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

Modal logics of almost sure validities in some classes of euclidean and transitive frames

Vladislav Sliusarev New Mexico State University, USA
and
Moscow Institute of Physics and Technology, Russia
Abstract

Given a class 𝒞\mathcal{C} of finite Kripke frames, we consider the uniform distribution on the frames from 𝒞\mathcal{C} with nn states. A formula is almost surely valid in 𝒞\mathcal{C} if the probability that it is valid in a random 𝒞\mathcal{C}-frame with nn states tends to 11 as nn tends to infinity. The formulas that are almost surely valid in CC form a normal modal logic.

We find complete and sound axiomatizations for the logics of almost sure validities in the classes of finite frames defined by the logics 𝐊𝐃𝟓\mathbf{KD5}, 𝐊𝐃𝟒𝟓\mathbf{KD45}, 𝐊𝟓𝐁\mathbf{K5B}, 𝐒𝟓\mathbf{S5}, 𝐆𝐫𝐳.3\mathbf{Grz.3}, and 𝐆𝐋.3.\mathbf{GL.3}.

MSC class: 03B45.

keywords:
modal logic, Kripke semantics, asymptotic probability, random graphs, euclidean relations, Grzegorczyk’s logic

The study of random structures for various logical systems is a significant research field within contemporary mathematics. A large body of work in this area concerns the first-order relational language. A key result in this field is the Zero-one law for first-order logic, which states that any first-order definable property of random relational structures, such as random graphs, has an asymptotic probability of either zero or one. This law was independently proved for the Erdős–Rényi model of random graph in [GKLT69] and [Fag76].

Exploring the behavior of random structures for logical systems that extend beyond the first-order language is a compelling research direction. The modal language, interpreted on relational structures using Kripke semantics, is a particularly important example, which has led to several noteworthy discoveries. Le Bars disproved the zero-one law for modal logic in Kripke frames [LB02]. Verbrugge [Ver18] proved the zero-one law in the finite models of of Grzegorczyk’s logic and weak Grzegorczyk’s logic, and later provided a valuable example of a modal logic that satisfies the zero-one law both in models and in frames, namely the provability logic GL [Ver21].

A related problem of interest involves the study of sentences that are true with a probability tending to one, termed as asymptotically almost surely true. Gaifman [Gai64] provided an axiomatization for almost sure truths in the Radó graph, a model of a countable random graph, in the first-order relational logic.

There are several results in the study of almost sure truths in the modal languages. The logic of almost sure truths in random Kripke models coincides with Carnap’s modal logic, according to Halpern and Kapron [HK94]. The logic of almost sure truths in transitive reflexive models is also described in [HK94]. Verbrugge [Ver18][Ver21] provided axiomatizations for almost sure truths in the models of GL,Grz,\textbf{GL},\,\textbf{Grz}, and wGrz, and almost sure validities in the frames of GL. Goranko [GK03] found a complete and sound axiomatization for the logic of almost sure validities in a countable frame. The paper [Gor20] utilizes this result to identify some of the almost sure validities in finite frames. However, the problem of complete axiomatization of almost sure validities in the class of finite frames remains open.

In this paper we discuss the logics of almost sure validities in various classes of finite frames. We generalize Goranko’s construction of the random finite frame: given a class of frames 𝒞,\mathcal{C}, we consider the uniform distribution on the labelled frames with nn states that belong to 𝒞.\mathcal{C}. Our general result states that the almost sure validities form a normal modal logic that extends Log𝒞.\operatorname{Log}\mathcal{C}. We achieve an important technical result of studying such logics: for a large class of logics, the almost sure validities in a random 𝒞\mathcal{C}-frame are also almost surely valid in a random connected 𝒞\mathcal{C}-frame (Theorem 2.9). Since the connected frames typically have a simpler combinatorial structure, this connection allows us to find upper bounds on the logic of almost sure validities in 𝒞.\mathcal{C}. We use this general theory to find finite axiomatizations for the logics of almost sure validities in the classes of frames defined by the logics 𝐊𝐃𝟓\mathbf{KD5}, 𝐊𝐃𝟒𝟓\mathbf{KD45}, 𝐊𝟓𝐁\mathbf{K5B}, 𝐒𝟓\mathbf{S5}, 𝐆𝐫𝐳.3,\mathbf{Grz.3}, and 𝐆𝐋.3.\mathbf{GL.3}.

1 Preliminaries

1.1 Modal syntax and semantics

We consider the basic modal language ML\mathrm{ML} of formulas in the alphabet that consists of a countable set PV={p0,p1,}\mathrm{PV}=\{p_{0},\,p_{1},\,\ldots\} of propositional variables, classical connectives ,,\to,\,\bot, and a unary operator .\Box. We use the standard abbreviations of connectives, in particular, φ¬¬φ.\Diamond\varphi\equiv\lnot\Box\lnot\varphi.

A set LMLL\subseteq\mathrm{ML} is a (normal modal) logic if LL contains all propositional tautologies, the normality axiom (pq)(pq)\Box(p\to q)\to(\Box p\to\Box q) and is closed under the rules of inference:

(MP)

If φ,φψL\varphi,\,\varphi\to\psi\in L, then ψL,\psi\in L,

(Gen)

If ϕL\phi\in L, then φL,\Box\varphi\in L,

(Sub)

If φL\varphi\in L, pPV,θMLp\in\mathrm{PV},\,\theta\in\mathrm{ML}, and the formula ψML\psi\in\mathrm{ML} is obtained from φ\varphi by replacing all instances of pp with θ,\theta, then ψL.\psi\in L.

By a frame we mean a Kripke frame F=(X,R),X,RX×X;F=(X,\,R),\,X\neq\varnothing,\,R\subseteq X\times X; we refer to the elements of XX as states of FF. The set of states, or domain, of FF is denoted domF.\operatorname{dom}F.

The notation FφF\models\varphi, where FF is a frame and φML\varphi\in\mathrm{ML} is a formula, means ‘φ\varphi is valid in FF’ with the standard definition (see, for example, [BdRV01, Definition 1.28]). For any set Γ\Gamma of modal formulas, FΓF\models\Gamma means φΓ(Fφ).\forall\varphi\in\Gamma\>\left(F\models\varphi\right). The set LogF\operatorname{Log}F of all formulas φML\varphi\in\mathrm{ML} that are valid in a frame FF is called the logic of F.F. This definition extends to classes of frames: if \mathcal{F} is a class of frames, then Log\operatorname{Log}\mathcal{F} is the set of all formulas that are valid in any frame FF\in\mathcal{F}. Given a set of formulas ΓML\Gamma\subseteq\mathrm{ML}, let FrΓ\operatorname{Fr}\Gamma denote the class of all frames FF such that FΓ.F\models\Gamma.

1.2 Operations on frames

For the convenience of the reader, we recall some basic notation and techniques of modal logic that we use in the article.

Let XX be a set. The diagonal relation on XX is IdX={(a,a)aX}.Id_{X}=\{(a,\,a)\mid a\in X\}. If RX×XR\subseteq X\times X is a relation, let R0=IdXR^{0}=Id_{X} and Ri+1=RRiR^{i+1}=R\circ R^{i} for any iω.i\in\omega. The inverse relation R1R{}^{-1} is defined as {(a,b)XbRa},\{(a,b)\in X\mid bRa\}, and Ri(R)1iR^{-i}\equiv{(R{}^{-1})}^{i} for any nω.n\in\omega.

For any UX,U\subseteq X, Rout[U]R_{\mathrm{out}}[U] denotes the set {aXuU(uRa)},\{a\in X\mid\exists u\in U\>\left(uRa\right)\}, and Rin[U]R_{\mathrm{in}}[U] denotes {aXuU(aRu)}.\{a\in X\mid\exists u\in U\>\left(aRu\right)\}. The notations Rout(a),Rin(a)R_{\mathrm{out}}(a),\,R_{\mathrm{in}}(a), where aX,a\in X, abbreviate Rout[{a}]R_{\mathrm{out}}[\{a\}] and Rin[{a}],R_{\mathrm{in}}[\{a\}], respectively.

For a relation RX×XR\subseteq X\times X, define the transitive closure R+=i1Ri,R^{+}=\bigcup_{i\geq 1}R^{i}, and the reflexive transitive closure R=IdXR+.R^{*}=Id_{X}\cup R^{+}.

Given RX×XR\subseteq X\times X and UX,U\subseteq X, the restriction of RR on UU is the relation RU=R(U×U).R{\upharpoonright}U=R\cap(U\times U).

If F=(X,R)F=(X,\,R) is a frame and UXU\subseteq X, the subframe of XX generated by UU is the frame FU=(Rout[U],RRout[U]).F{\uparrow}U=(R^{*}_{\mathrm{out}}[U],\,R{\upharpoonright}R^{*}_{\mathrm{out}}[U]). If aXa\in X, then FaF{\uparrow}a is a shorthand for F{a}.F{\uparrow}\{a\}. A frame FF is said to be point-generated if F=FaF=F{\uparrow}a for some adomF.a\in\operatorname{dom}F. The generated subframe preserves the validity of modal formulas: for any UdomFU\subseteq\operatorname{dom}F, LogFLogFU\operatorname{Log}{F}\subseteq\operatorname{Log}{F{\uparrow}U} [BdRV01, Proposition 2.6].

The disjoint sum iIFi\biguplus_{i\in I}F_{i} of the family of frames Fi=(Xi,Ri),iI,F_{i}=(X_{i},\,R_{i}),\,i\in I, where II is a nonempty set, is defined as (X,R)(X,\,R) where

X={(a,i)iI,aXi};(a,i)R(b,j)i=j and aRib.X=\{(a,i)\mid i\in I,\,a\in X_{i}\};\qquad(a,i)R(b,j)\iff i=j\text{ and }aR_{i}b.

The notation FGF\uplus G is a shorthand for i{1, 2}Fi\biguplus_{i\in\{1,\,2\}}F_{i} where F1=F,F2=G.F_{1}=F,\,F_{2}=G. It is well-known that LogiIFi=iILogFi\operatorname{Log}\biguplus_{i\in I}F_{i}=\bigcap_{i\in I}\operatorname{Log}F_{i} [BdRV01, Proposition 2.3].

Given a pair of frames F=(X,R)F=(X,\,R) and G=(Y,S),G=(Y,\,S), a p-morphism from FF to GG is a surjective map f:XYf:\>X\to Y such that Sout(f(a))=f(Rout(a))S_{\mathrm{out}}(f(a))=f(R_{\mathrm{out}}(a)) for any aX.a\in X. We write FGF\twoheadrightarrow G if there exists a p-morphism from FF to GG. The p-morphism preserves the validity of modal formulas: if FGF\twoheadrightarrow G, then LogFLogG\operatorname{Log}F\subseteq\operatorname{Log}G [BdRV01, Proposition 2.14]

A frame isomorphism between F=(X,R)F=(X,\,R) and G=(Y,S),G=(Y,\,S), is a bijection f:XYf:\>X\to Y such that aRbaRb iff f(a)Rf(b)f(a)Rf(b) for all a,bX.a,\,b\in X. It is straightforward that the existence of an isomorphism between FF and GG implies that LogF=LogG.\operatorname{Log}F=\operatorname{Log}G.

