by other language classes
$\mathcal {L}$
are discussed in [Reference Place33, Reference Place and Zeitoun35].) These separability results have been obtained using algebraic machinery from semigroup theory, which seems to be orthogonal to our model-theoretic approach to the IEP developed to deal with all modal logics of linear orders. However, for finite strict linear orders and the natural numbers, the algebraic approach also provides an upper bound for the size of interpolants.
It is also worth mentioning that, for these two frame classes, the smallest modal logic with the CIP is
$\mathsf {LTL}$
extended with fixed-point operators or, equivalently, monadic second-order logic (under very mild conditions on the definition of what a logic is) [Reference Gheerbrant, ten Cate, Grädel and Kahle15]. Thus, to ‘repair’ the CIP by extending the expressive power of the logic, we require the addition of second-order features.
2 Preliminaries
This section provides the basic definitions and facts that will be used later on in the article; consult [Reference Blackburn, de Rijke and Venema5, Reference Blackburn, van Benthem and Wolter6, Reference Chagrov and Zakharyaschev9, Reference Goldblatt16, Reference Goldblatt17] for more details.
2.1 Descriptive frames for normal modal logics
The formulas,
$\varphi $
, of propositional unimodal logics are built from propositional variables
$p_i \in \mathcal {V}$
, for some countably-infinite set
$\mathcal {V} = \{p_i \mid i < \omega \}$
, and constants
$\top $
,
$\bot $
using the Boolean connectives
$\neg $
,
$\land $
, and the unary possibility operator
$\Diamond $
. The other Booleans and the necessity operator
$\Box $
dual to
$\Diamond $
are defined as standard abbreviations. We also use
$\Diamond ^+\varphi = \varphi \lor \Diamond \varphi $
,
$\Box ^+\varphi =\varphi \land \Box \varphi $
, and
$\Diamond \Gamma =\{\Diamond \varphi \mid \varphi \in \Gamma \}$
, for a set
$\Gamma $
of formulas. By a signature we mean any set
$\sigma \subseteq \mathcal {V}$
, denoting by
$\textit {sig}(\varphi )$
the (finite) set of variables in a formula
$\varphi $
. If
$\textit {sig}(\varphi ) \subseteq \sigma $
, we call
$\varphi $
a
$\sigma $
-formula. We denote by
$\textit {sub}(\varphi )$
the set of subformulas of
$\varphi $
together with their negations, and let
$|\varphi |=|\textit {sub}(\varphi )|$
.
A (normal) modal logic, L, is any set of formulas that contains all Boolean tautologies, the modal axiom
$ \Box (p_0 \to p_1) \to (\Box p_0 \to \Box p_1) , $
and is closed under the rules of modus ponens, uniform substitution of formulas in place of variables, and necessitation
$\varphi /\Box \varphi $
. The smallest such logic goes by the moniker
$\mathsf {K}$
. Given a set
$\Gamma $
of formulas and a modal logic L, the smallest modal logic to contain L and
$\Gamma $
is denoted by
$L\oplus \Gamma $
. We write
$L\oplus \varphi $
for
$L\oplus \{\varphi \}$
. For example,
$$ \begin{align*} \mathsf{K4} & = \mathsf{K}\oplus \Box p_0 \to \Box\Box p_0, \\ \mathsf{K4.3} & = \mathsf{K4} \oplus \Box (\Box^+ p_0 \to p_1) \lor \Box (\Box^+ p_1 \to p_0),\\ \mathsf{GL.3} & = \mathsf{K4.3} \oplus \Box( \Box p_0 \to p_0) \to \Box p_0,\\ \mathsf{Log}\{(\mathbb N,<)\} & = \mathsf{K4.3} \oplus \Diamond\top \oplus \Box (\Box p_0 \to p_0) \to (\Diamond\Box p_0 \to \Box p_0). \end{align*} $$
All logics considered in this article are extensions of
$\mathsf {K4.3}$
.
We interpret formulas in (general) frames
$\mathfrak {F} = (W,R,\mathcal {P})$
, where R is a binary (accessibility) relation on a nonempty set W (of worlds or, more neutrally, points) and
$\mathcal {P} \subseteq 2^W$
contains
$\emptyset $
, W and is closed under
$\cap $
,
$\neg $
, and the operator
The structure
$\mathfrak F^+ = (\mathcal {P}, \cap ,\neg ,\emptyset , W, \Diamond ^{\mathfrak F})$
is a Boolean algebra
$(\mathcal {P}, \cap ,\neg ,\emptyset , W)$
with a normal and additive operator
$\Diamond ^{\mathfrak F}$
(BAO, for short). If
$\mathfrak F^+$
is generated by a set
$\mathcal X\subseteq \mathcal {P}$
as a BAO, we say that the frame
$\mathfrak F$
(or the set
$\mathcal {P}$
) is generated by
$\mathcal X$
. If
$|\mathcal X|=n$
, for some
$n<\omega $
, we call
$\mathfrak F n$
-generated or finitely generated. The elements of
$\mathcal {P}$
are called internal sets in
$\mathfrak F$
. If
$\mathcal {P} = 2^W$
,
$\mathfrak {F}$
is known as a Kripke frame; in this case, we drop
$\mathcal {P}$
and write
$\mathfrak {F}=(W,R)$
. A frame
$\mathfrak {F}=(W,R,\mathcal {P})$
is descriptive if the following conditions hold: for any
$x,y \in W$
and any
$\mathcal {X} \subseteq \mathcal {P}$
,
-
(dif)
$x = y$
iff
$\forall X \in \mathcal {P}\, (x \in X \leftrightarrow y \in X)$
, -
(tig)
$xRy$
iff
$\forall X\in \mathcal {P} \, (y \in X \to x \in \Diamond ^{\mathfrak F} X)$
, -
(com) if
$\mathcal {X} \subseteq \mathcal {P}$
has the finite intersection property (fip, for short)—that is,
$\bigcap \mathcal {X}' \ne \emptyset $
, for every finite
$\mathcal {X}' \subseteq \mathcal {X}$
—then
$\bigcap \mathcal {X} \ne \emptyset $
.
(Frames with (dif) are called differentiated, with (tig) tight, and with (com) compact.) Every BAO is isomorphic to
$\mathfrak F^+$
, for some descriptive frame
$\mathfrak F$
. A finite frame is descriptive iff it is a Kripke frame [Reference Chagrov and Zakharyaschev9, Section 8].
Given a signature
$\sigma $
, a
$\sigma $
-model based on a frame
$\mathfrak {F}=(W,R,\mathcal {P})$
is a pair
$\mathfrak {M}=(\mathfrak {F},\mathfrak {v})$
with a valuation
$\mathfrak {v} \colon \sigma \to \mathcal {P}$
. The atomic
$\sigma $
-type of
$x\in W$
in
$\mathfrak {M}$
is
We omit
$\sigma =\mathcal V$
, saying simply model and writing
$\textit {at}_{\mathfrak {M}}(x)$
. The value of a formula
$\varphi $
in
$\mathfrak M$
is the set
$\mathfrak {v}(\varphi ) \in \mathcal {P}$
computed inductively in the obvious way starting from
$\mathfrak v(p_i)$
,
$\mathfrak v(\top ) = W$
and
$\mathfrak v(\bot ) = \emptyset $
. A set
$X\subseteq W$
is definable in
$\mathfrak {M}$
if
$X=\mathfrak {v}(\varphi )$
, for some formula
$\varphi $
, in which case
$X \in \mathcal {P}$
. If every internal set
$X\in \mathcal {P}$
is definable in
$\mathfrak M$
, we say that
$\mathfrak {F}$
is
$\mathfrak {M}$
-generated. Every
$\mathfrak F$
with countable
$\mathcal {P}$
is clearly
$\mathfrak M$
-generated, for some model
$\mathfrak M$
.
A formula
$\varphi $
is true at x in
$\mathfrak M$
, written
$\mathfrak M,x \models \varphi $
, if
$x \in \mathfrak {v}(\varphi )$
. The
$\sigma $
-type of x in
$\mathfrak {M}$
is the set
$t^\sigma _{\mathfrak {M}}(x)$
of all
$\sigma $
-formulas that are true at x in
$\mathfrak M$
. For a set X of points in
$\mathfrak M$
, we let
$t_{\mathfrak {M}}^{\sigma }(X)=\big \{t_{\mathfrak {M}}^{\sigma }(x)\mid x\in X\big \}$
. As before, we drop
$\sigma =\mathcal V$
.
A set
$\Gamma $
of formulas is finitely satisfiable in
$\mathfrak M$
if, for every finite subset
$\Gamma ' \subseteq \Gamma $
, there is
$x' \in W$
such that
$\Gamma ' \subseteq t_{\mathfrak {M}}(x')$
;
$\Gamma $
is satisfiable in
$\mathfrak M$
if
$\Gamma \subseteq t_{\mathfrak {M}}(x)$
, for some
$x \in W$
. Using these definitions and notations, we can equivalently reformulate conditions (dif), (tig), and (com) for
$\mathfrak M$
-generated frames as follows: for any
$x,y \in W$
and any set
$\Gamma $
of formulas,
-
(dif)
$x = y$
iff
$t_{\mathfrak {M}}(x) = t_{\mathfrak {M}}(y)$
, -
(tig)
$xRy$
iff
$\Diamond t_{\mathfrak {M}}(y) \subseteq t_{\mathfrak {M}}(x)$
iff
$\{\varphi \mid \Box \varphi \in t_{\mathfrak {M}}(x)\}\subseteq t_{\mathfrak {M}}(y)$
, -
(com) if
$\Gamma $
is finitely satisfiable in
$\mathfrak M$
, then
$\Gamma $
is satisfiable in
$\mathfrak M$
.
A frame
$\mathfrak F$
satisfies
$\Gamma $
if there is a model
$\mathfrak M$
based on
$\mathfrak F$
satisfying
$\Gamma $
. Further,
$\varphi $
is valid in
$\mathfrak F$
, written
$\mathfrak F \models \varphi $
, if
$\mathfrak M,x \models \varphi $
, for any model
$\mathfrak M$
based on
$\mathfrak F$
and any
$x \in W$
. We call
$\mathfrak F$
a frame for a logic L and write
$\mathfrak F \models L$
if
$\mathfrak F \models \varphi $
, for all
$\varphi \in L$
. Conversely, any class
$\mathcal {S}$
of general frames determines the modal logic
$\mathsf {Log}\, \mathcal {S} = \{\varphi \mid \forall \mathfrak F \in \mathcal {S} \ \mathfrak F \models \varphi \}$
. We write
$\mathsf {Log}(\mathfrak F)$
for
$\mathsf {Log}(\{\mathfrak F\})$
.
A set
$\Gamma $
of formulas is L-consistent if
$(\bigwedge \Gamma '\to \bot )\notin L$
, for any finite
$\Gamma '\subseteq \Gamma $
. We require the following well-known fact (see, e.g., [Reference Chagrov and Zakharyaschev9, Section 8.6]).
Lemma 2.1. For any modal logic L and any finite signature
$\sigma $
, if
$\Sigma $
is an L-consistent set of
$\sigma $
-formulas, then
$\Sigma $
is satisfiable in a
$\sigma $
-model
$\mathfrak M$
based on a finitely
$\mathfrak M$
-generated descriptive frame for L.
By Lemma 2.1, every modal logic L is determined by the class of all descriptive frames for L. A logic L is Kripke complete if L is determined by the class of all Kripke frames for L. L is d-persistent (aka canonical) if
$(W,R,\mathcal {P}) \models L$
implies
$(W,R) \models L$
, for any descriptive frame
$(W,R,\mathcal {P})$
. L has the fmp if it is determined by its finite (Kripke) frames.
The smallest logic
$\mathsf {K4.3}$
we are interested in is d-persistent; its descriptive and Kripke frames
$\mathfrak F=(W,R,\mathcal {P})$
are transitive and weakly connected, that is,
$$ \begin{align*} & \forall x,y,z \in W\, (xRy \land yRz \to xRz),\\ & \forall x,y,z \in W\, (xRy\land xRz \to y = z \lor yRz \lor zRy). \end{align*} $$
$\mathsf {GL.3}$
, on the contrary, is not d-persistent yet has the fmp. In fact, all extensions of
$\mathsf {K4.3}$
are Kripke complete [Reference Fine12].
A frame
$\mathfrak F' = (W',R',\mathcal {P}')$
is a subframe of a frame
$\mathfrak F = (W,R,\mathcal {P})$
if
$W' \subseteq W$
,
$R' = {R}\mathop {\restriction }_{W'}=R\cap (W'\times W')$
, and
$\mathcal {P}' \subseteq \mathcal {P}$
. For every internal set
$V\in \mathcal {P}$
, the frame
${\mathfrak F}\mathop {\restriction }_{V}=\big (V,{R}\mathop {\restriction }_{V},{\mathcal {P}}\mathop {\restriction }_{V}\big )$
with
${\mathcal {P}}\mathop {\restriction }_{V}=\{V\cap X\mid X\in \mathcal {P}\}$
is a subframe of
$\mathfrak F$
. For a model
$\mathfrak M=(\mathfrak F,\mathfrak v)$
, we let
${\mathfrak M}\mathop {\restriction }_{V}=({\mathfrak F}\mathop {\restriction }_{V},{\mathfrak v}\mathop {\restriction }_{V})$
, where
${\mathfrak v}\mathop {\restriction }_{V}(p)=V\cap \mathfrak v(p)$
. Given a frame
$\mathfrak F = (W,R,\mathcal {P})$
with transitive R and a point
$x\in W$
, we define the frame
$\mathfrak F_x = (W_x,R_x,\mathcal {P}_x)$
by taking
$W_x = \{y \in W \mid xR^+y\}$
, where
$R^+$
is the reflexive closure of R (that is,
$R^+=R\cup \{(y,y) \mid y\in W\}$
),
$R_x = {R}\mathop {\restriction }_{W_x}$
, and
$\mathcal {P}_x = {\mathcal {P}}\mathop {\restriction }_{W_x}$
. We call
$\mathfrak {F}$
rooted if
$\mathfrak F = \mathfrak F_x$
, for some
$x \in W$
, in which case x is called a root of
$\mathfrak F$
. Note that
$\mathfrak F_x$
is not necessarily a subframe of
$\mathfrak F$
, but we have:
Indeed, suppose
$\mathfrak F = (W,R,\mathcal {P})$
is descriptive and
$x\in W$
. Conditions (dif) and (tig) for
$\mathfrak F_x$
are straightforward and left to the reader. To establish (com), consider any
$\mathcal {X}_x \subseteq \mathcal {P}_x$
with the fip. Then
also has the fip, and so
$\bigcap \mathcal {X} \ne \emptyset $
. To prove that
$\bigcap \mathcal {X}_x \ne \emptyset $
, it suffices to show that
$\bigcap \{V \in \mathcal {P} \mid W_x \subseteq V\} \subseteq W_x$
. To this end, suppose on the contrary that
$y \in \bigcap \{V \in \mathcal {P} \mid W_x \subseteq V\}$
and
$y \notin W_x$
. Then (dif) and (tig) give
$Z,Y \in \mathcal {P}$
such that
$x\in Z$
,
$y\notin Z$
,
$y \in Y$
, and
$x \in \Box \neg Y$
. It follows that
$Z\cup \neg Y \in \mathcal {P}$
,
$W_x \subseteq Z\cup \neg Y$
, and so
$y \in Z\cup \neg Y$
, which is a contradiction.
2.2 The structure of linear finitely-generated descriptive frames
From now on, all frames
$\mathfrak F=(W,R,\mathcal {P})$
are assumed to be rooted frames for
$\mathsf {K4.3}$
, so their relation R is always transitive and connected:
A cluster in
$\mathfrak F$
is any set of the form
$C(x) = \{x\} \cup \{ y \in W \mid xRy \land y R x\}$
with
$x\in W$
. If x is irreflexive, i.e.,
$xRx$
does not hold,
$C(x)$
is called a degenerate cluster and depicted as
$\bullet $
; a reflexive x (for which
$xRx$
) is depicted as
$\circ $
. A non-degenerate cluster with
$k \ge 1$
(reflexive) points is depicted as
. The next example will be used many times in what follows.
Example 2.2. Consider the frame
$\mathfrak F=(W_k,R_{k\bullet },\mathcal {P}_k)$
, where
$0 < k < \omega $
,
$$ \begin{align*} & W_k=A_k\cup\{b_n\mid n < \omega \},\qquad A_k= \{a_0,\dots, a_{k-1} \},\\ & xR_{k\bullet} y\ \text{ iff }\ \text{either } x = a_i \text{ or } x=b_n,\ y=b_m,\text{ and }m<n, \end{align*} $$
and
$\mathcal {P}_k$
is generated by the sets
$X_i = \{a_i\} \cup \{b_n\mid n < \omega ,\ n \equiv i \ (\text {mod}\ k) \}$
, for
$i <k$
, and
$\{b_n\}$
, for
$n < \omega $
. (For instance,
$\mathcal {P}_1$
consists of all finite subsets of
$\{b_n\mid n<\omega \}$
and their complements in
$W_1$
.) The underlying Kripke frame
$(W_k,R_{k\bullet })$
is shown in the picture below, where all
$\ast $
are
$\bullet $
.