1.3 Classes of frames and their logics

A relation RX×XR\subseteq X\times X is:

  1. (i)

    serial if aX(Rout(a));\forall a\in X\>\left(R_{\mathrm{out}}(a)\neq\varnothing\right);

  2. (ii)

    reflexive if aX(aRa);\forall a\in X\>\left(aRa\right);

  3. (iii)

    irreflexive if ¬aX(aRa);\lnot\exists a\in X\>\left(aRa\right);

  4. (iv)

    symmetric if a,bX(aRbbRa);\forall a,\,b\in X\>\left(aRb\implies bRa\right);

  5. (v)

    transitive if a,b,cX(aRb,bRcaRc);\forall a,\,b,\,c\in X\>\left(aRb,\,bRc\implies aRc\right);

  6. (vi)

    Euclidean if a,b,cX(aRb,aRcbRc);\forall a,\,b,\,c\in X\>\left(aRb,\,aRc\implies bRc\right);

  7. (vii)

    non-branching if a,b,cX(aRb,aRcbRc or cRb or c=b);\forall a,\,b,\,c\in X\>\left(aRb,\,aRc\implies bRc\text{ or }cRb\text{ or }c=b\right);

  8. (viii)

    Noetherian if there are no infinite chains a0Ra1Ra_{0}Ra_{1}R\ldots with aiai+1,iω.a_{i}\neq a_{i+1},\,i\in\omega.

A frame F=(X,R)F=(X,\,R) is called serial (reflexive, etc.) if the relation of FF is serial (reflexive, etc.)

In this paper we will consider the logics 𝐊𝐃𝟓,𝐊𝐃𝟒𝟓,𝐊𝟓𝐁,𝐒𝟓,\mathbf{KD5},\,\mathbf{KD45},\,\mathbf{K5B},\,\mathbf{S5}, 𝐆𝐋.3,𝐆𝐫𝐳.3\mathbf{GL.3},\,\mathbf{Grz.3}. Recall that the frame classes of these logics are:

  1. \normalshape(1)

    Fr𝐊𝐃𝟓={serial Euclidean frames};\operatorname{Fr}\mathbf{KD5}=\{\text{serial Euclidean frames}\};

  2. \normalshape(2)

    Fr𝐊𝐃𝟒𝟓={serial transitive Euclidean frames};\operatorname{Fr}\mathbf{KD45}=\{\text{serial transitive Euclidean frames}\};

  3. \normalshape(3)

    Fr𝐊𝟓𝐁={symmetric Euclidean frames};\operatorname{Fr}\mathbf{K5B}=\{\text{symmetric Euclidean frames}\};

  4. \normalshape(4)

    Fr𝐒𝟓={reflexive Euclidean frames)}\operatorname{Fr}\mathbf{S5}=\{\text{reflexive Euclidean frames})\}

  5. \normalshape(5)

    Fr𝐆𝐋.3={transitive irreflexive non-branching Noetherian frames}\operatorname{Fr}\mathbf{GL.3}=\{\text{transitive irreflexive non-branching Noetherian frames}\}

  6. \normalshape(6)

    Fr𝐆𝐫𝐳.3={transitive reflexive non-branching Noetherian frames}\operatorname{Fr}\mathbf{Grz.3}=\{\text{transitive reflexive non-branching Noetherian frames}\}

These logics have the finite model property, so each of them is the logic of all finite point-generated frames that satisfy the corresponding frame condition [BdRV01]. For instance, 𝐊𝐃𝟓\mathbf{KD5} is the logic of all finite point-generated serial Euclidean frames.

1.4 Random frames

For any 1nω,1\leq n\in\omega, let [n][n] denote the set {0,,n1}\{0,\,\ldots,\,n-1\}, and let n={([n],R)R[n]×[n]}\mathcal{F}_{n}=\{([n],\,R)\mid R\subseteq[n]\times[n]\} be the set of all frames with the set of states [n].[n].

Let 𝒞\mathcal{C} be a nonempty class of frames. For any 1nω,1\leq n\in\omega, let F^n(𝒞)\hat{F}_{n}(\mathcal{C}) be the uniformly distributed random element of the finite set n𝒞:\mathcal{F}_{n}\cap\mathcal{C}:

P(F^n(𝒞)A)=|n𝒞A||n𝒞|for any set An𝒞.\operatorname{P}(\hat{F}_{n}(\mathcal{C})\in A)=\frac{|\mathcal{F}_{n}\cap\mathcal{C}\cap A|}{|\mathcal{F}_{n}\cap\mathcal{C}|}\quad\text{for any set }A\subseteq\mathcal{F}_{n}\cap\mathcal{C}. (1)

Formally, we fix some measure space (Ω,𝒢,P),(\Omega,\,\mathcal{G},\,\operatorname{P}), where 𝒢\mathcal{G} is a σ\sigma-algebra on Ω\Omega and P:𝒢[0,1]\operatorname{P}:\>\mathcal{G}\to[0,1] is a measure, and define F^n(𝒞)\hat{F}_{n}(\mathcal{C}) to be a measurable map from Ω\Omega to n\mathcal{F}_{n} that satisfies (1), where P(F^n(𝒞)A)\operatorname{P}(\hat{F}_{n}(\mathcal{C})\in A) is a notation for  P{ωΩF^n(𝒞)(ω)A}.\operatorname{P}\{\omega\in\Omega\mid\hat{F}_{n}(\mathcal{C})(\omega)\in A\}. The values F^n(𝒞)(ω)\hat{F}_{n}(\mathcal{C})(\omega) for ωΩ\omega\in\Omega are called realizations of F^n(𝒞).\hat{F}_{n}(\mathcal{C}).

For any set of frames 𝒬n\mathcal{Q}\subseteq\mathcal{F}_{n}, we say that F^n(𝒞)\hat{F}_{n}(\mathcal{C}) belongs to 𝒬\mathcal{Q} asymptotically almost surely (a.a.s.) if

limnP(F^n(𝒞)Q)=1.\lim_{n\to\infty}\operatorname{P}(\hat{F}_{n}(\mathcal{C})\in Q)=1.

Sometimes we will refer to a set of frames 𝒬n\mathcal{Q}\subseteq\mathcal{F}_{n} as a property of frames. In this case ‘𝒬\mathcal{Q} holds in F^n(𝒞)\hat{F}_{n}(\mathcal{C}) a.a.s.’ means that F^n(𝒞)𝒬\hat{F}_{n}(\mathcal{C})\in\mathcal{Q} a.a.s.

Let Log𝐚𝐬(𝒞)\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) denote the set of formulas φML\varphi\in\mathrm{ML} such that F^n(𝒞)φ\hat{F}_{n}(\mathcal{C})\models\varphi asymptotically almost surely.

If LL is a normal modal logic, we write F^n(L)\hat{F}_{n}(L) for F^n(FrL)\hat{F}_{n}(\operatorname{Fr}L) and L𝐚𝐬L^{\mathbf{as}} for Log𝐚𝐬(FrL).\operatorname{Log}^{\mathbf{as}}(\operatorname{Fr}L).

By (1),

φLog𝐚𝐬(𝒞)limn|n𝒞Fr{φ}||n𝒞|=1.\varphi\in\operatorname{Log}^{\mathbf{as}}(\mathcal{C})\iff\lim_{n\to\infty}\frac{|\mathcal{F}_{n}\cap\mathcal{C}\cap\operatorname{Fr}\{\varphi\}|}{|\mathcal{F}_{n}\cap\mathcal{C}|}=1. (2)

The present work studies the sets Log𝐚𝐬(𝒞)\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) for some modally definable classes of frames: 𝒞=Fr(L)\mathcal{C}=\operatorname{Fr}(L) for some modal logic L.L.

Theorem 1.1.

For any class of frames 𝒞,\mathcal{C}, Log𝐚𝐬(𝒞)\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) is a normal modal logic and Log𝐚𝐬(𝒞)Log(𝒞).\operatorname{Log}^{\mathbf{as}}(\mathcal{C})\supseteq\operatorname{Log}(\mathcal{C}).

Proof 1.2.

Since the minimal normal modal logic 𝐊\mathbf{K} is valid in all frames, P(F^n(𝒞)𝐊)=1\operatorname{P}(\hat{F}_{n}(\mathcal{C})\models\mathbf{K})=1 for any nω,n\in\omega, hence 𝐊Log𝐚𝐬(𝒞).\mathbf{K}\subseteq\operatorname{Log}^{\mathbf{as}}(\mathcal{C}). Then Log𝐚𝐬\operatorname{Log}^{\mathbf{as}} contains all propositional tautogolies and the normality axiom.

Let us show that Log𝐚𝐬(𝒞)\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) is closed under 𝐌𝐏.\mathbf{MP}. Let φLog𝐚𝐬(𝒞)\varphi\in\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) and φψLog𝐚𝐬(𝒞).\varphi\to\psi\in\operatorname{Log}^{\mathbf{as}}(\mathcal{C}). Since the logic of any frame is a normal modal logic, Fr{φ,φψ}Fr{ψ}.\operatorname{Fr}\{\varphi,\,\varphi\to\psi\}\subseteq\operatorname{Fr}\{\psi\}. Then by (1)

P(F^n(𝒞)ψ)\displaystyle\operatorname{P}(\hat{F}_{n}(\mathcal{C})\models\psi) =|n𝒞Fr{ψ}||n𝒞||n𝒞Fr{φ,φψ}||n𝒞|\displaystyle=\frac{|\mathcal{F}_{n}\cap\mathcal{C}\cap\operatorname{Fr}\{\psi\}|}{|\mathcal{F}_{n}\cap\mathcal{C}|}\geq\frac{|\mathcal{F}_{n}\cap\mathcal{C}\cap\operatorname{Fr}\{\varphi,\,\varphi\to\psi\}|}{|\mathcal{F}_{n}\cap\mathcal{C}|}
=P(F^n(𝒞)φ and F^n(𝒞)φψ}))\displaystyle=\operatorname{P}(\hat{F}_{n}(\mathcal{C})\models\varphi\text{ and }\hat{F}_{n}(\mathcal{C})\models\varphi\to\psi\}))
1P(F^n(𝒞)⊧̸φ)P(F^n(𝒞)⊧̸φψ).\displaystyle\geq 1-\operatorname{P}(\hat{F}_{n}(\mathcal{C})\not\models\varphi)-\operatorname{P}(\hat{F}_{n}(\mathcal{C})\not\models\varphi\to\psi).