It is not hard to see that
and so
$A_k\notin \mathcal {P}_k$
. For every nonempty
$X\in \mathcal {P}_k$
, the set
$\Diamond ^{\mathfrak F}X$
is cofinite in
$W_k$
. Using these observations, it is readily checked that
$\mathfrak F$
is a descriptive frame; we denote it by
. Clearly,
is
$\mathfrak M$
-generated for
$\mathfrak M$
with
$\mathfrak {v}(p_i)=X_i$
if
$i<k$
, and
$\mathfrak {v}(p_i)=\emptyset $
otherwise. The descriptive frame
$(W_k,R_{k\circ },\mathcal {P}_k)$
with
$R_{k\circ }=R_{k\bullet }\cup \{ (b_n,b_n)\mid n<\omega \}$
is denoted by
;
$(W_k,R_{k\circ })$
looks like in the picture above, with all
$\ast =\circ $
. Note that
but
, cf. Example 2.10
$(a)$
.
The next lemma, originating in [Reference Fine12], will play a key role in our subsequent constructions. Let
$\mathfrak {M}$
be a model based on a rooted frame
$\mathfrak {F} = (W,R,\mathcal {P})$
for
$\mathsf {K4.3}$
, and let
$\Gamma $
be a set of formulas. A point
$x\in W$
is called
$\Gamma $
-maximal in
$\mathfrak {M}$
if
$\mathfrak {M},x \models \Gamma $
, and whenever
$xRy$
and
$\mathfrak {M},y\models \Gamma $
, then
$yRx$
. We denote by
$\max _{\mathfrak {M}} \Gamma $
the set of all
$\Gamma $
-maximal points in
$\mathfrak {M}$
.
Lemma 2.3. Suppose
$\Gamma $
is a set of formulas and
$\mathfrak M$
is a model based on a rooted descriptive frame
$\mathfrak {F}= (W,R,\mathcal {P})$
for
$\mathsf {K4.3}$
. Then the following hold:
(modal saturation)
if
$\mathfrak M, x\models \Diamond \bigwedge \Gamma '$
for every finite
$\Gamma ' \subseteq \Gamma $
, then there is y with
$xRy$
and
$\mathfrak M, y \models \Gamma $
;
(maximal points)
if there is x with
$\mathfrak M, x \models \Gamma $
, then
$\max _{\mathfrak {M}}\Gamma \ne \emptyset $
.
Given a rooted frame
$\mathfrak F = (W, R,\mathcal {P})$
for
$\mathsf {K4.3}$
, let
$R^s = \{(x,y) \in R \mid (y,x) \notin R\}$
be the strict R-accessibility in
$\mathfrak F$
. Sometimes it will be convenient to view
$(W,R)$
as a strict linear order
$\mathfrak F_c = (W_c, <_R)$
of clusters, where
$W_c = \{C(x) \mid x \in W\}$
and
$C(x) <_R C(y)$
iff
$xR^sy$
. A cluster C is final in
$\mathfrak F$
if there is no cluster
$C'$
with
$C<_R C'$
. A cluster C is a root cluster if there is no cluster
$C'$
with
$C'<_R C$
, in which case
$C<_R C'$
for every
$C'\ne C$
in
$\mathfrak F$
; the root cluster in
$\mathfrak F$
is unique. A cluster
$C'$
is an immediate successor of a cluster C in
$\mathfrak F$
if
$C<_R C'$
and there is no
$C"$
with
$C<_R C"<_R C'$
, in which case C is an immediate predecessor of
$C'$
. A sequence
$C_n$
,
$n<\omega $
, of clusters in
$\mathfrak F_c$
is an infinite ascending chain if
$C_n <_R C_{n+1}$
, for all
$n < \omega $
.
$\mathfrak F_c$
is converse well-founded if it has no infinite ascending chain of clusters.
The next lemma follows from, e.g., the more general [Reference Chagrov and Zakharyaschev9, Theorems 10.34 and 10.35].
Lemma 2.4. If
$\mathfrak F$
is a rooted n-generated descriptive frame for
$\mathsf {K4.3}$
, for some
$n<\omega $
, then:
-
(a)
$\mathfrak F_c$
is converse well-founded, and so the strict linear order
$\mathfrak F_c^{-1} = (W_c,>_R)$
is isomorphic to some ordinal; -
(b) every cluster in
$\mathfrak F$
has at most
$2^n$
points.
Proof. Let
$\mathfrak F=(W,R,\mathcal {P})$
, let
$\le _R$
be the reflexive closure of
$<_R$
, and let
$\mathcal G$
be a finite set generating
$\mathcal {P}$
with
$|\mathcal G| = n$
. For
$x,y\in W$
, we write
$x\sim _{\mathcal G} y$
in case
$x\in G$
iff
$y\in G$
, for all
$G\in \mathcal G$
, and denote by
$[x]_{\mathcal G}$
the
$\sim _{\mathcal G}$
-class of x. Clearly,
$\big |\big \{[x]_{\mathcal G}\mid x\in W\big \}\big |\le 2^{|{\mathcal G}|}=2^n$
.
$(a)$
Suppose on the contrary that
$C(x_i)$
,
$i<\omega $
, is an infinite ascending chain in
$\mathfrak F_c$
. Call
$x \in W$
a middle-point if
$C(x_0)\le _R C(x)\le _R C(x_i)$
, for some
$i<\omega $
. Let
$V_x=\{[y]_{\mathcal G}\mid y \text { a middle-point with } xR y\}$
. Since
$V_x\supseteq V_y$
whenever
$xR y$
and each
$V_x$
is finite, there is
$m<\omega $
such that
$V_y=V_{x_m}$
, for every middle-point y with
$C(x_m)\le _R C(y)$
. By induction on the construction of
$X\in \mathcal {P}$
from the generators in
$\mathcal G$
, it is readily seen that
$$ \begin{align} \text{if } y,z \text{ are middle-points, } C(x_m)\le_RC(y), & C(x_m)\le_RC(z), \text{ and } y\sim_{\mathcal G}z,\\ & \text{then } y\in X \text{ iff } z\in X, \text{ for all } X\in\mathcal{P}.\nonumber \end{align} $$
(Indeed, the only non-trivial case is when
$X=\Diamond ^{\mathfrak F}Y$
,
$yRz$
and
$y\in \Diamond ^{\mathfrak F}Y$
. Then there is
$x\in Y$
with
$yRx$
. If
$zRx$
, we are done. Otherwise, x is a middle-point. As
$V_y=V_z$
, there is a middle-point
$x'$
with
$zRx'$
and
$x\sim _{\mathcal G} x'$
. By IH,
$x'\in Y$
.) As there are finitely many
$\sim _{\mathcal G}$
-classes, there exist
$k\ne \ell \geq m$
such that
$x_k\sim _{\mathcal G}x_\ell $
, and so
$x_k\in X$
iff
$x_\ell \in X$
, for all
$X\in \mathcal {P}$
, by (4). But this contradicts (dif).
$(b)$
It is straightforward to show that if
$C(x)=C(y)$
and
$x\sim _{\mathcal G}y$
, then
$x\in X$
iff
$y\in X$
, for all
$X\in \mathcal {P}$
. So by (dif), every cluster in
$\mathfrak F$
has
$\le 2^{|{\mathcal G}|}$
points.
Note that the existence of maximal points (Lemma 2.3) in models based on rooted finitely generated descriptive frames for
$\mathsf {K4.3}$
also follows from Lemma 2.4. Another consequence is that such a frame
$\mathfrak F$
contains a unique final cluster, and any non-root cluster in
$\mathfrak F$
has an immediate predecessor. If
$\mathfrak F_c^{-1}= (W_c,>_R)$
is isomorphic to an ordinal
$\gamma $
and
$\alpha \le \gamma $
, we denote by
$C_\alpha ^{\mathfrak F}$
the cluster that is the image of
$\alpha $
under this isomorphism. If
$\alpha $
is a non-zero limit ordinal, we call
$C_\alpha ^{\mathfrak F}$
a limit cluster. A non-final cluster is a limit cluster iff it does not have an immediate successor. By (dif) and Lemma 2.4
$(b)$
, we also have the following.
Lemma 2.5. If
$\mathfrak F = (W,R,\mathcal {P})$
is a rooted finitely generated descriptive frame for
$\mathsf {K4.3}$
and
$C\in \mathcal {P}$
, for some cluster C, then
$\{x\}\in \mathcal {P}$
, for every
$x\in C$
.
Now, suppose
$\mathfrak M$
is a model based on a rooted finitely
$\mathfrak M$
-generated descriptive frame
$\mathfrak {F}= (W,R,\mathcal {P})$
for
$L\supseteq \mathsf {K4.3}$
. Given a formula
$\mu $
, a cluster C is called
$\mu $
-maximal in
$\mathfrak {M}$
if there is a point in C that is
$\{\mu \}$
-maximal in
$\mathfrak M$
. Further, C is maximal in
$\mathfrak {M}$
if it is
$\mu $
-maximal in
$\mathfrak M$
, for some
$\mu $
, and C is
$\sigma $
-maximal in
$\mathfrak {M}$
, for a signature
$\sigma $
, if there is such a
$\sigma $
-formula
$\mu $
. Every definable in
$\mathfrak M$
cluster is clearly maximal in
$\mathfrak M$
. The next lemma says that the converse is also true.
Lemma 2.6. Suppose
$\mathfrak {M}$
is a model based on a rooted finitely
$\mathfrak M$
-generated descriptive frame
$\mathfrak F = (W, R,\mathcal {P})$
for
$\mathsf {K4.3}$
. Then
-
(a) every degenerate cluster in
$\mathfrak F$
is maximal in
$\mathfrak {M}$
; -
(b) a cluster is maximal in
$\mathfrak {M}$
iff either it is final or has an immediate successor; -
(c) a cluster is definable in
$\mathfrak {M}$
iff it is maximal in
$\mathfrak {M}$
.
So limit clusters are not definable and not degenerate, while every other cluster is definable in
$\mathfrak M$
.
Proof.
$(a)$
If
$C(x)$
is degenerate, then
$\Diamond t_{\mathfrak M}(x) \not \subseteq t_{\mathfrak M}(x)$
by (tig). So there is a formula
$\mu $
with
$\mathfrak M,x\models \mu $
but
$\mathfrak M,x\not \models \Diamond \mu $
.
$(b,\Rightarrow )$
Let
$C(x)$
be maximal in
$\mathfrak M$
with
$\mathfrak M,x \models \mu $
and
$\mathfrak M,y \not \models \mu $
whenever
$xR^sy$
. Suppose
$C(x)$
is a limit cluster. Let
$S=\{C\in W_c\mid C(x)<_R C\}$
with
$y_C\in C$
, for
$C\in S$
. Consider
Clearly,
$\Gamma $
is finitely satisfiable in
$\mathfrak M$
, and so, by (com),
$\Gamma \subseteq t_{\mathfrak M}(y)$
, for some y. Thus, by (tig),
$x R y R y_C$
for all
$C\in S$
, and so
$yR^s y_C$
for all
$C\in S$
and
$yRx$
. But we also have
$\mathfrak M,y\models \Box \neg \mu $
, contrary to
$\mathfrak M,x\models \mu $
.
$(b,\Leftarrow )$
The (unique) final cluster is maximal in
$\mathfrak M$
for
$\top $
. Suppose
$C(y)$
is an immediate successor of
$C(x)$
. If
$C(y)$
is degenerate, then
$C(y)$
is maximal in
$\mathfrak M$
by
$(a)$
, and so there is
$\mu $
with
$\mathfrak M,y \models \mu \land \neg \Diamond \mu $
. It follows that
$C(x)$
is
$\Diamond (\mu \land \neg \Diamond \mu )$
-maximal in
$\mathfrak M$
. If
$C(y)$
is non-degenerate and
$C(x)$
is not maximal in
$\mathfrak M$
, then
$\Diamond t_{\mathfrak M}(x) \subseteq t_{\mathfrak M}(y)$
, and so
$yRx$
by (tig), contrary to
$xR^s y$
.
$(c,\Leftarrow )$
Let
$C(x)$
be
$\mu $
-maximal in
$\mathfrak M$
. If
$C(x)$
is degenerate, it is defined by
$\mu \land \neg \Diamond \mu $
. If
$C(x)$
is the non-degenerate root cluster, then
$\Diamond \mu $
defines
$C(x)$
. Otherwise, take the immediate predecessor
$C(y)$
of
$C(x)$
. By
$(b)$
,
$C(y)$
is
$\tau $
-maximal in
$\mathfrak M$
, for some
$\tau $
, so
$\Box ^+ \neg \tau \land \Diamond \mu $
defines
$C(x)$
.
$(c,\Rightarrow )$
is obvious.
We require a few important consequences of Lemmas 2.4 and 2.6.
Lemma 2.7. If
$\mathfrak F = (W,R,\mathcal {P})$
is a rooted finitely generated descriptive frame for
$\mathsf {K4.3}$
, then W is countable.
Proof. By Lemma 2.4, it suffices to show that the ordinal
$\gamma $
isomorphic to
$\mathfrak F_c^{-1}= (W_c,>_R)$
is countable. Let
$Z=\{\alpha +1\mid \alpha <\gamma ,\ \alpha +1\ne \gamma \}$
be the set of successor ordinals
$<\gamma $
. Then
$|Z|=|\gamma |$
and
$C_\beta ^{\mathfrak F}\in \mathcal {P}$
, for any
$\beta \in Z$
, by Lemma 2.6. As
$\mathfrak F$
is finitely generated,
$\mathcal {P}$
is countable, and so are Z and W.
Given a rooted finitely
$\mathfrak M$
-generated descriptive frame
$\mathfrak F = (W, R,\mathcal {P})$
for
$\mathsf {K4.3}$
, let
$m_{\mathfrak F}$
be the largest ordinal
$\le \omega $
with degenerate
$C_n^{\mathfrak F}$
for all
$n<m_{\mathfrak F}$
. We call the (possibly empty) interval
$Z=\bigcup _{n<m_{\mathfrak F}}C_n^{\mathfrak F}$
the tail of
$\mathfrak F$
. We may assume that
$Z=\{z_n\mid n< m_{\mathfrak F}\}$
, where all
$z_n$
are irreflexive and
$z_{n}R z_{n-1}$
,
$0<n<m_{\mathfrak F}$
. If Z is infinite, then
$Z\ne W$
(as
$\mathfrak F$
is rooted). If
$Z\ne W$
, we call
$C_{m_{\mathfrak F}}^{\mathfrak F}$
the head of Z. In particular, if
$Z=\emptyset $
, its head is the final (non-degenerate) cluster
$C_{0}^{\mathfrak F}$
; if
$Z\ne W$
and
$Z\ne \emptyset $
is finite, its head is the immediate predecessor of
$C_{m_{\mathfrak F}-1}^{\mathfrak F}=\{z_{m_{\mathfrak F}-1}\}$
; and if Z is infinite, its head is the limit cluster
$C_{\omega }^{\mathfrak F}$
. Thus, by Lemma 2.6,
2.3 Building linear models from pieces
Definition 2.8. The ordered sum
$\mathfrak F_{0} \lhd \dots \lhd \mathfrak F_{n-1} = (W, R,\mathcal {P})$
of rooted frames
$\mathfrak F_{i}= ( W_i,R_i,\mathcal {P}_i)$
,
$i<n$
, for
$\mathsf {K4.3}$
with pairwise disjoint
$W_i$
is defined by
It is not hard to see that if the
$\mathfrak F_i$
are descriptive, then
$\mathfrak F_{0} \lhd \dots \lhd \mathfrak F_{n-1}$
is also descriptive. If
$\mathfrak M_i = (\mathfrak F_i, \mathfrak v_i)$
, then
$\mathfrak M=\mathfrak M_{0} \lhd \dots \lhd \mathfrak M_{n-1}$
is the model based on
$\mathfrak F_{0} \lhd \dots \lhd \mathfrak F_{n-1}$
with the valuation
$\mathfrak v(p) = \bigcup _{ i < n} \mathfrak v_i(p)$
, for any
$p \in \mathcal {V}$
. We call the
$\mathfrak M_i \lhd $
-components of
$\mathfrak M$
.
Now, let
$\mathfrak F = (W, R,\mathcal {P})$
be a rooted frame for
$\mathsf {K4.3}$
. An interval in
$\mathfrak {F}$
is any subset
$I\subseteq W$
such that
$xRyRz$
and
$x,z \in I$
imply
$y \in I$
, for all
$x,y,z \in W$
. If
$I\cap C\ne \emptyset $
, for a cluster C, then clearly
$C\subseteq I$
. An interval I is closed if there are clusters
$C,C'$
such that
$I=C\cup C'\cup \bigcup \{ D\mid C<_R D <_R C'\}$
, in which case we write
$I=[C,C']$
. Given two closed intervals
$I,I'$
in
$\mathfrak F$
, we write
$I\prec _{\mathfrak F}I'$
if I and
$I'$
are disjoint and
$xRx'$
, for all
$x\in I$
,
$x'\in I'$
. Notice that if I is a closed internal interval in
$\mathfrak F$
, then
${\mathfrak F}\mathop {\restriction }_{I}$
is also a rooted frame for
$\mathsf {K4.3}$
. Also, if
$\mathfrak F$
is descriptive, then
${\mathfrak F}\mathop {\restriction }_{I}$
is descriptive as well. And if
$\mathfrak F$
is finitely
$\mathfrak M$
-generated, for some model
$\mathfrak M$
, then
${\mathfrak F}\mathop {\restriction }_{I}$
is finitely
${\mathfrak M}\mathop {\restriction }_{I}$
-generated. We clearly have the following.
Lemma 2.9. Suppose
$\mathfrak F= (W, R,\mathcal {P})$
is a rooted frame for
$\mathsf {K4.3}$
and W is partitioned as
$\{I_j\mid j<n\}$
,
$n<\omega $
, with closed intervals
$I_j\in \mathcal {P}$
and
$I_j\prec _{\mathfrak F}I_k$
iff
$j<k$
. Then
-
(a)
$\mathfrak F={\mathfrak F}\mathop {\restriction }_{I_0}\lhd \dots \lhd {\mathfrak F}\mathop {\restriction }_{I_{n-1}}$
; -
(b) if
$\mathfrak M$
is a model based on
$\mathfrak F$
, then
$\mathfrak M={\mathfrak M}\mathop {\restriction }_{I_0}\lhd \dots \lhd {\mathfrak M}\mathop {\restriction }_{I_{n-1}}$
.
2.4 Canonical formulas
To check whether a frame validates a given finitely axiomatisable logic, we use the canonical formulas of [Reference Bezhanishvili and Bezhanishvili4, Reference Chagrov and Zakharyaschev9, Reference Wolter38, Reference Zakharyaschev and Alekseev43] whose basic properties are summarised below in the context of
$\mathsf {K4.3}$
; for more details consult [Reference Chagrov and Zakharyaschev9, Section 16.3]. Every logic
$L \supseteq \mathsf {K4.3}$
can be represented in the form
where each
$\alpha (\mathfrak G_j, \mathfrak D_j,\bot )$
is a canonical formula based on a finite rooted Kripke frame
$\mathfrak G_j = (V_j, S_j)$
for
$\mathsf {K4.3}$
and a (possibly empty) set
$\mathfrak D_j\subseteq V_j$
of irreflexive non-root points in
$\mathfrak G_j$
. If L is finitely axiomatisable, its canonical axiomatisation (6) with finite
$J_L$
can be constructed effectively, given any finite set of axioms.
Let
$\mathfrak F = (W,R,\mathcal {P})$
be any rooted finitely generated descriptive frame for
$\mathsf {K4.3}$
. By Theorem 2.4,
$\mathfrak F$
contains a unique final cluster, and any non-root cluster in
$\mathfrak F$
has an immediate predecessor. The formulas
$\alpha (\mathfrak G_j, \mathfrak D_j,\bot )$
are defined so that
$\mathfrak F\not \models \alpha (\mathfrak G_j, \mathfrak D_j,\bot )$
iff there is an injection
$f \colon V_j \to W$
such that the following conditions hold: for all
$x,y \in V_j$
,
-
(cf1)
$x S_j y$
iff
$f(x) R f(y)$
(so x is irreflexive iff
$f(x)$
is); -
(cf2) if
$C(x)$
is the final cluster in
$\mathfrak G_j$
, then
$C(f(x))$
is the final cluster in
$\mathfrak F$
; -
(cf3) if
$x \in \mathfrak D_j$
and
$C(y)$
is the immediate predecessor of
$C(x) = \{x\}$
in
$\mathfrak G_j$
, then
$C(f(y))$
is the immediate predecessor of
$C(f(x))=\{f(x)\}$
in
$\mathfrak F$
; -
(cf4)
$\{f(x)\} \in \mathcal {P}$
.
Intuitively, every frame
$\mathfrak F$
with
$\mathfrak F \not \models \alpha (\mathfrak G_j, \mathfrak D_j,\bot )$
can be obtained by inserting certain chains of clusters immediately before some clusters
$C(x)$
in
$\mathfrak G_j$
, provided that
$x \notin \mathfrak D_j$
, and by enlarging some non-degenerate clusters in
$\mathfrak G_j$
.
Canonical formulas of the form
$\alpha (\mathfrak G, \emptyset ,\bot )$
axiomatise exactly the cofinal subframe logics whose frames are closed under taking cofinal subframes. We remind the reader [Reference Chagrov and Zakharyaschev9] that a subframe
$\mathfrak F' = (W',R',\mathcal {P}')$
of a frame
$\mathfrak F = (W,R,\mathcal {P})$
is called cofinal if
$W'$
is cofinal in
$\mathfrak F$
in the sense that, for any
$x \in W'$
and
$y \in W$
, whenever
$x R y$
then either
$y \in W'$
or there is
$z \in W'$
with
$y R z$
. Cofinal subframe logics enjoy the fmp, and so are decidable if finitely axiomatisable [Reference Zakharyaschev42]. Example 2.10 shows the canonical axioms of some extensions of
$\mathsf {K4.3}$
.
Example 2.10.
$(a)$
We prove that
Let
$\mathfrak F = (W,R,\mathcal {P})$
be a rooted finitely generated descriptive frame for
$\mathsf {K4.3}$
. By Lemma 2.7, W is countable, and so
$\mathfrak F$
is
$\mathfrak M$
-generated, for some model
$\mathfrak M=(\mathfrak {F},\mathfrak {v})$
. We claim that the following are equivalent:
-
1.
$\mathfrak M\not \models \mathsf {GL.3}$
; -
2. there is a formula
$\psi $
with a non-degenerate
$\psi $
-maximal cluster in
$\mathfrak M$
; -
3. there is a non-degenerate non-limit cluster in
$\mathfrak F$
; -
4.
$\mathfrak F\not \models \alpha (\circ , \emptyset , \bot ) \land \alpha (\circ \lhd \bullet , \emptyset , \bot )$
.
1.
$\Rightarrow $
2. Suppose
$\mathfrak M,x\not \models \Box (\Box \varphi \to \varphi )\to \Box \varphi $
, for some formula
$\varphi $
. Then the
$\neg \big (\Box (\Box \varphi \to \varphi )\to \Box \varphi \big )$
-maximal cluster C in
$\mathfrak M$
is non-degenerate.
2.
$\Leftrightarrow $
3. by Lemma 2.6.
2.
$\Rightarrow $
4. Suppose the
$\psi $
-maximal cluster
$C_\psi $
in
$\mathfrak M$
is non-degenerate, for some
$\psi $
. If
$C_\psi $
is the final cluster of
$\mathfrak F$
, then the injection f mapping
$\circ $
to a point in
$C_\psi $
satisfies
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
, and so
$\mathfrak F\not \models \alpha (\circ , \emptyset , \bot )$
. If
$C_\psi $
is not the final cluster, then the
$\neg \psi $
-maximal cluster
$C_{\neg \psi }$
is the final cluster in
$\mathfrak F$
. If
$C_{\neg \psi }$
is non-degenerate, then again
$\mathfrak F\not \models \alpha (\circ , \emptyset , \bot )$
; otherwise
$\mathfrak F\not \models \alpha (\circ \lhd \bullet , \emptyset , \bot )$
as witnessed by f sending
$\bullet $
to the point in the final cluster and
$\circ $
to a point in
$C_\psi $
.
4.
$\Rightarrow $
1. If
$\mathfrak F\not \models \alpha (\circ \lhd \bullet , \emptyset , \bot )$
, then take an injection f from
$\circ \lhd \bullet $
to
$\mathfrak F$
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
. By
$\mathbf {(cf_4)}$
and Lemma 2.6,
$\{f(\circ )\}=\mathfrak {v}(\varphi )$
, for some
$\varphi $
. As
$f(\circ )R f(\circ )$
by
$\mathbf {(cf_1)}$
, it is easy to see that
$\mathfrak M,f(\circ )\not \models \Box (\Box \neg \varphi \to \neg \varphi )\to \Box \neg \varphi $
. The case when
$\mathfrak F\not \models \alpha (\circ , \emptyset , \bot )$
is similar.
$(b)$
Similarly, we can prove that
$$ \begin{align*} \mathsf{Log}\{(\mathbb N,<)\} & = \mathsf{K4.3} \oplus \Diamond\top \oplus \Box (\Box p_0 \to p_0) \to (\Diamond\Box p_0 \to \Box p_0)\\ & = \mathsf{K4.3} \oplus \alpha (\bullet, \emptyset, \bot) \oplus \alpha (\circ \lhd \circ, \emptyset, \bot) \end{align*} $$
by showing that, for every
$\mathfrak M$
and
$\mathfrak F$
as above, the following are equivalent:
-
–
$\mathfrak M\not \models \mathsf {Log}\{(\mathbb N,<)\}$
; -
– either the final cluster in
$\mathfrak F$
is degenerate or there is a non-degenerate non-limit cluster different from the final cluster in
$\mathfrak F$
; -
–
$\mathfrak F\not \models \alpha (\bullet , \emptyset , \bot ) \land \alpha (\circ \lhd \circ , \emptyset , \bot )$
.
$(c)$
A prominent example of a non-cofinal subframe logic is
$\mathsf {K4.3} \oplus \Diamond p \to \Diamond \Diamond p$
with dense frames, whose canonical axioms
forbid any two consecutive degenerate clusters in finitely generated descriptive frames for the logic (see also Lemma 5.6).
3 Craig interpolant existence: Warming up
In this section, we first give a model-theoretic, bisimulation-based criterion of interpolant non-existence, then apply it to design a coNP-algorithm deciding the IEP in any finitely axiomatisable d-persistent cofinal subframe logic containing
$\mathsf {K4.3}$
. Finally, we illustrate by examples that a way more involved approach is needed to tackle arbitrary finitely axiomatisable extensions of
$\mathsf {K4.3}$
.
A formula
$\iota $
is called a Craig interpolant of formulas
$\varphi _1$
and
$\varphi _2$
in a logic L if
$\textit {sig}(\iota ) \subseteq \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
and both
$\varphi _1 \to \iota $
and
$\iota \to \varphi _2$
are in L. We say that L has the CIP if an interpolant for
$\varphi _1$
and
$\varphi _2$
exists whenever
$(\varphi _1\rightarrow \varphi _2) \in L$
.
Many standard modal logics have the CIP, including
$\mathsf {K}$
,
$\mathsf {K4}$
, and
$\mathsf {S4}$
. In fact, there are a continuum of logics containing
$\mathsf {K4}$
with the CIP. However, none of the continuum-many extensions of
$\mathsf {K4.3}$
with frames of unbounded depth has the CIP, and very few—not more than 37—out of the continuum-many logics containing
$\mathsf {S4}$
enjoy the CIP (deciding whether a finitely axiomatisable logic above
$\mathsf {S4}$
has the CIP is in coNExpTime and PSpace-hard). The reader can find proofs of these results and further references in [Reference Chagrov and Zakharyaschev9, Reference Gabbay and Maksimova14 (see also Example 3.6).
We now introduce the model-theoretic notions and tools that are needed in our non-uniform approach to deciding interpolant existence in modal logics.
Given two models
$\mathfrak {M}_i$
,
$i=1,2$
, based on
$\mathfrak F_i = (W_i,R_i,\mathcal {P}_i)$
with
$x_i \in W_i$
, we write
$\mathfrak {M}_1,x_1 \equiv _{\sigma } \mathfrak {M}_2,x_2$
, for a signature
$\sigma $
, if
$t_{\mathfrak {M}_1}^{\sigma }(x_1)=t_{\mathfrak {M}_2}^{\sigma }(x_2)$
. The equivalence relation
$\equiv _{\sigma }\subseteq W_1\times W_2$
can be characterised in terms of bisimulations. Namely, a relation
$\boldsymbol {\beta } \subseteq W_1 \times W_2$
is called a
$\sigma $
-bisimulation between
$\mathfrak M_1$
and
$\mathfrak M_2$
if the following conditions hold whenever
$x_1 \boldsymbol {\beta } x_2$
:
-
(atom)
${at}_{\mathfrak {M}_1}^{\sigma }(x_1) ={at}_{\mathfrak {M}_2}^{\sigma }(x_2)$
; -
(move) if
$x_1R_1y_1$
, then there is
$y_2$
such that
$x_2R_2y_2$
and
$y_1 \boldsymbol {\beta } y_2$
; and, conversely, if
$x_2R_2 y_2$
, then there is
$y_1$
with
$x_1R_1 y_1$
and
$y_1 \boldsymbol {\beta } y_2$
.
If there is such
$\boldsymbol {\beta }$
with
$z_1\boldsymbol {\beta } z_2$
, we write
$\mathfrak {M}_1,z_1 \sim _{\sigma } \mathfrak {M}_2,z_2$
. We call
$\boldsymbol {\beta }$
global if, for every
$x_1\in W_1$
, there is
$x_2\in W_2$
with
$x_1\boldsymbol {\beta } x_2$
, and, for every
$x_2\in W_2$
, there is
$x_1\in W_1$
with
$x_1\boldsymbol {\beta } x_2$
. In this case, we say that
$\mathfrak {M}_1$
and
$\mathfrak {M}_2$
are globally
$\sigma $
-bisimilar and write
$\mathfrak {M}_1\sim _{\sigma } \mathfrak {M}_2$
.
We employ the following characterisation of
$\equiv _{\sigma }$
(see [Reference Goranko and Otto20] for a further discussion of the relationship between bisimulations and modal equivalence).
Lemma 3.1. For any signature
$\sigma $
, any models
$\mathfrak {M}_i$
,
$i = 1,2$
, based on descriptive frames
$\mathfrak F_i = (W_i,R_i,\mathcal {P}_i)$
, and any
$x_i \in W_i$
,
The implication
$(\Leftarrow )$
holds for arbitrary models.
Proof.
$(\Rightarrow )$
We show that
$\{(y_1,y_2)\in W_1\times W_2\mid t_{\mathfrak {M}_1}^{\sigma }(y_1)=t_{\mathfrak {M}_2}^{\sigma }(y_2) \}$
is a
$\sigma $
-bisimulation between
$\mathfrak M_1$
and
$\mathfrak M_2$
. Condition (atom) is obvious. For (move), suppose
$y_1R_1z_1$
and
$t_{\mathfrak {M}_1}^{\sigma }(y_1)=t_{\mathfrak {M}_2}^{\sigma }(y_2)$
. Let
$\Gamma =t_{\mathfrak {M}_1}^{\sigma }(z_1)$
. Then, for every finite
$\Gamma ' \subseteq \Gamma $
, we have
$\mathfrak M_1,y_1\models \Diamond \bigwedge \Gamma '$
, and so
$\mathfrak M_2,y_2\models \Diamond \bigwedge \Gamma '$
as well. Since
$\mathfrak F_2$
is descriptive, Lemma 2.3 gives us
$z_2$
with
$y_2R_2z_2$
and
$\mathfrak M_2, z_2 \models \Gamma $
. It follows that
$t_{\mathfrak {M}_1}^{\sigma }(z_1)=t_{\mathfrak {M}_2}^{\sigma }(z_2)$
, as required. The implication
$(\Leftarrow )$
is straightforward.
Note that if
$\mathcal {B}$
is a set of
$\sigma $
-bisimulations between
$\mathfrak M_1$
and
$\mathfrak M_2$
, then
$\bigcup _{\boldsymbol {\beta }\in \mathcal {B}} \boldsymbol {\beta }$
is also a
$\sigma $
-bisimulation between
$\mathfrak M_1$
and
$\mathfrak M_2$
. It follows that there is always a largest
$\sigma $
-bisimulation between
$\mathfrak M_1$
and
$\mathfrak M_2$
(which is
$\equiv _\sigma $
if both
$\mathfrak M_i$
are based on descriptive frames).
Variations of the following criterion of interpolant (non-)existence are implicit in various (dis-)proofs of the CIP in modal logics [Reference Goranko and Otto20, Reference Marx32].
Theorem 3.2. Formulas
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in a modal logic L iff there are models
$\mathfrak M_i$
,
$i=1,2$
, based on finitely
$\mathfrak M_i$
-generated descriptive frames
$\mathfrak {F}_i = (W_{i},R_{i},\mathcal {P}_{i})$
for L with points
$x_i \in W_i$
such that
If
$L \supseteq \mathsf {K4}$
, we may assume that
$x_i$
is the root of the descriptive frame
$\mathfrak F_i$
,
$i=1,2$
.
Proof.
$(\Leftarrow )$
is straightforward (and holds for arbitrary frames for L). For
$(\Rightarrow )$
, consider the signature
$\delta = \textit {sig}(\varphi _1) \cup \textit {sig}(\varphi _2)$
and the set
of
$\delta $
-formulas. As
$\varphi _1$
and
$\varphi _2$
have no interpolant in L,
$\Sigma $
is L-consistent, and so, by Lemma 2.1, there exists a
$\delta $
-model
$\mathfrak M_2$
based on a finitely
$\mathfrak M_2$
-generated descriptive frame
$\mathfrak F_2$
and a point
$x_2$
with
$\mathfrak M_2,x_2\models \Sigma $
. Let
$\Sigma '=t_{\mathfrak M_2}^\sigma (x_2)\cup \{\varphi _1\}$
. As
$\Sigma '$
is an L-consistent set of
$\delta $
-formulas, Lemma 2.1 gives a
$\delta $
-model
$\mathfrak M_1$
based on a finitely
$\mathfrak M_1$
-generated descriptive frame
$\mathfrak F_1$
and an
$x_1$
in
$\mathfrak M_1$
such that
$\mathfrak M_1,x_1\models \Sigma '$
. We clearly have
$t_{\mathfrak M_1}^\sigma (x_1)=t_{\mathfrak M_2}^\sigma (x_2)$
, and so
$\mathfrak {M}_{1},x_{1} \sim _{\sigma } \mathfrak {M}_{2},x_{2}$
by Lemma 3.1. In case
$L \supseteq \mathsf {K4}$
, (1) allows us to make
$x_i$
the root of
$\mathfrak F_i$
.
The next lemma refines Theorem 3.2; it is used in the proof of Lemma 4.21.
Lemma 3.3. If
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in a logic
$L \supseteq \mathsf {K4.3}$
, then there are rooted models
$\mathfrak M_i,x_i$
,
$i=1,2$
, satisfying the criterion Theorem 3.2 such that
$C(x_i)$
is not a limit cluster in
$\mathfrak M_i$
, for
$i=1,2$
.
Proof. Suppose
$\mathfrak M,x$
is a rooted
$\delta $
-model, for some finite signature
$\delta $
, that is based on a finitely
$\mathfrak M$
-generated descriptive frame
$\mathfrak F=(W,R,\mathcal {P})$
such that
$\mathfrak M,x\models \varphi $
, for some
$\varphi $
with
$\textit {sig}(\varphi )\subseteq \delta $
, and
$C(x)$
is a root limit cluster in
$\mathfrak F$
. Pick a fresh variable
$q\notin \delta $
. For
$\ast \in \{\bullet , \circ \}$
, take the frames
$\mathfrak F^\ast = \ast \lhd \mathfrak F$
, denote the root point of
$\mathfrak F^\ast $
by
$x^\ast $
, and consider the
$\delta \cup \{q\}$
-models
$\mathfrak M^\ast $
based on
$\mathfrak F^\ast $
, which coincide with
$\mathfrak M$
on
$\mathfrak F$
and have
$\mathfrak M^\ast , x^\ast \models p$
iff
$\mathfrak M, x\models p$
, for
$p\in \delta $
, and
$\mathfrak M^\ast , x^\ast \models q$
. To prove the lemma, it suffices to show that there is
$\ast \in \{\bullet , \circ \}$
with
$(i) \mathfrak M^\ast ,x^\ast \models \varphi $
,
$(ii) \mathfrak M^\ast ,x^\ast \sim _\sigma \mathfrak M,x$
, for any
$\sigma \subseteq \textit {sig}(\varphi )$
, and
$(iii) \mathsf {Log}(\mathfrak F)\subseteq \mathsf {Log}(\mathfrak F^\ast )$
.
As the limit cluster
$C(x)$
is non-degenerate by Lemma 2.6, we have
$(i)$
and
$(ii)$
. To show
$(iii)$
, suppose on the contrary that, for each
$\ast \in \{\bullet , \circ \}$
, there is a canonical formula
$\alpha (\mathfrak G^\ast , \mathfrak D^\ast , \bot )$
with
$\mathfrak F\models \alpha (\mathfrak G^\ast , \mathfrak D^\ast , \bot )$
and
$\mathfrak F^\ast \not \models \alpha (\mathfrak G^\ast , \mathfrak D^\ast , \bot )$
. Let
$f^\ast $
be an injection from
$\mathfrak G^\ast $
to
$\mathfrak F^\ast $
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
for
$\alpha (\mathfrak G^\ast , \mathfrak D^\ast , \bot )$
, and let
$C(r^\ast )$
be the root-cluster in
$\mathfrak G^\ast $
and
$C(y^\ast )$
its immediate successor in
$\mathfrak G^\ast $
. By assumption,
$f^\ast $
is not an injection from
$\mathfrak G^\ast $
to
$\mathfrak F$
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
, so
$f^\ast (r^\ast )=x^\ast $
and
$f^\ast (y^\ast ) \in W$
. As
$\{f^\ast (y^\ast )\}\in \mathcal {P}$
by
$\mathbf {(cf_4)}$
and
$C(x)$
is a limit cluster, it follows from Lemma 2.6 that
$f^\ast (y^\ast ) \notin C(x)$
, and so
$y^\ast \notin \mathfrak D^\ast $
. Suppose, for definiteness, that
$f^\circ (y^\circ )R f^\bullet (y^\bullet )$
or
$f^\circ (y^\circ )=f^\bullet (y^\bullet )$
. Let C be the immediate predecessor of
$C(f^\circ (y^\circ ))$
in
$\mathfrak F$
. Then C is a non-limit cluster. By Lemma 2.6,
$C\in \mathcal {P}$
and, by Lemma 2.5,
$\{z\}\in \mathcal {P}$
, for every
$z\in C$
. If C is non-degenerate, then we modify
$f^\circ $
by taking
$f^\circ (r^\circ )\in C$
; otherwise, we modify
$f^\bullet $
by taking
$f^\bullet (r^\bullet ) \in C$
. In either case, the modified
$f^\ast $
is an injection from
$\mathfrak G^\ast $
to
$\mathfrak F$
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
, a contradiction.
We begin our study of the IEP by showing how the criterion of Theorem 3.2 can be used to decide whether given formulas have an interpolant in any fixed d-persistent cofinal subframe logic
$L \supseteq \mathsf {K4.3}$
(defined in Section 2.4). Suppose that
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in L. Let
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
. By Theorem 3.2, there exist models
$\mathfrak M_i$
,
$i=1,2$
, based on descriptive frames
$\mathfrak {F}_i = (W_{i},R_{i},\mathcal {P}_{i})$
for L with roots
$x_i \in W_i$
such that
$\mathfrak {M}_{1},x_{1} \sim _{\sigma } \mathfrak {M}_{2},x_{2}$
,
$\mathfrak {M}_{1},x_{1} \models \varphi _1$
and
$\mathfrak {M}_{2}, x_{2} \models \neg \varphi _2$
. We may assume that
$\boldsymbol {\beta }$
is the largest
$\sigma $
-bisimulation
$\equiv _\sigma $
between
$\mathfrak M_1$
and
$\mathfrak M_2$
(for which
$x_1 \boldsymbol {\beta } x_2$
, of course). We show how to extract from the
$\mathfrak M_i$
polynomial-size models
$\mathfrak M^{\prime }_i$
that still witness that
$\varphi _1$
and
$\varphi _2$
lack an interpolant in L. We proceed in two steps.
-
Step 1: For each
$i=1,2$
and each
$\tau \in \textit {sub}(\varphi _i)$
satisfied in
$\mathfrak {M}_i$
, we take a
$\{\tau \}$
-maximal point
$y_{\tau }\in W_i$
(which exists by Lemma 2.3), and denote the set of all these
$y_\tau $
by
$\boldsymbol {M}_{i} \subseteq W_i$
. Note that
$\boldsymbol {M}_{i}$
is cofinal in
$\mathfrak F_i$
because each point in
$W_i \setminus \boldsymbol {M}_{i}$
has a
$\{\varphi _i\}$
- or
$\{\neg \varphi _i\}$
-maximal
$R_i$
-successor. Set (7)
$$ \begin{align} T=\big\{ t_{\mathfrak{M}_{1}}^{\sigma}(x) \mid x\in \{x_{1}\}\cup \boldsymbol{M}_{1}\big\} \cup\big\{ t_{\mathfrak{M}_{2}}^{\sigma}(x) \mid x\in \{x_{2}\}\cup \boldsymbol{M}_{2}\big\}. \end{align} $$
-
Step 2: As
$\mathfrak {M}_{1},x_{1} \sim _{\sigma } \mathfrak {M}_{2},x_{2}$
and
$\boldsymbol {\beta }$
is the largest
$\sigma $
-bisimulation, each
$t\in T$
is satisfied in both
$\mathfrak M_i$
. For
$i=1,2$
, we take a smallest set
$\boldsymbol {S}_{i} \subseteq W_i$
containing a t-maximal point
$z_t$
in
$\mathfrak M_i$
(which exists by Lemma 2.3), for each
$t\in T$
.
Now, let
$W_{i}'=\{x_{i}\}\cup \boldsymbol {M}_{i} \cup \boldsymbol {S}_{i}$
,
$R^{\prime }_i = {R_i}\mathop {\restriction }_{W^{\prime }_i}$
,
$\mathfrak F^{\prime }_i = (W^{\prime }_i,R^{\prime }_i)$
, and let
$\mathfrak {M}_{i}'$
be the restriction of
$\mathfrak {M}_{i}$
to
$\mathfrak {F}_{i}'$
. We let
Clearly,
$|W_i'|\leq \boldsymbol {k}(\varphi _1,\varphi _2)$
, so the size of
$\mathfrak M_i$
is
$\mathcal {O}\big (\max (|\varphi _1|,|\varphi _2|)\big )$
. As L is d-persistent,
$(W_i,R_i) \models L$
. By construction,
$\mathfrak F^{\prime }_i$
is a cofinal subframe of
$(W_i, R_i)$
, and so
$\mathfrak F^{\prime }_i \models L$
as L is a cofinal subframe logic. Finally, we define
$\boldsymbol {\beta }'$
as the restriction of
$\boldsymbol {\beta }$
to
$W_{1}'\times W_{2}'$
:
$x_1' \boldsymbol {\beta }' x_2'$
iff
$t_{\mathfrak {M}_1}^{\sigma }(x_1')=t_{\mathfrak {M}_2}^{\sigma }(x_2')$
, for all
$x_1'\in W_1'$
,
$x_2'\in W_2'$
.
Lemma 3.4.
$(a)\ \mathfrak {M}_{1}',x_{1}\models \varphi _1$
,
$\mathfrak {M}_{2}',x_{2}\models \neg \varphi _2$
and
$(b)\ \boldsymbol {\beta }'$
is a
$\sigma $
-bisimulation between
$\mathfrak {M}_{1}'$
and
$\mathfrak {M}_{2}'$
with
$x_1 \boldsymbol {\beta }' x_2$
.
Proof.
$(a)$
follows from the fact that, for any
$\tau \in \textit {sub}(\varphi _i)$
and
$x\in W^{\prime }_i$
,
$\mathfrak M_i,x\models \tau $
iff
$\mathfrak M^{\prime }_i,x\models \tau $
, which can be established by a straightforward induction on the construction of
$\varphi _1$
and
$\varphi _2$
. We only show
$(\Rightarrow )$
for
$\tau =\Diamond \psi $
. If
$\mathfrak M_i,x\models \Diamond \psi $
, then there is
$y\in W_i$
with
$xR_iy$
and
$\mathfrak M_i,y\models \psi $
. Take
$y_\psi \in \boldsymbol {M}_{i}\subseteq W_i'$
. By the
$\{\psi \}$
-maximality of
$y_\psi $
, either
$y=y_\psi $
or
$yR_iy_\psi $
, and so
$xR_i'y_\psi $
and
$\mathfrak M^{\prime }_i,x\models \Diamond \psi $
.
$(b)$
Condition (atom) follows from the definition. To establish (move), assume
$x \boldsymbol {\beta }' x'$
and
$x R^{\prime }_1 y$
. Let
$t=t^\sigma _{\mathfrak M_1}(y)$
. Then
$t\in T$
, and so there is a t-maximal
$z_t\in \boldsymbol {S}_{2}\subseteq W_2'$
in
$\mathfrak M_2$
. In particular,
$t^\sigma _{\mathfrak M_2}(z_t)=t$
, and so
$y \boldsymbol {\beta }' z_t$
. As
$x \boldsymbol {\beta } x'$
and
$\boldsymbol {\beta }$
is the largest
$\sigma $
-bisimulation, there is
$z\in W_2$
with
$x'R_2z$
and
$t^\sigma _{\mathfrak M_2}(z)=t$
. It follows from the t-maximality of
$z_t$
that
$z=z_t$
or
$zR_2 z_t$
, and so
$x'R_2'z_t$
, as required.
Thus, the fact that
$\varphi _1$
and
$\varphi _2$
have no interpolant in L can always be witnessed (in the sense of Theorem 3.2) by models
$\mathfrak M_i$
of size polynomial in
$\max (|\varphi _1|,|\varphi _2|)$
, and so we can say that L has the polysize bisimilar model property. This gives the first claim of the following theorem.
Theorem 3.5.
$(a)$
All d-persistent cofinal subframe logics
$L \supseteq \mathsf {K4.3}$
have the polysize bisimilar model property.
$(b)$
If such an L is consistent and finitely axiomatisable, then the IEP for L is coNP-complete.
Proof. We show that
$(a) \Rightarrow (b)$
(cf. Theorem 4.9 in Section 4). Indeed, suppose L is given by (6) (with
$\mathfrak G_j=(V_j,S_j)$
and
$\mathfrak D_j= \emptyset $
, for all j in the finite index set
$J_L$
). To decide whether formulas
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in L, we guess polynomial-size pointed models
$\mathfrak M_i,x_i$
based on Kripke frames
$\mathfrak F_i=(W_i,R_i)$
for
$\mathsf {K4.3}$
and restricted to the variables in
$\varphi _1$
and
$\varphi _2$
. The conditions
$\mathfrak M_1,x_1\models \varphi _1$
and
$\mathfrak M_2,x_2\models \neg \varphi _2$
are clearly polynomially checkable; that
$\mathfrak {M}_{1},x_{1} \sim _{\sigma } \mathfrak {M}_{2},x_{2}$
, for
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
, can be established in polynomial time using a standard technique from [Reference Baier and Katoen2, Chapter 7]. Finally, to check whether
$\mathfrak F_i \models \alpha (\mathfrak G_j, \emptyset ,\bot )$
, for each
$j\in J_L$
, we simply enumerate all injective functions from
$\mathfrak G_j$
to
$\mathfrak F_i$
, whose number does not exceed
$|W_i|^{|V_j|}$
, and verify that at least one of them satisfies
$\mathbf {(cf_1)}$
and
$\mathbf {(cf_2)}$
, which can obviously be done in time polynomial in
$|W_i|$
. (Condition
$\mathbf {(cf_3)}$
holds vacuously, and
$\mathbf {(cf_4)}$
always holds as
$\mathfrak F_i$
is a Kripke frame.)
We now give two examples illustrating that the construction above does not work for logics that are not d-persistent, even for logics with the fmp. Prominent examples of such logics are
$\mathsf {GL.3}$
and
$\mathsf {Log}\{(\mathbb N,<)\}$
(see Example 2.10). We show that, for these logics, establishing model-theoretically (using Theorem 3.2) that some formulas do not have an interpolant requires a pair of models that are based on infinite descriptive (non-Kripke) frames.
Example 3.6.
$(a)$
Consider the following formulas
$\varphi _1$
and
$\varphi _2$
:
$$ \begin{align} \nonumber & \varphi_1 = \Diamond( p_{1} \wedge \Diamond^+ \neg q_{1}) \wedge \Box (p_{2} \rightarrow \Box^+ q_{1}),\\ & \varphi_2 = \neg [ \Diamond( p_{2} \wedge \Diamond^+\neg q_{2}) \land \Box (p_{1} \rightarrow \Box^+ q_{2}) ]. \end{align} $$
To show that
$(\varphi _1 \rightarrow \varphi _2) \in \mathsf {K4.3} \subseteq \mathsf {GL.3}$
, suppose otherwise. Then there exists a model
$\mathfrak M$
based on a frame
$\mathfrak F = (W,R)$
for
$\mathsf {K4.3}$
and
$z \in W$
with
$\mathfrak M, z \models \varphi _1 \land \neg \varphi _2$
. So we have
$x,x',y,y' \in W$
with
$z R x R^+ x'$
,
$z R y R^+ y'$
,
$\mathfrak M, x \models p_1$
,
$\mathfrak M, x' \models \neg q_1$
,
$\mathfrak M, y \models p_2$
, and
$\mathfrak M, y' \models \neg q_2$
. Since
$\mathfrak F$
is a frame for
$\mathsf {K4.3}$
, either
$x'=y'$
or
$x'Ry'$
or
$y'Rx'$
. However, none of these is possible because of the boxed conjuncts of
$\varphi _1$
and
$\neg \varphi _2$
.
We now use Theorem 3.2 to show that
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in
$\mathsf {GL.3}$
. Let
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2) = \{p_1,p_2\}$
. Observe that any models
$\mathfrak M_i$
meeting the conditions of Theorem 3.2 cannot be based on a Kripke frame
$\mathfrak F_i=(W_i,R_i)$
for
$\mathsf {GL.3}$
. Indeed, let
$\boldsymbol {\beta }$
be the corresponding bisimulation. Then
$\mathfrak {M}_{1},x_{1} \models \varphi _1$
implies that there is
$x^1_1 \in W_1$
with
$x_1 R_1 x^1_1$
and
$\mathfrak {M}_{1},x^1_{1} \models p_1$
; we must also have
$\mathfrak {M}_{1},y_{1} \models \neg q_1$
, for some
$y_1$
with
$x^1_1 R^+_1 y_1$
. Similarly,
$\mathfrak {M}_{2},x_{2} \models \neg \varphi _2$
implies that there is
$x^1_2 \in W_2$
with
$x_2 R_1 x^1_2$
and
$\mathfrak {M}_{2},x^1_{2} \models p_2$
, and we also have
$\mathfrak {M}_{2},y_{2} \models \neg q_2$
, for some
$y_2$
with
$x^1_2 R^+_2 y_2$
. As
$x_{1} \boldsymbol {\beta } x_{2}$
and
$x_1 R_1 x^1_1$
, (move) gives
$x^2_2$
with
$x_2 R_2 x^2_2$
and
$x^1_{1} \boldsymbol {\beta } x^2_{2}$
. But then
$\mathfrak {M}_{2},x^2_{2} \models p_1$
, and so
$x_2 R_2 x^1_2 R_2^+ y_2R_2 x^2_2$
since
$\mathfrak F_2$
is a frame for
$\mathsf {K4.3}$
and in view of
$\neg \varphi _2$
’s second conjunct. Symmetrically, we find
$x^2_1$
with
$x_1 R_1 x^1_1 R_1^+y_1 R_1 x^2_1$
and
$x^2_1 \boldsymbol {\beta } x^1_2$
. Using (move), we construct infinite ascending chains of not necessarily distinct points as shown in the picture below.