Take the limit of both sides as n.n\to\infty. By assumption, P(F^n(𝒞)⊧̸φ)0\operatorname{P}(\hat{F}_{n}(\mathcal{C})\not\models\varphi)\to 0 and P(F^n(𝒞)⊧̸φψ)0,\operatorname{P}(\hat{F}_{n}(\mathcal{C})\not\models\varphi\to\psi)\to 0, so limnP(F^n(𝒞)ψ)1.\lim_{n\to\infty}\operatorname{P}(\hat{F}_{n}(\mathcal{C})\models\psi)\geq 1. Then ψLog𝐚𝐬(𝒞)\psi\in\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) by the definition.

Since Fr(φ)Fr(φ)\operatorname{Fr}(\Box\varphi)\subseteq\operatorname{Fr}(\varphi) for any φML,\varphi\in\mathrm{ML}, Log𝐚𝐬(𝒞)\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) is closed under 𝐆𝐞𝐧\mathbf{Gen} by (2). A similar argument applies for 𝐒𝐮𝐛.\mathbf{Sub}.

Finally, if φLog(𝒞)\varphi\in\operatorname{Log}(\mathcal{C}), then φ\varphi is valid in any possible value of F^n(𝒞).\hat{F}_{n}(\mathcal{C}). Then for any nω,P(F^n(𝒞)φ)=1n\in\omega,\,\operatorname{P}(\hat{F}_{n}(\mathcal{C})\models\varphi)=1, so φLog𝐚𝐬(𝒞).{\varphi\in\operatorname{Log}^{\mathbf{as}}(\mathcal{C})}.

It follows directly from the theorem that LL𝐚𝐬L\subseteq L^{\mathbf{as}} for any logic L.L.

Proposition 1.3.

If LL is a modal logic, then (L𝐚𝐬)𝐚𝐬=L𝐚𝐬.(L^{\mathbf{as}})^{\mathbf{as}}=L^{\mathbf{as}}.

Proof 1.4.

By Theorem 1.1 L𝐚𝐬(L𝐚𝐬)𝐚𝐬.L^{\mathbf{as}}\subseteq(L^{\mathbf{as}})^{\mathbf{as}}. For the other direction, consider any φ(L𝐚𝐬)𝐚𝐬\varphi\in(L^{\mathbf{as}})^{\mathbf{as}}. By the definition, there exist the limits:

limn|nFrL𝐚𝐬Fr{φ}||nFrL𝐚𝐬|=P(F^n(L𝐚𝐬)φ)=1,\displaystyle\lim_{n\to\infty}\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\cap\operatorname{Fr}\{\varphi\}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\right|}=\operatorname{P}(\hat{F}_{n}(L^{\mathbf{as}})\models\varphi)=1,
limn|nFrL𝐚𝐬||nFrL|=P(F^n(L)L𝐚𝐬)=1.\displaystyle\lim_{n\to\infty}\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L\right|}=\operatorname{P}(\hat{F}_{n}(L)\models L^{\mathbf{as}})=1.

Since LL𝐚𝐬L\subseteq L^{\mathbf{as}}, we have FrL𝐚𝐬FrL\operatorname{Fr}L^{\mathbf{as}}\subseteq\operatorname{Fr}L , so for any φ(L𝐚𝐬)𝐚𝐬\varphi\in(L^{\mathbf{as}})^{\mathbf{as}},

P(F^n(L)φ)\displaystyle\operatorname{P}(\hat{F}_{n}(L)\models\varphi) =limn|nFrLFr{φ}||nFrL|\displaystyle=\lim_{n\to\infty}\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L\cap\operatorname{Fr}\{\varphi\}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L\right|}
limn|nFrL𝐚𝐬Fr{φ}||nFrL|\displaystyle\geq\lim_{n\to\infty}\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\cap\operatorname{Fr}\{\varphi\}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L\right|}
=limn|nFrL𝐚𝐬Fr{φ}||nFrL𝐚𝐬||nFrL𝐚𝐬||nFrL|\displaystyle=\lim_{n\to\infty}\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\cap\operatorname{Fr}\{\varphi\}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\right|}\cdot\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L\right|}
=limn|nFrL𝐚𝐬Fr{φ}||nFrL𝐚𝐬|limn|nFrL𝐚𝐬||nFrL|=11=1.\displaystyle=\lim_{n\to\infty}\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\cap\operatorname{Fr}\{\varphi\}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\right|}\lim_{n\to\infty}\frac{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L^{\mathbf{as}}\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L\right|}=1\cdot 1=1.

1.5 Asymptotics

Let f,g:ωωf,\,g:\>\omega\to\omega be some functions. Then we write

  1. \normalshape(1)

    fg,f\sim g, if limnf(n)g(n)=1;\lim_{n\to\infty}\frac{f(n)}{g(n)}=1;

  2. \normalshape(2)

    f=o(g),f=o(g), if there exists α:ω\alpha:\>\omega\to\mathbb{R} such that f(n)=α(n)g(n)f(n)=\alpha(n)g(n) for all nωn\in\omega and limnα(n)=0;\lim_{n\to\infty}\alpha(n)=0;

  3. \normalshape(3)

    f=O(g),f=O(g), if there exists a real number C>0C>0 and nωn\in\omega such that f(n)Cg(n);f(n)\leq Cg(n);

  4. \normalshape(4)

    f=Ω(g),f=\Omega(g), if there exists a real number C>0C>0 and nωn\in\omega such that f(n)Cg(n).f(n)\geq Cg(n).

1.6 Set partitions and combinatorial numbers

Let XX be a set. A family of subsets 𝒰𝒫(X){}\mathcal{U}\subseteq\mathcal{P}\left(X\right)\setminus\{\varnothing\} is a partition of XX if the elements of 𝒰\mathcal{U} are pairwise disjoint and X=𝒰.X=\bigcup\mathcal{U}.

The Bell number BnB_{n} is defined as the number of distinct partitions of the set [n]={0,,n1}[n]=\{0,\,\ldots,\,n-1\}. Equivalently, BnB_{n} is the number of distinct equivalence relations on [n].[n]. The growth rate of the Bell number is described by the asymptotic expression [dB58, Section 6.2]

lnBn=n(lnnlnlnn1+o(1)),n.\ln B_{n}=n(\ln n-\ln\ln n-1+o(1)),\quad n\to\infty. (3)

The Bell numbers satisfy

limnBnBn+10.\lim_{n\to\infty}\frac{B_{n}}{B_{n+1}}\to 0. (4)

We give the proof of (4) in Appendix.

Given rω,r\in\omega, the number of partitions 𝒰\mathcal{U} of [n][n] such that |U|r|U|\leq r for any U𝒰U\in\mathcal{U} is denoted Gn,r.G_{n,r}. The asymptotic and combinatorial behavior of Gn,rG_{n,r} is discussed in [MMW58].

In this paper we will use the following estimation, which we prove in Appendix. For any constant r,kω,r,\,k\in\omega,

Gn,r2kn=o(Bn),nG_{n,r}2^{kn}=o(B_{n}),\,n\to\infty (5)

Given nωn\in\omega and mn,m\leq n, the binomial coefficient (nm)\binom{n}{m} is defined as the number of distinct mm-element subsets of the set [n].[n]. The following estimation holds for any nωn\in\omega and mnm\leq n [SF14, Section 5.4]:

(nm)(nn/2)2nπn/2,n\binom{n}{m}\leq\binom{n}{\lfloor n/2\rfloor}\sim\frac{2^{n}}{\sqrt{\pi n/2}},\quad n\to\infty (6)

2 Connected frames

The frame classes of different modal logics can demonstrate a very intricate combinatorial behavior that complicates the direct computation of probabilities by (2). However, it turns out that under certain conditions finding Log𝐚𝐬(𝒞)\operatorname{Log}^{\mathbf{as}}(\mathcal{C}) can be reduced to a much simpler problem of finding the almost sure validities in the connected frames of 𝒞.\mathcal{C}.

Let F=(X,R)F=(X,\,R) be a frame. Let =(RR)1{\sim}={(R\cup R{}^{-1})^{*}}. Then \sim is an equivalence relation on X.X. The elements of X/X/{\sim} are called the connected components of X.X. If XX has exactly one connected component, FF is called connected.

For a class of frames 𝒞,\mathcal{C}, we denote Con𝒞\operatorname{Con}\mathcal{C} the class of all connected frames in 𝒞.\mathcal{C}.

Proposition 2.1.

Let LL be a modal logic such that nConFrL\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}L is nonempty for any nω.n\in\omega. Then |nFrL|Bn|\mathcal{F}_{n}\cap\operatorname{Fr}L|\geq B_{n} for any n<ω.n<\omega.

Proof 2.2.

Let nωn\in\omega and let 𝒰\mathcal{U} be a partition of [n].[n]. For any U𝒰U\in\mathcal{U} there exists a frame FU|U|ConFrL.F_{U}\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L. Construct a frame F𝒰=([n],R𝒰)F_{\mathcal{U}}=([n],R_{\mathcal{U}}) by putting a copy of FUF_{U} on UU for any U𝒰.U\in\mathcal{U}. Then F𝒰F_{\mathcal{U}} is isomorphic to U𝒰FU,\biguplus_{U\in\mathcal{U}}F_{U}, and F𝒰LF_{\mathcal{U}}\models L since FULF_{U}\models L for any U𝒰.U\in\mathcal{U}. Thus F𝒰nFrL.F_{\mathcal{U}}\in\mathcal{F}_{n}\cap\operatorname{Fr}L.

Observe that 𝒰\mathcal{U} is exactly the set of connected components of F𝒰.F_{\mathcal{U}}. Thus if 𝒰,𝒱\mathcal{U},\,\mathcal{V} are partitions of XX and F𝒰=F𝒱F_{\mathcal{U}}=F_{\mathcal{V}}, then 𝒰=𝒱.\mathcal{U}=\mathcal{V}. Then 𝒰F𝒰\mathcal{U}\mapsto F_{\mathcal{U}} is an injective mapping of the partitions of [n][n] into nFrL,\mathcal{F}_{n}\cap\operatorname{Fr}L, so |nFrL|Bn.|\mathcal{F}_{n}\cap\operatorname{Fr}L|\geq B_{n}.

Definition 2.3.

Let rω.r\in\omega. Denote by nr\mathcal{F}_{n}^{\leq r} the set of frames in n\mathcal{F}_{n} whose connected components have cardinality at most r.r.

Proposition 2.4.

For any rω,r\in\omega, |nr|=o(Bn),n.\left|\mathcal{F}_{n}^{\leq r}\right|=o(B_{n}),\,n\to\infty.

Proof 2.5.