It follows that the
$\mathfrak F_i$
are not frames for
$\mathsf {GL.3}$
(see any of [Reference Blackburn, de Rijke and Venema5, Reference Chagrov and Zakharyaschev9, Reference Goldblatt19] for details).
We now give a descriptive frame for
$\mathsf {GL.3}$
that can be used to show that
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in
$\mathsf {GL.3}$
. Take the descriptive frame
defined in Example 2.2 and construct
(see Definition 2.8), which is a frame for
$\mathsf {GL.3}$
by property
$(iii)$
in Example 2.10
$(a)$
. Consider the rooted models
$\mathfrak M_i, x_i$
,
$i=1,2$
, shown in Figure 1, both of which are based on a frame isomorphic to
$\mathfrak F$
. It is readily checked that
$\mathfrak {M}_{1},x_1 \models \varphi _1$
,
$\mathfrak {M}_{2},x_2 \models \neg \varphi _2$
, and the depicted relation
$\boldsymbol {\beta }$
is a
$\sigma $
-bisimulation between
$\mathfrak M_1$
and
$\mathfrak M_2$
with
$x_1\boldsymbol {\beta } x_2$
.

Figure 1
$\sigma $
-bisimilar models based on a descriptive frame for
$\mathsf {GL.3}$
.
In fact, the argument above shows that none of the logics L in the interval
$\mathsf {K4.3} \subseteq L \subseteq \mathsf {GL.3}$
has the CIP.
$(b)$
Consider next the logic
$\mathsf {Log}\{(\mathbb N,<)\}$
and show that the formulas
and
$\varphi _2$
given by (9) do not have an interpolant in it, though
$(\varphi _1 \rightarrow \varphi _2) \in \mathsf {K4.3}$
, and so
$(\varphi _1' \rightarrow \varphi _2) \in \mathsf {K4.3} \subseteq \mathsf {Log}\{(\mathbb N,<)\}$
.
As in
$(a)$
above, any models
$\mathfrak M_i$
,
$i=1,2$
, satisfying the conditions of Theorem 3.2 for
$\varphi _1'$
and
$\varphi _2$
cannot be based on Kripke frames, however the reason for this is slightly different. Suppose
$\boldsymbol {\beta }$
is a bisimulation witnessing these conditions. Then the models
$\mathfrak {M}_i$
must contain infinite ascending chains such as those in Example 3.6
$(a)$
. Also, the model
$\mathfrak {M}_1$
with
$\mathfrak {M}_1,x_1 \models \varphi _1'$
must contain a point z such that
$x_1R_1z$
and
$\mathfrak {M}_1,z \models r \land \Box \neg p_1$
, which means that z is located after all of the
$x_1^j$
,
$j < \omega $
. But then the Kripke frame
$\mathfrak F_1$
underlying
$\mathfrak {M}_1$
is not a frame for
$\mathsf {Log}\{(\mathbb N,<)\}$
, as it refutes its axiom
$\Box (\Box p \to p) \to (\Diamond \Box p \to \Box p)$
if we make p true everywhere after the initial ascending chain in
$\mathfrak F_1$
and false elsewhere.
The picture below shows models
$\mathfrak M_1$
and
$\mathfrak M_2$
based on
and satisfying the conditions of Theorem 3.2 for
$\varphi _1'$
and
$\varphi _2$
. That this frame is a frame for
$\mathsf {Log}\{(\mathbb N,<)\}$
follows from Example 2.10
$(b)$
.

4 Interpolant existence in logics above
$\mathsf {K4.3}$
We now generalise Theorem 3.5 to all finitely axiomatisable logics containing
$\mathsf {K4.3}$
. It turns out that, even though these logics do not have the polysize bisimilar model property in general, the structure of the models required in Theorem 3.2 is perfectly understandable. We show that one can assemble a pair of bisimilar models witnessing the absence of an interpolant for
$\varphi _1$
and
$\varphi _2$
in any
$L \supseteq \mathsf {K4.3}$
as the ordered sum of finitely-many ‘nice’ models, which are either finite or infinite but finitely ‘presentable’. Hence, we say that all
$L\supseteq \mathsf {K4.3}$
have the ‘quasi-finite bisimilar model property’. Moreover, if L is finitely axiomatisable, we can replace ‘finite’ by ‘polynomial in
$\boldsymbol {c}_L$
and
$\max (|\varphi _1|,|\varphi _2|)$
’, for some constant
$\boldsymbol {c}_L$
depending on L only. In this case, we say that L has the ‘quasi-polysize bisimilar model property’.
Section 4 is organised as follows. In Section 4.1, we formulate our main results (Theorems 4.5–4.7 and 4.9), and show how Theorem 4.6 implies Theorem 4.9. In Sections 4.2 and 4.3, we prove Theorem 4.5. Then, in Section 4.4, we show how to fine-tune the proof of Theorem 4.5 and obtain proofs of Theorems 4.6 and 4.7. Finally, in Section 4.5, we formulate and prove an interesting consequence of our methods for cofinal subframe logics (Theorem 4.25).
4.1 The quasi-polysize bisimilar model property
Given a finite signature
$\delta $
, a
$\delta $
-model
$\mathfrak M=(\mathfrak F,\mathfrak w)$
is called simple if either
$\mathfrak F$
is finite or
, for
$0<k<\omega $
and
$\ast \in \{\bullet ,\circ \}$
, and, for every
$p \in \delta $
, there is
$A_p\subseteq \{0,\dots ,k-1\}$
with
$\mathfrak {w}(p)=\bigcup _{i\in A_p}X_i$
, where the
$X_i$
are the infinite generators of the internal sets in
defined in Example 2.2. Thus, even though
is infinite, any simple
$\delta $
-model based on it is fully determined by the finitary information provided by the sets
$A_p$
,
$p\in \delta $
, that is, by the atomic
$\delta $
-types of the points in the
-cluster. A
$\delta $
-model is called quasi-finite if it is the ordered sum of finitely-many simple models.
Definition 4.1. A logic
$L \supseteq \mathsf {K4.3}$
is said to have the quasi-finite bisimilar model property if, for any formulas
$\varphi _1$
,
$\varphi _2$
without an interpolant in L, there are rooted quasi-finite
$\delta $
-models
$\mathfrak N_1,x_1$
and
$\mathfrak N_2,x_2$
satisfying conditions
$(a)$
–
$(c)$
below, for
$\delta = \textit {sig}(\varphi _1) \cup \textit {sig}(\varphi _2)$
and
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
:
-
(a)
$\mathfrak N_1,x_1\models \varphi _1$
and
$\mathfrak N_2,x_2\models \neg \varphi _2$
; -
(b)
$\mathfrak N_1$
and
$\mathfrak N_2$
are based on frames for L; -
(c)
$\mathfrak {N}_{1},x_{1} \sim _{\sigma } \mathfrak {N}_{2},x_{2}$
.
Our first result is as follows.
Theorem 4.2. All
$L \supseteq \mathsf {K4.3}$
have the quasi-finite bisimilar model property.
We actually prove a stronger Theorem 4.5 that prescribes more structure for the pair of quasi-finite models witnessing the lack of an interpolant, which makes it easy to deduce the existence of a
$\sigma $
-bisimulation between the models. The prescribed structure is easily checkable, which is used in the proof of the main Theorem 4.9. To formulate our ‘structural’ theorem, we require a few definitions.
For
$0 < m < \omega $
, let
$m^{<} = \underbrace {\bullet \lhd \dots \lhd \bullet }_m$
. An atomic frame takes one of the forms

The size
$\|{\mathfrak {F}}\|$
of an atomic
$\mathfrak {F}$
is defined by taking
$\|{m^{<}}\|=m$
,
, and
. If
$\mathfrak M=\mathfrak M_0\lhd \dots \lhd \mathfrak M_{n-1}$
, for some
$0<n<\omega $
and simple
$\delta $
-models
$\mathfrak M_j$
based on atomic frames
$\mathfrak F_j$
,
$j<n$
, then we set
$\|{\mathfrak {M}}\|=\|{\mathfrak F_0\lhd \dots \lhd \mathfrak F_{n-1}}\|= \|{\mathfrak F_0}\|+\dots +\|{\mathfrak F_{n-1}}\|$
.
Definition 4.3. Suppose
$\mathfrak N_i$
,
$i=1,2$
, is the ordered sum of finitely-many simple
$\delta $
-models based on atomic frames. The pair
$(\mathfrak N_1,\mathfrak N_2)$
is called
$\sigma $
-matching if it satisfies one of the following conditions
$({a})$
–
$({c})$
:
-
(a)
$\mathfrak N_1$
and
$\mathfrak N_2$
are simple models based on the same atomic frame
$\mathfrak H$
with
$\textit {at}^\sigma _{\mathfrak N_1}(y)=\textit {at}^\sigma _{\mathfrak N_2}(y)$
, for every point y in
$\mathfrak H$
; -
(b) the final clusters
$C_i$
of
$\mathfrak N_i$
,
$i=1,2$
, are non-degenerate and, for every point
$y_1$
in
$\mathfrak N_1$
, there is
$y_2 \in C_2$
with
$\textit {at}^\sigma _{\mathfrak N_1}(y_1)=\textit {at}^\sigma _{\mathfrak N_2}(y_2)$
, and, for every point
$y_2$
in
$\mathfrak N_2$
, there is
$y_1 \in C_1$
with
$\textit {at}^\sigma _{\mathfrak N_2}(y_2) = \textit {at}^\sigma _{\mathfrak N_1}(y_1)$
; - (c)
-
1. the last
$\lhd $
-components of the
$\mathfrak N_1$
and
$\mathfrak N_2$
are based on the same atomic frame
$\mathfrak G$
of the form
or
, with
$0<k\le 2^{|\delta |}$
; -
2.
$\textit {at}^\sigma _{\mathfrak N_1}(y)=\textit {at}^\sigma _{\mathfrak N_2}(y)$
, for every point y in the root
-cluster
$A_k$
of
$\mathfrak G$
; -
3. for every point
$y_1$
in any non-last
$\lhd $
-component of
$\mathfrak N_1$
, there is
$y_2 \in A_k$
with
$\textit {at}^\sigma _{\mathfrak N_1}(y_1)=\textit {at}^\sigma _{\mathfrak N_2}(y_2)$
and, for every point
$y_2$
in any non-last
$\lhd $
-component of
$\mathfrak N_2$
, there is
$y_1\in A_k$
with
$\textit {at}^\sigma _{\mathfrak N_2}(y_2)=\textit {at}^\sigma _{\mathfrak N_1}(y_1)$
.
-
If
$(\mathfrak N_1,\mathfrak N_2)$
satisfies condition
$({x})$
, for
$x=a,b,c$
, we say that it is of type
$({x})$
.

The following lemma justifies this definition.
Lemma 4.4. Suppose
$\mathfrak N_i=\mathfrak N_i^{0}\lhd \dots \lhd \mathfrak N_i^{N-1}$
, for
$i=1,2$
and
$0<N<\omega $
, and
$x_i$
is a root of
$\mathfrak N_i^0$
. If
$\textit {at}^\sigma _{\mathfrak N_1}(x_1) = \textit {at}^\sigma _{\mathfrak N_2}(x_2)$
and, for every
$\ell <N$
, the pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is
$\sigma $
-matching, then
$\mathfrak N_1,x_1\sim _\sigma \mathfrak N_2,x_2$
.
Proof. First, we show that, for every
$\ell <N$
, there is a global
$\sigma $
-bisimulation between
$\mathfrak N_1^\ell $
and
$\mathfrak N_2^\ell $
. This is clear for
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
of type
$({a})$
, in which case the identity function on
$\mathfrak H$
is a
$\sigma $
-bisimulation.
If
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is of type
$({b})$
, then the final clusters
$C_i$
of
$\mathfrak N_i^\ell $
,
$i=1,2$
, are non-degenerate. Thus,
$\boldsymbol {\beta }_1\cup \boldsymbol {\beta }_2$
is a global
$\sigma $
-bisimulation between
$\mathfrak N_1^\ell $
and
$\mathfrak N_2^\ell $
,
$$ \begin{align*} & \boldsymbol{\beta}_1=\big\{(y_1,y_2) \mid {y_1 \text{ in } \mathfrak N_1^\ell, y_2 \text{ in } C_2, {at}^\sigma_{\mathfrak N_1^\ell}(y_1)={at}^\sigma_{\mathfrak N_2^\ell}(y_2)} \big\},\\ & \boldsymbol{\beta}_2=\big\{(y_1,y_2) \mid {y_2 \text{ in } \mathfrak N_2^\ell, y_1 \text{ in } C_1, {at}^\sigma_{\mathfrak N_1^\ell}(y_1)={at}^\sigma_{\mathfrak N_2^\ell}(y_2)} \big\}. \end{align*} $$
If
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is of type
$({c})$
, suppose
$\mathfrak N_i^\ell =\mathfrak N_i^{0,\ell }\lhd \dots \lhd \mathfrak N_i^{\boldsymbol {n}_i-1,\ell }$
, for
$0<\boldsymbol {n}_i<\omega $
and
$i=1,2$
. By
$({c})$
.1,
$\mathfrak N_1^{\boldsymbol {n}_1-1,\ell }$
and
$\mathfrak N_2^{\boldsymbol {n}_2-1,\ell }$
are simple models based on the same atomic frame of the form
or
. As in Example 2.2, let
$A_k=\{a_s\mid s<k\}$
and
$W_k=A_k\cup \{b_n\mid n<\omega \}$
(containing all the points of
). We claim that
Indeed, suppose
$n<\omega $
and let
$s<k$
be such that
$n\equiv s$
(mod k). As
$\mathfrak N_i^{\boldsymbol {n}_i-1,\ell }$
is a simple model, we have
and so (11) follows from
$({c})$
.2. Now let
$$ \begin{align*} & \boldsymbol{\beta}_1=\big\{(y_1,y_2) \mid {y_1 \text{ in } \mathfrak N_1^{0,\ell}\lhd\dots\lhd\mathfrak N_1^{\boldsymbol{n}_1-2,\ell}, y_2\in A_k, {at}^\sigma_{\mathfrak N_1^\ell}(y_1)={at}^\sigma_{\mathfrak N_2^\ell}(y_2)} \big\},\\ & \boldsymbol{\beta}_2=\big\{(y_1,y_2) \mid {y_2 \text{ in } \mathfrak N_2^{0,\ell}\lhd\dots\lhd\mathfrak N_2^{\boldsymbol{n}_2-2,\ell}, y_1\in A_k, {at}^\sigma_{\mathfrak N_1^\ell}(y_1)={at}^\sigma_{\mathfrak N_2^\ell}(y_2)} \big\}. \end{align*} $$
By (11) and
$({c})$
.2–3,
$\boldsymbol {\beta }_1\cup \boldsymbol {\beta }_2\cup \big \{(b_n,b_n)\mid n<\omega \big \}$
is a global
$\sigma $
-bisimulation between
$\mathfrak N_1^\ell $
and
$\mathfrak N_2^\ell $
. Finally, if
$\boldsymbol {\beta }^0$
is a global
$\sigma $
-bisimulation between
$\mathfrak N_1^0$
and
$\mathfrak N_2^0$
, then
$\boldsymbol {\beta }^0\cup \{(x_1,x_2)\}$
is also a global
$\sigma $
-bisimulation between
$\mathfrak N_1^0$
and
$\mathfrak N_2^0$
because
$\textit {at}^\sigma _{\mathfrak N_1}(x_1) = \textit {at}^\sigma _{\mathfrak N_2}(x_2)$
. The union of the constructed global bisimulations is a (global) bisimulation
$\boldsymbol {\beta }$
between
$\mathfrak N_1$
and
$\mathfrak N_2$
with
$x_1\boldsymbol {\beta } x_2$
, as required.
The following strengthening of Theorem 4.2 will be proved in Sections 4.2 and 4.3.
Theorem 4.5. For any logic
$L \supseteq \mathsf {K4.3}$
and formulas
$\varphi _1$
,
$\varphi _2$
without an interpolant in L, there are rooted
$\delta $
-models
$\mathfrak N_1,x_1$
and
$\mathfrak N_2,x_2$
satisfying
$(a)$
–
$(d)$
below, for
$\delta = \textit {sig}(\varphi _1) \cup \textit {sig}(\varphi _2)$
and
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
:
-
(a)
$\mathfrak N_1,x_1\models \varphi _1$
and
$\mathfrak N_2,x_2\models \neg \varphi _2$
; -
(b) each
$\mathfrak N_i$
,
$i=1,2$
, is based on a frame for L; -
(c)
$\textit {at}^\sigma _{\mathfrak N_1}(x_1) = \textit {at}^\sigma _{\mathfrak N_2}(x_2)$
; -
(d) there is
$N=\mathcal {O}\big (\max (|\varphi _1|,|\varphi _2|)\big )$
such that
$\mathfrak N_i=\mathfrak N_i^0\lhd \dots \lhd \mathfrak N_i^{N-1}$
,
$i=1,2$
, and, for any
$\ell <N$
,-
1. each
$\mathfrak N_i^\ell $
is the ordered sum of
$\mathcal {O}\big (\max (|\varphi _1|,|\varphi _2|)\big )$
-many simple
$\delta $
-models based on atomic frames; -
2. the pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is
$\sigma $
-matching.
-
Observe that the models provided by Theorem 4.5 are ordered sums of polynomially-many simple models. However, the sizes of these simple models are not necessarily polynomial in
$\max (|\varphi _1|,|\varphi _2|)$
. Our second main result shows that all finitely axiomatisable logics
$L \supseteq \mathsf {K4.3}$
have the stronger quasi-polysize bisimilar model property: the lack of an interpolant can be witnessed by a pair quasi-finite models of polynomial size. More precisely, suppose L is given by its canonical axioms as
$L = \mathsf {K4.3} \oplus \{\alpha (\mathfrak G_j,\mathfrak D_j,\bot ) \mid j\in J_L\}$
, for some finite set
$J_L$
and
$\mathfrak G_j=(V_j,S_j)$
. Let
$\boldsymbol {c}_L=\max _{j\in J_L}|V_j|$
. An atomic frame in (10) is called L-bounded if it is of the form
$m^{<}$
or
with
$m\le \boldsymbol {c}_L+1$
, or it has one of the three remaining forms with
for the polynomial number
$\boldsymbol {k}(\varphi _1,\varphi _2)$
defined in (8). In Section 4.4, we prove the following.
Theorem 4.6. For any finitely axiomatisable logic
$L \supseteq \mathsf {K4.3}$
and formulas
$\varphi _1$
,
$\varphi _2$
without an interpolant in L, there are rooted
$\delta $
-models
$\mathfrak N_1,x_1$
and
$\mathfrak N_2,x_2$
satisfying
$(a)$
–
$(d)$
from Theorem 4.5, in which condition
$(d).1$
is strengthened to
-
1. each
$\mathfrak N_i^\ell $
,
$i=1,2$
, is the ordered sum of
$\mathcal {O}\big (\max (|\varphi _1|,|\varphi _2|)\big )$
-many simple
$\delta $
-models based on L-bounded atomic frames.
In Section 4.4, we also show the following.
Theorem 4.7. All finitely axiomatisable
$L \supseteq \mathsf {K4.3}$
have the quasi-polysize bisimilar model property, with the size of witnessing models bounded by
Remark 4.8. As a consequence we obtain that each finitely axiomatisable logic
$L \supseteq \mathsf {K4.3}$
has the quasi-polysize model property:
$\varphi \in L$
iff
$\varphi $
is true in all models
$\mathfrak M$
that are
$(i)$
ordered sums of simple models and
$(ii)$
are based on a frame for L of size
$\mathcal {O}(|\varphi |^2)$
(cf. [Reference Litak and Wolter30, Reference Zakharyaschev and Alekseev43]).
In the remainder of Section 4.1, we show how Theorem 4.6 implies the following.
Theorem 4.9. The IEP for any fixed finitely axiomatisable logic
$L \supseteq \mathsf {K4.3}$
is coNP-complete.
Proof. We describe an NP-algorithm deciding the complement of the IEP for L given by its canonical axioms (6). Given
$\varphi _1$
and
$\varphi _2$
, let
$\delta =\textit {sig}(\varphi _1)\cup \textit {sig}(\varphi _2)$
. We guess polynomial-size N. Then, for each
$\ell <N$
, we guess
$z_\ell \in \{a,b,c\}$
, and if
$z_\ell =a$
, we let
$\boldsymbol {n}_1^\ell =\boldsymbol {n}_2^\ell =1$
; otherwise, we guess polynomial-size
$\boldsymbol {n}_i^\ell $
, for
$i=1,2$
; we also guess simple
$\delta $
-models
$\mathfrak N_i^{j,\ell }$
, for
$\ell <N$
,
$i=1,2$
,
$j<\boldsymbol {n}_i^\ell $
, based on L-bounded atomic frames that are either of the form
,
, or
, for some
$k\le \boldsymbol {p}_L(\varphi _1,\varphi _2)$
, or of the form
$m^{<}$
or
, for some
$m\le \boldsymbol {c}_L+1$
, and respective roots
$x_i$
in
$\mathfrak N_i^{0,0}$
. We then let
$\mathfrak {N_i^\ell }=\mathfrak N_i^{0,\ell }\lhd \dots \lhd \mathfrak N_i^{\boldsymbol {n}_i^\ell -1,\ell }$
, for
$\ell <N$
,
$i=1,2$
, and
$\mathfrak N_i=\mathfrak N_i^0\lhd \dots \lhd \mathfrak N_i^{N-1}$
. Checking
$(c)$
and
$(d).2$
in Theorem 4.6 can clearly be done in time polynomial in
$\|{\mathfrak N_i}\|$
(which is polynomial in
$\max (|\varphi _1|,|\varphi _2|)$
). For
$(a)$
, we use the following.
Lemma 4.10. Checking whether
$\mathfrak M_0\lhd \dots \lhd \mathfrak M_{n-1},x\models \varphi $
, for simple
$\textit {sig}(\varphi )$
-models
$\mathfrak M_j$
,
$j<n$
, based on atomic frames with root x in
$\mathfrak M_0$
, can be done in time polynomial in
$|\varphi |$
and
$\|{\mathfrak M_0}\|+\dots +\|{\mathfrak M_{n-1}}\|$
.
Proof. Let
$\mathfrak M = \mathfrak M_0\lhd \dots \lhd \mathfrak M_{n-1}$
. Suppose
$\mathfrak M_j$
is based on the frame
defined in Example 2.2 with points
$a_s$
,
$s < k$
, and
$b_\ell $
,
$\ell < \omega $
. Using the definition of a simple model, it is readily shown by structural induction that any formula
$\psi \in \textit {sub}(\varphi )$
is satisfiable in
$\mathfrak M_j$
iff there is
$\ell <k+\textit {md}(\psi )$
with
$\mathfrak M_j,b_\ell \models \psi $
, where
$\textit {md}(\psi )$
, the modal depth of
$\psi $
, is the maximal number of nested modal operators in
$\psi $
. The required algorithm is now obvious.
Suppose
$L = \mathsf {K4.3} \oplus \{\alpha (\mathfrak G_j,\mathfrak D_j,\bot ) \mid j \in J_L\}$
with finite
$J_L$
and
$\mathfrak G_j=(V_j,S_j)$
. To check condition
$(b)$
in Theorem 4.6, we require the following.
Lemma 4.11. If
$\mathfrak F=\mathfrak F_0\lhd \dots \lhd \mathfrak F_{n-1}$
with atomic frames
$\mathfrak F_\ell $
,
$\ell <n$
, then checking whether
$\mathfrak F\models \alpha (\mathfrak G_j, \mathfrak D_j,\bot )$
, for all
$j \in J_L$
, can be done in time polynomial in
Proof. Let
$\mathfrak F=(W,R,\mathcal {P})$
. Given any
$\alpha (\mathfrak G_j, \mathfrak D_j,\bot )$
, we construct the Kripke frame
$\mathfrak H_j = (W_j,R_j)$
, where
$R_j = {R}\mathop {\restriction }_{W_j}$
and
$W_j\subseteq W$
comprises
-
– the underlying sets of all finite
$\lhd $
-components
$\mathfrak F_\ell $
of
$\mathfrak F$
; -
– the last
$|V_j|+1$
-many points
$b_{|V_j|},\dots ,b_0$
in
, where
$\ast \in \{\bullet ,\circ \}$
and
$b_{|V_j|}$
is ‘painted’ blue (see Example 2.2 for the notation).
Then
$\mathfrak H_j$
is a subframe of
$\mathfrak F$
because all finite subsets of
$\{b_n\mid n<\omega \}$
are internal in
. We show below that there is an injection
$f \colon V_j \to W$
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
in
$\mathfrak F$
iff there is an injection
$h \colon V_j \to W_j$
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
in
$\mathfrak H_j$
and having no blue points in
$h(V_j)$
. Note that the latter is checkable in time polynomial in
$n _{\mathfrak F,j}$
: just enumerate all (at most
$n _{\mathfrak F,j}^{|V_j|}$
-many) injections
$V_j \to W_j$
and verify that at least one of them meets the required conditions.
$(\Leftarrow )$
Suppose
$h \colon V_j \to W_j$
is an injection satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
in
$\mathfrak H_j$
and having no blue points in
$h(V_j)$
. We claim that h satisfies
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
in
$\mathfrak F$
. Indeed,
$\mathbf {(cf_1)}$
and
$\mathbf {(cf_4)}$
hold since
$\mathfrak H_j$
is a subframe of
$\mathfrak F$
, and
$\mathbf {(cf_2)}$
holds because the final cluster of
$\mathfrak H_j$
is the final cluster of
$\mathfrak F$
by definition. To show that h also meets
$\mathbf {(cf_3)}$
, observe that, as
$h(x)$
is not blue for any
$x\in V_j$
, the immediate predecessor cluster of
$C\big (h(x)\big )$
in
$\mathfrak H_j$
is also the immediate predecessor of
$C\big (h(x)\big )$
in
$\mathfrak F$
.
$(\Rightarrow )$
Let
$f \colon V_j \to W$
be an injection satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
. To obtain h, we modify those
$f(x)$
that belong to infinite
$\lhd $
-components
. Suppose the intersection of
$f(V_j)$
with such an
$\mathfrak F_\ell $
is not empty. By (3) in Example 2.2 and
$\mathbf {(cf_4)}$
,
$f(V_j) \cap \{a_0,\dots , a_{k-1}\} = \emptyset $
, and so the intersection of
$f(V_j)$
with
$\mathfrak F_\ell $
is
$\{b_{i_0},\dots ,b_{i_{m_\ell -1}}\}$
, for some
$m_\ell \le |V_j|$
. It is readily seen that by taking
$h(x)=b_z$
if
$f(x)=b_{i_z}$
, for
$z<m_\ell $
, and
$h(y) = f(y)$
, for
$f(y)$
in finite
$\lhd $
-components, we obtain an injection
$h \colon V_j \to W_j$
with
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
and no blue points in
$h(V_j)$
.
If all checks are positive, then, by Lemma 4.4,
$\mathfrak N_1,x_1$
and
$\mathfrak N_2,x_2$
satisfy the conditions of Theorem 3.2, and so
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in L.
4.2 Partitioning models into globally
$\sigma $
-bisimilar intervals
In this section, we start proving Theorem 4.5. In a nutshell, our plan is as follows. Given
$\varphi _1$
and
$\varphi _2$
without an interpolant in
$L \supseteq \mathsf {K4.3}$
, the criterion of Theorem 3.2 supplies models
$\mathfrak M_i$
,
$i=1,2$
, based on finitely
$\mathfrak M_i$
-generated descriptive frames
$\mathfrak {F}_i = (W_{i},R_{i},\mathcal {P}_{i})$
with roots
$x_i \in W_i$
such that:
-
–
$\mathfrak {M}_{1},x_{1} \models \varphi _1$
and
$\mathfrak {M}_{2},x_{2} \models \neg \varphi _2$
; -
– each
$\mathfrak M_i$
,
$i=1,2$
, is based on a frame for L; -
–
$\mathfrak {M}_{1},x_{1} \sim _{\sigma } \mathfrak {M}_{2},x_{2}$
, where
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
.
To prove Theorem 4.5, we need to turn the
$\mathfrak M_i,x_i$
to some
$\mathfrak N_i,x_i$
with the required structure and still satisfying these three conditions. In view of Example 3.6, extracting the roots
$x_i$
and the sets
$\boldsymbol {M}_{i}$
,
$\boldsymbol {S}_{i}$
of maximal points from
$\mathfrak M_i$
(similarly to the proof of Theorem 3.5
$(a)$
) is not enough now, so we need to develop a more involved construction. We proceed in two steps:
-
– First, we analyse the
$\sigma $
-types in the
$\mathfrak M_i$
and partition them into internal closed intervals
$\mathcal I_i=\{I_i^\ell \mid \ell <N\}$
, for the same
$N=\mathcal {O}\big (\max (|\varphi _1|,|\varphi _2|)\big )$
, such that
${\mathfrak M_1}\mathop {\restriction }_{I_1^\ell }$
and
${\mathfrak M_2}\mathop {\restriction }_{I_2^\ell }$
are globally
$\sigma $
-bisimilar, for every
$\ell <N$
. By Lemma 2.9,
$\mathfrak M_i = ({\mathfrak M_i}\mathop {\restriction }_{I_i^0})\lhd \dots \lhd ({\mathfrak M_i}\mathop {\restriction }_{I_i^{N-1}})$
,
$i=1,2$
. -
– Then, in Section 4.3, we complete the proof of Theorem 4.5 by transforming each pair
$\big ({\mathfrak M_1}\mathop {\restriction }_{I_1^\ell },{\mathfrak M_2}\mathop {\restriction }_{I_2^\ell }\big )$
,
$\ell <N$
, into a pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
of models with the required structure.
We begin with a simple observation on definable closed intervals.
Lemma 4.12. Suppose
$\mathfrak {M}$
is a model based on a rooted finitely
$\mathfrak M$
-generated descriptive frame
$\mathfrak F = (W, R,\mathcal {P})$
for
$\mathsf {K4.3}$
. Then every closed interval
$[C,C']$
in
$\mathfrak F$
with a non-limit cluster
$C'$
is definable in
$\mathfrak M$
.
Proof. By Lemma 2.6
$(a)$
–
$(c)$
, the non-limit
$C'$
is defined in
$\mathfrak M$
by some formula
$\gamma $
. Let
$\delta = \bot $
if C is the root cluster, and let
$\delta $
define the immediate predecessor of C in
$\mathfrak M$
otherwise, which exists by Lemma 2.4
$(a)$
and is definable by Lemma 2.6
$(a)$
–
$(c)$
. Then
$[C,C']$
is defined in
$\mathfrak M$
by
$\neg \Diamond ^+ \delta \land \Diamond ^+ \gamma $
.
Next, we look into the structure of
$\sigma $
-types in any model
$\mathfrak M$
based on a rooted finitely
$\mathfrak M$
-generated descriptive frame
$\mathfrak F=(W,R,\mathcal {P})$
for
$\mathsf {K4.3}$
. Given
$x\in W$
and a signature
$\sigma $
, we define the
$\sigma $
-block
${\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(x)$
of x in
$\mathfrak {M}$
by taking
$$\begin{align*}{\boldsymbol{b}}_{\mathfrak{M}}^{\sigma}(x) = \begin{cases} \{ y\in W \mid \Diamond t_{\mathfrak{M}}^{\sigma}(y) \subseteq t_{\mathfrak{M}}^{\sigma}(x), \ \Diamond t_{\mathfrak{M}}^{\sigma}(x) \subseteq t_{\mathfrak{M}}^{\sigma}(y)\}, & \text{if } \Diamond t_{\mathfrak{M}}^{\sigma}(x) \subseteq t_{\mathfrak{M}}^{\sigma}(x);\\ \{x\}, & \text{otherwise}; \end{cases} \end{align*}$$
in the latter case—when x must be an irreflexive point—the
$\sigma $
-block
${\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(x)$
is called degenerate. (It can happen that
$\{x\}$
is a degenerate cluster but
${\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(x)$
is not a degenerate
$\sigma $
-block.) We call a set
${\boldsymbol {b}} \subseteq W$
a
$\sigma $
-block in
$\mathfrak M$
if
${\boldsymbol {b}} = {\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(x)$
, for some x. It is readily seen that the relation
$x\approx y$
iff
${\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(x)={\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(y)$
is an equivalence relation on W, and every
$\sigma $
-block
${\boldsymbol {b}}$
is an interval in
$\mathfrak F$
. (See Example 4.14 below for an illustration.) Observe that
(block) for all
$\sigma $
-blocks
${\boldsymbol {b}}$
in
$\mathfrak M$
and
$y\in W$
, if
$y \notin {\boldsymbol {b}}$
, then
$t_{\mathfrak {M}}^{\sigma }(y) \notin t_{\mathfrak {M}}^{\sigma }({\boldsymbol {b}})$
.
For degenerate
$\sigma $
-blocks this follows from the definability of degenerate clusters (Lemma 2.6), and for other
$\sigma $
-blocks it is straightforward from the definitions.
Lemma 4.13. Suppose
$\mathfrak {M}$
is a model based on a rooted finitely
$\mathfrak M$
-generated descriptive frame
$\mathfrak F = (W, R,\mathcal {P})$
for
$\mathsf {K4.3}$
. For any
$\sigma $
-block
${\boldsymbol {b}}$
in
$\mathfrak {M}$
, there exist clusters
$C_{{\boldsymbol {b}}}^-$
,
$C_{{\boldsymbol {b}}}^+$
in
$\mathfrak F$
such that the following hold:
-
(a)
${\boldsymbol {b}} = \big [C_{{\boldsymbol {b}}}^-,C_{{\boldsymbol {b}}}^+\big ]$
; -
(b) if
$C_{{\boldsymbol {b}}}^+$
is maximal in
$\mathfrak M$
, then it is
$\sigma $
-maximal in
$\mathfrak M$
; -
(c) if
$C_{{\boldsymbol {b}}}^+$
is degenerate, then
${\boldsymbol {b}} = C_{{\boldsymbol {b}}}^+$
; -
(d)
${\boldsymbol {b}}$
is definable in
$\mathfrak M$
iff
$C_{{\boldsymbol {b}}}^+$
is not a limit cluster; -
(e)
$t_{\mathfrak {M}}^{\sigma }({\boldsymbol {b}}) = t_{\mathfrak {M}}^{\sigma }\big (C_{{\boldsymbol {b}}}^+\big )$
.
Proof.
$(a)$
Let
$\mathcal {X}_{{\boldsymbol {b}}}=\{C\mid C \text { be a cluster with } {\boldsymbol {b}}\cap C\ne \emptyset \}$
. As
${\boldsymbol {b}}$
is an interval,
${\boldsymbol {b}}=\bigcup \mathcal {X}_{{\boldsymbol {b}}}$
. By Lemma 2.4, there is a
$<_R$
-largest cluster
$C_{{\boldsymbol {b}}}^+$
in
$\mathcal {X}_{{\boldsymbol {b}}}$
. Also, there is a
$<_R$
-largest cluster D with
$D<_R C$
, for every
$C\in \mathcal {X}_{{\boldsymbol {b}}}$
. Suppose there is no
$<_R$
-smallest cluster in
$\mathcal {X}_{{\boldsymbol {b}}}$
. Then D is a limit cluster and if
$y\in D$
, then
$t_{\mathfrak {M}}^{\sigma }(y) \notin t_{\mathfrak {M}}^{\sigma }({\boldsymbol {b}})$
by (block). So there is a
$\sigma $
-formula
$\mu $
such that
$\mu \in t_{\mathfrak {M}}^{\sigma }(y)$
and
$\Diamond \mu \notin t_{\mathfrak {M}}^{\sigma }(x)$
for any
$x\in {\boldsymbol {b}}$
, and so for any x with
$yR^s x$
. As D is non-degenerate by Lemma 2.6, it follows that D is
$\Diamond \mu $
-maximal in
$\mathfrak M$
, contrary to Lemma 2.6
$(b)$
. Therefore, there is a
$<_R$
-smallest cluster
$C_{{\boldsymbol {b}}}^-$
in
$\mathcal {X}_{{\boldsymbol {b}}}$
, and so
${\boldsymbol {b}} = \big [C_{{\boldsymbol {b}}}^-,C_{{\boldsymbol {b}}}^+\big ]$
.
$(b)$
If
$C_{{\boldsymbol {b}}}^+$
is maximal in
$\mathfrak M$
, then either it is final or has an immediate successor, by Lemma 2.6
$(b)$
. If
$C_{{\boldsymbol {b}}}^+$
is final, then it is
$\top $
-maximal in
$\mathfrak M$
. So suppose that
$C(y)$
is an immediate successor of
$C_{{\boldsymbol {b}}}^+=C(x)$
. If
$C_{{\boldsymbol {b}}}^+$
is not degenerate, then
$\Diamond t_{\mathfrak {M}}^{\sigma }(x) \not \subseteq t_{\mathfrak {M}}^{\sigma }(y)$
follows from
$y\notin {\boldsymbol {b}}$
. So there is a
$\sigma $
-formula
$\mu $
such that
$\mathfrak M,x\models \mu $
and
$\mathfrak M,y\not \models \Diamond \mu $
. If
$\mathfrak M,y\models \mu $
, then
$C_{{\boldsymbol {b}}}^+$
is
$\Diamond \mu $
-maximal in
$\mathfrak M$
. And if
$\mathfrak M,y\not \models \mu $
, then
$C_{{\boldsymbol {b}}}^+$
is
$\mu $
-maximal in
$\mathfrak M$
. If
$C_{{\boldsymbol {b}}}^+$
is degenerate, we cannot have
$\Diamond t_{\mathfrak {M}}^{\sigma }(x) \subseteq t_{\mathfrak {M}}^{\sigma }(x)$
, for otherwise
$t_{\mathfrak {M}}^{\sigma }(x)\subseteq t_{\mathfrak {M}}^{\sigma }(y)$
, contrary to (block). Thus,
$\Diamond t_{\mathfrak {M}}^{\sigma }(x) \not \subseteq t_{\mathfrak {M}}^{\sigma }(x)$
, and so there is
$\sigma $
-formula
$\mu $
such that
$\mathfrak M,x\models \mu $
and
$\mathfrak M,x\not \models \Diamond \mu $
. Therefore,
$C_{{\boldsymbol {b}}}^+$
is
$\mu $
-maximal in
$\mathfrak M$
.
$(c)$
Suppose on the contrary that
$C_{{\boldsymbol {b}}}^+=\{x\}\ne {\boldsymbol {b}}$
. Then
$|{\boldsymbol {b}}|>1$
, and so
$\Diamond t_{\mathfrak {M}}^{\sigma }(x) \subseteq t_{\mathfrak {M}}^{\sigma }(x)$
follows from
${\boldsymbol {b}}={\boldsymbol {b}}^\sigma _{\mathfrak M}(x)$
. So, for every
$\sigma $
-formula
$\mu $
, if
$\mathfrak M,x\models \mu $
then
$\mathfrak M,x\models \Diamond \mu $
. On the other hand,
$C_{{\boldsymbol {b}}}^+$
is maximal in
$\mathfrak M$
by Lemma 2.6
$(a)$
, and so
$\sigma $
-maximal in
$\mathfrak M$
by
$(b)$
, which is a contradiction.
$(d,\Leftarrow )$
This is by
$(a)$
and Lemma 4.12.
$(d,\Rightarrow )$
Suppose that
${\boldsymbol {b}}$
is defined in
$\mathfrak M$
by some
$\psi $
. Then
$C_{{\boldsymbol {b}}}^+$
is
$\psi $
-maximal in
$\mathfrak M$
, and so cannot be a limit cluster by Lemma 2.6
$(b)$
.
$(e)$
If
$C_{{\boldsymbol {b}}}^+$
is degenerate, then this is obvious by
$(c)$
. So suppose
$C_{{\boldsymbol {b}}}^+=C(y)$
is non-degenerate and
$x\in {\boldsymbol {b}}$
. Then
$\Diamond t_{\mathfrak {M}}^{\sigma }(x) \subseteq t_{\mathfrak {M}}^{\sigma }(y)$
, and so
$\Diamond \bigwedge \Gamma \in t_{\mathfrak {M}}^{\sigma }(y)$
for every finite
$\Gamma \subseteq t_{\mathfrak {M}}^{\sigma }(x)$
. By Lemma 2.3, there is z such that
$yRz$
and
$t_{\mathfrak {M}}^{\sigma }(z)=t_{\mathfrak {M}}^{\sigma }(x)$
. By (block), we have
$z\in {\boldsymbol {b}}$
, and so
$z\in C_{{\boldsymbol {b}}}^+$
.
Example 4.14. The model
$\mathfrak M_1$
in Figure 1 from Example 3.6
$(a)$
is partitioned into the following
$\sigma $
-blocks (indicated by the brackets), for three different
$\sigma $
:

To show this for
$\sigma =\emptyset $
, observe that, for every
$n>0$
, we have
$\Diamond ^n \top \in t_{\mathfrak {M}_1}^{\sigma }(b_1^n)$
,
$\Diamond ^{n+1} \top \notin t_{\mathfrak {M}_1}^{\sigma }(b_1^n)$
,
$\neg \Diamond \top \in t_{\mathfrak {M}_1}^{\sigma }(b_1^0)$
, and
$\Diamond ^n \top \in t_{\mathfrak {M}_1}^{\sigma }(a_1^0)$
. The cluster
$C(a_1^0)$
is not maximal in
$\mathfrak M_1$
as any formula
$\alpha $
that is true at
$a_1^0$
or
$a_1^1$
is also true at
$b_1^n$
, for some
$n < \omega $
(which is seen by induction on the structure of
$\alpha $
). The model
$\mathfrak M_1$
in Example 3.6
$(b)$
has only one
$\emptyset $
-block comprising all of its points.
We now return to our models
$\mathfrak M_i$
,
$i=1,2$
, witnessing the lack of interpolants for
$\varphi _1$
and
$\varphi _2$
. By Lemma 4.13
$(a)$
,
$\sigma $
-blocks in each
$\mathfrak {M}_i$
are closed intervals that form a partition of
$W_i$
(with not all of them being necessarily definable in
$\mathfrak M_i$
). We show that there is a
$\prec _{\mathfrak F_i}$
-respecting bijection between the
$\sigma $
-blocks of the two models. Indeed, suppose that
$W_1$
is partitioned as
$\{{\boldsymbol {b}}^j\mid j\in F\}$
into
$\sigma $
-blocks in
$\mathfrak M_1$
, for some countable set F. For each
$j\in F$
, we let
Lemma 4.15. For all
$j\in F$
, the following hold:
-
(a)
$t^\sigma _{\mathfrak M_1}({\boldsymbol {b}}^j)=t^\sigma _{\mathfrak M_2}\big (\boldsymbol {\beta }({\boldsymbol {b}}^j)\big )$
; -
(b)
$\boldsymbol {\beta }({\boldsymbol {b}}^j)$
is a
$\sigma $
-block in
$\mathfrak M_2$
, and
${\boldsymbol {b}}^j$
is degenerate iff
$\boldsymbol {\beta }({\boldsymbol {b}}^j)$
is degenerate; -
(c)
$\{\boldsymbol {\beta }({\boldsymbol {b}}^j)\mid j\in F\}$
is a partition of
$W_2$
; -
(d)
${\boldsymbol {b}}^j\prec _{\mathfrak F_1} {\boldsymbol {b}}^k$
iff
$\boldsymbol {\beta }({\boldsymbol {b}}^j)\prec _{\mathfrak F_2}\boldsymbol {\beta }({\boldsymbol {b}}^k)$
, for
$j,k\in F$
; -
(e)
${\boldsymbol {b}}^j$
is definable in
$\mathfrak M_1$
iff
$\boldsymbol {\beta }({\boldsymbol {b}}^j)$
is definable in
$\mathfrak M_2$
.
Proof.
$(a)$
This follows from
$\mathfrak M_1,x_1\sim _\sigma \mathfrak M_2,x_2$
and Lemma 3.1.
$(b)$
Let
$j\in F$
. As
$\mathfrak M_1,x_1\sim _\sigma \mathfrak M_2,x_2$
,
$\boldsymbol {\beta }({\boldsymbol {b}}^j) \ne \emptyset $
. Take some
$y\in \boldsymbol {\beta }({\boldsymbol {b}}^j)$
. We show that
$\boldsymbol {\beta }({\boldsymbol {b}}^j)={\boldsymbol {b}}^\sigma _{\mathfrak M_2}(y)$
. Indeed, this is straightforward from the definitions if
$\Diamond t^\sigma _{\mathfrak M_2}(y)\subseteq t^\sigma _{\mathfrak M_2}(y)$
. If
$\Diamond t^\sigma _{\mathfrak M_2}(y)\not \subseteq t^\sigma _{\mathfrak M_2}(y)$
, then
${\boldsymbol {b}}^\sigma _{\mathfrak M_2}(y)=\{y\}$
. Take some
$x\in {\boldsymbol {b}}^j$
with
$t^\sigma _{\mathfrak M_1}(x)=t^\sigma _{\mathfrak M_2}(y)$
. Then
$\Diamond t^\sigma _{\mathfrak M_1}(x)\not \subseteq t^\sigma _{\mathfrak M_1}(x)$
, and so
${\boldsymbol {b}}^j=\{x\}$
. Thus,
$\boldsymbol {\beta }({\boldsymbol {b}}^j)=\{z\in W_2\mid t^\sigma _{\mathfrak M_2}(z)=t^\sigma _{\mathfrak M_2}(y)\}$
, and so there is a
$\sigma $
-formula
$\mu $
such that
$\mu \in t^\sigma _{\mathfrak M_2}(z)=t^\sigma _{\mathfrak M_2}(y)$
and
$\Diamond \mu \notin t^\sigma _{\mathfrak M_2}(z)=t^\sigma _{\mathfrak M_2}(y)$
. Suppose there is
$z\in \boldsymbol {\beta }({\boldsymbol {b}}^j)$
,
$z\ne y$
. Then either
$zR_2y$
or
$yR_2z$
, which is a contradiction.
$(c)$
As
$\boldsymbol {\beta }({\boldsymbol {b}}^j)$
and
$\boldsymbol {\beta }({\boldsymbol {b}}^{k})$
are disjoint for
$j\ne k$
by
$(a)$
and (block), the relation ‘
$y\approx y'$
iff there is
$j\in F$
with
$y,y'\in \boldsymbol {\beta }({\boldsymbol {b}}^j)$
’ is an equivalence relation on
$W_2$
.
$(d)$
This follows from
$\mathfrak M_1,x_1\sim _\sigma \mathfrak M_2,x_2$
,
$(a)$
and (block).
$(e)$
This follows from
$(b)$
–
$(d)$
and Lemma 4.13
$(a)$
and
$(d)$
.
So, from now on we assume that we have a strict linear order
$(F,\prec )$
such that each
$W_i$
,
$i=1,2$
, is partitioned as
$\{{\boldsymbol {b}}_i^j\mid j\in F\}$
into
$\sigma $
-blocks in
$\mathfrak M_i$
with
$j\prec k$
iff
${\boldsymbol {b}}_1^j\prec _{\mathfrak F_1}{\boldsymbol {b}}_1^k$
iff
${\boldsymbol {b}}_2^j\prec _{\mathfrak F_2}{\boldsymbol {b}}_2^k$
, for
$j,k\in F$
. (We write
$j\preceq k$
whenever
$j\prec k$
or
$j=k$
.) Observe that, by Lemmas 2.4
$(a)$
and 2.7,
$(F,\succ )$
is isomorphic to a countable ordinal. We say that
$j\in F$
is a
$\succ $
-limit iff it corresponds to a limit ordinal under this isomorphism. Thus, every
$j\in F$
has an immediate
$\prec $
-predecessor, and if j is not a
$\succ $
-limit, then it also has an immediate
$\prec $
-successor. Also, j is a
$\succ $
-limit iff
$C_{{\boldsymbol {b}}_i^j}^+$
is a limit cluster, for
$i=1,2$
.
Next, we analyse some properties of special
$\sigma $
-blocks. Recall that Steps 1 and 2 in the proof of Theorem 3.5
$(a)$
give us the sets
$\boldsymbol {M}_{i}$
containing the
$\{\psi \}$
-maximal points in
$\mathfrak M_i$
that satisfy each formula
$\psi $
in
$\textit {sub}(\varphi _i)$
that is satisfiable in
$\mathfrak M_i$
; the set T of the
$\sigma $
-types of points in
$\{x_1,x_2\}\cup \boldsymbol {M}_{1} \cup \boldsymbol {M}_{2}$
(cf. (7)); and also the sets
$\boldsymbol {S}_{i} \subseteq W_i$
of t-maximal points in
$\mathfrak M_i$
satisfying the
$\sigma $
-types t from T. Points in
$\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}$
are called relevant in
$\mathfrak M_i$
. A cluster or an interval is relevant in
$\mathfrak M_i$
if it contains a relevant point, and irrelevant otherwise. The number of relevant clusters (and of relevant
$\sigma $
-blocks) in
$\mathfrak M_i$
is clearly bounded by the number of relevant points, that is, by
$\boldsymbol {k}(\varphi _1,\varphi _2)$
(defined in (8)). Note that the root and final clusters of
$\mathfrak M_i$
are always relevant (the latter because
$\textit {sub}(\varphi _i)$
is closed under negation, so the final cluster always intersects with
$\boldsymbol {M}_{i}$
).
Example 4.16. For the models
$\mathfrak M_i$
shown in Figure 1 from Example 3.6
$(a)$
and
$\sigma = \{p_1,p_2\}$
, we have
$\boldsymbol {M}_{i} = \{x_i,y_i,b_i^1,b_i^0\}$
,
$\boldsymbol {S}_{1} = \{x_1,y_1,a_1^1,b_1^1,b_1^0\}$
, and
$\boldsymbol {S}_{2} = \{x_2,y_2,a_2^0,b_2^1,b_2^0\}$
, so only the first two and the last two
$\sigma $
-blocks in the
$\mathfrak M_i$
are relevant (cf. Example 4.14 for the
$\sigma $
-blocks).
The next lemma lists a few important properties of relevant
$\sigma $
-blocks.
Lemma 4.17. For all
$j\in F$
and
$i=1,2$
, the following hold:
-
(a)
$\boldsymbol {S}_{i}\cap {\boldsymbol {b}}_i^j=\boldsymbol {S}_{i}\cap C_{{\boldsymbol {b}}_i^j}^+$
; -
(b)
$\big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\cap C_{{\boldsymbol {b}}_i^j}^+=\boldsymbol {S}_{i}\cap C_{{\boldsymbol {b}}_i^j}^+$
; -
(c)
${\boldsymbol {b}}_i^j$
is relevant iff
$\boldsymbol {S}_{i}\cap C_{{\boldsymbol {b}}_i^j}^+\ne \emptyset $
; -
(d) there is a bijection
$f^-\colon \big (\boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}_1^j}^+\big )\to \big (\boldsymbol {S}_{2}\cap C_{{\boldsymbol {b}}_2^j}^+\big )$
with
$t_{\mathfrak M_1}^\sigma (y)=t_{\mathfrak M_2}^\sigma \big (f^-(y)\big )$
, for every
$y\in \boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}_1^j}^+$
; -
(e)
${\boldsymbol {b}}_1^j$
is relevant iff
${\boldsymbol {b}}_2^j$
is relevant.
Proof. Recall the following properties of the
$\boldsymbol {S}_{i}$
defined in Step 2 of the proof of Theorem 3.5:
-
1. if
$x\in \boldsymbol {S}_{i}$
, then x is
$t_{\mathfrak M_i}^\sigma (x)$
-maximal in
$\mathfrak M_i$
; -
2. if
$x\in \{x_i\}\cup \boldsymbol {M}_{i}$
and x is
$t_{\mathfrak M_i}^\sigma (x)$
-maximal in
$\mathfrak M_i$
, then
$x\in \boldsymbol {S}_{i}$
; -
3.
$t_{\mathfrak M_i}^\sigma \big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\subseteq t_{\mathfrak M_i}^\sigma (\boldsymbol {S}_{i})$
; -
4. there is a bijection
$f\colon \boldsymbol {S}_{1}\to \boldsymbol {S}_{2}$
with
$t_{\mathfrak M_1}^\sigma (y)=t_{\mathfrak M_2}^\sigma \big (f(y)\big )$
, for every
$y\in \boldsymbol {S}_{1}$
.
$(a)$
Let
$x\in \boldsymbol {S}_{i}\cap {\boldsymbol {b}}_i^j$
. By Lemma 4.13
$(e)$
, there is
$y\in C_{{\boldsymbol {b}}_i^j}^+$
with
$t_{\mathfrak M_i}^\sigma (y)=t_{\mathfrak M_i}^\sigma (x)$
. Then
$C(x)=C(y)$
follows from 1., and so
$x\in C_{{\boldsymbol {b}}_i^j}^+$
.
$(b)$
Take
$x\in \big (\{x_i\}\cup \boldsymbol {M}_{i}\big )\cap C_{{\boldsymbol {b}}_i^j}^+$
. By 3., there is
$y\in \boldsymbol {S}_{i}$
with
$t_{\mathfrak M_i}^\sigma (y)=t_{\mathfrak M_i}^\sigma (x)$
. By 1., y is
$t_{\mathfrak M_i}^\sigma (y)$
-maximal in
$\mathfrak M_i$
. Thus, by (block) and Lemma 4.13
$(e)$
,
$y\in C_{{\boldsymbol {b}}_i^j}^+$
. It follows that x is
$t_{\mathfrak M_i}^\sigma (x)$
-maximal in
$\mathfrak M_i$
, and so
$x\in \boldsymbol {S}_{i}$
by 2.
$(c)$
We show that
$t_{\mathfrak M_i}^\sigma \big (\big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\cap {\boldsymbol {b}}_i^j\big )\subseteq t_{\mathfrak M_i}^\sigma \big (\big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\cap C_{{\boldsymbol {b}}_i^j}^+\big )$
. Then
$(c)$
follows from
$(b)$
. To this end, take
$x\in \big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\cap {\boldsymbol {b}}_i^j$
. By 3., there is
$y\in \boldsymbol {S}_{i}$
with
$t_{\mathfrak M_i}^\sigma (y)=t_{\mathfrak M_i}^\sigma (x)$
. By 1., y is
$t_{\mathfrak M_i}^\sigma (y)$
-maximal in
$\mathfrak M_i$
. Thus, by (block) and Lemma 4.13
$(e)$
,
$y\in C_{{\boldsymbol {b}}_i^j}^+$
.
$(d)$
Let
$f^-={f}\mathop {\restriction }_{\boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}_1^j}^+}$
for the bijection f provided by 4. Then, for every
$x\in \boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}_1^j}^+$
,
$f^-(x)=f(x)\in \boldsymbol {S}_{2}$
with
$t_{\mathfrak M_2}^\sigma (f(x))=t_{\mathfrak M_1}^\sigma (x)$
. By Lemma 4.15
$(a)$
,
$t_{\mathfrak M_2}^\sigma (f(x))\in t_{\mathfrak M_2}^\sigma ({\boldsymbol {b}}_2^j)$
, so
$f(x)\in {\boldsymbol {b}}_2^j$
follows by (block). Thus,
$f(x)\in C_{{\boldsymbol {b}}_2^j}^+$
by
$(a)$
.
$(e)$
follows from
$(c)$
and
$(d)$
.
We are now in a position to partition each of the
$\mathfrak M_i$
into the same polynomial number N of closed intervals
$\mathcal I_i=\{I_i^\ell \in \mathcal {P}_i\mid \ell <N\}$
such that
${\mathfrak M_1}\mathop {\restriction }_{I_1^\ell }$
and
${\mathfrak M_2}\mathop {\restriction }_{I_2^\ell }$
are globally
$\sigma $
-bisimilar, for every
$\ell <N$
, even if there are infinitely many
$\sigma $
-blocks in each
$\mathfrak M_i$
and not all of them are definable in
$\mathfrak M_i$
.
Definition 4.18. We define the partitions
$\mathcal I_i$
of
$\mathfrak M_i$
,
$i=1,2$
, in three steps. In each step, we add interval-pairs
$(I_1,I_2)$
to
$\mathcal I_1\times \mathcal I_2$
in such a way that:
-
(a)
$I_i$
is a closed interval whose final cluster is a non-limit cluster, for
$i=1,2$
; -
(b) there are
$j,j'\in F$
such that
$I_i=\bigcup _{j\preceq k\preceq j'}{\boldsymbol {b}}_i^k$
, for
$i=1,2$
.
It follows then from
$(a)$
and Lemma 4.12 that all intervals in
$\mathcal I_i$
are definable in
$\mathfrak M_i$
. Also, it follows from
$(b)$
and Lemma 4.15
$(a)$
that
$$ \begin{align} \big\{(y_1,y_2)\in I_1^\ell\times I_2^\ell\mid t^\sigma_{\mathfrak M_1} & (y_1)=t^\sigma_{\mathfrak M_2}(y_2)\big\} \text{ is a global } \sigma\text{-bisimulation}\\& \text{between } {\mathfrak M_1}\mathop{\restriction}\nolimits _{I_1^\ell} \text{ and } {\mathfrak M_2}\mathop{\restriction}\nolimits _{I_2^\ell}, \text{ for every } \ell<N.\nonumber \end{align} $$
The three steps are as follows:
-
(s1) First, suppose
${\boldsymbol {b}}_1^j$
,
$j\in F$
, is a relevant
$\sigma $
-block that is definable in
$\mathfrak M_1$
. By Lemmas 4.15
$(e)$
and 4.17
$(e)$
,
${\boldsymbol {b}}_2^j$
is also a relevant
$\sigma $
-block definable in
$\mathfrak M_2$
. We put into
$\mathcal {I}_i$
all those relevant
$\sigma $
-blocks
${\boldsymbol {b}}_i^j$
that are definable in
$\mathfrak M_i$
, for
$i=1,2$
. Then
$(b)$
clearly holds, and
$(a)$
holds by Lemma 4.13
$(a)$
and
$(d)$
. -
(s2) Next, suppose
${\boldsymbol {b}}_1^j$
,
$j\in F$
, is a relevant
$\sigma $
-block that is not definable in
$\mathfrak M_1$
. By Lemmas 4.15
$(e)$
and 4.17
$(e)$
,
${\boldsymbol {b}}_2^j$
is also a relevant
$\sigma $
-block that is not definable in
$\mathfrak M_2$
. By Lemma 4.13
$(d)$
, each
$C_{{\boldsymbol {b}}_i^j}^+$
is a limit cluster in
$\mathfrak F_i$
, and so j is a
$\succ $
-limit. We pick some
$\ell \succ j$
such that the
$\sigma $
-blocks
${\boldsymbol {b}}_i^k$
, for
$j\prec k\preceq \ell $
, are all irrelevant, for
$i=1,2$
. Such an
$\ell $
must exist as j is a
$\succ $
-limit and the number of relevant points is finite, but this
$\ell $
is not unique. Let
$F^-=\{k\in F\mid j\preceq k\preceq \ell \}$
and
$\succ ^-={\succ }\mathop {\restriction }_{F^-}$
. By Lemmas 2.4
$(a)$
and 2.7, there is an isomorphism f from some countable ordinal
$\gamma $
to
$(F^-,\succ ^-)$
. As j is a
$\succ $
-limit,
$\gamma \ge \omega $
. Take
$f(n)$
,
$n<\omega $
. There are two cases:-
1. There exists m,
$0<m<\omega $
, such that
${\boldsymbol {b}}_1^{f(n)}$
is a degenerate
$\sigma $
-block for every n with
$m\le n<\omega $
. Then, by Lemma 4.15
$(b)$
,
${\boldsymbol {b}}_2^{f(n)}$
is a degenerate
$\sigma $
-block, for every n with
$m\le n<\omega $
. We set
$j'=f(m)$
. -
2. For every
$n<\omega $
, there is
$m_n$
,
$n\le m_n<\omega $
, such that
${\boldsymbol {b}}_1^{f(m_n)}$
is a non-degenerate
$\sigma $
-block. Then, by Lemma 4.15
$(b)$
,
${\boldsymbol {b}}_2^{f(m_n)}$
is a non-degenerate
$\sigma $
-block as well. Note that if
$n\ge 1$
, then
$f(m_n)$
is not a
$\succ $
-limit. Thus,
$C_{{\boldsymbol {b}}_i^{f(m_n)}}^+$
is not a limit cluster, and so it is definable in
$\mathfrak M_i$
by Lemma 2.6. We set
$j'=f(m_1)$
.
In both cases, we put the intervals
$\bigcup _{j\preceq k\preceq j'}{\boldsymbol {b}}_i^{k}$
into
$\mathcal I_i$
,
$i=1,2$
, and say that they extend the relevant non-definable
$\sigma $
-blocks
${\boldsymbol {b}}_i^j$
. Then
$(a)$
and
$(b)$
hold. -
-
(s3) Finally, suppose that, for
$i=1,2$
, the intervals
$I_i=\bigcup _{n_1\preceq k\preceq n_2}{\boldsymbol {b}}_i^k$
and
$J_i=\bigcup _{j_1\preceq k\preceq j_2}{\boldsymbol {b}}_i^k$
are such that there is k with
$n_2\prec k\prec j_1$
,
$I_i,J_i\in \mathcal I_i$
, and there is no interval in
$\mathcal I_i$
intersecting the ‘gap’ between
$I_i$
and
$J_i$
(that is, any
${\boldsymbol {b}}_i^k$
with
$n_2\prec k\prec j_1$
). By
$I_i\in \mathcal I_i$
and
$(i)$
,
$n_2$
is not a
$\succ $
-limit. Let
$n_2^+$
be the immediate
$\prec $
-successor of
$n_2$
and
$j_1^-$
the immediate
$\prec $
-predecessor of
$j_1$
. Then we put the (irrelevant) interval
$\bigcup _{n_2^+\preceq k\preceq j_1^-}{\boldsymbol {b}}_i^{k}$
into
$\mathcal I_i$
, for
$i=1,2$
. Then
$(b)$
clearly holds, and
$(a)$
holds as
$j_1^-$
is not a
$\succ $
-limit. By doing this for all the gaps, we end up with the required partition
$\mathcal I_i$
of
$\mathfrak M_i$
.
The number of intervals added in steps
$\mathbf {(s_{1})}$
and
${\mathbf {(s_{2})}}$
together cannot exceed the number of relevant
$\sigma $
-blocks, and so it is bounded by
$\boldsymbol {k}(\varphi _1,\varphi _2)$
. As the
$\prec _{\mathfrak F_i}$
-smallest and
$\prec _{\mathfrak F_i}$
-largest
$\sigma $
-blocks are relevant, the number of intervals added in step
${\mathbf {(s_{3})}}$
is bounded by
$\boldsymbol {k}(\varphi _1,\varphi _2)-1$
, so altogether the (same) number N of intervals in each
$\mathcal I_i$
does not exceed
$2\boldsymbol {k}(\varphi _1,\varphi _2)$
.
The following example illustrates Definition 4.18.
Example 4.19. For models
$\mathfrak M_i$
,
$i=1,2$
, from Example 3.6
$(a)$
and
$\sigma $
-blocks from Example 4.14 for
$\sigma = \{p_1,p_2\}$
, we can pick the intervals
$I_i^j$
,
$j \le 4$
, shown below, where
$I_i^2$
are irrelevant and all other intervals are relevant (cf. Example 4.16). The choice of the infinite intervals
$I_i^1$
extending the non-definable
$\sigma $
-blocks till
$b_i^4$
is arbitrary. We could make them shorter or, on the contrary, extend until
$b_i^2$
, in which case there would be no gap between these intervals (extending relevant non-definable
$\sigma $
-blocks) and the next relevant interval.