Let F=([n],R)nr.F=([n],R)\in\mathcal{F}_{n}^{\leq r}. If 𝒰\mathcal{U} is the set of connected components of FF, then (a,b)R(a,b)\not\in R for any aU,bVa\in U,\,b\in V where U,V𝒰,UV.U,\,V\in\mathcal{U},\,U\neq V. Then

RU𝒰U×U.R\subseteq\bigcup_{U\in\mathcal{U}}U\times U. (7)

Then R𝒫(U𝒰U×U).R\in\mathcal{P}\left(\bigcup_{U\in\mathcal{U}}U\times U\right). By the assumption, |U|r|U|\leq r for any U𝒰,U\in\mathcal{U}, so we can estimate

|U𝒰U×U|U𝒰r2=|𝒰|r2.\left|\bigcup_{U\in\mathcal{U}}U\times U\right|\leq\sum_{U\in\mathcal{U}}r^{2}=|\mathcal{U}|r^{2}. (8)

Let A𝒰nrA_{\mathcal{U}}\subseteq\mathcal{F}_{n}^{\leq r} denote the set of all frames in n\mathcal{F}_{n} whose set of connected components is 𝒰.\mathcal{U}. By (7) and (8), |A𝒰|2|𝒰|r2|A_{\mathcal{U}}|\leq 2^{|\mathcal{U}|r^{2}}. Therefore

|nr|𝒰|A𝒰|𝒰2r2|𝒰|𝒰2r2n,|\mathcal{F}_{n}^{\leq r}|\leq\sum_{\mathcal{U}}|A_{\mathcal{U}}|\leq\sum_{\mathcal{U}}2^{r^{2}|\mathcal{U}|}\leq\sum_{\mathcal{U}}2^{r^{2}n},

where the sum is taken over all partitions 𝒰\mathcal{U} of [n][n] into sets of cardinality at most r.r. The number of such partitions is Gn,r,G_{n,r}, so by (5)

|n|rGn,r2r2n=o(Bn).|\mathcal{F}_{n}|^{\leq r}\leq G_{n,r}2^{r^{2}n}=o(B_{n}).
Proposition 2.6.

Let LL be a modal logic such that nConFrL\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}L is nonempty for any nω.n\in\omega. Then for any fixed rω,r\in\omega, F^n(L)\hat{F}_{n}(L) has a connected component of cardinality greater than rr a.a.s.

Proof 2.7.

By Proposition 2.1 and Proposition 2.4,

P(F^n(L)nr)=|nrFrL||nFrL|o(Bn)Bn0,n.\operatorname{P}(\hat{F}_{n}(L)\in\mathcal{F}_{n}^{\leq r})=\frac{\left|\mathcal{F}_{n}^{\leq r}\cap\operatorname{Fr}L\right|}{\left|\mathcal{F}_{n}\cap\operatorname{Fr}L\right|}\leq\frac{o(B_{n})}{B_{n}}\to 0,\,n\to\infty.

To prove the following theorem, we consider the distributions of generated subframes of the random frame that have some fixed size mω.m\in\omega. To simplify the reasoning, we view them as random frames in m,\mathcal{F}_{m}, using the following definition.

Definition 2.8.

Let m,nω,mn.m,\,n\in\omega,\,m\leq n. If U[n]U\subseteq[n] and |U|=m,|U|=m, the monotone relabeling of UU is the unique bijection α:U[m]\alpha:\>U\to[m] such that α(a)α(b)\alpha(a)\leq\alpha(b) iff ab.a\leq b. Two frames F=(U,R)F=(U,R) and G=([m],S)G=([m],S) coincide up to monotone relabeling if the monotone relabeling α:U[m]\alpha:\>U\to[m] is a frame isomorphism between FF and G.G.

Theorem 2.9.

Let LL be a modal logic such that:

  1. \normalshape(1)

    nConFrL\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}L is nonempty for any nω.n\in\omega.

  2. \normalshape(2)

    For any φLog𝐚𝐬(ConFrL),\varphi\not\in\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}L),

    lim supnP(F^n(ConFrL)φ)<1.\limsup_{n\to\infty}\operatorname{P}(\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}L)\models\varphi)<1. (9)

Then L𝐚𝐬Log𝐚𝐬(ConFrL).L^{\mathbf{as}}\subseteq\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}L).

Proof 2.10.

Suppose that φLog𝐚𝐬(ConFrL),\varphi\not\in\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}L), then by (9)

lim supnP(F^n(ConFrL)φ)=lim supn|nConFrLFr{φ}||nConFrL|<1.\limsup_{n\to\infty}\operatorname{P}(\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}L)\models\varphi)=\limsup_{n\to\infty}\frac{|\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}L\cap\operatorname{Fr}\{\varphi\}|}{|\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}L|}<1.

Then for some rωr\in\omega and p>0,p>0,

|mConFrLFr{φ}||mConFrL|>pmr.\frac{|\mathcal{F}_{m}\cap\operatorname{Con}\operatorname{Fr}L\setminus\operatorname{Fr}\{\varphi\}|}{|\mathcal{F}_{m}\cap\operatorname{Con}\operatorname{Fr}L|}>p\quad\forall m\geq r. (10)

Define a random subset U^n[n]\hat{U}_{n}\subseteq[n] to be the connected component of F^n(L)\hat{F}_{n}(L) that has the maximal cardinality. If there are more than one such components U1,,UkU_{1},\,\ldots,\,U_{k}, let U^\hat{U} be the one that contains the state a=minj=1kUka=\min\bigcup_{j=1}^{k}U_{k}.

Any realization of the generated subframe F^n(L)U^n\hat{F}_{n}(L){\uparrow}\hat{U}_{n} is a connected frame that validates LL, so F^n(L)U^n\hat{F}_{n}(L){\uparrow}\hat{U}_{n} coincides, up to monotone relabeling of states, with some frame from |U^n|ConFrL\mathcal{F}_{|\hat{U}_{n}|}\cap\operatorname{Con}\operatorname{Fr}L almost surely.

We claim that for any fixed U[n]U\subseteq[n], the values of F^n(L)U^n\hat{F}_{n}(L){\uparrow}\hat{U}_{n} with U^n=U\hat{U}_{n}=U are distributed uniformly on |U|ConFrL.\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L. Informally, F^n(L)U^n\hat{F}_{n}(L){\uparrow}\hat{U}_{n} is independent of F^n(L)([n]U^n)\hat{F}_{n}(L){\uparrow}([n]\setminus\hat{U}_{n}).

Let U[n]U\subseteq[n] and denote by 𝒢n,U\mathcal{G}_{n,U} the set of frames from nFrL\mathcal{F}_{n}\cap\operatorname{Fr}L where UU is the maximal connected component. Let us consider any F1,F2|U|ConFrLF_{1},\,F_{2}\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L. For any G=([n],R)𝒢n,UG=([n],R)\in\mathcal{G}_{n,U} such that GUG{\uparrow}U equals F1F_{1} (up to monotone relabeling, which we assume hereinafter), we construct a frame GG^{\prime} by changing the relation of GG on UU in such a way that G([n]U)=G([n]U)G^{\prime}{\uparrow}([n]\setminus U)=G{\uparrow}([n]\setminus U) and GU=F2.G^{\prime}{\uparrow}U=F_{2}.

Let us show that G𝒢n,U.G^{\prime}\in\mathcal{G}_{n,U}. By the construction, UU is the maximal connected component of G.G^{\prime}. Moreover, GLG\models L implies that G([n]U)L,G{\uparrow}([n]\setminus U)\models L, so GLG^{\prime}\models L since GG([n]U)F2G^{\prime}\cong G{\uparrow}([n]\setminus U)\,\uplus\,F_{2} and F2L.F_{2}\models L. Thus GnFrL.G^{\prime}\in\mathcal{F}_{n}\cap\operatorname{Fr}L.

Then the mapping GGG\mapsto G^{\prime} is a bijection between {F𝒢n,U:FU=F1}\{F\in\mathcal{G}_{n,U}:\>F{\uparrow}U=F_{1}\} and {F𝒢n,U:FU=F2},\{F\in\mathcal{G}_{n,U}:\>F{\uparrow}U=F_{2}\}, so

|{F𝒢n,U:FU=F1}|=|{F𝒢n,U:FU=F2}|;\left|\{F\in\mathcal{G}_{n,U}:\>F{\uparrow}U=F_{1}\}\right|=\left|\{F\in\mathcal{G}_{n,U}:\>F{\uparrow}U=F_{2}\}\right|;

consequently, for any U[n]U\subseteq[n] and F1,F2|U|ConFrL,F_{1},\,F_{2}\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L,

P\displaystyle\operatorname{P} (F^n(L)U^n=F1|U^n=U)=P(F^n(L)U^n=F1 and U^n=U)P(U^n=U)\displaystyle\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}=F_{1}\,\Big{|}\,\hat{U}_{n}=U\right)=\frac{\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}=F_{1}\text{ and }\hat{U}_{n}=U\right)}{\operatorname{P}(\hat{U}_{n}=U)}
=|{F𝒢n,U:FU=F1}||𝒢n,U|=|{F𝒢n,U:FU=F2}||𝒢n,U|\displaystyle=\frac{\left|\{F\in\mathcal{G}_{n,U}:\>F{\uparrow}U=F_{1}\}\right|}{|\mathcal{G}_{n,U}|}=\frac{\left|\{F\in\mathcal{G}_{n,U}:\>F{\uparrow}U=F_{2}\}\right|}{|\mathcal{G}_{n,U}|}
=P(F^n(L)U^n=F2 and U^n=U)P(U^n=U)=P(F^n(L)U^n=F2|U^n=U)\displaystyle=\frac{\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}=F_{2}\text{ and }\hat{U}_{n}=U\right)}{\operatorname{P}(\hat{U}_{n}=U)}=\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}=F_{2}\,\Big{|}\,\hat{U}_{n}=U\right) (11)

Since F^n(L)U^n\hat{F}_{n}(L){\uparrow}\hat{U}_{n} is a connected frame that validates LL,

G|U|ConFrLP(F^n(L)U^n=G|U^n=U)=1.\sum_{G\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L}\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}=G\,\Big{|}\,\hat{U}_{n}=U\right)=1.

By (11) all terms in this sum are equal. Thus for any G|U|ConFrL,G\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L,

P(F^n(L)U^n=G|U^n=U)=1||U|ConFrL|.\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}=G\,\Big{|}\,\hat{U}_{n}=U\right)=\frac{1}{|\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L|}. (12)

Since F^n(L)U^n\hat{F}_{n}(L){\uparrow}\hat{U}_{n} is a generated subframe of F^n(L),\hat{F}_{n}(L), F^n(L)⊧̸φ\hat{F}_{n}(L)\not\models\varphi whenever F^n(L)U^n⊧̸φ.\hat{F}_{n}(L){\uparrow}\hat{U}_{n}\not\models\varphi. Then by the law of total probability [Gut13, Proposition 4.1]

P(F^n(L)⊧̸φ)\displaystyle\operatorname{P}(\hat{F}_{n}(L)\not\models\varphi) P(F^n(L)U^n⊧̸φ)\displaystyle\geq\operatorname{P}(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}\not\models\varphi)
=U[n]P(F^n(L)U^n⊧̸φ|U^n=U)P(U^n=U)\displaystyle=\sum_{U\subseteq[n]}\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}\not\models\varphi\,\Big{|}\,\hat{U}_{n}=U\right)P(\hat{U}_{n}=U)
|U|>rP(F^n(L)U^n⊧̸φ|U^n=U)P(U^n=U)\displaystyle\geq\sum_{|U|>r}\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}\not\models\varphi\,\Big{|}\,\hat{U}_{n}=U\right)\operatorname{P}(\hat{U}_{n}=U)
=|U|>rG|U|ConFrLG⊧̸φP(F^n(L)U^n=G|U^n=U)P(U^n=U)\displaystyle=\sum_{|U|>r}\ \sum_{\begin{subarray}{c}G\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L\\ G\not\models\varphi\end{subarray}}\operatorname{P}\left(\hat{F}_{n}(L){\uparrow}\hat{U}_{n}=G\,\Big{|}\,\hat{U}_{n}=U\right)\operatorname{P}(\hat{U}_{n}=U)
=(12)|U|>rG|U|ConFrLG⊧̸φ1||U|ConFrL|P(U^n=U)\displaystyle\overset{\eqref{eq:prob-of-G}}{=}\sum_{|U|>r}\ \sum_{\begin{subarray}{c}G\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L\\ G\not\models\varphi\end{subarray}}\frac{1}{|\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L|}\operatorname{P}(\hat{U}_{n}=U)
=|U|>rP(U^n=U)G|U|ConFrLG⊧̸φ1||U|ConFrL|\displaystyle=\sum_{|U|>r}\operatorname{P}(\hat{U}_{n}=U)\sum_{\begin{subarray}{c}G\in\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L\\ G\not\models\varphi\end{subarray}}\frac{1}{|\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L|}
=|U|>rP(U^n=U)||U|ConFrLFr{φ}|||U|ConFrL|\displaystyle=\sum_{|U|>r}\operatorname{P}(\hat{U}_{n}=U)\frac{|\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L\setminus\operatorname{Fr}\{\varphi\}|}{|\mathcal{F}_{|U|}\cap\operatorname{Con}\operatorname{Fr}L|}
=m>rP(|U^n|=m)|mConFrLFr{φ}||mConFrL|\displaystyle=\sum_{m>r}\operatorname{P}(|\hat{U}_{n}|=m)\frac{|\mathcal{F}_{m}\cap\operatorname{Con}\operatorname{Fr}L\setminus\operatorname{Fr}\{\varphi\}|}{|\mathcal{F}_{m}\cap\operatorname{Con}\operatorname{Fr}L|}
>(10)m>rP(|U^n|=m)p=pm>rP(|U^n|=m)=pP(|U^n|>r).\displaystyle\overset{\eqref{eq:prob-phi-false}}{>}\sum_{m>r}\operatorname{P}(|\hat{U}_{n}|=m)\cdot p=p\sum_{m>r}\operatorname{P}(|\hat{U}_{n}|=m)=p\operatorname{P}(|\hat{U}_{n}|>r).

Take the limit as n.n\to\infty. By Proposition 2.6, P(|U^n|>r)1,\operatorname{P}(|\hat{U}_{n}|>r)\to 1, so

limnP(F^n(L)⊧̸φ)plimnP(|U^n|>r)=p.\lim_{n\to\infty}\operatorname{P}(\hat{F}_{n}(L)\not\models\varphi)\geq p\lim_{n\to\infty}\operatorname{P}(|\hat{U}_{n}|>r)=p.

Then φL𝐚𝐬\varphi\not\in L^{\mathbf{as}} since

limnP(F^n(L)φ)=1limnP(F^n(L)⊧̸φ)1p<1.\lim_{n\to\infty}\operatorname{P}(\hat{F}_{n}(L)\models\varphi)=1-\lim_{n\to\infty}\operatorname{P}(\hat{F}_{n}(L)\not\models\varphi)\leq 1-p<1.

3 Euclidean frames

In this section we apply Theorem 2.9 to study the almost sure validities in Euclidean frames.

We begin with a simple observation about the structure of a Euclidean frame.

Proposition 3.1.

Let F=(X,R)F=(X,\,R) be Euclidean. Then there exists a unique subset UXU\subseteq X (possibly empty) such that RUR{\upharpoonright}U is an equivalence relation and RX×U.R\subseteq X\times U.

Proof 3.2.

Let U=aXRout(a).U=\bigcup_{a\in X}R_{\mathrm{out}}(a). For any uU,aRuu\in U,\,aRu for some aX,a\in X, so by the definition of Euclidean relation uRu.uRu. Then RUR{\upharpoonright}U is reflexive and Euclidean, hence an equivalence relation.

Every state aXa\in X with Rin(a)R_{\mathrm{in}}(a)\neq\varnothing is in UU, so RX×U.R\subseteq X\times U.

To prove the uniqueness, let us assume that VXV\subseteq X satisfies the conditions. If aVUa\in V\setminus U, then Rin(a)=,R_{\mathrm{in}}(a)=\varnothing, so RR is not reflexive on VU.V\setminus U. Since RVR{\upharpoonright}V is an equivalence relation, VU.V\subseteq U. If aUa\in U, then aRaaRa, so (a,a)X×V,(a,\,a)\in X\times V, hence aV.a\in V. Then U=VU=V, so UU is the unique subset of XX that satisfies the conditions.

Definition 3.3.

For any Euclidean frame F,F, let UFU_{F} denote the subset of domF\operatorname{dom}F defined by the conditions from Proposition 3.1.

Definition 3.4.

A frame F=(X,R)F=(X,\,R) is a cluster if R=X×X.R=X\times X.

Proposition 3.5.

A serial Euclidean frame FF is connected if and only if FUFF{\uparrow}U_{F} is a cluster.

Proof 3.6.

Let F=(X,R)F=(X,\,R) be connected. Since RR is an equivalence relation on UF,U_{F}, (R(R)1IdX)UF=RUF,(R^{*}\cup(R^{*}){}^{-1}\cup Id_{X}){\upharpoonright}U_{F}=R{\upharpoonright}U_{F}, so RUF=UF×UF,R{\upharpoonright}U_{F}=U_{F}\times U_{F}, thus FUFF{\uparrow}U_{F} is a cluster.

Conversely, suppose that RUFR{\upharpoonright}U_{F} is a cluster. For any aX,a\in X, Rout(a)UFR_{\mathrm{out}}(a)\subseteq U_{F} and Rout(a),R_{\mathrm{out}}(a)\neq\varnothing, so aRuaRu for some uU.u\in U. Then aR2vaR^{2}v for any vU.v\in U. Therefore Rout2[X]=U,R^{2}_{\mathrm{out}}[X]=U, so R2R2=X×X,R^{2}\cup R^{-2}=X\times X, and FF is connected.

Proposition 3.7.

|nConFr𝐊𝐃𝟓|=2n24+n+O(logn).\left|\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD5}\right|=2^{\frac{n^{2}}{4}+n+O(\log n)}.

Proof 3.8.

A frame F=([n],R)nConFr𝐊𝐃𝟓F=([n],\,R)\in\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD5} is uniquely determined by UFU_{F} and the family of subsets {Rout(a)a[n]UF}\{R_{\mathrm{out}}(a)\mid a\in[n]\setminus U_{F}\} where Rout(a)UFR_{\mathrm{out}}(a)\subseteq U_{F} and Rout(a)R_{\mathrm{out}}(a)\neq\varnothing for all a.a. Then

|nConFr𝐊𝐃𝟓|=U[n]U(2|U|1)n|U|=m=1n(nm)(2m1)nm.\left|\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD5}\right|=\sum_{\begin{subarray}{c}U\subseteq[n]\\ U\neq\varnothing\end{subarray}}\left(2^{|U|}-1\right)^{n-|U|}=\sum_{m=1}^{n}\binom{n}{m}\left(2^{m}-1\right)^{n-m}. (13)

We find the lower and the upper asymptotic bound for this sum using (6):

m=1n(nm)(2m1)nmm=1n(nm)2n24=2n24m=1n(nm)2n242n=2n24+n;\sum_{m=1}^{n}\binom{n}{m}(2^{m}-1)^{n-m}\leq\sum_{m=1}^{n}\binom{n}{m}2^{\frac{n^{2}}{4}}=2^{\frac{n^{2}}{4}}\sum_{m=1}^{n}\binom{n}{m}\leq 2^{\frac{n^{2}}{4}}\cdot 2^{n}=2^{\frac{n^{2}}{4}+n};
m=1n(nm)(2m1)nm(nn2)(2n21)nn2=2nπn/2(1+o(1))2n24(1+o(1))=2n24+n+O(logn).\sum_{m=1}^{n}\binom{n}{m}(2^{m}-1)^{n-m}\geq\binom{n}{\frac{n}{2}}\left(2^{\frac{n}{2}}-1\right)^{n-\frac{n}{2}}\\ =\frac{2^{n}}{\sqrt{\pi n/2}}(1+o(1))\cdot 2^{\frac{n^{2}}{4}}(1+o(1))=2^{\frac{n^{2}}{4}+n+O(\log n)}.
Proposition 3.9.

For any fixed rω,r\in\omega, |UF^n(ConFr𝐊𝐃𝟓)|>r|U_{\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})}|>r asymptotically almost surely.

Proof 3.10.

By (13), the number of connected 𝐊𝐃𝟓\mathbf{KD5}-frames with |UF|r|U_{F}|\leq r is

U[n],|U|r2|U|(n|U|)=m=1r(nm)2m(nm)rO(nr)2nr=O(2nr),n.\sum_{U\subseteq[n],\,|U|\leq r}2^{|U|(n-|U|)}=\sum_{m=1}^{r}\binom{n}{m}2^{m(n-m)}\leq r\cdot O(n^{r})\cdot 2^{nr}=O(2^{nr}),\,n\to\infty.

Then by Proposition 3.7,

P(|UF^n(ConFr𝐊𝐃𝟓)|r)=O(2nr)|nConFr𝐊𝐃𝟓|=O(2nr)2n24+O(n)0,n.\operatorname{P}(|U_{\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})}|\leq r)=\frac{O(2^{nr})}{\left|\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD5}\right|}=\frac{O(2^{nr})}{2^{\frac{n^{2}}{4}+O(n)}}\to 0,\quad n\to\infty.
Proposition 3.11.