4.3 Simplifying interval-based models
Consider again our
$\sigma $
-bisimilar
$\delta $
-models
$\mathfrak M_i$
,
$i=1,2$
, that are based on finitely
$\mathfrak M_i$
-generated descriptive frames
$\mathfrak {F}_i = (W_{i},R_{i},\mathcal {P}_{i})$
for L with roots
$x_i \in W_i$
and witness the lack of an interpolant for
$\varphi _1$
and
$\varphi _2$
, where
$\delta = \textit {sig}(\varphi _1) \cup \textit {sig}(\varphi _2)$
and
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
. In Definition 4.18, we determined
$N<2\boldsymbol {k}(\varphi _1,\varphi _2)$
, for the polynomial number
$\boldsymbol {k}(\varphi _1,\varphi _2)$
from (8), and constructed the partitions
$\mathcal I_i=\{I_i^\ell \in \mathcal {P}_i\mid \ell <N\}$
of
$\mathfrak M_i$
with
$I_i^0\prec _{\mathfrak F_i}\cdots \prec _{\mathfrak F_i}I_i^{N-1}$
satisfying (12). We now use these partitions to prove Theorem 4.5. First, in Lemma 4.21, we transform each pair
$({\mathfrak M_1}\mathop {\restriction }_{I_1^\ell },{\mathfrak M_2}\mathop {\restriction }_{I_2^\ell })$
,
$\ell <N$
, into a pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
of models meeting the list of requirements in Definition 4.20. Then, in Lemma 4.22, we show that these requirements ensure that
$\mathfrak N_i=\mathfrak N_i^0\lhd \dots \lhd \mathfrak N_i^{N-1}$
,
$i=1,2$
, satisfy all conditions in Theorem 4.5.
For all
$i=1,2$
and
$\ell <N$
, the frame
$\mathfrak H^\ell _i=(H^\ell _i,R_i^\ell ,\mathcal {P}^\ell _i)$
underlying
$\mathfrak N_i^\ell $
is such that
$H^\ell _i\subseteq I_i^\ell $
is definable in
$\mathfrak M_i$
and
$R_i^\ell ={R_i}\mathop {\restriction }_{H^\ell _i}$
, but
$\mathfrak H_i^\ell $
is not necessarily a subframe of
${\mathfrak F_i}\mathop {\restriction }_{I_i^\ell }$
. However, we require each
$\mathfrak H^\ell _i$
to meet some conditions making sure that the canonical formulas refuted in
$\mathfrak H_i=\mathfrak H_i^0\lhd \dots \lhd \mathfrak H_i^{N-1}$
are also refuted in
$\mathfrak F_i$
(and so
$L\subseteq \mathsf {Log}(\mathfrak F_i) \subseteq \mathsf {Log}(\mathfrak H_i)$
). Another feature of the construction is that the atomic type of some points in
$\mathfrak N_i$
could be different from their atomic type in
$\mathfrak M_i$
. We prove
$\mathfrak N_1,x_1\models \varphi _1$
and
$\mathfrak N_2,x_2\models \neg \varphi _2$
by ensuring that no new (compared to
$\mathfrak M_i$
) atomic types are introduced in
$\mathfrak N_i$
, and the distribution of old atomic types in
$\mathfrak N_i$
properly matches their distribution in
$\mathfrak M_i$
. We achieve this by introducing functions
$\mathsf {h}_i^\ell $
that assign to each point x in
$\mathfrak N_i^\ell $
a unique ‘parent’ point in
${\mathfrak M_i}\mathop {\restriction }_{I_i^\ell }$
whose
$\mathfrak M_i$
-behaviour x is intended to mimic in
$\mathfrak N_i$
.
Definition 4.20. Suppose
$I\in \mathcal {P}_i$
,
$i=1,2$
, is an interval in
$\mathfrak F_i$
. We say that a model
$\mathfrak N$
based on a frame
$\mathfrak H=(H,S,\mathcal {P}')$
is
$(I,i)$
-nice if the following hold:
$$ \begin{align} &\qquad \text{then } C \text{ is the root cluster incc} {\mathfrak F_i}\mathop{\restriction}\nolimits _{I}{;} \nonumber\\& \text{for every } x\in H, \text{ if } \{x\} \text{ is a degenerate non-root cluster in } \mathfrak H\end{align} $$
$$ \begin{align} &\qquad\text{and } C\subseteq I \text{ is the immediate predecessor of } \{x\} \text{ in } \mathfrak F_i, \text{ then} \nonumber\\&\qquad C\cap H \text{ is the immediate predecessor of } \{x\} \text{ in } \mathfrak H; \nonumber\\& \text{for every } x\in H, \text{ if } \{x\}\in\mathcal{P}', \text{ then } \{x\}\in\mathcal{P}_i; \end{align} $$
there is a function
$\mathsf {h}\colon H\to H$
such that:
Lemma 4.21. For all
$i=1,2$
and
$\ell <N$
, there exist models
$\mathfrak N_i^\ell $
based on frames
$\mathfrak H_i^\ell =(H_i^\ell ,S_i^\ell ,\mathcal {P}_i^\ell )$
, and numbers
$\boldsymbol {n}_i^\ell>0$
with
$\sum _{\ell <N}\boldsymbol {n}_i^\ell \le 3\boldsymbol {k}(\varphi _1,\varphi _2)-1$
such that the following hold:
-
(a)
$\mathfrak N_i^\ell $
is
$(I_i^\ell ,i)$
-nice; -
(b)
$\mathfrak N_i^\ell $
is the ordered sum of
$\boldsymbol {n}_i^\ell $
-many simple
$\delta $
-models based on atomic frames; -
(c) the pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is
$\sigma $
-matching.
Proof. We consider three Cases I–III, depending on the step the pair
$(I_1^\ell ,I_2^\ell )$
is added to
$\mathcal I_1\times \mathcal I_2$
in Definition 4.18.
Case I:
$(I_1^\ell ,I_2^\ell )$
is added in step
${\mathbf {(s{_3})}}$
, so
$I_i^\ell $
are irrelevant intervals. We let
$\boldsymbol {n}_1^\ell =\boldsymbol {n}_2^\ell =1$
and define
$\mathfrak N_1^\ell $
and
$\mathfrak N_2^\ell $
as follows. Let
$Z_i^\ell =\{z_i^j\mid j< m_i\}$
, for
$i=1,2$
, be the tail of
${\mathfrak F_i}\mathop {\restriction }_{I_i^\ell }$
, for some
$m_i\leq \omega $
, with
$z_{i}^jR_i^s z_i^{j-1}$
,
$0<j<m_i$
. By (12),
$\big \{(y_1,y_2)\in I_1^\ell \times I_2^\ell \mid t^\sigma _{\mathfrak M_1}(y_1)=t^\sigma _{\mathfrak M_2}(y_2)\big \}$
is a global
$\sigma $
-bisimulation between
${\mathfrak M_1}\mathop {\restriction }_{I_1^\ell }$
and
${\mathfrak M_2}\mathop {\restriction }_{I_2^\ell }$
. It is straightforward to see that because of this we must have
$|Z_1^\ell |=|Z_2^\ell |=m$
, for some
$m\leq \omega $
, and
$Z_1^\ell =I_1^\ell $
iff
$Z_2^\ell =I_2^\ell $
. Also, if
$Z_i^\ell \ne I_i^\ell $
, then there exist
$w_i^\ell $
in the head of
$Z_i^\ell $
with
$t^\sigma _{\mathfrak M_1}(w_1^\ell )=t^\sigma _{\mathfrak M_2}(w_2^\ell )$
. For
$i=1,2$
, let
$$\begin{align*}H_i^\ell=\left\{ \begin{array}{@{}ll@{}} Z_i^\ell, & \text{ if } Z_i^\ell=I_i^\ell,\\[3pt] \{w_i^\ell\}\cup Z_i^\ell, & \text{ otherwise}, \end{array} \right. \end{align*}$$
$S_i^\ell ={R_i}\mathop {\restriction }_{H_i^\ell }$
, and let
$\mathcal {P}_i^\ell $
consist of all finite subsets of
$Z_i^\ell $
and their complements in
$H_i^\ell $
. Then
$f\colon H_1^\ell \to H_2^\ell $
defined by
$f(z_1^j)=z_2^j$
,
$j<m$
, and
$f(w_1^\ell )=w_2^\ell $
is an isomorphism between the resulting frames
$\mathfrak H_1^\ell $
and
$\mathfrak H_2^\ell $
, which are isomorphic to
-
(i)
$m^<$
, when
$Z_i^\ell =I_i^\ell $
; -
(ii)
, when
$Z_i^\ell \ne I_i^\ell $
and
$m<\omega $
; -
(iii)
, when
$Z_i^\ell $
is infinite (as
$w_i^\ell R_i w_i^\ell $
by (5)).
This gives (13)–(18) for
$\mathfrak H=\mathfrak H_i^\ell $
and
$I=I_i^\ell $
(we have (18) because of (3) and Lemma 2.5). For
$p\in \delta $
, let
$\mathfrak {w}_i^\ell (p)=\mathfrak {v}_i(p)\cap H_i^\ell $
in cases
$(i)$
and
$(ii)$
, and
$$\begin{align*}\mathfrak w_i^\ell(p)=\left\{ \begin{array}{@{}ll@{}} H_i^\ell, & \text{ if } w_i^\ell\in\mathfrak v_i(p),\\[3pt] \emptyset, & \text{ otherwise} \end{array} \right. \end{align*}$$
in case
$(iii)$
. In all cases,
$\mathfrak w_i^\ell (p)\in \mathcal {P}_i^\ell $
and
$(b)$
holds for
$\mathfrak N_i^\ell = (\mathfrak H_i^\ell ,\mathfrak w_i^\ell )$
. Also, the pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is of type
$(a)$
in Definition 4.3, and so
$(c)$
of the lemma holds. Finally, for
$x\in H_i^\ell $
, we let
$\mathsf {h}_i^\ell (x)=x$
in cases
$(i)$
and
$(ii)$
, and
$\mathsf {h}_i^\ell (x)=w_i^\ell $
in case
$(iii)$
. It is straightforward to check that (20) and (21) hold for
$\mathsf {h}=\mathsf {h}_i^\ell $
. Note that (19) and (22) hold vacuously, as
$H_i^\ell \subseteq I_i^\ell $
and
$\big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big ) \cap I_i^\ell =\emptyset $
,
$i=1,2$
. Thus, we have
$(a)$
of the lemma.
Case II:
$(I_1^\ell ,I_2^\ell )$
is added in step
${\mathbf {(s{_1})}}$
. For
$i=1,2$
, let
${\boldsymbol {b}}_i$
be the relevant
$\sigma $
-blocks such that
$t^\sigma _{\mathfrak M_1}({\boldsymbol {b}}_1)=t^\sigma _{\mathfrak M_2}({\boldsymbol {b}}_2)$
and
$I_i^\ell ={\boldsymbol {b}}_i$
is definable in
$\mathfrak M_i$
. For
$\ell <N$
, let
$\boldsymbol {r}_i^\ell $
denote the number of relevant clusters in
$I_i^\ell $
, and let
$C_i^{\ell ,j}$
,
$j < \boldsymbol {r}_i^\ell $
, be the sequence (ordered by
$<_{R_i}$
) of all relevant clusters in
${\boldsymbol {b}}_i$
(that intersect with
$\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}$
). Then
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
is the final cluster
$C_{{\boldsymbol {b}}_i}^+$
of
${\boldsymbol {b}}_i$
.
Case II.1: Observe that, by Lemmas 4.13
$(c)$
and 4.15
$(b)$
,
$C_1^{\ell ,\boldsymbol {r}_1^\ell -1}=C_{{\boldsymbol {b}}_1}^+$
is degenerate iff
$C_2^{\ell ,\boldsymbol {r}_2^\ell -1}=C_{{\boldsymbol {b}}_2}^+$
is degenerate iff both
${\boldsymbol {b}}_1=C_{{\boldsymbol {b}}_1}^+$
and
${\boldsymbol {b}}_2=C_{{\boldsymbol {b}}_2}^+$
are degenerate
$\sigma $
-blocks, and so
$\boldsymbol {r}_i^\ell =1$
. So, in this case, we just set
$\boldsymbol {n}_i^\ell =1$
,
$\mathfrak H_i^\ell ={\mathfrak F_i}\mathop {\restriction }_{{\boldsymbol {b}}_i}$
,
$\mathfrak N_i^\ell ={\mathfrak M_i}\mathop {\restriction }_{{\boldsymbol {b}}_i}$
, and
$\mathsf {h}_i(z_i)=z_i$
for the only point
$z_i$
in
${\boldsymbol {b}}_i$
,
$i=1,2$
. It is straightforward to check that
$(a)$
–
$(c)$
of the lemma hold. In particular,
$(c)$
holds because the pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is of type
$(a)$
in Definition 4.3.
Case II.2: So, let
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}=C_{{\boldsymbol {b}}_i}^+$
be non-degenerate, for
$i=1,2$
. We may assume that, for any
$j< \boldsymbol {r}_i^\ell -1$
,
$C_i^{\ell ,j}$
is a non-limit cluster. (For
$j>0$
, this follows from Lemma 2.6, as
$C_i^{\ell ,j}\cap \boldsymbol {M}_{i}\ne \emptyset $
by Lemma 4.17
$(a)$
. However, if
$C_i^0$
is the root cluster in
$\mathfrak F_i$
, it can happen that
$(\{x_i\}\cup \boldsymbol {M}_{i})\cap C_i^0=\{x_i\}$
,
$x_i\notin \boldsymbol {M}_{i}$
and
$C_i^0$
is a limit cluster. We may exclude this case by Lemma 3.3.) Also, as
${\boldsymbol {b}}_i$
is definable in
$\mathfrak M_i$
,
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}=C_{{\boldsymbol {b}}_i}^+$
is a non-limit cluster by Lemma 4.13
$(d)$
. Below, we define sets
$A_i^\ell \subseteq C_{{\boldsymbol {b}}_i}^+$
, intervals
$J_i^{\ell ,j}\subseteq I_i^\ell $
, and models
$\mathfrak N_i^{\ell ,j}=(\mathfrak H_i^{\ell ,j},\mathfrak w_i^{\ell ,j})$
with
$\mathfrak H_i^{\ell ,j}=(H_i^{\ell ,j},{R_i}\mathop {\restriction }_{H_i^{\ell ,j}},\mathcal {P}_i^{\ell ,j})$
, for
$i=1,2$
and
$j<\boldsymbol {r}_i^\ell $
, such that the following hold:
$$ \begin{align} &\qquad \text{based on atomic frames, for } j<\boldsymbol{r}_i^\ell{;}\nonumber\\& \{J_i^{\ell,j}\mid j<\boldsymbol{r}_i^\ell\} \text{ is a partition of } I_i^\ell \text{ with } J_i^{\ell,0}\prec_{\mathfrak F_i}\cdots\prec_{\mathfrak F_i}J_i^{\ell,\boldsymbol{r}_i^\ell-1}; \end{align} $$
$$ \begin{align} &\displaystyle t_{\mathfrak M_1}^\sigma\Bigl(\bigcup_{j<\boldsymbol{r}_1^\ell-1}H_1^{\ell,j}\Bigr)\subseteq t_{\mathfrak M_2}^\sigma\big(A_2^\ell\big) \text{ and } \displaystyle t_{\mathfrak M_2}^\sigma\Big(\bigcup_{j<\boldsymbol{r}_2^\ell-1}H_2^{\ell,j}\Big)\subseteq t_{\mathfrak M_1}^\sigma\big(A_1^\ell\big).\end{align} $$
Then we show that (23)–(27) imply
$(a)$
–
$(c)$
for
$\mathfrak N_i^\ell =\mathfrak N_i^{\ell ,0}\lhd \dots \lhd \mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and some
$\boldsymbol {n}_i^\ell \le 2\boldsymbol {r}_i^\ell $
. In particular,
$(c)$
because
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is of type
$(b)$
in Definition 4.3.
To this end, we cover first the cases when
$j<\boldsymbol {r}_i^\ell -1$
and then, separately, the case
$j=\boldsymbol {r}_i^\ell -1$
. So suppose first that
$j<\boldsymbol {r}_i^\ell -1$
, and let
$J_i^{\ell ,j}=[D_i^{\ell ,j},C_i^{\ell ,j}]$
, where
$D_i^{\ell ,0}$
is the root cluster in
${\mathfrak F_i}\mathop {\restriction }_{I_i^\ell }$
and
$D_i^{\ell ,j}$
is the immediate successor of the non-limit cluster
$C_i^{\ell ,j-1}$
,
$0<j<\boldsymbol {r}_i^\ell -1$
. Observe that
$\big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\cap J_i^{\ell ,j}\subseteq C_i^{\ell ,j}$
. We consider four subcases
$(i)$
–
$(iv)$
, depending on the tail
$Z_i^{\ell ,j}$
of
${\mathfrak F_i}\mathop {\restriction }_{J_i^{\ell ,j}}$
.
-
(i)
$Z_i^{\ell ,j}=\emptyset $
, so
$C_i^{\ell ,j}$
is non-degenerate. Let
$H_i^{\ell ,j}=\big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\cap C_i^{\ell ,j}$
and
$\mathcal {P}_i^{\ell ,j}=2^{H_i^{\ell ,j}}$
. Then
$\mathfrak H_i^j$
is isomorphic to
, for
$k=|H_i^{\ell ,j}|$
. We set
$\mathsf {h}_i^{\ell ,j}(x)=x$
, for
$x\in H_i^{\ell ,j}$
, and
$\mathfrak {w}_i^{\ell ,j}(p)=\mathfrak {v}_i(p)\cap H_i^{\ell ,j}$
, for
$p\in \delta $
. -
(ii) If
$0<|Z_i^{\ell ,j}|=m<\omega $
and
$Z_i^{\ell ,j}=J_i^{\ell ,j}$
, then by taking
$H_i^{\ell ,j}=J_i^{\ell ,j}$
and
$\mathcal {P}_i^{\ell ,j}=2^{H_i^{\ell ,j}}$
we obtain
$\mathfrak H_i^{\ell ,j}$
isomorphic to
$m^<$
. We set
$\mathsf {h}_i^{\ell ,j}(x)=x$
, for
$x\in H_i^{\ell ,j}$
, and
$\mathfrak {w}_i^{\ell ,j}(p)=\mathfrak {v}_i(p)\cap H_i^{\ell ,j}$
, for
$p\in \delta $
. -
(iii) If
$0<|Z_i^{\ell ,j}|=m<\omega $
and
$Z_i^{\ell ,j}\ne J_i^{\ell ,j}$
, then setting
$H_i^{\ell ,j}=\{w_i^{\ell ,j}\}\cup Z_i^{\ell ,j}$
, for any
$w_i^{\ell ,j}$
in the head of
$Z_i^{\ell ,j}$
, and
$\mathcal {P}_i^{\ell ,j}=2^{H_i^{\ell ,j}}$
gives
$\mathfrak H_i^{\ell ,j}$
isomorphic to
. Let
$\mathsf {h}_i^{\ell ,j}(x)=x$
, for
$x\in H_i^{\ell ,j}$
, and
$\mathfrak {w}_i^{\ell ,j}(p)=\mathfrak {v}_i(p)\cap H_i^{\ell ,j}$
, for
$p\in \delta $
. -
(iv) If
$Z_i^{\ell ,j}$
is infinite, then let
$H_i^{\ell ,j}=\{w_i^{\ell ,j}\}\cup Z_i^{\ell ,j}$
, for any
$w_i^{\ell ,j}$
in the head of
$Z_i^{\ell ,j}$
, and
$\mathcal {P}_i^{\ell ,j}$
consist of all finite subsets of
$H_i^{\ell ,j}$
and their complements in
$H_i^{\ell ,j}$
. By (5), the resulting
$\mathfrak H_i^{\ell ,j}$
is isomorphic to
. In this case,
$C_i^{\ell ,j}=\{y_i^{\ell ,j}\}$
is a degenerate cluster for some
$y_i^{\ell ,j}\in \{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}$
. We set
$\mathsf {h}_i^{\ell ,j}(y_i^{\ell ,j})=y_i^{\ell ,j}$
and
$\mathsf {h}_i^{\ell ,j}(x)=w_i^{\ell ,j}$
for all
$x\in H_i^{\ell ,j}\setminus \{y_i^{\ell ,j}\}$
. For
$p\in \delta $
, let
$$\begin{align*}\mathfrak w_i^{\ell,j}(p)=\left\{ \begin{array}{@{}ll@{}} \big(H_i^{\ell,j}\setminus\{y_i^{\ell,j}\}\big)\cup\big(\mathfrak v_i(p)\cap\{y_i^{\ell,j}\}\big), & \text{ if } w_i^{\ell,j}\in\mathfrak v_i(p),\\ \mathfrak v_i(p)\cap\{y_i^{\ell,j}\}, & \text{ otherwise}. \end{array} \right. \end{align*}$$
Then it is not hard to check that, in all
$(i)$
–
$(iv)$
, we have
$\mathfrak w_i^{\ell ,j}(p)\in \mathcal {P}_i^{\ell ,j}$
, (24) for
$\mathfrak N_i^{\ell ,j}= (\mathfrak H_i^{\ell ,j},\mathfrak w_i^{\ell ,j})$
, and (13)–(22) hold for
$\mathfrak H=\mathfrak H_i^{\ell ,j}$
,
$\mathfrak N=\mathfrak N_i^{\ell ,j}$
,
$\mathsf {h}=\mathsf {h}_i^{\ell ,j}$
, and
$I=J_i^{\ell ,j}$
. In particular, in
$(i)$
–
$(iii)$
, we have (18) by Lemma 2.5. In
$(iv)$
, we also need (3) to obtain (18), and the fact that
$\boldsymbol {M}_{i}\cap H_i^{\ell ,j}=\{y_i^{\ell ,j}\}$
to obtain (22). Therefore, we have (23) for
$j<\boldsymbol {r}_i^\ell -1$
.
Now, consider
$j=\boldsymbol {r}_i^\ell -1$
. First, we let
$J_i^{\ell ,\boldsymbol {r}_i^\ell -1}=[D_i^{\ell ,\boldsymbol {r}_i^\ell -1},C_i^{\ell ,\boldsymbol {r}_i^\ell -1}]$
, where
$D_i^{\ell ,\boldsymbol {r}_i^\ell -1}=C_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
if
$\boldsymbol {r}_i^\ell =1$
and
$D_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
is the immediate successor of the non-limit cluster
$C_i^{\ell ,\boldsymbol {r}_i^\ell -2}$
otherwise. Then we have (25). We let
$Y_i^\ell =\bigcup _{j<\boldsymbol {r}_1^{\ell }-1}Y_i^{\ell ,j}$
, where
$Y_i^{\ell ,j}=H_i^{\ell ,j}$
in cases
$(i)$
–
$(iii)$
above, and
$Y_i^{\ell ,j}=\{w_i^{\ell ,j},z_i^{\ell ,j}\}$
in case
$(iv)$
. So
$Y_i^\ell $
is finite. Set
$\Theta =\big \{ t_{\mathfrak {M}_{1}}^{\sigma }(x) \mid x\in Y_1^\ell \big \} \cup \big \{ t_{\mathfrak {M}_{2}}^{\sigma }(x) \mid x\in Y_2^\ell \big \}$
. Let
$A_i^\ell $
be the smallest set such that
$\big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )\cap C_i^{\ell ,\boldsymbol {r}_i^\ell -1}\subseteq A_i^\ell \subseteq C_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and
$A_i^\ell $
contains a point
$z_t$
with
$t^\sigma _{\mathfrak M_i}(z_t)=t$
, for each
$t\in \Theta $
. As
$Y_i^\ell \subseteq I_i^\ell ={\boldsymbol {b}}_i$
,
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}=C_{{\boldsymbol {b}}_i}^+$
, and
$t^\sigma _{\mathfrak M_1}({\boldsymbol {b}}_1)=t^\sigma _{\mathfrak M_2}({\boldsymbol {b}}_2)$
, such
$A_i^\ell $
exist by Lemma 4.13
$(e)$
. Observe that not only
$t^\sigma _{\mathfrak M_1}(A_1^\ell )=t^\sigma _{\mathfrak M_2}(A_2^\ell )$
but, by Lemma 4.17
$(b)$
and
$(d)$
, we actually have (26). Then
$k:=\big |A_1^\ell \big |=\big |A_2^\ell \big | \le 2^{|\delta |}$
, by Lemma 2.4
$(b)$
, and also
By taking
$H_i^{\ell ,\boldsymbol {r}_i^\ell -1}=A_i^\ell $
and
$\mathcal {P}_i^{\ell ,\boldsymbol {r}_i^\ell -1}=2^{A_i^\ell }$
,
$i=1,2$
, we obtain
$\mathfrak H_1^{\ell ,\boldsymbol {r}_1^\ell -1}$
and
$\mathfrak H_2^{\ell ,\boldsymbol {r}_2^\ell -1}$
both isomorphic to
. (The sets
$A_i^\ell $
are used differently in Case III.) Then we have (27). For
$p\in \delta $
, set
$\mathfrak {w}_i^{\ell ,\boldsymbol {r}_i^\ell -1}(p)=\mathfrak {v}_i(p)\cap H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and
$\mathsf {h}_i^{\ell ,\boldsymbol {r}_i^\ell -1}(x)=x$
for all
$x\in H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
. Then we clearly have
$\mathfrak w_i^{\ell ,\boldsymbol {r}_i^\ell -1}(p)\in \mathcal {P}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
, (24) for
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}= (\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1},\mathfrak w_i^{\ell ,\boldsymbol {r}_i^\ell -1})$
, and (13)–(22) hold for
$\mathfrak H=\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
,
$\mathfrak N=\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
,
$\mathsf {h}=\mathsf {h}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and
$I=J_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
((18) is by Lemma 2.5). This gives (23) for
$j=\boldsymbol {r}_i^\ell -1$
.
Finally, we claim that
$(a)$
–
$(c)$
hold for
$\mathfrak N_i^\ell =\mathfrak N_i^{\ell ,0}\lhd \dots \lhd \mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and some
$\boldsymbol {n}_i^\ell $
with
$0<\boldsymbol {n}_i^\ell \le 2\boldsymbol {r}_i^\ell $
. Indeed,
$(b)$
is by the definition of
$\lhd $
and (24). For
$(c)$
: The final cluster in
$\mathfrak N_i^\ell =$
final cluster in
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}=$
the non-degenerate cluster
$A_i^\ell $
. So the requirements in Definition 4.3
$(b)$
follow from (26) and (27). For
$(a)$
: By (23), each
$\mathfrak N_i^{\ell ,j}$
is
$\big (J_i^{\ell ,j},i\big )$
-nice, for
$j<\boldsymbol {r}_i^\ell $
, that is, conditions (13)–(22) are satisfied for
$\mathfrak N=\mathfrak N_i^{\ell ,j}$
,
$\mathfrak H=\mathfrak H_i^{\ell ,j}$
,
$I=J_i^{\ell ,j}$
, and
$\mathsf {h}=\mathsf {h}_i^{\ell ,j}$
(as defined above). We claim that (13)–(22) are satisfied for
$\mathfrak N=\mathfrak N_i^\ell $
,
$\mathfrak H=$
the frame
$\mathfrak H_i^\ell $
underlying
$\mathfrak N_i^\ell $
,
$I=I_i^\ell $
and
$\mathsf {h}_i^\ell =\bigcup _{j<\boldsymbol {r}_i^\ell }\mathsf {h}_i^{\ell ,j}$
. Indeed, (13), (14), and (18)–(20) clearly follow from (25), the definition of
$\lhd $
, and the corresponding properties for
$\mathfrak N_i^{\ell ,j}$
,
$\mathfrak H_i^{\ell ,j}$
,
$J_i^{\ell ,j}$
, and
$\mathsf {h}_i^{\ell ,j}$
,
$j<\boldsymbol {r}_i^\ell $
; (15) follows from (15) for
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and
$J^{\ell ,\boldsymbol {r}_i^\ell -1}$
; and (16) follows from (16) for
$\mathfrak H_i^{\ell ,0}$
and
$J^{\ell ,0}$
. For (17): Suppose
$x\in H_i^\ell $
,
$\{x\}$
is a degenerate non-root cluster in
$\mathfrak H_i^\ell $
and
$C\subseteq I_i^\ell $
is the immediate predecessor of
$\{x\}$
in
$\mathfrak F_i$
. Let
$j<\boldsymbol {r}_i^\ell $
be such that
$x\in H_i^{\ell ,j}$
. If
$\{x\}$
is the root cluster in
$\mathfrak H_i^{\ell ,j}$
, then
$j>0$
and
$\{x\}$
is the root cluster in
${\mathfrak F_i}\mathop {\restriction }_{J_i^{\ell ,j}}$
by (16) for
$\mathfrak H_i^{\ell ,j}$
and
$J_i^{\ell ,j}$
. Thus, the final cluster
$C^-\subseteq H_i^{\ell ,j-1}\subseteq H_i^\ell $
of
$\mathfrak H_i^{\ell ,j-1}$
is a subset of C by (15) for
$\mathfrak H_i^{\ell ,j-1}$
and
$J_i^{\ell ,j-1}$
. If
$\{x\}$
is a non-root cluster in
$\mathfrak H_i^{\ell ,j}$
, then
$C\subseteq J_i^{\ell ,j}$
and (17) for
$\mathfrak H_i^\ell $
and
$I_i^\ell $
follows from
$H_i^{\ell ,j}\subseteq H_i^\ell $
and (17) for
$\mathfrak H_i^{\ell ,j}$
and
$J_i^{\ell ,j}$
. For (21): Suppose
$x,y\in H_i^\ell $
,
$xR_i y$
and let
$j\le j'<\boldsymbol {r}_i^\ell $
be such that
$x\in H_i^{\ell ,j}$
and
$y\in H_i^{\ell ,j'}$
. Then
$\mathsf {h}_i^\ell (x)R_i\mathsf {h}_i^\ell (y)$
follows by (21) for
$\mathsf {h}_i^{\ell ,j}$
when
$j=j'$
, and by the definition of
$\lhd $
when
$j<j'$
. For (22): Suppose
$x,y\in H_i^\ell $
,
$y\in \boldsymbol {M}_{i}$
,
$\mathsf {h}_i^\ell (x)R_i y$
, and let
$j\le j'<\boldsymbol {r}_i^\ell $
be such that
$x\in H_i^{\ell ,j}$
and
$y\in H_i^{\ell ,j'}$
. Then
$xR_i y$
follows by (22) for
$\mathsf {h}_i^{\ell ,j}$
when
$j=j'$
, and by the definition of
$\lhd $
when
$j<j'$
.
Case III:
$(I_1^\ell ,I_2^\ell )$
is added in step
${\mathbf {(s{_2})}}$
. For
$i=1,2$
, let
${\boldsymbol {b}}_i$
be the relevant
$\sigma $
-blocks such that
$t^\sigma _{\mathfrak M_1}({\boldsymbol {b}}_1)=t^\sigma _{\mathfrak M_2}({\boldsymbol {b}}_2)$
and
$I_i^\ell $
is extending
${\boldsymbol {b}}_i$
that is not definable in
$\mathfrak M_i$
. We use the notation from Case II. As explained in Case II, we may again assume that, for every
$j< \boldsymbol {r}_i^\ell -1$
,
$C_i^{\ell ,j}$
is a non-limit cluster. However, as now
${\boldsymbol {b}}_i$
is not definable in
$\mathfrak M_i$
,
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}=C_{{\boldsymbol {b}}_i}^+$
is a limit cluster by Lemma 4.13
$(d)$
. We again define sets
$A_i^\ell \subseteq C_{{\boldsymbol {b}}_i}^+$
, intervals
$J_i^{\ell ,j}\subseteq I_i^\ell $
, and models
$\mathfrak N_i^{\ell ,j}=(\mathfrak H_i^{\ell ,j},\mathfrak w_i^{\ell ,j})$
with
$\mathfrak H_i^{\ell ,j}=(H_i^{\ell ,j},{R_i}\mathop {\restriction }_{H_i^{\ell ,j}},\mathcal {P}_i^{\ell ,j})$
such that (23)–(27) hold. Then we show that
$(a)$
–
$(c)$
of the lemma hold for
$\mathfrak N_i^\ell =\mathfrak N_i^{\ell ,0}\lhd \dots \lhd \mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
. This time,
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is
$\sigma $
-matching because it is of type
$(c)$
in Definition 4.3.
To this end, for any
$i=1,2$
and
$j<\boldsymbol {r}_i^\ell -1$
, we define everything like in Case II.2. For
$j=\boldsymbol {r}_i^\ell -1$
, we set
$J_i^{\ell ,\boldsymbol {r}_i^\ell -1}=[D_i^\ell ,E_i^\ell ]$
, where
$D_i^\ell $
is the root cluster in
${\mathfrak F_i}\mathop {\restriction }_{I_i^\ell }$
if
$\boldsymbol {r}_i^\ell =1$
and the immediate successor of the non-limit cluster
$C_i^{\ell ,\boldsymbol {r}_i^\ell -2}$
if
$\boldsymbol {r}_i^\ell>1$
, and
$E_i^\ell $
is the final cluster in
$I_i^\ell $
. We clearly have (25) and can define the sets
$Y_i^\ell $
and
$A_i^\ell $
in the same way as in Case II.2. However, for property (18) to hold for
$\mathfrak H=\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
, we need to define
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
differently. We consider the two cases in step
${\mathbf {(s{_2})}}$
of Definition 4.18:
-
1. The tail of
${\mathfrak M_i}\mathop {\restriction }_{I_i^\ell }$
is
$\{b_i^n\in I_i^\ell \setminus {\boldsymbol {b}}_i\mid n<\omega \}$
with
$b_i^n R_i b_i^{n-1}$
,
$0<n<\omega $
. (Using the notation of Definition 4.18:
$\{b_i^n\}={\boldsymbol {b}}_i^{f(m+n)}$
,
$n<\omega $
.) -
2. There is a sequence of non-degenerate clusters
$D_i^n\subseteq I_i^\ell \setminus {\boldsymbol {b}}_i$
definable in
$\mathfrak M_i$
,
$n<\omega $
, with
$D_i^0$
being the final cluster in
${\mathfrak M_i}\mathop {\restriction }_{I_i^\ell }$
and
$D_i^n<_{R_i} D_i^{n-1}$
,
$0<n<\omega $
. (Using the notation of Definition 4.18:
$D_i^n=C_{{\boldsymbol {b}}_i^{f(m_{n+1})}}^+$
.) For
$n<\omega $
, we pick some
$b_i^n\in D_i^n$
.
In both cases, we set
$H_i^{\ell ,\boldsymbol {r}_i^\ell -1}=A_i^\ell \cup \{b_i^m\mid m<\omega \}$
. Take the
$k<\omega $
with
$|A_1^\ell |=|A_2^\ell |=k$
, and suppose
$A_i^\ell =\{a_i^0,\dots ,a_i^{k-1}\}$
,
$i=1,2$
. We let
$\mathcal {P}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
be generated in
$(H_i^{\ell ,\boldsymbol {r}_i^\ell -1},{R_i}\mathop {\restriction }_{H_i^{\ell ,\boldsymbol {r}_i^\ell -1}})$
by the sets
$\{b_i^n\}$
,
$n<\omega $
, and
$X_i^s$
,
$s<k$
, where
$X_i^s=\{a_i^s\} \cup \{b_i^n\mid n < \omega ,\ n \equiv s \ (\text {mod}\ k)\}$
(see Example 2.2). The resulting
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
are both isomorphic to
in case 1., and to
in case 2. For
$p\in \delta $
, we set
$\mathfrak w_i^{\ell ,\boldsymbol {r}_i^\ell -1}(p)=\bigcup _{a_i^s\in \mathfrak v_i(p)}X_i^s$
. For every
$x\in H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
, we set
$$\begin{align*}\mathsf{h}_i^{\ell,\boldsymbol{r}_i^\ell-1}(x)=\left\{ \begin{array}{@{}ll@{}} x, & \text{if } x=a_i^s, \text{ for } s<k,\\[3pt] a_i^s & \text{if } x=b_i^n, n<\omega \text{ and } n\equiv s\ (\text{mod } k). \end{array} \right. \end{align*}$$
Then clearly
$\mathfrak w_i^{\ell ,\boldsymbol {r}_i^\ell -1}(p)\in \mathcal {P}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and (24) holds for
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}= (\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1},\mathfrak w_i^{\ell ,\boldsymbol {r}_i^\ell -1})$
. It is not hard to check that (23) holds for
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
as well. In particular, (18) for
$\mathfrak H=\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
follows from (3) and Lemma 2.5. Also, as
$A_i^\ell \subseteq C_{{\boldsymbol {b}}_i}^+$
and
$C_{{\boldsymbol {b}}_i}^+$
is a limit cluster,
$\boldsymbol {M}_{i}\cap H_i^{\ell ,\boldsymbol {r}_i^\ell -1}=\boldsymbol {M}_{i}\cap A_i^\ell =\emptyset $
follows by Lemma 2.6, and so we also have (22) for
$H=H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and
$\mathsf {h}=\mathsf {h}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
.
Next, the arguments showing that
$(a)$
and
$(b)$
of the lemma hold for the models
$\mathfrak N_i^\ell =\mathfrak N_i^{\ell ,0}\lhd \dots \lhd \mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and some
$\boldsymbol {n}_i^\ell $
with
$0<\boldsymbol {n}_i^\ell \le 2\boldsymbol {r}_i^\ell $
are the same as in Case II.2. To establish
$(c)$
, we show that the pair
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
is of type
$(c)$
in Definition 4.3. Indeed, observe that the last
$\lhd $
-components of
$\mathfrak N_i^\ell $
are
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
whose underlying frames
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
are both isomorphic to the same atomic frame of the form
or
, with
$0<k\le 2^{|\delta |}$
. Also, the
-cluster in
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
is
$A_i^\ell $
, and so the requirements in Definition 4.3
$(c)$
follow from (26) and (27).
Finally, observe that
$\boldsymbol {n}_i^\ell =1$
if
$(I_1^\ell ,I_2^\ell )$
is added in step
${\mathbf {(s{_3})}}$
of Definition 4.18 (see Case I), and
$\boldsymbol {n}_i^\ell \le 2\boldsymbol {r}_i^\ell $
if
$(I_1^\ell ,I_2^\ell )$
is added in steps
${\mathbf {(s{_1})}}$
or
${\mathbf {(s{_2})}}$
(see Cases II and III). So
$\sum _{\ell <N}\boldsymbol {n}_i^\ell \le (\boldsymbol {k}(\varphi _1,\varphi _2)-1) + \sum _{\ell <N}2\boldsymbol {r}_i^\ell \le 3\boldsymbol {k}(\varphi _1,\varphi _2)-1$
, as required.
We now complete the proof of Theorem 4.5. In Definition 4.18, we partitioned the models
$\mathfrak M_1,x_1$
and
$\mathfrak M_2,x_2$
witnessing the lack of interpolants for
$\varphi _1$
,
$\varphi _2$
into the same polynomial number N of intervals. For each
$\ell <N$
, Lemma 4.21 gave us a pair of models
$(\mathfrak N_1^\ell ,\mathfrak N_2^\ell )$
. Let
$\mathfrak N_i=\mathfrak N_i^0\lhd \dots \lhd \mathfrak N_i^{N-1}$
, for
$i=1,2$
.
Lemma 4.22. Conditions
$(a)$
–
$(d)$
in Theorem 4.5 hold for
$\mathfrak N_1,x_1$
and
$\mathfrak N_2,x_2$
.
Proof. We use the notation of the proof of Lemma 4.21. By Lemma 4.21
$(a)$
, each
$\mathfrak N_i^\ell $
is
$(I_i^\ell ,i)$
-nice, that is, conditions (13)–(22) are satisfied for
$\mathfrak N=\mathfrak N_i^\ell $
,
$\mathfrak H=\mathfrak H_i^\ell $
,
$I=I_i^\ell $
, and
$\mathsf {h}=\mathsf {h}_i^\ell $
.
$(a)$
We show by induction that
$\mathfrak M_i,\mathsf {h}_i^\ell (x)\models \tau $
iff
$\mathfrak N_i,x\models \tau $
, for any
$i=1,2$
,
$\ell <N$
,
$\tau \in \textit {sub}(\varphi _i)$
, and
$x\in H_i^\ell $
. Then
$\mathfrak {N}_{1},x_{1}\models \varphi _1$
and
$\mathfrak {N}_{2},x_{2}\models \neg \varphi _2$
follow from
$\mathfrak {M}_{1},x_{1}\models \varphi _1$
and
$\mathfrak {M}_{2},x_{2}\models \neg \varphi _2$
, as we have
$x_i\in H_i^0$
and
$\mathsf {h}_i^0(x_i)=x_i$
by (14) and (19). For
$\tau =p \in \delta $
, the statement follows from (20). The Boolean cases are straightforward, so suppose
$\tau =\Diamond \psi $
.
$(\Rightarrow )$
If
$\mathfrak M_i,\mathsf {h}_i^\ell (x)\models \Diamond \psi $
, then there are
$k\geq \ell $
and
$y_\psi \in \boldsymbol {M}_{i}\cap I_i^k$
with
$\mathsf {h}_i^\ell (x)R_i y_\psi $
and
$\mathfrak M_i,y_\psi \models \psi $
. We have
$y_\psi \in H_i^k$
by (14), and so
$\mathfrak N_i,y_\psi \models \psi $
by (19) and IH. We claim that
$xR_i y_\psi $
, and so
$\mathfrak N_i,x\models \Diamond \psi $
. Indeed, for
$k>\ell $
, this follows from the definition of
$\lhd $
, and for
$\ell =k$
, by (22).
$(\Leftarrow )$
If
$\mathfrak N_i,x\models \Diamond \psi $
, then there are
$k\geq \ell $
and
$y\in H_i^k$
with
$xR_i y$
and
$\mathfrak N_i,y\models \psi $
. We have
$\mathfrak M_i,\mathsf {h}_i^k(y)\models \psi $
by IH, and
$\mathsf {h}_i^\ell (x)R_i\mathsf {h}_i^k(y)$
by the definition of
$\lhd $
when
$k>\ell $
, and by (21) when
$k=\ell $
. Hence
$\mathfrak M_i,\mathsf {h}_i(x)\models \Diamond \psi $
.
$(c)$
follows from (14), (19), (20) and
$\mathfrak M_1,x_1\sim _\sigma \mathfrak M_2,x_2$
.
$(d)$
It is shown in Definition 4.18 that
$0<N< 2\boldsymbol {k}(\varphi _1,\varphi _2)$
. The rest of
$(d)$
follows from Lemma 4.21
$(b)$
and
$(c)$
.
$(b)$
We use the refutation criteria for the canonical formulas to show that the frame
$\mathfrak H_i$
underlying
$\mathfrak N_i$
is a frame for L,
$i=1,2$
. To this end, we prove that,
$$ \begin{align} & \text{for any canonical formula } \alpha(\mathfrak G, \mathfrak D,\bot), \text{ if } f \text{ is an injection}\\[-2pt]\nonumber & \text{from } \mathfrak G \text{ to } \mathfrak H_i \text{ satisfying } \mathbf{(cf_1)}{-}\mathbf{(cf_4)}, \text{ then the same } f\\[-2pt]\nonumber & \text{is an injection from } \mathfrak G \text{ to } \mathfrak F_i \text{ satisfying } \mathbf{(cf_1)}{-}\mathbf{(cf_4)}. \end{align} $$
Indeed,
$\mathbf {(cf_1)}$
holds by (13) and the definition of
$\lhd $
;
$\mathbf {(cf_2)}$
holds, as the final cluster in
$\mathfrak H_i=$
final cluster in
$\mathfrak H_i^{N-1}\subseteq $
final cluster in
${\mathfrak F_i}\mathop {\restriction }_{I_i^{N-1}}=$
final cluster in
$\mathfrak M_i$
, by (15). Condition
$\mathbf {(cf_4)}$
holds by (18) and the definition of
$\lhd $
. For
$\mathbf {(cf_3)}$
, suppose
$x \in \mathfrak D$
,
$C(y)$
is the immediate predecessor of
$C(x) = \{x\}$
in
$\mathfrak G$
and
$C(f(y))$
is the immediate predecessor of
$C(f(x))=\{f(x)\}$
in
$\mathfrak H_i$
. Let
$\ell <N$
be such that
$x\in H_i^\ell $
. If
$\{x\}$
is the root cluster in
$\mathfrak H_i^\ell $
, then
$\ell>0$
and
$\{x\}$
is the root cluster in
${\mathfrak F_i}\mathop {\restriction }_{I_i^\ell }$
by (16). Thus,
$C(f(y))$
in
$\mathfrak H_i=$
final cluster in
$\mathfrak H_i^{\ell -1}\subseteq $
final cluster in
${\mathfrak F_i}\mathop {\restriction }_{I_i^{\ell -1}}=C(f(y))$
in
$\mathfrak F_i$
, by (15). If
$\{x\}$
is a non-root cluster in
$\mathfrak H_i^\ell $
, then
$C(f(y))$
in
$\mathfrak H_i=C(f(y))$
in
$\mathfrak H_i^\ell \subseteq C(f(y))$
in
${\mathfrak F_i}\mathop {\restriction }_{I_i^\ell }=C(f(y))$
in
$\mathfrak F_i$
, by (17). Now, (29) implies that
$L\subseteq \mathsf {Log}(\mathfrak F_i)\subseteq \mathsf {Log}(\mathfrak H_i)$
, as required.
4.4 Proofs of Theorems 4.6 and 4.7
Suppose the finitely axiomatisable logic L is given by its canonical axioms as
$L = \mathsf {K4.3} \oplus \{\alpha (\mathfrak G_j,\mathfrak D_j,\bot ) \mid j\in J_L\}$
, for some finite index set
$J_L$
and
$\mathfrak G_j=(V_j,S_j)$
,
$j\in J_L$
. Let
$\boldsymbol {c}_L=\max _{j\in J_L}|V_j|$
. Given formulas
$\varphi _1$
,
$\varphi _2$
without an interpolant in L, let
$0<N<2\boldsymbol {k}(\varphi _1,\varphi _2)$
and
$\mathfrak N_i=\mathfrak N_i^0\lhd \dots \lhd \mathfrak N_i^{N-1}$
with root
$x_i i=1,2$
, be the models satisfying the conditions of Theorem 4.5 and obtained via Lemma 4.21. In particular, the underlying frame
$\mathfrak H_i$
of each
$\mathfrak N_i$
is a frame for L. We show in Lemma 4.23 below that the proof of Lemma 4.21 can be refined to yield polynomial-size models
$\mathfrak N_i^{\star \ell }$
,
$\ell <N$
. However,
$\mathfrak N_i^{\star \ell }$
is no longer
$(I_i^\ell ,i)$
-nice, as conditions (16) and (17) in Definition 4.20 do not necessarily hold for
$\mathfrak H = \mathfrak H_i^{\star \ell }$
underlying
$\mathfrak N_i^{\star \ell }$
and
$I=I_i^\ell $
. Thus, we do not have (29) in the proof of Lemma 4.22 for the frames
$\mathfrak H_i^\star $
underlying
$\mathfrak N_i^\star =\mathfrak N_i^{\star 0}\lhd \dots \lhd \mathfrak N_i^{\star N-1}$
. We prove that
$\mathfrak H_i^\star $
,
$i=1,2$
, are frames for L (as required by Theorem 4.6
$(b)$
) by using Lemma 4.24 below instead.
Take the number N,
$0<N<2\boldsymbol {k}(\varphi _1,\varphi _2)$
, provided by Definition 4.18, the numbers
$\boldsymbol {n}_i^\ell>0$
with
$\sum _{\ell <N}\boldsymbol {n}_i^\ell \le 3\boldsymbol {k}(\varphi _1,\varphi _2)-1$
, and sets
$H_i^\ell $
,
$i=1,2$
,
$\ell <N$
, from Lemma 4.21.
Lemma 4.23. If
$L\supseteq \mathsf {K4.3}$
is finitely axiomatisable, then, for
$i=1,2$
,
$\ell <N$
, there exist sets
$H_i^{\star \ell }\subseteq H_i^\ell $
and models
$\mathfrak N_i^{\star \ell }$
based on frames
$\mathfrak H_i^{\star \ell }=(H_i^{\star \ell },S_i^{\star \ell },\mathcal {P}_i^{\star \ell })$
such that the following hold:
-
(a)
$\mathfrak N_i^{\star \ell }$
is ‘almost’
$(I_i^\ell ,i)$
-nice in the sense that (13)–(15) and (18)–(22) hold for
$\mathfrak N=\mathfrak N_i^{\star \ell }$
and
$I=I_i^\ell $
; -
(b)
$\mathfrak N_i^{\star \ell }$
is the ordered sum of
$\boldsymbol {n}_i^\ell $
-many simple
$\delta $
-models based on L-bounded atomic frames; -
(c) the pair
$(\mathfrak N_1^{\star \ell },\mathfrak N_2^{\star \ell })$
is
$\sigma $
-matching.
Proof. We go through Cases I–III in the proof of Lemma 4.21 and make the necessary modifications.
Case I:
$(I_1^\ell ,I_2^\ell )$
is added in step
${\mathbf {(s{_3})}}$
of Definition 4.18. An inspection of this part of the proof of Lemma 4.21 reveals that
$m^<$
or
is used in cases
$(i)$
and
$(ii)$
, and in both cases all the m elements of the finite non-empty tails
$Z_i^\ell $
of
${\mathfrak F_i}\mathop {\restriction }_{I_i^\ell }$
are put into the chosen subset
$H_i^\ell $
of
$I_i^\ell $
. Now, we choose a subset
$H_i^{\star \ell }\subseteq H_i^\ell $
with
$|H_i^{\star \ell }|\le \boldsymbol {c}_L+2$
as follows. Suppose
$Z_i^\ell =\{z_i^a\mid a< m\}$
with
$z_i^aR_i^s z_i^{a-1}$
,
$0<a<m$
, and let
$m'=\min (m,\boldsymbol {c}_L+1)$
. We set
$H_i^{\star \ell }=\{z_i^a\mid a< m'\}$
in case
$(i)$
, and
$H_i^{\star \ell }=\{w_i^\ell \}\cup \{z_i^a\mid a< m'\}$
, for the chosen
$w_i^\ell $
from the head of
$Z_i^\ell $
in case
$(ii)$
. In case
$(iii)$
, we let
$H_i^{\star \ell }=H_i^\ell $
. Then, in all cases
$(i)$
–
$(iii)$
, we let
$\mathfrak H_i^{\star \ell }={\mathfrak H_i^\ell }\mathop {\restriction }_{H_i^{\star \ell }}$
and
$\mathfrak N_i^{\star \ell }={\mathfrak N_i^\ell }\mathop {\restriction }_{H_i^{\star \ell }}$
. Observe that we have
$\mathsf {h}_i^\ell (x)\in H_i^{\star \ell }$
, for every
$x\in H_i^{\star \ell }$
, and so
$\mathsf {h}_i^{\star \ell }={\mathsf {h}_i^\ell }\mathop {\restriction }_{H_i^{\star \ell }}$
is an
$H_i^{\star \ell }\to H_i^{\star \ell }$
function. It is straightforward to check that (13)–(15) and (18)–(22) hold for
$\mathfrak N=\mathfrak N_i^{\star \ell }$
,
$\mathfrak H=\mathfrak H_i^{\star \ell }$
,
$\mathsf {h}=\mathsf {h}_i^{\star \ell }$
, and
$I=I_i^\ell $
. Note that all non-degenerate clusters in
$\mathfrak H_i^{\star \ell }$
are of the form
this case, and so
$\mathfrak N_i^{\star \ell }$
is a simple
$\delta $
-model based on an L-bounded atomic frame.
Cases II and III:
$(I_1^\ell ,I_2^\ell )$
is added in step
${\mathbf {(s{_1})}}$
or
${\mathbf {(s{_2})}}$
. An inspection of these parts of the proof of Lemma 4.21 reveals that
$m^<$
or
is used only when
${\boldsymbol {b}}_i$
is non-degenerate, in cases
$(ii)$
and
$(iii)$
of the definition of
$\mathfrak H_i^{\ell ,j}$
for some
$j<\boldsymbol {r}_i^\ell -1$
. (Recall that
$\boldsymbol {r}_i^\ell $
denotes the number of relevant clusters in
$I_i^\ell $
.) In both cases
$(ii)$
and
$(iii)$
, all the m elements of the finite non-empty tail
$Z_i^{\ell ,j}$
of
${\mathfrak F_i}\mathop {\restriction }_{J_i^{\ell ,j}}$
are put into the chosen subset
$H_i^{\ell ,j}$
of
$J_i^{\ell ,j}$
, for some subinterval
$J_i^{\ell ,j}$
of
$I_i^\ell $
. We repeat the trick from Case I above. Suppose
$Z_i^{\ell ,j}=\{z^a\mid a< m\}$
with
$z^aR_i^s z^{a-1}$
,
$0<a<m$
, and let
$m'=\min (m,\boldsymbol {c}_L+1)$
. We set
$H_i^{\star \ell ,j}=\{z^a\mid a< m'\}$
in case
$(ii)$
, and
$H_i^{\star \ell ,j}=\{w_i^{\ell ,j}\}\cup \{z^a\mid a< m'\}$
, for the chosen
$w_i^{\ell ,j}$
from the head of
$Z_i^{\ell ,j}$
in case
$(iii)$
. In cases
$(i)$
and
$(iv)$
of Cases II and III, we let
$H_i^{\star \ell ,j}=H_i^{\ell ,j}$
. Then, in all cases
$(i)$
–
$(iv)$
, we let
$\mathfrak H_i^{\star \ell ,j}={\mathfrak H_i^{\ell ,j}}\mathop {\restriction }_{H_i^{\star \ell ,j}}$
and
$\mathfrak N_i^{\star \ell ,j}={\mathfrak N_i^{\ell ,j}}\mathop {\restriction }_{H_i^{\star \ell ,j}}$
, for
$j<\boldsymbol {r}_i^\ell -1$
. One can see that
$\mathsf {h}_i^{\ell ,j}(x)\in H_i^{\star \ell ,j}$
, for every
$x\in H_i^{\star \ell ,j}$
, and so
$\mathsf {h}_i^{\star \ell ,j}={\mathsf {h}_i^{\ell ,j}}\mathop {\restriction }_{H_i^{\star \ell ,j}}$
is an
$H_i^{\star \ell ,j}\to H_i^{\star \ell ,j}$
function. It is straightforward to check that, for all
$j<\boldsymbol {r}_i^\ell -1$
, (13)–(15) and (18)–(22) hold for
$\mathfrak N=\mathfrak N_i^{\star \ell ,j}$
,
$\mathfrak H=\mathfrak H_i^{\star \ell ,j}$
,
$\mathsf {h}=\mathsf {h}_i^{\star \ell ,j}$
, and
$I=J_i^{\ell ,j}$
(but (16) and (17) do not necessarily hold). Note that the size of non-degenerate clusters in these
$\mathfrak H_i^{\star \ell ,j}$
is bounded by
$\boldsymbol {k}(\varphi _1,\varphi _2)$
, and so
$\mathfrak N_i^{\star \ell ,j}$
is the ordered sum of at most two simple
$\delta $
-models based on L-bounded atomic frames.
We also need to adjust the definitions of
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
,
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
, and
$\mathsf {h}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
. We define the sets
$Y_i^{\star \ell }\subseteq Y_i^\ell $
and
$A_i^{\star \ell }\subseteq A_i^\ell $
from
$H_i^{\star \ell ,j}$
,
$j<\boldsymbol {r}_i^\ell -1$
, in the same way as
$Y_i^\ell $
and
$A_i^\ell $
were defined from
$H_i^{\ell ,j}$
,
$j<\boldsymbol {r}_i^\ell -1$
, resulting in (26) and (27). Let
$k^\star =\big |A_1^{\star \ell }\big |=\big |A_2^{\star \ell }\big |$
. By (28),
$$ \begin{align*} k^\star\le & |Y_1^{\star\ell|}+|Y_2^{\star\ell}|+\boldsymbol{k}(\varphi_1,\varphi_2)\le\\ & 2(\boldsymbol{k}(\varphi_1,\varphi_2)-1)\cdot\max\big(\boldsymbol{c}_L+2,\boldsymbol{k}(\varphi_1,\varphi_2)\big)+\boldsymbol{k}(\varphi_1,\varphi_2)=\boldsymbol{p}_L(\varphi_1,\varphi_2). \end{align*} $$
In Case II, we set
$\mathfrak H_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}={\mathfrak H_i^{\ell ,\boldsymbol {r}_i-1}}\mathop {\restriction }_{H_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}}$
,
$\mathfrak N_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}={\mathfrak N_i^{\ell ,\boldsymbol {r}_i-1}}\mathop {\restriction }_{H_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}}$
, and
$\mathsf {h}_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}={\mathsf {h}_i^{\ell ,\boldsymbol {r}_i-1}}\mathop {\restriction }_{H_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}}$
. In Case III, the definitions of
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
,
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
, and
$\mathsf {h}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
need to be mimicked for
$k^\star $
in place of k to obtain
$\mathfrak H_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
,
$\mathfrak N_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
, and
$\mathsf {h}_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
. It is straightforward to check now that (13)–(15) and (18)–(22) hold for
$\mathfrak N=\mathfrak N_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
,
$\mathfrak H=\mathfrak H_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
,
$\mathsf {h}=\mathsf {h}_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
, and
$I=J_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
. Note that the size
$k^\star $
of the root cluster in
$\mathfrak H_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
is bounded by
$\boldsymbol {p}_L(\varphi _1,\varphi _2)$
and every other non-degenerate cluster in it is of the form
, so
$\mathfrak N_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
is a simple
$\delta $
-model based on an L-bounded atomic frame.
Therefore,
$\mathfrak N_i^{\star \ell }=\mathfrak N_i^{\star \ell ,0}\lhd \dots \lhd \mathfrak N_i^{\star \ell ,\boldsymbol {r}_i^\ell -1}$
, for
$i=1,2$
,
$\ell <N$
, is the ordered sum of
$\boldsymbol {n}_i^\ell $
-many simple
$\delta $
-models based on L-bounded atomic frames, for the same
$\boldsymbol {n}_i^\ell $
as in Lemma 4.21, and so we have
$(b)$
of the lemma. Finally, by the same arguments as in the proof of Lemma 4.21, we obtain
$(a)$
and
$(c)$
.
Lemma 4.24. For
$i=1,2$
,
$\ell <N$
, take the frames
$\mathfrak H_i^{\ell }$
and
$\mathfrak H_i^{\star \ell }$
provided by Lemmas 4.21 and 4.23. Let
$\mathfrak H_i=\mathfrak H_i^{0}\lhd \dots \lhd \mathfrak H_i^{N-1}$
and
$\mathfrak H_i^\star =\mathfrak H_i^{\star 0}\lhd \dots \lhd \mathfrak H_i^{\star N-1}$
. Then, for any
$j\in J_L$
, if there is an injection f from
$\mathfrak G_j$
to
$\mathfrak H_i^\star $
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
for
$\alpha (\mathfrak G_j,\mathfrak D_j,\bot )$
, then there is an injection
$f^\dagger $
from
$\mathfrak G_j$
to
$\mathfrak H_i$
also satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
for
$\alpha (\mathfrak G_j,\mathfrak D_j,\bot )$
. Thus,
$\mathfrak H_i\models L$
implies
$\mathfrak H_i^\star \models L$
.
Proof. Fix some
$i \in \{1,2\}$
and
$j\in J_L$
. Suppose that f is an injection from
$\mathfrak G_j=(V_j,S_j)$
to
$\mathfrak H_i^\star $
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
for
$\alpha (\mathfrak G_j,\mathfrak D_j,\bot )$
. For every atomic
$\lhd $
-component
$\mathfrak F^\star =(H^\star ,{R_i}\mathop {\restriction }_{H^\star })$
in
$\mathfrak H_i^\star $
such that:
-
1.
$\mathfrak F^\star $
is obtained from the atomic
$\lhd $
-component
$\mathfrak F=(H,{R_i}\mathop {\restriction }_{H})$
in
$\mathfrak H_i$
of the form
$m^<$
or
, and -
2. there is
$v\in V_j$
such that
$f(v_j)$
is an irreflexive point in
$\mathfrak F^\star $
,
we proceed as follows. Suppose
$H=\{y_0,\dots ,y_{m-1}\}$
or
$H=\{y,y_0,\dots ,y_{m-1}\}$
with
$yR_i y R_i y_{m-1}R_i\dots R_i y_0$
, and so
$H^\star =\{y_0,\dots ,y_{\boldsymbol {c}_L}\}$
or
$H^\star =\{y,y_0,\dots ,y_{\boldsymbol {c}_L}\}$
. Let
$V_j^-=\big \{v\in V_j\mid f(v)\in H^\star \setminus \{y\}\big \}$
. Then
$|V_j^-|\le |V_j|\le \boldsymbol {c}_L$
. Thus, by the pigeonhole principle, there is
$i\le \boldsymbol {c}_L$
with
$y_i\notin f(V_i)\cap (H^\star \setminus \{y\})$
. Suppose
$V_j^-=\{v_0,\dots ,v_{s-1}\}$
, for some
$s\le \boldsymbol {c}_L$
with
$v_{s-1}S_j\dots S_j v_0$
. Let a be the largest
$k< s$
with
$y_iR_i f(v_k)$
. As f satisfies
$\mathbf {(cf_3)}$
,
$v_a\notin \mathfrak D_j$
. Now, for
$k<s$
, we set
$$\begin{align*}f^\dagger(v_k)=\left\{ \begin{array}{@{}ll@{}} y_k, & \text{ if } k\le a,\\[3pt] y_{m-(s-k)}, & \text{ if } a+1\le k<s. \end{array} \right. \end{align*}$$