Let R^n\hat{R}_{n} denote the relation of F^n(ConFr𝐊𝐃𝟓)\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5}) and U^n=UF^n(ConFr𝐊𝐃𝟓).\hat{U}_{n}=U_{\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})}. Let rωr\in\omega be fixed. Then a.a.s. there exists a state a[n]U^na\in[n]\setminus\hat{U}_{n} such that r<|R^n(a)|<|U^n|r.r<|\hat{R}_{n}(a)|<|\hat{U}_{n}|-r.

Proof 3.12.

Let 𝒬n,r\mathcal{Q}_{n,r} be the subset of F=([n],R)nConFr𝐊𝐃𝟓F=([n],\,R)\in\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD5} consisting of all frames F=([n],R)F=([n],R) such that for any a[n]UF,a\in[n]\setminus U_{F}, |Rout(a)|r{|R_{\mathrm{out}}(a)|\leq r} or |Rout(a)||UF|r|R_{\mathrm{out}}(a)|\geq|U_{F}|-r. Let us estimate |𝒬n,r|.|\mathcal{Q}_{n,r}|. If we fix UF[n]U_{F}\subseteq[n] with |UF|=m|U_{F}|=m, then for any aUFa\not\in U_{F} the nonempty subset Rout(a)UFR_{\mathrm{out}}(a)\subseteq U_{F} can be chosen in k=1r((mk)+(mmk)1)\sum_{k=1}^{r}\left(\binom{m}{k}+\binom{m}{m-k}-1\right) ways. Note that

k=1r((mk)+(mmk)1)2r(mr)2rmrr!2mr.\sum_{k=1}^{r}\left(\binom{m}{k}+\binom{m}{m-k}-1\right)\leq 2r\binom{m}{r}\leq 2r\frac{m^{r}}{r!}\leq 2m^{r}.

Then by (13),

|𝒬n,r|m=1n(nm)(2mr)nmm=1n(nm)(2m)nr(2n)nrm=1n(nm)(2n)nr2n=O(2rnlog2n),n.|\mathcal{Q}_{n,r}|\leq\sum_{m=1}^{n}\binom{n}{m}\left(2m^{r}\right)^{n-m}\leq\sum_{m=1}^{n}\binom{n}{m}(2m)^{nr}\\ \leq(2n)^{nr}\sum_{m=1}^{n}\binom{n}{m}\leq(2n)^{nr}2^{n}=O\left(2^{rn\log_{2}n}\right),\quad n\to\infty.

Finally, we show that F^n(ConFr𝐊𝐃𝟓)𝒬n,r\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})\not\in\mathcal{Q}_{n,r} a.a.s. By Proposition 3.7,

P(UF^n(ConFr𝐊𝐃𝟓)𝒬n,r)=|𝒬n,r||nConFr𝐊𝐃𝟓|O(2rnlog2n)2n24+O(n)0,n.\operatorname{P}(U_{\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})}\in\mathcal{Q}_{n,r})=\frac{|\mathcal{Q}_{n,r}|}{\left|\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD5}\right|}\leq\frac{O(2^{rn\log_{2}n})}{2^{\frac{n^{2}}{4}+O(n)}}\to 0,\,n\to\infty.
Proposition 3.13.

For any φ𝐊𝐃𝟓,F^n(ConFr𝐊𝐃𝟓)⊧̸φ\varphi\not\in\mathbf{KD5},\,\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})\not\models\varphi a.a.s.

Proof 3.14.

Consider a formula φ𝐊𝐃𝟓.\varphi\not\in\mathbf{KD5}. There exists a finite point-generated 𝐊𝐃𝟓\mathbf{KD5}-frame G=(Y,S)=GcG=(Y,S)=G{\uparrow}c such that G⊧̸φ.G\not\models\varphi. Let r=|domG|.r=|\operatorname{dom}G|. Let F=(X,R)F=(X,\,R) be a connected 𝐊𝐃𝟓\mathbf{KD5}-frame such that

|UF|>r and aXUF(r<|Rout(a)|<|UF|r).|U_{F}|>r\text{ and }\exists a\in X\setminus U_{F}\>\left(r<|R_{\mathrm{out}}(a)|<|U_{F}|-r\right). (14)

We define a p-morphism f:FaG.{f:F{\uparrow}a\twoheadrightarrow G}. By Proposition 3.5, domFa=Rout(a)=UF{a}.\operatorname{dom}F{\uparrow}a=R^{*}_{\mathrm{out}}(a)=U_{F}\cup\{a\}. Let f(a)=c.f(a)=c. Observe that |Rout(a)|>r|UG|>|S(c)|,|R_{\mathrm{out}}(a)|>r\geq|U_{G}|>|S(c)|, so let ff map Rout(a)R_{\mathrm{out}}(a) surjectively onto S(c).S(c). Analogously, let fUFRout(a)f{\upharpoonright}_{U_{F}\setminus R_{\mathrm{out}}(a)} be a surjection onto UGS(c).U_{G}\setminus S(c).

By Propositions 3.9 and 3.11, F^n(ConFr𝐊𝐃𝟓)\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5}) has the property (14) a.a.s., so F^n(ConFr𝐊𝐃𝟓)G\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})\twoheadrightarrow G, hence F^n(ConFr𝐊𝐃𝟓)⊧̸φ\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})\not\models\varphi a.a.s.

Theorem 3.15.

𝐊𝐃𝟓𝐚𝐬=𝐊𝐃𝟓.\mathbf{KD5}^{\mathbf{as}}=\mathbf{KD5}.

Proof 3.16.

By Theorem 1.1, 𝐊𝐃𝟓𝐊𝐃𝟓𝐚𝐬\mathbf{KD5}\subseteq\mathbf{KD5}^{\mathbf{as}}. For the other direction, observe that the cluster ([n],[n]×[n])([n],\,[n]\times[n]) is a connected 𝐊𝐃𝟓\mathbf{KD5}-frame for all nωn\in\omega, thus nConFr𝐊𝐃𝟓\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD5}\neq\varnothing for all nω.n\in\omega. By Proposition 3.13lim supnP(F^n(ConFr𝐊𝐃𝟓)φ)=0<1.\limsup\limits_{n\to\infty}\operatorname{P}(\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})\models\varphi)=0<1. Then by Theorem 2.9, 𝐊𝐃𝟓𝐚𝐬Log𝐚𝐬(ConFr𝐊𝐃𝟓)𝐊𝐃𝟓,\mathbf{KD5}^{\mathbf{as}}\subseteq\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}\mathbf{KD5})\subseteq\mathbf{KD5}, where the latter inclusion also follows from Proposition 3.13.

Now we use our method to find 𝐊𝐃𝟒𝟓𝐚𝐬,𝐊𝟓𝐁𝐚𝐬,𝐒𝟓𝐚𝐬.\mathbf{KD45}^{\mathbf{as}},\,\mathbf{K5B}^{\mathbf{as}},\,\mathbf{S5}^{\mathbf{as}}. We follow mostly the same strategy, so we omit some technical steps.

Theorem 3.17.

𝐊𝐃𝟒𝟓𝐚𝐬=𝐊𝐃𝟒𝟓.\mathbf{KD45}^{\mathbf{as}}=\mathbf{KD45}.

Proof 3.18.

Let φ𝐊𝐃𝟒𝟓,\varphi\not\in\mathbf{KD45}, then G⊧̸φG\not\models\varphi for some finite point-generated frame G𝐊𝐃𝟒𝟓G\models\mathbf{KD45}. Let us show that F^n(ConFr𝐊𝐃𝟒𝟓)G\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45})\twoheadrightarrow G a.a.s. Notice that a Euclidean frame F=(X,R)F=(X,R) is serial, transitive and connected iff UFU_{F}\neq\varnothing and R=X×UFR=X\times U_{F}. Then such frame is uniquely determined by its cluster UFU_{F}\neq\varnothing. Therefore

|nConFr𝐊𝐃𝟒𝟓|=|{U[n]:U}|=2n1.|\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{KD45}|=\left|\{U\subseteq[n]:\>U\neq\varnothing\}\right|=2^{n}-1.

For any fixed rωr\in\omega we have r<|UF^n(ConFr𝐊𝐃𝟒𝟓)|<nr<|U_{\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45})}|<n a.a.s. Indeed, for any mnm\leq n there are (nm)\binom{n}{m} choices for a cluster of size mm in n,n, so

P(|UF^n(ConFr𝐊𝐃𝟒𝟓)|r)=m=1r(nm)2n1=O(nr)2n10,\displaystyle\operatorname{P}(|U_{\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45})}|\leq r)=\frac{\sum_{m=1}^{r}\binom{n}{m}}{2^{n}-1}=\frac{O(n^{r})}{2^{n}-1}\to 0, n,\displaystyle n\to\infty,
P(|UF^n(ConFr𝐊𝐃𝟒𝟓)|=n)=12n10,\displaystyle\operatorname{P}(|U_{\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45})}|=n)=\frac{1}{2^{n}-1}\to 0, n.\displaystyle n\to\infty.

There a.a.s. exists a p-morphism from some generated subframe of F^n(ConFr𝐊𝐃𝟒𝟓)\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45}) to G,G, so P(F^n(ConFr𝐊𝐃𝟒𝟓)φ)0\operatorname{P}(\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45})\models\varphi)\to 0 as n.n\to\infty.

Since φ𝐊𝐃𝟒𝟓\varphi\not\in\mathbf{KD45} was arbitrary, Log𝐚𝐬(ConFr𝐊𝐃𝟒𝟓)𝐊𝐃𝟒𝟓.\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45})\subseteq\mathbf{KD45}.

By Theorem 2.9 we have 𝐊𝐃𝟒𝟓𝐚𝐬Log𝐚𝐬(ConFr𝐊𝐃𝟒𝟓)𝐊𝐃𝟒𝟓.\mathbf{KD45}^{\mathbf{as}}\subseteq\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}\mathbf{KD45})\subseteq\mathbf{KD45}. The converse inclusion follows from Theorem 1.1.

Theorem 3.19.

𝐊𝟓𝐁𝐚𝐬=𝐊𝟓𝐁.\mathbf{K5B}^{\mathbf{as}}=\mathbf{K5B}.

Proof 3.20.

Notice that 𝐊𝟓𝐁\mathbf{K5B} is the logic of its finite point-generated frames, which are exactly the finite clusters and the irreflexive singletons. If φ𝐊𝟓𝐁\varphi\not\in\mathbf{K5B} for some formula φ,\varphi, then either ([r],[r]×[r])⊧̸φ([r],\,[r]\times[r])\not\models\varphi for some rωr\in\omega, or ({a},)⊧̸φ.(\{a\},\varnothing)\not\models\varphi.