We do this for every
$\mathfrak F^\star $
having 1. and 2. above, and set
$f^\dagger (x)=f(x)$
, for any other
$x\in V_j$
. It is straightforward to check that the resulting
$f^\dagger $
is an injection from
$\mathfrak G_j$
to
$\mathfrak H_i$
satisfying
$\mathbf {(cf_1)}$
–
$\mathbf {(cf_4)}$
for
$\alpha (\mathfrak G_j,\mathfrak D_j,\bot )$
.
This completes the proof of Theorem 4.6. We obtain Theorem 4.7 using Lemma 4.4 as we have, for
$i=1,2$
:
$$ \begin{align*} \|{\mathfrak N_i^\star}\|=\|{\mathfrak N_i^{\star 0}}\|+\cdots +\|{\mathfrak N_i^{\star N-1}}\| & \le\sum_{\ell<N}\boldsymbol{n}_i^\ell\cdot\max\big(\boldsymbol{c}_L+2,\boldsymbol{p}_L(\varphi_1,\varphi_2)\big)\le\\ & (3\boldsymbol{k}(\varphi_1,\varphi_2)-1)\cdot\max\big(\boldsymbol{c}_L+2,\boldsymbol{p}_L(\varphi_1,\varphi_2)\big). \end{align*} $$
4.5 Cofinal subframe logics
By Theorem 3.5
$(a)$
, all d-persistent cofinal subframe logics
$L \supseteq \mathsf {K4.3}$
have the polysize bisimilar model property, with the polynomial
$\boldsymbol {k}(\varphi _1,\varphi _2)$
(defined in (8)) not dependent on L. We show now that, for arbitrary, not necessarily d-persistent cofinal subframe L, it is enough to replace polysize in Theorem 3.5
$(a)$
by quasi-polysize.
Theorem 4.25. All cofinal subframe logics
$L \supseteq \mathsf {K4.3}$
have the quasi-polysize bisimilar model property, with the size of witnessing models bounded by
$\boldsymbol {k}(\varphi _1,\varphi _2)$
.
This follows from the following special case of the ‘structural’ Theorem 4.5.
Theorem 4.26. For any cofinal subframe logic
$L \supseteq \mathsf {K4.3}$
and formulas
$\varphi _1$
,
$\varphi _2$
without an interpolant in L, there are rooted
$\delta $
-models
$\mathfrak N_1,x_1$
and
$\mathfrak N_2,x_2$
satisfying
$(a)$
–
$(c)$
from Theorem 4.5 as well as conditions
$(d)$
and
$(e)$
below:
-
(d) there is
$M \leq \boldsymbol {k}(\varphi _1,\varphi _2)$
such that
$\mathfrak N_i=\mathfrak N_i^0\lhd \dots \lhd \mathfrak N_i^{M-1}$
and, for all
$j < M$
,-
1.
$\mathfrak N_i^j$
is the ordered sum of simple
$\delta $
-models based on atomic frames; -
2. the pair
$(\mathfrak N_1^j,\mathfrak N_2^j)$
is
$\sigma $
-matching;
-
-
(e)
$\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}$
coincides with the set of points in
$\mathfrak N_i$
,
$i=1,2$
that are not in the
$\{b_i^n\mid n<\omega \}$
-part of some
$\lhd $
-component based on a
.
It follows from
$(e)$
that
$\|{\mathfrak N_i}\|=|\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}|\le \boldsymbol {k}(\varphi _1,\varphi _2)$
.
Proof. As in the proof of Theorem 4.5, we take any
$\sigma $
-bisimilar witness models
$\mathfrak M_i,x_i$
,
$i=1,2$
, based on frames
$\mathfrak {F}_i = (W_{i},R_{i},\mathcal {P}_{i})$
for L. Let M be the number of relevant
$\sigma $
-blocks in
$\mathfrak M_1$
(or
$\mathfrak M_2$
, by Lemma 4.17
$(e)$
). For
$i=1,2$
, consider the partitions
$\mathcal I_i=\{I_i^\ell \in \mathcal {P}_i\mid \ell <N\}$
of
$\mathfrak M_i$
given by Definition 4.18, and let
$0=\ell _0<\dots <\ell _{M-1}=N-1$
be the list of indices such that the pair
$(I_1^{\ell _j},I_2^{\ell _j})$
is added to
$\mathcal I_1\times \mathcal I_2$
in step
${\mathbf {(s{_1})}}$
or
${\mathbf {(s{_2})}}$
, and
$I_i^{\ell _0}\prec _{\mathfrak F_i}\cdots \prec _{\mathfrak F_i}I_i^{\ell _{M-1}}$
. We define
$\mathfrak N_i^{\ell _z}$
,
$z<M$
, by choosing fewer points from
$I_i^{\ell _z}$
than in Cases II.2 and III in the proof of Lemma 4.21, and we also define functions
$\mathsf {h}_i^{\ell _z}$
. Let
$\ell =\ell _z$
, for
$z<M$
, let
$C_i^{\ell ,j}$
,
$j < \boldsymbol {r}_i^\ell $
, be the sequence (ordered by
$<_{R_i}$
) of all relevant clusters in
$I_i^\ell $
, and
$D_i^{\ell ,j}=C_i^{\ell ,j}\cap \big (\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}\big )$
. Three cases are possible now, the first of which coincides with Case II.1, while the other two select fewer points for
$\mathfrak N_i^{\ell _z}$
than Cases II.2 and III:
-
(i) If
$(I_1^{\ell },I_2^{\ell })$
is added in step
${\mathbf {\text {(s}{_1}\text {)}}}$
and
$I_i^{\ell }$
consists of a degenerate cluster, then, like in Case II.1, we let
$\mathfrak N_i^\ell ={\mathfrak M_i}\mathop {\restriction }_{I_i^\ell }$
and
$\mathsf {h}_i^\ell $
be the identity function on
${\mathfrak {N}}_i^\ell $
. -
(ii) If
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
is non-degenerate and
$(I_1^{\ell },I_2^{\ell })$
is added in step
${\mathbf {(s{_1})}}$
as in Case II.2, then
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
is definable in
$\mathfrak M_i$
. We let
$\mathfrak N_i^\ell ={\mathfrak M_i}\mathop {\restriction }_{D_i^{\ell ,0}}\lhd \dots \lhd {\mathfrak M_i}\mathop {\restriction }_{D_i^{\ell ,\boldsymbol {r}_i^\ell -1}}$
and
$\mathsf {h}_i^\ell $
be the identity function on
$\mathfrak N_i^\ell $
. -
(iii) If
$(I_1^{\ell },I_2^{\ell })$
is added in
${\mathbf {(s{_2})}}$
like in Case III, then
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
is a not definable in
$\mathfrak M_i$
. As shown in Case III, there is an infinite sequence of irrelevant points
$\{b_i^n\in I_i^\ell \mid n<\omega \}$
such that
$b_i^n R_i b_i^{n-1}$
,
$C_i^{\ell ,\boldsymbol {r}_i^\ell -1}<_{R_i} C(b_n^i)$
and
$C(b_n^i)\in \mathcal {P}_i$
,
$n<\omega $
, and the
$b_i^n$
are either 1) all irreflexive or 2) all reflexive. By Lemma 4.17, there is
$k\le \boldsymbol {k}(\varphi _1,\varphi _2)$
with
$|D_1^{\ell ,\boldsymbol {r}_1^\ell -1}|=|D_2^{\ell ,\boldsymbol {r}_2^\ell -1}|=k$
. Suppose
$D_i^{\ell ,\boldsymbol {r}_i^\ell -1}=\{a_i^0,\dots ,a_i^{k-1}\}$
. We let
$H_i^{\ell ,\boldsymbol {r}_i^\ell -1}=D_i^\ell \cup \{b_i^n\mid n<\omega \}$
and
$\mathcal {P}_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
be generated in
$(H_i^{\ell ,\boldsymbol {r}_i^\ell -1},{R_i}\mathop {\restriction }_{H_i^{\ell ,\boldsymbol {r}_i^\ell -1}})$
by the sets
$\{b_i^n\}$
,
$n<\omega $
, and
$X_i^s=\{a_i^s\} \cup \{b_i^n\mid n < \omega ,\ n \equiv s \ (\text {mod}\ k)\}$
,
$s<k$
(see Example 2.2). The resulting
$\mathfrak H_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
are both isomorphic to
in case 1), and to
in case 2). We then set
$\mathfrak w_i^{\ell ,\boldsymbol {r}_i^\ell -1}(p)=\bigcup _{a_i^s\in \mathfrak v_i(p)}X_i^s$
and
$\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}=\big ((H_i^{\ell ,\boldsymbol {r}_i^\ell -1},{R_i}\mathop {\restriction }_{H_i^{\ell ,\boldsymbol {r}_i^\ell -1}}, \mathcal {P}^{\ell ,\boldsymbol {r}_i^\ell -1}),\mathfrak w_i^{\ell ,\boldsymbol {r}_i^\ell -1}\big )$
. Finally, we set
$\mathfrak N_i^\ell ={\mathfrak M_i}\mathop {\restriction }_{D_i^{\ell ,0}}\lhd \dots \lhd {\mathfrak M_i}\mathop {\restriction }_{D_i^{\ell ,\boldsymbol {r}_i^\ell -2}}\mathop {\lhd }\mathfrak N_i^{\ell ,\boldsymbol {r}_i^\ell -1}$
and define
$\mathsf {h}_i^\ell $
as the identity on relevant points in
$\mathfrak N_i^\ell $
and
$\mathsf {h}_i^\ell (b_i^n)=a_i^s$
, for
$n<\omega $
with
$n\equiv s$
(mod k).
Clearly,
$(c)$
,
$(d).1,$
and
$(e)$
hold for
$\mathfrak N_i=\mathfrak N_i^{\ell _0}\lhd \dots \lhd \mathfrak N_i^{\ell _{M-1}}$
. Condition
$(a)$
is shown like in Lemma 4.22 using that (19)–(22) hold for
$\mathsf {h}=\mathsf {h}_i^{\ell _z}$
and
$H=H_i^{\ell _z}$
,
$z<M$
. Condition
$(b)$
is proved via (29):
$\mathbf {(cf_1)}$
clearly holds;
$\mathbf {(cf_2)}$
holds as the final cluster in
$\mathfrak F_i$
is always relevant; and
$\mathbf {(cf_4)}$
holds, as
$\{x\}$
being definable in
$\mathfrak N_i^{\ell _z}$
implies
$\{x\}\in \mathcal {P}_i$
, for all
$z<M$
and x in
$\mathfrak N_i^{\ell _z}$
. As L is a cofinal subframe logic,
$\mathfrak D=\emptyset $
, so
$\mathbf {(cf_3)}$
holds vacuously. Finally, to show
$(d).2$
, observe that
$(\mathfrak N_1^{\ell _z},\mathfrak N_2^{\ell _z})$
is
$\sigma $
-matching as it always meets one of the conditions in Definition 4.3: in case
$(i)$
, it meets
$(a)$
; in case
$(ii)$
, it meets
$(b)$
; and in case
$(iii)$
, it meets
$(c)$
.
Example 4.27. By Example 2.10
$(a)$
, given any formulas
$\varphi _1$
and
$\varphi _2$
without an interpolant in
$\mathsf {GL.3}$
, one can always find witnessing models
$\mathfrak N_i$
,
$i=1,2$
, of size
$\le \boldsymbol {k}(\varphi _1,\varphi _2)$
that are ordered sums of simple models based on
$m^<$
or
(see, e.g., the models depicted in Figure 1 in Example 3.6
$(a)$
).
We emphasise that the construction in the proof of Theorem 4.26 does not work for non-cofinal subframe logics, in which case
$\mathfrak D \ne \emptyset $
; see also the special treatment of the density axiom in the proof of Theorem 5.9 below.
5 The IEP for standard Priorean temporal logics
Priorean temporal logics [Reference Prior36] deal with the operators ‘sometime in the future’ denoted
$\Diamond _{\mathsf {F}}$
, ‘sometime in the past’ denoted
$\Diamond _{\mathsf {P}}$
, and their duals ‘always in the future’
$\Box _{\mathsf {F}}$
and ‘always in the past’
$\Box _{\mathsf {P}}$
. Temporal formulas—propositional bimodal formulas with these operators—are interpreted over general temporal frames of the form
$\mathfrak F = (W,R,R^-,\mathcal {P})$
representing various flows of time in such a way that
$(W,R)$
is transitive and connected (2), R is the ‘future-time’ accessibility relation for
$\Diamond _{\mathsf {F}}$
,
$\Box _{\mathsf {F}}$
, its inverse
$R^-$
is the ‘past-time’ accessibility relation for
$\Diamond _{\mathsf {P}}$
,
$\Box _{\mathsf {P}}$
, and the internal sets
$\mathcal {P} \subseteq 2^W$
are closed under the Booleans and the operators
To simplify notation, we omit
$R^-$
and write
$\mathfrak F = (W,R,\mathcal {P})$
. Also, as before, if
$\mathcal {P} = 2^W$
, we call
$\mathfrak F$
a Kripke frame and write
$\mathfrak F = (W,R)$
. The universal modality ‘always’ can be introduced as an abbreviation
$\Box \varphi = \varphi \land \Box _{\mathsf {F}} \varphi \land \Box _{\mathsf {P}} \varphi $
. Descriptive temporal frames are defined in the same way as in Section 2. Note that tightness condition (tig) for
$R^-$
actually follows from (tig) for R.
In fact, many results from Sections 2 and 3 straightforwardly generalise to the temporal setting. Let
$\mathfrak {M}$
be a temporal model—that is, a model based on some temporal frame
$\mathfrak {F} = (W,R,\mathcal {P})$
—and let
$\Gamma $
be a set of temporal formulas. A point
$x\in W$
is
$\Gamma $
-minimal in
$\mathfrak {M}$
if
$\mathfrak {M},x \models \Gamma $
and whenever
$x'R x$
and
$\mathfrak {M},x'\models \Gamma $
, then
$x R x'$
. Denote by
$\min _{\mathfrak {M}} \Gamma $
the set of all
$\Gamma $
-minimal points in
$\mathfrak {M}$
. (The definition of
$\max _{\mathfrak {M}} \Gamma $
remains the same.) In the temporal case, Lemma 2.3 generalises to the following lemma.
Lemma 5.1. Suppose
$\Gamma $
is a set of temporal formulas and
$\mathfrak M$
is a model based on a descriptive temporal frame
$\mathfrak {F}= (W,R,\mathcal {P})$
. Then the following hold:
(temporal saturation)
If
$\mathfrak M, x\models \Diamond _{\mathsf {F}} \bigwedge \Gamma '$
for every finite
$\Gamma ' \subseteq \Gamma $
, then there is y with
$xRy$
and
$\mathfrak M, y \models \Gamma $
. If
$\mathfrak M, x\models \Diamond _{\mathsf {P}} \bigwedge \Gamma '$
for every finite
$\Gamma ' \subseteq \Gamma $
, then there is y with
$xR^-y$
and
$\mathfrak M, y \models \Gamma $
.
(maximal and minimal points)
If there is x with
$\mathfrak M, x \models \Gamma $
, then
$\max _{\mathfrak {M}}\Gamma \ne \emptyset $
and
$\min _{\mathfrak {M}}\Gamma \ne \emptyset $
.
A relation
$\boldsymbol {\beta } \subseteq W_1 \times W_2$
is a temporal
$\sigma $
-bisimulation between temporal models
$\mathfrak M_1$
and
$\mathfrak M_2$
based on respective frames
$\mathfrak F_i = (W_i,R_i,\mathcal {P}_i)$
,
$i=1,2$
, if it satisfies (atom), (move) and its past-time couterpart: whenever
$x_1\boldsymbol {\beta } x_2$
, then
(move
$^-$
)
$x_1 R^-_1 y_1$
implies
$y_1\boldsymbol {\beta } y_2$
, for some
$y_2 \in W_2$
with
$x_2 R^-_2 y_2$
; conversely,
$x_2 R^-_2 y_2$
implies
$y_1\boldsymbol {\beta } y_2$
, for some
$y_1 \in W_1$
with
$x_1 R^-_1 y_1$
.
The relation
$\mathfrak {M}_1,x_1 \equiv _{\sigma } \mathfrak {M}_2,x_2$
, saying that temporal models
$\mathfrak {M}_1$
and
$\mathfrak {M}_2$
satisfy the same temporal
$\sigma $
-formulas at
$x_1$
and
$x_2$
, respectively, is characterised in terms of temporal
$\sigma $
-bisimulations: it is readily seen that, with this modification, Lemma 3.1 and Theorem 3.2 continue to hold for all Priorean temporal logics. (As temporal frames are transitive and connected, any of their points can be regarded as a root with respect to the relation
$R\cup R^-$
.)
In this article, we consider the Priorean temporal logics of five most popular classes of temporal Kripke frames [Reference Bull7]:
-
$\mathsf {Lin} = \{ \varphi \mid \mathfrak F \models \varphi , \ \mathfrak F = (W,R) \text { is any temporal Kripke frame} \}$
${}\!\!\!\! = \mathsf {K4}_2 \oplus p \to \Box _{\mathsf {F}}\Diamond _{\mathsf {P}} p \oplus p \to \Box _{\mathsf {P}}\Diamond _{\mathsf {F}} p \oplus \Diamond _{\mathsf {F}}\Diamond _{\mathsf {P}} p \lor \Diamond _{\mathsf {P}}\Diamond _{\mathsf {F}} p \to p \lor \Diamond _{\mathsf {F}} p \lor$
${}\qquad \Diamond _{\mathsf {P}} p$
; -
$\mathsf {Lin}_{\mathbb {Q}} = \{ \varphi \mid (\mathbb Q,<) \models \varphi \}$
${}= \mathsf {Lin} \oplus \Diamond _{\mathsf {F}} \top \oplus \Diamond _{\mathsf {P}} \top \oplus \Diamond _{\mathsf {F}} p \to \Diamond _{\mathsf {F}}\Diamond _{\mathsf {F}} p$
; -
$\mathsf {Lin}_{\mathbb {R}} = \{ \varphi \mid (\mathbb R,<) \models \varphi \}$
${}= \mathsf {Lin}_{\mathbb {Q}} \oplus \Box (\Box _{\mathsf {P}} p \to \Diamond _{\mathsf {F}}\Box _{\mathsf {P}} p) \to (\Box _{\mathsf {P}} p \to \Box _{\mathsf {F}} p)$
; -
$\mathsf {Lin}_{<\omega } = \{ \varphi \mid \mathfrak {F} \models \varphi , \ \mathfrak {F} = (W,<)\text { any } \mathit{finite} \text { strict linear order} \}$
${} = \mathsf {Lin} \oplus \big \{ \Box _{\mathsf {X}} (\Box _{\mathsf {X}} p \to p) \to \Box _{\mathsf {X}} p \mid \mathsf {X} \in \{\mathsf {F}, \mathsf {P}\}\big \}$
; -
$\mathsf {Lin}_{\mathbb {Z}} = \{ \varphi \mid (\mathbb Z,<) \models \varphi \}$
${} = \mathsf {Lin} \oplus \Diamond _{\mathsf {F}} \top \oplus \Diamond _{\mathsf {P}} \top \oplus \big \{ \Box _{\mathsf {X}}(\Box _{\mathsf {X}} p \to p) \to (\Diamond _{\mathsf {X}}\Box _{\mathsf {X}} p \to \Box _{\mathsf {X}} p) \mid \mathsf {X} \in $
${}\qquad \{\mathsf {F}, \mathsf {P}\}\big \}$
,
where
$\mathsf {K4}_2$
is the bimodal version of
$\mathsf {K4}$
(with
$\Diamond _{\mathsf {F}}$
and
$\Diamond _{\mathsf {P}}$
). None of these five logics (and any other temporal logic with frames of unbounded depth) has the CIP [Reference Chagrov and Zakharyaschev9, Reference Gabbay and Maksimova14], and our aim in this section is to prove that the IEP for each of them is decidable in coNP. The following example illustrates the new semantic phenomena of temporal logics compared to modal logics containing
$\mathsf {K4.3}$
that we need to address in order to achieve this aim.
Example 5.2.
$(a)$
Consider the formulas
$\varphi _1$
and
$\varphi _2$
from Example 3.6
$(a)$
in the context of
$\mathsf {Lin}_{<\omega }$
in place of
$\mathsf {GL.3}$
, reading
$\Diamond $
as
$\Diamond _{\mathsf {F}}$
and
$\Box $
as
$\Box _{\mathsf {F}}$
:
$$ \begin{align*} & \varphi_1 = \Diamond_{\mathsf{F}}( p_{1} \wedge \Diamond_{\mathsf{F}}^+ \neg q_{1}) \wedge \Box_{\mathsf{F}} (p_{2} \rightarrow \Box_{\mathsf{F}}^+ q_{1}) \wedge \Box_{\mathsf{F}}(p_{1} \rightarrow \neg p_{2}),\\ & \varphi_2 = \neg [ \Diamond_{\mathsf{F}}( p_{2} \wedge \Diamond_{\mathsf{F}}^+\neg q_{2}) \land \Box_{\mathsf{F}} (p_{1} \rightarrow \Box_{\mathsf{F}}^+ q_{2}) ]. \end{align*} $$
We clearly have
$(\varphi _1 \rightarrow \varphi _2) \in \mathsf {Lin}_{<\omega }$
. Using Theorem 3.2, we show that
$\varphi _1$
and
$\varphi _2$
have no interpolant in
$\mathsf {Lin}_{<\omega }$
. The argument from Example 3.6
$(a)$
shows that any models
$\mathfrak M_i$
meeting the criterion of Theorem 3.2 cannot be based on a Kripke frame for
$\mathsf {Lin}_{<\omega }$
. However, the descriptive frame
we employed for
$\mathsf {GL.3}$
in Example 3.6
$(a)$
does not help now, because it refutes
$\Box _{\mathsf {P}} (\Box _{\mathsf {P}} p \to p) \to \Box _{\mathsf {P}} p$
at any point save the first two under the valuation below:

To fix this issue, we modify
by making it symmetric in both directions. Consider the frame
$\mathfrak F_k = (W_k',R_{\bullet k\bullet },\mathcal {P}_k')$
,
$k> 0$
, in which the points in
are ordered as shown in the picture below

or, more formally,
$xR_{\bullet k\bullet } y$
iff (
$x=b_n^{L}$
,
$y=b_m^{L}$
for
$n<m$
), or (
$x=b_n^{L}$
,
$y=a_{i}$
), or (
$x=b_n^{L}$
,
$y=b_m^{R}$
), or (
$x=a_{i}$
,
$y=a_{j}$
) or
$(x=a_{i}$
,
$y= b_n^{R}$
), or (
$x = b_n^{R}$
,
$y = b_m^{R}$
, for
$n>m$
). The internal sets in
$\mathcal {P}_k$
are generated by
Observe that
$\{b_n^{L}\},\{b_n^{R}\}\in \mathcal {P}_k'$
, for all
$n<\omega $
. It is not hard to see that
$\mathfrak F_k$
is a descriptive frame; we denote it by
. As an exercise, the reader can check that, for any natural numbers
$k, l, \dots , m, n> 0$
,
where
is the mirror image of
(see also Lemma 5.6).
The picture below shows models
$\mathfrak M_1$
and
$\mathfrak M_2$
based on
and satisfying the conditions of Theorem 3.2 for
$\varphi _1$
and
$\varphi _2$
:

By (31),
, so
$\varphi _1$
and
$\varphi _2$
do not have an interpolant in
$\mathsf {Lin}_{<\omega }$
.
$(b)$
Consider next the temporal version of the implication
$\varphi _1'\rightarrow \varphi _2$
from Example 3.6
$(b)$
, which is clearly valid in
$\mathsf {Lin}_{\mathbb {Z}}$
. To demonstrate that
$\varphi _1'$
and
$\varphi _2$
have no interpolant in
$\mathsf {Lin}_{\mathbb {Z}}$
, we can use
, which is a frame for
$\mathsf {Lin}_{\mathbb {Z}}$
by (32). The models
$\mathfrak M_1$
and
$\mathfrak M_2$
depicted below