Let rωr\in\omega and ([r],[r]×[r])⊧̸φ([r],\,[r]\times[r])\not\models\varphi. The connected components of a 𝐊𝟓𝐁\mathbf{K5B}-frame are clusters and irreflexive singletons. Then F^n(𝐊𝟓𝐁)\hat{F}_{n}(\mathbf{K5B}) a.a.s. contains a cluster of size greater than rr by Proposition 2.6, so there is a p-morphism from a generated subframe of F^n(ConFr𝐊𝟓𝐁))\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{K5B})) to ([r],[r]×[r])([r],[r]\times[r]), so P(F^n(ConFr𝐊𝟓𝐁)φ)0\operatorname{P}(\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{K5B})\models\varphi)\to 0 as n.n\to\infty.

Next we consider a formula φ\varphi such that ({a},⊧̸φ)(\{a\},\varnothing\not\models\varphi). A 𝐊𝟓𝐁\mathbf{K5B}-frame with set of states nn is uniquely determined by a subset E[n]E\subseteq[n] that consists of the irreflexive singletons and an equivalence relation on nE.n\setminus E. Then

|n𝐊𝟓𝐁|=E[n]Bn|E|=m=0n(nm)Bnm=m=0n(nm)Bn=Bn+1.\left|\mathcal{F}_{n}\cap\mathbf{K5B}\right|=\sum_{E\subseteq[n]}B_{n-|E|}=\sum_{m=0}^{n}\binom{n}{m}B_{n-m}=\sum_{m=0}^{n}\binom{n}{m}B_{n}=B_{n+1}.

A 𝐊𝟓𝐁\mathbf{K5B}-frame has no irreflexive singletons iff its relation is an equivalence relation, so there are exactly BnB_{n} frames without irreflexive singletons in nFr𝐊𝟓𝐁.{\mathcal{F}_{n}\cap\operatorname{Fr}\mathbf{K5B}}. Then by (4)

P(F^n(𝐊𝟓𝐁) has no irreflexive singletons)=Bn|n𝐊𝟓𝐁|=BnBn+10\operatorname{P}(\hat{F}_{n}(\mathbf{K5B})\text{ has no irreflexive singletons})=\frac{B_{n}}{\left|\mathcal{F}_{n}\cap\mathbf{K5B}\right|}=\frac{B_{n}}{B_{n+1}}\to 0

as n.n\to\infty. Then F^n(ConFr𝐊𝟓𝐁)\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{K5B}) contains a generated subframe isomorphic to ({a},)(\{a\},\varnothing) a.a.s., so P(F^n(ConFr𝐊𝟓𝐁)φ)0\operatorname{P}(\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{K5B})\models\varphi)\to 0 as n.n\to\infty.

By Theorem 2.6, 𝐊𝟓𝐁𝐚𝐬Log𝐚𝐬(ConFr𝐊𝟓𝐁)𝐊𝟓𝐁.\mathbf{K5B}^{\mathbf{as}}\subseteq\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}\mathbf{K5B})\subseteq\mathbf{K5B}. The converse inclusion is true by Theorem 1.1.

Theorem 3.21.

𝐒𝟓𝐚𝐬=𝐒𝟓.\mathbf{S5}^{\mathbf{as}}=\mathbf{S5}.

Proof 3.22.

Observe that nConFr𝐒𝟓\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{S5} consists of one frame (n,n×n).(n,\,n\times n). The finite point-generated frames of 𝐒𝟓\mathbf{S5} are the finite clusters, which are p-morphic images of (n,n×n)(n,\,n\times n) for nn sufficiently large. Then P(F^n(ConFr𝐒𝟓)φ)0\operatorname{P}(\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{S5})\models\varphi)\to 0 for any φ𝐒𝟓.\varphi\not\in\mathbf{S5}. Then the statement of this theorem follows from Theorem 1.1 and Theorem 2.9.

4 Transitive frames

In this section we discuss the logics of almost sure validities in the finite frames of 𝐆𝐋.3\mathbf{GL.3} and 𝐆𝐫𝐳.3.\mathbf{Grz.3}.

Definition 4.1.

A frame F=(X,R)F=(X,\,R) is an inverse tree if there is a unique element a0a_{0} such that Rout(a0)=R_{\mathrm{out}}(a_{0})=\varnothing, and for any aX{a0},a\in X\setminus\{a_{0}\}, |Rout(a)|=1|R_{\mathrm{out}}(a)|=1 and aRa0.aR^{*}a_{0}. We denote TnT_{n} the set of all inverse trees over the set of states [n].[n].

Definition 4.2.

The distance between states aa and bb of a frame F=(X,R)F=(X,\,R) is the number dR(a,b)=min{nω:aRnb}d_{R}(a,\,b)=\min\{n\in\omega:\>aR^{n}b\} for any a,b[n].a,\,b\in[n].

Definition 4.3.

The height of a finite inverse tree is the maximum of the distance between the states in that tree.

Definition 4.4.

Given a transitive Noetherian relation RR on a set X,X, the transitive reduction of RR is defined as R=RR2.R^{-}=R\setminus R^{2}.

It is straightforward to see that for a transitive Noetherian relation R,R, its transitive reduction RR^{-} is the smallest relation on XX such that R(R){R\subseteq(R^{-})^{*}} [AGU72]. If (X,R)ConFr𝐆𝐋.3(X,\,R)\in\operatorname{Con}\operatorname{Fr}\mathbf{GL.3}, then R=(R)+R=(R^{-})^{+}. Similarly, if (X,R)(X,\,R) is a frame in ConFr𝐆𝐫𝐳.3\operatorname{Con}\operatorname{Fr}\mathbf{Grz.3}, then R=(R).R=(R^{-})^{*}. Therefore the mapping RRR\mapsto R^{-} has an inverse on ConFr𝐆𝐋.3\operatorname{Con}\operatorname{Fr}\mathbf{GL.3} and on ConFr𝐆𝐫𝐳.3.\operatorname{Con}\operatorname{Fr}\mathbf{Grz.3}.

Proposition 4.5.

For any nω,n\in\omega, the sets of frames Tn,T_{n}, nConFr𝐆𝐋.3,\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{GL.3}, and nConFr𝐆𝐫𝐳.3,\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{Grz.3}, are bijective. Moreover, the bijection preserves the distance between any pair of states.

Proof 4.6.

Follows directly from the discussion above.

Proposition 4.7.

Consider the random inverse tree with uniform distribution on TnT_{n} and let h^n\hat{h}_{n} be its height.

There exists a constant q>0q>0 such that for any fixed rω,{r\in\omega},

limnP(h^n>r)q.\lim_{n\to\infty}\operatorname{P}(\hat{h}_{n}>r)\geq q.
Proof 4.8.

The asymptotic expressions for the expected value and the variance of hnh_{n} are given in [RS67]:

E(h^n)2πn;Var(h^n)π(π3)n3,n.\operatorname{E}(\hat{h}_{n})\sim\sqrt{2\pi n};\quad\operatorname{Var}(\hat{h}_{n})\sim\frac{\pi(\pi-3)n}{3},\quad n\to\infty.

By Chebyshev’s inequality [Gut13, Theorem 1.4],

P(h^nr)P(|h^nEh^n|Eh^nr)Var(h^n)(Eh^nr)2π(π3)n3(2πnr)2π(π3)3(2πrn)2π36<1,n.\operatorname{P}(\hat{h}_{n}\leq r)\leq\operatorname{P}\left(|\hat{h}_{n}-\operatorname{E}\hat{h}_{n}|\geq\operatorname{E}\hat{h}_{n}-r\right)\leq\frac{\operatorname{Var}(\hat{h}_{n})}{(\operatorname{E}\hat{h}_{n}-r)^{2}}\\ \sim\frac{\pi(\pi-3)n}{3(\sqrt{2\pi n}-r)^{2}}\sim\frac{\pi(\pi-3)}{3(\sqrt{2\pi}-\frac{r}{\sqrt{n}})^{2}}\sim\frac{\pi-3}{6}<1,\quad n\to\infty.

Let q=1π36>0,q=1-\frac{\pi-3}{6}>0, then limnP(h^n>r)=1limnP(h^nr)q.\lim_{n\to\infty}\operatorname{P}(\hat{h}_{n}>r)=1-\lim_{n\to\infty}\operatorname{P}(\hat{h}_{n}\leq r)\geq q.

Theorem 4.9.

𝐆𝐋.3𝐚𝐬=𝐆𝐋.3.\mathbf{GL.3}^{\mathbf{as}}=\mathbf{GL.3}.

Proof 4.10.

Let us recall that the logic 𝐆𝐋.3\mathbf{GL.3} is complete w.r.t. the finite irreflexive chains. Let φ𝐆𝐋.3{\varphi\not\in\mathbf{GL.3}}. Then φ\varphi is falsified in some irreflexive chain FF of a finite cardinality r.r.

Let R^n\hat{R}_{n} denote the relation of F^n(ConFr𝐆𝐋.3).\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{GL.3}). By Proposition 4.5 and Proposition 4.7,

limnP(a[n]:|R^n(a)|r)=limnP(h^n>r)q.\lim_{n\to\infty}\operatorname{P}\left(\exists a\in[n]:\>|\hat{R}_{n}(a)|\geq r\right)=\lim_{n\to\infty}\operatorname{P}(\hat{h}_{n}>r)\geq q.

With an asymptotic probability q>0q>0 there exists a[n]a\in[n] such that |R^n(a)|r.{|\hat{R}_{n}(a)|\geq r}. In this case F^n(𝐆𝐋.3)a\hat{F}_{n}(\mathbf{GL.3}){\uparrow}a is an irreflexive chain of cardinality at least rr, so FF is isomorphic to some generated subframe of F^n(𝐆𝐋.3)a.\hat{F}_{n}(\mathbf{GL.3}){\uparrow}a.

Then lim supnP(F^n(ConFr𝐆𝐋.3)φ)q<1.\limsup_{n\to\infty}\operatorname{P}({\hat{F}_{n}(\operatorname{Con}\operatorname{Fr}\mathbf{GL.3})\models\varphi})\leq q<1.

Since φ𝐆𝐋.3\varphi\not\in\mathbf{GL.3} was arbitrary, Log𝐚𝐬(ConFr𝐆𝐋.3)=𝐆𝐋.3.\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}\mathbf{GL.3})=\mathbf{GL.3}.

Since ([n],<)([n],<) is a connected 𝐆𝐋.3\mathbf{GL.3}-frame, nConFr𝐆𝐋.3\mathcal{F}_{n}\cap\operatorname{Con}\operatorname{Fr}\mathbf{GL.3}\neq\varnothing for all nωn\in\omega, thus by Theorem 2.9 𝐆𝐋.3𝐚𝐬𝐆𝐋.3\mathbf{GL.3}^{\mathbf{as}}\subseteq\mathbf{GL.3}. The converse is true by Theorem 1.1.

Theorem 4.11.

𝐆𝐫𝐳.3𝐚𝐬=𝐆𝐫𝐳.3.\mathbf{Grz.3}^{\mathbf{as}}=\mathbf{Grz.3}.

Proof 4.12.

Analogous to the previous theorem.