satisfy the conditions of Theorem 3.2 for
$\varphi _1'$
and
$\varphi _2$
.
As illustrated by Example 5.2, the temporal frames
$\mathfrak F = (W,R,\mathcal {P})$
we need for checking the criterion of Theorem 3.2 may contain both infinite descending and ascending chains of clusters (and so the
$\mathfrak F_c^{-1}$
are not necessarily isomorphic to ordinals). Accordingly, we now have R-final and
$R^-$
-final clusters as well as two types of limit clusters: an R-limit cluster is a non-
$R^-$
-final cluster without an immediate
$R^-$
-successor and an
$R^-$
-limit cluster is a non-R-final cluster without an immediate R-successor. Some clusters can be both R- and
$R^-$
-limit clusters.
We say that a set
$S \ne \emptyset $
of clusters in
$\mathfrak F$
is R-unbounded (
$R^-$
-unbounded) if there is no
$C \in S$
such that
$C' \le _R C$
(respectively,
$C \le _R C'$
), for all
$C' \in S$
. A cluster C is the R-limit of an R-unbounded set S if
$C' <_R C$
for all
$C' \in S$
and there is no cluster
$C"$
with
$C' <_R C"<_R C$
for all
$C' \in S$
; the
$R^-$
-limit of an
$R^-$
-unbounded set S is defined symmetrically by replacing R with
$R^-$
. It is straightforward to see that each R-limit cluster C is the R-limit of the R-unbounded set
$\{C'\mid C'<_R C\}$
, and each
$R^-$
-limit cluster D is the
$R^-$
-limit of the
$R^-$
-unbounded set
$\{D'\mid D<_R D'\}$
. For any cluster C, we let
$(C,+\infty )=\{x\mid C<_R C(x)\}$
and
$(-\infty ,C)=\{x\mid C(x)<_R C\}$
.
Lemma 5.3. Suppose
$\mathfrak F=(W,R,\mathcal {P})$
is a temporal n-generated descriptive frame, for some
$n<\omega $
. Then
-
(a) every cluster in
$\mathfrak F$
has at most
$2^n$
points; -
(b) every R-unbounded (
$R^-$
-unbounded) set of clusters in
$\mathfrak F$
has an R-limit (
$R^-$
-limit) in
$\mathfrak F$
, and so
$\mathfrak F$
contains both R- and
$R^-$
-final clusters.
Proof.
$(a)$
is proved similarly to Lemma 2.4
$(b)$
.
$(b)$
Suppose
$\mathfrak F$
is
$\mathfrak M$
-generated, for some model
$\mathfrak M$
. Let S be an R-unbounded set of clusters in
$\mathfrak F$
with
$y_C\in C$
,
$C\in S$
, and let
Clearly,
$\Gamma $
is finitely satisfiable in
$\mathfrak M$
, and so by (com) and Lemma 5.1, there is a
$\Gamma $
-minimal point x in
$\mathfrak M$
. By (tig),
$y_C R x$
for all
$C\in S$
. Now suppose that y is such that
$y_CRy$
, for all
$C\in S$
, and
$yRx$
. Then
$\Gamma \subseteq t_{\mathfrak M}(y)$
, and so
$xRy$
by the
$\Gamma $
-minimality of x. Thus,
$C(x)$
is the R-limit of S. The existence of
$R^-$
-limits of
$R^-$
-unbounded S is symmetric.
A cluster C is called minimal (maximal) in a temporal model
$\mathfrak {M}$
if there is a formula
$\mu $
such that
$C\cap \min _{\mathfrak M}\{\mu \}\ne \emptyset $
(
$C\cap \max _{\mathfrak M}\{\mu \}\ne \emptyset $
). If there is such a
$\sigma $
-formula
$\mu $
, for some signature
$\sigma $
, we call
$C \sigma $
-minimal (
$\sigma $
-maximal) in
$\mathfrak {M}$
.
Lemma 5.4. Suppose
$\mathfrak {M}$
is a model based on a finitely
$\mathfrak M$
-generated temporal descriptive frame
$\mathfrak F$
. Then
-
(a) every degenerate cluster in
$\mathfrak F$
is both maximal and minimal in
$\mathfrak {M}$
; -
(b) a cluster is maximal (minimal) in
$\mathfrak {M}$
iff either it is R-final (respectively,
$R^-$
-final) or has an immediate R-successor (respectively,
$R^-$
-successor); -
(c) a cluster is definable in
$\mathfrak {M}$
iff it is both maximal and minimal in
$\mathfrak {M}$
.
It follows that the R- and
$R^-$
-limit clusters are not definable and not degenerate; all other clusters are definable in
$\mathfrak M$
. We also have that
-
(d) for any clusters
$C <_R C'$
in
$\mathfrak F$
, the interval
$[C, C']$
contains a maximal cluster and also a minimal one; -
(e) if C is not an R-limit cluster and
$C'$
is not an
$R^-$
-limit cluster, then the closed interval
$[C,C']$
is definable in
$\mathfrak M$
.
Proof. Items
$(a)$
–
$(c)$
are proved in the same way as Lemma 2.6. Item
$(d)$
follows from (tig), which gives formulas
$\varphi $
and
$\psi $
with
$\mathfrak M,x \not \models \Box _{\mathsf {F}}\varphi $
,
$\mathfrak M,y \models \Box _{\mathsf {F}}\varphi $
and
$\mathfrak M,x \models \Box _{\mathsf {P}}\psi $
,
$\mathfrak M,y \not \models \Box _{\mathsf {P}}\psi $
, and so
$[C(x), C(y)]$
contains a
$\Box _{\mathsf {F}}\varphi $
-minimal cluster and a
$\Box _{\mathsf {P}}\psi $
-maximal one. Item
$(e)$
: by
$(b)$
, C is
$\lambda $
-minimal and
$C'$
is
$\mu $
-maximal for some
$\lambda ,\mu $
. Then
$[C,C']$
is defined in
$\mathfrak M$
by
$\Diamond _{\mathsf {P}}^+\lambda \land \Diamond _{\mathsf {F}}^+\mu $
.
The following temporal analogue is harder to prove than Lemma 2.7.
Lemma 5.5. If
$\mathfrak F = (W,R,\mathcal {P})$
is a finitely generated temporal descriptive frame, then W is countable.
Proof. By Lemma 5.3
$(a)$
, it suffices to show that
$\mathfrak F_c = (W_c, <_R)$
is countable. Suppose
$\mathfrak F$
is
$\mathfrak {M}$
-generated, for some
$\delta $
-model
$\mathfrak M=(\mathfrak F,\mathfrak v)$
and finite signature
$\delta $
. First, observe that, by Lemma 5.4
$(b)$
, each non-R-limit cluster C is
$\mu _C$
-minimal in
$\mathfrak M$
for some
$\mu _C$
. Thus, the internal set
$X_C=\mathfrak v(\Diamond _{\mathsf {P}}^+\mu _C)$
distinguishes C from every D with
$D<_R C$
, and so
$X_C\ne X_D$
whenever
$C\ne D$
. As
$\mathcal {P}$
is countable, the number of non-R-limit clusters in
$\mathfrak F_c$
is countable. Similarly, there are countably-many non-
$R^-$
-limit clusters in
$\mathfrak F_c$
. So it is enough to show that the number of clusters in
$\mathfrak F_c$
that are both R- and
$R^-$
-limits is countable. We refer to such clusters as simply limit clusters. Call an interval
$[C^-,C^+]$
a neighbourhood of a limit cluster C if
$C^-<_{R}C <_R C^+$
. By Lemma 5.4, every limit cluster C has a nice neighbourhood
$N_C=[C^-,C^+]$
with non-limit clusters
$C^-$
and
$C^+$
. As the number of different nice
$N_C$
is countable, it follows that
$$ \begin{align} \text{every uncountable interval } & [D,D'] \text{ contains a limit cluster } C\\ & \text{all of whose neighbourhoods are uncountable}\nonumber \end{align} $$
(otherwise all limit clusters in
$[D,D']$
would belong to the countable union of the countable intervals
$N_C$
, and so
$[D,D']$
were countable).
By an atomic type we mean any
$\textit {at}_{\mathfrak M}^\delta (x)$
with
$x \in W$
. For any cluster C, we set
$\textit {at}(C)=\{\textit {at}_{\mathfrak M}^\delta (x)\mid x\in C\}$
. Let C be an R-limit cluster. We say that an atomic type
$\textit {a}$
occurs infinitely R-close to C if, for every
$C'<_R C$
, there is
$C"$
such that
$C'<_R C"<_R C$
and
$\textit {a}\in \textit {at}(C")$
. Similarly,
$\textit {a}$
occurs infinitely
$R^-$
-close to an
$R^-$
-limit cluster C if whenever
$C<_R C'$
, then there is
$C"$
such that
$C<_R C"<_R C'$
and
$\textit {a}\in \textit {at}(C")$
. We claim that
Indeed, let S be an R-unbounded set of clusters with R-limit C and
$y_D\in D$
,
$D\in S$
, and let
If
$\textit {a}$
occurs infinitely R-close to C, it can be shown similarly to the proof of Lemma 5.3
$(b)$
that there is a
$\Gamma _{\textit {a}}$
-minimal point
$x\in C$
, so
$\textit {a}=\textit {at}_{\mathfrak M}^\delta (x)\in \textit {at}(C)$
.
The converse of (34) also holds:
Indeed, suppose there is
$C'<_R C$
with
$\textit {a}\notin \textit {at}(C")$
, for any
$C"$
in the interval
$C'<_R C"<_R C$
. By Lemma 5.4
$(d)$
, there is a cluster
$C"$
in
$[C',C]$
that is
$\mu $
-minimal in
$\mathfrak M$
for some formula
$\mu $
. But then C is
$\Diamond _{\mathsf {P}}\mu \land \bigwedge \textit {a}$
-minimal, contrary to Lemma 5.4
$(b)$
. Symmetric variants of (34) and (35) hold for
$R^-$
-limit clusters.
Call non-degenerate clusters
$C' <_R C"$
twins if
$\textit {at}(C')=\textit {at}(C")$
and, for every C in
$[C',C"]$
, we have
$\textit {at}(C)\subseteq \textit {at}(C')=\textit {at}(C")$
. We claim that
Indeed, suppose
$C'$
,
$C"$
are twins. By induction on the construction of a
$\delta $
-formula
$\alpha $
, we see that if
$x,y\in [C',C"]$
with
$xRy$
and
$\textit {at}_{\mathfrak M}^\delta (x) = \textit {at}_{\mathfrak M}^\delta (y)$
, then
$\mathfrak M,x \models \alpha $
iff
$\mathfrak M,y \models \alpha $
. We only consider one of the nontrivial cases. Let
$\mathfrak M,x \models \Diamond _{\mathsf {F}}\alpha $
. Then there is z with
$xRz$
and
$\mathfrak M,z \models \alpha $
. If
$yRz$
, then clearly
$\mathfrak M,y \models \Diamond _{\mathsf {F}}\alpha $
. Otherwise,
$z\in [C',C"]$
, so
$\textit {at}_{\mathfrak M}^\delta (z) = \textit {at}_{\mathfrak M}^\delta (z')$
, for some
$z' \in C"$
. Thus, by IH,
$\mathfrak M,z' \models \alpha $
, which implies
$\mathfrak M,y \models \Diamond _{\mathsf {F}}\alpha $
as
$C"$
is non-degenerate. It follows that there are
$x \in C'$
and
$y \in C"$
with
$t_{\mathfrak M}(x) = t_{\mathfrak M}(y)$
, contrary to (dif).
We can now prove that
$\mathfrak F_c$
is countable. Suppose
$\mathfrak F_c$
is uncountable. By (33) and Lemma 5.3
$(b)$
,
$\mathfrak F_c$
contains a limit cluster C whose neighbourhoods are all uncountable. Let C be such a cluster with a minimal
$\textit {at}(C)$
. As
$\delta $
is finite, C has a neighbourhood N such that, for any
$D\in N$
with
$D<_R C$
, every
$\textit {a}\in \textit {at}(D)$
occurs infinitely R-close to C, and, for any
$D\in N$
with
$C<_R D$
, every
$\textit {a}\in \textit {at}(D)$
occurs infinitely
$R^-$
-close to C. We call such N a close proximity of C. As N is uncountable, either
$[C^-,C)$
or
$(C,C^+]$
is uncountable. We only consider the former case, as the latter is similar. We claim that
Indeed, take such
$C'$
. As
$[C^-,C']$
is contained in the close proximity N, for every limit cluster D in
$[C^-,C']$
, we have
$\textit {at}(D)\subsetneq \textit {at}(C)$
, by (34) and (36). So by the
$\textit {at}(C)$
-minimality of C among limit clusters with only uncountable neighbourhoods, every limit cluster D in
$[C^-,C']$
has a countable neighbourhood. Thus,
$[C^-,C']$
is countable by (33).
By (35), there is a countably infinite ascending chain
$C_1 <_R C_2 <_R \dots $
of clusters in
$[C^-,C)$
such that, for every
$\textit {a}\in \textit {at}(C)$
and every
$n<\omega $
, there is m with
$n<m<\omega $
and
$\textit {a}\in \textit {at}(C_m)$
. Let
$C'$
be the R-limit of the R-unbounded set
$\{C_n\mid n<\omega \}$
(which exists by Lemma 5.3
$(b)$
). Then
$C'\leq _R C$
. Also, every
$\textit {a}\in \textit {at}(C)$
occurs infinitely R-close to
$C'$
, and so
$\textit {at}(C)\subseteq \textit {at}(C')$
by (34). We cannot have
$C'<_R C$
since otherwise (as
$C'$
belongs to the close proximity N of C) every
$\textit {a}\in \textit {at}(C')$
occurred infinitely R-close to C, resulting in
$\textit {at}(C)=\textit {at}(C')$
by (34), and so
$C'$
and C were twins, contrary to (36). It follows that
$C'=C$
, and so
$[C^-,C)=\bigcup _{n<\omega } [C^-,C_n]$
. As each
$[C^-,C_n]$
is countable by (37),
$[C^-,C)$
is also countable, which is a contradiction.
Using Lemmas 5.3 and 5.4, we can also obtain elegant characterisations of descriptive frames for
$\mathsf {Lin}_{\mathbb {Q}}$
,
$\mathsf {Lin}_{\mathbb {R}}$
,
$\mathsf {Lin}_{<\omega }$
, and
$\mathsf {Lin}_{\mathbb {Z}}$
(cf. [Reference Burgess, Gabbay and Guenthner8, Reference Goldblatt19, Reference Wolter37, Reference Wolter, Zakharyaschev, Blackburn, Benthem and Wolter40]).
Lemma 5.6. Let
$\mathfrak F = (W,R,\mathcal {P})$
be any finitely generated temporal descriptive frame. Then
-
Lin ℚ :
$\mathfrak F \models \mathsf {Lin}_{\mathbb {Q}}$
iff
$\mathfrak F$
is serial in both directions—i.e., the R- and
$R^-$
-final clusters in
$\mathfrak F$
are both non-degenerate, and
$\mathfrak F$
is dense—i.e., there is a non-degenerate cluster between any two distinct degenerate ones; -
Lin ℝ :
$\mathfrak F \models \mathsf {Lin}_{\mathbb {R}}$
iff
$\mathfrak F$
is serial, dense, and Dedekind-complete in the sense that there is a degenerate cluster between any two distinct non-degenerate ones; -
Lin <ω :
$\mathfrak F \models \mathsf {Lin}_{<\omega }$
iff
$\mathfrak F$
does not contain a non-degenerate cluster C such that
$(-\infty ,C) \in \mathcal {P}$
or
$(C, +\infty ) \in \mathcal {P} ($
in particular, the R- and
$R^-$
-final clusters in
$\mathfrak F$
are degenerate
$)$
; -
Lin ℤ :
$\mathfrak F \models \mathsf {Lin}_{\mathbb {Z}}$
iff
$\mathfrak F$
is serial and does not contain a non-degenerate cluster C with
$\emptyset \ne (-\infty ,C) \in \mathcal {P}$
or
$\emptyset \ne (C, +\infty ) \in \mathcal {P} ($
a single non-degenerate cluster is a frame for
$\mathsf {Lin}_{\mathbb {Z}}$
but not for
$\mathsf {Lin}_{<\omega })$
.
Proof. We only show the
$(\Rightarrow )$
-directions, leaving the converses to the reader. Suppose
$\mathfrak F$
is
$\mathfrak M$
-generated, for some model
$\mathfrak M=(\mathfrak F,\mathfrak v)$
.
$\mathsf {Lin}_{\mathbb {Q}}$
: As
$\mathfrak F\models \Diamond _{\mathsf {F}} \top $
(
$\mathfrak F\models \Diamond _{\mathsf {P}}\top $
), Lemma 5.1 gives a
$\{\Diamond _{\mathsf {F}}\top \}$
-maximal (
$\{\Diamond _{\mathsf {P}}\top \}$
-minimal) point x in
$\mathfrak M$
with R-final (
$R^-$
-final) and non-degenerate
$C(x)$
. Thus,
$\mathfrak F$
is serial. Suppose
$\{x\}$
,
$\{y\}$
are degenerate clusters with
$xRy$
. Lemma 5.4 gives formulas
$\psi _x$
and
$\psi _y$
defining
$\{x\}$
and
$\{y\}$
in
$\mathfrak M$
. As
$\mathfrak M,x\models \Diamond _{\mathsf {F}}\psi _y$
and
$\mathfrak F\models \Diamond _{\mathsf {F}}\psi _y\to \Diamond _{\mathsf {F}}\Diamond _{\mathsf {F}}\psi _y$
, the formula
$\Diamond _{\mathsf {F}}\psi _y\land \Diamond _{\mathsf {P}}\psi _x$
is satisfiable in
$\mathfrak M$
. Let z be
$\{\Diamond _{\mathsf {F}}\psi _y\land \Diamond _{\mathsf {P}}\psi _x\}$
-maximal in
$\mathfrak M$
. Then
$xRzRy$
. As
$\mathfrak M,z\models \Diamond _{\mathsf {F}}(\Diamond _{\mathsf {F}}\psi _y\land \Diamond _{\mathsf {P}}\psi _x)$
by
$\mathfrak F\models \Diamond _{\mathsf {F}}\psi _y\to \Diamond _{\mathsf {F}}\Diamond _{\mathsf {F}}\psi _y$
, the cluster
$C(z)$
is non-degenerate.
$\mathsf {Lin}_{\mathbb {R}}$
: Non-degenerate
$C(x) <_R C(y)$
cannot be
$<_R$
-consecutive because otherwise, by Lemma 5.4,
$C(x)$
were
$\psi $
-maximal in
$\mathfrak M$
for some formula
$\psi $
, and so
$\mathfrak M,x\not \models \Box (\Box _{\mathsf {P}}\Diamond _{\mathsf {F}}\psi \to \Diamond _{\mathsf {F}}\Box _{\mathsf {P}} \Diamond _{\mathsf {F}}\psi ) \to (\Box _{\mathsf {P}} \Diamond _{\mathsf {F}}\psi \to \Box _{\mathsf {F}} \Diamond _{\mathsf {F}}\psi )$
, contrary to
$\mathfrak F\models \mathsf {Lin}_{\mathbb {R}}$
. Thus, there is z with
$C(x) <_R C(z) <_R C(y)$
. If z is irreflexive, we are done. Otherwise, by (tig), there is some formula
$\chi $
with
$\Box _{\mathsf {F}}\chi \in t_{\mathfrak M}(y)$
and
$\chi \notin t_{\mathfrak M}(z)$
, and so
$\mathfrak M,z\models \Diamond _{\mathsf {F}}\neg \chi $
. Let
$z'$
be a
$\Diamond _{\mathsf {F}}\neg \chi $
-maximal point in
$\mathfrak M$
. Clearly,
$C(x) <_R C(z') <_R C(y)$
. If
$z'$
is irreflexive, we are done. Otherwise, we take the immediate R-successor
$z"$
of
$z'$
, which exists by Lemma 5.4. As
$\mathfrak M,z"\models \Box _{\mathsf {F}}\chi \land \neg \chi $
, point
$z"$
is irreflexive and
$C(z")<_R C(y)$
.
$\mathsf {Lin}_{<\omega }$
: If there existed a non-degenerate cluster
$C(x)$
and a formula
$\psi $
with
$(-\infty ,C(x))=\mathfrak v(\psi )$
, then
$\mathfrak M,x\not \models \Box _{\mathsf {P}} (\Box _{\mathsf {P}}\psi \to \psi ) \to \Box _{\mathsf {P}} \psi $
, contrary to
$\mathfrak F\models \mathsf {Lin}_{<\omega }$
.
$\mathsf {Lin}_{\mathbb {Z}}$
: If there existed a non-degenerate cluster
$C(x)$
and some formula
$\psi $
with
$\emptyset \ne (C(x),+\infty )=\mathfrak v(\psi )$
, then
$\mathfrak M,x\not \models \Box _{\mathsf {F}}(\Box _{\mathsf {F}} \psi \to \psi ) \to (\Diamond _{\mathsf {F}}\Box _{\mathsf {F}} \psi \to \Box _{\mathsf {F}} \psi )$
, contrary to
$\mathfrak F\models \mathsf {Lin}_{\mathbb {Z}}$
.
Note that
$\mathsf {Lin}$
and
$\mathsf {Lin}_{\mathbb {Q}}$
are d-persistent while the other three logics are not [Reference Wolter37].
Example 5.7. The descriptive frame
$\mathfrak F = (W_2,R_{\circ \bullet },\mathcal {P}_2)$
with
$(W_2,R_{\circ \bullet })$
depicted below and
$\mathcal {P}_2$
defined in Example 2.2 is serial, dense, and Dedekind-complete, so
$\mathfrak F \models \mathsf {Lin}_{\mathbb {R}}$
.
It is readily seen, however, that
$(W_2,R_{\circ \bullet })\not \models \mathsf {Lin}_{\mathbb {R}}$
, so
$\mathsf {Lin}_{\mathbb {R}}$
is not d-persistent.
The notion of
$\sigma $
-block from Section 4.2 also needs a modification for temporal models. Namely, a set
${\boldsymbol {b}} \subseteq W$
is a
$\sigma $
-block in a temporal model
$\mathfrak M$
based on
$\mathfrak F = (W,R,\mathcal {P})$
if
${\boldsymbol {b}} = {\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(x)$
, for some
$x \in W$
, where
if both
$\Diamond _{\mathsf {F}} t_{\mathfrak {M}}^{\sigma }(x) \subseteq t_{\mathfrak {M}}^{\sigma }(x)$
and
$\Diamond _{\mathsf {P}} t_{\mathfrak {M}}^{\sigma }(x) \subseteq t_{\mathfrak {M}}^{\sigma }(x)$
hold; otherwise
${\boldsymbol {b}}_{\mathfrak {M}}^{\sigma }(x) = \{x\}$
. Then we have the following temporal analogue of Lemma 4.13.
Lemma 5.8. Suppose
$\mathfrak {M}$
is a model based on a finitely
$\mathfrak M$
-generated temporal descriptive frame
$\mathfrak F = (W, R,\mathcal {P})$
. Then, for any
$\sigma $
-block
${\boldsymbol {b}}$
in
$\mathfrak {M}$
, there exist clusters
$C_{{\boldsymbol {b}}}^-$
and
$C_{{\boldsymbol {b}}}^+$
in
$\mathfrak F$
such that the following hold:
-
(a)
${\boldsymbol {b}} = \big [C_{{\boldsymbol {b}}}^-, C_{{\boldsymbol {b}}}^+\big ]$
; -
(b) if cluster
$C_{{\boldsymbol {b}}}^-$
(cluster
$C_{{\boldsymbol {b}}}^+$
) is minimal (respectively, maximal) in
$\mathfrak M$
, then it is
$\sigma $
-minimal (respectively,
$\sigma $
-maximal) in
$\mathfrak M$
; -
(c) if
${\boldsymbol {b}}$
is non-degenerate, then both
$C_{{\boldsymbol {b}}}^-$
and
$C_{{\boldsymbol {b}}}^+$
are non-degenerate; -
(d)
${\boldsymbol {b}}$
is definable in
$\mathfrak M$
iff
$C_{{\boldsymbol {b}}}^-$
is not an R-limit cluster and
$C_{{\boldsymbol {b}}}^+$
is not an
$R^-$
-limit cluster; -
(e)
$t_{\mathfrak {M}}^{\sigma }({\boldsymbol {b}}) = t_{\mathfrak {M}}^{\sigma }\big (C_{{\boldsymbol {b}}}^-) = t_{\mathfrak {M}}^{\sigma }\big (C_{{\boldsymbol {b}}}^+\big )$
.
Proof. This can be proved similarly to Lemma 4.13, using Lemmas 5.3, 5.4, and 5.1, in place of Lemmas 2.4, 2.6, and 2.3, respectively.
Given
$\sigma $
-bisimilar models
$\mathfrak M_i$
,
$i=1,2$
, based on finitely
$\mathfrak M_i$
-generated temporal frames, we can adapt Lemma 4.15 to the temporal setting to show that
$\sigma $
-blocks in
$\mathfrak M_1$
and
$\mathfrak M_2$
always come in
$\sigma $
-bisimilar pairs
${\boldsymbol {b}}$
,
$\boldsymbol {\beta }({\boldsymbol {b}})$
. Being equipped with these modifications, we show first how to extend the selection procedure from the proof of Theorem 3.5 to
$\mathsf {Lin}$
,
$\mathsf {Lin}_{\mathbb {Q}}$
, and
$\mathsf {Lin}_{\mathbb {R}}$
.
Theorem 5.9. Each
$L \in \{\mathsf {Lin}, \mathsf {Lin}_{\mathbb {Q}}, \mathsf {Lin}_{\mathbb {R}}\}$
has the polysize bisimilar model property, and the IEP for L is coNP-complete.
Proof. Suppose
$\varphi _1$
and
$\varphi _2$
have no interpolant in L,
$\sigma = \textit {sig}(\varphi _1) \cap \textit {sig}(\varphi _2)$
, and
$\delta = \textit {sig}(\varphi _1) \cup \textit {sig}(\varphi _2)$
. By Theorem 3.2, there are
$\delta $
-models
$\mathfrak M_i$
, for
$i=1,2$
, based on
$\mathfrak M_i$
-generated temporal descriptive frames
$\mathfrak {F}_i = (W_{i},R_{i},\mathcal {P}_{i})$
for L with
$\mathfrak {M}_{1},x_{1} \sim _{\sigma } \mathfrak {M}_{2},x_{2}$
,
$\mathfrak {M}_{1},x_{1} \models \varphi _1$
and
$\mathfrak {M}_{2}, x_{2} \models \neg \varphi _2$
. Let
$\boldsymbol {\beta }$
be the largest
$\sigma $
-bisimulation between
$\mathfrak M_1$
and
$\mathfrak M_2$
, that is,
$y_1\boldsymbol {\beta } y_2$
iff
$t_{\mathfrak {M}_1}^{\sigma }(y_1)=t_{\mathfrak {M}_2}^{\sigma }(y_2)$
, for all
$y_i\in W_i$
. We show that there exist such
$\mathfrak M_i$
of polynomial size in
$\max (|\varphi _1|,|\varphi _2|)$
.
For any
$i=1,2$
and
$\tau \in \textit {sub}(\varphi _i)$
satisfied in
$\mathfrak {M}_i$
, we take one
$\{\tau \}$
-maximal and one
$\{\tau \}$
-minimal point in
$\mathfrak M_i$
. Let
$\boldsymbol {M}_{i}$
be the set of all selected points and let
For each
$t\in T$
, we take a smallest set
$\boldsymbol {S}_{i} \subseteq W_i$
containing one t-maximal and one t-minimal point in
$\mathfrak M_i$
.
Let
$W_{i}'=\{x_{i}\}\cup \boldsymbol {M}_{i} \cup \boldsymbol {S}_{i}$
,
$R^{\prime }_i = {R_i}\mathop {\restriction }_{W^{\prime }_i}$
,
$\mathfrak F^{\prime }_i = (W^{\prime }_i,R^{\prime }_i)$
, let
$\mathfrak {M}_{i}'$
be the restriction of
$\mathfrak {M}_{i}$
to
$\mathfrak {F}_{i}'$
, and let
$x_1'\boldsymbol {\beta }' x_2'$
iff
$t_{\mathfrak {M}_1}^{\sigma }(x_1')=t_{\mathfrak {M}_2}^{\sigma }(x_2')$
, for all
$x_1'\in W_1'$
,
$x_2'\in W_2'$
. Following the proof of Lemma 3.4, we see that
$\mathfrak {M}_{1}',x_{1}\models \varphi _1$
,
$\mathfrak {M}_{2}',x_{2}\models \neg \varphi _2$
, and
$\boldsymbol {\beta }'$
is a
$\sigma $
-bisimulation between
$\mathfrak {M}_{1}'$
and
$\mathfrak {M}_{2}'$
with
$x_1\boldsymbol {\beta }' x_2$
. Clearly,
$\mathfrak F_i' \models \mathsf {Lin}$
and the
$\mathfrak M_i$
are of polynomial size in
$\max (|\varphi _1|,|\varphi _2|)$
.
For
$L = \mathsf {Lin}_{\mathbb {Q}}$
, we do not necessarily have
$\mathfrak {F}_{i}' \models L$
. To fix this, we add some extra points from
$W_i$
to
$W_i'$
. As
$\mathfrak F_i\models \mathsf {Lin}_{\mathbb {Q}}$
, the R- and
$R^-$
-final clusters in
$\mathfrak F_i$
are non-degenerate and, as observed in the selection procedure from Section 3,
$W_i'$
contains some points from these final clusters. Thus,
$\mathfrak {F}_{i}' \not \models \mathsf {Lin}_{\mathbb {Q}}$
iff
$\mathfrak {F}_{i}'$
contains an irreflexive point x with an immediate irreflexive
$R^{\prime }_i$
-successor y. We call such pair
$x,y$
an irr-defect in
$\mathfrak {F}_{i}'$
. We are going to ‘cure’ one irr-defect after the other without introducing new irr-defects in either frame.
Given an irr-defect
$u_1,v_1$
in
$\mathfrak {F}_{1}'$
, we find an
$R_1$
-reflexive
$z_1$
with
$u_1 R_1 z_1 R_1 v_1$
, which exists by
$\mathfrak F_1\models \mathsf {Lin}_{\mathbb {Q}}$
and Lemma 5.6. Let
$t=t_{\mathfrak {M}_{1}}^{\sigma }(z_1)$
and
${\boldsymbol {b}}={\boldsymbol {b}}_{\mathfrak {M}_{1}}^{\sigma }(z_1)$
. As
$\Diamond _{F}t\subseteq t$
and
$\Diamond _{P}t\subseteq t$
,
${\boldsymbol {b}}$
is a non-degenerate
$\sigma $
-block in
$\mathfrak M_1$
. By Lemma 5.8, there are t-minimal and t-maximal points
$z_1^-$
and
$z_1^+$
in the non-degenerate clusters
$C_{{\boldsymbol {b}}}^-$
and
$C_{{\boldsymbol {b}}}^+$
. As
$\boldsymbol {\beta }({\boldsymbol {b}})$
is a non-degenerate
$\sigma $
-block in
$\mathfrak M_2$
by Lemma 4.15, there are t-minimal and t-maximal points
$z_2^-$
and
$z_2^+$
in the non-degenerate clusters
$C_{\boldsymbol {\beta }({\boldsymbol {b}})}^-$
and
$C_{\boldsymbol {\beta }({\boldsymbol {b}})}^+$
. By adding
$z_1$
,
$z_1^-$
,
$z_1^+$
to
$W_1'$
and
$z_2^-$
,
$z_2^+$
to
$W_2'$
we cure the irr-defect
$u_1,v_1$
without creating a new irr-defect in either frame. Let
$W_i"$
,
$i=1,2$
, be the sets we obtain after curing all irr-defects in both frames in this way,
$R^{\prime \prime }_i = {R_i}\mathop {\restriction }_{W^{\prime \prime }_i}$
,
$\mathfrak F^{\prime \prime }_i = (W^{\prime \prime }_i,R^{\prime \prime }_i)$
, let
$\mathfrak {M}_{i}"$
be the restriction of
$\mathfrak {M}_{i}$
to
$\mathfrak {F}_{i}"$
, and
$x_1'\boldsymbol {\beta }" x_2'$
iff
$t_{\mathfrak {M}_1}^{\sigma }(x_1')=t_{\mathfrak {M}_2}^{\sigma }(x_2')$
, for all
$x_1'\in W_1"$
,
$x_2'\in W_2"$
. Then
$\mathfrak F_i"\models \mathsf {Lin}_{\mathbb {Q}}$
, by Lemma 5.6, and
(minmax) for all
$x\in W_1"\cup W_2"$
and
$i=1,2$
, the set
$W_1"$
contains
$t_{\mathfrak M_i}^\sigma (x)$
-minimal and
$t_{\mathfrak M_i}^\sigma (x)$
-maximal points in
$\mathfrak M_1$
, and
$W_2"$
contains
$t_{\mathfrak M_i}^\sigma (x)$
-minimal and
$t_{\mathfrak M_i}^\sigma (x)$
-maximal points in
$\mathfrak M_2$
.
So it is readily seen (similarly to the proof of Lemma 3.4) that
$\mathfrak {M}_{1}",x_{1}\models \varphi _1$
,
$\mathfrak {M}_{2}",x_{2}\models \neg \varphi _2$
, and
$\boldsymbol {\beta }"$
is a
$\sigma $
-bisimulation between
$\mathfrak {M}_{1}"$
and
$\mathfrak {M}_{2}"$
with
$x_1\boldsymbol {\beta }" x_2$
.
Finally, let
$L = \mathsf {Lin}_{\mathbb {R}}$
. Since
$\mathsf {Lin}_{\mathbb {Q}} \subseteq \mathsf {Lin}_{\mathbb {R}}$
, we first cure the irr-defects in the
$\mathfrak F_i'$
,
$i=1,2$
, as described above. Let
$\mathfrak F_i"$
,
$i=1,2$
, be the resulting serial and dense frames. Thus,
$\mathfrak F_i" \not \models L$
iff
$\mathfrak {F}_{i}"$
contains two
$<_{R^{\prime \prime }_i}$
-consecutive non-degenerate clusters
$C(x) \ne C(y)$
. We call such
$x,y$
a ref-defect in
$\mathfrak {F}_{i}"$
. We show that the ref-defects can also be cured in a step-by-step manner without introducing new defects of either type, while maintaining (minmax).
If
$u_1,v_1$
is a ref-defect in
$\mathfrak F^{\prime \prime }_1$
, Lemma 5.6 provides an irreflexive
$z_1 \in W_1$
with
$u_1 R_1 z_1 R_1 v_1$
. Let
$t=t_{\mathfrak M_1}^\sigma (z_1)$
. The insertion of extra points into
$W_1"$
depends on whether
$u_1$
and
$v_1$
are in the same
$\sigma $
-block in
$\mathfrak M_1$
or not.
Case 1:
$u_1,v_1 \in {\boldsymbol {b}}$
, for some
$\sigma $
-block
${\boldsymbol {b}}$
in
$\mathfrak M_1$
. By Lemma 5.8,
${\boldsymbol {b}}$
is non-degenerate, and there are t-minimal and t-maximal points
$z_1^-$
and
$z_1^+$
in the non-degenerate clusters
$C_{{\boldsymbol {b}}}^-$
and
$C_{{\boldsymbol {b}}}^+$
. By Lemma 4.15,
$\boldsymbol {\beta }({\boldsymbol {b}})$
is a non-degenerate
$\sigma $
-block in
$\mathfrak M_2$
, so there are t-minimal and t-maximal points
$z_2^-$
and
$z_2^+$
in the non-degenerate clusters
$C_{\boldsymbol {\beta }({\boldsymbol {b}})}^-$
and
$C_{\boldsymbol {\beta }({\boldsymbol {b}})}^+$
. By adding
$z_1$
,
$z_1^-$
,
$z_1^+$
to
$W_1"$
and
$z_2^-$
,
$z_2^+$
to
$W_2"$
we cure the ref-defect
$u_1,v_1$
in
$\mathfrak F_1"$
and maintain (minmax). Also, as (minmax) held in
$\mathfrak F_i"$
, by Lemma 5.8 we already had some points from
$C_{{\boldsymbol {b}}}^-$
and
$C_{{\boldsymbol {b}}}^+$
in
$W_1"$
and some points from
$C_{\boldsymbol {\beta }({\boldsymbol {b}})}^-$
and
$C_{\boldsymbol {\beta }({\boldsymbol {b}})}^+$
in
$W_2"$
. So we did not create new defects in either frame, and the property (minmax) is maintained.
Case 2:
$u_1 \in {\boldsymbol {b}}^{u_1}$
,
$v_1 \in {\boldsymbol {b}}^{v_1}$
, for
$\sigma $
-blocks
${\boldsymbol {b}}^{u_1} \ne {\boldsymbol {b}}^{v_1}$
in
$\mathfrak M_1$
. By the definition of
$W_1"$
and
$C(u_1)$
,
$C(v_1)$
being
$<_{R_1"}$
-consecutive,
$C(u_1) = C^+_{{\boldsymbol {b}}^{u_1}}$
and
$C(v_1) = C^-_{{\boldsymbol {b}}^{v_1}}$
, so
$z_1\notin {\boldsymbol {b}}^{u_1}$
. We claim that there is an irreflexive
$z \in W_1$
such that
$u_1 R_1 z R_1 v_1$
and z is either
$t_{\mathfrak {M}_{1}}^{\sigma }(z)$
-maximal or
$t_{\mathfrak {M}_{1}}^{\sigma }(z)$
-minimal. Indeed, as
$u_1 R_1 z_1$
, we have
$\Diamond _{\mathsf {F}} t_{\mathfrak {M}_{1}}^{\sigma }(z_1) \subseteq t_{\mathfrak {M}_{1}}^{\sigma }(u_1)$
and
$\Diamond _{\mathsf {P}} t_{\mathfrak {M}_{1}}^{\sigma }(u_1) \subseteq t_{\mathfrak {M}_{1}}^{\sigma }(z_1)$
. As
$z_1\notin {\boldsymbol {b}}^{u_1}$
, there can be two cases: either
$(i) \Diamond _{\mathsf {F}} t_{\mathfrak {M}_{1}}^{\sigma }(u_1) \not \subseteq t_{\mathfrak {M}_{1}}^{\sigma }(z_1)$
or
$(ii) \Diamond _{\mathsf {P}} t_{\mathfrak {M}_{1}}^{\sigma }(u_1) \not \subseteq t_{\mathfrak {M}_{1}}^{\sigma }(z_1)$
. In case
$(i)$
, there is a
$\sigma $
-formula
$\chi $
with
$\mathfrak {M}_1,u_1\models \Diamond _{\mathsf {F}}\chi $
but
$\mathfrak {M}_1,z_1\not \models \Diamond _{\mathsf {F}}\chi $
. Take a
$\{\Diamond _{\mathsf {F}}\chi \}$
-maximal point
$z'$
. Clearly,
$u_1 R_1 z' R_1 v_1$
. If
$z'$
is irreflexive, we set
$z=z'$
as it is
$t_{\mathfrak {M}_{1}}^{\sigma }(z')$
-maximal. Otherwise, Lemma 5.4 gives an immediate degenerate
$<_{R_1}$
-successor
$C(z)$
of
$C(z')$
such that z is
$t_{\mathfrak {M}_{1}}^{\sigma }(z)$
-maximal. In case
$(ii)$
, there is a
$\sigma $
-formula
$\chi $
with
$\mathfrak {M}_1,u_1\not \models \Diamond _{\mathsf {P}}\chi $
but
$\mathfrak {M}_1,z_1\models \chi $
, and so
$\mathfrak {M}_1,v_1\models \Diamond _{\mathsf {P}}\chi $
. Take a
$\{\Diamond _{\mathsf {P}}\chi \}$
-minimal point
$z'$
. Clearly,
$u_1 R_1 z' R_1 v_1$
. If
$z'$
is irreflexive, we set
$z=z'$
as it is
$t_{\mathfrak {M}_{1}}^{\sigma }(z')$
-minimal. Otherwise, Lemma 5.4 gives an immediate degenerate
$<_{R_1}$
-predecessor
$C(z)$
of
$C(z')$
such that z is
$t_{\mathfrak {M}_{1}}^{\sigma }(z)$
-minimal.
Let
${\boldsymbol {b}}={\boldsymbol {b}}_{\mathfrak M_1}^\sigma (z)$
. Then
${\boldsymbol {b}}$
is a degenerate
$\sigma $
-block in
$\mathfrak M_1$
by Lemma 5.8. By Lemma 4.15,
$\boldsymbol {\beta }({\boldsymbol {b}})$
is a degenerate
$\sigma $
-block in
$\mathfrak M_2$
with
$\boldsymbol {\beta }({\boldsymbol {b}}^{u_1})\prec _{\mathfrak F_2}\boldsymbol {\beta }({\boldsymbol {b}})\prec _{\mathfrak F_2}\boldsymbol {\beta }({\boldsymbol {b}}^{v_1})$
. Also, by (minmax) in
$\mathfrak F_i"$
,
$C_{\boldsymbol {\beta }({\boldsymbol {b}}^{u_1})}^+$
and
$C_{\boldsymbol {\beta }({\boldsymbol {b}}^{v_1})}^-$
are
$<_{R_2"}$
-consecutive non-degenerate clusters. Therefore, by adding z to
$W_1"$
and
$z_2$
with
$C(z_2)=\boldsymbol {\beta }({\boldsymbol {b}})$
to
$W_2"$
, we cured the ref-defect
$u_1,v_1$
in
$\mathfrak F_1"$
and we did not create new defects of either kind in either frame while maintaining (minmax). So again it is readily seen (similarly to the proof of Lemma 3.4) that, after fixing all defects, we end up with a pair of models as required that are based on frames for
$\mathsf {Lin}_{\mathbb {R}}$
by Lemma 5.6.
This establishes the polysize bisimilar model property of
$L \in \{\mathsf {Lin}, \mathsf {Lin}_{\mathbb {Q}}, \mathsf {Lin}_{\mathbb {R}}\}$
. We show that the IEP for L is in coNP using the description of finite frames for L in Lemma 5.6.
The finitary selection construction in the proof above does not work for logics
$L\in \{\mathsf {Lin}_{<\omega },\mathsf {Lin}_{\mathbb {Z}}\}$
. In fact, these logics do not have the polysize bisimilar model property. However, below we show that they still have a kind of quasi-finite bisimilar model property similar to Definition 4.1 in the following sense. We can always witness the lack of an interpolant for
$\varphi _1$
,
$\varphi _2$
in L by a pair of temporal models that are based on frames for L, and assembled from
$\mathcal {O}\big (\max (|\varphi _1|,|\varphi _2|)\big )$
-many ‘simple’ models (like those in Example 5.2) that are based atomic descriptive frames of the forms
$m^<$
,
,
,
, and
, for
$m,k=\mathcal {O}\big (\max (|\varphi _1|,|\varphi _2|)\big )$
,
$k>0$
.
Given
$\varphi _i$
,
$\mathfrak M_i$
,
$x_i$
, for
$i=1,2$
, as above, let
$\boldsymbol {M}_{i}$
,
$\boldsymbol {S}_{i}$
, and
$W^{\prime }_i=\{x_i\}\cup \boldsymbol {M}_{i}\cup \boldsymbol {S}_{i}$
be as defined in the proof of Theorem 5.9. As before, we call the points from
$W^{\prime }_i$
relevant in
$\mathfrak M_i$
. A cluster or a
$\sigma $
-block in
$\mathfrak M_i$
is relevant if it contains a relevant point in
$\mathfrak M_i$
. Given any pair
${\boldsymbol {b}}$
,
$\boldsymbol {\beta }({\boldsymbol {b}})$
of
$\sigma $
-bisimilar
$\sigma $
-blocks in
$\mathfrak M_1$
and
$\mathfrak M_2$
, we can now have the temporal analogue of Lemma 4.17, dealing not only with
$\boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}}^+$
and
$\boldsymbol {S}_{2}\cap C_{\boldsymbol {\beta }({\boldsymbol {b}})}^+$
but also with
$\boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}}^-$
and
$\boldsymbol {S}_{2}\cap C_{\boldsymbol {\beta }({\boldsymbol {b}})}^-$
. In particular,
$$ \begin{align} & \text{there are } \sigma\text{-type preserving bijections } f^-\colon \boldsymbol{S}_{1}\cap C_{{\boldsymbol{b}}}^-\to \boldsymbol{S}_{2}\cap C_{\boldsymbol{\beta}({\boldsymbol{b}})}^-\\\nonumber &\qquad\qquad\qquad\qquad\qquad\qquad\qquad \text{and } f^+\colon \boldsymbol{S}_{1}\cap C_{{\boldsymbol{b}}}^+\to \boldsymbol{S}_{2}\cap C_{\boldsymbol{\beta}({\boldsymbol{b}})}^+{;} \end{align} $$
Theorem 5.10. The IEPs for
$\mathsf {Lin}_{<\omega }$
and
$\mathsf {Lin}_{\mathbb {Z}}$
are both coNP-complete.
Proof. Let
${\boldsymbol {b}}_1^0,\dots ,{\boldsymbol {b}}_1^N$
be all the relevant
$\sigma $
-blocks in
$\mathfrak M_1$
ordered by
$\prec _{\mathfrak F_1}$
, for some
$N=\mathcal {O}(\big (\max (|\varphi _1|,|\varphi _2|)\big )$
. By (40) and Lemma 4.15, the
$\prec _{\mathfrak F_2}$
-ordered list of all relevant
$\sigma $
-blocks in
$\mathfrak M_2$
is
${\boldsymbol {b}}_2^0,\dots ,{\boldsymbol {b}}_2^N$
, where
${\boldsymbol {b}}_2^j=\boldsymbol {\beta }({\boldsymbol {b}}_1^j)$
, for
$j\leq N$
. By (38) and (39), for every
$j\leq N$
there is
$k^j>0$
with
$k^j=|\boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}_1^j}^-|=|\boldsymbol {S}_{1}\cap C_{{\boldsymbol {b}}_1^j}^+|=|\boldsymbol {S}_{2}\cap C_{{\boldsymbol {b}}_2^j}^-|=|\boldsymbol {S}_{2}\cap C_{{\boldsymbol {b}}_2^j}^+|$
. Also, by Lemma 4.15,
${\boldsymbol {b}}_1^j$
is degenerate iff
${\boldsymbol {b}}_2^j$
is degenerate, for
$j\leq N$
.
Case
$L = \mathsf {Lin}_{<\omega }$
: By Lemmas 5.6 and 5.8,
${\boldsymbol {b}}_i^0$
and
${\boldsymbol {b}}_i^N$
,
$i=1,2$
, are degenerate. By Lemmas 5.4, 5.6, and 5.8, if
${\boldsymbol {b}}_i^j$
is non-degenerate, then
$C_{{\boldsymbol {b}}_i^j}^-$
and
$C_{{\boldsymbol {b}}_i^j}^+$
are
$R^-$
- and R-limit clusters, and
$C\cap \boldsymbol {M}_{i}=\emptyset $
, for every non-degenerate cluster C in
${\boldsymbol {b}}_i^j$
. (It can happen that
$x_i$
is in a non-degenerate cluster in
${\boldsymbol {b}}_i^j$
different from
$C_{{\boldsymbol {b}}_i^j}^-$
,
$C_{{\boldsymbol {b}}_i^j}^+$
.)
For all
$i=1,2$
and
$j\leq N$
, we let
$m_i^j=\big |\big ((\{x_i\}\cup \boldsymbol {M}_{i})\cap {\boldsymbol {b}}_i^j\big )\setminus (C_{{\boldsymbol {b}}_i^j}^-\cup C_{{\boldsymbol {b}}_i^j}^+)\big |$
and define an atomic frame
$\mathfrak H_i^j=(H_i^j,R_i^j,\mathcal {P}_i^j)$
by taking

Note that
$m_1^j$
and
$m_2^j$
might be different, and
$(\{x_i\}\cup \boldsymbol {M}_{i})\cap {\boldsymbol {b}}_i^j=\emptyset $
(and so
$m_i^j=0$
) can happen even when
$C_{{\boldsymbol {b}}_i^j}^-\ne C_{{\boldsymbol {b}}_i^j}^+$
. Let
$\mathfrak H_i=(H_i,R_i',\mathcal {P}_i')= \mathfrak H_i^0\lhd \dots \lhd \mathfrak H_i^N$
. It is readily seen that
$\mathfrak H_i$
is a frame for
$\mathsf {Lin}_{<\omega }$
, for
$i=1,2$
. Next, we define a ‘parent’ function
$\mathsf {h}_i\colon H_i\to W_i'$
such that, for all
$x\in H_i$
,
Finally, for
$j\leq N$
, we define a model
$\mathfrak N_i^j$
based on
$\mathfrak H_i^j$
by taking, for all
$x\in H_i^j$
,
and let
$\mathfrak N_i= \mathfrak N_i^0\lhd \dots \lhd \mathfrak N_i^N$
.
Instead of giving the general definitions of
$\mathsf {h}_i$
and
$\mathfrak N_i$
, we illustrate the construction in the picture below, where
$\mathfrak M_i$
has three degenerate
$\sigma $
-blocks
${\boldsymbol {b}}_i^0,{\boldsymbol {b}}_i^2$
, and
${\boldsymbol {b}}_i^3$
and one non-definable non-degenerate
$\sigma $
-block
${\boldsymbol {b}}_i^1$
; the relevant points in
$\mathfrak M_i$
are underlined;
$k^0=k^2=k^3=1$
,
$k^1=2,$
and
$m_i^1=3$
.

It is readily seen that this way (41)–(43) hold and
$\mathfrak N_i^j$
is based on
$\mathfrak H_i^j$
, for
$j\leq N$
. Thus,
$\mathfrak H_i^0\lhd \dots \lhd \mathfrak H_i^N$
,
$i=1,2$
, is a frame for
$\mathsf {Lin}_{<\omega }$
by Lemma 5.6. Using (41)–(43), a proof similar to that of Lemma 4.22
$(a)$
shows that each point x in
$\mathfrak N_i$
makes true exactly the same formulas in
$\textit {sub}(\varphi _i)$
as its parent
$\mathsf {h}_i(x)$
in
$\mathfrak M_i$
. It follows that
$\mathfrak {N}_{1},x_{1}' \models \varphi _1$
and
$\mathfrak {N}_{2},x_{2}' \models \neg \varphi _2$
, where
$x_i=\mathsf {h}_i(x^{\prime }_i)$
.
Further, the construction and (38) guarantee that each pair
$(\mathfrak N_1^j,\mathfrak N_2^j)$
, for
$j\leq N$
, satisfies an obvious condition similar to Definition 4.1
$(a)$
or
$(c)$
. Then a proof similar to that of Lemma 4.4 shows that
$\mathfrak N_1^j$
and
$\mathfrak N_2^j$
are globally
$\sigma $
-bisimilar for every
$j\leq N$
, and so
$\mathfrak {N}_{1},x^{\prime }_{1} \sim _{\sigma } \mathfrak {N}_{2},x^{\prime }_{2}$
.
Case
$L = \mathsf {Lin}_{\mathbb {Z}}$
: While the definitions of
$\mathfrak H_i^j$
, for
$0<j<N$
, are the same as above, for
$j=0,N$
we need new ones. Now, by Lemmas 5.6 and 5.8,
${\boldsymbol {b}}_i^0$
and
${\boldsymbol {b}}_i^N$
are non-degenerate, for
$i=1,2$
. Also, by Lemmas 5.4, 5.6, and 5.8, the
$R^-$
-final cluster
$C_{{\boldsymbol {b}}_i^0}^-$
in
$\mathfrak F_i$
is an
$R^-$
-limit cluster, and the R-final cluster
$C_{{\boldsymbol {b}}_i^N}^+$
in
$\mathfrak F_i$
is an R-limit cluster, for
$i=1,2$
. There are several cases. If
$N=0$
(that is,
${\boldsymbol {b}}_i^0=W_i$
) and
$C_{{\boldsymbol {b}}_i^0}^-=C_{{\boldsymbol {b}}_i^0}^+$
, then we let
. If
$N=0$
and
$C_{{\boldsymbol {b}}_i^0}^-\ne C_{{\boldsymbol {b}}_i^0}^+$
, then we let
. If
$N>0$
then

and

This way, by Lemma 5.6,
$\mathfrak H_i= \mathfrak H_i^0\lhd \dots \lhd \mathfrak H_i^N$
is a frame for
$\mathsf {Lin}_{\mathbb {Z}}$
, for
$i=1,2$
. Apart from these modifications, everything is similar to the
$\mathsf {Lin}_{<\omega }$
case.
A coNP-algorithm deciding interpolant existence in
$\mathsf {Lin}_{<\omega }$
or
$\mathsf {Lin}_{\mathbb {Z}}$
is an obvious adaptation of the algorithm detailed in the proof of Theorem 4.9.
We conjecture that the IEP for every consistent finitely axiomatisable Priorean temporal logic is coNP-complete.
6 Outlook and open problems
We have turned the lack of the CIP into a research question by asking whether deciding interpolant existence becomes harder than validity for modal logics without the CIP. As argued in [Reference Place33, Reference Place and Zeitoun35] for the closely related problem of separability of disjoint regular languages using a smaller language class (such as first-order definable languages), this question can be understood as a generalisation of satisfiability that provides new insights into the expressivity of the logic in question. We have shown that, in contrast to modal logics with nominals, the product modal logic
$\mathsf {S5}\times \mathsf {S5}$
, and the guarded and two-variable fragments of first-order logic, the complexity of deciding interpolant existence in finitely axiomatisable modal logics of linear frames is in coNP and, therefore, of the same complexity as validity. This appears to be the first general result about Craig interpolants for logics lacking the CIP. It gives rise to many further questions of which we mention only a few:
-
Q1: Is there a decidable modal logic above
$\mathsf {GL}$
,
$\mathsf {K4}$
, or
$\mathsf {K}$
with the undecidable IEP? Currently, the only known example of a decidable logic with the undecidable IEP is the two-variable fragment of first-order logic with two equivalence relations [Reference Wolter, Zakharyaschev, Giordano, Jung and Ozaki41]. -
Q2: Do all d-persistent (cofinal) subframe logics above
$\mathsf {K4}$
have the finite bisimilar model property? Can one show a quasi-finite bisimilar model property for all (cofinal) subframe logics above
$\mathsf {K4}$
and use it to prove that interpolant existence is decidable for all finitely axiomatisable ones? -
Q3: What is the situation with the IEP for propositional superintuitionistic (aka intermediate) logics and (super)intuitionistic modal logics without the CIP? Note that the Gödel translation reduces the IEP for propositional superintuitionistic logics to the IEP for (certain fragments of) modal logics above
$\mathsf {S4}$
(see the proof of [Reference Chagrov and Zakharyaschev9, Theorem 14.9]). -
Q4: Our proof is not constructive in the sense that is does not provide a non-trivial algorithm for computing interpolants if they exist (beyond exhaustive search) nor any upper bounds on their size. It would be of great interest to develop such algorithms. First steps towards computing interpolants in description logics without CIP are presented in [Reference Jung, Kołodziejski and Wolter23].
Descriptive frames have been crucial for our proofs. It would therefore be interesting and in line with the modal logic tradition to characterise logics for which descriptive frames can be replaced by Kripke (or even finite) frames in Theorem 3.2. While d-persistence is clearly a sufficient condition,
$\mathsf {Lin}_{\mathbb {R}}$
shows that it is not a necessary one (see Example 5.7). It is known, however, that
$\mathsf {Lin}_{\mathbb {R}}$
is strongly complete [Reference Wolter37], which suggests the conjecture that, in Theorem 3.2, descriptive frames for L can be replaced by Kripke frames iff L is strongly Kripke complete (in the sense that every L-consistent set of formulas is satisfiable in a Kripke frame for L). Note that a logic is strongly Kripke complete iff the corresponding variety of modal algebras is complex [Reference Goldblatt18, Reference Wolter37].
Acknowledgements
We are grateful to the anonymous reviewer whose comments and suggestions helped us to improve the presentation and terminology.