5 Results and discussion

We developed several general results about the almost sure validities in random Kripke frames. Theorem 1.1 states that the almost sure validities in the random frame F^(𝒞)\hat{F}(\mathcal{C}) is a normal modal logic that extends 𝒞\mathcal{C} for any class of frames 𝒞\mathcal{C}. Theorem 2.9, which applies for many commonly studied modal logics, establishes an important inclusion L𝐚𝐬Log𝐚𝐬(ConFrL)L^{\mathbf{as}}\subseteq\operatorname{Log}^{\mathbf{as}}(\operatorname{Con}\operatorname{Fr}L).

We established the axiomatizations for the considered classes of frames:

𝐊𝐃𝟓𝐚𝐬=𝐊𝐃𝟓;\displaystyle\mathbf{KD5}^{\mathbf{as}}=\mathbf{KD5}; 𝐊𝐃𝟒𝟓𝐚𝐬=𝐊𝐃𝟒𝟓;\displaystyle\mathbf{KD45}^{\mathbf{as}}=\mathbf{KD45}; 𝐊𝟓𝐁𝐚𝐬=𝐊𝟓𝐁;\displaystyle\mathbf{K5B}^{\mathbf{as}}=\mathbf{K5B};
𝐒𝟓𝐚𝐬=𝐒𝟓;\displaystyle\mathbf{S5}^{\mathbf{as}}=\mathbf{S5}; 𝐆𝐫𝐳.3𝐚𝐬=𝐆𝐫𝐳.3;\displaystyle\mathbf{Grz.3}^{\mathbf{as}}=\mathbf{Grz.3}; 𝐆𝐋.3𝐚𝐬=𝐆𝐋.3.\displaystyle\mathbf{GL.3}^{\mathbf{as}}=\mathbf{GL.3}.

Interestingly, all these logics share some desirable properties, such as finite axiomatization, finite model property, decidability, etc. This stands in contrast to the known results on 𝐊𝐚𝐬\mathbf{K}^{\mathbf{as}} [Gor20] and GL𝐚𝐬\textbf{GL}^{\mathbf{as}} [Ver21] that imply that these logics lack the finite axiomatizability.

Our results on 𝐊𝐃𝟓,𝐊𝐃𝟒𝟓,𝐊𝟓𝐁,𝐒𝟓,𝐆𝐫𝐳.3,𝐆𝐋.3\mathbf{KD5},\,\mathbf{KD45},\,\mathbf{K5B},\,\mathbf{S5},\,\mathbf{Grz.3},\,\mathbf{GL.3} provide examples of logics that are equal to their ‘almost sure’ counterparts. Finding a criterion that characterizes the logics with this property is an interesting direction for future research.

The computational method we use in this paper seems to be able to yield more general results, such as a classification of logics of almost sure validities of the frame classes of all logics above 𝐊𝟓.\mathbf{K5}. We also conjecture that many of such logics obey the zero-one law.

References

  • [AGU72] A. V. Aho, M. R. Garey, and J. D. Ullman, The transitive reduction of a directed graph, SIAM Journal on Computing 1 (1972), no. 2, 131–137.
  • [BdRV01] Patrick Blackburn, Maarten de Rijke, and Yde Venema, Modal logic, Studia Logica 76 (2001), no. 1, 142–148.
  • [CY94] Beifang Chen and Yeong-Nan Yeh, Some explanations of Dobiński’s formula, Studies in Applied Mathematics 92 (1994), 191–199.
  • [dB58] N.G. de Bruijn, Asymptotic methods in analysis, Bibliotheca mathematica, North-Holland Publishing Co., 1958.
  • [Fag76] Ronald Fagin, Probabilities on finite models, Journal of Symbolic Logic 41 (1976), no. 1, 50–58.
  • [Gai64] Haim Gaifman, Concerning measures in first order calculi, Israel Journal of Mathematics 2 (1964), 1–18.
  • [GK03] Valentin Goranko and Bruce Kapron, The modal logic of the countable random frame, Archive for Mathematical Logic 42 (2003), no. 3, 221–243.
  • [GKLT69] Yu. V. Glebskii, D. I. Kogan, M. I. Liogon’kii, and V. A. Talanov, Range and degree of realizability of formulas in the restricted predicate calculus, Cybernetics 5 (1969), 142–154.
  • [Gor20] V. Goranko, The modal logic of almost sure frame validities in the finite, AiML, 2020.
  • [Gut13] A. Gut, Probability: A graduate course, Springer Texts in Statistics, Springer, 2013.
  • [HK94] Joseph Y. Halpern and Bruce Kapron, Zero-one laws for modal logic, Annals of Pure and Applied Logic 69 (1994), no. 2-3, 157–193.
  • [LB02] J.-M Le Bars, The 0-1 law fails for frame satisfiability of propositional modal logic, Proceedings 17th Annual IEEE Symposium on Logic in Computer Science, 02 2002, pp. 225 – 234.
  • [MMW58] F.L. Miksa, L. Moser, and M. Wyman, Restricted partitions of finite sets, Canadian Mathematical Bulletin 1 (1958), no. 2, 87–96.
  • [RS67] Alfréd Rényi and Gy. Szekeres, On the height of trees, Journal of The Australian Mathematical Society 7 (1967), 497–507.
  • [SF14] J. H. Spencer and L. Florescu, Asymptopia, Student mathematical library, vol. 71, American Mathematical Society, 2014.
  • [Ver18] Rineke Verbrugge, Zero-one laws with respect to models of provability logic and two grzegorczyk logics, Advances in Modal Logic 2018: Accepted Short Papers, 08 2018, pp. 115–120.
  • [Ver21]  , Zero-one laws for provability logic: Axiomatizing validity in almost all models and almost all frames, 2021 36th Annual ACM/IEEE Symposium on Logic in Computer Science, LICS 2021, Proceedings - Symposium on Logic in Computer Science, IEEE Xplore, 06 2021.

6 Appendix

6.1 Proof of (4)

The Bell numbers satisfy Dobiński’s formula [CY94]

Bn=ek=01knk!,kω.B_{n}=e{}^{-1}\sum_{k=0}^{\infty}\frac{k^{n}}{k!},\quad k\in\omega. (15)

Let X^\hat{X} be a random variable of with the standard Poisson distribution:

P(X^=k)=e1k!.\operatorname{P}(\hat{X}=k)=\frac{e{}^{-1}}{k!}.

Then by (15) the nthn^{th} moment of X^\hat{X} is Bn:B_{n}:

E(X^n)=k=0kne1k!=Bn\operatorname{E}(\hat{X}^{n})=\sum_{k=0}^{\infty}k^{n}\frac{e{}^{-1}}{k!}=B_{n}

Since the function φ:[0,)[0,),φ(x)=xn+1n\varphi:\>[0,\infty)\to[0,\infty),\,\varphi(x)=x^{\frac{n+1}{n}} is convex, Jensen’s inequality [Gut13, Theorem 5.1] holds: φ(EX^n)E(φ(X^n))\varphi(\operatorname{E}\hat{X}^{n})\leq\operatorname{E}(\varphi(\hat{X}^{n})), so

Bnn+1nE(X^n+1n)=k=0(kn)n+1ne1k!=ek=0n1kn+1k!=Bn+1,{B_{n}}^{\frac{n+1}{n}}\leq\operatorname{E}\left(\hat{X}^{\frac{n+1}{n}}\right)=\sum_{k=0}^{\infty}\left(k^{n}\right)^{\frac{n+1}{n}}\frac{e{}^{-1}}{k!}=e{}^{-1}\sum_{k=0}^{n}\frac{k^{n+1}}{k!}=B_{n+1},

therefore

BnBn+1BnBnn+1n=Bn1n.\frac{B_{n}}{B_{n+1}}\leq\frac{B_{n}}{B_{n}^{\frac{n+1}{n}}}=B_{n}^{-\frac{1}{n}}.

Recall that by (3Bn=enlnn(1+o(1))B_{n}=e^{n\ln n(1+o(1))}, so Bn1n=elnn(1+o(1))0B_{n}^{-\frac{1}{n}}=e^{-\ln n(1+o(1))}\to 0 as n.n\to\infty. We conclude that

limnBnBn+1=0.\lim_{n\to\infty}\frac{B_{n}}{B_{n+1}}=0.

6.2 Proof of (5)

An asymptotic expression of Gn,rG_{n,r} for fixed rωr\in\omega and nn\to\infty is provided in [MMW58]:

Gn,r(nRn,r)nr12exp(nRn,r+Rn,rrr!n1),G_{n,r}\sim\left(\frac{n}{R_{n,r}}\right)^{n}r^{-\frac{1}{2}}\exp\left(\frac{n}{R_{n,r}}+\frac{R_{n,r}{}^{r}}{r!}-n-1\right), (16)

where Rn,rR_{n,\,r} is the positive root of the equation

Rn,r+Rn,r21!+Rn,r32!++Rn,rr(r1)!=n.R_{n,r}+\frac{R_{n,r}{}^{2}}{1!}+\frac{R_{n,r}{}^{3}}{2!}+\ldots+\frac{R_{n,r}{}^{r}}{(r-1)!}=n. (17)

Note that Rn,rR_{n,r}\to\infty as n.n\to\infty. Then Rn,rk=o(Rn,rr){R_{n,r}}^{k}=o(R_{n,r}^{r}) for any k<r,k<r, so the equation (17) yields

Rn,rr(r1)!(1+o(1))=n,\frac{R_{n,r}{}^{r}}{(r-1)!}(1+o(1))=n,

therefore

Rn,r=(n(r1)!)1r(1+o(1).R_{n,r}=\left(n(r-1)!\right)^{\frac{1}{r}}(1+o(1).

Then we may estimate lnRn,r=1rlnn(1+o(1));nRn,r+Rn,rrr!n1=O(n),\ln R_{n,r}=\frac{1}{r}\ln n(1+o(1));\;\frac{n}{R_{n,r}}+\frac{R_{n,r}{}^{r}}{r!}-n-1=O(n), so (16) implies

lnGn,r=nlnn1rnlnn+O(n).\ln G_{n,r}=n\ln n-\frac{1}{r}n\ln n+O(n).

Therefore for any kωk\in\omega, by the asymptotic expression (3) for BnB_{n} we get:

ln(Gn,r2knBn)=nlnn1rnlnn+O(n)+knln2nlnn(1+o(1))=1rnlnn(1+o(1));\ln\left(\frac{G_{n,r}2^{kn}}{B_{n}}\right)=n\ln n-\frac{1}{r}n\ln n+O(n)+kn\ln 2-n\ln n(1+o(1))\\ =-\frac{1}{r}n\ln n(1+o(1))\to-\infty;

then by exponentiating we get the desired estimation:

Gn,r2knBn0,n.\frac{G_{n,r}2^{kn}}{B_{n}}\to 0,\quad n\to\infty.